1: %\documentclass[preprint2,11pt]{aastex}
2: %\documentclass[manuscript,11pt]{aastex}
3: \documentclass[preprint,12pt]{aastex} %--original for ApJ
4:
5: %\usepackage{natbib}
6: \def\cm{\ifmmode {\rm cm}^{-1} \else cm$^{-1}$ \fi}
7: \def\s{\ifmmode {\rm s}^{-1} \else s$^{-1}$ \fi}
8: \def\cc{\ifmmode {\rm cm}^{-3} \else cm$^{-3}$ \fi}
9: \def\cs{\ifmmode {\rm cm}^{-2} \else cm$^{-2}$ \fi}
10: \def\g{\ifmmode \gamma \else $\gamma$\fi}
11: \def\G{\ifmmode \Gamma \else $\Gamma$\fi}
12: \def\Gs{\ifmmode \Gamma~ \else $\Gamma~$\fi}
13: \def\Ka{K$\alpha$~}
14: \def\Kaa{K$\alpha$}
15:
16: \def\gc{\ifmmode \gamma_{\rm c} \else $\gamma_{\rm c}$ \fi}
17: \def\sw{Schwarzschild~}
18: \def\gsim{\mathrel{\raise.5ex\hbox{$>$}\mkern-14mu
19: \lower0.6ex\hbox{$\sim$}}}
20:
21: \def\lsim{\mathrel{\raise.3ex\hbox{$<$}\mkern-14mu
22: \lower0.6ex\hbox{$\sim$}}}
23:
24: \def\simless{\mathbin{\lower 3pt\hbox
25: {$\rlap{\raise 5pt\hbox{$\char'074$}}\mathchar"7218$}}} %< or of order
26: \def\simmore{\mathbin{\lower 3pt\hbox
27: {$\rlap{\raise 5pt\hbox{$\char'076$}}\mathchar"7218$}}} %> or of order
28: \def\Msun{M_\odot} % solar masses
29: \def\4u{4U 1728--34}
30:
31: \received{} \accepted{}
32: %\journalid{337}{15 January 1989}
33: %\articleid{11}{14}
34: \slugcomment{Resubmitted to ApJ, June 27, 2007}
35:
36: \lefthead{et al.} \righthead{\4u}
37:
38: \shorttitle{Shock-Driven Outflows from Hot Accretion Flows}
39: \shortauthors{Fukumura \& Kazanas}
40:
41: \begin{document}
42:
43: \title{Mass Outflows from Dissipative Shocks in Hot Accretion Flows }
44:
45: %\date{\today}
46:
47: %
48: % author list tbd - include nikolai IF we use the Gamma-QPO correlation
49: %
50: \author{Keigo Fukumura \& Demosthenes Kazanas}
51: \affil{Astrophysics Science Division, NASA Goddard Space Flight
52: Center, Code 663, Greenbelt, MD 20771}
53: %\email{fukumura@milkyway.gsfc.nasa.gov}
54: \email{fukumura@milkyway.gsfc.nasa.gov, Demos.Kazanas-1@nasa.gov}
55: %
56: %\and
57: %
58: %\author{P. Reig and I. E. Papadakis}
59: %\affil{IESL, Foundation for Research and Technology,
60: %711 10 Heraklion, Crete, Greece}
61: %\affil{Physics Department, University of
62: %Crete, PO Box 2208, 710 03 Heraklion, Crete, Greece}
63: %\email{pau@physics.uoc.gr,jhep@physics.uoc.gr}
64: %
65: %
66: %
67:
68: \begin{abstract}
69:
70: \baselineskip=15pt
71:
72: We consider stationary, axisymmetric hydrodynamic accretion flows in
73: Kerr geometry. As a plausible means of efficiently separating a
74: small population of nonthermal particles from the bulk accretion
75: flows, we investigate the formation of standing dissipative shocks,
76: i.e. shocks at which fraction of the energy, angular momentum and
77: mass fluxes do not participate in the shock transition of the flow
78: that accretes onto the compact object but are lost into collimated
79: (jets) or uncollimated (winds) outflows.
80:
81: The mass loss fraction (at a shock front) is found to vary over a
82: wide range ($0\% - 95\%$) depending on flow's angular momentum and
83: energy. On the other hand, the associated energy loss fraction
84: appears to be relatively low ($\lesssim 1\%$) for a flow onto a
85: non-rotating black hole case, whereas the fraction could be an order
86: of magnitude higher ($\lesssim 10\%$) for a flow onto a
87: rapidly-rotating black hole. By estimating the escape velocity of
88: the outflowing particles with a mass-accretion rate relevant for
89: typical active galactic nuclei, we find that nearly 10\% of the
90: accreting mass could escape to form an outflow in a disk around a
91: non-rotating black hole, while as much as 50\% of the matter may
92: contribute to outflows in a disk around a rapidly-rotating black
93: hole. In the context of disk-jet paradigm, our model suggests that
94: shock-driven outflows from accretion can occur in regions not too
95: far from a central engine.
96: %
97: %(within 2-40 gravitational radii). Radial density profile for upstream flow
98: %is found to be $\sim r^{-3/2}$ while that for downstream flow can be as
99: %steep as $r^{-3/2}$ to $r^{-3}$.
100: %
101:
102: Our results imply that a shock front under some conditions could
103: serve as a plausible site where (nonthermal) seed particles of the
104: outflows (jets/winds) are efficiently decoupled from bulk accretion.
105:
106:
107:
108:
109: \end{abstract}
110:
111:
112: \keywords{accretion, accretion disks --- black hole physics ---
113: hydrodynamics --- shock waves --- galaxies: jets }
114:
115: \baselineskip=15pt
116:
117: \section{Introduction}
118:
119:
120: %It has been widely known that the solutions to inviscid accreting
121: %gas around black holes may exhibit both continuous (shock-free) and
122: %discontinuous (shock-included) flows simultaneously.
123:
124: It has been established by now that a large body of astrophysical
125: objects hosting supermassive black holes (e.g.,
126: quasars, galactic black hole candidates or microquasars) exhibit
127: collimated, powerful jets/winds \citep[e.g.,][for
128: review]{BBR84,Livio99}. In particular, strong outflows generally
129: occur in the radio-loud active galactic nuclei (AGNs). It is now
130: widely accepted that the outflows observed from a large class of
131: objects as jets or winds have their origin in accretion flows, which
132: at the same time power the radiation emission associated with these
133: objects \citep[e.g.,][]{BP82,Fender04}. Many astrophysical
134: systems apparently manage to transfer the energy from inflowing
135: accretion flows to outflowing jets/winds of a small population of
136: nonthermal particles. For instance, \cite{Junor99} in radio
137: observations of the nearby active galaxy M87 found a remarkably
138: broad jet with strong collimation occurring already at $\sim 30
139: -100$ \sw radii from a central engine. Recently, \citet{Kataoka07}
140: discussed the disk-jet connection of the radio galaxy 3C~120
141: observed with {\it Suzaku}. Such a population of relativistic
142: outflowing particles is thought to produce a subsequent synchrotron
143: emission that is observed in several sources
144: \citep[e.g.,][]{Mirabel99}. This issue clearly points to the
145: importance of investigating a fundamental connection between the
146: accreting flows and outflows in regions not too far from the central
147: engines -- presumably within $\sim 100$ \sw radii.
148: %
149: Because the emission of radiation requires the dissipation of the
150: kinetic energy of the accretion flow, it is not unreasonable to
151: suggest that the two phenomena, i.e. the dissipation/emission of
152: radiation and the presence of outflows, are related to the same
153: generic process, which manifests itself with different guises at the
154: diverse sites that these phenomena are observed.
155:
156: Given that an accretion flow may very well become turbulent or, at a
157: very minimum hot due to adiabatic compression, it is not surprising
158: to anticipate such a flow to emit radiation, with luminosity and
159: spectrum depending strongly on the specific character of the
160: accretion process (e.g. quasi-spherical or disk). However, the
161: ubiquitous presence of outflows in the same objects presents a
162: different problem altogether: An outflow requires that a fraction of
163: the accreted matter be endowed with velocity higher than the escape
164: velocity associated with the specific radius at which the outflow is
165: launched. Given that the only free energy available to the accreting
166: gas is that of the gravitational field, hydrodynamic dissipation of
167: the accretion kinetic energy can never produce an outflow since this
168: process simply converts the kinetic energy of a gravitationally
169: bound flow into thermal, with the energy per unit mass (thermal plus
170: kinetic) never higher than that of the local gravitational
171: potential.
172:
173: The launch of an outflow such as those observed requires the
174: presence of an ``engine", i.e. a mechanism that expends mechanical
175: work (i.e. low entropy energy) to transfer a fraction of the
176: available energy to an even smaller fraction of the available mass,
177: thereby imparting to it specific energy greater than the local
178: gravitational potential; it is then expected that in its further
179: evolution the excess energy will be converted into directed motion
180: and the fraction of the mass will escape to infinity in a collimated
181: (jet) or uncollimated (wind) outflow.
182:
183: The well known models of outflows, usually involving the action of
184: magnetic fields such as the magnetocentrigual jet models
185: \citep[e.g.,][]{BP82,Contopoulos94,Konigl94,Vlahakis00}, are
186: specific examples as to what may constitute such an engine. These
187: models assume the presence of a thin Keplerian disk ``threaded" by a
188: poloidal magnetic field; the Keplerian rotation of the magnetic
189: field line footpoints and the magnetic tension transfer energy and
190: angular momentum to the disk plasma which can escape to infinity.
191: These models are consistent in that they solve simultaneously for
192: the poloidal field geometry and the flow velocity (under certain
193: simplifying assumptions, i.e. the self-similarity of the solutions).
194: \citet{Pelletier92} further developed the general theory of
195: non-self-similar solutions of hydromagnetic disk winds. In these
196: models the transfer of excess energy to the escaping particles is
197: made at the expense of the rotational energy of the matter in the
198: disk and it is mediated by the magnetic field.
199:
200:
201: While this type of model gained popularity, the issue of jet/outlfow
202: formation took a different turn with the introduction of
203: advection-dominated accretion flow (ADAF)
204: \citep[e.g.,][]{Narayan94,Manmoto97}. These radiatively inefficient
205: accretion flows (RIAF) were found to have positive Bernoulli
206: integral of the flow and could therefore fulfill the condition
207: necessary for the launching of jet/wind outflows; as such they
208: present potentially interesting sites for the origin of such
209: outflows. The positivity of the Bernoulli integral has been
210: discussed and analysed by \citet{BB99} who pointed out that it is
211: due to the combination of energy transfer by the viscous torques
212: from the inner to the outer sections of the flow (the gas of the
213: flow becomes bounded at its inner edge) and the local dissipation of
214: the flow's azimuthal kinetic energy which is not radiated away but
215: stored in the fluid to increase its internal energy. The latter
216: authors then argued that the excess energy can be carried away to
217: infinity (along with some fraction of the accreting mass and angular
218: momentum) to produce continuous outflows from all radii to infinity
219: while leaving the remaining flow with negative Bernoulli constant to
220: naturally accrete onto the compact object: advection-dominated
221: inflow-outflow solution (ADIOS). In this case, while the necessary
222: excess energy is transferred by the viscous torques from the flow's
223: more highly bound inner section, the necessary separation of mass to
224: components with positive and negative total energy is still left
225: unspecified.
226:
227: An altogether different model that offers a simplified picture of
228: such a separation was presented by \citet{Subramanian99} who
229: proposed that in the tenuous, collisionless plasma of an ADAF
230: particles (protons) could be accelerated via a second-order Fermi
231: acceleration by the shear motions of the underlying quasi-Keplerian
232: azimuthal flow. They then argued that if sufficiently large pressure
233: is built in the accelerated particle proton population (the
234: electrons generally lose energy on time scales short compared to
235: their transit time through the system and cannot build an energy
236: density that could be dynamically important) and for favorable
237: geometries of the disk magnetic field (large scale poloidal loops
238: that open up above the disk) the relativistic particle population
239: could naturally (through the action of the gravitational field)
240: segregate itself from the non-relativistic one, carrying off to
241: infinity only the accelerated ($E \gsim m_pc^2$) portion of the disk
242: plasma. In this case the engine is a combination of the particle
243: acceleration and the action of the gravitational field.
244:
245: Finally, a model along the same lines was proposed by
246: \citet{Contopoulos95} who suggested that even in the case of a
247: completely turbulent magnetic field, a separation of the
248: relativistic and non-relativistic particle populations is possible
249: through the production of relativistic neutrons in the collisions of
250: the relativistic protons with the ambient plasma and the ensuing
251: production of relativistic neutrons. The subsequent decay of
252: neutrons back into protons produces then a proton fluid in regions
253: of space devoid of inertia whose energy-to-mass ratio (and hence its
254: asymptotic Lorentz factor) depends only on the ratio $R/c \tau_{\rm
255: n}$, where $R$ is the size of the system and $\tau_{\rm n}$ the
256: neutron life time and can lead to highly relativistic flows for
257: black hole masses $M \gsim 10^8 \Msun$.
258:
259: In the present note we follow a similar simplified view to study
260: outflows in objects powered by accretion: We consider the presence
261: of (2-dimensional) shocks as a means of dissipation of the accretion
262: kinetic energy in a fashion similar to that considered by
263: \citet{Cha90} and collaborators. That is we consider the transonic
264: accretion of matter with the proper angular momentum to produce a
265: standing shock at a radius close to the horizon, which, subsequently
266: accretes onto the black hole after passing through a downstream
267: sonic point. Previous steady-state analysis found that it is
268: possible to judiciously choose the specific angular momentum of the
269: flow, so that the outer transonic one could be connected through a
270: shock transition to an inner transonic one which, passing through an
271: inner sonic point, accretes onto the black hole. The location of the
272: shock in this situation is determined by finding a radial position
273: at which the density and velocity of the two flow sections were
274: those demanded by the dissipative Rankine-Hugoniot conditions across
275: a shock. So far, the formation of standing shocks in hydrodynamic
276: accretion has been extensively studied by a number of authors for
277: both inviscid flows
278: \citep[e.g.,][]{Cha90,Sponholz94,Cha96,Lu98,FT04} and viscous flows
279: \citep[e.g.,][]{Cha90,Lu99,Cha04}, considering either dissipative or
280: non-dissipative shock jump conditions. Recently, standing shocks in
281: the presence of poloidal magnetic fields [i.e., magnetohydrodynamic
282: (MHD) shocks] around a black hole was also studied for various
283: parameter dependence (see Das \& Chakrabarti 2007 for
284: pseudo-Newtonian geometry; Takahashi et al. 2002, Fukumura et al.
285: 2007 for Kerr geometry). In the context of the particle acceleration
286: via the first-order Fermi mechanism across a shock front, the
287: production of shock-accelerated relativistic protons were discussed
288: in spherical accretion \citep[][]{Kazanas83,Kazanas86}, while other
289: authors have explored the relativistic outflows in ADAF with shocks
290: \citep[][]{Becker04,Becker05}. Similar attempts have been made to
291: make physical connections between the shocked-accretion and
292: outflows. For instance, mass outflow rate were estimated from
293: adiabatic shocked-flow region in Newtonian gravity
294: \citep[e.g.,][]{Cha99,Das00}. \citet{Das99} took a similar approach
295: to study the shock-generated outflows with little energy dissipation
296: in pseudo-Newtonian geometry. Independently, from general
297: relativistic MHD simulations, the formation of jets
298: (magnetically-driven and gas-pressure driven jets) is found in the
299: high pressure regions due to the shock/adiabatic compression
300: \citep[e.g.,][]{Koide99,Nishikawa05}. They concluded that the jets
301: are mainly produced by the gas-pressure gradient which is greatly
302: enhanced by the shock front at around $r \sim 6$ gravitational radii
303: (note that this feature was not seen in the Newtonian calculations).
304: These studies also suggest an essential connection between the
305: shocked accretion flows and the jets; i.e., the shock front may
306: serve as a base of the outflows.
307:
308:
309: The novelty of our approach lies in considering the possibility of
310: shock formation (i.e. obeying the general relativistic, dissipative
311: Rankine-Hugoniot conditions at a shock front) in which part of the
312: mass, angular momentum and energy fluxes escape in the z-direction
313: and do not participate in the shock transition. We then examine the
314: energy per unit mass of the escaping matter which we compare to the
315: escape velocity at the shock radius; if it is greater than the
316: latter we conclude that this scenario can produce an outflow with
317: its outflow rate $\dot m$ and luminosity that are calculable and can
318: be compared to those of the entire accretion to obtain a measure of
319: the efficiency of our ``engine" in producing outflows. Some
320: population of energetic nonthermal particles, produced via a shock
321: acceleration, may then be well separated from the equatorial
322: accretion flows (which consists primarily of thermal particles).
323:
324:
325: %
326: %In particular, in order to understand the underluminous systems such
327: %as our Galactic center source (Sgr A$^\ast$), an advection-dominated
328: %accretion flow (ADAF) appears to be a very promising model in which
329: %most of the viscously dissipated energy is advected inward with the
330: %flow \citep[e.g.,][]{Narayan95,Manmoto97}.
331: %
332:
333:
334: %Although the exact processes triggering strong outflows from the
335: %surface of an accretion disk are still not completely understood,
336: %some theoretical models have been proposed. \citet{BB99} proposed an
337: %advection-dominated inflow-outflow solution (ADIOS), which is a
338: %modified version of an advection-dominated accretion flow (ADAF)
339: %model \citep[e.g.,][]{Narayan94,Manmoto97}. In the ADIOS model, some
340: %fraction of the accreting gas (including energy, mass and angular
341: %momentum) is dissipated away from the flow as strong winds before
342: %reaching the black hole horizon such that the leftover gas possesses
343: %a negative Bernoulli parameter, thus gravitationally bound. Their
344: %model therefore naturally explains a physical necessity of powerful
345: %outflows. \citet{Becker01} extended the original ADIOS model to
346: %incorporate a pseudo-Newtonian gravity which modifies the dynamics
347: %of the flow in the inner region of ADIOS.
348:
349:
350: %
351: %The ADIOS model, when applied to Sgr A$\ast$ data with {\it
352: %Chandra}, also seems to produce a better predicted spectrum compared
353: %with that by ADAF model \citep{Yuan02}.
354: %
355:
356: %It is also discussed, as another plausible mechanisms for producing
357: %relativistic outflowing particles, that the formation of shocks is
358: %physically relevant for accelerating the electrons up to
359: %relativistic regime \citep[e.g.,][]{Blandford78,Becker04,Becker05}.
360:
361:
362: More specifically, since we are interested in the formation of
363: outflows through shocks in the inner disk region relatively close to
364: a presumably rotating black hole at a center (say, $r \lesssim 30$
365: gravitational radii), it is important to include the strong gravity
366: and frame-dragging effects described by general relativity.
367: %Importantly, due to the fact that many black holes at the center of
368: %the astrophysical systems have been considered to be rotating and
369: %that shocks can occur in the innermost accretion disk (say, within
370: %ten gravitational radii), it is also important to accurately
371: %describe the relativistic dynamics of the accreting gas.
372: To the best of our knowledge, no relevant work in the literature
373: incorporates such mass and energy loss in the shock jump conditions.
374: %
375: %
376: %However, no explicit work has been done in the literature, in this
377: %perspective, on incorporating mass and energy loss at dissipative
378: %shocks in the jump conditions.
379: %
380: In the framework of our model presented here, mass loss is coupled
381: to energy loss via the shock jump conditions, and therefore it must
382: be considered simultaneously. It is this point that motivates us to
383: explore, for the first time, the formation of outflowing particles
384: as a consequence of the formation of dissipative standing shocks in
385: accretion in a fully relativistic treatment. The formalism of our
386: current model is partly based on the previous works
387: \citep[][]{Yang95,Lu98,FT04}. Our main objective in this study is
388: therefore to explore in details a possibility that the formation of
389: outflows (jets/winds) can occur at a dissipative shock front in
390: transonic accreting flows: i.e., a connection between
391: shocked-accreting gas and the outflowing particles.
392:
393:
394: The structure of this paper is as follows. In \S 2 we review and
395: explain our simplified model that simultaneously considers both
396: shocks and outflows with appropriate jump conditions. Main parameter
397: dependence of the shock-outflow solutions is explored in \S 3, where
398: we show the nature of the shock-outflow solutions and the
399: corresponding global accretion solutions. Our primary goal in these
400: analysis is to examine the coupling between the shock-outflow
401: solutions. In \S 4 we discuss our results and make some observation
402: implications. Brief summary and concluding remarks are given there
403: as well.
404:
405:
406:
407:
408: \section{Model Assumptions \& Basic Equations}
409:
410: In black hole accretion, accreting gas must be transonic. After
411: passing through a first sonic radius, the gas is slowed down, and a
412: shock may develop. For causality, the shocked gas must become
413: supersonic again before crossing the event horizon. Below, we will
414: explain the details of our simplified model.
415:
416: \subsection{Accreting Flows around a Black Hole}
417:
418: We consider a steady-state, axisymmetric accreting flows in Kerr
419: geometry. The spacetime metric is expressed by the Boyer-Lindquist
420: coordinates as
421: \begin{eqnarray}
422: ds^2 &=& -\left(1-\frac{2 m r}{\Sigma}\right) dt^2 - \frac{4 m r a
423: \sin^2 \theta}{\Sigma} dt d\phi \nonumber \\ & & + \frac{A \sin^2
424: \theta}{\Sigma} d\phi^2 + \frac{\Sigma}{\Delta} dr^2 + \Sigma
425: d\theta^2 \ , \label{eq:BL}
426: \end{eqnarray}
427: where $\Delta \equiv r^2-2mr+a^2$, $\Sigma \equiv r^2+a^2 \cos^2
428: \theta$, $A \equiv (r^2+a^2)^2-a^2 \Delta \sin^2 \theta$. $m$ and
429: $a$ are mass and specific angular momentum (or Kerr parameter) of a
430: black hole. Following the standard geometrized units, we have taken
431: $G=c=1$ in equation~(\ref{eq:BL}). Thus, the length (distance) $r$
432: and $a$ are measured in units of $m$. Since we are interested in the
433: equatorial flows, we set $\theta=\pi/2$ throughout this paper. The
434: black hole horizon is then expressed by $r_{\rm{H}} \equiv
435: m+\sqrt{m^2-a^2}$. The self-gravity of the flow is ignored, and we
436: do not include the effects of magnetic fields for simplicity.
437:
438:
439:
440: We follow the earlier works on relativistic shock formation as
441: follows; accretion time is assumed to be shorter than that of energy
442: diffusion, thus treating the flows as adiabatic except at a shock
443: front where a fraction of fluid energy, angular momentum, and mass
444: are dissipated. To prescribe thermodynamic quantities we adopt a
445: polytropic form as
446: \begin{eqnarray}
447: P = K \rho_0^{1+1/N} \ , \label{eq:polytropic}
448: \end{eqnarray}
449: where $P$ and $\rho_0$ are the thermal pressure and rest-mass
450: density of the flow, which are locally measured in the fluid frame.
451: Here, $N$ denotes the polytropic index. The entropy of the fluid is
452: characterized by $K$ which is related to the entropy $S$ by $S
453: \equiv c_v \log K$ where $c_v$ denotes a specific volume heat of the
454: flow. Because of stationarity ($\partial t=0$) and axial symmetry
455: ($\partial \phi=0$), there exist two conserved quantities along a
456: fluid stream line; namely, specific energy $E$ and axial angular
457: momentum $L$ of the fluid, defined by
458: \begin{eqnarray}
459: E &\equiv& -\mu u_t \ , \label{eq:energy}
460: \\
461: L &\equiv& \mu u_\phi \ , \label{eq:L}
462: \end{eqnarray}
463: where $\mu = (P+\rho)/\rho_0$ is the relativistic enthalpy of the
464: fluid, and $\rho = \rho_0 + N P$ is the net baryon mass-energy
465: density (including internal energy $N P$). The number density of the
466: constituent baryon $n$ is given by $\rho_0 \equiv n m_p$ where $m_p$
467: is the baryon mass in the flow. We assume that energy $E$ and
468: angular momentum $L$ are both conserved along the flow except at a
469: shock location ($r=r_{\rm{sh}}$) where they are partially dissipated
470: (will be explained in detail in the next section).
471:
472:
473: From the four-velocity normalization of the flow ($u_\alpha u^\alpha
474: =-1$), we get
475: \begin{eqnarray}
476: 1+u_r u^r + (u^t)^2 V_{\rm{eff}}(r,\lambda)=0 \ , \label{eq:V}
477: \end{eqnarray}
478: where the effective potential has been introduced by
479: $V_{\rm{eff}}(r,\lambda) \equiv g^{tt}- 2\lambda g^{t
480: \phi}+\lambda^2 g^{\phi\phi}$ with $g^{\alpha \beta}$ being the
481: inverse metric components for $g_{\alpha \beta}$ in
482: equation~(\ref{eq:BL}). Here, the specific angular momentum of the
483: flow $\lambda$ is defined by
484: \begin{eqnarray}
485: \lambda \equiv \frac{L}{E} = -\frac{u_\phi}{u_t} \ , \label{eq:ell}
486: \end{eqnarray}
487: which is assumed to be conserved along the whole flow in our model.
488: %This assumption in general does not have to be imposed, and in fact
489: %it may not always valid in the actual astrophysical accretion.
490: %However, our main purpose of this work is to systematically explore
491: %a characteristic nature of the energy and mass loss associated with
492: %the shock formation. Without the above assumption,
493: From equation~(\ref{eq:V}) we obtain
494: \begin{eqnarray}
495: u_t(r,\lambda) = \left[\frac{1+u_r u^r}{-V_{\rm
496: eff}(r,\lambda)}\right]^{1/2} \ . \label{eq:ut}
497: \end{eqnarray}
498:
499:
500: From the definition of enthalpy and the polytropic relation given by
501: equation~(\ref{eq:polytropic}) we can rewrite the enthalpy as
502: \begin{eqnarray}
503: \mu = 1+(N+1)K \rho_0^{1/N} \ . \label{eq:mu}
504: \end{eqnarray}
505: Local adiabatic sound speed $c_s$ is defined as
506: \begin{eqnarray}
507: c_s^2 \equiv \left(\frac{\partial P}{\partial \rho}\right)_{\rm{ad}}
508: = \frac{(1+1/N) P}{\mu \rho_0} \ . \label{eq:sound}
509: \end{eqnarray}
510: %We define flow's Mach number as
511: %\begin{eqnarray}
512: %{\cal{M}} \equiv \frac{v_{\rm CRF}}{c_s} \ , \label{eq:Mach}
513: %\end{eqnarray}
514: %where $v_{\rm CRF} \equiv (u_r u^r)/(1+u_r u^r)$ is the flow speed
515: %measured in the corotating reference frame (CRF).
516: Combining the equations~(\ref{eq:polytropic}), (\ref{eq:mu}) and
517: (\ref{eq:sound}) we can express $\mu$ in terms of the sound speed
518: $c_s$ as
519: \begin{eqnarray}
520: \mu = \frac{1}{1-N c_s^2} \ . \label{eq:mu2}
521: \end{eqnarray}
522: Accordingly, the energy of the flow $E$ in
523: equation~(\ref{eq:energy}) can be explicitly rewritten as a function
524: of $c_s$
525: \begin{eqnarray}
526: E = \left(\frac{1+u_r u^r}{-V_{\rm eff}}\right)^{1/2} / \left(1-N
527: c_s^2\right) \ , \label{eq:energy2}
528: \end{eqnarray}
529: which is conserved along the flow except at a shock front. That is,
530: across a shock front, the energy $E$ and the angular momentum $L$
531: will both decrease in such a way that the ratio $\lambda \equiv L/E$
532: is continuous across the shock. Using equation~(\ref{eq:mu}) and
533: (\ref{eq:mu2}) the baryon rest-mass density can be rewritten as
534: \begin{eqnarray}
535: \rho_0 = \left[\frac{c_s^2}{\left(1+1/N\right)\left(1-N
536: c_s^2\right)}\right]^N \frac{1}{K^N} \ . \label{eq:rho}
537: \end{eqnarray}
538: We define the mass-accretion rate $\dot{M}$ as
539: \begin{eqnarray}
540: \dot{M} \equiv -4\pi r H u^r \rho_0 \ , \label{eq:mdot1}
541: \end{eqnarray}
542: where $H$ represents the vertical scale-height of the flow defined
543: as
544: \begin{eqnarray}
545: H \equiv \frac{c_s}{\Omega_K} \ , \label{eq:H}
546: \end{eqnarray}
547: from the conventional hydrostatic equilibrium assumption. Here,
548: $\Omega_K(r) \equiv m^{1/2}/(r^{3/2}+a m^{1/2})$ is the Keplerian
549: angular velocity. Note that for accretion we have $u^r<0$. As often
550: assumed, we only consider a constant mass-accretion rate in this
551: paper, although a variable accretion rate has been discussed in the
552: literature \citep[e.g.,][]{BB99}. Eliminating $\rho_0$ from
553: equations~(\ref{eq:rho}) and (\ref{eq:mdot1}) we rewrite $\dot{M}$
554: as
555: \begin{eqnarray} \dot{M} = -4\pi r
556: m^{1/2} \left(r^3/2+a m^{1/2}\right) u^r
557: \left[\frac{c_s^{2N+1}}{(1+1/N)(1-N c_s^2)^N}\right] \frac{1}{K^N} \
558: . \label{eq:mdot2}
559: \end{eqnarray}
560: Similarly to $E$ and $L$, the mass-accretion rate $\dot{M}$ is also
561: conserved along the flow except at a shock location. After defining
562: all the physical quantities necessary to solve for black hole
563: accretion, we will describe below the transonic properties of the
564: physical accretion solutions.
565:
566:
567:
568:
569: \subsection{Regularity Conditions}
570:
571: For accretion to continue on to the event horizon, it is required
572: that the accreting flows become supersonic at a sonic radius. After
573: taking the derivatives of the equations~(\ref{eq:energy2}) and
574: (\ref{eq:mdot2}) with respect to $r$ (note that $E$ and $\dot{M}$
575: are both constants), $dc_s/dr$ can be eliminated. Finally we obtain
576: \begin{eqnarray}
577: \frac{du^r}{dr} = \frac{\cal{N}}{\cal{D}} \ , \label{eq:critical}
578: \end{eqnarray}
579: where
580: \begin{eqnarray}
581: {\cal{D}} &\equiv& 2N c_s^2 + u_r u^r \left\{N \left(3c_s^2-2
582: \right)-1 \right\} \ , \label{eq:D}
583: \\
584: {\cal{N}} &\equiv& \zeta \left(1+u_r u^r \right) \left(r^{3/2}+a
585: m^{1/2}\right) \frac{dV_{\rm eff}}{dr} \nonumber \\ & & - \left[a
586: \left(4Nc_s^2+\eta \right)+r^{3/2} \left\{2Nc_s^2
587: \left(5+3g_{rr}\right)+\eta \right\} \right] V_{\rm eff} \ ,
588: \label{eq:N}
589: \end{eqnarray}
590: and
591: \begin{eqnarray}
592: \eta &\equiv& 4N g_{rr} c_s^2 + \zeta (u^r)^2 \frac{d g_{rr}}{dr} \
593: . \label{eq:eta}
594: \\
595: \zeta &\equiv& r \left\{N \left(c_s^2-2 \right)-1 \right\} \ .
596: \label{eq:zeta}
597: \end{eqnarray}
598: In equation~(\ref{eq:critical}), the fact that the velocity gradient
599: $du^r/dr$ is finite at $r=r_c$ requires that ${\cal{D}}(r=r_c)=0$
600: and ${\cal{N}}(r=r_c)=0$ simultaneously (i.e., regularity
601: conditions), which will allow us to find the critical radius $r_c$
602: for a given parameter set. A physically valid accretion solution,
603: therefore, must pass through a critical point at $r=r_c$ before
604: reaching the horizon whether or not the shock formation is possible.
605: In the presence of a shock, a global shock-included accretion
606: solution must go through a critical point on both sides of the shock
607: location (i.e., before and after the shock forms). This fact
608: requires multiple critical points. It is widely known that multiple
609: critical points (up to three at most) can exist in general for a
610: certain flow parameter space. Much work has been done on examining
611: the topological behaviors (i.e., saddle, node, center and spiral
612: points) of the critical points \citep[e.g.,][]{Cha90}, thus we will
613: not repeat this. Because our main goal of this paper is to explore a
614: possibility of outflows that are coupled to the shocks in the global
615: accretion flows, we will only investigate the upstream flows passing
616: through the outer critical points $r_c^{\rm out}$ while the
617: downstream flows passing through the inner critical points $r_c^{\rm
618: in}$ (the middle one is known to be unphysical).
619:
620:
621:
622: So far, for a specified flow parameter set, one can obtain accretion
623: solutions. Next, let us impose the shock conditions that connects
624: one solution (i.e. upstream flow) to another (i.e. downstream flow),
625: taking into account energy, angular momentum, and mass loss at a
626: shock location.
627:
628:
629:
630:
631: \subsection{Shock Formation with Energy and Mass Loss}
632:
633: Accreting flows around a black hole are generally subject to a
634: number of ``invisible'' obstacles that can decelerate the flow: (1)
635: centrifugal barrier due to the fluid's angular momentum, (2) gas
636: pressure-gradient, (3) radiation pressure-gradient, and (4) magnetic
637: forces (i.e., pressure-gradient and/or tension force). Although our
638: model is purely adiabatic and hydrodynamic [thus (3) and (4) are
639: absent], flows are still under the influence of the deceleration
640: mechanisms (1) and (2).
641:
642:
643: \begin{figure}[t]% ------------------------------------- Figure~1
644: \centering
645: \epsscale{0.5}
646: \includegraphics[angle=0, width=3.5in]{f1.eps}
647: \caption{Schematic picture of mass outflows originating from a standing shock front
648: in hydrodynamic accretion. $\bigotimes$ indicates azimuthal motion of the flow.
649: At the shock, a fraction of the accreting gas is lost
650: carrying energy $\Delta E$, angular momentum $\Delta L$ and mass
651: outflow rate
652: $\Delta \dot{M}$.
653: } \label{fig:schematic}
654: \end{figure} %-----------------------------------------------
655:
656:
657:
658: Following the previous works \citep[see][]{Yang95,Lu98}, let us
659: assume jump conditions that allow energy, angular momentum and mass
660: loss at a standing shock front. Figure~\ref{fig:schematic}
661: illustrates a schematic description of our model. From
662: equation~(\ref{eq:energy2}) we have
663: \begin{eqnarray}
664: E_1 &=& \left[\frac{1+g_{rr} (u_1^r)^2}{-V_{\rm eff}}\right]^{1/2} /
665: \left(1-N c_{s1}\right) \ , \label{eq:E1}
666: \\
667: E_2 &=& \left[\frac{1+g_{rr} (u_2^r)^2}{-V_{\rm eff}}\right]^{1/2} /
668: \left(1-N c_{s2}\right) \ , \label{eq:E2}
669: \end{eqnarray}
670: where the subscripts ``1'' and ``2'' denote the quantities for
671: upstream and downstream flows evaluated at a shock location
672: ($r=r_{\rm sh}$), respectively. We require $E_1 > E_2$ and define
673: the energy dissipation and its fraction as
674: \begin{eqnarray}
675: \Delta E \equiv E_1 - E_2 ~~~~ {\rm and} ~~~~ f_E \equiv
676: \frac{\Delta E}{E_1} \ , \label{eq:DeltaE}
677: \end{eqnarray}
678: where $0 < f_E < 1$. The associated angular momentum carried by the
679: outflows is then given by $\Delta L \equiv L_1 - L_2 = \lambda
680: \Delta E$.
681:
682:
683: In the baryon mass conservation across the shock front, we also
684: consider some mass loss (equivalently the loss of mass-accretion
685: rate) associated with the outflows that are blown away as
686: winds/jets. From equation~(\ref{eq:mdot2}) we have
687: \begin{eqnarray}
688: \dot{M}_1 = -4\pi r_{\rm{sh}} \left(r_{\rm{sh}}^3/2+a m^{1/2}\right)
689: u_1^r \left[\frac{c_{s1}^{2N+1}}{(1+1/N)\left(1-N
690: c_{s1}^2\right)^N}\right] \frac{1}{K_1^N} \ , \label{eq:Mdot1}
691: \\
692: \dot{M}_2 = -4\pi r_{\rm{sh}} \left(r_{\rm{sh}}^3/2+a m^{1/2}\right)
693: u_2^r \left[\frac{c_{s2}^{2N+1}}{(1+1/N)(1-N c_{s2}^2)^N}\right]
694: \frac{1}{K_2^N} \ . \label{eq:Mdot2}
695: \end{eqnarray}
696: Clearly, the mass-accretion rate depends on the measurement of
697: entropy $K$ which must increase across the shock because of the heat
698: generated (second law of thermodynamics). However, in the presence
699: of both energy dissipation and mass loss at the shock front, a
700: fraction of total energy (including mass and thermal energies) can
701: be released from the flow surface, quickly reducing the rise of $K$.
702: This hypothesis is justifiable when cooling processes are very
703: efficient, although it is beyond the scope of this work to discuss
704: the details of these mechanisms. Hence, for simplicity, we assume
705: that the entropy will roughly remain unchanged at the shock front as
706: a result from partial heat loss of the shocked flow. Thus, we set
707: $K_1 = K_1 \equiv K_0$ and require $\dot{M}_1 \ge \dot{M}_2$.
708: Similarly to energy dissipation, let us define the mass loss and its
709: fraction as
710: \begin{eqnarray}
711: \Delta \dot{M} \equiv \dot{M}_1 - \dot{M}_2 ~~~~ {\rm and} ~~~~
712: f_{\dot{M}} \equiv \frac{\Delta \dot{M}} {\dot{M}_1} \ ,
713: \label{eq:DeltaMdot}
714: \end{eqnarray}
715: where $0 \le f_{\dot{M}} < 1$. That is, $f_{\dot{M}}=0$ corresponds
716: to no mass outflows from shocks.
717:
718:
719: The momentum flux density $T^{\alpha \beta}$ for an ideal fluid is
720: given by
721: \begin{eqnarray}
722: T^{\alpha \beta} = (\rho+P)u^{\alpha}u^{\beta} + P g^{\alpha \beta}
723: \ . \label{eq:momentum}
724: \end{eqnarray}
725: Therefore, the radial component $T^{rr}$ is
726: \begin{eqnarray}
727: T^{rr} = \rho_0 \mu u^r \left[u^r+\frac{c_s^2}{(1+1/N)u_r}\right] \
728: . \label{eq:Trr}
729: \end{eqnarray}
730: From the mass and momentum conservations in radial direction, we
731: finally obtain
732: \begin{eqnarray}
733: \mu_1 c_{s2} \left[u_1^r + \frac{c_{s1}^2}{(1+1/N)g_{rr}
734: u_1^r}\right] = \mu_2 c_{s1} \left(1-f_{\dot{M}}\right) \left[u_2^r
735: + \frac{c_{s2}^2}{(1+1/N)g_{rr} u_2^r}\right] \ ,
736: \label{eq:mom-cons}
737: \end{eqnarray}
738: which completes a series of our dissipative shock conditions.
739:
740: The strength of shocks is measured by the local compression ratio of
741: the flow, $n_2/n_1$, which is expressed as
742: \begin{eqnarray}
743: \frac{n_2}{n_1} = \frac{u^r_1}{u^r_2} \frac{c_{s1}}{c_{s2}}
744: (1-f_{\dot{M}}) = \left[ \left(\frac{c_{s2}}{c_{s1}}\right)^2
745: \left(\frac{1-N c_{s1}^2}{1-N c_{s2}^2}\right) \right]^N \ .
746: \label{eq:n12}
747: \end{eqnarray}
748: Hence, in the absence of mass loss ($f_{\dot{M}}=0$), the
749: compression ratio must be greater than unity when shocks occur.
750: However, in the presence of mass loss ($0 < f_{\dot{M}} <1$), the
751: compression ratio becomes a product of the velocity ratio and the
752: mass loss fraction. Therefore, there can be a case where rarefaction
753: or decompression with $n_2/n_1 \le 1$ might take place even if the
754: fluid velocity (and sound speed) abruptly decreases across a very
755: strong shock. To avoid further complications, we will focus our
756: attention on compression shock waves only.
757:
758:
759: \subsection{Dynamical Stability of Shocks}
760:
761: In terms of the radial momentum balance across the shock front, some
762: standing shocks can be dynamically unstable. That is, the shock may
763: decay away (either radially inward or outward) as a result of a
764: small (radial) perturbation of its position. In order to examine the
765: stability of the obtained shock solutions, we perturb the radial
766: momentum flux density $T^{rr}$ (equivalent to pressure) by invoking
767: an infinitesimally small variation of the shock location $\delta
768: r_{\rm sh}$. If a shock front shifts back to its original location
769: to retain the momentum equilibrium there, it is dynamically stable.
770: The criteria is expressed as
771: \begin{eqnarray}
772: \left(\delta T^{rr}_2 - \delta T^{rr}_1\right)_{\rm sh} =
773: \left(\frac{d T^{rr}_2}{dr} - \frac{d T^{rr}_1}{dr}\right)_{\rm sh}
774: \delta r_{\rm sh} \equiv \kappa(r_{\rm sh}) \delta r_{\rm sh} \ ,
775: \label{eq:stability}
776: \end{eqnarray}
777: where $\kappa(r_{\rm sh})$ is a function of $r_{\rm sh}$ alone, thus
778: possible to be numerically evaluated. In our calculations we take
779: advantage of the known fact that $\kappa(r_{\rm sh})<0$ always
780: guarantees the stable standing shocks
781: \citep[see][]{Yang95,Lu98,FT04}.
782:
783:
784: \section{Numerical Results}
785:
786: Following the formalism outlined above we calculate the fractions of
787: energy and mass losses for various model parameters. There are
788: essentially three primary variables ($E_1,\lambda;r_{\rm sh}$) that
789: determine the solutions for a given geometry ($a,\theta$). In this
790: paper, we restrict ourselves to the flows corotating with the black
791: hole ($a \lambda > 0$) at the equator ($\theta=\pi/2$).
792:
793:
794: The topology of the accreting solutions is normally classified as
795: ``x-type" or ``$\alpha$-type" depending on whether the shock-free
796: solution is global or not \citep[e.g., see][for definition]{Lu98}.
797: Although this is also one of the important aspects of the studies of
798: standing shocks, we do not consider the distinction here since it is
799: not crucial to our current investigations.
800:
801:
802: In our calculations we set $N=3$ and choose several representative
803: values of the upstream flow energy: $E_1=1.003,1.004,1.005$ for
804: $a/m=0$ (\sw case) and $0.99$ (Kerr case) to illustrate the
805: frame-dragging effect.
806: %We
807: %have explored the parameter space that allows both dissipative
808: %shocks and mass loss, from which the values have been chosen in our
809: %calculations.
810:
811: %It should be noted that for a given $\lambda$ there exists the
812: %degeneracy of shock-outflow solutions: i.e., for a single value of
813: %angular momentum $\lambda$ there allowed a finite range of the
814: %possible shock locations $r^{min}_{sh} < r_{\rm sh} < r_{\rm sh}^{\rm max}$
815: %corresponding to respective outflows $f_{\dot{M}}^{min} \le
816: %f_{\dot{M}} \le f_{\dot{M}}^{max}$ where the indices ``min" and
817: %``max" denote minimum and maximum values, respectively.
818: %
819: %
820: %In the context of our model, one can not uniquely remove this
821: %degeneracy, although it should be removed by additional physical
822: %conditions (e.g., viscosity, magnetic fields, for instance).
823: %
824:
825:
826: \subsection{Dependence of Energy and Mass Loss}
827:
828: We first present in Figures~\ref{fig:rsh-a0} and \ref{fig:rsh-a099}
829: the mass loss fraction $f_{\dot{M}}$ as a function of shock location
830: $r_{\rm sh}$ for (a) $E_1=1.003$, (b) 1.004, and (c) 1.005, with
831: $a/m=0$ (Figure~\ref{fig:rsh-a0}) and $0.99$
832: (Figure~\ref{fig:rsh-a099}). Stable shocks are represented by solid
833: curves while (dynamically) unstable shocks are represented by dotted
834: curves. Stable standing shocks, according to the criterion of
835: equation~(\ref{eq:stability}), can form in most cases in regions
836: relatively close to the black hole ($r_{\rm sh}/m \lesssim 80$). For
837: a rotating black hole case, the shock location can be considerably
838: closer to the hole ($r_{\rm sh}/m \gtrsim 2-3$), results similar to
839: no mass loss cases \citep[e.g.,][]{Sponholz94,Lu98}, due to the fact
840: that the horizon shifts more inward. There appears to be a smooth
841: transition between the stable and unstable ones. Filled circles
842: denote the {\it maximum} stable shock location, also corresponding
843: to the {\it weakest} shock (i.e., smallest $n_2/n_1$), while open
844: circles denote the {\it minimum} stable shock location, also
845: corresponding to the {\it strongest} shocks (i.e., largest
846: $n_2/n_1$). In other words, the stronger dissipative shocks can
847: develop at smaller radii, in agreement with the previous result in
848: the absence of mass loss \citep[e.g.,][in which
849: $f_{\dot{M}}=0$]{Lu98,FT04}. Unstable shocks start to develop from
850: where stable shocks become strongest. At both ends of the solution
851: the curves are restricted by the dissipative shock conditions
852: [equations~(\ref{eq:E1}), (\ref{eq:E2}), (\ref{eq:Mdot1}),
853: (\ref{eq:Mdot2}), (\ref{eq:mom-cons})] and the transonic properties
854: [equation~(\ref{eq:critical})]. It is seen that smaller angular
855: momentum is required for shock (and accretion) to take place when
856: the fluid energy is larger, also consistent with previous studies of
857: standing shocks \citep[][]{Cha96,Lu97,Lu98,FT04}. Only stable shocks
858: alone are allowed when angular momentum becomes sufficiently large
859: (see Fig.~\ref{fig:rsh-a0}c and Fig.~\ref{fig:rsh-a099}). To
860: simplify our discussion, we will focus primarily on stable shocks
861: alone from this point. As the angular momentum of the flow $\lambda$
862: increases, the shock location tends to (but not always) shift
863: radially outward for a given energy $E_1$ as expected because the
864: centrifugal force correspondingly increases at a given radius.
865: Hence, accreting flow subject to more outward force must decelerate
866: at larger distance, forcing the shock location to shift outward, as
867: seen in our results. This trend appears to be more clear for the
868: $a/m=0.99$ case. It is also noted that in these figures larger
869: angular momentum can allow for a larger mass outflow fraction
870: $f_{\dot{M}}$.
871:
872: In our model, mass loss fraction $f_{\dot{M}}$ can vary over a wide
873: range ($0\% \lesssim f_{\dot{M}} \lesssim 95\%$) depending on the
874: shock location and angular momentum. That is, mass outflows are not
875: suppressed by relativistic effects and energy dissipation \citep[for
876: comparion with psuredo-Newtonian case with no energy dissipation,
877: see, e.g.][]{Das99}. For a fixed angular momentum, on the other
878: hand, a higher value of $f_{\dot{M}}$ is expected from stronger
879: shocks occurring in the inner regions, and this seems to be the case
880: more in rotating black hole cases (see Fig.~\ref{fig:rsh-a099}). We
881: shall explain the shaded regions in these figures in the Discussion
882: section.
883:
884:
885: \begin{figure}[t]% ------------------------------------- Figure~2
886: \epsscale{0.3} \plotone{f2a.eps}\plotone{f2b.eps}\plotone{f2c.eps}
887: \caption{$f_{\dot{M}}$ vs. $r_{\rm sh}$ for various flow energies
888: (a) $E_1=1.003$, (b) 1.004, and (c) 1.005 with $a/m=0$. Stable
889: (unstable) shocks are denoted by solid (dotted) curves. In each
890: branch, filled circles denote the maximum stable shock location also
891: corresponding to the weakest shock (smallest $n_2/n_1$), while open
892: circles denote the minimum stable shock location also corresponding
893: to the strongest shocks (largest $n_2/n_1$). A labeled value of
894: angular momentum $\lambda$ is fixed for each curve. Unbound outflow
895: solutions are indicated by shaded boxes. } \label{fig:rsh-a0}
896: \end{figure} % -------------------------------------
897:
898:
899:
900: \begin{figure}[t]% ------------------------------------- Figure~3
901: \epsscale{0.3} \plotone{f3a.eps}\plotone{f3b.eps}\plotone{f3c.eps}
902: \caption{Same as Fig.~\ref{fig:rsh-a0} but for $a/m=0.99$ case.}
903: \label{fig:rsh-a099}
904: \end{figure} % -------------------------------------
905:
906:
907:
908:
909: Figures~\ref{fig:E-a0} ($a/m=0$) and \ref{fig:E-a099} ($a/m=0.99$)
910: display the same mass loss fraction $f_{\dot{M}}$ as a function of
911: the energy loss fraction $f_E$, corresponding to the solutions in
912: Figure~\ref{fig:rsh-a0} and \ref{fig:rsh-a099}, respectively. At a
913: first glance, there is an explicit positive correlation between
914: $f_E$ and $f_{\dot{M}}$. As the angular momentum increases, maximum
915: value of $f_E$ is clearly reduced. In these figures the upper left
916: portion corresponds to the outflow solutions with larger
917: $f_{\dot{M}}$ and smaller $f_E$ (i.e., less energetic outflows)
918: while the lower right portion represents the solutions with smaller
919: $f_{\dot{M}}$ and larger $f_E$ (i.e., more energetic outflows).
920: Therefore, the results above suggest that more energetic outflowing
921: particles may be separated from a shock front when the upstream flow
922: possesses smaller angular momentum regardless of black hole spin
923: $a$. Mass loss fraction $f_{\dot{M}}$ can become as high as
924: $\lesssim 95\%$ for both $a/m=0$ and $a/m=0.99$ cases regardless of
925: energy $E_1$, whereas energy loss fraction is only $f_E < 1\%$ for
926: $a/m=0$ case but $f_E \lesssim 10\%$ for $a/m=0.99$ case. By direct
927: comparison of these figures, we note that rotation of a black hole
928: $a$ is also effective such that $d f_E / d f_{\dot{M}}(a/m=0.99) > d
929: f_E / d f_{\dot{M}}(a/m=0)$, i.e. an increase in mass loss would
930: allow more energy loss for $a/m=0.99$ case. Therefore, more
931: energetic outflows can be expected from a shock around a rotating
932: black hole.
933: %
934: %Lastly, we note that more input energy $E_1$ consistently yields
935: %smaller energy loss $f_E$ but larger mass loss $f_{\dot{M}}$; i.e.,
936: %less energetic outflows.
937:
938:
939: \begin{figure}[t]% ------------------------------------- Figure~4
940: \epsscale{0.3} \plotone{f4a.eps}\plotone{f4b.eps}\plotone{f4c.eps}
941: \caption{$f_{\dot{M}}$ vs. $f_E$ corresponding to the solutions in
942: Fig.~\ref{fig:rsh-a0} for $a/m=0$ case. } \label{fig:E-a0}
943: \end{figure} % -------------------------------------
944:
945:
946: \begin{figure}[t]% ------------------------------------- Figure~5
947: \epsscale{0.3} \plotone{f5a.eps}\plotone{f5b.eps}\plotone{f5c.eps}
948: \caption{Same as Fig.~\ref{fig:E-a0} but for $a/m=0.99$ case.}
949: \label{fig:E-a099}
950: \end{figure} % -------------------------------------
951:
952:
953:
954:
955:
956:
957: The upstream flow energy $E_1$ also affects the mass loss fraction.
958: It is noted that the energy-dependence of $f_{\dot{M}}$ is strong;
959: higher energy $E_1$ can lead to larger mass outflow fraction for a
960: given angular momentum $\lambda$ and a shock location $r_{\rm sh}$.
961: For instance, with $\lambda = 3.5$ for $a/m=0$ case (in
962: Fig.~\ref{fig:E-a0}), $f_{\dot{M}} < 20\%$ when $E_1=1.003$, $20-30
963: \%$ when $E_1=1.004$ and as large as $\sim 40\%$ when $E_1=1.005$.
964: With $\lambda=2.16$ for $a/m=0.99$ (in Fig.~\ref{fig:E-a099}),
965: $f_{\dot{M}} \sim 20-85 \%$ when $E_1=1.003$, $40-85\%$ when
966: $E_1=1.004$, and it is as high as $60-85\%$ when $E_1=1.005$. As a
967: reference, we also examine the energy-dependence of the shock
968: strength in Figure~\ref{fig:n12} where the (local) compression ratio
969: $n_2/n_1$ is plotted against $r_{\rm sh}$ for different energies:
970: $E_1=1.003,~1.004$ and $1.005$ as used before. Angular momentum
971: $\lambda$ is fixed in each case to see energy-dependence alone
972: (although different values of $\lambda$ for different spin $a$ must
973: be chosen to obtain the solutions). We find that our dissipative
974: shocks (coupled to mass loss) become stronger with decreasing
975: energy, a behavior similar to the typical types of shock formation
976: (e.g., see Lu et al. 1997 for adiabatic shocks; Lu \& Yuan 1998,
977: Fukumura \& Tsuruta 2004 for isothermal shocks).
978:
979:
980:
981: \begin{figure}[t]% ------------------------------------- Figure~6
982: \epsscale{1}\plottwo{f6a.eps}{f6b.eps} \caption{$n_2/n_1$ vs.
983: $r_{\rm sh}$ for (a) $a/m=0$ and (b) $0.99$ cases. We choose
984: $E_1=1.003,~1.004,~1.005$ from top to bottoms curves. Other
985: notations are the same as before. } \label{fig:n12}
986: \end{figure} % -------------------------------------
987:
988:
989:
990: To exclusively illustrate the black hole spin dependence $a$ of the
991: mass loss efficiency $f_{\dot{M}}$ we fix all other parameters
992: ($E_1, \lambda, r_{\rm sh}$) except for $a$. Figure~\ref{fig:spin}
993: shows $f_{\dot{M}}$ against $a$ and $f_E$ for $\lambda=3.45$ and
994: $r_{\rm sh}/m=30$. As seen in the earlier results, black hole
995: rotation alone can clearly enhance the efficiency of mass outflows
996: $f_{\dot{M}}$ from $\sim 3\%$ (for $a/m=0$) up to $\sim 95\%$ (for
997: $a/m=0.35$). On the other hand, the corresponding energy loss
998: efficiency $f_E$ remains as low as $\sim 0.02 - 0.1\%$. Note here
999: that $\lambda$ would have to be properly adjusted in order to obtain
1000: the solutions for higher black hole spin $a$.
1001:
1002:
1003:
1004:
1005: \begin{figure}[t]% ------------------------------------- Figure~7
1006: \epsscale{0.5} \plotone{f7.eps} \caption{Mass loss efficiency
1007: $f_{\dot{M}}$ as a function of black hole spin $a$ and energy loss
1008: efficiency $f_E$ for a set of fixed parameters. We choose
1009: $\lambda=3.45$ and $r_{\rm sh}/m=30$. Solution curve is projected on
1010: each plane as shown. } \label{fig:spin}
1011: \end{figure} % -------------------------------------
1012:
1013:
1014:
1015:
1016:
1017: We have chosen above some representative values for the flow energy
1018: for parametric purpose. Weakly viscous/inviscid accretion in general
1019: is a good model for some limited specific cases, like our Galactic
1020: center, for example. For such specific cases, the realistic choice
1021: of energy should be very small. From this perspective we examine to
1022: see whether low energy flows can still produce shock-driven
1023: outflows. Figure~\ref{fig:low-E} shows mass loss efficiency
1024: $f_{\dot{M}}$ as a function of $r_{\rm sh}$ for $a/m=0$. We set
1025: $E_1=1.000001$ and $\lambda=3.73$. Mass outflows can indeed be
1026: produced with $f_{\dot{M}}$ ranging from $\sim 1\%$ up to $\sim
1027: 65\%$. Both unstable and stable shocks are present as in the earlier
1028: cases but not continuously connected (no shock regions between the
1029: two). The range of shock location is much narrower in radius in this
1030: case, over which the mass loss efficiency can significantly change
1031: as mentioned above. We will discuss this more in the Discussion
1032: section.
1033:
1034:
1035: \begin{figure}[t]% ------------------------------------- Figure~8
1036: \epsscale{0.5} \plotone{f8.eps} \caption{Mass loss efficiency
1037: $f_{\dot{M}}$ vs. $r_{\rm sh}$ for $a/m=0$. We set $E_1=1.000001$
1038: and $\lambda=3.73$. Notations are the same as in
1039: Figure~\ref{fig:rsh-a0}. } \label{fig:low-E}
1040: \end{figure} % -------------------------------------
1041:
1042:
1043:
1044: The major correlations we find among the primary parameters are
1045: summarized in Table~\ref{tab:tbl-1}. Table~\ref{tab:tbl-2} shows
1046: various correlations with shock strength. Compression ratio
1047: $n_2/n_1$ is strongly correlated with ($r_{\rm sh}, f_E,
1048: f_{\dot{M}}$); stronger shocks are expected in regions very close to
1049: the central engine particularly around a rotating black hole. Strong
1050: shocks in principle are accompanied by high energy and mass loss
1051: fractions.
1052:
1053:
1054: \begin{deluxetable}{c|cccc}%[h]% ------------------------------- Table~1
1055: \tabletypesize{\scriptsize} \tablecaption{Correlation among primary
1056: quantities. \label{tab:tbl-1}} \tablewidth{0pt} \tablehead{Parameter
1057: & $r_{\rm sh}$ & $f_E$ & $f_{\dot{M}}$ & $n_2/n_1$ } \startdata
1058: $\lambda$ & $+$ & $--$ & $++$ & $\leftrightarrow$ \\
1059: $E_1$ & $\leftrightarrow$ & $--$ & $+$ & $--$ \\
1060: \enddata
1061: \tablecomments{Strong positive (negative) correlation is denoted by
1062: $++$ ($--$) while relatively weak positive (negative) correlation is
1063: denoted by $+$ ($-$). No significant correlation is shown by
1064: $\leftrightarrow$.}
1065: \end{deluxetable}
1066:
1067:
1068:
1069: \begin{deluxetable}{c|ccc}%[h]% ------------------------------- Table~2
1070: \tabletypesize{\scriptsize} \tablecaption{Correlation with shock
1071: strength $n_2/n_1$ among primary quantities. \label{tab:tbl-2}}
1072: \tablewidth{0pt} \tablehead{Parameter & $r_{\rm sh}$ & $f_E$ &
1073: $f_{\dot{M}}$ } \startdata
1074: $n_2/n_1$ & $--$ & $++$ & $++$ \\
1075: \enddata
1076: \tablecomments{Notations are the same as in Table~\ref{tab:tbl-1}.}
1077: \end{deluxetable}
1078:
1079:
1080: To sum up, our results show that strongest stable shocks generally
1081: develop at the smallest radii (closer to the black hole),
1082: accompanied by the largest mass loss $f_{\dot{M}}$ (and the largest
1083: energy loss $f_E$ as well). The rotation of the black hole
1084: apparently amplifies the shock strength by more than a factor of
1085: two, also extending the outflowing site significantly inward (i.e.,
1086: $r_{\rm sh}/m \gtrsim 2-3$ when $a/m=0.99$ while $r_{\rm sh}/m
1087: \gtrsim 12$ when $a/m=0$). We will make some implications of the
1088: obtained shock-outflow solutions in the last section, \S 4.
1089:
1090:
1091:
1092:
1093: \subsection{Global Accreting Flows}
1094:
1095: Samples of the global run of the physical parameters of accretion
1096: flows that include shocks are given in Figures~\ref{fig:global-1}
1097: ($a/m=0$) and \ref{fig:global-2} ($a/m=0.99$). In
1098: Tables~\ref{tab:tbl-3} and \ref{tab:tbl-4} we provide the specifics
1099: of the shocks associated respectively with the flows of
1100: Figures~\ref{fig:global-1} and \ref{fig:global-2}. These plots have
1101: been made assuming that all flows accrete at 1\% of the Eddington
1102: rate, i.e. $\dot{m} \equiv \dot{M} / \dot{M}_{\rm Edd} = 0.01$ and
1103: that the black hole mass is $m = 10^7 \Msun$, typical of Seyfert
1104: nuclei. Each panel shows (a) the radial velocity $|u^r|$, (b) the
1105: angular velocity $\Omega$, (c) the electron scattering optical depth
1106: defined by $\tau \equiv n \sigma_T H$, (d) flow temperature $T$ [K],
1107: (e) the density $\rho_0$ [g~cm$^{-3}$], and (f) the ratio of the
1108: vertical scale-height to the radius $H/r$. Vertical lines denote the
1109: positions of shocks connecting the upstream and downstream values of
1110: the corresponding quantities of each flow. Dotted curve in panel (b)
1111: shows the Keplerian angular velocity $\Omega_{K}$. Note that for
1112: clarity purposes we only show three of the representative shock
1113: solutions obtained, although the shock can occur at any radius
1114: between the outermost and innermost solutions (with different values
1115: of $f_{\dot{M}}$ and $f_E$).
1116: %
1117: %
1118: %Sample solutions for shock-included, global accreting flows are
1119: %displayed in Figures~\ref{fig:global-1} and \ref{fig:global-2}.
1120: %
1121: %\assume the radiative efficiency to be 1\% {\sf ???}[i.e.,
1122: %$\eta_{\rm eff} \equiv L_{Edd}/(m_p c^2) = 0.01$ where $L_{Edd}$ is
1123: %the Eddington luminosity]. We also take dimensionless mass-accretion
1124: %rate to be $\dot{m} \equiv \dot{M}/\dot{M}_{\rm Edd} = 0.01$,
1125: %relevant for typical Seyfert nuclei.
1126: %
1127: %
1128: %Figure~\ref{fig:global-1} shows a global solution for the $a/m=0$
1129: %case with $E_1=1.003, \ell=3.5$. We find $r_{sh}/m=21$ and
1130: %$f_{\dot{M}}=0.064$. Each panel shows (a) the radial speed
1131: %$|v^r/c|$, (b) the angular velocity $\Omega$, (c) the scattering
1132: %optical depth $\tau \equiv n \sigma_T H$, (d) teh flow temperature
1133: %$\log T$ [K], (e) the number density $\log n$ [cm$^{-3}$], and (f)
1134: %the normalized vertical scale-height $H/r$. Vertical line denote a
1135: %shock connecting upstream and downstream flows. Dotted curves
1136: %represent the sound speed in panel (a) while Keplerian angular
1137: %velocity $\Omega_K$ in panel (b).
1138:
1139: %[Demos: modified below by KF as of 3/21]
1140:
1141: We find that in both the non-rotating (Fig.~\ref{fig:global-1}) and
1142: rotating (Fig.~\ref{fig:global-2}) cases the upstream flow density
1143: scales as $\rho_{0,1}(r) \sim r^{-3/2}$, as in the ADAF self-similar
1144: solution. On the other hand, as we explain below, the downstream
1145: flow density generally has a slightly steeper power-law slope,
1146: $\rho_{0,2}(r) \sim r^{-3/2} - r^{-3}$ unless the shock location is
1147: too close to the horizon. As seen in Figures~\ref{fig:global-1}a and
1148: \ref{fig:global-2}a, this is primarily because of the decreasing
1149: (radial) downstream flow speed $|u^r_2(r)|$, whose feature is more
1150: obvious around a rotating black hole. After the shock transition,
1151: frame-dragging of the rotating black hole forces the accreting flow
1152: to accelerate more in the toroidal direction (i.e., corotate) rather
1153: than radial direction. Under adiabatic assumption, at the same time,
1154: the flow temperature continues to rise (Fig.~\ref{fig:global-1}d and
1155: \ref{fig:global-2}d) due to $p dV$ compression of the flow. Since
1156: $c_s \propto T^{1/2}$, sound speed also increases. The resulting
1157: (thermal) pressure gradient ($dP/dr<0$) prevents the incoming flow
1158: from speeding up and in fact decelerate the downstream flow (for a
1159: while) until the flow falls deep inside the gravitational potential
1160: well to radially speed up again. Consequently, the fluid motion is
1161: governed mainly by increasing toroidal velocity rather than radial
1162: one (compare Fig.~\ref{fig:global-1}a with
1163: Fig.~\ref{fig:global-2}a). This results in a larger radial velocity
1164: gradient $d |u^r_2| / dr>0$ in Figure~\ref{fig:global-2}a.
1165:
1166:
1167: In differentially-rotating flows, the Keplerian frequency increases
1168: with decreasing $r$ even more rapidly than the sound speed, leading
1169: to a decreasing scale-height $H$ [see equation~(\ref{eq:H})]. The
1170: dependence of these two quantities, $\Omega_K$ and $c_s$, on the
1171: radius in the downstream flow could lead to geometrically slim/thick
1172: all the way to the horizon [see, e.g., $H/r \gtrsim 0.2 - 0.3$ in
1173: Figs.~\ref{fig:global-1}f and \ref{fig:global-2}f].
1174:
1175:
1176: Recalling $\rho_0 \propto 1 /(r H |u^r|)$ from
1177: equation~(\ref{eq:mdot1}), the above fact that both $|u^r_2|$ and
1178: $H$ decrease allows a very steep downstream density profile
1179: $\rho_{0,2}(r)$ after the shock transition (and this is more so in a
1180: rotating black hole case because of larger drop in $|u^r_2|$ as
1181: mentioned earlier). The downstream flows become slightly more opaque
1182: to electron scattering because of the shock compression, yet they
1183: continue to remain optically-thin ($\tau < 1$).
1184: %
1185: %
1186: %{\sf [ Keigo the explanation you give does not quite work. The
1187: %dependence of the density on the radius has more do to with the
1188: %radial velocity which you do not show in the figs. the speed of
1189: %sound is proportional to the square root of the temperature which
1190: %increases with $r$ as expected. However the radial velocity
1191: %DECREASES for a while after the shock and that is the gives rise to
1192: %the much steeper density profile. This seems to be the case mainly
1193: %in the $a=0.99$ case. We need to rephrase this section. the velocity
1194: %presumably decreases because of the increase in the temperature due
1195: %to pdV compression of the flow. Also the geometric cross section of
1196: %the flow maybe different than spherical near the horizon]}.
1197: %
1198: %
1199: %Because the radial speed (equivalent to Mach number {\sf [not
1200: %really; as i pointed out above, the sound speed also increases
1201: %inward so that the decrease in the Mach number is partly due to the
1202: %increase in the temperature. it is this that apparently chokes the
1203: %flow and slows it down until eventually the infinite grav. potential
1204: %of the BH wins over and sucks the fluid in}]) after the shock tends
1205: %to decrease (for a while) whereas Keplerian angular velocity keeps
1206: %rising, density profile can be significantly steep in the downstream
1207: %region (sound speed only slightly increases).
1208: %
1209: %
1210:
1211:
1212: \begin{deluxetable}{c|cccc}%[h]% ------------------------------- Table~3
1213: \tabletypesize{\scriptsize} \tablecaption{Sets of Parameters for the
1214: Global Solutions in Figure~\ref{fig:global-1}. \label{tab:tbl-3}}
1215: \tablewidth{0pt} \tablehead{Model & $r_{\rm sh}/m$ & $n_2/n_1$ &
1216: $f_E$ (\%) & $f_{\dot{M}}$ (\%) } \startdata
1217: (1) & $15$ & $4.2$ & $0.319$ & $52.4$ \\
1218: (2) & $31$ & $3.8$ & $0.136$ & $47.3$ \\
1219: (3) & $54$ & $3.1$ & $0.00253$ & $43.4$ \\
1220: \enddata
1221: \tablecomments{Note that $a/m=0,~E_1=1.003,~\lambda=3.6$.}
1222: \end{deluxetable}
1223:
1224:
1225:
1226: \begin{deluxetable}{c|cccc}%[h]% ------------------------------- Table~4
1227: \tabletypesize{\scriptsize} \tablecaption{Sets of Parameters for the
1228: Global Solutions in Figure~\ref{fig:global-2}. \label{tab:tbl-4}}
1229: \tablewidth{0pt} \tablehead{Model & $r_{\rm sh}/m$ & $n_2/n_1$ &
1230: $f_E$ (\%) & $f_{\dot{M}}$ (\%) } \startdata
1231: (4) & $2.4$ & $10.2$ & $8.3$ & $89$ \\
1232: (5) & $10$ & $7.9$ & $2.6$ & $51.0$ \\
1233: (6) & $78$ & $2.73$ & $0.002$ & $17.6$ \\
1234: \enddata
1235: \tablecomments{Note that $a/m=0.99,~E_1=1.003,~\lambda=2.16$.}
1236: \end{deluxetable}
1237:
1238:
1239: \begin{figure}[h]% ------------------------------------- Figure~9
1240: \epsscale{1}\plotone{f9.eps} \caption{Global shock-included
1241: solutions, (1)-(3), for $a/m=0$ case. See Table~\ref{tab:tbl-3} for
1242: the parameters. (a) radial velocity $|u^r|$, (b) angular velocity
1243: $\Omega$, (c) scattering optical depth $\tau$, (d) temperature $T$,
1244: (e) density $\rho_0$, and (f) aspect ratio $H/r$. Vertical lines
1245: denote shocks while filled circles indicate outer/inner critical
1246: radii. In (b) Keplerian profile $\Omega_{K}$ is denoted by a dotted
1247: curve. In (e) dotted gray lines show the slopes of $-3/2$ and $-3$.
1248: } \label{fig:global-1}
1249: \end{figure} % -------------------------------------
1250:
1251:
1252:
1253: \begin{figure}[h]% ------------------------------------- Figure~10
1254: \epsscale{1}\plotone{f10.eps} \caption{Same as
1255: Figure~\ref{fig:global-1} but $a/m=0.99$ case. See
1256: Table~\ref{tab:tbl-4} for the parameters.} \label{fig:global-2}
1257: \end{figure} % -------------------------------------
1258:
1259:
1260: \section{Discussion \& Conclusions}
1261:
1262: In this work we have examined the structure of accretion flows that
1263: includes shocks of more general character than those discussed so
1264: far in the literature \citep[e.g.,][]{Cha90}. In particular we have
1265: examined whether it is possible to have flows with shocks at which
1266: part of the mass and/or energy fluxes are lost and do not
1267: participate in the shock transition (this maybe the case in
1268: multidimensional shocks or in shocks that involve acceleration of
1269: particles that escape from the shock region). Such generalized
1270: shocks obey jump conditions more general than those of
1271: Rankine-Hugoniot and it is not {\it a priori} certain if accretion
1272: flows can allow for such shocks. We have also examined the degree of
1273: mass and energy loss allowed that are at the same time consistent
1274: with the continuation of the downstream flow through another sonic
1275: transition onto the black hole. In this respect we have focused our
1276: study on shocks in which the energy per particle for those escaping
1277: the shock transition is greater than that of the local gravitational
1278: potential. In such a case one can argue that these particles can
1279: escape to infinity producing the jets/winds observed in many
1280: accretion-powered systems. We have found that there are indeed flows
1281: with global parameters (i.e. $\lambda$ and $E$) that allow for such
1282: shocks. On the other hand, we have also obtained solutions at which
1283: the energy per particle of the escaping matter is smaller than the
1284: local gravitational potential, indicating that in all likelihood
1285: these particles will stagnate and eventually join the rest of the
1286: flow onto the compact object.
1287: %
1288: %{\sl \{KF: We examined the mass loss efficiency that are coupled to
1289: %the shock solutions for various flows. Most importantly, in order
1290: %for the outflowing particles to leave the shocked flow surface, it
1291: %must posses sufficiently large kinetic energy exceeding some
1292: %threshold which is determined by equating outward force against
1293: %inward gravitational pull. Otherwise, the released particles might
1294: %stagnate around the shocked flow surface or is advected with the
1295: %downstream flow, not being ejected away\}}.
1296: %
1297: However, a deeper understanding of the precise mechanisms that can
1298: lead to the particles that do not participate in the shock
1299: transition, as we conjectured above, requires the knowledge of the
1300: detailed microphysics of the shock (i.e. the acceleration
1301: efficiency, the fraction of energy put into relativistic particles
1302: at the shock and their transport in and around the shock geometry);
1303: these are beyond the scope of the present paper.
1304:
1305:
1306: To get a crude estimate of the energetics of these outflowing
1307: particles, we compute a threshold energy of the escaping particles
1308: under a set of assumed parameters. The total energy flux carried by
1309: the escaping particles (of energy $\Delta E$) is given by $\Delta E
1310: \times 4\pi r_{\rm sh} \bar{H} \bar{n} \bar{u}^r$ [erg~s$^{-1}$]
1311: where $\bar{H} \equiv (H_1+H_2)/2, \bar{n} \equiv (n_1+n_2)/2$ and
1312: $\bar{u}^r \equiv (u^r_1+u^r_2)/2$ are evaluated at the shock
1313: location ($r=r_{\rm sh}$). The minimum kinetic energy necessary for
1314: the separated particles (of mass outflow rate $\Delta \dot{M}$) to
1315: escape the bulk flow is given by $\Delta \dot{M} v_{c}^2 /2$
1316: [erg~s$^{-1}$] where $v_{c}$, to be determined below, is the
1317: required critical velocity of the particles. By equating these, we
1318: solve for $v_{c}$ to obtain
1319: \begin{eqnarray}
1320: v^2_{c} \sim 8 \pi r_{\rm sh} \bar{H} \bar{n} \bar{u}^r m_p c^2
1321: \Delta E / \Delta \dot{M} \ .
1322: \end{eqnarray}
1323: For the outflowing particles to be effectively decoupled from the
1324: bulk flow, kinetic energy associated with this critical velocity
1325: should exceed or be at least comparable to the corresponding
1326: gravitational binding energy at the shock location. This condition
1327: reads
1328: \begin{eqnarray}
1329: \frac{1}{2} v_c^2 \gtrsim \frac{G m}{r_{\rm sh}} \ ,
1330: \label{eq:condition-1}
1331: \end{eqnarray}
1332: or
1333: \begin{eqnarray}
1334: \Delta E \gtrsim E_c \ , \label{eq:condition-2}
1335: \end{eqnarray}
1336: where $E_c \equiv G m \Delta \dot{M} / (4 \pi r_{\rm sh}^2 \bar{H}
1337: \bar{n} \bar{u}^r m_p c^2)$.
1338: %
1339: %
1340: %The energy flux carried with the escaping particles is given by
1341: %$\Delta E \times 4\pi r_{\rm sh} \bar{H} \bar{n} \bar{u}^r$
1342: %erg~s$^{-1}$ where $\bar{H} \equiv (H_1+H_2)/2, \bar{n} \equiv
1343: %(n_1+n_2)/2$ and $\bar{u}^r \equiv (u^r_1+u^r_2)/2$ are evaluated at
1344: %the shock location $r_{\rm sh}$. The corresponding mass-energy flux
1345: %of the escaping particles is $\Delta \dot{M} v^2_{esc}$ erg~s$^{-1}$
1346: %where $v_{esc}$ is the escape velocity of the particles. By equating
1347: %these, we obtain
1348: %%
1349: %\begin{eqnarray}
1350: %v^2_{esc} \sim 4\pi r_{\rm sh} \bar{H} \bar{n} \bar{u}^r m_p c^2
1351: %\Delta E / \Delta \dot{M} \ .
1352: %\end{eqnarray}
1353: %
1354: %{\sf Then the condition for these particles to escape reads}
1355: %
1356: %\begin{eqnarray}
1357: %\Delta E \gtrsim E_c \ ,
1358: %\end{eqnarray}
1359: %%
1360: %where $E_c \equiv G m \Delta \dot{M} / (2\pi r_{\rm sh}^2 \bar{H}
1361: %\bar{n} \bar{u}^r m_p c^2)$.
1362: %
1363: %
1364: Since all these parameters ($\bar{H}, \bar{n}, \bar{u}^r$) are
1365: functions of shock location $r_{\rm sh}$ which depends on
1366: $f_{\dot{M}}$, $E_c$ is obtained once we specify $f_{\dot{M}}$. Now,
1367: we can review the shock solutions obtained earlier from this new
1368: perspective: In Figures~\ref{fig:rsh-a0}-\ref{fig:E-a099} the
1369: allowed shock solutions satisfying $\Delta E \ge E_c$, which is
1370: relevant for outflows that may escape the shocked flow, are
1371: displayed by shaded regions, while the rest of the solutions
1372: correspond to the outflows presumably bound to the bulk flow.
1373: According to these estimates, an outflow that expels a certain
1374: amount of the preshock accreting matter is possible, over a rather
1375: broad range of radii in the vicinity of the central object;
1376: %
1377: %{\sl \{ In this estimate certain degrees of outflows could be
1378: %possible over a wide range of radius not too far from the central
1379: %engine\} }
1380: %
1381: up to nearly 10\% of matter may participate in producing seed
1382: particles for jets/winds within $\sim 20 - 40$ gravitational radii
1383: around a non-rotating black hole, while as much as 50\% of preshock
1384: matter could contribute to forming jets/winds within $\sim 2 - 30$
1385: gravitational radii around a rapidly-rotating black hole. More
1386: interestingly, our solutions also allow for dissipative shocks ($f_E
1387: \ne 0$) with negligible mass loss ($f_{\dot{M}} \sim 0$). Such
1388: solutions that involve the loss of infinitesimal mass and finite
1389: amount of energy can lead to relativistic outflows from the
1390: corresponding shocks.
1391: %
1392: %{\sl \{ lost and These solutions correspond to a case where all the
1393: %liberated energy is transferred to a small population of nonthermal
1394: %particles \} }.
1395: %
1396: Perhaps, outflows via this type of shocks may produce kinematically
1397: strong jets (with high Lorentz factor) rather than moderate velocity
1398: winds. Thus, shock-driven outflows may provide clues on the origin
1399: of the jets/winds from the inner accretion regions, for instance, in
1400: some active galaxies, e.g., M87 and 3C120.
1401:
1402: %\hspace{10in}
1403:
1404: In this simple model, shocks in principle could drive axisymmetric
1405: outflowing matter.
1406: %
1407: Near the base of the outflow, as illustrated in
1408: Figure~\ref{fig:schematic}, its shape would be that of a hollow
1409: cone. However, as it propagates to longer distances we do not expect
1410: that this shape would be preserved all the way.
1411: %
1412: %Because we treat the thickness of a shock as infinitesimally small
1413: %(i.e. mathematical discontinuity), it is difficult to discuss the
1414: %actual (particle) acceleration mechanisms at a shock front,
1415: %necessary for constraining the geometry of the outflow.
1416: %
1417: Within the framework of our current scenario it can be speculated
1418: that the most of the outflows discussed here could be in a diffuse
1419: form, while some strong outflows with low $f_{\dot{M}}$ and high
1420: $f_E$ might possess a relatively more collimated geometry (perhaps
1421: in the poloidal direction) due to its high Lorentz factor. In the
1422: presence of large-scale magnetic fields the dissipated plasma would
1423: stream along the field lines, producing collimated outflows.
1424: However, it is beyond the scope of our model to further discuss the
1425: exact geometry of the outflows.
1426:
1427: %When efficient shock acceleration is in operation at a shock front,
1428: %the estimated velocity $\bar{u}^r$ above could be much larger. In
1429: %that case, the kinetic energy of the outflows would easily exceed
1430: %the threshold value in the present estimate. Thus our estimate here
1431: %may put a tighter constraint on the allowed shock-related quantities
1432: %presented in Figures~\ref{fig:rsh-a0}-\ref{fig:E-a099}.
1433:
1434:
1435: It should be noted that for a given $\lambda$ there exists the
1436: degeneracy of shock-outflow solutions: i.e., for a single value of
1437: angular momentum $\lambda$ there is a finite range of possible shock
1438: locations allowed ($r^{\rm min}_{\rm sh} < r_{\rm sh} < r_{\rm
1439: sh}^{\rm max}$) that corresponds respectively to outflows with
1440: $f_{\dot{M}}^{\rm min} \le f_{\dot{M}} \le f_{\dot{M}}^{\rm max}$
1441: where the indices ``min" and ``max" denote minimum and maximum
1442: values, respectively. In the context of our model, one can predict
1443: the shock location $r_{\rm sh}$ by the (observational) knowledge of
1444: $(\lambda, E_1, f_{\dot{M}})$, although the obtained parameter space
1445: can be altered by additional physical ingredients (e.g., viscosity,
1446: magnetic fields, for instance).
1447:
1448:
1449: Another issue to be addressed is the fact that in general critical
1450: points are not necessarily equivalent to sonic points depending on
1451: the flow geometry and equation of state used
1452: \citep[e.g.,][]{Das07b}. That is, there could be a potential danger
1453: that a false shock could occur in subsonic region between a critical
1454: radius and an actual sonic radius, in which case the obtained shock
1455: would be unphysical. To ensure that a physically valid shock forms
1456: in supersonic regions, we have checked the validity of our shock
1457: solutions by computing a three-velocity component of the flow
1458: measured by a suitable local observer at the shock location. We
1459: calculated the (radial) flow velocity $v^r$ measured by a locally
1460: stationary observer in the corotating reference frame
1461: \citep[e.g.,][]{Lu86, Lu98}
1462: \begin{eqnarray}
1463: v^r \equiv \frac{u_r u^r}{1+u_r u^r} \ , \label{eq:vr}
1464: \end{eqnarray}
1465: and compared this velocity to the local sound velocity $c_s$ given
1466: by equation~(\ref{eq:sound}). All the shock solutions presented here
1467: are found to form in supersonic regions (i.e., $|v^r| > c_s$ at
1468: $r=r_{\rm sh}$).
1469:
1470:
1471:
1472: We showed that low energy flows can still produce mass outflows with
1473: suitable angular momenta. Although continuous accreting solutions
1474: (i.e. shock-free solutions) are persistently present even for
1475: smaller energy $E_1$ (as expected), we do not find shocked flow
1476: solutions (or perhaps they are present but in much narrower
1477: parameter space). It can be speculated that maybe by prohibiting our
1478: jump condition in energy (energy dissipation at a shock front) we
1479: may obtain shock-driven outflows for smaller flow energy $E_1$,
1480: which could be a case similar to \citet{Das99} where they considered
1481: little energy loss and found mass outflows in pseudo-Newtonian
1482: geometry. However, powerful outflows should carry away a significant
1483: amount of (kinetic) energy. Hence, outflow solutions should in
1484: principle be coupled to energy dissipation as treated here.
1485:
1486:
1487: We have explored a coupling between shock solutions in accretion and
1488: mass/energy losses (fractions) under a scenario that the
1489: shock-driven outflowing particles may participate in forming a base
1490: of jets/winds. For various flow parameters with a given black hole
1491: spin, we have shown, by steady-state, axisymmetric hydrodynamic
1492: calculations, that the dissipative shock front could be a plausible
1493: site where a fraction of the accreting matter can be decoupled as
1494: jets/winds from the bulk accretion flows.
1495:
1496:
1497:
1498: \acknowledgments
1499:
1500: K.F. thanks John Cannizzo for his useful comments. We are also
1501: grateful to the anonymous referee for several useful suggestions and
1502: comments that clarified the manuscript. This research was supported
1503: in part by an appointment to the NASA Postdoctoral Program at the
1504: Goddard Space Flight Center, administered by Oak Ridge Associated
1505: Universities through a contract with NASA.
1506:
1507:
1508:
1509: %\clearpage
1510:
1511:
1512: \begin{thebibliography}{999}
1513: %\bibitem[Becker et al.(2001)]{Becker01} Becker, P. A., \& Subramanian, P., \& Kazanas,
1514: %D. 2001, \apj, 552, 209
1515: \bibitem[Begelman et al.(1984)]{BBR84} Begelman, M. C., Blandford, R. D., \& Rees, M. J. 1984, Rev. Mod.
1516: Phys., 56, 255
1517: \bibitem[Blandford \& Begelman(1999)]{BB99} Blandford, R. D., \& Begelman, M. C. 1999, \mnras, 303,
1518: L1
1519: %\bibitem[Blandford \& Ostriker(1978)]{Blandford78} Blandford, R. D., \& Ostriker, J. P. 1978,
1520: %\apj, 221, L29
1521: \bibitem[Blandford \& Payne(1982)]{BP82} Blandford, R. D., \& Payne, D. G. 1982, \mnras, 199,
1522: 883
1523: \bibitem[Chakrabarti(1990)]{Cha90} Chakrabarti, S. K. 1990, Theory of Transonic Astrophysical Flows (World Scientific,
1524: Singapore)
1525: \bibitem[Chakrabarti(1996)]{Cha96} Chakrabarti, S. K. 1996, \mnras, 283, 325
1526: \bibitem[Chakrabarti(1999)]{Cha99} Chakrabarti, S. K. 1999, \mnras, 351, 185
1527: \bibitem[Chakrabarti \& Das(2004)]{Cha04} Chakrabarti, S. K., \& Das, S. 2004, \mnras, 349, 649
1528:
1529: \bibitem[Contopoulos \& Lovelace(1994)]{Contopoulos94} Contopoulos, J., \& Lovelace, R. V. E. 1994, \apj, 429,
1530: 139
1531: \bibitem[Contopoulos \& Kazanas(1995)]{Contopoulos95} Contopoulos, J., \& Kazanas, D. 1995, \apj, 441, 521
1532: \bibitem[Das \& Chakrabarti(1999)]{Das99} Das, T. K., \& Chakrabarti, S. K. 1999, Classical
1533: Quantum Gravity, 16, 3879
1534: \bibitem[Das(2000)]{Das00} Das, T. K. 2000, \mnras, 318, 294
1535: \bibitem[Das \& Chakrabarti(2007)]{Das07} Das, T. K., \& Chakrabarti, S. K. 2007, \mnras, 374,
1536: 729
1537: \bibitem[Das(2007)]{Das07b} Das, T. K. 2007 (astro-ph/0704.3618)
1538: \bibitem[Fender et al.(2004)]{Fender04} Fender, R. P., Belloni, T. M., \& Gallo, E.
1539: 2004 , \mnras, 355, 1105
1540: \bibitem[Fukumura \& Tsuruta(2004)]{FT04} Fukumura, K., \& Tsuruta, S. 2004, \apj, 611, 964
1541: \bibitem[Fukumura et al.(2007)]{Fukumura07} Fukumura, K., Takahashi, M., \& Tsuruta, S. 2007,
1542: \apj, 657, 415
1543: \bibitem[Junor et al.(1999)]{Junor99} Junor, W., Biretta, J. A., \& Livio, M. 1999, Nature, 401, 891
1544: \bibitem[Kataoka et al.(2007)]{Kataoka07} Kataoka et al., 2007, (astro-ph/0612754)
1545: \bibitem[Protheroe \& Kazanas(1983)]{Kazanas83} Protheroe, R. J., \& Kazanas, D. 1983, \apj, 265,
1546: 620
1547: \bibitem[Kazanas \& Ellison(1986)]{Kazanas86} Kazanas, D., \& Ellison, D. C. 1986, \apj, 304, 178
1548: \bibitem[Koide et al.(1999)]{Koide99} Koide, S., Shibata, K., \& Kudoh, T. 1999, \apj, 522,
1549: 727
1550: \bibitem[K\"{o}nigl \& Kartje(1994)]{Konigl94} K\"{o}nigl, A., \& Kartje, J. F. 1994, \apj, 434, 446
1551: \bibitem[Le \& Becker(2004)]{Becker04} Le, T., \& Becker, P. A. 2004, \apj, 617,
1552: L25
1553: \bibitem[Le \& Becker(2005)]{Becker05} Le, T., \& Becker, P. A. 2005, \apj, 632, 476
1554: \bibitem[Livio(1999)]{Livio99} Livio, M. 1999, Phys. Rep., 311, 225
1555: \bibitem[Lu et al.(1997)]{Lu97} Lu, J.-F., Yu, K. N., Yuan, F., \&
1556: Young, E. C. M. 1997, \aap, 321, 665
1557: \bibitem[Lu(1986)]{Lu86} Lu, J.-F. 1986, General Relativity and Gravitation,
1558: 18,?
1559:
1560: \bibitem[Lu \& Yuan(1998)]{Lu98} Lu, J.-F., \& Yuan, F. 1998, \mnras, 295, 66
1561: \bibitem[Lu et al.(1999)]{Lu99} Lu, J.-F., Gu, W.-M., \& Yuan, F. 1999, \apj, 523, 340
1562: \bibitem[Mirabel et al.(1999)]{Mirabel99} Mirabel, I. F., \& Rodriguez, L. F. 1999, ARA\&A, 37, 409
1563: \bibitem[Manmoto et al.(1997)]{Manmoto97} Manmoto, T., Mineshige, S., \& Kusunose, M.
1564: 1997, \apj, 476, 49
1565: \bibitem[Narayan \& Yi(1994)]{Narayan94} Narayan, R., \& Yi, I. 1994, \apj, 428, L13
1566: \bibitem[Nishikawa et al.(2005)]{Nishikawa05} Nishikawa, K.-I., Richardson, G., Koide, S.,
1567: S hibata, K., Kudoh, T., Hardee, P., \& Fishman, G. J. 2005, \apj,
1568: 625, 60
1569:
1570: \bibitem[Pelletier \& Pudritz(1992)]{Pelletier92} Pelletier, G., \& Pudritz, R. E. 1992, \apj, 394, 117
1571:
1572: \bibitem[Subramanian et al.(1999)]{Subramanian99} Subramanian, P., Becker, P. A., \& Kazanas, D. 1999,
1573: \apj, 523, 203
1574: \bibitem[Sponholz \& Molteni(1994)]{Sponholz94} Sponholz, H., \& Molteni, D.
1575: 1994, \mnras, 271, 233
1576: \bibitem[Takahashi et al.(2002)]{Takahashi02} Takahashi, M., Rilett, D., Fukumura, K., \& Tsuruta, S. 2002, \apj, 572,
1577: 950
1578: \bibitem[Vlahakis et al.(2000)]{Vlahakis00} Vlahakis, N., Tsinganos, K., Sauty, C., \& Trussoni, E.
1579: 2000, \mnras, 318, 417
1580: \bibitem[Yang \& Kafatos(1995)]{Yang95} Yang, R., \& Kafatos, M. 1995, \aap, 295, 238
1581: \end{thebibliography}
1582:
1583:
1584:
1585:
1586:
1587: \end{document}
1588: