astro-ph0411801/ms.tex
1: %
2: % Le & Becker
3: %
4: 
5: % aastex produces a one-column, double-spaced document:
6: 
7: % \documentclass{aastex}
8: 
9: % preprint produces a one-column, single-spaced document:
10: 
11:  \documentclass[12pt,preprint]{aastex}
12: 
13: % preprint2 produces a double-column, single-spaced document:
14: 
15: % \documentclass[preprint2]{aastex}
16: 
17: % emulateapj emulates a double-column, single-spaced ApJ paper:
18: 
19: % \documentclass{emulateapj}
20: 
21:  \usepackage{epsfig}
22: 
23: % \eqsecnum
24: 
25: % \newcommand{\apj}{ApJ}
26: 
27: \newcommand{\vdag}{(v)^\dagger}
28: \newcommand{\myemail}{skywalker@galaxy.far.far.away}
29: 
30: \newcommand{\kinenergy}{\varepsilon}
31: \newcommand{\rs}{r_{_{\rm S}}}
32: \newcommand{\rg}{r_{_{\rm g}}}
33: \newcommand{\enrate}{\dot E}
34: \newcommand{\momrate}{\dot J}
35: \newcommand{\rc}{r_c}
36: \newcommand{\vff}{v_{\rm ff}}
37: \newcommand{\asym}{\doteq}
38: \newcommand{\ellK}{\ell_{\rm K}}
39: \newcommand{\OmegaK}{\Omega_{\rm K}}
40: \newcommand{\vc}{v_c}
41: \newcommand{\gapprox}{\lower.4ex\hbox{$\;\buildrel >\over{\scriptstyle\sim}\;$}}
42: \newcommand{\lapprox}{\lower.4ex\hbox{$\;\buildrel <\over{\scriptstyle\sim}\;$}}
43: \newcommand{\begeq}{\begin{equation}}
44: \newcommand{\fineq}{\end{equation}}
45: \newcommand{\begeqarray}{\begin{eqnarray}}
46: \newcommand{\fineqarray}{\end{eqnarray}}
47: \newcommand{\green}{f_{_{\rm G}}}
48: \newcommand{\Qjet}{L_{\rm jet}}
49: \newcommand{\rjet}{r_{\rm jet}}
50: \newcommand{\Qtheory}{L}
51: \newcommand{\SgrA}{Sgr~A$^*$}
52: 
53: \def\vr{v}
54: \def\ellprime0{\ell'_0}
55: 
56: %\voffset=0.5truein
57: 
58: %\slugcomment{accepted by ApJ Letters}
59: 
60: \shorttitle{Shocks in ADAF Disks}
61: 
62: \shortauthors{Le \& Becker}
63: 
64: \begin{document}
65: 
66: \title{A Self-Consistent Model for the Formation of
67: Relativistic Outflows in Advection-Dominated Accretion
68: Disks with Shocks}
69: 
70: \author{Truong Le\altaffilmark{1} and
71: Peter A. Becker\altaffilmark{2}$^,$\altaffilmark{3}}
72: 
73: \affil{Center for Earth Observing and Space Research, George Mason
74: University, Fairfax, VA 22030-4444, USA}
75: 
76: \altaffiltext{1}{tler@gmu.edu}
77: \altaffiltext{2}{pbecker@gmu.edu}
78: \altaffiltext{3}{also Department of Physics and Astronomy,
79: George Mason University, Fairfax, VA 22030-4444, USA}
80: 
81: \begin{abstract}
82: In this Letter, we suggest that the relativistic protons powering the
83: outflows emanating from radio-loud systems containing black holes are
84: accelerated at standing, centrifugally-supported shocks in hot,
85: advection-dominated accretion disks. Such disks are ideal sites for
86: first-order Fermi acceleration at shocks because the gas is tenuous, and
87: consequently the mean free path for particle-particle collisions
88: generally exceeds the thickness of the disk. The accelerated particles
89: are therefore able to avoid thermalization, and as a result a small
90: fraction of them achieve very high energies and escape from the disk. In
91: our approach the hydrodynamics and the particle acceleration are coupled
92: and the solutions are obtained self-consistently based on a rigorous
93: mathematical treatment. The theoretical analysis of the particle
94: transport parallels the early studies of cosmic-ray acceleration in
95: supernova shock waves. We find that particle acceleration in the
96: vicinity of the shock can extract enough energy to power a relativistic
97: jet. Using physical parameters appropriate for M87 and \SgrA, we confirm
98: that the jet kinetic luminosities predicted by the theory agree with the
99: observational estimates.
100: 
101: \end{abstract}
102: 
103: % The different journals have different requirements for keywords.  The
104: % keywords.apj file, found on aas.org in the pubs/aastex-misc directory,
105: % contains a list of keywords used with the ApJ and Letters.  These are
106: % usually assigned by the editor, but authors may include them in their
107: % manuscripts if they wish.
108: 
109: \keywords{accretion, accretion disks --- hydrodynamics --- black hole
110: physics --- galaxies: jets}
111: 
112: \section{INTRODUCTION}
113: 
114: Radio-loud active galactic nuclei (AGNs) are thought to contain supermassive
115: central black holes surrounded by hot, two-temperature,
116: advection-dominated accretion flows (ADAFs) with significantly
117: sub-Eddington accretion rates. In these disks, the ion temperature $T_i
118: \sim 10^{12}\,$K greatly exceeds the electron temperature $T_e \sim
119: 10^{10}\,$K (e.g., Narayan, Kato, \& Honma 1997; Becker \& Le 2003). The
120: observed correlation between high radio luminosities and the presence of
121: outflows suggests that the hot ADAF disks are efficient sites for the
122: acceleration of the relativistic particles powering the jets. This
123: motivated Blandford \& Begelman (1999) to consider the possibility of
124: strong outflows of matter and energy from such systems, under the
125: assumption of self-similarity and Newtonian gravity. The same scenario
126: was extended to the general relativistic case by Becker, Subramanian, \&
127: Kazanas (2001). However, a single, global, self-consistent model for the
128: structure of the disk/jet that includes a specific microphysical
129: acceleration mechanism has not yet appeared in the literature. In
130: this Letter, we explore the consequences of the presence of a
131: shock in an ADAF disk for the acceleration of the observed nonthermal jet
132: particles.
133: 
134: It is well known that for certain sets of outer boundary conditions
135: describing the specific angular momentum and the specific energy of the
136: gas supplied to the accretion flow, both shocked and shock-free
137: solutions for the global disk structure are possible, although there is
138: an upper limit on the amount of viscosity beyond which no shocks can
139: form. However, it is clear that shocks can exist in both dissipative and
140: inviscid flows (e.g., Chakrabarti \& Das 2004). When both shocked and
141: shock-free solutions are possible for the same set of upstream
142: conditions, one must incorporate additional physical principles in order
143: to determine which flow solution actually occurs. Since in general the
144: shock solution possesses a higher entropy content than the shock-free
145: solution, we argue based on the second law of thermodynamics that shocks
146: are preferred (see Becker \& Kazanas 2001). Particle acceleration has
147: not been studied in shocked disks in either the viscid or the inviscid
148: cases, and therefore we concentrate here on the simpler situation of
149: inviscid (adiabatic) flow. In this Letter we focus on disks containing
150: isothermal shocks because the jump in the energy flux at such a shock
151: can be identified with the energy carried away from the disk by the
152: relativistic particles in the outflow.
153: 
154: \section{DYNAMICAL MODEL}
155: 
156: The dynamical structure of the accretion flows considered here follows
157: the model studied by Chakrabarti (1989a) and Abramowicz \& Chakrabarti
158: (1990), which comprises an inviscid disk that either contains an
159: isothermal shock, or else is shock-free. The model is based on the
160: standard one-dimensional conservation equations for advection-dominated
161: accretion flows, including utilization of the pseudo-Newtonian
162: gravitational potential to simulate the effects of general relativity
163: (Paczy\'nski \& Wiita 1980). In general, the flow is assumed to be
164: adiabatic and in vertical hydrostatic equilibrium, although the entropy
165: can jump at the shock. Under the inviscid flow assumption, the
166: disk/shock model depends on two parameters, namely the energy transport
167: rate per unit mass, $\epsilon$, and the specific angular momentum,
168: $\ell$. The value of $\epsilon$ is constant in ADAF type flows since
169: there are no radiative losses, although it will jump at the location of
170: the isothermal shock if there is one in the disk. On the other hand, the
171: value of $\ell$ remains constant since the disk is inviscid. The energy
172: transport rate per unit mass can be written as
173: %
174: \begeq
175: \epsilon = {1 \over 2} \, {\ell^2 \over r^2} + {1 \over 2} \, v^2
176: + {1 \over \gamma-1} \, a^2 - {GM \over r-\rs} \ ,
177: \label{eq1}
178: \fineq
179: %
180: where $v$ denotes the radial velocity (defined to be positive for
181: inflow), $\rs=2GM/c^2$, and the adiabatic sound speed is given by $a =
182: (\gamma P/\rho) ^{1/2}$ for a gas with pressure $P$, mass density
183: $\rho$, and specific heat ratio $\gamma$. Note that $\epsilon$ is
184: defined to be positive for an inward flow of energy into the black hole.
185: Furthermore, in this semi-classical approach, $\epsilon$ does not
186: include the rest-mass contribution to the energy.
187: 
188: Dissipation and radiative losses are unimportant in inviscid,
189: advection-dominated disks, and therefore the flow is in general
190: adiabatic and isentropic, except at the shock location. In this
191: situation, the ``entropy parameter,''
192: %
193: \begeq
194: K \equiv v \, a^{(\gamma+1)(\gamma-1)} \, r^{3/2} \, (r-\rs) \ ,
195: \label{eq2}
196: \fineq
197: %
198: is conserved throughout the flow except at the location of a shock
199: (Becker \& Le 2003). When a shock is present, we shall use the
200: subscripts ``-'' and ``+'' to refer to quantities measured just upstream
201: and just downstream from the shock, respectively. In the isothermal
202: shock model, $K$ and $\epsilon$ have smaller values in the post-shock
203: region ($K_{+}$, $\epsilon_{+}$) compared with the pre-shock region
204: ($K_{-}$, $\epsilon_{-}$). The entropy and energy lost from the disk are
205: carried off by the outflowing high-energy particles, and the total
206: entropy of the combined (disk/outflow) system is higher when a shock is
207: present, as expected according to the second law of thermodynamics. By
208: combining equations~(\ref{eq1}) and (\ref{eq2}) and differentiating with
209: respect to radius, we can obtain a ``wind equation'' of the form $(1/v)
210: \, (dv/dr) = N / D$, where
211: %
212: \begeq
213: N \equiv - \, {\ell^2 \over r^3} + {GM \over (r-\rs)^2}
214: + {a^2 \, (3 \rs - 5 r) \over (\gamma+1) \, r \, (r-\rs)} \ , \ \ \ \ \
215: D \equiv {2 \, a^2 \over \gamma+1} - v^2
216: \ .
217: \label{eq2b}
218: \fineq
219: %
220: Critical points occur where the numerator $N$ and the denominator $D$
221: vanish simultaneously (see Figure~1{\it b} for a specific example). The
222: associated critical conditions can be combined with the energy
223: equation~(\ref{eq1}) to express the critical velocity $v_c$, the
224: critical sound speed $a_c$, and the critical radius $r_c$ as functions
225: of $\epsilon$ and $\ell$. The value of $K$ is then obtained by
226: substituting $r_c$, $v_c$, and $a_c$ into equation~(\ref{eq2}). In
227: general, one obtains four solutions for the critical radius, denoted by
228: $r_{c1}$, $r_{c2}$, $r_{c3}$, and $r_{c4}$. Only $r_{c1}$ and $r_{c3}$
229: are physically acceptable sonic points, although the types of accretion
230: flows that can pass through them are different. Specifically, $r_{c3}$
231: is an X-type critical point, and therefore a smooth, global, shock-free
232: solution always exists in which the flow is transonic at $r_{c3}$ and
233: then remains supersonic all the way to the event horizon. On the other
234: hand, $r_{c1}$ is an $\alpha$-type critical point, and therefore any
235: accretion flow originating at a large distance that passes through this
236: point must display a shock transition below $r_{c1}$ in order to cross
237: the event horizon (Abramowicz \& Chakrabarti 1990).
238: 
239: The radius of the isothermal shock, denoted by $r_*$, is determined
240: self-consistently along with the structure of the disk by satisfying the
241: velocity and energy jump conditions (Chakrabarti 1989a)
242: %
243: \begeq
244: R_*^{-1} \equiv {v_+ \over v_-} = {1 \over \gamma \, {\cal M}_-^2}
245: \ , \ \ \ \ \
246: \Delta\epsilon \equiv \epsilon_+-\epsilon_- = {v_+^2 - v_-^2 \over 2}
247: \ ,
248: \label{eq2d}
249: \fineq
250: %
251: where ${\cal M}_- \equiv v_-/a_-$ is the upstream Mach number at the
252: shock location and $R_*$ is the shock compression ratio. Due to the
253: escape of energy at the isothermal shock, the energy transport rate
254: $\epsilon$ drops from the upstream value $\epsilon_-$ to the downstream
255: value $\epsilon_+$, and consequently $\Delta \epsilon < 0$. The flow
256: must therefore pass through a new inner critical point located at $\hat
257: r_{c3} < r_*$, which is computed using the downstream value
258: $\epsilon_+$. This structure represents the transonic solution of
259: interest here. The kinetic power lost from the disk at the isothermal
260: shock location is related to observable parameters by writing
261: %
262: \begeq
263: \Qjet \equiv \dot M c^2 \triangle \epsilon \ ,
264: \label{eq3}
265: \fineq
266: %
267: where $\Qjet$ and $\dot M$ are the observed kinetic power and the
268: accretion rate for a specific source, respectively. Hence we conclude
269: that $\triangle \epsilon = \Qjet/(\dot M c^2)$. For given observational
270: values of $M$, $\dot M$, and $\Qjet$, only certain combinations of the
271: specific angular momentum $\ell$ and the upstream specific energy
272: transport rate $\epsilon_{-}$ will yield the correct value for
273: $\triangle \epsilon$. Hence we can view $\epsilon_{-}$ as the
274: fundamental free parameter and determine $\ell$ using the energy
275: constraint $\triangle \epsilon=\Qjet/(\dot M c^2)$ once $M$, $\dot M$,
276: and $\Qjet$ are specified for a particular source. The two sources of
277: interest here are M87 and \SgrA, which correspond to models~A and B,
278: respectively.
279: 
280: \begin{figure}
281: \begin{center}
282: \epsfig{file=f1.eps,height=8.0cm}
283: \end{center}
284: \caption{Results obtained using the M87 values for $M$, $\dot M$, and
285: $\Qjet$. In panel ({\it a}) we plot $\kappa_0$ ({\it dashed line}) and
286: $\Gamma_\infty$ ({\it solid line}) as functions of the upstream energy
287: transport rate $\epsilon_-$ such that the energy conservation condition
288: $\dot N_{\rm esc} E_{\rm esc} = \Qjet$ is satisfied. In panel ({\it b})
289: the flow velocity $v(r)$ ({\it solids lines}) and sound speed $a(r)
290: [2/(\gamma+1)]^{1/2}$ ({\it dashed lines}) are plotted as functions of
291: the radius for the model~A parameters, which are indicated by the
292: letter ``A'' in panel ({\it a}). In panel ({\it b}), the thin lines
293: represent the shocked solutions and the heavy lines denote the
294: shock-free solutions. See the discussion in the text.}
295: \end{figure}
296: 
297: \section{PARTICLE TRANSPORT EQUATION}
298: 
299: The model for the diffusive particle transport employed here is similar
300: to the scenario for the acceleration of cosmic rays first suggested by
301: Blandford \& Ostriker (1978), although in our situation the shock
302: maintains a fixed location in the disk and the seed particles for the
303: acceleration process are provided by the accretion flow itself. The
304: energy of the injected seed particles is given by $E_0 = 0.002\,$erg,
305: which corresponds to an injected Lorentz factor $\Gamma_0 \equiv
306: E_0/(m_p \, c^2) \sim 1.3$, where $m_p$ is the proton mass. Starting
307: with values for $\epsilon_-$ and $\ell$ associated with a particular
308: source, we can compute the injection rate for the seed particles, $\dot
309: N_0$, by setting $\dot N_0 E_0 = \dot M c^2 \triangle \epsilon = \Qjet$
310: (see eq.~[\ref{eq3}]), which ensures that the energy injection rate for
311: the relativistic seed particles is equal to the energy loss rate for the
312: background gas at the isothermal shock location. The particle injection
313: rate is therefore given by $\dot N_0 = \Qjet / E_0$, which represents
314: one of the fundamental self-consistency requirements for the model.
315: 
316: The particle Green's function $\green(\epsilon,r)$ describes the energy
317: and spatial distribution of the relativistic particles injected with energy
318: $E_0$ at the shock location $r_*$. The associated number and energy densities
319: of the relativistic particles are given by $n_r=\int_0^\infty 4 \pi E^2 \,
320: \green \, dE$ and $U_r=\int_0^\infty 4 \pi E^3 \, \green \, dE$, respectively.
321: Including the escape of the relativistic particles at the shock location,
322: the energy moments $I_n \equiv \int_0^\infty 4 \pi E^n \, \green \, dE$
323: of the Green's function satisfy the one-dimensional steady-state transport
324: equation (Le \& Becker 2004)
325: %
326: \begeq
327: {d \over dr}\left(4 \pi r H F_n \right)
328: = 4 \pi r H \left[{(2-n)v\over 3} \, {dI_n \over dr}
329: + {\dot N_0 E_0^{n-2} \over 4 \pi r_* H_*} \, \delta(r-r_*)
330: -c A_0 I_n \, \delta(r-r_*)\right]
331: \ ,
332: \label{eq4}
333: \fineq
334: %
335: where $c$ is the speed of light, $H_*$ is the disk half-thickness
336: at the shock location, and the flux $F_n$ is defined by
337: %
338: \begeq
339: F_N \equiv - \, {(n+1)v \over 3} \, I_n - \kappa \, {dI_n \over dr}
340: \ ,
341: \label{eq4b}
342: \fineq
343: %
344: which is obtained by integrating over the vertical structure of the
345: disk. The terms on the right-hand side of equation~(\ref{eq4}) represent
346: first-order Fermi acceleration, the particle source, and the escape of
347: particles from the disk at the shock location, respectively. The terms
348: on the right-hand side of equation~(\ref{eq4b}) describe the diffusion
349: and advection of particles, respectively, and $\kappa$ denotes the
350: spatial diffusion coefficient. The dimensionless parameter $A_0$ is a
351: constant that represents the microphysical processes controlling the
352: vertical escape of energetic particles from the disk in the vicinity of
353: the shock, as discussed below.
354: 
355: To close the system, we must also specify the radial variation of the
356: spatial diffusion coefficient $\kappa$ in equation~(\ref{eq4b}). Since
357: the radial dependence of $\kappa$ is not known precisely in this
358: physical situation, we shall adopt the general form
359: %
360: \begeq
361: \kappa(r) = \kappa_0 \, v(r) \, \rs \left({r \over \rs} - 1\right)^\alpha
362: \ ,
363: \label{eq5}
364: \fineq
365: %
366: where $\kappa_0$ is a dimensionless positive constant. Close to the
367: event horizon, inward advection at the speed of light must dominate over
368: outward diffusion. Conversely, in the outer region, we expect that
369: diffusion will dominate over advection as $r \to \infty$. A careful
370: analysis of the differential equation for the relativistic particle
371: number density $n_r$ obtained by setting $n=2$ in equation~(\ref{eq4})
372: reveals that these conditions are both satisfied if $\alpha > 1$, and in
373: our work we shall set $\alpha=2$.
374: 
375: The spatial diffusion coefficient $\kappa$ is related to the magnetic
376: mean free path via the usual expression $\kappa = c \, \lambda_{\rm mag}
377: / 3$. Analysis of the three-dimensional random walk of the escaping
378: particles then yields
379: %
380: \begeq
381: A_0 = \left(3 \, \kappa_* \over c \, H_*\right)^2 \ ,
382: \label{eq6}
383: \fineq
384: %
385: where $\kappa_* \equiv (\kappa_- + \kappa_+)/2$. We need to specify the
386: value of $\kappa_0$ in order to compute $\kappa_-$ and $\kappa_+$ using
387: equation~(\ref{eq5}). Since $A_0$ is independent of the particle energy,
388: it follows that all of the relativistic particles have the same escape
389: probability per unit time in the vicinity of the shock, and therefore
390: the mean energy of the escaping particles is given by $E_{\rm esc} =
391: U_*/n_*$, where $n_* \equiv n_r(r_*)$ and $U_* \equiv U_r(r_*)$ denote
392: the relativistic particle number and energy densities at the shock
393: location. Note that $E_{\rm esc}$ is proportional to $E_0$ and
394: independent of $\dot N_0$. To obtain a self-consistent result, the
395: product of the calculated escape energy $E_{\rm esc}$ and the particle
396: escape rate $\dot N_{\rm esc}$ at the shock location must be equal to
397: the total energy lost from the background in the dynamical model, and
398: therefore we set $\dot N_{\rm esc} E_{\rm esc} = \dot M c^2 \triangle
399: \epsilon = \Qjet$. This condition is not automatically satisfied in
400: general and therefore it constrains the free parameters $\epsilon_-$ and
401: $\kappa_0$ as explained below.
402: 
403: \section{APPLICATIONS TO M87 AND \SgrA}
404: 
405: In our numerical examples, the disk structures for M87 and \SgrA are
406: determined based on observational estimates for the black hole mass $M$,
407: the mass accretion rate $\dot M$, and the jet kinetic power $\Qjet$.
408: From observations of M87, we obtain $M \sim 3 \times 10^9 {\rm \
409: M_{\odot}}$ (e.g., Ford et al. 1994), $\dot M \sim 1.3 \times 10^{-1}
410: {\rm \ M_{\odot} \ yr^{-1}}$ (Reynolds et al. 1996), and $\Qjet \sim
411: 10^{43-44}\,{\rm erg \ s^{-1}}$ (Reynolds et al. 1996; Bicknell \&
412: Begelman 1996; Owen, Eilek, \& Kassim 2000). We shall adopt the average
413: value $\Qjet$ = $5.5 \times 10^{43}\,{\rm erg \ s^{-1}}$. For \SgrA,
414: observations indicate that $M \sim 2.6 \times 10^6 {\rm \ M_{\odot}}$
415: (e.g., Schodel et al. 2002) and $\dot M \sim 8.8 \times 10^{-7} {\rm \
416: M_{\odot} \ yr^{-1}}$ (e.g., Yuan, Markoff, \& Falcke 2002; Quataert
417: 2003). The kinetic luminosity of the jet in \SgrA is rather uncertain
418: (see, e.g., Yuan 2000; Yuan et al. 2002). In our work, we will adopt the
419: value quoted by Falcke \& Biermann (1999), $\Qjet \sim 5 \times 10^{38}
420: {\rm erg \ s^{-1}}$.
421: 
422: Our goal here is to compute self-consistently the mean energy of the
423: escaping particles, $E_{\rm esc}$, while matching the observational
424: values of $M$, $\dot M$, and $\Qjet$ for a specific source. We shall
425: utilize natural gravitational units, with $GM=c=1$ and $\rs=2$, except
426: as noted. Following Narayan, Kato, \& Honma (1997), we assume
427: approximate equipartition between the gas and magnetic pressures, and
428: therefore we set $\gamma=1.5$. Once the values of $M$, $\dot M$, and
429: $\Qjet$ have been specified for the source, we select a value for the
430: fundamental free parameter $\epsilon_{-}$ and then compute $\ell$ based
431: on the energy conservation condition $\triangle \epsilon=\Qjet/(\dot M
432: c^2)$ (see \S~2). The velocity profile $v(r)$ is determined by
433: integrating numerically the wind equation $(1/v) \, (dv/dr) = N / D$,
434: where $N$ and $D$ are given by equation~(\ref{eq2b}). The associated
435: solution for the isothermal sound speed $a(r)$ can then be computed
436: using equation~(\ref{eq2}) based on the fact that $K$ is constant in the
437: adiabatic flow (except at the shock if there is one).
438: 
439: Once the velocity profile $v(r)$ has been obtained, the transport
440: equation~(\ref{eq4}) can be solved numerically to determine the number
441: and energy density distributions for the relativistic particles in the
442: disk, $n_r(r) = I_2(r)$ and $U_r(r) = I_3(r)$. The energy conservation
443: requirement $\dot N_{\rm esc} E_{\rm esc} = \Qjet$ then allows us to
444: solve for the corresponding value of the diffusion coefficient constant
445: $\kappa_0$. Based on the observational values for $M$, $\dot M$, and
446: $\Qjet$ associated with M87, in Figure~1{\it a} we plot the results
447: obtained for $\kappa_0$ as a function of $\epsilon_-$. In general, one
448: obtains two roots for $\kappa_0$ if $\epsilon_- < \epsilon_{\rm max}$,
449: where $\epsilon_{\rm max} = 0.001527$ for the M87 parameters and
450: $\epsilon_{\rm max} = 0.001229$ for the \SgrA parameters. No
451: self-consistent solutions exist if $\epsilon_- > \epsilon_{\rm max}$,
452: and a single root for $\kappa_0$ is obtained if $\epsilon_- =
453: \epsilon_{\rm max}$ (see Fig.~1{\it a}). For illustrative purposes, we
454: shall develop two detailed models by setting $\epsilon_- = \epsilon_{\rm
455: max}$ for M87 and \SgrA.
456: 
457: The parameter values for the two models are $\epsilon_- = 0.001527$,
458: $\ell=3.1340$, $\kappa_0=0.02044$, $\dot N_0 = 2.75 \times 10^{46} \,
459: {\rm s}^{-1}$, $\kappa_* = 0.427877$, $A_0 = 0.0124$, $n_* = 1.19 \times
460: 10^{44}\, {\rm cm}^{-3}$, $U_* = 1.42 \times 10^{42} \, {\rm erg \
461: cm}^{-3}$, $\dot N_{\rm esc} = 4.61 \times 10^{45} \, {\rm
462: s}^{-1}$,$E_{\rm esc} = 0.0119\,$erg, $r_* = 21.654$, $\epsilon_{+} =
463: -0.005746$, $r_{c1} = 98.524$, $r_{c3} = 5.379$, $\hat r_{c3} = 5.659$,
464: ${\cal M}_- = 1.125$, $R_* = 1.897$, and $H_* = 11.544$ for model~A
465: (M87). For model~B (\SgrA), we have $\epsilon_- = 0.001229$,
466: $\ell=3.1524$, $\kappa_0=0.02819$, $\dot N_0 = 2.51 \times 10^{41}\,{\rm
467: s}^{-1}$, $\kappa_* = 0.321414$, $A_0 = 0.0158$, $n_* = 1.93 \times
468: 10^{39}\, {\rm cm}^{-3}$, $U_* = 2.10 \times 10^{37}\,{\rm erg \
469: cm}^{-3}$, $\dot N_{\rm esc} = 4.56 \times 10^{40}\,{\rm s}^{-1}$,
470: $E_{\rm esc} = 0.0109\,$erg, $r_* = 15.583$, $\epsilon_{+} = -0.008749$,
471: $r_{c1} = 131.874$, $r_{c3} = 5.329$, $\hat r_{c3} = 5.723$, ${\cal M}_-
472: = 1.146$, $R_* = 1.970$, and $H_* = 7.672$. The corresponding dynamical
473: solutions for $v(r)$ and $a(r)$ (with and without a shock) are plotted
474: in Figure~1{\it b} for model~A.
475: 
476: The terminal (asymptotic) Lorentz factor of the jet can be computed
477: using $\Gamma_\infty = E_{\rm esc}/(m_p c^2)$. In Figure~1{\it a} we
478: plot $\Gamma_\infty$ as a function of $\epsilon_-$ for the M87
479: parameters. Since in general we obtain two roots for $\kappa_0$ for each
480: value of $\epsilon_-$, we also obtain two $\Gamma_\infty$ values, with
481: the larger one corresponding to the smaller $\kappa_0$ root. For
482: models~A (M87) and B (\SgrA), we obtain $\Gamma_\infty=7.92$ and
483: $\Gamma_\infty=7.26$, respectively. The associated values for the mean
484: energy boost ratio, $E_{\rm esc}/E_0$, are given by $E_{\rm esc}/E_0 =
485: 5.95$ and $E_{\rm esc}/E_0 = 5.45$, respectively. Conversely, we find
486: that in the shock-free models with the same values for $\epsilon_-$,
487: $\ell$, and $\kappa_0$, the energy is boosted by only a factor of $\sim
488: 1.4 - 1.5$. This clearly establishes the essential role of the shock in
489: efficiently accelerating particles up to very high energies, well above
490: the energy required to escape from the disk. The self-consistency
491: conditions $\dot{N}_0 E_0 = \Qjet$ and $\dot N_{\rm esc} E_{\rm esc} =
492: \Qjet$ can be combined to conclude that $\dot N_{\rm esc} = \dot N_0 E_0
493: / E_{\rm esc}$, which implies that $\lapprox 20\%$ of the injected
494: particles escape (vertically) through the surface of the disk. The
495: remainder of the particles are either advected into the black hole or
496: else diffuse outward (horizontally) through the disk. Our predictions
497: for the shock/jet location and the asymptotic Lorentz factor agree
498: rather well with the observations of M87 reported by Biretta, Junor, \&
499: Livio (2002) and Biretta, Sparks, \& Macchetto (1999). The shock
500: location predicted by the model in the case of \SgrA is fairly close to
501: the value suggested by Yuan (2000). However, no reliable observational
502: estimate for the jet Lorentz factor in \SgrA is currently available.
503: 
504: \section{CONCLUSIONS}
505: 
506: In this Letter, we have demonstrated for the first time the physical
507: consequences of a standing shock in an inviscid ADAF accretion disk for
508: the acceleration of relativistic particles and the production of
509: outflows. The global model presented here provides a single, coherent
510: explanation for the disk/outflow structure based on the well-understood
511: concept of first-order Fermi acceleration in shock waves. The theory is
512: based on an exact mathematical approach to the solution of the combined
513: hydrodynamical and particle transport equations. Our work may help to
514: explain the observational fact that the brightest X-ray AGNs do not possess
515: strong outflows, whereas the sources with low X-ray luminosities but
516: high levels of radio emission do. We suggest that the gas in the
517: luminous X-ray sources is too dense to allow efficient Fermi
518: acceleration of a relativistic particle population, and therefore in
519: these systems, the gas simply heats as it crosses the shock. Conversely,
520: in the tenuous ADAF accretion flows studied here, the relativistic
521: particles are able to avoid thermalization due to the long collisional
522: mean free path, resulting in the development of a significant nonthermal
523: component in the particle distribution which powers the jets and
524: produces the strong radio emission.
525: 
526: We have presented detailed results that confirm that the general
527: properties of the jets observed in M87 and \SgrA can be understood
528: within the context of our disk/shock/outflow model. Although the
529: specific numerical results discussed here are limited to the case of an
530: inviscid disk, we believe that the inclusion of viscosity will not
531: change the basic features of the inflow/outflow solutions obtained here
532: because efficient first-order Fermi acceleration will also occur in
533: viscous disks, provided they contain shocks. In addition to viscosity,
534: future work must also address several other important considerations
535: such as the effects of general relativity and radiative transport on the
536: disk/shock structure.
537: 
538: The authors are grateful to the anonymous referee for several
539: useful suggestions that improved the manuscript.
540: 
541: \begin{thebibliography}{}
542: 
543: \bibitem{1} Abramowicz, M. A. \& Chakrabarti, S. K. 1990, \apj, 350, 281
544: 
545: \bibitem{2} Becker, P. A., Kazanas, D. 2001, \apj, 546, 429
546: 
547: \bibitem{3} Becker, P. A., Le, T. 2003, \apj, 588, 408
548: 
549: \bibitem{4} Becker, P. A., Subramanian, P., \& Kazanas, D. 2001, \apj,
550: 552, 209
551: 
552: \bibitem{5} Bicknell, G. V., \& Begelman, M. C. 1996, \apj, 467, 597
553: 
554: \bibitem{6} Biretta, J. A., Junor, W., \& Livio, M. 2002, NewAR, 46, 239
555: 
556: \bibitem{7} Biretta, J. A., Sparks, W. B., \& Macchetto, F. 1999, \apj,
557: 520, 621
558: 
559: \bibitem{8} Blandford, R.~D., \& Begelman, M.~C. 1999, \mnras, 303, L1
560: 
561: \bibitem{9} Blandford, R. D., Ostriker, J. P. 1978, ApJ, 221, L29
562: 
563: \bibitem{10} Chakrabarti, S. K. 1989, Publ. Astron. Soc. Japan, 41, 1145
564: 
565: \bibitem{11} Chakrabarti, S. K., \& Das, S. 2004, \mnras, 349, 649
566: 
567: \bibitem{12} Falcke, H., \& Biermann, P. L. 1999, A\&A, 342, 49
568: 
569: \bibitem{13} Ford, H. C. et al. 1994, \apj, 435, L27
570: 
571: \bibitem{14} Le, T., \& Becker, P. A. 2004, in preparation
572: 
573: \bibitem{15} Narayan, R., Kato, S., \& Honma, F. 1997, \apj, 476, 49
574: 
575: \bibitem{16} Owen, F. N., Eilek, J. A., \& Kassim, N. E. 2000, \apj,
576: 543, 611
577: 
578: \bibitem{17} Paczy\'nski, B., \& Wiita, P. J. 1980, A\&A, 88, 23
579: 
580: \bibitem{18} Quataert, E. 2003, astro-ph/0304099
581: 
582: \bibitem{19} Reynolds, C. S. et al. 1996, MNRAS, 283, L111
583: 
584: \bibitem{20} Schodel, R. et al. 2002, Nature, 419, 694
585: 
586: \bibitem{21} Yuan, F. 2000, MNRAS, 319, 1178
587: 
588: \bibitem{22} Yuan, F., Markoff, S., \& Falcke, H. 2002, A\&A, 383, 854
589: 
590: \end{thebibliography}
591: 
592: \end{document}
593: 
594: 
595: