0810.0203/ms.tex
1: %\documentclass[11pt,preprint]{aastex}
2: %\documentclass{aastex}
3: %\usepackage{emulateapj5,epsfig}
4: \documentclass[useAMS, usenatbib, onecolumn]{mn2e}
5: \usepackage{epsfig}
6: \usepackage{amsmath}
7: %\usepackage{setspace}
8: %\doublespacing
9: \voffset -1.3cm
10: \def\be{\begin{equation}}
11: \def\ee{\end{equation}}
12: \def\ba{\begin{eqnarray}}
13: \def\ea{\end{eqnarray}}
14: \def\go{\mathrel{\raise.3ex\hbox{$>$}\mkern-14mu
15:              \lower0.6ex\hbox{$\sim$}}}
16: \def\lo{\mathrel{\raise.3ex\hbox{$<$}\mkern-14mu
17:              \lower0.6ex\hbox{$\sim$}}}
18: \def\bB{{\bf B}}
19: \def\bw{{\bf w}}
20: \def\bv{{\bf v}}
21: \def\bu{{\bf u}}
22: \def\br{{\bf r}}
23: \def\hate{{\hat{\bf e}}}
24: \def\bfe{{\bf e}}
25: \def\cB{{\cal B}}
26: \def\cA{{\cal A}}
27: \def\cR{{\cal R}}
28: \def\d{\partial}
29: 
30: \def\barR{{\bar R}}
31: \def\barz{{\bar z}}
32: \def\tomega{\tilde\omega}
33: \def\tomega{{\tilde{\omega}}}
34: 
35: \begin{document}
36: \title[Modes in Black Hole Accretion Discs]
37: {Corotational Instability of Inertial-Acoustic Modes in Black Hole
38: Accretion Discs and Quasi-Periodic Oscillations}
39: \author[D. Lai and D. Tsang]
40: {Dong Lai$^{1}$\thanks{Email:
41: dong@astro.cornell.edu; dtsang@astro.cornell.edu} 
42: and David Tsang$^{1,2}$\footnotemark[1] \\ 
43: $^1$Department of Astronomy, Cornell University, Ithaca, NY 14853, USA \\
44: $^2$Department of Physics,
45: Cornell University, Ithaca, NY 14853, USA \\}
46: 
47: \label{firstpage}
48: \maketitle
49: 
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: \begin{abstract}
52: We study the global stability of non-axisymmetric p-modes
53: (also called inertial-acoustic modes) trapped in the inner-most regions
54: of accretion discs around black holes.  We show that the lowest-order
55: (highest-frequency) p-modes, with frequencies $\omega=(0.5-0.7)
56: m\Omega_{\rm ISCO}$ (where $m=1,2,3,\cdots$ is the azimuthal wave
57: number, $\Omega_{\rm ISCO}$ is the disc rotation frequency at the
58: Inner-most Stable Circular Orbit, ISCO), can be overstable due to 
59: general relativistic effects, according to which the radial epicyclic
60: frequency $\kappa$ is a non-monotonic function of radius near the black
61: hole. The mode is trapped inside the corotation resonance radius $r_c$ 
62: (where the wave pattern rotation speed $\omega/m$
63: equals the disc rotation rate $\Omega$) and carries a negative energy. 
64: The mode growth arises primarily from wave absorption at the
65: corotation resonance, and the sign of the wave
66: absorption depends on the gradient of the disc vortensity,
67: $\zeta=\kappa^2/(2\Omega\Sigma)$ (where $\Sigma$ is the surface
68: density). When the mode frequency $\omega$ is
69: sufficiently high, such that $d\zeta/r>0$ at $r_c$,
70: positive wave energy is absorbed at the corotation, leading to the
71: growth of mode amplitude. The mode growth is further enhanced by
72: wave transmission beyond the corotation barrier.  We also study how
73: the rapid radial inflow at the inner edge of the disc affects the mode
74: trapping and growth. Our analysis of the behavior of the fluid perturbations
75: in the transonic flow near the ISCO indicates that, while
76: the inflow tends to damp the mode, the damping effect is sufficiently
77: small under some conditions (e.g., when the disc density decreases
78: rapidly with decreasing radius at the sonic point) so that net mode growth can 
79: still be achieved. We further clarify the role of the Rossby wave
80: instability and show that it does not operate for black hole accretion
81: discs with smooth-varying vortensity profiles.
82: Overstable non-axisymmetric p-modes driven by the
83: corotational instability provide a plausible explanation for the
84: high-frequency ($\go 100$~Hz) quasi-periodic oscillations (HFQPOs) 
85: observed from a number of black-hole X-ray binaries in the very high 
86: state. The absence of HFQPOs in the soft (thermal) state may result from 
87: mode damping due to the radial infall at the ISCO.
88: \end{abstract}
89: 
90: \begin{keywords}
91: accretion, accretion discs -- hydrodynamics -- waves -- 
92: -- black hole physics -- X-rays: binaries
93: \end{keywords}
94: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
95: \section{Introduction}
96: 
97: \subsection{Models of Quasi-Periodic X-Ray Oscillations: A Brief Review}
98: 
99: Rapid X-ray variabilities from Galactic compact binary systems have
100: been studied for decades (e.g. van del Klis 2006).  In recent years,
101: our knowledge of quasi-periodic oscillations (QPOs) in black-hole
102: X-ray binaries has greatly improved (see Remillard \& McClintock 2006
103: for a review), thanks in large part to NASA's {\it Rossi X-ray Timing
104: Explorer} (Swank 1999). The low-frequency QPOs (about 0.1-50~Hz) are
105: common, observable when the systems are in the hard state and the
106: steep power-law state (also called ``very high state''; see Done,
107: Gierlinski \& Kubota 2007), and they typically have high amplitudes
108: and high coherence ($Q>10$), and can vary in frequency on short
109: timescales (minutes).  However, it is the weaker, transient High-Frequency QPOs
110: (HFQPOs, 40--450~Hz) that have attracted more attention, since their
111: frequencies do not vary significantly in response to sizable (factors
112: of 3-4) luminosity changes and are comparable to the orbital
113: frequencies at the Innermost Stable Circular Orbit (ISCO) of black
114: holes with mass $M\sim 10 M_{\odot}$. As such, HFQPOs potentially
115: provide a probe to study the effects of strong gravity. HFQPOs are
116: usually observed in the very high state of the X-ray binaries, 
117: and have low amplitudes
118: ($0.5-2\%$ rms at 2-60~keV) and low coherence ($Q\sim 2-10$). Out of
119: the seven black-hole binaries from which HFQPOs have been reported, four
120: %(GRO J1655-40, XTE J1550-564, GRS 1915+105 and H1743-322) 
121: show pairs of QPOs (first discovered in GRO J1655-40; Strohmayer 2001)
122: with frequency ratios close to $2:3$ (300 and
123: 450~Hz in GRO J1655-40, 184 and 276~Hz in XTE J1550-564, 113 and
124: 168~Hz in GRS 1915+105, 165 and 240~Hz in H1743-322; note that GRS
125: 1915+105 also has a second pair of QPOs with $f=41$ and $67$~Hz).
126: 
127: It is worth noting that QPO (with period of $\sim 1$~hour) in X-ray emission
128: has recently been detected in the active galaxy RE J1034+396 
129: (Gierlinski et al.~2008). This could be the ``supermassive'' analog of the
130: HFQPOs detected in black-hole X-ray binaries.
131: 
132: Despite the observational progress, the origin of the HFQPOs remain 
133: unclear. A number of possibilities/models have been suggested or
134: studied to various degrees of sophistication. We comment on some of these
135: below:
136: 
137: -- Stella, Vietri \& Morsink (1999) and others (see Schnittman \&
138: Bertschinger 2004, Schnittman 2005) suggested that orbiting hot spots
139: (blobs) in the disc oscillating with epicyclic frequencies may provide
140: variability in the X-ray emission.  However the radial positions of
141: such blobs are free parameters, which must be tuned to match the
142: observed QPO frequencies, and it is also not clear that the blobs can
143: survive the differential rotation of the disc.
144: 
145: -- Abramowicz \& Kluzniak (2001) suggested that HFQPOs involve certain
146: nonlinear resonant phenomenon in the disc (e.g., coupling between the
147: radial and vertical epicyclic oscillations of the disc fluid element;
148: Kluzniak \& Abramowicz 2002).  This was motivated by the observed
149: stability of the QPO frequencies and the commensurate frequency ratio.
150: However, so far analysis has been done based only on toy models involving coupled harmonic oscillators (e.g. Rebusco 2004; Horak \& Karas 2006) and no fluid dynamical model producing these resonances
151: has been developed (see Abramowicz et al 2007 and Rebusco 2008 for
152: recent reviews). Petri (2008) considered the resonant oscillation of a
153: test mass in the presence of a spiral density wave, but the origin of
154: the wave is unclear.
155: 
156: -- Acoustic oscillation modes in pressure-supported accretion tori
157: have been suggested as a possible source of the observed QPOs
158: (Rezzolla et al. 2003; Lee, Abramowicz \& Kluziniak 2004; see also
159: Blaes, Arras \& Fragile 2006; Schnittman \& Rezzolla 2006, Blaes 
160: et al 2007, Sramkova et al 2007).  In this
161: model, the commensurate mode frequencies arise from matching the
162: radial wavelength to the size of the torus. Note that the QPO
163: frequencies are determined mainly by the radial boundaries of the
164: torus, which must be tuned to match the observed QPO frequencies. It
165: is also not clear that the accretion flow in the very high state (in
166: which HFQPOs are observed) is well represented by such a torus (e.g. Done et al 2007).
167: 
168: -- Li \& Narayan (2004) considered the dynamics of the
169: interface between the accretion
170: disc and the magnetosphere of a central compact object (see also
171: Lovelace \& Romanova 2007). 
172: The interface is generally Rayleigh-Tayor unstable and may also be
173: Kelvin-Helmholtz unstable. While such an interface is clearly relevant
174: to accreting magnetic neutron stars, Li \& Narayan suggested that it
175: may also be relevant to accreting black holes and that the strongly
176: unstable interface modes may give rise to QPOs with commensurate
177: frequencies.
178: 
179: -- Perhaps the theoretically most appealing is the relativistic
180: diskoseismic oscillation model, according to which general
181: relativistic (GR) effects produce trapped oscillation modes in the
182: inner region of the disc (Kato \& Fukue 1980; Okazaki et al.~1987;
183: Nowak \& Wagoner 1991; see Wagoner 1999; Kato 2001 for reviews; see
184: also Tassev \& Bertschinger 2007 for the kinematic description of some
185: of these wave modes).  A large majority of previous studies have
186: focused on disc g-modes (also called inertial modes or
187: inertial-gravity modes, whose wavefunctions -- such as the pressure
188: perturbation, contain nodes in the vertical direction), because the
189: trapping of the g-mode does not require a reflective inner/outer disc
190: boundary.  Kato (2003a) and Li, Goodman \& Narayan (2003) showed that
191: the g-mode that contains a corotation resonance (where the wave
192: patten frequency equals the rotation rate of the background flow) in
193: the wave zone is heavily damped. Thus the only nonaxisymmetric ($m\neq
194: 0$) g-modes of interest are those trapped around the maximum of
195: $\Omega+\kappa/m$ (where $\Omega$ is the rotational frequency,
196: $\kappa$ is the radial epicyclic frequency and $m$ is the azimuthal
197: mode number; see Fig.~1 below). Unfortunately, the frequencies of such
198: modes, $\omega\simeq m\Omega_{\rm ISCO}$, are too high (by a
199: factor of 2-3) compared to the observed values, given the measured
200: mass and the estimated spin parameter of the black hole (Silbergleit
201: \& Wagoner 2007; see also Tassev \& Bertschinger 2007).  Axisymmetric
202: g-modes ($m=0$) may still be viable in the respect, and recent studies
203: showed that they can be resonantly excited by global disc deformations
204: through nonlinear effects (Kato 2003a,2008; Ferreira \& Ogilvie 2008).
205: Numerical simulations (Arras, Blaes \& Turner 2006; Reynolds \& Miller
206: 2008), however, indicated that while axisymmetric g-mode oscillations
207: are present in the hydrodynamic disc with no magnetic field, they
208: disappear in the magnetic disc where MHD turbulence develops. Also, Fu
209: \& Lai (2008) carried out an analytic study of the effect of magnetic
210: fields on diskoseismic modes and showed that even a weak (sub-thermal)
211: magnetic field can ``destroy'' the self-trapping zone of disc g-modes,
212: and this may (at least partly) explain the disappearance of the
213: g-modes in the MHD simulations.
214: 
215: -- Tagger and collaborators (Tagger \& Pellat 1999; Varniere \& Tagger
216: 2002; Tagger \& Varniere 2006; see Tagger 2006 for a review) developed
217: the theory of accretion-ejection instability for discs threaded by
218: strong (of order or stronger than equipartition), large-scale poloidal
219: magnetic fields.  They showed that such magnetic field provides a
220: strong coupling between spiral density waves and Rossby waves at the
221: corotation, leading to the growth of the waves and energy ejection to
222: disc corona. Tagger \& Varniere (2006) suggested that normal modes
223: trapped in the inner region of the disc become strongly unstable by a
224: combination of accretion-ejection instability and an MHD form of the
225: Rossby wave instability (see Lovelace et al.~1999; Li et al.~2000; see section 6
226: below). The Tagger model has the appealing feature that the instability leads
227: to energy ejection into the disc corona, and thus explains why HFQPOs
228: manifest mainly as the variations of the nonthermal (power-law)
229: radiation from the systems.
230: 
231: 
232: 
233: %\cite{Tsang:2008p4205} explored the effect of the corotation
234: %resonance on the reflection coefficient for non-axisymmetric waves
235: %incident on the corotation barrier for perturbations with no vertical
236: %nodes. It was found that incoming waves could be super-reflected
237: %(reflection coefficient greater than unity) due to wave absorption at
238: %the corotation singularity depending on the slope of the vortensity
239: %($\zeta = \kappa^2/2\Omega\Sigma$), and explicit WKB values for the
240: %Reflection, Transmission and Absorption coefficients were calculated
241: %analytically. This absorption effect was found to dominate the
242: %outgoing flux of angular momentum that determines the growth rate of
243: %the classic corotation amplifier.
244: 
245: \subsection{This Paper}
246: 
247: In this paper we study the global corotational instability of
248: nonaxisymmetric p-modes (also called inertial-acoustic modes) trapped
249: in the inner-most region of the accretion disc around a black
250: hole. The p-modes do not have vertical structure (i.e., the
251: wavefunctions have no node in the vertical direction). We focus on
252: these modes because their basic wave properties (e.g. propagation
253: diagram) are not affected qualitatively by disc magnetic fields (Fu \&
254: Lai 2008) and they are probably robust under hydromagnetic effects
255: and disc turbulence (see Reynolds \& Miller 2008).
256: 
257: The corotational instability of p-modes studied in this paper relies
258: on the well-known GR effect of test-mass orbit around a black hole:
259: Near the black hole, the radial epicyclic frequency $\kappa$ reaches a
260: maximum (at $r=8GM/c^2$ for a Schwarzschild black hole) and goes to
261: zero at the ISCO ($r_{\rm ISCO}=6GM/c^2$).  This causes non-monotonic
262: behavior in the fluid vortensity, $\zeta=\kappa^2/(2\Sigma\Omega)$
263: (assuming the surface density $\Sigma$ is relatively smooth), such
264: that $d\zeta/dr>0$ for $r<r_{\rm peak}$ (where $r_{\rm peak}$ is the
265: radius where $\zeta$ peaks) and $d\zeta/dr<0$ for $r>r_{\rm peak}$.
266: The vortensity gradient $d\zeta/dr$ plays an important role in wave
267: absorption at the corotation resonance (Tsang \& Lai 2008a; see
268: also Goldreich \& Tremaine 1979 for corotational wave absorption due to
269: external forcing). We show that the p-modes with frequencies such that
270: the corotation radii lie inside the vortensity peak can grow in
271: amplitude due to absorption at the corotation resonance, and that the
272: overstability can be achieved for several modes with frequencies
273: closely commensurate with the azimuthal wavenumber $m$.  Tagger \&
274: Varniere (2006) have studied similar modes in discs threaded by strong
275: magnetic fields, but in our analysis magnetic fields play no role.
276: 
277: The trapping of the p-modes requires the existence of a (partially)
278: reflecting boundary at the disc inner edge, close to the ISCO. One may
279: suspect that the rapid radial inflow at the ISCO will diminish any
280: potential instabilities in the inner accretion disc (see Blaes 1987
281: for the case of thick accreting tori). Our analysis of the
282: wave perturbations in the transonic accretion flow (see section 5)
283: suggests that waves are partially reflected at the sonic point, and
284: global overstable p-modes may still be produced under certain 
285: conditions (e.g., when the surface density of the flow
286: varies on sufficiently small length scale around the sonic point).
287: Even better mode trapping (and therefore larger mode growth)
288: may be achieved when the system is an accretion
289: state such that the inner disc edge does not behave as a zero-torque
290: boundary (see section 7 for discussion and references).
291: 
292: Our paper is organized as follows. After summarizing the basic
293: fluid equations for our problem in section 2, we give a physical
294: discussion of the origin of the corotational instability of disc 
295: p-modes in section 3. We present in section 4
296: our calculations of the growing p-modes with simple reflective
297: inner disc boundary conditions. Section 5 contains our analysis of
298: the effect of the transonic radial inflow at the ISCO 
299: on the p-mode growth rate. In section 6, we discuss the role of the Rossby 
300: wave instability and show that it is not effective in typical 
301: accretion discs under consideration.
302: In section 7 we discuss the application
303: of our results to HFQPOs in black hole X-ray binaries.
304: 
305: 
306: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
307: \section{Setup and Basic equations}
308: 
309: We consider a geometrically thin disc and adopt cylindrical
310: coordinate system $(r,\phi,z)$. The flow is assumed to be barotropic,
311: so that the vertically integrated pressure, $P=\int p~dz$, depends
312: only on the surface density, $\Sigma=\int \rho~dz$.  We use
313: the pseudo-Newtonian potential of Paczynski \& Wiita (1980) 
314: \be
315: \Phi = -{GM\over r-r_S},
316: \ee
317: with $r_S = 2GM/c^2$ the Schwarzschild radius.
318: The free-particle (Keplerian) orbital and radial epicyclic (angular)
319: frequencies are
320: \be
321: \Omega_K = \left(\frac{1}{r}\frac{d\Phi}{dr}\right)^{1/2} =
322: \sqrt{\frac{GM}{r}}\frac{1}{r-r_S}~, \qquad \kappa =
323: \left[\frac{2\Omega_K}{r} \frac{d}{dr}(r^2\Omega_K)\right]^{1/2} = \Omega_K
324: \sqrt{\frac{r-3r_S}{r-r_S}}.
325: \label{eq:OmegaK}
326: \ee
327: The function $\kappa$ peaks at $r=(2+\sqrt{3})r_S$ and declines to zero at 
328: $r_{\rm ISCO}=3r_S$ (while for a Schwarzschild black hole in GR, 
329: $\kappa$ peaks at $r=4r_S$). 
330: The unperturbed flow has velocity $\bu_0=(u_r,r\Omega,0)$.
331: Since pressure is negligible for thin discs, we have
332: $\Omega\simeq \Omega_K$.
333: 
334: Neglecting the self-gravity of the disc we have the linear
335: perturbation equations:
336: \ba
337: && {\partial\over\partial t}\delta\Sigma+\nabla\cdot(\Sigma\,\delta \bu
338: +\bu_0\,\delta\Sigma)=0,\label{eq:rho}\\
339: &&{\partial \over\partial t}\delta\bu+(\bu_0\cdot\nabla)\delta\bu
340: +(\delta\bu\cdot\nabla)\bu_0
341: =-\nabla\delta h, \label{eq:u}
342: \ea
343: where $\delta\Sigma,~\delta \bu$ and $\delta h=\delta P/\Sigma$
344: are the (Eulerian) perturbations of surface density, velocity and enthalpy,
345: respectively. For barotropic flow, $\delta h$ and $\delta\Sigma$ are 
346: related by
347: \be
348: \delta h=c_s^2{\delta\Sigma\over\Sigma},
349: \ee
350: where $c_s$ is the sound speed, with $c_s^2=dP/d\Sigma$.
351: We assume all perturbed quantities to be of the form
352: $e^{im\phi - i\omega t}$, 
353: where $m$ is a positive integer, and $\omega$ is the 
354: wave (angular) frequency.
355: The perturbation equations then become
356: \ba
357: && -i\tomega {\Sigma\over c_s^2}\delta h + \frac{1}{r} \frac{\d}{\d r}
358: (\Sigma r \delta u_r) + \frac{im}{r}\Sigma \delta u_\phi 
359: +{1\over r}{\partial\over\partial r}(ru_r\delta\Sigma)
360: = 0,\label{eq:sig}\\
361: && -i\tomega \delta u_r - 2\Omega \delta u_\phi 
362: +{\partial\over\partial r}(u_r\delta u_r)= -{\partial\over\partial r}
363: \delta h,\label{eq:ur}\\
364: && -i\tomega \delta u_\phi + \frac{\kappa^2}{2\Omega}\delta u_r 
365: +{u_r\over r}{\partial\over\partial r}(r\delta u_\phi)
366: =-\frac{im}{r} \delta h,\label{eq:uphi}
367: \ea
368: where 
369: \be
370: \tomega = \omega - m \Omega,
371: \ee
372: is the wave frequency in the frame corotating with the unperturbed fluid.
373: 
374: Except very near the inner edge of the disc, $r_{\rm in}\simeq r_{\rm ISCO}$,
375: the unperturbed radial velocity is small, $|u_r|\ll r\Omega$.
376: In our calculations of the disc modes, we will neglect $u_r$ and
377: set the last terms on the left-hand sides of 
378: equations (\ref{eq:sig})-(\ref{eq:uphi}) to zero (However,
379: $u_r$ plays an important role in determining the inner boundary 
380: condition of the fluid perturbations at $r_{\rm in}$; see section 5).
381: Eliminating the velocity perturbations in favor of the enthalpy,
382: we obtain our master equation
383: \be
384: \left[\frac{d^2}{dr^2} -
385: \frac{d}{dr}\left(\ln\frac{D}{r\Sigma}\right)\frac{d}{dr} -
386: \frac{2m\Omega}{r\tomega}\left(\frac{d}{dr}\ln\frac{\Omega\Sigma}{D}\right)
387: - \frac{m^2}{r^2} - \frac{D}{c_s^2}\right]\delta h = 0,
388: \label{perturbeq1}
389: \ee
390: where 
391: \be
392: D = \kappa^2 - \tomega^2.
393: \ee
394: For concreteness we assume the surface density to have a the power-law form 
395: \be
396: \Sigma \propto r^{-p},
397: \ee
398: where $p$ is the density index.
399: 
400: The above equations adequately describe disc p-modes (also called
401: inertial-acoustic modes), which do not have vertical structure.
402: Other disc modes (g-modes and c-modes) involve the vertical
403: degree of freedom [see Kato 2001 for a review; also see Fig.~1 of
404: Fu \& Lai (2008) for a quick summary],
405: and their stability properties are studied by
406: Kato (2003a), Li et al.~(2003) and Tsang \& Lai (2008b).
407: 
408: To determine the global modes of the disc, appropriate 
409: boundary conditions must be specified. These are discussed in 
410: sections 4 and 5.
411: 
412: 
413: 
414: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
415: \section{P-modes and Their Growth Due
416: to Corotation Resonance: A Physical Discussion}
417: 
418: A WKB analysis of the wave equation (\ref{perturbeq1})
419: yields the dispersion 
420: relation for the local plane wave 
421: $\delta h\propto \exp\left[i\int^r\!k(s) ds\right]$:
422: \be
423: k^2+{D\over c_s^2}+
424: \frac{2m\Omega}{r\tomega}\left(\frac{d}{dr}\ln\frac{\Omega\Sigma}{D}\right)
425: \simeq 0.
426: \label{eq:disper}\ee
427: Far from the singularity ($\tomega=0$) at the corotation radius $r_c$, 
428: this reduces to the 
429: well-known dispersion relation of spiral density wave with no self-gravity
430: (e.g., Shu 1992), $k^2\simeq -D/c_s^2$, or 
431: \be
432: \tomega^2 = \kappa^2 + k^2 c_s^2~.
433: \ee
434: Density waves (p-modes) can propagate inside the inner Lindblad resonance
435: radius $r_{\rm IL}$ (defined by $\tomega = -\kappa$), and outside the
436: outer Lindblad resonance radius $r_{\rm OL}$ (defined by $\tomega =
437: \kappa$), i.e., in the region where
438: $\omega/m<\Omega-\kappa/m$ and 
439: $\omega/m > \Omega+\kappa/m$, respectively (see Fig.~1). 
440: Between $r_{\rm IL}$ and $r_{\rm OL}$, waves are 
441: evanescent except that a very narrow Rossby wave zone exists 
442: around the corotation radius. Indeed, 
443: in the vicinity of $\tomega=0$, equation (\ref{eq:disper}) reduces to 
444: \be
445: \tomega\simeq {2m\Omega\over r (k^2+\kappa^2/c_s^2)}
446: \left({d\ln\zeta\over dr}\right)_{r_c},
447: \ee
448: where 
449: \be
450: \zeta={\kappa^2\over 2\Omega\Sigma}
451: \ee
452: is the vortensity of the (unperturbed) flow.
453: % and the subscript ``c'' implies tha the quantity is evaluated at $r=r_c$.
454: For $(d\zeta/dr)_{r_c}>0$, the Rossby wave zone lies between
455: $r_c$ and $r_c+\Delta r_R$, where $\Delta r_R=(2c_s/\kappa)|\nu|$ and
456: $\int_{r_c}^{r_c+\Delta r_R}k\,dr=\pi\nu$, with the number of wavelengths in the
457: Rossby zone given by (Tsang \& Lai 2008a)
458: \be
459: \nu = \left(\frac{c_s}{q\kappa}\frac{d\ln \zeta}{dr} \right)_{r_c}
460: ={c_s\over q\kappa}\left[{d\over dr}\ln\left({\kappa^2\over\Omega}\right)
461: +{p\over r}\right]_{r_c},
462: \label{eq:nu}\ee
463: where $q \equiv -\left(d\ln \Omega/d \ln r\right)_{\rm r_c}$, and the second
464: equality assumes $\Sigma\propto r^{-p}$. 
465: For $(d\zeta/dr)_{r_c}<0$, the Rossby wave zone lies inside $r_c$, between
466: $r_c-|\Delta r_R|$ and $r_c$ (see Fig.~1). Note that since $\nu\sim c_s/(\kappa r)\sim
467: H/r\ll 1$, no standing Rossby wave can exist in the Rossby zone (see also section 6).
468: 
469: 
470: 
471: %===================================
472: \begin{figure}
473: \centering
474: \includegraphics[width=13.5cm]{f1.eps}
475: \caption{Wave propagation diagram for non-axisymmetric p-modes in thin
476: accretion discs around black holes.  In the upper panel, the three
477: solid curves depict the disc rotation profile $\Omega(r)$ and
478: $\Omega\pm\kappa/m$ (where $\kappa$ is the radia epicyclic frequency);
479: note that the three curves join each other at the disc inner radius
480: $r_{\rm in}=r_{\rm ISCO}$ (the inner-most circular orbit) since
481: $\kappa(r_{\rm ISCO})=0$.  The wavy lines (of height $\omega/m$)
482: indicate the propagation zones for inertial-acoustic waves. Disc
483: p-modes are trapped between $r_{\rm in}$ and the inner Lindblad
484: resonance radius (where $\omega/m = \Omega - \kappa/m$), but can
485: tunnel through the corotation barrier. The lower panel depicts the
486: disc vortensity profile, $\zeta=\kappa^2/(2\Omega\Sigma)$, which has a
487: maximum at the radius $r_{\rm peak}$ (as long as the surface density
488: $\Sigma$ does not vary too strongly with $r$).  P-modes with
489: $\omega/m>\Omega_{\rm peak}=\Omega(r_{\rm peak})$ (the upper wavy
490: line) are overstable due to wave absorption at the corotation
491: resonance radius $r_c$ (where $\omega/m=\Omega$) since
492: $(d\zeta/dr)_{r_c}>0$. P-modes with $\omega/m<\Omega_{\rm peak}$ (the
493: lower wavy line) tend to be damped by wave absorption at $r_c$ since
494: $(d\zeta/dr)_{r_c}<0$. Note that a narrow Rossby wave zone (labeled by
495: thick horizontal bars) exists just outside or inside the corotation
496: radius --- the location of this Rossby zone determines the sign of
497: the corotational wave absorption.  }
498: \label{fig1}
499: \end{figure}
500: %=======================================
501: 
502: Assuming that there exists a reflecting boundary at the inner 
503: disc radius $r_{\rm in}\simeq r_{\rm ISCO}$ (see sections 4.3 and 5),
504: normal modes can be produced, with the waves partially trapped between 
505: $r_{\rm in}$ and 
506: $r_{\rm IL}$ -- these are the p-modes that we will focus on in this paper.
507: The mode eigen-frequency $\omega=\omega_r+i\omega_i$ is generally complex,
508: with the real part $\omega_r$ determined approximately by the Sommerfeld 
509: ``quantization'' condition
510: \be
511: \int_{r_{\rm in}}^{r_{\rm IL}} {1 \over c_s} \sqrt{\tomega_r^2
512: - \kappa^2} dr = n\pi +\varphi, 
513: \ee 
514: where $\tomega_r=\omega_r-m\Omega$, 
515: $n$ is an integer and $\varphi$ (of order unity) is a phase factor
516: depending on the details of the (inner and outer) boundary conditions.
517: The overstability of the p-mode is directly related to the 
518: reflectivity of the corotation barrier between $r_{\rm IL}$ and 
519: $r_{\rm OL}$. In the WKB approximation, the imaginary part of the 
520: mode frequency is given by (Tsang \& Lai 2008a; see also Narayan et al.~1987,
521: who considered shearing-sheet model)
522: \be
523: \omega_i = \left(\frac{|{\cal R}| -1}{|{\cal R}|+1}\right)
524: \left(\int_{r_{\rm in\ }}^{r_{\rm IL}}
525: {|\tomega_r|\over c_s\sqrt{\tomega_r^2 - \kappa^2}} dr\right)^{-1},  
526: \label{eq:omegai}\ee
527: where ${\cal R}$ is the reflectivity (see below).
528: Thus the mode becomes overstable ($\omega_i > 0$) for $|{\cal R}| > 1$
529: (termed ``super-reflection'')  and stable
530: ($\omega_i < 0$) for $|{\cal R}| < 1$.
531: 
532: Super-reflection in fluid discs arises because
533: the waves inside the corotation radius and those outside carry 
534: energy or angular momentum of different signs: 
535: Since the wave inside $r_c$ has pattern speed $\omega_r/m$ less than the 
536: fluid rotation rate $\Omega(r)$, it carries negative energy; outside
537: $r_c$, we have $\omega_r/m>\Omega(r)$, the wave carries positive 
538: energy. Consider an incident wave $\delta h\propto \exp(-i\int^r k\,dr)$, 
539: carrying energy of the amount $(-1)$, propagating from small radii
540: toward the corotation barrier
541: \footnote{Note that since the group velocity 
542: of the wave has opposite sign as the phase velocity for $r<r_{\rm IL}$,
543: the wave of the form $\exp(-i\int^rk\,dr)$ (with $k>0$)
544: is outward propagating.}. The wave reflected at $r_{\rm IL}$
545: takes the form $\delta h\propto {\cal R}\exp(i\int^r k\,dr)$, and the 
546: transmitted wave in the region $r>r_{\rm OL}$ is
547: $\delta h\propto {\cal T}\exp(i\int^r k\,dr)$. 
548: Because of the corotation singularity, the wave energy can also be
549: transferred to the background flow and dissipated at the corotation 
550: radius.   Energy conservation then gives
551: $-1=(-1)|{\cal R}|^2+|{\cal T}|^2+{\cal D}_c$, or
552: \be
553: |{\cal R}|^2=1+|{\cal T}|^2+{\cal D}_c, \label{eq:Econs}
554: \ee
555: where ${\cal D}_c$ is the wave energy dissipated at the corotation. 
556: 
557: Tsang \& Lai (2008a) derived the analytical expressions (in the WKB
558: approximation) for ${\cal T}$, ${\cal R}$ and ${\cal D}_c$. 
559: Two effects determine the reflectivity.
560: (i) The transmitted wave (corresponding to the $|{\cal T}|^2$ term)
561: always carries away positive energy and thus increases 
562: $|{\cal R}|^2$.
563: (ii) Wave absorption at the corotation can have both signs, depending on 
564: $\nu$: For $\nu>0$, the Rossby wave zone lies outside $r_c$, positive wave
565: energy is dissipated and we have ${\cal D}_c>0$; for $\nu<0$, the Rossby zone
566: lies inside $r_c$ and we have ${\cal D}_c<0$ (see Fig.~1). 
567: Tsang \& Lai (2008a) showed explicitly that under most conditions,
568: $|{\cal D}_c|\gg |{\cal T}|^2$ (except when $\nu\simeq 0$, for which
569: ${\cal D}_c\simeq 0$). In the limit of $|\nu|\ll 1$, we have
570: \be
571: |{\cal T}|^2 \simeq \exp(-2\Theta_{\rm II}),\qquad
572: {\cal D}_c\simeq 2\pi\nu\exp(-2\Theta_{\rm IIa}),
573: \ee
574: where
575: \be
576: \Theta_{\rm II}\equiv \int_{r_{\rm IL}}^{r_{\rm OL}}
577: \frac{\sqrt{\kappa^2 - \tomega_r^2}}{c_s}, \qquad \Theta_{\rm IIa}\equiv
578: \int_{r_{\rm IL}}^{r_{c}} \frac{\sqrt{\kappa^2 - \tomega_r^2}}{c_s}
579: \ee
580: [these expressions are valid for $\Theta_{\rm II},\Theta_{\rm IIa}\gg 1$; see 
581: Tsang \& Lai (2008a) for more general expressions]. Thus super-reflectivity
582: ($|{\cal R}|^2>1$) and growing modes ($\omega_i>0$) are achieved when 
583: \be
584: \nu>\nu_{\rm crit}=-{1\over 2\pi}\exp(-2\Theta_{\rm IIb}),\qquad
585: {\rm with}\quad\Theta_{\rm IIb}=\Theta_{\rm II}-\Theta_{\rm IIa}
586: =\int_{r_c}^{r_{\rm OL}}
587: \frac{\sqrt{\kappa^2 - \tomega_r^2}}{c_s}.
588: \label{eq:nucrit}\ee
589: Note that typically $|\nu_{\rm crit}|\ll 1$; if $|{\cal T}|^2$ is neglected
590: compared to ${\cal D}_c$, then $\nu_{\rm crit}=0$.
591: 
592: 
593: As mentioned before, since $\kappa$ is non-monotonic near the black
594: hole, the vortensity $\zeta$ is also non-monotonic, attaining a peak
595: value at $r=r_{\rm peak}$ before dropping to zero at the
596: ISCO. Therefore, p-modes with frequencies such that the corotation
597: radius $r_c$ lies inside $r_{\rm peak}$ are expected to be overstable
598: by the corotational instability discussed above. In other words, when
599: $\omega_r/m> \Omega_{\rm peak}\equiv \Omega(r_{\rm peak})$, the
600: corotation resonance acts to grow the mode.  Note that $\Omega_{\rm
601: peak}$ depends on the surface density profile as well as the spacetime
602: curvature around the black hole (see Fig.~2).  On the other hand, when
603: $\omega_r/m< \Omega_{\rm peak}$ ($\nu<0$), the corotational wave
604: absorption acts to damp the mode.  However, when $\omega_r/m$ is only
605: slightly smaller than $\Omega_{\rm peak}$ ($\nu_{\rm crit}<\nu<0$)
606: mode growth can still be obtained due to wave leakage beyond the outer
607: Lindblad resonance, though the growth rate will be small (see section
608: 4.4 for examples).
609: 
610: 
611: %===========================================
612: \begin{figure}
613: \centering
614: \includegraphics[width=13cm]{f2.eps}
615: \caption{Critical mode frequency for corotational instability as a function 
616: of the disc surface density index $p$ (with $\Sigma\propto r^{-p}$).
617: Wave absorption at the corotation resonance acts to grow the mode only if the
618: corotation occurs in the region of positive vortensity gradient, i.e., 
619: if the mode pattern frequency 
620: $\omega/m>\Omega_{\rm peak}$ (see Fig.~1). P-mode trapping also
621: requires $\omega/m<\Omega_{\rm ISCO}=\Omega(r_{\rm ISCO})$.
622: %Thus the growth rates can only become large if the pattern
623: %frequency occurs between $\Omega_{\rm peak}$ and $\Omega_{ISCO}$. 
624: Note that weak mode growth can still occur when $\omega/m$ is slightly 
625: below $\Omega_{\rm peak}$ due to wave leakage beyond the outer Lindblad resonance
626: radius. See text for detail.
627: %however  this effect is only significant for small negative $\nu$.}
628: }
629: \label{fig2}
630: \end{figure}
631: %===========================================
632: 
633: 
634: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
635: \section{Calculations of Trapped, Overstable P-modes}
636: 
637: To determine the eigenvalues $\omega_r$ and $\omega_i$ of the trapped
638: modes, we solve equations (\ref{eq:sig})-(\ref{eq:uphi})
639: (with $u_r=0$) or equation \eqref{perturbeq1} subjected to 
640: appropriate boundary conditions at $r_{\rm in}$ and $r_{\rm out}$. 
641: 
642: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
643: \subsection{``Landau'' Integration Contour}
644: 
645: When solving eigenvalue problem using the standard method (e.g. the
646: shooting method as described in Press et al 1998), we encountered a
647: conundrum: For $\nu>0$, we could find both a growing mode and
648: a decaying mode, with almost the same $\omega_r$ but opposite
649: $\omega_i$.
650: This appears to contradict our discussion in section 3. 
651: This conundrum arises because our 
652: numerical integration is confined to the real $r$ axis. However,
653: analogous to Landau's analysis of wave damping in a plasma (e.g.,
654: Lifshitz \& Pitaevskii 1981), care must
655: be taken in defining appropriate contour of integration across the
656: corotation resonance. Indeed, at corotation, equation 
657: \eqref{perturbeq1} contains a singular term, proportional to 
658: \be
659: {1\over \tilde\omega} \propto {1\over r-R_c},
660: \ee
661: where $R_c \equiv r_c - ir_c\omega_i/(q\omega_r)$ is the complex
662: pole, $r_c$ is determined by $\omega_r=m\Omega(r_c)$ and 
663: $q=-(d\ln\Omega/d\ln r)_c>0$.
664: 
665: As discussed in Lin (1955) in the context of hydrodynamical shear flows,  
666: to obtain physically relevant solutions of the fluid system, it is 
667: necessary that the integration contour lies above the pole.
668: This is the Landau contour. In essence, only by adopting such
669: a Landau contour can one obtain the correct wave absorption (dissipation)
670: at the corotation. For growing modes ($\omega_i>0$), Im$(R_c)<0$,
671: our numerical integration along the real $r$ axis constitutes
672: the correct Landau contour. On the other hand, 
673: for decaying modes ($\omega_i<0$), the real $r$ axis is not the
674: correct Landau contour as Im$(R_c)>0$. Instead, to obtain physical
675: solutions for these decaying modes, 
676: the integration contour must be deformed so that $R_c$ lies
677: below it (see Fig.~3). As we are primarily interested in over-stable modes
678: in this paper, it is adequate to integrate along the real $r$ axis
679: in our calculation.
680:  
681: %We thus follow the procedure outlined by Lin (1955) to obtain the
682: %physically relevant branch for the given system, which is equivalent
683: %to crossing the singular points with $\omega_i > 0$ (a similar result
684: %is achieved by treating the initial value problem via Laplace
685: %transforms). Thus when integrating over the singularity to identify
686: %growing modes ($\omega_i > 0$) it is sufficient to perform the
687: %numerical integration with $\omega = \omega_r + i \omega_i$.  Such an
688: %integration is not physical for shrinking modes ($\omega_i < 0$),
689: %instead the contour must be deformed to take the same branch as the
690: %growing case (see Figure \ref{fig3}), however we are primarily
691: %interested in over-stable case.
692: 
693: \begin{figure}
694: \centering
695: \includegraphics[width=9cm]{f3.eps}
696: \caption{``Landau'' contour for integration across the corotation resonance.
697: To calculate the growing mode ($\omega_i>0$), 
698: it is adequate to integrate the fluid perturbation 
699: equations along the real $r$ axis (upper panel). To obtain the physical solution
700: for the shrinking mode ($\omega_i<0$),
701: the integration contour must deformed so that the corotational pole 
702: $R_c$ lies below the contour. 
703: %Above: The contour for integrating over the corotation singularity at
704: %$R_c$ for a growing mode. Below: The same but for a shrinking mode,
705: %taking the physically relevant branch.
706: } 
707: \label{fig3}
708: \end{figure}
709: 
710: 
711: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
712: \subsection{Outer Boundary Condition}
713: 
714: As we are interested in self-excited modes in the inner region of the disc,
715: we adopt the radiative outer boundary condition. Specifically,
716: far from the outer Lindblad resonance ($r > r_{\rm OL}$) we demand that 
717: only an outgoing wave exists:
718: \be
719: \delta h \propto A \exp\left(i\int^r k~dr \right)~,\quad
720: {\rm with}~~A=\left({D\over r\Sigma k}\right)^{1/2},
721: \ee
722: where $k = \sqrt{-D/c_s^2}$ (see
723: Tsang \& Lai 2008a). This gives the boundary condition at some 
724: $r_{\rm out} > r_{\rm OL}$: 
725: \be
726: \delta h'(r_{\rm out}) 
727: = \delta h(r_{\rm out}) 
728: %\left(ik + \frac{S'}{2S} - \frac{k'}{2k}\right)_{r_{\rm out}}.
729: \left(ik + {1\over A}{dA\over dr}\right)_{r_{\rm out}}.
730: \label{eq:outer}\ee
731: In practice, we find that $r_{\rm out}\sim 2 r_{\rm OL}$ would yield sufficiently 
732: accurate results.
733: 
734: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
735: \subsection{Inner Boundary Conditions}
736: 
737: To obtain global trapped modes, at least partial wave reflection must occur 
738: at $r_{\rm in}$. To focus on the effect of corotational instability discussed
739: in section 3, in this section we consider two simple inner boundary 
740: conditions. We defer our analysis of the effect of radial inflow
741: on the p-modes to section 5.
742: 
743: (i) At the ISCO, the flow plunges into the black hole, we expect a sudden
744: decrease in the surface density of the disc. Thus, it is reasonable to 
745: consider the free surface boundary condition, i.e.,
746: the Lagrangian pressure perturbation $\Delta P=0$.
747: Using $\delta u_r=-i\tomega\xi_r$ (where $\xi_r$ is the radial Lagrangian
748: displacement), and $\Delta P=\delta P+\xi_r dP/dr$, we have
749: \be
750: \left({\Delta P\over\Sigma}\right)_{\rm r_{\rm in}}=
751: \left(\delta h - pc_s^2\frac{i\delta u_r}{r \tomega}\right)_{r_{\rm 
752: in}}=0,
753: \label{eq:DeltaP}\ee 
754: where we have used
755: $dP/dr = (dP/d\Sigma)(d\Sigma/dr) = -pc_s^2\Sigma/r$
756: for barotropic, power-law discs ($\Sigma\propto r^{-p}$).
757: 
758: (ii) We assume that 
759: the radial velocity perturbation vanishes at the inner boundary, i.e., 
760: $\delta u_r = 0$. This was adopted by Tagger \& Varniere (2006)
761: in their calculations of overstable global modes due to accretion-ejection 
762: instability.
763: 
764: Both of these boundary conditions correspond to zero loss of wave energy
765: at the inner boundary: If a wave from large radii impinges toward 
766: $r_{\rm in}$, the reflected wave will have the same amplitude.
767: However, the phase shifts due to reflection
768: differ in the two cases, and the resulting mode frequencies $\omega_r$ 
769: are different. Since the corotational wave amplification depends on
770: $\omega_r$, the mode growth rate $\omega_i$ will also be different.
771: 
772: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
773: \subsection{Numerical Results}
774: 
775: %====================================================
776: \begin{figure}
777: \centering
778: \begin{tabular}{ccc}
779: %\epsfig{file=wf1.eps,width=0.5\linewidth,clip=} &
780: %\epsfig{file=wf2.eps,width=0.5\linewidth,clip=}
781: \epsfig{file=f4a.eps,width=0.5\linewidth,clip=} &
782: \epsfig{file=f4b.eps,width=0.5\linewidth,clip=}
783: \end{tabular}
784: \caption{Example wavefunctions for disc p-modes. The upper and 
785: middle panels show $\delta h$ and $i\delta u_r$ (the solid lines for
786: the real part and dashed lines for the imaginary part),
787: the lower panels show the angular momentum flux, all in arbitrary 
788: units [with $\delta h(r_{\rm out})=1$]. The radius $r$ is in units of
789: $GM/c^2$. The disc sound speed is $c_s =0.1r\Omega$, and the $m=2$ 
790: modes are obtained using the inner boundary condition $\Delta 
791: P(r_{\rm ISCO}) = 0$. 
792: The left panels show the p-mode for the disc with 
793: a density profile $\Sigma \propto r^{-1}$, with 
794: the eigenvalues $\omega_r = 0.467 m\Omega_{\rm ISCO},~ \omega_i/\omega_r
795: = 0.0029$ [where $\Omega_{\rm ISCO}
796: =\Omega(r_{\rm ISCO})$]; the right panels show the mode for 
797: the disc with constant surface density profile ($p=0$), with
798: eigenvalues $\omega_r =0.464 m\Omega_{\rm ISCO}, ~
799: \omega_i/\omega_r = 0.00073$. 
800: Note that the model shown on the left panels has $r_c<r_{\rm peak}$
801: (the radius of peak vortensity) and thus $F(r_c+)<F(r_c-)$, 
802: while the model shown on the right panels has $r_c>r_{\rm peak}$
803: and $F(r_c+)>F(r_c-)$.
804: In both models there is a positive flux for $r > r_c$ due to the outward
805: propagating wave. The inserts on the left panels show the blowups
806: of the real wavefunctions near the corotation radius.
807: }
808: \label{fig4}
809: \end{figure}
810: 
811: %====================================================
812: \begin{figure}
813: \centering
814: \begin{tabular}{ccc}
815: \epsfig{file=f5a.eps,width=0.5\linewidth,clip=} &
816: \epsfig{file=f5b.eps,width=0.5\linewidth,clip=}
817: \end{tabular}
818: \caption{The real and imaginary frequencies of disc p-modes (with 
819: azimuthal wave numbers $m=1,2,3$)
820: as a function of the surface density index $p$ (where $\Sigma
821: \propto r^{-p}$). The modes are calculated assuming the inner 
822: boundary condition $\Delta P(r_{\rm ISCO}) = 0$.
823: The left panels are for discs with $c_s=0.1r\Omega$ and the right
824: panels for $c_s=0.2r\Omega$. 
825: The dotted lines denote the lower bound $\omega_r/m = \Omega_{\rm
826: peak}$ for which the corotational wave absorption 
827: acts to enhance mode growth.}
828: \label{fig5}
829: \end{figure}
830: 
831: %====================================================
832: \begin{figure}
833: \centering
834: \epsfig{file=f6.eps,width=0.5\linewidth,clip=}
835: \caption{The real and imaginary frequencies of disc p-modes
836: (with $m=2,3$) as a function of the normalized sound speed
837: $c_s/(r\Omega)$. The disc is assumed to have a constant
838: density profile ($p = 0$), and the inner boundary condition
839: is $\Delta P(r_{\rm ISCO}) = 0$. The bottom panel shows
840: the $\nu$ parameters for the modes.}
841: \label{fig6}
842: \end{figure}
843: 
844: 
845: %====================================================
846: \begin{figure}
847: \centering
848: \begin{tabular}{ccc}
849: \epsfig{file=f7a.eps,width=0.5\linewidth,clip=} &
850: \epsfig{file=f7b.eps,width=0.5\linewidth,clip=}
851: \end{tabular}
852: \caption{The real and imaginary frequencies of disc p-modes (with 
853: azimuthal wave numbers $m=1,2,3$)
854: as a function of the surface density index $p$ (where $\Sigma
855: \propto r^{-p}$). The modes are calculated assuming the inner 
856: boundary condition $\delta u_r(r_{\rm ISCO}) = 0$.
857: The left panels are for discs with $c_s=0.1r\Omega$ and the right
858: panels for $c_s=0.2r\Omega$. 
859: The dotted lines denote the lower bound $\omega/m = \Omega_{\rm
860: peak}$ for which the corotational wave absorption 
861: acts to enhance mode growth.}
862: \label{fig7}
863: \end{figure}
864: %====================================================
865: 
866: We solve for the complex eigen-frequency $\omega = \omega_r + i
867: \omega_i$ using the shooting method, with a 
868: fifth-order Runge-Kutta integrator (Press et al.~1992). As
869: discussed in section 4.1 we only calculate the growing modes ($\omega_i>0$). We
870: consider disc models with different surface density profile (characterized
871: by the index $p$), sound speed $c_s$, and inner boundary
872: conditions. For a given set of disc parameters 
873: and the azimuthal mode wavenumber $m$, the 
874: lowest order (highest frequency) mode has the best chance
875: of being overstable. This is easily understood from our discussion in
876: section 3 (see Fig.~1): a low-frequency wave has to penetrate a wider 
877: evanescent barrier for the corotational amplifier to be effective,
878: and when $\omega_r<m\Omega_{\rm peak}$ the corotation resonance acts to
879: damp the mode. For most disc models we have considered, 
880: the lowest order mode (of a given $m$)
881: is the only mode that has $\omega_i>0$.
882: 
883: Our numerical results are presented in Figures \ref{fig4}--\ref{fig7}.
884: Figure 4 gives two examples of the eigenfunctions of overstable trapped 
885: p-modes, obtained with the inner boundary condition $\Delta P=0$.
886: In addition to $\delta h$ and $\delta u_r$, we also plot the
887: angular momentum flux carried by the wave across the disc
888: (e.g., Goldreich \& Tremaine 1979; Zhang \& Lai 2006)
889: \be
890: F(r) = \pi r^2\Sigma\, {\rm Re} \left(\delta u_r \delta u_\phi^\ast\right)=
891: \frac{\pi m r \Sigma}{D}\,{\rm Im}\left(\delta h\frac{d\delta
892: h^*}{dr} \right),
893: \label{eq:flux}\ee
894: where the second equality follows from equations 
895: (\ref{eq:sig})-(\ref{eq:uphi}) (with $u_r=0$).
896: We see from Fig.~4 that outside the corotation radius ($r_c$), 
897: $F$ is nearly constant since only
898: the outgoing wave exists in this region and the wave action is conserved
899: in the limit of $\omega_i\ll \omega_r$. Inside $r_c$, the interference between
900: the ingoing and outgoing waves gives rise to the variation of $F$. At 
901: $r_{\rm ISCO}$, $F$ approaches zero since no wave action is lost through 
902: the disc inner boundary when $\Delta P=0$.\footnote{
903: Note that equation (\ref{eq:flux}) is the time averaged flux and is defined
904: for waves with real $\omega$.
905: Using equation (\ref{eq:DeltaP}) and 
906: $D\delta u_r=i\tomega d\delta h/dr-(2im\Omega/r)\delta h$ (obtained from 
907: eqs.~[\ref{eq:ur}]-[\ref{eq:uphi}]) it is easy to
908: show that $F(r_{\rm in})=0$ exactly for real $\omega$.}
909: Figure 4 also shows a flux jump across the corotation.
910: In the limit of $\omega_i\ll \omega_r$, $d\delta h/dr$,
911: $\delta u_r$ and $\delta u_\phi$
912: are discontinuous across $r_c$
913: (although $\delta h$ is continuous), giving rise to the flux discontinuity
914: (see Tsang \& Lai 2008a):
915: \be
916: F(r_c+)-F(r_c-)=-{2\pi^2 m r\Sigma \nu \kappa\over c_s D}|\delta h|^2
917: \Bigl|_{r_c}.
918: \ee
919: This discontinuity signifies the corotational wave absorption, the
920: sign of which depends on $\nu\propto (d\zeta/dr)_{r_c}$, as discussed in
921: section 3. Thus, the model
922: shown on the left panels of Fig.~4 has $r_c<r_{\rm peak}$ and
923: $\nu>0$, and the mode growth is primarily driven by wave absorption 
924: at the corotation. The model shown on the right panels of Fig.~4, 
925: on the other hand, has $r_c>r_{\rm peak}$ and $\nu<0$, thus the corotational
926: wave absorption acts to damp the mode, and the overall growth of the mode
927: is due to the outgoing wave beyond $r_c$, and the growth rate 
928: is much smaller than the model shown on the left panels.
929: 
930: Figures 5--6 show the frequencies of the
931: fundamental (no node/highest frequency) growing 
932: p-modes (with $m=1,2,3$) for various disc parameters, again obtained with 
933: the inner boundary condition $\Delta P=0$. We consider 
934: $p$ in the range between $-1.5$ and 1.5, and $c_s$ 
935: up to $0.3r\Omega$. 
936: %As seen in Fig.~1, the modes are trapped between the inner boundary
937: %and the inner Lindblad resonance. 
938: For a given sound speed,
939: the real mode frequency $\omega_r$ depends very weakly on $p$, but
940: the growth rate $\omega_i$ increases with $p$ (see Fig.~5) since
941: a larger value of $p$ leads to a larger $\nu$ and enhanced wave
942: absorption at the corotation [see equation (\ref{eq:nu})].
943: In general, as the sound speed 
944: increases, the effective wavelength of the mode increases, and 
945: $\omega_r$ decreases in order ``fit in'' the trapping zone between
946: $r_{\rm in}$ and $r_{\rm IL}$ (see Fig.~6). 
947: The mode growth rates $\omega_i$ depends 
948: on $c_s$ in a non-monotonic way because of two competing effects:
949: As $c_s$ increases, less attenuation occurs in the evanescent zone, 
950: and more wave energy can be absorbed at the corotation and
951: propagate to the outer edge of the disc; these tend to increase
952: $\omega_i$. On the other hand, increasing $c_s$ also leads to
953: smaller $\omega_r$, which shifts the corotation resonance to 
954: a larger radius and leads to decreasing $\nu$ and $\omega_i$.
955: 
956: Note that the growing modes shown in Fig.~5 extend below the 
957: $\omega_r/m=\Omega_{\rm peak}$ boundary due to the propagation of 
958: waves beyond the corotation radius, as discussed in section 3
959: [see eqs.~(\ref{eq:nucrit})].
960: %up to the point that $\nu \approx -
961: %\frac{1}{2\pi} e^{-2\Theta_{\rm IIb}}$, where $\Theta_{\rm IIb} \equiv
962: %\int_{r_{\rm c}}^{r_{\rm OL}} \sqrt{\kappa^2 - \tomega_r^2}/c$. 
963: Such modes (with $\nu_{\rm crit}<\nu<0$) grow
964: significantly slower than the modes with $\nu>0$ 
965: as the flux is attenuated by the entire barrier between $r_{\rm IL}$ and $r_{\rm OL}$.
966: 
967: For comparison, Figure 7 shows the disc mode
968: frequencies and growth rates
969: when the inner boundary condition $\delta u_r=0$ is
970: adopted. The different boundary condition leads to a different phase shift
971: $\varphi$ and higher mode frequency, but 
972: the results are similar to those illustrated in Fig.~5.
973: In particular, as $p$ increases, $\omega_r$ remains approximately
974: constant while $\omega_i$ increases.
975: 
976: 
977: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
978: \section{Effect of Radial Inflow on the P-Mode Growth Rate}
979: 
980: Our mode calculations presented in section 4 neglect 
981: the radial velocity of the accretion flow and assume
982: a loss-less inner disc boundary condition (either $\Delta P=0$ or 
983: $\delta u_r=0$ at $r_{\rm ISCO}$). In real discs, the radial 
984: inflow velocity $u_r$ is not negligible as $r$ approaches 
985: $r_{\rm ISCO}$, and the flow goes through a transonic point (where
986: $u_r=-c_s$) at a radius very close to $r_{\rm ISCO}$. We expect 
987: that part of the fluid perturbations may be advected into the
988: black hole and the inner disc boundary will not be completely
989: loss-less. Here we study the effect of the transonic flow
990: on the p-mode growth rate.
991: 
992: We note that just as the accretion disc is not laminar but 
993: turbulent, the accretion flow around $r_{\rm ISCO}$
994: is complicated. General relativistic MHD simulations in 3D
995: are only beginning to shed light on the property of the black hole
996: accretion flow (e.g., Beckwith, Hawley \& Krolik 2008; Shafee et al.~2008;
997: Noble, Krolik \& Hawley 2008), and many uncertainties remain 
998: unresolved. Here, to make analytic progress, we adopt a simple viscous
999: transonic flow model, which qualitatively describes the inner accretion flow
1000: of the black hole as long as the flow remains 
1001: geometrically thin (see Afshordi \& Paczynski 2003).
1002: 
1003: 
1004: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1005: \subsection{Boundary Condition at the Sonic Point}
1006: 
1007: We rewrite equations (\ref{eq:sig})-(\ref{eq:ur}) as
1008: \ba
1009: &&{u_r\over c_s^2}\delta h'+\delta u_r'=\left[{i\tomega\over c_s^2}
1010: -u_r\left(c_s^{-1}\right)'\right]\delta h+{u_r'\over u_r}\,\delta u_r-{im\over r}
1011: \,\delta u_\phi\equiv {\cal A}_1,\\
1012: &&\delta h'+u_r\,\delta u_r'=(i\tomega-u_r')\,\delta u_r+2\Omega\,\delta u_\phi
1013: \equiv {\cal A}_2,
1014: \ea
1015: where $'$ stands for $d/dr$ and we have used $r\Sigma u_r=$ constant for 
1016: the background flow. Solving for $\delta h'$ and $\delta u_r'$ we have
1017: \ba
1018: &&\delta h'={{\cal A}_2-{\cal A}_1u_r\over 1-u_r^2/c_s^2},\\
1019: &&\delta u_r'={{\cal A}_1-(u_r/c_s^2){\cal A}_2 \over 1-u_r^2/c_s^2}.
1020: \ea
1021: Clearly, in order for the perturbation to be regular at the sonic
1022: point $r_s$, where $u_r=-c_s$, we require
1023: \be
1024: {\cal A}_2+c_s{\cal A}_1=0\quad {\rm at}~~r=r_s.
1025: \label{eq:calA}\ee
1026: For definiteness, we characterize the variations of $\Sigma$ and $c_s$ at the
1027: sonic point $r_s\simeq r_{\rm ISCO}$ by the two length scales: 
1028: \be
1029: \left({\Sigma'\over \Sigma}\right)_{r_s}={1\over L_\Sigma},\qquad
1030: \left({c_s'\over c_s}\right)_{r_s}={1\over L_c}.
1031: \label{eq:Lc}\ee
1032: From $r\Sigma u_r=$ constant, we also have $u_r'=c_s (r_s^{-1}+L_\Sigma^{-1})$ at $r=r_s$.
1033: Then equation (\ref{eq:calA}) becomes
1034: \be
1035: \left({i\tomega\over c_s}-{2\over L_c}\right)\delta h
1036: +\left[i\tomega-2c_s\left({1\over r}+{1\over L_\Sigma}\right)\right]\delta u_r
1037: +\left(2\Omega-{imc_s\over r}\right)\delta u_\phi=0\quad {\rm at}~~r=r_s.
1038: \label{eq:bc}\ee
1039: This is the boundary condition for the fluid perturbations at the sonic point.
1040:  
1041: 
1042: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1043: \subsection{Properties of the Transonic Flow}
1044: 
1045: Before exploring the effect the radial inflow on the disc modes, we first 
1046: estimate the length scale for the surface density variation, 
1047: $L_\Sigma$, using the viscous slim disc model (e.g.,
1048: Muchotrzeb \& Paczynski 1982; Matsumoto et al.~1984;
1049: Abramowicz et al.~1988)
1050: 
1051: The basic steady-state slim disc equations are
1052: \ba
1053: &&\dot M=-2\pi r\Sigma u_r,\\
1054: &&u_ru_r'=-c_s^2{\Sigma'\over\Sigma}+(\Omega^2-\Omega_K^2)r,\\
1055: &&{\dot M} l_0={\dot M} l+2\pi\nu_{\rm vis} r^3\Sigma\Omega',\label{eq:l0}
1056: \ea
1057: where $l=r^2\Omega$ is the specific angular momentum of the flow, 
1058: $\Omega$ is the actual rotation rate,
1059: $\Omega_K$ is given by equation (\ref{eq:OmegaK}), $\nu_{\rm vis}$ 
1060: is the kinetic viscosity,
1061: and $l_0$ is the eigenvalue that must be solved so that flow pass through the 
1062: sonic point smoothly. We shall use the $\alpha$-disc model,
1063: so that $\nu_{\rm vis}=\alpha Hc_s$, with $H\simeq c_s/\Omega_K$.
1064: 
1065: To estimate $L_\Sigma$, we assume $l(r)\simeq r^2\Omega_K(r)$ for $r\go r_{\rm ISCO}$
1066: and $l_0\simeq l_K(r_{\rm ISCO})$. Equation (\ref{eq:l0})
1067: gives 
1068: \be
1069: u_r(1-l_0/l)\simeq -3\nu_{\rm vis}/(2r)=-3\alpha Hc_s/(2r),
1070: \ee
1071: valid for $r\go r_{\rm ISCO}$. At the radius $r=r_{\rm ISCO}+\Delta r$, we have
1072: $u_r(\Delta r/r_{\rm ISCO})^2\simeq -4\alpha Hc_s/r_{\rm ISCO}$.
1073: The sonic point ($u_r=-c_s$) is at $\Delta r\simeq 2\sqrt{\alpha Hr_{\rm ISCO}}$,
1074: and $u_r=-c_s/2$ at $\Delta r\simeq 2\sqrt{2\alpha Hr_{\rm ISCO}}$. Thus
1075: $u_r'(r_s)\sim c_s/L_\Sigma$, with 
1076: \be
1077: L_\Sigma=\left({\Sigma\over\Sigma'}\right)_{r_s}\sim
1078: 2\sqrt{\alpha Hr_s}=2\sqrt{\alpha\beta}\,r_s,
1079: \ee
1080: where $\beta=c_s/(r\Omega)$. This should be compared to the disc thickness
1081: $H=\beta r$: depending on the value of $\alpha$, both $L_\Sigma<H$ and $L_\Sigma>H$ are
1082: possible.
1083: 
1084: The value and sign of $L_c$ depend on the thermodynamical and radiative
1085: properties of the flow, and cannot be estimated in a simple way.
1086: It is reasonable to expect $|L_c|\sim r_s$.
1087: 
1088: 
1089: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1090: \subsection{Reflectivity at the Sonic Point}
1091: 
1092: 
1093: \begin{figure}
1094: \centering
1095: \begin{tabular}{ccc}
1096: \epsfig{file=f8a.eps,width=0.5\linewidth,clip=} &
1097: \epsfig{file=f8b.eps,width=0.5\linewidth,clip=}
1098: \end{tabular}
1099: \caption{The wave reflectivity at the transonic point of the inner disc
1100: as a function of the parameter $L_\Sigma/H$, for $r_s/L_c=0$ (the left panel)
1101: or $r_s/L_c=-3$ (the right panel). In both panels, the heavier lines are for
1102: $c_s=0.1r\Omega$ and the lighter lines for $c_s=0.05r\Omega$. The 
1103: short-dashed, solid and long-dashed lines are for $m=1,2,3$, respectively.
1104: The wave frequency is set to be $\omega=0.7m\Omega(r_s)$.
1105: }
1106: \label{fig8}
1107: \end{figure}
1108: 
1109: 
1110: We can understand qualitatively the effect of the transonic boundary
1111: condition on the p-mode by calculating the reflectivity $\cR_s$
1112: of the inner boundary.
1113: 
1114: Consider a density wave $\delta h\propto \exp(i\int^r k\,dr)$
1115: in the wave zone $r_{\rm in}=r_s<r<r_{\rm IL}$, traveling toward
1116: the inner disc boundary.
1117: \footnote{Note that since the group velocity 
1118: of the wave has opposite sign as the phase velocity for $r<r_{\rm IL}$,
1119: the wave of the form $\exp(i\int^r k\,dr)$ (with $k>0$) is inward propagating.}. 
1120: Upon reflection,
1121: the wave becomes $\delta h\propto \cR_s\exp(-i\int^r k\,dr)$. 
1122: Including the correct WKB amplitude (see Tsang \& Lai 2008a),
1123: the wave outside the sonic point can be written as (up to a constant 
1124: prefactor)
1125: \be
1126: \delta h=A\left[\exp\left(i\int^r_{r_s}k\,dr\right)
1127: +\cR_s \exp\left(-i\int^r_{r_s} k\,dr\right)\right],
1128: \qquad (r_s<r<r_{\rm IL})
1129: \label{eq:deltah}\ee
1130: where 
1131: \be
1132: k={(-D)^{1/2}\over c_s},\quad
1133: A=\left({D\over r\Sigma k}\right)^{1/2}.
1134: \ee
1135: To apply the boundary condition (\ref{eq:bc})
1136: to equation (\ref{eq:deltah}), we neglect
1137: $u_r$ in equations (\ref{eq:sig})--(\ref{eq:uphi}) at $r=r_s+\varepsilon$,
1138: with $\varepsilon\ll r_s$ and $r_s\simeq r_{\rm ISCO}$. Implicit in
1139: this procedure is the assumption that the fluid perturbations
1140: do not vary significantly between $r_s$ and $r_s+\varepsilon$. 
1141: We then obtain
1142: \be
1143: \cR_s={ik+L_A^{-1}+K-2m\Omega/(r\tomega)\over
1144: ik-L_A^{-1}-K+2m\Omega/(r\tomega)}\Biggl|_{r_s},
1145: \ee
1146: where
1147: \be 
1148: L_A^{-1}=\left({A'\over A}\right)_{r_s},
1149: \ee
1150: and
1151: %\be K=\left({\tomega^2L_\Sigma\over 2c_s^2}\right)
1152: %{1-\left({mc_s\over r\tomega}\right)^2-i\left({2mc_s\Omega\over 
1153: %r\tomega^2}-{2c_s\over \tomega L_c}\right)\over
1154: %1+{L_\Sigma\over r}-i{\tomega L_\Sigma\over 2c_s}}\Biggl|_{r_s}. \ee
1155: \be
1156: K=\left({\tomega^2L_\Sigma\over 2c_s^2}\right)
1157: {1-\left({mc_s/r\tomega}\right)^2-i\left[{2mc_s\Omega/ 
1158: (r\tomega^2)}-{2c_s/(\tomega L_c)}\right]\over
1159: 1+({L_\Sigma/r})-i({\tomega L_\Sigma/2c_s})}\Biggl|_{r_s}.
1160: \label{eq:K}\ee
1161: 
1162: When considering the damping of the p-mode 
1163: %(trapped between $r_{\rm in}=r_s\simeq r_{\rm ISCO}$ and $r_{\rm IL}$) 
1164: due to the transonic flow, the quantity
1165: $|\cR_s|^2-1$ is the most relevant (see section 5.4 below). Let 
1166: \be
1167: K=|K|\exp(-i\psi),
1168: \ee
1169: we have
1170: \be
1171: |\cR_s|^2-1=-{4\,k\,|K|\sin\psi\over (L_A^{-1}+|K|\cos\psi)^2
1172: +(k+|K|\sin\psi)^2}\Biggl|_{r_s}.
1173: \ee
1174: Using $\beta =c_s/(r\Omega)$, $\tomega=-\hat\omega m\Omega_{\rm ISCO}$ (where
1175: $0<{\hat\omega}<1$), we find from equation (\ref{eq:K}) 
1176: that 
1177: \be
1178: \psi=\tan^{-1}\left({2\beta\over m{\hat\omega}^2}{1+r_s\hat\omega/L_c
1179: \over 1-\beta^2/\hat\omega^2}\right)
1180: +\tan^{-1}\left({m\hat\omega L_\Sigma\over 2\beta r_s}{1\over 1+L_\Sigma/r_s}
1181: \right).
1182: \ee
1183: 
1184: Figure 8 shows how the reflectivity depends on various parameters 
1185: of the disc inner edge. In particular, for small $L_\Sigma/H=L_\Sigma/(\beta r_s)$,
1186: i.e., when the surface density of the disc decreases rapidly at the sonic point, 
1187: $|\cR_s|^2$ is only slightly smaller than unity and the wave loss
1188: at the inner edge of the disc is small.
1189: 
1190: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1191: \subsection{Mode Growth Rate in the WKB Approximation}
1192: 
1193: Consider the p-mode trapped between $r_{\rm in}=r_s\simeq r_{\rm ISCO}$ and
1194: $r_{\rm IL}$. With the reflectivity
1195: at $r_{\rm IL}$ given by $\cR$ (see section 3), we can write the wave amplitude 
1196: for $r<r_{\rm IL}$ as\footnote{Note that this definition of $\cR$
1197: differs from that in Tsang \& Lai (2008a) by a phase factor of
1198: $\exp(i\pi/4)$.}
1199: \be
1200: \delta h \propto \left({D\over r\Sigma k}\right)^{1/2}\left[
1201: \exp\left(-i\int^r_{r_{\rm IL}}k\,dr\right)
1202: +\cR \exp\left(i\int^r_{r_{\rm IL}} k\,dr\right)\right],
1203: \qquad (r_s<r<r_{\rm IL})
1204: \label{eq:deltah2}\ee
1205: On the other hand, with the reflectivity at 
1206: $r_{\rm in}=r_s$ given by $\cR_s$, the wave can also be expressed as
1207: (\ref{eq:deltah}). For stationary waves we therefore require
1208: \be
1209: \exp(2i\Theta)=\cR\cR_s,\qquad {\rm with}~~\Theta=\int_{r_{\rm in}}^{r_{\rm IL}}k\,dr
1210: =\Theta_r+i\Theta_i,
1211: \ee
1212: where $\Theta_r$ and $\Theta_i$ are real. The real eigen-frequency 
1213: $\omega_r$ is given by
1214: \be
1215: \Theta_r=\int_{r_{\rm in}}^{r_{\rm IL}}k_r\,dr=
1216: \int_{r_{\rm in}}^{r_{\rm IL}}{\sqrt{\tomega_r^2-\kappa^2}\over c_s}\,dr=
1217: n\pi+{\varphi\over 2},
1218: \ee
1219: where $\cR\cR_s=|\cR\cR_s|\exp(i\varphi)$, and $n$ is an integer. The mode growth
1220: rate $\omega_i$ is determined by
1221: $|\cR\cR_s|=\exp(-2\Theta_i)$, or
1222: \be
1223: \tanh\Theta_i=-\left({|\cR\cR_s|-1\over |\cR\cR_s|+1}\right).
1224: \ee
1225: For $\Theta_i=\int_{r_{\rm in}}^{r_{\rm IL}}k_i\,dr\ll 1$ and
1226: $k_i\simeq \omega_i\tomega_r/(c_s\sqrt{\tomega_r^2-\kappa^2})$, we obtain
1227: \be
1228: \omega_i = \left(\frac{|\cR\cR_s| -1}{|\cR\cR_s|+1}\right)
1229: \left[\int_{r_{\rm in\ }}^{r_{\rm IL}}
1230: {|\tomega_r|\over c_s\sqrt{\tomega_r^2 - \kappa^2}} dr\right]^{-1},  
1231: \label{eq:omegai2}\ee
1232: where we have assumed $\tomega_r<0$.
1233: Equation (\ref{eq:omegai2}) is to be compared with (\ref{eq:omegai}), 
1234: where perfect reflection at $r_{\rm in}$ is assumed.
1235: Clearly, to obtain growing modes we require $|\cR\cR_s|>1$. For a given
1236: $|\cR|>1$, growing modes are possible only when the loss
1237: at the sonic point is sufficiently small (i.e., $|\cR_s|$ is sufficiently
1238: close to unity).
1239: 
1240: 
1241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1242: \subsection{Numerical Results}
1243: 
1244: 
1245: \begin{figure}
1246: \centering
1247: \epsfig{file=f9.eps,width=0.5\linewidth,clip=}
1248: \caption{Wavefunctions for a disc p-mode. The 
1249: notations are the same as in Fig.~4.
1250: The disc has sound speed $c_s =0.1r\Omega$ and constant density profile
1251: ($p=0$), and the $m=2$ mode is calculated using 
1252: the transonic inner boundary condition (\ref{eq:bc}) with 
1253: $L_\Sigma/H=0.25$ and $L_c=\infty$.
1254: The eigenvalues are $\omega_r = 0.725m\Omega_{\rm ISCO}$ and
1255: $\omega_i/\omega_r=0.00267$.
1256: }
1257: \label{fig9}
1258: \end{figure}
1259: 
1260: 
1261: \begin{figure}
1262: \centering
1263: \begin{tabular}{ccc}
1264: \epsfig{file=f10a.eps,width=0.5\linewidth,clip=} &
1265: \epsfig{file=f10b.eps,width=0.5\linewidth,clip=}
1266: \end{tabular}
1267: \caption{The real and imaginary frequencies of disc p-modes (with 
1268: azimuthal wave numbers $m=1,2,3$)
1269: as a function of $L_\Sigma/H$ [see eq.~(\ref{eq:Lc})]
1270: The modes are calculated using the transonic inner 
1271: boundary condition (\ref{eq:bc}) with $L_c=\infty$.
1272: The disc has a constant surface density profile and the sound 
1273: speed is $c_s=0.1r\Omega$ (left panels) or $0.05r\Omega$ (right panels).
1274: }
1275: \label{fig10}
1276: \end{figure}
1277: 
1278: 
1279: \begin{figure}
1280: \centering
1281: \begin{tabular}{ccc}
1282: \epsfig{file=f11a.eps,width=0.5\linewidth,clip=} &
1283: \epsfig{file=f11b.eps,width=0.5\linewidth,clip=}
1284: \end{tabular}
1285: \caption{The real and imaginary frequencies of disc p-modes (with 
1286: azimuthal wave numbers $m=1,2,3$).
1287: The modes are calculated using the transonic inner 
1288: boundary condition (\ref{eq:bc}). 
1289: The disc has a constant surface density profile and the sound 
1290: speed is $c_s=0.1r\Omega$. The left panels show the cases with
1291: $L_c=-r_s/3$ by varying $L_\Sigma/H$, and right panels show the
1292: cases with $L_\Sigma/H=0.25$ by varying $r_s/L_c$ 
1293: [see eq.~(\ref{eq:Lc})].
1294: }
1295: \label{fig11}
1296: \end{figure}
1297: 
1298: We solve equations (\ref{eq:sig})-(\ref{eq:uphi}) (with $u_r=0$)
1299: subjected to the radiative outer boundary condition (\ref{eq:outer})
1300: and the transonic inner boundary condition (\ref{eq:bc}).
1301: 
1302: Figure 9 depicts an example of the p-mode wavefunctions.
1303: Again, the discontinuity in the angular momentum flux $F$ at 
1304: $r_c$ signifies wave absorption; since $r_c<r_{\rm peak}$, 
1305: this leads to mode growth.
1306: Comparing with Fig.~4, here the angular momentum flux at $r_{\rm in}$ is
1307: significantly nonzero, indicating wave loss through the sonic point.
1308: Nevertheless, the corotational instability is sufficiently strong to
1309: overcome the loss and makes the mode grow.
1310: 
1311: Figures 10--11 show the fundamental p-mode frequencies and growth rates
1312: as a function of the disc parameters. Consistent with the result of
1313: section 5.3 (see Fig.~8), growing modes are obtained for sufficiently
1314: small $L_\Sigma$. Negative $L_c$ also tends to reduce wave loss
1315: at $r_s$ and make the growing modes possible.
1316: Such values of $L_\Sigma$ and $L_c$ are not unreasonable for 
1317: black hole accretion discs.
1318: 
1319: It is important to note that while the mode growth rates $\omega_i$ depend
1320: sensitively on the inner disc parameters, particularly the physical
1321: property of the transonic flow near the ISCO, the real mode
1322: frequencies $\omega_r$ show only weak dependence on
1323: the inner disc parameters (e.g., $\omega_r$ decreases with increasing
1324: sound speed; see Fig.~6).
1325:  Thus we may expect that kHz
1326: QPOs appear only in certain accretion states of the black hole,
1327: and the frequencies do not vary much as the accretion rate changes.
1328: 
1329: 
1330: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1331: \section{The Role of Rossby Wave Instability}
1332: 
1333: Lovelace et al.~(1999) (see also Li et al.~2000) have shown that 
1334: when the vortensity $\zeta=\kappa^2/(2\Omega\Sigma)$ has an extremum
1335: at a certain radius ($r_{\rm peak}$) in the disc\footnote{Lovelace et al.
1336: considered non-barotropic flows, so the ``generalized vortensity'' depends
1337: on the entropy profile of the disc.}, it is possible to form 
1338: normal Rossby modes around $r_{\rm peak}$. If the trapped Rossby waves
1339: can propagate on both sides of the corotation, a standing pattern of waves of 
1340: opposite energies are formed, making the mode unstable --- This is the
1341: ``Rossby wave instability''. Tagger \& Varniere (2006) have considered
1342: the MHD version of the instability and suggested that it played a role
1343: in the diskoseismic modes around black holes (see also Tagger 2006).
1344: 
1345: We do not find any trapped Rossby modes in our calculation. To clarify the
1346: issue in light of works by Lovelace et al. and by Tagger \& Varniere,
1347: let us consider equation (\ref{perturbeq1}) and define the effective potential
1348: \be
1349: V_{\rm eff}(r)=
1350: \frac{2m\Omega}{r\tomega}\left(\frac{d}{dr}\ln\frac{\Omega\Sigma}{D}\right)
1351: +\frac{m^2}{r^2}+ \frac{D}{c_s^2}.
1352: \ee
1353: The wave equation can be approximated by $(d^2/dr^2-V_{\rm eff})\delta h\simeq 0$
1354: (see Tsang \& Lai 2008a). We will focus on modes with $r_c$ very close to $r_{\rm peak}$ 
1355: (i.e., $|r_c-r_{\rm peak}|\ll r_c$, so that 
1356: $\omega_r\simeq m\Omega_{\rm peak}$; see Fig.~1). For 
1357: $|r-r_{\rm peak}|\ll r_{\rm peak}$ in a thin disc (so that $m^2/r^2$ can be
1358: neglected compared to $\kappa^2/c_s^2$), the effective potential becomes
1359: \be
1360: V_{\rm eff}(r)\simeq -\frac{2m\Omega}{r\tomega}\left(\frac{d}{dr}\ln\zeta
1361: \right)+{\kappa^2\over c_s^2}
1362: \simeq {2\over qL_\zeta^2}\left({r-r_{\rm peak}\over r-R_c}\right)
1363: +{\kappa^2\over c_s^2},
1364: \label{eq:veff2}\ee
1365: where in the second equality we have used 
1366: $R_c=r_c-i(r_c\omega_i/q\omega_r)$, $\Omega\propto r^{-q}$, and 
1367: defined $L_\zeta$ via
1368: \be
1369: {d\ln\zeta\over dr}=-{r-r_{\rm peak}\over L_\zeta^2},\qquad ({\rm for}~
1370: |r-r_{\rm peak}|\ll r_{\rm peak}).
1371: \ee
1372: Consider the case $r_c<r_{\rm peak}$ and assume $\omega_i\ll \omega_r$. 
1373: The Rossby wave zone (where $V_{\rm eff}<0$) lies between $r_c$ and $r_c+\Delta r_R$, 
1374: with 
1375: \be
1376: \Delta r_R\simeq {r_{\rm peak}-r_c\over (q/2)(L_\zeta/H)^2},
1377: \ee
1378: where $H\simeq c_s/\kappa$, and we have used $L_\zeta/H\sim r/H\gg 1$. The
1379: number of wavelengths in the Rossby zone is 
1380: \be
1381: \int_{r_c}^{r_c+\Delta r_R}\!\!k\,dr=
1382: \int_{r_c}^{r_c+\Delta r_R}\!\!(-V_{\rm eff})^{1/2}\,dr
1383: \sim {4H (r_{\rm peak}-r_c)\over qL_\zeta^2}
1384: \ee
1385: Two points should be noted: (i) Since $\int_{r_c}^{r_c+\Delta r_R}\!k\,dr\ll 1$
1386: for $L_\zeta\sim r_{\rm peak}$, no stationary wave can form in the Rossby zone;
1387: (ii) since the Rossby zone lies only on one side of the corotation radius,
1388: even if the mode can be trapped it will not grow by the Rossby wave instability
1389: mechanism. Similar result can be obtained for the $r_c>r_{\rm peak}$ case.
1390: We conclude that for the ``smooth'' vortensity maximum (with length scale
1391: $L_\zeta\sim r$; see the lower panel of Fig.~1) considered in this paper,
1392: there is no trapped Rossby mode around $r_{\rm peak}$ 
1393: and the Rossby wave instability is ineffective.
1394: 
1395: In the hypothetical situation where the vortensity $\zeta$ has a {\it minimum}
1396: at $r=r_{\rm min}$, equation (\ref{eq:veff2}) should be replaced by
1397: \be
1398: V_{\rm eff}(r)\simeq -{2\over qL_\zeta^2}\left({r-r_{\rm min}\over r-R_c}\right)
1399: +{\kappa^2\over c_s^2},
1400: \ee
1401: where we have used 
1402: \be
1403: {d\ln\zeta\over dr}={r-r_{\rm min}\over L_\zeta^2},\qquad ({\rm for}~
1404: |r-r_{\rm min}|\ll r_{\rm min}).
1405: \ee
1406: In this case, for a mode with $\omega_r=m\Omega(r_{\rm min})$
1407: (or $r_c=r_{\rm min}$), we find 
1408: $V_{\rm eff}(r_c)\simeq -2/(qL_\zeta^2)+1/H^2$ (for $\omega_i\ll\omega_r$). 
1409: When $V_{\rm eff}(r_c)<0$, or when
1410: \be
1411: L_\zeta<\left({2\over q}\right)^{1/2}\!\!H,
1412: \ee
1413: Rossby waves can propagate on both sides of 
1414: the corotation, leading to mode growth
1415: --- this is the Rossby wave instability.
1416: Thus, the Rossby wave instability would operate if there existed 
1417: a ``sharp'' vortensity {\it minimum} in the disc (with $\zeta$ varying on the 
1418: lengthscale comparable or less than the disc thickness)
1419: --- this is not the case for typical black hole accretion discs
1420: considered in this paper.
1421: 
1422: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1423: \section{Discussion}
1424: 
1425: High-frequency QPOs (HFQPOs) in black-hole X-ray binaries have been studied
1426: observationally for more than a decade now and they provide a
1427: potentially important tool for studying the strong gravitational
1428: fields of black holes (see Remillard \& McClintock 2006).
1429: Despite much theoretical effort,
1430: the physical mechanisms that generate these QPOs remain unclear
1431: (see section 1.1 for a brief review of existing theoretical models).
1432: Ultimately, numerical simulations of realistic accretion discs
1433: around black holes may provide the answer.
1434: However, such simulations are still at their early stage of
1435: development and have their own limitations (e.g., De Villiers \& Hawley 2003; 
1436: Machida \& Matsumoto 2003, 2008; Arras et al.~2006; Fragile et al.~2007;
1437: Reynolds \& Miller 2008; Beckwith et al.~2008; Shafee et al.~2008;
1438: Noble et al.~2008), semi-analytical study remains a useful,
1439: complementary approach in order to identify the key physics involved.
1440: 
1441: In this paper, we have studied the global instability of the
1442: non-axisymmetric p-modes in black-hole accretion discs.  These modes
1443: have frequencies $\omega\sim (0.5-0.7) m\Omega_{\rm ISCO}$ (where $m$
1444: is the azimuthal wave number, $\Omega_{\rm ISCO}$ is the disc rotation
1445: frequency at the inner-most stable circular orbit), where the
1446: pre-factor (0.5-0.7) depends on the inner disc structure.  Recent
1447: works (Arras et al.~2006; Reynolds \& Miller 2008; Fu \& Lai 2008)
1448: suggested that, unlike other diskoseismic modes (g-modes and c-modes),
1449: the p-modes may be robust in the presence of disc magnetic fields and
1450: turbulence. Our linear analysis showed that due to GR effects, the
1451: p-modes may grow in amplitude due to wave absorptions at the corotation
1452: resonance. For a given $m$, only the lowest-order p-mode
1453: has sufficiently high frequency ($\omega>m\Omega_{\rm peak}$; see Fig.~1) 
1454: to be driven overstable by the corotational instability, while high-order
1455: (lower frequency) modes are damped by the corotational wave absorption.
1456: 
1457: The greatest uncertainty of our calculation of the p-mode growth rate
1458: concerns the boundary condition at the inner disc edge near the ISCO.
1459: In particular, the rapid radial inflow at the ISCO has the tendency to
1460: damp the mode (see Blaes 1986). While our analysis in section 5
1461: indicates that this damping does not completely suppress the mode
1462: growth under certain disc conditions, it suggests that mode growth may
1463: not always be achieved in real black-hole accretion 
1464: discs. Observationally, it is of interest to note that HFQPOs
1465: are observed only when the X-ray binaries are in the steep
1466: power-law state, while they do not appear in other spectral states
1467: (Remillard \& McClintock 2006). In particular, HFQPOs are absent
1468: in the thermal (soft-high) state, believed to correspond to geometrically thin 
1469: discs extending down to the ISCO. It is reasonable to expect that 
1470: in this state p-modes are damped due to the rapid radial inflow.
1471: 
1472: Our current understanding of the steep power-law state (also called
1473: very high state) of black-hole X-ray binaries is rather limited.  A
1474: thermal-radiation-emitting disc is suggested by spectral modelings,
1475: but it is not clear whether the disc is truncated at the ISCO or
1476: slightly larger radius (see Done et al.~2007). The observed power-law
1477: radiation component requires a significant corona that Compton
1478: up-scatters the disc thermal radiation. It is possible that in the
1479: steep power-law state, the inner disc behaves as a more reflective
1480: boundary (modeled in section 4) than a transonic flow (modeled in section 5), 
1481: and thus more robust p-mode growth can be achieved. 
1482: One possibility is that a significant magnetic field
1483: flux can accumulate in the inner disc when the disc accretion rate is
1484: sufficiently high (see Bisnovtyi-Kogan \& Lovelace 2007; Rothstein \& Lovelace 2008
1485: and references therein). Such a magnetic field may also
1486: enhance the corotational instability and induce variability 
1487: in the power-law radiation flux (see Tagger \& Varniere 2006).
1488: 
1489: Although the p-mode growth rates depend sensitively on a number of
1490: (uncertain) disc parameters (particularly those related to the inner
1491: disc boundary), the mode frequencies are more robust (see Figs.~5-7,
1492: 10-11). More precisely, the real mode frequency can be written as
1493: $\omega_r={\bar\omega}m\Omega_{\rm ISCO}$, where $\bar\omega<1$
1494: depends weakly on $m$ and has only modest dependence on 
1495: disc parameters (e.g. sound speed). This implies a
1496: commensurate frequency ratio as observed in HFQPOs (note that in some
1497: of our models, the $m=2,3$ modes have the largest growth rates; see
1498: Fig.~5). The fact that $\bar\omega<1$ would also make the numerical values
1499: of the p-mode frequencies more compatible with the measurements of the QPO
1500: frequencies and black hole masses. 
1501: 
1502: We note that our calculations in this work are done with a pseudo-Newtonian potential. For direct comparison with observations a fully general relativistic calculation\footnote{Previous work on relativistic diskoseismic g-modes has been done by Perez et al. (1992) and Silbergleit \& Wagoner (2008) while the c-mode was studied by Silbergleit et al. (2001). Axisymmetric p-modes were studied using a general relativistic formalism by Ortega-Rodriguez et al. (2002), but these do not include the effect of the corotation singularity.} is needed including a careful treatment of the corotation singularity. Including the effect of black hole spin would likely increase the value of $\omega$ by modifying $r_{\rm ISCO}$ and $\Omega_{\rm ISCO}$, while $\bar\omega$ will likely remain similar to the non-spinning case discussed above. We plan to study these effects in future work.
1503: 
1504: 
1505: \section*{Acknowledgments}
1506: 
1507: We thank Richard Lovelace for useful discussion. 
1508: This work has been supported in part by NASA Grant NNX07AG81G, NSF
1509: grants AST 0707628, and by {\it Chandra} grant TM6-7004X
1510: (Smithsonian Astrophysical Observatory).
1511: 
1512: 
1513: %\bibliographystyle{mn.bst}
1514: %\bibliography{pmodes}
1515: 
1516: %%%%%%%%%%%%%%%%%%%%%%
1517: \begin{thebibliography}{99}
1518: 
1519: \bibitem[]{}
1520: Afshordi, N., Paczynski, B. 2003, ApJ, 592, 354
1521: % Geometrically Thin Disk Accreting into a Black Hole
1522: 
1523: \bibitem[]{}
1524: Abramowicz, M.A., Czerny, B., Lasota, J. P., Szuszkiewicz, E. 1988,
1525: ApJ, 332, 646
1526: 
1527: \bibitem[]{}
1528: Abramowicz, M.A., Kluzniak, W. 2001, A\&A, 374, L19
1529: 
1530: \bibitem[]{}
1531: Abramowicz, M.A. et al 2007, Rev. Mexicana Astron. Astrofísica, 27, 8
1532: 
1533: \bibitem[]{}
1534: Arras, P., Blaes, O.M. \& Turner, N. J., 2006, ApJ, 645, L65
1535: 
1536: \bibitem[]{}
1537: Beckwith, K., Hawley, J.F., Krolik, J.H. 2008, MNRAS, in press
1538: (arXiv:0801.2974)
1539: %Where is the Radiation Edge in Magnetized Black Hole Accretion discs?
1540: 
1541: \bibitem[]{}
1542: Blaes, O.M. 1987, MNRAS, 227, 975
1543: 
1544: \bibitem[]{}
1545: Blaes, O.M., Arras, P., Fragile, P.C. 2006, MNRAS, 369, 1235
1546: %Oscillation modes of relativistic slender tori
1547: 
1548: \bibitem[]{}
1549: Blaes, O.M., Sramkova, E., Abramowicz, M. A., Kluzniak, W., Torkelsson, U., ApJ, 665, 642
1550: 
1551: \bibitem[]{}
1552: Balbus, S.A., Hawley, J.F. 1998, Rev. Mod. Phys., 70, 1.
1553: 
1554: \bibitem[]{}
1555: Bisnovatyi-Kogan, G.S., Lovelace, R.V.E. 2007, ApJ, 667, L167
1556: % Large-Scale B-Field in Stationary Accretion Disks
1557: 
1558: \bibitem[]{}
1559: De Villiers, J.-P., Hawley, J.F. 2003, ApJ, 592, 1060
1560: 
1561: \bibitem[]{}
1562: Done, C., Gierlinski, M., Kubota, A. 2007, Astron. Astrophys. Review, 15, 1
1563: 
1564: \bibitem[]{}
1565: Gierlinski, M., Middleton, M., Ward, M., Done, C. 2008, Nature, 455, 369
1566: % A periodicity of 1 hour in X-ray emission from the active galacy 
1567: % RE J1034_396
1568: 
1569: \bibitem[]{}
1570: Goldreich, P., Tremaine, S. 1979, ApJ, 233, 857
1571: 
1572: \bibitem[]{}
1573: Ferreira, B. T. \& Ogilvie, G. I., 2008, astro-ph/08031671
1574: 
1575: \bibitem[]{}
1576: Fragile, P.C., Blaes, O., Anninos, P., Salmonson, J.D. 2007, ApJ, 668, 417
1577: % Global General Relativistic Magnetohydrodynamic Simulation of a Tilted 
1578: % Black Hole Accretion Disk
1579: 
1580: \bibitem[]{}
1581: Fu, W., Lai, D. 2008, ApJ, in press (arXiv:0806.1938)
1582: 
1583: \bibitem[]{}
1584: Horak, J., Karas, V., 2006, A\&A, 451, 377
1585: 
1586: \bibitem[]{}
1587: Kato, S. \& Fukue, J., 1980, PASJ, 32, 377
1588: 
1589: \bibitem[]{}
1590: Kato, S., 1990, PASJ, 42, 99
1591: 
1592: \bibitem[]{}
1593: Kato, S.,  2001, PASJ, 53, 1
1594: 
1595: \bibitem[]{}
1596: Kato, S., 2003a, PASJ, 55, 257
1597: % g-mode damping
1598: 
1599: \bibitem[]{}
1600: Kato, S., 2003b, PASJ, 55, 801
1601: % Excitation by disk deformation
1602: 
1603: \bibitem[]{}
1604: Kato, S., 2008, PASJ, 60, 111
1605: 
1606: \bibitem[]{}
1607: Kluzniak, W. \& Abramowicz, M. A., 2002, astro-ph/0203314
1608: 
1609: \bibitem[]{}
1610: Lee, W. H., Abramowicz, M. A. \& Kluziniak, W., 2004, ApJ, 603, L93
1611: %Resonance in Forced Oscillations of an Accretion Disk and Kilohertz
1612: %Quasi-periodic Oscillations
1613: 
1614: \bibitem[]{}
1615: Li, H., Finn, J.M., Lovelace, R.V.E., Colgate, S.A. 2000, ApJ, 533, 1023
1616: 
1617: \bibitem[]{}
1618: Li, L., Goodman, J., Narayan, R., 2003, ApJ, 593, 980
1619: 
1620: \bibitem[]{}
1621: Li, L., Narayan, R., 2004, ApJ, 601, 414
1622: %Quasi-periodic Oscillations from Rayleigh-Taylor and Kelvin-Helmholtz
1623: %Instability at a Disk-Magnetosphere Interface
1624: 
1625: \bibitem[]{}
1626: Lifshitz, E.M., Pitaevskii, L.P. 1981, Physical Kinetics (Pergamon Press: Oxford)
1627: 
1628: \bibitem[]{}
1629: Lin, C.C. 1995, The Theory of Hydrodynamic Stability (Cambridge Univ. Press), 
1630: Chap.~8
1631: 
1632: \bibitem[]{}
1633: Lovelace, R.V.E., Li, H., Colgate, S.A., Nelson, A.F. 1999,
1634: ApJ, 513, 805
1635: 
1636: \bibitem[]{}
1637: Lovelace, R.V.E., Romanova, M.M., 2007, ApJ, 670, L13
1638: %Three Disk Oscillation Modes of Rotating Magnetized Neutron Stars
1639: 
1640: \bibitem[]{}
1641: Machida, M., Matsumoto, R., 2003, ApJ, 585, 429
1642: 
1643: \bibitem[]{}
1644: Machida, M., Matsumoto, R., 2008, PASJ, 60, 613
1645: 
1646: \bibitem[]{}
1647: Matsumoto, R., Kato, S., Fukue, J., Okazaki, A. T., 1984, PASJ, 36, 71
1648: 
1649: \bibitem[]{}
1650: Muchotrzeb, B., Paczynski, B. 1982, Acta Astron., 32, 1
1651: 
1652: \bibitem[]{}
1653: Narayan, R., Goldreich, P., Goodman, J. 1987, MNRAS, 228, 1
1654: 
1655: \bibitem[]{}
1656: Noble, S.C., Krolik, J.H., Hawley, J.F. 2008, ApJ, submitted 
1657: (arXiv:0808.3140)
1658: 
1659: \bibitem[]{}
1660: Nowak, M. A. \& Wagoner, R. V., 1991, ApJ, 378, 656
1661: 
1662: \bibitem[]{}
1663: Nowak, M. A. \& Wagoner, R. V., 1992, ApJ, 393, 697
1664: 
1665: \bibitem[]{}
1666: Okazaki, A. T., Kato, S. \& Fukue, J., 1987, PASJ, 39, 457
1667: 
1668: \bibitem[]{}
1669: Ortega-Rodriguez, M., Silbergleit, A. S. \& Wagoner, R. V., 2002, ApJ, 567, 1043
1670: 
1671: \bibitem[]{}
1672: Ortega-Rodriguez, M., Silbergleit, A. S. \& Wagoner, R. V., 2006, astro-ph/0611101
1673: 
1674: \bibitem[]{}
1675: Paczynski, B., Wiita, P.J. 1980, A\&A, 88, 23
1676: 
1677: \bibitem[]{}
1678: Petri, J. 2008, Astrophys. Space Science, in press (arXiv:0809.3115)
1679: 
1680: \bibitem[]{}
1681: Perez, C.A., Silbergleit, A.S., Wagoner, R.V., \& Lehr, D.E., 1997, ApJ, 476, 589
1682: 
1683: \bibitem[]{}
1684: Press, W.H., et al. 1998, Numerical Recipes (Cambridge Univ. Press)
1685: 
1686: \bibitem[]{}
1687: Rebusco, P., 2004, PASJ, 56, 553
1688: 
1689: \bibitem[]{}
1690: Rebusco, P., 2008, astro-ph/08013658
1691: 
1692: \bibitem[]{}
1693: Remillard, R. A. \& McClintock, J. E., 2006, ARAA, Vol. 44, pp. 49-92
1694: 
1695: \bibitem[]{}
1696: Reynolds, C. S. \& Miller, M. C., 2008, astro-ph/08052950
1697: 
1698: \bibitem[]{}
1699: Rezzolla, L., Yoshida, S'i., Maccarone \& Zanotti, O., 2003, MNRAS, 344, L37-L41
1700: %A new simple model for high-frequency quasi-periodic oscillations in 
1701: %black hole candidates
1702: 
1703: \bibitem[]{}
1704: Rothstein, D.M., Lovelace, R.V.E. 2008, ApJ, 677, 1221
1705: %Advection of Magnetic Fields in Accretion Disks: Not So Difficult After All
1706: 
1707: \bibitem[]{}
1708: Shafee, R., et al. 2008, ApJL, submitted (arXiv:0808.2860)
1709: %Three-Dimensional Simulations of Magnetized Thin Accretion 
1710: %Disks around Black Holes: Stress in the Plunging Region
1711: 
1712: \bibitem[]{}
1713: Silbergleit, A. S., Wagoner, R. V. \& Ortega-Rodriguez, M., 2001, ApJ, 548, 335
1714: 
1715: \bibitem[]{}
1716: Silbergleit, A. S. \& Wagoner, R. V., 2008, ApJ, 680, 1319
1717: 
1718: \bibitem[]{}
1719: Schnittman, J.D. 2005, ApJ, 621, 940 
1720: 
1721: \bibitem[]{}
1722: Schnittman, J.D., Bertschinger, E. 2004, ApJ, 606, 1098 
1723: 
1724: \bibitem[]{}
1725: Schnittman, J.D., Rezzolla, L. 2006, ApJ, 637, L113
1726: %Quasi-periodic Oscillations in the X-Ray Light Curves from Relativistic Tori
1727: 
1728: \bibitem[]{}
1729: Shu, F.H. 1992, The Physics of Astrophysics II: Gas Dynamics (University
1730: Science Books), Chap.~12
1731: 
1732: \bibitem[]{}
1733: Sramkova, E., Torkelsson, U., Abramowicz, M. A., 2007, A\&A, 467, 641
1734: 
1735: \bibitem[]{}
1736: Stella, L., Vietri, M., Morskink, S.M. 1999, ApJ, 524, L63 
1737: 
1738: \bibitem[]{}
1739: Strohmayer, T.E. 2001, ApJ, 552, L49
1740: % Discovery of a 450 HZ Quasi-periodic Oscillation from the Microquasar 
1741: % GRO J1655-40 with the Rossi X-Ray Timing Explorer
1742: 
1743: \bibitem[]{}
1744: Swank, J. 1999, Nucl. Phys. B, Proc. Suppl., 69, 12 (astro-ph/9802188)
1745: 
1746: \bibitem[]{}
1747: Tagger, M. 2006, arXiv:astro-ph/0612175
1748: %Global MHD instabilities: from Low Frequency to High Frequency QPOs, 
1749: %and to Sgr A*
1750: 
1751: \bibitem[]{}
1752: Tagger, M. 2006, in the proceedings of the VI Microquasar Workshop 
1753: ``Microquasars and beyond'', ed. T. Belloni (arXiv:astro-ph/0612175)
1754: 
1755: \bibitem[]{}
1756: Tagger, M., Pellat, R. 1999, A\&A, 349, 1003
1757: 
1758: \bibitem[]{}
1759: Tagger, M., Varniere, P. 2006, ApJ, 652, 1457
1760: 
1761: \bibitem[]{}
1762: Tassev, S.V., Bertschinger, E. 2007, ApJ, submitted (arXiv:0711.0065)
1763: 
1764: \bibitem[]{}
1765: Tsang, D. \& Lai, D., 2008a, MNRAS, 387, 446
1766: 
1767: \bibitem[]{}
1768: Tsang, D. \& Lai, D., 2008b, MNRAS, submitted
1769: % The c-mode paper
1770: 
1771: \bibitem[]{}
1772: Varniere, P., Tagger, M. 2002, A\&A, 394, 329
1773: % Accretion-Ejection Instability in magnetized disks: Feeding the corona 
1774: % with Alfven waves
1775: 
1776: \bibitem[]{}
1777: van der Klis, M. 2006, in Compact Stellar X-ray Sources, ed. 
1778: W.H.G. Lewin and M. van der Klis (Cambridge Univ. Press) 
1779: (astro-ph/0410551)
1780: 
1781: \bibitem[]{}
1782: Wagoner, R. V., 1999, Phys. Rep., 311, 259
1783: 
1784: \bibitem[]{}
1785: Zhang, H., Lai, D., MNRAS, 2006, 368, 917
1786: 
1787: \end{thebibliography}
1788: 
1789: 
1790: \end{document}
1791: