1: \documentclass[apj]{emulateapj}
2: %\documentclass[12pt,preprint]{aastex}
3: \usepackage{amsmath}
4: \bibliographystyle{apj}
5:
6: \newcommand \p{\partial}
7: \newcommand \lya{Ly$\alpha$ }
8: \def\dim#1{\mbox{\,#1}}
9: \newcommand \chimps{\ h^{-1} \dim{Mpc}}
10:
11: \begin{document}
12:
13: \title{Resolving Gas Dynamics in the Circumnuclear Region of a Disk Galaxy
14: in a Cosmological Simulation}
15:
16: \author{Robyn Levine\altaffilmark{1,2,3}}
17: \author{Nickolay Y. Gnedin\altaffilmark{3,4,5}}
18: \author{Andrew J. S. Hamilton\altaffilmark{1,2}}
19: \author{Andrey V. Kravtsov\altaffilmark{4,5,6}}
20: \altaffiltext{1}{JILA, University of Colorado, Boulder, CO 80309;
21: robyn.levine@colorado.edu}
22: \altaffiltext{2}{Department of Astrophysical \& Planetary Sciences,
23: University of Colorado, Boulder, CO 80309, USA}
24: \altaffiltext{3}{Particle Astrophysics Center, Fermi National
25: Accelerator Laboratory, Batavia, IL 60510, USA}
26: \altaffiltext{4}{Kavli Institute for Cosmological Physics, The
27: University of Chicago, Chicago, IL 60637, USA}
28: \altaffiltext{5}{Department of Astronomy \& Astrophysics, The
29: University of Chicago, Chicago, IL 60637, USA}
30: \altaffiltext{6}{Enrico Fermi Institute, The University of Chicago,
31: Chicago, IL 60637, USA}
32:
33: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
34:
35: \begin{abstract}
36: Using a hydrodynamic adaptive mesh refinement code, we simulate
37: the growth and evolution of a galaxy, which could potentially host
38: a supermassive black hole, within a cosmological volume. Reaching
39: a dynamical range in excess of 10 million, the simulation follows
40: the evolution of the gas structure from super-galactic scales all
41: the way down to the outer edge of the accretion disk. Here, we
42: focus on global instabilities in the self-gravitating, cold,
43: turbulence-supported, molecular gas disk at the center of the
44: model galaxy, which provide a natural mechanism for angular
45: momentum transport down to sub-pc scales. The gas density profile
46: follows a power-law $\propto r^{-8/3}$, consistent with an
47: analytic description of turbulence in a quasi-stationary
48: circumnuclear disk. We analyze the properties of the disk which
49: contribute to the instabilities, and investigate the significance
50: of instability for the galaxy's evolution and the growth of a
51: supermassive black hole at the center.
52: \end{abstract}
53:
54: \keywords{galaxies: evolution---galaxies: high-redshift---galaxies:
55: ISM---galaxies: nuclei---galaxies: structure }
56:
57:
58: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
59:
60: \section{INTRODUCTION AND MOTIVATION}
61:
62: It is increasingly evident that the growth histories of different
63: components of galaxies: stars, gas, and supermassive black holes,
64: are intricately connected. Observations indicate a relationship
65: between the masses of supermassive black holes (SMBHs) and various
66: properties of their host galaxies, such as the spheroid mass
67: \citep{Magorrian98} and the velocity dispersion of stars in the
68: bulge \citep{FM00, Geb00, Trem02}. Recent simulations have shown
69: that feedback from accreting SMBHs can regulate the growth of the
70: black holes as well as the evolution of their host galaxies, making
71: feedback a potentially important piece of the theory of SMBH-host
72: galaxy co-evolution \citep[e.g.][]{DiMatteo05, Springel05,
73: DiMatteo07, Sijacki07}.
74:
75: Building up the mass of a galaxy and driving AGN feedback requires a
76: continuous replenishment of fuel in the center of the galaxy. A
77: comprehensive understanding of such fueling is not possible without
78: detailed knowledge about matter transport from large scales to the
79: vicinity of the black hole. That transport helps determine, in
80: particular, both the amount of material available for accretion,
81: feedback, and star formation, as well as the strength and duration
82: of the fueling. Large scale gravitational tidal fields created
83: during major mergers of galaxies are thought to be effective at
84: funneling matter toward the centers of galaxies. Indeed, simulations
85: and semi-analytic modeling of hierarchical growth scenarios have
86: shown that a combination of accretion and black hole mergers can
87: effectively build-up the local black hole population
88: \citep{KaufHae00,YooME04,VolRees05,Malbonetal06,VolRees06,Lietal07},
89: as well as assemble the $\sim 10^9$ M$_{\sun}$ black holes already
90: observed to be present in quasars at $z\approx6$
91: \citep{Fan03,Lietal07}.
92:
93: Secular evolution of galaxies, particularly in the absence of major
94: merger events, can drive fuel down to the center as well. Global bar
95: instabilities are thought to be efficient at transporting angular
96: momentum, allowing material in the disk to move toward the center of
97: the galaxy
98: \citep{Robertsetal79,Simkinetal80,Noguchi88,barswinbars,KormKenn04,RegTeu04}.
99: In the ``bars within bars'' scenario of \citet{barswinbars}, a large
100: scale galactic bar drives gas inward where it forms a
101: self-gravitating disk. As this disk becomes unstable, a smaller
102: secondary bar forms, driving material down to scales of order $10$
103: pc, at which point other physics can take over and transport the
104: material the rest of the way toward the black hole
105: \citep[e.g.][]{ShlosRev90}.
106:
107: Given the complexity of transporting matter from cosmological scales
108: all the way down to the circumnuclear region of a galaxy, and
109: ultimately to the vicinity of a SMBH, it is clear that a thorough
110: understanding of the relationship between SMBHs and galaxy formation
111: requires modeling of a wide range of scales. Recently, cosmological
112: SPH simulations have been combined with smaller scale simulations of
113: a SMBH host galaxy environment to study the effects of merger driven
114: fueling and AGN feedback on supermassive black hole growth and
115: demography \citep{DiMatteo07, Sijacki07}. Reaching a large dynamic
116: range in N-body+SPH simulations, \citet{Mayeretal07} have studied
117: galaxy dynamics during mergers resulting in a supermassive black
118: hole binary. Small scale simulations have addressed gas dynamics in
119: sub-galactic scale disks with high resolution \citep{Fukuda00,
120: Wada01, WadaNorman01,Escala07,WadaNorman07}, finding the
121: development of a turbulent, multi-phase interstellar medium (ISM).
122:
123: The study we present here follows the evolution of the circumnuclear
124: region in a typical galaxy environment rather than a dramatic, rare
125: event such as a major merger, in a self-consistent cosmological
126: simulation. We have specifically chosen a simulated galaxy that will
127: evolve into a typical $L_*$ galaxy at $z=0$. The simulation follows
128: the galaxy during a phase of its evolution in which it has not
129: undergone any major mergers within several dynamical time scales, in
130: order to follow the development of instabilities in the
131: circumnuclear disk, which may drive the transport of matter and
132: angular momentum from large to small scales. The cosmological
133: simulations use the adaptive mesh refinement (AMR) technique to
134: self-consistently model the gas dynamics in a single galaxy at high
135: resolution (sub-pc resolution in the center of the galaxy). A large
136: dynamic range ($> 10$ million), achievable with AMR technique,
137: allows us to bridge cosmological scales to scales relevant for
138: molecular cloud formation (the birthplace of stars) and AGN
139: fueling. After studying the physics in this basic model galaxy, we
140: can begin to include physical processes that are directly relevant
141: to the problem of SMBH growth in the context of galaxy evolution,
142: such as AGN feedback.
143:
144: It is a complex task to implement mergers, feedback and secular
145: evolution in large, cosmological simulations all at once. Our
146: approach is to split the problem into pieces to be addressed one at
147: a time, ultimately building a more realistic simulation. While the
148: present paper focuses on the structure of the galaxy, and the
149: development of instabilities which may be responsible for driving
150: matter and angular momentum transport within the galaxy, we plan to
151: explore the time evolution of such transport, specifically the
152: evolution of the accretion rates of mass and angular momentum, in a
153: subsequent paper. The nature of the circumnuclear region and a
154: quantitative analysis of accretion rates each have important
155: consequences for galaxy evolution.
156:
157: The organization of the paper is as follows. In Section
158: \ref{sec:sim}, we describe the details of the simulation and the
159: adopted ``zoom-in'' method. In Section \ref{sec:res}, we describe
160: the features of the highly-resolved galaxy in a single zoom-in
161: episode at $z=4$, including a stability analysis of the disk and the
162: potential role of physics not included in the current
163: simulation. Section \ref{sec:miss} examines the potential role of
164: physics missing from the simulations. Finally, Section
165: \ref{sec:disc} contains a summary of our conclusions and their
166: interpretation.
167:
168: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
169:
170: \section{SIMULATION}
171: \label{sec:sim}
172:
173: Sections \ref{subsection:art} and \ref{subsection:ref} briefly
174: explain details of the simulation and of our methods. Additional
175: details regarding tests of angular momentum conservation, and the
176: treatment of dark matter particle discreteness effects will be
177: available in the PhD thesis of the author Levine (also see the
178: discussion of angular momentum conservation in Appendix
179: \ref{sec:con}).
180:
181: \subsection{Simulation Code}
182: \label{subsection:art}
183:
184: It is unfeasible to follow the evolution of a galaxy all the way
185: down to the scale of a black hole hole accretion disk over
186: cosmological times. Such a simulation spans a large range of scales
187: (with a dynamic range $> 10^7$), and each spatial scale has a
188: different temporal scale. As a result, the Courant condition for
189: numerical stability requires very small time steps in the most
190: highly refined regions, making it computationally expensive to
191: follow the evolution of the galaxy over long periods of time. It is
192: much more feasible to start with a lower resolution cosmological
193: simulation, and zoom in to the small-scale region with increasingly
194: smaller cell sizes and time steps. As the simulation evolves on
195: small scales, the large scale portion of the simulation does not
196: undergo much evolution, and does not need as high resolution.
197:
198: The cosmological simulations presented in this paper are conducted
199: with the Adaptive Refinement Tree (ART) code \citep{Kravtsovetal97,
200: KravtsovPhD, Kravtsovetal02}. The ART code includes the technique of
201: adaptive mesh refinement (AMR), allowing high resolution of a galaxy
202: residing in a small region of the cosmological simulation, while
203: following the rest of the simulated region with lower
204: resolution. The technique is appropriate for studying the structure
205: and evolution of a galaxy over a large dynamical range and in a
206: cosmological context.
207:
208: The ART code includes a range of physics for modeling dark matter,
209: stars, and gas dynamics. Gas is converted into stars in cells with
210: densities greater than $\rho_{\textrm{SF}}$ and temperatures less
211: than $T_{\textrm{SF}}$, where $\rho_{\textrm{SF}} = 1.64
212: \dim{M}_{\sun} \dim{pc}^{-3}$ and $T_{\textrm{SF}} = 9000$ K
213: \citep[see][for more details]{Kravtsov03}, resulting in a star
214: formation efficiency consistent with a Kennicutt law on kiloparsec
215: scales \citep{Kennicutt98} and with observations on $100$ pc scales
216: as well \citep[e.g.][]{Youngetal96, WongBlitz02}. The code follows
217: ISM physics, such as molecular hydrogen formation, and gas cooling
218: by heavy elements and dust under the assumption of collisional
219: ionization equilibrium. The cooling and heating rates are tabulated
220: as functions of gas density, temperature, metallicity, and redshift
221: over the temperature range $10^2<T<10^9$ K using CLOUDY
222: \citep{Cloudy98}, which accounts for the metallicity of the gas, and
223: formation of molecular hydrogen and cosmic dust. In future studies,
224: we plan to include the ART code's radiative transfer capabilities,
225: which will be necessary for implementing AGN feedback.
226:
227:
228: %%%%%%%%%%%%%%%%%%%%%%%%%%%
229:
230: \subsection{``Zooming-In'' to the Center of a Galaxy}
231: \label{subsection:ref}
232:
233: We begin with a cosmological simulation, evolved from a realization
234: of a random Gaussian density field at $z=50$, with periodic boundary
235: conditions, measuring $6 h^{-1}$ comoving Mpc across. The simulation
236: was first run with low-resolution in order to select a galactic-mass
237: halo for subsequent study. A Lagrangian region the size of five
238: virial radii of the halo at $z=0$ was then identified at $z=50$, and
239: re-sampled and run with higher resolution to $z=2.8$ \citep[see][for
240: a description of the technique]{Klypinetal01}. The Lagrangian region
241: of the simulation was automatically refined three levels, as a
242: minimum. There are $2.64 \times 10^6$ dark matter particles in the
243: Lagrangian region, each with mass $9.18 \times 10^5 h^{-1}
244: \dim{M}_{\sun}$. Subsequent refinement and de-refinement in this
245: region followed a dark matter mass criterion, in which a cell was
246: refined if its total dark matter mass is greater than $1.8 \times
247: 10^6 h^{-1} \dim{M}_{\sun}$. At $z=4$, the highest matter density
248: peaks in the simulation have a maximum resolution of $\approx 50$ pc
249: (in physical units; corresponding to $9$ levels of refinement on top
250: of a $64^3$ root grid). The largest halo has a total mass of $2
251: \times 10^{11} \dim{M}_{\sun}$ at $z=4$, and it contains the
252: progenitor of an $L^*$ spiral galaxy. This initial cosmological
253: simulation was run including metal enrichment and energy feedback
254: from supernova, and radiative transfer in addition to the physics
255: described in Section \ref{subsection:art}.
256:
257: %\clearpage
258: \begin{figure*}
259: \center{\includegraphics[width=1.9\columnwidth]{f1.eps}}
260: \caption{\label{fig:pan} Volume rendering of the $3$-dimensional
261: gas density on several different scales at $z=4$, from $200$ kpc
262: in the upper left panel, to $2$ pc in the lower right panel. The
263: color bar shows the density scale in units of
264: $\dim{cm}^{-3}$. Note the presence of the spiral arms and
265: instabilities on a wide range of scales. (This figure is best
266: viewed in color.)}
267: \end{figure*}
268: %\clearpage
269:
270: This particular $z=4$ cosmological simulation was chosen for this
271: study because it contains a galaxy that has not been disturbed by a
272: major merger since it merged with a galaxy with $25\%$ its mass at
273: $z\approx6$. The time since this dynamically active period ($\approx
274: 600$ million years) is significantly longer than the dynamical time of
275: the galactic disk at $z\approx4$ ($\approx 15$ million years). The
276: galaxy is, however, far from quiescent, having grown in mass by
277: $\approx 25\%$ from $z=5$ to $z=4$ via minor mergers and accretion
278: from its environment.
279:
280: %\clearpage
281: \begin{figure*}
282: \center{\includegraphics[width=1.9\columnwidth]{f2.eps}}
283: \caption{\label{fig:star} The stellar component of the galaxy at
284: $z=4$ on a scale of $3$ kpc. The stars in the image are divided
285: into two populations, with yellow stars corresponding to the oldest
286: stars in the galaxy, and blue stars corresponding to the youngest
287: stars. The youngest stars are found primarily in the disk of the
288: galaxy. (This figure is best viewed in color.)}
289: \end{figure*}
290: %\clearpage
291:
292: In the cosmological simulation, we allow further refinement (at $z
293: \approx 4$), one level at a time, in a $1.5$ kpc region centered on
294: the galaxy, effectively zooming in to the center of the galaxy with
295: increasing resolution. The ``zoom-in'' technique is similar to the
296: one used by \citet{Abeletal02} in simulations of the formation of
297: the first star, which required an even larger dynamical
298: range. Because we are starting simply in this first pilot study, and
299: building a more realistic simulation in pieces, radiative transfer
300: and stellar feedback have been switched off in the zoom-in
301: portion of the cosmological simulation (they will be examined in
302: future studies). The high-resolution portion of the simulation
303: employs additional refinement criteria, refining according to a
304: level-dependent mass criterion on levels $11$ and below\footnote{It
305: should be noted that we adopt the convention of referring to level
306: $0$ as the ``top'' level, and the maximum (most-refined) level as
307: the ``bottom'' level, so that the terms ``above'' and ``below''
308: refer to lower and higher resolution, respectively.} in the zoom-in
309: region. The refinement criterion is defined so that the finer the
310: resolution, the more aggressively the mesh refines, ensuring that
311: there are enough highly refined cells to resolve structures on small
312: scales. Specifically, the mesh refinement is super-Lagrangian in the
313: circumnuclear region of the galaxy, and cells are marked for
314: refinement if the gas mass in a cell is
315: $m_{\textrm{r}}^{\textrm{level}-10}$ times the Lagrangian mass
316: criterion ($9 \times 10^7 h^{-1} \dim{M}_{\sun}$), where
317: $m_{\textrm{r}}=0.7$ is our fiducial value. The factor $0.7$ was
318: chosen through experimentation, so as to make the refinement in the
319: central region more aggressive, but not so aggressive that the
320: simulation becomes too computationally expensive.
321:
322: During the zoom-in period, we increase the maximum level of
323: refinement gradually, one level at a time, allowing the simulation
324: to reach a quasi-stationary state on each level before moving to the
325: next. The slow initial refinement allows the ART code to resolve
326: spatial scales within the simulated galaxy while avoiding transient
327: effects. Since the time steps depend on the sound speed of the gas,
328: they are therefore related to the dynamical time, and a fixed
329: minimum number of time steps on each level forces the mesh to evolve
330: for several dynamical times. We have run parallel simulations with
331: $100$, $300$, and $1000$ minimum steps on each level, and have
332: determined that a requirement of a minimum of $300$ on each level
333: sufficiently reduces transient numerical effects that result from
334: refining too quickly.
335:
336: Throughout the initial zoom-in, we take precautions to avoid
337: numerical artifacts. As the resolution of the simulation increases
338: and throughout its subsequent evolution, the density of dark matter
339: is smoothed on scales corresponding to level $9$ resolution and
340: above. This ensures that each cell contains $\approx 15-20$ dark
341: matter particles, avoiding effects resulting from their
342: discreteness. It is well known that poor resolution in
343: hydrodynamics codes can lead to artificial fragmentation in the
344: gas. \citet{Truelove97} investigated these numerical instabilities
345: and found that a simulation must resolve the Jeans length to avoid
346: artificial fragmentation. In addition to the mass criteria described
347: above, the ART code's refinement scheme meets the Jeans condition of
348: \citet{Truelove97} on the maximum level of refinement (which
349: increases gradually during the initial zoom-in episode), requiring
350: that $\Delta x/\lambda_{\textrm{J}} < 0.25$, where $\Delta x$ is the
351: resolution, or cell size, and $\lambda_{\textrm{J}}$ is the Jeans
352: length. Additionally, as the mesh refines, rapidly collapsing
353: structures can cause numerical instabilities at steep density
354: gradients. The ART code implements artificial pressure support (as
355: described in \citet{Machacek01}) on the maximum level of refinement,
356: in order to avoid the over-collapsing of structures.
357:
358: After the initial zoom-in, the highest level cells are $0.03$ pc
359: across (corresponding to $20$ levels of refinement), allowing us to
360: determine the location for a supermassive black hole test particle
361: with high precision, but without resolving the black hole accretion
362: disk, which would require additional physics (such as
363: magneto-hydrodynamics, or MHD) not currently included in the ART
364: code. Our results extend reliably down to a scale of at least $0.12$
365: pc, or the size of four level $20$ cells, which we consider to be
366: our resolution limit. After reaching this maximum resolution (level
367: $20$), we replace $3\times10^7 \dim{M}_{\sun}$ of the gas from the
368: center of the simulated galaxy with a black hole point mass of equal
369: mass and momentum. At present, we focus on the dynamical properties
370: of the circumnuclear disk on scales where the gravity of the gas
371: dominates that of the black hole. The addition of the black hole
372: point mass here simply allows us to follow the black hole's location
373: and velocity as a reference point. However the successful
374: introduction of the black hole particle will be essential in
375: subsequent simulations involving physics associated with the black
376: hole (such as AGN feedback).
377:
378: After the initial zoom-in and the successful introduction of the
379: black hole test particle, the simulation then evolves at high
380: resolution for approximately one dynamical time at the $100
381: \dim{pc}$ scale, showing a highly resolved galactic disk. For a
382: quasi-Keplerian disk, the number of orbital periods,
383: $N_{\textrm{orb}}$, undergone by the simulation at radii less than
384: $100 \dim{pc}$, is $(R/100\dim{pc})^{-3/2}$. Figure \ref{fig:pan}
385: shows the $3$-dimensional gas density in an approximately face-on
386: view of the galactic disk, on several different spatial scales, for
387: a single time step, demonstrating the large dynamic range of the
388: simulation. The distribution of stars is shown in Figure
389: \ref{fig:star}, in both a face-on and an edge-on view of the galaxy
390: on a scale of $3$ kpc. The two different colors correspond to old
391: and young stars (yellow and blue, respectively), with the youngest
392: stars in the galaxy located primarily in the disk, and the oldest
393: stars distributed more isotropically, populating the galaxy's bulge.
394:
395: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
396:
397: \section{RESULTS}
398: \label{sec:res}
399: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
400: \subsection{Structure of the Circumnuclear Disk}
401: \label{subsection:disk}
402:
403: Volume rendered images of the gas density in the $z=4$ simulation,
404: such as those in Figure \ref{fig:pan}, clearly show the spiral disk
405: structure of the galaxy. The gas disk extends inward all the way to
406: sub-pc scales, which are the smallest scales resolved in our
407: simulation\footnote{In particular, the simulated disk does not
408: contain the commonly observed toroidal structure in the inner few pc
409: of the galaxy. The failure of our simulation to reproduce the AGN
410: torus may be the result of additional physics still missing in the
411: simulation, as we will discuss in \S \ref{subsection:torus}.}. The
412: geometric structure of the circumnuclear region is best illustrated
413: by the eigenvalues of the inertia tensor, calculated in spherical
414: shells around the black hole. The inertia tensor is given by
415:
416: \begin{equation}\label{eq:iner}
417: I_{jk} = \sum_{\alpha=1}^N m_{\alpha}x_j^{\alpha}x_k^{\alpha},
418: \end{equation}
419:
420: \noindent where $m_{\alpha}$ is the gas mass of cell $\alpha$ at radius
421: $r_{\alpha} = |\bf{x}_{\alpha}|$ inside a spherical shell containing
422: $N$ cells, and centered on the black hole. The eigenvalues of the
423: inertia tensor define the principal axes of each shell, and their
424: ratios describe the shape. Figure \ref{fig:shape} shows the average
425: radial profile of the ratios of the inertia tensor's
426: eigenvalues. Above $0.1$ pc, $c$ is significantly smaller than $b$
427: over several orders of magnitude in radius, indicating a disk
428: structure. However, the fact that $b/a$ is significantly less than
429: one in parts of the disk indicates that the disk is not axially
430: symmetric, and that it has large-scale structures, such as spiral
431: waves and bars.
432:
433: %\clearpage
434: \begin{figure}[t]
435: % \centering \epsscale{1.}
436: \centering \epsscale{1.2}
437: \plotone{f3.eps}
438: \caption{\label{fig:shape} Ratios of eigenvalues of the inertia
439: tensor in spherical shells centered on the black hole. The ratio
440: $c/a \ll b/a$, indicating a disk structure.}
441: \end{figure}
442: %\clearpage
443:
444: The gas disk in the simulation is fully rotationally supported, with
445: the tangential component of velocity dominating over the radial
446: component by at least an order of magnitude, as is illustrated in
447: Figure \ref{fig:vel}. The top panel of Figure \ref{fig:vel} shows
448: the ratio of the radial to the tangential component of velocity,
449: which is small throughout much of the disk, as the motion of the gas
450: is almost entirely rotational. The average radial velocity is
451: negative, indicating the inflow of gas. The bottom panel shows the
452: tangential component of velocity in units of a quasi-Keplerian
453: velocity, $\sqrt{GM(r)/r}$, determined by the interior total mass,
454: $M(r)$, at each radius. The velocity is close to being Keplerian,
455: but since the mass distribution in the disk resembles a flattened
456: ellipsoid, rather than a spherical distribution, the rotation is
457: slightly super-Keplerian.
458:
459: %\clearpage
460: \begin{figure}[t]
461: % \centering \epsscale{1.}
462: \centering \epsscale{1}
463: \plotone{f4.eps}
464: \caption{\label{fig:vel} {\it Top:} Ratio of radial and tangential
465: velocity components of the gas, each averaged over $550{,}000$
466: years. {\it Bottom:} Average tangential velocity component in units
467: of the average quasi-Keplerian velocity.}
468: \end{figure}
469: %\clearpage
470:
471: A remarkable feature of the circumnuclear disk is that the average
472: gas density profile, measured within spherical shells centered on
473: the black hole, follows an almost perfect power-law with little
474: evolution in time. Figure \ref{fig:dens} shows the gas density
475: profile from two different snapshots of the simulation, as well as
476: the average of the profile over a $\approx 550{,}000$ year
477: period. Both the snapshots and the average profiles of the gas
478: density increase by $\approx 8$ orders of magnitude in the inner
479: $100$ pc of the simulation, obeying a steep power-law with slope
480: $-8/3$. The stability of the gas density profile indicates that the
481: disk is in a quasi-stationary state on timescales of several hundred
482: thousand years. Figure \ref{fig:dens} also shows the dark matter and
483: stellar mass density profiles. In the central kiloparsec of the
484: galaxy, gas comprises $\sim 62 \%$ of the mass ($\sim 2.3 \times
485: 10^{10} h^{-1} \dim{M}_{\sun}$ total in the central kpc), dominating
486: the dynamics of the disk. In contrast, the stellar and dark matter
487: populations comprise $\sim 13$ and $\sim 25 \%$ of the galaxy's mass
488: inside the central kiloparsec. The disk is extremely gas rich at
489: $z=4$, because the galaxy is still actively growing and has not yet
490: formed all of its stars. Interestingly, we find that both the
491: stellar and dark matter profiles measured in the central $\sim 200$
492: pc of the galaxy follow simple $\propto r^{-2}$ power-laws. The
493: profiles match those predicted for the adiabatic contraction of dark
494: matter from an NFW profile \citep{NFW97}, using the model of
495: \citet{Gnedinetal04}, all the way down to the $1\dim{pc}$ scale.
496:
497: %\clearpage
498: \begin{figure}[t]
499: % \centering \epsscale{1.}
500: \centering \epsscale{1.2}
501: \plotone{f5.eps}
502: \caption{\label{fig:dens} Radial profiles of the gas, dark matter,
503: and stellar mass density. Snapshots of the gas density are shown
504: at $100{,}000$ and $500{,}000$ years after the initial
505: refinement ({\it thin solid lines; green and red respectively,
506: in the color version}), in addition to a $550{,}000$ year
507: average ({\it thick solid line}) . The snapshots do not show
508: much variation from the averaged profile, demonstrating that on
509: timescales of several hundred thousand years, the disk is in a
510: quasi-stationary state. The dotted line shows a power-law with
511: slope $-8/3$ for comparison, which matches the gas density slope
512: well. The dark matter and stellar mass density profiles are
513: shown by the long-dashed ({\it blue in the color version}) and
514: short-dashed ({\it magenta in the color version}) lines,
515: respectively. In the inner $\sim 200$ pc, the slope of the dark
516: matter and stellar density profiles is $\sim -2$ consistent with
517: the adiabatic contraction model.}
518:
519: \end{figure}
520: %\clearpage
521:
522: Figure \ref{fig:vcmp} shows that the RMS velocity dispersion of the
523: gas (measured between neighboring cells, and therefore depending on
524: the resolution) greatly exceeds the sound speed in the inner $100$
525: pc, indicating supersonic turbulence in the disk. Supersonic
526: turbulence decays on a dynamical time scale, so the persistence of
527: the turbulence in the circumnuclear disk of the galaxy indicates a
528: driving mechanism, which will be addressed in the following
529: sections. The mean sound speed of the gas indicates a cold,
530: molecular gas disk within the central $100 \dim{pc}$ of the
531: simulated galaxy. The floor in the sound speed shown in Figure
532: \ref{fig:vcmp} is determined by the minimum temperature to which our
533: adopted cooling rates (computed using the CLOUDY package) apply. The
534: mean sound speed is computed for cells with $T<20{,}000$ K, because
535: the spherical averages shown in Figure \ref{fig:vcmp} include the
536: low density gas, infalling perpendicular to the plane of the
537: galactic disk, and shock-heated to high temperatures ($\gtrsim 10^6$
538: K). Additionally, turbulence inside the disk produces localized
539: shocks that briefly heat individual cells to similarly high
540: temperatures. The inclusion of shock-heated gas produces a broad
541: temperature distribution and raises the mean temperature of the gas
542: so that it does not effectively describe the typical sound speed of
543: the gas in the disk. Therefore, the mean is computed over a
544: temperature range that selects cells in the disk, while excluding
545: atypical cells, resulting in a more representative quantity for
546: describing the thermal properties of the disk.
547:
548: %\clearpage
549: \begin{figure}[t]
550: \centering
551: % \epsscale{1.}
552: \epsscale{1.2}
553: \plotone{f6.eps}
554: \caption{\label{fig:vcmp} A comparison of the radial profiles of
555: the tangential velocity component of the gas ({\it solid}), its
556: RMS velocity dispersion, or turbulent velocity ({\it dashed}),
557: and the mean sound speed, computed in cells with $T<20{,}000$ K
558: ({\it dotted}), time averaged over $550{,}000$ years. The sound
559: speed is low, indicating cold, molecular gas within the central
560: $20 \dim{pc}$.}
561: \end{figure}
562: %\clearpage
563:
564: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
565: \subsection{Angular Momentum Transport}
566: \label{subsection:ang}
567:
568: The gas density profile of Figure \ref{fig:dens} motivates a simple
569: analytic description of angular momentum transport in the
570: disk. Averaged over a sufficiently long time, the disk can be
571: considered to be approximately azimuthally symmetric. Additionally,
572: the rotational velocity of the disk dominates over other velocity
573: components: over the local velocity dispersion, and over the sound
574: speed, as demonstrated in Figure \ref{fig:vcmp}. Therefore, the
575: circumnuclear disk in the simulation is well approximated by a thin,
576: rotationally supported, viscous, molecular gas disk. Conservation of
577: angular momentum in such a disk is described by the following
578: equation in cylindrical coordinates \citep{Pringle81}:
579:
580: \begin{equation}\label{eq:ang}
581: \frac{\p}{\p t}(J_{\textrm{z}}) + \frac{1}{R}\frac{\p}{\p R}(R
582: v_{\textrm{R}} J_z) = \frac{1}{2\pi R}\frac{\p G}{\p R},
583: \end{equation}
584:
585: \noindent where the angular momentum density normal to the disk is
586: given by $J_{\textrm{z}} = \Sigma R^2 \Omega$, where
587: $\Omega=v_{\textrm{t}}/R$. The surface density of the gas is
588: $\Sigma$, and $v_{\textrm{R}}$ is the radial component of velocity,
589: which is small compared to the rotational velocity of the gas. The
590: right hand side of equation (\ref{eq:ang}) describes the effects of
591: a viscous torque, $G$, generated by turbulent motions of the
592: gas. \citet{Pringle81} parameterizes the viscous torque $G$, which
593: necessarily vanishes in solid body rotation, $\p \Omega / \p R = 0$,
594: as
595:
596: \begin{equation}\label{eq:torqa}
597: G(R,t) = 2 \pi \nu \Sigma R^3 \frac{\p \Omega}{\p R}
598: \end{equation}
599:
600: \noindent where $\nu$ is the coefficient of turbulent
601: viscosity. During the initial zoom-in, the gas disk quickly reaches
602: the power-law density profile shown in Figure \ref{fig:dens}. On
603: short time scales of a few hundred thousand years, traced by a
604: single high-resolution zoom-in episode of our simulation,
605: time-averaged radial motions through the disk are small compared to
606: the rotational motion, as shown in Figure \ref{fig:vel}. Over longer
607: time scales, angular momentum is transported outward by a viscous
608: torque driven by the turbulent motions of the gas, allowing
609: accretion to slowly feed the black hole. The left hand side of
610: equation (\ref{eq:ang}) can therefore be considered to be small on
611: time scales of a few hundred thousand years. It then follows that
612: the right hand side of equation (\ref{eq:ang}) should also be small,
613: and that the torque $G$ is approximately constant with radius.
614:
615: Given that the gas density $\rho$ follows a power law (Figure
616: \ref{fig:dens}), it is reasonable to approximate the surface density
617: of the gas $\Sigma$ by a power law with the radius $R$:
618:
619: \begin{equation}\label{eq:surf}
620: \Sigma \propto R^{-\beta}.
621: \end{equation}
622:
623: \noindent Since the gravitational potential $\Phi$ satisfies
624: $\nabla^2 \Phi = 4 \pi G \rho$ and $\Sigma = 2\rho R$, it follows
625: that
626:
627: \begin{equation}\label{eq:phi}
628: \Phi \propto R^{1-\beta}.
629: \end{equation}
630:
631: \noindent Since the disk is almost entirely rotationally supported,
632: the tangential velocity $v_{\textrm{t}}$ is effectively
633: the circular velocity $v_{\textrm{c}}$, defined by $v_{\textrm{c}}^2
634: = R\ d\Phi/dR$. Thus, the angular speed is
635:
636: \begin{equation}
637: \Omega = \frac{1}{R} \left(R \frac{d\Phi}{dR}\right)^{1/2} \propto
638: R^{-(1+\beta)/2},
639: \end{equation}
640:
641: \noindent and the viscous torque
642:
643: \begin{equation}\label{eq:torqb}
644: G \propto R^{3(1-\beta)/2}\nu.
645: \end{equation}
646:
647: \noindent If $\nu \propto R$ (which, as we will show in Section
648: \ref{subsection:stab}, is the case in our simulations), then
649: equation (\ref{eq:torqb}) implies a power-law slope of $\beta
650: \approx 5/3$. It then follows that the volume density of the gas
651: should scale as $\rho \propto R^{-8/3}$, consistent with the slope
652: measured in Figure \ref{fig:dens}. Therefore, the simple description
653: of angular momentum transport, maintained by turbulence in the disk,
654: potentially provides a consistent interpretation of the power-law
655: slope of the gas density profile of Figure \ref{fig:dens}.
656:
657:
658: %%%%%%%%%%%%%%%%%%%%%%%%%%%
659:
660: \subsection{Disk Stability and the Source of Turbulence}
661: \label{subsection:stab}
662:
663: In Section \ref{subsection:disk}, we showed that the simulated
664: galaxy contains a self-gravitating, turbulent, cold, molecular gas
665: disk within the central $\approx 100 \dim{pc}$. Such disks are
666: susceptible to instabilities and to fragmentation. In the snapshots
667: shown in Figure \ref{fig:evol}, the presence of instabilities is
668: illustrated by waves moving through the simulated disk. The
669: snapshots trace the evolution of disk structure on scales of
670: $\approx 125$ pc, and $\approx 12.5$ pc (inset). The rotational
671: period of the disk at the outer radius is such that the disk has
672: undergone at least one full rotation (several more at smaller radii)
673: over the time scales followed by the present simulation. The panels
674: show the somewhat chaotic formation and re-formation of spiral
675: structures on these scales, in contrast with the more ordered spiral
676: structure seen on kiloparsec scales (Figure \ref{fig:pan}). This
677: section includes an analysis of the behavior of structures in the
678: disk on the scales shown in Figure \ref{fig:evol}.
679:
680: %\clearpage
681: \begin{figure*}
682: \center{\includegraphics[width=1.9\columnwidth]{f7.eps}}
683: \caption{\label{fig:evol} The evolution of global instabilities in
684: the disk over time, on a scale of $\approx 125$ pc and on a scale
685: of $\approx 12.5$ pc (inset). Each panel shows a volume rendering
686: of the gas density (in $\dim{cm}^{-3})$ at a different
687: epoch. The similarities between the appearance of the disk on
688: the two different scales indicate the hierarchical structure of
689: the disk. (This figure is best viewed in color.)}
690: \end{figure*}
691: %\clearpage
692:
693: The Toomre $Q$ parameter describes the stability of the disk against
694: linear perturbations, and is given by $Q = \sigma_{\textrm{turb}}
695: \kappa / \pi G\Sigma$, where $\kappa$ is the epicyclic frequency,
696: and $\Sigma$ is the surface density of the gas, both determined
697: locally \citep{Toomre64,GoldLB65}. Although the disk is cold, and
698: the sound speed is low, the RMS velocity dispersion is substantially
699: higher than the sound speed inside $100$ pc, indicating that the gas
700: is turbulent. Therefore, in place of the sound speed typically used
701: in the definition of the Toomre $Q$ parameter, we have substituted a
702: turbulent velocity, given by the quadrature sum of the sound speed
703: and the RMS velocity dispersion of the gas. The disk is, for the
704: most part, quasi-Keplerian (as described in Section
705: \ref{subsection:disk}), so that the epicyclic frequency $\kappa$ is
706: proportional to the angular speed of the disk,
707: $\Omega=v_{\textrm{t}}/r$.
708:
709: The top panel of Figure \ref{fig:Q} shows the Toomre $Q$ parameter
710: for the simulated disk, corresponding to the region shown in Figure
711: \ref{fig:evol}. Where $Q<1$, the disk is susceptible to local
712: gravitational instabilities resulting from axisymmetric
713: perturbations. For $Q>1$ the disk is likely to be stable against
714: axisymmetric perturbations, but is still susceptible to
715: instabilities arising from non-axisymmetric perturbations, which are
716: less stable than radial perturbations \citep{Polyachenko}. For the
717: density profile shown in Figure \ref{fig:dens}, the disk becomes
718: stable for $Q\gtrsim2$. The shaded region in Figure \ref{fig:Q}
719: ($1<Q<2$) therefore represents marginally stable values of the $Q$
720: parameter, where the disk still might become unstable. Figure
721: \ref{fig:Q} suggests that the disk in the simulated galaxy lies
722: mostly in the region of marginal stability inside $\approx
723: 1\dim{kpc}$. In the case of axisymmetric perturbations, the fastest
724: growing unstable mode corresponds to the scale
725: $\lambda_{\textrm{fast,r}}$, at which $d\omega^2 / dk = 0$ (where
726: $\omega$ is the angular frequency of waves in the disk), given by
727:
728: \begin{equation}\label{eq:fastr}
729: \lambda_{\textrm{fast,r}} = \frac{2 \sigma_{\textrm{turb}}^2}{G
730: \Sigma}.
731: \end{equation}
732:
733: \noindent Using the condition for marginal stability from
734: \citet{Polyachenko}, the fastest growing mode for all modes
735: (axisymmetric and non-axisymmetric), $\lambda_{\textrm{fast,all}}$,
736: is given by
737:
738: \begin{equation}\label{eq:fasta}
739: \lambda_{\textrm{fast,all}} = \frac{\lambda_{\textrm{fast,r}}}{2}
740: = \frac{ \sigma_{\textrm{turb}}^2}{G \Sigma}.
741: \end{equation}
742:
743: \noindent The bottom panel of Figure \ref{fig:Q} shows the ratio
744: $\lambda / r$ for each of the fastest growing modes. In the inner
745: $100$ pc, the scales of the fastest growing modes are smaller than
746: the radius by only a factor of a few, implying that the disk is
747: stable on scales $\lambda \ll R$ (at least in the linear
748: regime). This is an indication that perturbations in the disk
749: operate on a range of scales, but always on scales that are an
750: appreciable fraction of the size of the system, driving global
751: instabilities, which generate turbulence on smaller scales. The disk
752: remains locally stable all the way into a region less than $1$ pc
753: from the black hole. There is no catastrophic fragmentation into
754: clumps small enough to form stars, so that angular momentum
755: transport may continue uninterrupted by bursts of star formation.
756:
757: %\clearpage
758: \begin{figure}[t]
759: % \epsscale{1.}
760: \epsscale{1.2}
761: \centering
762: \plotone{f8.eps}
763: \caption{\label{fig:Q} {\it Top:} Toomre $Q$ parameter. For $Q<1$
764: the disk is unstable, and for $Q\gtrsim2$ the disk is stable. In
765: the regime $1<Q<2$, the disk is marginally stable (stable
766: against axisymmetric modes, but not non-axisymmetric
767: modes). {\it Bottom:} Average values of the the fastest growing
768: unstable mode for axisymmetric perturbations,
769: $\lambda_{\textrm{fast,r}}$ ({\it solid}), and the fastest
770: growing unstable mode for all perturbations,
771: $\lambda_{\textrm{fast,all}}$ ({\it dashed}), each divided by
772: radius. The ratio $\lambda/r \gtrsim 0.1$ throughout the
773: circumnuclear disk, so there is no catastrophic fragmentation
774: down to small scales.}
775: \end{figure}
776: %\clearpage
777:
778: By dimensional analysis, the turbulent kinematic viscosity $\nu$
779: discussed in Section \ref{subsection:ang} can be described by
780: $\lambda \sigma_{\textrm{turb}}$, where $\lambda$ is a
781: characteristic scale for turbulence and $\sigma_{\textrm{turb}}$,
782: the turbulent velocity shown in Figure \ref{fig:vcmp}, is a
783: characteristic velocity. It is natural to identify the
784: characteristic length scale with the length scale corresponding to
785: the fastest growing unstable mode, $\lambda_{\textrm{fast,r}}$, so
786: that $\nu \sim \lambda_{\textrm{fast,r}}
787: \sigma_{\textrm{turb}}$. The top panel of Figure \ref{fig:visc}
788: shows that the turbulent viscosity $\nu$, given approximately by
789: $\lambda_{\textrm{fast,r}} \sigma_{\textrm{turb}}$, scales linearly
790: with the radius over much of the disk, so that $\nu \propto
791: R$. Although the behavior of the turbulent viscosity $\nu$ can
792: potentially change with time, the viscous torque $G$ should remain
793: approximately constant in radius for the above interpretation to be
794: valid. The bottom panel of Figure \ref{fig:visc} shows the average
795: viscous torque $G$, normalized to its value at $10$ pc. The
796: normalized value is close to unity over most of the circumnuclear
797: region, which is indeed the condition needed to explain the $-8/3$
798: slope of the spherically-averaged gas density profile of Figure
799: \ref{fig:dens}.
800:
801: %\clearpage
802: \begin{figure}[t]
803: \centering
804: % \epsscale{1.}
805: \epsscale{1.2}
806: \plotone{f9.eps}
807: \caption{\label{fig:visc} {\it Top:} Ratio of kinematic viscosity
808: $\nu$, given approximately by $\lambda_{\textrm{fast,r}}
809: \sigma_{\textrm{turb}}$, to radius $R$. The ratio $\nu / R$ is
810: roughly constant in radius. {\it Bottom:} The viscous torque $G$
811: normalized to its $10$ pc value. The viscous torque is constant in
812: radius, consistent with the power-law gas density slope $\rho
813: \propto R^{-8/3}$.}
814: \end{figure}
815: %\clearpage
816:
817: \noindent Thus, a simple description of the structure of the
818: circumnuclear disk based on equations (\ref{eq:ang}-\ref{eq:torqb})
819: and Fig.\ \ref{fig:visc} provides a good match to the simulation
820: results, supporting the idea that global instabilities in the disk
821: are responsible for generating turbulence in the gas, resulting in
822: the power-law slope described in Sections \ref{subsection:disk} and
823: \ref{subsection:ang}.
824:
825: The conditions in the circumnuclear disk are potentially conducive
826: to gaseous bar formation, the conditions for which have been studied
827: extensively for different geometries and physical conditions,
828: resulting in a variety of criteria
829: \citep[e.g.][]{OP73,Efstat82,Christo95_1,Christo95_2,Bottema03,Wyse04}.
830: A comparison with the criterion of \citet{OP73}, which characterizes
831: stability in terms of the ratio $t=T/|W|$ of kinetic to
832: gravitational potential energy, indicates that the disk is both
833: secularly and dynamically unstable to bar formation on all scales $r
834: \gtrsim 0.1 \dim{pc}$ adequately resolved by the simulation. Because
835: of the chaotic and transient behavior of the instabilities, the
836: structures shown in Figure \ref{fig:evol} do not resemble a single,
837: well defined bar, but rather a highly perturbed ``bar.''
838:
839: Since the Toomre criterion applies only to small linear
840: perturbations and the Ostriker-Peebles criterion describes global
841: modes, the question remains whether non-linear effects lead to
842: fragmentation on smaller scales, below the resolution of the
843: simulation. The top panel of Figure \ref{fig:jeans} shows the ratio
844: of the local Jeans length of the gas, $\lambda_{\textrm{J}}$, to
845: cell size, $\Delta x$, for all simulation cells in the circumnuclear
846: disk (with temperatures less than $10^3$ K). In most cells, the
847: Jeans length is larger than the cell size, preventing the gas from
848: fragmenting into sub-cell sized clouds, ultimately leading to star
849: formation. The Truelove criterion for preventing numerical
850: fragmentation is enforced on the maximum level of refinement (level
851: $20$) only, so there are no numerical restrictions to prevent gas
852: from collapsing all the way to level $20$. However, only the central
853: sub-pc part of the disk reaches levels $19$ and $20$, as shown by
854: the histogram in the lower panel of Figure \ref{fig:jeans},
855: indicating that the disk is stable to non-linear effects. While some
856: cells in the disk have $\lambda_{\textrm{J}} < \Delta x$ and may
857: form stars, Figure \ref{fig:jeans} demonstrates that most of the
858: disk is stable against collapse, and that there is no widespread
859: burst of star formation. The few cells with
860: $\lambda_{\textrm{J}}/\Delta x < 1$ may correspond to resonances
861: where the pattern speed of waves in the disk is the same as the
862: rotational velocity, leading to higher gas densities, and possible
863: star formation. However, a detailed analysis of the behavior of
864: these resonances with time is beyond the scope of this paper.
865:
866: %\clearpage
867: \begin{figure}[t]
868: \centering
869: % \epsscale{1.}
870: \epsscale{1.2}
871: \plotone{f10.eps}
872: \caption{\label{fig:jeans} {\it Top:} Scatter plot showing the
873: ratio of the local Jeans length, $\lambda_{\textrm{J}}$, to the
874: cell size, $\Delta x$ at $300{,}000$ years after the initial
875: refinement. Only cool cells with temperatures less than $10^3$ K
876: are shown. {\it Bottom:} Histogram showing the volume of level
877: $20$ and of level $19$ cells as a function of radius. The
878: simulation only refines to the maximum levels in the center of
879: the circumnuclear disk.}
880: \end{figure}
881: %\clearpage
882:
883: %%%%%%%%%%%%%%%%%%%%%%%%%%%
884:
885: \subsection{The Nature of the Turbulence}
886: \label{subsection:turb}
887:
888: Turbulence has been widely studied in modeling of the ISMs of
889: galaxies and star forming regions \citep[for recent reviews,
890: see][]{MacLowKlessen04,McKeeOst07}. It is important to understand
891: the nature of the turbulence in the present simulated galaxy as it
892: plays a key role in the disk properties. The global instabilities
893: arising in the simulated circumnuclear disk generate turbulence on a
894: range of scales, maintaining the quasi-steady state of the disk.
895:
896: Turbulence dissipates energy in such a way that the turbulent
897: velocity scales as $\sigma_{\textrm{turb}} \propto \lambda^q$, where
898: $q=1/3$ for incompressible, sub-sonic (Kolmogorov) turbulence and
899: $q=1/2$ for compressible turbulence in the zero-pressure limit
900: (Burgers turbulence). The supersonic turbulence in the simulated
901: disk falls between these two limits. In AMR simulations, it is
902: straightforward to measure the turbulent velocity on different
903: scales because the information on different levels of refinement is
904: readily available. However, the measurements are only accurate to
905: within a factor of two in scale because the cell size decreases by a
906: factor of two with each refinement. The approximate scaling of the
907: turbulence is shown in the top panel of Figure \ref{fig:spec}. The
908: figure shows the turbulent velocity (given by the RMS velocity
909: dispersion between neighboring cells) at two different radii for two
910: different simulation runs. The first is the fiducial run described
911: in the previous sections, and the second is a short portion of the
912: fiducial run, with a more aggressive refinement criterion on levels
913: $13$ and below, given by $m_{\textrm{r}}^{\textrm{level}-10}
914: \textrm{max}[0.5^{\textrm{level}-12},\ 0.125]$ (where
915: $m_{\textrm{r}}$ is an empirically determined parameter defined in
916: Section \ref{subsection:ref}). The more aggressive run probes
917: smaller scales at a given radius, in order to capture the scale of
918: the turbulence. The turbulent velocities shown in Figure
919: \ref{fig:spec} demonstrate the scaling of the turbulence down to
920: small scales. It is difficult to determine the precise slope of the
921: scaling using measurements from the present simulation because of
922: the limitation imposed by the refinement scheme. While a more
923: precise characterization of the turbulence spectrum is beyond the
924: scope of the present simulation, Figure \ref{fig:spec} shows that
925: the slope approximately falls between the Kolmogorov and Burgers
926: turbulence limits, as expected.
927:
928: %\clearpage
929: \begin{figure}[t]
930: \centering
931: \epsscale{1.}
932: \plotone{f11.eps}
933: \caption{\label{fig:spec} {\it Top:} Turbulent velocity as a
934: function of scale (or cell size) at $3$ and $8$ pc, for the
935: fiducial and aggressive refinement runs. The more aggressive
936: refinement run resolves the scaling of the turbulent down to
937: smaller scales. {\it Bottom:} A comparison of the smallest
938: resolved scale of turbulence and the fastest growing unstable
939: mode, $\lambda_{\textrm{fast,all}}$ (where
940: $\lambda_{\textrm{fast,all}}$ is the same as in Figure
941: \ref{fig:Q}).}
942: \end{figure}
943: %\clearpage
944:
945: At scales below the resolution at a given location in the disk, the
946: turbulent velocities in Figure \ref{fig:spec} level off at constant
947: values because the turbulence scaling can only be measured down to
948: the resolution limit. The more aggressive refinement run ({\it
949: triangles} in Figure \ref{fig:spec}) probes smaller scales for a
950: given radius, and therefore levels off at smaller scales. The bottom
951: panel of Figure \ref{fig:spec} shows the approximate scale at which
952: the flattening of the spectrum occurs for several different radii in
953: the circumnuclear disk of the fiducial run. For comparison, the
954: fastest growing unstable mode, $\lambda_{\textrm{fast,all}}$ is
955: shown as well. The turbulent velocity measured at the resolution of
956: the simulation is a numerical quantity, but it roughly corresponds
957: to the turbulence on the scale of $\lambda_{\textrm{fast,r}}$, which
958: is a physical quantity. Figure \ref{fig:spec} demonstrates that the
959: simulation resolves all scales that are Toomre $Q$ unstable,
960: supporting the resulting interpretation that instability driven
961: turbulence maintains the quasi-stationary state of the circumnuclear
962: disk.
963:
964: The density structure of the disk is also consistent with the
965: description of turbulence in the disk. Supersonic turbulence
966: typically imposes a log-normal density distribution on the gas, as
967: demonstrated by models of turbulence in molecular clouds
968: \citep[e.g.][]{Vazquez94, PassotVaz98, Wada01, Kritsuketal07,
969: McKeeOst07}, such as
970:
971: \begin{equation}\label{eq:pdf}
972: \frac{dV}{d\rho} = \frac{1}{\rho \sqrt{2\pi\sigma^2}} \
973: \textrm{exp}\left[\frac{-(\textrm{ln} \rho - \mu)^2}{2
974: \sigma^2}\right],
975: \end{equation}
976:
977: \noindent where the mean, $\mu$, is defined as
978: $\overline{\textrm{ln}\ \rho}$, and $\sigma$ is the
979: dispersion. Figure \ref{fig:pdf} shows the volume-weighted
980: probability distribution function (PDF), time-averaged over $\sim
981: 250,000$ years, for a shell of gas at a radius of $30$ pc. The PDF is
982: well fit by a log-normal distribution over at least $4$ orders of
983: magnitude in density, consistent with models of supersonic turbulence.
984:
985: %\clearpage
986: \begin{figure}[t]
987: \centering
988: % \epsscale{1.}
989: \epsscale{1.2}
990: \plotone{f12.eps}
991: \caption{\label{fig:pdf} A volume-weighted PDF of the gas density
992: in cells at a scale of $\sim 30$ pc, with temperatures $< 10^3$
993: K (in order to exclude the hot corona outside the disk),
994: averaged over $\sim 550,000$ years. The PDF is well fit by a
995: log-normal distribution over the range $10^2 \lesssim \rho
996: \lesssim 10^6 \dim{M}_{\sun} \dim{pc}^{-3}$ with a mean density
997: of $10^{3.9} \dim{M}_{\sun} \dim{pc}^{-3}$ ({\it line}).}
998: \end{figure}
999: %\clearpage
1000:
1001:
1002: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1003:
1004: \section{Possible Effects of Missing Physics}
1005: \label{sec:miss}
1006:
1007: The simulations presented here do not include all of the potentially
1008: important physics for galaxy evolution. Therefore, in this section
1009: we discuss the potential effect that the physics missing from the
1010: simulations might have on our results.
1011:
1012: \subsection{Optically Thick Cooling}
1013: % \label{subsubsection:torus}
1014: \label{subsection:torus}
1015:
1016: Cooling in the high density, central region of the galaxy is not
1017: treated entirely correctly by the ART code. The column density of
1018: the molecular gas in the center is so high that the gas is expected
1019: to become optically thick to its own cooling radiation
1020: \citep[e.g.][]{RipAbel04}. Additionally, the presence of dust grains
1021: in this region traps radiation and halts cooling further
1022: \citep[see][]{DraineLee84, Ossenkopf94}. Figure \ref{fig:col} shows
1023: the column density vertically through the circumnuclear gas
1024: disk. Horizontal lines are shown for comparison, corresponding to
1025: the column densities at which dust and H$_2$ each become optically
1026: thick to their own cooling radiation ($\sim 10^{26} \dim{cm}^{-2}$
1027: and $\sim 10^{27} \dim{cm}^{-2}$, respectively, at $0.1$ solar
1028: metallicity, which is the metallicity of the circumnuclear disk in
1029: the simulation at $z=4$). In the inner $\approx 10$ pc, the column
1030: density of the gas is large enough that the opacity of dust and
1031: molecular gas must be accounted for to accurately describe cooling
1032: in the simulation.
1033:
1034: %\clearpage
1035: \begin{figure}[t]
1036: \centering \epsscale{1.2} \plotone{f13.eps}
1037: \caption{\label{fig:col} Column density of the gas. The horizontal
1038: lines show the column densities at which dust and H$_2$ become
1039: optically thick to their own cooling radiation.}
1040: \end{figure}
1041: %\clearpage
1042:
1043: The simulation runs presented here do not include optically thick
1044: radiative transfer. This may explain why the gas remains in a thin
1045: disk all the way in to sub-pc scales, and does not resemble the
1046: obscuring tori observed in the inner few parsecs of AGN. Should the
1047: opacity of the gas to its own cooling radiation be included, the gas
1048: around the mid-plane of the disk would not be able to cool and would
1049: heat to temperatures corresponding to the energy dumped into the gas
1050: by the turbulence. These higher temperatures may be able to provide
1051: a substantial vertical pressure support in the central few parsecs
1052: of the disk, resulting in a thicker, more toroidal-like
1053: structure\footnote{Notice, that the outer, optically thin layers of
1054: such a disk would continue cooling efficiently, covering the hot
1055: interior with a cold molecular ``skin''. Such a skin may become
1056: Rayleigh-Taylor unstable, leading to fragmentation of the torus into
1057: individual clouds.}. However, it has been suggested that
1058: hydromagnetic disk winds, and not hydrostatic pressure support, may
1059: be responsible for sustaining the optically and geometrically thick
1060: obscuration region, or ``obscuring torus,'' in the nuclei of
1061: galaxies \citep[e.g.][]{KonigKartje94,ElitzShlos06}. In which case,
1062: the inclusion of optically thick radiative transfer may not be
1063: sufficient to create a ``torus'' in our simulations.
1064:
1065: In future simulations we plan to remedy this limitation of the
1066: present simulation by accounting for the opacity of the high
1067: density region in the center, which will allow us to consistently
1068: incorporate radiative feedback from a central source and to test the
1069: above hypothesis.
1070:
1071: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1072: \subsection{Magnetic Fields}
1073: \label{subsection:mag}
1074:
1075: The ART code does not include magnetic fields, whereas MHD is
1076: certainly important for accretion disk physics \citep[see][and
1077: references therein]{BalbusHawley98}. It is for this reason that the
1078: present study has been restricted to scales larger than $\approx 0.1
1079: \dim{pc}$, corresponding to about $10^4$ Schwarzschild radii for a
1080: black hole of mass $3 \times 10^7$ M$_{\sun}$.
1081:
1082: At scales larger than the accretion disk, the absence of magnetic
1083: fields in the ART code is probably not important. An estimate of the
1084: strength that an equipartition magnetic field would need in order to
1085: affect the disk on small scales is given by $B_{\textrm{eq}}^2
1086: \approx 4\pi \rho \sigma_{\textrm{turb}}^2$. In order to affect the
1087: large-scale dynamics of the gas, the magnetic field would have to
1088: have an even larger strength of order $B_{\textrm{dyn}}^2 \approx
1089: 4\pi \rho v_{\textrm{t}}^2$. Figure \ref{fig:B} shows estimates of
1090: $B_{\textrm{eq}}$ and $B_{\textrm{dyn}}$, in order to demonstrate
1091: how high the magnetic field would need to be to significantly
1092: influence gas dynamics in the simulated galaxy. For most of the
1093: galactic disk, the above estimates for the magnetic field are far
1094: higher than the few $\mu$G fields observed in real galaxies
1095: \citep[e.g.][]{ZweibHeil97,Beck01}. Even in the sub-pc region, where
1096: water maser observations indicate stronger fields of a few tens of
1097: mG \citep[e.g.][]{Modjazetal05,Vlemmetal07}, the magnetic field is
1098: still too low to affect the gas dynamics in the model galaxy. We
1099: therefore conclude that magnetic fields, which are not included in
1100: our simulations, will have a negligible effect on the dynamical
1101: state of this disk unless their strength greatly exceeds
1102: observational measurements.
1103:
1104: %\clearpage
1105: \begin{figure}[t]
1106: \centering
1107: \epsscale{1.2}
1108: \plotone{f14.eps}
1109: \caption{\label{fig:B} Equipartition ({\it solid}) and dynamical
1110: ({\it dashed}) magnetic fields needed to influence the gas
1111: dynamics in the disk.}
1112: \end{figure}
1113: %\clearpage
1114:
1115: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1116: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1117:
1118: \section{DISCUSSION AND CONCLUSIONS}
1119: \label{sec:disc}
1120:
1121: We have used a large-dynamic range cosmological simulation to study
1122: gas dynamics in the circumnuclear disk of typical mass galaxy at
1123: $z\approx4$ (evolving into an $L_*$ galaxy at $z=0$), resolving the
1124: distribution of matter from megaparsec scales all the way down to
1125: sub-pc scales (with $20$ levels of refinement). The simulation
1126: reveals a cold, fully molecular, self-gravitating, and turbulent
1127: rotationally-supported gas disk, which is globally unstable but
1128: locally stable.
1129:
1130: The global instability, operating on a range of scales comparable to
1131: the size of the system, generates turbulence down to the smallest
1132: scale resolved in the simulation. On small scales, the turbulence
1133: supports most of the disk against gravitational fragmentation and
1134: collapse. The disk, therefore, remains locally stable and reaches a
1135: quasi-stationary state. In that state, global instability drives
1136: bar-like and spiral-wave-like structures on time scales of the order
1137: of $100{,}000$ years at $100\dim{pc}$ scales (and on proportionally
1138: shorter time scales at smaller radii), but on a $500{,}000$ year
1139: time-scale the structure of the disk remains quasi-steady.
1140:
1141: The disk develops a power-law gas density profile with a well
1142: defined slope of $-5/3$ in surface density and $-8/3$ in
1143: spherically-averaged volume density. This slope is a natural
1144: consequence of the dynamical state of the disk: the turbulence,
1145: generated by the global instability, redistributes the angular
1146: momentum in the disk on a dynamical time scale, reducing the
1147: gradient of the viscous torque. The resulting (quasi-)steady state
1148: uniquely determines the gas density profile we find in the
1149: simulation and the distribution of the angular momentum with gas
1150: mass.
1151:
1152: The turbulence in the disk drives the local outward transport of
1153: angular momentum and the inward flow of gas toward the supermassive
1154: black hole on time scales of at least 10 million years, which are
1155: too long to follow in a single zoom-in episode of the
1156: simulation. Thus, we only capture a single snapshot in the
1157: cosmological life of the supermassive black hole. In follow-up
1158: work, we plan to consider several snapshots taken at different
1159: cosmological times in order to describe the system on longer time
1160: scales.
1161:
1162: The dynamical state of the disk that we find in our simulation
1163: appears to be consistent with the results of previous simulations of
1164: isolated circumnuclear disks in galaxies \citep{Fukuda00,
1165: Wada01, WadaNorman01, Escala07}. A distinctive feature of our
1166: approach is that we follow the dynamics of the circumnuclear disk
1167: within cosmological simulations. While most of the volume of the
1168: simulation evolves little on time scales relevant for the dynamics
1169: of the circumnuclear disk, cosmological scales provide realistic
1170: boundary conditions for the dynamics on sub-kpc scales. Since we
1171: find that the circumnuclear disk rapidly reaches a quasi-stationary
1172: state, its evolution is entirely governed by the boundary
1173: conditions. For the same reason, the fact that the cosmological
1174: simulation that was used for the initial conditions did not resolve
1175: the scale of the circumnuclear disk, does not compromise our
1176: results: the quasi-stationary state of the disk does not depend on
1177: the initial conditions, and so the lack of power on scales below
1178: about $100\dim{pc}$ in the initial cosmological simulation is not
1179: important.
1180:
1181: The adopted approach of this work is comparable to the recent study
1182: by \citet{Mayeretal07}, who used a cosmological simulation to model the
1183: coalescence of two supermassive black holes. While the detailed
1184: treatment of gas physics and the scientific questions answered by
1185: the two studies are different, many of our results are consistent
1186: with those of \citet{Mayeretal07}. For example, they find a similar
1187: slope for the spherically averaged density profile, although they do
1188: not elaborate on the physical mechanisms for the formation of this
1189: profile.
1190:
1191: The local stability of the disk, supported by highly supersonic
1192: turbulence, implies that the disk is capable of continuously feeding
1193: a central black hole, uninterrupted by catastrophic bursts of star
1194: formation, which could have consumed the available fuel were the
1195: disk locally unstable. This interesting dynamical state of the disk
1196: provides a potential solution to the problem of how AGN fueling is
1197: maintained by self-gravitating gas disks \citep[for a related
1198: discussion, see e.g.][]{ShlosBeg89,Riceetal05,NayakshinKing07}.
1199:
1200: The circumnuclear disk in the simulation extends all the way to the
1201: maximally resolved scale of $\approx0.1\dim{pc}$, which corresponds
1202: to the outer part of the black hole accretion disk. We find no
1203: toroidal-like structures on several parsec scales, which are
1204: commonly inferred to exist around AGN. A possible reason that the
1205: disk does not form an AGN torus on the appropriate scales (if it
1206: indeed, should) is the limitation imposed by our implementation of
1207: gas cooling in the code. The ART code, like all existing
1208: cosmological codes, assumes that the cooling radiation from cosmic
1209: gas escapes freely into the IGM. This assumption breaks down at the
1210: densities and temperatures reached by the simulation in the inner
1211: $10\dim{pc}$. At this scale the disk becomes optically thick to its
1212: own cooling radiation from dust and molecular hydrogen. In that
1213: regime the disk may to heat up and acquire a substantial
1214: amount of pressure support, which may result in a puffier, more
1215: toroidal-like configuration for the inner several parsecs of the
1216: disk. A consistent treatment of a putative AGN torus will require
1217: simulations that incorporate optically-thick cooling.
1218:
1219: \acknowledgements
1220:
1221: The authors thank Oleg Gnedin for his analysis of the mass profiles
1222: from the simulations. The authors also thank Mitch Begelman, Moshe
1223: Elitzur, Lucio Mayer, Brant Robertson, Isaac Shlosman, Volker
1224: Springel, Marta Volonteri, \& Keiichi Wada for comments on the
1225: paper. This work was supported in part by the DOE and the NASA grant
1226: NAG 5-10842 at Fermilab and by the NSF grants AST-0134373 and
1227: AST-0507596. Supercomputer simulations were run on the IBM P690
1228: array at the National Center for Supercomputing Applications and San
1229: Diego Supercomputing Center (under grant AST-020018N).
1230:
1231:
1232: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1233:
1234: \bibliography{ms}
1235:
1236: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1237:
1238: \appendix
1239:
1240: \section{Measuring the Viscosity}
1241: \label{sec:mvisc}
1242:
1243: Here we address the validity of our estimate of the turbulent
1244: kinematic velocity parameter $\nu$, which motivates the
1245: interpretation that turbulent transport of angular momentum
1246: maintains local stability and the quasi-steady state of the
1247: circumnuclear disk. Instead of a direct measurement of the
1248: viscosity, we have assumed that the viscosity depends on
1249: characteristic velocity and length scales, given by
1250: $\sigma_{\textrm{turb}}$ and $\lambda_{\textrm{fast,r}}$,
1251: respectively. The accuracy of this estimate can be tested by
1252: measuring the transport of gas due to turbulent motions.
1253:
1254: The ART code does not follow individual parcels of gas across cell
1255: boundaries, but rather advected quantities describing the
1256: gas. Therefore, in order to follow the turbulent motions of the gas,
1257: we introduce a passive scalar into the simulation which has a value
1258: of unity inside a spherical region centered on the black hole
1259: particle, and a value of zero outside this region, effectively
1260: ``painting'' the gas. As the simulation evolves, turbulent motions
1261: of the gas cause the profile of the scalar quantity to deviate from
1262: its original spherical form. The evolution of the profile depends on
1263: the turbulent kinematic viscosity, $\nu$, and in the early stages of
1264: diffusion, while the deviations from the initial profile are small,
1265: the profile approximately satisfies the linear diffusion equation
1266: for initial conditions,
1267:
1268: \begin{equation}
1269: P(r,t=0) =
1270: \begin{cases}
1271: 1, & \mbox{if }r<r_0, \\
1272: 0, & \mbox{if }r>r_0,
1273: \end{cases}
1274: \end{equation}
1275:
1276: \noindent (where $r_0$ is the radius of the initial sphere) allowing
1277: a more direct calculation of the viscosity. The solution to the
1278: linear diffusion equation for the above initial conditions, is
1279:
1280: \begin{equation}\label{eq:prof}
1281: P(x,q) = \frac{1}{2}\left[ \textrm{erf}\left(\frac{x+1}{\sqrt{q}}\right) -
1282: \textrm{erf}\left(\frac{x-1}{\sqrt{q}}\right)\right] - \frac{q}{x \sqrt{\pi
1283: q}}\ e^{-(1+x^2)/q}\ \sinh\left(\frac{2x}{q}\right),
1284: \end{equation}
1285:
1286: \noindent where
1287:
1288: \begin{equation}
1289: x = \frac{r}{r_0}\ \ \&\ \ q = \frac{4 \nu t}{r_0^2}.
1290: \end{equation}
1291:
1292: Figure \ref{fig:paint} shows samples of the profile evolving from a
1293: $30$ pc sphere for different times ($q$). The measured profiles
1294: agree well with the solution given in Equation
1295: \ref{eq:prof}. Starting from several painted regions at $1, 3, 10,
1296: 30, \textrm{and}\ 100$ pc radii, we followed the evolution of the
1297: paint-weighted density profiles and measured the turbulent
1298: viscosity, $\nu$, as a function of radius. Figure \ref{fig:vpaint}
1299: shows the best-fit measurements of $\nu/r$ in relation to the
1300: estimate given by $\sigma_{\textrm{turb}} \lambda_{\textrm{fast,r}} /
1301: r$. The lines in Figure \ref{fig:vpaint} show individual snapshots
1302: of $\sigma_{\textrm{turb}} \lambda_{\textrm{fast,r}} / r$
1303: corresponding to the timescale over which the diffusion was followed
1304: (whereas Figure \ref{fig:visc} showed a time-averaged estimate).
1305: The best-fit measurements match the estimates for $\nu/r$ well, thus
1306: lending justification to the estimate for $\nu$ used in the analytic
1307: arguments of Section \ref{subsection:stab}.
1308:
1309: %\clearpage
1310: \begin{figure}[t]
1311: \centering
1312: \epsscale{1.}
1313: \plotone{fa1.eps}
1314: \caption{\label{fig:paint} Mass-weighted profiles of the passive
1315: scalar as a function of radius, and evolving over time from a $30$
1316: pc sphere. The symbols are measurements from the simulation, and
1317: the lines are the best fits to Equation \ref{eq:prof}.}
1318: \end{figure}
1319: %\clearpage
1320:
1321: %\clearpage
1322: \begin{figure}[t]
1323: \centering
1324: \epsscale{1.}
1325: \plotone{fa2.eps}
1326: \caption{\label{fig:vpaint} Ratio of kinematic viscosity $\nu$, to
1327: radius $R$. The lines show estimates for $\nu/R$ given
1328: approximately by $\lambda_{\textrm{fast,r}}
1329: \sigma_{\textrm{turb}}$, for three different snapshots. The points
1330: are estimates from a second method measuring the advection of gas
1331: at five different radii in the simulation. The snapshots
1332: correspond to the profile fits at different times, so the $1$ pc
1333: point should be compared to the solid line, the next point to the
1334: dashed line, and the last three to the dotted line.}
1335: \end{figure}
1336: %\clearpage
1337:
1338: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1339:
1340: \section{Angular Momentum Conservation}
1341: \label{sec:con}
1342:
1343: Anomalous numerical transport of angular momentum is sometimes
1344: presented as a concern for adaptive mesh refinement simulations
1345: using interpolation schemes for calculating velocities on the
1346: mesh. We have tested conservation of angular momentum in the code by
1347: conducting a separate set of simulations of the collapse of an
1348: isothermal sphere (the details are included in the PhD thesis of the
1349: author Levine). The test was conducted for a mesh which refined
1350: along with the collapsing sphere, and for a pre-refined mesh, in
1351: order to test the conservation of angular momentum both as the mesh
1352: refines and as gas moves through the mesh. The sphere collapses into
1353: a thin disk ($h/r << 1$), which conserves angular momentum over
1354: several rotation periods.
1355:
1356: In Figure \ref{fig:jvm} we demonstrate the conservation of angular
1357: momentum in the present simulation. As the maximum resolution of the
1358: simulation increases, the angular momentum profile initially
1359: evolves, as the viscous torque, $G$, described in Section
1360: \ref{subsection:ang} redistributes angular momentum within the
1361: disk. But after reaching the maximum, level $20$ resolution (the
1362: point at which we introduce the black hole particle into the
1363: simulation and let it evolve), the profile remains rather steady
1364: with time, because the disk has reached a quasi-stationary
1365: state. Figure \ref{fig:jvm} shows the angular momentum as a function
1366: of enclosed gas mass at several different times during a single
1367: zoom-in episode of the simulation. The figure shows the evolution of
1368: the angular momentum distribution from $200,000$ years before the
1369: introduction of the black hole particle, when the simulation has
1370: refined to level $11$, to $500,000$ years after the introduction of
1371: the black hole particle, when the simulation is fully refined, and
1372: has evolved for several hundred thousand years in a quasi-steady
1373: state.
1374:
1375: On the scale of the circumnuclear disk, the simulation has made
1376: several thousand time steps between $t=100$ kyrs and $t=500$ kyrs,
1377: meaning that the simulation has undergone significant evolution on
1378: this spatial scale. Numerical effects, accumulating with each time
1379: step, would cause deviations in the angular momentum profile. Over
1380: longer time scales, the slow inward transport of matter will alter
1381: the angular momentum profile. However, on the hundred thousand year
1382: time scale we expect the profile to remain rather steady, reflecting
1383: the quasi-stationary state of the disk, unless there is anomalous
1384: numerical transport of angular momentum. Figure \ref{fig:jvm} shows
1385: no systematic deviations in the angular momentum profile after the
1386: initial refinement and the corresponding re-distribution of the
1387: angular momentum, demonstrating that there is no sign of unphysical
1388: numerical angular momentum transport on the time scales of the
1389: simulation. On time scales much longer than those simulated in the
1390: present paper, however, these effects may become important.
1391:
1392: %\clearpage
1393: \begin{figure}[t]
1394: \centering
1395: \epsscale{1.}
1396: \plotone{fb1.eps}
1397: \caption{\label{fig:jvm} Angular momentum as a function of
1398: enclosed gas mass for several different times. The times are measured
1399: from the introduction of the black hole. The vertical dashed lines
1400: correspond to enclosed masses on scales of $\approx 10$ and $\approx
1401: 1000$ pc. At each of these scales, the simulation has made
1402: approximately $230{,}000$ and $3{,}600$ steps, respectively
1403: (corresponding to the resolution at each scale), between $t=100$
1404: and $t=500$ kyrs.}
1405: \end{figure}
1406: %\clearpage
1407:
1408:
1409:
1410:
1411: \end{document}
1412: