1:
2: \documentstyle[preprint,psfig]{aastex}
3: %\documentstyle[manuscript]{aastex}
4:
5: \def\lesssim{\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$<$}}}}
6: \def\gtrsim{\mathrel{\hbox{\rlap{\hbox{\lower4pt\hbox{$\sim$}}}\hbox{$>$}}}}
7:
8:
9: \begin{document}
10:
11: \title{Global MHD Simulation of the Inner Accretion Disk in a
12: Pseudo-Newtonian Potential}
13:
14: \author{John F. Hawley}
15: \affil{Department of Astronomy, University of Virginia,
16: Charlottesville VA 22903}
17:
18: \author{Julian H. Krolik}
19: \affil{Physics and Astronomy Department, Johns Hopkins University,
20: Baltimore, MD 21218}
21:
22: \shorttitle{Global MHD Simulation}
23:
24: \begin{abstract}
25:
26: We present a detailed three dimensional magnetohydrodynamic (MHD)
27: simulation describing the inner region of a disk accreting onto a black
28: hole. To avoid the technical complications of general relativity, the
29: dynamics are treated in Newtonian fashion using the pseudo-Newtonian
30: Pacz\'ynski-Wiita potential. The disk evolves due to angular momentum
31: transport that is produced naturally from MHD turbulence generated by
32: the magnetorotational instability. We find that the resulting stress
33: is continuous across the marginally stable orbit, in contradiction
34: with the widely-held assumption that the stress should go to
35: zero there. As a consequence,
36: the specific angular momentum of the matter accreted into the hole is
37: smaller than the specific angular momentum at the marginally stable
38: orbit. The disk exhibits large fluctuations in almost every quantity,
39: both spatially and temporally. In particular, the ratio of stress to
40: pressure (the local analog of the Shakura-Sunyaev $\alpha$ parameter)
41: exhibits both systematic gradients and large fluctuations; from $\sim
42: 10^{-2}$ in the disk midplane at large radius, it rises to $\sim 10$
43: both at a few gas density scale heights above the plane at large radius,
44: and near the midplane well inside the plunging region. Driven in part
45: by large-amplitude waves excited near the marginally stable orbit, both
46: the mass accretion rate and the integrated stress exhibit large
47: fluctuations whose Fourier power spectra are smooth ``red" power-laws
48: stretching over several orders of magnitude in timescale.
49:
50: \end{abstract}
51:
52: \keywords{accretion, accretion disks, MHD, black holes}
53:
54: \section{Introduction}
55:
56: For many years, the physical nature of angular momentum transport
57: in accretion disks has been one of the great outstanding
58: questions of high-energy astrophysics. As a result of nearly a decade
59: of effort, it is now becoming increasingly clear that the mechanism is
60: turbulent magnetic stress in which the energy for the magnetic field is
61: largely drawn from the free energy of orbital shear (as reviewed by
62: Balbus \& Hawley 1998). Theoretical grounds for this conclusion were
63: first found in the rediscovery of the magneto-rotational instability
64: (MRI) by Balbus \& Hawley (1991). Increasingly detailed simulations of
65: local ``shearing boxes'' have bolstered our confidence that this is
66: indeed the process found in nature (e.g., Hawley, Gammie \& Balbus
67: 1995, 1996; Brandenburg et al. 1995; Stone et al. 1996;
68: Miller \& Stone 2000).
69:
70: What has been missing until recently is studies of the radial structure
71: of disks accreting under the influence of magnetohydrodynamic (MHD)
72: turbulent stresses. Hitherto, only a few such three-dimensional global
73: simulations have been presented (Armitage 1998; Matsumoto 1999; Hawley
74: 2000; Machida, Hayashi, \& Matsumoto 2000).
75: In global simulations it is difficult to obtain adequate
76: resolution over the wide range of length scales involved, namely the
77: MHD turbulent length scales within the disk, the disk scale height $H$,
78: and the radial distance from the central object $R$. These length
79: scales are generally quite disparate, but near the inner boundary of an
80: accretion disk they should become comparable. In this paper we present
81: a global simulation of an accretion disk in which we focus on this
82: inner region where adequate numerical resolution can be obtained.
83:
84: With the results of this new simulation, we can begin to answer a
85: number of questions of central interest to the mechanics of accretion.
86: For example, the most common view of the inner portions of black hole
87: accretion disks is based on the work of Novikov \& Thorne (1973) (e.g.
88: Abramowicz \& Zurek 1981; Muchotrzeb \& Paczy\'nski 1982; Matsumoto et
89: al. 1984). In this picture, the disk is (nearly) time-steady and
90: axisymmetric. Page \& Thorne (1974) argued on heuristic grounds that
91: the $R$-$\phi$ component of the stress (the one responsible for angular
92: momentum transport) should go to zero at the marginally stable orbit;
93: later work by Abramowicz and Kato (1989) showed that if the stress is a
94: (constant) small fraction of the local pressure, it automatically
95: approaches zero near that location.
96:
97: However, because the MHD turbulence is the result of a general, local,
98: rapidly-growing instability, assumptions such as stationarity and
99: axisymmetry are likely inappropriate. One might also question whether
100: there is any reason for the stress to diminish at the disk's inner
101: edge, given that the growth rate of the MRI does not appreciably
102: diminish near the marginally stable orbit (cf. Krolik 1999; Gammie 1999;
103: see also the footnote in Page \& Thorne 1974 in which they speculate
104: that their heuristic arguments might fail if strong magnetic fields were
105: present). Further, because magnetic
106: stresses do not automatically maintain a fixed ratio to the pressure,
107: conclusions based upon $\alpha$ parameterizations (e.g., Abramowicz \&
108: Kato 1989) may not directly apply. This issue is important because
109: continued stress at the marginally stable orbit could alter both the
110: energy and angular momentum with which matter arrives at the black
111: hole; that is, this boundary condition determines the efficiency of
112: accretion and the rate at which the black hole is spun up.
113:
114: Because the greatest luminosity should be released near the inner edge
115: of the accretion disk, the detailed behavior and structure of the
116: region near the marginally stable orbit is crucial to
117: understanding the observations of black hole systems. For example, the
118: luminosity of essentially all accreting black holes exhibits sizable
119: fluctuations with a very broad-band distribution of fluctuation power
120: with timescale (e.g., as discussed by Sunyaev \& Revnivtsev 2000); can we
121: identify the specific dynamics that drive these variations?
122: Occasionally, the lightcurve exhibits
123: quasi-periodic oscillations (QPOs); can we either identify the
124: mechanism responsible for these, or test some of those mechanisms that
125: have been proposed?
126:
127: In the following section we will present a technical description of the
128: three-dimensional global simulation we report. We will then set out
129: our results in \S 3, with their qualitative implications outlined in \S
130: 4. In \S 5 we will discuss the degree to which limitations of our
131: simulation hamper direct application of our results, and suggest in
132: which directions the greatest improvement might soon be made.
133: We summarize our conclusions in \S 6.
134:
135:
136: \section{Numerical Method}
137:
138: \subsection{Equations}
139:
140: An investigation of the inner edge of a disk orbiting a black hole
141: would be best carried out with a fully general relativistic
142: simulation. However, we have not yet written a three dimensional
143: general relativistic MHD code. In the present study we instead make
144: use of the existing three dimensional Newtonian MHD code (Hawley 2000).
145: The code evolves the equations of ideal MHD, i.e.,
146: \begin{equation}\label{mass}
147: {\partial\rho\over \partial t} + \nabla\cdot (\rho {\bf v}) = 0
148: \end{equation}
149: \begin{equation}\label{mom}
150: \rho {\partial{\bf v} \over \partial t}
151: + (\rho {\bf v}\cdot\nabla){\bf v} = -\nabla\left(
152: P + {\mathcal Q} +{B^2\over 8 \pi} \right)-\rho \nabla \Phi +
153: \left( {{\bf B}\over 4\pi}\cdot \nabla\right){\bf B}
154: \end{equation}
155: \begin{equation}\label{ene}
156: {\partial\rho\epsilon\over \partial t} + \nabla\cdot (\rho\epsilon
157: {\bf v}) = -(P+{\mathcal Q}) \nabla \cdot {\bf v}
158: \end{equation}
159: \begin{equation} \label{ind}
160: {\partial{\bf B}\over \partial t} =
161: \nabla\times\left( {\bf v} \times {\bf B} \right)
162: \end{equation}
163: where $\rho$ is the mass density, $\epsilon$ is the specific internal
164: energy, ${\bf v}$ is the fluid velocity, $P$ is the pressure, $\Phi$ is
165: the gravitational potential, ${\bf B}$ is the magnetic field vector,
166: %%%NEW
167: and ${\mathcal Q}$ is an explicit artificial viscosity of
168: the form described by Stone \& Norman (1992a).
169: %%%
170: The global disk code is written in cylindrical coordinates, $(R,\phi,z)$.
171: To model important effects associated with the relativistic Schwarzschild
172: metric, we employ the pseudo-Newtonian potential of Paczy\'nski \& Wiita
173: (1980). The pseudo-Newtonian potential has the form
174: \begin{equation}\label{pwp}
175: \Phi = - {G M \over r-r_g}
176: \end{equation}
177: where $r$ is spherical radius,
178: and $r_g \equiv 2GM/c^2$ is the ``gravitational radius,''
179: akin to the black hole horizon. For this potential,
180: the Keplerian specific angular momentum (i.e., that
181: corresponding to a circular orbit) is
182: \begin{equation}\label{pwl}
183: \ell_{kep} = (GMr)^{1/2} {r \over r-r_g} ,
184: \end{equation}
185: and the angular frequency $\Omega = \ell/R^2$. The innermost marginally
186: stable circular orbit is located at $r_{ms}=3r_g$.
187:
188: We use an adiabatic equation of state, $P=\rho\epsilon(\Gamma
189: -1) = K\rho^\Gamma$, with $\Gamma = 5/3$, and ignore radiation
190: transport and losses. Since there is no explicit resistivity or
191: physical viscosity, the gas can heat only through adiabatic
192: compression or by the action of the artificial viscosity
193: which acts in shocks.
194: Equations (\ref{mass})-(\ref{ind}) are solved using time-explicit
195: Eulerian finite differencing. The numerical algorithm is that employed
196: by the ZEUS code for hydrodynamics (Stone \& Norman 1992a) and MHD
197: (Stone \& Norman 1992b; Hawley \& Stone 1995). We adopt units where
198: $GM=1$ and $r_g = 1$ (so that $c = \sqrt{2}$).
199:
200:
201: In this paper we repeat---with significantly better
202: resolution---simulation GT4 from Hawley (2000), whose initial state was
203: a thick torus with an angular velocity distribution $\Omega \propto
204: R^{-1.68}$, slightly steeper than Keplerian. The angular momentum
205: within the torus is equal to the Keplerian value at $R=10$, which
206: determines the location of the pressure maximum in the torus. In
207: somewhat arbitrary units, the pressure at this point is equal to 0.036
208: and the density $\rho_{max}=34$. The initial magnetic field is
209: constructed by setting the toroidal component of the vector potential
210: %%%NEW
211: $A_\phi (R,z) = \rho (R,z) -
212: \rho_{min}$, for all $\rho$ greater than a minimum value, $\rho_{min} =
213: 0.1$. Poloidal field components are then constructed from ${\bf B} =
214: \nabla \times {\bf A}$. This procedure produces large-scale poloidal
215: field lines that follow the equal-density contours inside the torus.
216: The magnetic field is initially zero everywhere $\rho \leq \rho_{min}$.
217: As discussed in
218: Hawley (2000), this configuration is motivated by the desire to have
219: well-resolved unstable modes of the poloidal field instability, and to
220: have the initial field fully contained within the torus. The magnetic
221: field level is renormalized so that the total magnetic energy in the
222: torus is 0.01 the total integrated gas pressure. This corresponds to an
223: overall value of $\beta=P_{\rm gas}/P_{\rm mag} = 100$.
224: %%%%
225:
226: %NEW
227: The aim of this initial condition is not so much to
228: follow the evolution of this particular torus, as it is to construct a
229: disk that is nearly Keplerian with an inner boundary that is initially
230: close to, but still outside, the marginally stable orbit. To that end,
231: the torus is perturbed with low-amplitude random adiabatic pressure
232: fluctuations. These seed the magneto-rotational instability whose
233: rapid evolution leads to the desired state. Further disk evolution
234: results in inflow through $r_{ms}$;
235: %
236: it is this resulting accretion flow that will be studied in detail.
237:
238: The computational grid has $128\times 128 \times 128$ zones. This is the
239: same number of zones as used in GT4 of Hawley (2000), but here,
240: to increase the resolution within the disk itself and in the inflow
241: region, we locate more of the zones near the marginally stable orbit
242: and around the equator. The center of the coordinate system is
243: excised, i.e., the radial coordinate begins at a nonzero
244: $R_{\rm min} =1.5$; this choice avoids both the coordinate singularity
245: associated with the axis and the singularity at $r_g$. The outer
246: radius is set at $R=31.5$. The $z$ coordinate is centered on the
247: equatorial plane, and runs from -10 to $+10$. The angle $\phi$ is
248: periodic and covers the full $2\pi$. In the radial direction, there
249: are 30 equally spaced zones in $R$ from 1.5 to 4; the radial zones are
250: then graded logarithmically from 4 to the outer boundary at 31.5. In
251: the vertical direction, half of the $z$ zones are equally spaced around
252: the equator from $-2$ to 2, then graded from those locations out to the
253: vertical boundaries at $z = \pm 10$. The grid cells are equally spaced
254: in $\phi$. The simulation GT4 from Hawley (2000) used equally spaced
255: zones in $R$ and $z$, and the outer radial boundary was set at 21.5.
256: We have also run this initial torus at lower resolution for comparison,
257: using $64^3$ grid zones.
258:
259: The boundary conditions on the grid are simple zero-gradient outflow
260: conditions; no flow into the computational domain is permitted. The
261: magnetic field boundary condition is set by requiring the transverse
262: components of the field to be zero outside the computational domain,
263: while the perpendicular component satisfies the divergence-free
264: constraint. Although this produces a nonzero stress at the boundary,
265: it works well in preventing an artificial buildup of field at the
266: boundary. Most of the time, for $|z| \leq 1$ along the inner radial
267: boundary and $R \leq 2$ in the equatorial plane, the poloidal velocity
268: is greater than the magnetosonic speed, so that the influence of the
269: boundary condition on the flow upstream is limited.
270:
271: Timescales of interest in the simulation are set by the circular
272: orbital period $P_{orb} = 2\pi \Omega^{-1} = 2\pi r^{3/2} (r-r_g)/r$ at
273: significant points within the disk. In units where $GM=r_g=1$,
274: $P_{orb}$ at the pressure maximum ($R=10$) is $179$, and at the
275: marginally stable orbit, where the orbital frequency $\Omega =
276: 1/(2\sqrt 3)$, the orbital period is 21.8. The local accretion
277: timescale is set by the effective stress, namely $t_{acc} = R/\langle
278: v_R \rangle \sim \ell/W_{R\phi} ( \gg P_{orb})$, where $W_{R\phi}$ is
279: the $R$--$\phi$ component of the stress, $W_{R\phi} = \langle \delta
280: v_R \, \delta v_\phi - {v_{AR} v_{A\phi} \rangle}$, and $v_A$ is an
281: Alfv\'en velocity. The goal is to evolve long enough to observe a
282: number of accretion times near the marginally stable orbit. Since the
283: code is time-explicit and therefore Courant-limited, this is difficult
284: to achieve over large radial extents. Here we ran the simulation out
285: to time $t=1500$, which is 69 orbits at $r_{ms}$.
286:
287: \subsection{Diagnostics}
288:
289: The quantity of data associated with a three dimensional simulation is
290: daunting, and this makes analyzing the outcome a challenge. Diagnostics
291: can be computed during the simulation and in post-processing. The best
292: procedures for this are still being developed. Here we briefly describe
293: some of the diagnostic procedures used in the present simulation.
294:
295: The time-evolution of the disk can be visualized through animation
296: sequences. Animation frames of the $\phi = 0$ slice and the equatorial
297: ($z=0$) slice are dumped every 4 units in time. Full 3D density
298: information is saved every 10 units of time, and these are used to make
299: an animation of a three dimensional volumetric rendering.
300:
301: The time-evolution can also be studied in a manageable way through
302: space-time $(R,t)$ studies of azimuthally- and vertically-integrated quantities.
303: Examples include the averaged mass density
304: \begin{equation}\label{sigma}
305: \langle \rho\rangle = {\int \rho R d\phi dz \over \int R d\phi dz},
306: \end{equation}
307: the average angular momentum,
308: \begin{equation}\label{angm}
309: \langle \rho \ell \rangle = {\int \rho \ell R d\phi dz \over \int R
310: d\phi dz},
311: \end{equation}
312: and average specific angular momentum,
313: \begin{equation}\label{el}
314: \langle \ell \rangle= {\langle \rho \ell \rangle \over \langle\rho\rangle }.
315: \end{equation}
316: The net radial mass flux is
317: \begin{equation}\label{mdot}
318: \langle \dot M\rangle = \langle -\rho v_R \rangle = \int -\rho v_R R d\phi dz,
319: \end{equation}
320: here defined to be positive for net accretion (inflow).
321: Similarly, the angular momentum flux,
322: \begin{equation}\label{lflux}
323: \langle \rho v_R \ell \rangle = \int \rho v_R \ell R d\phi dz,
324: \end{equation}
325: can be computed.
326:
327: Radial angular momentum transport in the disk is due to
328: the $R$--$\phi$ component of the stress,
329: \begin{equation}\label{stress}
330: T^{R\phi} = \rho \delta v_R \, \delta v_\phi - {B_R B_\phi \over 4\pi},
331: \end{equation}
332: where $\delta v_R$ and $\delta v_\phi$ are the {\it turbulent} portions
333: of the velocity, that is, the departures from the mean flow. The
334: kinematic portion is the Reynolds stress and the magnetic portion is
335: the Maxwell stress. In the simulation we have no way of
336: knowing what the mean flow is. For example, the typical radial
337: velocity is much larger than the time-averaged net radial drift
338: velocity, and there are significant differences between the
339: instantaneous angular velocity and the value associated with a circular
340: orbit (which one might assume should be the mean orbital flow). As a
341: result, there is no unique prescription to compute the Reynolds
342: component of the total stress. Hawley (2000) used a definition for the
343: perturbed orbital velocity $\delta v_\phi$ in terms of the difference
344: between the total instantaneous angular momentum flux, and the mass
345: flux times the average specific angular momentum. The difference then
346: represents the excess or deficit angular momentum transport due to
347: orbital velocity fluctuations compared to the mean. Specifically,
348: \begin{equation}\label{rstress}
349: \langle \rho \delta v_R \delta v_\phi \rangle =
350: \langle \rho v_R v_\phi \rangle
351: - \langle \rho v_R \rangle \langle \ell \rangle / R .
352: \end{equation}
353:
354: In contrast, we can easily compute the stress due solely to magnetic
355: forces, the second term on the right-hand-side of equation
356: (\ref{stress}); we denote this quantity by $M^{R\phi}$. The
357: vertically- and azimuthally-averaged Maxwell stress is
358: \begin{equation}\label{maxwell}
359: \langle M^{R\phi}\rangle = {\int (-B_R B_\phi / 4\pi) R d\phi dz
360: \over \int Rd\phi dz} .
361: \end{equation}
362: %%%NEW
363: We do not distinguish the total magnetic field from the fluctuating field
364: because, as we shall show later, where the field is dynamically important,
365: the magnitude of the fluctuating portion is large compared to any time-averaged
366: mean field.
367: %%%%
368:
369: In addition to these vertically- and azimuthally-averaged quantities,
370: we examine full data dumps of the simulation at specific moments in
371: time. From these we can examine not only the full three dimensional
372: structure, but also azimuthally-averaged and vertically-averaged
373: slices in $(R,z)$ and $(R,\phi)$.
374:
375: \section{Results}
376:
377: \subsection{Quasi-stationary state}
378:
379: The initial torus was in hydrodynamical equilibrium, but it evolves
380: rapidly as the simulation proceeds. The overall evolution is very
381: similar to that of GT4 in Hawley (2000; see Fig. 13).
382: The radial field is sheared by the differential rotation, creating
383: toroidal field. Turbulence develops within the disk with the onset of
384: the magnetorotational instability. The resulting Maxwell stresses
385: drive angular momentum transfer; the disk begins to accrete into the
386: central hole and expands radially outward. Relatively rapidly the
387: inner portion of the disk achieves an approximate steady-state. By $t
388: \simeq 700$, the mass accretion rate reaches roughly the long-time
389: mean; after $t \simeq 1000$, the shape of the disk's average radial
390: density distribution $\langle \rho\rangle (R)$ no longer changes
391: appreciably. Thus, we observe the disk in a quasi-stationary state
392: for several dozen orbital periods at the marginally stable orbit.
393: In much of what follows, we will present a detailed analysis of the
394: state of the disk during this time frame, focusing particularly on
395: the end of the simulation, using that ``snapshot" as a fairly typical
396: sample of conditions in the quasi-stationary disk.
397: However, it must be borne in mind that there are always
398: significant fluctuations around all these mean states.
399:
400: \subsection{Gas density}
401:
402: A contour plot of the azimuthally-averaged gas density at the initial
403: and final times is presented in Figure~1. As might be expected, the
404: density diminishes inside the marginally stable orbit and outside the
405: initial outer radial boundary of the disk, while also decreasing with
406: increasing altitude away from the equatorial plane. Although the
407: vertical heights of the highest contour levels at the final time are
408: roughly independent of radius within the disk proper, the
409: azimuthally-averaged exponential scale-height of the pressure, $H$,
410: increases very nearly linearly with radius from $R \simeq 2$ out to $R
411: \simeq 20$; at the marginally stable orbit ($R = 3$), $H \simeq 0.4$,
412: while at $R = 10$, $H \simeq 1.5$. This particular disk shape is
413: almost certainly a consequence of the equation of state, in which all
414: fluid elements retain the same specific entropy except for shock
415: heating.
416: %NEW
417: As a result of this special choice of equation of state,
418: the temperature rises inward as the gas is compressed; by coincidence,
419: this occurs at such a rate as to keep the ratio of sound speed to
420: orbital speed, and therefore $H/R$, almost constant at $\simeq 0.15$.
421: %
422: In a real disk, one in which turbulent dissipation introduces
423: heat and radiation vents it, the shape might be rather different.
424:
425: Density gradients in radius and altitude are determined primarily
426: by the overall dynamical effects of approximate hydrostatic equilibrium.
427: No such requirements exist for azimuthal fluctuations, whose
428: amplitudes characterize the turbulence. We
429: quantify the azimuthal fluctuation level by defining
430: \begin{equation}\label{fluctuation}
431: {\delta \rho \over \rho}\left(R,z\right) =
432: {1 \over \langle \rho\rangle_\phi} \left\{{1 \over 2\pi} \int \, d\phi \,
433: \left[\rho - \langle \rho \rangle_\phi \right]^2 \right\}^{1/2} ,
434: \end{equation}
435: where ${\langle X \rangle}_\phi$ refers to the azimuthal average at
436: fixed $R$ and $z$ of the quantity $X$. Measured in this way,
437: $\delta\rho/\rho \simeq 0.2$ -- 0.6 over most of the problem volume,
438: with smaller values typically nearer the equatorial plane. In a few
439: places, $\delta \rho/\rho$ rises to be greater than unity. Thus, quite
440: sizable density fluctuations are generic.
441:
442: These fluctuations are primarily sheared strips whose radial extent is
443: $\lesssim 0.1 R$, but extend $\sim 1$~radian in azimuthal angle (Figure~2).
444: This pattern (tightly wrapped spiral features) is characteristic of
445: fluctuations in almost all quantities. Part of their nature is
446: revealed by constructing spacetime diagrams in which the surface
447: density at a specific radius is plotted as a function of $t$ and $\phi$.
448: From this visualization we are able to measure the angular speed of
449: these features; it is generally quite close to the local orbital frequency
450: $\Omega$. We can also measure their persistence at any particular radius;
451: it is usually less than an orbital period. In this sense, they can be
452: regarded as fluctuations that merely ``ride" with the local orbital motion,
453: but see further discussion in \S 3.5.
454:
455: \subsection{Magnetic field}
456:
457: The magnetic field distribution is rather more complicated than the
458: density distribution. A poloidal projection (i.e., averaged
459: azimuthally) shows the magnetic pressure to have its largest value
460: %%%NEW
461: immediately
462: %%%%
463: outside the dense regions of the torus, with large variations inside
464: the torus (Figure~3). In the shearing-box simulations of Miller \&
465: Stone (2000), the field strength likewise peaked away from the midplane,
466: but we find somewhat greater contrast. Field-strength
467: fluctuations are also larger than
468: those for density: the {\it rms} fractional fluctuations in
469: %%%NEW
470: $|\vec B|$ in
471: %%%
472: the azimuthal direction are $\simeq 1$ -- 1.5 in the disk body, falling to
473: $\simeq 0.25$ several gas scale-heights immediately above and below the
474: disk.
475: %%%NEW
476: Another measure of the magnitude of the fluctuations is
477: $|\langle B_i \rangle_\phi |/\langle |B_i | \rangle_\phi$, the ratio of the
478: mean value of an individual field component to the mean of its absolute
479: value. This quantity is typically $\sim 0.2$ for all components
480: wherever the gas density
481: is significant, indicating that the field is extremely variable as a
482: function of azimuth (and therefore as a function of time at fixed azimuth,
483: as orbital rotation carries the field around).
484: Only where the gas density is very low does the field
485: become relatively smooth and steady
486: %NEW
487: (a result also qualitatively similar to the findings of Miller
488: \& Stone 2000).
489: %
490: %%%
491: Like the density, the characteristic shape of fluctuations is
492: sheared arcs that run for $\sim 1$~radian before losing coherence.
493: However, the peaks in magnetic field strength tend to lie adjacent to,
494: rather than coincident with, peaks in gas density.
495:
496: %%%NEW
497: To illustrate some of the field structure, we show in Figure~4
498: poloidal slices taken at $\phi = \pi/2$. Panel a) is a close-up of the inner,
499: high-resolution region; panel b) shows (almost) the entire disk.
500: This view, of course, suppresses the toroidal component (which is
501: typically larger than the poloidal component by a factor $\sim 1$ -- 10);
502: in addition, in order to make the figures visually comprehensible, we have
503: omitted small-scale field variations (smaller than 0.125 length
504: units in panel a), smaller than 0.5 length units in panel b). Even
505: with this smoothing, it is apparent that the field is very tangled. The
506: only field property exhibiting any stability is the roughly dipolar
507: pattern seen in the low density regions. This is a remnant of the
508: initial condition. Because it only persists in these very low density
509: regions, and they do not contribute much to the stress (see \S 3.6),
510: we do not regard this remnant artifact as dynamically significant.
511: %%%
512:
513: As one expects from magnetic fields in shearing plasma,
514: $B_R$ and $B_\phi$ are correlated in the sense that the mean ratio of
515: the Maxwell stress to the magnetic pressure, $\langle -2 B_R B_\phi
516: /B^2 \rangle > 0$ and has (approximately) constant magnitude $\simeq
517: 0.2$ -- 0.3 in the body of the disk. Inside the
518: marginally stable orbit, this ratio increases slowly in magnitude as
519: radial accretion ``combs'' the field out into radial loops that are
520: strongly sheared, increasing the correlation between $B_R$ and
521: $B_\phi$. In fact, the Maxwell stress is so correlated with the
522: average magnetic pressure $\langle B^2/8\pi\rangle$, that a figure
523: showing the stress is almost indistinguishable from one showing the
524: magnetic pressure.
525:
526: \subsection{Accretion rate}
527:
528: Perhaps the single most basic quantity of interest is the accretion
529: rate, $\langle \dot M\rangle$, defined in equation (\ref{mdot}).
530: The disk is never perfectly time-steady; $\rho v_R$ varies
531: substantially from place to place and from time to time. In fact, disk
532: quantities tend to be quite nonaxisymmetric, and much of the final
533: accretion inside $r_{ms}$ takes place in a spiral flow. However,
534: time-averaging can remove much of this variability, and
535: Figure 5 shows the time-averaged accretion rate as a function of $R$
536: along with the instantaneous values of $\dot M (R)$ at $t= 1000$
537: and 1500. The time-averaged $\dot M(R)$ is nearly constant as a
538: function of radius inside $R=10$, with $\dot M = 5$. We shall adopt
539: this as the fiducial steady state accretion rate for the inner disk.
540: Between $R=9$ and 14 there is inflow at a decreasing rate; outside
541: $R=14$ the net flux is outward. In this simulation, the initial disk
542: is entirely contained within the grid, and the outer part of the disk
543: must move outward in order to ``soak up" the angular momentum received
544: from the inner disk material. Over the course of the simulation the
545: total disk mass decreases by 14\%; about 90\% of this has gone into the
546: black hole.
547:
548: Figure~6 shows a detailed time-history of the accretion rate through
549: the inner radial boundary. As can be seen from this figure, the
550: accretion rate shows significant time variations after its initial
551: growth phase. The nature of the time-variability can be studied using
552: the Fourier power spectrum of the accretion rate
553: during the latter portion of the simulation (beyond $t=600$, i.e., after the
554: initial growth phase is over). This is shown in Figure~7. It is
555: most simply described as a ``red" power-law; roughly speaking,
556: $P(f) \propto f^{-1.7}$. However, there is also some curvature, in the
557: sense that it steepens with increasing frequency. Because the accretion
558: rate does appear to have a well-defined mean value, $d\ln P/d\ln f$ must
559: be $> -1$ at the lowest frequencies. At the high frequency end,
560: the spectrum is largely featureless, extending almost to the inverse
561: free-fall time at the innermost boundary. Other than a smooth
562: steepening, there is no special feature at the orbital frequency
563: of the marginally stable orbit (marked on the figure). The flow is
564: unsteady everywhere, even well inside the plunging region.
565:
566: Figure~6 provides some idea of how nonsteady the accretion is; the mass
567: flux is highly spatially inhomogeneous as well. To gain some sense of
568: the magnitude of the spatial variations, we show the accretion
569: rate as a function of position on a cylinder at $R= 5$ (Figure~8). As
570: this figure shows, the fluctuations are very large. If we suppose that
571: the mean accretion rate of 5 is evenly distributed over the range of
572: altitudes $-1 \leq z \leq 1$, the mean mass flux would be $\sim 0.1$.
573: As the figure makes plain, there are several places where the mass flux
574: is five times as great as this; there are also several places where the
575: mass flux is $\simeq 0.3$ -- 0.4 {\it outward}. Relatively large
576: fluctuations compared to the mean are to be expected when the transport of
577: angular momentum in the disk is by turbulence.
578:
579: \subsection{Propagation of fluctuations}
580:
581: As we have stressed all along, one of the most striking features
582: of this simulation is the presence of large amplitude nonaxisymmetric
583: modulations in the density, mass flux, and magnetic field strength.
584: We have already discussed how $(\phi,t)$ spacetime plots enable us to
585: measure their angular velocity. It is possible to learn more
586: by constructing analogous plots in $(R,t)$. For example, in Figure~9
587: we show the history of the azimuthally- and vertically-averaged
588: mass flux (eq. \ref{mdot}) in this simulation. The diagonal stripes
589: show the radial motion through the disk of fluctuations in the mass
590: flux, their slopes indicating their radial group speeds. Typically
591: these disturbances begin near $R \simeq 5$ and travel both outward and
592: inward. In the outward direction, they move at a very nearly constant
593: speed $\simeq 0.07$, while in the inward direction, their speed is at
594: least three times faster. Because the outward speed very nearly
595: matches the magnetosonic speed, we infer that these disturbances
596: propagate as magnetosonic waves in the outward direction. On the other
597: hand, inward-going mass flux perturbations are simply advected along
598: with the dynamical inflow.
599:
600: We conjecture that their origin lies in the way in which material
601: enters the region of unstable orbits. When a fluid element begins to
602: accelerate inward, it creates a rarefaction wave that moves outward at
603: the magnetosonic speed, while traveling azimuthally at the (faster)
604: orbital speed. On the other hand, the radial inflow speed
605: near and inside the marginally stable orbit is comparable to the
606: magnetosonic speed, so fluctuations travel inward rather more quickly.
607: These fluctuations are never smoothed out into a completely steady flow
608: because individual fluid elements move into the plunging region as a
609: result of the continually changing torques exerted by the MHD
610: turbulence (see the next subsection). Thus, at this level of detail,
611: there is no way to achieve a stationary state.
612:
613: Although strong nonaxisymmetric pressure waves were generated by MHD
614: turbulence in the shearing box simulations, they had no
615: significant impact on the evolution of those simulations. In this
616: global simulation, however, as Figure~9 shows, they play
617: a significant role in the creation of large fluctuations in the
618: accretion rate.
619:
620: \subsection{$R$--$\phi$ Stress}
621:
622: The total stress must satisfy the equation of angular momentum
623: conservation. In a steady-state disk, this means that locally the
624: torque must be equal to the mass accretion rate (which is constant)
625: times the difference between the local specific angular momentum $\ell$
626: and the specific angular momentum carried into the black hole
627: $\ell(r_{g})$. Thus, the stress is equal to
628: \begin{equation}\label{steadystress}
629: \bar
630: S = {\dot M \Omega(R) \over 2\pi} \left[1 - {\ell(r_g)
631: \over \ell(R)}\right],
632: \end{equation}
633: where $\dot M$ is the total mass accretion rate, $\Omega(R)$ the
634: rotational frequency (for the pseudo-Newtonian potential, $\Omega(R) =
635: 1/[R^{1/2}(r-1)]$), and $\ell$ is the specific angular momentum ($\ell
636: \equiv R^2 \Omega$).
637: In the time-steady, axisymmetric ``Novikov-Thorne" disk, it is assumed
638: that $\ell(r_g) = \ell(r_{ms})$ because the stress is assumed to be zero
639: for all $R \leq r_{ms}$.
640:
641:
642: For the purpose of studying angular momentum transport, it is useful to
643: look at the azimuthally- and vertically-integrated magnetic stress,
644: $\langle \int dz M^{R\phi}\rangle_\phi$, as this is the quantity most directly
645: related to transfer of the $z$-component of angular momentum. In particular,
646: we compare this quantity to the total stress per unit area
647: that would be expected in the steady state ``Novikov-Thorne" disk. As
648: discussed above, in the inner part of the disk the mean accretion rate
649: is about 5 in code units. In Figure~10 $\bar S$ (for $\dot M=5$ and
650: $\ell(r_g) = \ell(r_{ms})$) is
651: contrasted with $\langle \int \, dz \, M^{R\phi}\rangle_\phi$ averaged
652: from $t=750$ to $t=1500$. As that figure
653: shows, in the main body of the disk (approximately $4 \leq R \leq 15$),
654: the average magnetic stress is almost exactly equal to the expected stress.
655: At large radius the two curves diverge because
656: equation (\ref{steadystress}) assumes a time-steady disk, whereas ours
657: has only a finite mass, so the behavior at its radial exterior is quite
658: different. At the inner edge of the disk, in contrast to the
659: prediction of the conventional model, the azimuthally- and vertically-averaged
660: magnetic stress does not
661: diminish, but, if anything, increases slightly inward (the final upward
662: spike is an artifact of the boundary condition).
663: Note, however, that the local value of the ratio
664: $|M^{R\phi}/(\rho v_R v_\phi)|$ in the plunging region varies widely:
665: from as much as $\sim 10$ several scale-heights out of the equatorial
666: plane to as little as $\sim 10^{-4}$ at some locations in the plane.
667:
668: In the context of time-steady and axisymmetric disks, the stress is often
669: discussed in terms of the dimensionless parameter introduced by Shakura
670: \& Sunyaev (1973),
671: \begin{equation}
672: \alpha_{SS} \equiv \langle {\int \, dz \, T^{R\phi} \over \int \, dz \, p}
673: \rangle_\phi .
674: \end{equation}
675: %%%NEW
676: In many papers following this seminal work, the ``$\alpha$ model" is
677: generalized so that the {\it local} ratio of stress to pressure
678: is measured by the quantity $\alpha \equiv |T^{R\phi}|/p$. Often,
679: it is also assumed that the stress is due to some variety of
680: turbulent viscosity. However,
681: here we follow the spirit of the original Shakura \& Sunyaev work
682: by supposing that $\alpha$ is merely a convenient dimensionless
683: way of measuring the stress in units of the pressure. We do not
684: particularly care whether the stress is due to chaotic or regular
685: mechanisms. In this spirit,
686: we define $\alpha \equiv |M^{R\phi}|/p$.
687: %%%
688: Unlike $\alpha_{SS}$, which, in
689: most studies, is taken to be a constant independent of time and location,
690: $\alpha$ in this simulation is a dynamical and highly variable quantity
691: that is determined by local and transient magnetic dynamics. For this
692: reason, the absolute value is significant; although $M^{R\phi}$ is
693: positive most places, there are counter-examples.
694:
695: To give a sense of its scale and range of variation, we first examine
696: $\langle \alpha \rangle_\phi$, the azimuthally-averaged version of this
697: ratio, as found in the late-time snapshot (Figure~11). In the main
698: portion of the disk, $\langle \alpha \rangle_\phi$ ranges between $\sim
699: 10^{-2}$ and $\sim 10^{-1}$. However, it rises dramatically (to
700: $\sim10$) at several scale-heights above and below the mid-plane. The
701: same tendency has been seen (with much better vertical resolution)
702: in stratified shearing box simulations (Miller \& Stone 2000).
703: $\langle \alpha \rangle_\phi$ also increases sharply at small radius.
704: In the mid-plane, $\langle \alpha \rangle_\phi \simeq 0.2$ at $R=3$,
705: but rises to $\sim 3$ at the inner radial edge of the simulation.
706: Moreover, the vertical
707: rise in $\langle\alpha\rangle_\phi$ also becomes sharper at small
708: radius, so that at $R=3$, $\langle \alpha \rangle_\phi \sim 10$ at
709: $|z|$ as small as 0.5.
710:
711: In addition to these systematic trends, $\alpha$ is also subject to
712: sizable fluctuations. Measured in the same way as for the magnetic
713: energy density and the gas density (i.e., in terms of fluctuations in
714: azimuth at fixed $R$ and $z$), the {\it rms} fractional fluctuation
715: amplitude in the bulk of the disk ranges from $\simeq 1.5$ to $\simeq
716: 3$, but at large altitude above the disk midplane, the fluctuations can
717: be much greater. The character of these fluctuations is illustrated by
718: a different projection of $\alpha$, which we call $\bar{\alpha}$:
719: \begin{equation}
720: \bar{\alpha} \equiv {\int \, dz \, \rho \alpha \over \int \, dz \, \rho},
721: \end{equation}
722: the density-weighted vertical average of $\alpha$. In this projection
723: (Figure~12), we clearly see the nature of the azimuthal fluctuations:
724: just as for the density fluctuations (shown on smaller scale in Figure~2),
725: the magnetic fluctuations are sheared filaments of limited angular coherence
726: length.
727:
728: These results highlight the limitations of the primary assumption made
729: in many previous studies of disk structure, that $\alpha_{SS}$ is a
730: universal constant, independent of time and location. By contrast, we
731: find that the ratio of stress to pressure varies strongly from place to
732: place, both in the sense of having large fluctuations and in the sense
733: of possessing strong systematic gradients. In many places the gas
734: pressure is relatively smooth while the magnetic stress varies
735: considerably. In other places the variation in $\alpha$ is simply due
736: to changes in the background pressure. For example, the magnetic
737: stress does not drop with height as rapidly as the pressure does.
738:
739: For purposes of comparison with standard disk models, it is possible to
740: compute (albeit with some trepidation) the value of $\alpha_{SS}$ found
741: in this simulation. This quantity, defined in terms of the Maxwell
742: stress, is shown as a function of radius in Figure~13 for the disk at
743: the end of the simulation. It is $\sim 0.1$ through most of the disk,
744: but, starting just outside the marginally stable orbit, it rises inward.
745: It reaches a peak $\sim 1$ near the inner radial boundary of the
746: simulation. The rise is primarily due to the rapid drop in gas
747: pressure as the accretion flow accelerates inward. Also shown is the
748: ratio of the Maxwell stress to the magnetic pressure (dashed line).
749: This $\alpha_{mag}$ is 0.2--0.3 throughout most of the disk, and
750: increases inside $r_{ms}$.
751:
752: Whether $\alpha_{SS}$ is meaningful depends on how it is used. As the
753: previous discussion emphasized, the local ratio of magnetic stress to
754: pressure has strong vertical gradients that are completely obscured
755: when the vertically-integrated version is used. Similarly, there are
756: large fluctuations in azimuth and as a function of time. For these reasons,
757: $\alpha_{SS}$ is useless as a local dynamical variable. However,
758: the approximate constancy of $\alpha_{SS}$ as a function of radius {\it
759: in the disk proper} means that it can be used as a rough estimator of
760: the surface density in the body of the disk, provided its use is not
761: extended to local properties nor is it thought to be accurate at better
762: than ``factor of a few" accuracy. This conclusion is entirely
763: consistent with many standard uses of $\alpha$, and within the spirit
764: with which it was originally introduced by Shakura and Sunyaev (1973)
765: as a scaling parameter, not a fundamental variable.
766:
767: \subsection{Accreted angular momentum and energy}
768:
769: Another significant diagnostic of the flow is the mean specific angular
770: momentum distribution. There are a variety of ways to characterize the
771: specific angular momentum; one is by vertical and
772: azimuthal average, $\langle \ell \rangle$, as defined in equation
773: (\ref{el}). This function was discussed in Hawley (2000) for the model
774: GT4 (see Fig.~14 in Hawley 2000); the present model is very similar.
775: At late time the average specific angular momentum is slightly
776: sub-Keplerian throughout most of the disk. The most striking feature
777: is that $\langle \ell \rangle$ continues to decline inside $r_{ms}$
778: rather than leveling off. If all stress ceased at the marginally
779: stable orbit, material would travel inward from there retaining
780: precisely its energy and angular momentum. Conversely, the degree to
781: which fluid changes its specific energy and angular momentum is a
782: measure of the stresses that exist in the plunging region. The mean
783: specific angular momentum diminishes from 2.60 at $r_{ms}$ to 2.48 at
784: the innermost radius. Thus, on average in this snapshot, the accreted
785: matter has lost $\sim$5\% of the angular momentum it had when it
786: was at the marginally stable orbit. It is likely (for reasons
787: explained in \S 5) that this number is an underestimate, and in reality
788: the specific angular momentum at the innermost radius is rather
789: smaller.
790:
791: The total rate at which angular momentum is carried across the inner
792: radial boundary is slightly smaller than these numbers would suggest
793: because there is also a small outward electromagnetic angular momentum
794: flux. This electromagnetic flux is due to a generic property of
795: magnetic fields in shearing plasma: when the orbital frequency
796: decreases outward, the field is swept back so as to transport angular
797: momentum outward, i.e., $\langle B_R B_\phi \rangle < 0$. However,
798: near the inner radial boundary this is a small effect, the
799: electromagnetic angular momentum flux there is only about 2\% of the
800: angular momentum flux carried by the matter.
801:
802: Further complexity is revealed by a closer examination of the $(R,z)$
803: distribution of the late time azimuthally-averaged angular momentum
804: inside the plunging region. In fact, the specific angular momentum
805: in the cells lying along the equatorial plane
806: is nearly constant inside $r_{ms}$. However, the value of
807: $\ell$ drops with $z$ above and below the equator. This is consistent
808: with the distribution of magnetic field which, of course, accounts for
809: the nonzero stress. The low value of field near the equator is in part
810: a consequence of the assumed symmetry of the initial distribution of
811: field. The initial field loops are symmetric with respect to the
812: equator ($B_R$ and $B_\phi$ are zero there), and throughout the run the
813: strongest fields tend to be located above and below the equatorial
814: plane.
815:
816: Just as for every other quantity, there is also considerable spiral structure.
817: Figure~14 shows the density-weighted vertically-averaged specific
818: angular momentum. It makes plain how non\-axisymmetric the distribution of
819: torque is in the plunging region.
820:
821: For a number of reasons, the accreted energy is harder to determine
822: than the accreted angular momentum (see further discussion in \S 5.2).
823: To introduce this topic, we begin by displaying (Figure~15) the
824: specific energy $e$ in rest-mass units,
825: \begin{equation}\label{specenergy}
826: ec^2 = {1\over 2} v^2 + {(\Gamma -1)P\over \rho }
827: + {B^2\over8\pi \rho } - {1\over (r-r_g)},
828: \end{equation}
829: in the equatorial plane at late time. The specific energy for a
830: circular orbit precisely at $r_{ms}$ is -0.0625; here, $\langle e
831: \rangle_\phi$ is a bit greater than that ($\simeq -0.055$) because of
832: the thermal, magnetic, and turbulent kinetic energy contributions to
833: the total energy. In the equatorial plane, $\langle e \rangle_\phi$ is
834: almost constant from $R=3$ to just outside $R = 2$, but actually {\it
835: rises} as $R$ decreases from there to $R=1.6$, reaching a peak that is
836: very nearly unbound: -0.01. The azimuthally-averaged energy in the
837: equatorial plane then falls sharply to -0.07. Even this wild swing
838: underestimates the fluctuations: the smallest specific energy found in
839: the equatorial plane is -0.19. Thus, we are confident in asserting
840: that there is a large amount of energy transfer within the plunging
841: region. However, given these large fluctuations, we do not feel that
842: we can compute a well-defined mean energy for the accreted matter. It
843: is likely, moreover, that at least some of the very large fluctuations
844: in the innermost radial zones are not physical, but due to the inner
845: boundary condition.
846:
847: In fact, the situation is somewhat worse than this. The reason is that
848: dissipation is crucial for computing the accreted energy, and our
849: simulation contains no real dissipation physics. Absent numerical
850: losses (see \S 5.1, 5.2), energy taken from plunging matter may go
851: several places. One possibility is to build magnetic field energy in
852: the plunging region itself. If the field energy is advected into the
853: black hole without dissipation and radiative losses, the transfer of
854: energy from fluid to magnetic field makes no difference to the accreted
855: energy. It is also possible for magnetic stresses to transfer energy
856: all the way back to the disk proper (and indeed we see significant
857: magnetic stress throughout the plunging region). Again, without
858: dissipation, the energy transferred in that way stays in the fluid,
859: whether in the form of kinetic or magnetic energy. Only in the event
860: that the energy is transferred all the way out to non-accreting matter
861: would it be possible for that energy to avoid being carried back in.
862: Thus, to the degree that this simulation avoids numerical losses, it is
863: very difficult for it to show any significant transfer of energy out of
864: the plunging region. By contrast, in a real disk, dissipation changes
865: energy into heat, which may then be radiated if the thermal timescale
866: is shorter than the inflow timescale; in our simulation, neither
867: physical dissipation nor radiation is present.
868:
869: \section{Discussion}
870:
871: \subsection{Black hole spin-up}
872:
873: We are now ready to begin answering some of the questions posed in the
874: Introduction. Most significantly, this simulation has shown that the
875: stress due to MHD turbulence does not automatically cease at the
876: marginally stable orbit. In fact, it continues to be present
877: throughout the plunging region. As an immediate consequence, matter
878: continues to lose angular momentum as it travels from the marginally
879: stable orbit toward the event horizon. Because it then enters the
880: black hole with smaller angular momentum than if no stresses had been
881: present, the rate at which the black hole is spun up is slower than it
882: would have been otherwise (Agol \& Krolik 2000; Gammie 1999). It
883: further follows that estimates of the black hole spin distribution are
884: shifted toward slower rotation (e.g., Moderski \& Sikora 1996).
885: Although in this simulation the magnitude of this effect is small
886: ($\simeq 5\%$), it is possible that more realistic conditions will
887: increase its size (see \S 5.1).
888:
889: \subsection{Additional dissipation in the inner disk}
890:
891: We have already remarked on the fact that our simulation does not
892: directly compute the dissipation rate. In this section we attempt to
893: estimate what dissipation might be associated with dynamics of the sort
894: seen here. In conventional disk models, the local dissipation rate is
895: directly tied to the local stress:
896: \begin{equation}\label{dissipation}
897: Q = \left|{d\Omega \over d\ln r}\right| \int \, dz T^{R\phi},
898: \end{equation}
899: where $Q$ is the dissipation rate per unit area. This relationship is
900: predicated upon the notion that the same stress producing local
901: transport also produces local dissipation. This remains true for the
902: mean flow dynamics of MHD turbulence (Balbus \& Papaloizou 1999); in
903: fact, numerical dissipation mimicking the physical dissipation at the
904: short wavelength end of the turbulent cascade is the way in which
905: matter in this simulation loses energy as it moves slowly inward in the
906: body of the disk. However, this identification is not necessarily
907: correct as a prediction for the instantaneous rate of dissipation in
908: specific fluid elements. It is especially questionable in the plunging
909: region. The stress is not necessarily turbulent, and there may not be
910: enough time for the energy deposited in the fluid by the stress to be
911: transferred to small enough lengthscales for it to be dissipated.
912: Nonetheless, in this subsection we adopt the speculative assumption that
913: the vertically-integrated dissipation rate is proportional to the
914: vertically-integrated stress in order to speculate about the
915: consequences for disk radiation that may result.
916:
917: One possible consequence is an altered radial distribution of the disk
918: surface brightness. As shown in Figure~10, the azimuthally-averaged
919: stress in the simulation is nearly constant inside $R\simeq 5$. If the
920: relation of equation (\ref{dissipation}) holds, the total dissipation
921: rate outside $R=3$ in this simulation would be $\simeq 50\%$ greater
922: than in a disk with the same accretion rate, but zero stress at the
923: marginally stable orbit.
924:
925: If this were to happen, there would be a number of effects of
926: considerable phenomenological interest (Agol \& Krolik 2000). Because
927: this additional heat is released in a relatively small area of the
928: disk, it leads to relatively high temperatures, and therefore
929: contributes to the highest-frequency portion of the spectrum. Another
930: consequence stems from the fact that the dissipation is located in the
931: innermost part of the disk, where relativistic effects are strongest.
932: The light that is radiated due to this extra heating should be
933: comparatively strongly beamed into the equatorial plane by a
934: combination of special relativistic Doppler beaming driven by the
935: orbital motion and general relativistic trajectory-bending.
936: Trajectory-bending also forces a large fraction of the
937: photons to return to the disk at larger radius, to be reprocessed there
938: into lower-frequency light.
939:
940: Whether or not the enhanced stresses in the plunging region
941: lead to additional radiative losses
942: depends on the dissipation and photon diffusion rates. In the disk
943: proper, dissipation takes place in the turbulent cascade, in small
944: lengthscale fluctuations that are short-lived, exhibit little spatial
945: coherence, but are smoothly distributed in a statistical sense. In the
946: time-varying accretion of the plunging region this turbulent cascade
947: may not have time to go to completion. On the other hand, there might
948: be dissipative events, mediated by reconnection for example, that
949: involve structures with relatively large spatial coherence lengths and
950: which are not necessarily smoothly distributed throughout the region.
951: As we will discuss in \S 5.1, our limited spatial resolution makes it
952: difficult for us to estimate fairly the rate of such events in the
953: plunging region. Moreover, assessing the radiative properties of such
954: dissipative events is beyond the the capabilities of this simulation.
955: However, if events of this sort do occur, significant luminosity might
956: result because the reduced optical depth of the plunging matter (as
957: compared to the disk proper) also greatly reduces the thermal time.
958: Note that this mechanism results in an enhanced radiative efficiency
959: without the transfer of any energy from the plunging region to the disk
960: outside the marginally stable orbit.
961:
962: Finally, although the azimuthally-averaged stress matches conventional
963: expectations in much of the disk, there are large azimuthal variations
964: in the form of sheared filaments (Figure~12). Translated into a
965: dissipation rate, these spiral waves of enhanced (or diminished) stress
966: become spirals of enhanced (or diminished) heating. Whether this
967: concentration of heating has an effect on the emitted radiation depends
968: on the characteristic timescale of these fluctuations {\it as seen in a
969: specific fluid element}. If a particular fluid element has an elevated
970: heating rate for as long as its thermal timescale, the effective
971: radiating area of the disk is reduced to the area of the regions where
972: the heating is greatest, and the typical effective temperature is
973: raised accordingly. On the other hand, if over a thermal timescale
974: individual fluid elements see both positive and negative fluctuations,
975: the effects will average out. We have already estimated that the
976: characteristic speed of fluctuations through the disk is of order the
977: magnetosonic speed; on the other hand, the thermal timescale is $\sim
978: \alpha_{SS}^{-1}$ times the orbital period. Consequently, the ratio of
979: the coherence time of fluctuations to the thermal time should be, very
980: roughly, $\sim \alpha_{SS} v_{orb}/(v_{A}+c_s)$; in this simulation this
981: ratio is $\sim 1$.
982:
983: \subsection{Fluctuations in light output}
984:
985: We have emphasized throughout this paper the highly non-stationary
986: behavior that is intrinsic to disk behavior as found in this simulation.
987: One form these fluctuations can take is the coherent waves that
988: are continually radiated outward through the disk from the region of
989: the marginally stable orbit. Even if, on average, dynamics in the
990: plunging region change the mean accretion efficiency by only a small
991: amount, the large fluctuations we observe in the simulation should cause
992: substantial time-dependent variations in the light output.
993:
994: Once again, because there is no direct connection between the dynamics
995: traced by this simulation and dissipation, we cannot directly predict the time
996: variation of the disk luminosity. In fact, to do so would also require
997: calculating the photon diffusion rate from inside the (optically thick)
998: disk. However, some indication of what may occur may be gleaned from
999: Figures~6 and 7. Some of the fluctuations in the accretion rate
1000: (particularly those at the highest frequencies) are due to dynamics in
1001: the plunging region; these may or may not lead to dissipation (see \S
1002: 4.2), and therefore may or may not contribute to fluctuations in the
1003: light output. For this reason, and following the {\it ansatz} of
1004: equation (\ref{dissipation}), in Figure~7 we also plot the Fourier
1005: power spectrum of the volume-integrated magnetic stress. Whether the
1006: accretion rate or the stress is more closely related to the dissipation
1007: rate, the actual output is the convolution of fluctuations in the
1008: heating rate with the probability distribution for the photon escape
1009: time. In frequency space, this amounts to a (position-dependent)
1010: low-pass filter. When the heating occurs deep inside the disk, the
1011: cut-off frequency for this filter is $\sim \alpha_{SS}\Omega$;
1012: fluctuations in the light due to heating closer to the surface are
1013: subject to a less stringent frequency cut-off (in the optically thin
1014: limit, there is no cut-off at all). Consequently, we expect the
1015: Fourier power spectrum of the disk luminosity to be rather strongly cut
1016: off at frequencies above $\sim 10^{-2}$, but the power spectrum of
1017: fluctuations associated with coronal dynamics may be different.
1018:
1019: The observation of QPOs in black hole candidates has
1020: motivated a number of studies of possible sources of systematic time
1021: variability in accretion disks. In a series of papers (Nowak \&
1022: Wagoner 1991, 1992; Perez et al. 1997) Wagoner and collaborators have
1023: explored the linear theory of fluid oscillations in the relativistic
1024: portions of stationary accretion disks. In these disks only
1025: hydrodynamic effects are considered, and angular momentum transport is
1026: modeled by a pseudo-viscosity parameterized by a constant
1027: $\alpha$. These disks possess a variety of normal modes, including
1028: $p$-modes that are trapped between $r_{ms}$ and the radius where the
1029: epicyclic frequency is maximum, and $g$-modes, whose greatest amplitude
1030: is found under the peak of the epicyclic frequency curve. In the
1031: Paczy\'nski-Wiita potential we use, the position of the maximum
1032: epicyclic frequency is $R \simeq 3.6$; with the additional correction
1033: terms introduced by Nowak and Wagoner (1991), it moves outward to
1034: $R=4$. We would not expect to find exactly the same mode frequencies
1035: as they because of the difference in the potentials used; however, the
1036: mode frequencies should in general be $\sim \Omega(r_{ms})$. A perusal
1037: of Figure~7 fails to find any evidence for special frequencies in this
1038: range.
1039:
1040: There are several reasons why trapped disk oscillations do not appear
1041: here. First, the normal modes discussed in the previous paragraphs are
1042: predicted on the basis of a purely {\it hydrodynamic} analysis; no
1043: account is made of magnetic forces. Here, however, magnetic forces
1044: play a major role in determining the dynamics of fluid elements. For
1045: example, if one repeated the derivation of the ``diskoseismic"
1046: dispersion relation, but allowed for a weak magnetic field, one would
1047: rediscover the magneto-rotational instability, thus vitiating the
1048: hydrodynamic ``normal modes". More fundamentally, our simulation
1049: indicates that there is no underlying quiet flow to support coherent
1050: oscillations. The disk is so turbulent that there is no steady
1051: equilibrium against which linear perturbations can develop. Put
1052: another way, we find that the typical radial motion timescale in the
1053: vicinity of $r_{ms}$ (where these normal modes are supposed to be
1054: concentrated) is comparable to the predicted mode frequencies, so the
1055: fundamental assumption that these modes are perturbations to circular
1056: orbits is inappropriate.
1057:
1058: \section{Limitations}
1059:
1060: The results presented here and in Hawley (2000) obviously represent
1061: only the first steps in detailed simulations of global accretion
1062: disks. In this section we outline how the approximations we have made
1063: may impact our conclusions. Of course the only definitive way to
1064: investigate the effects of these approximations would be with
1065: additional simulations, and this discussion should be viewed as a guide
1066: for future work.
1067:
1068: \subsection{Effects of resolution}
1069:
1070: Foremost among the limitations is the numerical resolution of the
1071: simulation. Although we placed 30 radial zones between $R=1.5$ and
1072: $R=4$, that is probably insufficient. There are two reasons for
1073: believing this to be the case. First, we may apply the standard test
1074: for convergence of comparing two simulations with the same physics but
1075: different resolution, and ask whether they show significant
1076: differences. Simulation GT4, reported in Hawley (2000), and a new
1077: $64^3$ run (hereafter referred to as run LR for Lower Resolution) serve
1078: as good comparisons. Both GT4 and LR share identical initial and
1079: boundary conditions. GT4 had equally spaced grid zones over a distance
1080: of 20 in both $R$ and $z$ for $\Delta R = \Delta z = 0.156$.
1081: Simulation LR used equally spaced radial zones over a range of 30 for
1082: $\Delta R = 0.47$, and used half of the $z$ grid zones centered around
1083: the equator within $z = \pm 2$, for $\Delta z= 0.125$. By contrast,
1084: here $\Delta R = 0.0833$ and $\Delta z = 0.0625$ in the inner region.
1085:
1086: All three simulations are qualitatively similar in global properties,
1087: as measured by volume integration, but differ in certain other respects
1088: that are important to physical interpretation. The time-averaged
1089: accretion rate in GT4 is about 40\% smaller than in the work reported
1090: here (although the time-averaged accretion rate in LR is only about 2\%
1091: smaller). However, the most striking contrast is in magnetic field
1092: strength. Higher resolution results in larger magnetic fields,
1093: particularly in the poloidal components. Contrasting maps of the
1094: magnetic field intensity at similar late time for GT4 and our new
1095: simulation (Figure~16), we find that the field is consistently stronger
1096: in the new simulation everywhere inside $R=5$, rising to typically a
1097: factor of 3 greater in magnitude inside $R=4$. Significantly, outside
1098: $R=5$, the field is generally stronger in the old simulation (sometimes
1099: by more than a factor of 10), but that, too, is in keeping with the
1100: sense of the resolution contrast: GT4 has better resolution in the
1101: outer regions.
1102:
1103: The field intensity is controlled by a trade-off between the turbulent
1104: MHD dynamo driven by the magneto-rotational instability and numerical
1105: dissipation (recall that there is no explicit resistivity in this
1106: simulation). Whatever the real physics of resistivity in accretion
1107: disks may be, it is unlikely that it is as large as the effective
1108: numerical resistivity induced by our modest spatial resolution. The
1109: contrast between the field intensity in the two simulations is
1110: therefore very likely explained by the smaller numerical resistivity of
1111: the better-resolved simulation.
1112:
1113: Another piece of evidence supporting the belief that the magnetic field
1114: is weaker than it would be with better resolution is presented in
1115: Figure~17, where we display the magnetic field structure in the
1116: equatorial plane within $R=4$, plotted for every fourth grid zone.
1117: The planar component of the field is
1118: organized into long sheared loops. Because the loops are wrapped
1119: around one another, there are many places where the field direction
1120: reverses within 4 -- 8 grid cells. This means that with a modest
1121: amount of flow convergence numerical effects could cause the field to
1122: decay on a relatively short timescale.
1123:
1124: Finally there is the issue of the computational domain employed, in
1125: particular the location of the boundaries. Very little outflow
1126: leaves through the $z$ boundaries (typically only a few
1127: percent of the accretion rate). The outer radial boundary was
1128: extended in the present simulation compared to GT4, which had the
1129: effect of keeping more of the disk on the grid, but otherwise had no
1130: obvious effect on the evolution. We expect that there is also little
1131: dependence upon the location of the inner boundary because the
1132: flow across that boundary is generally supermagnetosonic.
1133: Although we made no tests here of the effect of varying $r_{in}$, tests
1134: of this sort were made on the similar simulations in Hawley (2000),
1135: and it appeared to make little difference.
1136:
1137: \subsection{Energy conservation and physical dissipation rates}
1138:
1139: The artificial dissipation caused by insufficient resolution also has
1140: consequences for energy conservation. To very good accuracy, angular
1141: momentum is conserved in this simulation. The primary source of
1142: numerical dissipation, the truncation error associated with transport
1143: terms in the equation, does not alter the total angular momentum,
1144: although it may artificially change its distribution.
1145: Total energy, however, is different because the energy equation that is
1146: solved is the internal energy equation. Numerical dissipation can
1147: destroy kinetic energy in regions of strong velocity gradients and
1148: magnetic energy in regions of strong magnetic field gradients, but
1149: there is no provision in this simulation for putting that energy back
1150: into heat (except for dissipation in shocks through the use of an
1151: artificial viscosity). Both sorts of dissipation can also occur in
1152: real systems, but at rates that might be very different from what
1153: happens in the simulation.
1154:
1155: In the body of the disk, the problems that numerical dissipation can
1156: cause are limited. The single largest items in the energy budget are
1157: potential energy and rotational kinetic energy; neither one is
1158: significantly affected by numerical effects. The thermal energy is
1159: only a small fraction of those energies; this is a prime motivation
1160: for solving an internal rather than total energy equation.
1161: Artificial energy losses occur in the dissipation of relatively small
1162: amounts of energy bound up in small-scale random fluid motions and in
1163: magnetic field. In fact, in this regard, numerical dissipation acts in
1164: a way that is consistent with the physical dissipation of energy at the short
1165: lengthscale end of the turbulent cascade. In both cases, the losses
1166: occur on the shortest lengthscales, and in both cases (assuming the
1167: disk radiates efficiently) energy is lost
1168: from the disk on timescales short compared to the inflow time.
1169:
1170: However, near and inside the marginally stable orbit, detailed
1171: questions of energy transfer become significant. The contributions to
1172: the energy budget due to both turbulent motions and magnetic field
1173: grow. At the same time, their importance also grows. For example, it
1174: would be of great interest to determine whether plunging fluid elements
1175: can, by exerting magnetic stresses, transfer energy back to the disk
1176: proper (cf. \S 3.7). As we have seen, significant magnetic stresses
1177: exist throughout this region, so there is some possibility that this happens.
1178: However, the rate at which this may occur is regulated by the magnetic field
1179: strength and topology, and these are subject to significant artificial
1180: numerical dissipation. In addition, any energy that is delivered is
1181: likely to take the form of MHD turbulence; whether this turbulence is
1182: dissipated (and the energy radiated) quickly or slowly compared to the
1183: inflow is critical for evaluating the radiative efficiency that
1184: results. Because the inflow, thermal, and dissipation timescales all
1185: change significantly in the vicinity of $r_{ms}$, it is unlikely that
1186: in this region numerical dissipation serves as a successful stand-in
1187: for physical dissipation and radiative losses in the way it does at
1188: larger radii.
1189:
1190: \subsection{Equation of state}
1191:
1192: Dissipation and radiation are, of course, key elements in the gas
1193: thermodynamics. In the present simulation, we describe the gas's
1194: internal energy in only an approximate way: we assume it is
1195: adiabatic except for heating due to artificial viscosity.
1196: Explicit treatment of dissipation and radiation would fill in the
1197: missing terms in the gas energy equation.
1198:
1199: In the absence of a full treatment of the energy equation, we can
1200: at best make educated speculations about the impact of our assumption
1201: that the gas behaves adiabatically. Some consequences were already
1202: discussed in the previous subsection. There are several more, not
1203: as closely related to resolution issues.
1204:
1205: As we have remarked earlier (\S 3.7), one of the most dramatic features
1206: of this simulation is the coherent disturbances that propagate as
1207: magnetosonic waves from the vicinity of the marginally stable orbit
1208: outward through the disk. Their behavior depends strongly, of course,
1209: on the sound speed in the disk. However, in this simulation, in which
1210: all gas is given the same specific entropy initially and then evolved
1211: adiabatically, the sound speed could be quite different from what it
1212: would be in a more realistic simulation that includes heating from
1213: dissipation of MHD turbulence and cooling via radiation. We
1214: expect, therefore, that properties that depend on pressure forces (such
1215: as the propagation of these waves, or the vertical thickness of the
1216: disk) are not well-described in this simulation.
1217:
1218: Another consequence of our assumed equation of state, combined with
1219: the particular value of specific entropy we assigned the gas, is that
1220: the disk is modestly thick. As a result, it begins from a non-Keplerian,
1221: partially pressure-supported state and radial pressure gradients
1222: drive some of its initial evolution. Magnetic stresses
1223: rapidly change the angular momentum distribution within the torus to
1224: nearly Keplerian, but without cooling, the internal pressure continues to
1225: push matter away from the pressure maximum. Some accretion,
1226: therefore, is driven not by magnetic stresses but simply by the pressure
1227: gradient. At late time, however, the inner disk establishes a
1228: quasi-steady state and nearly all the accretion can be accounted for by
1229: the Maxwell stress. It is likely that the observed dynamics are
1230: reasonably representative of generic inner disk accretion.
1231:
1232: Finally, although the present disk is moderately thin, with $H/R
1233: \simeq 0.15$ in the inner region, it is conceivable that new
1234: phenomena might appear in a still thinner disk. Although our results
1235: provide no particular reason to predict this, it is possible that in a
1236: disk with still smaller $H/R$ there might be modes with wavelength
1237: large compared to $H$ yet still small compared to $R$. Questions such
1238: as these could be investigated by contrasting the present work with
1239: simulations of the inner regions of more extended, thinner, initially
1240: Keplerian disks.
1241:
1242: A possible example of this sort of comparison is provided by the work
1243: of Armitage et al. (2000). Like us, they also simulated the inner
1244: region of an accretion disk in a pseudo-Newtonian potential using a
1245: three dimensional
1246: MHD code. They find a smaller reduction of $\ell$ inside $r_{ms}$ than
1247: we find in the simulation reported here, an effect they conjecture
1248: could be due to the smaller temperature and scale height in their
1249: disk. However, their scale height is $\simeq 0.08$, not much smaller
1250: than our typical value of $\simeq 0.15$. Moreover, there are other
1251: differences between the two simulations that may also account for
1252: quantitative contrasts: in their stratified simulation, only a limited
1253: vertical extent is simulated; their equation of state is isothermal;
1254: and they simulate a $\pi/3$ wedge, rather than a full $2\pi$ in
1255: azimuth. Because in our simulation fluctuations with azimuthal
1256: wavelength $\sim 1$~radian or more have considerable amplitude, the
1257: limitation to a wedge could have significant consequences. In any
1258: event, as they emphasize, the differences between their simulations and
1259: ours are simply quantitative, not qualitative. They, too, find a
1260: nonzero Maxwell stress inside $r_{ms}$, but its amplitude is lower.
1261: Further work is required to determine what sets the stress levels in a
1262: variety of different situations.
1263:
1264: Thus, for all these reasons, it will be important for future simulations
1265: to incorporate improved treatment of gas heating and cooling. It
1266: should not be difficult to integrate a total energy equation along with
1267: the gas momentum (i.e., force) equation. Heating could be treated in a
1268: controlled way by inserting an explicit resistivity into the magnetic
1269: field evolution equation and an explicit viscosity into the fluid force
1270: equation, and placing corresponding terms in the internal energy equation.
1271: Radiative losses could be described with either an escape probability
1272: formalism, or, in the longer term, solution of the radiation transfer
1273: problem.
1274:
1275: A simulation with both heating and radiation would simulate
1276: the behavior of thin, radiative disks. There has been no previous
1277: work of this sort; the significance of such an effort is obvious.
1278:
1279: Heating alone would, in effect, lead to a simulation of the popular
1280: advection dominated accretion flow (ADAF) model. Some previous
1281: simulations of this model have used a phenomenological prescription for
1282: viscosity (e.g., Igumenshchev \& Abramowicz 1999; Igumenshchev,
1283: Abramowicz, \& Narayan
1284: 2000; Stone, Pringle \& Begelman 1999). Others have computed the
1285: stress from a global magnetic field (in axisymmetry: Uchida \& Shibata
1286: 1985 and Stone \& Norman 1994; in three dimensions: Matsumoto 1999).
1287: Still another approach has been to study the stress created by MHD
1288: turbulence including resistive dissipation, but in axisymmetry (Stone
1289: \& Pringle 2000). Addition of explicit dissipation to the simulation
1290: presented here would, in effect, extend the Stone \& Pringle work to
1291: three dimensions.
1292:
1293: \subsection{Duration}
1294:
1295: The timestep of the simulation is limited by the Courant condition.
1296: The most stringent limitation comes from the high orbital velocities
1297: and the radial plunging speeds in the inner part of the grid where the
1298: resolution is reasonably fine. The total duration of the simulation
1299: was 1500 time units, requiring almost 400,000 timesteps. As discussed
1300: in \S3.1, this is long enough for a quasi-steady state to be
1301: established in the inner part of the disk. Since $\alpha \sim 0.1$,
1302: the simulation covers several accretion times in the inner part of the
1303: disk. Whether or not the results obtained are truly representative,
1304: must, however, be tested with simulations of longer duration. An
1305: evolution beginning with a nearly Keplerian disk that extends out to
1306: large radius could allow the inner disk region to establish itself
1307: self-consistently from a long-term accretion flow.
1308:
1309: \subsection{Field topology}
1310:
1311: What are the limitations due to the choice of initial disk and field
1312: topology? The initial poloidal field loops used here are amenable to
1313: rapid development of the MRI on large scales. Further, the presence of
1314: significant initial radial field quickly generates through simple shear
1315: amplification a toroidal field of near-equipartition strength. Thus,
1316: the initial field topology guarantees magnetically-driven evolution
1317: within a few orbits. In earlier work (Hawley 2000) it was shown that
1318: in this regard (generating magnetically-driven inflow after a few orbits),
1319: initially poloidal and initially toroidal field configurations
1320: are essentially
1321: equivalent. However, the symmetry inherent in this
1322: choice of initial field topology is partly responsible
1323: for the low field intensity near the equatorial plane and the weak
1324: dipolar structure characteristic of the low-density regions (\S 3.3).
1325:
1326: Whether the field is contained entirely within the problem area,
1327: or, equivalently, whether there is zero net flux through the box,
1328: has a greater qualitative impact. Shearing box simulations (Hawley,
1329: Gammie \& Balbus 1995, 1996) also found strong similarities between
1330: simulations with initial toroidal field and simulations with random
1331: initial fields. However, they found dramatically different behavior
1332: when there was net vertical field. Such field topologies tend to
1333: produce larger net angular momentum transport and more violent
1334: evolution. These effects have also been observed in
1335: global simulations (Matsumoto 1999).
1336:
1337: \subsection{Displacement current}
1338:
1339: The equations of MHD as we have written them (see especially
1340: equation~2) omit the displacement current term. This approximation is
1341: valid so long as the Alfv\'en speed $v_A < c$. This condition is
1342: satisfied almost everywhere in this simulation, with the only exception
1343: a small region near the inner radial boundary ($1.5 \leq R \leq 2$ and
1344: $1 \leq |z| \leq 2$). In the equatorial plane, where most of the mass
1345: flows, the largest value of $v_A \simeq 0.3$, and that occurs only at
1346: the inner radial boundary; everywhere outside $R = 2$ in the midplane,
1347: $v_A \leq 0.15$. Future simulations in which the displacement current
1348: term is included should be somewhat more accurate, but we doubt that
1349: this correction will cause any major changes.
1350:
1351: \subsection{Relativity}
1352:
1353: Obviously, it would be preferable to run the simulation using genuine
1354: general relativistic dynamics, rather than the approximation of
1355: Newtonian physics in a Paczy\'nski-Wiita potential. We might hope that
1356: this approximation is qualitatively reasonable when attempting to mimic
1357: dynamics around a non-rotating black hole; after all, the most
1358: important aspect of this case is the existence of dynamically unstable
1359: orbits inside a critical radius and enforced inward radial motion
1360: within a still smaller distance from the center. Further, Gammie \&
1361: Popham (1998) compute the growth rate of the MRI in full relativity and
1362: have shown that it remains undiminished $r_{ms}$. The
1363: Paczy\'nski-Wiita potential, however, has no representation of the
1364: frame-dragging created by a rotating black hole, so it cannot be a good
1365: approximation to accretion dynamics around a Kerr black hole. Gammie
1366: (1999) and Gammie \& Popham (1998) examine some of the interesting
1367: possibilities created by the Kerr geometry, but examination of those
1368: effects must await a code with true relativistic dynamics.
1369:
1370: \section{Conclusions}
1371:
1372: The simulation reported here represents another step towards more
1373: realistic simulations of accretion disk dynamics. By improving the
1374: grid resolution in the inner region, we have shown that magnetic
1375: effects become increasingly important near and inside the radius of the
1376: marginally stable orbit. Contrary to the most widely-held assumption,
1377: the $R$--$\phi$ component of the stress tensor does not go to zero at
1378: $r_{ms}$; averaged over azimuth, height, and time, it is almost
1379: constant as a function of $R$ for $R \leq 2 r_{ms}$. As a result,
1380: matter in the region of unstable orbits does not follow simple energy-
1381: and angular momentum-conserving free-fall trajectories, although we are
1382: not yet ready to make quantitative predictions about how much the
1383: energy and angular momentum change.
1384:
1385: The ratio $T^{R\phi}/P \equiv \alpha$ is commonly used as a
1386: measure of the magnitude of the stress, and almost as commonly is
1387: thought to be a constant, independent of location and time. Our
1388: simulation showed clearly that while it can provide a good
1389: qualitative measure of the stress when averaged over several orbits
1390: in time, at any particular instant $\alpha$ has large systematic
1391: gradients (as a function of both $R$ and $z$) and substantial
1392: temporal fluctuations. On average the value of $\alpha$ in the
1393: disk is $\sim 0.1$.
1394:
1395: This simulation has shown that substantial time- and space
1396: variations are the rule. The inner regions of accretion disks around
1397: black holes are highly turbulent, non-steady systems. Large-amplitude
1398: fluctuations sheared into spiral fragments come and go, many of them
1399: initiated in the region immediately outside the marginally stable
1400: orbit. As a result, there are continuing fluctuations in the accretion
1401: rate; we expect that some of these fluctuations will be mirrored in the
1402: light curves of accreting black holes.
1403:
1404: Future simulations, guided by these results, will probe the beyond the
1405: limitations of the present simulation. Within the context of the same
1406: pseudo-Newtonian potential and the present adiabatic equation of state,
1407: we can explore the effects of different initial disk configurations and
1408: magnetic field strengths and topologies, at higher resolution and
1409: for longer time. Beyond this, it will be important to incorporate
1410: improved physics into the models. These include general relativistic
1411: dynamics and improved treatments of the equation of state and
1412: dissipative processes. This future work should bring us still closer
1413: to a quantitative understanding of the dynamics of accretion onto black
1414: holes.
1415:
1416: \begin{acknowledgments}
1417:
1418: JHK would like to thank Eric Agol for numerous helpful conversations,
1419: and Nancy Levenson for invaluable instruction in the wiles of IDL.
1420: He would also like to thank Giora Shaviv for comments on the difference
1421: between heating rate fluctuations and luminosity fluctuations.
1422: JFH thanks Charles Gammie for useful comments on the manuscript.
1423: This work was supported by NSF grant AST-0070979, and NASA grants
1424: NAG5-9266 and NAG5-7500 to JFH, and NSF Grant AST-9616922 and NASA
1425: grant NRA-99-01-ATP-031 to JHK. Simulations were carried out on the
1426: Cray T3E system of the San Diego Supercomputer Center of the National
1427: Partnership for Advanced Computational Infrastructure, funded by the
1428: NSF.
1429:
1430: \end{acknowledgments}
1431:
1432: \begin{thebibliography}{}
1433:
1434: \bibitem[Abramowicz \& Kato]{ak89} Abramowicz, M.A., \& Kato, S. 1989, ApJ,
1435: 336, 304
1436: \bibitem[Abramowicz \& Zurek 1981]{az81} Abramowicz, M.A., \& Zurek, W.H. 1981,
1437: ApJ, 246, 314
1438: \bibitem[Agol \& Krolik 2000]{ak00} Agol, E., \& Krolik, J.H. 2000,
1439: ApJ, 528, 161
1440: \bibitem[Armitage 1998]{a98} Armitage, P.J. 1998, ApJ, 501, L189
1441: \bibitem[Armitage 2000]{a00} Armitage, P.J., Reynolds, C.S., \& Chiang,
1442: J. 2000, ApJ, submitted
1443: \bibitem[Balbus \& Hawley 1991]{bh91} Balbus, S.A., \& Hawley, J.F. 1991, ApJ,
1444: 376, 214
1445: \bibitem[Balbus \& Hawley 1998]{bh98} Balbus, S.A., \& Hawley, J.F. 1998, Rev
1446: Mod Phys, 70, 1
1447: \bibitem[Balbus \& Papaloizou 1999]{bp99} Balbus, S.A., \& Papaloizou,
1448: J. C. B. 1999, ApJ, 521, 650
1449: \bibitem[Brandenburg et al. 1996]{b96} Brandenburg, A., Nordlund, \AA,
1450: Stein, R.~F., \& Torkelsson, U. 1995, ApJ, 446, 741
1451: \bibitem[Gammie 1999]{g99}Gammie, C.F. 1999, ApJ, 522, L57
1452: \bibitem[Gammie 1998]{gp98}Gammie, C. F., \& Popham, R. 1998, ApJ, 498, 313
1453: \bibitem[Hawley 2000]{h00} Hawley, J.F. 2000, ApJ, 528, 462
1454: \bibitem[Hawley, Gammie \& Balbus 1995]{hgb95} Hawley, J.F., Gammie, C.
1455: F., \& Balbus, S.A. 1995, ApJ, 440, 742
1456: \bibitem[Hawley, Gammie \& Balbus 1995]{hgb96} Hawley, J.F., Gammie, C.
1457: F., \& Balbus, S.A. 1996, ApJ, 464, 690
1458: \bibitem[Hawley \& Stone 1995]{hs95}Hawley, J. F., \& Stone, J. M. 1995,
1459: Comp Phys Comm, 89, 127
1460: \bibitem[Igumenshchev \& Abramowicz 1999]{ia99}
1461: Igumenshchev, I. V., \& Abramowicz, M. A. 1999, MNRAS, 303, 309
1462: \bibitem[Igumenshchev et al. 2000]{i00}
1463: Igumenshchev, I. V., Abramowicz, M. A., \& Narayan, R. 2000, ApJ, 537, 27
1464: \bibitem[Krolik 1999]{h99}Krolik, J.H. 1999, ApJ, 515, L73
1465: \bibitem[Machida et al. 2000]{mhm00} Machida, M., Hayashi, M.~R., \&
1466: Matsumoto, R. 2000, ApJ, 532, L67
1467: \bibitem[Matsumoto 1999]{m99} Matsumoto, R. 1999, in Numerical Astrophysics,
1468: eds. Miyama, S., Tomisaka, K. \& Hanawa, T (Kluwer: Dordrecht), 195
1469: \bibitem[Matsumoto et al. 1984]{m84} Matsumoto, R., Kato, S., Fukue, J.,
1470: \& Okazaki, A.T. 1984, PASJ, 36, 71
1471: \bibitem[Miller \& Stone 2000]{ms00} Miller, K.A., \& Stone, J.M. 2000, ApJ,
1472: 534, 398
1473: \bibitem[Moderski \& Sikora 1996]{ms96} Moderski, R., \& Sikora, M. 1996,
1474: MNRAS, 283, 854
1475: \bibitem[Muchotrzeb\& Pacy\'nski 1982]{mp82} Muchotrzeb, B., \& Paczy\'nski, B.
1476: 1982, Acta. Astr., 32, 1
1477: \bibitem[Novikov \& Thorne 1973]{nt73} Novikov, I.D., \& Thorne, K.S. 1973, in
1478: Black Holes, eds. C. de Witt and B. de Witt (New York: Gordon \& Breach), 343
1479: \bibitem[Nowak \& Wagoner 1991]{nw91} Nowak, M., \& Wagoner, R.V. 1991,
1480: ApJ, 378, 656
1481: \bibitem[Nowak \& Wagoner 1992]{nw92} Nowak, M., \& Wagoner, R.V. 1992,
1482: ApJ, 393, 697
1483: \bibitem[Paczy\'nski \& Wiita 1980]{pw80}Paczy\'nski, B., \& Wiita, P. J.
1484: 1980, A\&A, 88, 23
1485: \bibitem[Page \& Thorne 1974]{pt74} Page, D.N., \& Thorne, K.S. 1974,
1486: ApJ 191, 499
1487: \bibitem[Perez et al. 1997]{p97} Perez, C.A., Silbergleit, A.S., Wagoner,
1488: R.V., \& Lehr, D.E. 1997, ApJ, 476, 589
1489: \bibitem[Shakura \& Sunyaev 1973]{ss73} Shakura, N.~I., \& Sunyaev, R.~A.
1490: 1973, A\&A, 24, 337
1491: \bibitem[Stone et al. 1996]{s96} Stone, J. M., Hawley, J. F., Gammie, C.
1492: F., \& Balbus, S. A. 1996, ApJ, 463, 656
1493: \bibitem[Stone \& Norman 1992a]{sn92a} Stone, J. M., \& Norman, M. L.
1494: 1992a, ApJS, 80, 753
1495: \bibitem[Stone \& Norman 1992b]{sn92b} Stone, J. M., \& Norman, M. L.
1496: 1992b, ApJS, 80, 791
1497: \bibitem[Stone \& Norman 1994]{sn94} Stone, J. M., \& Norman, M. L. 1994,
1498: ApJ, 433, 746
1499: \bibitem[Stone \& Pringle 2000]{sp00} Stone, J. M, \& Pringle, J. E.
1500: 2000, MNRAS, in press
1501: \bibitem[Stone et al. 1999]{spb99} Stone, J. M, Pringle, J. E., \&
1502: Begelman, M. C. 1999, MNRAS, 310, 1002
1503: \bibitem[Sunyaev \& Revnivtsev 2000]{sr00} Sunyaev, R., \& Revnivtsev,
1504: M. 2000, A\&A, 358, 617
1505: \bibitem[Uchida \& Shibata 1985]{us85} Uchida, Y., \& Shibata, K. 1985,
1506: PASJ, 37, 515
1507: \end{thebibliography}
1508:
1509: \clearpage
1510: \centerline{Figure Captions}
1511:
1512: \figcaption{Azimuthally-averaged gas density distribution at the beginning
1513: ($t=0$: panel a) and the end ($t=1500$: panel b) of the simulation.
1514: There are 14 evenly spaced contours. The maximum density at $t=0$ is 34.2;
1515: at $t=1500$, $\rho_{max}=14.3$.}
1516:
1517: \figcaption{Density in the equatorial plane for $R \leq 10$ at $t=1500$.}
1518:
1519: \figcaption{Azimuthally-averaged magnetic pressure, $\langle
1520: B^2/8\pi\rangle_\phi$, plotted on a logarithmic scale from $10^{-5}$
1521: to $10^{-2}$. The mean value of magnetic pressure
1522: within the region where $\rho > 0.01$ is $7.1\times 10^{-4}$.
1523: Using the mean gas pressure this corresponds to an averge
1524: $\beta = 13$ in the disk.
1525: }
1526:
1527:
1528: \figcaption{Poloidal magnetic field structure in the slice at $\phi = \pi/2$,
1529: overlaid on density contours. Panel a) shows the inner, high-resolution
1530: region; panel b) shows the disk body. The length of each arrow
1531: is proportional to the magnitude of the poloidal field, but the scales
1532: are different in the two panels.}
1533:
1534: \figcaption{Accretion rate (in code units) as a function of
1535: radius at times $t=1000$
1536: and $t=1500$ (solid lines). The dashed curve is the time-average of the
1537: accretion rate between those two times. Note that
1538: negative accretion rate means outward flow.}
1539:
1540: \figcaption{Mass accretion rate $\langle \dot M \rangle$ through the
1541: inner radial boundary as a function of time. To capture the
1542: time-variability accurately, the accretion rate was measured every 10
1543: timesteps in the simulation, corresponding to an average interval of
1544: 0.0037 in time.}
1545:
1546: \figcaption{
1547: Fourier power density of the accretion rate through the inner boundary
1548: (solid curve) and volume-integrated Maxwell stress (dashed curve).
1549: Frequency is in code units; the orbital frequency (i.e., the inverse of
1550: the orbital period $\Omega(r_{ms})/2\pi$) at the marginally stable
1551: orbit is marked with an arrow.}
1552:
1553: \figcaption{ Accretion rate at cylindrical radius $R=5$; inward mass flow
1554: is given a positive sign. Note that the
1555: {\it local} mass flux is almost as likely to be outward as inward.}
1556:
1557: \figcaption{Vertically- and azimuthally-averaged mass flux as a function of
1558: radius $R$ and time. The diagonal stripes denote radially-propagating
1559: waves. Reds and yellows correspond to accretion, blues to outflow.
1560: }
1561:
1562: \figcaption{Azimuthally-averaged and vertically-integrated magnetic
1563: stress averaged from $t=750$ until the end of the simulation
1564: (solid curve) and nominal mean stress per unit area (dashed curve).}
1565:
1566: \figcaption{Absolute value of $\langle \alpha \rangle_\phi$, in a
1567: logarithmic scale in the snapshot at $t=1500$.
1568: This quantity increases with decreasing radius, and
1569: with increasing distance above or below the disk midplane.}
1570:
1571: \figcaption{Absolute value of the vertically-averaged $\bar{\alpha}$, plotted
1572: on a logarithmic scale for the snapshot at $t=1500$. Note the
1573: large fluctuations, even when values are compared at fixed radius.
1574: $\bar{\alpha}$ is large at both large and small $R$, but for different
1575: reasons (see text).}
1576:
1577: \figcaption{Shakura-Sunyaev $\alpha_{SS}$ as a function of radius
1578: (solid line) constructed from vertical and azimuthally averaging the
1579: Maxwell stress and the gas pressure. Also shown is the averaged
1580: $\alpha_{mag}$ as a function of radius (dashed line), constructed using
1581: the magnetic pressure in place of the gas pressure. $\alpha_{mag}$ is
1582: relatively constant in the disk, whereas $\alpha_{SS}$ varies with the
1583: gas pressure.
1584: }
1585:
1586: \figcaption{Density-weighted vertically-averaged specific angular momentum
1587: in the inner portion of the accretion flow.}
1588:
1589: \figcaption{Specific energy of matter (in rest-mass units) in the
1590: equatorial plane within $R=4$ (the high-resolution segment of the
1591: simulation). Note the large amplitude spiral fluctuations. The sharp
1592: features immediately surrounding the inner radial boundary are
1593: artifacts.}
1594:
1595: \figcaption{Logarithm of the ratio of the density-weighted mean magnetic
1596: field intensity in the new simulation to the same quantity in GT4 at
1597: analogous late times. Only the region inside $R=10$ is shown.
1598: Because these are different simulations, one cannot expect features
1599: to match up precisely. However, there is a clear general
1600: tendency for the field in the new simulation to be stronger than in the
1601: old simulation inside $R=5$ and weaker outside.}
1602:
1603: \figcaption{Planar component of the magnetic field in the equatorial
1604: slice of the high-resolution region of the simulation. The length of
1605: each arrow is proportional to the intensity of the magnetic field. For
1606: clarity, the field is shown in every fourth cell.}
1607:
1608: \end{document}
1609:
1610:
1611: