1: \documentclass[12pt,preprint]{aastex}
2: %\usepackage{graphics}
3:
4: \newcommand{\bm}[1]{\mbox{\boldmath{$#1$}}}
5:
6: \newcommand{\at}{{\rm Athena}}
7: \newcommand{\del}{{\bf \nabla}}
8:
9: \newcommand{\etot}{E_{\rm tot}}
10: \newcommand{\md}{\dot{M}}
11: \newcommand{\kd}{\dot{K}}
12: \newcommand{\td}{\dot{T}}
13: \newcommand{\gd}{\dot{G}}
14: \newcommand{\qk}{\dot{Q}_{\rm k}}
15: \newcommand{\qm}{\dot{Q}_{\rm m}}
16: \newcommand{\ein}{E_{\rm in}}
17: \newcommand{\alf}{{\rm Alfv\acute{e}n}}
18: \newcommand{\pr}{P_m}
19:
20: \newcommand{\maxwell}{\langle\langle-B_xB_y\rangle\rangle}
21: \newcommand{\reynolds}{\langle\langle \rho v_x\delta v_y\rangle\rangle}
22: \newcommand{\magnetic}{\langle\langle B^2/2 \rangle\rangle}
23: \newcommand{\magx}{\langle\langle B_x^2/2 \rangle\rangle}
24: \newcommand{\magy}{\langle\langle B_y^2/2 \rangle\rangle}
25: \newcommand{\magz}{\langle\langle B_z^2/2 \rangle\rangle}
26: \newcommand{\kinetic}{\langle\langle \rho v^2/2\rangle\rangle}
27: \newcommand{\pertkinetic}{\langle\langle \rho \delta v^2/2\rangle\rangle}
28: \newcommand{\kinx}{\langle\langle \rho v_x^2/2\rangle\rangle}
29: \newcommand{\pertkiny}{\langle\langle \rho \delta v_y^2/2\rangle\rangle}
30: \newcommand{\kinz}{\langle\langle \rho v_z^2/2\rangle\rangle}
31:
32: \newcommand{\bstar}{\widetilde{{\bm B}}^{\bm \ast}({\bm k})}
33: \newcommand{\bfour}{\widetilde{{\bm B}}({\bm k})}
34: \newcommand{\vstar}{\widetilde{\sqrt{\rho}{\bm v}^{\bm \ast}}({\bm k})}
35: \newcommand{\vfour}{\widetilde{\sqrt{\rho}{\bm v}}({\bm k})}
36: \newcommand{\vpfour}{\widetilde{\sqrt{\rho}{\bm \delta}{\bm v}}({\bm k})}
37: \newcommand{\imi}{{\it i}}
38: \newcommand{\vsh}{V_{{\rm sh}}}
39: \newcommand{\vt}{{\bm v_{\rm t}}}
40: \newcommand{\dv}{{\rm d^3}{\bm x}}
41: \newcommand{\sr}{\sqrt{\rho}}
42: \newcommand{\fterm}{{\rm e}^{-\imi {\bm k} \cdot {\bm x}} \dv}
43: \newcommand{\etaeff}{\eta_{\rm eff}}
44: \newcommand{\nueff}{\nu_{\rm eff}}
45: \newcommand{\reeff}{Re_{\rm eff}}
46: \newcommand{\rmeff}{Rm_{\rm eff}}
47: \newcommand{\pmeff}{P_{m,{\rm eff}}}
48:
49: \newcommand{\tbb}{T_{bb}}
50: \newcommand{\tdivv}{T_{{\rm div}v}}
51: \newcommand{\tbv}{T_{bv}}
52: \newcommand{\tvv}{T_{vv}}
53: \newcommand{\tcomp}{T_{\rm comp}}
54: \newcommand{\tvb}{T_{vb}}
55: \newcommand{\tpress}{T_{\rm press}}
56: \newcommand{\tcor}{T_{\rm cor}}
57: \newcommand{\tphi}{T_\phi}
58: \newcommand{\dmag}{D_{\rm mag}}
59: \newcommand{\dkin}{D_{\rm kin}}
60: \newcommand{\tnu}{T_\nu}
61: \newcommand{\teta}{T_\eta}
62:
63: \newcommand{\zonesix}{{\rm SZ16}}
64: \newcommand{\zthree}{{\rm SZ32}}
65: \newcommand{\zsix}{{\rm SZ64}}
66: \newcommand{\zone}{{\rm SZ128}}
67: \newcommand{\nonesix}{{\rm NZ16}}
68: \newcommand{\nthree}{{\rm NZ32}}
69: \newcommand{\nsix}{{\rm NZ64}}
70: \newcommand{\none}{{\rm NZ128}}
71:
72: \newcommand{\zdonesix}{{\rm SZD16}}
73: \newcommand{\zdthree}{{\rm SZD32}}
74: \newcommand{\zdsix}{{\rm SZD64}}
75: \newcommand{\zdone}{{\rm SZD128}}
76: \newcommand{\ndonesix}{{\rm NZD16}}
77: \newcommand{\ndthree}{{\rm NZD32}}
78: \newcommand{\ndsix}{{\rm NZD64}}
79: \newcommand{\ndone}{{\rm NZD128}}
80:
81: \newcommand{\narone}{{\rm NZAR1}}
82: \newcommand{\nartwo}{{\rm NZAR2}}
83: \newcommand{\narthree}{{\rm NZAR3}}
84: \newcommand{\narfour}{{\rm NZAR4}}
85: \newcommand{\narfive}{{\rm NZAR5}}
86: \newcommand{\narsix}{{\rm NZAR6}}
87:
88: \begin{document}
89:
90: \title{Simulations of Magnetorotational Turbulence with a Higher-Order Godunov Scheme}
91:
92: \author{Jacob B. Simon, John F. Hawley, Kris Beckwith}
93: \affil{Department of Astronomy \\ University of Virginia \\ P.O. Box 400325 \\
94: Charlottesville, VA 22904-4325}
95:
96: \begin{abstract}
97: We apply a new, second-order Godunov code, Athena, to studies of the
98: magnetorotational instability (MRI) using unstratified shearing box
99: simulations with a uniform net vertical field and a sinusoidally
100: varying zero net vertical field. The Athena results agree well with
101: similar studies that used different numerical algorithms, including
102: the observation that the turbulent energy decreases with increasing
103: resolution in the zero net field model. We conduct analyses to study the
104: flow of energy from differential rotation to turbulent fluctuations
105: to thermalization. A study of the time-correlation between the rates
106: of change of different volume-averaged energy components shows that
107: energy injected into turbulent fluctuations dissipates on a timescale
108: of $\Omega^{-1}$, where $\Omega$ is the orbital frequency of the local
109: domain. Magnetic dissipation dominates over kinetic dissipation, although
110: not by as great a factor as the ratio of magnetic to kinetic energy.
111: We Fourier-transform the magnetic and kinetic energy evolution equations
112: and, using the assumption that the time-averaged energies are constant,
113: determine the level of numerical dissipation as a function of length
114: scale and resolution. By modeling numerical dissipation as if it were
115: physical in origin, we characterize numerical resistivity and viscosity
116: in terms of effective Reynolds and Prandtl numbers. The resulting
117: effective magnetic Prandtl number is $\sim 2$, independent of resolution
118: or initial field geometry. MRI simulations with effective Reynolds and
119: Prandtl numbers determined by numerical dissipation are not equivalent to
120: those where these numbers are set by physical resistivity and viscosity.
121: These results serve, then, as a baseline for future shearing box studies
122: where dissipation is controlled by the inclusion of explicit viscosity
123: and resistivity.
124: \end{abstract}
125:
126: \keywords{Black holes - magnetohydrodynamics - stars:accretion}
127:
128: \section{Introduction}
129: \label{introduction}
130:
131: The process of accretion powers a wide range of astrophysical systems,
132: from protostars to quasars. In accretion disks, gravitational energy is
133: converted into other forms including bulk outflows, heat, and radiation.
134: In the traditional time-stationary thin disk model of \cite{shak73}, the
135: $r,\phi$ component of the stress, $\tau_{r\phi}$, is proportional to the
136: local pressure, $\tau_{r\phi} = \alpha P$. The $\alpha$ model assumes
137: that the accretion energy is deposited as heat locally and radiated
138: rapidly, providing a relation between disk emissivity and accretion rate.
139: While the $\alpha$ model has proven valuable in interpreting many aspects
140: of accretion systems, advancing beyond it will require a more detailed
141: understanding of the stress that produces angular momentum
142: transport as well as the physical processes involved in the subsequent
143: thermalization and radiation of the orbital energy released by those
144: stresses.
145:
146: It is now understood that magnetohydrodynamic (MHD) turbulence generated
147: by the magnetorotational instability (MRI) \cite[]{balb91,balb98} produces
148: significant Maxwell stresses, $-B_rB_\phi/4\pi$, and Reynolds stresses,
149: $\rho \delta v_r \delta v_\phi$, that account for transport within
150: accretion disks. The absence of an analytic theory for MHD turbulence,
151: however, means that direct numerical simulations play an essential role
152: in investigating accretion physics. In this regard, local simulations,
153: which reduce the problem to the simplest form that can sustain MRI-driven
154: turbulence, have proven very useful. The ``shearing box'' model is
155: a representation of a small patch of the disk constructed by boosting to
156: a local co-rotating Cartesian frame that ignores geometric curvature but
157: retains all rotational forces. MRI shearing box simulations were
158: introduced by \cite{haw95} and have been extensively used since then
159: both without \cite[e.g.,][]{haw96,balb98} and with vertical stratification
160: \cite[e.g.,][]{bran95,stone96,hir06}.
161:
162:
163:
164: Shearing box simulations can investigate several key questions including
165: the functional dependence of the stress on disk properties
166: and the turbulent energy flow that leads to dissipation as heat.
167: These simulations have made it increasingly clear, for example, that
168: the basic $\alpha$ stress parameterization is not only too simplistic,
169: it is actually misleading. Shearing boxes have provided ample evidence
170: that stress is {\it not} determined by pressure, at least in the usual
171: manner of the $\alpha$ disk \cite[][]{haw95, sano04}. Early studies showed
172: instead that stress is (in some cases) proportional to the {\it magnetic} pressure,
173: but the magnetic energy is not itself directly determined by the gas and
174: radiation pressure. \cite{black08} recently reviewed a large number of
175: shearing box results and found that this result holds across the full
176: ensemble of simulations with only small differences in the constant of
177: proportionality from one run to another. The implications of these results are
178: significant. For example, recent local simulations using stratified
179: shearing boxes and radiation transport \cite[][]{bla07, kro07}
180: have found no evidence of the thermal instability long believed to be
181: present in radiation-pressure supported $\alpha$ disks.
182:
183:
184:
185: If stress is proportional to magnetic rather than total pressure,
186: what determines the magnetic pressure in a disk? Apart from the
187: expectation that the field will remain subthermal, this remains uncertain.
188: The simplest shearing box simulations using ideal MHD have a limited
189: range of significant parameters; this is both a strength and a weakness
190: of that model. The magnetic energy in the saturated state could depend
191: upon such factors as box size, the amplitude and geometry of the imposed initial
192: magnetic field, and the ratio of the gas pressure to magnetic pressure
193: (the plasma $\beta$ value). \cite{haw95} and \cite{haw96} studied the
194: effect of initial magnetic field topology on the resulting stress and
195: found that although the MRI leads to turbulence regardless of the initial
196: field, simulations that had an imposed net vertical field produce higher
197: turbulence levels than an imposed toroidal field or a simulation that
198: began with zero net magnetic flux within the domain. \cite{haw95} found
199: that the total magnetic energy and the resulting stress in the saturated
200: turbulent state was a function of the initial plasma $\beta$ with a
201: uniform vertical field, namely that larger $\beta$ (i.e., weaker fields)
202: leads to smaller saturation levels. Other initial field configurations
203: do not yield so direct a correlation between background field strength
204: and saturation. Many simulations have failed to find any noticeable
205: correlation between mean turbulent magnetic energy and the gas pressure.
206: A comprehensive parameter study by \cite{sano04} observed at best only
207: a very weak gas pressure dependence.
208:
209:
210:
211:
212: Since the mean magnetic energy at saturation is presumably a balance
213: between continued driving by the MRI and loss due to magnetic dissipation
214: and reconnection, there has been interest in going beyond ideal MHD to
215: include explicit physical dissipation in the form of kinematic viscosity,
216: $\nu$, and Ohmic resistivity, $\eta$. Both of these properties have been shown to be
217: important in determining the mean energies and stresses in MRI turbulence.
218: Simulations by \cite{haw96}, \cite{sano98}, \cite{flem00}, \cite{sano01},
219: \cite{zieg01}, and \cite{sano02b} have investigated the impact of a
220: nonzero $\eta$. The main result of these studies is that increasing
221: the resistivity leads to a decrease in turbulence, independent of the
222: initial field configuration. In zero net field models, the effect of
223: resistivity on the turbulence is larger than one might expect from the
224: linear MRI relation \cite[]{flem00}. On the other hand, \cite{haw96}
225: found that increasing the viscosity increased the magnetic energy in the
226: saturated state. Recent work has clarified the situation by demonstrating
227: a dependence of the saturation level on both $\eta$ and $\nu$ in terms
228: of the magnetic Prandtl number, $\pr = \nu/\eta$. In particular, the
229: level of angular momentum transport increases with increasing $\pr$ for
230: simulations initiated with a uniform as well as vanishing mean magnetic
231: field in the vertical direction \cite[]{from07b,lesur07}.
232:
233:
234:
235: Determining the stress levels in MRI turbulence is only one aspect of
236: the problem; another is exploring how that turbulence is dissipated
237: into heat. This question has direct relevance to phenomenological disk
238: models as well as observations. The $\alpha$ model
239: assumes that the accretion energy is deposited as heat locally and
240: rapidly, and \cite{balb99} showed that this property should hold for
241: the energetics of MHD turbulence as well.
242: In the simulations, we can determine the rate at which turbulent energy
243: is thermalized and the path that energy takes as it moves from the free
244: energy of the shear flow to turbulence and then to heat. Such issues
245: were briefly touched on by \cite{bran95} who found that the turbulent
246: magnetic energy was $\sim 6$ times greater than the perturbed kinetic
247: energy, but dissipational heating resulted from roughly equal contributions of
248: magnetic and kinetic energy dissipation. This result led them to suggest
249: that there was a net transfer of magnetic energy to turbulent kinetic energy.
250: \cite{sano01} studied
251: energy flow in the context of MRI channel modes, which are strong radial
252: streaming motions that result from the linear growth of the vertical
253: field MRI \cite[]{haw92,balb98}. Their work included Ohmic resistivity
254: (but not viscosity) and showed that resistive heating dominated the
255: thermalization of energy stored in these channel modes. Dissipational
256: heating also plays an important role in radiative effects and determining
257: disk structure, both of which may be observable properties of disks \cite[e.g.,][]{beck08}.
258:
259:
260:
261: In any study that depends on simulations, there remain
262: factors which cannot be overlooked: the effects due to numerics and
263: finite resolution. The majority of the results to-date were obtained
264: with numerical codes based on the finite-difference ZEUS algorithm
265: \cite[]{stone92a,stone92b}, carried out at relatively low resolution.
266: ZEUS is effectively first-order in asymptotic convergence, and in
267: its most widely used form, evolves the internal rather than the
268: total energy equation. There have been improvements in both the
269: available computational power, which makes higher resolutions and longer
270: evolution times possible, and in the algorithms for compressible MHD.
271: In this work, we will reexamine the properties of MHD turbulence in the
272: shearing box using a higher-order, Godunov scheme.
273:
274:
275:
276: The new code, $\at$, \cite[see][]{stone08} represents an improvement
277: over ZEUS in several ways including true second-order convergence,
278: increased effective resolution \cite[see][]{stone05}, accurate shock capturing,
279: and conservation of total energy. The energy-conserving properties
280: of $\at$ allow us to study energy flow and
281: dissipation within the shearing box in greater detail than allowed for
282: by the ZEUS algorithm. The version of $\at$ we use in this paper does
283: not include explicit resistivity or viscosity and instead relies on
284: numerical dissipation to thermalize the turbulent energy. Nevertheless,
285: this work will serve as a starting point for planned studies of nonideal
286: effects, including the influence of $\pr$ on the turbulence
287: \cite[][]{from07b, lesur07}. As an important
288: part of establishing a baseline of simulations, we will characterize
289: the numerical resistivity and viscosity of $\at$ for the shearing box
290: problem. To do so, we will follow the recent work of \cite{from07a} who
291: studied the numerical effects of ZEUS on the saturated state of MRI shearing
292: box simulations that begin with zero net field. They found that the
293: amplitude of the turbulence decreases with increasing resolution and
294: developed several useful diagnostics with which to quantify the effective
295: numerical resistivity and viscosity in the problem.
296:
297:
298:
299: The structure of the paper is as follows. In \S~\ref{method},
300: we describe the algorithm employed and our simulations. In
301: \S~\ref{general_properties}, we reexamine some of the
302: results from previous MRI studies and provide a comparison with these
303: studies. In \S~\ref{energy_fluctuations}, we present the first of two
304: diagnostics used to study turbulent energy flow and dissipation. The
305: second of these diagnostics is applied in \S~\ref{trans_funcs}. Finally,
306: we discuss our results and summarize our conclusions in \S~\ref{conclusions}.
307:
308: \section{Numerical Simulations}
309: \label{method}
310:
311: The code used for all of our simulations is $\at$, a second-order accurate
312: Godunov scheme for solving the equations of ideal MHD in conservative
313: form. The equations are solved using the dimensionally unsplit corner
314: transport upwind (CTU) method of \cite{col90} coupled with the third-order in space
315: piecewise parabolic method (PPM) of \cite{col84} and a constrained
316: transport (CT) algorithm for preserving the $\del \cdot {\bm B}$~=~0
317: constraint. Details of the algorithm are described in \cite{gard05a},
318: \cite{gard08}, and \cite{stone08}. The $\at$ code has been extensively
319: tested against various hydrodynamic and MHD tests \cite[]{stone08}.
320:
321:
322:
323:
324: We employ the shearing box formalism, in which our computational domain
325: is corotating with the fluid flow at some radius in the disk. The domain
326: size is small compared to this radius, allowing us to expand the equations
327: of motion in Cartesian form, as described in detail by \cite{haw95}.
328:
329:
330:
331: In the ideal MHD approximation, the evolution of the fluid in the
332: shearing box is described by:
333:
334: \begin{equation}
335: \label{cont}
336: \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho {\bm v}) = 0,
337: \end{equation}
338: \begin{equation}
339: \label{momentum}
340: \frac{\partial \rho {\bm v}}{\partial t} + \nabla \cdot (\rho {\bm v}{\bm v} - {\bm B}{\bm B}) +
341: \nabla (P + \frac{1}{2} B^2) = 2 q \rho \Omega^2 {\bm x} - 2 {\bm \Omega} \times \rho {\bm v},
342: \end{equation}
343: \begin{equation}
344: \label{induction}
345: \frac{\partial {\bm B}}{\partial t} + \nabla \cdot ({\bm v}{\bm B} - {\bm B}{\bm v}) = 0,
346: \end{equation}
347: \begin{equation}
348: \label{energy_eqn}
349: \frac{\partial E}{\partial t} + \nabla \cdot [(E + P + \frac{1}{2} B^2) {\bm v}
350: - {\bm B}({\bm B} \cdot {\bm v})] = 2 q \Omega^2 \rho {\bm v} \cdot {\bm x},
351: \end{equation}
352:
353: \noindent
354: where $\rho$ is the mass density, $\rho {\bm v}$ is the
355: momentum density, ${\bm B}$ is the magnetic field, $P$ is the gas
356: pressure, $E$ is the total energy density, and $q$ is the shear
357: parameter, defined as $q = -d$ln$\Omega/d$ln$R$. $\Omega$ is the angular velocity of the
358: center of the shearing box. Note that our system
359: of units has the magnetic permeability $\mu = 1$. We use $q = 3/2$,
360: appropriate for a Keplerian disk. The first source
361: term on the right-hand side of equation~(\ref{momentum}) and the term on
362: the right-hand side of equation~(\ref{energy_eqn}) correspond to
363: tidal forces (gravity and centrifugal) in the corotating frame.
364: The second source term in
365: equation~(\ref{momentum}) is the Coriolis force. The total energy density
366: is the sum of the thermal, kinetic, and magnetic energy
367: densities
368:
369: \begin{equation}
370: \label{energy}
371: E = \epsilon + \frac{1}{2} \rho v^2 + \frac{1}{2} B^2
372: \end{equation}
373:
374: \noindent
375: where $\epsilon$ is thermal energy density. The equation of
376: state is that of an ideal gas, $\epsilon = P / (\gamma-1)$,
377: where the adiabatic index is $\gamma = 5/3$ in all simulations. The
378: terms on the right hand sides of equations~(\ref{momentum}) and
379: (\ref{energy_eqn}) are added to the MHD equations in a directionally
380: unsplit manner, consistent with the CTU algorithm. Note that we have
381: neglected vertical stratification.
382:
383: An important component of shearing box simulations is the shearing
384: periodic boundary conditions at the $x$ boundaries, which are implemented
385: as described in \cite{haw95} with a few modifications for $\at$. First,
386: as in \cite{haw95}, the $y$ momentum is adjusted to account for the
387: shear across the $x$ boundaries as fluid moves out one boundary and
388: enters at the other. Since $\at$ evolves the total energy, however,
389: this energy must also be adjusted to account for the difference in $y$
390: momentum across the boundaries. Second, following the description in
391: \cite{haw95}, quantities are linearly reconstructed in the
392: ghost zones from appropriate zones in the physical domain that have
393: been shifted along $y$ to account for the shear across the boundary.
394: However, we have found that the precise conservation of a quantity
395: depends on how this reconstruction is performed; the fluxes of a
396: particular conserved quantity must be reconstructed to conserve
397: the quantity to roundoff level.
398: For example, consider the conservation of magnetic flux through the computational domain.
399: For the magnetic flux through the box to be
400: conserved to machine precision, the line integral of the electromotive
401: force (EMF), ${\cal E} = -{\bm v} \times {\bm B}$, along the boundaries
402: must remain zero. The $y$ and $z$ boundary conditions are periodic,
403: and therefore, the line integrated EMFs along these boundaries cancel.
404: This is not the case with the shearing periodic boundaries, however. Consider
405: the net $B_z$ flux through the grid, which will be conserved
406: if ${\cal E}_y = v_z B_x - v_x B_z$ is zero when integrated along both
407: $x$ boundaries. Computing the EMF using ghost zone variables $v_z,
408: B_x, v_x,$ and $B_z$ after reconstruction introduces a truncation error,
409: and the $B_z$ flux is not conserved. This is avoided if we instead
410: perform the shearing-periodic reconstruction step on ${\cal E}_y$ itself.
411: A similar argument applies to mass conservation; one needs to reconstruct
412: the density flux in the shearing boundaries instead of the density itself.\footnote{In principle,
413: the same argument applies to momentum and energy conservation, but
414: these equations are not conserved to machine precision due to the
415: existence of source terms.}
416:
417: In the code, we only perform this EMF/flux reconstruction for ${\cal E}_y$.
418: We have found that conservation of the $B_z$ flux is essential, owing
419: to the strong effect on the turbulence due to a net vertical field.
420: The perfect conservation of $B_y$ is not as important, and as ensuring
421: its precise conservation involves a more complex procedure, we allow
422: the $B_y$ flux to be conserved only to the truncation level. Similarly,
423: the precise conservation of mass has minimal impact on the behavior
424: of the turbulence, and we allow the mass to be conserved to truncation
425: level. We would like to note, however, that because of this, mass is lost
426: during the MRI evolution in our simulations. To quantify the level of
427: mass loss, the total percentage of mass lost over 100 orbits of evolution
428: is $\sim$ 2\% for our highest resolution simulations (see below for a description of our
429: simulations) and $\sim$ 10\% for our lowest resolution simulations;
430: we observe convergence of mass conservation with resolution.
431:
432: Although $\at$ conserves total energy, the shearing
433: boundaries do work on the fluid and represent a significant energy
434: source.
435: As was shown in \cite{haw95}, one can integrate the total energy plus
436: gravitational potential energy, $E + \rho \Phi$, where $\Phi = q \Omega^2
437: (\frac{L_x^2}{12}-x^2)$, over the domain to obtain
438:
439: \begin{equation}
440: \label{ein}
441: \frac{\partial \langle E+\rho \Phi\rangle}{\partial t} = \frac{q \Omega}{L_y L_z} \int_X (\rho v_x \delta v_y - B_x B_y) dy dz,
442: \end{equation}
443:
444: \noindent
445: where $L_x$, $L_y$, and $L_z$ are the domain sizes in the $x$, $y$, and $z$ directions respectively (see below),
446:
447: \begin{equation}
448: \label{pert_vy}
449: \delta v_y \equiv v_y+q\Omega x,
450: \end{equation}
451:
452: \noindent
453: and the integral is calculated over one of the $x$ boundaries.
454: In our simulations, equation~(\ref{ein}) is satisfied to truncation
455: level with the error coming from the tidal potential source term
456: in equation~(\ref{energy_eqn}). It is possible to rewrite this source
457: term to guarantee that equation~(\ref{ein}) is satisfied to roundoff
458: level \cite[see][]{gard05b}, but we have found that this makes very
459: little difference to how the total energy evolves.
460:
461: \cite{gard05b} point out that the source terms in the momentum equation
462: cannot be written in a purely conservative form and that the $x$ and $y$
463: momenta are tightly coupled through these terms. In the hydrodynamic
464: limit the source terms account for epicyclic oscillations, and if the
465: epicyclic kinetic energy (see their equation~8) is not conserved to
466: machine precision, coupling between long wavelength modes and epicyclic
467: oscillation modes can result from truncation error. Over time this
468: coupling can artificially increase the kinetic energy. To ensure the
469: conservation of epicyclic energy, \cite{gard05b} evolved the angular
470: momentum fluctuations directly rather than the $y$ momentum, casting
471: the equations into a form consistent with uniform epicyclic motion.
472: They then employed a Crank-Nicholson scheme to evolve the source terms
473: that govern the evolution of the mometum fluctuations. In MHD, however,
474: oscillatory epicyclic motion is replaced by unstable, growing MRI modes.
475: Epicyclic kinetic energy is not conserved and these special techniques
476: are not required. Therefore, we use the standard Athena algorithm
477: \cite[e.g.,][]{stone08} to evolve the momentum equations.
478:
479: As was done in the original shearing box simulations \cite[]{haw95} our
480: standard shearing box has a radial size $L_x = 1$, an azimuthal size $L_y
481: = 2\pi$, and a vertical size $L_z = 1$. We initialize a velocity flow with
482: ${\bm v} = -q\Omega x \hat{{\bm y}}$, with $q$~=~3/2, $\Omega$~=~0.001,
483: and $-L_x/2 \leq x \leq L_x/2$. In an isothermal disk, the sound speed
484: is $c_{\rm s} \sim \Omega H$ where $H$ is the scale height. With $L_z = H$,
485: we have $c_{\rm s}$~=~$L_z\Omega$, and we define the initial pressure as $P =
486: \rho \Omega^2 L_z^2$. With $\rho$~=~1, we have $P$~=~$10^{-6}$. In this
487: paper, we consider two initial magnetic field geometries that are commonly
488: used in shearing box studies. Models labeled NZ (for Net Z-field) have
489: an initial uniform vertical magnetic field, $B_z$, and models labeled SZ
490: (for Sine Z-field) begin with a sinusoidal distribution of $B_z$ and
491: have zero net flux through the box. Specifically, we initialize the NZ
492: runs with ${\bm B} = \sqrt{2P/\beta}\,\hat{{\bm z}}$, and the SZ runs
493: with ${\bm B} = \sqrt{2P/\beta}\,{\rm sin[}(2\pi/L_x) x{\rm]}\hat{{\bm
494: z}}$. In both cases, we set $\beta =1600$. This determines the ratio
495: of the vertical box size to the fastest growing linear MRI wavelength as
496: $L_z/\lambda_c \sim$~4, where $\lambda_c = 2\pi \sqrt{16/15}|v_{\rm A}|/\Omega$,
497: and $v_{\rm A}$ is the $\alf$ speed. To seed the MRI, we introduce random
498: adiabatic perturbations to $P$ and $\rho$ with amplitude $\delta
499: P/P$~=~0.01.
500:
501:
502:
503: For both of these initial field configurations, we have run a full range of
504: grid resolutions, from $N_x$~=~16, $N_y$~=~32, $N_z$~=~16 to the highest
505: resolution used in this study, $N_x$~=~128, $N_y$~=~256, $N_z = 128$,
506: proceeding by factors of two. All of the simulations were run for a total
507: of 100~orbits.
508:
509:
510:
511: In addition to the standard shearing box simulations, we have run some
512: additional experiments designed to further investigate magnetic and
513: kinetic energy dissipation. First, we perform a set of simulations
514: in which we remove the velocity shear and the tidal and Coriolis force
515: terms, thus removing the energy source that maintains the turbulence.
516: The purpose of these simulations is to investigate energy flow and
517: dissipation in the absence of the shear, which is the driving force for
518: the turbulence. We perform these simulations by restarting each of the
519: standard shearing box runs at a time when the shearing periodic boundaries
520: are strictly periodic. These ``periodic points" are given by $t_n =
521: n\,L_y/q\,\Omega\,L_x$, with $n = 0, 1, 2, ...,$ \cite[see][]{haw95}. We
522: choose the restart time to be 40~orbits. We then evolve the system to
523: follow the decay of the kinetic and magnetic energies.
524:
525:
526:
527: Finally, we run a set of low resolution simulations with varying aspect
528: ratio to examine the effect of secondary parasitic modes on the channel
529: solution (see \S~\ref{channel_solution}). These simulations have the
530: same initial conditions as the net flux simulation with $N_x = 32$,
531: $N_y = 64$, and $N_z = 32$ but with varying domain size in the $x$
532: and $y$ dimensions. The grid cell size (e.g., $L_x/N_x$ in the $x$
533: direction) in each dimension is kept constant. All simulations are
534: summarized in Table~\ref{tbl:runs}.
535:
536:
537: \clearpage
538: \begin{deluxetable}{l|cccc}
539: \tabletypesize{\scriptsize}
540: \tablewidth{0pc}
541: \tablecaption{MRI Simulations with $\at$ \label{tbl:runs}}
542: \tablehead{
543: \colhead{Label}&
544: \colhead{Initial Field Geometry}&
545: \colhead{Resolution ($N_x \times N_y \times N_z$)}&
546: \colhead{Domain ($L_x \times L_y \times L_z$)}&
547: \colhead{Description} }
548: \startdata
549: $\nonesix$ & net flux & 16 $\times$ 32 $\times$ 16 & $1 \times 2\pi \times 1$ & -- \\
550: $\nthree$ & net flux & 32 $\times$ 64 $\times$ 32 & $1 \times 2\pi \times 1$ & -- \\
551: $\nsix$ & net flux & 64 $\times$ 128 $\times$ 64 & $1 \times 2\pi \times 1$ & -- \\
552: $\none$ & net flux & 128 $\times$ 256 $\times$ 128 & $1 \times 2\pi \times 1$ & fiducial run - net flux \\
553: %$\ndonesix$ & net flux & 16 $\times$ 32 $\times$ 16 & $1 \times 2\pi \times 1$ & decaying turbulence \\
554: %$\ndthree$ & net flux & 32 $\times$ 64 $\times$ 32 & $1 \times 2\pi \times 1$ & decaying turbulence \\
555: %$\ndsix$ & net flux & 64 $\times$ 128 $\times$ 64 & $1 \times 2\pi \times 1$ & decaying turbulence \\
556: $\ndone$ & net flux & 128 $\times$ 256 $\times$ 128 & $1 \times 2\pi \times 1$ & decaying turbulence \\
557: $\zonesix$ & zero net flux & 16 $\times$ 32 $\times$ 16 & $1 \times 2\pi \times 1$ & -- \\
558: $\zthree$ & zero net flux & 32 $\times$ 64 $\times$ 32 & $1 \times 2\pi \times 1$ & -- \\
559: $\zsix$ & zero net flux & 64 $\times$ 128 $\times$ 64 & $1 \times 2\pi \times 1$ & -- \\
560: $\zone$ & zero net flux & 128 $\times$ 256 $\times$ 128 & $1 \times 2\pi \times 1$ & fiducial run - zero net flux \\
561: %$\zdonesix$ & zero net flux & 16 $\times$ 32 $\times$ 16 & $1 \times 2\pi \times 1$ & decaying turbulence \\
562: %$\zdthree$ & zero net flux & 32 $\times$ 64 $\times$ 32 & $1 \times 2\pi \times 1$ & decaying turbulence \\
563: %$\zdsix$ & zero net flux & 64 $\times$ 128 $\times$ 64 & $1 \times 2\pi \times 1$ & decaying turbulence \\
564: $\zdone$ & zero net flux & 128 $\times$ 256 $\times$ 128 & $1 \times 2\pi \times 1$ & decaying turbulence \\
565: $\narone$ & net flux & 16 $\times$ 64 $\times$ 32 & $\frac{1}{2} \times 2\pi \times 1$ & varied aspect ratio \\
566: $\nartwo$ & net flux & 64 $\times$ 64 $\times$ 32 & $2 \times 2\pi \times 1$ & varied aspect ratio \\
567: $\narthree$ & net flux & 32 $\times$ 32 $\times$ 32 & $1 \times \pi \times 1$ & varied aspect ratio \\
568: $\narfour$ & net flux & 16 $\times$ 32 $\times$ 32 & $\frac{1}{2} \times \pi \times 1$ & varied aspect ratio \\
569: $\narfive$ & net flux & 64 $\times$ 32 $\times$ 32 & $2 \times \pi \times 1$ & varied aspect ratio \\
570: $\narsix$ & net flux & 128 $\times$ 32 $\times$ 32 & $4 \times \pi \times 1$ & varied aspect ratio \\
571: \enddata
572: \end{deluxetable}
573: \clearpage
574:
575: \section{General Properties of MRI Turbulence}
576: \label{general_properties}
577:
578: This work represents the first detailed study of the MRI with $\at$,
579: which has an algorithm significantly different from that used in ZEUS.
580: To begin, we will reexamine many of the shearing box models and the
581: results already documented in the literature. Any significant differences
582: between $\at$ results and those previously published could indicate
583: where numerical effects (algorithm, resolution) have an influence.
584: Since $\at$ is an energy-conserving, shock-capturing algorithm it has at
585: least the potential to produce somewhat different results. Conversely,
586: agreement between $\at$ and other codes would support the robustness of
587: the shearing box results to date.
588:
589:
590:
591: In this section, we describe some of the general properties of MRI
592: turbulence as simulated with $\at$ and compare our results with those in
593: the literature. These properties will also serve as a starting point for
594: further analysis presented in the following sections. In what follows,
595: the highest resolution runs $\none$ and $\zone$ will serve as our fiducial
596: simulations for each initial field geometry. We study resolution effects
597: for each field geometry using the lower resolution simulations.
598:
599: \subsection{Characteristics of Saturation}
600: \label{sat_characteristics}
601:
602: Figures~\ref{turb_n}~and~\ref{turb_z} show the development of the MRI
603: and the subsequent evolution of the resulting MHD turbulence for the
604: fiducial $\none$ and $\zone$ runs respectively. The MRI saturates
605: before orbit 5 and the MHD turbulent state lasts for the remainder of
606: the 100 orbit simulation. Along with these figures, we list several
607: time- and volume-averaged quantities from the fiducial runs
608: in Table~\ref{tbl:sat_char}. The time average is done from orbits
609: 20 to 100, and the errors are given by one standard deviation over
610: this period. Volume-averaged values are indicated by the single-angled
611: bracket notation (e.g., $\langle B^2 \rangle$), and time- and volume-
612: averaged values are denoted by double-angled brackets (e.g., $\langle
613: \langle B^2 \rangle \rangle$). In both fiducial
614: runs, the toroidal field magnetic energy dominates with $\langle
615: B_y^2/2\rangle > \langle B_x^2/2\rangle > \langle B_z^2/2\rangle$.
616: Examining the components of the kinetic energy and perturbed kinetic
617: energy, which is $(\rho/2)(v_x^2+\delta v_y^2+v_z^2)$ with $\delta v_y$
618: given by equation~(\ref{pert_vy}), we find they are closer to each other
619: in value than are the components of the magnetic energy. The relative
620: ordering is similar except that the $x$ kinetic energy is larger than the
621: perturbed $y$ kinetic energy, $\rho \delta v_y^2/2$, in $\zone$. Another feature
622: of note is the greater saturation level and fluctuation amplitude of
623: the $\none$ run compared to that of $\zone$. As in past studies, the
624: Maxwell stress dominates over the Reynolds; the ratio of the Maxwell to
625: Reynolds stress oscillates between 1 and 10. Similarly, past studies
626: have shown a tight correlation between Maxwell (and total) stress and
627: the magnetic energy density \cite[see, e.g.,][]{black08}. Here the ratio
628: of the Maxwell stress to the magnetic energy density is roughly 1/2.
629: These values and the overall observations are generally consistent with
630: the results of \cite{haw95}, \cite{haw96}, and \cite{sano04}.
631:
632:
633:
634: One major difference from past ZEUS simulations is
635: the evolution of the total ($E+\rho\Phi$) and thermal ($\epsilon$)
636: energy densities, shown in the lower right plot of
637: Figs.~\ref{turb_n}~and~\ref{turb_z} for the $\none$ and $\zone$
638: runs respectively. Since we evolve an adiabatic equation of state
639: and there is no cooling term in the energy equation, the total energy
640: increases with time at a rate given by equation~(\ref{ein}). The total
641: energy increases because the free energy of the shearing fluid is being
642: thermalized by the turbulence, but the shearing box boundary conditions
643: continuously reinforce that shear. The stresses at the radial boundaries
644: therefore constitute a source term. Equation~(\ref{ein}) also explains
645: why the total energy reaches a higher value at the end of the simulation
646: in $\none$ compared to $\zone$. Since the volume-averaged stress (which
647: is roughly equal to the stress at the radial boundaries) is higher in
648: $\none$, the energy injection rate will be larger. These plots also
649: show that the thermal energy follows the total energy very closely.
650: That is, the injected energy ends up as thermal energy a short time
651: later \cite[]{gard05b}.
652: We will further study the thermalization of injected energy in
653: \S\ref{energy_fluctuations} and \S\ref{trans_funcs}.
654:
655:
656:
657: Does the significant increase in thermal energy affect the turbulence
658: in any way? This question was examined by \cite{sano04} in an extensive
659: series of simulations. They found evidence of a very weak dependence of
660: the time-averaged Maxwell stress on the gas pressure. Such an increase
661: is not apparent from a first look at Figs.~\ref{turb_n} and \ref{turb_z},
662: but short timescale fluctuations are a dominant feature of these
663: volume-averaged quantities. We examined the long term behavior of
664: the Maxwell stress using time-averaging procedures to smooth away
665: the fluctuations (which do not appear to change over long timescales).
666: We found marginal evidence for a weak dependence of the Maxwell stress on
667: the gas pressure in some, but not all, of the data. While it is possible
668: that longer evolution times and a wider exploration of parameter space
669: could be useful to address this question further, it is clear the stress
670: has barely changed despite an increase in thermal pressure by a factor
671: of order 100 in run $\none$. Thus if there is any dependence of the
672: stress on the pressure, it is very weak and does not significantly affect
673: the characteristics of local MRI turbulence.
674:
675:
676:
677: We study the effect of resolution through a series of lower resolution
678: simulations (see Table~\ref{tbl:runs}). Figure~\ref{turb_res} shows the
679: time- and volume-averaged magnetic and perturbed kinetic energies as a function of grid
680: resolution for both the net flux and zero net flux initial conditions.
681: The time average is calculated from orbits 20 to 100; the error bars
682: indicate one standard deviation. For the net flux simulation, there
683: appears to be a slight trend of increasing energy with resolution,
684: as observed in \cite{haw95}. Resolution has a more obvious effect
685: on the zero net flux initial condition. The turbulent energies
686: {\it decrease} with increasing resolution. This resolution effect
687: was previously reported for zero net field initial conditions in other
688: simulations \cite[]{from07a,pess07} using different numerical algorithms.
689: With $\at$, the time- and volume-averaged total magnetic energy density
690: decreases by roughly a factor of two for each factor of two resolution
691: increase. The amplitude of the fluctuations in the total magnetic
692: energy density decreases by roughly a factor of two to four for each
693: resolution increment. At all resolutions, the $y$ magnetic energy
694: density continues to be the largest, followed by the $x$ energy, and
695: then the $z$ energy. As was the case for $\none$, $\rho \delta v_y^2/2$
696: dominates for all net flux simulations, followed by $\rho v_x^2/2$, and
697: then $\rho v_z^2/2$. In the zero net flux simulations, the $x$ kinetic
698: energy density is greater than the perturbed $y$ kinetic energy density.
699: These components of the perturbed kinetic energy density are close in
700: value, and it is often the case that the $x$ and $y$ components are
701: within one standard deviation of each other. The ratio of time- and
702: volume-averaged Maxwell stress to time- and volume-averaged magnetic
703: energy density is constant with resolution. The ratio of time- and
704: volume-averaged Maxwell stress to time- and volume-averaged Reynolds
705: stress has a slight increase with resolution in the net flux simulations
706: and a slight decrease with resolution in the zero net flux simulations.
707: However, we point out that the observed trends in the ratio of stresses
708: are subject to considerable uncertainty given the large error bars
709: calculated for the various quantities.
710:
711:
712: \subsection{Channel Solution}
713: \label{channel_solution}
714:
715: One of the interesting aspects of the vertical field MRI in a shearing
716: box is that the fastest growing mode leads to axisymmetric
717: radial streaming motions, dubbed ``channel solutions'' \cite[]{haw92}.
718: \cite{good94} pointed out that for the vertical field in an
719: unstratified box, the linear MRI eigenmode is also a nonlinear solution in the
720: incompressible limit. They further show that the nonlinear channel solution is itself
721: unstable to ``parasitic modes.'' These modes require radial and
722: azimuthal wavelengths larger than the vertical wavelength of the
723: channel solution and will disrupt the channel flow if the box is
724: large enough \cite[]{balb98}.
725:
726: In the present simulations, the initial vertical field is sufficiently
727: weak that the fastest growing vertical wavelength is less than the radial
728: and azimuthal dimensions of the box, and any initial tendency toward
729: the channel solution at the end of the linear growth phase is quickly
730: disrupted. However, we find that the large fluctuations in the magnetic
731: energy density for $\none$ are a result of recurring channel solutions.\footnote{
732: The recurrence of the channel solution presumably results from the fact that the net vertical
733: magnetic field can never be destroyed or removed from the domain, given the periodic
734: boundary conditions and the strict conservation of $z$ magnetic flux.}
735: Figure~\ref{channel} shows the azimuthally-averaged velocities at
736: several times during the amplification and subsequent decay of one such
737: fluctuation. The flow organizes itself into a two-channel solution,
738: which becomes more well-defined as the magnetic energy increases.
739: The channel solution is eventually destroyed via secondary, parasitic
740: instabilities \cite[see][]{good94}, which coincides with a decrease in
741: magnetic energy. The same channel solution appears during other instances
742: of large magnetic energy fluctuation in $\none$ and does not appear in
743: $\zone$. Furthermore, the recurring channel flows appear in the lower
744: resolution net magnetic flux simulations. As observed previously, the
745: channel solution and large magnetic energy fluctuations are a property
746: of simulations with a uniform $B_z$ field \cite[]{sano01}.
747:
748:
749:
750: Since the channel solution is subject to parasitic modes that depend
751: on the available wavelengths that can fit in the box, we expect
752: that this behavior is influenced by the domain aspect ratio employed.
753: To verify this, we have run several low resolution simulations (labelled
754: $\narone-\narsix$, see \S~\ref{method}) using different aspect ratios.
755: We found that for large enough $L_x$, the intermittent channel modes
756: no longer occur; this behavior was also observed by \cite{bod08}.
757: The prominence of intermittent channel flows is a consequence of the
758: restrictions introduced by the domain size. However, we use this
759: property in \S~\ref{energy_fluctuations}, where the large fluctuations
760: in turbulent energy created by the channel solutions provide a clear
761: marker of energy injection by the boundaries. We can then track the
762: subsequent thermalization of that energy.
763:
764: \subsection{Energy Power Spectra}
765: \label{energy_spec}
766:
767: The nature of MRI-driven MHD turbulence can be characterized in part
768: by the power spectrum of kinetic and magnetic energies. To obtain such
769: power spectra, we do a full 3D Fourier transform on the simulation data
770: employing the procedures outlined in \cite{haw95} to account for the
771: shearing-periodic boundaries. Briefly, the shearing periodic boundary
772: conditions in the $x$ direction allow the domain to be strictly periodic
773: in the $x$ direction only at certain times, called periodic points $t_n$
774: (described in \S\ref{method}). To perform a standard fast Fourier
775: transform (FFT) at some time $t$ that is not equal to $t_n$, we transform
776: the data into a frame where the $x$ boundaries are strictly periodic.
777: We then calculate the FFT in this frame and remap to the original frame.
778:
779:
780:
781: The turbulent magnetic, kinetic, and perturbed kinetic energy densities
782: in Fourier space are defined as
783:
784: \begin{equation}
785: \label{mag_fourier}
786: \frac{1}{2}|\bfour|^2 \equiv \frac{1}{2}\left[|\widetilde{B_x}({\bm k})|^2+|\widetilde{B_y}({\bm k})|^2+|\widetilde{B_z}({\bm k})|^2\right],
787: \end{equation}
788:
789: \begin{equation}
790: \label{ke_fourier}
791: \frac{1}{2}|\vfour|^2 \equiv \frac{1}{2}\left[|\widetilde{\sqrt{\rho}v_x({\bm k})}|^2+|\widetilde{\sqrt{\rho}v_y({\bm k})}|^2+|\widetilde{\sqrt{\rho}v_z}({\bm k})|^2\right],
792: \end{equation}
793:
794: \begin{equation}
795: \label{pke_fourier}
796: \frac{1}{2}|\vpfour|^2 \equiv \frac{1}{2}\left[|\widetilde{\sqrt{\rho}v_x}({\bm k})|^2+|\widetilde{\sqrt{\rho}\delta v_y({\bm k})}|^2+|\widetilde{\sqrt{\rho}v_z({\bm k})}|^2\right],
797: \end{equation}
798:
799: \noindent
800: where $\widetilde{f}$ means the Fourier transform of $f$ defined by
801:
802: \begin{equation}
803: \label{four_trans}
804: \widetilde{f({\bm k})} = \int \int \int f({\bm x}) \fterm.
805: \end{equation}
806:
807: \noindent Note that for the kinetic energies, we include the density along
808: with the velocity when calculating the Fourier transform, resulting in
809: the appearance of $\sqrt{\rho}$ in the above equations. To obtain these
810: quantities as a function of length scale and to improve statistics, we
811: average our data over shells of constant $k = |{\bm k}|$. For further
812: improvement of statistics, we average each of these terms over 161 frames
813: (i.e., from orbit 20 to 100 in increments of 0.5 orbits).
814:
815:
816:
817: Figure~\ref{power_spec} shows the power spectra
818: of these energy densities for the net flux and zero net flux runs.
819: The figure shows resolution effects as different lines in each plot.
820: In all cases, the largest scales account for most of the energy. The
821: general shape of the energy power spectra agrees with previous
822: studies \cite[e.g.,][]{haw95,from07a}. For the net flux simulations,
823: the magnetic energy dominates over the kinetic and perturbed kinetic
824: energies at all scales, independent of resolution. As the resolution
825: is increased, the power spectra extend to higher $k$, but the general
826: shape remains constant.
827: At some values for $k$, the uncertainty in
828: energy (not plotted), represented by one temporal standard deviation around the mean,
829: is large enough to overlap with other energy components, making it difficult
830: to conclusively say which energy dominates at these particular scales.
831:
832:
833:
834: We calculated a power law index in Fourier space for each energy density
835: and at each resolution. This slope was determined by a linear fit to
836: the energy densities in log space from $k L/(2\pi) = 1$ to the maximum
837: scale for the given resolution. There is some uncertainty in this measurement
838: because the power spectra are not strictly linear in log space (see Fig.~\ref{power_spec}).
839: In $\none$, the energy density is
840: proportional to $[k L/(2\pi)]^n$ with $n \approx -4$ for every energy
841: density. This index is approximately constant with resolution, but
842: there is evidence that $n$ becomes more negative at higher resolutions.
843: In determining an error in the value of $n$, we found that this error is
844: often dominant. Thus, such a resolution dependence is somewhat tentative.
845:
846:
847:
848: There is a noticeable resolution dependence in the zero net
849: flux simulations. First, as resolution is increased, the magnetic
850: energy density decreases at all scales. This effect was discussed in
851: \S~\ref{sat_characteristics}; the power spectra are consistent with the
852: power spectrum analysis of \cite{from07a}. The same resolution dependence
853: is observed for the perturbed kinetic energy density. The magnetic
854: energy density at small $k$ decreases faster with resolution than does
855: the perturbed kinetic energy density. The total kinetic energy density (i.e.,
856: including shear) remains constant with resolution, which simply results
857: from the fact that the shear velocity, which dominates the kinetic energy,
858: is constant with resolution. The uncertainty in each energy component
859: appears to be smaller than in the net flux simulations. However, there
860: are still some values of $k$ at which the calculated errors overlap.
861:
862:
863:
864: We calculated a power law index in Fourier space for each energy density
865: and resolution for the zero net flux simulations. The procedure we
866: used was the same as for the net flux simulations. For the kinetic and
867: perturbed kinetic energy densities, we found that $n$ lies between -3.5
868: and -4, whereas for the magnetic energy density, $n$ lies between -3
869: and -3.5. There does not appear to be any resolution dependence in $n$
870: for the magnetic energy density, but there is a tentative decrease in $n$
871: (similar to the net flux case) with increasing resolution for the kinetic
872: and perturbed kinetic energy densities.
873:
874:
875: \clearpage
876: \begin{deluxetable}{c|cc}
877: \tabletypesize{\scriptsize}
878: \tablewidth{0pc}
879: \tablecaption{Saturation Characteristics \label{tbl:sat_char}}
880: \tablehead{
881: \colhead{Quantity}&
882: \colhead{$\none$}&
883: \colhead{$\zone$} }
884: \startdata
885: $\maxwell/P_o$ & 0.216 $\pm$ 0.116 & $(6.55\pm1.15) \times 10^{-3}$ \\
886: $\reynolds/P_o$ & 0.028 $\pm$ 0.019 & $(1.91\pm0.76) \times 10^{-3}$ \\
887: $\magnetic/P_o$ & 0.488 $\pm$ 0.262 & 0.014 $\pm$ 0.003 \\
888: $\magx/P_o$ & 0.071 $\pm$ 0.027 & $(2.01\pm0.38) \times 10^{-3}$ \\
889: $\magy/P_o$ & 0.388 $\pm$ 0.231 & 0.011 $\pm$ 0.002 \\
890: $\magz/P_o$ & 0.029 $\pm$ 0.011 & $(7.98\pm1.57) \times 10^{-4}$ \\
891: $\pertkinetic/P_o$ & 0.145 $\pm$ 0.060 & $(7.69\pm1.81) \times 10^{-3}$ \\
892: $\kinx/P_o$ & 0.046 $\pm$ 0.024 & $(3.73\pm1.27) \times 10^{-3}$ \\
893: $\pertkiny/P_o$ & 0.078 $\pm$ 0.035 & $(2.68\pm0.60) \times 10^{-3}$ \\
894: $\kinz/P_o$ & 0.021 $\pm$ 0.011 & $(1.28\pm0.21) \times 10^{-3}$ \\
895: $\maxwell/\reynolds$ & 7.60 $\pm$ 6.47 & 3.43 $\pm$ 1.49 \\
896: $\maxwell/\magnetic$ & 0.443 $\pm$ 0.336 & 0.462 $\pm$ 0.116 \\
897: \enddata
898: \end{deluxetable}
899: \clearpage
900:
901: \section{Energy Fluctuations}
902: \label{energy_fluctuations}
903:
904: $\at$ evolves the equation for total energy, the volume-average
905: of which will change only due to the Maxwell and Reynold stresses
906: at the radial boundaries (equation (\ref{ein})). As was discussed in
907: \S\ref{general_properties}, the individual volume-averaged magnetic and
908: kinetic energies are highly variable throughout the evolution as energy is continuously
909: transferred between magnetic, kinetic and thermal components. We can
910: study these energy flow processes by tracking the energy injected at the boundaries as
911: it is subsequently thermalized in the turbulence. For this purpose, the
912: existence of the recurring channel solution in the net magnetic field
913: simulation is very useful; the sudden increase in stress provides a
914: clear injection of energy that can be traced using several diagnostics.
915: Having developed these diagnostics we can then apply them to the zero
916: net magnetic flux simulations. Finally, to gain additional insight
917: into dissipation in the turbulence, we conduct an experiment in which
918: the shear flow and gravity terms have been removed.
919:
920:
921: \subsection{Sustained Turbulence}
922: \label{sustained}
923:
924: The total energy density, including the gravitational potential energy density,
925: is defined as
926:
927: \begin{equation}
928: \label{total_energy}
929: \etot = E + \rho \Phi = \epsilon + \frac{1}{2} \rho v^2 + \frac{1}{2} B^2 + \rho \Phi
930: \end{equation}
931:
932: \noindent
933: where $\Phi$ is given in \S\ref{method}. Averaging
934: equation~(\ref{total_energy})
935: over the entire domain, taking the time derivative, and
936: rearranging the terms, we obtain,
937:
938: \begin{equation}
939: \label{dtedt}
940: \td = \ein - \kd - \md - \gd .
941: \end{equation}
942:
943: \noindent
944: where $\ein \equiv \partial \langle \etot\rangle/\partial t$ is the energy
945: injection rate due to stress at the boundaries (see equation~(\ref{ein})),
946: $\td \equiv \partial \langle \epsilon\rangle /\partial t$ is the rate of
947: change of thermal energy density, $\kd \equiv \partial \langle \frac{1}{2}\rho
948: v^2 \rangle / \partial t$ is the rate of change of the kinetic energy
949: density, $\md \equiv \partial \langle \frac{1}{2} B^2\rangle /
950: \partial t$ is the rate of change of the magnetic energy density,
951: and $\gd \equiv \partial \langle \rho \Phi\rangle / \partial t$
952: is the time derivative of the tidal potential energy density.
953: The brackets indicate a
954: volume-average over the simulation domain. $\gd$ is the change
955: in a fluid element's gravitational energy as it moves within the domain.
956: We expect the contribution of the tidal potential
957: term to be insignificant, an expectation borne out by direct computation.
958: We will ignore this term in most of the subsequent discussion. The stress
959: terms at the radial boundaries are generally positive, which means
960: energy is being injected into the box via the work done by this stress
961: ($\ein >$ 0). The remaining terms in equation~(\ref{dtedt}) can be
962: either positive or negative.
963:
964:
965:
966: The lower right plot in Figure~\ref{turb_n} shows that the thermal
967: energy density closely follows the total energy density, but with
968: a short time delay. This can be better seen in Figure~\ref{dedt_n},
969: which shows the individual terms from equation~(\ref{dtedt})
970: for a 20 orbit period in the $\none$ simulation. There is a clear time
971: delay of less than one orbit between significant changes in the energy
972: injection rate and the thermal energy derivative, suggesting a comparable delay before
973: the injected energy is thermalized, a property noted in \cite{sano01}
974: as well as in \cite{gard05b}. These features in the energy derivatives
975: result from the creation and destruction of channel flows. During this time interval, the magnetic
976: and kinetic energies are also changing. By examining the maxima in
977: the thermal energy derivative and the corresponding features in the kinetic
978: and magnetic energy derivatives, it appears that the magnetic energy
979: dissipation dominates the thermalization process.
980:
981:
982:
983:
984: It is useful to define a temporal correlation function for the various
985: energy components by writing
986:
987: \begin{equation}
988: \label{corr_func}
989: C_{AB} \equiv \left\{ \begin{array}{ll}
990: \frac{\displaystyle\frac{1}{N-|L|}\sum_{i=0}^{N-|L|-1} A_{i+|L|} B_i}{\displaystyle\frac{1}{N}\sum_{i=0}^{N-1} A_i} & \quad
991: \mbox{if $L <$ 0} \\
992: \frac{\displaystyle\frac{1}{N-|L|}\sum_{i=0}^{N-L-1} A_i B_{i+L}}{\displaystyle\frac{1}{N}\sum_{i=0}^{N-1} A_i} & \quad
993: \mbox{if $L \ge$ 0}
994: \end{array} \right.
995: \end{equation}
996:
997: \noindent
998: where $A$ and $B$ are two time-series datasets $N$ elements
999: in length. The quantity $L$ is the number of elements over which to
1000: shift $A$ and $B$ with respect to each other to calculate the correlation
1001: coefficient. We apply equation~(\ref{corr_func}) to the energy rates
1002: by setting $A$ = $\td$, and $B$ = $\kd$, $\md$, or $\ein$. This allows
1003: us to correlate the energy injection rate and the change in kinetic and
1004: magnetic energies against the change in thermal energy over certain timescales. Since
1005: $\td >$ 0, if the correlation between $\td$ and $\kd$ (or $\md$) is
1006: negative, then kinetic energy (or magnetic energy) must be decreasing,
1007: and a strong negative correlation would suggest that kinetic energy
1008: (or magnetic energy) is being thermalized.
1009:
1010: Figure~\ref{corr_n} is the correlation function for $B = \ein, \kd,$ and
1011: $\md$ calculated over orbits 20 to 100.
1012: The $x$-axis is the correlation timescale in units of orbits.
1013: We only look at correlation times of $\lesssim$ 1 orbit as the degree to which
1014: the thermal energy evolution follows that of the total energy (see Figures~\ref{turb_n}
1015: and \ref{turb_z}) indicate that thermalization happens over that timescale. To examine
1016: the correlation function on longer timescales would be misleading since peaks in the function
1017: would suggest a correlation between two events that are not causally related (e.g., the injection
1018: of energy for one channel event being correlated with the thermalization of energy for another channel
1019: event). The left plot of the figure shows that $\ein$ is strongly correlated with
1020: $\td$ on a timescale of $\Delta t \sim$ -0.2 orbits. This correlation is
1021: exactly what we observed in Fig.~\ref{dedt_n}. The energy injected by
1022: the stress at the boundaries ends up as heat less than one orbit later.
1023: The negative sign on this value of $\Delta t$ simply means that the
1024: injection happens before the thermalization. In the right plot,
1025: both $\kd$ and $\md$ are negatively correlated with $\td$ suggesting that
1026: magnetic and kinetic energy are being thermalized. The stronger magnetic correlation further
1027: suggests that magnetic dissipation contributes more to thermalization than
1028: kinetic dissipation. The positive correlation between $\kd$ and $\md$
1029: against $\td$ at negative $\Delta t$ values is a result of the magnetic
1030: and kinetic energies increasing along with the energy injection into
1031: the box. That is, the stress at the boundaries increases the magnetic
1032: and kinetic energies which are dissipated a short time later.
1033:
1034:
1035:
1036: An interesting feature is evident in Fig.~\ref{corr_n}: the negative
1037: peak in the magnetic and kinetic correlation functions occur for $\Delta
1038: t$ slightly greater than zero. Similarly, in Fig.~\ref{dedt_n} one
1039: can see that peaks in the magnetic and kinetic energy derivatives are
1040: offset with respect to the energy injection and thermalization peaks.
1041: For example, the maximum rate for magnetic energy loss occurs after
1042: the maximum rate for thermal energy gain. Of course, these are plots of
1043: the time derivative of the energy, so a peak simply indicates where the
1044: second derivative is zero. The magnetic energy is both losing energy
1045: to dissipation while gaining energy from the shear at the boundaries.
1046: When the energy injection rate peaks decline, the thermalization rate
1047: is still growing and the magnetic energy rate also peaks and begins
1048: to decline. Similarly, the slope of the magnetic energy loss rate will
1049: change sign after the thermalization rate has peaked and when the energy
1050: injection rate is no longer itself in decline.
1051:
1052:
1053:
1054: As a test, we performed this correlation analysis on
1055: the lower resolution net-flux simulations and find that energy injection
1056: precedes thermalization by $\sim$ 0.2 orbits, independent of resolution.
1057: Furthermore, magnetic dissipation dominates over kinetic dissipation for all
1058: net flux simulations.
1059:
1060:
1061:
1062: The analysis so far has only examined the rate of change in the energy
1063: terms, not specifically how they change. For example, does a ``dip"
1064: in $\md$ correspond to direct thermalization of magnetic energy, or is
1065: there a transfer of energy from magnetic to kinetic? To examine the
1066: energy flow in more detail, we focus on orbits 50 to 52 in $\none$, for which we ran the $\none$ simulation at
1067: high temporal resolution. This high time resolution allows us to resolve short timescale features, but also
1068: generates many large data files. Therefore, we restrict this part of the analysis to the two orbit period mentioned above.
1069: Consider the evolution equation for the volume-averaged kinetic energy
1070: given by
1071:
1072: \begin{eqnarray}
1073: \label{ke_eqn}
1074: \kd & = & -\left\langle \del \cdot \left[{\bm v}\left(\frac{1}{2}\rho v^2+\frac{1}{2}B^2 + P + \rho \Phi \right) - {\bm B} ({\bm v} \cdot {\bm B})\right]\right\rangle \nonumber \\
1075: & & + \left\langle\left(P + \frac{1}{2} B^2\right)\del \cdot {\bm v}\right\rangle-\left\langle{\bm B} \cdot ({\bm B} \cdot \del {\bm v})\right\rangle - \gd - \qk,
1076: \end{eqnarray}
1077:
1078: \noindent
1079: where $\qk$ is the volume-averaged (numerical) kinetic energy dissipation rate.
1080: The evolution equation for the volume-averaged magnetic energy is given by
1081:
1082: \begin{equation}
1083: \label{be_eqn}
1084: \md = - \left\langle \del \cdot \left(\frac{1}{2}B^2{\bm v} \right)\right\rangle - \left\langle \frac{1}{2} B^2 \del \cdot {\bm v}\right\rangle + \left\langle{\bm B} \cdot ({\bm B} \cdot \del {\bm v})\right\rangle- \qm
1085: \end{equation}
1086:
1087: \noindent
1088: where $\qm$ is the volume-averaged (numerical) magnetic energy
1089: dissipation rate. We have calculated each term in these equations over
1090: the two orbit period and find that the dominant terms are $-\langle\del
1091: \cdot (\frac{1}{2}\rho v^2){\bm v}\rangle$, $\langle\del \cdot [{\bm
1092: B}({\bm v} \cdot {\bm B})]\rangle$, $\langle{\bm B} \cdot ({\bm B} \cdot
1093: \del {\bm v})\rangle$, $\qk$, and $\qm$. $\qk$ and $\qm$ are what remain
1094: after calculating all other terms in the energy equations at a particular
1095: instant in time. Calculating the volume-averages
1096: of the first two terms yields the radial boundary Reynolds and Maxwell
1097: stresses in Equation~(\ref{ein}) \cite[]{haw95}, namely the energy
1098: injection rate by the shearing periodic boundaries. The third of
1099: the dominant terms is the transfer rate of kinetic to magnetic energy via
1100: field line stretching. Figure~\ref{energies_n} plots the time-history of
1101: this term (pink line) along with $\td$ (black line), the energy injection
1102: rate $\ein$ (blue line), and $-\qk$ and $-\qm$ (green and red lines, respectively).
1103: As energy is injected into the grid, a significant fraction of this
1104: energy is transferred to the magnetic field via field line stretching,
1105: presumably through the shear flow. Thermalization follows 0.2 orbits
1106: later and is marked by increases in the
1107: absolute value of $\qk$ and $\qm$, with $|\qm| > |\qk|$. The ratio of
1108: kinetic to magnetic dissipation is approximately constant in time over
1109: this period, with $\qk/\qm \approx 0.6$. This suggests that the details
1110: of the thermalization do not vary with intermittent increases in $\ein$
1111: that occur when the fluid experiences a channel flow.
1112:
1113:
1114:
1115: As discussed, the recurring channel modes in the net flux simulations
1116: create distinguishable points of energy injection that make it
1117: straightforward to follow the subsequent thermalization. Such modes do
1118: not exist in the zero net flux simulations, which makes the identification
1119: of specific correlations slightly more difficult. The situation is
1120: further complicated by the overall reduced levels of the turbulence which
1121: causes the time derivative of the thermal energy to be dominated by very high frequency oscillations due to
1122: propagating spiral density waves \cite[][]{gard05b}. We have determined
1123: that these waves are created by compressibility and have very little effect
1124: on the dissipational heating within the box.
1125: To remove their dominance in the energy derivatives,
1126: we rebin the time data using a ``neighborhood"
1127: averaging procedure in which the rebinned data points are calculated from
1128: averages of a specified number of original data points. We then apply
1129: equation~(\ref{corr_func}) between $\ein$ and $\td$; the result is shown
1130: in Fig.~\ref{corr_z}. The correlation curve has several narrow peaks,
1131: which result from residual effects of the rebinning process. The curve
1132: has a broader peak near $\Delta t\sim$ -0.2~orbits, which agrees with the
1133: same curves for $\none$ (Fig.~\ref{corr_n}). The correlation
1134: function for $\zone$ is not as sharply peaked as that for $\none$,
1135: which is most likely a result of the lower amplitude variability in the
1136: rebinned $\zone$ data. Applying this analysis to the lower resolution
1137: zero net flux simulations, we find that the correlation function always
1138: has a broad peak at $\Delta t\sim$ -0.2~orbits. Thus, as was the case in
1139: the net flux simulations, the energy injection/thermalization timescale
1140: is independent of resolution.
1141:
1142:
1143:
1144: Finally, we note that the saturated state of $\zone$ is too complex to
1145: obtain correlations between $\md$, $\kd$, and $\td$, such as was done
1146: for $\none$. In the net flux simulations, the recurring channel modes
1147: lead to the build up and thermalization of magnetic energy. The creation
1148: and thermalization of magnetic energy are events that are well-separated
1149: in time, making it easy to study the flow of energy between various
1150: components. In the zero net flux simulations, however, the average
1151: properties of the turbulence remain more constant in time. We will
1152: further investigate the dissipation of magnetic and kinetic energy for the
1153: zero net flux geometry in \S\ref{decaying} and \S\ref{zero_net_flux}.
1154:
1155: \subsection{Decaying Turbulence}
1156: \label{decaying}
1157:
1158: As noted by \cite{haw95}, the MHD turbulence decays without differential
1159: rotation to sustain the MRI. We make use of this to observe how
1160: rapidly thermalization occurs when there is no further input of energy.
1161: This analysis should provide some additional
1162: insight into the thermalization process for each field geometry.
1163: We remove the net shear flow and the Coriolis and tidal forces from a
1164: state taken from the sustained MRI turbulence in the fiducial models.
1165: These runs are labeled ``NZD'' and ``SZD'' in Table~\ref{tbl:runs} and are described
1166: in more detail in \S\ref{method}. Figure~\ref{turb_decay} shows
1167: the subsequent magnetic and kinetic energy decay for both runs.
1168: In the figure, the kinetic and magnetic energies have been normalized to
1169: their values at the starting time of $t=40$ orbits.
1170:
1171: In $\ndone$, the ratio of total magnetic to kinetic energy at $t=40$~orbits is 3.4. The figure shows that
1172: the magnetic energy decays more rapidly than the kinetic energy at early
1173: times, losing almost half its initial value within 0.2 orbits.
1174: In $\zdone$, the ratio of total magnetic to
1175: kinetic energy at $t=40$~orbits is 1.4. The kinetic energy shows
1176: high frequency oscillations about an average value that decays in time.
1177: These oscillations are due to the same compressive, spiral waves that
1178: exist in the sustained turbulence simulations. The magnetic energy
1179: is unaffected by these waves. The average decay of kinetic energy,
1180: calculated from smoothing away the oscillations, is also shown in the
1181: figure. Both the kinetic and magnetic energies decay quickly over time.
1182: Again, almost half the magnetic energy is lost within 0.2 orbits.
1183: The high frequency oscillations also decay in amplitude over time. As
1184: was the case in $\ndone$, the magnetic dissipation rate is initially
1185: faster than that for the kinetic energy. After about one orbit, the decay
1186: rates become comparable.
1187:
1188: Finally, we checked the contributions from the various terms in equations~(\ref{ke_eqn})
1189: and (\ref{be_eqn}). In both $\ndone$ and $\zdone$, there is some transfer from magnetic
1190: to kinetic energy during the decay. However, the transfer rate is small compared to the decay
1191: rate of the magnetic energy and is such that the numerical dissipation of magnetic energy
1192: dominates over that of kinetic energy.
1193:
1194: \section{Transfer Functions}
1195: \label{trans_funcs}
1196:
1197: In their investigation of convergence of zero net flux shearing box
1198: simulations, \cite{from07a} carried out an analysis based on the evolution
1199: of magnetic energy in Fourier space. This analysis shows how magnetic
1200: energy is created, transferred from one scale to another, and finally lost
1201: due to numerical dissipation. Their study used the ZEUS code and assumed
1202: an isothermal equation of state. Here we repeat and expand upon their
1203: analysis to understand dissipation as a function of length scale in $\at$.
1204:
1205: We note several differences between our work and that of \cite{from07a}.
1206: First, they focus on magnetic energy evolution and did not provide a
1207: comparable calculation for the kinetic energy. Second, recognizing
1208: that the $y$ direction is dominated by the largest scales, they restricted
1209: their analysis to axisymmetric modes, namely $k_y = 0$. Finally, as they
1210: were primarily interested in how poloidal field could be regenerated as
1211: part of a dynamo process, a portion of their analysis concentrated
1212: on the poloidal components rather than the full magnetic field. We have
1213: chosen to extend the \cite{from07a} analysis more generally to include a
1214: kinetic energy density evolution, nonaxisymmetric effects, and the effects
1215: of a nonzero toroidal field.
1216:
1217:
1218: Following \cite{from07a}, we decompose the velocity field of the flow
1219: into the mean flow, ${\bm V}_{\rm sh}$, and the turbulent velocity, $\vt$, via
1220:
1221: \begin{equation}
1222: \label{turb_velocity}
1223: {\bm v} = {\bm V}_{\rm sh} + \vt.
1224: \end{equation}
1225:
1226: \noindent
1227: The mean flow is defined as
1228:
1229: \begin{equation}
1230: \label{vsh_define}
1231: {\bm V}_{\rm sh} = \vsh \hat{\bm y} = \frac{\hat{\bm y}}{L_y L_z} \int \int v_y(x,y,z){\rm d}y
1232: {\rm d}z.
1233: \end{equation}
1234:
1235:
1236:
1237: Turning next to the induction equation, we substitute
1238: equation~(\ref{turb_velocity}) for the velocity, take the Fourier
1239: transform, and dot the result with the complex conjugate of $\bfour$,
1240: which is defined by equation~(\ref{four_trans}) with $f = {\bm B}$.
1241: All Fourier transforms are done via equation~(\ref{four_trans}) using a
1242: standard FFT and replacing $f$ with the appropriate quantity. The data
1243: is mapped into a frame in which the $x$ boundaries are periodic and then
1244: remapped into the original frame after performing the FFT.
1245:
1246:
1247:
1248: The result of this calculation is an equation describing the magnetic
1249: energy density evolution in Fourier space,
1250:
1251: \begin{equation}
1252: \label{bfour_evolution}
1253: \frac{1}{2}\frac{\partial|\bfour|^2}{\partial t} = A + S + \tbb + \tdivv + \tbv + \dmag,
1254: \end{equation}
1255:
1256: \noindent
1257: where
1258:
1259: \begin{equation}
1260: \label{aterm}
1261: A = -Re \left[\bstar \cdot \int \int \int \vsh \frac{\partial{\bm B}}{\partial y} \fterm \right],
1262: \end{equation}
1263:
1264: \begin{equation}
1265: \label{sterm}
1266: S = +Re \left[\widetilde{B}^\ast_{\it y}({\bm k}) \cdot \int \int \int B_x \frac{\partial \vsh}{\partial x} \fterm \right],
1267: \end{equation}
1268:
1269: \begin{equation}
1270: \label{t_bb}
1271: \tbb = -Re \left[\bstar \cdot \int \int \int (\vt \cdot \del){\bm B} \fterm \right],
1272: \end{equation}
1273:
1274: \begin{equation}
1275: \label{t_divv}
1276: \tdivv = -Re \left[\bstar \cdot \int \int \int (\del \cdot \vt){\bm B} \fterm \right],
1277: \end{equation}
1278:
1279: \begin{equation}
1280: \label{t_bv}
1281: \tbv = +Re \left[\bstar \cdot \int \int \int ({\bm B} \cdot \del)\vt \fterm \right].
1282: \end{equation}
1283:
1284: \noindent
1285: The $\dmag$ term has no analytic expression; it is simply
1286: what is left over and accounts for numerical losses of magnetic energy
1287: \cite[]{from07a}. In the present simulations, there is no physical
1288: resistivity. The other terms have the following meanings: $A$ is the
1289: transfer of magnetic energy between scales by the shear flow, $S$ is the
1290: creation of magnetic energy from this shear flow, $\tbb$ is the advection of
1291: magnetic energy between scales by the turbulent velocity field, $\tdivv$
1292: results from the turbulent compressibility, and $\tbv$ describes the
1293: creation of magnetic field by the turbulent velocity fluctuations.
1294: In each case, $Re$ signifies the real part of the transform.
1295:
1296:
1297:
1298: One can follow a similar procedure using the momentum equation to
1299: determine the evolution of the kinetic energy density in Fourier space.
1300: As described previously, we include the density in our Fourier transforms.
1301: Consider the time derivative of $\sr{\bm v}$ given by
1302:
1303: \begin{equation}
1304: \label{rho_v_deriv1}
1305: \frac{\partial\sr{\bm v}}{\partial t} = \sr \frac{\partial{\bm v}}{\partial t} + \frac{{\bm v}}{2\sr}\frac{\partial \rho}{\partial t}.
1306: \end{equation}
1307:
1308: \noindent
1309: Note that here, for simplicity, we do not decompose the velocity
1310: into mean and turbulent components. Using a combination of the continuity
1311: and momentum equations, this equation can be written as
1312:
1313: \begin{eqnarray}
1314: \label{rho_v_deriv2}
1315: \frac{\partial\sr{\bm v}}{\partial t} & = & \sr \left[-{\bm v} \cdot \del{\bm v} - \frac{1}{\rho}\del (P + \frac{1}{2}B^2) + \frac{1}{\rho}({\bm B} \cdot \del {\bm B}) - 2 {\bm \Omega} \times {\bm v} + 2 q \Omega^2 x \hat{{\bm x}} \right] \nonumber \\
1316: & & + \frac{{\bm v}}{2 \sr} \left[-\rho(\del \cdot {\bm v}) - {\bm v} \cdot \del \rho \right],
1317: \end{eqnarray}
1318:
1319: \noindent
1320: If we take the Fourier transform of this equation and dot the
1321: result with the complex conjugate of
1322:
1323: \begin{equation}
1324: \label{vfour}
1325: \vfour = \int \int \int \sr({\bm x}) {\bm v}({\bm x}) \fterm,
1326: \end{equation}
1327:
1328: \noindent
1329: we arrive at
1330:
1331: \begin{equation}
1332: \label{vfour_evolution}
1333: \frac{1}{2}\frac{\partial|\vfour|^2}{\partial t} = \tvv + \tcomp + \tvb + \tpress + \tcor + \tphi + \dkin,
1334: \end{equation}
1335:
1336: \noindent
1337: where
1338:
1339: \begin{equation}
1340: \label{kin_fourier}
1341: \frac{1}{2}|\vfour|^2 \equiv \frac{1}{2}\left[|\widetilde{\sqrt{\rho}v_x({\bm k})}|^2+|\widetilde{\sqrt{\rho}v_y({\bm k})}|^2+|\widetilde{\sqrt{\rho}v_z}({\bm k})|^2\right],
1342: \end{equation}
1343:
1344: \begin{equation}
1345: \label{t_vv}
1346: \tvv = -Re \left[\vstar \cdot \int \int \int [\sr({\bm v} \cdot \del){\bm v} + \frac{{\bm v}}{2 \sr}({\bm v} \cdot \del)\rho] \fterm \right],
1347: \end{equation}
1348:
1349:
1350: \begin{equation}
1351: \label{t_comp}
1352: \tcomp = -Re \left[\vstar \cdot \int \int \int \frac{\sr{\bm v}}{2} (\del \cdot {\bm v}) \fterm \right],
1353: \end{equation}
1354:
1355: \begin{equation}
1356: \label{t_vb}
1357: \tvb = +Re \left[\vstar \cdot \int \int \int \frac{1}{\sr}({\bm B} \cdot \del) {\bm B} \fterm \right],
1358: \end{equation}
1359:
1360: \begin{equation}
1361: \label{t_press}
1362: \tpress = -Re \left[\vstar \cdot \int \int \int \frac{1}{\sr}\del(P + \frac{1}{2}B^2) \fterm \right],
1363: \end{equation}
1364:
1365: \begin{equation}
1366: \label{t_cor}
1367: \tcor = -Re \left[\vstar \cdot \int \int \int (2{\bm \Omega} \times \sr {\bm v}) \fterm \right],
1368: \end{equation}
1369:
1370: \begin{equation}
1371: \label{t_phi}
1372: \tphi = +Re \left[\widetilde{\sqrt{\rho}v_x^{\bm \ast}}({\bm k}) \cdot \int \int \int 2 \sr q \Omega^2 x \fterm \right],
1373: \end{equation}
1374:
1375: \noindent
1376: and $\dkin$ accounts for the dissipation of kinetic energy. Again,
1377: this dissipation is numerical as we have not included an explicit
1378: viscosity term in our equations. Equation~(\ref{vfour_evolution})
1379: describes the evolution of the kinetic energy density in Fourier space.
1380: $\tvv$ is a term that describes the transfer of kinetic energy between
1381: scales by the velocity field (both the mean and turbulent velocity),
1382: $\tcomp$ results from turbulent compressibility, $\tvb$ describes how
1383: kinetic energy changes from magnetic tension, $\tpress$ represents the
1384: effect of both gas and magnetic pressure on the kinetic energy, and $\tphi$
1385: is the effect of the tidal potential on the kinetic energy. Note that $\tcor$
1386: is analytically equal to zero, and it is not included in any of
1387: the following analysis or discussion.
1388:
1389:
1390:
1391: In the saturated state of the MRI, the magnetic and kinetic energy
1392: densities should be in a steady state on average (although they
1393: do show strong fluctuations over short periods of time). If we
1394: consider the time-averages of equations~(\ref{bfour_evolution}) and
1395: (\ref{vfour_evolution}), then we can set the left hand sides to zero.
1396: We then rewrite these equations as
1397:
1398: \begin{equation}
1399: \label{vfour_zero}
1400: \tvv + \tcomp + \tvb + \tpress + \tphi + \dkin = 0,
1401: \end{equation}
1402:
1403: \begin{equation}
1404: \label{bfour_zero}
1405: A + S + \tbb + \tdivv + \tbv + \dmag = 0,
1406: \end{equation}
1407:
1408: \noindent
1409: where each of these terms is now a time-average. Here we average
1410: over 161 snapshots from orbit 20 to 100 in increments of 0.5 orbits.
1411: Each of these terms is a function of $k_x, k_y,$ and $k_z$,
1412: and in what follows we average the terms on shells of constant $k = |{\bm
1413: k}|$ as was done in \cite{from07a}.\footnote{In our analysis, the average over shells of
1414: constant $k$ was done before the temporal average.} Note that unlike the averaging
1415: described in that paper, we include $k_y$ in the calculation of $k$.
1416:
1417: \subsection{Zero Net Magnetic Flux}
1418: \label{zero_net_flux}
1419:
1420: \subsubsection{Fiducial Run}
1421: \label{fid_z}
1422:
1423: In this section, we focus on the Fourier transfer
1424: functions for the fiducial zero net magnetic flux simulation.
1425: Figure~\ref{magtfz} plots the magnetic transfer functions
1426: defined in equations~(\ref{aterm})-(\ref{t_bv}) as a function of length
1427: scale for $\zone$, and Fig.~\ref{kintfz} plots the kinetic
1428: transfer functions defined in equations~(\ref{t_vv})-(\ref{t_phi}).
1429: The dashed lines correspond to plus or minus one standard deviation around
1430: the mean value of the time average. Most of the transfer functions show
1431: large variation at small $k$ values which may be due to poor statistics
1432: at small $k$ and relatively large time variability. Because the transfer
1433: functions approach zero rapidly, we plot the ranges $1 < k L/(2\pi) < 20$
1434: and $20 < k L/(2\pi) < 64$ in the same figure, but with different $y$ scalings.
1435:
1436:
1437:
1438: The shear term $S$ is positive at all scales, as observed in
1439: \cite{from07a}, meaning that $B_y$ is created by the shear flow at
1440: all scales. $A$ is small at all scales, supporting the assumption made
1441: in \cite{from07a} that $A \approx 0$. $\tbv$ is primarily negative
1442: at the largest scales, although there are large fluctuations, and
1443: becomes positive for $k L/(2\pi) \gtrsim 35$. The turbulent velocity
1444: fluctuations seem to be creating magnetic energy at the smallest scales,
1445: but at larger scales, the magnetic field appears to lose energy via this
1446: interaction with the turbulence. $\tbb$ is negative for $k$ smaller
1447: than $k L/(2\pi) \sim 20$, meaning that the turbulence is transferring
1448: magnetic energy away from these scales. Although this analysis doesn't
1449: determine the direction of this cascade, at the largest scale (i.e.,
1450: the box size) the energy
1451: can only cascade to smaller scales. In terms of absolute value, $S$
1452: and $\tbb$ are dominant on the largest scales, while on small scales,
1453: $\tbb > \tbv > S > 0$.
1454:
1455: It is difficult to say anything conclusive about the kinetic transfer
1456: functions on the largest scales as they are subject to considerable
1457: uncertainty, although $\tvb < 0$ appears reasonably well constrained
1458: at these scales. At smaller scales, the two dominant terms are $\tvv$
1459: and $\tvb$, with $\tvb > \tvv > 0$; kinetic energy is being transferred
1460: to these scales by the turbulence, and being created by magnetic field.
1461:
1462:
1463:
1464: Equations~(\ref{vfour_zero})~and~(\ref{bfour_zero}) have been set to
1465: zero from the assumption that the magnetic and kinetic energies are in a
1466: time-averaged steady state. The dissipation terms $\dmag$ and $\dkin$
1467: are simply what is left over after the other transfer functions have
1468: been computed. The top plots in Fig.~\ref{dis_z} are the kinetic and
1469: magnetic dissipation and the ratio $\dkin/\dmag$ as a function of $k$
1470: for $20 < k L/(2\pi) < 64$; the scatter at small $k$ is large and there
1471: is considerable uncertainty in the dissipation values. At small scales,
1472: magnetic dissipation dominates kinetic dissipation by a factor of roughly
1473: three. The kinetic and magnetic dissipation rate increases in magnitude
1474: towards larger scales.
1475:
1476:
1477:
1478: Following \cite{from07a}, we can determine an effective resistivity and
1479: viscosity as a function of length scale by assuming that the numerical
1480: effects behave as if they were physical resistivity and viscosity.
1481: For example, with a constant Ohmic resistivity, the induction equation
1482: would have an additional term proportional to $\nabla^2 B$, with the
1483: constant of proportionality being the resistivity. If we take the Fourier
1484: transform of this term and dot it with the complex conjugate of $\bfour$,
1485: the real part is
1486:
1487: \begin{equation}
1488: \label{t_eta}
1489: \teta = +Re \left[\bstar \cdot \int \int \int \nabla^2{\bm B} \fterm \right] = -k^2|\bfour|^2.
1490: \end{equation}
1491:
1492: \noindent
1493: We can then define an effective resistivity as a function of $k$ by
1494:
1495: \begin{equation}
1496: \label{eta_eff}
1497: \etaeff(k) \equiv \frac{\dmag(k)}{\teta(k)}.
1498: \end{equation}
1499:
1500: Similarly, a constant kinematic shear viscosity would add a term
1501: proportional to \\ $\sr [\nabla^2{\bm v} + \frac{1}{3}\del(\del \cdot {\bm v})]$ to equation~(\ref{rho_v_deriv2}), with the constant of
1502: proportionality being the viscosity. Note that we only consider shear
1503: viscosity here for simplicity. We take the Fourier transform of the
1504: viscous term, dot it with the complex conjugate of equation~(\ref{vfour}),
1505: and take the real part. The result is
1506:
1507: \begin{equation}
1508: \label{t_nu_first}
1509: \tnu = +Re \left[\vstar \cdot \int \int \int \sr [\nabla^2{\bm v} + \frac{1}{3}\del(\del \cdot {\bm v})] \fterm \right].
1510: \end{equation}
1511:
1512: \noindent
1513: This equation can be made simpler by realizing that the
1514: second term of the integrand, related to the divergence of ${\bm v}$,
1515: is negligible. We can also assume that the density is relatively
1516: constant, and arrive at
1517:
1518: \begin{equation}
1519: \label{t_nu}
1520: \tnu = -k^2|\vpfour|^2 .
1521: \end{equation}
1522:
1523: \noindent
1524: We have substituted the perturbed velocity here because it
1525: is the only velocity that can lead to {\it numerical} dissipation of
1526: kinetic energy. That is, a pure shear flow will not encounter any numerical
1527: viscosity, and we can subtract off this flow. We define an effective
1528: viscosity by
1529:
1530: \begin{equation}
1531: \label{nu_eff}
1532: \nueff(k) \equiv \frac{\dkin(k)}{\tnu(k)}.
1533: \end{equation}
1534:
1535: We can also characterize the effective
1536: resistivity and viscosity in terms of a Reynolds number,
1537:
1538: \begin{equation}
1539: \label{re}
1540: \reeff(k) \equiv \frac{c_o H}{\nueff(k)},
1541: \end{equation}
1542:
1543: \noindent
1544: and magnetic Reynolds number,
1545:
1546: \begin{equation}
1547: \label{rm}
1548: \rmeff(k) \equiv \frac{c_o H}{\etaeff(k)},
1549: \end{equation}
1550:
1551: \noindent
1552: where we have used the initial isothermal sound speed, $c_o = 0.001$, as
1553: a characteristic velocity, and $H = L_z$ is a characteristic length.
1554: These numbers quantify the numerical dissipation coefficients
1555: in a dimensionless manner.
1556:
1557:
1558: Finally, we define an effective Prandtl number by
1559:
1560: \begin{equation}
1561: \label{pmeff}
1562: \pmeff(k) \equiv \frac{\nueff(k)}{\etaeff(k)}
1563: \end{equation}
1564:
1565:
1566: The effective viscosity and resistivity as well as the effective Prandtl
1567: number are shown in the bottom plots of Fig.~\ref{dis_z}.
1568: The viscosity and resistivity are fairly constant at large $k$.
1569: The effective Reynolds numbers are on the order of $\reeff \sim 12000$, and $\rmeff \sim 20000$ at large $k$.
1570: The Prandtl number is also relatively flat at these scales, and $\pmeff
1571: \sim 1.6$. This result agrees with \cite{from07a}, where $\pmeff >
1572: 1$ for ZEUS. While the numerical dissipation of $\at$ is not physical,
1573: the ``flatness" of $\nueff$ and $\etaeff$ suggests a resemblance to
1574: physical dissipation at small scales.
1575:
1576: Finally, note that although the Prandtl number is greater than unity,
1577: the magnetic dissipation dominates over kinetic dissipation. Evidently,
1578: $\teta$ is larger than $\tnu$ because there is more magnetic energy than
1579: kinetic energy at a given scale. In particular,
1580:
1581: \begin{equation}
1582: \label{energy_ratio1}
1583: \frac{\teta}{\tnu} = \frac{|\bfour|^2}{|\vpfour|^2}.
1584: \end{equation}
1585:
1586: \noindent
1587: Since there is more magnetic energy than perturbed kinetic energy at a
1588: given scale, magnetic dissipation dominates.
1589:
1590: \subsubsection{Resolution Effects}
1591: \label{res_effects_z}
1592:
1593: To gauge the effect of resolution on these various quantities, we
1594: perform the same analysis on the lower resolution runs, $\zonesix$,
1595: $\zthree$, and $\zsix$. We focus, in particular, on the small scales
1596: (i.e., large $k$) where our quantities are statistically more well-determined.
1597: Figure~\ref{nu_eta_res_z} shows $\nueff$, $\etaeff$, $\pmeff$, and the
1598: ratio of $\dkin$ to $\dmag$ as a function of $x$ resolution, $N_x$. The
1599: data points are calculated by averaging the quantity of interest over $k$
1600: in the regions of $k$-space where the error on the quantity is less than
1601: its mean value.\footnote{There are some quantities for which the error is
1602: never less than the mean. In these cases, we average over regions where
1603: the mean is greater than 80\% of the error.} The displayed error bars are
1604: the propagation of the errors from the temporal statistics. At these
1605: large values of $k$, $\nueff$, $\etaeff$, $\pmeff$, and $\dkin/\dmag$
1606: are relatively flat, varying by a factor of at most 2. Consequently,
1607: these averages should be representative at small scales.
1608:
1609:
1610:
1611: The numerical viscosity and resistivity decrease as a function of
1612: resolution. The dashed lines in the two upper panels of the figure
1613: show the line $\nueff, \etaeff \propto N_x^{-2}$. The viscosity and
1614: resistivity decrease slower than this with increasing $N_x$; we measured
1615: $\nueff, \etaeff \propto N_x^{-1.6}$. The figure also shows that both the
1616: effective Prandtl number and the ratio of kinetic to magnetic dissipation
1617: are constant with resolution to within the error bars.
1618:
1619: \subsubsection{Comparison with Previous Results}
1620: \label{trans_compare_section}
1621:
1622: \cite{from07a} were interested in the transfer function for the poloidal
1623: field, as the regeneration of this field is key to a self-sustaining
1624: dynamo. They found that the magnetic dissipation of ZEUS for the
1625: poloidal magnetic field departs from the physical dissipation model
1626: at small $k$ and could even be a nonphysical ``positive" dissipation.
1627: We repeat the same analysis as performed in that paper, but with $\zone$,
1628: for comparison. First, we examine the magnetic dissipation for the full
1629: 3D Fourier analysis described above. Second, we do the same procedure but
1630: setting $B_y = 0$ to focus on the effect of only including poloidal field.
1631: Finally, we perform the procedure with $B_y = 0$ and in the plane $k_y
1632: = 0$ (i.e., axisymmetry). These simplifications allow us to reproduce
1633: the poloidal field analysis of \cite{from07a}.
1634:
1635:
1636:
1637: The results are shown in Fig.~\ref{trans_compare}. The left two plots
1638: correspond to the Fourier analysis in which only $B_y = 0$ is assumed.
1639: The right plots assume $B_y = 0$ and $k_y = 0$. The black lines in the
1640: bottom two plots correspond to the magnetic dissipation for the full 3D
1641: Fourier analysis with $B_y \neq 0$ and $k_y \neq 0$. It is apparent that
1642: when $B_y = 0$ is assumed in the calculations, the magnetic dissipation
1643: becomes positive at large scales. However, when $B_y$ is included, the
1644: magnetic dissipation remains negative. Whether or not $k_y = 0$ is
1645: assumed seems to make very little difference, supporting the notion that
1646: small $k_y$ dominates. Since $\at$ and ZEUS both find positive $\dmag$ at
1647: small $k$, it is unlikely that this effect can be attributed to algorithmic
1648: limitations specific to ZEUS. Since $\dmag$ is not a derived quantity but
1649: simply what remains after all the transfer functions are calculated,
1650: it seems likely that the positive $\dmag$ values for the poloidal field analysis
1651: are due to incomplete statistics at large scales, or other inadequacies of the
1652: analysis when applied solely
1653: to the poloidal field. At small $k$, the standard deviations of the
1654: quantities (dashed lines) are considerable. The standard deviation on
1655: $\dmag$ when $B_y \neq 0$ is significantly larger than when one sets
1656: $B_y = 0$. This reflects the large variability of $\langle B_y^2/2
1657: \rangle$ compared to the other components of magnetic energy (see e.g.,
1658: Fig.~\ref{turb_z}). At any given time, $\dmag$ can be positive;
1659: the assumption of time-stationarity does not hold at any point in time.
1660: But when the data are time-averaged, $\dmag < 0$.
1661:
1662:
1663: Finally, we compare the numerical magnetic Reyolds number calculated with
1664: equation~(\ref{rm}) but with the $B_y = 0$ and $k_y = 0$ assumptions.
1665: For $\zone$, we find that $\rmeff \sim$ 11000, and for $\zsix$, $\rmeff
1666: \sim$ 3500. \cite{from07a} find $\rmeff \sim$ 30000 for their $N_x =
1667: 128$ run, and $\rmeff \sim$ 10000 for their $N_x = 64$ run; both of their
1668: calculated effective Reynolds numbers are larger than those calculated
1669: for $\at$. This result seems to suggest that ZEUS is actually less
1670: dissipative than $\at$. However, there are several points to consider.
1671: First, numerical dissipation is a nonlinear function of resolution,
1672: sharply increasing as the number of zones per wavelength decreases
1673: (high wavenumbers). The effective Reynolds number is obtained by
1674: measuring dissipation at the high $k$ end of the spectrum. As reported
1675: by \cite{shen06} $\at$ appears to have higher dissipation than ZEUS
1676: for poorly resolved waves, as evidenced by the ability of $\at$ to
1677: avoid the aliasing errors seen with ZEUS for hydrodynamic shearing
1678: box waves. They further point out that for wavelengths larger than
1679: 16 grid points $\at$ is less dissipative. Further, 2D simulations of
1680: decaying turbulence have demonstrated that when saturation amplitude is reached,
1681: the decay time is longer in $\at$ than in ZEUS, consistent with
1682: $\at$ having a higher effective resolution \cite[]{stone05}.
1683: In the present context, we find that the time- and volume-averaged total
1684: stresses in our simulations are larger than those calculated in the
1685: simulations of \cite{from07a}. Stronger turbulence leads to larger kinetic
1686: and magnetic turbulent fluctuations, which in turn enhances dissipation
1687: via grid-scale effects. Finally, we reemphasize that assuming $B_y = 0$
1688: may have a significant impact on the measurement of effective magnetic
1689: dissipation via this analysis.
1690:
1691: \subsection{Net Magnetic Flux}
1692: \label{net_flux}
1693:
1694: \subsubsection{Fiducial Run}
1695: \label{fid_n}
1696:
1697: We perform the same transfer function analysis on the
1698: fiducial net magnetic flux run, $\none$. The various
1699: transfer function terms as a function of $k$ are shown in
1700: Figs.~\ref{magtfn}--\ref{kintfn}. As was the
1701: case in the zero net flux simulation, $S$ is positive at all scales
1702: and dominates at small $k$; $A$ is relatively small throughout. $\tbv$
1703: and $\tbb$ are negative at large scales and positive at small scales,
1704: with $\tbb > 0$ for $k L/(2\pi) \gtrsim 5$, and $\tbv >0$ for $k L/(2\pi)
1705: \gtrsim 20$. At small scales, $\tbb > \tbv > S > 0$. Of the kinetic
1706: terms, $\tvv$ and $\tvb$ dominate with $\tvb > \tvv > 0$. These results
1707: are in general agreement with $\zone$, except that the magnitude of
1708: the various terms is larger for $\none$ than for $\zone$, and $\tbb$
1709: and $\tbv$ become positive at smaller $k$ values compared to $\zone$.
1710:
1711:
1712:
1713: As before, we calculate the kinetic and magnetic dissipation as well as
1714: effective values for the viscosity and resistivity. Figure~\ref{dis_n}
1715: shows these quantities for $\none$ at the smallest scales. As was the
1716: case for $\zone$, the mean magnetic dissipation dominates over kinetic
1717: dissipation by a factor of roughly three at these scales. Note, however,
1718: the large error bars associated with these plots, which encompass values
1719: of $\dkin/\dmag > 1$. Again, the error bars are the temporal
1720: standard deviation of the transfer functions.
1721: Since $\none$ has a larger temporal variability, larger error bars
1722: are expected. The mean value for $\dkin/\dmag$ is on the order of
1723: 0.6-0.7, which is consistent with the analysis in \S~\ref{sustained} in which
1724: we found $\qk/\qm \sim 0.6$.
1725:
1726:
1727:
1728: The effective viscosity and resistivity show the same basic result as
1729: in the $\zone$ case. $\nueff$, $\etaeff$, and $\pmeff$ change by a
1730: factor of order unity at large $k$. The effective
1731: Reynolds numbers are on the order of $\reeff \sim 4000$, and $\rmeff \sim 8000$ at large $k$.
1732: $\pmeff$ has a mean value of $\sim 1.9$. Again, there is considerable uncertainty in these values
1733: due to the large amplitude fluctuations in the turbulence. The error
1734: bars encompass values of $\pmeff$ less than unity. As a result, it
1735: is more difficult to conclusively say that the dissipation behaves the
1736: same way in $\none$ as in $\zone$. However, in an average sense, the
1737: two simulations agree well qualitatively.
1738:
1739: \subsubsection{Resolution Effects}
1740: \label{res_effects_n}
1741:
1742: We can again look at the effect of resolution on these various dissipation
1743: quantities. Figure~\ref{nu_eta_res_n} shows this effect for the net flux
1744: simulations ($\nonesix$, $\nthree$, $\nsix$, and $\none$). The procedure
1745: by which to average over $k$ is the same as described in
1746: \S~\ref{res_effects_z}. The displayed error bars are the propagation of
1747: the errors from the temporal statistics. At these large values of $k$,
1748: $\nueff$, $\etaeff$, $\pmeff$, and $\dkin/\dmag$ are relatively flat,
1749: varying by a factor of at most 2.
1750:
1751:
1752:
1753: The numerical viscosity and resistivity decrease as a function of
1754: resolution. The dashed lines in the two upper panels of the figure
1755: show the line $\nueff, \etaeff \propto N_x^{-2}$. The viscosity and
1756: resistivity decrease slower than this with increasing $N_x$; we measured
1757: $\nueff, \etaeff \propto N_x^{-1.3}$. The figure shows that the effective
1758: Prandtl number is constant with resolution to within the error bars. There
1759: appears to be a slight increase in $\dkin/\dmag$ with resolution, but
1760: this trend is not definitive given the large uncertainties on the data.
1761:
1762:
1763:
1764: One might expect $\nueff$ and $\etaeff$ to decrease with increasing
1765: resolution since these terms arise from truncation error. Linear wave
1766: advection test problems with $\at$ have shown that the truncation error
1767: converges at second order \cite[e.g.,][]{stone08}. On this basis, one
1768: would expect $\nueff, \etaeff \propto N_x^{-2}$. We find a shallower
1769: decrease with $N_x$, but MRI turbulence is a fully nonlinear system,
1770: and one should not necessarily expect the same convergence behavior as
1771: in a linear system.
1772:
1773:
1774: \section{Discussion and Conclusions}
1775: \label{conclusions}
1776:
1777: We have carried out a series of local, unstratified shearing box
1778: simulations with the recently developed $\at$ code to study the characteristics
1779: of MRI driven turbulence. $\at$ uses a second-order,
1780: conservative, compressive MHD algorithm, which is significantly different
1781: from the algorithms employed in many of the previous MRI studies.
1782: In our work, we have run several standard models for comparison with
1783: previous work, and characterized the numerical dissipation of the $\at$
1784: code for the shearing box problem. Furthermore, we have exploited the energy
1785: conservation property of $\at$ to carry out a study of energy flow within
1786: MRI-driven turbulence.
1787:
1788: To compare with previous numerical results, we have investigated the
1789: effects of different initial field geometries (uniform or sinusoidal
1790: $B_z$), varying domain aspect ratio, and numerical resolution. In all
1791: of our simulations, the MRI is initiated and sustained over many orbits.
1792: The time- and volume-averaged properties of the resulting turbulent flow,
1793: such as stress levels and magnetic and kinetic energies, are consistent
1794: with previous results. As in previous work, we find that boxes containing
1795: net vertical field saturate at higher amplitudes compared to those
1796: without net fields. The total stress is proportional to the magnetic
1797: pressure with a constant of proportionality $\sim 0.5$, but is independent of the
1798: gas pressure. In the net field simulation, the gas pressure increases by
1799: a factor of 100, due to thermalization of the turbulence, without affecting
1800: the stress. The consistency of these results with past work indicate that
1801: these properties do not result from details of the employed algorithm.
1802:
1803: Fourier analysis of the turbulence shows that the largest scales in the
1804: box dominate the energetics. In the presence of a net field, the amplitude
1805: of the spatial power spectra is largely independent of
1806: resolution on the largest scales. This is not true for the zero net
1807: flux simulations however. For those simulations, the amplitude decreases as resolution increases,
1808: which is consistent with the overall resolution behavior.
1809: For net field simulations, the averaged turbulent magnetic and kinetic
1810: energies increase slightly with resolution, whereas for the zero net field
1811: simulations, the energies decrease with increasing resolution roughly in
1812: proportion to the grid zone size. This apparent lack of convergence for
1813: the zero net field shearing box simulations was previously demonstrated
1814: by \cite{from07a} using the ZEUS code.
1815:
1816: The net field simulation shows intermittent channel flows which cause
1817: temporary increases in stress through amplification of large-scale
1818: MRI modes. The parasitic modes described by \cite{good94} destroy the
1819: channel flow within about one orbit of time, but the rapid increase in
1820: stress produces a subsequent increase in thermal energy. The presence
1821: of these discrete channel flow events is a consequence of the box
1822: size---larger boxes do not experience
1823: them---but we use their presence to study the subsequent energy flow
1824: following a rapid increase in stress.
1825:
1826: Because $\at$ evolves the total energy equation, magnetic and kinetic energy
1827: losses due to numerical grid-scale effects are added to the internal
1828: energy. This makes $\at$ well suited to examining the turbulent energy flow and
1829: subsequent dissipation. The recurring channel flows
1830: in the net flux model provide a sudden injection of energy into the
1831: box by increasing the stress operating on the shearing boundaries of
1832: the box. The injected energy appears as heat after $\sim 0.2$ orbits.
1833: This corresponds to a timescale $\Omega^{-1}$, which equals $L_z/c_{\rm s}$
1834: where $c_{\rm s}$ is the initial soundspeed. This timescale determines the amplitude
1835: of the $\alf$ speed, $v_{\rm A}$, and its fundamental MRI wavelength, $\lambda_{\rm MRI}$; $L_z/c_{\rm s} \sim
1836: \lambda_{\rm MRI}/v_{\rm A}$. The timescale is thus on the order of the eddy turnover time, indicating that
1837: dissipational heating is a local process and that energy is not carried
1838: over large distances before it is thermalized.
1839:
1840: In the fiducial zero net magnetic flux simulation, $\zone$, there are no
1841: recurring channel modes, making it more difficult to trace the flow of
1842: injected energy. The analysis is further complicated by the presence
1843: of compressive waves that dominate the time derivative of the thermal
1844: energy, $\td$. These waves are also present in the net field simulations,
1845: but their amplitude is smaller relative to the larger turbulent kinetic
1846: energy found with a net field. A detailed examination of the components
1847: of the internal energy equation indicate that the compressive waves do not
1848: appear to contribute significantly to irreversible heating. By averaging
1849: $\td$ for the zero net flux simulation, we find a correlation of $\td$
1850: with $\ein$ on the same timescale of $\sim$ 0.2~orbits.
1851:
1852: In the net field simulation, the dissipation of magnetic energy is larger
1853: than that for the kinetic energy, not unexpected as the ratio of the average
1854: magnetic to perturbed kinetic energy is $\sim 3.4$. But the ratio of the magnetic
1855: to kinetic dissipation rate is roughly constant at $\sim 1.7$. The fact that the
1856: ratio of dissipation rates does not equal the ratio of energies
1857: may result from a couple of possibilities. First, there could be a
1858: net transfer of magnetic to {\it perturbed} kinetic energy as was
1859: suggested in \cite{bran95}.\footnote{\S~\ref{sustained} shows that
1860: there is in fact a net transfer of kinetic to magnetic energy. However,
1861: this kinetic energy includes the shear flow, and thus, this result
1862: tells us nothing of the energy transfer between magnetic and {\it perturbed} kinetic energy.}
1863: Second, the difference in the ratios could arise from the effective Prandtl number
1864: being larger than one. In particular, if $\qk \propto \nueff \delta v^2/2$ and
1865: $\qm \propto \etaeff B^2/2$, then $(B^2/\delta v^2)(\qk/\qm) \sim \pmeff$. With the above
1866: values for the energy and dissipation ratios, we find $(B^2/\delta v^2)(\qk/\qm) \sim 2$,
1867: which is consistent with the determination of $\pmeff$ from the Fourier analysis (see
1868: discussion below). The agreement
1869: between the two separate calculations of $\pmeff$ may be coincidental, but it is suggestive
1870: of $\qk \propto \nueff \delta v^2/2$ and $\qm \propto \etaeff B^2/2$.
1871:
1872: The turbulence is sustained by the continued action of the MRI
1873: in extracting energy from the differential rotation. This can
1874: be removed from the simulations allowing us to study the decay
1875: of the turbulence in detail (simulations $\ndone$ and $\zdone$).
1876: Figure~\ref{turb_decay} shows that magnetic
1877: losses dominate over kinetic losses during this decay. In both
1878: simulations nearly 50\% of the magnetic energy and 20\% of the kinetic
1879: energy has been dissipated after 0.2~orbits. By one orbit into the decay,
1880: most of the magnetic and kinetic energy has been lost. Although these
1881: decay timescales arise in a turbulent flow that lacks power input from
1882: the MRI, the results are consistent with the conclusion that turbulent
1883: energy dissipation occurs on a rapid timescale of order $\Omega^{-1}$.
1884:
1885: \cite{from07a} used a detailed Fourier analysis
1886: (\S~\ref{trans_funcs}) to study magnetic energy flow and thermalization
1887: as a function of length scale in the shearing box. In this analysis, the
1888: individual terms in the evolution equation for the magnetic energy are
1889: examined in Fourier space. Averaging over time and assuming that the
1890: magnetic energy is in a statistical steady state, one sets the sum of
1891: these terms equal to a remainder, which is credited to numerical effects.
1892: These numerical losses can then be modeled as an effective resistivity
1893: (and viscosity for the kinetic energy), allowing one to characterize
1894: the numerical dissipation in the simulation.
1895:
1896: We repeated their analysis with $\at$ and extended it to the kinetic
1897: energy. The dominant effect at large scales is the generation of
1898: magnetic field by the background shear. This energy is transferred to
1899: other scales by the turbulence. Net positive field creation
1900: by the turbulent flow and energy gains by the transfer between scales only
1901: happens at small wavelengths. This point of transition from loss to gain
1902: happens at smaller scales for the zero net field simulation compared to the
1903: net field model. Magnetic dissipation dominates over kinetic dissipation
1904: at small scales (i.e., $k L/(2\pi) \gtrsim 20$). Modeling these as
1905: an effective resistivity $\eta$ and viscosity $\nu$ shows that $\eta$
1906: and $\nu$ drop with increasing resolution with a power that lies between
1907: first- and second-order in grid resolution. The effective Prandtl number,
1908: on the other hand, is nearly constant as a function of resolution with
1909: a value between $\sim 1.5$ and 2.
1910:
1911: \cite{from07a} observed what they described as ``negative" resistivity
1912: in an analysis restricted to the poloidal field alone. In repeating their
1913: exact analysis with $\at$, we also observed such an ``anti-dissipation"
1914: at large scales. This indicates that this effect is not associated with
1915: a numerical algorithm limitation associated with ZEUS. More likely,
1916: it arises from the statistical uncertainty at large scales and from
1917: the failure of the assumptions that go into the definition of the
1918: dissipation term. We note that the inclusion of the toroidal field $B_y$
1919: in the analysis shows net dissipation at all scales, although again the
1920: statistical variation is large at large scales.
1921:
1922: In conclusion, what do these results imply for shearing box simulations
1923: and the MRI? First, as observed by \cite{from07a}, the scales over
1924: which turbulent energy generation occurs are not well-separated from
1925: those where there is significant dissipation; the MRI operates over a
1926: wide range of scales. The MRI grows at a rate $\sim k v_{\rm A}$ for all $k$
1927: less than $\Omega/v_{\rm A}$. At large scales, a weak field will grow more
1928: slowly than the timescale over which energy is transferred between
1929: scales, between magnetic and kinetic forms, and ultimately thermalized. If a field is
1930: chopped up by reconnection, it may be reduced to small scales where the
1931: MRI no longer operates. In the presence of a net field, there will always
1932: be a significant driving term at the scales set by that imposed field.
1933: In the absence of such a field, however, the outcome will be determined
1934: by the complex interplay of loss due to dissipation and amplification by
1935: the MRI. In the numerical simulations with zero net field, increasing
1936: the resolution causes an overall decrease in the saturation energies.
1937: \cite{from07a} attribute this to higher resolution enabling the MRI
1938: to operate at intermediate scales which facilitates the transfer
1939: of energy to small scales and promotes reconnection and dissipation.
1940: What is perhaps surprising is that resolving the MRI at these scales leads
1941: to greater field dissipation than would otherwise be accomplished by the
1942: numerical losses that would occur if those scales were underresolved.
1943: Because the same effect is observed with both $\at$ and ZEUS, it seems
1944: likely that this ability of the MRI to transfer energy away from the
1945: largest scales in the shearing box and to increase the total dissipation
1946: is a physical rather than numerical effect.
1947:
1948: In related work, \cite{from07b} and \cite{lesur07} studied the effect of
1949: varying the physical (not numerical) magnetic Prandtl number, $P_m$, on
1950: the turbulence. They found that the saturation amplitudes were increased
1951: with increased $P_m$. \cite{from07b} found evidence that there exists
1952: a critical $P_m > 1$ below which zero net field simulations would die
1953: out rather than achieve a steady turbulent state. Our results in this
1954: investigation show that this Prandtl number dependence is a distinct
1955: effect from the observed dependence of the turbulence on resolution.
1956: We find the numerical $P_m$ to be largely independent of resolution
1957: in $\at$. Taken together, however, the dependence on physical $P_m$
1958: and the dependence on resolution point to the importance of small and
1959: intermediate scale magnetic dissipation and reconnection to establishing
1960: saturation amplitudes in MRI-driven turbulence.
1961:
1962: As discussed by \cite{from07a}, numerical dissipation can deviate
1963: significantly from physical dissipation. In \S~\ref{fid_z}, we showed
1964: that $\etaeff$ and $\nueff$ are relatively flat at small scales,
1965: suggesting a resemblance to physical dissipation. However, consider the
1966: numerical Reynolds number as calculated from equation~(\ref{re}) for our
1967: zero net flux simulations. For $N_x$ = 128, we found $\reeff \sim$ 12000,
1968: and $\pmeff \sim$ 1.6 for all of zero net flux simulations. From the
1969: parameter space studies of \cite{from07b}, these values for the Reynolds
1970: and Prandtl numbers correspond to marginal MRI turublence; that is,
1971: they lie very close to the critical line between sustained and decaying
1972: turbulence. For $N_x$ = 64, $\reeff \sim$ 4100, and the Reynolds number
1973: is even smaller for the lower resolutions. These values are well within
1974: the decaying turbulence regime, but we find active MRI turbulence in all
1975: of our simulations. These results show that the effective Reynolds and
1976: Prandtl numbers of $\at$ as measured at large wavenumbers does not apply
1977: at smaller $k$ values where there are many grid zones per wavelength.
1978: Thus, the Reynolds numbers and Prandtl numbers that we calculate should
1979: be taken as a measure of the effective numerical dissipation of the code
1980: and not equated to a flow with the same Reynolds and Prandtl number as
1981: determined by a simple physical resistivity and viscosity.
1982:
1983: This result highlights an uncertainty associated with any MRI simulation
1984: that depends only on numerical rather than physical dissipation.
1985: It is apparent that the numerical Prandtl number can play an important
1986: role in determining the ratio of magnetic to kinetic dissipation.
1987: More speculatively, the Prandtl number may also play a role in the
1988: timescale over which thermalization occurs. In the present study, we
1989: found that both the thermalization timescale and the effective numerical
1990: Prandtl number were largely independent of resolution. However, the
1991: turbulent energy thermalization timescales and properties we measure may
1992: be subject to change when explicit dissipation is included. It will be
1993: a very important next step in this work to include physical dissipation
1994: and verify these results.
1995:
1996: This work is only the first step in applying $\at$
1997: to the problem of the energetics of MRI turbulence. The present
1998: study provides a calibration of the numerical dissipation, which will
1999: be important in future studies that include explicit resistivity and
2000: viscosity. Furthermore, the unstratified shearing box has the virtue
2001: of simplicity and allows a detailed study of MRI turbulence without
2002: too many confounding factors, but it also may prove too limited for
2003: predictive application to accretion flows. The inclusion of vertical
2004: stratification and radiative cooling are both straightforward extensions
2005: to the present study. The detailed diagnostics developed and applied
2006: in this study should prove valuable in this planned work.
2007:
2008:
2009: \acknowledgments
2010:
2011: We thank Jim Stone, Steve Balbus, and Sebastien Fromang for useful discussions and suggestions
2012: regarding this work. We also thank the anonymous referee whose
2013: comments and suggestions improved this paper.
2014: This material is based upon work supported by NASA
2015: Grants NNG04GK77G and NAG5-13288. The simulations were run on the NCSA
2016: TeraGrid IA-64 Linux Cluster (Mercury) and the University of Virginia
2017: Astronomy Department Beowulf Cluster.
2018:
2019: \begin{thebibliography}{34}
2020: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
2021:
2022: \bibitem[{{Balbus} \& {Hawley}(1991)}]{balb91}
2023: {Balbus}, S.~A., \& {Hawley}, J.~F. 1991, \apj, 376, 214
2024:
2025: \bibitem[{{Balbus} \& {Hawley}(1998)}]{balb98}
2026: ---. 1998, Reviews of Modern Physics, 70, 1
2027:
2028: \bibitem[{{Balbus} \& {Papaloizou}(1999)}]{balb99}
2029: {Balbus}, S.~A., \& {Papaloizou}, J.~C.~B. 1999, \apj, 521, 650
2030:
2031: \bibitem[{{Beckwith} {et~al.}(2008){Beckwith}, {Hawley}, \& {Krolik}}]{beck08}
2032: {Beckwith}, K., {Hawley}, J.~F., \& {Krolik}, J.~H. 2008, \mnras, 1003
2033:
2034: \bibitem[{{Blackman} {et~al.}(2008){Blackman}, {Penna}, \&
2035: {Varni{\`e}re}}]{black08}
2036: {Blackman}, E.~G., {Penna}, R.~F., \& {Varni{\`e}re}, P. 2008, New Astronomy,
2037: 13, 244
2038:
2039: \bibitem[{{Blaes} {et~al.}(2007){Blaes}, {Hirose}, \& {Krolik}}]{bla07}
2040: {Blaes}, O., {Hirose}, S., \& {Krolik}, J.~H. 2007, \apj, 664, 1057
2041:
2042: \bibitem[{{Bodo} {et~al.}(2008){Bodo}, {Mignone}, {Cattaneo}, {Rossi}, \&
2043: {Ferrari}}]{bod08}
2044: {Bodo}, G., {Mignone}, A., {Cattaneo}, F., {Rossi}, P., \& {Ferrari}, A. 2008, \aap, 487, 1
2045:
2046: \bibitem[{{Brandenburg} {et~al.}(1995){Brandenburg}, {Nordlund}, {Stein}, \&
2047: {Torkelsson}}]{bran95}
2048: {Brandenburg}, A., {Nordlund}, A., {Stein}, R.~F., \& {Torkelsson}, U. 1995,
2049: \apj, 446, 741
2050:
2051: \bibitem[{{Colella} \& {Woodward}(1984)}]{col84}
2052: {Colella}, P., \& {Woodward}, P.~R. 1984, Journal of Computational Physics, 54, 174
2053:
2054: \bibitem[{{Colella}(1990)}]{col90}
2055: {Colella}, P. 1990, Journal of Computational Physics, 87, 171
2056:
2057: \bibitem[{{Fleming} {et~al.}(2000){Fleming}, {Stone}, \& {Hawley}}]{flem00}
2058: {Fleming}, T.~P., {Stone}, J.~M., \& {Hawley}, J.~F. 2000, \apj, 530, 464
2059:
2060: \bibitem[{{Fromang} \& {Papaloizou}(2007)}]{from07a}
2061: {Fromang}, S., \& {Papaloizou}, J. 2007, \aap, 476, 1113
2062:
2063: \bibitem[{{Fromang} {et~al.}(2007){Fromang}, {Papaloizou}, {Lesur}, \&
2064: {Heinemann}}]{from07b}
2065: {Fromang}, S., {Papaloizou}, J., {Lesur}, G., \& {Heinemann}, T. 2007, \aap,
2066: 476, 1123
2067:
2068: \bibitem[{{Gardiner} \& {Stone}(2005{\natexlab{a}})}]{gard05a}
2069: {Gardiner}, T.~A., \& {Stone}, J.~M. 2005{\natexlab{a}}, Journal of
2070: Computational Physics, 205, 509
2071:
2072: \bibitem[{{Gardiner} \& {Stone}(2005{\natexlab{b}})}]{gard05b}
2073: {Gardiner}, T.~A., \& {Stone}, J.~M. 2005{\natexlab{b}}, in American Institute
2074: of Physics Conference Series, Vol. 784, Magnetic Fields in the Universe: From
2075: Laboratory and Stars to Primordial Structures., ed. E.~M. {de Gouveia dal
2076: Pino}, G.~{Lugones}, \& A.~{Lazarian}, 475--488
2077:
2078: \bibitem[{{Gardiner} \& {Stone}(2008)}]{gard08}
2079: ---. 2008, Journal of Computational Physics, 227, 4123
2080:
2081: \bibitem[{{Goodman} \& {Xu}(1994)}]{good94}
2082: {Goodman}, J., \& {Xu}, G. 1994, \apj, 432, 213
2083:
2084: \bibitem[{{Hawley} \& {Balbus}(1992)}]{haw92}
2085: {Hawley}, J.~F., \& {Balbus}, S.~A. 1992, \apj, 400, 595
2086:
2087: \bibitem[{{Hawley} {et~al.}(1995){Hawley}, {Gammie}, \& {Balbus}}]{haw95}
2088: {Hawley}, J.~F., {Gammie}, C.~F., \& {Balbus}, S.~A. 1995, \apj, 440, 742
2089:
2090: \bibitem[{{Hawley} {et~al.}(1996){Hawley}, {Gammie}, \& {Balbus}}]{haw96}
2091: ---. 1996, \apj, 464, 690
2092:
2093: \bibitem[{{Hirose} {et~al.}(2006){Hirose}, {Krolik}, \& {Stone}}]{hir06}
2094: {Hirose}, S., {Krolik}, J.~H., \& {Stone}, J.~M. 2006, \apj, 640, 901
2095:
2096: \bibitem[{{Krolik} {et~al.}(2007){Krolik}, {Hirose}, \& {Blaes}}]{kro07}
2097: {Krolik}, J.~H., {Hirose}, S., \& {Blaes}, O. 2007, \apj, 664, 1045
2098:
2099: \bibitem[{{Lesur} \& {Longaretti}(2007)}]{lesur07}
2100: {Lesur}, G., \& {Longaretti}, P.-Y. 2007, \mnras, 378, 1471
2101:
2102: \bibitem[{{Pessah} {et~al.}(2007){Pessah}, {Chan}, \& {Psaltis}}]{pess07}
2103: {Pessah}, M.~E., {Chan}, C.-k., \& {Psaltis}, D. 2007, \apjl, 668, L51
2104:
2105: \bibitem[{{Sano} \& {Inutsuka}(2001)}]{sano01}
2106: {Sano}, T., \& {Inutsuka}, S.-i. 2001, \apjl, 561, L179
2107:
2108: \bibitem[{{Sano} {et~al.}(1998){Sano}, {Inutsuka}, \& {Miyama}}]{sano98}
2109: {Sano}, T., {Inutsuka}, S.-I., \& {Miyama}, S.~M. 1998, \apjl, 506, L57
2110:
2111: \bibitem[{{Sano} {et~al.}(2004){Sano}, {Inutsuka}, {Turner}, \&
2112: {Stone}}]{sano04}
2113: {Sano}, T., {Inutsuka}, S.-i., {Turner}, N.~J., \& {Stone}, J.~M. 2004, \apj,
2114: 605, 321
2115:
2116: \bibitem[{{Sano} \& {Stone}(2002)}]{sano02b}
2117: {Sano}, T., \& {Stone}, J.~M. 2002, \apj, 577, 534
2118:
2119: \bibitem[{{Shakura} \& {Syunyaev}(1973)}]{shak73}
2120: {Shakura}, N.~I., \& {Syunyaev}, R.~A. 1973, \aap, 24, 337
2121:
2122: \bibitem[{{Shen}, {Stone}, \& {Gardiner}(2006)}]{shen06}
2123: {Shen}, Y., {Stone}, J.~M. \& {Gardiner}, T.~A. 2006, \apj, 653, 513
2124:
2125: \bibitem[{{Stone} \& {Gardiner}(2005)}]{stone05}
2126: {Stone}, J.~M., \& {Gardiner}, T.~A. 2005, in American Institute of Physics
2127: Conference Series, Vol. 784, Magnetic Fields in the Universe: From Laboratory
2128: and Stars to Primordial Structures., ed. E.~M. {de Gouveia dal Pino},
2129: G.~{Lugones}, \& A.~{Lazarian}, 16--26
2130:
2131: \bibitem[{{Stone} {et~al.}(2008){Stone}, {Gardiner}, {Teuben}, {Hawley}, \&
2132: {Simon}}]{stone08}
2133: {Stone}, J.~M., {Gardiner}, T.~A., {Teuben}, P., {Hawley}, J.~F., \& {Simon},
2134: J.~B. 2008, \apjs, 178, 137
2135:
2136: \bibitem[{{Stone} {et~al.}(1996){Stone}, {Hawley}, {Gammie}, \&
2137: {Balbus}}]{stone96}
2138: {Stone}, J.~M., {Hawley}, J.~F., {Gammie}, C.~F., \& {Balbus}, S.~A. 1996,
2139: \apj, 463, 656
2140:
2141: \bibitem[{{Stone} \& {Norman}(1992{\natexlab{a}})}]{stone92a}
2142: {Stone}, J.~M., \& {Norman}, M.~L. 1992{\natexlab{a}}, \apjs, 80, 753
2143:
2144: \bibitem[{{Stone} \& {Norman}(1992{\natexlab{b}})}]{stone92b}
2145: ---. 1992{\natexlab{b}}, \apjs, 80, 791
2146:
2147: \bibitem[{{Ziegler} \& {R{\"u}diger}(2001)}]{zieg01}
2148: {Ziegler}, U., \& {R{\"u}diger}, G. 2001, \aap, 378, 668
2149:
2150: \end{thebibliography}
2151:
2152: \clearpage
2153: \begin{figure}
2154: \plotfiddle{f1.eps}{1in}{90.}{325}{475}{-10}{-100}
2155: \vspace{-1in}
2156: \caption{Volume-averaged energy densities and stresses normalized to the initial gas pressure versus time for the $\none$ simulation.
2157: In the upper two plots, the black line is the total energy density, the green line is the component of the energy density in the $x$ direction, the red line is the $y$ direction component,
2158: and the blue line is the $z$ direction component.
2159: The upper left plot shows the volume-averaged magnetic energy density, the upper right plot shows the perturbed kinetic energy density (i.e., with the shear subtracted off of $v_y$), and the lower left plot is the volume-averaged total stress (black), Maxwell stress (pink), and Reynolds stress (blue). The lower right plot is the total energy density, including gravitational energy (solid line), and the thermal energy density (dashed line). The $y$ axes have the same range for all plots except for the total/thermal energy density plot.
2160: \label{turb_n}}
2161: \end{figure}
2162:
2163: \begin{figure}
2164: \plotfiddle{f2.eps}{1in}{90.}{325}{475}{-10}{-100}
2165: \vspace{-1in}
2166: \caption{Volume-averaged energy densities and stresses normalized to the initial gas pressure versus time for the $\zone$ simulation.
2167: In the upper two plots, the black line is the total energy density, the green line is the component of the energy density in the $x$ direction, the red line is the $y$ direction component,
2168: and the blue line is the $z$ direction component.
2169: The upper left plot shows the volume-averaged magnetic energy density, the upper right plot shows the perturbed kinetic energy density (i.e., with the shear subtracted off of $v_y$), and the lower left plot is the volume-averaged total stress (black), Maxwell stress (pink), and Reynolds stress (blue). The lower right plot is the total energy density, including gravitational energy (solid line), and the thermal energy density (dashed line). The $y$ axes have the same range for all plots except for the total/thermal energy density plot.
2170: \label{turb_z}}
2171: \end{figure}
2172:
2173: \begin{figure}
2174: \plotfiddle{f3.eps}{1in}{90.}{325}{475}{-10}{-100}
2175: \vspace{-1in}
2176: \caption{Time- and volume-averaged energy densities normalized to the initial pressure for various resolutions.
2177: The two upper plots correspond to the net flux simulations, and the two lower plots correspond to the
2178: zero net flux simulations. The left plots are the averaged magnetic energy densities, and the right plots
2179: are the averaged perturbed kinetic energy densities (i.e., with shear subtracted off of $v_y$). In all plots, the
2180: black symbols are the total energy density, the green symbols are the $x$ component of the energy density, the
2181: red symbols are the $y$ component, and the blue symbols are the $z$ component.
2182: The time averages are done from orbit 20 to 100, and the error bars indicate one standard deviation
2183: over this period.
2184: \label{turb_res}}
2185: \end{figure}
2186:
2187: \begin{figure}
2188: \plotfiddle{f4.eps}{1in}{90.}{325}{475}{-10}{-100}
2189: \vspace{-1in}
2190: \caption{The development and destruction of a channel flow during the $\none$ simulation. The upper left plot shows a fluctuation in the volume-averaged magnetic energy density from $t~=~80$~orbits to $t~=~85$~orbits. The remaining plots show the $y$-averaged perturbed $y$ velocity (colors) and $v_x$ and $v_z$ (vectors).
2191: The upper right plot occurs at $t = 82.5$~orbits, the lower left plot occurs at $t = 83$~orbits, and the lower right plot occurs at $t = 84$~orbits. These times are indicated on the upper left plot by the arrows.
2192: At $t = 82.5$~orbits, one can see the development of a two-channel flow, in which one channel has $v_x < 0$ and $\delta v_y < 0$, and the other channel has $v_x > 0$ and $\delta v_y > 0$.
2193: At $t = 83$~orbits, this channel flow is even more developed as the perturbations to the $y$ velocity have
2194: become even stronger and $v_x$ dominates over $v_z$ everywhere. The development of this channel flow coincides with an increase in volume-averaged magnetic energy density. By $t = 84$~orbits, the channel flow has been destroyed, coinciding with the decrease in magnetic energy density.
2195: \label{channel}}
2196: \end{figure}
2197:
2198: \begin{figure}
2199: \plotfiddle{f5a.eps}{0in}{90.}{125}{175}{20}{-100}
2200: \vspace{-2.1in}
2201: \plotfiddle{f5b.eps}{0in}{90.}{125}{175}{255}{-100}
2202: \plotfiddle{f5c.eps}{0in}{90.}{125}{175}{20}{-100}
2203: \vspace{-2.1in}
2204: \plotfiddle{f5d.eps}{0in}{90.}{125}{175}{255}{-100}
2205: \plotfiddle{f5e.eps}{0in}{90.}{125}{175}{20}{-100}
2206: \vspace{-2.1in}
2207: \plotfiddle{f5f.eps}{0in}{90.}{125}{175}{255}{-100}
2208: \caption{Spatial power spectra of various energy densities in the saturated state of the standard net flux (left panels) and zero net flux simulations (right panels). The spectra were obtained via an average over 161 frames in the saturated state and an average over shells of constant modulus $|{\bm k}|$. In each column, the first plot shows magnetic energy density, the second shows kinetic energy density, and the third shows perturbed kinetic energy density (as defined in the text). The effect of resolution is shown in each individual plot; the dotted line corresponds to the resolution with $N_x = 16$, the dot-dashed line corresponds to $N_x = 32$, the dashed line corresponds to $N_x = 64$, and the solid line corresponds to $N_x = 128$.
2209: All energy densities have been normalized to the initial gas pressure and are plotted against a dimensionless wave number ($L$ is the length of the smallest dimension of the box).
2210: \label{power_spec}}
2211: \end{figure}
2212:
2213: \begin{figure}
2214: \begin{center}
2215: \plotfiddle{f6.eps}{1in}{90.}{200}{350}{-100}{-100}
2216: \vspace{-1in}
2217: \end{center}
2218: \caption{Time derivative of various volume-averaged energy densities for a 20 orbit period in the $\none$ simulation.
2219: The time derivative of the energy densities have been multiplied by an orbital time over the initial gas pressure.
2220: The dark blue line is the energy injection rate, $\ein$, the black line is the thermal energy density derivative, $\td$, the green line is the kinetic energy density derivative, $\kd$, and the red line is the magnetic energy density derivative, $\md$. The dotted line indicates zero.
2221: \label{dedt_n}}
2222: \end{figure}
2223:
2224: \begin{figure}
2225: \plotfiddle{f7a.eps}{1in}{90.}{200}{250}{-30}{-100}
2226: \vspace{-3.12in}
2227: \plotfiddle{f7b.eps}{0in}{90.}{200}{250}{235}{-100}
2228: \caption{Correlation coefficients calculated over orbits 20 to 100 in the $\none$ simulation.
2229: The plot on the left was calculated by correlating the energy injection rate, $\ein$, against the thermal energy time derivative, $\td$.
2230: The $x$-axis is the correlation length in time, and the $y$-axis is the coefficient multiplied by an orbital period over the initial gas pressure.
2231: The plot on the right was calculated by correlating the magnetic energy derivative, $\md$, (solid line) and
2232: kinetic energy derivative, $\kd$, (dashed line) against the thermal energy derivative. The dotted line indicates $C_{AB} = 0$. Note that
2233: the two plots have different $y$ scales.
2234: \label{corr_n}}
2235: \end{figure}
2236:
2237: \begin{figure}
2238: \plotfiddle{f8.eps}{1in}{90.}{325}{475}{-10}{-100}
2239: \vspace{-1in}
2240: \caption{Various terms in the volume-averaged magnetic, kinetic, and thermal energy density evolution equations over a two orbit period of $\none$.
2241: The energy terms are $\td$ (black),
2242: $\ein$ (blue), -$\qk$ (green), -$\qm$ (red), and the volume-averaged transfer rate from kinetic to magnetic energy (pink). All of these terms
2243: are defined in the text and have been multiplied by an orbital period over the initial gas pressure.
2244: \label{energies_n}}
2245: \end{figure}
2246:
2247: \begin{figure}
2248: \plotfiddle{f9.eps}{1in}{90.}{325}{475}{-10}{-100}
2249: \vspace{-1in}
2250: \caption{Correlation coefficient calculated over the saturated state of the $\zone$ simulation.
2251: The coefficient was calculated by correlating the energy injection rate against the thermal energy density derivative.
2252: The $x$-axis is the correlation length in time, and the $y$-axis is the coefficient multiplied by an orbital period over the initial gas pressure.
2253: The narrow peaks in the curve correspond to residual effects from rebinning the energy derivatives (described in the text). The broader peak in the correlation
2254: function occurs at $\Delta t \sim$ -0.2 orbits.
2255: \label{corr_z}}
2256: \end{figure}
2257:
2258: \begin{figure}
2259: \plotfiddle{f10a.eps}{1in}{90.}{200}{250}{-30}{-100}
2260: \vspace{-3.12in}
2261: \plotfiddle{f10b.eps}{0in}{90.}{200}{250}{235}{-100}
2262: \caption{Volume-averaged magnetic and kinetic energy densities in the first 1.5~orbits of $\ndone$ (left) and $\zdone$ (right). In both plots, the upper curves correspond to the kinetic energy density and the lower curves correspond to the magnetic energy density. In $\zdone$, high frequency oscillations appear in the kinetic energy evolution. To smooth away these oscillations, a moving window average was applied to the kinetic energy density. The unsmoothed kinetic energy is shown by the dotted line, while the
2263: smoothed kinetic energy is the solid line. The magnetic energy density in the $\zdone$ plot has also been smoothed for consistency.
2264: Both the kinetic and magnetic energy densities have been normalized to their respective (unsmoothed) values at $t =$~40~orbits.
2265: \label{turb_decay}}
2266: \end{figure}
2267:
2268: \begin{figure}
2269: \plotfiddle{f11.eps}{0.1in}{90.}{475}{300}{80}{-500}
2270: \caption{Magnetic Fourier transfer functions versus a dimensionless wave number ($L$ is the length of the smallest dimension in the box) for $\zone$. Each plot
2271: is displayed in two components; the left part shows the data for $1 < k L/(2\pi) < 20$, and the right part shows the data for $20 < k L/(2\pi) < 64$ by
2272: changing the $x$ and $y$ axis scaling. In all plots, the solid line is the average value for the transfer function. This average was obtained over 161 frames in the saturated state and shells of constant $|{\bm k}|$. The upper (lower) dashed line that matches color with the solid line correspond to the transfer function plus (minus) one temporal standard deviation. From top to bottom, the plots show $S$ (red) and $A$ (black), $\tbb$, $\tdivv$, and $\tbv$.
2273: \label{magtfz}}
2274: \end{figure}
2275:
2276: \clearpage
2277:
2278: \begin{figure}
2279: \plotfiddle{f12.eps}{0.1in}{90.}{475}{300}{80}{-500}
2280: \caption{Kinetic Fourier transfer functions versus a dimensionless wave number ($L$ is the length of the smallest dimension in the box) for $\zone$. Each plot
2281: is displayed in two components; the left part shows the data for $1 < k L/(2\pi) < 20$, and the right part shows the data for $20 < k L/(2\pi) < 64$ by
2282: changing the $x$ and $y$ axis scaling. In all plots, the solid line is the average value for the transfer function. This average was obtained over 161 frames in the saturated state and shells of constant $|{\bm k}|$. The upper (lower) dashed line that matches color with the solid line correspond to the transfer function plus (minus) one temporal standard deviation. From top to bottom, the plots show $\tvv$, $\tvb$, $\tpress$ (red) and $\tcomp$ (black), and $\tphi$.
2283: \label{kintfz}}
2284: \end{figure}
2285:
2286: \begin{figure}
2287: \plotfiddle{f13.eps}{1in}{90.}{325}{475}{-10}{-100}
2288: \vspace{-1in}
2289: \caption{Numerical dissipation quantities plotted against a dimensionless wave number ($L$ is the length of the smallest dimension in the box). These plots correspond to data from $\zone$. The upper left plot shows the dissipation rate of kinetic energy (green) and magnetic energy (red) in Fourier space. The upper right plot shows the ratio of these two dissipation rates. The lower left plot shows the effective numerical viscosity (green) and resistivity (red). The lower right plot shows the ratio of the viscosity to resistivity (i.e., the effective Prandtl number). In all plots, the solid line is the average value for the quantity of interest. For $\dkin$ and $\dmag$, this average was obtained from averaging over shells of constant $|{\bm k}|$ and over 161 frames in the saturated state. The averaged viscosity and resistivity values were calculated as described in the text. The upper and lower dashed lines correspond to the error propagated from one temporal standard deviation.
2290: \label{dis_z}}
2291: \end{figure}
2292:
2293: \begin{figure}
2294: \plotfiddle{f14.eps}{1in}{90.}{325}{475}{-10}{-100}
2295: \vspace{-1in}
2296: \caption{Averaged dissipation related quantities as a function of grid resolution. These plots correspond to data from the zero net flux simulations, $\zonesix$, $\zthree$, $\zsix$, and $\zone$. The upper left plot shows the effective viscosity versus $x$ resolution. The dashed line shows $\nueff \propto N_x^{-2}$. The upper right plot shows the effective resistivity versus $x$ resolution. Again, the dashed line shows $\etaeff \propto N_x^{-2}$. The lower left plot shows the ratio of kinetic to magnetic dissipation versus $x$ resolution. The lower right plot shows the effective Prandtl number versus $x$ resolution. For each resolution, the data point was obtained from averaging the quantity as a function of $k$ over values of $k$ where the error in this quantity is not much larger than the mean value. The error bars represent the propagated errors from
2297: the temporal statistics.
2298: \label{nu_eta_res_z}}
2299: \end{figure}
2300:
2301: \begin{figure}
2302: \plotfiddle{f15.eps}{1in}{90.}{325}{475}{-10}{-100}
2303: \vspace{-1in}
2304: \caption{Magnetic dissipation rate from the $\zone$ simulation for three versions of the transfer function analysis. The upper left plot and the red lines in the lower left plot correspond to the analysis in which $B_y = 0$ was assumed. The upper right plot and the red lines in the lower right plot correspond to the analysis in which both $B_y = 0$ and $k_y = 0$ were assumed. The black lines in the lower plots result from relaxing both of these assumptions. The solid lines in the upper plots correspond to $\dmag$ whereas the dashed lines correspond to $\eta \teta$ with $\eta = 10^{-7}$ chosen to provide a reasonable match to $\dmag$ at large $k$. The dashed lines in the lower plots correspond to one standard deviation above and below the quantity represented by the solid line of the same color. A horizontal line at zero is shown in all plots as the blue dotted line. Note the difference in $y$-axis scale between the upper and lower plots.
2305: \label{trans_compare}}
2306: \end{figure}
2307:
2308: \begin{figure}
2309: \plotfiddle{f16.eps}{0.1in}{90.}{475}{300}{80}{-500}
2310: \caption{Magnetic Fourier transfer functions versus a dimensionless wave number ($L$ is the length of the smallest dimension in the box) for $\none$. Each plot
2311: is displayed in two components; the left part shows the data for $1 < k L/(2\pi) < 20$, and the right part shows the data for $20 < k L/(2\pi) < 64$ by
2312: changing the $x$ and $y$ axis scaling. In all plots, the solid line is the average value for the transfer function. This average was obtained over 161 frames in the saturated state and shells of constant $|{\bm k}|$. The upper (lower) dashed line that matches color with the solid line correspond to the transfer function plus (minus) one temporal standard deviation. From top to bottom, the plots show $S$ (red) and $A$ (black), $\tbb$, $\tdivv$, and $\tbv$.
2313: \label{magtfn}}
2314: \end{figure}
2315:
2316: \begin{figure}
2317: \plotfiddle{f17.eps}{0.1in}{90.}{475}{300}{80}{-500}
2318: \caption{Kinetic Fourier transfer functions versus a dimensionless wave number ($L$ is the length of the smallest dimension in the box) for $\none$. Each plot
2319: is displayed in two components; the left part shows the data for $1 < k L/(2\pi) < 20$, and the right part shows the data for $20 < k L/(2\pi) < 64$ by
2320: changing the $x$ and $y$ axis scaling. In all plots, the solid line is the average value for the transfer function. This average was obtained over 161 frames in the saturated state and shells of constant $|{\bm k}|$. The upper (lower) dashed line that matches color with the solid line correspond to the transfer function plus (minus) one temporal standard deviation. From top to bottom, the plots show $\tvv$, $\tvb$, $\tpress$ (red) and $\tcomp$ (black), and $\tphi$.
2321: \label{kintfn}}
2322: \end{figure}
2323:
2324:
2325: \begin{figure}
2326: \plotfiddle{f18.eps}{1in}{90.}{325}{475}{-10}{-100}
2327: \vspace{-1in}
2328: \caption{Numerical dissipation quantities plotted against a dimensionless wave number ($L$ is the length of the smallest dimension in the box). These plots correspond to data from $\none$. The upper left plot shows the dissipation rate of kinetic energy (green) and magnetic energy (red) in Fourier space. The upper right plot shows the ratio of these two dissipation rates. The lower left plot shows the effective numerical viscosity (green) and resistivity (red). The lower right plot shows the ratio of the viscosity to resistivity (i.e., the effective Prandtl number). In all plots, the solid line is the average value for the quantity of interest. For $\dkin$ and $\dmag$, this average was obtained from averaging over shells of constant $|{\bm k}|$ and over 161 frames in the saturated state. The averaged viscosity and resistivity values were calculated as described in the text. The upper and lower dashed lines correspond to the error propagated from one temporal standard deviation.
2329: \label{dis_n}}
2330: \end{figure}
2331:
2332: \begin{figure}
2333: \plotfiddle{f19.eps}{1in}{90.}{325}{475}{-10}{-100}
2334: \vspace{-1in}
2335: \caption{Averaged dissipation related quantities as a function of grid resolution. These plots correspond to data from the net flux simulations, $\nonesix$, $\nthree$, $\nsix$, and $\none$. The upper left plot shows the effective viscosity versus $x$ resolution. The dashed line shows $\nueff \propto N_x^{-2}$. The upper right plot shows the effective resistivity versus $x$ resolution. Again, the dashed line shows $\etaeff \propto N_x^{-2}$. The lower left plot shows the ratio of kinetic to magnetic dissipation versus $x$ resolution. The lower right plot shows the effective Prandtl number versus $x$ resolution. For each resolution, the data point was obtained from averaging the quantity as a function of $k$ over values of $k$ where the error in this quantity is not much larger than the mean value. The error bars represent the propagated errors from
2336: the temporal statistics.
2337: \label{nu_eta_res_n}}
2338: \end{figure}
2339:
2340:
2341: \end{document}
2342: