0706.4303/ms.tex
1: %---------------------------------------------------------------------------
2: %Use this line for submission to ApJ
3: %\documentclass[12pt,preprint]{aastex}
4: %Use this line for a more compact single-column version to submit to LANL
5: \documentclass[10pt,preprint]{aastex}
6: %Use this line for a compact two-column version
7: %\documentclass[10pt,preprint2]{aastex}
8: 
9: %\usepackage{emulateapj5}
10: %\usepackage{amsmath,natbib,graphicx,euscript} %,epsfig}
11: 
12: \newcommand{\myemail}{fragilep@cofc.edu}
13: \slugcomment{To be Submitted to ApJ}
14: 
15: %-----------------------------------------------------------------------------
16: \begin{document}
17: %------------------------------------------------------------------------------
18: 
19: \title{ Global General Relativistic MHD Simulation of a Tilted Black-Hole Accretion Disk }
20: 
21: \author{P. Chris Fragile}
22: \affil{Department of Physics and Astronomy, College of Charleston,
23: Charleston, SC 29424; fragilep@cofc.edu}
24: \author{Omer M. Blaes}
25: \affil{Department of Physics, University of California, Santa
26: Barbara, CA 93106}
27: \author{Peter Anninos and Jay D. Salmonson}
28: \affil{University of California, Lawrence Livermore National
29: Laboratory, Livermore CA 94550}
30: 
31: %-----------------------------------------------------------------------------
32: 
33: \begin{abstract}
34: This paper presents a continuation of our efforts to numerically
35: study accretion disks that are misaligned (tilted) with respect to
36: the rotation axis of a Kerr black hole. Here we present results of a
37: global numerical simulation which fully incorporates the effects of
38: the black hole spacetime as well as magnetorotational turbulence
39: that is the primary source of angular momentum transport in the
40: flow. This simulation shows dramatic differences from comparable
41: simulations of untilted disks. Accretion onto the hole occurs
42: predominantly through two opposing plunging streams that start from
43: high latitudes with respect to both the black-hole and disk
44: midplanes. This is due to the aspherical nature of the gravitational
45: spacetime around the rotating black hole. These plunging streams
46: start from a larger radius than would be expected for an untilted
47: disk. In this regard the tilted black hole effectively acts like an
48: untilted black hole of lesser spin. Throughout the duration of the
49: simulation, the main body of the disk remains tilted with respect to
50: the symmetry plane of the black hole; thus there is no indication of
51: a Bardeen-Petterson effect in the disk at large. The torque of the
52: black hole instead principally causes a global precession of the
53: main disk body. In this simulation the precession has a frequency of
54: $3 (M_\odot/M)$ Hz, a value consistent with many observed
55: low-frequency quasi-periodic oscillations. However, this value is
56: strongly dependent on the size of the disk, so this frequency may be
57: expected to vary over a large range.
58: \end{abstract}
59: 
60: \keywords{accretion, accretion disks --- black hole physics ---
61: galaxies: active --- MHD --- relativity --- X-rays: stars}
62: 
63: 
64: \section{Introduction}
65: \label{sec:intro}
66: 
67: Black-hole accretion has long been postulated to power the energetic
68: emissions seen from quasars, active galactic nuclei (AGN), and many
69: galactic X-ray sources; there is now ample observational evidence to
70: support such claims \citep[e.g. ][]{kro99,mcc05}.  Black-hole
71: accretion flows are also of interest as laboratories to test
72: predictions of general relativity. However, the nature of such flows
73: is complex, involving time-dependent, multi-dimensional dynamics
74: with generically little symmetry.  Hence numerical simulations play
75: an integral role in advancing our understanding.
76: 
77: Many simulations of black-hole accretion flows have been carried out
78: over the past three decades, both in the hydrodynamic
79: \citep[e.g.][]{wil72,haw84b,haw91} and magnetohydrodynamic (MHD)
80: \citep[e.g.][]{koi99,gam03a,dev03b} regimes. A common assumption in
81: nearly all of the work to date has been that the symmetry plane of
82: the central black hole is aligned with the midplane of the accretion
83: flow, at least in some averaged sense. However, there is compelling
84: observational evidence in several black-hole X-ray binaries (BHBs),
85: e.g. GRO J1655-40 \citep{oro97} and XTE J1550-564
86: \citep{han01,oro02}, and AGN, e.g. NGC 3079 \citep{kon05}, NGC 1068
87: \citep{cap06}, and NGC 4258 \citep{cap07}, suggesting that
88: misaligned (or tilted) black holes may be common \citep[see
89: also][]{mac02}. This claim relies on the observation of relativistic
90: bipolar jets (thought to be aligned with the spin axis of the black
91: hole) that are not perpendicular to the plane of the accretion disk
92: observed at large scales.
93: 
94: There are also compelling theoretical arguments that many black
95: holes should be tilted. First, the formation avenues for many
96: black-hole - disk systems favor, or at least allow for, a tilted
97: configuration \citep{fra01a}. In stellar mass binaries, the
98: orientation of the outer disk is fixed by the binary orbit, whereas
99: the orientation of the black hole is determined by how it became
100: part of the system, whether through a supernova explosion or
101: multi-body interaction. If the black hole formed from a member of a
102: preexisting binary through a supernova, then the black hole could be
103: tilted if the explosion were asymmetric. If the black hole joined
104: the binary through multi-body interactions, such as binary capture
105: or replacement, then there would have been no preexisting symmetry,
106: so the resulting system would nearly always harbor a tilted black
107: hole. This same argument can be extended to AGN in which merger
108: events reorient the central black hole or its fuel supply and result
109: in repeated tilted configurations.
110: 
111: If an accretion disk is misaligned or tilted, it will be subject to
112: Lense-Thirring precession.  For an ideal test particle in a slightly
113: tilted orbit at a radius $r$ around a black hole of mass $M$ and
114: specific angular momentum $a$, this precession occurs at an angular
115: frequency $\Omega_{\rm LT}\approx2aM/r^3$.  Close to the black hole,
116: this is comparable to the orbital angular frequency
117: $\Omega=(M/r^3)^{1/2}/[1+a(M/r^3)^{1/2}]\approx\Omega_{\rm Kep}$.
118: However, because of its strong radial dependence, Lense-Thirring
119: precession becomes much weaker far from the hole. Therefore, a disk
120: will experience a differential precession that will tend to twist
121: and warp it.
122: 
123: A warping disturbance can be communicated through a disk in either a
124: diffusive or wave-like manner. In the diffusive case, the warping is
125: limited by secular (i.e. ``viscous'') responses within the disk. In
126: such a case, Lense-Thirring precession is expected to dominate out
127: to a unique, nearly constant transition radius \citep{bar75,kum85},
128: inside of which the disk is expected to be flat and aligned with the
129: black-hole midplane, and outside of which the disk is also expected
130: to be flat but in a plane determined by the angular momentum vector
131: of the gas reservoir. This is what we term a ``Bardeen-Petterson''
132: configuration. Interestingly, data for the two black-hole X-ray
133: binaries previously mentioned are best fit by disk components with
134: inclinations that differ from their binary measurements. The
135: best-fit inclinations are more consistent with inclination
136: constraints derived from the radio jets \citep{dav06}, possibly
137: suggesting Bardeen-Petterson configurations. \citet{cap06} also
138: claim that the observations of NGC 1068 are consistent with the
139: predictions of the Bardeen-Petterson effect. Confirmation could come
140: through observations of relativistically broadened reflection
141: features \citep{fra05c}.
142: 
143: The Bardeen-Petterson result is expected to apply for Keplerian
144: disks whenever the dimensionless stress parameter $\alpha$
145: \citep{sha73} is larger than the ratio of the disk semi-thickness
146: $H$ to the radius $r$ at all radii.  Given that $\alpha$ is usually
147: considered to be significantly less than one, this implies very
148: geometrically ``thin'' disks. Unfortunately, current computational
149: limitations prevent us from conducting global simulations of disks
150: that are this thin.  On the other hand, the Bardeen-Petterson regime
151: may not be that common in real disks. Neglecting relativistic
152: correction factors, the innermost, radiation pressure and electron
153: scattering dominated portions of radiatively efficient accretion
154: disks satisfy
155: \begin{equation}
156: \frac{H}{r}\sim\varepsilon^{-1}\left(\frac{L}{L_{\rm Edd}}\right)
157: \left(\frac{r}{r_G}\right)^{-1}, \label{eqhoverr}
158: \end{equation}
159: where $\varepsilon\sim0.1$ is the radiative efficiency, $L/L_{\rm
160: Edd}$ is the luminosity in units of Eddington, and $r_G=GM/c^2$ is
161: the gravitational radius.  Note that equation (\ref{eqhoverr}) is
162: independent of whether the stress is chosen to be proportional to
163: gas pressure, radiation pressure, or some combination of the two. We
164: therefore conclude that the Bardeen-Petterson regime will be
165: relevant in radiatively efficient disks near the black hole only for
166: very small Eddington ratios $L/L_{\rm
167: Edd}\lesssim\alpha\varepsilon<<1$. Moreover, radiatively less
168: efficient, geometrically slim and thick flows will clearly not be in
169: the Bardeen-Petterson regime.
170: 
171: Global simulations of tilted disks that have $H/r>\alpha$ {\em are}
172: computationally feasible. In this regime Lense-Thirring precession
173: is expected to produce warps that propagate in a wave-like manner
174: \citep{pap95a}. In \citet{fra05b} we presented results from the
175: first fully general relativistic three-dimensional hydrodynamic
176: numerical studies of tilted thick-disk accretion onto rapidly
177: rotating (Kerr) black holes. We found that, although Lense-Thirring
178: precession did cause the disk to warp, the warping only occurred
179: inside a radius in the disk at which the precession time became
180: comparable to other dynamical timescales, primarily the azimuthal
181: sound-crossing time. After the differential warping ended and the
182: evolution became quasi-static, the disks underwent near solid-body
183: precession at rates consistent with some low-frequency
184: quasi-periodic oscillations (QPOs).
185: 
186: In this paper we extend the results of \citet{fra05b} to include
187: magnetic fields. The inclusion of magnetic fields is important
188: because it is now widely believed that local stresses within
189: black-hole accretion disks are generated by turbulence that results
190: from the magnetorotational instability \citep[MRI; ][]{bal91}. Here
191: we report on our first global general relativistic MHD (GRMHD)
192: simulation of a tilted accretion disk around a moderately rapidly
193: rotating black hole ($a/M=0.9$). The simulation is initialized
194: starting from the analytic solution for an axisymmetric torus around
195: a rotating black hole. A weak poloidal magnetic field is added to
196: the torus to seed the MRI. After the torus is initialized, the black
197: hole is tilted by an angle $\beta_0=15^\circ$ relative to the disk
198: through a transformation of the metric. The system is then allowed
199: to evolve. This paper reports the results as follows: In \S
200: \ref{sec:methods} we describe the numerical procedures used in this
201: GRMHD simulation. In \S \ref{sec:results} we present the results of
202: this simulation. In \S \ref{sec:discussion} we summarize our
203: findings and draw conclusions.
204: 
205: 
206: \section{Numerical Methods}
207: \label{sec:methods}
208: 
209: This work is carried out using the Cosmos++ astrophysical
210: magnetohydrodynamics code \citep{ann05}. Similar to our predecessor
211: code Cosmos \citep{ann03a}, Cosmos++ includes several schemes for
212: solving the GRMHD equations. The fluid equations can be solved using
213: a traditional artificial viscosity scheme, non-oscillatory central
214: difference methods, or a new hybrid dual energy (internal and total)
215: method. For this work, we use the artificial viscosity formulation,
216: mainly because of its speed and robustness. With the magnetic fields
217: we solve the induction equation in an advection-split form and apply
218: a hyperbolic divergence cleanser to maintain an approximately
219: divergence-free magnetic field. For clarity and notation sake, we
220: present the full evolution equations for mass, internal energy,
221: momentum, and magnetic induction as solved in this work. Throughout
222: this paper we use units where $G=c=1$ and the metric signature is
223: ($-$,$+$,$+$,$+$). We use the standard notation in which four- and
224: three-dimensional tensor quantities are represented by Greek and
225: Latin indices, respectively.
226: 
227: The evolution equations are
228: \begin{eqnarray}
229:  \partial_t D + \partial_i (DV^i) &=& 0 ~,  \label{eqn:av_de} \\
230:  \partial_t E + \partial_i (EV^i) &=&
231:     - P \partial_t W - \left(P + Q\right) \partial_i (WV^i) ~,
232:     \label{eqn:av_en} \\
233:  \partial_t S_j + \partial_i (S_j V^i) &=&
234:       \frac{1}{4\pi} \partial_t (\sqrt{-g} B_j B^0)
235:     + \frac{1}{4\pi} \partial_i (\sqrt{-g} B_j B^i)
236:     \label{eqn:av_mom} \nonumber \\
237:     & & {} + \left( \frac{S^\mu S^\nu}{2S^0} - \frac{\sqrt{-g}}{8\pi}
238:              B^\mu B^\nu \right) \partial_j g_{\mu\nu}
239:     - \sqrt{-g}~\partial_j \left( P + P_B + Q \right) ~, \\
240:  \partial_t \mathcal{B}^j + \partial_i (\mathcal{B}^j V^i) &=&
241:     \mathcal{B}^i \partial_i V^j + g^{ij}~\partial_i \psi ~,
242:       \label{eqn:av_ind} \\
243:  \partial_t \psi + c_h^2 \partial_i \mathcal{B}^i &=&
244:  -\frac{c_h^2}{c_p^2} \psi ~, \label{eqn:div_clean}
245: \end{eqnarray}
246: where $g_{\mu\nu}$ is the 4-metric, $g$ is the 4-metric determinant,
247: $W=\sqrt{-g} u^0$ is the relativistic boost factor, $D=W\rho$ is the
248: generalized fluid density, $V^i=u^i/u^0$ is the transport velocity,
249: $u^\mu = g^{\mu \nu} u_\nu$ is the fluid 4-velocity, $S_\mu = W(\rho
250: h + 2P_B) u_\mu$ is the covariant momentum density,
251: $E=We=W\rho\epsilon$ is the generalized internal energy density, $P$
252: is the fluid pressure, $Q$ is the artificial viscosity used for
253: shock capturing, and $c_h$ and $c_p$ are coefficients to determine
254: the strength of the hyperbolic and parabolic pieces of the
255: divergence cleanser. There are two representations of the magnetic
256: field in these equations: $B^\mu$ is the rest frame magnetic
257: induction used in defining the stress tensor
258: %
259: \begin{equation}
260: T^{\mu\nu} = \left(\rho h + 2P_B\right) u^\mu u^\nu + \left(P +
261: P_B\right)g^{\mu\nu} - \frac{1}{4\pi}B^\mu B^\nu
262: \end{equation}
263: %
264: and
265: \begin{equation}
266: \mathcal{B}^\mu = W(B^\mu - B^0 V^\mu)
267: \end{equation}
268: is the divergence-free ($\partial \mathcal{B}^i / \partial x^i =
269: 0$), spatial ($\mathcal{B}^0=0$) representation of the field. The
270: time component of the magnetic field $B^0$ is recovered from the
271: orthogonality condition $B^\mu u_\mu = 0$
272: \begin{equation}
273: B^0 = -\frac{W}{g} \left(g_{0i} \mathcal{B}^i + g_{ij} \mathcal{B}^j
274:         V^i \right ) ~. \label{eqn:B0}
275: \end{equation}
276: The relativistic enthalpy is
277: %
278: \begin{equation}
279: h = 1 + \frac{\Gamma P}{(\Gamma-1) \rho} + \frac{Q}{\rho}~,
280: \end{equation}
281: %
282: where we have assumed an equation of state of the form
283: $P=(\Gamma-1)\rho \epsilon$. Finally, $P_B = \vert\vert B
284: \vert\vert^2/8\pi = g_{\mu\nu}B^\mu B^\nu/8\pi$ is the magnetic
285: pressure. We use the scalar $Q$ from \citet{ann05} with $k_q=2.0$
286: and $k_l=0.3$. We fix the divergence cleanser coefficients to be
287: $c_h = c_{\rm cfl} \Delta x_{\rm min}/\Delta t$ and $c_p^2 = c_h$,
288: where $c_{\rm cfl} = 0.7$ is the Courant coefficient, $\Delta x_{\rm
289: min}$ is the minimum covariant zone length, and $\Delta t$ is the
290: evolution timestep. For simplicity, we hold the timestep fixed at
291: $\Delta t = c_{\rm cfl} \Delta x_{\rm min}$ throughout the
292: simulation.
293: 
294: These GRMHD equations are evolved in a ``tilted'' Kerr-Schild polar
295: coordinate system $({t},{r},{\vartheta},{\varphi})$. This coordinate
296: system is related to the usual (untilted) Kerr-Schild coordinates
297: $({t},{r},{\theta},{\phi})$ through a simple rotation about the
298: ${y}$-axis by an angle $\beta_0$, such that
299: \begin{equation}
300: \left( \begin{array}{c} \sin{{\vartheta}}\cos{{\varphi}} \\
301:                         \sin{{\vartheta}}\sin{{\varphi}} \\
302:                         \cos{{\vartheta}} \end{array} \right)
303:  = \left( \begin{array}{ccc} \cos{\beta_0} & 0 & -\sin{\beta_0} \\
304:                                  0 & 1 & 0 \\
305:                   \sin{\beta_0} & 0 & \cos{\beta_0} \end{array} \right)
306:        \left( \begin{array}{c} \sin{{\theta}}\cos{{\phi}} \\
307:                         \sin{{\theta}}\sin{{\phi}} \\
308:                         \cos{{\theta}} \end{array} \right) ~.
309: \label{eqn:tiltarray}
310: \end{equation}
311: The full tilted metric terms are provided in \citet{fra05b} [see
312: also \citet{fra07b}]. The computational advantages of the
313: ``horizon-adapted'' Kerr-Schild form of the Kerr metric were first
314: described in \citet{pap98} and \citet{fon98b}. The primary advantage
315: is that, unlike Boyer-Lindquist coordinates, there are no
316: singularities in the metric terms at the event horizon, so the
317: computational mesh can extend into the hole's interior. In
318: principle, this should keep the inner boundary causally disconnected
319: from the flow, although numerically there is still some
320: communication.
321: 
322: The simulation is carried out on a spherical polar mesh with nested
323: resolution layers. The base grid contains $32^3$ mesh zones and
324: covers the full $4\pi$ steradians. Varying levels of refinement are
325: added on top of this base layer; each refinement level doubles the
326: resolution relative to the previous layer. The main simulation,
327: referenced as Model 915h, has two levels of refinement, thus
328: achieving a peak resolution equivalent to a $128^3$ simulation. For
329: comparison we also discuss results from an equivalent untilted
330: simulation (Model 90h) with the same resolution. As an argument that
331: our results are reasonably well converged, we also include results
332: from two other tilted simulations: one with a single refinement
333: layer and an equivalent resolution of $64^3$ (Model 915m) and
334: another that starts from a base grid of $24\times24\times32$ and
335: adds three layers of refinement for an equivalent resolution of
336: $192\times192\times256$ (Model 915vh). The evolution times for these
337: simulations differ as described below. In all cases, the full
338: refinement covers the region $r_{\rm min}\le r \le r_{\rm max}$,
339: $0.075\pi =\vartheta_1 \le \vartheta \le \vartheta_2 = 0.925\pi$, $0
340: \le \varphi \le 2\pi$, where $r_{\rm min}=0.98r_{\rm BH}=1.41r_G$
341: and $r_{\rm max}=120r_G$ are the inner and outer boundaries of the
342: grid, respectively, and $r_{\rm BH}=1.43r_G$ is the black-hole
343: horizon radius. The primary motivation for using a nested grid is to
344: allow us to maintain a reasonable Courant-limited timestep without
345: sacrificing any spatial resolution within the disk nor completely
346: excluding the region near the pole. The gain in computational
347: efficiency is significant since, for a polar mesh, the timestep
348: scales as $\Delta t \sim r_{\rm min} \sin \vartheta_{\rm min} \Delta
349: \varphi$. By underresolving the polar region, we gain by increasing
350: both $\vartheta_{\rm min}$ and $\Delta \varphi$. With 2 levels of
351: refinement, we are able to use a timestep that is a factor of 11.8
352: larger than what we could use if our most refined layer extended all
353: the way to the pole.
354: %Note, however, that any polar jets that are
355: %produced in the simulation may not be well resolved.
356: The main drawback of this approach is that we are unable to resolve
357: the region in which jets are expected to form.
358: 
359: In the radial direction we use a logarithmic coordinate of the form
360: $\eta \equiv 1.0 + \ln (r/r_{\rm BH})$. The spatial resolution near
361: the black-hole horizon is $\Delta r \approx 0.05 r_G$; near the
362: initial pressure maximum of the torus, the resolution is $\Delta r
363: \approx 0.5 r_G$. Both are considerably smaller than the initial
364: characteristic MRI wavelength $\lambda_\mathrm{MRI} \equiv 2\pi
365: v_\mathrm{A}/\Omega \approx 2.5 r_G$. This also gives us a large
366: number of zones inside the plunging region. In the angular
367: direction, in addition to the nested grids, we use a concentrated
368: latitude coordinate $x_2$ of the form $\vartheta = x_2 + \frac{1}{2}
369: (1 - h) \sin (2 x_2)$ with $h = 0.5$, which concentrates resolution
370: toward the midplane of the disk. As a result $r_{\rm center} \Delta
371: \vartheta = 0.3 r_G$ near the midplane while it is a factor of $\sim
372: 3$ larger for the fully refined zones near the pole. The grid used
373: in Models 915h and 90h is shown in Figure \ref{fig:grid}.
374: 
375: %\clearpage
376: \begin{figure}
377: %\begin{center}
378: %\includegraphics[scale=0.5]{torus3d.m.915h_grid.ps}
379: %\end{center}
380: \plotone{f1.eps} \caption{Plot of the grid geometry used for the
381: main simulation (Model 915h). The initial torus is aligned in the
382: symmetry plane of the grid, while the black hole is not.
383: \label{fig:grid}}
384: \end{figure}
385: %\clearpage
386: 
387: Since we cover the full $4\pi$ steradians, the only ``external''
388: boundaries are the inner and outer radial boundaries, where we apply
389: outflow conditions: Fluid variables are set the same in the external
390: boundary zone as in the neighboring internal zone, except for
391: velocity, which is chosen to satisfy
392: \begin{equation}
393: V^r_\mathrm{ext} = \left\{ \begin{array}{cc}
394:           V^r_\mathrm{int} & \mathrm{when~} V^r \mathrm{~points~off~the~grid}~, \\
395:          -V^r_\mathrm{int} & \mathrm{when~} V^r \mathrm{~points~onto~the~grid}~.
396:          \end{array} \right.
397: \end{equation}
398: In the azimuthal direction we apply periodic boundaries at
399: $\varphi=0$ and $2\pi$. Since Cosmos++ is a zone-centered code, we
400: do not have to treat the pole ($\vartheta=0$ or $\pi$) directly.
401: Instead unboosted scalar quantities, such as the gas pressure $P$,
402: in the ``ghost'' zones across the pole are filled with real data
403: from the corresponding zone located $180^\circ$ away in azimuth.
404: Unboosted vector quantities, such as velocity $V^i$, are similarly
405: filled with data from appropriate real zones, albeit with the signs
406: reversed for the $\vartheta$ and $\varphi$ components to maintain a
407: consistent sense of coordinate direction across the pole. Boosted
408: quantities, since they contain the metric determinant $\sqrt{-g}$,
409: are reflected across the pole so they extrapolate to zero there.
410: This treatment differs from the pure reflecting boundaries used in
411: other works \citep[e.g.][]{dev03c,mck06} in its treatment of the
412: unboosted variables. For untilted black holes the difference is
413: relatively minor. However, for tilted black holes, our approach
414: makes the pole more transparent to the fluid.
415: 
416: We initialize these simulations starting from the analytic solution
417: for an axisymmetric torus around a rotating black hole
418: \citep{cha85}. To provide a link with an untilted model already in
419: the literature, we start with identical torus conditions as model
420: KDP of \citet{dev03c}, which is the relativistic analog of model GT4
421: of \citet{haw00}. In our initialization, the torus is defined by:
422: the black-hole spacetime, specifically the spin of the black hole;
423: the inner radius of the torus $r_{in}$; the radius of the pressure
424: maximum of the torus $r_{\rm center}$; and the power-law exponent
425: $q$ used in defining the specific angular momentum distribution,
426: %
427: \begin{equation}
428: \ell = -u_\phi/u_t = k \Lambda^{2-q} ~.
429: \end{equation}
430: As in model KDP, $a/M=0.9$, $r_{\rm in}=15 r_G$, $r_{\rm center}=25
431: r_G$, and $q=1.68$. Knowledge of $r_{\rm center}$ leads directly to
432: a determination of $\ell_{\rm center}$ by setting it equal to the
433: geodesic value at that radius. The numerical value of $k$ comes
434: directly from the choice of $q$ and the determination of
435: $\Lambda_{\rm center}$, where
436: %
437: \begin{equation}
438: \frac{1}{\Lambda^2} = -\frac{g_{t \phi}+\ell g_{tt}}{\ell g_{\phi
439: \phi} + \ell^2 g_{t \phi}}~. \label{eqn:Lambda}
440: \end{equation}
441: Finally, having chosen $r_{in}$ we can obtain $u_{in}=u_t (r_{in})$,
442: the surface binding energy of the torus, from $u_t^{-2} =
443: g^{tt}-2\ell g^{t\phi} +\ell^2 g^{\phi\phi}$.
444: 
445: The solution of the torus variables can now be specified. The
446: internal energy of the torus is \citep{dev03c}
447: %
448: \begin{equation}
449: \epsilon(r,\theta) = \frac{1}{\Gamma} \left[ \frac{u_{in}
450: f(\ell_{in})}{u_t(r,\theta)f(\ell (r,\theta))} \right] ~,
451: \end{equation}
452: where $\ell_{in}=\ell(r_{in})$ is the specific angular momentum of
453: the fluid at the surface and
454: %
455: \begin{equation}
456: f(\ell) = \left|1-k^{2/n}\ell^\alpha\right|^{1/\alpha} ~,
457: \end{equation}
458: where $n=2-q$ and $\alpha=(2n-2)/n$. Assuming an isentropic equation
459: of state $P=\rho \epsilon(\Gamma-1)=\kappa \rho^\Gamma$, the density
460: is given by $\rho = \left[ \epsilon(\Gamma-1)/\kappa
461: \right]^{1/(\Gamma-1)}$. As in model KDP, we take $\Gamma=5/3$ and
462: $\kappa=0.01$ (arbitrary units). Finally, the angular velocity of
463: the fluid is specified by
464: %
465: \begin{equation}
466: \Omega = V^\phi = -\frac{g_{t\phi}+\ell g_{tt}}{g_{\phi \phi} + \ell
467: g_{t \phi}} ~.
468: \end{equation}
469: 
470: The dependence of $\Lambda$ on $\ell$ in equation (\ref{eqn:Lambda})
471: for Kerr black holes means that the solution requires an iterative
472: procedure. However, we can get an approximate solution by taking the
473: Schwarzschild form (i.e. ignoring $g_{t \phi}$)
474: %
475: \begin{equation}
476: \Lambda^2 = -\frac{g_{\phi \phi}}{g_{tt}} ~.
477: \end{equation}
478: The error introduced by doing so is small and only affects the
479: initial torus configuration, which will already be unstable to the
480: MRI due to the seed magnetic fields being added. Thus, this slightly
481: simplified treatment has no real consequence for the evolution. We
482: note that the same procedure is followed in \citet{dev03c}.
483: 
484: Once the torus is constructed, it is seeded with a weak magnetic
485: field in the form of poloidal loops along the isobaric contours
486: within the torus. The initial magnetic field vector potential is
487: \citep{dev03a}
488: \begin{equation}
489: A_\varphi = \left\{ \begin{array}{ccc}
490:           b(\rho-\rho_{\rm cut}) & \mathrm{for} & \rho\ge\rho_{\rm cut}~, \\
491:           0                  & \mathrm{for} & \rho<\rho_{\rm cut}~.
492:          \end{array} \right.
493: \label{eq:torusb}
494: \end{equation}
495: The non-zero spatial magnetic field components are then
496: $\mathcal{B}^r = - \partial_\vartheta A_\varphi$ and
497: $\mathcal{B}^\vartheta =
498: \partial_r A_\varphi$. The parameter $\rho_{\rm cut}=0.5*\rho_{\rm max,0}$ is used to keep the field a suitable
499: distance inside the surface of the torus, where $\rho_{\rm max,0}$
500: is the initial density maximum within the torus. Using the constant
501: $b$ in equation (\ref{eq:torusb}), the field is normalized such that
502: initially $\beta_{\rm mag} =P/P_B \ge \beta_{\rm mag,0}=10$
503: throughout the torus. This initialization is slightly different than
504: \citet{dev03b}, who use a volume integrated $\beta_{\rm mag}$ to set
505: the field strength; the difference is such that $\beta_{\rm
506: mag,0}=100$ in their work is roughly comparable to $\beta_{\rm
507: mag,0}=10$ here.
508: 
509: In the background region not specified by the torus solution, we set
510: up a rarefied non-magnetic plasma accreting into the black hole
511: \citep{kom06}. The density and pressure have the form
512: %
513: \begin{equation}
514: \rho = 10^{-3} \rho_{\rm max,0} \exp \left( \frac{-3r}{r_{\rm
515: center}} \right) ~~,~~ P = \kappa \rho^\Gamma ~.
516: \end{equation}
517: The radial velocity has the form
518: %
519: \begin{equation}
520: V^r = \frac{g^{tr}}{g^{tt}} \left[ 1 - \left( \frac{r_G}{r}
521: \right)^4 \right] ~.
522: \end{equation}
523: This introduces inflow through the horizon without creating large
524: velocity jumps at the torus surface. This background is initially
525: more dense than the static background used by \citet{dev03c}.
526: However, since this background reservoir is not replenished at the
527: outer boundary, it is rapidly depleted and has virtually no
528: long-term dynamical impact on the problem. Numerical floors are
529: placed on $\rho$ and $e$ at approximately $10^{-10}$ and $10^{-16}$
530: of their initial maxima, respectively. These floors are very seldom
531: applied once the initial background is replaced by evolved disk
532: material.
533: 
534: The final step of the initialization is to tilt the black hole by an
535: angle $\beta_0=15^\circ$ relative to the disk (and the grid) by
536: transforming the Kerr metric. The full transformation is provided in
537: \citet{fra05b} [see also \citet{fra07b}]. Thus, while the torus is
538: responding to the action of the MRI, it will also experience a
539: gravitomagnetic torque from the tilted black hole.
540: 
541: \section{Results}
542: \label{sec:results}
543: 
544: In the main simulation (915h) the torus is evolved for a total of 10
545: orbital periods ($10t_{\rm orb}$) as measured at $r=r_{\rm center}$,
546: which corresponds to $\sim 350$ orbits near $r_{\rm ISCO}=2.32 r_G$,
547: the coordinate radius of the innermost stable circular orbit (for
548: prograde orbits in the symmetry plane of the black hole). The very
549: high resolution simulation (915vh) is only run for half as long ($5
550: t_{\rm orb}$), while the lower resolution simulation (915m) is run
551: for twice as long ($20 t_{\rm orb}$). Figure \ref{fig:rho_sequence}
552: shows snapshots of the disk from Model 915h at times $t=0$, 1, 2, 4,
553: 7, and $10t_{\rm orb}$. The first orbit is dominated by winding of
554: the magnetic field lines and nonlinear growth of the MRI. Both of
555: these cause rapid redistributions of disk material and angular
556: momentum. The initial torus is stretched radially and material
557: begins to accrete onto the hole and is also carried out to large
558: radii. A strong current sheet forms in the initial symmetry plane of
559: the disk through differential winding.
560: 
561: %\clearpage
562: \begin{figure}
563: %\begin{center}
564: %\includegraphics[scale=0.4]{rho_0.eps}
565: %\includegraphics[scale=0.4]{rho_1.eps}
566: %\includegraphics[scale=0.4]{rho_2.eps}
567: %\includegraphics[scale=0.4]{rho_4.eps}
568: %\includegraphics[scale=0.4]{rho_7.eps}
569: %\includegraphics[scale=0.4]{rho_10.eps}
570: %\end{center}
571: \plotone{f2_comp.eps}\caption{Volume visualization of the logarithm
572: of density (scaled from $0.008 \rho_{\rm max,0}$ to $0.8 \rho_{\rm
573: max,0}$) at ({\em a}) $t=0$, ({\em b}) 1, ({\em c}) 2, ({\em d}) 4,
574: ({\em e}) 7, and ({\em f}) $10t_{\rm orb}$. Half of the disk has
575: been cut away to reveal the cross section. The black hole spin axis
576: is oriented vertically in each frame so that the initial torus is
577: tilted $15^\circ$ to the right. \label{fig:rho_sequence}}
578: \end{figure}
579: %\clearpage
580: 
581: From orbits 1-2, MRI driven turbulence begins to grow in the inner
582: parts of the disk. At the same time, some bending of the disk due to
583: the differential precession from the hole becomes apparent. The MRI
584: is fully developed through most of the disk around orbit 2.
585: 
586: By about orbit 7-8, the disk has reached a quasi-steady state. In
587: the remainder of this section we detail the properties of the
588: resultant structure. We follow an ``inside-out'' track, starting
589: from key features of the flow near the hole and working toward
590: larger radii. Where practical, we draw attention to similarities and
591: differences between the quasi-steady structure that results in this
592: simulation and the untilted simulations of \citet{dev03c}. In
593: particular, we draw attention to the fact that some features, such
594: as the inner torus and plunging region, are significantly altered,
595: while others, such as the main body and coronal envelope, show very
596: similar properties. Again, because of the varying levels of
597: refinement along the poles, we do not discuss the evacuated funnel
598: or funnel-wall jet in this paper.
599: 
600: \subsection{Global Structure}
601: 
602: \subsubsection{Plunging Streams}
603: \label{sec:plunging}
604: 
605: Perhaps the most striking feature in the tilted disk at late times
606: are the two opposing streams that start from high latitudes both
607: with respect to the black-hole symmetry plane and the disk midplane
608: \citep{fra07a}. Figure \ref{fig:plunge} shows a zoomed-in view of
609: the region around the black hole including these streams. Note that
610: stream 1 remains entirely above the black-hole symmetry plane, while
611: stream 2 remains below. Clearly the material in each stream is in a
612: plunging orbit into the black hole. Hence, we refer to these
613: features as the ``plunging streams.''
614: 
615: %\clearpage
616: \begin{figure}
617: %\begin{center}
618: %\includegraphics[scale=0.75]{rho_10_zoom.eps}
619: %\end{center}
620: \plotone{f3_comp.eps}\caption{Zoomed in view of the inner $10 r_G$
621: of the accretion flow revealing two opposing, high-latitude streams
622: of material connecting the disk to the horizon (indicated by
623: arrows). Data is taken from the last frame of the simulation ($t=10
624: t_{\rm orb}$). To emphasize the plunging streams, the scaling in
625: this figure is adjusted from that used in Fig.
626: \ref{fig:rho_sequence} by adding a density isosurface at $\rho =
627: 0.024 \rho_{\rm max,0}$. The figure is oriented as in Fig.
628: \ref{fig:rho_sequence} with the black-hole spin axis vertical. The
629: black-hole symmetry plane ({\em black line}) and initial disk
630: midplane ({\em blue line}) are marked for reference. Note that
631: stream 1 remains entirely above both planes while stream 2 remains
632: below. \label{fig:plunge}}
633: \end{figure}
634: %\clearpage
635: 
636: Figure \ref{fig:plunge2} captures the plunging streams from a
637: different perspective. This image is a view looking down the angular
638: momentum axis of the black hole onto a single isodensity surface.
639: The two opposing streams are clearly visible in the interior region
640: of the disk as well as two relatively evacuated lobes.
641: 
642: %\clearpage
643: \begin{figure}
644: %\begin{center}
645: %\includegraphics[scale=0.5]{torus3d.m.915h_rho_surf.eps}
646: %\end{center}
647: \plotone{f4_comp.eps}\caption{Isodensity contour at
648: $\rho=0.1\rho_{\rm max,0}$ from the same time slice as Fig.
649: \ref{fig:plunge} ($t=10t_{\rm orb}$) viewed down the angular
650: momentum axis of the black hole. The initial disk angular momentum
651: axis (and polar axis of the grid) is tilted $15^\circ$ to the right
652: in this image. One plunging stream (indicated by solid arrow) starts
653: near the left edge of the figure and connects to the hole on the
654: right. This stream lies entirely above the black-hole symmetry plane
655: and corresponds to stream 1 in Fig. \ref{fig:plunge}. The opposing
656: stream (stream 2) remains below the black-hole symmetry plane and is
657: seen connecting with the horizon on the left. \label{fig:plunge2}}
658: \end{figure}
659: %\clearpage
660: 
661: As material passes through the plunging streams it undergoes strong
662: differential precession. As we show below, the precession totals
663: approximately $180^\circ$, accounting for how the material in the
664: plunging streams is able to enter the black hole from the opposite
665: azimuth from which it began its plunge without ever passing through
666: the symmetry plane of the hole.
667: 
668: Two very important points to make about these streams is that they
669: appear to be stable and stationary. They begin forming as early as
670: $t=7t_{\rm orb}$ and last until the end of the simulation. During
671: this time their azimuthal location does not change appreciably. The
672: interesting questions are {\em why} do these opposing plunging
673: streams form and why do they start from such high latitude with
674: respect to the black-hole symmetry plane and disk midplane? The
675: answers, of course, are related and the fundamental cause is the
676: aspherical nature of the gravitational spacetime around the rotating
677: black hole. This is best illustrated by considering the dependence
678: of $r_{\rm ISCO}$ on inclination for orbits that are circular in the
679: sense that they have constant coordinate radius. Briefly, $r_{\rm
680: ISCO}$ is the radius at which the quantity
681: \begin{equation}
682: R\equiv A^2 \left(\frac{\mathrm{d}r}{\mathrm{d}\tau} \right)^2 =
683: \left[ E(r^2+a^2)-aL_z \right]^2 - \Delta \left[ r^2 + (L_z-aE)^2 +Q
684: \right] \label{eq:R}
685: \end{equation}
686: and its first two derivatives equal zero, i.e. $R=R'=R''=0$, where
687: $E$, $L_z$, and $Q$ are the energy, angular momentum, and Carter
688: constant, respectively, describing orbits around Kerr black holes
689: \citep{hug01} and $A = r^2 +a^2\cos^2 \theta$ and
690: $\Delta=r^2-2Mr+a^2$. Following \citet{hug01}, we can eliminate $Q$
691: in favor of the inclination $i$ defined as
692: \begin{equation}
693: \cos i = \frac{L_z}{(L_z+Q)^{1/2}} ~.
694: \end{equation}
695: Figure \ref{fig:isco} illustrates this dependence for a few selected
696: cases of $a$. The key point of the formula and the plot is that
697: orbital stability around a rotating black hole is strongly dependent
698: on the inclination of the orbit. Notice that the unstable region
699: increases monotonically for increasing inclination.
700: 
701: %\clearpage
702: \begin{figure}
703: %\begin{center}
704: %\includegraphics[scale=0.5]{isco2.eps}
705: %\end{center}
706: \plotone{f5.eps}\caption{Plot of the inclination dependence of
707: $r_{ISCO}$ for black-hole spins $a=0$, 0.5, 0.9, and 0.998.
708: Inclinations $0 \le i \le 90^\circ$ represent prograde orbits,
709: whereas inclinations $90^\circ \le i \le 180^\circ$ represent
710: retrograde orbits. \label{fig:isco}}
711: \end{figure}
712: %\clearpage
713: 
714: We can make better use of the information in Figure \ref{fig:isco}
715: by converting it to a polar plot (using only the prograde orbits)
716: and overlaying it onto a plot of data from the simulation, as is
717: done in Figure \ref{fig:rhoIsco}. Such a polar plot creates a
718: representation of the prograde ``ISCO surface'' (symmetric about the
719: spin axis of the black hole), which gives a clear indication of
720: where the most unstable regions of the spacetime are. Note that the
721: plunging orbits highlighted previously start near where the disk
722: first encounters the ISCO surface. More precisely, the streams start
723: near the largest cylindrical radius ($r\cos\vartheta$) of the ISCO
724: surface, measured with respect to the angular momentum axis of the
725: disk. This explains why the plunging streams start at such high
726: inclinations relative to the black-hole symmetry plane and the disk
727: midplane and why there are only two streams. The plunging region is
728: no longer azimuthally symmetric from the perspective of the disk.
729: 
730: %\clearpage
731: \begin{figure}
732: %\begin{center}
733: %\includegraphics[scale=0.5]{torus3d.m.915h_rho_slice.eps}
734: %\end{center}
735: \plotone{f6.eps}\caption{Meridional plot ($\varphi=0$) through the
736: final dump ($t=10 t_{\rm orb}$) of the simulation showing a
737: pseudocolor representation of the logarithm of density (scaled from
738: $0.008 \rho_{\rm max,0}$ to $0.8 \rho_{\rm max,0}$ as in previous
739: figures) and an isocontour of density at $\rho = 0.024 \rho_{\rm
740: max,0}$ (red curve). Unlike previous figures, this one is shown
741: oriented in the sense of the grid, so that the black hole is tilted
742: $15^\circ$ to the left. The plot is overlaid with a polar plot of
743: the ``ISCO surface'' for prograde orbits about an $a=0.9$ black hole
744: (white curve). This surface is symmetric about the spin axis of the
745: hole. Notice that the plunging streams from Figs. \ref{fig:plunge}
746: and \ref{fig:plunge2} start near the largest cylindrical radius
747: ($r\cos\vartheta$) of this surface (indicated by white arrows) and
748: connect with the horizon approximately $180^\circ$ away in azimuth
749: (indicated by black arrows). \label{fig:rhoIsco}}
750: \end{figure}
751: %\clearpage
752: 
753: Another point to take away from Figures \ref{fig:isco} and
754: \ref{fig:rhoIsco} is that $r_{\rm ISCO}$ is larger for larger
755: inclinations. Thus, for a given black-hole spin, plunging orbits
756: will always start further away from the hole for more tilted disks.
757: The tilted black hole effectively acts like an untilted black hole
758: of lower spin, which would likewise have a larger $r_{\rm ISCO}$.
759: 
760: \subsubsection{Inner Torus}
761: 
762: In our tilted simulation, the plunging streams appear to connect
763: directly to the main disk body without a clearly identifiable
764: intermediate ``inner torus''. This appears to be a particular result
765: of the tilted simulation and not, for instance, due to the
766: differences in the coordinates used in our simulation (Kerr-Schild)
767: versus those used in \citet{dev03c} (Boyer-Lindquist) or numerical
768: techniques. We base this statement on the fact that our own untilted
769: simulation in Kerr-Schild coordinates shows an inner torus very
770: similar to the one described in \citet{dev03c}. For instance, Figure
771: \ref{fig:avgdensity} shows the shell-averaged density and pressure
772: as a function of radius for our tilted and untilted simulations.
773: Shell averaged quantities are computed over the most refined grid as
774: follows:
775: %
776: \begin{equation}
777: \langle\mathcal{Q}\rangle_A(r,t) = \frac{1}{A} \int^{2\pi}_0
778: \int^{\vartheta_2}_{\vartheta_1} \mathcal{Q} \sqrt{-g}
779: \mathrm{d}\vartheta \mathrm{d}\varphi ~,
780: \end{equation}
781: %
782: where $A = \int^{2\pi}_0 \int^{\vartheta_2}_{\vartheta_1} \sqrt{-g}
783: \mathrm{d}\vartheta \mathrm{d}\varphi$ is the surface area of the
784: shell. The data in Figure \ref{fig:avgdensity} has also been
785: time-averaged over the final orbit, $9t_{\rm orb} = t_{\rm min} \le
786: t \le t_{\rm max} = 10t_{\rm orb}$, where time averages are defined
787: as
788: %
789: \begin{equation}
790: \langle\mathcal{Q}\rangle_t = \frac{1}{t_{\rm max} - t_{\rm min}}
791: \int^{t_{\rm max}}_{t_{\rm min}} \mathcal{Q} \mathrm{d}t ~.
792: \end{equation}
793: %
794: In the untilted simulation, both the density and the pressure show
795: local maxima near $4.5 r_G$, indicating an inner torus. The tilted
796: simulation, on the other hand, shows only marginal evidence for
797: local maxima near $10 r_G$.
798: 
799: %\clearpage
800: \begin{figure}
801: %\begin{center}
802: %\includegraphics[scale=0.3]{torus3d.m.915h_Rho_av.eps}
803: %\includegraphics[scale=0.3]{torus3d.m.90h_Rho_av.eps}
804: %\end{center}
805: \plottwo{f7a.eps}{f7b.eps}\caption{Plot of
806: $\langle\langle\rho\rangle_A\rangle_t$ ({\em solid line}) and
807: $\langle\langle P\rangle_A\rangle_t$ ({\em dashed line}) as a
808: function of radius for equivalent ({\em a}) tilted
809: $\beta_0=15^\circ$ (915h) and ({\em b}) untilted $\beta_0=0^\circ$
810: (90h) simulations. For both simulations, the data has been
811: time-averaged over the interval $t=9$ to $10t_{\rm orb}$. The
812: density and pressure have been normalized by their respective maxima
813: at $t=0$, which are the same in both simulations.
814: \label{fig:avgdensity}}
815: \end{figure}
816: %\clearpage
817: 
818: Another check of the presence of an inner torus is to look at the
819: distribution of specific angular momentum in the disk. Because the
820: inner torus is partially supported by pressure gradients, some
821: portion of the flow must be locally super-geodesic. In Figure
822: \ref{fig:angmom} we plot the density-weighted shell average of the
823: specific angular momentum $\langle\ell\rangle_A = \langle\rho
824: \ell\rangle_A/\langle\rho\rangle_A$ as a function of radius, again
825: time-averaged over the interval $t=9$ to $10t_{\rm orb}$. We compare
826: this against the specific angular momentum distribution of circular
827: orbits with inclinations of $15^\circ$ and $0^\circ$. These are
828: calculated from the following expression
829: %
830: \begin{equation}
831: \ell = \frac{N_1 + \Delta (Mr)^{1/2} N_2^{1/2} \cos i}{D} ~,
832: \end{equation}
833: %
834: where
835: %
836: \begin{equation}
837: N_1 = -aMr \left(3r^2 + a^2 - 4Mr \right) \cos^2 i ~,
838: \end{equation}
839: %
840: \begin{equation}
841: N_2 = r^4 + a^2 \sin^2 i \left(a^2 + 2r^2 - 4Mr \right) ~,
842: \end{equation}
843: %
844: and
845: %
846: \begin{equation}
847: D = a^2 \left( 2r^2 + a^2 - 3Mr \right) \sin^2 i + r^4 + 4M^2r^2 -
848: 4r^3M - Mra^2 ~,
849: \end{equation}
850: %
851: which comes from noting that for circular orbits $R=R'=0$ from
852: equation (\ref{eq:R}) and from the definition $\ell=L_z/E$. Both
853: simulations show a nearly geodesic angular momentum distribution
854: through most of the disk with a small region of super-geodesic flow
855: inside $10 r_G$. This region clearly corresponds to the inner torus
856: in the untilted simulation. It also suggests that there should be an
857: inner torus in the tilted simulation, though, again, this is not as
858: evident in the plots of density and pressure.
859: 
860: %\clearpage
861: \begin{figure}
862: %\begin{center}
863: %\includegraphics[scale=0.3]{torus3d.m.915h_Ell_av.eps}
864: %\includegraphics[scale=0.3]{torus3d.m.90h_Ell_av.eps}
865: %\end{center}
866: \plottwo{f8a.eps}{f8b.eps}\caption{Plot of the density-weighted
867: time- and shell-averaged specific angular momentum
868: $\langle\langle\ell\rangle_A\rangle_t$ ({\em thick line}) as a
869: function of radius for equivalent ({\em a}) tilted
870: $\beta_0=15^\circ$ (915h) and ({\em b}) untilted $\beta_0=0^\circ$
871: (90h) simulations. For both simulations, the data has been
872: time-averaged over the interval $t=9$ to $10t_{\rm orb}$. In each
873: plot a comparison is provided with the specific angular momentum of
874: circular orbits with inclinations of $15^\circ$ and $0^\circ$,
875: respectively ({\em dashed line}). For reference we also include the
876: initial angular momentum distribution in the midplane of the torus
877: ({\em thin line}). \label{fig:angmom}}
878: \end{figure}
879: %\clearpage
880: 
881: Another indication that the inner torus is less prominent in the
882: tilted simulation than the untilted one comes from comparing the
883: total rest mass in the near-hole region ($r<r_{\rm cut} = 10r_G$).
884: This is done in Figure \ref{fig:torusmass}, where we plot the time
885: histories of the total (volume-integrated) rest mass
886: %
887: \begin{equation}
888: \left\langle\rho u^0\right\rangle_V = \int^{2\pi}_0 \int^{\pi}_0
889: \int^{r_{\rm cut}}_{r_{\rm min}} D \mathrm{d}r \mathrm{d}\vartheta
890: \mathrm{d}\varphi ~.
891: \end{equation}
892: %
893: At $t=10t_{\rm orb}$, the inner torus is 42\% less massive in Model
894: 915h.
895: 
896: %\clearpage
897: \begin{figure}
898: %\begin{center}
899: %\includegraphics[scale=0.6]{Massvst.eps}
900: %\end{center}
901: \plotone{f9.eps}\caption{Total rest mass in the near-hole region
902: ($r<10r_G$) as a function of time for the tilted (915h \& 915vh) and
903: untilted (90h) simulations. The mass and time are normalized by the
904: initial mass and orbital period of the torus, respectively.
905: \label{fig:torusmass}}
906: \end{figure}
907: %\clearpage
908: 
909: When present, the inner torus usually performs two functions:
910: regulating the accretion of matter into the black hole and serving
911: as the launching point for the funnel-wall jet. Therefore, we may
912: expect a weaker funnel-wall jet (to be discussed in future work) and
913: a higher mass accretion rate in our tilted-disk simulation relative
914: to the untilted simulation due to the less prominent inner torus in
915: the former. We compute the mass accretion rate
916: \begin{equation}
917: \dot{M}(r) = \int^{2\pi}_0 \int^\pi_0 D V^r \mathrm{d}\vartheta
918: \mathrm{d}\varphi
919: \end{equation}
920: 100 times per $t_{\rm orb}$ (about every $8M$) at each of the
921: external grid boundaries. Figure \ref{fig:accretion}{\em a} shows a
922: plot comparing $\dot{M}(r_{\rm min})$ for our equivalent tilted and
923: untilted simulations. When averaged over the quasi-steady state of
924: each simulation (from $t=7$ to $10 t_{\rm orb}$),
925: $\langle\dot{M}\rangle_t$ into the hole for the tilted simulation
926: (915h) is $7.2\times10^{-6}$, while for the untilted one (90h), it
927: is $4.9\times10^{-6}$. There is a clear tendancy toward a higher
928: $\dot{M}$ in the tilted-disk simulation.
929: 
930: %\clearpage
931: \begin{figure}
932: %\begin{center}
933: %\includegraphics[scale=0.3]{massFluxA.eps}
934: %\includegraphics[scale=0.3]{massFluxB.eps}
935: %\end{center}
936: \plottwo{f10a.eps}{f10b.eps}\caption{({\em a}) Plot of the mass
937: accretion history from Model 915h with $\beta_0=15^\circ$ ({\em
938: thick line}) and Model 90h with $\beta_0=0^\circ$ ({\em thin line}).
939: The accretion rate and time are normalized by the initial mass and
940: orbital period of the torus, respectively. ({\em b}) Plot of mass
941: accretion rate, comparing our medium (915m), high (915h), and very
942: high (915vh) resolution tilted disk simulations. The very high
943: resolution simulation was only run to $t=5t_{\rm orb}$.
944: \label{fig:accretion}}
945: \end{figure}
946: %\clearpage
947: 
948: Figure \ref{fig:accretion}{\em b} compares $\dot{M}$ of the tilted
949: disk simulation at three different resolutions. Due to the chaotic
950: nature of the mass accretion we do not expect the individual peaks
951: to match; yet we are encouraged that the overall shape and magnitude
952: of the two high-resolution models (915h and 915vh) are very
953: consistent, suggesting we are reasonably well converged. The medium
954: resolution simulation (Model 915m), on the other hand, is clearly
955: underresolved.
956: 
957: 
958: \subsubsection{Main Disk Body \& Coronal Envelope}
959: 
960: The main disk body does not differ substantially between the tilted
961: and untilted simulations, except in the notable fact that the tilted
962: disk precesses (as discussed in \S \ref{sec:precession} below).
963: Likewise, the coronal envelope, which extends above and below the
964: disk, shows very similar properties in all our simulations.
965: % except in the circumpolar region which we describe next.
966: The material in the coronal envelope is characterized by low density
967: and rough magnetic equipartition ($\beta_{\rm mag} \approx 1$). By
968: contrast the main body of the disk is generally gas-pressure
969: dominated ($\beta_{\rm mag} \ll 1$). Therefore, a plot of
970: $\beta_{\rm mag}$ and $\rho$, such as Figure \ref{fig:beta_mag},
971: provides a convenient means to identify these two regions. As found
972: in \citet{dev03c}, the material in the coronal envelope moves mostly
973: radially outward, yet has ($-hu_t <1$). This suggests that the
974: material may be gravitationally bound, in which case it must
975: circulate back to the disk at large radii. However, we point out
976: that this definition of binding energy ignores the contribution of
977: the magnetic fields, so some of this material may in fact escape the
978: system. We plan to examine outflows from tilted disks more
979: thoroughly in future work.
980: 
981: %\clearpage
982: \begin{figure}
983: %\begin{center}
984: %\includegraphics[scale=0.5]{torus3d.m.915h_beta_slice.eps}
985: %\end{center}
986: \plotone{f11.eps}\caption{Azimuthal slice through the simulation
987: along $\varphi=0$ taken from the final dump ($t=10 t_{\rm orb}$).
988: The ratio of magnetic pressure to gas pressure ($\beta_{\rm
989: mag}^{-1}$) is represented as a pseudocolor plot. The colors are
990: scaled logarithmically and cover the range $10^{-2} \le \beta_{\rm
991: mag} \le 10^2$. The gas density is given by isocontours at $\rho =
992: 10^{-2}$, $10^{-1.5}$, $10^{-1}$, and $10^{-0.5} \rho_{\rm max,0}$.
993: As with Fig. \ref{fig:rhoIsco}, this figure is oriented in the sense
994: of the grid, so that the black hole is tilted $15^\circ$ to the
995: left. The apparent tilt of the disk is actually due its precession
996: about the black-hole spin axis, such that the angular momentum axis
997: of the disk is no longer in the plane of this image; the disk has
998: not actually realigned with the hole. We remind the reader that the
999: region near the poles is not sufficiently resolved, so caution
1000: should be used when interpreting results there.
1001: \label{fig:beta_mag}}
1002: \end{figure}
1003: %\clearpage
1004: 
1005: Because the disk is precessing, its angular momentum axis does not
1006: remain aligned with the grid. Therefore, an azimuthal slice through
1007: the disk at late times, such as Figure \ref{fig:beta_mag}, may give
1008: the impression that the disk has aligned with the symmetry plane of
1009: the black hole when indeed this is not the case. We now turn to the
1010: question of disk alignment and precession.
1011: 
1012: \subsection{Results Specific to A Tilted Disk}
1013: 
1014: \subsubsection{Tilt}
1015: One key diagnostic for describing the global response of a tilted
1016: disk subject to Lense-Thirring precession is the tilt between the
1017: angular momenta of the black hole and disk as a function of radius
1018: and time. For example, in the Bardeen-Petterson solution, no time
1019: variability is observed, and the tilt transitions from nearly zero
1020: close to the black hole to a non-zero asymptote at large radii.
1021: 
1022: As in \citet{fra05b}, we recover the tilt from the simulation data
1023: using the definition
1024: \begin{equation}
1025: \beta(r) = \arccos\left[ \frac{\mathbf{J}_{\rm BH} \cdot
1026: \mathbf{J}_{\rm Disk}(r)} {\vert\mathbf{J}_{\rm BH} \vert
1027: \vert\mathbf{J}_{\rm Disk}(r) \vert} \right] ~,
1028: \end{equation}
1029: where
1030: \begin{equation}
1031: \mathbf{J}_{\rm BH} = \left( -a M \sin\beta_0 \hat{x}, 0, a M
1032: \cos\beta_0 \hat{z} \right)
1033: \end{equation}
1034: is the angular momentum vector of the black hole and
1035: \begin{equation}
1036: \mathbf{J}_{\rm Disk}(r) = \left[ (J_{\rm Disk})_1 \hat{x}, (J_{\rm Disk})_2
1037: \hat{y}, (J_{\rm Disk})_3 \hat{z} \right]
1038: \end{equation}
1039: is the angular momentum vector of the disk in an asymptotically flat
1040: space. This is given by
1041: \begin{equation}
1042: (J_{\rm Disk})_\rho = \frac{\epsilon_{\mu \nu \sigma \rho}
1043:                          L^{\mu \nu} S^\sigma}
1044:                        {2 \sqrt{-S^\alpha S_\alpha}} ~,
1045: \end{equation}
1046: where
1047: \begin{equation}
1048: L^{\mu \nu} = \int \left( x^\mu T^{\nu 0} - x^\nu T^{\mu 0} \right)
1049: \mathrm{d}^3 x ,
1050: \end{equation}
1051: and $S^\sigma = \int T^{\sigma 0} \mathrm{d}^3 x$. The equations for
1052: $L^{\mu \nu}$ and $S^\sigma$ are integrated over concentric radial
1053: shells of the most-refined grid layer, e.g.
1054: %
1055: \begin{equation}
1056: S^\sigma (r) = \int^{2\pi}_0 \int^{\vartheta_2}_{\vartheta_1}
1057: T^{\sigma 0} \sqrt{-g} \Delta r \mathrm{d}\vartheta
1058: \mathrm{d}\varphi ~.
1059: \end{equation}
1060: %
1061: The unit vector $\hat{y}$ points along the axis about which the
1062: black hole is initially tilted and $\hat{z}$ points along the
1063: initial angular momentum axis of the disk.
1064: 
1065: In Figure \ref{fig:beta}, we show the radial profile of $\beta$ time
1066: averaged over the interval $9 t_{\rm orb} \le t \le 10 t_{\rm orb}$.
1067: Recall $\beta_0=15^\circ$ for this simulation. This profile remains
1068: fairly consistent over many orbital times once the quasi-steady
1069: state is reached, so the time-averaged data gives a good
1070: representation for all $t\gtrsim 7 t_{\rm orb}$. The variability
1071: from this time-averaged profile is generally $\lesssim 20\%$ and is
1072: generally carried by moderate amplitude waves traveling through the
1073: disk. The increase in tilt at $r \lesssim 10r_G$ is attributable to
1074: the high latitude plunging streams described in
1075: \S\ref{sec:plunging}.
1076: 
1077: %\clearpage
1078: \begin{figure}
1079: %\begin{center}
1080: %\includegraphics[scale=0.35]{BetavsR_av.eps}
1081: %\end{center}
1082: \plotone{f12.eps}\caption{Plot of the tilt $\langle{\beta}\rangle_t$
1083: as a function of radius through the disk. The data for this plot has
1084: been time averaged from $t=9$ to $10t_{\rm orb}$. The initial tilt
1085: was $\beta_0=15^\circ$. \label{fig:beta}}
1086: \end{figure}
1087: %\clearpage
1088: 
1089: One very obvious characteristic of the profile in Figure
1090: \ref{fig:beta} is that $\beta$ does {\em not} approach zero except
1091: perhaps very close to the hole. Thus we do not see evidence for the
1092: Bardeen-Petterson effect in this simulation. This is not surprising
1093: since the Bardeen-Petterson solution is only expected for thin disks
1094: ($H/r < \alpha$). This is not the applicable regime for this
1095: simulation, as we illustrate in Figure \ref{fig:stress}, which shows
1096: $H/r$ and $\alpha$ plotted as functions of $r$. The scale height
1097: $H(r)$ is defined in each radial shell as one-half the distance
1098: ($0.5 r \Delta \vartheta$) between the two points where $\rho =
1099: \rho_{max}/e$, where we use the time-averaged density along the
1100: $\varphi=0$ azimuthal slice. The dimensionless stress parameter
1101: $\alpha$ in the disk is taken to be
1102: %
1103: \begin{equation}
1104: \alpha = \left\langle \frac{ \vert u^r u^\varphi \vert\vert B
1105: \vert\vert^2 - B^r B^\varphi \vert}{4 \pi P} \right\rangle_A ~.
1106: \end{equation}
1107: We restrict the calculation of $\alpha$ to only bound material
1108: ($-hu_t <1$). Using these definitions we find $H/r \sim 0.2$ and
1109: $\alpha \lesssim 0.01$ through most of the disk.
1110: 
1111: %\clearpage
1112: \begin{figure}
1113: %\begin{center}
1114: %\includegraphics[scale=0.35]{scale_height.eps}
1115: %\end{center}
1116: \plotone{f13.eps}\caption{Plot of the scale height $\langle
1117: H\rangle_t/r$ and magnetic stress parameter
1118: $\langle\alpha\rangle_t$, time averaged over the interval $7 t_{\rm
1119: orb} \le t \le 10 t_{\rm orb}$. This plot shows that this simulation
1120: falls into the thick-disk limit $H/r > \alpha$. \label{fig:stress}}
1121: \end{figure}
1122: %\clearpage
1123: 
1124: Since warps in slim disks are expected to propagate as bending
1125: waves, it may seem unusual at first that we see little evidence for
1126: such waves in Figure \ref{fig:beta}. For instance, \citet{lub02}
1127: provides an analysis of the theory of bending waves in nearly
1128: Keplerian, weakly inclined disks and predicts that the tilt $\beta$
1129: should be a {\em time-independent, oscillatory} function of radius
1130: \citep[see also][]{mar98}. However, using equation (16) of
1131: \citet{lub02}, we estimate the wavelength of such oscillations for
1132: our simulation to be
1133: %
1134: \begin{equation}
1135: \lambda \approx \frac{\pi r^{9/4}}{(6a)^{1/2}} \left( \frac{H}{r}
1136: \right) \sim 50 M
1137: \end{equation}
1138: at $r=10 r_G$. This is strongly radially dependent ($\lambda \propto
1139: r^{9/4}$ with $H/r\sim\mathrm{constant}$), so oscillations of
1140: $\beta$ are essentially absent outside $r=10 r_G$, consistent with
1141: what is shown in Figure \ref{fig:beta}.
1142: 
1143: The same conclusion, that $\beta$ is not expected to oscillate
1144: outside $r=10 r_G$ for this simulation, is also reached by
1145: considering equation (22) of \citet{lub02}. That equation defines a
1146: dimensionless variable
1147: %
1148: \begin{equation}
1149: x=\left( \frac{24a}{\epsilon^2}\right)^{1/2}
1150: \frac{r^{-(h+1/4)}}{h+1/4} ~,
1151: \end{equation}
1152: which is used to identify the transition radius between oscillatory
1153: behavior and asymptotic behavior, where $h$ and $\epsilon$ are used
1154: to parameterize the radial dependence of the disk scale height
1155: $H/r=\epsilon r^{h-1}$. Whenever $x>>1$ (small $r$), oscillations
1156: should be prominent, whereas whenever $x<<1$ (large $r$), $\beta$
1157: tends to the outer boundary value. For our simulation, with
1158: $\epsilon \approx 0.2$ and $h\approx1$, $x=1$ at $r\approx10 r_G$.
1159: Thus, from both approaches, it is clear that our simulation does not
1160: satisfy the criteria to develop large oscillations in $\beta$ within
1161: the main body of the disk.
1162: 
1163: Inside $r=10 r_G$, the density of the disk drops off rapidly and the
1164: dynamics are dominated by the plunging streams, which are not
1165: accounted for in the model of \citet{lub02}. Nevertheless, we appear
1166: to capture one-half of one wavelength of a bending wave oscillation
1167: inside $r=10r_G$, based on Figure \ref{fig:beta}. Thus, overall our
1168: results seem to be generally consistent with the predictions of
1169: \citet{lub02}.
1170: 
1171: 
1172: \subsubsection{Precession}
1173: \label{sec:precession}
1174: 
1175: A second useful diagnostic for tilted disks is the twist $\gamma$ of
1176: the disk as a function of radius and time. We define the precession
1177: angle (twist) as
1178: 
1179: %
1180: \begin{equation}
1181: \gamma(r) = \arccos\left[ \frac{\mathbf{J}_{\rm BH} \times
1182: \mathbf{J}_{\rm Disk}(r)} {\vert \mathbf{J}_{\rm BH} \times
1183: \mathbf{J}_{\rm Disk}(r) \vert} \cdot \hat{y}\right] ~, \label{eq:twist}
1184: \end{equation}
1185: From this definition, $\gamma(r) = 0$ throughout the disk at $t=0$.
1186: In order to capture twists larger than $180^\circ$, we also track
1187: the projection of $\mathbf{J}_{\rm BH} \times \mathbf{J}_{\rm Disk}(r)$ onto
1188: $\hat{x}$, allowing us to break the degeneracy in $\arccos$. A
1189: time-averaged plot of $\gamma$ is provided in Figure
1190: \ref{fig:gamma}.
1191: 
1192: %\clearpage
1193: \begin{figure}
1194: %\begin{center}
1195: %\includegraphics[scale=0.35]{GammavsR_av.eps}
1196: %\end{center}
1197: \plotone{f14.eps}\caption{Plot of the twist $\langle\gamma\rangle_t$
1198: as a function of radius through the disk. The data for this plot has
1199: been time averaged from $t=9$ to $10t_{\rm orb}$. Initially the
1200: twist was zero throughout the disk. The disk matter has precessed
1201: roughly $\sim 180^\circ$ by the time it reaches the hole. The shape
1202: of this twist profile remains fairly constant throughout the
1203: simulation. \label{fig:gamma}}
1204: \end{figure}
1205: %\clearpage
1206: 
1207: As described in our previous work \citep{fra05b}, we expect
1208: differential Lense-Thirring precession to dominate whenever the
1209: precession timescale $t_{\rm LT} = \Omega_{\rm LT}^{-1} =
1210: g^{tt}/g^{t\phi}$ is shorter than local dynamical timescales in the
1211: disk \citep{bar75,kum85}. We consider three possible limiting
1212: timescales: the mass accretion timescale $t_{\rm acc} =
1213: r/\overline{V}^r$, where $\overline{V}^r=\langle \langle \rho V^r
1214: \rangle_A/\langle \rho \rangle_A \rangle_t$ is the density-weighted
1215: average inflow velocity; the sound-crossing time
1216: $t_{cs}=r/\overline{c}_s$, where $\overline{c}_s = \langle \langle
1217: \rho c_s \rangle_A/\langle \rho \rangle_A \rangle_t$ is a
1218: density-weighted average of the local sound speed; and the Alfv\'en
1219: crossing time $t_{A}=r/\overline{V}_A$, where $\overline{V}_A$ is a
1220: density-weighted average of the local Alfv\'en speed. The local
1221: sound speed is recovered from the fluid state through the relation
1222: $c_s^2=\Gamma (\Gamma-1)P/[(\Gamma-1)\rho + \Gamma P]$. The Alfv\'en
1223: speed is
1224: %
1225: \begin{equation}
1226: v_A =  \sqrt{\frac{\vert\vert B \vert\vert^2}{4 \pi \rho h +
1227: \vert\vert B \vert\vert^2}} ~.
1228: \end{equation}
1229: %
1230: Since $c_s$ and $v_A$ are defined in the frame of the fluid, it is
1231: not strictly accurate to compare $t_{cs}$ and $t_A$ to quantities
1232: defined using the coordinate time (such as $t_{\rm LT}$ and
1233: $\Omega^{-1}$). However, we are mostly concerned with the timescales
1234: in the main body of the disk where such discrepancies are small.
1235: From Figure \ref{fig:timescales}, we can see that the Lense-Thirring
1236: precession timescale is longer than the sound-crossing time at
1237: virtually all radii.
1238: 
1239: %\clearpage
1240: \begin{figure}
1241: %\begin{center}
1242: %\includegraphics[scale=0.35]{timescales.eps}
1243: %\end{center}
1244: \plotone{f15.eps}\caption{Plot comparing various timescales within
1245: the disk, including the Lense-Thirring precession timescale $t_{\rm
1246: LT}$, the accretion timescale $t_{\rm acc}$, the sound-crossing time
1247: $t_{\rm cs}$, and the Alfv\'en crossing time $t_{\rm A}$. All
1248: timescales are normalized by the local orbital period in the
1249: midplane of the black hole, $\Omega^{-1}$. The data for this plot
1250: has been time averaged from $t=9t_{\rm orb}$ to $t=10t_{\rm orb}$.
1251: \label{fig:timescales}}
1252: \end{figure}
1253: %\clearpage
1254: 
1255: Since the sound-crossing time is short compared to the precession
1256: timescale throughout the bulk of the disk, pressure waves strongly
1257: couple the disk material. The disk, thus, responds as a single
1258: entity to the torque of the black-hole and precesses as a global
1259: structure. Such global precession has been noted before in low Mach
1260: number hydrodynamic disks \citep{nel00,fra05b}. To estimate the
1261: precession period, we have plotted $\gamma$, averaged over the bulk
1262: of the disk ($20 \le r/r_G \le 50$), as a function of time in Figure
1263: \ref{fig:precession}. A linear fit to this plot yields a precession
1264: period of $T_{\rm prec} \approx 0.3 (M/M_\odot)$ s, which
1265: corresponds to about $80t_{\rm orb}$. This is longer than the
1266: evolution time of all of our models, so we have had to extrapolate
1267: the full precession period. However, Model 915m is run to $20t_{\rm
1268: orb}$ and shows a nearly linear growth of precession over the full
1269: simulation.
1270: 
1271: %\clearpage
1272: \begin{figure}
1273: %\begin{center}
1274: %\includegraphics[scale=0.35]{gammavst.eps}
1275: %\end{center}
1276: \plotone{f16.eps}\caption{Plot of the twist $\gamma$, averaged over
1277: the bulk of the disk ($20 \le r/r_G \le 50$), as a function of time.
1278: The slope of this plot can be used to estimate the precession period
1279: of the disk as a whole, which is $0.3 (M/M_\odot)$ s.
1280: \label{fig:precession}}
1281: \end{figure}
1282: %\clearpage
1283: 
1284: Classically, we expect the precession period for a solid-body
1285: rotator with angular momentum $J$ subject to a torque $\tau$ to be
1286: $T_{\rm prec} = 2\pi (\sin \beta) (J/\tau)$ \citep{liu02}. Assuming
1287: a radial dependence to the surface density of the form
1288: $\Sigma=\Sigma_i (r/r_i)^{-\zeta}$ and ignoring higher order general
1289: relativistic corrections, we have $J=2\pi M^{1/2} \Sigma_i r_i^\zeta
1290: r_0^{5/2-\zeta}[1-(r_i/r_o)^{5/2-\zeta}]/(5/2-\zeta)$ and $\tau=4\pi
1291: (\sin \beta) aM^{3/2}
1292: \Sigma_i[1-(r_i/r_o)^{1/2+\zeta}]/[r_i^{1/2}(1/2+\zeta)]$, where
1293: $r_i$ and $r_o$ are the inner and outer radii of the evolved disk,
1294: respectively. Therefore,
1295: %
1296: \begin{equation}
1297: T_{\rm prec} = \frac{\pi (1+2\zeta)}{(5-2\zeta)}
1298: \frac{r_o^{5/2-\zeta} r_i^{1/2+\zeta} \left[1-(r_i/r_o)^{5/2-\zeta}
1299: \right]} { aM \left[1-(r_i/r_o)^{1/2+\zeta}\right]} ~.
1300: \label{eqn:precession}
1301: \end{equation}
1302: %
1303: For $r_i=10 r_G$, $r_o=50 r_G$, and $\zeta=0$ (the value we find in
1304: our simulation), equation (\ref{eqn:precession}) predicts $T_{\rm
1305: prec}=0.3 (M/M_\odot)$ s, which is the same as the observed value in
1306: the simulation. Note that equation (\ref{eqn:precession}) differs
1307: from the test particle Lense-Thirring precession period because
1308: $T_{\rm prec}$ depends on the total torque integrated over the
1309: entire disk.
1310: 
1311: 
1312: \section{Discussion}
1313: \label{sec:discussion}
1314: 
1315: In this paper we studied the evolution of an MRI turbulent disk that
1316: was tilted with respect to the spin axis of a modestly fast rotating
1317: black hole. Although this prescription can lead to a
1318: Bardeen-Petterson configuration for some disk parameters, we did not
1319: see evidence for this in this simulation, as alignment of the disk
1320: with the equatorial plane of the black hole did not occur. This is
1321: not surprising since this simulation was carried out in the
1322: thick-disk regime where $H/r > \alpha$ and warps produced in the
1323: disk propagate as waves \citep{pap95a}, rather than diffusively as
1324: in the Bardeen-Petterson case. Since the expected bending wavelength
1325: \citep{lub02} turned out to be longer than the radial extent of the
1326: disk in the simulation, little warping of the disk was observed.
1327: Instead the unwarped disk precessed uniformly. The extrapolated
1328: precession period $T_{\rm prec} \approx 0.3 (M/M_\odot)$ s equates
1329: to periods of $\approx 3$ s and $\approx 3$ d for black holes of
1330: mass $M=10M_\odot$ and $M=10^6 M_\odot$, respectively. Such global
1331: disk precession could explain certain variability features observed
1332: from accreting black holes, such as low-frequency QPOs (LFQPOs)
1333: \citep{ste99,liu02,sch06}, since the observer's viewing angle of the
1334: inner, X-ray emitting region of the disk would vary periodically.
1335: % or the 106 day variability observed from Sgr A* \citep{liu02}.
1336: 
1337: If the inner disk is optically thick enough to produce
1338: relativistically-broadened reflection features, such as an iron
1339: K$\alpha$ line, then such precession should also be observable
1340: through periodic changes in both the shape and strength of the lines
1341: \citep{fra05c}. These changes should be correlated with the phase of
1342: the corresponding LFQPO. Such a correlation has been observed in GRS
1343: 1915+105 \citep{mil05}, although only between line strength and QPO
1344: phase; those data were not sufficiently resolved to determine the
1345: line shape.
1346: 
1347: Generally, we expect the precession period to be given by equation
1348: (\ref{eqn:precession}), which has a strong dependence on the radial
1349: distribution of the disk ($\propto r_o^{5/2-\zeta}r_i^{1/2+\zeta}$).
1350: One idea to consider is that the outer radius may correspond to the
1351: truncation radius proposed to explain the hard state of black hole
1352: X-ray binaries (e.g. \citet{esi97}, but see also \citet{ryk07}). In
1353: this case our simulated disk would represent the hot, geometrically
1354: thick flow that fills the region inside the truncation radius. The
1355: LFQPO would then correspond to the precession frequency of this
1356: inner flow, in which case it should scale as $r_o^{-5/2+\zeta}$.
1357: \citet{sob00} explored the dependence of the LFQPO frequency on
1358: spectral fitting parameters, including what would be the truncation
1359: radius in the context of the suggested hard state model. They
1360: studied two sources, XTE~J1550-564 and GRO~J1655-40, and found
1361: opposite trends between frequency and radius. For XTE~J1550-564 the
1362: observed frequency was $\nu_{\rm LFQPO}\sim 5$ Hz, and the observed
1363: truncation radius was $r_o/r_G= 2.7 (10M_\odot/M)(D/6
1364: \rm{~kpc})(\cos\theta)^{-1/2}$. From equation (\ref{eqn:precession})
1365: we would expect
1366: \begin{equation}
1367: \frac{r_o}{r_G}=\left[\frac{5-2\zeta}{\pi(1+2\zeta)}\right]^{2/(5-\zeta)}
1368: \left(\frac{a}{M}\right)^{2/(5-\zeta)}
1369: \left(\frac{r_i}{r_G}\right)^{-(1+2\zeta)/(5-2\zeta)}\left(\nu
1370: M\right)^{-2/(5-\zeta)} ~.
1371: \end{equation}
1372: In our simulation we found $\zeta\approx0$, which gives $r_o \approx
1373: 33 r_G$ for $M=10 M_\odot$ and $\nu = 5$~Hz. This is considerably
1374: larger than the observed value. However, some of the discrepancy may
1375: be attributable to the large uncertainties in the parameters used to
1376: describe this source, including its distance, mass, and inclination.
1377: Also, if the surface density in XTE J1550-564 depends strongly on
1378: radius, which was not the case for our simulated disk, then our
1379: prediction would change significantly. Further observational studies
1380: along this line are needed to test this prediction more thoroughly.
1381: 
1382: Although the main body of the disk was not significantly altered by
1383: the tilt, we did find significant differences in the inner regions
1384: of the flow when compared with untilted simulations. First, a tilted
1385: disk encounters the generalized ISCO surface at a larger radius than
1386: an untilted disk. This causes the plunging region to start further
1387: out. The binding energy of the innermost material in the disk is
1388: therefore less than it would be for an aligned disk, and the overall
1389: radiative efficiency should then be reduced.
1390: 
1391: On the other hand, tilting the disk appears to produce a higher
1392: overall mass accretion rate \citep[shown here in Figure 10a; also
1393: discussed in][]{lod06}. A tilted accretion disk will therefore have
1394: a lower surface density than an untilted disk with the same
1395: accretion rate. This may affect the emergent spectrum, especially
1396: for hot, optically thin flows.  On the other hand for flows that are
1397: effectively optically thick, \citet{dav05} found that the emergent
1398: spectra are remarkably independent of the overall stress and surface
1399: density.
1400: 
1401: We also found that the plunging region is not axially symmetric.
1402: Instead, accretion onto the hole in the tilted-disk case occurs
1403: through two discrete streams of material that leave the disk at high
1404: latitudes with respect to the black-hole and disk symmetry planes.
1405: %The magnetic linkage between the plunging region and the rest of the
1406: %disk occurs through these discrete high latitude streams;
1407: This may affect the magnitude of magnetic torques exerted by the
1408: plunging region on the disk. An interesting question for future work
1409: is how these streams vary on the timescale of the precession of the
1410: disk. We intend to explore the detailed properties of the plunging
1411: region and innermost disk in a future paper.
1412: 
1413: The tilted disk also seems not to have formed a clearly identifiable
1414: inner torus. This could be significant because the inner torus
1415: serves as a launching point for the matter-dominated, funnel-wall
1416: jet. The absence of a prominent inner torus may lead to a weaker
1417: matter jet. However, the present simulation is not suited to
1418: addressing this issue because of the poor and varying resolution
1419: used near the pole. Instead, we plan to explore jets and outflows
1420: from tilted disks in future work.
1421: 
1422: In many respects the tilted disk simulation exhibited properties
1423: consistent with an untilted disk around a black hole of lower spin.
1424: These included the larger plunging radius, higher mass accretion
1425: rate, and less prominent inner torus. Thus black-hole tilt could
1426: hamper efforts to estimate black-hole spin based on such properties.
1427: Indeed, it is commonly stated that astrophysical black hole
1428: spacetimes depend on just two parameters:  mass and spin.  But it
1429: should be remembered that the observed properties of black hole
1430: accretion disks also depend on their inclinations with respect to
1431: the spin axes of their central black holes. This inclination should
1432: be a target of future observational programs that use accretion
1433: disks as surrogates to study properties of black holes.
1434: 
1435: \acknowledgements We would like to recognize Chris Lindner for his
1436: contributions to this work. We would also like to thank Shane Davis,
1437: Julian Krolik, and the anonymous referee for their suggestions to
1438: improve this manuscript. PCF gratefully acknowledges the support of
1439: a Faculty R\&D grant from the College of Charleston and a REAP grant
1440: from the South Carolina Space Grant Consortium. This work was
1441: supported in part by the National Science Foundation under grants
1442: PHY99-0794 and AST03-07657, under the auspices of the U.S.
1443: Department of Energy by University of California Lawrence Livermore
1444: National Laboratory under contract W-7405-ENG-48, and under the
1445: following NSF programs: Partnerships for Advanced Computational
1446: Infrastructure, Distributed Terascale Facility (DTF) and Terascale
1447: Extensions: Enhancements to the Extensible Terascale Facility.
1448: 
1449: 
1450: 
1451: 
1452: %\clearpage
1453: \bibliographystyle{apj}
1454: \bibliography{myrefs}
1455: 
1456: 
1457: \end{document}
1458: