1: % Template article for preprint document class `elsart'
2: % with harvard style bibliographic references
3: % SP 2006/04/26
4:
5: \documentclass{elsart3p}
6:
7: % Use the option doublespacing or reviewcopy to obtain double line spacing
8: % \documentclass[doublespacing]{elsart}
9:
10: % the natbib package allows both number and author-year (Harvard)
11: % style referencing;
12: \usepackage{natbib}
13:
14: % if you use PostScript figures in your article
15: % use the graphics package for simple commands
16: % \usepackage{graphics}
17: % or use the graphicx package for more complicated commands
18: \usepackage{graphicx}
19: % or use the epsfig package if you prefer to use the old commands
20: % \usepackage{epsfig}
21:
22: % The amssymb package provides various useful mathematical symbols
23: \usepackage{amssymb}
24:
25: % The lineno packages adds line numbers. Start line numbering with
26: % \begin{linenumbers}, end it with \end{linenumbers}. Or switch it on
27: % for the whole article with \linenumbers.
28: % \usepackage{lineno}
29:
30: % \linenumbers
31: \begin{document}
32:
33: \def\lta{\lower2pt\hbox{$\buildrel {\scriptstyle <}
34: \over {\scriptstyle\sim}$}}
35: \def\gta{\lower2pt\hbox{$\buildrel {\scriptstyle >}
36: \over {\scriptstyle\sim}$}}
37: \def\msun{M_\odot}
38:
39: \begin{frontmatter}
40:
41: % Title, authors and addresses
42:
43: % use the thanksref command within \title, \author or \address for footnotes;
44: % use the corauthref command within \author for corresponding author footnotes;
45: % use the ead command for the email address,
46: % and the form \ead[url] for the home page:
47: \title{Advection-Dominated Accretion and the Black Hole Event Horizon}
48: \author{Ramesh Narayan},
49: \ead{rnarayan@cfa.harvard.edu}
50: \author{Jeffrey E. McClintock}
51: \ead{jmcclintock@cfa.harvard.edu}
52: \address{Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, U.S.A.}
53:
54: \begin{abstract}
55: As the luminosity of an accreting black hole drops to a few percent of
56: Eddington, the spectrum switches from the familiar soft state to a
57: hard state that is well-described by a distended and tenuous
58: advection-dominated accretion flow (ADAF). An ADAF is a poor
59: radiator, and the ion temperature can approach $10^{12}$~K near the
60: center, although the electrons are cooler, with their temperature
61: typically capped at $\sim 10^{9-11}$~K. The foundational papers
62: predicted that the large thermal energy in an ADAF would drive strong
63: winds and jets, as later observed and also confirmed in computer
64: simulations. Of chief interest, however, is the accreting gas that
65: races inward. It carries the bulk of the accretion energy as stored
66: thermal energy, which vanishes without a trace as the gas passes
67: through the hole's event horizon. One thus expects black holes in the
68: ADAF regime to be unusually faint. Indeed, this is confirmed by a
69: comparison of accreting stellar-mass black holes and neutron stars,
70: which reside in very similar transient X-ray binary systems. The
71: black holes are on average observed to be fainter by a factor of $\sim
72: 100-1000$. The natural explanation is that a neutron star must
73: radiate the advected thermal energy from its surface, whereas a black
74: hole can hide the energy behind its event horizon. The case for an
75: event horizon in Sagittarius A$^*$, which is immune to caveats on jet
76: outflows and is furthermore independent of the ADAF model, is
77: especially compelling. These two lines of evidence for event horizons
78: are impervious to counterarguments that invoke strong gravity or
79: exotic stars.
80: \end{abstract}
81:
82: \begin{keyword}
83: % keywords here, in the form: keyword \sep keyword
84:
85: % PACS codes here, in the form: \PACS code \sep code
86:
87: \end{keyword}
88:
89: \end{frontmatter}
90:
91: % main text
92: \section{Introduction}
93: \label{Intro}
94:
95: The foundations of our present understanding of {\it
96: advection-dominated accretion} were laid out in a series of papers by
97: Narayan \& Yi (1994, 1995a,b, hereafter NY94, NY95a, NY95b),
98: Abramowicz et al. (1995) and Chen et al. (1995), although some ideas
99: were anticipated much earlier by Ichimaru (1977). The specific
100: abbreviation ADAF, which stands for {\it advection-dominated accretion
101: flow}, was introduced by Lasota\footnote{It gives the authors great
102: pleasure to highlight Jean-Pierre Lasota's research on ADAFs in this
103: contribution to his Festschrift.}(1996) and has become standard in the
104: field. We review here the application of ADAFs to accreting black
105: holes (BHs), and in \S4 we also consider their application to
106: accreting neutron stars.
107:
108: \section{Accretion Regimes}
109:
110: The energy equation per unit volume of an accretion flow can be
111: written compactly in the form
112: \begin{equation}
113: \rho T{dS\over dt} \equiv
114: \rho T \left({\partial S\over \partial t} +
115: \vec{v}\cdot\vec{\nabla}S\right) = q_+ - q_-,
116: \end{equation}
117: where $\rho$ is the density, $T$ is the temperature, $S$ is the
118: entropy per unit mass, $t$ is time, $\vec{v}$ is the flow velocity,
119: and $q_+$ and $q_-$ are the heating and cooling rates per unit volume.
120: This equation states that the rate at which the entropy per unit
121: volume of the gas increases is equal to the heating rate minus the
122: cooling rate. Since any entropy stored in the gas is advected with
123: the flow, the left-hand side of equation (1) may be viewed as the
124: effective {\it advective cooling} rate $q_{\rm adv}$. Equation (1)
125: can then be rearranged to give
126: \begin{equation}
127: q_+ = q_- + q_{\rm adv},
128: \end{equation}
129: which states that the heat energy released by viscous dissipation is
130: partially lost by radiative cooling and partially by advective
131: cooling.
132:
133: The standard thin accretion disk model (Shakura \& Sunyaev 1973;
134: Novikov \& Thorne 1973; Frank, King \& Raine 2002) corresponds to the
135: case when the accreting gas is {\it radiatively efficient}, so that we have
136: \begin{equation}
137: {\rm Thin~Disk}: \quad q_- \gg q_{\rm adv}, \quad L \sim 0.1\dot{M}c^2.
138: \end{equation}
139: Since the gas cools efficiently, the sound speed is much less than the
140: local Keplerian speed $v_K$ and the disk is geometrically thin. Also,
141: the disk radiates about a tenth of the rest mass energy of the
142: accreting gas, the precise fraction depending on the BH spin (see
143: Shapiro \& Teukolsky 1983).
144:
145: An ADAF corresponds to the opposite regime. Here the gas is {\it
146: radiatively inefficient} and the accretion flow is underluminous.
147: Thus, an ADAF is defined by the condition
148: \begin{equation}
149: {\rm ADAF}: \quad q_{\rm adv} \gg q_-, \quad L \ll 0.1\dot{M}c^2.
150: \label{ADAFdef}
151: \end{equation}
152: Some papers in the literature define an ADAF as a flow that
153: corresponds exactly to a self-similar solution described in NY94 (see
154: \S3.1). This is a needless restriction. In our opinion, it is more
155: fruitful to employ the general definition of an ADAF as given in
156: equation (\ref{ADAFdef})\footnote{Actually, at mass accretion rates
157: for which the flow is close to the boundary between the thin disk and
158: ADAF solutions, we have (by continuity) $q_{\rm adv} \ \gta\ q_-$, and
159: the flow is only marginally radiatively inefficient (see Fig. 4).
160: Nevertheless, even here, the two solutions are very distinct from each
161: other.}.
162:
163: There are two distinct regimes of advection-dominated accretion. The
164: first, the one that we focus on in this article is when the accreting
165: gas is very tenuous and has a long cooling time (NY94; NY95b;
166: Abramowicz et al. 1995). This regime is sometimes referred to as a
167: RIAF --- a ``radiatively inefficient accretion flow'' --- and is
168: defined by the condition
169: \begin{equation}
170: {\rm ADAF/RIAF}: \quad t_{\rm cool} \gg t_{\rm acc}.
171: \end{equation}
172: Here $t_{\rm cool}$ is the cooling time of the gas and $t_{\rm acc}$
173: is the accretion time.
174:
175: The second regime of advection-dominated accretion is when the
176: particles in the gas have no trouble cooling, but the scattering
177: optical depth of the accretion flow is so large that the radiation is
178: unable to diffuse out of the system. This radiation-trapped regime
179: was briefly discussed by Begelman (1979) and was developed in detail
180: by Abramowicz et al. (1988) in their ``slim disk'' model. The
181: defining condition for this regime of ADAFs is
182: \begin{equation}
183: {\rm ADAF/Slim~Disk}: \quad t_{\rm diff} \gg t_{\rm acc},
184: \end{equation}
185: where $t_{\rm diff}$ is the diffusion time for photons.
186:
187: The present review is devoted exclusively to the ADAF/RIAF form of
188: accretion (see Narayan, Mahadevan \& Quataert 1998b; Kato, Fukue \&
189: Mineshige 1998; Lasota 1999a,b; Quataert 2001; Narayan 2002, 2005;
190: Igumenshchev 2004; Done, Gierlinsky \& Kubota 2007 for reviews
191: emphasizing various aspects of ADAFs). We will henceforth drop the
192: modifier RIAF and refer to these flows simply as ADAFs.
193:
194: As explained above, an ADAF is by definition very different from the
195: standard thin accretion disk. Correspondingly, it is characterized by
196: very distinct observational signatures. Both ADAFs and thin disks
197: have well-known counterparts in nature.
198:
199: \section{Properties of ADAFs}
200:
201: \subsection{Geometry and Kinematics}
202:
203: In an ADAF, since most of the energy released by viscous dissipation
204: is retained in the gas, the pressure is large, and so is the sound
205: speed: $c_s \sim v_K$, where the Keplerian velocity $v_K(R)$ is equal
206: to $c/\sqrt{2r}$ with $r=R/R_S$ being the radius in Schwarzschild
207: units, $R_S=2GM/c^2$. The large pressure has several immediate
208: consequences. First, the accretion flow becomes geometrically thick,
209: with a vertical height $H$ of order the radius $R$ (an ADAF may be
210: viewed as the viscous rotating analog of spherical Bondi accretion).
211: Second, the flow has considerable pressure-support in the radial
212: direction, so the angular velocity becomes sub-Keplerian. Third, the
213: radial velocity of the gas is relatively large: $v \sim \alpha
214: v_K(H/R)^2 \sim \alpha c/r^{1/2}$, where $\alpha\sim0.1-0.3$ (see
215: below) is the standard dimensionless viscosity parameter (Shakura \&
216: Sunyaev 1973). Fourth, the large radial velocity leads to a short
217: accretion time: $t_{\rm acc} =R/v \sim t_{\rm ff}/\alpha$, where
218: $t_{\rm ff}=(2GM/R^3)^{1/2}$ is the free-fall time. Finally, the
219: large velocity and large scale height cause the gas density to be very
220: low, and so the cooling time is very long and the medium is optically
221: thin.
222:
223: The above properties are nicely illustrated in the self-similar ADAF
224: solution derived by NY94, in which all quantities in the accretion
225: flow behave as power-laws in radius. (A similar solution was obtained
226: by Spruit et al. 1987 in a different context.) NY94 obtained a
227: general solution for arbitrary viscosity parameter $\alpha$, adiabatic
228: index $\gamma$, and advection parameter\footnote{By this definition,
229: $f= 0$ corresponds to a fully cooling-dominated (no advection) flow
230: and $f=1 $ corresponds to a fully advection-dominated (no radiative
231: cooling) flow.} $f \equiv q_{\rm adv}/q_+$. In the limit
232: $\alpha^2\ll1$ (a good approximation) and $f\to1$ (radiatively very
233: inefficient flow), the solution simplifies to
234: \begin{eqnarray}
235: v &=& -\alpha \left[{\gamma-1\over\gamma-5/9}\right]v_K =
236: -0.53 \alpha v_K, \\
237: \Omega &=& \left[{2(5/3-\gamma)\over3(\gamma-5/9)}\right]^{1/2}\Omega_K =
238: 0.34 \Omega_K, \\
239: c_s &=& \left[{2(\gamma-1)\over3(\gamma-5/9)}\right]^{1/2}v_K =
240: 0.59 v_K.
241: \end{eqnarray}
242: The numerical coefficients on the right correspond to $\gamma=1.5$, a
243: reasonable choice (see Quataert \& Narayan 1999).
244:
245: The great virtue of the above self-similar solution is that it is
246: analytic and provides an easy and transparent way of understanding all
247: the key properties of an ADAF. Its biggest deficiency is that it is
248: scale-free, which means that it is inappropriate near the inner or
249: outer boundary of the flow. Therefore, for detailed work, one must
250: use global solutions of the ADAF equations that satisfy appropriate
251: boundary conditions (e.g., Abramowicz et al. 1996; Narayan, Kato \&
252: Honma 1997c; Chen, Abramowicz \& Lasota 1997; Manmoto, Mineshige \&
253: Kusunose 1997; Popham \& Gammie 1998; Manmoto 2000). Calculating
254: global solutions is somewhat involved and one may wish to employ some
255: short-cuts (e.g., Yuan, Ma \& Narayan 2008). On the other hand, even
256: the exact global solutions are somewhat limited since they solve a set
257: of height-integrated equations. For greater realism, one might wish
258: to work directly with numerical simulations (e.g., Goldston, Quataert
259: \& Igumenshchev 2005; Noble et al. 2007).
260:
261: \subsection{Thermal Properties}
262:
263: The ADAF solution is gas pressure dominated. Since $c_s\sim v_K$,
264: this means the gas temperature is nearly virial. Under normal
265: conditions, gas at such a high temperature will radiate copiously,
266: especially at small radii where the temperature can approach $10^{12}$
267: K. Thus, in order to have an ADAF, the accreting gas generally has to
268: be a {\it two-temperature plasma} (at least at small radii), with
269: electron temperature $T_e$ much less than the ion temperature $T_i$
270: (NY95b). (The only way to avoid this condition is by having an
271: extremely low accretion rate below about $10^{-6}$ of the Eddington
272: rate.) Typical ADAF models have the two temperatures scaling as
273: \begin{equation}
274: T_i \sim 10^{12}{\rm K}/r, \quad T_e \sim
275: {\rm Min}\, (T_i,\, 10^{9-11} {\rm K}).
276: \end{equation}
277:
278: In order for gas in an ADAF to be two-temperature, there must be weak
279: coupling between electrons and ions. Models generally assume that the
280: coupling occurs via Coulomb collisions, which is inefficient at the
281: densities under consideration. Begelman \& Chiueh (1988) investigated
282: whether plasma instabilities might enhance the coupling and drive the
283: plasma rapidly to a single temperature; this would be problematic for
284: the ADAF solution. However, they were unable to identify a clear
285: mechanism. For normal mass accretion rates, the electrons in a
286: two-temperature ADAF will have a thermal energy distribution (but not
287: necessarily the ions, see Mahadevan \& Quataert 1997).
288:
289: Early work on two-temperature ADAFs assumed that viscous heating acts
290: primarily on the ions; for instance, NY95b took the ratio of electron
291: heating to total heating, $\delta$, to be zero, while Esin, McClintock
292: \& Narayan (1997) assumed $\delta \sim m_e/m_p \sim10^{-3}$. However,
293: an ADAF does not {\it require} $\delta$ to be this small. Because of
294: a degeneracy in model parameters (Quataert \& Narayan 1999), it is
295: possible to have a viable ADAF model with larger values of $\delta$,
296: provided the mass loss parameter $s$, defined in \S3.6, is adjusted.
297: More recent ADAF models (e.g., Yuan, Quataert \& Narayan 2003)
298: typically assume $\delta\sim0.3-0.5$.
299:
300: Various attempts have been made to estimate the value of $\delta$ from
301: first principles by considering the effects of reconnection
302: (Bisnovatyi-Kogan \& Lovelace 1997; Quataert \& Gruzinov 1999) or MHD
303: turbulence (Quataert 1998; Blackman 1999; Medvedev 2000). These
304: studies do not agree on a single value of $\delta$, but generally
305: suggest that $\delta$ is likely to be much larger than $10^{-3}$.
306: Recently, Sharma et al. (2007) considered heating by the dissipation
307: of pressure anisotropy and showed that $\delta\sim (T_e/T_i)^{1/2}/3$.
308:
309: It should be noted that, even if electrons and ions receive equal
310: amounts of the dissipated energy (i.e., $\delta=0.5$), the plasma can
311: still be two-temperature. This is because a large part of the heating
312: in an ADAF is by compression (since the density increases inward).
313: Once $kT \ \gta\ m_e c^2$, which is the case for $r\ \lta$ a few
314: hundred, the electrons become relativistic and have an adiabatic index
315: $\gamma_e\sim4/3$, whereas the ions continue to be non-relativistic
316: with $\gamma_i\sim5/3$. Since adiabatic heating by compression causes
317: the temperature to scale as $T \propto \rho^{\gamma-1}$, the electrons
318: heat up only as $T_e \sim \rho^{1/3}$ whereas the ions heat up as $T_i
319: \sim \rho^{2/3}$. Thus, even in the limiting case of $\delta\sim0.5$,
320: ADAFs naturally become two-temperature at small radii. For instance,
321: in the ADAF model of Sagittarius A$^*$ (Sgr A$^*$) proposed by Yuan et
322: al. (2003), the authors obtain $T_e \sim 0.1T_i$ close to the BH even
323: though they assumed $\delta=0.55$.
324:
325: An important property of the ADAF solution is that it is thermally
326: stable (NY95b; Wu \& Li 1996; Kato et al. 1997; Wu 1997). The
327: demonstration of this property was a crucial advance. Two decades
328: earlier, in a seminal paper, Shapiro, Lightman \& Eardley (1976), and
329: after them Rees et al. (1982; 'ion tori'), introduced the idea of a
330: two-temperature accretion flow and derived a hot two-temperature
331: accretion flow solution, the SLE solution. However, that solution
332: turned out to be thermally unstable (Piran 1978). Until the
333: development of the ADAF solution, and the recognition that it is
334: different from the SLE solution, no stable, hot, optically thin
335: solution was available to model the many accretion systems whose
336: spectra (especially in the hard state, see below) demand such a flow.
337:
338: Chen et al. (1995) and Yuan (2003) have explored the relationships
339: among the ADAF/RIAF, SLE, ADAF/Slim Disk and Thin Disk solutions.
340:
341: \subsection{Models and Spectra}
342:
343: The ADAF is a full accretion solution which incorporates consistent
344: dynamics, thermal balance, radiation physics, etc. Therefore, the
345: radial profiles of all gas properties can be calculated
346: self-consistently once we know the values of certain parameters: BH
347: mass $M$, accretion rate $\dot{M}$, viscosity parameter $\alpha$,
348: pressure parameter $\beta\equiv P_{\rm gas}/(P_{\rm gas}+P_{\rm
349: mag})$, adiabatic index $\gamma$ (usually $\sim1.5$), viscous heating
350: parameter $\delta$, advection parameter $f$.
351:
352: Actually, apart from the system-specific parameters $M$ and $\dot{M}$
353: (which may be estimated through observation), most of the other
354: parameters are constrained. Under the nearly collisionless conditions
355: expected in an ADAF, the viscosity parameter is moderately enhanced
356: relative to a collisional gas (Sharma et al. 2006). In the case of
357: dwarf novae in the hot state, Smak (1999) estimates $\alpha\sim0.2$,
358: while numerical simulations of the magneto-rotational instability give
359: $\alpha\ \lta\ 0.1$ (Hawley, Gammie \& Balbus 1996). Thus we expect
360: $\alpha\sim0.1-0.3$ for an ADAF (typical values used in models are
361: $0.2-0.3$). Numerical simulations further suggest that magnetic
362: fields are generally subthermal, with $P_{\rm mag}/P_{\rm gas}\sim
363: 0.1$ (e.g., Hawley et al. 1996), so we expect $\beta\sim0.9$. By
364: calculating the energy loss via radiation from the hot accretion flow
365: (synchrotron, bremsstrahlung, Compton scattering), $f$ can be obtained
366: self-consistently (e.g., NY95b; Narayan, Barret \& McClintock 1997a;
367: Esin et al. 1997; Yuan et al. 2003). Thus, we have only one poorly
368: constrained parameter: $\delta$. The recent work of Sharma et
369: al. (2007) provides a serviceable prescription even for this
370: parameter. At large radii, where the plasma is effectively
371: one-temperature, their formula gives $\delta\sim0.3$, while in the
372: energetically important inner region, where the plasma is
373: two-temperature, they find $\delta\sim 0.01-0.1$, depending on model
374: details. The final two parameters are discussed later: the transition
375: radius $R_{\rm tr}$, \S3.5, and the wind parameter $s$, \S3.6.
376:
377: \begin{figure}
378: \includegraphics[width=3.3in,clip]{J0422}
379:
380: \caption{Combined TTM (2--20 keV), HEXE (20--200 keV), and OSSE
381: (50--600 keV) spectrum of GRO J0422+32 observed between 1992 August 29
382: and September 2. The solid line is an ADAF model which reproduces the
383: overall luminosity of the source, the power-law slope below the peak
384: and the shape of the high-energy cutoff. The model has a mass
385: accretion rate $\sim 0.1\dot{M}_{\rm Edd}$. (From Esin et al. 1998)}
386: \end{figure}
387:
388: Since the earliest days of X-ray astronomy, it has been clear that BH
389: binaries (BHBs) have a number of distinct spectral states (see
390: Zdziarski \& Gierlinski 2004; McClintock \& Remillard 2006; Done et
391: al. 2007).
392: %A dramatic example is the black hole binary (BHB) source
393: %GX 339-4, which exhibits several states,
394: The most notable among these are the luminous {\it high soft state},
395: or {\it thermal state}, the slightly less luminous {\it low hard
396: state}, and the very under-luminous {\it quiescent state}.
397:
398: The thermal state is well described by the thin disk model, and a
399: multi-color disk (MCD) blackbody model (e.g., {\it diskbb}, Mitsuda et
400: al. 1984; {\it ezdiskbb}, Zimmerman et al. 2005; both available in
401: XSPEC, Arnaud et al. 1996) has been successfully used for years to
402: model the X-ray spectra of sources in this state. Recently, fully
403: relativistic versions of the MCD model for arbitrary BH spin ({\it
404: kerrbb}, Li et al. 2005; {\it bhspec}, Davis \& Hubeny 2006) have been
405: developed, based on the relativistic thin disk model of Novikov \&
406: Thorne (1973). These models provide excellent fits to the X-ray
407: spectra of BHBs in the thermal state (e.g., McClintock et al. 2006;
408: Davis, Done \& Blaes 2006).
409:
410: Whereas a satisfactory theoretical model, viz., the thin disk model,
411: was established early on for the thermal state, the hard state was for
412: many years a mystery. Figure 1 shows the spectrum of a typical BHB,
413: GRO J0422+32, in the hard state (Esin et al. 1998). The observations
414: indicate that the accreting gas is very hot, $kT\ \gta\ 100$ keV. The
415: gas must also be optically thin, since optically thick blackbody
416: emission at a temperature of 100 keV would correspond to a luminosity
417: $L=\sigma T^4 A\ \gta\ 10^{46} ~{\rm erg\,s^{-1}}$ for any reasonable
418: estimate of the radiating area $A$.
419:
420: The most natural explanation of the emission in J0422 and other BHBs
421: in the hard state is that it is produced by thermal Comptonization.
422: Until the ADAF model was established, no accretion model could
423: reproduce such a spectrum. (The SLE solution could, but it was
424: unstable.) Indeed, astronomers were reduced to using empirical
425: Comptonization models (Sunyaev \& Titarchuk 1980) in which they
426: postulated a Comptonizing cloud with some arbitrary geometry and
427: parameterized the cloud with an adjustable temperature and an optical
428: depth (e.g., Zdziarski et al. 1996, 1998; Gierlinski et al. 1997).
429:
430: The situation changed with the recognition of the ADAF solution. This
431: model turned out to have the precise properties --- density, electron
432: temperature, stability --- needed to explain the hard state. The
433: solid line in Fig. 1 shows an ADAF model of J0422 (Esin et al. 1998)
434: in which the accretion rate has been adjusted to fit the spectrum; the
435: required rate is about a tenth of the Eddington mass accretion rate
436: $\dot{M}_{\rm Edd}$, where $\dot{M}_{\rm Edd}=L_{\rm Edd}/(0.1c^2)$,
437: i.e., it is the mass accretion rate at which a disk with radiative
438: efficiency 0.1 would radiate at the Eddington luminosity. It is
439: gratifying that both the temperature (which determines the position of
440: the peak) and the Compton $y$-parameter (which determines the
441: power-law slope below the peak) are reproduced well. Esin et
442: al. (1997, 1998, 2001) present models of other BHBs in the hard state.
443:
444: \begin{figure}
445: \vskip-0.2in
446: \includegraphics[width=3.7in,clip]{V404}
447: \vskip -1.5in
448: \caption{The spectrum of the BHB source V404 Cyg in quiescence. The
449: three dots on the left represent optical fluxes (Narayan et al. 1996),
450: the error box in the X-ray band corresponds to $2\sigma$ limits from
451: ASCA (Narayan et al. 1997a), and the upper limit in the EUV is derived
452: from the absence of a HeII $\lambda4686$ line. The solid line
453: corresponds to an ADAF model with $\dot{M}\sim 0.005\dot{M}_{\rm
454: Edd}$. The dashed line is a thin disk model whose $\dot{M}$ has been
455: adjusted to fit the optical data. This model fits poorly in the X-ray
456: band and is inconsistent with the EUV limit. (From Narayan et
457: al. 1997a)}
458: \end{figure}
459:
460: \begin{figure}
461: \includegraphics[width=3in,clip]{Sgr}
462:
463: \caption{The quiescent spectrum of Sgr A$^*$. The radio data are from
464: Falcke et al. (1998; open circles) and Zhao et al. (2003; filled
465: circles), the IR data are from Serabyn et al. (1997) and Hornstein et
466: al. (2002), and the two ``bow-ties'' in the X-ray band correspond to
467: the quiescent (lower) and flaring (higher) data from Baganoff et
468: al. (2001, 2003). The solid line is an ADAF model of Sgr A$^*$ in the
469: quiescent state. The mass accretion rate is
470: $\dot{M}\sim10^{-6}\dot{M}_{\rm Edd}$ near the BH. (From Yuan et al. 2003)}
471: \end{figure}
472:
473: A typical accreting BH observed in the hard state has a luminosity on
474: the order of a few percent of Eddington. At much lower luminosities,
475: we have the quiescent state, where the spectrum becomes noticeably
476: different. Figures 2 and 3 show observations of a quiescent BHB, V404
477: Cyg, and a quiescent supermassive BH, the Galactic Center source Sgr
478: A$^*$. Although these spectra look very different from the one shown
479: in Fig. 1, the ADAF model is able to fit these and other observations
480: of quiescent systems (Narayan, McClintock \& Yi 1996; Narayan et
481: al. 1997a; Yuan et al. 2003). All it requires is a lower value of
482: $\dot{M}$, as appropriate for the lower luminosity. The qualitative
483: features of the spectrum, e.g., a softening of the X-ray power-law
484: index (see Corbel, Tomsick \& Kaaret 2006), follow naturally.
485: However, a caveat is in order: In some cases, the X-ray emission in
486: quiescence may be from a jet lauched from the ADAF (\S3.6), rather
487: than from the ADAF itself (e.g., Yuan \& Cui 2005).
488:
489: \subsection{Radiative Efficiency}
490:
491: \begin{figure}
492: \includegraphics[width=3.3in,clip]{radeff2}
493:
494: \caption{Radiative efficiency $\eta$ of an accretion flow around a BH,
495: as a function of the Eddington-scaled mass accretion rate (Hopkins,
496: Narayan \& Hernquist 2006b). The horizontal segment between 0.01 and
497: 0.1 of the Eddington accretion rate shows a transition regime in which
498: a part of the accretion flow is an ADAF, but the radiative efficiency
499: is still large. It might correspond to the Intermediate State, and
500: perhaps the upper end of the Hard State (\S3.5, Fig. 5). Although
501: this plot is based on calculations shown in Fig. 11 of NY95b and
502: Fig. 13 of Esin et al. (1997), it is still very approximate.}
503: \end{figure}
504:
505: As expressed in equation (\ref{ADAFdef}), the defining property of an
506: ADAF is that it is radiatively inefficient,
507: \begin{equation}
508: {\rm ADAF}: \qquad \eta\equiv L/\dot{M}c^2 \ll 0.1 .
509: \label{eta}
510: \end{equation}
511: Calculations show that the ADAF solution is possible only for low mass
512: accretion rates. Specifically, only when $L\ \lta\ 0.1L_{\rm Edd}$ is
513: the gas density low enough to permit a two-temperature plasma (NY95b;
514: Narayan 1996; Esin et al. 1997). Near the critical luminosity $L_{\rm
515: crit}$ or critical mass accretion rate $\dot{M}_{\rm crit}$ at which
516: the ADAF solution first becomes viable, the radiative efficiency is,
517: by continuity, not very different from that of a thin disk:
518: $\eta\sim0.1$. However, with decreasing $\dot{M}$, the efficiency
519: decreases. Very roughly, we estimate (Fig. 4)
520: \begin{equation}
521: \eta \sim 0.1\left(\dot{M}/0.01\dot{M}_{\rm Edd}\right), \quad
522: \dot{M} < 0.01\dot{M}_{\rm Edd}.
523: \end{equation}
524: %Here $\dot{M}_{\rm Edd}=L_{\rm Edd}/ (0.1c^2)$, i.e., it is the mass
525: %accretion rate at which a disk with radiative efficiency 0.1 would
526: %radiate at the Eddington luminosity.
527: (We note that for the different
528: prescription for $\delta$ used by Sharma et al. 2006, $\eta$ falls
529: rapidly only for $\dot{M} \ \lta\ 10^{-4} \dot{M}_{\rm Edd}$.)
530:
531: The extreme inefficiency of an ADAF at very low accretion rates is
532: critical for understanding the peculiar properties of accreting BHs in
533: the quiescent state (as first discussed by Narayan et al. 1996 for
534: BHBs and Narayan et al. 1995 for supermassive BHs). We do not discuss
535: this topic here, but point the reader to other reviews for details
536: (e.g., Quataert 2001; Narayan 2002, 2005). The quiescent state has
537: also played a major role in our efforts to test for the presence of an
538: event horizon in BHBs. This is a key topic of \S4.
539:
540: \subsection{Spectral Regimes}
541:
542: \begin{figure}
543: \includegraphics[width=3.7in,clip]{states}
544:
545: \caption{Configuration of the accretion flow in different spectral
546: states, shown schematically as a function of the mass accretion rate
547: $\dot{M}$ (based on Esin et al. 1997). The ADAF is represented by the
548: hatched ellipses, with the intensity of hatching indicating the
549: density of the hot gas. The horizontal lines represent a standard
550: thin disk. The lowest panel shows the quiescent state (QS), which
551: corresponds to a very low mass accretion rate (say $<10^{-3}
552: \dot{M}_{\rm Edd}$), a weak jet (\S3.6), a low radiative efficiency
553: (Fig. 4), and a large transition radius (Fig. 6). The second panel
554: from the bottom shows the hard state (HS), where the mass accretion
555: rate is higher ($\sim 10^{-3}-10^{-1.5}\dot{M}_{\rm Edd}$) but still
556: below the critical rate $\dot{M}_{\rm crit}$, the jet is stronger, the
557: radiative efficiency is somewhat larger, and the transition radius is
558: smaller. The second panel from the top shows the intermediate state
559: (IS), where $\dot{M}\sim 10^{-1.5}-10^{-1}\dot{M}_{\rm Edd}\ \lta\
560: \dot{M}_{\rm crit}$, the jet is even stronger, the radiative
561: efficiency is high $\sim0.1$, and the transition radius is fairly
562: close to the ISCO. Finally, the top panel shows the thermal state
563: (TS), where there is no ADAF, the thin disk extends down to the ISCO,
564: $\eta\sim0.1$, and there is no jet. Esin et al. (1997) included a
565: tentative proposal for the so-called very high state (see also Done et
566: al. 2007), now called the steep power-law state (SPL, McClintock \&
567: Remillard 2006), but we do not include this still-mysterious state
568: here.}
569: \end{figure}
570:
571: Based on the properties of the ADAF solution discussed above, Narayan
572: (1996) proposed a simple model for understanding the spectral states
573: of accreting BHs. This picture was developed in detail by Esin et
574: al. (1997). The basic idea is illustrated in Fig. 5.
575:
576: The key parameter that determines the spectral state of an accreting
577: BH is $\dot{M}$. When $\dot{M}>\dot{M}_{\rm crit}$, only the thin
578: disk solution is available, and so the accretion flow is in the form
579: of a thin disk all the way down to the innermost stable circular orbit
580: (ISCO). The system is then in the thermal state, and its spectrum is
581: well-described by the MCD model.
582:
583: Once $\dot{M}$ falls below $\dot{M}_{\rm crit}$, both the thin disk
584: and ADAF solutions become viable (at least at small radii). Now
585: accretion continues as a thin disk at radii larger than a transition
586: radius, $R > R_{\rm tr}$, but the flow switches to an ADAF at smaller
587: radii. When the transition from a pure thin disk (thermal state) to a
588: disk-plus-ADAF configuration first occurs, i.e., when $\dot{M}$ is
589: just below $\dot{M}_{\rm crit}$, the ADAF is very small in size and we
590: have a more-or-less radiatively efficient flow, with $\eta\sim0.1$.
591: This corresponds to the so-called {\it intermediate state}. Then, at
592: a somewhat lower $\dot{M}$, the ADAF expands a bit and $\eta$ is
593: modestly lower and we have the classic hard state. Finally, when
594: $\dot{M}$ is much lower than $\dot{M}_{\rm crit}$, the ADAF becomes
595: much larger, $R_{\rm tr} \sim10^3-10^4R_S$, with $\eta\ll0.1$ (Narayan
596: et al. 1996, 1997a; Menou, Narayan \& Lasota 1999b). This is the
597: quiescent state.
598:
599: The above paradigm has proved durable (see Zdziarski \& Gierlinski
600: 2004; Done et al. 2007). In particular, considerable evidence has
601: accumulated that the thin disk retreats from the innermost stable
602: circular orbit (ISCO) to a large radius in the quiescent state. The
603: evidence is strongest in transient BHBs and CVs. The spectra of
604: quiescent BHBs show absolutely no sign of any soft blackbody-like
605: X-ray emission from a thin disk at small radii (Narayan et al. 1996,
606: 1997a; McClintock, Narayan \& Rybicki 2004). Timing properties also
607: indicate a large a large size for the ADAF (Hynes et al. 2003; Shahbaz
608: et al. 2005). In additional, theoretical arguments indicate that the
609: thermal-viscous disk instability, which causes the transient behavior
610: in these systems, is incompatible with observations unless the disk is
611: severely truncated in the quiescent state (Lasota, Narayan \& Yi
612: 1996b; Hameury, Lasota \& Dubus 1999; Lasota 2001, 2008; Dubus,
613: Hameury \& Lasota 2001; Yungelson et al. 2006). There are also
614: indications from the time delay between the optical and X-ray
615: outbursts in the BHB GRO J1655--40 (Orosz et al. 1997; Hameury et
616: al. 1997) that the cool disk is truncated at a large radius in the
617: quiescent state.
618:
619: In the case of supermassive BHs in the quiescent state, there is no
620: feature in the spectrum that might be associated with a thin disk,
621: suggesting that there is no disk at all; examples are Sgr A$^*$
622: (Narayan et al. 1998a; Narayan 2002) and M87 (Di Matteo et al. 2000,
623: 2003). Intermediate luminosity AGN do exhibit optical emission from a
624: disk, but the ``big blue bump'' is much less pronounced than in
625: high-luminosity AGN (Ho 1999); this suggests that the disk is
626: truncated at a radius $\sim10-100R_S$ and the interior is filled with
627: an ADAF (Gammie, Narayan \& Blandford 1999; Quataert et al. 1999).
628: Incidentally, the thermal-viscous disk instability does not appear to
629: operate in AGN disks (Menou \& Quataert 2001; Hameury, Lasota \&
630: Viallet 2007).
631:
632: In the more luminous hard state, again, there is considerable spectral
633: evidence that the disk is truncated at a transition radius outside the
634: ISCO, and that the inside is filled with an ADAF-like hot flow. The
635: most spectacular example is the BHB XTE J1118+480, for which
636: observations carried out in the hard state had unprecedented spectral
637: coverage. The observations are fit well with an ADAF model, with a
638: transition radius at $\sim 50 R_S$ (Esin et al. 2001). The model even
639: fits the complicated timing behavior of the source (Yuan, Cui \&
640: Narayan 2005). It is hard to imagine that the same data could be
641: explained with any model in which a cool disk extends down to the
642: ISCO. Done et al. (2007) review spectral observations of a number of
643: other BHBs where again the data require a truncated disk. Nemmen et
644: al. (2006) show that both the spectrum and the double-peaked Balmer
645: line profile of the LINER source NGC 1097 are consistent with a disk
646: truncated at a few hundred $R_S$.
647:
648: In an interesting study of Cyg X--1, Gilfanov, Churazov \& Revnivtsev
649: (1999; see Cui et al. 1999; Done et al. 2007; for discussions of other
650: sources) found that, as the characteristic frequency in the
651: variability spectrum of the source increases, the power-law tail in
652: the spectral energy distribution steepens and the amplitude of the
653: reflection component in the spectrum increases. This is exactly what
654: one expects when the transition radius between the outer cool disk and
655: the inner hot ADAF varies. With decreasing transition radius, (i) the
656: noise frequency (which is likely related in some fashion to the
657: Keplerian frequency at the transition radius) should increase, (ii)
658: the hot medium should be cooled more effectively by soft photons from
659: the disk, giving a steeper power-law tail, and (iii) the solid angle
660: subtended by the cool disk at the ADAF should increase, and there
661: should be a larger reflection component. Zdziarski, Lubinski \& Smith
662: (1999) have noted that a correlation between spectral slope and
663: reflection is commonly seen in both BHBs and AGN.
664:
665: A few BHBs in the hard state have been found to show a soft
666: blackbody-like component in their spectra (Balucinska-Church et
667: al. 1995; Di Salvo et al. 2001; Miller et al. 2006a,b; Ramadevi \&
668: Seetha 2007; Rykoff et al. 2007). This could be interpreted as
669: evidence that the thin disk extends all the way down to the
670: ISCO\footnote{However, Di Salvo et al. (2001) estimate from their
671: observations of Cyg X--1 that the disk inner edge is located at tens
672: of $R_S$, not at the ISCO}. Some authors have noted that it is
673: possible for a thin disk to evaporate to an ADAF at a relatively large
674: radius and for the hot gas to then re-condense into a thin disk at
675: small radii (Rozanska \& Czerny 2000; Liu et al. 2007; Mayer \&
676: Pringle 2007)\footnote{Models of this kind may give hard spectra
677: without a significant decrease in radiative efficiency. This would be
678: consistent with the transition region between mass accretion rates of
679: 0.01 and 0.1 of Eddington shown in Fig. 4, where we have suggested an
680: ADAF with high radiative efficiency may be present.}. Such models
681: might explain the occurrence of a soft spectral component in some hard
682: state sources. However, we note that the soft component typically has
683: only 10\% of the total observed luminosity. It is hard to understand
684: how a radiatively efficient thin disk located at the ISCO could be
685: such a minor contributor to the emitted radiation.
686:
687: This difficulty is highlighted in the work of D'Angelo et al. (2008).
688: They point out that, in any model of the hard state, there will be
689: considerable interaction between the hot gas which produces the hard
690: X-rays and cool gas that may be present in a conventional disk. The
691: interaction will occur via X-ray irradiation (unless one has large
692: outward beaming, which seems unlikely, e.g., Narayan \& McClintock
693: 2005) as well as particle bombardment. Using a prototype model for
694: ion bombardment (Deufel et al. 2002; Spruit \& Deufel 2002; Dullemond
695: \& Spruit 2005), D'Angelo et al. (2008) show that a weak soft
696: component in the X-ray spectrum arises quite naturally when the thin
697: disk is truncated at $\sim15-20 R_S$\footnote{This is a reasonable
698: location for the transition radius in the hard state (Esin et
699: al. 1998; Done et al. 2007)}. The model fits the observations
700: surprisingly well; in fact, the same model would predict a much
701: stronger soft component, and would strongly disagree with the
702: observations, if the cool disk were to extend down to the ISCO.
703:
704: \begin{figure}
705: \includegraphics[width=3.1in,clip]{rtrans}
706:
707: \caption{Estimated location of the dimensionless transition radius
708: $R_{\rm tr}/R_S$, plotted as a function of the Eddington-scaled
709: accretion luminosity $L/L_{\rm Edd}$, for a sample of BHBs and
710: low-luminosity AGN (Yuan \& Narayan 2004). The individual estimates
711: of $R_{\rm tr}$ are obtained by fitting spectral observations and are
712: very uncertain. Nevertheless, there seems to be a trend of increasing
713: transition radius with decreasing luminosity.}
714: \end{figure}
715:
716: One of the most vexing problems in the theory of ADAFs is that we do
717: not have a robust method of estimating the location of the transition
718: radius $R_{\rm tr}$. From the earliest studies (e.g., NY95b; Meyer \&
719: Meyer-Hofmeister 1994), it has been plausibly argued that $R_{\rm tr}$
720: increases with decreasing $\dot{M}$. However, reliable predictions of
721: the exact dependence have proved difficult, although a number of
722: studies have come up with semi-quantitative results (Meyer \&
723: Meyer-Hofmeister 1994; Honma 1996; Rozanska \& Czerny 2000; Meyer, Liu
724: \& Meyer-Hofmeister 2000; Spruit \& Deufel 2002; Mayer \& Pringle
725: 2007). Yuan \& Narayan (2004) tried to use observations to deduce the
726: run of $R_{\rm tr}$ with accretion luminosity. Figure 6 shows their
727: results.
728:
729: Another issue that has been discussed recently is hysteresis in the
730: transition between the thermal state and the hard state.
731: Specifically, with increasing $\dot{M}$, the hard state survives up to
732: fairly high luminosities $\sim0.1-0.2L_{\rm Edd}$, whereas with
733: decreasing $\dot{M}$, the transition from the thermal state to the
734: hard state occurs at a much lower luminosity $\sim 0.02-0.05L_{\rm
735: Edd}$ (Miyamoto et al. 1995; Maccarone \& Coppi 2003; Zdziarski et
736: al. 2004; Done et al. 2007). Meyer-Hofmeister, Liu \& Meyer (2005)
737: have provided a plausible explanation. Their disk evaporation model,
738: coupled with Compton-cooling of the hot electrons in the ADAF,
739: naturally produces a hysteresis in the location of the transition
740: radius.
741:
742: The brightest systems in the hard state ($L\ \gta\ 0.1L_{\rm Edd}$)
743: are difficult to model with the standard ADAF model. A variant of the
744: ADAF solution --- a natural extension of the model --- called the
745: luminous hot flow (LHAF; Yuan 2001, 2003; Yuan \& Zdziarski 2004; Yuan
746: et al. 2007; see also Machida, Nakamura \& Matsumoto 2006), looks
747: promising for modeling these sources.
748:
749: \subsection{Outflows and Jets}
750:
751: NY94, NY95a discovered an unexpected property of the ADAF solution:
752: {\it the accreting gas in an ADAF has a positive Bernoulli
753: parameter}\footnote{Technically, it should be called the Bernoulli
754: ``function'' since the quantity varies with radius, but we use
755: Bernoulli ``parameter'' since that was the term used in NY94}, which
756: is defined as the sum of the kinetic energy, potential energy and
757: enthalpy,
758: \begin{equation}
759: Be = v^2/2 -GM/R +w.
760: \end{equation}
761: A positive Bernoulli constant means that the gas is not bound to the
762: BH. (This is not surprising, since the gas is not losing energy
763: through radiation.) Therefore, the above authors suggested that ADAFs
764: should be associated with strong winds and jets\footnote{There has
765: been a tendency in the literature to ignore these early papers and to
766: credit Blandford \& Begelman (1999) for establishing the connection
767: between ADAFs and outflows. We note that the abstract of the first
768: paper on ADAFs, NY94, states: ``Further, the Bernoulli parameter is
769: positive, implying that advection-dominated flows are susceptible to
770: producing outflows...We suggest that advection-dominated accretion may
771: provide an explanation for...the widespread occurrence of outflows and
772: jets in accreting systems.'' In the second paper, NY95a, the title
773: itself mentions outflows, ``Advection-Dominated Accretion:
774: Self-Similarity and Bipolar Outflows,'' and an entire paragraph of the
775: abstract is devoted to the topic.} (see also Meier 2001).
776:
777: Strong outflows have been seen in numerical simulations of ADAFs. The
778: first indications came from 2D and 3D hydrodynamic simulations (Stone,
779: Pringle \& Begelman 1999; Igumenshchev \& Abramowicz 2000;
780: Igumenshchev, Abramowicz \& Narayan 2000), but it was soon confirmed
781: in MHD simulations as well (Stone \& Pringle 2001; Hawley \& Balbus
782: 2002; Igumenshchev, Narayan \& Abramowicz 2003; Pen, Matzner \& Wong
783: 2003; Machida, Nakamura \& Matsumoto 2004; McKinney \& Gammie 2004;
784: Igumenshchev 2004).
785:
786: Apart from producing a gas-dominated, large-scale outflow, MHD
787: simulations of ADAFs also have a second distinct outflow component
788: along the axis in the form of a collimated, Poynting-dominated,
789: relativistic jet (McKinney 2005, 2006). The original suggestion of
790: NY94 that ADAFs would have outflows and jets has thus been confirmed
791: by these simulations. Nevertheless, some authors have disputed a
792: connection between a positive Bernoulli parameter and an outflow since
793: it is possible to come up with explicit models that have a positive
794: Bernoulli parameter but no outflow (Paczy\'nski 1998; Abramowicz et
795: al. 2000).
796:
797: Observational evidence for the association of nonthermal relativistic
798: jets with ADAFs has accumulated in recent years with the discovery of
799: radio emission in virtually every BHB in the hard state (Corbel et
800: al. 2000; Fender 2001; Fender, Belloni \& Gallo 2004; Fender \&
801: Belloni 2004). The radio emission is generally too bright to be
802: produced by thermal electrons in the accretion flow. It is therefore
803: very likely to come from nonthermal electrons in a jet; in fact, radio
804: VLBI imaging has revealed a resolved jet in a few sources. Thus, it
805: is now observationally well-established that the hard state/ADAF is
806: associated with relativistic jets. According to the discussion in
807: \S3.5, the quiescent state also has an ADAF and should have a
808: (weaker) jet. Indeed, radio emission has been seen from two quiescent
809: systems, V404 Cyg (Hjellming et al. 2000; Gallo, Fender \& Hynes 2005)
810: and A0620--00 (Gallo et al. 2006), confirming this expectation.
811:
812: How much of the X-ray emission in an ADAF system comes from the
813: accretion flow and how much from the jet? A strong correlation has
814: been seen between radio and X-ray luminosity in the hard state and
815: quiescent state (Corbel et al. 2003; Gallo, Fender \& Pooley 2003).
816: At first sight this might suggest that the X-ray must also come from
817: the jet (e.g., Falcke, K\"ording \& Markoff 2004). However, since the
818: jet flows out of the ADAF and is thus highly coupled to it, any model
819: in which the X-ray emission is from the ADAF and radio is from the jet
820: is equally compatible with the observations. In recent times, several
821: authors have come down in favor of an ADAF origin for the X-rays
822: (e.g., Heinz \& Sunyaev 2003; Merloni, Heinz \& Di Matteo 2003; Heinz
823: 2004; Heinz et al. 2005; Yuan et al. 2005).
824:
825: Additional evidence on this issue comes from hard state spectra
826: indicating thermal emission (e.g., see Fig. 1), which is very
827: different from the nonthermal power-law emission one expects from a
828: jet. Markoff, Falcke \& Fender (2001), among others, have argued that
829: the X-ray emission is due to synchrotron emission from a carefully
830: tuned distrubution of nonthermal electrons. However, Zdziarski et
831: al. (2003) and Zdziarski \& Gierlinski (2004) showed that the
832: high-energy cutoff in the spectra of BHBs in the hard state, which
833: arises naturally in a thermal ADAF-like model, is very difficult to
834: reproduce in a nonthermal synchrotron model. A variant of the jet
835: model assigns the X-ray emission to thermal electrons in the ``base of
836: the jet'' (Markoff, Nowak \& Wilm 2005), but the discussion then
837: becomes semantic. The base of the jet is surely embedded in the
838: underlying ADAF and it is not clear that one gains anything by simply
839: relabeling the radiation from the ADAF as jet emission. The issue is
840: discussed in greater depth in Narayan (2005), who advocates a
841: `jet-ADAF' model (as described in Yuan et al. 2005; Malzac, Merloni \&
842: Fabian 2004), in which the high energy X-ray emission in the hard
843: state comes from the ADAF and the low-energy radio (and infrared)
844: emission comes from a jet.
845:
846: The situation is less clear-cut in the quiescent state, where
847: nonthermal emission from the jet might dominate even in X-rays (Yuan
848: \& Cui 2005; Wu, Yuan \& Cao 2007). It is interesting, however, that
849: the most quiescent system we know, Sgr A$^*$, shows no sign of a jet.
850: Radio images with a resolution of $\sim15R_S$ appear to be jet-free
851: (Shen et al. 2005), while the quiescent X-ray emission is spatially
852: resolved and appears to be from thermal gas near the Bondi radius at
853: $\sim10^5R_S$, not from a jet (Baganoff et al. 2001, 2003).
854:
855: Turning our attention now to the extended wind that is predicted from
856: an ADAF, and seen in numerical simulations, one consequence of the
857: wind is that, at each radius, a fraction of the accreting gas is lost
858: from the system. Thus, the accretion rate itself varies with radius.
859: This is usually parameterized with an index $s$ such that $\dot{M}(r)
860: \propto r^{s}$. Operationally, this means that, in the self-similar
861: regime, the gas density varies with radius as $\rho\propto r^{-3/2+s}$
862: rather than as $r^{-3/2}$ in the NY94 model. Unfortunately, apart
863: from the rather general constraint that $s$ should lie between 0 (no
864: mass loss) and 1 (the limit of a CDAF, \S3.7), there is no good
865: theoretical estimate of the value of $s$. We have to resort to
866: numerical simulations or observations. The former tends to give
867: somewhat larger values, e.g., $s\sim0.7$ (Pen et al. 2003), while the
868: one case where the latter approach has been tried, Sgr A$^*$, yields a
869: smaller value, $s\sim0.3$ (Yuan et al. 2003).
870:
871: \subsection{Convection}
872:
873: Another general property of ADAFs was highlighted by NY94 (see also
874: Begelman \& Meier 1982): ADAFs have entropy increasing inward and
875: should be convectively unstable by the Schwarzschild criterion.
876: Numerical hydrodynamic simulations of ADAFs confirmed the presence of
877: strong convection (Igumenshchev et al. 2000; Narayan, Igumenshchev \&
878: Abramowicz 2000) and led to the development of an analytical
879: self-similar model called the convection-dominated accretion flow
880: (CDAF; Narayan et al. 2000; Quataert \& Gruzinov 2000).
881:
882: The CDAF model employs a simplified one-dimensional treatment of the
883: accretion flow, in which all fluxes are assumed to be in the radial
884: direction. The resulting density varies as $\rho \propto r^{-1/2}$.
885: In the language of \S3.6, a CDAF corresponds to $s=1$. However,
886: technically, an idealized CDAF has no mass outflow, only an outward
887: flow of energy by convection; the energy is assumed to flow out into a
888: surrounding medium. In practice, the flow is never perfectly
889: one-dimensional; 2D effects intrude and one expects substantial winds
890: and mass loss as well. A real ADAF thus involves a complicated
891: interplay between convection and winds.
892:
893: Magnetic fields introduce further complexity, and there has been some
894: discussion in the literature on whether MHD ADAFs do or do not have
895: real convection (Machida, Matsumoto \& Mineshige 2001; Balbus \&
896: Hawley 2002; Narayan et al. 2002b; Pen et al. 2003; Igumenshchev
897: 2004). The question has no practical significance, however, since
898: there is general agreement that ADAFs have both unstable entropy
899: gradients and strong magnetic stresses.
900:
901: In the particular case of a spherical, non-rotating, MHD flow (the
902: magnetic analog of the Bondi problem), some complications are absent
903: and the problem becomes relatively clean. Here one finds that
904: convection and magnetic fields strongly influence the accretion
905: physics, and the resulting flow is very different from the standard
906: Bondi solution (Igumenshchev \& Narayan 2002; Igumenshchev 2004,
907: 2006). This result is likely to have significant implications for
908: astrophysics; for instance, isolated neutron stars accreting from the
909: interstellar medium will be very much dimmer than one might expect
910: based on the Bondi solution (Perna et al. 2003).
911:
912: \subsection{ADAFs Around Supermassive Black Holes}
913:
914: The ADAF solution is essentially independent of the mass $M$ of the
915: central BH. That is, if length and time are scaled by $M$ and the
916: accretion rate is scaled by the Eddington rate (also proportional to
917: $M$), then the same solution is valid for any BH mass. Therefore, any
918: successful application of the ADAF model to a BHB system has immediate
919: consequences for supermassive BHs (SMBHs) in an equivalent state, and
920: vice versa. This close connection between ADAFs in BHBs and ADAFs in
921: AGN was highlighted in Narayan (1996); it is also empirically obvious
922: from Fig. 6, which combines observations and models of BHBs and AGN.
923:
924: The first SMBH to be modeled as an ADAF was the Galactic Center source
925: Sgr A$^*$ (Narayan et al. 1995). This was soon followed by Fabian \&
926: Rees (1995), who suggested that quiescent nuclei in nearby giant
927: ellipticals (e.g., M87) must be accreting via ADAFs, and by Lasota et
928: al. (1996a), who argued that Low Ionization Nuclear Emission-line
929: Region sources (LINERs, e.g., NGC 4258) and low-luminosity active
930: galactic nuclei (LLAGN) must have ADAFs. Both suggestions have turned
931: out to be correct (e.g., Reynolds et al. 1996; Mahadevan 1997; Gammie
932: et al. 1999; Quataert et al. 1999; Di Matteo et al. 2000, 2003;
933: Loewenstein et al. 2001; Ulvestad \& Ho 2001; Nemmen et al. 2006).
934:
935: Other authors have suggested that all of the following systems have
936: ADAFs: FRI sources (Baum, Zirbel \& O'Dea 1995; Reynolds et al. 1996;
937: Begelman \& Celotti 2004), BL Lac sources (Maraschi \& Tavecchio
938: 2003), X-ray Bright Optically Normal Galaxies (XBONGs; Yuan \& Narayan
939: 2004), and even some Seyferts (Chiang \& Blaes 2003). All of these
940: sources are relatively low-luminosity AGN, where an ADAF is likely to
941: be present (Figs. 5, 6).
942:
943: As discussed in \S3.6, a feature of the ADAF model is the presence of
944: outflows and jets. This implies that ADAF systems should generally be
945: radio-loud. This is certainly the case for the FRI and BL Lac sources
946: mentioned above. Ho (2002; see also Nagar et al. 2000; Falcke et
947: al. 2000) presents in Fig. 5b of his paper a very interesting plot of
948: radio loudness, defined as the ratio of the flux at 6 cm to the flux
949: in the optical B band versus Eddington-scaled luminosity, for a
950: sample of galactic nuclei. Virtually every source in the plot with an
951: Eddington-ratio below about 0.01 is radio loud, which is perfectly
952: consistent with our expectation that all such sources should have
953: ADAFs. Moreover, the degree of radio loudness increases with
954: decreasing Eddington ratio, again consistent with our expectation that
955: the ADAF should become more and more dominant with decreasing
956: $\dot{M}$. The majority of sources in Ho's plot that have $L/L_{\rm
957: Edd} > 0.1$ are radio-quiet, as we would expect from Fig. 5 if these
958: systems have cool disks. However, a small minority of these bright
959: sources do show powerful jets (they are generally FRII sources). The
960: exact nature of the accretion in these sources is unclear. A more
961: complete plot, with updated data, can be found in Sikora, Stawarz \&
962: Lasota (2007).
963:
964: The jet and the extended wind from an ADAF carry with them substantial
965: kinetic energy. This energy will be dumped into the external medium
966: and will have important consequences. In the context of galaxy
967: formation, there has been discussion recently of the so-called ``radio
968: mode'' of accretion (Croton et al. 2006) in which outflowing energy
969: from an accreting SMBH in the galactic nucleus heats up the
970: surrounding medium. This kind of AGN feedback can lead to various
971: effects such as shutting off accretion and/or star formation (Di
972: Matteo et al. 2005; Hopkins et al. 2006a).
973:
974: It is worth noting that this radio mode is nothing other than the ADAF
975: mode of accretion reviewed in this article. Investigators of AGN
976: feedback may find it profitable to study the considerable work that
977: has been done on ADAFs over the last fifteen years.
978:
979: \section{The Black Hole Event Horizon}
980:
981: The first BH, Cygnus X-1, was identified and established in 1972 via a
982: measurement of its mass, which was shown to be too large for a neutron
983: star (NS). The surest evidence for the existence of BHs continues to
984: be through dynamical mass measurements. We now know of 20 additional
985: compact binary X-ray sources (McClintock \& Remillard 2006; Orosz et
986: al. 2007) with primaries that are too massive to be a NS or any stable
987: assembly of cold degenerate matter, assuming that GR is valid.
988: Similarly, dynamical data have established the existence of
989: supermassive BHs, most notably in the nucleus of our Milky Way Galacy
990: (Sch\"odel et al. 2002; Ghez et al. 2005a) and in NGC 4258 (Miyoshi et
991: al. 1995).
992:
993: Are these compact objects genuine BHs -- pockets of fully collapsed
994: matter that are walled off from sight by self gravity and that, like a
995: shadow, reveal no detail --- or are they exotic objects that have no
996: event horizons but manage to masquerade as BHs? Most astrophysicists
997: believe that they are genuine BHs. There are several reasons for this
998: confidence, the most important being that BHs are an almost inevitable
999: prediction of GR. However, this argument is circular because it
1000: presumes that GR is the correct theory of gravity. Furthermore, it
1001: ignores the many exotic alternatives to BHs that have been suggested.
1002: Thus, the collapsed objects that we refer to throughout as ``black
1003: holes'' are strictly speaking dynamical BH candidates. The current
1004: evidence for BHs is not decisive, nor can dynamical measurements be
1005: expected to make it so. We now consider some approaches aimed at
1006: establishing that these dynamical BHs are genuine.
1007:
1008: The defining property of a BH is its event horizon. Demonstrating the
1009: existence of this immaterial surface would be the certain way to prove
1010: the reality of BHs. Unfortunately, unlike any ordinary astronomical
1011: body such as a planet or a star of any kind, it is quite impossible to
1012: detect radiation from the event horizon's surface of infinite
1013: redshift. (Hawking radiation is negligibly weak for massive
1014: astrophysical BHs.) Nevertheless, despite the complete absence of any
1015: emitted radiation, it is possible to marshal strong {\it
1016: circumstantial evidence} for the reality of the event horizon. The
1017: fruitful approaches described below are based on comparing X-ray
1018: binary systems that contain BH primaries with very similar systems
1019: that contain NS primaries. Such investigations are motivated by the
1020: simple fact that the termination of an accretion flow at the hard
1021: surface of a NS has observational consequences.
1022:
1023: The earliest and strongest evidence for the event horizon is based on
1024: the faintness in quiescence of BH transient systems relative to
1025: comparable NS systems. In \S4.1, we discuss the physical arguments
1026: that underpin this evidence, and in \S4.2 we present the comparative
1027: luminosity data for BHs and NSs. Alternative explanations for the
1028: lower luminosities of the BH systems, which do not involve the event
1029: horizon, are considered in \S4.3. In \S4.4, we present three
1030: independent and additional arguments for the existence of the event
1031: horizon, which are again rooted in comparing BH and NS X-ray binaries.
1032: In \S4.5, we discuss the extreme faintness and properties of Sgr A*
1033: and the evidence that this supermassive BH has an event horizon, and
1034: in \S4.6 we argue that even very exotic objects with enormously strong
1035: surface gravity (e.g., gravastars) cannot escape our arguments for the
1036: event horizon.
1037:
1038: \subsection{Accretion and the Event Horizon}
1039:
1040: A test particle in a circular orbit at radius $R$ around a mass $M$
1041: has, in the Newtonian limit, kinetic energy per unit mass equal to
1042: $GM/2R$, and potential energy equal to $-GM/R$. In the context of a
1043: gaseous accretion disk, this means that a gas blob at radius $R$
1044: retains $50\%$ of its potential energy as kinetic energy. The
1045: remaining 50\% was transformed into thermal energy during the viscous
1046: accretion of the blob to its current radius (Frank et al. 2002).
1047:
1048: We now turn to consider accretion on to the material surface of a NS.
1049: If the mass $M$ at the center of an accretion disk has a surface, then
1050: the kinetic energy in the accreting gas will be converted to thermal
1051: energy in a viscous boundary layer and radiated (Frank et al. 2002).
1052: In addition, the residual thermal energy that the gas possesses when
1053: it reaches the inner edge of the disk will also be radiated from the
1054: surface.
1055:
1056: If accretion occurs via an ADAF, the luminosity from the accretion
1057: disk $L_{\rm acc}$ and that from the stellar surface $L_{\rm surf}$
1058: will satisfy
1059: \begin{eqnarray}
1060: L_{\rm acc} &\ll& GM\dot{M}/2R_*, \\
1061: L_{\rm surf} &\approx& GM\dot{M}/R_* \gg L_{\rm acc},
1062: \label{ADAFsurf}
1063: \end{eqnarray}
1064: where $R_*$ is the radius of the stellar surface. The accretion
1065: luminosity is small because the flow is radiatively inefficient.
1066: Therefore, essentially all the potential energy of the accreting gas
1067: remains in the gas in the form of thermal and kinetic energy. If the
1068: central object has a surface, e.g., it is a NS, the total luminosity
1069: we observe will be equal to $L_{\rm surf}$. However, if the object
1070: has an event horizon, i.e., it is a BH, there will be no radiation
1071: from a stellar surface and the luminosity we observe will only be
1072: equal to $L_{\rm acc}$. Thus, we expect
1073: \begin{equation}
1074: {\rm ADAF}: \quad L_{\rm NS} \gg L_{\rm BH}.
1075: \label{ADAFlum}
1076: \end{equation}
1077:
1078: \begin{figure}
1079: \includegraphics[width=3.3in,clip]{radeff}
1080:
1081: \caption{Predicted luminosities of accreting BHs and NSs as a function
1082: of mass accretion rate. This plot is based on the accretion
1083: efficiency prescription shown in Fig. 4. Because a NS has a surface,
1084: the total luminosity is always $\sim GM_{\rm NS}\dot{M}/R_{\rm NS}
1085: \sim 0.1\dot{M}c^2$, regardless of whether the accretion flow is
1086: radiatively efficient or not. In the case of a BH, however, the
1087: radiative efficiency $\eta$ of the accretion flow is of paramount
1088: importance, and the observed luminosity becomes $\ll0.1\dot{M}c^2$
1089: when the accretion rate is highly sub-Eddington.}
1090: \end{figure}
1091:
1092: Figure 7 shows schematically the luminosity difference we predict
1093: between a NS and a BH. This plot is based on the accretion efficiency
1094: estimate shown in Fig. 4. Especially at low mass accretion rates, say
1095: $\dot{M}\ \lta\ 10^{-4}\dot{M}_{\rm Edd}$ (deep quiescent state), we
1096: expect a huge luminosity difference between NSs and BHs. Efforts to
1097: test this prediction are described in \S4.2.
1098:
1099: If the accretion disk is radiatively efficient, then the gas has
1100: little thermal energy when it reaches the inner edge of the disk. In
1101: this case, we expect\footnote{The result given here is for a
1102: non-spinning accretor. If the central mass is spinning, the boundary
1103: layer luminosity from the surface will be somewhat smaller. However,
1104: it requires very fine tuning to make $L_{\rm surf} \ll L_{\rm acc}$.
1105: Moreover, if the central mass spins too close to the ``break-up''
1106: limit, the extra luminosity due to the surface can actually exceed
1107: $GM\dot{M}/2R_*$ by a large factor (Popham \& Narayan 1991).}
1108: \begin{eqnarray}
1109: L_{\rm acc} &\approx& GM\dot{M}/2R_*, \\
1110: L_{\rm surf} &\approx& GM\dot{M}/2R_* \sim L_{\rm acc}.
1111: \label{surf2}
1112: \end{eqnarray}
1113: The above result is based on a Newtonian analysis and is okay so long
1114: as the central object has a radius $R_*\ \gta\ R_{\rm ISCO}$, the
1115: radius of the ISCO. For more compact objects, we must allow for the
1116: fact that, inside the ISCO, the accreting gas no longer spirals in by
1117: viscosity but free-falls in the gravitational potential of the central
1118: mass. Now we have
1119: \begin{eqnarray}
1120: L_{\rm acc} &\approx& 0.1\dot{M}c^2, \\
1121: L_{\rm surf} &\approx& GM\dot{M}/R_*- L_{\rm acc} > L_{\rm acc}.
1122: \label{surf3}
1123: \end{eqnarray}
1124: The accretion luminosity is limited by the binding energy of the gas
1125: at the ISCO, which gives a radiative efficiency $\eta\sim0.1$. On the
1126: other hand, the total energy budget is $GM\dot{M}/R_*$, which means
1127: that a larger fraction of the luminosity is released at the surface.
1128: (The free-falling gas inside the ISCO, crashes on the surface and
1129: releases its energy in a shock.) In the limit when $R_*$ is
1130: arbitrarily close to $R_S$ (cf., the discussion of gravastars in
1131: \S4.6), the total luminosity is equal to $\dot{M}c^2$, i.e., the
1132: entire rest mass energy of the accreting gas is converted to radiation
1133: (100\% radiative efficiency). Note that all luminosities discussed
1134: here refer to measurements by an observer at infinity.
1135:
1136: Combining the results in equations (\ref{ADAFsurf}), (\ref{surf2}) and
1137: (\ref{surf3}), we see that, regardless of the radius of the accretor
1138: and whether the accretion flow is radiatively efficient or
1139: inefficient, we expect
1140: \begin{equation}
1141: {\rm Any~Accretion~Flow:} \qquad L_{\rm surf} \ \gta\ L_{\rm acc}.
1142: \label{Lsurfacc}
1143: \end{equation}
1144: Although this relation is weaker than (\ref{ADAFsurf}), it can still
1145: be used in favorable cases to test for the presence of an event
1146: horizon (\S4.5).
1147:
1148: \subsection{Evidence for the Event Horizon: Luminosities of Quiescent X-ray
1149: Binaries}
1150:
1151: BH and NS transient binary systems are very similar in many respects,
1152: and it is reasonable to expect that their mass accretion rates and
1153: luminosities would be comparable under similar conditions. There is,
1154: however, one important qualitative difference between the two kinds of
1155: object --- NSs have surfaces whereas BHs do not. Often, this
1156: difference is not important. However, as discussed in \S4.1 (see
1157: eq. \ref{ADAFlum}), when accretion occurs via an ADAF, a NS binary
1158: should be much more luminous than a BH binary (NY95b). The difference
1159: will be especially large in the quiescent state, when the accretion
1160: flow is radiatively extremely inefficient (Fig. 7).
1161:
1162: Narayan, Garcia \& McClintock (1997b) and Garcia, McClintock \&
1163: Narayan (1998) collected available X-ray data on quiescent NS and BH
1164: transients and showed that BH systems are consistently fainter than NS
1165: systems. This was the first indication that the objects that
1166: astronomers call ``black holes'' are indeed genuine BHs with
1167: event horizons.
1168:
1169: \begin{figure}
1170: \includegraphics[width=3.3in,clip]{nsbh1}
1171: \caption{Eddington-scaled luminosities (0.5--10 keV) of BH transients
1172: ({\it filled circles}) and NS transients ({\it open circles}) versus
1173: the orbital period (see Garcia et al. 2001; McClintock et al. 2004).
1174: Only the lowest quiescent detections or {\it Chandra/XMM} upper limits
1175: are shown. The plot shows all systems with known orbital periods,
1176: which have optical counterparts and good distance estimates. The
1177: diagonal hatched areas delineate the regions occupied by the two
1178: classes of sources and indicate the observed dependence of luminosity
1179: on orbital period. Note that the BH systems are on average nearly 3
1180: orders of magnitude fainter than the NS systems with similar orbital
1181: periods. }
1182: \end{figure}
1183:
1184: \begin{figure}
1185: \includegraphics[width=3.3in,clip]{nsbh2}
1186: \caption{Same as Fig.\ 8 except that shown here are the observed
1187: luminosities without Eddington scaling. In this representation, the
1188: BH systems are on average a factor of $\sim 100$ fainter than NS
1189: systems with similar orbital periods. }
1190: \end{figure}
1191:
1192: Lasota \& Hameury (1998) made the important point that it is necessary
1193: to compare quiescent NS and BH transients at similar mass accretion
1194: rates, and the surest way to achieve this is to plot luminosities as a
1195: function of the binary orbital period $P_{\rm orb}$ (the relevant
1196: arguments are outlined in Menou et al. 1999b; Lasota 2000, 2008).
1197: Since 1999, this has been the standard way of plotting the data (Menou
1198: et al. 1999a; Garcia et al. 2001; Narayan, Garcia \& McClintock 2002a;
1199: Hameury et al. 2003; McClintock et al. 2004). Figure 8 shows the
1200: current status of the comparison, with the luminosities of quiescent
1201: BH and NS transients plotted in (most-appropriate) Eddington units.
1202: Fig. 9 shows the same data without the luminosities being scaled. It
1203: is clear that, for comparable orbital periods, the BH systems are 2 to
1204: 3 orders of magnitude fainter than their NS cousins.
1205:
1206: As discussed in \S4.2, a radiatively inefficient ADAF provides a
1207: natural explanation for the large luminosity deficit of the BH
1208: systems. In this model, the gas approaches the center with a large
1209: amount of thermal energy. A BH is dim because the bulk of this
1210: thermal energy is trapped in the advective flow, passes through the
1211: event horizon, and is lost from sight. On the other hand, a NS is
1212: bright because the thermal energy is radiated from its surface.
1213:
1214: \subsection{Other Explanations for the Relative Faintness of Quiescent
1215: Black Hole Binaries}
1216:
1217: A number of alternative models have been put forward in an attempt to
1218: rationalize the luminosity differences between quiescent NS and BH
1219: systems. Notice below that some of these models generate X-ray
1220: emission without invoking accretion at all. We first consider a
1221: recent challenge centered on the discovery of the extremely
1222: low-luminosity NS transient 1H 1905+00. Next, we discuss the
1223: possibility that the bulk of the accretion power is channeled into a
1224: steady jet rather than into X-ray emission. We then consider several
1225: diverse models that are discussed in further detail in Narayan et
1226: al. (2002a).
1227:
1228: {\it The case of 1H 1905+00:} Jonker et al.\ (2007) claim that the
1229: extremely low X-ray luminosity of the NS transient and type I burst
1230: source 1H 1905+00 (hereafter, H1905) undermines the evidence
1231: summarized in \S4.2 for the existence of the event horizon. The
1232: luminosity of this NS ($\approx 2 \times 10^{30} ~{\rm erg\,s^{-1}}$;
1233: $D = 10$~kpc), is the same as the luminosity of A0620--00 and several
1234: other short-period ($P_{\rm orb} \approx 4-12$~hr) BH systems. Based
1235: on this result, Jonker et al.\ assert that the evidence for event
1236: horizons is ``unproven.'' However, they ignore a key point of our
1237: argument: Namely, as discussed above and illustrated in Figures 8 and
1238: 9, the case for event horizons depends critically on comparing BH and
1239: NS systems with {\it similar orbital periods}.
1240:
1241: We first note that the orbital period of H1905 is unknown. More
1242: importantly, this unknown period is believed to be very short -- so
1243: short that there are no BH systems with comparable periods. This
1244: conclusion is based on deep optical imaging data and the lack of a
1245: counterpart. Jonker et al.\ (2007) conclude that the secondary ``can
1246: only be a brown or a white dwarf,'' and that the system is probably an
1247: ultracompact binary. Hence the orbital period is likely to be tens of
1248: minutes, far less than the shortest BH binary period of 4.1~hr.
1249:
1250: Thus, there is no BH system comparable to H1905. And there is no way
1251: to usefully predict the mass accretion rate, as Jonker et al. conclude
1252: in the paper's final sentence. The rough trend of luminosity with
1253: period indicated in Figure 9 for NSs might imply a very low luminosity
1254: for H1905, as observed, but such an extrapolation is unwarranted. See
1255: Lasota (2007, 2008) for a more detailed discussion of this and other
1256: issues.
1257:
1258: Jonker et al.\ (2007) additionally argue that an unknown amount of
1259: mass transfered from the secondary might be lost from the system in
1260: winds or jets. Their one example of a wind, viz., an extraordinary
1261: mass-loss episode observed in GRO J1655-40 (Miller et al. 2006c) when
1262: the source was in outburst and had a luminosity $\sim 10^6$ times the
1263: quiescent level, is quite inappropriate. Also, their mention of
1264: possible mass expulsion for quiescent NSs via the propeller mechanism
1265: (Lasota \& Hameury 1998; Menou et al. 1999a), if effective, would
1266: reduce the luminosities of NSs and would only strengthen our argument.
1267: %On the other hand, if shown to be
1268: %significant, the preferential mass loss from BHs via their stronger
1269: %jets would weaken the above evidence for event horizons, a subject
1270: %that we discuss next.
1271:
1272: {\it Jet outflows:} Fender et al. (2003) argue that transient BHs at
1273: low accretion rates ($L/L_{\rm Edd} < 10^{-4}$) ``should enter
1274: `jet-dominated' states,'' in which the majority of the accretion power
1275: drives a radiatively-inefficient jet. As we have seen in \S3.6, there
1276: is good evidence for this. For instance, the presence of a radio jet
1277: has been reasonably well established in quiescence for A0620--00
1278: (Gallo et al. 2006), the closest and one of the least luminous of the
1279: BHs in question. Fender et al. (2003) appeal to the empirical result
1280: that BHs are $\sim 30$ times as `radio loud' as NSs at similar
1281: accretion rates to deduce that quiescent BHs should be $\sim 100$
1282: times {\it less} luminous in X-rays than quiescent NSs. Their
1283: argument is a little convoluted since, if the radio and X-ray emission
1284: in quiescence come from the jet (see \S3.6), one would think the BHs
1285: would be $\sim 30$ times {\it more} luminous, not 100 times less
1286: luminous. Therefore, naively, the jet argument only makes the
1287: discrepancy in Figs. 8 and 9 more severe. Fender et al. (2003) get
1288: around this difficulty by postulating, in addition to different jet
1289: efficiencies, also different origins for the X-rays seen in quiescent
1290: BHs and NSs.
1291: %This ad hoc argument does not require any significant
1292: %advection of energy through the event horizon and, if true, provides
1293: %an alternative explanation for the faintness of BHs. Fender et al.\
1294: %conclude that the difference in the quiescent X-ray luminosities of
1295: %BHs and NS can be mostly, if not entirely, explained by a difference
1296: %in the efficiency of jet production between the two types of sources.
1297:
1298: It is unsatisfying that the reason for the difference in jet activity
1299: between BHs and NSs ``remains unclear.'' At bottom, the authors are
1300: suggesting that in quiescence the BHs are in the `jet-dominated'
1301: regime and the NSs are, ``if not jet-dominated, close to the
1302: transition to this regime.'' A key uncertainty in this suggestion is
1303: whether the scaling relation that has been established for BHs between
1304: radio luminosity and jet luminosity ($L_{\rm radio} \propto L_{\rm
1305: x}^{0.7}$) is also valid for NSs, as is assumed.
1306: %Fender et al's.\ explanation is also
1307: %based on a purely theoretical relation for optically thick jets that
1308: %implies a radio luminosity that scales as the fourth root of the jet
1309: %power [**Ramesh: can their eqn 3 be criticized as uncertain for
1310: %stellar BHs?**].
1311: %While there are a number of uncertainties in the above argument, we
1312: %note that
1313: %($L/L_{\rm Edd} ~ 10^{-8.5}$).
1314: %Is the bulk of the accretion power in
1315: %this system carried off by its jet while only a small fraction of the
1316: %power is tapped off in X-rays, as Fender et al.\ argue? This is
1317: %possible. However, in addition to the uncertainties noted above,
1318: In addition, Occam's razor suggests that the relative faintness of BHs
1319: is unlikely to be attributable to their stronger jets: In \S4.5, we
1320: show that Sagittarius A$^*$, which is radiating in quiescence at the
1321: same level of Eddington-scaled luminosity as A0620-00, does not
1322: possess an energetically-important jet.
1323:
1324: {\it Coronal emission from BH secondaries:} The quiescent X-ray
1325: luminosity of the BH systems has been attributed to a
1326: rotationally-enhanced stellar corona in the secondary star (Bildsten
1327: and Rutledge 2000, but see Lasota 2000). However, as discussed in
1328: detail by Narayan et al. (2002a), the luminosities of three of the
1329: systems plotted in Figures 8 and 9 exceed by a factor of 6--60 the
1330: maximum predicted luminosity of the coronal model; likewise the three
1331: systems with adequate data quality show X-ray spectra that are
1332: harder/hotter than that typically seen in stellar coronae. Finally,
1333: if stellar coronae do contribute at some level, then the accretion
1334: luminosities of the BHs are even lower than our estimates, which would
1335: further strengthen the evidence for event horizons.
1336:
1337: {\it Incandescent neutron stars:} The quiescent luminosity of NS
1338: transients has been attributed to heating of the star's crust during
1339: outburst followed by cooling in quiescence. This model likewise has
1340: problems. (For details and references, see Narayan et al. 2002a.)
1341: Briefly, the rapid variability of the prototypical NS transient system
1342: Cen X-4 is not expected in a cooling model and implies that no more
1343: than about a third of the quiescent luminosity is due to crustal
1344: cooling. Furthermore, power-law tails that carry about half the total
1345: luminosity are observed for many NS transients (e.g., Cen X-4 and Aql
1346: X-1). These are unlikely to arise from a cooling NS surface, but
1347: could easily be produced by accretion. Finally, strong evidence for
1348: continued accretion in quiescence comes from optical variability,
1349: which is widely observed for these quiescent systems.
1350:
1351: {\it Pulsar wind/shock emission:} In another accretionless model, the
1352: NS transient switches to a radio pulsar-like mode in quiescence
1353: (Campana \& Stella 2000). The X-ray luminosity is expected to be of
1354: order the ``pulsar shock'' luminosity $\sim 2 \times 10^{32} ~{\rm
1355: erg\,s^{-1}}$, which is close to the observed level. Both the
1356: observed power-law and thermal components of emission (see the
1357: previous paragraph) are naturally explained by this model: the former
1358: component is produced by the shock and the latter is radiated from the
1359: NS surface. The lack of any significant periodicity in the quiescent
1360: emission (in any electromagnetic band) could be a problem for this
1361: model. Millisecond X-ray pulsars such as SAX J1808.4-3658 do show
1362: periodicities in outburst, but that emission is clearly the result of
1363: accretion, not a pulsar wind/shock.
1364:
1365: {\it Optical/UV luminosity:} Campana \& Stella (2000) note that the
1366: comparison shown in Figs. 8 and 9 assumes the X-ray luminosity is an
1367: accurate measure of the accretion rate near the BH or NS. They argue
1368: that the optical and UV luminosity also originates near the central
1369: object and should therefore be included in the comparison. The
1370: non-stellar optical/UV luminosity is much greater than the X-ray
1371: luminosity. When it is included, the difference between the BH and NS
1372: systems largely disappears. However, as detailed in Narayan et
1373: al. (2002a), there are some problems with this argument. The level of
1374: optical/UV emission generated in the inner region of ADAFs is strongly
1375: suppressed by winds and convection, as evidenced by a comparison of
1376: the luminosities of NSs and white dwarfs (Loeb, Narayan \& Raymond
1377: 2001). It thus appears unlikely that the optical and UV emission is
1378: generated in the hot gas close to the accretor (see Shahbaz et
1379: al. 2005). In the case of CVs, it has been established that a large
1380: fraction of the optical/UV emission comes from the ``hot spot,'' and
1381: it is quite reasonable to expect that this is true for BH and NS
1382: systems as well (there is some evidence to support this; Narayan et
1383: al. 2002a).
1384: %If true, the similar optical/UV luminosities of quiescent BH and NS
1385: %binaries with comparable orbital periods provides observational
1386: %confirmation of our interpretation of Figure 1.
1387: Further observations are needed to determine the origin of optical/UV
1388: emission in these systems.
1389:
1390: \subsection{Further Evidence for the Event Horizon in BHBs}
1391:
1392: In addition to the arguments discussed above, we briefly summarize
1393: three additional lines of evidence for the existence of event
1394: horizons. All three are based on comparisons between BH and NS X-ray
1395: binaries. As in the examples above, the first argument considers
1396: quiescent systems, whereas the latter two consider active states of
1397: accretion.
1398:
1399: In quiescence, a soft component of thermal emission is very commonly
1400: observed from the surfaces of accreting NSs, which is widely
1401: attributed to either deep crustal heating (Brown, Bildsten \& Rutledge
1402: 1998) or to accretion. No such component is present in the spectrum
1403: of the quiescent BHB XTE J1118+480 (hereafter J1118), as one would
1404: expect if the compact X-ray source is a bona fide BH that possesses an
1405: event horizon (McClintock et al.\ 2004). Because of the remarkably
1406: low column density to J1118 ($N_{\rm H} \approx 1.2 \times
1407: 10^{20}$~cm$^{-2}$) the limit on a hypothetical thermal source is very
1408: strong ($kT_{\infty}<$ 0.011 keV); it is in fact a factor of $\sim 25$
1409: lower in flux than the emission predicted by the theory of deep
1410: crustal heating, assuming that J1118 has a material surface analogous
1411: to that of NSs. Likewise, there is no evidence that accretion is
1412: occurring in quiescence onto the surface of J1118, which is the
1413: mechanism often invoked to explain the far greater thermal
1414: luminosities of NSs. The simplest explanation for the absence of any
1415: thermal emission is that J1118 lacks a material surface and possesses
1416: an event horizon.
1417:
1418: Type I X-ray bursts are very common in NS X-ray binaries, but no type
1419: I burst has been seen among the BH systems. A model developed by
1420: Narayan \& Heyl (2002, 2003), which reproduces the gross observational
1421: trends of bursts in NS systems, shows that, if the dynamical BHs have
1422: surfaces, they should exhibit instabilities similar to those that lead
1423: to type I bursts on NSs. Remillard et al.\ (2006), following earlier
1424: work by Tournear et al. (2003), measured the rates of type I X-ray
1425: bursts from a sample of 37 nonpulsing X-ray transients observed with
1426: {\it RXTE} during 1996--2004. Among the NS sources, they found 135
1427: type I bursts in 3.7 Ms of PCA exposures (13 sources) with a burst
1428: rate function consistent with the Narayan \& Heyl model. However, for
1429: the BH group (18 sources), they found no confirmed type I bursts in
1430: 6.5 Ms of exposure. Their upper limit on the incidence of burst
1431: activity in these sources is inconsistent with the model predictions
1432: at a high statistical significance if the accretors in BHBs have solid
1433: surfaces. The results provide strong indirect evidence for BH event
1434: horizons, and it would appear that the evidence can be refuted only by
1435: invoking rather exotic physics.
1436:
1437: Likewise drawing on the extensive archive of {\it RXTE} data, Done and
1438: Gierlinski (2003) have examined the patterns of X-ray spectral
1439: evolution of active BH and NS sources and identified a distinct type
1440: of soft spectrum that is occasionally observed only in the BH sources.
1441: They attribute this spectrum to thermal emission from the inner
1442: accretion disk (corresponding to the high state discussed earlier).
1443: They then argue that NSs with a similar accretion rate cannot exhibit
1444: such a simple, low-temperature spectrum because they would have a
1445: second component in the emission from the boundary layer where
1446: accreting matter impacts the stellar surface. They present a
1447: thermal/nonthermal Comptonization model for the boundary layer
1448: emission which has significant uncertainty, given that so many details
1449: of the accretion physics are complex and poorly understood.
1450: Nevertheless, Done \& Gierlinski (2003) do appear to have identified a
1451: systematic difference in the X-ray spectra of accreting BHs and NSs,
1452: and it is surely worth pursuing this signature in the effort to amass
1453: evidence for the reality of event horizons.
1454:
1455: %Sunyaev \& Revnivtsev (2000) analyzed {\it RXTE} timing data for a
1456: %sample of 9 BHs and 9 weakly-magnetized NSs in the low hard state. They
1457: %find that the power spectra of the NSs contain significant power at
1458: %frequencies above $\sim 500$~kHz, whereas the power spectra of the BHs
1459: %decline at much lower frequencies. They attribute this difference to NS
1460: %surface effects that occur in a thin radiation-dominated ``spreading
1461: %layer.'' It is suggested that via turbulence this layer generates
1462: %resonant sound waves and plasma instabilities with characteristic
1463: %frequencies well above the maximum Keplerian frequency and reaching up
1464: %to several kilohertz. The empirical result that NSs are noisier at high
1465: %frequencies is potentially useful for distinguishing them from BHs.
1466: %However, achieving a useful model of the extreme, noise-generating
1467: %accretion flow near the NS surface is a formidable task, and...
1468:
1469: \subsection{Event Horizon in Sagittarius A$^*$}
1470:
1471: The supermassive BH in Sgr A$^*$ has a mass $M\sim4\times10^6 M_\odot$
1472: (Sch\"odel et al. 2002; Ghez et al. 2005a) but an accretion luminosity
1473: of only $\sim 10^{36} ~{\rm erg\,s^{-1}}$; thus the source is highly
1474: sub-Eddington, $L\ \lta\ 10^{-8}L_{\rm Edd}$. From our earlier
1475: discussion, the accretion flow must be in the form of an ADAF, i.e.,
1476: the radiating gas must be extremely hot and optically thin. This
1477: expectation is confirmed by the $10^{10}$ K brightness temperature of
1478: the radio/millimeter emission (Shen et al. 2005) and the fact that
1479: most of the emission is in this band rather than at frequencies $\nu
1480: \sim kT/h$ (X-ray/$\gamma$-ray band).
1481:
1482: The ADAF model has been successfully applied to Sgr A$^*$ (Narayan et
1483: al. 1995, 1998a; Manmoto et al. 1997; Mahadevan 1998; Yuan et al. 2003;
1484: to mention a few). If we assume that the accretion flow is
1485: radiatively very inefficient, then it is easy to make a strong case
1486: for Sgr A$^*$ not having a surface (Narayan et al. 1998a). However, as
1487: Broderick \& Narayan (2006, 2007) showed, we can argue for the
1488: presence of an event horizon even without assuming an ADAF.
1489:
1490: If Sgr A$^*$ does not have an event horizon, but has a surface, then
1491: any emission from the surface will be blackbody-like. This is because
1492: we expect the surface to be optically thick. However, the
1493: radio/millimeter emission mentioned above cannot be from this surface
1494: because of its incredibly high $10^{10}$ K brightness temperature.
1495: Thus, the radio/millimeter radiation is from the accretion flow, and
1496: its luminosity gives a strict lower bound on $L_{\rm acc}$. By
1497: equation (\ref{Lsurfacc}) then, we expect a luminosity of at least
1498: $L_{\rm surf}\sim 10^{36} ~{\rm erg\,s^{-1}}$ from the surface. The
1499: surface radiation would be thermal and blackbody-like and should come
1500: out in the infrared (as easily shown, given the luminosity and the
1501: likely area of the surface). However, there is no sign of it!
1502:
1503: \begin{figure}
1504: \includegraphics[width=3.3in,clip]{SgrEH}
1505:
1506: \caption{The four solid lines show independent upper limits on the
1507: mass accretion rate at the surface of Sgr A$^*$ (assuming the source
1508: has a surface) as a function of the surface radius $R$. Each limiting
1509: curve is derived from a limit on the quiescent flux of Sgr A$^*$ in an
1510: infrared band. The hatched area at the top labeled ``Typical RIAF
1511: Range'' corresponds to the mass accretion rate in typical ADAF models
1512: of Sgr A$^*$ (e.g., Yuan et al. 2003). The horizontal dashed line
1513: represents the minimum accretion rate needed to power the bolometric
1514: luminosity of Sgr A$^*$. (From Broderick \& Narayan 2006)}
1515: \end{figure}
1516:
1517: Figure 10 shows constraints on the mass accretion rate in Sgr A$^*$.
1518: The horizontal dashed line is the minimum accretion rate $\dot{M}_{\rm
1519: acc}$ in Sgr A$^*$; this is the rate needed, even with a radiatively
1520: efficient flow ($\eta\sim0.1$), just to power the observed
1521: radio/millimeter radiation. The four solid lines show the maximum
1522: mass accretion rate $\dot{M}_{\rm surf}$ allowed if Sgr A$^*$ has a
1523: spherical surface of radius $R$. Each curve corresponds to a measured
1524: limit on the steady quiescent infrared flux in a particular band
1525: (Stolovy et al. 2003; Clenet et al. 2004; Ghez et al. 2005b) and
1526: provides an independent upper limit on $\dot{M}_{\rm surf}$. For
1527: radii less than $15GM/c^2$ (the upper limit on the size of Sgr A$^*$;
1528: Shen et al. 2005), we see that all four bands give upper limits on
1529: $\dot{M}_{\rm surf}$ that are far below the minimum $\dot{M}_{\rm
1530: acc}$; in fact, the discrepancy is larger than a factor of 100 in the
1531: $3.8 ~\mu{\rm m}$ band. Note that the discrepancy would be much
1532: larger if we assumed that the accretion flow is radiatively
1533: inefficient (corresponding to the hatched region in Fig. 10).
1534:
1535: The only way to avoid the large discrepancy illustrated in Fig. 10 is
1536: to give up the assumption that Sgr A$^*$ has a surface. Obviously, if
1537: the source has an event horizon, then we do not expect any surface
1538: radiation, and there is no problem.
1539:
1540: What if Sgr A$^*$ ejects all the accreting mass via a jet before the
1541: gas reaches the surface? Is this a viable explanation for the lack of
1542: surface emission? We can easily rule out this possibility. Recall
1543: that the source of energy in an accretion system is gravity. At a
1544: bare minimum, we know that matter is being accreted at a rate
1545: $\dot{M}_{\rm acc}$ (shown by the horizontal dashed line in Fig. 10)
1546: in order to produce the observed radiation. All of this mass {\it
1547: has} to fall into the potential well in order to release its energy.
1548: If Sgr A$^*$ has a jet with a kinetic luminosity $L_{\rm jet}$, the
1549: energy for this must also come from accretion. We will then require
1550: an even larger $\dot{M}_{\rm acc}$, and hence a larger $\dot{M}_{\rm
1551: surf}$, since the accreting gas now has to power both the radiation
1552: and the jet. The discrepancy with the observed limits on the
1553: quiescent infrared flux would then be even larger.
1554:
1555:
1556: \subsection{Effects of Very Strong Gravity}
1557:
1558: The arguments for the event horizon presented in \S\S4.2, 4.5 are
1559: based on Newtonian ideas. Some authors (e.g., Abramowicz, Kluzniak \&
1560: Lasota 2002) have questioned whether the arguments might be
1561: substantially modified by strong gravity in the vicinity of the
1562: compact accretor.
1563:
1564: Buchdahl (1959) showed, for a wide class of reasonable equations of
1565: state, that the smallest radius allowed for a compact non-rotating
1566: object is $R_{\rm min}= (9/8)R_S$. An object with this limiting
1567: radius has a gravitational redshift from its surface $z_{\rm gr}=2$.
1568: At such modest redshifts, we do not expect the effects of strong
1569: gravity to be particularly large, and so the arguments presented in
1570: \S\S4.2, 4.5 will continue to hold.
1571:
1572: Recently, however, a new class of solutions has been discussed, which
1573: goes variously under the name of ``gravastar'' and ``dark energy
1574: star'' (Mazur \& Mottola 2001; Chapline et al. 2003; Visser \&
1575: Wiltshire 2004; Carter 2005; Lobo 2006). In this model, the radius
1576: $R$ of an object of mass $M$ is allowed to be arbitrarily close to
1577: $R_S$: $\Delta R \equiv R-R_S \ll R_S$. The surface redshift can
1578: then be arbitrarily large:
1579: \begin{equation}
1580: 1+z_{\rm gr}\approx (R_S/\Delta R)^{1/2} \gg 1.
1581: \end{equation}
1582: Although the gravastar model is very artificial\footnote{Broderick \&
1583: Narayan first submitted their 2007 paper to Phys. Rev. Lett., but the
1584: paper was not even sent out for review because, in the words of the
1585: editor, ``the gravastar model of Mazur and Mottola is not considered
1586: by the community to be a viable alternative.'' B\&N then submitted
1587: the paper to Class. Quantum Grav., where most papers on gravastars and
1588: dark energy stars are published.}, it is nevertheless interesting to
1589: ask whether such a model, which has no event horizon, can explain the
1590: observations described in \S\S4.2, 4.5 (Abramowicz et al. 2002).
1591: Broderick \& Narayan (2006, 2007) show that it cannot. We briefly
1592: summarize the arguments here.
1593:
1594: A large value of $z_{\rm gr}$ means that any radiation emitted from
1595: the surface is redshifted greatly before it reaches the observer at
1596: infinity. This looks like an easy way of hiding the surface
1597: luminosity. However, a simple energy conservation argument shows
1598: otherwise. Assuming steady state, in the limit as $\Delta R/R_S\to
1599: 0$, the total luminosity observed at infinity, $L_{\rm acc} +L_{\rm
1600: surf}$, {\it must} be equal to $\dot{M}c^2$ (the rate of accretion of
1601: rest mass energy). Moreover, the effective radius of the source as
1602: viewed by a distant observer is $(3\sqrt{3}/2)R_S=2.6R_S$, and so the
1603: surface radiation will be in the X-ray band for accreting BHBs
1604: (\S4.2) and in the infrared for Sgr A$^*$ (\S4.5), exactly as in the
1605: Newtonian analysis.
1606:
1607: Could the large gravitational redshift cause a large delay in the
1608: signals from the surface of the star, and could this be why we do not
1609: see the radiation? The extra delay due to relativity is easily
1610: estimated by considering null geodesics in the Schwarzschild metric:
1611: \begin{equation}
1612: \Delta t = (R_S/c)\ln(R_S/\Delta R).
1613: \end{equation}
1614: Since the dependence is only logarithmic, the extra delay is not
1615: significant. For instance, even if we take $\Delta R$ to be
1616: comparable to the Planck length $\sim 10^{-33}$ cm (the smallest
1617: length we can legitimately consider), the delay is only $\sim 10$ ms
1618: for a BHB and $\sim 1000$ s for Sgr A$^*$.
1619:
1620: We assumed that the radiation from the surface would have a
1621: blackbody-like spectrum. Is this likely? Actually, it is virtually
1622: guaranteed as $\Delta R/R_S\to0$. When the radius of an object is
1623: very close to the Schwarzschild radius, most rays emitted from the
1624: surface are bent back on to the surface, and only a tiny fraction of
1625: rays escapes to infinity. The solid angle corresponding to escape is
1626: $\Omega_{\rm esc} \sim \Delta R/R_S \ll 1$. In this limit, the
1627: surface behaves just like a furnace with a pinhole (the textbook
1628: example of a blackbody!), and so the escaping radiation is certain to
1629: have a nearly perfect blackbody spectrum.
1630:
1631: Could the surface luminosity escape in the form of particles rather
1632: than radiation? By the blackbody argument given above, whatever
1633: escapes must be in thermodynamic equilibrium at the temperature
1634: $T_\infty$ observed at infinity. For the sources of interest to us,
1635: $kT_\infty \sim$ eV (Sgr A$^*$) and keV (BHBs). In thermodynamic
1636: equilibrium, the only particles with any significant number density at
1637: these temperatures are neutrinos. Even if we include all three
1638: species of neutrinos, the radiation flux is reduced by a factor of
1639: only 8/29 (Broderick \& Narayan 2007). This is a small correction
1640: compared to the large discrepancies $\gta\ 100$ that we described in
1641: \S\S4.2, 4.5.
1642:
1643: Finally, we note that our arguments for the event horizon are based on
1644: a steady state assumption. Specifically, we assume that the surface
1645: luminosity is proportional to the average mass accretion rate on the
1646: surface. In the gravastar model, in particular, one can imagine
1647: scenarios in which steady state is not reached. However, Broderick \&
1648: Narayan (2007) show that this explanation can be ruled out, at least
1649: with current gravastar models.
1650:
1651: We thus conclude that strong gravity effects are unable to weaken our
1652: arguments for the presence of an event horizon in quiescent BHBs and
1653: in Sgr A$^*$. We must look to more conventional astrophysical
1654: explanations.
1655: %One possibility is that the surface is somehow able to radiate in a
1656: %band completely different from the blackbody band that we assume.
1657: %(For this to work, obviously, we cannot have an extremely compact
1658: %object with $\Delta R\ll R_S$, since we argued above that, in that
1659: %limit, a blackbody is guaranteed.) We are not aware of any reasonable
1660: %radiation model that can do this, but it is possible in principle.
1661: The only idea that we find somewhat plausible is that the radiation we
1662: observe from quiescent X-ray binaries and Sgr A$^*$ is not produced by
1663: accretion at all, but by some other process. This would undercut all
1664: our arguments since we assume, as an article of faith, that the
1665: radiation we observe is powered by gravity through accretion. Given
1666: the success of the accretion paradigm in explaining a vast body of
1667: observations on a variety of BH systems in many different spectral
1668: states, it seems rather extreme to abandon the idea of accretion in
1669: just those particular sources where we find evidence for the presence
1670: of an event horizon.
1671:
1672: %Other explanations such as those discussed in \S4.3 are not promising
1673: %since they work only in some sources and not in others. For instance,
1674: %one could invoke mass outflow in a jet to explain why quiescent BHBs
1675: %are underluminous (Gallo et al. ??), but the same jet explanation will
1676: %not work in the case of Sgr A$^*$ (\S4.5).
1677:
1678: \section{Conclusion}
1679:
1680: The aim of this article is two-fold: To give an updated and current
1681: account of the ADAF model, with an emphasis on applications to
1682: accreting BHs (\S3), and to review the considerable body of evidence
1683: presently available for the presence of event horizons in
1684: astrophysical BHs (\S4).
1685:
1686: Observations of BH binaries (BHBs) and active galactic nuclei (AGN)
1687: indicate that, at luminosities of a few percent or less of Eddington,
1688: accretion occurs via a very different mode than the standard thin
1689: accretion disk. At these luminosities, sources have hard X-ray
1690: spectra, quite unlike the soft blackbody-like spectra seen in more
1691: luminous sources. Prior to the establishment of the ADAF model,
1692: observations in the hard state were modeled in an ad hoc way using
1693: empirical thermal Comptonization models. In the mid-1990s, the ADAF
1694: solution was shown to have precisely the densities, temperatures,
1695: radiative (in)efficiencies -- and stability -- required to provide a
1696: physical description of the observations. Moreover, since the ADAF
1697: model is essentially mass-independent, observations of BHBs and AGN
1698: are both explained with more-or-less the same ADAF model. The model
1699: gives satisfactory results for a wide range of luminosity, from about
1700: $10^{-1}$ down to $10^{-8}$ of Eddington (below which there are no
1701: observations).
1702:
1703: The defining characteristic of an ADAF is that the gas is not
1704: radiatively efficient. A significant fraction of the energy released
1705: by viscous dissipation is retained in the accreting gas and advected
1706: with the flow. The trapped thermal energy causes the accreting gas to
1707: be weakly bound to the BH. Based on this property, it was predicted
1708: already in the earliest ADAF papers that sources in the ADAF state
1709: would have strong winds and jets. Nonthermal radio emission has been
1710: detected in recent years from BHBs in the hard state and quiescent
1711: state, as well as from all AGN with luminosities below $\sim1\%$ of
1712: Eddington. These sources have spectra consistent with the ADAF model,
1713: thus confirming a strong connection between ADAFs and jets/outflows.
1714: A currently active topic of research is the role of AGN feedback on
1715: galaxy formation. Research in this area can now draw on the extensive
1716: literature on ADAFs.
1717:
1718: Of special interest in this work are those accreting objects for which
1719: there exist estimates or constraints on the two system specific
1720: parameters: BH mass $M$ and mass accretion rate $\dot M$.
1721: Specifically, we are referring here to Saggitarius A$^*$ and a
1722: selected sample of quiescent BH and NS X-ray binaries. We have
1723: featured these two examples because they provide strong observational
1724: evidence for the existence of the event horizon.
1725:
1726: In the case of the binaries, one can make a plausible argument that,
1727: for comparable orbital periods, the mass accretion rates and
1728: luminosities of both types of systems should be comparable.
1729: Remarkably, however, the BH systems are observed to be dimmer by a
1730: factor of $\sim100-1000$. We review a wide range of unsatisfactory
1731: attempts to explain this large luminosity difference. In contrast,
1732: the ADAF model provides an entirely straightforward explanation for
1733: the faintness of quiescent BHs: A NS must radiate the trapped thermal
1734: radiation that is advected with the accretion flow and rains down on
1735: its surface, whereas the BH hides the energy behind its event horizon.
1736: Additional evidence for the event horizon is provided in further
1737: comparisons of NSs and BHs in outburst: The BHs lack type I X-ray
1738: bursts, and they lack a distinctive boundary-layer component of
1739: emission. Both of these properties are expected if the BHs possess
1740: event horizons, but hard to explain otherwise.
1741:
1742: The supermassive BH in Sgr A$^*$ is extraordinarily quiescent. The
1743: observations strongly support the existence of an ADAF, and it is
1744: consequently easy to provide a compelling argument for an event
1745: horizon. However, for Sgr A$^*$, one can make an even stronger
1746: argument for the lack of a hypothetical material surface without at
1747: all invoking an ADAF. The large radio/millimeter accretion luminosity
1748: of Sgr A$^*$, which has a brightness temperature $>10^{10}$ K,
1749: obviously cannot be emitted by an optically-thick surface, and so it
1750: must be radiated from the accretion flow. This establishes a hard
1751: lower limit on the accretion luminosity and $\dot{M}$. Meanwhile,
1752: high angular resolution radio observations constrain the radius of the
1753: surface to be $< 15R_{\rm S}$. This constraint and the lower limit on
1754: $\dot M$ predict a near-IR flux of thermal, blackbody-like surface
1755: emission that is far above the observed limits. The obvious
1756: explanation is that there is no material surface, only an event
1757: horizon. As an added bonus, it is energetically impossible to explain
1758: away the lack of surface emission by appealing to mass loss in a jet.
1759:
1760: The case for event horizons in both Sgr A$^*$ and in the stellar-mass
1761: BHs is robust against appeals to strong gravity or the leading models
1762: of exotic stars. First, GR effects will be mild for nearly all
1763: conventional models of degenerate stars, whose radii are restricted to
1764: be $\ge 9/8R_{\rm S}$ and whose surface redshifts are therefore $\le
1765: 2$. Secondly, even for an exotic star (a gravastar) with an
1766: extraordinary surface redshift of a million or more, the full
1767: accretion luminosity from its surface will be delivered as X-rays (in
1768: binaries) or infrared (in Sgr A$^*$) to a distant observer. Thus,
1769: extreme redshifts have practically no effect on the argument for the
1770: event horizon.
1771:
1772: To play on Carl Sagan's famous comment, there will always be an
1773: absence of {\it direct} evidence for the event horizon, but this
1774: surely cannot be taken as evidence of its absence in nature. On the
1775: contrary, and with a pun in mind, many indicators show that the event
1776: horizon is an inescapable reality.
1777:
1778: \medskip\noindent The authors thank Bozena Czerny, Jean-Pierre Lasota
1779: and Feng Yuan for helpful comments.
1780:
1781: % The Appendices part is started with the command \appendix;
1782: % appendix sections are then done as normal sections
1783: % \appendix
1784:
1785: % \section{}
1786: % \label{}
1787:
1788: % Bibliographic references with the natbib package:
1789: % Parenthetical: \citep{Bai92} produces (Bailyn 1992).
1790: % Textual: \citet{Bai95} produces Bailyn et al. (1995).
1791: % An affix and part of a reference:
1792: % \citep[e.g.][Ch. 2]{Bar76}
1793: % produces (e.g. Barnes et al. 1976, Ch. 2).
1794:
1795: \begin{thebibliography}{}
1796:
1797: % \bibitem[Names(Year)]{label} or \bibitem[Names(Year)Long names]{label}.
1798: % (\harvarditem{Name}{Year}{label} is also supported.)
1799: % Text of bibliographic item
1800:
1801: \bibitem[Abramowicz et al.(1996)]{abr96}
1802: Abramowicz, M. A., Chen, X., Granath, M., \& Lasota, J. P. 1996,
1803: ApJ, 471, 762
1804:
1805: \bibitem[Abramowicz et al.(1995)]{abr95}
1806: Abramowicz, M. A., Chen, X., Kato, S., Lasota, J. P., \& Regev, O.
1807: 1995, ApJ, 438, L37
1808:
1809: \bibitem[Abramowicz et al.(1988)]{abr88}
1810: Abramowicz, M. A., Czerny, B., Lasota, J. P., \& Szuszkiewicz,
1811: E. 1988, ApJ, 332, 646
1812:
1813: \bibitem[Abramowicz et al.(2002)]{abr02}
1814: Abramowicz, M. A., Kluzniak, W., \& Lasota, J. P. 2002, A\&A, 396, L31
1815:
1816: \bibitem[Abramowicz et al.(2000)]{ali00}
1817: Abramowicz, M. A., Lasota, J. P., \& Igumenshchev, I. V.
1818: 2000, MNRAS, 314, 775
1819:
1820: \bibitem[Arnaud(1996)]{arn96}
1821: Arnaud, K. A. 1996, in ASP Conf. Ser. 191, Astronomical Data Analysis and
1822: Systems V, ed. G. H. Jacoby \& J. Barnes (San Francisco: ASP) p17
1823:
1824: \bibitem[Baganoff(2001)]{bag01}
1825: Baganoff, F. K., et al. 2001, Nature, 413, 45
1826:
1827: \bibitem[Baganoff(2003)]{bag03}
1828: Baganoff, F. K., et al. 2003, ApJ, 591, 891
1829:
1830: \bibitem[Balbus \& Hawley(2002)]{bal02}
1831: Balbus, S. A., \& Hawley, J. F. 2002, ApJ, 573, 749
1832:
1833: \bibitem[Balucinska-Church et al.(1995)]{bal95}
1834: Balucinska-Church, M., Belloni, T., Church, M. J., \&
1835: Hasinger, G. 1995, A\&A, 302, L5
1836:
1837: \bibitem[Baum et al.(1995)]{bau95}
1838: Baum, S. A., Zirbel, E. L., \& O'Dea, C. P. 1995, ApJ, 451, 88
1839:
1840: \bibitem[Begelman(1979)]{beg79}
1841: Begelman, M. C. 1979, MNRAS, 187, 237
1842:
1843: \bibitem[Begelman \& Celotti(2004)]{beg04}
1844: Begelman, M. C., \& Celotti, A. 2004, MNRAS, 352, L45
1845:
1846: \bibitem[Begelman \& Chiueh(1988)]{bc88}
1847: Begelman, M. C., \& Chiueh, T. 1988, ApJ, 332, 872
1848:
1849: \bibitem[Begelman \& Meier(1982)]{beg82}
1850: Begelman, M. C., \& Meier, D. L. 1982, ApJ, 253, 873
1851:
1852: \bibitem[Bildsten \& Rutledge(2000)]{bil00}
1853: Bildsten, L., \& Rutledge, R. E. 2000, ApJ, 541, 908
1854:
1855: \bibitem[Bisnovatyi-Kogan \& Lovelace(1997)]{bl97}
1856: Bisnovatyi-Kogan, G., \& Lovelace, R. V. E. 1997, ApJ, 486, L43
1857:
1858: \bibitem[Blackman(1999)]{bla99}
1859: Blackman, E. G. 1999, MNRAS, 302, 723
1860:
1861: \bibitem[Blandford \& Begelman(1999)]{bb99}
1862: Blandford, R. D., \& Begelman, M. C. 1999, MNRAS, 303, L1
1863:
1864: \bibitem[Broderick \& Narayan(2006)]{bn06}
1865: Broderick, A. E., \& Narayan, R. 2006, ApJ, 638, L21
1866:
1867: \bibitem[Broderick \& Narayan(2007)]{bn07}
1868: Broderick, A. E., \& Narayan, R. 2007, Class. Quantum Grav., 24, 659
1869:
1870: \bibitem[Brown et al.(1998)]{bbr98}
1871: Brown, E. F., Bildsten, L., \& Rutledge, R. E. 1998, ApJ, 504, L95
1872:
1873: \bibitem[Buchdahl(1959)]{buc59}
1874: Buchdahl, H. A. 1959, Phys. Rev., 116, 1027
1875:
1876: \bibitem[Campana \& Stella(2000)]{cam00}
1877: Campana, S., \& Stella, L. 2000, ApJ, 541, 849
1878:
1879: \bibitem[Carter(2005)]{car05}
1880: Carter, B. M. N. 2005, Class. Quantum Grav., 22, 4551
1881:
1882: \bibitem[Chapline et al.(2003)]{cha03}
1883: Chapline, G., et al. 2003, Int. J. Mod. Phys. A, 18, 3587
1884:
1885: \bibitem[Chen et al.(1997)]{che97}
1886: Chen, X., Abramowicz, M. A., \& Lasota, J. P. 1997, ApJ, 476, 61
1887:
1888: \bibitem[Chen et al.(1995)]{che95}
1889: Chen, X., Abramowicz, M. A., Lasota, J. P., Narayan, R., \& Yi, I. 1995,
1890: ApJ, 443, L61
1891:
1892: \bibitem[Chiang \& Blaes(2003)]{chi03}
1893: Chiang, J., \& Blaes, O. 2003, ApJ, 586, 97
1894:
1895: \bibitem[Clenet et al.(2004)]{cle04}
1896: Clenet, Y., et al. 2004, A\&A, 424, L21
1897:
1898: \bibitem[Corbel et al.(2000)]{cor00}
1899: Corbel, S., et al. 2000, A\&A, 359, 251
1900:
1901: \bibitem[Corbel et al.(2003)]{cor03}
1902: Corbel, S., Nowak, M. A., Fender, R. P., Tzioumis, A. K., \& Markoff, S.
1903: 2003, A\&A, 400, 1007
1904:
1905: \bibitem[Corbel et al.(2006)]{cor06}
1906: Corbel, S., Tomsick, J. A., \& Kaaret, P. 2006, ApJ, 636, 971
1907:
1908: \bibitem[Croton et al.(2006)]{cro06}
1909: Croton, D. J., et al. 2006, MNRAS, 365, 11
1910:
1911: \bibitem[Cui et al.(1999)]{cui99}
1912: Cui, W., Zhang, S. N., Chen, W., \& Morgan, E. H. 1999, ApJ, 512, L43
1913:
1914: \bibitem[D'Angelo et al.(2008)]{dan08}
1915: D'Angelo, C., Giannios, D., Dullemond, C., \& Spruit, H. 2008,
1916: A\&A, submitted
1917:
1918: \bibitem[Davis et al.(2006)]{ddb06}
1919: Davis, S. W., Done, C., \& Blaes, O. M. 2006, ApJ, 647, 525
1920:
1921: \bibitem[Davis \& Hubeny(2006)]{dav06}
1922: Davis, S. W., \& Hubeny, I. 2006, ApJS, 164, 530
1923:
1924: \bibitem[Deufel et al.(2002)]{deu02}
1925: Deufel, B., Dullemond, C. P., \& Spruit, H. C. 2002, A\&A, 387, 907
1926:
1927: %\bibitem[Deufel \& Spruit(2000)]{deu00}
1928: %Deufel, B., \& Spruit, H. C. 2000, A\&A, 362, 1
1929:
1930: \bibitem[Di Matteo et al.(2003)]{dim03}
1931: Di Matteo, T., Allen, S. W., Fabian, A. C., Wilson, A. S., \& Young, A. J.
1932: 2003, ApJ, 582, 133
1933:
1934: \bibitem[Di Matteo et al.(2000)]{dim00}
1935: Di Matteo, T., Quataert, E., Allen, S. W., Fabian, A. C., \& Narayan, R.
1936: 2000, MNRAS, 311, 507
1937:
1938: \bibitem[Di Matteo et al.(2005)]{dim05}
1939: Di Matteo, T., Springel, V., \& Hernquist, L. 2005, Nature, 433, 604
1940:
1941: \bibitem[Di Salvo et al.(2001)]{dis01}
1942: Di Salvo, T., Done, C., Zycki, P. T., Burderi, L., \& Robba, N. R. 2001,
1943: ApJ, 547, 1024
1944:
1945: \bibitem[Done \& Gierlinski(2003)]{dg03}
1946: Done, C., \& Gierlinski, M. 2003, MNRAS, 342, 1041
1947:
1948: \bibitem[Done et al.(2007)]{don07}
1949: Done, C., Gierlinski, M., \& Kubota, A. 2007, A\&AR, 15, 1
1950:
1951: \bibitem[Dubus et al.(2001)]{dub01}
1952: Dubus, G., Hameury, J. M., \& Lasota, J. P. 2001, A\&A, 373, 251
1953:
1954: \bibitem[Dullemond \& Spruit(2005)]{dul05}
1955: Dullemond, C. P., \& Spruit, H. C. 2005, A\&A, 434, 415
1956:
1957: \bibitem[Esin et al.(1997)]{esi97}
1958: Esin, A. A., McClintock, J. E., \& Narayan, R. 1997, ApJ, 489, 865
1959:
1960: \bibitem[Esin et al.(1998)]{esi98}
1961: Esin, A. A., Narayan, R., Cui, W., Grove, J. E., \& Zhang, S. N. 1998,
1962: ApJ, 505, 854
1963:
1964: \bibitem[Esin et al.(2001)]{esi01}
1965: Esin, A. A., et al. 2001, ApJ, 555, 483
1966:
1967: \bibitem[Fabian \& Rees(1995)]{fab95}
1968: Fabian, A. C., \& Rees, M. J. 1995, MNRAS, 277, L5
1969:
1970: \bibitem[Falcke et al.(1998)]{fal98}
1971: Falcke, H., Goss, W. M., Matsuo, H., Teuben, P., Zhao, J. H., \& Zylka,
1972: R. 1998, ApJ, 499, 731
1973:
1974: \bibitem[Falcke et al.(2000)]{fal00}
1975: Falcke, H., Nagar, N. M., Wilson, A. S., \& Ulvestad, J. S. 2000,
1976: ApJ, 542, 197
1977:
1978: \bibitem[Fender(2001)]{fen01}
1979: Fender, R. P. 2001, MNRAS, 322, 31
1980:
1981: \bibitem[Fender \& Belloni(2004)]{fb04}
1982: Fender, R. P., \& Belloni, T. 2004, ARAA, 42, 317
1983:
1984: \bibitem[Fender et al.(2004)]{fbg04}
1985: Fender, R. P., Belloni, T. M., \& Gallo, E. 2004, MNRAS, 355, 1105
1986:
1987: \bibitem[Fender et al.(2003)]{fen03}
1988: Fender, R. P., Gallo, E., \& Jonker, P. G. 2003, MNRAS, 343, L99
1989:
1990: \bibitem[Frank et al.(2002)]{fkr02}
1991: Frank, J., King, A., \& Raine, D. J. 2002, Accretion Power in Astrophysics
1992: (Cambridge Univ. Press)
1993:
1994: \bibitem[Gallo et al.(2005)]{gal05}
1995: Gallo, E., Fender, R. P., \& Hynes, R. I. 2005, MNRAS, 356, 1017
1996:
1997: \bibitem[Gallo et al.(2003)]{gal03}
1998: Gallo, E., Fender, R. P., \& Pooley, G. G. 2003, MNRAS, 344, 60
1999:
2000: \bibitem[Gallo et al.(2006)]{gal06}
2001: Gallo, E., et al. 2006, MNRAS, 370, 1351
2002:
2003: \bibitem[Gammie et al.(1999)]{gam99}
2004: Gammie, C. F., Narayan, R., \& Blandford, R. 1999, ApJ, 516, 177
2005:
2006: \bibitem[Garcia et al.(1998)]{gar98}
2007: Garcia, M. R., McClintock, J. E.,
2008: \& Narayan, R. 1998, ASP Conf. Series, Vol. 137, ed. S. Howell,
2009: E. Kuulkers \& C. Woodward, p.506
2010:
2011: \bibitem[Garcia et al.(2001)]{gar01}
2012: Garcia, M. R., McClintock, J. E., Narayan, R., Callanan, P.,
2013: Barret, D., \& Murray, S. S. 2001, ApJ, 553, L47
2014:
2015: \bibitem[Ghez et al.(2005a)]{ghe05a}
2016: Ghez, A. M., et al. 2005a, ApJ, 620, 744
2017:
2018: \bibitem[Ghez et al.(2005b)]{ghe05b}
2019: Ghez, A. M., et al. 2005b, ApJ, 635, 1087
2020:
2021: \bibitem[Gierlinski et al.(1997)]{gie97}
2022: Gierlinski, M., et al. 1997, MNRAS, 288, 958
2023:
2024: \bibitem[Gilfanov et al.(1999)]{gil99}
2025: Gilfanov, M., Churazov, E., \& Revnivtsev, M. 1999, A\&A, 352, 182
2026:
2027: \bibitem[Goldston et al.(2005)]{gol05}
2028: Goldston, J. E., Quataert, E., \& Igumenshchev, I. V. 2005, ApJ, 621, 785
2029:
2030: \bibitem[Hameury et al.(1999)]{ham99}
2031: Hameury, J. M., Lasota, J. P., \& Dubus, G. 1999, MNRAS, 303, 39
2032:
2033: \bibitem[Hameury et al.(1997)]{ham97}
2034: Hameury, J. M., Lasota, J. P., McClintock, J. E., \& Narayan, R.
2035: 1997, ApJ, 489, 234
2036:
2037: \bibitem[Hameury et al.(2003)]{ham03}
2038: Hameury, J. M., et al. 2003, A\&A, 399, 631
2039:
2040: \bibitem[Hameury et al.(2007)]{ham07}
2041: Hameury, J. M., Lasota, J. P., \& Viallet, M. 2007, in Black Holes from
2042: Stars to Galaxies, ed. V. Karas \& G. Matt, Cambridge Univ. Press, p297
2043:
2044: \bibitem[Hawley \& Balbus(2002)]{haw02}
2045: Hawley, J. F., \& Balbus, S. A. 2002, ApJ, 573, 738
2046:
2047: \bibitem[Hawley et al.(1996)]{haw96}
2048: Hawley, J. F., Gammie, C. F., \& Balbus, S. A. 1996, ApJ, 464, 690
2049:
2050: \bibitem[Heinz(2004)]{hei04}
2051: Heinz, S. 2004, 355, 835
2052:
2053: \bibitem[Heinz et al.(2005)]{hei05}
2054: Heinz, S., Merloni, A., Di Matteo, T., \& Sunyaev, R. 2005, Astrophys.
2055: Sp. Sci., 300, 15
2056:
2057: \bibitem[Heinz \& Sunyaev(2003)]{hei03}
2058: Heins, S., \& Sunyaev, R. 2003, MNRAS, 343, L59
2059:
2060: \bibitem[Hjellming et al.(2000)]{hje00}
2061: Hjellming, R. M., Rupen, M. P., Mioduszewski, A. J., \& Narayan, R.
2062: 2000, ATel, \#54
2063:
2064: \bibitem[Ho(1999)]{ho99}
2065: Ho, L. C. 1999, ApJ, 516, 672
2066:
2067: \bibitem[Ho(2002)]{ho02}
2068: Ho, L. C. 2002, ApJ, 564, 120
2069:
2070: \bibitem[Honma(1996)]{hon96}
2071: Honma, F. 1996, PASJ, 48, 77
2072:
2073: \bibitem[Hopkins et al.(2006a)]{hop06a}
2074: Hopkins, P. F., et al. 2006a, ApJS, 163, 1
2075:
2076: \bibitem[Hopkins et al.(2006b)]{hop06b}
2077: Hopkins, P. F., Narayan, R., \& Hernquist, L. 2006b, ApJ, 643, 641
2078:
2079: \bibitem[Hornstein et al.(2002)]{hor02}
2080: Hornstein, S. D., Ghez, A. M., Tanner, A., Morris, M., Becklin, E. E.,
2081: \& Wizinowich, P. 2002, ApJ, 577, 738
2082:
2083: \bibitem[Hynes et al.(2003)]{hyn03}
2084: Hynes, R. I., Charles, P. A., Casares, J., Haswell, C. A., Zurita, C.,
2085: \& Shahbaz, T. 2003, MNRAS, 340, 447
2086:
2087: \bibitem[Ichimaru(1977)]{ich77}
2088: Ichimaru, S. 1977, ApJ, 214, 840
2089:
2090: \bibitem[Igumenshchev(2004)]{igu04}
2091: Igumenshchev, I. V. 2004, Prog. Theor. Phys. Suppl., 155, 87
2092:
2093: \bibitem[Igumenshchev(2006)]{igu06}
2094: Igumenshchev, I. V. 2006, ApJ, 649, 361
2095:
2096: \bibitem[Igumenshchev \& Abramowicz(2000)]{ia00}
2097: Igumenshchev, I. V., \& Abramowicz, M. A. 2000, ApJS, 130, 463
2098:
2099: \bibitem[Igumenshchev et al.(2000)]{igu00}
2100: Igumenshchev, I. V., Abramowicz, M. A., \& Narayan, R. 2000,
2101: ApJ, 537, L27
2102:
2103: \bibitem[Igumenshchev \& Narayan(2002)]{igu02}
2104: Igumenshchev, I. V., \& Narayan, R. 2002, ApJ, 566, 137
2105:
2106: \bibitem[Igumenshchev et al.(2003)]{igu03}
2107: Igumenshchev, I. V., Narayan, R., \& Abramowicz, M. A. 2003,
2108: ApJ, 592, 2042
2109:
2110: \bibitem[Jonker et al.(2007)]{jon07}
2111: Jonker, P. G., Steeghs, D., Chakrabarty, D., \& Juett, A. M. 2007,
2112: ApJ, 665, L147
2113:
2114: \bibitem[Kato et al.(1998)]{kfm98}
2115: Kato, S., Fukue, J., \& Mineshige, S. 1998, Black Hole Accretion Disks
2116: (Kyoto Univ. Press)
2117:
2118: \bibitem[Kato et al.(1997)]{kat97}
2119: Kato, S., Yamasaki, T., Abramowicz, M. A., \& Chen, X. 1997, PASJ, 49, 221
2120:
2121: \bibitem[Lasota(1996)]{las96}
2122: Lasota, J. P. 1996, in Physics of Accretion Disks, ed. S. Kato,
2123: J. Fukue \& S. Mineshige (Gordon \& Breach) p85
2124:
2125: \bibitem[Lasota(1999a)]{las99a}
2126: Lasota, J. P. 1999a, Phys. Rep., 311, 247
2127:
2128: \bibitem[Lasota(1999b)]{las99b}
2129: Lasota, J. P. 1999b, Sci. Am., 280, 30
2130:
2131: \bibitem[Lasota(2000)]{las00}
2132: Lasota, J. P. 2000, A\&A, 360, 575
2133:
2134: \bibitem[Lasota(2001)]{las01}
2135: Lasota, J. P. 2001, in Black Hole Binaries and Galactic Nuclei,
2136: ed. L. Kaper, E. P. J. van den Heuvel \& P. A. Woudt, Springer-Verlag, p149
2137:
2138: \bibitem[Lasota(2007)]{las07}
2139: Lasota, J. P. 2007, Comptes Rendus Physique, 8, 45
2140:
2141: \bibitem[Lasota(2008)]{las08}
2142: Lasota, J. P. 2008, New Astron. Rev., ed. M. A. Abramowicz
2143: (astro-ph/0801.0490)
2144:
2145: \bibitem[Lasota et al.(1996a)]{las96a}
2146: Lasota, J. P., Abramowicz, M. A., Chen, X., Krolik, J., Narayan, R., \&
2147: Yi, I. 1996a, ApJ, 462, 142
2148:
2149: \bibitem[Lasota \& Hameury(1998)]{las98}
2150: Lasota, J. P., \& Hameury, J. P. 1998, AIP conf. Series, 431, 351
2151:
2152: \bibitem[Lasota et al.(1996b)]{las96b}
2153: Lasota, J. P., Narayan, R., \& Yi, I. 1996b, A\&A, 314, 813
2154:
2155: \bibitem[Li et al.(2005)]{li05}
2156: Li, L. X., Zimmerman, E. R., Narayan, R., \& McClintock, J. E.
2157: 2005, ApJS, 157, 335
2158:
2159: \bibitem[Liu et al.(2007)]{liu07}
2160: Liu, B. F., Taam, R. F., Meyer-Hofmeister, E., \& Meyer, F. 2007,
2161: ApJ, 671, 695
2162:
2163: \bibitem[Lobo(2006)]{lob06}
2164: Lobo, F. S. N. 2006, Class. Quantum Grav., 23, 1525
2165:
2166: \bibitem[Loeb et al.(2001)]{lnr01}
2167: Loeb, A., Narayan, R., \& Raymond, J. C. 2001, ApJ, 547, L151
2168:
2169: \bibitem[Loewenstein et al.(2001)]{loe01}
2170: Loewenstein, M., Mushotzky, R. F., Angelini, L., Arnaud, K., \&
2171: Quataert, E. 2001, ApJ, 555, L21
2172:
2173: \bibitem[Maccarone \& Coppi(2003)]{mac03}
2174: Maccarone, T. C., \& Coppi, P. S. 2003, MNRAS, 338, 189
2175:
2176: \bibitem[Machida et al.(2001)]{mch01}
2177: Machida, M., Matsumoto, R., \& Mineshige, S. 2001, PASJ, 53, L1
2178:
2179: \bibitem[Machida et al.(2004)]{mch04}
2180: Machida, M., Nakamura, K., \& Matsumoto, R. 2004, PASJ, 56, 671
2181:
2182: \bibitem[Machida et al.(2006)]{mch06}
2183: Machida, M., Nakamura, K., \& Matsumoto, R. 2006, PASJ, 58, 193
2184:
2185: \bibitem[Mahadevan(1997)]{mah97}
2186: Mahadevan, R. 1997, ApJ, 477, 585
2187:
2188: \bibitem[Mahadevan(1998)]{mah98}
2189: Mahadevan, R. 1998, Nature, 394, 651
2190:
2191: \bibitem[Mahadevan \& Quataert(1997)]{mq97}
2192: Mahadevan, R., \& Quataert, E. 1997, ApJ, 490, 605
2193:
2194: \bibitem[Malzac et al.(20040]{mal04}
2195: Malzac, J., Merloni, A., \& Fabian, A. C. 2004, MNRAS, 351, 253
2196:
2197: \bibitem[Manmoto(2000)]{man00}
2198: Manmoto, T. 2000, ApJ, 534, 734
2199:
2200: \bibitem[Manmoto et al.(1997)]{man97}
2201: Manmoto, T., Mineshige, S., \& Kusunose, M. 1997, ApJ, 489, 791
2202:
2203: \bibitem[Maraschi \& Tavecchio(2003)]{mar03}
2204: Maraschi, L., \& Tavecchio, F. 2003, ApJ, 593, 667
2205:
2206: \bibitem[Markoff et al.(2001)]{mar01}
2207: Markoff, S., Falcke, H., \& Fender, R. P. 2001, A\&A, 372, L25
2208:
2209: \bibitem[Markoff et al.(2005)]{mar05}
2210: Markoff, S., Nowak, M. A., \& Wilms, J. 2005, ApJ, 635, 1203
2211:
2212: \bibitem[Mayer \& Pringle(2007)]{may07}
2213: Mayer, M., \& Pringle, J. E. 2007, MNRAS, 376, 435
2214:
2215: \bibitem[Mazur \& Mottola(2001)]{maz01}
2216: Mazur, P. O., \& Mottola, E. 2001, gr-qc/0109035
2217:
2218: \bibitem[McClintock, J. E. et al.(2004)]{mcc04}
2219: McClintock, J. E., Narayan, R., \& Rybicki, G. B. 2004,
2220: ApJ, 615, 402
2221:
2222: \bibitem[McClintock \& Remillard(2007)]{mr07}
2223: McClintock, J. E., \& Remillard, R. A. 2006, in Black Hole Binaries,
2224: ed. W. Lewin \& M. van der Klis (Cambridge Univ. press) p157
2225:
2226: \bibitem[McClintock et al.(2006)]{mcc06}
2227: McClintock, J. E., Shafee, R., Narayan, R., Remillard, R. A., Davis,
2228: S. W., \& Li, L. X. 2006, ApJ, 652, 518
2229:
2230: \bibitem[McKinney \& Gammie(2004)]{mck04}
2231: McKinney, J. C., \& Gammie, C. F. 2004, ApJ, 611, 977
2232:
2233: \bibitem[McKinney(2005)]{mck05}
2234: McKinney, J. C. 2005, ApJ, 630, L5
2235:
2236: \bibitem[McKinney(2006)]{mck06}
2237: McKinney, J. C. 2006, MNRAS, 368, 1561
2238:
2239: \bibitem[Medvedev(2000)]{med00}
2240: Medvedev, M. V. 2000, ApJ, 541, 811
2241:
2242: \bibitem[Meier(2001)]{mei01}
2243: Meier, D. L. 2001, ApJ, 548, L9
2244:
2245: \bibitem[Menou et al.(1999a)]{men99a}
2246: Menou, K., Esin, A. A., Narayan, R., Garcia, M. R., Lasota, J. P.,
2247: \& McClintock, J. E. 1999a, ApJ, 520, 276
2248:
2249: \bibitem[Menou et al.(1999b)]{men99b}
2250: Menou, K., Narayan, R., \& Lasota, J. P. 1999b, ApJ, 513, 811
2251:
2252: \bibitem[Menou \& Quataert(2001)]{men01}
2253: Menou, K., \& Quataert, E. 2001, ApJ, 552, 204
2254:
2255: \bibitem[Merlone et al.(2003)]{mer03}
2256: Merloni, A., Heinz, S., \& Di Matteo, T. 2003, MNRAS, 345, 1057
2257:
2258: \bibitem[Meyer et al.(2000)]{MLM00}
2259: Meyer, F., Liu, B. F., \& Meyer-Hofmeister, E. 2000, A\&A, 361, 175
2260:
2261: \bibitem[Meyer \& Meyer-Hofmeister(1994)]{mm94}
2262: Meyer, F., \& Meyer-Hofmeister, E. 1994, A\&A, 361, 175
2263:
2264: \bibitem[Meyer-Hofmeister et al.(2005)]{MLM05}
2265: Meyer-Hofmeister, E., Liu, B. F., \& Meyer, F. 2005, A\&A, 432, 181
2266:
2267: \bibitem[Miller et al.(2006a)]{mil06a}
2268: Miller, J. M., Homan, J., \& Miniutti, G. 2006a, ApJ, 652, L113
2269:
2270: \bibitem[Miller et al.(2006b)]{mil06b}
2271: Miller, J. M., et al. 2006b, ApJ, 653, 525
2272:
2273: \bibitem[Miller et al.(2006c)]{mil06c}
2274: Miller, J. M., et al. 2006c, Nature, 441, 953
2275:
2276: \bibitem[Mitsuda et al.(1984)]{mit84}
2277: Mitsuda, K., et al. 1984, PASJ, 36, 741
2278:
2279: \bibitem[Miyamoto et al.(1995)]{miy95}
2280: Miyamoto, S., et al. 1995, ApJ, 442, L13
2281:
2282: \bibitem[Miyoshi et al.(1995)]{mi95}
2283: Miyoshi, M., et al. 1995, Nature, 373, 127
2284:
2285: \bibitem[Nagar et al.(2000)]{nag00}
2286: Nagar, N. M., Falcke, H., Wilson, A. S., \& Ho, L. C. 2000,
2287: ApJ, 542, 186
2288:
2289: \bibitem[Narayan(1996)]{nar96}
2290: Narayan, R. 1996, ApJ, 461, 136
2291:
2292: \bibitem[Narayan(2002)]{nar02}
2293: Narayan, R. 2002, in Lighthouses of the Universe, ed. M. Gilfanov,
2294: R. Sunyaev (Springer) p405
2295:
2296: \bibitem[Narayan(2005)]{nar05}
2297: Narayan, R. 2005, Astrophys. Sp. Sci., 300, 177
2298:
2299: \bibitem[Narayan et al.(1997a)]{nbm97a}
2300: Narayan, R., Barret, D., \& McClintock, J. E. 1997a, ApJ, 482, 448
2301:
2302: \bibitem[Narayan et al.(1997b)]{nar97b}
2303: Narayan, R., Garcia, M. R., \& McClintock, J. E. 1997b, ApJ, 478, L79
2304:
2305: \bibitem[Narayan et al.(2002a)]{ngm02a} Narayan, R., Garcia, M. R.,
2306: \& McClintock, J. E. 2002a, in The Ninth Marcel Grossmann Meeting,
2307: ed. V. G. Gurzadyan, R. T. Jantzen \& R. Ruffini, World Scientific, p405
2308:
2309: \bibitem[Narayan \& Heyl(2002)]{nh02}
2310: Narayan, R., \& Heyl, J. S. 2002, ApJ, 574, L139
2311:
2312: \bibitem[Narayan \& Heyl(2003)]{nh03}
2313: Narayan, R., \& Heyl, J. S. 2003, ApJ, 599, 419
2314:
2315: \bibitem[Narayan \& Igumenshchev(2000)]{ni00}
2316: Narayan, R., \& Igumenshchev, I. V. 2000, ApJ, 539, 798
2317:
2318: \bibitem[Narayan, Kato \& Honma(1997c)]{nkh97c}
2319: Narayan, R., Kato, S., \& Honma, F. 1997c, ApJ, 476, 49
2320:
2321: \bibitem[Narayan et al.(1998a)]{nar98a}
2322: Narayan, R., Mahadevan, R., Grindlay, J. E., Popham, R. G., \&
2323: Gammie, C. 1998a, ApJ, 492, 554
2324:
2325: \bibitem[Narayan et al.(1998b)]{nar98b} Narayan, R., Mahadevan, R., \&
2326: Quataert, E. 1998b, in Theory of Black Hole Accretion Disks,
2327: ed. M. A. Abramowicz, G. Bjornsson, \& J. E. Pringle (Cambridge
2328: Univ. Press) p148
2329:
2330: \bibitem[Narayan \& McClintock(2005)]{nmc05}
2331: Narayan, R., \& McClintock, J. E. 2005, ApJ, 623, 1017
2332:
2333: \bibitem[Narayan et al.(1996)]{nmy96}
2334: Narayan, R., McClintock, J. E., \& Yi, I. 1996, ApJ, 457, 821
2335:
2336: \bibitem[Narayan et al.(2002b)]{nqia02b}
2337: Narayan, R., Quataert, E., Igumenshchev, I. V., \& Abramowicz, M. A.
2338: 2002b, ApJ, 577, 295
2339:
2340: \bibitem[Narayan \& Yi(1994)]{NY94}
2341: Narayan, R., \& Yi, I. 1994, ApJ, 428, L13 (NY94)
2342:
2343: \bibitem[Narayan \& Yi(1995a)]{NY95a}
2344: Narayan, R., \& Yi, I. 1995a, ApJ, 444, 231 (NY95a)
2345:
2346: \bibitem[Narayan \& Yi(1995b)]{NY95b}
2347: Narayan, R., \& Yi, I. 1995b, ApJ, 452, 710 (NY95b)
2348:
2349: \bibitem[Narayan et al.(1995)]{nar95}
2350: Narayan, R., Yi, I., \& Mahadevan, R. 1995, Nature, 374, 623
2351:
2352: \bibitem[Nemmen et al.(2006)]{nem06}
2353: Nemmen, R. S., et al. 2006, ApJ, 643, 652
2354:
2355: \bibitem[Noble et al.(2007)]{non07}
2356: Noble, S. C., Leung, P. K., Gammie, C. F., \& Book, L. G.
2357: 2007, Class. Quant. Grav., 24, S259
2358:
2359: \bibitem[Novikov \& Thorne(1973)]{nt73}
2360: Novikov, I. D., \& Thorne, K. S. 1973, in Blackholes, ed. C. DeWitt,
2361: \& B. DeWitt (Gordon \& Breach) p343
2362:
2363: \bibitem[Orosz et al.(1997)]{oro97}
2364: Orosz, J. A., Remillard, R. A., Bailyn, C. D., \& McClitnock, J. E.
2365: 1997, ApJ, 478, L83
2366:
2367: \bibitem[Orosz et al.(2007)]{oro07}
2368: Orosz, J. A., et al. 2007, Nature, 449, 872
2369:
2370: \bibitem[Paczy\'nski(1998)]{pac98}
2371: Paczy\'nski, B. 1998, Acta Astronomica, 48, 667
2372:
2373: \bibitem[Pen et al.(2003)]{pen03}
2374: Pen, U. L., Matzner, C. D., \& Wong, S. 2003, ApJ, 596, L207
2375:
2376: \bibitem[Perna et al.(2003)]{per03}
2377: Perna, R., Narayan, R., Rybicki, G., Stella, L., \& Treves, A.
2378: 2003, ApJ, 594, 936
2379:
2380: \bibitem[Piran(1978)]{pir78}
2381: Piran, T. 1978, ApJ, 221, 652
2382:
2383: \bibitem[Popham \& Gammie(1998)]{pop98}
2384: Popham, R. G., \& Gammie, C. F. 1998, ApJ, 504, 419
2385:
2386: \bibitem[Popham \& Narayan(1991)]{pop91}
2387: Popham, R. G., \& Narayan, R. 1991, ApJ, 370, 604
2388:
2389: \bibitem[Quataert(2001)]{qua01}
2390: Quataert, E. 2001, in Probing the Physics of Active Galactic Nuclei by
2391: Multiwavelength Monitoring, ed. B. M. Peterson, R. S. Polidan \& R. W.
2392: Pogge (Astr. Soc. Pacific) p71
2393:
2394: \bibitem[Quataert(1998)]{qua98}
2395: Quataert, E. 1998, ApJ, 500, 978
2396:
2397: \bibitem[Quataert et al.(1999)]{q99}
2398: Quataert, E., Di Matteo, T., Narayan, R., \& Ho, L. C. 1999,
2399: ApJ, 525, L89
2400:
2401: \bibitem[Quataert \& Gruzinov(1999)]{qg99}
2402: Quataert, E., \& Gruzinov, A. 1999, ApJ, 520, 248
2403:
2404: \bibitem[Quataert \& Gruzinov(1999)]{qg00}
2405: Quataert, E., \& Gruzinov, A. 2000, ApJ, 539, 809
2406:
2407: \bibitem[Quataert \& Narayan(1999)]{qn99}
2408: Quataert, E., \& Narayan, R. 1999, ApJ, 520, 298
2409:
2410: \bibitem[Ramadevi \& Seetha(2007)]{rs07}
2411: Ramadevi, M. C., \& Seetha, S. 2007, MNRAS, 378, 182
2412:
2413: \bibitem[Rees et al.(1982)]{ree82}
2414: Rees, M. J., Begelman, M. C., Blandford, R. D., \& Phinney, E. S.
2415: 1982, Nature, 295, 17
2416:
2417: \bibitem[Remillard et al.(2006)]{rem06}
2418: Remillard, R. A., Lin, D., Cooper, R. L., \& Narayan, R. 2006,
2419: ApJ, 646, 407
2420:
2421: \bibitem[Reynolds et al.(1996)]{rey96}
2422: Reynolds, C. S., Di Matteo, T., Fabian, A. C., Hwang, U., \& Canizares,
2423: C. R. 1996, MNRAS, 283, L111
2424:
2425: \bibitem[Rozanska \& Czerny(2000)]{rc00}
2426: Rozanska, A., \& Czerny, B. 2000, A\&A, 360, 1170
2427:
2428: \bibitem[Rykoff et al.(2007)]{ryk07}
2429: Rykoff, E. S., Miller, J. M., Steeghs, D., \& Torres, M. A. P. 2007,
2430: ApJ, 666, 1129
2431:
2432: \bibitem[Sch\"odel et al.(2002)]{sch02}
2433: Sch\"odel, R., et al. 2002, Nature, 419, 694
2434:
2435: \bibitem[Serabyn et al.(1997)]{ser97}
2436: Serabyn, E., Carlstrom, J., Lay, O., Lis, D. C., Hunter, T. R.,
2437: \& Lacy, J. H. 1997, ApJ, 490, L77
2438:
2439: \bibitem[Shahbaz et al.(2005)]{sha05}
2440: Shahbaz, T., et al. 2005, MNRAS, 362, 975
2441:
2442: \bibitem[Shakura \& Sunyaev(1973)]{ss73}
2443: Shakura, N. I., \& Sunyaev, R. A. 1973, A\&A, 24, 337
2444:
2445: \bibitem[Shapiro et al.(1976)]{sle76}
2446: Shapiro, S. L., Lightman, A. P., \& Eardley, D. M. 1976, ApJ, 204, 187
2447:
2448: \bibitem[Shapiro \& Teukolsky(1983)]{sha83}
2449: Shapiro, S. L., \& Teukolsky, 1983, Black Holes, White Dwarfs, and
2450: Neutron Stars: The Physics of Compact Objects (Wiley-Interscience)
2451:
2452: \bibitem[Sharma et al.(2007)]{sha07}
2453: Sharma, P., Quataert, E., Hammett, G. W., \& Stone, J. M. 2007, ApJ,
2454: 667, 714
2455:
2456: \bibitem[Sharma et al.(2006)]{sha06}
2457: Sharma, P., Hammett, G. W., Quataert, E., \& Stone, J. M. 2006, ApJ,
2458: 637, 952
2459:
2460: \bibitem[Shen et al.(2005)]{she05}
2461: Shen, Z. Q., Lo, K. Y., Liang, M. C., Ho, P. T., \& Zhao, J. H.
2462: 2005, Nature, 438, 62
2463:
2464: \bibitem[Sikora et al.(2007)]{sik07}
2465: Sikora, M., Stawarz, L., \& Lasota, J. P. 2007, ApJ, 658, 815
2466:
2467: \bibitem[Smak(1999)]{sma99}
2468: Smak, J. 1999, Acta Astronomica, 49, 391
2469:
2470: \bibitem[Spruit \& Deufel(2002)]{sd02}
2471: Spruit, H. C., \& Deufel, B. 2002, A\&A, 387, 918
2472:
2473: \bibitem[Spruit et al.(1987)]{spr87}
2474: Spruit, H. C., Matsuda, T., Inoue, M., \& Sawada, K. 1987,
2475: MNRAS, 229, 517
2476:
2477: \bibitem[Stolovy et al.(2003)]{sto03}
2478: Stolovy, S., Melia, F., McCarthy, D., \& Yusef-Zadeh, F. 2003,
2479: Astron. Nachr., 324, 419
2480:
2481: \bibitem[Stone \& Pringle(2001)]{sto01}
2482: Stone, J. M., \& Pringle, J. E. 2001, MNRAS, 322, 461
2483:
2484: \bibitem[Stone et al.(1999)]{sto99}
2485: Stone, J. M., Pringle, J. E., \& Begelman, M. C. 1999, MNRAS, 310, 1002
2486:
2487: \bibitem[Sunyaev \& Titarchuk(1980)]{st80}
2488: Sunyaev, R. A., \& Titarchuk, L. G. 1980, A\&A, 86, 121
2489:
2490: %\bibitem[Tomsick(2006)]{tom06}
2491: %Tomsick, J. A. 2006, Adv. Sp. Res., 38, 2805
2492:
2493: \bibitem[Tournear et al.(2003)]{tou03}
2494: Tournear, D., et al. 2003, ApJ, 595, 1058
2495:
2496: \bibitem[Ulvestad \& Ho(2001)]{ulv01}
2497: Ulvestad, J. S., \& Ho, L. C. 2001, ApJ, 562, L133
2498:
2499: \bibitem[Visser \& Wiltshire(2004)]{vis04}
2500: Visser, M., \& Wiltshire, D. L. 2004, Class. Quantum Grav., 21, 1135
2501:
2502: \bibitem[Wu(1997)]{wu97}
2503: Wu, X. 1997, MNRAS, 292, 113
2504:
2505: \bibitem[Wu \& Li(1996)]{wu96}
2506: Wu, X., \& Li, Q. 1996, ApJ, 469, 776
2507:
2508: \bibitem[Wu et al.(2007)]{wu07}
2509: Wu, Q., Yuan, F., \& Cao, X. 2007, ApJ, 669, 96
2510:
2511: \bibitem[Yuan(2001)]{yua01}
2512: Yuan, F. 2001, MNRAS, 324, 119
2513:
2514: \bibitem[Yuan(2003)]{yua03}
2515: Yuan, F. 2003, ApJ, 594, L99
2516:
2517: \bibitem[Yuan \& Cui(2005)]{yua05}
2518: Yuan, F., \& Cui, W. 2005, ApJ, 629, 408
2519:
2520: \bibitem[Yuan et al.(2005)]{ycn05}
2521: Yuan, F., Cui, W., \& Narayan, R. 2005, ApJ, 620, 905
2522:
2523: \bibitem[Yuan et al.(2008)]{yua08}
2524: Yuan, F., Ma, R., \& Narayan, R. 2008, ApJ, in press
2525: (astro-ph/08021679)
2526:
2527: \bibitem[Yuan \& Narayan(2004)]{yn04}
2528: Yuan, F., \& Narayan, R. 2004, ApJ, 612, 724
2529:
2530: \bibitem[Yuan et al.(2003)]{yqn03}
2531: Yuan, F., Quataert, E., \& Narayan, R. 2003, ApJ, 598, 301
2532:
2533: \bibitem[Yuan \& Zdziarski(2004)]{yua04}
2534: Yuan, F., \& Zdziarski, A. 2004, MNRAS, 354, 953
2535:
2536: \bibitem[Yuan et al.(2007)]{yua07}
2537: Yuan, F., Zdziarski, A., Xue, Y., \& Wu, X. B. 2007, ApJ, 659, 541
2538:
2539: \bibitem[Yungelson et al.(2006)]{yun06}
2540: Yungelson, L. R., et al. 2006, A\&A, 454, 559
2541:
2542: \bibitem[Zdziarski \& Gierlinski(2004)]{zg04}
2543: Zdziarski, A. A., \& Gierlinski, M. 2004, Prog. Theor. Phys. Suppl.,
2544: 155, 99
2545:
2546: \bibitem[Zdziarski et al.(1999)]{zdz99}
2547: Zdziarski, A. A., Lubinski, P., \& Smith, D. A. 1999, MNRAS, 303, L11
2548:
2549: \bibitem[Zdziarski et al.(2003)]{zdz03}
2550: Zdziarski, A. A., Lubinski, P., Gilfanov, M., \& Revnivtsev, M. 2003,
2551: MNRAS, 342, 355
2552:
2553: \bibitem[Zdziarski et al.(2004)]{zdz04}
2554: Zdziarski, A. A., et al. 2004, MNRAS, 351, 791
2555:
2556: \bibitem[Zdziarski et al.(1996)]{zdz96}
2557: Zdziarski, A. A., Johnson, W. N., \& Magdiarz, P. 1996, MNRAS, 283, 193
2558:
2559: \bibitem[Zdziarski et al.(1998)]{zdz98}
2560: Zdziarski, A. A., Poutanen, J., Mikolajewska, J., Gierlinski, M., Ebisawa,
2561: K., \& Johnson, W. L. 1998, MNRAS, 301, 435
2562:
2563: \bibitem[Zhao et al.(2003)]{zha03}
2564: Zhao, J. H., et al. 2003, ApJ, 586, L29
2565:
2566: \bibitem[Zimmerman et al.(2005)]{Zim05}
2567: Zimmerman, E. R., Narayan, R., McClintock, J. E., \& Miller, J. M.
2568: 2005, ApJ, 618, 832
2569:
2570: \end{thebibliography}
2571:
2572: \end{document}
2573:
2574: