1: %\documentclass[12pt,preprint]{aastex}
2: \documentclass[12pt]{emulateapj}
3: %\usepackage{graphicx}
4: %\usepackage{caption2}
5: %\usepackage{amssymb}
6: %\usepackage{times}
7: %\usepackage{emulateapj}
8: %\usepackage{aasms4}
9: \usepackage{epsfig}
10:
11: %\topmargin=1cm
12:
13: \def\AaA{{\em Astr.~Astrophys.}}
14: \def\AJ{{\em Astr.~J.}}
15: \def\AN{{\em Astron.~Nachr.}}
16: \def\ApJ{{\em Astrophys.~J.}}
17: \def\ApJL{{\em Astrophys.~J.~Lett.}}
18: \def\ApJS{{\em Astrophys.~J.~Suppl.}}
19: \def\ARAA{{\em Ann.~Rev.~Astr.~Astrophys.}}
20: \def\MN{{\em Mon.~Not.~R.~astr.~Soc.}}
21: \def\Nat{{\em Nature}}
22: \def\PASJ{{\em Publ.~astr.~Soc.~Japan}}
23: \def\PASP{{\em Publ.~astr.~Soc.~Pacif.}}
24: \def\PhD{{\em PhD thesis}}
25: \def\Pre{Preprint}
26: \def\PS{{\em Phys.~Scr.}}
27: \def\RMP{{\em Rev.~Mod.~Phys.}}
28: \def\SSR{{\em Space Sci.~Rev.}}
29: \def\etal{{et al.\thinspace}}
30: \def\cf{{\em cf.\ }}
31: \def\eg{{\em e.g.\ }}
32: \def\etc{{\em etc.\ }}
33: \def\ie{{\em i.e.\ }}
34: \def\spose#1{\hbox to 0pt{#1\hss}}
35:
36: \let\approxlt=\lesssim
37: \let\approxgt=\gtrsim
38: \def\multcent#1{\halign {\centerline{##}\cr #1}}
39: \def\multleft#1{\hbox to size{\vbox {\halign {\lft{##}\cr #1}}\hfill}\par}
40: \def\multright#1{\hbox to size{\vbox {\halign {\rt{##}\cr #1}}\hfill}\par}
41: \def\Mdot{\hbox{$\dot M$}}
42: \def\degmark{^\circ}
43: \def\boxit#1{\vbox{\hrule\hbox{\vrule\kern3pt\vbox{\kern3pt
44: #1 \kern3pt}\kern3pt\vrule}\hrule}}
45: \font\big=cmr10 scaled\magstep2
46: \font\bigbf=cmbx10 scaled\magstep2
47: \font\bigit=cmti10 scaled\magstep2
48:
49: % Simple units
50:
51: \def\cm{{\rm\thinspace cm}}
52: \def\dyn{{\rm\thinspace dyn}}
53: \def\erg{{\rm\thinspace erg}}
54: \def\eV{{\rm\thinspace eV}}
55: \def\g{{\rm\thinspace g}}
56: \def\ga{{\rm\thinspace gauss}}
57: \def\K{{\rm\thinspace K}}
58: \def\keV{{\rm\thinspace keV}}
59: \def\km{{\rm\thinspace km}}
60: \def\kpc{{\rm\thinspace kpc}}
61: \def\Lsun{\hbox{$\rm\thinspace L_{\odot}$}}
62: \def\m{{\rm\thinspace m}}
63: \def\Mpc{{\rm\thinspace Mpc}}
64: \def\Msun{\hbox{$\rm\thinspace M_{\odot}$}}
65: \def\pc{{\rm\thinspace pc}}
66: \def\ph{{\rm\thinspace ph}}
67: \def\s{{\rm\thinspace s}}
68: \def\yr{{\rm\thinspace yr}}
69: \def\sr{{\rm\thinspace sr}}
70: \def\Hz{{\rm\thinspace Hz}}
71: \def\chisq{\hbox{$\chi^2$}}
72: \def\delchi{\hbox{$\Delta\chi$}}
73:
74: % Compound units
75:
76: \def\cmps{\hbox{$\cm\s^{-1}\,$}}
77: \def\cmsq{\hbox{$\cm^2\,$}}
78: \def\cmcu{\hbox{$\cm^3\,$}}
79: \def\pcmcu{\hbox{$\cm^{-3}\,$}}
80: \def\pcmcuK{\hbox{$\cm^{-3}\K\,$}}
81: \def\dynpcmsq{\hbox{$\dyn\cm^{-2}\,$}}
82: \def\ergcmcups{\hbox{$\erg\cm^3\ps\,$}}
83: \def\ergpcmps{\hbox{$\erg\cm^{-3}\s^{-1}\,$}}
84: \def\ergpcmsqps{\hbox{$\erg\cm^{-2}\s^{-1}\,$}}
85: \def\ergpcmsqpspA{\hbox{$\erg\cm^{-2}\s^{-1}$\AA$^{-1}\,$}}
86: \def\ergps{\hbox{$\erg\s^{-1}\,$}}
87: \def\gpcm{\hbox{$\g\cm^{-3}\,$}}
88: \def\gpcmps{\hbox{$\g\cm^{-3}\s^{-1}\,$}}
89: \def\gps{\hbox{$\g\s^{-1}\,$}}
90: \def\kmps{\hbox{$\km\s^{-1}\,$}}
91: \def\Lsunppc{\hbox{$\Lsun\pc^{-3}\,$}}
92: \def\Msunpc{\hbox{$\Msun\pc^{-3}\,$}}
93: \def\Msunpkpc{\hbox{$\Msun\kpc^{-1}\,$}}
94: \def\Msunppc{\hbox{$\Msun\pc^{-3}\,$}}
95: \def\Msunppcpyr{\hbox{$\Msun\pc^{-3}\yr^{-1}\,$}}
96: \def\Msunpyr{\hbox{$\Msun\yr^{-1}\,$}}
97: \def\pcm{\hbox{$\cm^{-3}\,$}}
98: \def\pcmsq{\hbox{$\cm^{-2}\,$}}
99: \def\pcmK{\hbox{$\cm^{-3}\K$}}
100: \def\phpcmsqps{\hbox{$\ph\cm^{-2}\s^{-1}\,$}}
101: \def\pHz{\hbox{$\Hz^{-1}\,$}}
102: \def\pmpc{\hbox{$\Mpc^{-1}\,$}}
103: \def\pmpccu{\hbox{$\Mpc^{-3}\,$}}
104: \def\ps{\hbox{$\s^{-1}\,$}}
105: \def\psqcm{\hbox{$\cm^{-2}\,$}}
106: \def\psr{\hbox{$\sr^{-1}\,$}}
107: \def\pyr{\hbox{$\yr^{-1}\,$}}
108: \def\kmpspMpc{\hbox{$\kmps\Mpc^{-1}$}}
109: \def\Msunpyrpkpc{\hbox{$\Msunpyr\kpc^{-1}$}}
110:
111: \makeatletter
112:
113: \makeatother
114:
115: \begin{document}
116:
117: \title{On the time variability of geometrically-thin black hole
118: accretion disks I : the search for modes in simulated disks.}
119:
120: \author{Christopher~S.~Reynolds\altaffilmark{1} and
121: M.~Coleman~Miller\altaffilmark{1}}
122:
123: \altaffiltext{1}{Department of Astronomy and the Maryland Astronomy
124: Center for Theory and Computation, University of Maryland, College
125: Park, MD 20742-2421}
126:
127: \begin{abstract}
128: We present a detailed temporal analysis of a set of hydrodynamic and
129: magnetohydrodynamic (MHD) simulations of geometrically-thin ($h/r\sim
130: 0.05$) black hole accretion disks. The black hole potential is
131: approximated by the Paczynski-Wiita pseudo-Newtonian potential. In
132: particular, we use our simulations to critically assess two widely
133: discussed models for high-frequency quasi-periodic oscillations,
134: global oscillation modes (diskoseismology) and parametric resonance
135: instabilities. We find that initially disturbed hydrodynamic disks
136: clearly display the trapped global g-mode oscillation predicted by
137: linear perturbation theory. In contrast, the sustained turbulence
138: produced in the simulated MHD disks by the magneto-rotational
139: instability does not excite these trapped g-modes. We cannot say at
140: present whether the MHD turbulence actively damps the hydrodynamic
141: g-mode. Our simulated MHD disks also fail to display any indications
142: of a parametric resonance instability between the vertical and radial
143: epicyclic frequencies. On the other hand, we do see characteristic
144: frequencies at any given radius in the disk corresponding to local
145: acoustic waves. We also conduct a blind search for {\it any}
146: quasi-periodic oscillation in a proxy lightcurve based on the
147: instantaneous mass accretion rate of the black hole, and place an
148: upper limit of 2\% on the total power in any such feature. We
149: highlight the importance of correcting for secular changes in the
150: simulated accretion disk when performing temporal analyses.
151: \end{abstract}
152:
153: \keywords{{accretion: accretion disks --- black hole physics ---
154: magnetic fields --- X-ray: binaries}}
155:
156: \section{Introduction}
157:
158: Rapid X-ray variability is a ubiquitous characteristic of accretion
159: onto black holes. Aperiodic X-ray fluctuations are seen from both
160: Galactic Black Hole Binaries (GBHBs) and Active Galactic Nuclei (AGN)
161: and, accounting for the inverse scaling of all relevant frequencies
162: with black hole mass, appear to have similar characteristics
163: (Uttley, McHardy \& Vaughan 2005; McHardy et al. 2006). While it is
164: highly tempting to relate this variability to the magnetohydrodynamic
165: (MHD) turbulence that is believed to drive the accretion process
166: (Balbus \& Hawley 1991, 1998), the exact physical
167: processes underlying the observed fluctuations remain mysterious.
168:
169: GBHBs also display quasi-periodic oscillations (QPOs) in their X-ray
170: lightcurves (see review by McClintock \& Remillard 2003)\footnote{Very
171: recently, the first convincing case of a QPO in an AGN was reported by
172: Gierlinski et al. (2008).}. Of particular interest are the
173: high-frequency quasi-periodic oscillations (HFQPOs) that are seen in
174: the very-high (or steep power law; McClintock \& Remillard 2003) state
175: of GBHBs. The HFQPOs have quality factors of few-to-$10$, centroid
176: frequencies of order $100\Hz$ and appear to be imprinted on the hard
177: X-ray tail of the spectrum rather than the thermal disk emission. The
178: fact that their frequencies are stable and at least loosely comparable
179: to the orbital frequency at the innermost stable circular orbit (ISCO)
180: around the black hole suggests that their properties are set by the
181: relativistic portions of the gravitational potential. This gives them
182: enormous promise as a diagnostic of black hole mass and spin.
183:
184: However, the utility of HFQPOs to relativistic astrophysics is
185: severely limited by the lack of a compelling theoretical framework in
186: which to interpret measurements of the frequencies, quality factors
187: and root-mean-square (rms) powers. There exist well defined geodesic
188: frequencies (i.e., the orbital, radial epicyclic and vertical
189: epicyclic frequencies) at any given radius in the accretion disk.
190: However, these frequencies (as well as all non-trivial linear
191: combinations) change with radius, and it is not clear why the
192: frequencies of any one particular radius would be preferentially
193: displayed in the overall power-spectrum.
194:
195: HFQPOs are commonly found in pairs with an approximate 3:2 frequency
196: ratio, and this has been used to suggest that a particular radius is
197: picked out due to a resonance. In the parametric resonance model
198: (Abramowicz \& Klu\'zniak 2001, 2003; Abramowicz et al. 2002, 2003),
199: there is a resonance in the disk at the radius where the radial
200: epicyclic frequency and vertical epicyclic frequency are in small
201: integer ratios. As discussed below, the strongest resonance occurs
202: when these frequencies are in 3:2 ratio, at least in the simplest
203: manifestation of this model. Other resonance models have been
204: examined by Rezzolla et al. (2003a,b) and Kato (2004a,b,c).
205:
206: Another interesting possibility is that the HFQPOs are global
207: oscillation modes of the accretion disk (i.e., ``diskoseismic''
208: modes). Global modes have been examined analytically (using linear
209: theory) on a hydrodynamic background in spacetimes that are
210: pseudo-Newtonian (Okazaki, Kato, \& Fukue 1987; Nowak \& Wagoner 1991,
211: 1992, 1993; Markovi\'c \& Lamb 1998), Schwarzschild (Kato \& Fukue
212: 1980), and Kerr (Kato 1990, 1991, 1993; Kato \& Honma 1991; Perez et
213: al. 1997; Silbergleit et al. 2001; Wagoner et al. 2001;
214: Ortega-Rodriguez et al. 2001). Three classes of mode are recovered
215: corresponding to pressure modes (p-modes), inertial modes
216: (conventionally referred to as g-modes even though the restoring force
217: results from rotation or inertia, depending on your frame of reference;
218: J.Pringle priv. communication) and warping/corrugation modes
219: (c-modes). For plausible masses and spins, the fundamental ($m=0$)
220: g-mode was quickly identified as a good candidate for the first HFQPO
221: discovered, the 67\,Hz oscillation found in the system GRS~1915$+$105
222: (Nowak et al. 1997). Current diskoseismology theory does not provide
223: a natural explanation of HFQPO pairs with small-integer ratios;
224: however, all present analyses are conducted using linear theory
225: whereas these HFQPOs pairs would likely arise from mode-coupling that
226: would only be revealed by a non-linear analysis.
227:
228: Clearly, many open questions concerning the physics of X-ray
229: variability remain, including the correct interpretation of HFQPOs.
230: The current dominant paradigm for understanding black hole accretion
231: is that the magnetorotational instability (MRI; Balbus \& Hawley 1991)
232: drives powerful MHD turbulence, and correlated Maxwell stresses within
233: this turbulence mediate the outward transport of angular momentum
234: that allows accretion to proceed. However, the connection between the
235: MHD turbulence paradigm and models for the aperiodic and
236: quasi-periodic variability remains highly uncertain. For example, can
237: the MHD turbulence naturally produce the rms-flux relation noted in
238: most black hole X-ray lightcurves (Uttley \& McHardy 2001) and/or the
239: log-normal flux distribution found in Cygnus X-1 (Uttley, McHardy \&
240: Vaughan 2005)? Are diskoseismic modes excited by turbulent
241: fluctuation (Nowak \& Wagoner 1993), or does the turbulence act to
242: damp such modes (Arras, Blaes \& Turner 2006)? Do the Maxwell
243: stresses couple radial and vertical motions in such a way as to excite
244: parametric resonance instabilities of the type identified by
245: Abramowicz \& Klu\'zniak (2001) or any other resonant phenomena? Does
246: the fact that the HFQPOs are imprinted on the high-energy tail provide
247: a fundamental clue to their origin, or is it a generic consequence of
248: any oscillating thermal accretion disk surrounded by a Comptonizing
249: corona (Lehr, Wagoner \& Wilms 2000)?
250:
251: In this paper, we use a set of global hydrodynamic and MHD simulations
252: of geometrically-thin accretion disks in a pseudo-Newtonian potential
253: to begin an exploration of these issues. Our canonical MHD simulation
254: represents a thinner disk, and is run for more orbits, than any
255: previously published well-resolved 3-d MHD disk simulation. This
256: allows us to conduct a more extensive study of the temporal
257: variability of such disks than has previously been attempted. In
258: \S~2, we give a brief review of the theory of both local and global
259: hydrodynamic modes of black hole accretion disks, as well as the
260: parametric resonance instability model for HFQPOs. \S~3 presents our
261: study of ideal (inviscid) hydrodynamic disks, both with imposed
262: axisymmetry and in full 3-dimensions. We find prominent trapped
263: g-modes in the axisymmetric simulations which remain (albeit with
264: diminished amplitude) in the full 3-d case. We then study the MHD
265: case in \S~4, where we find that the turbulence excites neither the
266: diskoseismic modes nor the parametric resonances discussed above.
267: Instead, we find that the turbulence excites local hydrodynamic waves
268: of the type elucidated by Lubow \& Pringle (1993). We discuss our
269: results, including a comparison to previous work, in \S~5 and conclude
270: in \S~6.
271:
272: \section{Theoretical Preliminaries}
273:
274: Here we provide a brief review of some previously established
275: theoretical results that are pertinent to this paper.
276:
277: \subsection{Local oscillations and waves in accretion disks}
278: \label{sec:local_osc}
279:
280: There is a very extensive literature on oscillations and waves in
281: accretion disks. Here we focus on just those aspects of the field
282: that turn out to be relevant for the interpretation of our simulation
283: which have been elucidated most clearly by Lubow \& Pringle (1993).
284:
285: Lubow \& Pringle (1993; hereafter LP93) studied three dimensional wave
286: propagation in accretion disks ignoring self-gravity. In the case
287: where one ignores vertical motions, they show that radial waves obey
288: the well-known dispersion relation
289: \begin{equation}
290: \omega^2=\kappa^2+c_s^2k^2,
291: \label{eq:disk_dispersion}
292: \end{equation}
293: where $c_s$ is the sound speed (assumed, in this case, to be purely a
294: function of $r$) and $\kappa$ is the radial epicyclic frequency given
295: by $\kappa^2=4\Omega^2+r\,\partial \Omega^2/\partial r$ (also see
296: Binney \& Tremaine 1987). Here, $\Omega(r)$ is the angular frequency
297: of the background Keplerian flow. LP93 proceed to study the
298: propagation of axisymmetric waves in the case where the atmosphere has
299: a locally isothermal vertical structure. They find two types of
300: waves. There are low-frequency gravity waves for which
301: \begin{equation}
302: 0<\omega<\Omega.
303: \end{equation}
304: There are also high-frequency acoustic waves which have
305: \begin{equation}
306: \omega^2>(n\gamma+1)\Omega^2,
307: \end{equation}
308: where $n=0,1,2,\ldots$ and $\gamma$ is the adiabatic index.
309: In the special case of purely vertical
310: perturbations, the inequality becomes an equality,
311: \begin{equation}
312: \omega^2=(n\gamma+1)\Omega^2.
313: \label{eq:vert_modes}
314: \end{equation}
315: The $n=0$ mode corresponds to a bulk vertical displacement of the disk
316: and subsequent vertical oscillation at the vertical epicyclic
317: frequency which, in the analysis of LP93 and in all analyses performed
318: in this paper, coincides with the orbital frequency. In general, the
319: $n-th$ mode has $n$ vertical nodes (i.e., locations where the vertical
320: velocity perturbation vanishes), and is either even or odd depending
321: on whether $n$ is even or odd, respectively.
322:
323: As discussed in \S~\ref{sec:mhd}, our simulations demonstrate the
324: effectiveness with which MHD turbulence excites these local acoustic
325: waves.
326:
327: \subsection{Global oscillation modes of an accretion disk}
328:
329: As discussed in the Introduction, several groups have studied the
330: global oscillation of black hole accretion disks using linear
331: perturbation theory, identifying three classes of normal mode
332: ($g$-modes, $p$-modes and $c$-modes). Trapped $g$-modes have received
333: particular attention as a possible source of HFQPO, although the other
334: families of modes may well be relevant. Here we review some of the
335: basic results of these analyses, following the approach of Nowak \&
336: Wagoner (1991, 1992; hereafter NW91 and NW92). The NW91 and NW92
337: analyses are not fully relativistic, instead employing a
338: pseudo-Newtonian potential. Thus, these analyses can be readily
339: compared with our pseudo-Newtonian simulations. We also note that
340: full general relativistic MHD simulations have typically yielded
341: results for slowly rotating black holes that are very similar to those
342: obtained with pseudo-Newtonian potentials (e.g., Gammie, McKinney, \&
343: T\'oth 2003; De~Villiers \& Hawley 2003).
344:
345: NW91 and NW92 use a Lagrangian formalism (Friedman \& Schutz 1978) and
346: a WKBJ approximation to derive the linearized equations describing
347: perturbations of an inviscid hydrodynamic thin accretion accretion
348: disk about a pure Keplerian background state. They also initially
349: examined the special case of purely radial oscillations and found the
350: standard dispersion relation of disk theory
351: (eqn.~\ref{eq:disk_dispersion}). Given that the gravity-modified
352: $p$-modes described by this dispersion relation become evanescent when
353: $\omega^2<\kappa^2$, global $p$-modes can be trapped between the ISCO
354: (where $\kappa=0$) and the radius at which $\omega=\kappa$. In
355: practice, however, the ``leaky'' nature of the ISCO would seem to make
356: the trapping of these modes ineffective.
357:
358: The more general case, including perturbations that have vertical as
359: well as radial motions, yields more promising results. NW92 examine
360: the linearized equations describing the behavior of the scalar potential
361: \begin{equation}
362: \delta u\equiv \delta P/\rho,
363: \end{equation}
364: where $\delta P$ is the Eulerian variation in the pressure. They show
365: that the linearized equations are approximately separable into radial
366: and vertical equations, with the separation constant being a slowly
367: varying function of $r$, $\Upsilon(r)$. The general dispersion
368: relation for these modes becomes
369: \begin{equation}
370: [\omega^2-\gamma \Upsilon(r)\Omega^2](\omega^2-\kappa^2)=\omega^2c_s^2k^2\; .
371: \end{equation}
372: Assuming that $\gamma \Upsilon\Omega^2>\kappa^2$, non-evanescent
373: solutions exist for $\omega^2>\gamma\Upsilon\Omega^2$ (predominantly
374: radial $p$-modes) or $\omega^2<\kappa^2$ (predominantly vertical
375: $g$-modes). Through this analysis, NW92 identify a class of global
376: $g$-modes that are trapped between two evanescent regions, $r<r_-$ and
377: $r>r_+$, where $\kappa(r_\pm)=\omega$. In other words, these modes
378: are trapped under the peak of the epicyclic frequency. They
379: principally focus on the $m=0$ (axisymmetric) modes, and show that the
380: mode frequency is only {\it slightly} smaller than the maximum radial
381: epicyclic frequency $\kappa_{\rm max}$. Radial harmonics of these
382: modes are very closely spaced. Thus, the inner radius at which the
383: mode becomes evanescent is still a finite distance (and, in plausible
384: settings, several vertical scale heights) from the ISCO. This raises
385: the interesting possibility of having appreciable power in such modes
386: without significant leakage across the ISCO.
387:
388: A major issue, however, is the effect of the turbulent MHD background
389: state on these modes. The diskoseismic mode frequencies are
390: comparable to the frequencies characterizing the expected MHD
391: turbulent fluctuations (which is very different to the situation in
392: the Sun, for example, where the observed helioseismic modes have
393: frequencies that are four orders of magnitude higher than the
394: turbulent turnover frequency). Thus, an MHD turbulent disk is likely
395: to be a hostile environment for any diskoseismic modes. Furthermore,
396: magnetic forces can lead to a rather gradual transition in flow
397: properties around the ISCO, potentially worsening the leakage of the
398: trapped g-modes. Considering these mode destruction/suppression
399: mechanisms, it is reasonable to suppose that mode survival becomes
400: easier in thinner disks since (1) the ratio of the typical turbulent
401: cell size to the radial extent of the resonant cavity will decrease
402: with disk thickness and (2) there are suggestions that the transition
403: in flow properties around the ISCO is sharper in thinner disks
404: (Reynolds \& Fabian 2008; Shafee et al. 2008) thereby producing less
405: mode leakage. This raises the possibility that a $g$-mode can be
406: sustained against (or even fed by; Nowak \& Wagoner 1993) the
407: turbulence in a sufficiently thin disk.
408:
409: Previously published global MHD disk simulations (e.g., Hawley \&
410: Krolik 2001) have modeled flows as thin as $h/r\sim 0.1$ and have not
411: reported diskoseismic modes. However, these authors did not conduct a
412: directed search for such modes and hence it is not possible to say
413: that the modes were really not present. Careful examination of local
414: ``shearing-box'' simulations have indeed failed to find trapped
415: $g$-modes associated with MHD turbulence (Arras, Blaes \& Turner
416: 2006), but this issue has yet to be examined in a global thin-disk
417: setting. Searching for and characterizing these trapped $g$-modes in
418: global thin-disk simulations will be a major theme of our paper.
419:
420:
421: \subsection{Parametric resonance}
422:
423: From the point of view of explaining HFQPOs, the need for global
424: oscillation modes is diminished if some process does indeed select
425: special radii in the accretion disk. As discussed in the
426: Introduction, the discovery of pairs of HFQPOs with small integer
427: ratios has prompted several groups to examine resonance
428: models. In particular, we shall briefly review the parametric
429: resonance model of Abramowicz \& Klu\'zniak (Abramowicz \& Klu\'zniak
430: 2001, 2003; Abramowicz et al. 2002, 2003).
431:
432: We begin by considering an accretion disk in which the flow deviates
433: only slightly from Keplerian so that the position of a fluid element
434: (in spherical polar coordinates) is given by
435: \begin{equation}
436: r(t)=r_0+\delta r(t), \ \theta(t)=\frac{1}{2}\pi+\delta\theta(t),\
437: \phi(t)=\Omega t.
438: \end{equation}
439: We have specialized to the case of axisymmetric perturbations. By
440: expanding the resulting equations of motion to third order (and
441: wrapping up the non-gravitational forces into two unspecified force
442: functions), one finds a Mathieu-type equation of motion
443: \begin{equation}
444: \delta\theta_{,\,tt}+\Omega^2[1+a\cos(\kappa t)]\delta\theta +
445: \lambda\,\delta\theta_{,\,t}=0,
446: \label{eq:mathieu}
447: \end{equation}
448: where $a$ is a small constant that describes the coupling between the
449: vertical and radial perturbations, and $\lambda$ is a (small) damping
450: constant. Here, we have specialized the equations of Abramowicz et
451: al. (2003; hereafter A2003) to the case where the vertical epicyclic
452: frequency is the same as the orbital frequency. This is valid for a
453: non-spinning black hole and, in particular, the pseudo-Newtonian
454: potential that we use for the simulations in this paper.
455:
456: One expects a system described by eqn.~\ref{eq:mathieu} to exhibit a
457: parametric resonance instability when $\kappa=2\Omega/n$, where $n$ is
458: a non-zero positive integer. Given that black hole potentials always
459: have $\kappa<\Omega$, the smallest value of $n$ for which the
460: resonance condition is obeyed is $n=3$, i.e., $\kappa=2\Omega/3$.
461: This is expected to be the strongest of the set of resonances. It is
462: interesting that the fundamental ``test-particle'' frequencies at this
463: resonant radius have 3:2 frequency ratio in agreement with
464: observations of HFQPO pairs. For the Paczynski-Wiita pseudo-Newtonian
465: potential we use in our simulations (Paczynski \& Wiita 1980, hereafter PW),
466: \begin{equation}
467: \Phi=-\frac{GM}{r-2r_g},\hspace{1cm}r_g\equiv \frac{GM}{c^2},
468: \label{eq:pw_pot}
469: \end{equation}
470: we have
471: \begin{equation}
472: \Omega=\frac{1}{r-2r_g}\sqrt{\frac{GM}{r}}
473: \end{equation}
474: and,
475: \begin{equation}
476: \kappa=\sqrt{\frac{GM(r-6r_g)}{r(r-2r_g)^3}}.
477: \end{equation}
478: Hence, the 3:2 resonance occurs at $r=9.2r_g$. It must be noted,
479: however, that a full integration of a toy-problem by A2003 reveals
480: that higher-order effects shift the location of the resonance, making
481: the ratio of the epicyclic frequencies extremely sensitive to the
482: strength of the coupling between the $r$ and $\theta$ perturbations.
483: While A2003 suggest that this sensitivity is a strength of the model,
484: allowing application to HFQPO pairs in neutron star systems which do
485: not have simple integer ratios, it inevitably diminishes the power of
486: this model to explain black hole systems. In addition, there is
487: currently no physical model of the coupling between the radial
488: and vertical motions. The simulations described in this paper allow
489: us to assess whether magnetic forces couple these motions in such a
490: way as to drive a parametric resonance instability.
491:
492: \section{Hydrodynamic disks}
493:
494:
495: \subsection{Initial comments}
496: \label{sec:initial_hydro}
497:
498: In the remainder of this paper, we construct and analyze numerical
499: simulations of thin disks in order to examine their variability
500: properties, focusing on the presence of local and global modes, as
501: well as parametric resonances. Of course, real accretion disks are
502: believed to require at least an MHD-level description in order to
503: capture the MRI-driven turbulence that transports angular momentum and
504: drives accretion. MHD simulations are addressed in the next section.
505: However, it is useful to begin with a discussion of ideal hydrodynamic
506: models in order to help isolate the various physical effects present
507: in these complex systems. That will be the focus of this section.
508:
509: In order to allow us to perform a set of simulations with
510: modest-to-high resolution, we begin with 2-d (axisymmetric)
511: hydrodynamic simulations. We expect (and indeed show) that these
512: axisymmetric models are well suited for studying the fundamental $m=0$
513: g-mode. However, once we move to MHD, Cowling's anti-dynamo theorem
514: (Cowling 1957) leads to qualitatively different behavior in
515: axisymmetric compared with full 3-d simulations. Hence, all of the
516: MHD models that we shall describe are performed in 3-dimensions.
517: Since we will be directly comparing results (e.g., $g$-mode
518: amplitudes) between the hydrodynamic and MHD simulations, we also need
519: to perform a ``bridging'' 3-d hydrodynamic simulation.
520:
521: \subsection{Basic simulation set-up}
522: \label{sec:hydro_setup}
523:
524: To make the problem tractable despite the severe resolution
525: requirements imposed by the geometrical thinness of the accretion
526: disk, we choose to focus on only the most essential aspects of the
527: physics. From the discussion in \S~2, it is clear that the essential
528: aspect of the relativistic potential that must be captured is the
529: nature of the radial epicyclic frequency (e.g., that it goes to zero
530: at some finite radius and hence produces an ISCO at that radius). In
531: this sense, the PW potential (eqn.~\ref{eq:pw_pot}) is a good
532: approximation for the gravitational field of a non-rotating black
533: hole; its ISCO (at $6r_g$) and marginally bound orbit (at $4r_g$) are
534: both at the same radius as in the Schwarzschild geometry. We also
535: simplify the simulations by neglecting all radiation processes
536: (radiative heating, radiative cooling, radiative transfer, and the
537: dynamical effects of radiation pressure). In place of a full energy
538: equation, the gas is given an adiabatic equation of state with
539: $\gamma=5/3$.
540:
541: \begin{figure*}
542: \hbox{
543: \hspace{-1cm}
544: \psfig{figure=f1a.ps,width=0.55\textwidth}
545: \hspace{-1cm}
546: \psfig{figure=f1b.ps,width=0.55\textwidth}
547: }
548: \hbox{
549: \hspace{-1cm}
550: \psfig{figure=f1c.ps,width=0.55\textwidth}
551: \hspace{-1cm}
552: \psfig{figure=f1d.ps,width=0.55\textwidth}
553: }
554: \caption{Snapshots of evolution of the canonical axisymmetric
555: hydrodynamic simulation (HD2d\_1) at $t=0.5T_{\rm isco}$ (top-left),
556: $t=1.0T_{\rm isco}$ (top-right), $t=5T_{\rm isco}$ (bottom-left) and
557: $t=50T_{\rm isco}$ (bottom-right), where $T_{\rm isco}$ is the orbital
558: period at the ISCO. Both the color table and contours show the
559: logarithmic density structure of the disk cross-section, with 10
560: contours per decade of density. A range of densities spanning three
561: orders of magnitude are shown. The curved line to the left of each
562: frame represents the event horizon.}
563: \label{fig:hydro_evol}
564: \end{figure*}
565:
566: \begin{table*}[t]
567: \begin{center}
568: {\small
569: \begin{tabular}{cccccccc}\hline
570: Run & dim & $h_2/r_{\rm isco}$ & r-domain ($r_g$)& z-domain ($r_g$) & $\phi$-domain & $n_r\times n_z (\times n_\phi)$ & $T_{\rm stop}$ ($r_g/c$) \\
571: (1) & (2) & (3) & (4) & (5) & (6) & (7) & (8) \\\hline
572: HD2d\_1 & 2 & 0.05$^*$ & (4,16) & (-1.5,1.5) & -- & $256\times 128$ & 12320 \\
573: HD2d\_1hr & 2 & 0.05$^*$ & (4,28) & (-1.5,1.5) & -- & $1024\times 256$ & 12320 \\
574: HD2d\_2 & 2 & 0.025$^*$ & (4,16) & (-0.75,0.75) & -- & $512\times 128$ & 12320 \\
575: HD2d\_3 & 2 & 0.1$^*$ & (4,16) & (-3,3) & -- & $256\times 128$ & 12320 \\
576: HD3d\_1 & 3 & 0.05$^*$ & (4,16) & (-1.5,1.5) & $(0,\pi/6)$ & $240\times 128\times 32$ & 12320 \\\hline
577: MHD\_1 & 3 & 0.05 & (4,16) & (-3,3) & $(0,\phi/6)$ & $240\times 256\times 32$ & 38800 \\
578: MHD\_2 & 3 & 0.05 & (4,16) & (-1.5,1.5) & $(0,\phi/6)$ & $240\times 128\times 32$ & 38800 \\
579: MHD\_2hr & 3 & 0.05 & (4,16) & (-1.5,1.5) & $(0,\phi/3)$ & $480\times 256\times 64$ & 5236 \\
580: MHD\_3 & 3 & 0.05 & (4,28) & (-1.5,1.5) & $(0,\phi/6)$ & $960\times 128\times 32$ & 11400 \\\hline
581: \end{tabular}
582: \caption{Summary of the hydrodynamic (HD) and MHD simulations
583: discussed in this paper. Column (1) gives the designation of the
584: simulation. Column (2) lists the dimensionality of the simulation.
585: Column (3) gives the fractional disk thickness at the ISCO; the
586: asterisk ($*$) denotes that the simulation was started with an initial
587: vertical structure that is slightly out of equilibrium in order to
588: seed subsequent oscillation modes (as described in the text). Columns
589: (4), (5) and (6) list the $r$-, $z$- and $\phi$- domains of the
590: simulation box. Column (7) gives the number of computational zones
591: within the domain. Column (8) lists the run time of the simulated
592: disk.}}
593: \end{center}
594: \end{table*}
595:
596:
597: All simulations are performed in cylindrical polar coordinates
598: $(r,z,\phi)$. The initial condition consists of a disk with a
599: constant mid-plane density ($\rho_0=1$) for $r>r_{\rm isco}$. The
600: initial density scale-height of the disk is assumed to be constant
601: with radius, implying a sound speed which falls off with radius as
602: approximately $r^{-3/2}$. There are two motivations for choosing to
603: model a ``constant-h'' disk; (1) such a choice is well suited to the
604: cylindrical symmetry of our coordinate grid and (2) according to the
605: standard model of Shakura \& Sunyaev (1973), the radiation-pressure
606: dominated disks of real accreting black holes are likely to maintain
607: an approximately constant scale-height in their innermost regions.
608: In more detail, the vertical structure of the disk is given by
609: \begin{eqnarray}
610: \rho(r,z)&=&\rho_0\,\exp\left(-\frac{z^2}{2h_1^2}\right),\\
611: p(r,z)&=&\frac{GMh_2^2}{(R-2r_g)^2R}\,\rho(r,z),
612: \end{eqnarray}
613: where $r$ is the cylindrical radius, $z$ is the vertical height above
614: the disk midplane and $R=\sqrt{r^2+z^2}$. This corresponds to an
615: isothermal atmosphere which, when $h_1=h_2$, is in vertical
616: hydrostatic equilibrium in the PW potential. In order to give the
617: disk an initial vertical kick, we set $h_1=1.2h_2$ (the values of
618: $h_2$ for all of our runs are detailed in Table~1). Thus, the initial
619: disk temperature is $\sim 20\%$ too cold for the density and pressure
620: run, leading to a vertical collapse and bounce of the disk. The
621: initial density is set to zero for $r<r_{\rm isco}$. The initial
622: velocity field is everywhere set to
623: \begin{equation}
624: v_\phi=r\Omega=\frac{\sqrt{GMr}}{r-2r_g},\hspace{1cm}v_r=v_z=0,
625: \end{equation}
626: corresponding to rotation on cylinders and pure Keplerian motion for
627: material on the mid-plane. We impose zero-gradient outflow boundary
628: conditions on both the radial and vertical boundaries of the
629: simulation, i.e., the fluid quantities in the ghost zones are set to
630: the values of the neighboring active zone, and a ``diode'' condition
631: is imposed on the component of the velocity perpendicular to the
632: boundary which allows outflow but disallows inflow.
633:
634:
635:
636: Table~1 details our hydrodynamic simulations. We perform four
637: simulations in which strict axisymmetry is imposed at all times
638: ($\partial/\partial \phi=0$). These axisymmetric simulations were
639: performed using the serial ZEUS-3D MHD code (Stone \& Norman 1992a,b)
640: in its pure hydrodynamic axisymmetric mode. In our canonical
641: axisymmetric run (HD2d\_1), we set $h_2=0.3r_g$ (corresponding to
642: $h_2/r=0.05$ at the ISCO). To model the dynamics of the disk in a
643: robust manner, our vertical domain must cover many scale-heights; in
644: our canonical run, the vertical domain is $z\in (-5h_2,+5h_2)$. We
645: place 128 uniformly spaced grid cells in this vertical direction,
646: giving 13 cells per scale-height. This allows us to resolve
647: hydrodynamic waves with wavelengths of $\sim 0.5h_2$ or greater. The
648: vertical resolution requirements force us to consider a limited radial
649: domain. In the canonical simulation, the radial domain extends from
650: $4r_g$ (i.e., well within the plunge region) to $16r_g$ and should
651: correspond to the range of radii where trapped diskoseismic modes or
652: A2003-type resonances occur. Tolerating a grid-cell aspect ratio of
653: $\sim 2$, we place 256 uniformly spaced radial cells in this radial
654: domain. The simulation was evolved for a time $200T_{\rm isco}$ where
655: \begin{equation}
656: T_{\rm isco}\approx 61.6\,GM/c^3,
657: \end{equation}
658: is the orbital period at the ISCO.
659:
660: In order to test the robustness of the results discussed below to
661: resolution and the limited radial domain, we performed a second
662: simulation (HD2d\_1hr) in which the spatial resolution was doubled
663: (i.e., the size of each voxel was halved in both the radial and
664: vertical dimensions) and the radial domain extended out to $28r_g$.
665: We also performed two additional axisymmetric simulations employing
666: the same set-up as the canonical simulation but with disks that are
667: half (HD2d\_2) and twice (HD2d\_3) the thickness (including
668: appropriate modifications to the vertical domain and resolution; see
669: Table~1).
670:
671: As discussed above, we perform a 3-dimensional hydrodynamic simulation
672: in order to aid the later interpretation of the 3-d MHD simulations.
673: The set-up of our 3-d run (HD3d\_1) is identical to that for the
674: canonical 2-d run except that the computational domain has a
675: $\phi$-dimension. To reduce computational expense while capturing the
676: essential physics, we simulate only a $\Delta\phi=30^\circ$ wedge of
677: the disk using 32 uniformly spaced grid cells, imposing periodic
678: boundary conditions on the $\phi$-boundaries. This 3-dimensional
679: simulation was performed using an MPI-parallelized version of ZEUS
680: kindly supplied to us by Eve Ostriker (and similar to the ZEUS-MP code
681: of Vernaleo \& Reynolds 2006).
682:
683: \subsection{Axisymmetric hydrodynamic models and the recovery of trapped
684: $g$-modes}
685: \label{sec:axisym}
686:
687:
688: We now discuss the evolution of the axisymmetric hydrodynamic
689: simulations, beginning with the canonical simulation. Starting from
690: the initial condition, the disk undergoes dynamical timescale
691: variability because of pressure gradients which push matter inside of
692: the ISCO. The strong transients close to the ISCO launch
693: outward-radially directed waves into the disk which break rapidly to
694: become rolls. This behavior can be seen in Fig.~\ref{fig:hydro_evol}.
695: These high amplitude transients are short lived, however, lasting only
696: $\sim 10T_{\rm isco}$. At subsequent times, the disk settles into a
697: stationary state apart from small amplitude (and decaying) internal
698: oscillations and a very weak accretion stream driven by the numerical
699: viscosity.
700:
701: \begin{figure}[t]
702: \centerline{
703: \psfig{figure=f2.ps,width=0.55\textwidth}
704: }
705: \caption{Change of the quantity $K=\int_{\cal D} \rho v_z^2\,dV$ with
706: time. The integration domain ${\cal D}$ is the annulus $r\in (7,14)$. The
707: bottom (black) line is for the canonical-2d run (HD2d\_1),
708: lower-middle (red) line is for the high-resolution 2-d run
709: (HD2d\_1hr), the upper-middle (blue) line is for the canonical 3-d run
710: (HD3d\_1), and the top (green) line is for the canonical MHD run
711: (MHD\_1). The distinct ``ringing'' seen in runs HD2d\_1 and HD2d\_1hr
712: corresponds to the axisymmetric g-mode (see Section~\ref{sec:axisym}).
713: In run HD3d\_1, non-axisymmetric aperiodic structures mask the
714: underlying g-mode (see Section~\ref{sec:3dhydro}). The fact that
715: $K(t)$ for the run MHD\_1 lies an order of magnitude above that for
716: HD3d\_1 demonstrates that the MRI-driven turbulence completely
717: overwhelms the hydrodynamic disturbances seen in the analogous 3-d
718: hydrodynamic simulation (see Section~\ref{sec:basicmhdevol}).}
719: \label{fig:hydro_decay}
720: \end{figure}
721:
722:
723: In order to study the decay of the hydrodynamic fluctuations in a more
724: quantitative manner, we start by computing the quantity,
725: \begin{equation}
726: K=\int_{\cal D} \rho v_z^2\,dV.
727: \label{eq:decay}
728: \end{equation}
729: This quantity is particularly well suited to diagnose vertical
730: oscillations of the disk, and will vanish once the hydrodynamic
731: configuration has established a stationary state. In order to
732: diagnose the state of the body of the disk (i.e., side-stepping issues
733: of the outer radial boundary or the plunge region), we choose to
734: compute this integral over a restricted domain ${\cal D}$ consisting
735: of the annulus $r\in (7r_g,14r_g)$. The time-dependence of
736: $K/\max(K)$ for the canonical axisymmetric simulation (HD2d\_1) and its
737: high-resolution counterpart (HD2d\_1hr) is shown in
738: Fig.~\ref{fig:hydro_decay}. The rapid initial decline of $K(t)$ is
739: very similar for these two simulations and corresponds to the strong
740: transients described above. At long times (after about $t\sim 4\times
741: 10^3\,GM/c^3\approx 70\,T_{\rm isco}$) the behavior of these
742: simulations begins to deviate. HD2d\_1 continues to decay in an
743: approximately exponential manner $K(t)\propto e^{-t/\tau_0}$, where
744: $\tau_0\approx 4\times 10^3\,GM/c^3$. Superposed on this decay is a
745: distinct oscillation. This corresponds to (twice) the frequency of
746: the trapped $g$-mode that we shall discuss below. The high-resolution
747: version of this simulation HD2d\_1hr shows very similar behavior
748: except that the exponential decay time is longer, $\tau_0\approx
749: 6\times 10^3\,GM/c^3$. This suggests that the decay of these small
750: perturbations is due to numerical dissipation which scales
751: approximately as the square root of the size of the simulation grid
752: cells.
753:
754:
755: We now study the spatio-temporal variability of the disk and, in
756: particular, seek the diskoseismic modes predicted by linear theory.
757: Figure~\ref{fig:vr_rt} shows the mid-plane value of $v_r$ on the
758: $(r,t)$-plane, i.e., the value of the function $v_r(r,z=0;t)$, for run
759: HD2d\_1. Note that, outside of the ISCO, the average value is
760: $\langle v_r\rangle\ll 0.001c$ so that this Figure essentially plots the
761: fluctuation of $v_r$ from its mean value. At early times, we see
762: strong wave-like disturbances which are generated in the inner parts
763: of the disk (at $r\approx 8r_g$) and propagate to both large and small
764: radii. Although the outer radial boundary condition is zero-gradient
765: outflow, impedance mismatching results in some reflection of these
766: initial strong waves. After these initial transients have died out,
767: it can be seen that the highest amplitude fluctuations are limited to
768: a rather narrow range of radii in the approximate range $r=7-9r_g$.
769: The fact that these perturbations are essentially vertical on the
770: $(t,r)$-plane indicates that they are coherent across this radial
771: range, as would be expected if we are seeing a global mode.
772:
773: \begin{figure}[b]
774: \centerline{
775: \psfig{figure=f3.ps,width=0.55\textwidth}
776: }
777: \caption{Radial component of velocity $v_r$ on the midplane of the
778: disk as a function of radius and time for the the canonical
779: axisymmetric hydrodynamic simulation (HD2d\_1). The linear color
780: table extends from radial velocities of $v_r=-0.001c$ (black) to
781: $v_r=+0.001c$ (white). Note that, outside of the ISCO, the average
782: value is $|v_r|\ll 0.001c$ so that this diagram essentially plots the
783: fluctuation of $v_r$ from its mean value.}
784: \label{fig:vr_rt}
785: \end{figure}
786:
787: % FIG 4 HERE
788: \begin{figure*}
789: \hbox{
790: \psfig{figure=f4a.ps,width=0.55\textwidth}
791: \psfig{figure=f4b.ps,width=0.55\textwidth}
792: }
793: \caption{Midplane PSD for pressure fluctuations in run HD2d\_1 (left
794: panel) and its high-resolution counterpart HD2d\_1hr (right panel).
795: Note that we show the radial range $r\in (4,16)$ which is coincident
796: with the full computational domain of HD2d\_1 but only the inner half
797: of the domain for HD2d\_1hr (whose full radial domain extends from
798: $4r_g$ to $28r_g$). Also shown are the radial epicyclic frequency
799: (solid white line), orbital frequency (dashed) and the $n=1,2,3$ pure
800: vertical p-modes (from left to right dot-dashed lines). The absolute
801: scaling of the PSD, as indicated by the color-bar, is arbitrary.}
802: \label{fig:hd_psd}
803: \vspace{0.5cm}
804: \end{figure*}
805:
806: This temporal variability can be explored in more detail using the
807: power spectral density (PSD), defined as $P(\nu)=\alpha |\tilde
808: f(\nu)|^2$, where $\alpha$ is some normalization constant and $\tilde
809: f(\nu)$ is the Fourier transform of the time-sequence $f(t)$ under
810: consideration,
811: \begin{equation}
812: \tilde f(\nu)=\int f(t)e^{-2\pi i\nu t}\,dt.
813: \end{equation}
814: Note that, for ease of interpretation, most of the PSDs presented in
815: this paper will be in terms of the usual frequency, $\nu$, rather than
816: angular frequency, $\omega$. Figure~\ref{fig:hd_psd} shows the PSD of
817: the mid-plane pressure as a function of $r$ and frequency for HD2d\_1
818: and HD2d\_1hr. This is computed using approximately the final half
819: ($\Delta t=102.4T_{\rm isco}\approx 6308\,GM/c^3$) of the simulation
820: in order to avoid the initial strong transients. These PSDs differ in
821: the range $r=12-16r_g$; run HD2d\_1hr has significantly less low
822: frequency noise than run HD2d\_1 in this radial range. We attribute
823: this to effects related to the outer radial boundary which is at
824: $r=16r_g$ in run HD2d\_1 but substantially further out ($r=28r_g$) in
825: run HD2d\_1hr.
826:
827:
828: \begin{figure*}
829: \hbox{
830: \hspace{-1cm}
831: \psfig{figure=f5a.ps,width=0.55\textwidth}
832: \hspace{-1cm}
833: \psfig{figure=f5b.ps,width=0.55\textwidth}
834: }
835: \caption{{\it Left panel} : Midplane pressure PSD for run HD2d\_1
836: summed up over a range of radii $\Delta r=0.5r_g$ centered on
837: $r=8r_g$. The dashed line shows the maximum radial epicyclic frequency.
838: {\it Right panel }: Integral of the midplane pressure PSD of those
839: frequency bins that exceed $5\times 10^{-14}$ in the left panel, as a
840: function of radius. Vertical dashed lines show the range of radii for
841: which the mean frequency of this peak is less than the radial
842: epicyclic frequency.}
843: \vspace{0.5cm}
844: \end{figure*}
845:
846:
847: \begin{figure*}
848: \hbox{
849: \hspace{-1cm}
850: \psfig{figure=f6a.ps,width=0.55\textwidth}
851: \psfig{figure=f6b.ps,width=0.55\textwidth}
852: }
853: \hbox{
854: \hspace{-1cm}
855: \psfig{figure=f6c.ps,width=0.55\textwidth}
856: \psfig{figure=f6d.ps,width=0.55\textwidth}
857: }
858: \caption{Maps of period-folded pressure deviation (i.e., the
859: difference between the instantaneous pressure and the time-averaged
860: pressure) for run HD2d\_1hr. Only the last 102.4 orbits of data have
861: been included in order to avoid the transient behavior associated with
862: the initial conditions. The folding period corresponds to the peak of
863: the PSD see in Fig.~\ref{fig:gmode}. Phases of 0.0 (top-left), 0.25
864: (top-right), 0.5 (bottom-left) and 0.75 (bottom-right) are shown.
865: This, in essence, gives us a direct view of the eigenmode.}
866: \label{fig:folding}
867: \end{figure*}
868:
869:
870: Inside of $r=12r_g$, however, the PSDs of these two simulations are
871: very similar and we can trust that neither the resolution nor the
872: outer radial boundary affect the results significantly. We note that
873: simulations in which the initial radial density profile of the disk is
874: truncated before reaching the outer boundary also produce essentially
875: identical results, again giving us confidence that noise infiltration
876: from the outer boundary is not an important issue. In addition to
877: very low frequency noise, the most prominent feature of the inner-disk
878: PSD is a vertical ridge of enhanced power at $\nu\approx (4-5)\times
879: 10^{-3}\,c^3/GM$ extending from $r\approx 6.5r_g$ out to $r\approx
880: 9.5r_g$. The one dimensional cuts through this ridge in
881: Fig.~\ref{fig:gmode} demonstrate it to have all of the expected
882: properties of a trapped $g$-mode. Firstly, it exists in a rather
883: narrow range of frequencies, $\nu\approx (4-5)\times 10^{-3}\,c^3/GM$,
884: just below the maximum radial epicyclic frequency ($\kappa_{\rm
885: max}\approx 5.52\times 10^{-3}\,c^3/GM$). Secondly, it is spatially
886: bounded by the radii at which the mode frequency becomes equal to the
887: radial epicyclic frequency. There is, however, some leaking of the
888: mode down to the ISCO.
889:
890: We can visualize the eigenmode by producing maps of pressure
891: deviations that have been ``period-folded'' on the period
892: corresponding to the peak power in this mode. More precisely, we use
893: the last half of the simulation to produce maps of the difference
894: between the instantaneous pressure and the time-averaged pressure with
895: a sampling rate of $0.2T_{\rm isco}$. We then sort these maps into 16
896: phase bins (based on the period corresponding to the peak power in
897: this mode) and average together all maps within a given phase bin.
898: The result is shown in Fig.~\ref{fig:folding}. In addition to the
899: g-mode itself, these maps reveal a striking ``chevron'' pattern at
900: large radii. Time-sequences of maps reveal these features to be
901: outward-radially traveling acoustic waves driven by the global
902: g-mode, refracted into the upper layers of the disk atmosphere as they
903: propagate.
904:
905: % FIG 7 HERE
906: \begin{figure}[t]
907: \centerline{
908: \psfig{figure=f7.ps,width=0.55\textwidth}
909: }
910: \caption{Frequency of the g-modes as a function of mid-plane sound
911: speed at $r=8r_g$ for runs HD2d\_2, HD2d\_1, and HD2d\_3 (from left to
912: right). The vertical bars indicate the range of frequencies over
913: which enhanced power is seen.}
914: \label{fig:f_vs_cs}
915: \vspace{0.5cm}
916: \end{figure}
917:
918:
919: As a final demonstration that we have properly identified trapped
920: $g$-modes in our axisymmetric hydrodynamic simulations, we use our set
921: of runs to study the dependence of the mode frequency on the sound
922: speed in the disk. Figure~\ref{fig:f_vs_cs} shows the dependence of
923: the mode frequency on midplane sound speed of the disk (and hence disk
924: thickness) at $r=8r_g$. As expected from analytic theory (see
925: Appendix~A), the difference between the mode frequency and the maximum
926: epicyclic frequency depends linearly on the sound speed. There is,
927: however, a discrepancy in the slope of this linear relationship obtained
928: by the simulations and expected from the analytic theory. Note that
929: the analytic treatment assumes that the gas is strictly isothermal
930: across the whole region of interest, a condition that is clearly
931: violated in the simulated disk. Given the sensitivity of the mode to
932: the radial structure of the disk (as demonstrated by the factor of 2.5
933: difference in the analytic results between the two pseudo-Newtonian
934: potentials examined in Appendix~A), the non-isothermality of the gas
935: in the simulated disk can readily shift the mode frequency away from
936: the analytic value.
937:
938: \subsection{Extension to 3-dimensional hydrodynamic models}
939: \label{sec:3dhydro}
940:
941: We begin our analysis of the 3-dimensional hydrodynamic simulations
942: (HD3d\_1) by examining the decay of hydrodynamic perturbations in our
943: canonical 3-d run (HD3d\_1) using the quantity $K(t)$ as defined in
944: eqn.~\ref{eq:decay}. The $K(t)$ behavior for HD3d\_1 is somewhat
945: different from the axisymmetric simulations at late times, reaching a
946: quasi-steady state at $K/\max(K)\sim 3\times 10^{-3}$ with aperiodic
947: fluctuations rather than continuing a ringing exponential decay
948: (Fig.~\ref{fig:hydro_decay}). Numerical dissipation must be just as
949: effective in HD3d\_1 as compared with HD2d\_1, hence these
950: perturbations must be driven by some instability. While it is thought
951: that free Keplerian accretion disks are stable to linear hydrodynamic
952: perturbations (see Balbus \& Hawley 1998; Hantao et al. 2006), the
953: presence of the simulation boundaries can introduce true instabilities
954: (associated with reflection from the boundaries) as well as
955: uncontrolled numerical noise, and these might explain these low level
956: sustained fluctuations. Visual inspection of density slices in the
957: $(r,z)$ and $(r,\phi)$ planes also suggests that non-axisymmetric
958: wave-like perturbations interacting with the boundaries are
959: responsible for perturbing the 3-d hydrodynamic disk. However, a
960: detailed study of the sustained fluctuations in the 3-d hydrodynamic
961: disks is beyond the scope of this paper. Since these perturbations
962: exist at a low level (particularly compared with the MHD turbulence
963: discussed in \S~4; see Fig.~\ref{fig:hydro_decay} for a comparison of
964: the sustained fluctuations in the 3-d hydrodynamic run compared with
965: the canonical MHD run) and are likely to be driven by our boundaries
966: (hence, are not of astrophysical relevance), they are not of
967: importance for the principal focus of our study.
968:
969: % FIG 8 HERE
970: \begin{figure}[b]
971: \centerline{
972: \psfig{figure=f8.ps,width=0.55\textwidth}
973: }
974: \caption{
975: Midplane PSD of azimuthally-averaged pressure for run HD3d\_1 summed
976: up over a range of radii $\Delta r=0.5r_g$ centered on $r=8r_g$. The
977: dashed line shows the maximum epicyclic frequency.
978: }
979: \label{fig:hd3d_psd}
980: \end{figure}
981:
982:
983: The principal issue to be addressed here is the impact of the
984: transition to 3-dimensions on the presence of the trapped $m=0$
985: g-modes in the simulated disks. We might expect the power in these
986: axisymmetric modes to be reduced in the 3-d case as the free energy is
987: shared with non-axisymmetric modes. This is indeed the
988: case. Figure~\ref{fig:hd3d_psd} shows the PSD at $r=8r_g$ of the last
989: $\Delta t=102.4T_{\rm isco}$ of the canonical 3-d run HD3d\_1. A
990: region of enhanced power is clearly seen in the correct range of
991: frequencies $\nu\approx (4-5)\times 10^{-3}\,c^3/GM$ to be identified
992: with the axisymmetric g-modes studied in \S~\ref{sec:axisym}.
993: Furthermore, examination of the radially-resolved PSD shows that this
994: region of enhanced power is bounded by the radial epicyclic frequency
995: in precisely the manner expected for trapped g-modes. A comparison of
996: the absolute values of the PSD across the mode does reveal, however,
997: that the mode contains almost an order of magnitude less power than in
998: the axisymmetric case.
999:
1000: We note the existence of a narrow but large amplitude spike just above
1001: a frequency of $7\times 10^{-3}\,c^3/GM$ in the $r=8r_g$ PSD of this
1002: run. The identification of this feature is not clear; it is not at
1003: the frequency of any expected global g- or p-mode. It is, however,
1004: confined to a single frequency bin, contains only a small amount of
1005: power and only shows up over a narrow range of radii. It seems likely
1006: that this is a noise spike.
1007:
1008: \section{Magnetohydrodynamic disks}
1009: \label{sec:mhd}
1010:
1011: Having gained an understanding of the thin hydrodynamic disks, we move
1012: onto the more astrophysically relevant case of MHD disks. In this
1013: section, we construct MHD simulations of thin accretion disks in a PW
1014: potential. We then examine the properties of the broad-band noise and
1015: search for modes in the resulting turbulent MHD disks.
1016:
1017: \subsection{Simulation set-up}
1018:
1019: We simulate geometrically-thin MHD accretion disks by building upon
1020: our hydrodynamic computational set-up described in
1021: \S~\ref{sec:hydro_setup}. As discussed in \S~\ref{sec:initial_hydro},
1022: we only consider 3-dimensional MHD simulations.
1023:
1024: Table~1 details our set of MHD simulations. Most of our discussion
1025: will center around run MHD\_1 (which we shall refer to as our
1026: canonical MHD run). In this run, the hydrodynamic variables are
1027: set-up as for the canonical hydrodynamic run (see
1028: \S~\ref{sec:hydro_setup}) except that we begin the disk in vertical
1029: hydrostatic equilibrium ($h_1=h_2=0.3r_g$ corresponding to
1030: $h_1/r=h_2/r=0.05$ at the ISCO). An initially weak magnetic field is
1031: introduced in the form of poloidal field loops specified in terms of
1032: their vector potential ${\bf A}=(A_r,A_z,A_\phi)$ in order to ensure
1033: that the initial field is divergence free. We choose the explicit
1034: form for the vector potential,
1035: \begin{eqnarray}
1036: A_\phi=A_0\,f(r,z)\,p^{1/2}\,\sin \left(\frac{2\pi r}{5h_1}\right),\hspace{0.5cm}A_r
1037: =A_z=0,
1038: \end{eqnarray}
1039: where $A_0$ is a normalization constant and $f(r,z)$ is an envelope
1040: function that is unity in the body of the disk and then smoothly goes
1041: to zero at $r=r_{\rm isco}$, $r=r_{\rm out}$ and at a location three
1042: pressure scale heights away from the midplane of the disk. The use of
1043: $f(r,z)$ keeps the initial field configuration well away from either
1044: the radial boundaries of the initial disk configuration or the
1045: vertical boundaries of the calculation domain. The final
1046: multiplicative term produces field reversals with a radial wavelength
1047: of $5h$. This results in a number of distinct poloidal loops
1048: throughout the disk. The normalization constant $A_0$ is set to give
1049: an initial ratio of gas-to-magnetic pressure of $\beta=10^3$ in the
1050: inner disk. In our canonical MHD simulation, this initial condition
1051: is evolved for a duration of $630T_{\rm isco}$ ($38800\,GM/c^3$) using
1052: the MPI-parallelized version of ZEUS described above.
1053:
1054: We supplement the ideal MHD algorithms of ZEUS in two ways. Firstly,
1055: it is necessary to impose a floor to the density field of $10^{-5}$
1056: times the initial maximum density in order to prevent the numerical
1057: integration from producing negative densities. This essentially
1058: amounts to a subtle distributed mass source. The density only reaches
1059: this floor close to the $z$-boundary (i.e., many scale heights above
1060: and below the disk plane). Secondly, we implement the prescription of
1061: Miller \& Stone (2000) to include some effects of the displacement
1062: current, principally forcing the Alfv\'en speed to correctly limit to
1063: the speed of light as the magnetic fields becomes strong. We note
1064: that this ``Alfv\'en speed limiter'' only plays a role within small
1065: patches of the tenuous magnetized atmosphere that forms at large
1066: vertical distances above and below the disk; it never plays a
1067: significant role in the body of the accretion disk.
1068:
1069: Periodic boundary conditions were imposed on the $\phi$-boundaries,
1070: and zero-gradient outflow boundary conditions were imposed at both
1071: the inner and outer radial boundaries (Stone \& Norman 1992a,b).
1072: However, the choice of the $z$-boundary condition for this kind of
1073: simulation is notoriously problematic. The most physically motivated
1074: choice would be a free outflow boundary. However, as described in
1075: Stone et al. (1996), field-line ``snapping'' at these free boundaries
1076: can halt such a simulation. Indeed, our own test simulations
1077: employing zero-gradient outflow boundary conditions on the
1078: $z$-boundaries were found to be subject to these difficulties, as well
1079: as occasional numerical instabilities appearing to result from an
1080: interplay of the imposition of the density floor and the free
1081: boundary. Furthermore, these tests showed that the tenuous matter
1082: high above the disk mid-plane generally flows slowly across these
1083: boundaries at very sub-sonic and sub-Alfv\'enic speeds; strictly, this
1084: invalidates the use of such boundary conditions anyways (since the
1085: flow on the other side of the boundary should be able to act back on
1086: the simulation domain).
1087:
1088: We adopt the solution of Stone et al. (1996) and choose to employ
1089: periodic boundary conditions in the $z$-directions. While this is
1090: obviously unphysical in the sense that matter cannot leave the
1091: simulation domain in the vertical direction, it does guarantee
1092: mathematically reasonable behavior at the boundary (eliminating
1093: numerical instabilities) and, more importantly, appears to have no
1094: effect on the dynamics of the accretion disk itself (as diagnosed
1095: through comparisons with our vertical-outflow test runs). In order to
1096: further isolate the simulated disk from the vertical boundaries, we
1097: expand the vertical domain (compared to the canonical hydrodynamic
1098: simulation) to $z\in(-3,3)$ (i.e., $\pm 10h_1)$.
1099:
1100: We perform four additional simulations aimed at demonstrating the
1101: robustness of the canonical simulation. In run MHD\_2, we restrict
1102: the vertical domain back to $z\in(-1.5,1.5)$, allowing us to gauge the
1103: importance of the location of the $z$-boundaries. Run MHD\_2hr has an
1104: identical set-up to MHD\_2 except that the radial and vertical
1105: resolution is doubled compared with the canonical run (i.e., the voxel
1106: size is reduced by a factor of two in each of the radial and vertical
1107: directions), and the azimuthal domain is doubled to
1108: $\Delta\phi=60^\circ$ at fixed resolution. Due to the factor of 8
1109: increase in number of computational cells (and the decrease in the
1110: timestep), this run was only integrated for a duration of $85T_{\rm
1111: isco}$ ($5236\,GM/c^3$). While the run time is insufficient to
1112: conduct a detailed temporal study, a comparison with run MHD\_2 does
1113: allow us to investigate the effect of both the radial/vertical
1114: resolution and the extent of the $\phi$-domain on the turbulent state
1115: (see below). As discussed below, we find that the properties of the
1116: simulated disks are very similar these two runs.
1117:
1118: Run MHD\_3 is similar to MHD\_2 except that the outer radius is pushed
1119: to $r=28r_g$ (at fixed resolution), doubling the size of the radial
1120: domain. Comparing MHD\_2 and MHD\_3, we find no evidence that the
1121: outer radial boundary is affecting the inner disk ($r<12r_g$) in any
1122: way.
1123:
1124:
1125: \subsection{Basic evolution of the MHD disks}
1126: \label{sec:basicmhdevol}
1127:
1128: % FIG 9 HERE
1129: \begin{figure*}
1130: \hbox{
1131: \hspace{-1cm}
1132: \psfig{figure=f9a.ps,width=0.55\textwidth}
1133: \hspace{-1cm}
1134: \psfig{figure=f9b.ps,width=0.55\textwidth}
1135: }
1136: \hbox{
1137: \hspace{-1cm}
1138: \psfig{figure=f9c.ps,width=0.55\textwidth}
1139: \hspace{-1cm}
1140: \psfig{figure=f9d.ps,width=0.55\textwidth}
1141: }
1142: \caption{Snapshots of evolution of the canonical MHD simulation
1143: (MHD\_1) at $t=0$ (top-left), $t=1T_{\rm isco}$ (top-right),
1144: $t=10T_{\rm isco}$ (bottom-left) and $t=100T_{\rm isco}$
1145: (bottom-right), where $T_{\rm isco}$ is the orbital period at the
1146: ISCO. Both the color table and contours show the logarithmic density
1147: structure of the disk cross-section, with 10 contours per decade of
1148: density. A range of densities spanning three orders of magnitude are
1149: shown. The curved line to the left of each frame represents the event
1150: horizon.}
1151: \label{fig:mhd_evol}
1152: \vspace{0.5cm}
1153: \end{figure*}
1154:
1155:
1156: We now discuss the evolution and general properties of these MHD
1157: disks, centering our discussion around run MHD\_1. At very early
1158: times ($t<5T_{\rm isco}$) strong hydrodynamic transients dominate the
1159: evolution as radial pressure forces drive mass into the region within
1160: the ISCO. Similarly to the hydrodynamic cases, these transients
1161: launch outwardly directed axisymmetric waves that break to form rolls.
1162: These strong hydrodynamic transients largely damp away in all but the
1163: outermost parts of the disk within $10T_{\rm isco}$. Concurrently,
1164: the MRI amplifies the initial magnetic field until the (domain wide)
1165: volume-averaged ratio of gas-to-magnetic pressure peaks at
1166: $\langle\beta\rangle \sim 5$ (at $t\approx 10T_{\rm isco}$), at which
1167: point most of the flow has become turbulent. The entire flow (outside
1168: of the ISCO) becomes turbulent by $t\approx 20T_{\rm isco}$.
1169: Figure~\ref{fig:mhd_evol} displays the density field across a vertical
1170: slice through the accretion disk at various times.
1171:
1172: Between $t=10T_{\rm isco}$ and $t=20T_{\rm isco}$, the total magnetic
1173: energy in the computational domain declines until $\langle\beta\rangle
1174: \sim 20$. After this, a quasi-steady state seems to be achieved where
1175: the magnetic field generation by the MRI-driven MHD turbulence is
1176: balanced by the removal of magnetic field energy due to numerical
1177: reconnection and magnetic buoyancy. The fact that buoyancy is playing
1178: an important role is revealed by examining time-sequences of the
1179: strengths of B-field components in $(r,z)$ cross-sections of the disk.
1180: One clearly sees highly magnetized structures being generated close to
1181: the midplane of the disk which then propagate vertically away from the
1182: midplane\footnote{In principle, one could address the importance of
1183: magnetic buoyancy in removing magnetic energy from the mid-plane
1184: regions of the disk by comparing the vertical Poynting flux with
1185: (numerical) reconnection losses. However, energy losses due to
1186: numerical reconnection are cannot be tracked in our simulation (due to
1187: the non-conservative nature of ZEUS algorithm) and, hence, a rigorous
1188: study of this issue is not possible.}. From this time until the end
1189: of the simulation at $t=630T_{\rm isco}$, MHD turbulence and hence
1190: accretion is sustained. Over the course of the simulation, the disk
1191: loses a little more than 60\% of its mass (Fig.~\ref{fig:mass_loss}).
1192:
1193: % FIG 10 HERE
1194: \begin{figure}
1195: \centerline{
1196: \hspace{-1cm}
1197: \psfig{figure=f10.ps,width=0.55\textwidth}
1198: }
1199: \caption{Normalized total mass as a function of time for run MHD\_1.}
1200: \label{fig:mass_loss}
1201: \vspace{0.5cm}
1202: \end{figure}
1203:
1204: \begin{figure*}
1205: \hbox{
1206: \hspace{-1cm}
1207: \psfig{figure=f11a.ps,width=0.55\textwidth}
1208: \hspace{-1cm}
1209: \psfig{figure=f11b.ps,width=0.55\textwidth}
1210: }
1211: \caption{{\it Left panel : }Total (domain integrated) magnetic and
1212: thermal energies for run MHD\_2 (black solid line) and its
1213: high-resolution counterpart (MHD\_2hr; red dashed line). {\it Right
1214: panel : }Azimuthally-averaged plasma-$\beta$ parameter (i.e., the
1215: ratio of the thermal to magnetic pressure) as a function of vertical
1216: height in the disk at $r=8\,r_g$. To obtain this plot, data from
1217: $t=30-35T_{\rm isco}$ have been averaged together. The solid (black)
1218: line shows run MHD\_2 whereas the dashed (red) line shows its
1219: high-resolution counterpart (MHD\_2hr).}
1220: \label{fig:mhd_compare}
1221: \end{figure*}
1222:
1223: The total magnetic energy and thermal energy undergo a slow decline as
1224: mass is drained out of the simulated disk. During this decline, the
1225: volume-averaged plasma-$\beta$ parameter within the domain remains in
1226: the range $\langle\beta\rangle\sim 20-30$. However, as expected, the
1227: $\beta$ parameter within the high-density body of the disk is
1228: appreciably higher, reaching values of $\beta\approx 70-100$. While
1229: this is significantly larger than the $\beta$ found in global
1230: simulations of thicker disks (e.g., Hawley \& Krolik 2001, 2002), it
1231: is in line with what might be expected for thin accretion disks with
1232: zero net field as diagnosed through local shearing-box simulations
1233: both with and without vertical stratification (Stone et al. 1996;
1234: Hawley, Gammie, \& Balbus 1996; Miller \& Stone 2000).
1235:
1236: A generic concern in this class of simulation is the affect of the
1237: $z$-boundaries. Once it achieves its quasi-steady state, our
1238: canonical MHD simulation displays a 2--3 order of magnitude drop in
1239: magnetic pressure, and a 5 order of magnitude drop in gas pressure,
1240: between the disk and the $z$-boundary. Thus, the disk boundary seems
1241: to be well isolated from the boundary. Further confidence is gained
1242: from an examination of run MHD\_2 in which the $z$-boundaries have
1243: been brought in from $z=\pm3$ to $z=\pm1.5$. Despite the fact that
1244: the magnetic pressure now only drops by 1 order of magnitude from the
1245: disk to the boundary, all of the results from the canonical MHD run
1246: discussed in this paper are reproduced by MHD\_2. More precisely, (1)
1247: the PSD of the fluid variables (\S\ref{sec:mhdfluidpsd}) are
1248: essentially indistinguishable, failing to show any evidence for global
1249: modes but displaying prominent local $p$-modes, (2) the PSD of the
1250: mass accretion rate has a broken power-law form with indices that
1251: differ from those found in MHD\_1 by $\Delta\approx 0.1$
1252: (comparable to the 1$\sigma$ error bars) and break frequencies that
1253: differ by $\Delta (\log \nu_{\rm break})\approx 0.05$ (again,
1254: comparable to the 1$\sigma$ error bars).
1255:
1256: Another generic concern with thin disk simulations is whether the
1257: resolution in the vertical direction is adequate. The canonical MHD
1258: simulation (MHD\_1) has approximately 13 grid cells spanning a
1259: vertical range $\Delta z=h_2$, implying that we cannot follow any
1260: modes with a wavelength smaller than $\lambda\sim h_2/2$. This is
1261: just sufficient to follow the fastest growing MRI mode (with
1262: wavelength $\lambda\sim 2\pi h/\beta^{1/2}$) if $\beta\approxlt 100$.
1263: To ensure that we have, in fact, achieved adequate resolution, we
1264: compare runs MHD\_2 and its high resolution counterpart, run MHD\_2hr.
1265: For example, Fig.~\ref{fig:mhd_compare} compares the time-dependence
1266: of the thermal and magnetic energies, as well as the height dependence
1267: of the plasma-$\beta$ parameter. While there is some indication of
1268: increased buoyancy driven escape of magnetic fields from the
1269: high-resolution simulation (as is apparent from the higher value of
1270: $\beta$ at intermediate heights), the two simulations generally
1271: compare very well. Thus, we conclude that we have achieved adequate
1272: numerical resolution.
1273:
1274: \subsection{Temporal power spectra of the basic fluid variables}
1275:
1276: \subsubsection{The importance of correcting for secular changes
1277: during the simulation}
1278: \label{sec:decay_corr}
1279:
1280: The length of our canonical MHD run makes it well suited for a detailed
1281: study of temporal variability. In particular, the long stream of
1282: simulation data facilitates the construction of PSDs. However, as we
1283: now discuss, significant complications arise in the analysis of these
1284: MHD simulation as compared with the pure hydrodynamic simulations.
1285:
1286: In the case of the hydrodynamic simulations, the background (i.e.,
1287: unperturbed) flow achieves almost a stationary state once the large
1288: transients caused by the initial conditions have died out. In
1289: particular, the lack of angular momentum transport within the
1290: hydrodynamic disk (other than that due to the very small numerical
1291: viscosity) allows the disk to achieve a non-accreting state. Density
1292: and pressure fluctuations about this background state can then be
1293: studied via the straightforward construction of the PSD. The constant
1294: background level does not contribute to the power spectrum and hence
1295: the PSD faithfully characterizes the fluctuations of interest.
1296:
1297: However, even once the initial transients have been dissipated, MHD
1298: disks never achieve this kind of stationary background flow. MHD
1299: turbulence leads to continued accretion that depletes mass from the
1300: simulated disks. This decline in total mass leads to secular changes
1301: (with an approximately exponential form) in the density and pressure
1302: of the background flow. Unless corrected for, even a rather slow
1303: exponential decay can have a significant influence on the PSD of the
1304: pressure or density fluctuations, severely affecting attempts to
1305: characterize the properties of the astrophysically relevant
1306: fluctuations (i.e., the fluctuations that would be present in the
1307: ideal case of a steady-state disk in which the mass was replaced from
1308: a large reservoir).
1309:
1310: To see this, consider some quantity whose time-series $f(t)$ we extract
1311: from the simulation (for example, this could be the mid-plane
1312: density or pressure at some given radius). Let us assume that this
1313: can be decomposed as
1314: \begin{equation}
1315: f(t)=\epsilon(t)[1+g(t)],
1316: \label{eq:fluc_decomp}
1317: \end{equation}
1318: where $g(t)$ is the time-series of the astrophysically-interesting
1319: fluctuations about some mean state (i.e., $\langle g\rangle=0$), and
1320: $\epsilon(t)$ is a decay function that describes the secular change in
1321: the background state due to the draining of mass from the simulation.
1322: It must be noted that the decomposition given in
1323: eq.~\ref{eq:fluc_decomp} is not completely general; this assumes that
1324: the amplitude of the ``real'' fluctuations is proportional to the mean
1325: background value, and that the properties of the fluctuations
1326: otherwise remain invariant as the background state decays. This
1327: decomposition would be valid if the decay of the simulated disk simply
1328: amounted to a gradual decline in the density scale of the simulation
1329: while the temperature and velocity structures remained unchanged. We
1330: shall refer to such (simulated) disks as {\it density-invariant
1331: disks}. This does indeed appear to describe our simulated MHD disks
1332: (i.e., there is little or no corresponding secular change in disk
1333: thickness or characteristic turbulent velocities). In this case, the
1334: mass accretion rate will be proportional to the density and the decay
1335: will hence have an exponential form,
1336: $\epsilon=\epsilon_0\,e^{-t/t_0}$.
1337:
1338: In the uninteresting case where there are no fluctuations ($g(t)=0,
1339: \forall t$), the observed signal is just $f(t)=\epsilon(t)$, leading
1340: to a Fourier Transform (FT) and PSD given by
1341: \begin{equation}
1342: \tilde{\epsilon}(\omega)=\frac{A_1}{(1/t_0)+i\omega}, \hspace{0.5cm}{\rm and}\hspace{0.5cm}P_\epsilon(\omega)=\frac{A_2}{(1/t_0)^2+\omega^2},
1343: \end{equation}
1344: where $A_1$ and $A_2$ are uninteresting normalization constants.
1345: Thus, for $\omega\gg 1/t_0$, the PSD of the exponential decay goes as
1346: $P_\epsilon(\omega)\propto\omega^{-2}$.
1347:
1348: In the more interesting case of non-zero fluctuations, the FT of the
1349: observed signal is,
1350: \begin{equation}
1351: \tilde{f}(\omega)=\tilde{\epsilon}(\omega)+\int
1352: \tilde{\epsilon}(\omega^\prime)\tilde{g}(\omega-\omega^\prime)\,d\omega^\prime,
1353: \label{eq:combined_ft}
1354: \end{equation}
1355: i.e., the sum of the FT of the exponential decay with the convolution
1356: of the FTs of the decay and the interesting fluctuations. When the
1357: fluctuations are small compared with the (decaying) background state,
1358: as is the case for the density and pressure, one can see that the PSD
1359: of the observed signal will be dominated by the $1/\omega^2$
1360: associated with the decay. Even in the case where the fluctuations
1361: are large compared with the decay, the PSD will still be influenced by
1362: the exponential decay due to the convolution term in
1363: eqn.\ref{eq:combined_ft}. In particular, regions of the ``real'' PSD
1364: which are steeper than $\omega^{-2}$ (including the high-frequency
1365: wing of any QPO or regions above a high-frequency break) will tend to
1366: get filled.
1367:
1368: Clearly, we must correct for this decay of the background state, and
1369: be cognizant of manifestations of any remaining uncorrected effects of
1370: this decay. This procedure plays the same role as the
1371: ``pre-whitening'' employed by Schnittman, Krolik, \& Hawley (2006;
1372: hereafter SKH). We proceed by dividing the observed time series by a
1373: ``best-fitting'' exponential decay function. Here, the time-constant
1374: of the exponential decay $t_0$ is estimated via two methods. Firstly,
1375: we can set $t_0$ by the requirement that the starting and final values
1376: of the observed series are equal [$f(t=0)=f(t=T)$, where $t=T$
1377: corresponds to the end of the time-series]. This is the procedure
1378: adopted by Schnittman et al. (2006) except that they employ a linear
1379: form for the decay function. Secondly, we can set $t_0$ via a
1380: least-square fit of an exponential form to the observed time-series.
1381: These methods give very similar values of $t_0$ and similar corrected
1382: PSDs when applied to density or pressure fluctuations, as we shall now
1383: see.
1384:
1385: \subsubsection{Power spectra of fluid variables in disk mid-plane}
1386: \label{sec:mhdfluidpsd}
1387:
1388: % FIG 12 HERE
1389: \begin{figure*}
1390: \begin{center}
1391: \hbox{
1392: \psfig{figure=f12a.ps,width=0.55\textwidth}
1393: \psfig{figure=f12b.ps,width=0.55\textwidth}
1394: }
1395: \hbox{
1396: \psfig{figure=f12c.ps,width=0.55\textwidth}
1397: \psfig{figure=f12d.ps,width=0.55\textwidth}
1398: }
1399: \end{center}
1400: \caption{PSD for azimuthally-averaged midplane fluid quantities from
1401: in run MHD\_1. Panels show PSDs for radial velocity (top left),
1402: vertical velocity (top right), density (bottom left) and pressure
1403: (bottom right). The density and pressure variables have been divided
1404: by the least-square best-fitting exponential function to correct for
1405: the secular decay of the simulation, as discussed in
1406: \S~\ref{sec:decay_corr}. Also shown are the radial epicyclic
1407: frequency (solid), orbital frequency (dashed) and the $n=1,2,3$ pure
1408: vertical p-modes (from left to right dot-dashed lines). The absolute
1409: scaling of the PSDs, as indicated by the color-bar, is arbitrary. }
1410: \label{fig:mhd_psd}
1411: \end{figure*}
1412:
1413: We now examine the PSD of the azimuthally averaged fluid variables
1414: (velocities, pressure and density) for the simulated MHD disks, as we
1415: did with the hydrodynamic disks in \S~\ref{sec:axisym}. The top
1416: panels of Fig.~\ref{fig:mhd_psd} show the PSD for the radial and
1417: vertical components of velocity in the midplane of the disk, $v^{\rm
1418: mid}_r(r,t)$ and $v^{\rm mid}_z(r,t)$ for a duration lasting $\Delta
1419: t=409.6T_{\rm isco}$ starting at $t=100T_{\rm isco}$ (i.e., well after
1420: all of the initial transients have dissipated and a quasi-steady
1421: turbulent state has been established). Note that, outside of the
1422: ISCO, these velocity components are themselves first-order fluctuating
1423: quantities (i.e., the radial and vertical velocity of the background
1424: state is approximately zero) and, within the density-invariant disk
1425: assumption, will have characteristic fluctuation amplitudes that
1426: remain constant throughout the simulation. Hence, we do not need to
1427: correct these PSDs for the effect of the density decay in the
1428: simulation.
1429:
1430: It is readily seen that the PSD of the midplane radial velocity
1431: $v_r^{\rm mid}$ is dominated by the radial epicyclic frequency from
1432: $r\sim 7-8r_g$ out to the outer radial boundary. Inside of
1433: $r\approx 7-8r_g$, the PSD shows broad band power associated with
1434: the transition from turbulent to plunging flow.
1435:
1436: The PSD of the midplane vertical velocity $v^{\rm mid}_z$ is quite
1437: different and can be interpreted in the light of the {\it local} fluid
1438: oscillations discussed in \S~\ref{sec:local_osc}. Below the radial
1439: epicyclic frequency, there is a broad spectrum of g-modes. Above the
1440: orbital frequency, the distinct modes described by
1441: eqn.~\ref{eq:vert_modes} become apparent; the $n=0$ mode
1442: ($\omega>\Omega$) and $n=2$ mode ($\omega>2.08\Omega$) can be picked
1443: out as distinct tracks, and there are hints of higher frequency modes
1444: as well. The fact that only even-$n$ modes are seen is readily
1445: understood given that the odd-$n$ modes have $v_z$-nodes at the
1446: midplane, as is evident from the power deficit centered on the $n=1$
1447: frequency in the $v_z$ PSD.
1448:
1449: The bottom panels of Fig.~\ref{fig:mhd_psd} show the corresponding PSD
1450: for the density and pressure from the canonical simulation. The
1451: displayed PSDs have been corrected for the decay of the background
1452: state by dividing through by an exponential curve that best-fits (in a
1453: least square sense) the domain-averaged density or pressure. Very
1454: similar results are obtained using an exponential determined from just
1455: the end points. The radial banding of the low-frequency noise is
1456: largely the effect of the pre-whitening procedure; the full
1457: (non-corrected) low-frequency power is somewhat greater and the
1458: bands correspond to the residuals between the real decay and the
1459: simple exponential model. In addition to this low frequency noise
1460: (which is of secondary interest to the investigation presented in this
1461: paper), both PSDs show an enhancement corresponding to the $n=1$
1462: vertical p-mode (with $\omega>1.63\Omega$). As expected, there is no
1463: enhancement in the density or pressure PSDs corresponding to the $n=0$
1464: or $n=2$ p-modes; these have pressure and density nodes at the
1465: midplane.
1466:
1467: % FIG 13 HERE
1468: \begin{figure}
1469: \centerline{
1470: \psfig{figure=f13.ps,width=0.55\textwidth}
1471: }
1472: \caption{PSD of the midplane (decay-corrected) pressure for $\Delta
1473: r=0.5r_g$ wide zones centered on $r=7r_g$ (top, red curve),
1474: $r=8r_g$ (middle, black curve), and $r=9.2r_g$ (bottom, green
1475: curve). The thin vertical dashed line marks the maximum radial
1476: epicyclic frequency. A line-segment with a slope of $-2$ is shown for
1477: reference. Also shown (heavy vertical blue lines) are the radial and
1478: vertical epicyclic frequencies at $r=9.2r_g$ where the simplest
1479: form of the parametric instability model would predict resonances.}
1480: \label{fig:mhd_1dpsd}
1481: \vspace{0.5cm}
1482: \end{figure}
1483:
1484:
1485: In stark contrast to the hydrodynamic simulation, the
1486: (decay-corrected) midplane pressure PSD does not show the
1487: $\kappa$-bounded vertical ``ridge'' on the frequency-radius plane that
1488: is characteristic of trapped g-modes. The absence of an excited
1489: trapped g-mode is also confirmed by examining the PSD at $r=8r_g$
1490: (Fig.~\ref{fig:mhd_1dpsd}, normalized such that a direct comparison
1491: can be made with the 3-d hydrodynamic results of
1492: Fig.~\ref{fig:hd3d_psd}). It is important to note, however, that a
1493: trapped g-mode of the strength seen in our 3-d hydrodynamic simulation
1494: would not stand out from the background turbulence. Thus, while it is
1495: apparent that the fundamental trapped g-mode is not strongly excited
1496: by the turbulence (in the way that the local p-modes are, for
1497: example), we cannot say whether the g-mode is actively damped by the
1498: turbulence.
1499:
1500: Finally, we note that there is no indication that the parametric
1501: resonance instability of Abramowicz \& Kluzniak is at work, at least
1502: in its simplest form. As discussed in \S~2, this model predicts its
1503: strongest resonance at the location where the radial and vertical
1504: epicyclic frequencies are in a 2:3 ratio; this occurs at $r=9.2r_g$
1505: in the PW potential. As shown in Fig.~\ref{fig:mhd_1dpsd}, the PSD of
1506: the pressure fluctuations at $r=9.2r_g$ shows no structure
1507: associated with the local epicyclic frequencies. We have verified
1508: that a similar conclusion holds true for the PSD of the other
1509: variables. While it is possible that non-linear effects have shifted
1510: the location of the resonance inwards from $r=9.2r_g$ (A2003), we
1511: note that the PSDs of Fig.~\ref{fig:mhd_psd} show no obvious radius at
1512: which the power at the radial epicyclic and the orbital frequencies
1513: appears to be locally enhanced.
1514:
1515: \subsection{Temporal properties of the instantaneous black hole accretion rate}
1516: \label{sec:lightcurves}
1517:
1518: The previous section addressed the temporal properties of the
1519: fundamental fluid variables through the body of the disk. Of course,
1520: real observations of accretion disks measure the electromagnetic
1521: radiation generated by the accretion flow. While our simulations
1522: miss all of the physics relevant for making a first-principles
1523: prediction of the observed lightcurve, it is still instructive to
1524: analyze a simple scalar quantity that can be generated from the
1525: simulation and may be related to the observed hard X-ray radiation
1526: (since it is the hard X-rays that carry the HFQPO signal).
1527:
1528: % FIG 14 HERE
1529: \begin{figure}[b]
1530: \centerline{
1531: \psfig{figure=f14.ps,width=0.55\textwidth}
1532: }
1533: \caption{The instantaneous mass accretion rate onto the black hole
1534: from our canonical MHD simulation (run MHD\_1).}
1535: \label{fig:lightcurve}
1536: \end{figure}
1537:
1538:
1539: Here we consider one such proxy for the observed lightcurve, the
1540: {\it instantaneous mass accretion rate} into the black hole,
1541: \begin{equation}
1542: \dot{M}=\int_{\partial {\cal R}\,_i}r(-v_r)\rho\,dS,
1543: \end{equation}
1544: where the integral is performed over the surface $\partial {\cal
1545: R}\,_i$ defining the inner radial boundary of the computational
1546: domain. Figure~\ref{fig:lightcurve} shows the raw lightcurve (i.e.,
1547: not corrected for the exponential decay of the disk). In the
1548: remainder of this section we address the temporal properties of this
1549: lightcurve. Of particular interest is the presence of breaks or QPOs
1550: in the PSD of this light curve. Hence, we need a quantitative
1551: approach by which the significance of such features in the PSD can be
1552: assessed. We begin by discussing our general statistical approach,
1553: which differs from the one advocated by SKH.
1554:
1555: \subsubsection{Analysis method}
1556:
1557: Our approach to PSD fitting is predicated on the assumption (also made
1558: by SKH) that at a given frequency the power density has an exponential
1559: probability distribution; if the mean is $p_0$, then the probability
1560: of measuring a power between $p$ and $p+dp$ is
1561: \begin{equation}
1562: P(p)dp={1\over p_0}e^{-p/p_0}dp\; .
1563: \end{equation}
1564: Suppose that we have a model that predicts a power density
1565: $p_{\rm mod,i}$ for frequency bin $i$, and that in our MHD
1566: simulation we actually observe a power density $p_{\rm obs,i}$
1567: in that bin. The likelihood of the data given the model in
1568: that bin is then
1569: \begin{equation}
1570: {\cal L}_i=(1/p_{\rm mod,i})\exp(-p_{\rm obs,i}/p_{\rm mod,i}).
1571: \end{equation}
1572: The likelihood of the whole power density spectrum given the
1573: model is the product of the individual likelihoods, but the
1574: log likelihood is typically more useful:
1575: \begin{equation}
1576: \ln{\cal L}=\sum_i \left[-\ln p_{\rm mod,i}-p_{\rm obs,i}/p_{\rm mod,i}
1577: \right]\; .
1578: \end{equation}
1579: This is the figure of merit for a given model. It can therefore be
1580: used for standard tasks such as parameter estimation and model
1581: comparison (e.g., determining if a QPO or break is required by the
1582: data).
1583:
1584: Note that to maximize the information content, the bin size should be
1585: the smallest possible, in this case the frequency resolution of the
1586: raw PSD. As a result, this method does not require rebinning to
1587: coarser resolution. Our approach therefore yields an accurate
1588: evaluation of one or more precisely specified models, as opposed to
1589: the broader but less sensitive method of trying to detect a signal in
1590: a model-independent way.
1591:
1592: We apply a Markov Chain Monte Carlo method to search for best fits and
1593: establish confidence regions. As discussed in
1594: \S~\ref{sec:decay_corr}, we correct for secular changes. Given the
1595: large amplitude fluctuations in the $\dot{M}$ and $S_{\rm tot}$
1596: curves, the end-point method discussed in \S~\ref{sec:decay_corr} is
1597: not appropriate. Hence, we employ the least-square method described
1598: in \S4.3.1 in order to determine and then divide out the best-fitting
1599: exponential decay.
1600:
1601: \subsubsection{Results}
1602: \label{sec:mdot_psd}
1603:
1604: % FIG 15 HERE
1605: \begin{figure*}
1606: \hbox{
1607: \psfig{file=f15a.ps,width=0.45\textwidth}
1608: \psfig{file=f15b.ps,width=0.45\textwidth}
1609: }
1610: \caption{ Comparison of the PSD and Fourier phase of the mass
1611: accretion rate without (left panel) and with (right panel)
1612: multiplication of the time series by an exponential in time designed
1613: to compensate for the slow loss of mass from the disk. The similarity
1614: of the two implies that at these frequencies the variability is
1615: dominated by intrinsic fluctuations instead of by secular changes in
1616: the disk.}
1617: \label{fig:mdotcomp}
1618: \end{figure*}
1619:
1620:
1621: The mass accretion rate $\dot{M}$ has sufficient intrinsic variability
1622: that secular changes in the disk properties have only minor effects on
1623: the PSD. This is evident from Fig.~\ref{fig:mdotcomp}, which compares
1624: the PSD and Fourier phase of the raw $\dot{M}$ curve as a function of
1625: frequency (left panel) with these after multiplication by the
1626: exponential in time that minimizes the overall rms amplitude (right
1627: panel). The lack of significant differences suggests that inferences
1628: drawn from the PSD are robust. The comparison shown in
1629: Fig.~\ref{fig:mdotcomp} is for the second quarter of run MHD\_1
1630: (approximately $t=150T_{\rm isco}$ to $t=310T_{\rm isco}$); analyses
1631: of the third and fourth quarters also reveal a lack of sensitivity to
1632: the secular decay of the disk.
1633:
1634: Having established that the secular decay of the disk is unimportant
1635: for the PSD of ${\dot M}$, we will work directly with the raw time
1636: series from run MHD\_1, without multiplication by an exponential
1637: function. We analyze separately the second, third, and fourth quarter
1638: of the data (each encompassing a period of $\Delta t\approx 160T_{\rm
1639: isco}$) to look for trends or stability in the PSD, while discarding
1640: the first quarter as potentially biased by initial conditions. The
1641: results are summarized in Table~\ref{tab:tabmdot}, where we compare
1642: single power law models for the PSD with models involving a broken
1643: power law. We use the method described in the previous section; note
1644: that only differences in the log likelihood $\ln{\cal L}$ are
1645: important rather than the absolute magnitude of $\ln{\cal L}$.
1646:
1647: \begin{table*}
1648: \centering
1649: \begin{tabular}{lccr}\hline
1650: Data segment&Model&Parameters and uncertainties&Maximum $\ln{\cal
1651: L}$\\\hline
1652: 1&Single PL&$\Gamma=2.50\pm 0.05$&9131\\
1653: &Broken PL&$\Gamma_1=0.82\pm 0.09$, $\Gamma_2=2.96\pm 0.07$,&9238\\
1654: &&$\log_{10}(\nu_{\rm break})=-1.94\pm 0.04$\\\hline
1655: 2&Single PL&$\Gamma=2.70\pm 0.08$&6508\\
1656: &Broken PL&$\Gamma_1=1.45\pm 0.11$, $\Gamma_2=2.91\pm 0.10$,&6609\\
1657: &&$\log_{10}(\nu_{\rm break})=-1.65\pm 0.07$\\\hline
1658: 3&Single PL&$\Gamma=2.16\pm 0.04$&9561\\
1659: &Broken PL&$\Gamma_1=1.47\pm 0.07$, $\Gamma_2=2.89\pm 0.09$,&9630\\
1660: &&$\log_{10}(\nu_{\rm break})=-1.60\pm 0.04$\\\hline
1661: \end{tabular}
1662: \caption{Model comparisons for PSD of mass accretion rate $\dot{M}$.
1663: $\Gamma$ is the power law index of a single power-law model for the
1664: PSD; $P(\nu)\propto \nu^{-\Gamma}$. $\Gamma_1$ and $\Gamma_2$ are
1665: the
1666: two indices obtained from a broken power law; $P\propto
1667: \nu^{-\Gamma_1}$ for $\nu<\nu_{\rm break}$, $P\propto
1668: \nu^{-\Gamma_2}$
1669: for $\nu>\nu_{\rm break}$. All error bars are one standard
1670: deviation.}
1671: \label{tab:tabmdot}
1672: \end{table*}
1673:
1674: From this analysis, we find compelling evidence for a break in the
1675: power law characterizing the PSD of ${\dot M}$. Compared to a single
1676: power law, the broken power fit is better by $\Delta\ln{\cal
1677: L}=76-90$, hence the maximum likelihood ratio is at least
1678: $\exp(76)=10^{33}$ in all three data segments independently. The
1679: break frequency and power law slopes are consistent between the second
1680: and third data segments, but these do not match the first data
1681: segment. The break frequencies are within a factor of two of, but not
1682: consistent with, the orbital frequency at the ISCO, $\log_{10}(\nu_{\rm
1683: ISCO})=-1.79$. Therefore, although there is a clear steepening in the
1684: power density spectrum, it is not possible at this point to assign a
1685: specific physical meaning to the break. We note that this general
1686: form of the PSD, i.e., an approximate power-law with curvature or a
1687: break at frequencies close to the ISCO orbital frequency, has been
1688: previously seen in the mass accretion rate of global disk simulations
1689: (Hawley \& Krolik 2001, 2002).
1690:
1691: We see no indications of QPOs in the $\dot{M}$ PSD. Quantitatively,
1692: we add a Lorentzian QPO to the broken power-law PSD model in which the
1693: QPO centroid frequency, full-width half-maximum, and amplitude are
1694: allowed to be free parameters, with the one restriction that the
1695: quality factor of the QPO must exceed 2. We find that the peak power
1696: of any QPO cannot exceed 2\% of the continuum power measured at the
1697: centroid of the QPO at the 99\% confidence level.
1698:
1699: \section{Discussion}
1700:
1701: \subsection{Comparison with previous numerical results}
1702:
1703: In recent years, several groups have reported temporal analyses of MHD
1704: disk simulations. Here, we briefly compare our work with some of
1705: these published results.
1706:
1707: Probably the most relevant previous work is that of Arras, Blaes \&
1708: Turner (2006; hereafter ABT). These authors perform a local, shearing
1709: box MHD simulation of a patch of an accretion disk; this provides a
1710: controlled environment in which fluid modes can be characterized. ABT
1711: find that the MHD turbulence excites a spectrum of distinct acoustic
1712: modes as well as radial epicyclic motions. However, they note a lack
1713: of distinct inertial modes (g-modes) and use this fact to argue
1714: against the excitation of trapped g-modes in global accretion disks.
1715: Our findings are completely in line with those of ABT, and represent
1716: an extension of ABT's conclusions to global simulations of thin
1717: accretion disks.
1718:
1719: There have been QPOs reported from global simulations. Kato (2004d)
1720: performed a global MHD disk simulation in a PW potential and presented
1721: an analysis of quantities derived from the mass accretion rate.
1722: Through visual inspection of the resulting PSDs from four periods of
1723: the (long) simulation, he reports a pair of transient QPOs and a pair of
1724: QPOs that are labeled as persistent (although it is not clear that they
1725: are present in the PSD of all data segments, and the statistical
1726: significance of the features is unclear). The QPOs are attributed
1727: to resonances between the vertical and radial epicyclic frequencies,
1728: and it is found that these QPO pairs have frequency ratios that are
1729: {\it approximately} 3:2. We find no evidence of these resonances in
1730: our simulations. In another interesting difference, an inspection of
1731: the radially-resolved PSDs in Kato (2004d) reveal no signs of the
1732: local p-modes that seem to feature so prominently in our PSDs. While
1733: the reason for these discrepancies is unclear, we do note that the
1734: Kato (2004d) simulations have an order of magnitude less resolution in
1735: both the azimuthal and (more importantly) the vertical direction,
1736: although his simulations do have a significantly larger computational
1737: domain. It is possible that the Kato (2004d) simulations have failed
1738: to adequately resolve the vertical dynamics of the thin disk.
1739:
1740: Chan et al. (2006) have also performed global MHD disk simulations in
1741: a PW potential using the pseudo-spectral algorithm of Chan, Psaltis,
1742: \& \"Ozel (2005, 2006). Their study included a post-processing of the
1743: simulation to include detailed radiative transfer, and was explicitly
1744: targeted at understanding the variability (including the large
1745: amplitude flaring) of the hot accretion flow around the black hole at
1746: the center of the Galaxy. They found that the turbulence of the
1747: quiescent flow could only produce a factor of two modulations in the
1748: observed luminosity. To model the large amplitude flares, they
1749: introduce large density perturbations into the flow. After being
1750: perturbed, the disk displays a QPO with a frequency equal to the
1751: orbital frequency at the magnetosonic point. Given that the Chan et
1752: al. study is exploring a rather different regime of accretion than our
1753: present study (i.e., hot, thick accretion flows versus thin, cold
1754: accretion flows), it is hard to make a direct comparison.
1755:
1756: Finally, SKH have performed detailed analyses of General Relativistic
1757: MHD (GRMHD) simulations of disks performed using the code of
1758: De~Villiers \& Hawley (2003). In particular, they have studied a long
1759: ($6000GM/c^3$) simulation of a disk around a Schwarzschild black hole,
1760: focusing on the temporal behavior of proxy-lightcurves (rather than
1761: the underlying fluid properties discussed in this paper). It is
1762: unclear from their analysis whether their simulated accretion disk has
1763: excited local p-modes of the type that we find in this current work.
1764: SKH do find, though, that a proxy-lightcurve based on radiative
1765: transfer through the disk assuming black body emission and free-free
1766: absorption displays QPOs with an approximate 3:2 ratio. However,
1767: these QPOs are transient, only appearing at certain times and certain
1768: viewing inclinations.
1769:
1770: \subsection{Comments on 3:2 frequency ratios}
1771:
1772: As described above, several authors have reported transient QPO pairs
1773: from MHD simulations with frequency ratio 3:2. It is tempting to
1774: interpret these as resonance phenomena. Here, we note that there are
1775: several ways that approximate 3:2 ratios can be generated that do not
1776: necessarily involve resonances and, hence, one should guard against
1777: over-interpreting QPO pairs. Indeed, we must remember that some
1778: sources have QPO frequency ratios that are inconsistent with 3:2
1779: (e.g., the 67~Hz and 41~Hz QPOs from GRS~1915+105; see Strohmayer
1780: 2001).
1781:
1782: For example, consider the local vertical p-modes of the disk
1783: (discussed in \S~\ref{sec:local_osc}). At a given radius, the $n=1$
1784: vertical pressure mode has a frequency of $\omega_{\rm
1785: vert,1}=\sqrt{\gamma+1}\,\Omega$, where $\Omega$ is the orbital
1786: frequency. For a gas pressure dominated disk in which $\gamma=5/3$,
1787: $\omega_{\rm vert,1}/\Omega=1.63$, and for a radiation pressure
1788: dominated disk in which $\gamma=4/3$, $\omega_{\rm
1789: vert,1}/\Omega=1.53$. If some unspecified physical process enhances
1790: emission from a given ring of the disk, one can imagine a situation
1791: where QPOs are generated at the orbital frequency and the $n=1$
1792: vertical pressure mode, thereby giving frequency ratios compatible
1793: with the measured ratios in several sources. Alternatively, the next
1794: lowest vertical mode that has a vertical velocity node in the midplane
1795: (and thus maximum variation of pressure and density there) has a
1796: frequency of $\omega_{\rm vert,3}=\sqrt{3\gamma+1}\,\Omega$. As a
1797: result, when $\gamma=5/3$ we have $\omega_{\rm vert,3}/\omega_{\rm
1798: vert,1}=1.5$, and when $\gamma=4/3$ the ratio is 1.46. Once again, if
1799: nature picks out a specific radius and the emission is modulated by
1800: the $n=1$ and $n=3$ modes, we would see QPOs with frequency ratios
1801: entirely consistent with the observations.
1802:
1803: As another example, suppose that the disk emission is modulated at the
1804: vertical and radial epicyclic frequencies, and that the emission is
1805: distributed in radius according to a standard Page \& Thorne (1974)
1806: disk. The radial distribution of the emission then peaks close to the
1807: radius where the radial epicyclic frequency is a maximum (and hence is
1808: slowly changing with radius). One might then expect to see a pair of
1809: QPOs corresponding to these two epicyclic frequencies. The frequency
1810: ratio depends only weakly on the spin parameter, ranging from 1.46 at
1811: $a/M=0$ to 1.7 at $a/M=0.9$.
1812:
1813: \section{Conclusions}
1814:
1815: The origin of HFQPOs remains elusive. Our simulations of
1816: geometrically-thin accretion disks have shown that MRI-driven MHD
1817: turbulence does not excite the trapped g-modes of Nowak \& Wagoner
1818: (1992), even when those modes definitively exist in the equivalent
1819: hydrodynamic disk. We have also shown that MHD turbulence does not
1820: excite the parametric resonance instability of Abramowicz \&
1821: Klu\'zniak (2001). Instead, the only distinct modes found in our
1822: simulated MHD disks are local vertical p-modes and radial epicyclic
1823: oscillations.
1824:
1825: Clearly, the failure of all simulations to date to produce stable QPO
1826: pairs of the type seen in GBHBs suggests that either the QPOs are
1827: too weak to be detected in the simulations or the models are missing
1828: some important physical ingredients. It is an open question whether
1829: the global disk modes or the parametric instabilities discussed in
1830: this paper can be excited once one includes the full effects of GR
1831: close to rapidly spinning black holes and/or radiation physics.
1832: Indeed, the fact that HFQPOs are only seen in rather special spectral
1833: states (the soft intermediate state; e.g., see Belloni 2006) when the
1834: accretion rate is thought to be comparable to the Eddington limit
1835: suggests that radiation physics, in particular, may well be important
1836: to the HFQPO mechanism. It is also noteworthy that the transition
1837: from the hard intermediate state into the soft intermediate state is
1838: associated with powerful relativistic ejection events. Thus, another
1839: interesting possibility is that HFQPO production occurs in the black
1840: hole magnetosphere (i.e., the base of the jet) and not the accretion
1841: disk at all.
1842:
1843: \section*{Acknowledgments}
1844:
1845: We thank Eve Ostriker for insightful discussions as well as allowing
1846: us the use of her MPI-parallelized version of ZEUS. We are also
1847: grateful to Andy Fabian, Gordon Ogilvie, Sean O'Neill, Jim Pringle,
1848: Jon Miller, Phil Uttley, Simon Vaughan, and Bob Wagoner for comments
1849: and discussion that significantly improved this paper. All
1850: simulations described in this paper were performed on the Beowulf
1851: cluster (``The Borg'') supported by the Center for Theory and
1852: Computation (CTC) in the Department of Astronomy at the University of
1853: Maryland College Park. CSR and MCM thank the National Science
1854: Foundation for support under grant AST~06-07428. CSR also gratefully
1855: acknowledges the University of Maryland's Graduate Research Board
1856: Semester Award Program which supported the early phases of this work.
1857:
1858: \section*{References}
1859:
1860: \noindent Abramowicz, M.~A., Almergren, G.~J.~E., Klu\'zniak, W.,
1861: Thampan, A.~V., \& Wallinder, F. 2002, Class. Quant. Grav., 19, L57
1862:
1863: \noindent Abramowicz, M.~A., Karas, V., Klu\'zniak, W., Lee, W.~H., \&
1864: Rebusco, P. 2003, PASJ, 55, 467 (A2003)
1865:
1866: \noindent Abramowicz, M.~A., \& Klu\'zniak, W. 2001, A\&A, 374, L19
1867:
1868: \noindent Abramowicz, M.~A., \& Klu\'zniak, W. 2003, Gen. Rel. Grav.,
1869: 35, 69
1870:
1871: \noindent Arras P., Blaes O.~M., \& Turner N.~J. 2006, ApJ, 645, L65 (ABT)
1872:
1873: \noindent Balbus, S.~A., \& Hawley, J.~F. 1991, ApJ, 376, 214
1874:
1875: \noindent Balbus, S.~A., \& Hawley, J.~F. 1998, Rev. Mod. Phys., 70, 1
1876:
1877: \noindent Belloni, T. 2006, Ad. Space Res., 38, 2801
1878:
1879: \noindent Binney, J., \& Tremaine, S. 1987, Galactic Dynamics (Princeton:
1880: Princeton Univ. Press), 359
1881:
1882: \noindent Chan, C.~K., Psaltis, D., \& \"Ozel, F. 2005, ApJ, 628, 353
1883:
1884: \noindent Chan, C.~K., Psaltis, D., \& \"Ozel, F. 2006, ApJ, 645, 506
1885:
1886: \noindent Chan, C.~K., et al. 2006, ApJ, submitted (astro-ph/0611269)
1887:
1888: \noindent Churazov, E., Gilfanov, M., \& Revnivtsev, M. 2001, MNRAS,
1889: 321, 759
1890:
1891: \noindent Cowling, T.~G. 1957, Quart. J. Mech. Appl. Math., 10, 129
1892:
1893: \noindent De~Villiers, J.-P., \& Hawley J.~F. 2003, ApJ, 592, 1060
1894:
1895: \noindent Friedman, J.~L., \& Schutz, B.~F. 1978, ApJ, 221, 937
1896:
1897: \noindent Gammie, C.~F., McKinney, J.~C., \& T\'oth, G. 2003,
1898: ApJ, 589, 444
1899:
1900: \noindent Gierlinski M., Middleton M., Ward M., Done C., 2008, Nature,
1901: in press
1902:
1903: \noindent Gleissner T., Wilms J., Pottschmidt K., Uttley P., Nowak
1904: M.A., Staubert R., 2004, A\&A, 391, 875.
1905:
1906: \noindent Hawley J.F., Krolik J.H., 2001, ApJ, 548, 348
1907:
1908: \noindent Hawley J.F., Krolik J.H., 2002, ApJ, 566, 164
1909:
1910: \noindent Hawley, J.~F., Gammie, C.~F., \& Balbus, S.~A. 1996, ApJ, 464, 690
1911:
1912: \noindent Hantao, J., Burin, M., Schartman, E., \& Goodman, J. 2006, Nature,
1913: 444, 343
1914:
1915: \noindent Kato, S. 1990, PASJ, 42, 99
1916:
1917: \noindent Kato, S. 1991, PASJ, 43, 557
1918:
1919: \noindent Kato, S. 1993, PASJ, 45, 219
1920:
1921: \noindent Kato, S. 2004a, PASJ, 56, 559
1922:
1923: \noindent Kato, S. 2004b, PASJ, 56, 905
1924:
1925: \noindent Kato, S. 2004c, PASJ, 56, L25
1926:
1927: \noindent Kato, S., \& Fukue, J. 1980, PASJ, 322, 377
1928:
1929: \noindent Kato, S., \& Honma, F. 1991, PASJ, 43, 95
1930:
1931: \noindent Kato, Y. 2004d, PASJ, 56, 931
1932:
1933: \noindent Krolik, J. 1999, ApJ, 515, L73
1934:
1935: \noindent Lehr, D., Wagoner, R.~V., \& Wilms, J., 2000, astro-ph/0004211
1936:
1937: \noindent Lubow, S.~H., \& Pringle, J.~E. 1993, ApJ, 409, 360 (LP93)
1938:
1939: \noindent Lyubarskii, Yu.~E. 1997, MNRAS, 292, 679
1940:
1941: \noindent Markoff, S., Nowak, M.~A., Corbel, S., Fender, R.~P., \& Falcke, H.
1942: 2003, New AR, 47, 491
1943:
1944: \noindent Markovi\'c, D., \& Lamb, F.~K. 1998, ApJ, 507, 316
1945:
1946: \noindent McClintock, J.~E., \& Remillard, R.~A. 2003, astro-ph/0306213
1947:
1948: \noindent McHardy, I.~M., K\"ording, E., Knigge, C., Uttley, P., \& Fender,
1949: R.~P. 2006, Nature, 444, 730
1950:
1951: \noindent Miller K.A., Stone J.M., 2000, ApJ, 534, 398
1952:
1953: \noindent Nowak, M.~A., \& Wagoner, R.~V. 1991, ApJ, 378, 656 (NW91)
1954:
1955: \noindent Nowak, M.~A., \& Wagoner, R.~V. 1992, ApJ, 393, 697 (NW92)
1956:
1957: \noindent Nowak, M.~A., \& Wagoner, R.~V. 1993, ApJ, 418, 187
1958:
1959: \noindent Nowak, M.~A., Wagoner, R.~V., Begelman, M.~C., \& Lehr, D.~E.
1960: 1997, ApJ, 477, L91
1961:
1962: \noindent Okazaki, A.~T., Kato, S., \& Fukue, J. 1987, PASJ, 39,
1963: 457
1964:
1965: \noindent Ortega-Rodriguez, M., Silbergleit, A.~S., \& Wagoner, R.~V.
1966: 2001, ApJ, 567, 1043
1967:
1968: \noindent Paczynski, B., \& Wiita, P.~J. 1980, A\&A, 88, 23 (PW)
1969:
1970: \noindent Page, D.~N., \& Thorne, K.~S. 1974, ApJ, 191, 499
1971:
1972: \noindent Perez, C.~A., Silbergleit, A.~S., Wagoner, R.~V., \&
1973: Lehr, D.~E. 1997, ApJ, 476, 589
1974:
1975: \noindent Reynolds, C.~S., \& Fabian, A.~C., 2008, ApJ, in press
1976:
1977: \noindent Rezzolla, L., Yoshida, S.'i., Maccarone, T.~J., \& Zanotti, O.
1978: 2003a, MNRAS, 344, L37
1979:
1980: \noindent Rezzolla, L., Yoshida, S.'i., \& Zanotti, O. 2003b, MNRAS, 344, 978
1981:
1982: \noindent Schnittman, J.~D., Krolik, J.~H., \& Hawley, J.~F.
1983: 2006, ApJ, 651, 1031 (SKH)
1984:
1985: \noindent Shakura N.I., Sunyaev R.A., 1973, A\&A, 24, 337
1986:
1987: \noindent Shafee R., McKinney J.C., Narayan R., Tchekhovkoy A., Gammie
1988: C.F., McClintock J.E., 2008, ApJL, in press (arXive:0808.0860)
1989:
1990: \noindent Silbergleit, A.~S., Wagoner, R.~V., \& Ortega-Rodriguez, M.
1991: 2001, ApJ, 548, 335
1992:
1993: \noindent Stone, J.~M., \& Norman, M.~L. 1992a, ApJS, 80, 753
1994:
1995: \noindent Stone, J.~M., \& Norman, M.~L. 1992b, ApJS, 80, 791
1996:
1997: \noindent Strohmayer, T.~E. 2001, ApJ, 554, L169
1998:
1999: \noindent Uttley, P., \& McHardy, I.~M. 2001, MNRAS, 323, L26
2000:
2001: \noindent Uttley, P., McHardy, I.~M., \& Vaughan, S. 2005, MNRAS, 359, 345
2002:
2003: \noindent Vaughan, S., \& Uttley, P. 2005, MNRAS, 362, 235
2004:
2005: \noindent Vernaleo, J.~C., \& Reynolds, C.~S. 2006, ApJ, 645, 83
2006:
2007: \noindent Wagoner, R.~V., Silbergleit, A.~S., \& Ortega-Rodriguez, M.
2008: 2001, ApJ, 559, L25
2009:
2010:
2011: \appendix
2012:
2013: \section{Fundamental g-mode frequency for very small sound speeds}
2014:
2015: For very small sound speeds, there are simplifications that allow the
2016: frequency of the fundamental trapped g-mode in a hydrodynamic
2017: accretion disk to be obtained analytically. Here we base our
2018: analysis on the equations and formalism of NW92.
2019:
2020: NW92 examine the linearized equations describing the behavior of the
2021: scalar potential
2022: \begin{equation}
2023: \delta u\equiv \delta P/\rho,
2024: \end{equation}
2025: where $\delta P$ is the Eulerian variation in the pressure. The
2026: radial equation for the perturbation is
2027: \begin{equation}
2028: \omega^2 c_s^2 {\partial^2\delta u\over{\partial r^2}}=
2029: -\left(\omega^2-\gamma\Upsilon\Omega^2\right)
2030: \left(\omega^2-\kappa^2\right)\delta u\; ,
2031: \label{eq:diff_eqn}
2032: \end{equation}
2033: where $\omega$ is the mode angular frequency, $\Omega$ is the orbital
2034: angular frequency, $\kappa$ is the radial epicyclic angular frequency,
2035: and $\gamma$ is the usual adiabatic index $\gamma=5/3$. $\Upsilon$ is
2036: determined by the quantization condition
2037: \begin{equation}
2038: {A+1/2\over{(1-4B)^{1/2}}}=j+1/2\; ,
2039: \label{eq:quantization}
2040: \end{equation}
2041: where $A=\Upsilon-\zeta$ and
2042: \begin{equation}
2043: B={\zeta\over\gamma}{(\omega^2-
2044: \gamma\Upsilon\Omega^2)\over{\omega^2}}
2045: \end{equation}
2046: with $\zeta=(\gamma-1)/\gamma= 2/5$. For the fundamental mode, we
2047: have $j=0$ and eq.~\ref{eq:quantization} can be solved to obtain
2048: \begin{equation}
2049: \Upsilon=0.4{\Omega^2\over \omega^2}-0.2.
2050: \end{equation}
2051:
2052: Now we concentrate on low sound speeds, $c_s\ll 1$ (within this
2053: Appendix, speeds will be given in units of $c$; frequencies in units
2054: of $c^3/(GM)$; lengths in units of $r_g$). This implies that the
2055: mode frequency $\omega\approx\kappa_{\rm max}$, where $\kappa_{\rm
2056: max}$ is the maximum radial epicyclic frequency. In this limit, we
2057: note that the factor ($\omega^2-\gamma\Upsilon\Omega^2$) within
2058: eqn.~\ref{eq:diff_eqn} is close to constant with radius. Let us
2059: define $D\equiv -(\omega^2-\gamma\Upsilon\Omega^2)>0$, where we
2060: evaluate $D$ by assuming $\omega=\kappa_{\rm max}$.
2061:
2062: We also note that near the maximum, the radial epicyclic
2063: frequency has a parabolic form, $\kappa^2=\kappa_{\rm max}^2-
2064: E(r-r_{\rm max})^2$, where $r_{\rm max}$ is the radius at which
2065: $\kappa=\kappa_{\rm max}$. The differential equation then becomes
2066: \begin{equation}
2067: {\partial^2\delta u\over{\partial r^2}}=\left[
2068: -{D\over c_s^2}\left(\kappa_{\rm max}^2/\omega^2-1\right)
2069: +{D\over c_s^2}{E\over\omega^2}\left(r-r_{\rm max}\right)^2\right]\delta u\; .
2070: \end{equation}
2071: This has the form of a harmonic oscillator, so we try a solution
2072: of the type
2073: \begin{equation}
2074: \delta u\propto e^{-{1\over 2}C(r-r_{\rm max})^2}
2075: \end{equation}
2076: meaning that
2077: \begin{equation}
2078: {\partial^2\delta u\over{\partial r^2}}=
2079: \left[C^2(r-r_{\rm max})^2-C\right]e^{-{1\over 2}C(r-r_{\rm max})^2}\; .
2080: \end{equation}
2081: This yields the conditions
2082: \begin{eqnarray}
2083: C&=&{D\over c_s^2}\left(\kappa^2_{\rm max}/\omega^2-1\right)\\
2084: C^2&=&{D\over c_s^2}{E\over\omega^2}\; .\\
2085: \end{eqnarray}
2086: Defining $x\equiv 1/\omega^2$, these two conditions can be combined
2087: (eliminating $C$) to yield
2088: \begin{equation}
2089: {D^2\over c_s^4}\left(\kappa_{\rm max}^4x^2-2\kappa_{\rm max}^2x+1
2090: \right)={D\over c_s^2}Ex\; .
2091: \end{equation}
2092: Solving for $x$ gives
2093: \begin{equation}
2094: x={1\over{2\kappa_{\rm max}^4}}\left[2\kappa_{\rm max}^2+c_s^2E/D
2095: \pm\sqrt{4\kappa_{\rm max}^2c_s^2E/D+c_s^4E^2/D^2}\right]\; .
2096: \end{equation}
2097: Since $\omega<\kappa_{\rm max}$, we choose the positive sign.
2098: For small $c_s$, the first term in the square root dominates,
2099: and the lowest order in $c_s$ gives
2100: \begin{equation}
2101: x\approx {1\over{\kappa_{\rm max}^2}}\left(1+{c_s\over{\kappa_{\rm max}}}
2102: \sqrt{E\over D}\right)\; .
2103: \end{equation}
2104: This implies finally that
2105: \begin{equation}
2106: {\kappa_{\rm max}-\omega\over{\kappa_{\rm max}}}=
2107: {c_s\over{2\kappa_{\rm max}}}\sqrt{E/D}\; .
2108: \end{equation}
2109: This clearly demonstrates that the fractional difference of the mode
2110: frequency from the radial epicyclic maximum is linear in $c_s$.
2111:
2112: Evaluating this expression for the PW potential, we obtain
2113: $\kappa^2_{\rm max}=1.202\times 10^{-3}$
2114: (at $r=r_{\rm max}=4+2\sqrt{3}=7.464$), and
2115: \begin{equation}
2116: {\kappa_{\rm max}-\omega\over{\kappa_{\rm max}}}=2.269c_s\; .
2117: \end{equation}
2118: In contrast, the Nowak \& Wagoner pseudo-Newtonian potential,
2119: \begin{equation}
2120: \Phi_{\rm NW}=-{1\over r}+{3\over r^2}-{12\over r^3}\; .
2121: \end{equation}
2122: yields $\kappa^2_{\rm max}=8.607\times 10^{-4}$ (at $r_{\rm max}=7.746$) and
2123: \begin{equation}
2124: {\kappa_{\rm max}-\omega\over{\kappa_{\rm max}}}=5.621c_s\; .
2125: \end{equation}
2126: For both potentials, the analytic frequencies agree extremely well
2127: with direct numerical solutions to eqn.~\ref{eq:diff_eqn}.
2128:
2129: \end{document}
2130:
2131: