0806.4543/ms.tex
1: \documentclass[emulateapj,epsfig]{article}
2: \usepackage[onecolumn]{emulateapj}
3: \usepackage[]{epsfig}
4: \usepackage{graphicx}
5: \def\perc{\%\,\,}
6: \hoffset -1.5truecm \lefthead{Menci et al.}
7: \righthead{}
8: \def\newline{\hfil\break}
9: \begin{document}
10: \title{The Blast Wave Model for AGN Feedback: Effects on AGN Obscuration}
11: \author{N. Menci$^1$, F. Fiore$^1$, S. Puccetti$^{1,2}$, A. Cavaliere$^3$}
12: \affil{$^1$INAF - Osservatorio Astronomico di Roma, via di Frascati
13: 33, I-00040 Monteporzio, Italy}
14: \affil{$^2$ ASI SDC, c/o ESRIN via Galileo Galilei, 00044 Frascati, Italy}
15: \affil{$^3$ Dip. Fisica,
16: Universita'di Roma Tor Vergata, via Ricerca Scientifica 1, 00133, Roma, Italy}
17: 
18: \smallskip
19: 
20: \begin{abstract}
21: We compute the effect of the galactic absorption on AGN emission in a
22: cosmological context by including a physical model for AGN feeding and feedback
23: in a semi-analytic model of galaxy formation. This is based on galaxy
24: interactions as triggers for AGN accretion, and on expanding blast waves as a
25: mechanism to propagate outwards the AGN energy injected into the interstellar
26: medium at the center of galaxies.
27: 
28: We first test our model against the observed number density of AGNs with
29: different intrinsic luminosity as a function of redshift. The model
30: yields a ''downsizing'' behavior in close agreement with
31: the observed one for $z\lesssim  2$. At higher redshifts, the model
32: predicts an overall abundance of AGNs (including Compton-thick sources)
33: larger than the observed Compton-thin sources by a factor $\approx 2$ for
34: $z\gtrsim 2$ and $L_X\leq 10^{44}$ erg/s. Thus, we expect that at such
35: luminosities and redshifts about $1/2$ of the total AGN population is contributed
36: by Compton-thick sources.
37: 
38: We then investigate the dependence of the absorbing column density $N_H$
39: associated to cold galactic gas (and responsible for the Compton-thin component
40: of the overall obscuration) on the AGN luminosity and
41: redshift. We find that the absorbed fraction of AGNs with $N_H\geq 10^{22}$
42: cm$^{-2}$ decreases with luminosity for  $z\leq 1$; in addition, the total
43: (integrated over luminosity) absorbed fraction increases with redshift up to
44: $z\approx 2$, and saturates to the value $\approx 0.8$ at higher redshifts. Finally, we predict the luminosity dependence of the absorbed fraction of AGNs
45: with $L_X\leq 3\,10^{44}$ erg/s to weaken with increasing redshift.
46: 
47: We compare our results with recent observations,
48: and discuss their implications in the context of cosmological models
49: of galaxy formation.
50: 
51: \end{abstract}
52: 
53: \keywords{galaxies: active --- galaxies: formation --- galaxies: evolution}
54: 
55: \section{Introduction}
56: 
57: The accretion which built up the supermassive black holes (SMBHs) now
58: hosted in many local galaxies is widely thought to be associated with a
59: sequence of output episodes observed as Active Galactic Nuclei
60: (AGNs). Since only a minority ($\sim 10^{-2}$, see Richstone et al.  1998)
61: of local galaxies host a currently active AGN, the corresponding  lifetimes
62: are estimated to be close to $\tau\sim 10^8$ yrs. Several
63: observations indicate the accretion episodes to be fundamentally
64: related to the galaxy growth; one such indication is provided by the
65: narrow scatter in the observed correlation of the SMBH mass $M_{BH}$
66: with its host galaxy mass $M$, when the former is in the range
67: $10^7\leq M_{BH}/M_{\odot}\leq 10^9$ (Ferrarese \& Merritt 2000,
68: Gebhardt et al. 2000). The emissions of AGNs thus may conceivably
69: constitute a probe for the history of accretion and growth of SMBHs,
70: and for its interplay with the galaxy building process.
71: 
72: The co-evolution of galaxies and AGNs and their so called
73: ``downsizing'' (faster evolution for more luminous objects)
74: depends also on feedback between nuclear and other galactic
75: activities. In fact, the density of the high luminosity QSOs is peaked at high
76: redshift and declines strongly toward us; similarly, massive galaxies
77: are characterized by a star formation history peaked at high
78: redshifts. Luminous AGNs are efficient in "sterilizing" their massive
79: host galaxies by heating the interstellar matter through winds, shocks,
80: and high energy radiation, see Granato et al. (2004);  
81: Murray, Quataert, Thompson (2005); Hopkins et al. (2006); 
82: Bower et al. (2006); Menci et al. (2006). Intriguingly,
83: the latter authors found
84: that the bimodal color distribution of galaxies at z$\gtrsim 1.5$ can only
85: be explained if AGN feedback is considered.  In this picture an AGN
86: phase precedes the phase when a galaxy is caught in a passive state
87: with red optical-UV colors, most of the star-formation having been
88: inhibited by the AGN activity. Indeed, Pozzi et al. (2007) using
89: Spitzer photometry found that that a sample of optically obscured QSO
90: at z=1--2 are mainly hosted by red passive galaxies, suggesting a later stage
91: in their evolution.
92: 
93: On the other hand, at low redshift many weak AGNs have
94: been found in star-forming galaxies (Salim et al. 2007). In these
95: cases feedback from less powerful AGNs (the so-called ''radio mode'') 
96: is probably acting to
97: self-regulate accretion and star-formation, and cold gas is left
98: available for both processes for a much longer time (Croton et
99: al. 2006). The same cold gas can intercept the line of sight to the nucleus. 
100: Indeed, Compton-thin absorbers (with column densities $N_H\leq 10^{24}$ cm$^{-2}$)
101: may well be located in the galactic disk (Malkan, Gorjian, Tam 1998; Matt 2000; 
102: see also Ballantyne, Everett, Murray 2006). 
103: Therefore a natural expectation in this scenario is the fraction of obscured
104: AGNs to be large at low AGN luminosities. It is well known since the
105: pioneering work done with the {\it Einstein} satellite (Lawrence \&
106: Elvis 1982) and with optically and radio-selected AGNs (Lawrence 1991) 
107: that this fraction strongly decreases with increasing AGN
108: luminosity (see Ueda et al. 2003; La Franca et al. 2005; Gilli, Comastri, Hasinger
109: 2007; Triester, Krolik, Dullemond 2008; Hasinger 2008).  A widely
110: shared view holds that the luminosity dependence of the obscured
111: fraction is related to the energy fed back by the AGNs onto the
112: surrounding gas that constitutes the interstellar medium (ISM). Given that
113: the AGN emission is proportional to the fraction of such a gas
114: available for accretion, a \textit{positive} correlation between
115: luminosity and absorption would be expected instead in the absence of
116: an energy feedback depleting the ISM after the onset of the AGN activity.
117: 
118: A mounting body of observations cogently indicates that strong nuclear feedback
119: is present in galaxies hosting AGNs (see for a review Elvis 2006 and
120: references therein).  On small (sub-pc) scales, the
121: observed X-ray absorption lines indicate the presence of outward winds
122: with velocities up to some $ 10^4$ km/s (Weymann 1981; 
123: Turnshek et al. 1988; Creenshaw et al. 2003; 
124:  Chartas et al. 2002; Pounds
125: et al. 2003, 2006; Risaliti et al. 2005b).  These likely 
126: originate from the acceleration of disk outflows due to the AGN
127: radiation field (Proga 2007 and references therein). On
128: larger scales, broad absorption lines in about 10\% of optically
129: luminous QSOs indicate fast outflows (up to 30,000 km/s).  Massive
130: (10-50 $M_{\odot}/yr$) flows of neutral gas with speed $\sim1000$ km/s
131: are observed through 21-cm absorption of radio-loud AGNs (see
132: Morganti, Tadhunter, Oosterloo 2005), indicating that AGNs have a
133: major effect on the circumnuclear gas in the central kiloparsec region
134: around  AGNs. On even larger scales of some $10^{2}$ kpc, the presence of
135: AGN-induced outflows is revealed by X-ray observations of the
136: intra-cluster medium (see McNamara \& Nulsen 2007 for a review)
137: showing cavities and expanding shocks with Mach numbers ranging from
138: $\approx 1.5$ to $\sim 8$. How the outflows produced in the innermost
139: regions of the active galaxies are transported outwards to affect such
140: large scales is still matter of investigation; buoyant bubbles (see
141: Reynolds, Heinz \& Begelman 2001, Churazov 2001) and expanding blast
142: waves (see Cavaliere, Lapi \& Menci 2002; Lapi, Cavaliere \& Menci 2005)
143: constitute viable mechanisms for such a transport.
144: 
145: Nuclear obscuration is directly linked to AGN feedback, since the same
146: gas (and dust) which feed the AGN output may well be
147: responsible for its obscuration. Therefore, modelling AGN obscuration
148: is essential to connect the observed AGN properties to the
149: accretion history of SMBHs over cosmological time. This is a significant
150: theoretical challenge as it requires not only connecting the AGN
151: evolution to the galaxy formation and growth, but also implementing in
152: a model a self-consistent description of the AGN feedback on the
153: galactic gas. Indeed, very few attempts have been
154: carried out in this direction so far. 
155: An attempt to include AGN absorption into an {\it ab initio}
156: galaxy formation model has been proposed by Nulsen \& Fabian
157: (2000), by relating both the SMBH fueling and the AGN
158: absorption to the cooling flows associated with the hot gas pervading
159: the growing dark matter haloes, but this assumption did not lead to a 
160: full description of the statistical distribution of QSO luminosity.
161: 
162: Here we develop our semi-analytic model of hierachical galaxy
163: formation and AGN evolution (see Menci 2006) to self-consistently
164: include the absorption of AGNs, with the aim of investigating the
165: cosmic evolution of the latter and its dependence on AGN
166: properties like luminosity and redshifts.
167: The model is suited to our scope as it includes a detailed 
168: treatment of the feedback on the interstellar gas. Our aim is
169: to investigate the dependence of AGN {\it absorption} on luminosity $L$ and
170: redshift $z$ arising in hierarchical galaxy formation scenarios that
171: include an effective description of the evolution of AGNs and of their
172: feedback.
173: 
174: The plan of the paper is as follows: in Sect. 2 we describe our model
175: for galaxy formation and the associated AGN evolution. Sect. 3 is
176: focussed on describing our treatment of the AGN feedback onto the
177: interstellar gas and  its effects on the AGN absorption. In Sect. 4
178: we test our feedback-inclusive model for AGN evolution by comparing
179: its outcomes with the redshift distribution of the number density of
180: AGNs for different X-ray luminosities. Our results on the luminosity
181: and redshift dependence of the absorbed fraction of AGNs are shown and
182: discussed in Sect. 5. Sect. 6 is devoted to summarize our conclusions.
183: 
184: 
185: \section{The Model}
186: 
187: The semi-analytic model we develop and use  connects, within a cosmological framework,
188: the accretion onto SMBHs and the ensuing AGN activities with the  evolution of galaxies.
189: 
190: \subsection{Hierarchical Galaxy Evolution}
191: 
192: Galaxy formation and evolution is driven by the collapse and growth of dark
193: matter (DM) haloes, which originate by gravitational instability of  overdense
194: regions in the primordial DM density field. This is taken to be a random,
195: Gaussian  density field with Cold Dark Matter (CDM) power spectrum within the
196: ''concordance cosmology" (Spergel et al. 2006) for which we adopt round
197: parameters  $\Omega_{\Lambda}=0.7$, $\Omega_{0}=0.3$, baryonic density
198: $\Omega_b=0.04$ and Hubble constant (in units of 100 km/s/Mpc) $h=0.7$. The
199: normalization of the spectrum is taken to be $\sigma_8=0.9$ in terms of the variance
200: of the field smoothed over regions of 8 $h^{-1}$ Mpc.
201: 
202: As  cosmic time increases, larger and larger regions of the density field
203: collapse, and  eventually lead to the formation of groups and clusters of
204: galaxies; previously formed, galactic size  condensations are enclosed. In
205: closer detail , the process implies  not only smooth mass inflow, but also
206: merging and coalescence of smaller condensations. The corresponding merging rates
207: of the DM haloes are provided by the Extended Press \& Schechter formalism (see
208: Bond et al. 1991; Lacey \& Cole 1993).  The clumps included into larger DM haloes
209: may survive as satellites, or merge to form larger galaxies due to binary
210: aggregations,  or coalesce into the central dominant galaxy due to dynamical
211: friction; these processes take place over timescales that grow longer over
212: cosmic time, so the number of satellite galaxies increases as the DM host haloes
213: grow from groups to clusters. All the above processes are implemented in our
214: model  as described in detail in Menci et al. (2005, 2006), based on canonical
215: prescriptions of semianalytic modeling.
216: 
217: The radiative gas cooling, the ensuing star formation and the
218: Supernova events with the associated feedback occurring  in the growing
219: DM haloes (with mass $m$ and circular velocity $v$) are described in
220: our previous papers (e.g., Menci et al. 2005). The cooled gas with mass
221: $m_c$ settles into a rotationally supported disk with radius $r_d$
222: (typically ranging from $1$ to $5 $ kpc), rotation velocity $v_d$
223: and dynamical time $t_d=r_d/v_d$. The gas gradually condenses  into
224: stars at a rate $\dot m_*\propto m_c/t_d$; the stellar ensuing feedback
225: returns part of the cooled gas to the hot gas phase
226: with mass $m_h$ at the virial temperature of the halo. An additional
227: channel for star formation implemented in the model is provided by
228: interaction-driven starbursts, triggered not only by merging but
229: also by fly-by events between galaxies; such a star formation mode
230: provides an important contribution to the early formation of stars
231: in massive galaxies, as described in detail in Menci et al. (2004,
232: 2005).
233: 
234: \subsection{Accretion onto SMBHs and AGN emission}
235: 
236: The model also includes a treatment of SMBHs growing at the centre of galaxies
237: by interaction-triggered inflow of cold gas, following the physical model proposed
238: by Cavaliere \& Vittorini (2000) and implemented in Menci et al. (2006). The
239: accretion of cold gas is triggered by galaxy encounters (both of fly-by and
240: of merging kind), which destabilizes part of the cold gas available by inducing
241: loss af angular momentum. In fact, small scale (0.1-a few $kpc$) regions are likely to have 
242: disk geometry if they are to efficiently remove angular momentum and convey to 
243: the (unresolved) pc scales the gas provided on larger scales (these 
244: may be isotropized by head-on, major merging events). 
245: 
246: The rate of such  interactions is given by Menci et
247: al. (2003) in the form
248: \begin{equation}
249: \tau_r^{-1}=n_T\,\Sigma (r_t,v,V_{rel})\,V_{rel}.
250: \end{equation}
251: Here $n_T$ is the number density of galaxies in the same halo and
252: $V_{rel}$ is their relative velocity. Encounters effective for
253: angular momentum transfer require i) the interaction time to be
254: comparable with the internal dynamical times of the partner galaxies
255: (a resonance condition), ii) the orbital specific energy of the
256: partners not to exceed the sum of their specific internal
257: gravitational energies. The cross section $\Sigma$ for such
258: encounters is given by Saslaw (1985) in terms of the tidal radius
259: $r_t$ associated to a galaxy with given circular velocity $v$ (see
260: Menci et al. 2003, 2004).
261: 
262: The fraction of cold gas accreted by the BH in an interaction event is computed
263: in terms the  variation $\Delta j$ of the specific angular momentum $j\approx
264: Gm/v_d$ of the gas to read (Menci et al. 2003)
265: \begin{equation}
266: f_{acc}\approx 10^{-1}\,
267: \Big|{\Delta j\over j}\Big|=
268: %{1\over 6}\Big\langle {Gm'\,r_d\over V\,b}/{G\,m\over v_d}\Big\rangle  =
269: 10^{-1}\Big\langle {m'\over m}\,{r_d\over b}\,{v_d\over V_{rel}}\Big\rangle\, .
270: \end{equation}
271: Here $b$ is the impact parameter, evaluated as the average distance of the
272: galaxies in the halo. Also, $m'$ is the mass of the  partner galaxy in the
273: interaction,  and the average runs over the probability of finding such a galaxy
274: in the same halo where the galaxy with mass $m$ is located.
275: The values of the quantities involved in the average are computed from the 
276: semi-analytic model recalled in Sect. 2.1, and yield values of $f_{acc}\lesssim 10^{-2}$. 
277: For minor merging events and for the encounters among galaxies with very unequal mass ratios
278: $m'\ll m$, dominating the statistics in all hierarchical models of galaxy formation, the 
279: accreted fraction takes values $10^{-3}\lesssim f_{acc}\lesssim 10^{-2}$. 
280: 
281: The average amount of cold gas accreted during an accretion episode is thus
282: $\Delta m_{acc}=f_{acc}\,m_c$, and the duration of an accretion episode, i.e.,
283: the timescale for the QSO or AGN to shine, is assumed to be the crossing time
284: $\tau=r_d/v_d$ for the destabilized cold gas component.
285: 
286: The  time-averaged bolometric luminosity so produced by a QSO hosted in a given galaxy 
287: is then provided  by
288: \begin{equation}
289: L={\eta\,c^2\Delta m_{acc}\over \tau} ~.
290: \end{equation}
291: We adopt an energy-conversion efficiency $\eta= 0.1$ (see Yu \&
292: Tremaine 2002), and derive the X-ray luminosities $L_X$ in the 2-10
293: keV band  from the bolometric corrections given in Marconi et al.
294: (2004). The SMBH mass $m_{BH}$ grows mainly through accretion
295: episodes as described above, besides  coalescence with other SMBHs
296: during galaxy merging. As initial condition, we assume small seed
297: BHs of mass $10^2\,M_{\odot}$ (Madau \& Rees 2001) to be initially
298: present in all galaxy progenitors; our results are insensitive to
299: the specific value as long as it is smaller than some
300: $10^5\,M_{\odot}$.
301: 
302: In our Monte Carlo model, at each time step we assign to a  galaxy
303: the interaction probability corresponding to the rate given in eq. (1).
304: According to it,  we assign to the galaxy an active SMBH accretion event of
305: duration $\tau$.  Then we compute the accreted cold gas and the
306: associated AGN emission through equations (2) and (3).
307: 
308: \section{The AGN feedback and the column density of absorbing gas}
309: 
310: The model for AGN evolution presented here -- relative  to previous
311: implementation by Menci et al. (2003-2004) -- is complemented with
312: the feedback from AGNs onto the surrounding ISM. This affects: a)
313: the available cold gas left over  in the galactic disk after each
314: AGN event, that determines the subsequent gas accretion history; b)
315: the density distribution of the galactic gas during each AGN event,
316: that determines  the AGN absorption.
317: 
318: \subsection[]{The Blast Wave Model for the AGN Feedback}
319: As mentioned in the Introduction, fast winds with velocity up to
320: $10^{-1}c$ are observed in the central regions of AGNs; they likely
321: originate from the acceleration of disk outflows by the AGN
322: radiation field (see Begelman 2003 for a review), and affect the
323: environment in the host galaxy and beyond, leaving imprints out to
324: large scales of some $10^2$ kpc  in the intra-cluster medium (ICM).
325: 
326: A detailed model for the transport of energy from the inner, outflow
327: region to the large scales has been developed by Cavaliere, Lapi \&
328: Menci (2002), and Lapi, Cavaliere  \& Menci (2005). Central, highly
329: supersonic outflows compress the gas into a blast wave terminated by
330: a leading shock front, which  moves outwards with a lower but still
331: supersonic speed and sweeps  out the surrounding medium. Eventually,
332: this is expelled from the galaxy;  in the case of powerful shocks,
333: it is expelled  even  from a  group or poor cluster hosting the galaxy.
334: 
335: The key quantity determining all  shock properties is the total energy $\Delta E$
336: injected by AGNs into the surrounding gas. This is computed as
337: \begin{equation}
338: \Delta E = \epsilon_{AGN}\,\eta\,c^2\,\Delta m_{acc}
339: \end{equation}
340: (see Sect. 2.2) for each SMBH accretion episode in our Monte Carlo
341: simulation; the value of the energy feedback efficiency for coupling with
342: the surrounding gas is taken  as
343: $\epsilon_{AGN}=5\, 10^{-2}$, consistent with the values required to
344: match the X-ray properties of the ICM  in clusters of galaxies (see
345: Cavaliere, Lapi \& Menci 2002). This is also consistent with the
346: observations of wind speeds up to $v_w\approx 0.1\,c$ in the central
347: regions, that yield $\epsilon_{AGN}\approx v_w/2c\approx 0.05$ by
348: momentum conservation between photons and particles
349: (see Chartas 2002, Pounds 2003); this value has been also adopted
350: in a number of simulations (e.g., Di
351: Matteo, Springel  \& Hernquist 2005) and semi-analytic models
352:  of galaxy formation (e.g., Menci et al. 2006).
353: 
354: The blast expands  into the ISM or the ICM  as  described by the
355: hydrodynamical equations; these are to include the effects  not
356: only   of  initial density gradient, but also those of  upstream
357: pressure and DM gravity,  clearly important quantities in galaxies,
358: as discussed by Lapi, Cavaliere, Menci (2005);
359: the solutions show in detail how the
360: perturbed gas is confined to an expanding  shell bounded by an outer
361: shock at the  radius $R_s(t)$ which sweeps out the gas surrounding
362: the AGN. The overall effects have been emulated in the simulations
363: by Di Matteo et al. (2005), performed for specific  cases of major
364: mergers. Our treatment covers also the less energetic but more
365: frequent events, with   accretion rates and AGN energy inputs
366: $\Delta E$ provided by eqs. (1) to (4).  On the other hand, 
367: our model assumes spherical symmetry; 
368: note however that this approximatively holds in the central 
369: regions (the ones mainly affecting our results, as we shall show below) 
370: due the exponential decline of the gas density. 
371: In addition, both 
372: the effects discussed in points 1) and 2) of the Appendix (see below Sect. 3.2) 
373: make the average absorption even less dependent on the disk outside the central region.
374: As for the outflow geometry,  our model also applies to segments of spherical shells expanding outwards, originated by winds driven by a spherically symmetric AGN radiation field; 
375: these segments may not cover uniformly the central source, as in the conical distribution of Elvis et al. (2000). While its is true that random line-of-sights may not intercept some of these segments, the random orientation of the internal BH accretion disk with respect to the outer, galactic disk (Gallimore,  Baum,  O'Dea 1997; Nagar \& Wilson 1999; Thean et al. 2001) makes our results realistic on average. Finally, we do not consider clumpy outflows; however, these are not expected to affect the dependence of the column density distribution on the AGN luminosity and redshift, though they could broaden the column density distributions that we derive from our model.
376:  
377: Lapi et al. (2005) showed in detail how  the hydrodynamical
378: equations for the finite amplitude gas perturbation constituting the
379: blast wave, supplemented with the Rankine-Hugoniot boundary
380: conditions at the shock, still admit  self-similar solutions (see
381: Sedov 1959) when  gravity,  initial gradients and pressure are
382: included. These solutions at the distance $r$ from the center
383: are expressed in terms of $r/R_s$, where
384: $R_s(t)$ is the shock position after a time
385: $t$ from an  AGN outburst; this is given by the eq. (C7) in Lapi et al.
386: (2005), that we recast in the form
387: \begin{equation}
388: R_s(t)=v_d\,t_d\,\Big[{5\,\pi \omega^2 \over 24 \pi (\omega-1)}\Big]^{1/\omega}
389: {\mathcal M}^{2/\omega}\Big[{t\over t_d}\Big]^{2/\omega},
390: \end{equation}
391: where ${\mathcal M}$ is the Mach number ${\mathcal M}=v/c_s(R_s(t))$.
392: 
393: Here the initial gas density has a radial
394: profile given by  a power law $\rho\propto r^{\omega}$  with the exponent in the
395: realistic  range $2\leq \omega <2.5$. While self-similar, analytical solutions
396: for the shock expansion law can only be found  for spherical (or in general 1D)
397: symmetry and power law gradients, we shall apply eq. (5) also to blasts
398: expanding in the inner regions of a galactic disk, where the actual  density
399: profile follows an exponential law $\rho=(1/2\,h)\,\sigma_o\,exp(-r/r_d)$ ($h$
400: is the disk thickness and $\sigma_0$ is the central surface density), with a
401: cutoff in the vertical direction corresponding to the thickness of the disk. To
402: use the above self-similar blast we construct  piecewise power-law approximations of
403: the above exponential density by subdividing the radial disk coordinate into a
404: sequence of shells; within each shell the exponential decline is approximated
405: with a power-law with a different exponent $\omega$.  The expansion of the blast
406: in each shell will be computed on using eq. (5) with the appropriate value of
407: $\omega$.
408:  
409: Self-similarity imply a definite time behavior for the AGN energy injection
410: which is determined by the exponent $\omega$ defining the density decline, normalized  
411: to reach its total value $\Delta E$ by the end of the accretion episode. 
412: In turn, this implies a well-defined form for the time decay of the AGN luminosity, 
413: as shown in the Appendix. However, in the self-similar solutions the ratio 
414: $\Delta E/E$ between the energy injected up to a time $t$ and the energy of the ISM 
415: within the shock radius $R_s(t)$ is shown to be independent of time and position
416: (see Appendix), and thus constitutes the basic quantity marking the strength 
417: of the shock. In fact. 
418:  the Mach number is shown to be simply related to $\Delta E/E$ which we compute 
419: at the final time of the AGN active phase. 
420: In the following,  we shall take ${\mathcal M}^2=1+\Delta
421: E/E$ as derived by  Lapi et al. (2005).
422: 
423: In terms of the rescaled variable $r/R_s$, the previous authors  derived the
424: density distribution;  this is found to be confined to  a shell close to $R_s$
425: between the outer shock and an inner "piston"  at a position $\lambda\,R_s(t)$
426: ($\lambda < 0.7$) with a width  $(1-\lambda)\,R_s$ weakly dependent on the Mach
427: number. Inside  the shell little  gas is  left over  to absorb the AGN
428: radiation.
429: 
430: \subsection[]{The Effect of a Blastwave on the Column Density of the Absorbing Galactic Gas}
431: 
432: The amount of unshocked gas absorbing the AGN emission will be given, for a
433: given line-of sight, by the time-dependent fraction still unperturbed outside
434: the shock located at  $r= R_s(t)$ (see fig. 1). The faster the shock, the lower
435: will be the fraction of still unperturbed gas outside $R_s(t)$ after a time
436: $t$ from an  AGN outburst.
437: 
438: The dependence of $R_s(t)$ in eq. (5) on the AGN luminosity will then affect the
439: absorption by  the unperturbed gas;  in particular, for the  case $\omega=2$
440: (corresponding to the standard isothermal sphere) the relation $R_s(t)\propto
441: {\mathcal M}\,t\propto (\Delta E/E)^{1/2}\,t$ holds, yielding {\it lower} absorbing
442: gas fraction outside $R_s(t)$ for {\it brighter} AGNs at a fixed time $t$. 
443: 
444: An additional contribution to the absorption will be given by the gas compressed
445: by the shock in  the  shell of   width $(1-\lambda)\,R_s$ (see  Lapi, Cavaliere
446: \& Menci 2005), where $\lambda$ can  be also expressed in terms of the relevant
447: ratio $\Delta E/E$; however, we shall show  this second contribution to  depend
448: only weakly  on the AGN luminosity.
449: 
450: 
451: \begin{center}
452: \vspace{-0.cm}
453: \scalebox{0.73}[0.73]{{\includegraphics{f1.pdf}}}
454: \end{center} {\footnotesize \vspace{-0.4cm }
455: Fig. 1. - A schematic representation of the
456: effect of the blast wave induce by AGN feedback on the density
457: distribution of the interstellar gas. The shock radius $R_s(t)$
458: expands outwards, compressing the swept gas into a thin shell
459: (represented in darker colour) with width $R_s(1-\lambda)$, and
460: leaving a cavity inside. Such a density distribution $\rho(r)$ is
461: also plotted at the bottom. A line-of-sight (at an
462: angle $\theta$ with respect to the plane of the galactic disk) is
463: also represented in the scheme, along with the length $\ell$ of the
464: line of sight intercepting the unperturbed gas outward of the shock
465: front.
466:  \vspace{0.4cm}}
467: 
468: We implement the above model for AGN feedback and for absorption in
469: our SAM as follows. \newline \newline 0) For each galaxy in our
470: Monte Carlo simulations, with disk exponential scale length $r_d$ and
471: total disk mass $m_c$ computed as described in \S  2.1, we assume
472: for the density distribution of the  unperturbed gas the simplified form
473: $\rho=\rho_o\,exp(-r/r_d)$ (where $r$ is the distance from the
474: centre of the galaxy) with a cutoff in the vertical direction
475: (perpendicular to the disk) at $r=h$ corresponding to the disk
476: thickness. \newline The value of the vertical scale height $h$ for
477: the {\it gaseous} disk is taken to be $r_d/15$ (see Narayan \& Jog
478: 2002),  corresponding to $\sim 200$ pc for a typical $L_*$ galaxy.
479: With such a low value of $h$, our assumed density for the gaseous
480: disk closely follows a constant vertical density distribution with a
481: radial exponential profile $exp(-x/r_d)$, $x$ being the distance
482: from the centre in the plane of the disk; the deviations are  $\leq
483: 14 \% $ in the central region,  and $\leq 1 \%$ at $x=r_d$. These
484: are  both smaller than the present uncertainty on the vertical
485: distribution of the density of the gaseous disks; on the other hand,
486: the above schematic  form for the gas density distribution allows us
487: to apply a simple  anaytical description for  the evolution of the
488: blast. The density distribution is normalized as to recover the
489: total gas mass $m_c$ when integrated over the disk volume.
490:  \newline
491:  \newline
492: 1) We construct a piecewise power-law approximation of the above
493: exponential density distribution, by subdividing the radial
494: coordinate into shells; in each shell we approximate the exponential
495: shape  $exp(-r/r_d)$ with a power-law $A\,q^{\omega}\,s^{-\omega}$
496: where  $s=r+q$, where $q$, $A$, and $\omega$  are parameters that we
497: adjust to optimize the fit in each radial shell. We find that, on
498: retaining  $q=1.8\,r_d$ at all radii,  the following set of radial
499: shells and  parameters provides a good fit to the exponential law:
500: \newline $A=1$, $\omega=2$ for $r_m\leq r\leq r_d/2$,
501: \newline $A=1.23$, $\omega=2.3$ for $r_d/2\leq r\leq 3\,r_d/2$,
502: \newline $A=1$, $\omega=2.5$ for
503: $r\geq 3\,r_d/2$; \newline
504: here the minimal  radius considered is
505: $r_m=50$ pc. With the above choice, the r.m.s. deviation in the fits
506: are $\Delta\leq 3\%$ within the inner region ($r\leq 10 r_d$, the
507: dominant  range for the AGN abosrption),   and $\Delta\leq 30\%$ for
508: $r\geq 10 r_d$ (a region irrelevant to our computation of
509: the AGN absorbtion). The temperature of the gas in the disk is
510: assumed to be $T_d=10^4$ K.
511: \newline
512:  \newline
513: 2) When a  SMBH starts its accretion phase triggered by a close
514: interaction of the host galaxy with a  companion, we compute the
515: corresponding AGN total energy injection $\Delta E$. For the sake of 
516: simplicity, we shall label each AGN with a constant luminosity $L$ after eq. (3), 
517: although a full consistency with the self-similar solutions adopted for the 
518: blastwave expansion would require a time-decay of the luminosity related to the 
519: exponent $\omega$. However, we show in the Appendix that assuming a constant $L$ 
520: yields results not distinguishable from those obtained by assigning the AGNs the exact, 
521: time-dependent luminosity.
522: \newline
523: \newline
524: 3) During the time interval $\tau$ we follow the expansion of the shock given
525: by eq. (5) for each radial shell  defined under  1), on  using the appropriate
526: value of $\omega$ given by the fitting procedure described above. We compute the
527: amount of gas still unperturbed outwards of  the shock position $R_s(t)$ at a
528: time $t$ within the  interval $\tau$; the unshocked gas distribution is  given
529: by the density profile  $\rho(r)$ defined under  (1), with $r\geq R_s(t)$ (see
530: fig. 1).
531: \newline
532: \newline
533: 4) We extract a random line-of-sight angle $\theta$ defining the disk
534: inclination to the observer; at a time $t$ within the interval $\tau$
535: corresponding to the active AGN,  we compute the column density corresponding to
536: the gas outside the shock position along the selected line-of-sight as
537: \begin{equation}
538: N_H=\int_{R_s(t)}^{h/sin{\theta}} \rho(r)\,d\ell ~.
539: \end{equation}
540: 5) If for the chosen line-of-sight the shock position is within the
541: disk, we add the absorption column density corresponding to the gas
542: compressed by the shock (see fig. 1).  Lapi et al. (2005) showed
543: that the density of the gas swept by the blast  and  compressed to a
544: shell of width $(1-\lambda)\,R_s$ rises toward the center as a
545: function $D(r/R_s(t))$ diverging weakly (but still integrating to a
546: limited mass) at the piston position $\lambda\,R_s(t)$. On adopting
547: the behaviour
548: $D(r/R_s(t))=[(r/\lambda\,R_s)-1]^{(\omega-6)/3(7-\omega)}$
549: (strictly valid  on approaching the piston position), and
550: integrating over $r$ within the shocked shell, the corresponding
551: column density is
552: \begin{equation}
553: N_{H, shock}={15-5\omega\over 3(7-2\omega)}\,\Big[{1\over \lambda}-1\Big]^{15-5\omega\over 3(7-2\omega)}\,{n_2\over n_1}\,n_1\,R_s(t)~.
554: \end{equation}
555: Here $n_2/n_1$ is the density jump at the shock, and $n_1=\rho (R_s)$ is the
556: unperturbed gas density just outside the shock front. Both the density jump
557: $n_2/n_1$ and the $\lambda$ grow slowly  with increasing Mach number
558: $\mathcal{M}$ (in turn a function of $\Delta E/E$), as given in Lapi et al.
559: (2005); both  quantities saturate to constant values ($n_2/n_1=4$ and
560: $\lambda\approx 0.7$) in the limit of very strong shocks ($\mathcal{M}\gg 1$).
561: 
562: In sum, the blastwave has two effects: on the one hand, it depletes
563: the internal regions    thus decreasing the column density of the
564: absorber (as computed at point 4 above); on the other hand, it
565: increases the density inside a thin layer close to the piston (an
566: effect computed at point 5 above). The first effect however
567: dominates, since the gas initially distributed in the
568: denser central region is spread out by the blast into an expanding
569: spherical shell; thus, the integration along
570: the line of sight intercepts only a tiny fraction of the gas
571: initially concentrated in the core of the distribution. 
572: In addition, the minor contribution to the column 
573: density from eq. (7) represents an upper limit, since 
574: the gas temperature in the blast wave is
575: increased to values (given in eq. C8 of Lapi, Cavaliere, Menci 2005)
576: which generally exceed the ionization temperature of the gas and
577: thus limit the absorption.
578: At the end of the AGN accretion episode (i.e., at
579: $t=\tau$) only a fraction of the initial cold gas mass $m_c$ will be
580: left in the galaxy, as discussed in detail in Cavaliere et al.
581: (2002) and Lapi et al. (2005). Note that, the brighter is the AGN,
582: the larger is the injected energy  $\Delta E$, the faster is the
583: shock espansion $R_s(t)$ (see eq. 5), the lower will be the
584: resulting column density at a given time after the start of the
585: outburst (see eq. 6).
586: 
587: Thus the evolution of the gas distribution and hence of the gas absorption
588: during the AGN event is best  described in terms of the key ratio $\Delta E/E$
589: which determines all  the relevant  quantities of the blast: the expansion of
590: the shock front $R_s(t)$, the thickness of the shocked shell $(1-\lambda)$, and
591: the density jump at the shock. Such a ratio is computed in the model for each
592: galaxy when an active accretion episode is triggered by a close encounter.
593: 
594: The effects of the above detailed model for the AGN feedback on the evolution of
595: the cold galactic gas and on the galactic absorption of the AGN emission are
596: given  next.
597: 
598: \section{Results}
599: 
600: Here we present our results concerning the evolution of the AGN number density
601: and the obscuration by galactic interstellar gas.
602: We do not provide a description of absorption
603: due to the ''clouds'' within the central regions of AGNs, 
604: so we cannot disentangle Compton-thick sources (where the obscuration is
605: presumably provided by gas in regions $r<10$ $pc$) from the absorbed
606: population we obtain from our model. Indeed, 
607: the shock may originate at any point within our resolution scale (50 $pc$), 
608: and it may lie inside or outside the Compton-thick region. 
609: Therefore we cannot make any strong statement about the dependency of 
610: Compton-thick absorption on the outflow and hence on the AGN luminosity.
611: 
612: Before addressing the problem of the luminosity and redshift dependence of
613: AGN absorption, we test the basic predictions of our model with recent data.
614: To this aim, we first compare in fig. 2 the predicted evolution of the cosmic
615: density of AGN with different {\it intrinsic} luminosity $L$ with
616: data from La Franca et al. (2005). Note that the model reproduces the
617: marked {\it downsizing} of AGNs, i.e., the faster redshift evolution of brighter AGNs
618: compared to the fainter sources (a behaviour already obtained in  semi-analytic models, 
619: see Menci et al. 2004; Fontanot et al. 2006). In the model, such an
620: effect is related to the faster consumption of gas in the
621: progenitors of massive galaxies at high redshifts $z\gtrsim 3$; in
622: fact, these are formed in biased high regions of the cosmological 
623: density field, where the early cooling of galactic gas provides
624: early reservoirs for BH accretion, and the galaxy merging and
625: interactions are enhanced. When such progenitors assemble to form a
626: massive galaxy (and correspondingly a bright AGN when the BH is
627: actively accretiing) at $z\lesssim 2$, the available galactic
628: gas has largely been converted into stars or accreted onto the BH at
629: higher redshifts. Conversely, low-mass galaxies retain a large
630: fraction of their gas down to low redshifts, and this results into
631: prolonged star formation (yielding blue colors as observed) 
632: allowing for effective BH accretion down to $z\approx 0$.
633: 
634: \begin{center}
635: \vspace{0.2cm}
636: \scalebox{0.5}[0.5]{{\includegraphics{f2.pdf}}}
637: \end{center} {\footnotesize \vspace{0.cm }
638: Fig. 2. - 
639: The evolution of the cosmic density of AGNs with X-ray luminosities
640: (in the $2-10$ keV band)
641: in 5 different luminosity bins: $42\leq$ log $L$/erg s$^{-1}< 43$ (top, blak line),
642: $43\leq $ log $L$/erg s$^{-1}< 44$ (red line), $44\leq$ log $L$/erg s$^{-1}< 44.5$ (green line),
643: $44.5\leq$ log $L$/erg s$^{-1}< 45$ (blue line), log $L$/erg s$^{-1}\geq 45$ (bottom cyan line).
644: Note that the the plotted absorbed fraction includes the Compton-thick sources.
645:  \vspace{0.4cm}}
646: 
647: Note that at $z\gtrsim 2$ the model predictions lie above the observed points by a factor $\approx 2$ for low luminosity AGNs with $L_X\leq 10^{43}$ erg/s. This, at least in part, is because
648: the model predictions in fig. 2 concern all AGNs, regardless of whether they are Compton-thin or Compton-thick,since we cannot disentangle the two populations in our model, while the data we compare with include only  Compton-thin sources.  
649: A complementary check would be constituted by the 
650: X-ray background at its 30 keV peak (where observations are not affected by Copton-thick obscuration). However this does not actually constitute a strong constraint for the model, since it is mainly contributed by objects at $z\leq 2$ (Gilli, Comastri, Hasinger 2007) whose observed density is well reproduced by the model. We have repetedly checked since Menci et al. (2003; 2004)  that  the basic features of our model are consistent with the constraint set by the observed X-ray background (see Barcons, Mateos, Ceballos, 2000) and that the predicted global history of BH assembly is consistent with the local observed SMBH mass function 
651: (Marconi et al. 2004). 
652: 
653: Thus the comparison between model predictions and observation strongly suggests the existence
654: of a relevant fraction of Compton-thick sources at $z\gtrsim 1.5$ (as found by Martinez-Sansigre et al. 2005, 2007;
655: Fiore et al. 2008a,b). Note however that a complementary possibility, which could partly account for the model
656: overestimate of low-luminosity sources at $z\gtrsim 2$, is that the BH growth in
657: small-mass galaxy halos (DM masses $M\leq 10^9\,M_{\odot}$) be inhibited by some
658: process not included in the model; we shall comment further this point in the Conclusions.
659: 
660: Finally, we stress that the model predicts at $z\gtrsim 5$ a density of very luminous-objects  
661: ($L_X\geq 10^{45}$ erg/s) similar to that at $z\approx 1$. These are the X-ray counterparts of the
662: luminous $M_B\lesssim -27.2$ optical QSOs observed up to $z\approx 6.5$
663: (see Fan et al. 2004; Richards et al. 2006) and imply that fast 
664: building up of massive BH at early epochs is possible in hierarchical scenarios
665: (as long as a constant mass-energy conversion factor $\eta\approx 0.1$ holds, see
666: discussion in the final Section). In turn, this is a consequence of the speed up of
667: structure growth (enhanced merging rate) for haloes collapsed in biased, high-density
668: regions of the primordial density field, which constitutes a natural feature of hierarchical
669: scenarios.
670: 
671: We now turn to investigating the predictions of our model
672: concerning the luminosity and the redshift  dependence of the
673: absorption due to galactic gas. In fig. 3 we show
674: the predicted distribution of the column densities of the absorbers
675: for low redshift ($z\leq 1$) AGNs with different luminosity, and
676: compare it with data from La Franca et al. (2005), and with expectations based
677: on AGN synthesis models for the cosmic X-ray background (Gilli et al. 2007).
678: The rapid expansion of the blast wave produced by bright AGNs yields a low
679: fraction of absorbed ($N_H\geq 10^{23}$ cm$^{-2}$), bright ($L_X\geq
680: 10^{45.5}$ erg/s) AGNs, while at lower luminosity the $N_H$
681: distribution shifts to larger values.
682: Note that the above relation between the luminosity and the obscuring column density
683: holds only after averaging over {\it time} after the start of the AGN active phase and
684: over the {\it orientations} of the galactic disk, the two fundamental
685: parameters that define the absorption properties of any single AGN in our model.
686: The probability for  a line-of-sight to intercept a given amount of
687: galactic gas is independent of the AGN luminosity (although models predicting such a dependence
688: have been proposed, see Lamastra, Perola \& Matt 2006), while the probability to observe an AGN
689: at a time when the blast wave has depleted the gas content of the galaxy
690: is larger for luminous AGNs, since these produce faster blast waves.
691: 
692: \begin{center}
693: \vspace{0.cm}
694: \scalebox{0.6}[0.6]{\rotatebox{-90}{\includegraphics{f3.pdf}}}
695: \end{center} {\footnotesize \vspace{0.cm }
696: Fig. 3. - 
697: Left panel: The distribution of
698: column densities $N_H$ at $z\leq 1$ for AGN with different luminosities:
699: 42$\leq$ log $L_X$/erg s$^{-1}\leq$ 43 (green dotted line), 43$\leq$
700: log $L_X$/erg s$^{-1}\leq$ 44 (red dashed line), log $L_X$/erg
701: s$^{-1}\geq$ 44 (violet solid line). The lowest $N_H$ bin
702: actually includes the contribution from all absorbing column
703: densities $N_H \leq 10^{22}$ cm$^{-2}$. \newline Right panel: The
704: predicted fraction of absorbed AGNs with $N_H>10^{22}$ cm$^{-2}$ is
705: plotted as a function of their X-ray luminosity for $z<1$ (solid
706: line), and compared with data from La Franca et al. (2005, solid squares); the
707: latter have been corrected for the selection effects as described by the authors.
708: We plot also the expectations of the AGN synthesis
709: model for the cosmic X-ray background by Gilli et al. (2007, dots).
710: We also show as a dashed line the obscured fraction as a function of AGN luminosity
711: when no AGN feedback is considered; note that this has been computed with the same choice of
712: disk parameters and therefore it is normalized as to yield the same fraction of obscured
713: AGNs at low luminosities where the feedback is not effective.
714: \vspace{0.4cm}}
715: 
716: We stress that the inverse correlation between the fraction of obscured objects and luminosity constitutes a 
717: key {\it test} for AGN feedback as a source of the
718: luminosity-dependence of AGN obscuration; indeed, when such a feedback
719: is not included, an opposite correlation is found (as shown in fig. 3) since more luminous AGNs result from a 
720: larger fraction of cold gas available for both accretion and obscuration.
721: Note that qualitatively similar results are obtained for a range of AGN feedback efficiencies  
722: $2\,10^{-2}\leq \epsilon_{AGN}\leq 0.2$. Values outside this range would primarily  violate  
723: constraints set by galaxy colors, the $M_{BH}-\sigma$ correlation, and the specific star formation 
724: rate of massive galaxies at high redshifts. 
725: 
726: We expect the overall absorption to evolve strongly
727: with redshift, as shown in fig. 4 where we plot as a
728: function of $z$ the fraction of AGN (with any luminosity)
729: absorbed by gas with column density $N_H\geq 10^{23}$ cm$^{-2}$.
730: Such a fraction of absorbed AGNs rises appreciably with redshift, in
731: agreement with the observational findings by La Franca et al.
732: (2005). Although there is still some disagreement 
733: among observers on whether the AGN obscuration is purely luminosity dependent 
734: or both luminosity and redshift dependent (see Gilli et al. 2007 and references therein), 
735: the latter case constitutes 
736: a generic expectation of models in which obscuration is generated 
737: on host galaxy scales. This is due to
738: the larger fraction of cool galactic gas available for both
739: BH accretion and AGN absorption at high redshifts; in turn, this
740: is due to the rapid gas cooling corresponding to the higher
741: densities of the collapsed structures hosting the AGNs, and to the
742: frequent galaxy merging events which allow to continuously replenish
743: the cold gas to be turned into stars at high redshift.
744: Note that in our model the fraction of
745: obscured objects with $N_H\geq 10^{22}$ cm$^{-2}$ plotted in fig. 4
746: includes Compton-thick sources, since it does not resolve the
747: inner regions of AGNs. These are the regions where the Compton-thick absorbers are
748: presumably located according to several observational indications, from the local abundance of
749: Compton-thick sources (favoring large covering factors and consequently
750: compact sizes, see Risaliti, Maiolino, Salvati 1999),
751: to the direct limits from Chandra observations of the Circinus galaxy
752: (see Guainazzi et al. 2001 and references therein) to the rapid oscillation timescale
753: of the absorber in NGC1365 (Risaliti et al. 2005a).
754: In the fig. 4, we tentatively estimate the 
755: contribution of Compton-thin sources (dashed line) to the obscured fraction
756: by assuming that the difference between the predicted and the observed
757: density of AGNs shown in fig. 2 is entirely due to Compton-thick sources. 
758: Finally, in fig. 4 we give our predictions concerning the  
759: redshift evolution of the absorbed fraction of AGNs with different luminosities, 
760: which are shown by the thiner lines. Note how for $z\gtrsim 2$
761: the luminosity dependence of 
762: the absorbed fractions weakens for faint AGNs with $L_X\lesssim 3\, 10^{44}$ erg/s, while 
763: it exhibits a sharp transition to low values $\lesssim 0.5$ 
764: for brighter AGNs. 
765: 
766: \begin{center}
767: \vspace{0.cm}
768: \scalebox{0.45}[0.45]{{\includegraphics{f4.pdf}}}
769: \end{center} {\footnotesize \vspace{-0.1cm }
770: Fig. 4. - 
771: The evolution of the predicted
772: fraction of absorbed ($N_H\geq 10^{22}$ cm$-2$) AGNs for all sources
773: with  43$\leq$ log $L_X$/erg s$^{-1}\leq$ 46 (solid heavy line). The data from  La Franca
774: et al. (2005)  have been corrected for the selection effects.
775: The heavy dotted line represents the contribution to the fraction of absorbed AGNs
776: due to only Compton-thick sources estimated assuming that the
777: difference between the observed and the predicted AGN density in fig. 2 is entirely due
778: to the Compton-thick AGNs. 
779: The thin solid lines correspond to the absorbed fraction of AGNs within luminosity 
780: bins of width 0.25 dex centered on the following values: 
781: $L_X$/erg s$^{-1}=$ 43.25 (green, dashed line), $L_X$/erg s$^{-1}=$ 44.25 (blue, long 
782: dashed line), $L_X$/erg s$^{-1}=$ 44.7 (cyan, dot dashed line).
783: \vspace{0.4cm}}
784: 
785: Such a {\it combined} luminosity and redshift dependence of AGN
786: absorption, which constitutes a {\it specific prediction} of our model,  is 
787: shown in fig. 5 for a continuous, wide range of luminosities and redshift. 
788: As redshift increases, the overall fraction of absorbed
789: AGNs increases, but with a luminosity dependence much
790: weaker than holding at low $z$ (represented in fig. 2)  
791: for luminosities $L_X\lesssim 3\, 10^{44}$ erg/s. 
792: In fact,  the quantity of absorbing gas is predicted to be large enough (due
793: to the processes discussed above) as to make the expanding
794: blast wave induced by the AGN feedback ineffective to expell a major
795: fraction of the galactic gas, which is continuously replenished in
796: the galactic potential wells. For higher luminosities at $z\gtrsim 2$
797: a sharp transition occurs between a highly obscured population 
798: with $L_X\lesssim 3\,10^{44}$ erg/s and a nearly unobscured population 
799: with $L_X\gtrsim 10^{45}$ erg/s. 
800: Thus, a testable prediction of our model 
801: is that for intermediate
802: luminosities ($10^{43} \lesssim L_X/{\rm erg\,s^{-1}}\lesssim 10^{45}$) 
803: the ratio of optical to X-ray luminosity functions 
804: at redshifts $z\gtrsim 2$ should be much {\it lower} than that
805: at z=0, while for higher luminosities it remains {\it close} to the local value.
806: 
807: \begin{center}
808: \vspace{0.2cm}
809: \scalebox{0.8}[0.8]{{\includegraphics{f5.pdf}}}
810: \end{center} {\footnotesize \vspace{0.cm }
811: Fig. 5. - 
812: The combined luminosity and redshift dependence of the absorbed 
813: fraction of AGNs, represented by the color code shown on top.
814: \vspace{0.4cm}}
815: 
816: 
817: \section{Summary and Conclusions}
818: 
819: We have computed the effects of the galactic absorption on AGN emission in a
820: cosmological context, by including a physical model for AGN fueling and feedback
821: into a semi-analytic model of galaxy formation in the concordance cosmology.
822: The model is based on galaxy 
823: interactions as triggers for AGN accretion and on expanding blast waves as a
824: mechanism to propagate outwards the AGN energy injected into the interstellar
825: medium at the center of galaxies.
826: 
827: We have shown that an {\it inverse} dependence of AGN absorption on
828: luminosity (fig. 3) and a {\it direct} dependence on redshift (fig. 4) is
829: a natural outcome in such a context. The former arises from the
830: faster expansion of blast waves induced by feedback of energetic,
831: luminous AGNs onto the galactic interstellar gas; the rapid sweeping
832: of gas in the inner regions, where the density was initially
833: higher, results in the fast formation of a gas-depleted region with 
834: size larger for higher AGN intrinsic luminosities. Qualitatively 
835: similar conclusions were reached by Hopkins et al. (2005b) on the basis of
836: dedicated hydrodynamical N-body simulations.
837: On the other hand, the redshift dependence of the AGN absorption is due to
838: the larger amount of cold galactic gas available at high redshift, when the
839: higher densities allowed for fast cooling occurring in galactic
840: haloes. The quantitative {\it predictions} of our model are consistent
841: with existing observations concerning the fraction
842: of absorbed ($N_H\geq 10^{22}$ cm$^{-2}$) AGNs as a function of their luminosity
843: and redshift. Our model specifically predicts that 
844: for AGNs with $L_X\leq 3\,10^{44}$ erg/s
845: the luminosity dependence of the absorbed fraction  weakens
846: with increasing redshift (see fig. 5), while 
847: for the brightest objects with $L_X\gtrsim 3\,10^{44}$
848: the absorbed fraction quickly decreases with luminosity for $z\gtrsim 2.5$ (see fig. 5). 
849: 
850: Note that, after averaging over the line of sight, unobscured or mildly obscured
851: AGNs correspond to late stages of the feedback action; in particular, for a given
852: orientation of the line of sight, the observed column density depends on the time
853: elapsed since the start of the blast wave expansion. The faster expansion characterizing the blast wave of luminous AGNs thus corresponds to a {\it larger probability} to observe them when the blast has already swept out the central regions of the galaxy IGM.
854: This picture constitutes an {\it extension} of the unified picture for AGNs (see Antonucci 1993) beyond the canonical scheme based on the single orientation parameter, since the absorption
855: properties now depend on the combination of {\it orientation} and {\it time} needed to sweep the central regions of the galaxy disk. Note that our picture has also straightforward implications on the connection between star formation and obscuration properties
856: (proposed , since from our results we on average expect larger star formation in heavily obscured objects, while luminous, mildly absorbed AGNs should be generally associated to galaxies with lesser star formation (or  in transition to a passive state). 
857: The relevance of evolutionary effects of the kind modeled here in determining the 
858: absorption properties of AGNs is supported by several observational works. 
859: Stevens et al.  (2005) and Page et al. (2004) find that X-ray obscured QSOs have 
860: much higher submillimeter detection rates than X-ray unobscured
861: QSOs, suggesting strong star formation on-going in the host of  
862: obscured AGN only.  Sajina et al. (2007) and Martinez-Sansigre et al. (2008) report Spitzer IRS spectra dominated by AGN continuum but showing PAHs features in
863: emission, typical of starforming galaxies, in samples of ULIRGs and  
864: radio selected obscured QSOs at z$\sim2$. Lacy et al.  (2007) find evidence for
865: dust-obscured star formation in type-2 QSOs. From the analysis of the X-ray background, 
866: Baalntyne ey al. (2006) argue that the AGN obscuration is connected with the star formation in the 
867: host galaxy. Finally, Martinez-Sansigre et al. (2005, 2008) found little
868: or no Lyman-$\alpha$ emission in a sample of z$>1.7$ obscured QSOs,
869: suggesting large scale (kpc) dust distribution. All these findings
870: are in general agreement with the evolutionary picture, although 
871: some disagreeing works argue for a dominance of geometrical effect  (see Triester et al. 2008). A 
872: geometrical effect may be constituted by the gravitational bending of the 
873: interstellar gas due to the BH, as described in Lamastra et al. (2006); however, except for very massive 
874: BHs ($M_{BH}\sim 10^9\,M_{\odot}$), this  
875: will affect mainly regions below our resolution scale of 50 $pc$. Another class of geometrical models, 
876: based on the luminosity dependence of the sublimation radius of the BH accretion disk,  is that commonly referred to as 
877: the ''receeding torus'' picture (Lawrence 1991). These
878: however provide an appreciable luminosity dependence only at very high AGN luminosities $L_x\gtrsim 10^{44-45}$ erg/s 
879: and for dust located close to the sublimation radius ($1-10$ pc). Dust located in galaxy disk can hardly be affected directly by the AGN radiation. 
880: 
881: In our model we did not try to model the processes leading to
882: Compton-thick absorption with $N_H\geq 10^{24}$ cm$^{-2}$,
883: generally thought to be caused by gas directly associated with the
884: central regions of the AGNs. On the other hand, our model provides a hint
885: concerning luminosity and redshift dependence to be expected as for the abundance of
886: Compton-thick sources. Inspection of
887: fig. 2, where we compare the predicted global (including
888: Compton-thick sources) density of AGNs with different {\it
889: intrinsic} luminosities with data corrected for absorption (but not including
890: Compton-thick sources), shows
891: that our model predicts a number of low/intermediate-luminosity AGNs
892: ($L_X\leq 10^{44}$ erg/s$^{-2}$) larger than the observed
893: Compton-thin sources by a factor around 2 at $z\gtrsim 2$. 
894: While it is possible that our model overestimates the AGN fueling in this 
895: range of $L_X$ and $z$ (e.g., due to its specific modeling of 
896: interaction-driven destabilization for cold gas in galactic disks), 
897: the above excess could support the view that at such luminosities and redshifts 
898: a fraction around 1/2 of the total AGNs are Compton-thick.
899: A complementary process which may explain our overprediction  of
900: low-luminosity AGNs at $z\gtrsim 2$ is suppression of BH growth
901: in small mass galactic haloes (DM masses $M\leq 10^9\,M_{\odot}$). This may 
902: be provided by gravitational-rocket effect on the BHs due
903: to the recoil following the emission of gravitational waves
904: during the coalescence of BH binaries following galaxy mergers (see Madau \& Quataert 2004).
905: Such a recoil may produce BH velocities of order $10^2$ km/s,
906: sufficient to unbind the hole from galaxies with
907: DM velocity dispersion $\lesssim 50$ km/s.
908: In this respect, the observational selection of
909: Compton-thick sources by combining mid-infrared to near-infrared
910: and optical photometry of galaxies (Fiore et al. 2008a,b) will provide
911: crucial constrains on the
912: relative role of obscuration and BH depletion
913: in low-mass galactic halos.
914: 
915: Our results are also relevant for constraining the physical
916: mechanisms of AGN feedback onto the interstellar gas, a key issue
917: in recent developments of cosmological
918: galaxy formation models. Indeed, the recent realizations of such
919: models include AGN feedback as the key process to suppress gas cooling
920: in massive galactic haloes. Actually, two kinds of feedback are at
921: present implemented in this context: on the one hand, models based
922: on galaxy interactions as triggers of AGN activity and feedback
923: relate the latter to the bright, accretion
924: phase onto supermassive black holes, i.e., to the same phase
925: corresponding to AGN activity (see Menci et al. 2003, 2007; Di
926: Matteo et al. 2005; Hopkins et al. 2005a); on the other hand, other
927: authors associate the AGN feedback only to a quiescent (so called
928: ''radio'') phase of accretion, characterized by very low
929: accretion rates ($\lesssim 10^{-2}\,M_{\odot}$/yr) and not
930: observable as radiative AGNs (Bower et al. 2006, Croton et al. 2006). Since
931: the latter mode would result into a feedback activity continuing down
932: to low redshift, we expect such models to provide a much milder
933: dependence of AGN absorption on redshift;  so observational
934: results on the high-redshift absorption of AGN emission will
935: constitute an effective test for models of AGN feedback.
936: 
937: Finally, we note that the feedback related to the active AGN phase described
938: here is effective to decrease the galactic gas and the associated
939: absorption mainly at low redshift $z\lesssim 2$ and for bright
940: ($L_X\gtrsim 10^{44}$ erg/s) AGNs (see, e.g., Cavaliere \& Menci 2007).
941: This implies that at high
942: redshifts the effective cooling and the continuous replenishing of
943: galactic gas due to fequent merging events will override the depletion
944: due to the AGN feedback, so the latter can not suppress the
945: early growth of supermassive BH. Indeed, the predicted number
946: density of early ($z\geq 4$), bright ($L_x\geq 10^{45.5}$ erg/s)
947: AGNs shown in fig. 2 is consistent with that observed for bright
948: ($M_i\leq -27.5$) optical QSOs up to $z\approx 6$ (Hopkins et al.
949: 2006). Note however that such a result does not include the BH
950: spin-up which may occur during the growth due to accretion of gas 
951: endowed with angular momentum 
952: (for different modeling of such a process see, e.g., Volonteri
953: et al. 2005; King, Pringle \& Hofman 2007),
954: which in turn may yield larger radiative efficiencies
955: up to values $\eta\approx 0.3$ for a dominant
956: fraction of BHs; as noted by the above authors, the lower
957: mass accretion rate related to larger radiation efficiencies may
958: delay the mass assembly of massive BH at early epochs when included
959: in a cosmological model.
960: We shall investigate these issues in a following paper.
961: 
962: \acknowledgments We thank our referee for helpful comments which contributed to improve the paper. We acknowledge grants from ASI. 
963: 
964: \vspace{2cm}
965: \section*{APPENDIX A}
966: \setcounter{equation}{0} \renewcommand{\theequation}{A-\arabic{equation}}
967: 
968: 
969: The key quantity determining the blastwave evolution is energy deposition ratio $\Delta E/E$  of the 
970: energy injected up to the time $t$ over to the ISM energy out to the shock radius
971: at the same time. This is constant in self-similar solutions, which 
972: imply for the energy injection the law 
973: $\Delta E(t)\propto t^{2(5-2\omega)/\omega}$ (sect. 5 and Appendix C2 of Lapi, Cavaliere, Menci 2005), normalized to yield 
974: the total value $\Delta E=\eta\,\epsilon_{AGN}\,\Delta m_{acc}$ (eq. 4) 
975: at the end of the AGN active phase. Its time behavior is thus related to  the 
976: exponent $\omega$ defining the decline of the gas density distribution. 
977: 
978: Since the supersonic wave affects the plasma only 
979: out to the shock distance $R_s(t)$ where the unperturbed (initial) energy content is 
980: $E(<R_s(t))\propto R_s(t)^{5-2\omega}\propto t ^{2(5-2\omega)/\omega}$ (Lapi et al. 2005), the ratio $\Delta E/E$ is constant, and  constitutes the basic quantity determining the strength of the shock; in other words, 
981: $\Delta E/E$ remains constant while the wave expands, since both
982: the energy injected up to the considered time and the ISM energy content within the shock radius grows in time with the 
983: same law. Thus we choose to compute $\Delta E/E$ as the total  energy injected by the AGN divided into the total
984:  energy content of the ISM within the whole volume of the disk. 
985: 
986: The luminosity $L$ is the time derivative of $\Delta E$, and thus in the self-similar model 
987:  is to behave as $L= d \Delta E/dt\propto t^{5(2-\omega)/\omega}$; so its decay is  related to the exponent $\omega$ 
988:  defining the density decline (see Lapi, Cavaliere, Menci 2005, Sect.5 and Appendix C2). 
989: 
990: For our piecewise approximation of the density profile (point 1 in Sect. 3.2), $L$ is constant during the time when the 
991: wave sweeps the inner region $r\leq r_d/2$, declines as $L\propto t^{-0.65}$ when the blast has expanded into the region 
992: $r_d/2< r \leq 3\,r_d/2$, and then drops to zero when $r>3\,r_d/2$, where $\Delta E$ remains constant at the value attained 
993: at the border $r=3r_d/2$, consistently with the index $\omega=5/2$ holding in that region.
994: 
995: Full consistency with the self-similar blastwave model would require us to assigne each AGN a 
996: time-dependent luminosity, decaying as described above. However, 
997: the results are indistingishable from those obtained with the simple prescription $L$=constant 
998: (provided both luminosities are normalized to the same final total injection $\Delta E$) for two reasons.
999: 
1000: 1) As for the central $r\leq r_d/2$ region $L$ is indeed constant, 
1001: as we showed above; in the outer region the 
1002: self-similar behavior $L\propto t^{-0.65}$ results in a much slower decay of the luminosity 
1003: compared to the decline of the ISM density encountered by 
1004: the blast $\rho\propto R_s^{-\omega}$. In fact, 
1005: using eq. (5) for the time expansion of $R_s(t)\propto t^{2/\omega}$, 
1006: the density in front of the blast declines as $\rho\propto t^{-2}$. 
1007: Clearly, the regime when the blast expands into the outer region $r>3/2\,r_d$ is not important, since here the luminosity is null and the object is not recognized as an active AGN in computing
1008: the $N_H$ distributions shown in Sect. 4. 
1009: 
1010: 2) Most important, the scale height of the disk $h= r_d/15$ is much shorter than the radius $r_d/2$ enclosing the first radial shell we use to fit the exponential profile of the disk. This in turn means that nearly all 
1011: the random line-of-sights that we draw from the galaxy center only intercept the first radial shell, 
1012: where the luminosity is effectively constant according to the self-similar construction of the model.
1013: Quantitavely, the probability for a line-of-sight to intercepts the outer ($r>r_d/2$) disk 
1014: is $[arctan(2/15)]/(\pi/2)\approx 8\,10^{-2}$. 
1015: Beyond the details of our specific self-similar blastwave model, the physical point is that 
1016: the geometry of the disk is such that the vast majority of the line-of-sights is only affected by 
1017: the inner region; the latter is completely swept out by the blast early on, when the luminosity decay has not yet occurred. 
1018: \newpage
1019: 
1020: \begin{references}
1021: 
1022: \reference{} Antonucci, R. 1993, ARAA, 31, 473
1023: 
1024: \reference{} Ballantyne, D.R., Everett, J.E., Murray, N. 2006, ApJ, 639, 740
1025: 
1026: \reference{} Barcons, X., Mateos, S., \& Ceballos, M. T. 2000, MNRAS, 316, L13
1027: 
1028: \reference{} Begelman, M.C. 2003, in {\it Coevolution of Black Holes
1029: and Galaxies}, Carnegie Observatories Astrophysics Series, Vol. 1,
1030: ed. L.C. Ho (Cambridge: Cambrdidge University Press)
1031: 
1032: \reference{} Bond, J.R., Cole, S., Efstathiou, G., \& Kaiser, N.,
1033: 1991, ApJ, 379, 440
1034: 
1035: \reference{} Bower, R.G. et al. 2006, MNRAS, 370, 645
1036: 
1037: \reference{} Cavaliere, A.,  \& Vittorini, V.  2000, ApJ, 543, 599
1038: 
1039: \reference{} Cavaliere, A., Lapi, A., Menci, N., 2002, ApJ, 581, L1
1040: 
1041: \reference{} Cavaliere, A., Menci, N. 2007, ApJ, 664,47
1042: 
1043: \reference{} Chartas, G., Brandt, W. N., Gallagher, S. C., Garmire,
1044: G. P., 2002, ApJ, 579, 169
1045: 
1046: \reference{} Churazov, E., Br\"{u}ggen, M., Kaiser, C., B\"{o}ringer, H.,
1047: Forman, W. 2001, ApJ, 554, 261
1048: 
1049: \reference{} Crenshaw, D. M., Kraemer, S. B., \& George, I. M. 2003, ARA\&A, 41, 117
1050: 
1051: \reference{} Croton, D.J, Springel, V., White, S.D.M., De Lucia, G.,
1052: Frenk, C.S., Gao, L., Jenkins, A., Kauffmann, G., Navarro, J.F.,
1053: Yoshida, N., 2006, MNRAS, 365, 11
1054: 
1055: \reference{} Di Matteo, T., Springel, V., Hernquist, L., 2005,
1056: Nature, 433, 604
1057: 
1058: \reference{}Elvis, M. 2000, ApJ, 545, 63
1059: 
1060: \reference{} Elvis, M. 2006, Memorie della Societa Astronomica Italiana, v.77,
1061: p.573, Proceedings of the Workshop on "AGN and Galaxy Evolution",
1062: Specola Vaticana, Castel Gandolfo
1063: 
1064: \reference{} Fan X. et al. 2004, ApJ, 128, 515
1065: 
1066: \reference{} Ferrarese, L. Merritt, D., 2000, ApJ, 539, L9
1067: 
1068: \reference{} Fiore, F. et al. 2008a, ApJ, 672, 94
1069: 
1070: \reference{} Fiore, F. et al. 2008b, ApJ, submitted
1071: 
1072: \reference{} Forman, W. et al. 2005, ApJ, 635, 894
1073: 
1074: \reference{} Gallimore, J.F., Baum, S.A.,  O'Dea, C.P. 1997, Nature, 388, 852
1075: 
1076: \reference{} Gebhardt, K. et al., 2000, ApJ, 539, L13
1077: 
1078: \reference{} Gilli, R., Comastri, A., Hasinger, G. 2007, A\&A, 463, 79
1079: 
1080: \reference{} Granato, G.L, De Zotti, G., Silva, L., Bressan, A., Danese, L. 2004, ApJ, 600, 580
1081: 
1082: \reference{} Guainazzi, M., Fiore, F., Matt, G., Perola, G. C. 2001, MNRAS, 327, 323
1083: 
1084: \reference{} Hasinger, G. 2008, A\& A, submitted
1085: 
1086: \reference{} King, A.R., Pringle, J.E., Hofmann, J.A. 2008, preprint [astro-ph/08011564]
1087: 
1088: \reference{} La Franca, F. et al. 2005, ApJ, 635, 864
1089: 
1090: \reference{}Lacey, C., \& Cole, S., 1993, MNRAS, 262, 627 
1091: 
1092: \reference{} Lacy, M. et al. 2007, ApJ, 669, L61
1093: 
1094: \reference{} Hopkins, P.F., Hernquist, L., Cox, T.J., Di Matteo, T.,
1095: Martini,P., Robertson, B., Springel, V. 2005a, ApJ, 630, 705
1096: 
1097: \reference{} Hopkins, P.F., Hernquist, L., Cox, T.J., Di Matteo, T., 
1098: Robertson, B., Springel, V. 2005b, ApJ, 632, 81
1099: 
1100: \reference{} Hopkins, P.F. Hernquist, L., Cox, T.J., Di Matteo, T.,
1101:  Robertson, B., Springel, V. 2006, ApJS, 163, 1
1102: 
1103: \reference{} Lacey, C.G. and Cole, S. 1993, MNRAS,  262, 627
1104: 
1105: \reference{} Lamastra, A., Perola, G.C., Matt, G. 2006, A\&A, 449, 551
1106: 
1107: \reference{} Lapi, A., Cavaliere, A., \&  Menci, N. 2005, ApJ, 619, 60
1108: 
1109: \reference{} Lawrence, A., Elvis, M. 1982, ApJ, 256, 410
1110: 
1111: \reference{} Lawrence, A. 1991, MNRAS, 252, 586
1112: 
1113: \reference{} Madau, P., \& Rees, M.J., 2001, ApJ, 551, L27
1114: 
1115: \reference{} Madau, P., \& Quataert, E., 2004, ApJ, 606, L17
1116: 
1117: \reference{} Malkan M.A., Gorjian V., Tam R. 1998, ApJS, 117, 25
1118: 
1119: \reference{} Marconi, A., Risaliti, G., Gilli, R., Hunt, L.K.,Maiolino, R., Salvati, M. 2004, MNRAS, 351, 169
1120: 
1121: \reference{} Martinez-Sansigre A. et al. 2005, Nature, 436, 666
1122: 
1123: \reference{} Martinez-Sansigre A. et al. 2007, MNRAS, 379, L6
1124: 
1125: \reference{} Martinez-Sansigre, A. Lacy, M., Sajina, A., Rawlings, S., 2008, ApJ, 674, 676
1126: 
1127: \reference{} Matt, G. 2000, MNRAS, 335, L31
1128: 
1129: \reference{} McNamara, B.R., Nulsen, P.E.J. 2007, ARAA, 45, 117
1130: 
1131: \reference{} Menci, N., Cavaliere, A., Fontana, A., Giallongo, E., Poli, F., Vittorini, V. 2003, ApJ, 587, L63
1132: 
1133: \reference{} Menci, N., Cavaliere, A., Fontana, A., Giallongo, E., \& Poli, F. 2004,
1134: ApJ, 604, 12
1135: 
1136: \reference{} Menci, N., Fontana, A., Giallongo, E.,  Salimbeni, S.
1137: 2005, ApJ,  632, 49
1138: 
1139: \reference{} Menci, N., Fontana, A., Giallongo, E., Grazian, A.,
1140: Salimbeni, S. 2006, ApJ,  647, 753
1141: 
1142: \reference{}Fontanot, F., Monaco, P., Cristiani, S., Tozzi, P. 2006, MNRAS, 373, 1173
1143: 
1144: \reference{} Morganti, R., Tadhunter, C. N., Oosterloo, T. A., 2005,
1145: A\&A, 444, L9
1146: 
1147: \reference{} Murray, N., Quataert, E., Thompson, T.A. 2005, ApJ, 618, 569 
1148: 
1149: \reference{} Narayan, C.A., Jog, C.J. 2002, A\&A, 394, 89
1150: 
1151: \reference{} Nagar, N.M., Wilson, A.S. 1999, ApJ, 516, 97
1152: 
1153: \reference{} Nulsen, P.E.J, Fabian, A.C. 2000, MNRAS, 311, 346
1154: 
1155: \reference{} Page, M.J., Stevens, J.A., Ivison, R.J, Carrera, F.J 2004, ApJ, 611, L85
1156: 
1157: \reference{} Pounds, K., King, A.R., Page, K.L., O'Brien, P.T 2003,
1158: MNRAS, 346, 1025
1159: 
1160: \reference{} Pounds, K., Page, K.L., 2006, MNRAS, 372, 127
1161: 
1162: \reference{} Pozzi, F. et al. 2007, A\&A, 468, 603
1163: 
1164: \reference{} Proga, D. 2007, ApJ, 661, 693
1165: 
1166: \reference{} Reynolds, C.S., Heinz S., Begelman M.C. 2001. ApJ, 549,
1167: L179
1168: 
1169: \reference{} Richards, G.T. et al. 2006, ApJ, 131, 2766
1170: 
1171: \reference{} Richstone, D., et al. 1998, Nature, 395, 14
1172: 
1173: \reference{} Risaliti, G., Maiolino, R., Salvati, M. 1999, ApJ, 522, 157
1174: 
1175: \reference{} Risaliti, G., Elvis, M., Fabbiano, G., Baldi, A., Zezas, A. 2005a, ApJ, 623, 93
1176: 
1177: \reference{} Risaliti, G., Bianchi, S., Matt, G., Baldi, A., Elvis, M., Fabbiano, G.,Zezas, A. 2005b, ApJ, 630, L129
1178: 
1179: \reference{} Sajina et al. 2007, ApJ, 664, 713
1180: 
1181: \reference{} Salim, S. 2007, ApJS, 173, 267
1182: 
1183: \reference{} Saslaw, W.C., 1985, {\it  Gravitational Physics of Stellar and 
1184: Galactic Systems} (Cambridge: Cambridge Univ. Press) 
1185: 
1186: \reference{} Sedov, L.I. 1959, {\it Similarity and Dimensional
1187: Methods in Mechanics}, (New York: Academic)
1188: 
1189: \reference{} Spergel, D.N. et al. 2006, ApJ,  in  press (astro-ph/0603449)
1190: 
1191: \reference{} Stevens, J.A. et al.  2005, MNRAS, 360, 610 
1192: 
1193: \reference{} Triester, E., Krolik, J.H., Dullemond, C.  2008, preprint (astro-ph/08013849)
1194: 
1195: \reference{} Thean, A.H.C., Gillibrand, T. I., Pedlar, A., Kukula, M. J. 2001, MNRAS, 325, 737
1196: 
1197: \reference{} Turnshek, D.A., Grillmair, C.J., Foltz, C.B., Weymann, R.J. 1988, ApJ, 325, 651
1198: 
1199: \reference{} Ueda, Y., Akiyama, M., Ohta, K., Miyaji, T. 2003, ApJ, 598, 886
1200: 
1201: \reference{} Volonteri, M., Madau, P., Quataert, E., Rees, M.J. 2005, ApJ, 620, 69
1202: 
1203: \reference{} Weymann, R.J. 1981, ARAA, 19,41
1204: 
1205: \reference{} Yu, Q., \& Tremaine, S., 2002, MNRAS, 335, 965
1206: 
1207: \end{references}
1208: 
1209: 
1210: \end{document}
1211: 
1212: 
1213: 
1214: 
1215: 
1216: