0710.4488/ms.tex
1: %\documentclass[11pt,preprint]{aastex}
2: %\documentclass[12pt,preprint]{amsart}
3: %\usepackage{natbib}
4: %\usepackage{emulateapj5}
5: \documentclass{emulateapj}
6: %\usepackage{lscape}
7: %\usepackage{rotcapt}
8: %\usepackage{rotating}
9: %\usepackage[hang,small,bf]{caption}
10: %\setlength{captionmargin}{20pt}
11: \usepackage{apjfonts}
12: %\usepackage{natbib}
13: 
14: 
15: %\shorttitle{ } \shortauthors{ }
16: 
17: \begin{document}
18: 
19: \def\sarc{$^{\prime\prime}\!\!.$}
20: \def\arcsec{$^{\prime\prime}$}
21: \def\arcmin{$^{\prime}$}
22: \def\degr{$^{\circ}$}
23: \def\seco{$^{\rm s}\!\!.$}
24: \def\ls{\lower 2pt \hbox{$\;\scriptscriptstyle \buildrel<\over\sim\;$}}
25: \def\gs{\lower 2pt \hbox{$\;\scriptscriptstyle \buildrel>\over\sim\;$}}
26: 
27: \title{Self-Consistent Models of the AGN and Black Hole Populations:
28:    Duty Cycles, Accretion Rates, and the Mean Radiative Efficiency}
29: 
30: \author{Francesco Shankar\altaffilmark{1}, David H. Weinberg\altaffilmark{1}
31: and Jordi Miralda-Escud\'{e}\altaffilmark{2}}
32: \altaffiltext{1}{Astronomy Department, Ohio State University,
33:     Columbus, OH-43210, U.S.A.}
34: \altaffiltext{2}{Institut de Ci\`encies de L'Espai (IEEC-CSIC)/ICREA,
35:     Bellaterra, Spain}
36: 
37: %\begin{doublespace}
38: 
39: \begin{abstract}
40: We construct evolutionary models of the populations of active galactic nuclei
41: (AGN) and supermassive
42: black holes, in which the black hole mass function grows at the rate
43: implied by the observed luminosity function, given assumptions about the
44: radiative efficiency and the luminosity in Eddington units. We draw on a variety of recent
45: X-ray and optical measurements to estimate the bolometric AGN luminosity
46: function and compare to X-ray background data and the independent estimate
47: of Hopkins et al.\ (2007) to assess remaining systematic uncertainties.
48: The integrated AGN emissivity closely tracks the cosmic star formation
49: history, suggesting that star formation and black hole growth are closely
50: linked at all redshifts. We discuss observational uncertainties in the
51: local black hole mass function, which remain substantial, with estimates of the
52: integrated black hole mass density $\rho_{\bullet}$ spanning the range
53: $3 - 5.5 \times 10^5\, {\rm M_{\odot}\, Mpc^{-3}}$. We find good agreement with
54: estimates of the local mass function for a reference model where all active black holes have a fixed
55: efficiency $\epsilon = 0.065$ and $L_{\rm bol}/L_{\rm Edd} \approx 0.4$
56: (shifting to $\epsilon = 0.09$, $L_{\rm bol}/L_{\rm Edd} \approx 0.9$ for
57: the Hopkins et al.\ luminosity function). In our reference model, the duty
58: cycle of $10^9 M_\odot$ black holes declines from 0.07 at $z=3$ to 0.004 at
59: $z=1$ and $10^{-4}$ at $z=0$. The decline is shallower for less massive
60: black holes, a signature of ``downsizing'' evolution in which more massive
61: black holes build their mass earlier. The predicted duty cycles and AGN
62: clustering bias in this model are in reasonable accord with observational
63: estimates. If the typical Eddington ratio declines at $z<2$, then the
64: ``downsizing'' of black hole growth is less pronounced. Models with reduced
65: Eddington ratios
66: at low redshift or black hole mass predict fewer low mass black holes
67: ($M_{\bullet}\lesssim 10^8\, {\rm M_{\odot}}$) in the local universe, while models
68: with black hole mergers predict more black holes at $M_{\bullet}>10^9\, {\rm M_{\odot}}$.
69: Matching the integrated AGN emissivity to the local
70: black hole mass density implies $\epsilon = 0.075 \times
71: (\rho_{\bullet} / 4.5\times 10^5\, {\rm M_{\odot}\, Mpc^{-3}})^{-1}$
72: for our standard
73: luminosity function estimate, or 25\% higher for Hopkins et al.'s estimate.
74: It is difficult to reconcile current
75: observations with a model in which most black holes have the high efficiencies
76: $\epsilon \approx 0.16-0.20$ predicted by MHD simulations of disk accretion.
77: We provide electronic tabulations of our bolometric luminosity function and our reference model
78: predictions for black hole mass functions and duty cycles as a function of
79: redshift.
80: \end{abstract}
81: 
82: \keywords{cosmology: theory -- black hole: formation -- galaxies:
83: evolution -- quasars: general}
84: 
85: \section{Introduction}
86: \label{sec|intro}
87: 
88: The long-standing hypothesis that quasars are powered by accretion
89: onto supermassive black holes (Salpeter 1964; Lynden-Bell 1969; Rees
90: 1984) is now strongly supported by many lines of evidence, including
91: the apparent ubiquity of remnant black holes in the spheroids of
92: local galaxies (Richstone et al. 1998). The strong correlations
93: between the masses of central black holes and the luminosities,
94: dynamical masses, and velocity dispersions of their host spheroids
95: (e.g., Magorrian et al. 1998; Ferrarese \& Merritt 2000; McLure \&
96: Dunlop 2002; Marconi \& Hunt 2003; H\"{a}ring \& Rix 2004; Ferrarese
97: \& Ford 2005; Greene \& Ho 2006; Graham 2007; Hopkins et al. 2007b;
98: Shankar \& Ferrarese 2008; Shankar et al. 2008) imply that the
99: processes of black hole growth and bulge formation are intimately
100: linked. Theoretical models typically tie black hole growth to
101: episodes of rapid star formation, perhaps associated with galaxy
102: mergers, and ascribe the black hole-bulge correlations to energy or
103: momentum feedback from the black hole or to regulation of black hole
104: growth by the bulge potential (e.g., Silk \& Rees 1998; Kauffmann \&
105: Haehnelt 2000; Cavaliere \& Vittorini 2000; Granato et al. 2004,
106: 2006; Murray, Quataert, \& Thompson 2004; Cattaneo et al. 2005;
107: Miralda-Escud\`{e} \& Kollmeier 2005; Monaco \& Fontanot 2005;
108: Croton et al. 2006; Hopkins et al. 2006a, Malbon et al. 2006). These
109: correlations also make it possible to estimate the mass function of
110: black holes in the local universe (e.g., Salucci et al. 1999; Yu \&
111: Tremaine 2002; Marconi et al. 2004; Shankar et al. 2004, S04
112: hereafter). This local mass function provides important constraints
113: on the co-evolution of the quasar and black hole populations. The
114: most general and most well known of these constraints is the link
115: between the integrated emissivity of the quasar population, the
116: integrated mass density of remnant black holes, and the average
117: radiative efficiency of black hole accretion (So{\l}tan 1982; Fabian
118: \& Iwasawa 1999; Elvis et al. 2002).
119: 
120: In the paper we construct self-consistent models of the quasar population
121: using a method that can be considered a ``differential'' generalization
122: of So{\l}tan's (1982) argument.
123: Given assumed values of the radiative efficiency and the Eddington ratio
124: $L/L_{\rm Edd}$, the observed luminosity function of quasars at a given redshift can be
125: linked to the average growth rate of black holes of the corresponding mass,
126: and these growth rates can be integrated forward in time to track the evolution
127: of the black hole mass function. This modeling approach has been developed and applied
128: in a variety of forms by numerous authors, drawing on steadily improving observational data
129: (e.g. Cavaliere et al. 1973; Small \&
130: Blandford 1992; Salucci et al. 1999;
131: Cavaliere \& Vittorini 2000; Yu \& Tremaine 2002;
132: Steed \& Weinberg 2003, hereafter SW03; Hosokawa 2004; Yu \& Lu 2004; Marconi et al.
133: 2004; S04; Merloni 2004; Vittorini,
134: Shankar \& Cavaliere 2005; Tamura et al. 2006; Hopkins et al. 2007).
135: The consensus of recent studies is that evolutionary models with radiative efficiencies
136: of roughly 10\% and mildly sub-Eddington accretion rates yield a reasonable
137: match to the observational estimates of the remnant mass function. However, uncertainties
138: in the bolometric luminosity function of active galactic nuclei (AGN, a term we will use to describe both quasars and less
139: luminous systems powered by black hole accretion) and in the local black hole mass function remain
140: an important source of uncertainty in these conclusions.
141: 
142: We begin our investigation by constructing an estimate of the bolometric AGN luminosity function.
143: Our estimate starts from the model of Ueda et al. (2003, hereafter
144: U03), based
145: on data from several X-ray surveys, but we adjust its parameters based on more recent
146: optical and X-ray data that provide more complete coverage of luminosity and redshift. Comparison to the X-ray background
147: and to the independent luminosity function estimate of Hopkins, Richards \& Hernquist (2007; HRH07 hereafter)
148: gives an indication of the remaining uncertainties, associated mainly with bolometric corrections and the fraction
149: of obscured sources. In agreement with Marconi et al. (2004) and Merloni (2004), we find
150: similar trends in the evolution of AGN emissivity and the cosmic star formation rate, supporting
151: models in which the growth of black holes occurs together with the formation of stars in their
152: hosts (e.g., Sanders et al. 1988; Granato et al. 2004; Hopkins et al. 2006a). We also reassess current
153: estimates of the local black hole mass function, finding that different choices and calibrations of
154: the black hole-bulge correlation lead to significantly different results, with integrated mass densities
155: that span nearly a factor of two.
156: 
157: Our simplest models of the evolving black hole and AGN populations assume that all active black holes have a single radiative
158: efficiency $\epsilon$ and a single accretion rate $\dot{m}$ in Eddington units. The work of Kollmeier et al. (2006)
159: suggests that a constant value of $\dot{m}$ may be a reasonable approximation
160: for luminous quasars, but the observational evidence
161: on this point is mixed (e.g., McLure \& Dunlop 2004;
162: Vestergaard 2004; Babic et al. 2007; Bundy et al. 2007; Netzer \& Trakhtenbrot 2007; Netzer et al. 2007; Rovilos \& Gorgantopoulos 2007;
163: Shen et al. 2007b),
164: and for lower luminosity, local
165: AGN there is clearly a broad range of Eddington ratios (e.g., Heckmann et al. 2004).
166: We also consider models in which $\dot{m}$ depends on redshift or black hole mass, and we consider
167: a simple model of black hole mergers to assess their potential impact on the mass function.
168: We will examine models with distributions of $\dot{m}$ values and a more realistic treatment
169: of mergers in future work.
170: 
171: Our modeling allows a consistency test of the basic scenario in which the observed luminosity
172: of black hole accretion drives the growth of the underlying black hole population, and
173: comparison to the local black hole mass function yields constraints on the average
174: radiative efficiency and the typical accretion rate. For specified $\epsilon$ and $\dot{m}$,
175: the model predicts the black hole mass function as a function of time, and combination
176: with the observed luminosity function yields the duty cycle as a function of redshift and black hole mass.
177: These predictions can be tested against observations of AGN host galaxy properties and AGN clustering,
178: and they can be used as inputs for further modeling. While significant uncertainties
179: remain, we find that a simple model of the black hole and AGN populations achieves
180: a good match to a wide range of observational data.
181: 
182: \section{The AGN bolometric luminosity function}
183: \label{sec|AGNLF}
184: %To use equations~(\ref{eq|charact}) and ~(\ref{eq|mdotav})
185: In order to constrain the
186: accretion history of black holes, it is essential to know the shape
187: and evolution with redshift of the AGN bolometric LF. The
188: determination of this LF is not straightforward, as it must be made
189: using the AGN LF \emph{observed} in particular bands and converted
190: into bolometric estimates using empirically calibrated bolometric
191: corrections. The bolometric corrections have been shown by several
192: authors to be redshift independent, with some possible trend with
193: the intrinsic luminosity of the source (e.g., Elvis et al. 1994;
194: Marconi et al. 2004; Richards et al. 2006; Hopkins et al. 2007a). AGN
195: surveys have been carried out at several energy ranges, and in each
196: one the detection of sources can be highly affected by intrinsic or instrumental selection effects.
197: In different ranges of
198: redshift and luminosity, the best statistics come from
199: different wavebands or surveys, further complicating the effort to create a comprehensive
200: AGN LF.
201: 
202: An additional challenge in assembling a reliable and complete census
203: of the AGN population is obscuration of the central engine by gas
204: and dust, which may reside in the ``molecular torus'' of unified
205: models (Antonucci 1993) or in the interstellar medium of the
206: galactic host (Martinez-Sansigr\'{e} et al. 2005; Rigby et al.
207: 2006). Strong obscuration can eliminate sources from surveys
208: entirely, while uncorrected weak obscuration causes their
209: luminosities to be underestimated. For solar composition gas, the
210: absorption optical depth to X-ray photons is
211: $\tau=2.04(N_H/10^{22}\, {\rm cm^{-2}})(E/1\, {\rm keV})^{-2.4}$
212: (e.g., Kembhavi \& Narlikar 1999). For Galactic dust-to-gas ratio,
213: the extinction in the rest-frame visual band is $A_V=5.35{\rm mag}
214: \times \,(N_H/10^{22}\, {\rm cm^{-2}})$ (e.g., Binney \& Merrifield
215: 2000). Hard X-ray selection is thus the least affected by
216: obscuration, and deep X-ray surveys indeed reveal many faint AGNs
217: that are missed by traditional optical selection criteria (e.g.,
218: Hasinger et al. 2005). Numerous studies suggest that the incidence
219: of obscuration decreases towards high intrinsic luminosity, so that
220: X-ray and optical LFs agree at the bright end but diverge at low
221: luminosities (e.g., U03; Richards et al. 2005; La Franca et al.
222: 2005; Hasinger et al. 2005). Statistically speaking, there is fairly
223: good correspondence between X-ray AGNs with $\log N_H/{\rm cm^{-2}}
224: \le 22$ and broad-line optical AGNs (e.g., Tozzi et al. 2006). Even
225: 2-10 keV selection becomes highly incomplete for Compton-thick
226: sources, with $\log N_H/{\rm cm^{-2}} \ge {24}$, and the best
227: constraints on this population come from the normalization and
228: spectral shape of the X-ray background.
229: 
230: Given these considerations, we have chosen to base our bolometric LF
231: estimate principally on the work of U03, who compiled a vast sample
232: from \emph{Chandra}, \emph{ASCA}, and \emph{HEAO}-1 surveys and inferred
233: the absorption-corrected LF in the rest-frame 2-10 keV band out to $z\sim 3$.
234: We adopt the luminosity-dependent bolometric correction of Marconi
235: et al. (2004); we include a dispersion of 0.2 dex in this correction
236: (see, e.g., Elvis et al. 1994), but omitting the scatter would not
237: significantly affect our results. U03 fit their data with a parameterized
238: model of luminosity-dependent density evolution. We adopt the
239: model and adjust its parameters in some redshift ranges to better fit
240: other data sets, especially at high redshifts and at bright
241: luminosities at low redshifts.
242: We note that many other studies have reported
243: general trends of the LF evolution and obscuring columns similar to
244: those of U03 (e.g., La Franca et al. 2005), and that infrared surveys also
245: suggest a substantial population of obscured AGNs at low
246: luminosities (Treister et al. 2006; see also Ballantyne and Papovich
247: 2007).
248: 
249: The number of sources per unit volume per dex of luminosity $\log
250: L_X=\log L_{2-10\, {\rm keV}}$ is fitted by U03 using a double
251: power-law multiplied by an ``evolution'' term:
252: \begin{equation}
253: \Phi(L_X,z)=e(z,L_X)\frac{A}{\left(\frac{L_X}{L^*}\right)^{\gamma_1}+\left(\frac{L_X}{L^*}\right)^{\gamma_2}}\,
254: ,
255:     \label{eq|U03}
256: \end{equation}
257: where
258: \begin{equation}
259: e(L_X,z)=\left\{
260:   \begin{array}{ll}
261:     (1+z)^{p_1}                      & \hbox{if $z<z_c(L_X)$} \\
262:     (1+z_c)^{p_1}[(1+z)/(1+z_c(L_X))]^{p_2} & \hbox{if $z\ge z_c(L_X)$\, ,}
263:   \end{array}
264: \right.
265:     \label{eq|eLz}
266: \end{equation}
267: with
268: \begin{equation}
269: z_c(L_X)=\left\{
270:   \begin{array}{ll}
271:     z_c^*                      & \hbox{if $L_X\ge L_a$} \\
272:     z_c^*(L_X/L_a)^{0.335}    & \hbox{if $L_X< L_a$\, .}
273:   \end{array}
274: \right.
275:     \label{eq|zc}
276: \end{equation}
277: We tune the value of parameters in order to provide a good fit to
278: the overall set of data presented in Figure~\ref{fig|BolLF}. The
279: full list of parameters is given in Table~\ref{Table|LF}. Some of
280: the parameters assume different values in different redshift
281: intervals, in which case we apply a linear interpolation in $z$ between
282: these intervals, as reported in
283: Table~\ref{Table|LF}. %The parameter $p_1$ decreases from $z=4.3$ to
284: %$z=3$ linearly with redshift, i.e. $p_1(z)=m\times z+q$, where $m$
285: %and $q$ are determined by the points $(z=4,p_1=4.23)\,,
286: %(z=6,p_1=3)$. The bright-end slope $\gamma_2$ varies in the same
287: %manner, $\gamma_2(z)=m\times z+q$, between discontinuous decreasing
288: %redshift bins $(z=6,\gamma_2=2.8)\, (z=4,\gamma_2=2.32)$ and
289: %$(z=0.8,\gamma_2=2.32)\, (z=0.5,\gamma_2=2.6)$. Same for $\log L_a$,
290: %which varies linearly ($\log L_a=m\times z+q$) between $(z=6,\log
291: %L_a=44.8)\, (z=3,\log L_a=44.6)$.
292: The X-ray LF of equation~(\ref{eq|U03}) is converted into a bolometric
293: LF using the fit to the bolometric correction given by Marconi et
294: al. (2004), $\log L/L_X=1.54+0.24\zeta+0.012\zeta^2-0.0015\zeta^3$
295: with $\zeta=\log L/L_{\odot}-12$, and $L_{\odot}=4\times 10^{33}\
296: {\rm erg\, s^{-1}}$, then convolved with a Gaussian scatter of
297: dispersion $0.2\, {\rm dex}$ to account for dispersion in the bolometric
298: correction.
299: We find the bright-end slope $\gamma_2$ should vary with time
300: to match the data, steepening at
301: very low redshifts ($z\le 0.8$) and at high ($z\ge 4$) redshifts.
302: We caution that the ``evolution'' term $e(L_X,z)$ substantially modifies
303: the shape of the luminosity function below $L\approx 10^{45}\, {\rm erg\, s^{-1}}$.
304: For example, the effective LF slope in the range $10^{43}\, {\rm erg\, s^{-1}}\le L\le 10^{44.5}\, {\rm erg\, s^{-1}}$
305: changes from $-1.4$ at $z\sim 2$ to $-0.7$
306: at $z\sim0$ even though we keep $\gamma_1$
307: (appearing in equation~[\ref{eq|U03}]) fixed at 0.86.
308: 
309: Figure~\ref{fig|BolLF} compares our model of the AGN LF to a large
310: collection of data from optical surveys (Pei 1995; Wisotzki 1999;
311: Fan et al. 2001, 2004; Kennefick et al. 1994; Hunt et al. 2004; Wolf
312: et al. 2003; Richards et al. 2005, 2006; Jiang et al. 2006; Cool et
313: al. 2006; Bongiorno et al. 2007; Fontanot et al. 2007; Shankar \&
314: Mathur 2007) and X-ray surveys (Barger et al. 2003; Ueda et al.
315: 2003; Barger \& Cowie 2005; Barger et al. 2005; La Franca et al.
316: 2005; Nandra et al. 2005; Silverman et al. 2008). Optical data have
317: been converted from the $B$-band into bolometric quantities using
318: $L_{\nu_B}\nu_B C_B=L$, where $L_{\nu_B}$ is the monochromatic
319: luminosity at $\nu_B=6.8\times 10^{14}\, {\rm Hz}$ ($\sim 4400\,$
320: {\AA}). We use an average value of $C_B=10.4$ consistent with
321: Richards et al. (2006). Note that Marconi et al. (2004) and Hopkins
322: et al. (2007; HRH07 hereafter) suggest values 30\%-50\% lower
323: depending on luminosity, while Elvis et al. (1994) proposed a higher
324: value of 11.8.
325: %The LF is plotted in
326: %each panel at the median of the redshift bin of the data.
327: In Figure~\ref{fig|BolLF} we plot $L\Phi(L)$, instead of
328: simply $\Phi(L)$ itself, to highlight the luminosity bins where most of the
329: energy is emitted. For the same reason, we will later plot the black hole
330: mass function as $M_{\bullet}\Phi(M_{\bullet})$ to highlight the
331: mass bins in which most of the density resides.
332: 
333: The dot-dashed curves in Figure~\ref{fig|BolLF} show our model LF
334: including only sources with $\log N_H/{\rm cm^{-2}}\le 22$, which we
335: calculate using the luminosity-dependent column density distribution
336: functions of U03. The dashed curves show the contribution of all
337: sources with $\log N_H/{\rm cm^{-2}}\le 24$, while the solid curves
338: include higher column density (Compton-thick) systems. We assume that the number of Compton-thick
339: sources in each luminosity bin is equal to the number of sources in the column density
340: range $23\le \log N_H/{\rm cm^{-2}}\le 24$. We will
341: henceforth follow the X-ray convention of describing systems with
342: $\log N_H/{\rm cm^{-2}} < 22$ as ``Type I'' AGN and systems with
343: $22\le \log N_H/{\rm cm^{-2}} \le 24$ as Type II AGN. We roughly
344: expect the ``Type I'' curve (dot-dashed) to agree with optical survey
345: data points and the ``Type I+Type II'' curve to agree with X-ray
346: survey data points. The agreement between the model and the data is
347: overall fairly good, though in some regimes different data sets give
348: seemingly incompatible results, making it difficult to judge the
349: validity of the model itself. %In our model, the full LF is a factor
350: %about $2-3$ higher than the Type I LF at high luminosities (above
351: %the break luminosity); it is difficult to say whether the data
352: %support a gap of this magnitude between the optical and total LFs.
353: 
354: HRH07 have used their observationally constrained theoretical models to estimate
355: the fraction of AGN that are ``missing'' from observational samples because of obscuration, as a function
356: of waveband and luminosity. In Figure~\ref{fig|BolLFcorrected}, we plot data points corrected
357: for these obscured fractions, with the same model curves as Figure~\ref{fig|BolLF}.
358: If the data, the obscuration corrections,
359: and our model were all perfect, then all of the data points should line up with each other and with the solid
360: model curves.
361: The obscuration corrections do reduce the discrepancies among data sets, most notably in the
362: range $z\approx 0.7-3.2$, but they do not remove all of the differences. The upper envelope of the data points
363: is generally close to our model near the peak of $L\Phi(L)$, though at the highest luminosities the model
364: exceeds the data (principally SDSS) for $z\approx 0.7-2.0$.
365: 
366: In Figure~\ref{fig|XRBGs} we show the integrated intensity obtained
367: from our model AGN LF compared with all the available data on the
368: cosmic X-ray Background (XRBG) for energies above 1 keV. The data
369: are a collection of old and new results presented by Frontera et al.
370: (2007), plus the recent results from \emph{INTEGRAL} (Churazov et
371: al. 2006). While the shape of the XRBG is now well established at
372: low energies (2-10 keV band), different studies imply normalizations
373: that differ by up to 40\%. The minimum estimate for the
374: normalization was found by the very first \emph{HEAO} experiments
375: (e.g., Marshall et al. 1980; filled circles in
376: Figure~\ref{fig|XRBGs}), while the much more recent \emph{Chandra}
377: and \emph{XMM} results point towards higher values (shaded bands in
378: Figure~\ref{fig|XRBGs}; see Comastri 2004 for a review). Several
379: groups have assumed that the \emph{HEAO} measurements were correct
380: in shape but underestimated in normalization because of flux
381: calibration and/or instrumental background subtraction (Frontera et
382: al. 2007), implying a true XRBG spectrum that is $\sim 1.4$ times
383: the \emph{HEAO} spectrum at all energies. However, recent
384: measurements from \emph{INTEGRAL} (Churazov et al. 2006; open
385: circles in Figure~\ref{fig|XRBGs}) and from the \emph{PDS}
386: instrument aboard \emph{BeppoSAX} (Frontera et al. 2007) produce
387: results close to the original \emph{HEAO} measurements in the $\sim$
388: 20-100 keV range. Based on the agreement of multiple experiments in
389: overlapping energy ranges, we tentatively conclude that the true
390: XRBG spectrum approximately follows the \emph{INTEGRAL} points over
391: the range $\sim$ 5-100 keV.
392: 
393: The solid curve in Figure~\ref{fig|XRBGs} shows the XRBG predicted
394: by our AGN bolometric LF model with the U03 column density
395: distribution. Following U03, we use the PEXRAV code (Magdziarz \&
396: Zdziarski 1995) to generate spectra of families of AGN with
397: different column densities, also taking into account Compton
398: down-scattering for highly obscured sources (Wilman \& Fabian
399: 1999). We compute the spectra of Compton-thick ($\log N_H/{\rm cm^{-2}}>
400: 24$) sources assuming $\log N_H/{\rm cm^{-2}}=24.5$. Dashed, dot-dashed, and triple-dot-dashed curves show the
401: cumulative contributions of sources with $\log N_H/{\rm cm^{-2}}<22,
402: 23$, and 24, respectively. The Compton-thick sources increase the predicted XRBG by a factor of
403: about 1.3 at $E\gtrsim 20$ keV, with smaller contributions at lower
404: energies and at much higher energies. Our full model would be a good
405: match to the \emph{HEAO} spectrum renormalized by a constant factor
406: of $\sim 1.4$ as proposed in some earlier studies. However, it significantly overpredicts the
407: \emph{INTEGRAL} and \emph{SAX/PDS} results at $E> 10$ keV; these are
408: well matched by the contribution of $\log N_H/{\rm cm^{-2}}\le 24$
409: sources alone, with \emph{no} Compton-thick contribution. As noted
410: above, we employ U03's prescription for the luminosity dependence of
411: the Compton-thick fraction. If we use a simpler model in which
412: the number of Compton-thick sources is $40\%$ of the number of $\log
413: N_H/{\rm cm^{-2}}\le 24$ sources at all luminosities, then we obtain
414: very similar XRBG predictions (higher by a few percent near the peak
415: and even closer at other energies).
416: 
417: We conclude from
418: Figure~\ref{fig|XRBGs} that the Compton-thick fraction in our
419: standard model is probably an upper limit on the true fraction, and
420: in our calculations below we will also consider a model in which the
421: Compton-thick fraction is zero. However, Gilli et al. (2007) have
422: proposed a successful model to fit the $INTEGRAL$ measurements of the XRBG
423: that uses a combined contribution from
424: unobscured sources (which they define to be only those with $\log N_H/{\rm cm^{-2}}\le 21$)
425: a factor $\sim$ 2 below
426: our estimate and a higher contribution from obscured sources up to $\log
427: N_H/{\rm cm^{-2}}=26$, which they derive by assuming that these sources have a Gaussian distribution
428: of spectral indices around a mean slope $\langle \Gamma \rangle=1.9$.
429: With this result in mind, we will also explore models with a higher fraction of Compton-thick
430: sources (but our standard estimate of unobscured sources).
431: 
432: Reproducing the high background normalization
433: found by \emph{XMM} in the 2-5 keV range would require a
434: substantially higher contribution from Type I AGN, since obscured
435: AGN have little emission at these energies. Such an increase seems
436: implausible on other grounds. Following Schirber \& Bullock (2003)
437: and Miralda-Escud\'{e} (2003), we have estimated the average
438: hydrogen ionization rate produced by Type I X-ray AGN with optical
439: luminosity $M_B\le -15$ assuming UV spectral energy index $\alpha_{\rm
440: UV}=1.57$. Using the redshift-dependent X-ray-to-optical ratio by
441: Steffen et al. (2006), we find that the Type I AGNs produce an
442: emissivity that saturates the estimated UV background intensity at
443: $z\lesssim 1$. We therefore conclude
444: that the true AGN contribution to the XRBG is probably closer to our model prediction than to the $XMM$
445: band shown in Figure~\ref{fig|XRBGs}. If future missions confirm a high normalization of the XRBG in the 2-5 keV range,
446: it would probably mean that a substantial contribution from X-ray binaries in normal galaxies is present at these
447: energies. On the other hand, we note from Figure~\ref{fig|BolLF} that our
448: AGN LF defined for sources with $\log N_H/{\rm cm^{-2}}\le 22$
449: is consistent with, or even lower than, all optical survey estimates,
450: so we do not expect a much \emph{lower} contribution from these sources.
451: 
452: HRH07 have recently undertaken an exercise similar to ours,
453: attempting to construct a bolometric AGN LF that is consistent with
454: a wide range of data over a large span of energies and wavelengths.
455: Figure~\ref{fig|HopkinsLF} compares our model to theirs. We find
456: good agreement in the shape and overall normalization of the AGN LF
457: in the redshift range $z < 1$ and $3.5 \le z \le 4.5$, while at
458: intermediate redshifts our model LF is systematically lower. The
459: normalization difference arises because HRH07 adopt an X-ray
460: bolometric correction that is about $30\%$ higher (at typical
461: luminosities) than the Marconi et al. (2004) bolometric correction
462: used here. If we adopt the HRH07 bolometric correction, then we find
463: good agreement in the range $1\le z \le 3$, but our model LF is
464: higher at high and low redshifts. At $z \ge 4$ we adopt a steeper
465: bright-end LF slope based on the results of GOODs (Fontanot et al.
466: 2006), Cool et al. (2006) and Shankar \& Mathur (2007), who find
467: evidence for a steeper slope relative to the estimates of Richards
468: et al. (2006). However, at $1 \le z \le 2.5$ the HRH07 bright-end
469: slope is steeper than ours, since they effectively require a match
470: to the SDSS-based measurements of Richards et al. (2004) in this
471: redshift range, while we retain the shallower slope adopted by U03,
472: also supported by recent X-ray LF determination of Silverman et al.
473: (2008). Recent Spitzer studies (Hickox et al. 2007; Polletta et al.
474: 2008) suggest a significant fraction of obscured AGN even at high
475: luminosities, leaving some room for a slope difference between the
476: optical and bolometric LFs. We regard the differences between our
477: model LF and that of HRH07, derived from independent attempts to
478: match the full range of current data, as a reasonable representation
479: of the remaining systematic uncertainties in determination of the
480: bolometric AGN luminosity function. We will show in
481: \S~\ref{subsec|models} that the difference between these two
482: determinations does not alter our overall conclusions, but the
483: differences in normalization and shape at intermediate redshifts do
484: lead to different preferred accretion parameters for the black hole
485: population.
486: 
487: \section{The local black hole mass function and the integrated Soltan argument}
488: \label{sec|LMFSOLTAN}
489: 
490: \subsection{The local black hole mass function}
491: \label{subsec|LMF}
492: 
493: Marconi et al. (2004) and S04 found similar results for the local
494: black hole mass function, using somewhat different methods. Both
495: groups found that starting from the relation between black hole mass
496: and bulge velocity dispersion ($M_{\bullet}-\sigma$) or the relation
497: between black hole mass and bulge luminosity ($M_{\bullet}-L_{\rm
498: sph}$) led to similar local mass functions. However, a more recent
499: analysis by Lauer et al. (2007a) and Tundo et al. (2007) argues that
500: these two methods do not provide the same answer. We therefore
501: revisit some of the uncertainties in computing the local mass
502: function.
503: 
504: Figure~\ref{fig|LocalMFs} compares the local mass functions obtained
505: using different relations between black hole mass and host galaxy
506: properties. Following S04, we start with the galaxy LF in the $r^*$
507: band by Nakamura et al. (2002), who used light concentration
508: parameters to distinguish early and late type galaxies and derived
509: LFs for both. After correcting the Petrosian magnitudes to total magnitudes
510: by adding $-0.2$ mag, we convert the LF into a LF of spheroids, which we
511: compute as follows. First we compute the numerical fraction of
512: Elliptical-S0 (spirals Sab-Sbc-Scd) in the early-type (late-type)
513: galaxy LF, then correct each morphological galaxy type for its
514: respective spheroidal luminosity component, as given in Table 1 of
515: Fukugita et al. (1998) for the $r$-band (see also Yu \& Tremaine
516: 2002 and Marconi et al. 2004). Note that this method is
517: equivalent to correcting the luminosities themselves adopting an average
518: weighted fraction of $f_{\rm sph}=0.83-0.85$ for early-type galaxies, and $f_{\rm
519: sph}=0.27-0.3$ for late-type galaxies, as done in S04 (and references therein).
520: 
521: The solid line in Figure~\ref{fig|LocalMFs} shows our result
522: using the $M_{\bullet}-L_{\rm sph}$ relation calibrated by McLure \&
523: Dunlop (2002; their equation 6) for a sample of 18 black holes in
524: inactive galaxies, where we correct the magnitudes for our
525: cosmology. Using the calibration in equation (A9) of Tundo et al. (2006)
526: would give a similar result. These calculations yield
527: a black hole mass function $\sim$ 34\%
528: higher than that in S04 (shown with solid squares).
529: We include a Gaussian scatter of 0.3 dex around the mean
530: $M_{\bullet}-L_{\rm sph}$ relation in both cases. The intrinsic scatter to insert in the local
531: relations between black hole mass and host galaxy properties is
532: empirically uncertain because it is similar in magnitude to the
533: observational uncertainties themselves. However, most authors
534: estimate a scatter of about 0.3 dex, which is close to the
535: value predicted in the numerical simulations of Hopkins et al.
536: (2007b). In the following we will always adopt this value of the
537: scatter for all our computations of the local mass function.
538: 
539: If we instead use the galaxy velocity dispersion function and the
540: $M_{\bullet}-\sigma$ relation in equation (A5) of Tundo et al. (2006),
541: with a scatter of 0.3 dex, we get the short-dashed line in
542: Figure~\ref{fig|LocalMFs}, close to the central estimate of S04.
543: Here we use the velocity dispersion function by Sheth et al. (2003),
544: which includes their estimate for the contribution of the
545: bulges in spirals, which in turn accounts for $\sim 25\%$ of the total
546: estimated black hole mass density. The Sheth et
547: al. (2003) estimate of the velocity dispersion function is in very
548: good agreement with the velocity function estimated by S04 through
549: the bivariate relation of galaxy luminosity and velocity dispersion.
550: However, using the $M_{\bullet}-\sigma$ relation in Marconi et al.
551: (2004) yields the dot-dashed line, which is higher than the S04 determination
552: by about $1-\sigma$. The $M_{\bullet}-\sigma$ relation given in
553: Ferrarese \& Ford (2005) is steeper than the Tundo et al. (2007)
554: relation, and adopting it yields a similar mass density to S04, but
555: a local mass function shifted towards higher masses, shown with a
556: triple-dot dashed line in Figure~\ref{fig|LocalMFs} (see also Wyithe
557: 2004). Both of these estimates yield a local mass function
558: below the $M_{\bullet}-L_{\rm sph}$-based local mass function
559: estimate at $M_{\bullet}<10^{7.5}\, M_{\odot}$, as also noted
560: by S04.
561: 
562: %The $M_{\bullet}-L_{\rm sph}$ relation therefore appears to yield a higher estimate of the
563: %black hole MF than the $M_{\bullet}-\sigma$ relation
564: In the following we will adopt the grey band of Figure~\ref{fig|LocalMFs}, which spans the range of these estimates,
565: as representative of the mean and the systematic uncertainties of present estimates of the local mass function.
566: The integrated mass density of the local black hole population is $\rho_{\bullet}=(3.2-5.4)\times 10^5\, {\rm M_{\odot}\, Mpc^{-3}}$ (for $h=0.7$).
567: Figure~\ref{fig|LocalMFs} also presents three additional estimates of the local mass function. Stars show the results of combining
568: the Bell et al. (2003) galaxy baryonic mass function (corrected for spheroid fraction as above) with the
569: H\"{a}ring \& Rix (2004) estimate of the relation between black hole mass and spheroid stellar mass, again assuming
570: 0.3 dex intrinsic scatter. This result is in reasonable agreement with the $M_{\bullet}-\sigma$ estimates, though uncertainties
571: in the stellar mass-to-light ratio are a remaining source of systematic uncertainty. The dotted curve shows the estimate
572: of Hopkins et al. (2007b), based on a ``fundamental plane'' relation between black hole mass, galaxy velocity dispersion, and spheroid half-light
573: radius; it is more sharply peaked than the grey band and implies a higher integrated black hole mass density. Open circles
574: show Graham et al.'s (2007) estimate based on the correlation between black hole mass and the S\'{e}rsic light-profile index
575: of the galaxy spheroid. This estimate is similar to Hopkins et al.'s (2007b) near the peak, but it implies an even sharper fall off towards
576: low black hole masses. Further discussion of the discrepancies in local mass function estimates, and possible routes
577: to alleviate them, will appear in Shankar \& Ferrarese (in preparation).
578: 
579: \subsection{The integrated mass density}
580: \label{subsec|massdensity}
581: 
582: Before turning to the evolution of the differential black hole mass
583: function, the central theme of this paper, we briefly revisit the
584: classic So{\l}tan (1982) argument, which relates the integrated
585: black hole density to the integrated emissivity of the AGN
586: population. If the average efficiency of converting accreted mass
587: into bolometric luminosity is $\epsilon \equiv L/\dot{M}_{\rm
588: inflow}c^2$, where $\dot{M}_{\rm inflow}$ is the mass accretion
589: rate, then the actual accretion onto the central black hole is
590: $\dot{M}_{\bullet}=(1-\epsilon)\dot{M}_{\rm inflow}$, where the factor $1-\epsilon$ accounts for the fraction of
591: the incoming mass that is radiated away instead of being added to
592: the black hole. The rate at which mass is added to the black hole mass function
593: is then given by
594: \begin{equation}
595: \frac{d\rho_{\bullet}}{dt}=\frac{1-\epsilon}{\epsilon
596: c^2}\int_0^{\infty}\Phi(L)Ld\log L\, .
597:     \label{eq|soltan}
598: \end{equation}
599: The mass growth rate implied by equation~(\ref{eq|soltan}) and our
600: estimate of the AGN LF from \S 2 is shown by the solid line in
601: Figure~\ref{fig|SFR-Rhoz}a. We set the radiative efficiency to a
602: value of $\epsilon=0.075$, as it provides a cumulative mass density
603: in agreement with the median estimate of the local mass density discussed in
604: \S~\ref{subsec|LMF}. At each time
605: step we integrate equation~(\ref{eq|soltan}) down to the observed
606: faint-end cut in the $2-10$ keV AGN LF, which we parameterize as
607: \begin{equation}
608: \log L_{\rm MIN, 2-10\, {\rm keV}}(z)=\log L_{0,\, 2-10\, {\rm
609: keV}}+2.5\log(1+z)\, .
610:     \label{eq|Lbolz}
611: \end{equation}
612: We set $\log L_{2-10\, \rm{keV}}/({\rm erg\, s^{-1}})=41.5$, in agreement with
613: the faintest low redshift objects observed by U03 and
614: La Franca et al. (2005). For a typical $L_{\rm opt}/L_X$, equation~(\ref{eq|Lbolz})
615: yields an optical luminosity of $M_B\sim -22$ at $z\sim 6$,
616: comparable to the faintest AGN sources observed by Barger et al.
617: (2003) in the 2 Msec Chandra Deep Field North (see also
618: Figure~\ref{fig|BolLF} and Shankar \& Mathur 2007). At each time
619: step we compute the minimum observed luminosity given in
620: equation~(\ref{eq|Lbolz}) and convert it into a bolometric quantity
621: $L_{\rm MIN}(z)$ applying the adopted bolometric correction by Marconi et al. (2004).
622: 
623: Dashed and dot-dashed lines in Figure~\ref{fig|SFR-Rhoz}a show two
624: recent estimates of the cosmic star formation rate (SFR) as a
625: function of redshift, from Hopkins \& Beacom (2006), reported with
626: its 3-$\sigma$ uncertainty region (dark area), and Fardal et al.
627: (2007). We have multiplied both estimates by a redshift-independent factor of
628: $8\times 10^{-4}$. Since local
629: estimates imply a typical ratio $M_{\bullet}/M_{\rm star}\sim
630: 1.6\times 10^{-3}$ for spheroids (e.g., Har\`{\i}ng \& Rix 2004), this is a
631: reasonable scaling factor if roughly $50\%$ of star formation goes
632: into spheroidal components (see also Marconi et al. 2004 and Merloni
633: et al. 2004). The agreement between the inferred histories of black
634: hole growth and star formation suggests that the two processes are
635: intimately linked. In particular this association seems to hold down to the last several Gyrs,
636: even at $z\lesssim 1.5$ when
637: disk galaxies are expected to dominate the SFR. A possible link
638: between black hole growth and star formation in disks could arise from re-activations
639: induced by tidal interactions between satellite
640: and central galaxies (e.g., Vittorini et al. 2005). Also, bars could possibly funnel
641: gas into the central black hole, though empirical studies cast
642: some doubt on this mechanism as a primary trigger for black hole growth (Peeples
643: \& Martini 2006 and references therein).
644: 
645: Figure~\ref{fig|SFR-Rhoz}b presents the same comparison in
646: integrated form (see also De Zotti et al. 2006 and Hopkins et al.
647: 2006b). Solid squares show the black hole mass density
648: obtained by converting the $z=1$ and $z=2$ galaxy stellar mass
649: function into a black hole mass function by assuming a ratio
650: $M_{\bullet}/M_{\rm star}$ equal to the local one (i.e., $1.6\times 10^{-3}$). The galaxy
651: stellar mass function has been computed from the Caputi et al.
652: (2006) $K$-band galaxy luminosity function, assuming an average
653: mass-to-light ratio $M_{\rm star}/L_K$=$0.4$ at $z=1$ and
654: $M_{\rm star}/L_K$=$0.3$ at $z=2$. The latter values have been
655: obtained from the \emph{Pegase2} code (Fioc \& Rocca-Volmerange
656: 1997) by taking a short burst of star formation ($< 10^9$ yr) and a
657: Kennicutt double power-law stellar Initial Mass Function. The quoted
658: values for $M_{\rm star}/L_K$ can be taken as lower
659: limits, as other choices of the parameters in the code would tend to
660: increase their value. However, we also note that our result on the
661: stellar mass function is in good agreement with the recent
662: estimate by Fontana et al. (2006). Our scaling factor of $1.6\times 10^{-3}$ implicitly assumes
663: that all the stellar mass in the luminous galaxies probed by these high-redshift observations resides in
664: spheroidal components today, and is therefore associated with black hole mass.
665: 
666: Figure~\ref{fig|SFR-Rhoz} suggests that the ratio of black hole
667: growth to SFR is approximately the same at all redshifts, and
668: suggests a close link between black hole accretion and star
669: formation. If the average ratio of black hole mass to stellar mass
670: were much higher than the local value at $z=1-2$, then the squares
671: in Figure~\ref{fig|SFR-Rhoz}b would shift above the black hole mass
672: density curve $\rho_{\bullet}(z)$. Increasing $M_{\rm star}/L_K$
673: would move the squares higher still, while associating only a
674: fraction of the high-redshift stellar mass with present day
675: spheroids would move them lower. Our conclusions are in marginal
676: agreement with those of McLure et al. (2006) and Shields et al.
677: (2006) and in some disagreement with those of Peng et al. (2006),
678: who find $M_{\bullet}/M_{\rm star}$ in gravitationally lensed quasar
679: hosts at $z\sim 1-2$ a factor $\sim 3$ above the local value.
680: However, several observational biases may effect studies of the
681: $M_{\bullet}/M_{\rm star}$ ratio in high redshift quasar samples
682: (Lauer et al. 2007b). Uncertainties in the local normalization of
683: $\rho_{\bullet}$ and in the stellar mass-to-light ratios still leave
684: a fair amount of wiggle room, but our results in
685: Figure~\ref{fig|SFR-Rhoz} disfavor models in which black holes
686: ``grow first'' and spheroid star formation ``catches up'' at later
687: times. Alternatively, if the average efficiency $\epsilon$ of black
688: hole accretion is lower at high redshifts, then the
689: $\rho_{\bullet}(z)$ curve in Figure~\ref{fig|SFR-Rhoz}b would shift
690: upwards at these redshifts accordingly.
691: 
692: %To match Peng et al. (2006)'s
693: %results the mass accreted at $z\sim 2$ should be much higher than
694: %the one inferred by equation~(\ref{eq|soltan}). This can be achieved
695: %or through significant "dark" accretion during the evolution, and/or
696: %starting at $z=6$ with a considerable black hole mass density
697: %already in place. The latter seems however implausible, as the
698: %initial black hole MF should have a space density of orders of
699: %magnitude higher than the observed AGN LF.
700: %
701: %and still be consistent with the accreted mass So{\l}tan argument,
702: %the $z=6$  in which the $z=6$ accreted mass function has a mass density of
703: %$\rho_{\bullet}\gtrsim 10^5\, M_{\odot}\, {\rm Mpc^{-3}}$, which
704: %implies a very low initial duty-cycle $P_0(z=6)\sim
705: %10^{-4}-10^{-3}$. However such low duty-cycles at those high
706: %redshifts seem implausible.
707: %
708: %The comparison between the black hole accretion rate with the integrated
709: %SFR and the stellar MF, implies that the ratio $M_{\bullet}/M_{\rm
710: %STAR}\lesssim 1.6\times 10^{-3}$ at $z>0$, in marginal agreement
711: %with recent results from Mclure \& Dunlop 2006 and Shields et al.
712: %2006. These results are
713: %quite at odds with the empirical observations by, e.g., Peng et al.
714: %(2006), which would instead imply a black hole mass density
715: %$\rho_{\bullet}\gtrsim 3\times 10^5\, M_{\odot}\, {\rm Mpc^{-3}}$,
716: %at $z\sim 2$, if their average $M_{\bullet}/M_{\rm STAR}\sim 3$ was
717: %used in Figure~\ref{fig|SFR-Rhoz} to convert the stellar MF into a
718: %black hole MF. The only model we can think of that can match Peng et al.'s
719: %results and still be consistent with the So{\l}tan's argument, is
720: %one in which the $z=6$ accreted mass function has a mass density of
721: %$\rho_{\bullet}\gtrsim 10^5\, M_{\odot}\, {\rm Mpc^{-3}}$, which
722: %implies a very low initial duty-cycle $P_0(z=6)\sim
723: %10^{-4}-10^{-3}$. However such low duty-cycles at those high
724: %redshifts seem implausible.
725: %The comparison between the black hole accretion rate with the integrated
726: %SFR and the stellar MF, implies that the ratio $M_{\bullet}/M_{\rm
727: %STAR}\lesssim 1.6\times 10^{-3}$ at $z>0$, in marginal agreement
728: %with recent results from Mclure \& Dunlop 2006 and Shields et al.
729: %2006.
730: %
731: %To further investigate this issue, in the lower panel of
732: %Figure~\ref{fig|SFR-Rhoz} we have assumed that a decreasing fraction
733: %$C(z)$ of the SFR with redshift is actually associated with black hole
734: %growth. In particular we set this fraction to be 80\% at $z>2$ is
735: %correlated to SFR, and it linearly decreases down to 20\% at $z=0$.
736: %To investigate any possible redshift dependence of this ratio with
737: %time, we then compute the cumulative remnant stellar mass associated
738: %with this particular model of the SFR, which we scale by the
739: %recycling factor and the ratio $M_{\bullet}/M_{\rm STAR}=1.6\times
740: %10^{-3}\times (1+z)^{\alpha}$. We find that $\alpha$ must be close
741: %to zero in order to match the accreted mass from the AGN LF. Also
742: %this result is quite independent of the actual fractions used in the
743: %computation, provided that the integration of the SFR gives by
744: %$z\sim 0$ enough stellar mass available in spheroids in the ratio
745: %$M_{\bullet}/M_{\rm STAR}\sim 1.6\times 10^{-3}$.
746: %
747: \section{Evolving the black hole mass function}
748: \label{sec|BHevol}
749: 
750: \subsection{Method}
751: \label{subsec|method}
752: 
753: Our goal is to calculate the evolution of the black hole mass
754: function implied by the bolometric AGN LF described in \S 2, given
755: assumptions about the radiative efficiency and the typical accretion
756: rate. In the following we will use the symbol $\Phi(x)$ to denote mass and
757: luminosity functions in logarithmic units of $L$ or $M_{\bullet}$,
758: i.e.
759: \begin{equation}
760: \Phi(x)=n(x)x \ln(10)\, ,
761:     \label{eq|Phi_x}
762: \end{equation}
763: where $n(x)$ is the comoving space density of black holes in the
764: mass or luminosity range $x\rightarrow x+dx$, in units of ${\rm
765: Mpc^{-3}}$ for $h=0.7$.
766: 
767: %%the shape and evolution of the bolometric AGN LF. In this section we
768: %compute the total accreted mass from the observed AGN LF once a
769: %radiative efficiency is assumed (So{\l}tan's argument). Small \&
770: %Blandford (1992), Salucci et al. (1999), Yu \& Tremaine (2002), Yu
771: %\& Lu (2004), Marconi et al. (2004) and S04, have
772: %considered a monotonic relation between AGN Luminosity $L$ and black hole
773: %mass $M_{\bullet}$, $L=\lambda L_{\rm Edd}$, where $L_{\rm Edd}$ is
774: %the Eddington luminosity.
775: %, being $t_E$ the
776: %Salpeter time and $\lambda=L/L_{\rm Edd}$
777: We define the Eddington accretion rate to be
778: \begin{equation}
779: \dot{M}_{\rm Edd}\equiv \frac{L_{\rm Edd}}{0.1\, c^2}\simeq
780: 22\left(\frac{M_{\bullet}}{10^9\, M_{\odot}}\right)\, M_{\odot}\,
781: {\rm yr^{-1}}
782:     \label{eq|Medd}
783: \end{equation}
784: and the dimensionless accretion rate
785: $\dot{m}=\dot{M_{\bullet}}/\dot{M}_{\rm Edd}$, where $L_{\rm Edd}$
786: is the standard Eddington luminosity (for Thomson scattering opacity and pure hydrogen
787: composition) at mass $M_{\bullet}$. The black hole growth rate $\dot{M}_{\bullet}$ is related
788: to the large scale accretion rate by
789: $\dot{M}_{\bullet}=(1-\epsilon)\dot{M}_{\rm inflow}$, where
790: $L=\epsilon \dot{M}_{\rm inflow}c^2$, because a fraction $\epsilon$
791: of the mass is radiated away before entering the black hole. We define
792: $f=\epsilon/(1-\epsilon)$, and $f_{0.1}=f/0.1$, so that a black hole
793: of mass $M_{\bullet}$ growing at a dimensionless rate $\dot{m}$ has
794: bolometric luminosity
795: \begin{equation}
796: L=\epsilon \dot{M}_{\rm inflow}c^2=0.1 f_{0.1}
797: \dot{M}_{\bullet}c^2=f_{0.1}\dot{m}lM_{\bullet}\, ,
798:     \label{eq|Lbol}
799: \end{equation}
800: where $l\equiv L_{\rm Edd}/M_{\bullet}=1.26\times 10^{38}\, {\rm
801: erg\, s^{-1}\, M_{\odot}^{-1}}$. Note that we define $\dot{M}_{\rm
802: Edd}$ for $f=0.1$, so the \emph{mass} Eddington ratio $\dot{m}$ is
803: linked to the \emph{luminosity} Eddington ratio $\lambda$ by the
804: radiative efficiency, i.e.
805: \begin{equation}
806: \lambda\equiv \frac{L}{L_{\rm Edd}}=\dot{m}f_{0.1}\, .
807:     \label{eq|Lambda}
808: \end{equation}
809: A black hole
810: accreting at a constant value of $\dot{m}$ grows exponentially in
811: time with a timescale $t_{\rm
812: growth}=M_{\bullet}/\dot{M}_{\bullet}=t_s/\dot{m}=4.5\times 10^7\,
813: \dot{m}^{-1}\, {\rm yr}$, where $t_s$ is equal to the Salpeter
814: (1964) timescale for $f_{0.1}=1$.
815: 
816: The evolution of the black hole mass function $n(M_{\bullet},t)$ is
817: governed by a continuity equation (Cavaliere et al. 1973,
818: Small \& Blandford 1992)
819: \begin{equation}
820: \frac{\partial n}{\partial t}(M_{\bullet},t)=-\frac{\partial
821: (M_{\bullet}\langle \dot{m}\rangle n(M_{\bullet},t))}{\partial
822: M_{\bullet}}\,,
823:     \label{eq|conteq}
824: \end{equation}
825: where $\langle \dot{m}\rangle$ is the mean dimensionless accretion
826: rate (averaged over the active and inactive populations) of the black holes of mass $M_{\bullet}$ at time $t$. This
827: evolution is equivalent to the case in which every black hole grows constantly at the mean accretion rate $\langle \dot{m}\rangle$. In
828: practice, individual black holes turn on and off, and there may be a
829: dispersion in $\dot{m}$ values, but the mass function evolution
830: depends only on the mean accretion rate as a function of mass.
831: 
832: All models in this paper assume a single accretion rate
833: $\dot{m}=\dot{m}_0$. At any given time, a black hole is either
834: accreting at $\dot{m}_0$ or not accreting. In some models we allow the
835: characteristic accretion rate $\dot{m}_0$ to depend on $z$, or on $M_{\bullet}$. The assumption of a
836: single $\dot{m}$ is clearly not valid for low luminosity AGNs in the
837: nearby universe, which have a wide range of Eddington ratios (e.g., Heckman et al. 2004, Greene \& Ho 2007).
838: However, Kollmeier et al. (2005) find that luminous AGN at
839: $0.5<z<3.5$ have a narrow range of Eddington ratios, with a peak at
840: $\lambda\sim 1/4$ and a dispersion of 0.3 dex. Since this dispersion
841: includes contributions from random errors in black hole mass
842: estimates and bolometric corrections, the true dispersion should be
843: even smaller. Netzer et al. (2007) find a similar result, with a slightly
844: larger dispersion, from a sample centered at $z\sim 2.5$. We will consider models with multiple $\dot{m}$ values
845: in future work, but single-$\dot{m}$ models (also adopted by, e.g.,
846: Marconi et al. 2004 and S04) are a good starting point for understanding black hole
847: growth.
848: 
849: If there is a single accretion rate $\dot{m}_0$, then the duty cycle of black hole
850: activity (i.e., the probability that a black hole of mass
851: $M_{\bullet}$ is active at a particular time) is given by the ratio
852: of luminosity and mass functions,
853: \begin{equation}
854: P_0(M_{\bullet},z)=\frac{\Phi(L,z)\left|\frac{d\log L}{d\log
855: M_{\bullet}}\right|}{\Phi(M_{\bullet},z)}\le 1\, ,
856:     \label{eq|P0general}
857: \end{equation}
858: where $M_{\bullet}=L/(f_{0.1}\dot{m}_0 l)$ is the black hole mass
859: that corresponds to luminosity $L$. Models with constant
860: $\dot{m}_0$ and $\epsilon$ have $L\propto
861: M_{\bullet}$, making the Jacobian factor unity. A physically consistent
862: model must have $P_0(M_{\bullet},z)\le 1$; there must be enough
863: black holes to produce the observed luminous AGNs.
864: 
865: Our strategy is to start with an assumed black hole mass function
866: $n(M_{\bullet},z_i)$ at an initial redshift $z_i$, then track the
867: characteristic curves $M_{\bullet}(M_{\bullet, i},z)$ of
868: equation~(\ref{eq|conteq}) by direct integration
869: \begin{eqnarray}
870: M_{\bullet}(M_{\bullet, i},z)=\int_{z_i}^{z}\langle
871: \dot{m}_{\bullet}\rangle
872: M_{\bullet}(z')\frac{dt}{dz'}dz' \nonumber\\
873: =\int_{z_i}^{z}\dot{m}_0P_0(M_{\bullet},z')M_{\bullet}(z')\frac{dt}{dz'}dz'
874: \, .
875:     \label{eq|mdotav}
876: \end{eqnarray}
877: Here $P_0(M_{\bullet},z')$ is given by equation~(\ref{eq|P0general})
878: with the observed luminosity function, and the evolved
879: $n(M_{\bullet},z)$ is given by
880: \begin{equation}
881: n(M_{\bullet},z)=n(M_{\bullet,i},z_i)\frac{dM_{\bullet,i}}{dM_{\bullet}}\,
882: ,
883:     \label{eq|nmJac}
884: \end{equation}
885: where $M_{\bullet}$ is the black hole mass that corresponds to
886: initial mass $M_{\bullet,i}$. To put this calculation in more physical terms: we take advantage of the fact
887: that equation~(\ref{eq|conteq}) is equivalent to having every black
888: hole grow at the rate $\langle \dot{m}(M_{\bullet},z)\rangle$.
889: Starting with a set of logarithmically spaced initial values of
890: $M_{\bullet,i}$, we integrate the masses forward in time with a
891: mid-point scheme from redshift step $z_j$ to step
892: $z_{j+1}=z_j-\Delta z$
893: \begin{equation}
894: M_{\bullet,j+1}=M_{\bullet,j}+\dot{m}_0P_0(M_{\bullet,j+1/2},z_{j+1/2})M_{\bullet,j+1/2}\frac{dt}{dz}\Delta
895: z\, ,
896:     \label{eq|leapfrog}
897: \end{equation}
898: where the values at the mid-point $j+1/2$ are evaluated by
899: extrapolating black hole masses for a half step $\Delta z/2$ using
900: $dM_{\bullet}/dt$ at the beginning of the step and the duty cycle $P_0$ is computed
901: using the mid-step mass function and the luminosity function evaluated at
902: $z_{j+1/2}=z_j+\Delta z/2$.
903: We choose sufficiently small redshift steps $\Delta z$ such that the
904: total mass density added at each iteration matches the one obtained
905: by integration of the bolometric luminosity function
906: (equation~\ref{eq|soltan}), and we have confirmed that smaller $\Delta z$
907: yields essentially identical results. Since the number of black
908: holes is conserved, $n(M_{\bullet},z)\Delta
909: M_{\bullet}$=$n(M_{\bullet,i},z)\Delta M_{\bullet,i}$.
910: 
911: As an aside, we note that this approach is mathematically
912: equivalent to solving the continuity equation in the form
913: \begin{equation}
914: \frac{\partial n(M_{\bullet},t)}{\partial t}=-\frac{\dot{m}_0}{t_s
915: \ln(10)^2 M_{\bullet}}\frac{\partial \Phi(L,z)}{\partial \log L}\,
916:     \label{eq|POsingle}
917: \end{equation}
918: used by, e.g. Small \& Blandford (1992) and Marconi et al. (2004).
919: Equation~(\ref{eq|POsingle}) follows from equation~(\ref{eq|conteq})
920: with the assumption of a single $\dot{m}_0$ and $f_{0.1}$, and it is
921: also valid in the case of a redshift dependent $\dot{m}_0$. From
922: equation~(\ref{eq|POsingle}), it is explicitly clear that the
923: evolution of the black hole mass function depends only on the
924: initial conditions, the observed luminosity function, and the values of $\dot{m}_0$
925: and $f_{0.1}=L/(M_{\bullet}l \dot{m}_0)$, given
926: the assumptions made here. The So{\l}tan argument
927: (\S~\ref{subsec|massdensity}) relates the integrated black hole
928: mass density to the integrated emissivity of the AGN population. The
929: continuity equation yields the full evolution of the black hole mass
930: function in terms of the full luminosity function; in essence, it
931: applies the So{\l}tan argument as a function of mass. We have checked
932: that our method yields consistent results against direct integration
933: of the equation.
934: 
935: For our integrations, we generally start at $z_i=6$ and determine
936: $n(M_{\bullet,i},z_i)$ from equation~(\ref{eq|leapfrog}) with our
937: estimate of the luminosity function at $z=6$ and an assumed
938: $P_0=0.5$. This relatively high initial duty cycle implies we
939: start with nearly the minimal black hole mass function required to
940: reproduce the luminosity function. By $z\sim 3.5$ the
941: integration has essentially forgotten the initial value of $P_0$
942: unless we set it much lower (e.g., $P_0<0.01$), since the accreted mass is much larger
943: than the mass in the initial black hole population. The clustering of luminous
944: AGN at $z\sim 3-4$ implies duty cycles of at least several percent (Shen et al. 2007;
945: Shankar et al., in preparation). In our standard calculations, we set
946: $P_0$ to zero for masses $M_{\bullet, {\rm MIN}}(z)$ that would
947: produce AGN luminosities below the minimum observed luminosity at
948: that redshift (indicated by the cutoffs in Figure~\ref{fig|BolLF});
949: we do not extrapolate the luminosity function below these minimum
950: values. Since $L_{\rm MIN}(z)$ drops
951: towards lower $z$, we must have a supply of black holes in place to
952: provide AGNs of lower luminosity as these become visible. To ensure
953: this, we set our initial black hole mass function to
954: \begin{equation}
955: n(M_{\bullet,i},z_i)=8\times n(M_{\bullet,{\rm
956: MIN}}(z_i),z_i)\times\left(\frac{M_{\bullet}}{M_{\bullet,{\rm
957: MIN}}(z_i)}\right)^{-2.3}\, ,
958: %\Phi(M_{\bullet},z_i)= \left\{
959: %  \begin{array}{ll}
960: %    \Phi(L,z_i)\left|d\log L/d \log M_{\bullet}\right|/P_0(M_{\bullet},z_i) & \hbox{if $M_{\bullet} \ge M_{\bullet, \rm{MIN}}(z_i)$} \\
961: %    k\Phi(M_{\bullet, \rm{MIN}}(z_i),z_i)10^{[\alpha(\log M_{\bullet}-\log M_{\bullet, \rm{MIN}}(z_i))]} & \hbox{if $M_{\bullet} < M_{\bullet,\rm{MIN}}(z_i)$}\, ,
962:  %   \Phi(M_{\bullet},z_i)=\Phi(M_{\bullet}^{*},z_i)10^{[\gamma_2(\log M_{\bullet}-\log M_{\bullet}^{*})]} & \hbox{if $M_{\bullet} < M_{\bullet}^{*}$}\\
963: %  \end{array}
964: %\right.
965: \label{eq|InitialMF}
966: \end{equation}
967: i.e., below $M_{\bullet,\rm MIN}(z_i)$ we boost $n(M_{\bullet},z_i)$
968: by a factor of $8$ and add an $M_{\bullet}^{-2.3}$ rise (see
969: Figure~\ref{fig|bestFitModel}, below). This high initial abundance
970: of low mass black holes ensures that we are never forced to duty
971: cycles $P_0>1$. Once $M_{\bullet}$ significantly exceeds
972: $M_{\bullet,{\rm MIN}}(z)$, the black hole mass function is
973: dominated by growth rather than initial values, and
974: $n(M_{\bullet},z)$ is insensitive to the assumed
975: $n(M_{\bullet,i},z_i)$. We adopted this procedure so that our results would not depend
976: on the extrapolation of the luminosity function into regions where it is not observed.
977: In practice, we find that extrapolating the luminosity function as a power law down to very
978: faint luminosities yields similar results. We follow the $L_{\rm min}$ procedure for our standard models in \S\S~\ref{subsec|bestfit} and ~\ref{subsec|models}, and
979: use extrapolation of the LF in \S\S~\ref{subsec|etamdot}-\ref{subsec|merging} where it gives
980: more stable numerical solutions.
981: 
982: %The total average {\it visibility} time, i.e. the time spent above
983: %$L_{\rm MIN}(z)$, could be estimated as a function of relic mass as
984: %\begin{equation}
985: %\langle \tau_{\rm vis}\rangle (M_{\bullet}^0)=\int_{z_i}^0
986: %P_0(M_{\bullet}^0,z) \frac{dt}{dz}dz\, ,
987: %    \label{eq|tvis}
988: %\end{equation}
989: %where $M_{\bullet}^0$ is the relic mass at $z=0$ and $z_i$ is the
990: %redshift in which we set the initial condition. Previously, it was
991: %discussed that the integration of equation~(\ref{eq|tvis}) could
992: %provide significant results which could be compared by other
993: %empirical, independent studies, such as those by Heckman et al.
994: %(2004), who have shown that the local less massive black holes have very
995: %long lifetimes. However equation~(\ref{eq|tvis}) only provides lower
996: %limits, which are also strongly dependent on the time in which black holes
997: %start shining in the AGN LF, i.e. on $L_{\rm MIN}(z)$. We conclude
998: %that equation~(\ref{eq|tvis}) is not very constraining and it will
999: %not be further taken into account.
1000: %%%
1001: %Here we just point out that, in the single-accretion rate scenario,
1002: %a \emph{lower} limit (cfr. S04) to the visible timescale can be
1003: %settled by the direct comparison of the cumulative statistics of
1004: %AGNs and local black holes, which in turn converts into a lower limit on
1005: %the average duty cycle. Following Yu \& Tremaine (2002), Yu \& Lu
1006: %(2004), S04 we find
1007: %\begin{equation}
1008: %\tau_{\rm vis, MIN}(M_{\bullet}^0)=\frac{\int_{\bar{L}}^{\infty}
1009: %dL\int_{\infty}^0\frac{dt}{dz}dz\Phi(L,z)d\log
1010: %L}{\int_{\bar{M}_{\bullet}}^{\infty}dM_{\bullet}^0n(M_{\bullet}^0,z)}\leq
1011: %\tau_{\rm vis}(M_{\bullet}^0)\, .
1012: %    \label{eq|taulumFin}
1013: %\end{equation}
1014: %Equation~(\ref{eq|taulumFin}) provides values $\tau_{\rm vis,
1015: %MIN}(M_{\bullet}^0)\sim 5\times 10^6-5\times 10^7\, {\rm yr}$ for
1016: %$6\lesssim \log M_{\bullet}/M_{\odot}\lesssim 10$ which means
1017: %$\langle P_0\rangle \gtrsim 10^{-4}-10^{-3}$.
1018: %
1019: %Therefore a steep rise at the low mass end, as
1020: %expressed in equation~(\ref{eq|InitialMF}), provides a valuable tool
1021: %for ensuring continuity during the initial stages of BH evolution,
1022: %without the need of invoking for a strong extrapolation in the AGN
1023: %LF below $L_{\rm MIN}(z)$.
1024: %
1025: \subsection{The reference model}
1026: \label{subsec|bestfit}
1027: 
1028: Figure~\ref{fig|bestFitModel} shows the results of our calculations
1029: for a reference model with our standard estimate of the AGN bolometric
1030: luminosity function and accretion parameters $\dot{m}=0.60$ and $\epsilon=0.065$
1031: ($f_{0.1}\sim 0.7$), which yield good overall agreement with the
1032: average estimate of the local black hole mass function.
1033: Panel~\ref{fig|bestFitModel}a shows the evolution of the mass
1034: function starting from the initial condition at $z=6$ (solid
1035: triangles) and continuing through to $z=0$. Because of the
1036: luminosity-dependent density evolution in the observed luminosity
1037: function, the massive end of the black hole mass function builds up
1038: early, and the lower mass regime grows at later redshifts. For
1039: $M_{\bullet}>10^{8.5}\, {\rm M_{\odot}}$, the mass function is
1040: almost fully in place by $z=1$. Panel~\ref{fig|bestFitModel}b plots
1041: $M_{\bullet}\Phi(M_{\bullet})$, proportional to the fraction of
1042: black hole mass per logarithmic interval of $M_{\bullet}$, which
1043: allows better visual comparison to the observed local mass function
1044: and highlights the contribution of each black hole mass bin to
1045: the total mass density at each time. The accretion model agrees well
1046: with our estimate of the local mass function (grey band) up to $\sim 10^{9.3}\, {\rm
1047: M_{\odot}}$. At higher masses our model exceeds the observational
1048: estimate, but in this regime the estimate relies on extrapolation of
1049: the scaling relations, and it is sensitive to the assumed intrinsic scatter.
1050: In addition, the high-mass end of the predicted mass function is sensitive to the
1051: bright end of the LF at $z\lesssim 2$ (see
1052: \S~\ref{subsec|models}).
1053: The open circles with error bars
1054: show the estimate of the local mass function by Graham et al. (2007), which
1055: we cannot reproduce even approximately with constant $\dot{m}_0$.
1056: If this estimate is correct, then the low end of the
1057: luminosity function must be produced mainly by high mass black holes accreting at low $\dot{m}$, so that
1058: the predicted growth of low mass black holes is reduced (see \S~\ref{subsec|dmdtMbh}).
1059: 
1060: 
1061: Figure~\ref{fig|bestFitModel}c plots the duty cycle as a function of mass for different
1062: redshifts, as labeled. The duty cycle for $M_{\bullet}\sim 10^9\, {\rm M_{\odot}}$
1063: is $\sim 0.2$ at $z\sim 4-5$, falling to 0.03-0.08 at $z=2-3$, when quasar activity
1064: is at its peak, then dropping to 0.003 at $z=1$ and $\sim 10^{-4}$ at $z=0$. Below $z\sim 3$, the duty cycle
1065: rises towards low black hole masses. This ``downsizing'' evolution, in which high mass black holes
1066: complete their growth early but low mass black holes continue to grow at late times, is required
1067: by the observed LF evolution in any model with approximately constant $\dot{m}_0$. Above $10^9\, {\rm M_{\odot}}$ the duty
1068: cycle curves are generally flat, because the constant slope of the bright end of the LF drives the growth
1069: of a mass function with a parallel slope. Below $z=1$, the LF slope changes, and the duty cycle
1070: curves are no longer flat. By running several cases with varying initial conditions, we find that the duty cycles
1071: for $z\lesssim 3.5$ are insensitive to the assumed initial duty cycle
1072: (which determines the initial mass function) provided $0.01\lesssim P_0(z=6)\le 1$.
1073: 
1074: Several lines of independent observational evidence have set
1075: constraints on the duty cycle of AGNs at different redshifts.
1076: Porciani et al. (2004), Croom et al. (2005), Porciani and Norberg
1077: (2006), and da Angela et al. (2006) have analyzed large samples of
1078: optical AGNs in the Two Degree Field (2dF) QSO Redshift Survey,
1079: showing that their large-scale ($\gtrsim 1$ Mpc) clustering
1080: properties are consistent with quasars residing in very massive
1081: halos with $M_{\rm h}\sim 5\times 10^{12}-10^{13}\, {\rm
1082: M_{\odot}}$. Comparing the AGN abundance to the abundance of such halos
1083: implies duty cycles of order of $\sim (1-4)\%$ in the redshift range $z\sim 1-2$.
1084: Similar results have been confirmed by large-scale
1085: optical AGN clustering studies in the Sloan Digital Sky Survey
1086: (SDSS; e.g., Myers et al. 2007). Shen et al. (2007) have extended the SDSS
1087: results to $z=3-4$, obtaining similar results for halo masses and duty cycles.
1088: Similar or lower values for the duty cycle could be found directly by
1089: comparing the AGN number density to the galaxy luminosity function
1090: at comparable redshifts, the latter now well estimated and complete
1091: down to rather low luminosities (e.g., Pannella et al. 2006). Using \emph{Chandra} X-ray observations of the Hubble Field
1092: North, Nandra et al. (2002) and Steidel et al.
1093: (2002) find that about 3\% of a large spectroscopic survey of
1094: $\sim 1000$ Lyman Break Galaxies at the average redshift of $z\sim
1095: 3$ host an X-ray bright, moderately obscured AGN. Within the uncertainties of the
1096: comparison, these estimates of AGN duty cycles are in good agreement with the predictions
1097: in Figure~\ref{fig|bestFitModel}c. We will present predictions for AGN clustering bias in
1098: \S~\ref{sec|hosts} below, and we will carry out quantitative comparisons to observed clustering in future work.
1099: The occurrence of AGN in massive star-forming galaxies
1100: is much higher than these overall duty cycle estimates
1101: (Alexander et al. 2003; Borys et al. 2005), supporting the scenario in
1102: which star-formation in massive galaxies is closely related to black
1103: hole growth (e.g., Granato et al. 2006).
1104: %Such values are in
1105: %good agreement with those obtained in
1106: %Figure~(\ref{fig|bestFitModel})c, within a factor of $2-3$,
1107: %comparable to the systematics in the absolute duty-cycle values, due
1108: %to uncertainties in the high-redshift Compton-thick fraction (e.g.,
1109: %Treister \& Urry 2006) and the assumed radiative efficiency. Much
1110: %higher seems to be instead the chance of AGN appearance in massive
1111: %star-forming galaxies ,
1112: % performed X-ray
1113: %spectral analysis of bright SCUBA galaxies in the 2 Ms
1114: %\emph{Chandra} Deep Field North for submillimeter fluxes $f_{\rm
1115: %850\, \mu m}\geq 5$ mJy and a signal-to-noise $\geq 4$. They claim
1116: %that a high fraction $\geq 36\%$ of the submillimeter sources in
1117: %their sample hosts an X-ray AGN, most probably powered by black holes in
1118: %the growing phase.
1119: 
1120: Figure~\ref{fig|bestFitModel}d shows the black hole accretion
1121: histories as a function of relic mass and redshift, following
1122: equation~(\ref{eq|mdotav}). Tracing back one of these curves shows the mass that a
1123: ``typical'' black hole of present mass $M_{\bullet}$ had at earlier
1124: times, as predicted by our model. The dot-dashed line shows the minimum
1125: black hole mass associated with $L_{\rm MIN}(z)$ at each redshift.
1126: By construction, all growth curves are flat (dashed lines) before crossing
1127: $L_{\rm MIN}(z)$. Figure~\ref{fig|bestFitModel}d shows again that high mass black holes build their
1128: mass early and experience little growth at late times, while lower mass black holes grow rapidly at lower
1129: redshifts.
1130: 
1131: %In Table~\ref{Table|LF} we provide the \footnote{The full tables can be found
1132: %in electronic format at http://www.astronomy.ohio-state.edu/~shankar/Models.}
1133: %numerical values of our reference bolometric AGN LF at several bins of redshifts
1134: %from $z=0.02$ to $z\sim 6$ and, at each redshift,
1135: %interpolated it in the luminosity range $41\lesssim \log L \lesssim 48$. Also listed
1136: %in the fourth and fifth columns is the AGN LF with no Compton-thick and with Double
1137: %Compton-thick sources, respectively.
1138: %\begin{landscape}
1139: 
1140: %\end{landscape}
1141: %
1142: %In Table~\ref{Table|Fits} we instead provide
1143: %the numerical values of the outputs obtained using equation~(\ref{eq|POsingle})
1144: %for our reference model with $\epsilon=0.065$ and $\dot{m}_0=0.6$,
1145: %using the set of luminosity functions
1146: %listed in Table~\ref{Table|LF}. The outputs are presented
1147: %from $z=0.02$ to $z\sim 6$ and, at each redshift,
1148: %interpolated in the mass range $5\lesssim \log M_{\bullet}/{\rm M_{\odot}} \lesssim 9.6$.
1149: %
1150: %\begin{landscape}
1151: 
1152: %\end{landscape}
1153: %
1154: \subsection{Parameter variations}
1155: \label{subsec|models}
1156: 
1157: The top panels of Figure~\ref{fig|Models} show the effect of varying the two
1158: main model parameters, the
1159: radiative efficiency $\epsilon$ and accretion rate $\dot{m}_0$. Figure~\ref{fig|Models}a
1160: compares models with varying $\epsilon$ and accretion rate fixed to
1161: the reference model value $\dot{m}=0.60$. Lowering (raising) the radiative
1162: efficiency has the effect of shifting the accreted mass function up
1163: (down) and shifting the peak of the mass distribution to higher
1164: (lower) masses. Recall that the luminosity Eddington ratio is $\lambda=\dot{m}f_{0.1}$; if we
1165: held $\lambda$ fixed instead of $\dot{m}$, then curves in Figure~\ref{fig|Models}a would be
1166: vertically displaced with no horizontal shift.
1167: Figure~\ref{fig|Models}b shows models with varying $\dot{m}_0$,
1168: all with $\epsilon=0.065$. The So{\l}tan argument implies that $\rho_{\bullet}=\int M_{\bullet} \Phi(M_{\bullet})d\log M_{\bullet}$
1169: should be the same for models with the same $\epsilon$ that reproduce the same observed LF. The area
1170: under the curves in Figure~\ref{fig|Models}b is thus the same for all models, but the peak in the mass
1171: function shifts to lower $M_{\bullet}$ for higher $\dot{m}_0$ because a given observed luminosity
1172: is associated with a lower black hole mass. While
1173: there are still systematic uncertainties in the normalization and
1174: shape of the local mass function, the peak of the
1175: $M_{\bullet}\Phi(M_{\bullet})$ distribution (Figure~\ref{fig|LocalMFs})
1176: is in good agreement among all the
1177: estimates, and this, in turn, sets strong constraints on
1178: $\dot{m}_0$. Shifting the peak of the accreted mass distribution
1179: to $\log (M_{\bullet}/M_{\odot})\sim 8$ would require
1180: Super-Eddington accretion rates, while shifting it to $\log (M_{\bullet}/M_{\odot})\sim 9$
1181: requires $\dot{m}_0\sim 0.2-0.3$.
1182: Changes to $\epsilon$ and $\dot{m}_0$ shift the accreted
1183: mass function in different directions in the $M_{\bullet}\Phi(M_{\bullet})-M_{\bullet}$
1184: plane, but they do not alter its shape.
1185: 
1186: %For our estimate of the bolometric LF,
1187: %we find that a good match between the local and accreted mass
1188: %functions is obtained only for quite narrow ranges in
1189: %parameter values, i.e., $0.055\lesssim \epsilon\lesssim 0.07$ and
1190: %$0.6\lesssim \dot{m}_0\lesssim 0.8$. This result is also consistent
1191: %with previous studies (Marconi et al. [2004] and S04).
1192: 
1193: The lower panels of Figure~\ref{fig|Models} show the effect of changing the input LF. In panel~\ref{fig|Models}c,
1194: solid and dotted curves show models with no Compton-thick sources or a double fraction of Compton-thick sources, respectively,
1195: with radiative efficiency and accretion rate fixed at the reference model values $\epsilon=0.065$ and $\dot{m}_0=0.6$.
1196: The predicted black hole mass functions are similar in shape but offset in amplitude by about $\pm$ 20\%. These
1197: offsets are similar to our estimated uncertainty in the local mass function, shown by the grey band. We
1198: can restore agreement between the double Compton-thick model and the central estimate of the local mass function
1199: by raising $\epsilon$ to 0.08 (to produce more luminosity with the same black hole mass) and lowering
1200: $\dot{m}_0$ to 0.5 (to compensate the influence of higher $\epsilon$ on the peak location). Conversely,
1201: the no-Compton-thick model yields similar predictions if $\epsilon$ is lowered to 0.05 and $\dot{m}_0$
1202: is raised to 0.8. As noted in \S~\ref{sec|AGNLF}, we find better agreement with the observed
1203: X-ray background for no Compton-thick and worse agreement for double Compton-thick (see Figure~\ref{fig|XRBGs}).
1204: 
1205: Figure~\ref{fig|Models}d compares our reference model to the
1206: one obtained by integrating the AGN bolometric LF of HRH07.
1207: For the reference model values $\epsilon=0.065$ and $\dot{m}_0=0.6$,
1208: the predicted mass function shifts to higher normalization and higher peak mass (solid line).
1209: Principally this difference reflects the higher bolometric correction adopted by HRH07,
1210: which shifts their LF to higher normalization at $1\lesssim z \lesssim 3$ (Figure~\ref{fig|HopkinsLF}).
1211: The HRH07 estimate also has higher peak luminosity in this redshift range, leading to higher peak mass.
1212: Adopting a higher efficiency and accretion rate, $\epsilon=0.09$ and $\dot{m}_0=1$, largely
1213: compensates these two effects. As noted in \S~\ref{sec|AGNLF}, the HRH07 LF has a steeper bright-end slope
1214: at intermediate redshifts, and this leads to a steeper high-mass slope of the accreted mass function, improving
1215: agreement with observational estimates above $M_{\bullet}\sim 10^9\, {\rm M_{\odot}}$.
1216: 
1217: To present our model comparison in a more global way, we characterize the local mass function
1218: by four quantities that contain complementary information:
1219: \begin{itemize}
1220:   \item The integrated black hole mass density $\rho_{\bullet}=\int M_{\bullet}\Phi(M_{\bullet})d\log M_{\bullet}$.
1221:   \item The location $\log M_{\rm PEAK}$ at which the mass function peaks in the $M_{\bullet}\Phi(M_{\bullet})-M_{\bullet}$ plane.
1222:   \item The width of the mass function, characterized by $(\Delta \log M_{\bullet})_{1/2}$ such that twice
1223:   the integral from $\log M_{\rm PEAK}$ to $\log M_{\rm PEAK}+(\Delta \log M_{\bullet})_{1/2}$ contains half the total
1224:   mass density, $2\times \int_{\log M_{\rm PEAK}}^{\log
1225: \bar{M_{\bullet}}} \Phi(M_{\bullet})M_{\bullet}d\log
1226:   M_{\bullet}$$=1/2\times \rho_{\bullet}$. If $M_{\bullet}\Phi(M_{\bullet})$ were a Gaussian, $(\Delta \log M_{\bullet})_{1/2}$
1227:   would be $\sim 0.26$ times the full width at half maximum.
1228:   \item The asymmetry of the mass function, characterized by $\Delta
1229:   \rho_{\bullet}/\rho_{\bullet}$, the normalized difference in the mass density integrated above and below $\log M_{\rm PEAK}$.
1230: \end{itemize}
1231: 
1232: Figure~\ref{fig|multiparam} plots the predicted values of these four quantities
1233: as a function of the radiative efficiency $\epsilon$ for models with our standard
1234: AGN LF and accretion rates $\dot{m}_0=1, 0.6, 0.45, 0.3$. Grey horizontal bands
1235: show the observational estimates corresponding to the grey band in Figure~\ref{fig|LocalMFs}.
1236: Short-dashed and long-dashed horizontal lines show instead the parameters for the local mass functions inferred by
1237: Graham et al. (2007) and Hopkins et al. (2007b), respectively (see Figure~\ref{fig|LocalMFs}).
1238: As expected from the So{\l}tan argument, the predicted $\rho_{\bullet}$ is proportional
1239: to $(1-\epsilon)/\epsilon$ but independent of $\dot{m}_0$. Consistency with the
1240: grey band observational estimate requires $0.063\le \epsilon \le 0.1$.\footnote{Our reference model
1241: of \S~\ref{subsec|bestfit} has $\epsilon$ at the low end of this range, rather
1242: than the middle, because we chose parameters to match the mass function near its peak, where it is
1243: best determined. However, the reference model predicts $\rho_{\bullet}=5.24\times 10^5\, {\rm M_{\odot}\, Mpc^{-3}}$,
1244: above the middle of the grey band.} Reproducing the
1245: observed value of $\log M_{\rm PEAK}/{\rm M_{\odot}}\approx 8.5$ then requires $0.4\lesssim \dot{m}_0 \lesssim 0.8$,
1246: though for any given $\epsilon$ the value of $\dot{m}_0$ is determined to $\sim 10\%$.
1247: As noted in our discussion of Figure~\ref{fig|Models}a, the location of the mass function
1248: peak is determined by $\lambda=L/L_{\rm Edd}$ largely (but not completely) independent of $\epsilon$,
1249: and reproducing the observed $\log M_{\rm PEAK}$ implies $\lambda\approx 0.4-0.5$ over the range
1250: $0.05\lesssim \epsilon \lesssim 0.13$.
1251: 
1252: The higher $\rho_{\bullet}$ implied by the Hopkins et al. (2007b) local mass function requires a lower efficiency,
1253: $\epsilon\approx 0.06$ which in turn implies a higher $\dot{m}_0\approx 1$ to match $\log M_{\rm PEAK}$. (Note that this is
1254: \emph{not} the comparison in Figure~\ref{fig|Models}d, where we use the HRH07 \emph{luminosity} function
1255: but show our standard estimate of the local \emph{mass} function). Our inferred ranges for $\epsilon$ and
1256: $\dot{m}_0$ are consistent with the findings of Marconi et al. (2004) and S04, who obtained ($\epsilon$, $\lambda$)=(0.08, 0.5)
1257: and (0.09, 0.3), respectively.
1258: Note that S04 tends towards slightly higher values of the radiative efficiency mainly because they
1259: used an X-ray bolometric correction normalized to the optical bolometric correction of Elvis et al. (1994), i.e., $C_B$=11.8,
1260: while we have adopted the lower value suggested by recent work (see \S~\ref{sec|AGNLF}).
1261: 
1262: With $\epsilon$ and $\dot{m}_0$ fixed by matching $\rho_{\bullet}$ and $\log M_{\rm PEAK}$, there is no freedom
1263: in the model to adjust the predicted width and asymmetry of the mass function, but these prove remarkably
1264: insensitive to $\epsilon$ and $\dot{m}_0$ in any case. To substantially alter these predictions one must
1265: either change the input luminosity function or add a new physical ingredient to the model, such as
1266: mass- or redshift-dependent $\dot{m}_0$, mergers, or multiple accretion rates. For our minimal
1267: model and standard LF estimate, the predicted widths are larger than any of the observational
1268: estimates. This discrepancy is largely a consequence of the shallow high-mass slope of the predicted mass function
1269: (see Figure~\ref{fig|Models}), which in turn is sensitive to the bright-end slope of the luminosity function
1270: in the range $1\lesssim z \lesssim 3$, as discussed earlier. Note, however, that the observational estimates of
1271: $(\Delta \log M_{\bullet})_{1/2}$ depend on the extrapolation of the black hole-galaxy correlations above $M_{\bullet}\approx 10^9\, {\rm M_{\odot}}$,
1272: where they are quite uncertain, so the grey band in Figure~\ref{fig|multiparam} should not necessarily
1273: be taken as a true upper limit. The predicted asymmetries are in reasonable agreement with the observational estimates,
1274: except for the Graham et al. (2007) mass function, in which the population of low mass black holes is greatly reduced.
1275: 
1276: Figure~\ref{fig|multiparamLF} presents a similar comparison for models with different input luminosity functions.
1277: Here we fix $\dot{m}_0=0.75$ in all cases, noting that the value of $\dot{m}_0$ affects the predicted
1278: $\log M_{\rm PEAK}$ but has little impact on other quantities. Eliminating Compton-thick sources from
1279: our standard LF (dotted curve) lowers the efficiency required to match a given $\rho_{\bullet}$ by about
1280: $20\%$, to $\epsilon=0.062$ for $\rho_{\bullet}=4.26\times 10^5\, {\rm M_{\odot}\, Mpc^{-3}}$ (short-dashed line).
1281: Doubling the Compton-thick contribution (not shown) has a similar effect in the opposite direction,
1282: raising the required $\epsilon$ by 20\%. Changing the Compton-thick contribution has little
1283: impact on $\log M_{\rm PEAK}$ for a given $\epsilon$, or on the width or asymmetry of the predicted mass function.
1284: Eliminating \emph{all} obscured sources and retaining only Type I quasars drastically reduces the predicted $\rho_{\bullet}$,
1285: so such a model (which is implausible on direct observational grounds anyway) can only be reconciled
1286: with the local black hole population if the efficiency is very low or if the true
1287: value of $\rho_{\bullet}$ is significantly below our observational estimate. However,
1288: a model in which the full bolometric luminosity function is just three times that of Type I AGN --- two obscured
1289: sources for every unobscured source independent of luminosity and redshift --- yields similar predictions to our full model
1290: with the luminosity-dependent column density distribution of U03.
1291: 
1292: For the HRH07 luminosity function, we require
1293: $0.079\lesssim \epsilon \lesssim 0.125$ to reproduce the grey-band range of $\rho_{\bullet}$, $25\%$ higher than our standard
1294: LF at fixed $\rho_{\bullet}$. As discussed earlier, this difference mainly reflects the higher bolometric correction
1295: adopted by HRH07, which boosts the normalization of the LF at intermediate redshifts. The HRH07 LF also leads to a higher
1296: $M_{\rm PEAK}$ at a given $\epsilon$, and matching $\log M_{\rm PEAK}/{M_{\odot}}\approx 8.5$ implies
1297: $0.7\lesssim \dot{m}_0 \lesssim 1.1$ for the range of $\epsilon$ that matches $\rho_{\bullet}$. The HRH07
1298: LF predicts a somewhat narrower and more asymmetric mass function, mainly because of the steeper bright-end
1299: slope at intermediate redshifts, which steepens the high mass end of the black hole mass function. These results
1300: accord with those shown already in Figure~\ref{fig|Models}d. If we adopt \emph{both} the HRH07 luminosity function
1301: and the Hopkins et al. (2007b) black hole mass function, which has $\rho_{\bullet}=5.7\times 10^5\, {\rm M_{\odot}\, Mpc^{-3}}$
1302: and $\log M_{\rm PEAK}/{\rm M_{\odot}}\approx 8.4$, we find $\epsilon=0.075$ and $\dot{m}_0=1.3$.
1303: 
1304: The remaining uncertainties in the luminosity function and the local black hole mass function still leave considerable
1305: uncertainty in the radiative efficiency. If we assume that our no-Compton-thick LF and the HRH07 LF bracket
1306: the luminosity function uncertainty and our grey band brackets the $\rho_{\bullet}$ uncertainty, the allowed range
1307: is $0.05\lesssim \epsilon \lesssim 0.125$, and there is clearly some room for the luminosity function or $\rho_{\bullet}$
1308: to be outside this range. However, the high efficiencies predicted
1309: by MHD simulations of thick accretion disks around rapidly spinning black holes, $\epsilon\approx 0.15-0.2$
1310: (e.g., Gammie et al. 2004), can only represent the mean efficiency of black hole accretion if $\rho_{\bullet}$
1311: is considerably below our estimates or the luminosity function is substantially higher than even the HRH07 estimate. There are some
1312: $\rho_{\bullet}$ estimates as low as $2-3\times 10^5\, {\rm M_{\odot}\, Mpc^{-3}}$ (e.g., Wyithe et al. 2004). Our conclusions about efficiency are in agreement with most
1313: previous studies, though Elvis et al. (2002) reached a different conclusion mainly because they assumed that the sources producing
1314: the X-ray background have an effective mean redshift of $z\sim 2$, while subsequent data have shown that the peak contribution to the background
1315: is at $z\lesssim 1$ (e.g., U03; La Franca et al. 2005).
1316: 
1317: 
1318: \subsection{Redshift-dependent Eddington ratio}
1319: \label{subsec|etamdot}
1320: 
1321: So far we have assumed that all black holes at all times radiate at
1322: a constant, mildly sub-Eddington rate, an assumption supported by
1323: Kollmeier et al. (2005)'s results for luminous quasars in the AGES survey at $0.5\lesssim z \lesssim 4$.
1324: However,  Eddington ratios of local AGN have a wide distribution, with a large fraction of Seyferts radiating
1325: at $\lambda < 0.1$. Some studies also show evidence for
1326: mass-dependence of the typical Eddington ratio (e.g., Heckmann et al. 2004; Vestergaard 2004; McLure \& Dunlop 2004; Constantin \& Vogeley 2006;
1327: Dasyra et al. 2007; Shen et al.\ 2007b), though systematic uncertainties in the
1328: reverberation mapping techniques and extrapolations of empirical
1329: virial relations (e.g., Kaspi et al. 2000; Bentz et al. 2006)
1330: make it hard to draw firm conclusions.
1331: In this section and the one that follows, we
1332: discuss simple models with $\dot{m}_0$ values (and thus Eddington ratios)
1333: that depend on redshift or mass. More realistic models would incorporate a \emph{range}
1334: of Eddington ratios with the relative importance of high and low accretion rates depending on mass or
1335: redshift; we will investigate such models in future work. We note again that the integrated
1336: black hole mass density should depend only on $\epsilon$ (equation~\ref{eq|soltan}), but
1337: the shape of the black hole mass function can change if $\dot{m}_0$ is not constant.
1338: 
1339: Figure~\ref{fig|etamdot} compares the results of our reference model, which has $\epsilon=0.065$,
1340: $\dot{m}_0=0.6$, to a model with the same $\epsilon$ but
1341: \begin{equation}
1342: \dot{m}_0(z)=[1-\exp(-z/z_s)]\, ,
1343:     \label{eq|mdotz}
1344: \end{equation}
1345: with $z_s=2$. This implies values of $\dot{m}_0=0.78,0.63,0.39,0.22,0.095, 0.049$ at
1346: $z=3,2,1,0.5,0.2,0.1$, respectively, which are consistent with the
1347: average drop shown in Figure 6b of Vestergaard et al. (2004) for
1348: the more massive black holes, assuming the distribution starts at
1349: $\dot{m}_0\approx 1$ at very high redshifts. The decreasing $\dot{m}_0(z)$ significantly
1350: alters the shape of the black hole mass function (Figure~\ref{fig|etamdot}a) by shifting the low
1351: redshift growth of the mass density away from low mass black holes and towards higher mass black
1352: holes. This model still requires a substantial number of low-mass black holes residing in spiral galaxy
1353: bulges if it is to match the local mass function.
1354: 
1355: %The choice of
1356: %defining the Eddington accretion rate in units of a fixed radiative
1357: %efficiency is motivated by accretion disk physics, in which we
1358: %consider a 10\% radiative efficiency proper of a thin-disk. However
1359: %at very low accretion rates the structure of the in-falling gas and
1360: %the much lower viscosity around the black hole, could damp the
1361: %radiative emission, due to the less efficient coupling between
1362: %electrons and ions. Following Narayan et al. (1998) we therefore
1363: %reduce the radiative efficiency when the accretion is below the
1364: %critical value $\dot{m}_{\rm crit}=0.01$. However we note that even
1365: %neglecting equation~(\ref{eq|epsilon}) does not alter the main
1366: %conclusions of this section.
1367: %On the other
1368: %extreme, when the accretion rate is very large, in particular it is
1369: %super-Eddington, if the optical depth is not very large, radiation
1370: %can leak out of the gas as it moves inwards. The radiation drag can
1371: %then decrease the sound speed in the inner regions and limit the
1372: %subsequent accretion rate. For simplicity we will assume that the
1373: %radiative efficiency scales inversely with the accretion rate to
1374: %limit the emission to the Eddington value (e.g. Begelman 1978,
1375: %Kawaguchi et al. 2004), even if some leak of .
1376: %In Figure~\ref{fig|etamdot}a we compare the predictions of this
1377: %model with the reference model. In general, the net effect of the
1378: %$\dot{m}_0(z)$ strong decrease with redshift, is to limit the late
1379: %growth of black holes (equation~(\ref{eq|mdotav})), especially for
1380: %the less massive/luminous ones which start growing faster at later
1381: %times with respect to the more massive ones (see
1382: %Figure~\ref{fig|etamdot}c). The number density in the accreted mass
1383: %function starts decreasing for masses $M_{\bullet}\lesssim
1384: %10^{7.7}\, {\rm M_{\odot}}$, as it can be seen in
1385: %Figure~\ref{fig|etamdot}a by comparing the $z\sim 0$ output of this
1386: %model (solid line) with the reference model (squares). However, note
1387: %that even if these models significantly reduce the cumulative number
1388: %of low mass-black holes, they still require a substantial amount of
1389: %low-mass black holes to reside in the bulges of spirals in the local
1390: %mass function to preserve the statistical match.
1391: Figures~\ref{fig|etamdot}b compares the duty cycles produced by the
1392: $\dot{m}_0(z)$ model to those from the reference model, at several redshifts.
1393: The reference model duty cycles are
1394: strongly decreasing functions of redshift and black hole mass, dropping by up
1395: to two orders of magnitude in the lower redshift bins. The
1396: $\dot{m}_0(z)$-model ``damps'' the downsizing, leaving a much milder
1397: variation of the duty cycle over time and mass. As emphasized by SW03, the downward shift of the LF break luminosity at $z<2$ can be explained either
1398: by a strong suppression of activity in high mass black holes or by a drop in characteristic
1399: Eddington ratios. Our $\dot{m}_0(z)$ model is a compromise between these two explanations;
1400: higher mass black holes still build their mass earlier than low mass black holes, but the difference
1401: is smaller than in the reference model (Figure~\ref{fig|etamdot}c).
1402: 
1403: Greene \& Ho (2007) have
1404: recently estimated the local MF for active black holes through the
1405: virial relations mentioned above. When compared with the local mass function of
1406: relic black holes, their study supports a
1407: duty-cycle of $P_0\sim 10^{-2.35}$ for black holes with mass $\log M_{\bullet}/M_{\odot}\gtrsim 7$,
1408: consistent with the $z\sim 0$ duty-cycle inferred from this model,
1409: and about two orders of magnitude above the one corresponding to
1410: our reference model (Figure~\ref{fig|etamdot}b).
1411: 
1412: \subsection{Eddington ratio varying with black hole mass}
1413: \label{subsec|dmdtMbh}
1414: 
1415: In the previous section we studied a model in which the Eddington ratio steadily decreases
1416: with redshift and is fixed with black hole mass. Here we
1417: consider the complementary model in which the Eddington ratio distribution is
1418: constant with redshift but declines with black hole
1419: mass.
1420: 
1421: We adopt
1422: \begin{equation}
1423: \dot{m}_0(M_{\bullet})=0.445\times \left(\frac{M_{\bullet}}{10^9\,
1424: {\rm M_{\odot}}}\right)^{0.5}\,
1425:     \label{eq|EddMbh}
1426: \end{equation}
1427: with a constant radiative efficiency $\epsilon=0.075$.
1428: As discussed in \S~\ref{subsec|method},
1429: we evolve this model by integrating equation~(\ref{eq|POsingle}) with a fixed black hole mass grid, extrapolating
1430: the luminosity function below $L_{\rm MIN}(z)$. Relative to the reference model, the $\dot{m}_0(M_{\bullet})$ model associates
1431: the peak of the LF with higher mass black holes, especially at lower redshifts. The low redshift growth
1432: of low mass black holes is strongly suppressed, and the $z=0$ abundance of low mass black holes is much lower,
1433: consistent with the Graham et al. (2007) local mass function estimate, as shown in Figure~\ref{fig|dmdtMbh}.
1434: The dependence of duty cycle on mass is also much stronger in this model because
1435: of the paucity of low mass black holes; note that the duty cycle at low masses
1436: is high but the amount of growth is low because of the low accretion rates. Getting our mass-dependent
1437: $\dot{m}_0$ model to run at all requires a high initial black hole density, with a
1438: duty cycle $P_0(z=6)=0.01-0.02$; otherwise, the slow growth of low mass black holes leads to required
1439: duty cycles that exceed unity at intermediate redshifts. Alternatively, we can start with a higher initial
1440: duty cycle and only ``turn on'' the mass dependence of equation~(\ref{eq|EddMbh}) after an epoch
1441: of growth at constant $\dot{m}_0$.
1442: 
1443: More generally, we can vary the Eddington ratio both in redshift and mass,
1444: $\dot{m}_0(M_{\bullet},z)=T(z)B(M_{\bullet})$.
1445: The continuity equation can then be written
1446: \begin{eqnarray}
1447: \frac{\partial n(M_{\bullet},t)}{\partial t}=-\frac{T(z)}{(\ln 10)^2M_{\bullet}}\times\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\\
1448: \nonumber \frac{\partial}{\partial \log M_{\bullet}}
1449: \nonumber\left[B(M_{\bullet})\Phi(L,z)\left|\frac{d \log
1450: \nonumber L}{d\log M_{\bullet}}\right|\right]=
1451: \nonumber -\frac{T(z)}{\ln(10)^2M_{\bullet}}\times\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\\
1452: \nonumber  \left[B(M_{\bullet})\frac{\partial
1453: \nonumber \Phi(L,z)}{\partial \log L}\left(\frac{\partial\log L}{\partial\log
1454: \nonumber M_{\bullet}}\right)^2+\left|\frac{\partial \log L}{\partial \log M_{\bullet}}\right|
1455: \nonumber \Phi(L,z)\frac{\partial B(M_{\bullet})}{\partial \log M_{\bullet}}\right] \, ,
1456: \label{eq|MeddMbh}
1457: \end{eqnarray}
1458: where we have set to zero the \emph{second} derivative $\partial^2 \log L/(\partial \log M_{\bullet})^2$
1459: assuming a power-law relation like that in equation~(\ref{eq|EddMbh}).
1460: We have used equation~(19)
1461: to investigate models in which we vary the Eddington ratio with redshift and mass
1462: as given in equations~(\ref{eq|mdotz}) and ~(\ref{eq|EddMbh}).
1463: We find that we can get somewhat better fits to Graham et al.'s (2007) local
1464: mass function, since both the implemented trends
1465: act to decrease the numbers of low mass black holes.
1466: 
1467: %In Figure~\ref{fig|dmdtMbh}c we plot the output of a model in which
1468: %we set $T(z)$ as given in equation~(\ref{eq|mdotz}), and
1469: %$B(M_{\bullet})$ as given in equation~(\ref{eq|EddMbh}), with
1470: %$\beta=-0.3$ and $K_0=2.5$, and set $\epsilon=0.065$, fixed in time.
1471: %The solution is then tracked at each mass and at each time solving
1472: %equation~(\ref{eq|MeddMbh}). An higher absolute value of the
1473: %gradient $\beta$ is ruled out. This model provides a good match with
1474: %the local mass estimate by Graham et al. (2007) mass function.
1475: %Similarly as in Figure~\ref{fig|etamdot}b, this model predicts much
1476: %higher duty-cycles at low redshifts, as shown in
1477: %Figure~\ref{fig|dmdtMbh}d, due to the redshift dependence on the
1478: %Eddington ratio, but still a strong mass gradient at fixed redshift,
1479: %due to the mass dependence in the $B$ term.
1480: 
1481: \subsection{Black hole mergers}
1482: \label{subsec|merging}
1483: 
1484: Observations show, and hierarchical galaxy formation models predict, that a significant
1485: fraction of galaxies experience mergers with comparably massive galaxies during their
1486: lifetime. At least some of these galaxy mergers are likely to be accompanied by mergers of
1487: the central black holes that they contain, though the mechanisms that shrink
1488: black hole orbits to the scale where gravitational radiation can drive a final
1489: merger are not fully understood (see Merritt \& Milosavljevi\'{c} 2005).
1490: In a future paper, we will
1491: will combine our accretion model with theoretically
1492: predicted merger rates for cold dark matter subhalos (Yoo et al. 2007)
1493: to create models with realistic contributions of accretion and merger driven growth.
1494: Here we illustrate the potential impact
1495: of mergers on the black hole mass function using a simple mathematical model
1496: that assumes constant probability
1497: of equal mass mergers per Hubble time,
1498: similar to the models of Richstone et al. (1998) and SW03.
1499: %Galaxies are born and develop
1500: %within dark matter halos which do undergo severe merging during
1501: %cosmic time. Observationally, mergers within two galaxies, usually
1502: %star-forming, are clearly observed in the Mid-infrared bands,
1503: %showing strong signatures in their disturbed morphology of the past
1504: %or ongoing interaction (e.g., Dasyra et al. 2007). Also,
1505: %interferometric CO surveys have revealed double or multiple
1506: %starforming subclumps merging or interacting at high redshifts (e.g,
1507: %Greve et al. 2005, Tacconi et al. 2006). Statistically speaking, it
1508: %still not clear the role merging events play in the actual build up
1509: %of galaxies of different morphological types or, better, at what
1510: %stage star formation takes place, if sets in already in the
1511: %subclumps which are going to merge, or, predominantly, when they are
1512: %already assembled. In any event, black holes do reside at the center
1513: %of galaxies today and, if merging plays any role in building
1514: %galaxies, it must involve the central black holes, at some level.
1515: %However, it is also true that the merging rate of galaxies could be
1516: %quite different from that of black holes, due to the much longer and
1517: %also unknown timescales involved in getting black pairs close enough
1518: %to start runaway merging (see, e.g., Merritt \& Milosavljevi\'{c}
1519: %2005 for a review). While we reserve for a future paper a detailed
1520: %study in which we will combine the growth histories of black holes
1521: %with theoretically predicted merger rates (see also Volonteri et al.
1522: %2005), in this section we present a simplified merger model, similar
1523: %to that proposed by Steed \& Weinberg (2003), to probe the generic
1524: %effects merging have on the mass distribution of black holes within
1525: %the accretion models developed so far.
1526: %
1527: %After the results discussed in \S~\ref{sec|results}, we can
1528: %conclude that radiative accretion must have played a dominant role
1529: %in building up the local black hole population. Here we want to
1530: %probe to what extent models which include accretion \emph{and}
1531: %merging, are still capable of reproducing the match between the
1532: %local mass and accreted mass functions.
1533: %
1534: %black holes can grow in mass through merging and gas accretion. Hughes \&
1535: %Blandford (2003) studied the remnants of merging black holes of different
1536: %sizes, spins and orbital parameters. They concluded that black holes are
1537: %generally spun-down by mergers and only for a very narrow range of
1538: %orbital parameters merging is effective in rapidly spinning the
1539: %holes. Gammie et al. (2004) carried out numerical simulations
1540: %confirming the Hughes \& Blandford (2003) result that a black hole is
1541: %usually spun down by frequent minor mergers, while equal mass-major
1542: %mergers usually spun-up the spins to value $a/M_{\bullet}\gtrsim
1543: %0.8$, with a spin equilibrium value $a\sim 0.93$, i.e. $\epsilon
1544: %\sim 0.16$, for accretion from a thin disk (see also Volonteri et
1545: %al. 2006).
1546: We will ignore physically interesting complications such as ejection of black holes
1547: by gravitational radiation or three-body interactions or variations of radiative efficiency
1548: caused by the impact of mergers on the black hole
1549: spin distribution (e.g., Hughes \& Blandford 2003; Volonteri et
1550: al. 2003; Gammie et al. 2004; Islam et al. 2004; Yoo \& Miralda-Escud\'{e} 2004;
1551: Merritt \& Milosavljevi\'{c} 2005; Volonteri et
1552: al. 2006). These effects, along with unequal mergers and the mass and redshift-dependence
1553: of merger rates, should be considered in a complete model of the evolving
1554: black hole population. Mergers redistribute mass within the black hole population,
1555: but they do not change the integrated mass density, so they do not affect
1556: the integrated So{\l}tan (1982) argument. (In principle, gravitational radiation
1557: during mergers can \emph{reduce} the integrated mass density [see Yu \& Tremaine 2002], but we do
1558: not consider this effect here).
1559: 
1560: %In this model we allow black holes to accrete and merge, and assume
1561: %that merging events, on average, do not modify the spin distribution
1562: %of the black hole population at any time, therefore maintaining the
1563: %mean radiative efficiency constant throughout the evolution, in
1564: %agreement with several studies (e.g., Hughes \& Blandford 2003,
1565: %Gammie et al. 2004, Volonteri et al. 2005, 2006).
1566: %
1567: %In the following we will therefore keep the radiative efficiency
1568: %constant during the whole evolution, however pointing out that any
1569: %increase in its best-fit value would worsen the match with the local
1570: %mass function.
1571: %Volonteri et al. find that about 70\% of the whole final
1572: %distribution of black holes is
1573: %maximally rotating. %Shapiro (2005) finally conducted relativistic
1574: %MHD simulations of disk accretion onto Kerr black holes which showed that
1575: %the maximal equilibrium spin rate reachable by the system in the
1576: %presence of significant magnetic fields is $a/M_{\bullet}\gtrsim
1577: %0.95$, corresponding to $\epsilon \sim 0.2$.
1578: 
1579: In our model, a black hole of mass $M_{\bullet}$ has a probability
1580: $P_{\rm merg}$ of merging with an equal mass black hole in the
1581: Hubble time $t_H(z)$ (age of the Universe at redshift $z$).
1582: Therefore the fraction $F$ of black holes that merge in a timestep
1583: $\Delta t$ is
1584: \begin{equation}
1585: F=P_{\rm merg}\times\frac{\Delta t}{t_H(z)}\, .
1586:     \label{eq|mergF}
1587: \end{equation}
1588: At each time $t_1$, we first advance the mass function to time
1589: $t_2=t_1+\Delta t$ with accretion only, then add to each bin of the mass function
1590: an increment (which may be positive or negative)
1591: \begin{equation}
1592: \Delta \Phi(M_{\bullet},t_2)=\frac{F\times \Phi
1593: \left(\frac{M_{\bullet}}{2},t_2\right)}{2}-F\times
1594: \Phi(M_{\bullet},t_2)\, ,
1595:     \label{eq|merging}
1596: \end{equation}
1597: where the second term represents black holes lost from the bin by merging
1598: and $\Phi \left(\frac{M_{\bullet}}{2},t_2\right)$ is calculated by
1599: interpolation.
1600: 
1601: Figure~\ref{fig|merging} shows the evolution of the black hole mass function for a model
1602: with $P_{\rm merg}=0.5$ and our reference model values of
1603: $\epsilon=0.065$ and $\dot{m}_0=0.6$. Comparing the $z=0.02$
1604: output to that of the reference model shows that the net effect of
1605: merging is to slightly lower the abundance of black holes below the peak
1606: of the mass function and to significantly increase the abundance of very massive
1607: black holes, as expected
1608: (see also Malbon et al. 2006; Yoo et al. 2007).
1609: Although we start merging at $z=6$, comparison of Figure~\ref{fig|merging}
1610: to Figure~\ref{fig|bestFitModel}b shows that the mass function is nearly identical
1611: to that of the reference model down to $z=1$. Mergers at this level
1612: only change the mass function noticeably once
1613: accretion-driven evolution has slowed.
1614: 
1615: Since our reference model already produces an excess of massive
1616: black holes relative to local estimates of the black hole function,
1617: adding mergers only makes the match to observations worse. However,
1618: the impact of mergers is evident mainly at $M_{\bullet}>10^9\, {\rm
1619: M_{\odot}}$, where the local mass function estimates are most
1620: sensitive to the adopted scatter in the black hole-host scaling
1621: relations and to the extrapolations of these relations to the most
1622: luminous galaxies. Accommodating the predictions of our merger model
1623: would require an intrinsic scatter of $\sim 0.5$ dex at high masses,
1624: or a change in slope of the scaling relations. Either of these
1625: changes seems possible given the current observational
1626: uncertainties, but neither seems likely (see, e.g., discussions by
1627: Wyithe 2004; Batcheldor et al. 2006; Lauer et al. 2006; Tundo et al.
1628: 2007). If we adopt the HRH07 LF in place of our standard LF, then
1629: the predicted black hole mass function \emph{with} mergers is
1630: similar to that of our reference model \emph{without} mergers (see
1631: Figure~\ref{fig|merging}). The impact of mergers at this level is
1632: therefore comparable to the systematic uncertainties associated with
1633: the AGN luminosity function.
1634: 
1635: In fact, the value $P_{\rm merg}=0.5$ is high compared to
1636: theoretical predictions for massive galaxies (Maller et al. 2006),
1637: and theoretically predicted merger rates decline towards lower
1638: masses. Observational estimates of the galaxy merger rate and its
1639: mass dependence span a substantial range (see, e.g., Bell et al.
1640: 2006; Conroy et al. 2007; Masjedi et al. 2008, and references
1641: therein), and $P_{\rm merg}=0.5$ is roughly consistent with the high
1642: end of these estimates. We conclude that the impact of mergers on
1643: the black hole mass function is probably small compared to remaining
1644: uncertainties in accretion-driven growth, except perhaps for the
1645: rare, high mass black holes. The most interesting impact of black
1646: hole mergers may arise indirectly, through their effect on black
1647: hole spins and thus on radiative efficiencies (e.g., Volonteri et
1648: al. 2005).
1649: %In Figure~\ref{fig|merging}b we attempt to show a "full" model which
1650: %contains an Eddington ratio decreasing with redshift, as given in
1651: %eqaution~(\ref{eq|mdotz}), mass-dependent as in
1652: %equation~(\ref{eq|EddMbh}), with $\beta=-0.3$, $K_0=2.5$,
1653: %$\epsilon=0.065$ and $P_0(z=6)=0.5$ and with merging parameters
1654: %$P_{\rm merg}=0.5$ and activation redshift $z=3$. This choice of the
1655: %parameters is made in way to produce a good overall fit to the
1656: %Graham et al. (2007)'s local mass function, shown with filled and
1657: %open circles.
1658: %
1659: %We do not show here the accretion mass histories and duty-cycles of
1660: %black holes for these kind of models with accretion and merging.
1661: %This is because the outputs in this case are very similar for the
1662: %simple only-accretion models, within a maximum $20\%$ uncertainty.
1663: 
1664: \subsection{Tabulation of luminosity and mass functions}
1665: \label{subsec|Tables}
1666: 
1667: For the convenience of readers who may wish to use them, we provide
1668: electronic tables that list our estimate of the AGN bolometric
1669: LF and some of our model predictions of the black hole mass function and duty cycle, all
1670: as a function of redshift. Table~\ref{Table|FitsLF} lists
1671: our standard LF estimate, the LF after eliminating or doubling
1672: the number of Compton thick sources, and the HRH07 LF estimate evaluated at the same
1673: redshifts and luminosities for convenient comparison.
1674: Table~\ref{Table|Fits} lists the predicted black hole mass function at the same
1675: redshifts for our reference model values of $\epsilon=0.065$, $\dot{m}_0=0.60$,
1676: computed for each of the luminosity functions in Table~\ref{Table|FitsLF}.
1677: The final column lists the mass function predicted for the HRH07 LF and the parameter
1678: values $\epsilon=0.09$ and $\dot{m}_0=1$, which yield a good match to the local mass function
1679: for this LF. Table~\ref{Table|DutyCycles} lists instead the duty cycles corresponding to the same models
1680: reported in Table~\ref{Table|Fits}, computed at the same redshifts and black hole masses.
1681: [Prior to publication, the full tables can be found in electronic format at
1682: {\tt{http://www.astronomy.ohio-state.edu/\\
1683: $\sim$shankar/Models}}.]
1684: 
1685: \section{Space density and bias of AGN hosts}
1686: \label{sec|hosts}
1687: 
1688: Predictions of the black hole mass function at redshifts $z>0$ can be tested observationally
1689: if one assumes that black hole mass increases monotonically with the luminosity
1690: of the host galaxy or the mass of its parent halo. This assumption is unlikely
1691: to be perfect, but given the tight correlation between black hole mass
1692: and bulge mass observed today, it may be a reasonable approximation. Figure~\ref{fig|hosts}a shows,
1693: at several redshifts, the predictions of our reference model for the cumulative
1694: space density of galaxies that can host an AGN of luminosity $L$ or greater. Since
1695: the model assumes a single $\dot{m}_0$, this is simply equal to the cumulative
1696: space density of black holes of mass $M_{\bullet}>L/(f_{0.1}\dot{m}_0l)=1.9\times 10^8\, {\rm M_{\odot}}
1697: \times (L/{\rm 10^{46}\, {\rm erg\, s^{-1}}})$. If the monotonic assumption holds, then the observed
1698: space density of galaxies brighter than $L_{\rm host}(L_{\rm AGN})$ should equal
1699: the space density predicted in Figure~\ref{fig|hosts}a, where $L_{\rm host}(L_{\rm AGN})$
1700: is the luminosity of galaxies that host AGN of luminosity $L_{\rm AGN}$.
1701: 
1702: As emphasized by Haiman \& Hui (2001) and Martini \& Weinberg (2001), the clustering
1703: of AGN can be a powerful diagnostic for the duty cycle of black hole activity: a
1704: high duty cycle implies that halos of AGN are rare, hence high mass, hence strongly clustered.
1705: In the context of our models, we can predict the halo mass $M_h$ associated with black holes
1706: of mass $M_{\bullet}$ by matching the cumulative mass functions,
1707: %Martini \& Weinberg (2001) and Haiman \& Hui (2001) were the first
1708: %to discuss the importance of measuring AGN bias to infer their
1709: %duty-cycles. In this section we somewhat reverse their approach. As
1710: %discussed above our models do predict an output duty-cycles of black
1711: %holes as a function of mass and redshift. These in turn can be used
1712: %to predict the cumulative number of hosts above a given luminosity
1713: %$L$ which are predicted by any given successful model. The latter
1714: %information can then be simply translated into a measure of the
1715: %bias, as shown below, and then compared to existing data on AGN
1716: %clustering. The So{\l}tan argument has the great advantage of being
1717: %an empirical constraint on black hole evolution; however, it is able
1718: %to constrain the accretion parameters \emph{only} in a "cumulative"
1719: %sense, i.e., after integration over the whole lifetime of the
1720: %universe. The approach we are proposing in this section instead,
1721: %can be used to constrain the radiative efficiency and Eddington
1722: %ratio distributions at a fixed redshift bin, at a given environment and
1723: %luminosity range.
1724: \begin{equation}
1725: \Phi(>M_{\bullet},z)=\Phi(>M_h,z)\, .
1726:     \label{eq|nhosts}
1727: \end{equation}
1728: Here $\Phi(>M_{\bullet},z)$ is the space density of black holes more massive than $M_{\bullet}$
1729: at redshift $z$, in units of ${\rm Mpc^{-3}}$, as predicted by our evolutionary model,
1730: and $\Phi(>M_h,z)$ is the space density of halos more massive than $M_h$ expected for a $\Lambda$CDM
1731: cosmological model, which we compute using the Sheth \& Tormen (1999) halo mass function with
1732: cosmological parameters $\Omega_m=0.3, \Omega_{\Lambda}=0.7, h=0.7, \sigma_8=0.8$ and linear power spectrum
1733: taken from Smith et al. (2003) with $\Gamma=0.2$. This determination assumes both a monotonic
1734: relation between black hole mass and halo mass and one black hole per halo, but we have checked
1735: that allowing black holes to reside in subhalos does not substantially change our results if
1736: we adopt the subhalo statistics of Vale \& Ostriker (2004).
1737: Finally, we compute the bias of black holes of mass $M_{\bullet}$, and thus the bias of
1738: AGN of luminosity $L=f_{0.1}\dot{m}_0 l M_{\bullet}$, using
1739: Sheth et al.'s (2001) analytic model for halo bias,
1740: \begin{equation}
1741: b(L,z)=b(M_{\bullet},z)=b(M_h,z)\, .
1742:     \label{eq|bias}
1743: \end{equation}
1744: Figure~\ref{fig|hosts}b shows the predicted bias as a function of luminosity
1745: $L$ at several redshifts for the reference model of \S~\ref{subsec|bestfit}.
1746: The predicted bias is much stronger at high redshift because the comoving space
1747: density of black holes is lower (Figure~\ref{fig|bestFitModel}) and because
1748: the bias of halos of a given space density is higher. Points in Figure~\ref{fig|hosts}b
1749: show recent observational estimates of AGN clustering bias from the 2dF
1750: and SDSS quasar redshift surveys (da Angela et al. 2007; Myers et al. 2007; Porciani \& Norberg 2007)
1751: in the redshift range $0.8\lesssim z \lesssim 1.3$ and $1.7\lesssim z \lesssim 2.1$ (Porciani \& Norberg 2007)
1752: and from SDSS at $3\lesssim z \lesssim 3.5$
1753: (Shen et al. 2007a). Our reference model is quite successful at predicting
1754: the absolute level of bias (and thus AGN clustering strength) for quasars with $L\ge 10^{46}\, {\rm erg\, s^{-1}}$ at $z\sim 1$,
1755: $z\sim 2$ and $z\sim 3$, and, especially, at explaining the strong trend of bias among these samples.
1756: It appears to underpredict the clustering
1757: of lower luminosity AGN at $z\sim 1$, though Myers et al. (2007) note the including
1758: the possible impact of stellar contamination on their sample expands their error bars by a factor
1759: of $\sim 1.5$ (an effect we have not included in Figure~\ref{fig|hosts}b).
1760: 
1761: We reserve a detailed examination of AGN clustering constraints --- including an assessment of what models
1762: can be ruled out with present data or improved measurements --- for future work. Allowing scatter
1763: between black hole mass and halo mass or a range of Eddington ratios can significantly
1764: alter predictions both for AGN host density and for AGN clustering,
1765: so these classes of models will be especially interesting to explore.
1766: For example, the stronger clustering in the Myers et al. (2007)
1767: sample relative to our reference model predictions
1768: could indicate that a fraction of these lower luminosity AGN
1769: are associated with high mass black holes radiating at low Eddington ratios
1770: (e.g., Lidz et al. 2006).
1771: 
1772: %Our results are shown in Figure~\ref{fig|hosts}. In panel a we show the cumulative number of hosts predicted by
1773: %two models with same Eddington ratio and different radiative efficiency, one the reference value of $\epsilon=0.065$ and the other
1774: %with $\epsilon=0.30$. There is no apparent change in the number. This is somewhat unexpected as, following equation~(\ref{eq|nhosts}), one would
1775: %naively expect that an higher radiative efficiency would predict a lower accreted mass function at all times
1776: %and therefore lower number of hosts. However, this effect is nearly completely compensated by our definition of bolometric
1777: %luminosity given in equation~(\ref{eq|Lbol}), for which at fixed luminosity, to an higher radiative efficiency
1778: %is associated a smaller black hole mass and an higher number density. We then the "classical" definition
1779: %of Eddington ratio,
1780: %\begin{equation}
1781: %\dot{M}_{\rm Edd}\equiv \frac{(1-\epsilon)L_{\rm Edd}}{\epsilon\,
1782: %c^2}
1783: %    \label{eq|MeddCla}
1784: %\end{equation}
1785: %to regenerate the two outputs above and recomputed the number of hosts. The result is shown in Figure~\ref{fig|hosts}b,
1786: %in which is now possible to distinguish the expected trend of decreasing number density with increasing radiative efficiency (thick lines), and viceversa (thin lines).
1787: %Figure~\ref{sec|hosts}c instead shows that by fixing the radiative efficiency and lowering
1788: %the Eddington ratio, an higher black hole mass and therefore a lower number density of hosts is expected (thick lines), and viceversa (thin lines).
1789: %Figure~\ref{fig|hosts}d finally shows our results for a model in which we vary the Eddington ratio
1790: %in time as given in equation~(\ref{eq|mdotz}). Here the number of hosts is higher at high luminosities and at low redshifts, with respect
1791: %to our reference model. This is expected as the black hole mass function is much higher in these intervals
1792: %with respect to our reference model (see Figure~\ref{fig|etamdot}a).
1793: %Figure~\ref{fig|hosts} shows the predicted bias, as given by equation~(\ref{eq|bias}). The bias
1794: %obtained from equations~(\ref{eq|nhosts})-(\ref{eq|bias}) can be considered
1795: %as an independent way to test the validity of a given accretion model.
1796: %Also, by matching the local mass function alone may not be conclusive enough to extend the validity of
1797: %a given accretion model at redshifts $z \gtrsim 3$, as only $\lesssim 10\%$ of the final mass is accreted
1798: %until that moment (see Figure~\ref{fig|SFR-Rhoz}b).
1799: 
1800: \section{Discussion}
1801: \label{sec|DiscuConclu}
1802: 
1803: We have constructed self-consistent models for the evolution of the supermassive
1804: black hole population and the AGN population, in which black holes
1805: grow at the rate implied by the observed luminosity function given assumed values
1806: of the radiative efficiency $\epsilon$ and the characteristic
1807: accretion rate $\dot{m}_0=\dot{M}_{\bullet}/\dot{M}_{\rm Edd}$ (see Eqs.~\ref{eq|conteq}, ~\ref{eq|mdotav}, ~\ref{eq|POsingle}).
1808: These models can be tested against the mass distribution of black holes in the local universe, and they make predictions
1809: for the duty cycles of black holes as a function of redshift and mass, which
1810: can be tested against observations of quasar hosts and quasar clustering if one assumes
1811: an approximately monotonic relation between the masses of black holes and the masses of their host galaxies and halos.
1812: Our method is similar to that used previously by Cavaliere et al. (1971), Small \& Blandford (1992),
1813: Yu \& Tremaine (2002), SW03, Marconi et al. (2004), and S04. However, we have drawn on more recent
1814: data on the AGN luminosity function and the local black hole mass function, and we have considered a broader
1815: spectrum of models. This approach can be considered a ``differential'' generalization of the So{\l}tan (1982) argument
1816: relating the integrated emissivity of the quasar population to the integrated mass
1817: density of the local black hole population.
1818: 
1819: Our model for the bolometric AGN luminosity function starts from the model of U03, based on X-ray
1820: data from a variety of surveys, but we adjust its parameters as a function of redshift in light
1821: of more recent measurements and data at other wavelengths. In agreement with previous
1822: studies we find that the LF of optical AGNs is roughly consistent with that
1823: of X-ray AGN that have absorbing column densities $\log N_H/{\rm cm^{-2}}\le 22$ and that
1824: unobscured AGN dominate the bright end of the LF. We show that the latest constraints on the
1825: hard X-ray background ($E\sim 10-100$ keV) from \emph{INTEGRAL} and from the \emph{PDS} instrument
1826: on BeppoSax support a reduced normalization relative to extrapolations
1827: from other missions at lower energies.
1828: They therefore favor a lower contribution from very highly
1829: obscured AGN ($\log N_H/{\rm cm^{-2}}\gtrsim 24.5$) than some previous estimates
1830: (but see also Gilli et al. 2007). Our estimate of the bolometric
1831: AGN LF is independent in implementation but similar in spirit to that of HRH07.
1832: The most significant differences for purposes of this investigation
1833: are that HRH07 have a higher LF normalization at $z\sim 1-3$, principally because of their choice of bolometric correction,
1834: and they have a steeper bright-end slope at $z\sim 1-2.5$, where they adopt
1835: the slopes measured by Richards et al. (2006) from the SDSS and we use the slopes inferred
1836: by U03 from X-ray data. We regard the differences between our estimate and that of HRH07 as a reasonable
1837: indication of the remaining systematic uncertainties in the bolometric LF of AGNs.
1838: 
1839: With our LF estimate, the bolometric emissivity of the AGN population tracks recent estimates of the cosmic star
1840: formation rate as a function of redshift.
1841: (Comparisons based on the space density of high luminosity quasars [e.g., Richards et al. 2006,
1842: Osmer 2004 and references therein]
1843: reach a different conclusion because at low redshifts the bright
1844: end of the AGN LF drops much more rapidly with time than the overall emissivity.)
1845: The integrated black hole mass density implied by this emissivity
1846: is $\sim 8\times 10^{-4}$ of the stellar
1847: mass at all redshifts, or about half of the estimated ratio
1848: of black hole mass to bulge stellar mass in local galaxies.
1849: This tracking favors scenarios in which black holes and the stellar mass of bulges grow
1850: in parallel, with about $50\%$ of the star formation linked to black hole growth
1851: at all redshifts. These findings are hard to reconcile with any models where
1852: black hole growth substantially precedes stellar mass buildup, or with recent claims
1853: that the ratio of black hole mass to stellar mass is much larger at high redshift than the local value.
1854: However, our finding refers to integrated densities, so it does not indicate
1855: the relative timing of black hole and bulge growth on an object-by-object basis.
1856: 
1857: Observational estimates of the local black hole mass function still show substantial discrepancies
1858: among different authors, depending on the correlation used to derive it (e.g., $M_{\bullet}-\sigma$, $M_{\bullet}-L_{\rm bulge}$,
1859: $M_{\bullet}-M_{\rm star}$, $M_{\bullet}$-S\'{e}rsic index, or fundamental plane),
1860: the calibration of the correlation, and the intrinsic scatter of the correlation.
1861: Above $M_{\bullet}\sim 10^9\, {\rm M_{\odot}}$, estimates depend on extrapolation
1862: of the observed scaling relations, and below $M_{\bullet}\sim 10^{7.5}\, {\rm M_{\odot}}$
1863: they are sensitive to the treatment of spiral bulges. The grey band
1864: in Figure~\ref{fig|LocalMFs} (and subsequent figures) encompasses most
1865: estimates, but the fundamental plane (Hopkins et al. 2007b) and S\'{e}rsic
1866: index (Graham et al. 2007) methods imply more sharply peaked mass functions. The integrated
1867: mass densities of all of these estimates are in the range $\rho_{\bullet}\sim 3-5.5\times 10^5\, {\rm M_{\odot}\,Mpc^{-3}}$.
1868: 
1869: Our simplest models assume a single characteristic Eddington accretion rate $\dot{m}_0$,
1870: independent of redshift and black hole mass, and all of our models assume a single
1871: value of the radiative efficiency $\epsilon$. Matching the local
1872: black hole mass density requires $\epsilon=0.075\times(\rho_{\bullet}/4.5\times 10^5\, {\rm M_{\odot}\,Mpc^{-3}})^{-1}$
1873: for our standard estimate of the AGN LF, or $\epsilon \sim 0.094(\rho_{\bullet}/4.5\times 10^5\, {\rm M_{\odot}\,Mpc^{-3}})^{-1}$
1874: for the HRH07 luminosity function.\footnote{More precisely, it is $\epsilon/(1-\epsilon)$ that
1875: is proportional to $\rho_{\bullet}^{-1}$, but the difference from
1876: $\epsilon\propto \rho_{\bullet}^{-1}$ is tiny over the allowed range.}
1877: With $\epsilon$ thus fixed, the value of $\dot{m}_0$ determines
1878: the peak of the predicted local black hole mass function in the
1879: $M_{\bullet}\Phi(M_{\bullet})$ vs $M_{\bullet}$ plane. Note that
1880: with our definitions the Eddington luminosity ratio
1881: is $\lambda=L/L_{\rm Edd}\approx \dot{m}_0(\epsilon/0.1)$ (equation~\ref{eq|Lambda}).
1882: Matching the observed
1883: peak at $\log M_{\bullet}/{\rm M_{\odot}}\sim 8.5$
1884: implies $\dot{m}_0\approx 0.6$ ($\lambda \approx 0.45$) for our standard LF estimate and
1885: $\dot{m}_0\approx 1$ ($\lambda \approx 0.95$)
1886: for the HRH07 LF. Lower values, $\dot{m}_0=0.1-0.3$, shift
1887: the peak location to untenably high masses, $\log M_{\bullet}/{\rm M_{\odot}}\sim 8.9-9.3$
1888: or, for HRH07, $\log M_{\bullet}/{\rm M_{\odot}}\sim 9.1-9.6$.
1889: The single-$\dot{m}_0$ models achieve a reasonable match to our ``grey-band''
1890: observational estimates of the width and asymmetry of the local mass function, though
1891: for our LF estimate the predicted mass function is too high at $M_{\bullet}>10^9\, {\rm M_{\odot}}$
1892: and is therefore somewhat too broad. Single-$\dot{m}_0$ models cannot
1893: reproduce the more sharply peaked local mass functions estimated by Graham et al. (2007)
1894: or Hopkins et al. (2007b).
1895: 
1896: Our reference model, which has $\epsilon=0.065$, $\dot{m}_0=0.60$,
1897: and our standard LF estimate, predicts a duty cycle for activity of
1898: $10^9\, {\rm M_{\odot}}$ black holes that declines steadily from
1899: 0.15 at $z=4$ to 0.07 ($z=3$), 0.035 ($z=2$), 0.004 ($z=1$), and
1900: $10^{-4}$ ($z=0$). The decline in duty cycle for lower mass black
1901: holes is much shallower. Massive black holes therefore build their
1902: mass relatively early while low mass black holes grow later, the
1903: phenomenon often referred to as ``downsizing''. Our results on mean
1904: radiative efficiency and duty cycle evolution are also in
1905: qualitative agreement with those found by Haiman et al. (2004). The
1906: predicted duty cycles seem in reasonable accord with observational
1907: estimates, though these estimates have considerable uncertainty and
1908: do not, as yet, probe mass and redshift dependence in much detail.
1909: The electronic tables described in \S~\ref{subsec|Tables} provide
1910: tabulations as a function of redshift of our AGN bolometric LF
1911: estimate, the HRH07 LF, and black hole mass functions and duty
1912: cycles for single-$\dot{m}_0$ models that are in good agreement with
1913: the observed $z=0$ mass functions given these LF inputs.
1914: 
1915: We have examined models in which the Eddington ratio accretion rate
1916: $\dot{m}_0$ is reduced at low redshift or at low black hole mass.
1917: Declining redshift evolution of $\dot{m}_0$ damps ``downsizing,''
1918: reducing the dependence of duty cycles on black hole mass and
1919: redshift. This model produces a typical duty cycle $P_0\sim
1920: 10^{-2.5}$ at $z=0$, about two orders of magnitude higher than in
1921: the reference model and consistent with the local duty cycle
1922: inferred from observations by Greene \& Ho (2007). In general terms,
1923: the observed luminosity-dependent density evolution of the AGN LF
1924: can be explained by preferential suppression of activity in high
1925: mass black holes at low redshift, by a decline in the typical
1926: accretion rate at low redshift, or by some combination thereof. The
1927: mass-dependent $\dot{m}_0$ model associates more of the AGN
1928: emissivity to high mass black holes, so it predicts a $z=0$ mass
1929: function that is more sharply peaked, in better agreement with the
1930: estimates of Hopkins et al. (2007b) and Graham et al. (2007) but
1931: worse agreement with other estimates. This model predicts a stronger
1932: mass dependence of duty cycles than our reference model because it
1933: maps low mass black holes, whose abundance is already suppressed, to
1934: less luminous, and hence more common, AGN.
1935: 
1936: We have also considered a model in which each black hole has a $50\%$ probability per Hubble time of merging
1937: with another black hole of equal mass. Merger-driven growth in this
1938: model has little impact on the black hole mass function until $z<1$, when
1939: accretion-driven evolution has slowed. Low redshift mergers slightly
1940: depress the low mass end of the $z=0$ mass function and significantly
1941: enhance the high mass tail, worsening the agreement with observational
1942: estimates. Models incorporating theoretically predicted merger rates can allow more realistic
1943: calculations of the impact of mergers on the black hole population; this impact will
1944: probably be smaller than in the simplified model considered here.
1945: 
1946: We have calculated the clustering bias of AGN as a function of luminosity and redshift
1947: for our reference model, assuming a monotonic relation between black hole mass
1948: and halo mass. The predictions are in reasonable accord with observational
1949: estimates. We will examine AGN clustering predictions in more detail
1950: in future work, with attention to what models can be excluded by the data and what can
1951: be learned by matching the full AGN correlation function
1952: in addition to an overall bias factor.
1953: 
1954: MHD simulations (e.g., Gammie et al. 2004; Shapiro 2005) show that disk accretion
1955: onto Kerr black holes spins them up to an equilibrium spin rate $a\approx 0.95$
1956: (where $a=1$ is the angular momentum parameter for a maximally rotating black hole). The radiative
1957: efficiency in these models is $\epsilon\approx 0.16-0.2$.
1958: These high efficiencies would lead to black hole mass densities a factor of two or more
1959: below our central estimate, and below our estimated lower bound.
1960: Furthermore, our results show that models with $\epsilon$ in the range
1961: $0.06-0.11$ can achieve a good match to the overall shape of the
1962: black hole mass function near the peak in $M_{\bullet}\Phi(M_{\bullet})$,
1963: not just the value of $\rho_{\bullet}$, given
1964: plausible choices of $\dot{m}_0$. Systematic uncertainties
1965: in the AGN LF do not appear large enough to accommodate
1966: $\epsilon\gtrsim 0.15$. Accommodating these high efficiencies
1967: would instead require a substantial downward revision of recent
1968: estimates of the local black hole mass function, reducing the integrated mass density
1969: to $\rho_{\bullet}\sim 2\times 10^5\, {\rm M_{\odot}\,Mpc^{-3}}$. Our results
1970: are consistent with a scenario like the one
1971: of King \& Pringle (2006) in which chaotic accretion spins down
1972: black holes because of counter-alignment with the accretion disk angular momentum,
1973: or with other mechanisms that reduce efficiencies below
1974: the MHD-simulation predictions.
1975: 
1976: The assumption that all active black holes at a given mass and
1977: redshift have the same $\dot{m}_0$ is clearly an idealization, at
1978: least in the local universe where observations indicate a wide range
1979: of Eddington ratios (Heckman et al.\ 2005; Greene \& Ho 2007). Steed
1980: \& Weinberg (2003) and Yu \& Lu (2004) discussed continuity equation
1981: models evolved adopting a distribution of Eddington ratios. In
1982: particular, Yu \& Lu (2004) have derived the relation between the
1983: integrated number of AGNs shining at all times at a given luminosity
1984: $L$, the mean light curve of black holes, and the local black hole
1985: mass function. Following their equation (18), we find that a good
1986: match between the cumulative number of AGNs and of relic black holes
1987: can be achieved for $\epsilon\sim 0.07$ (required by the So{\l}tan
1988: argument) and a mean AGN light curve exponentially increasing with
1989: $\lambda=0.6$ and a negligible declining phase, similar to our
1990: reference model. Alternatively, we find that a good match can be
1991: obtained by assuming that black holes grow rapidly in a
1992: Super-Eddington phase with $\lambda\gtrsim 2$ and then have a long
1993: declining phase, qualitatively resembling our $\dot{m}(z)$ model
1994: discussed in \S~\ref{subsec|etamdot}. In future work we will
1995: investigate models that incorporate multiple $\dot{m}_0$-values and
1996: accretion modes, including the addition of modest log-normal scatter
1997: in $\dot{m}_0(M,z)$ (e.g., Kollmeier et al.\ 2005; Netzer et al.\
1998: 2007; Shen et al.\ 2007b) and sharper revisions in which some black
1999: holes accrete at super-Eddington or highly sub-Eddington rates,
2000: perhaps with reduced radiative efficiencies (Narayan, Mahadevan, \&
2001: Quataert 1998 and references therein).  We will also incorporate
2002: mergers at the rates predicted by theoretical models of cold dark
2003: matter subhalos and their associated black holes (Yoo et al.\ 2007).
2004: For appropriate parameter choices, we expect that many scenarios can
2005: be made consistent with the observed AGN LF and the local black hole
2006: mass function. However, direct measurements of Eddington ratio
2007: distributions and measurements of AGN clustering and host
2008: properties, all as a function of luminosity and redshift, should
2009: greatly narrow the field of viable models.  Within the (often
2010: substantial) uncertainties of existing data, a simple model in which
2011: all black holes grow by accreting gas at mildly sub-Eddington rates
2012: with a radiative efficiency $\epsilon \approx 0.06-0.1$ is
2013: surprisingly successful at reproducing a wide range of observations.
2014: 
2015: %\end{doublespace}
2016: 
2017: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2018: \begin{acknowledgements}
2019: This work was supported by NASA grant GRT000001640 and by
2020: Spanish ministry grants AYA2006-06341 and AYA2006-15623-C02-01.
2021: We thank
2022: Alister Graham and Philip Hopkins for providing data on their
2023: results on the local black hole mass function and John Silverman
2024: for providing data on the X-ray luminosity function.
2025: \end{acknowledgements}
2026: 
2027: 
2028: %;;;;;;;;;;;;;;;;;;;references;;;;;;;;;;;;;;;;;
2029: 
2030: \newpage
2031: \begin{thebibliography}{}
2032: \bibitem{} Alexander, D. M., et al. 2003, AJ, 126, 539
2033: \bibitem{} Antonucci, R. R. 1993, ARA\&A, 31, 473
2034: \bibitem{} Babic, A., Miller, L., Jarvis, M.J., Turner, T.J., Alexander, D.M., \& Croom, S.M. 2007, A\&A,
2035: 474, 755
2036: \bibitem{} Ballantyne, D. R., \& Papovich, C. 2007, ApJ, 660, 988
2037: \bibitem{} Barger, A. J., Cowie, L. L., Capak, P., Alexander, D. M.,
2038: Bauer, F. E., Brandt, W. N., Garmire, G. P. \& Hornschemeier, A. E.
2039: 2003, ApJ, 584, L61
2040: \bibitem{} Barger , A. J., \& Cowie, L. L. 2005, ApJ, 635, 115
2041: \bibitem{} Barger, A. J., et al. 2005, AJ, 129, 578
2042: \bibitem{} Batcheldor, D., Marconi, A., Merritt, D., \& Axon, D. J.
2043: 2007, ApJL, 663, 85
2044: \bibitem{} Bell, E. F., Phleps, S., Somerville, R. S., Wolf, C., Borch, A., \& Meisenheimer, K. 2006, ApJ, 652, 270
2045: %\bibitem{} Begelman, M. 1978, MNRAS, 243, 610
2046: %\bibitem{} Begelman, M. 2002, \apj, 568, L97
2047: \bibitem{} Bentz, M. C., Peterson, B. M., Pogge, R. W., Vestergaard, M., \& Onken, C.
2048: A. 2006, ApJ, 644, 133
2049: \bibitem{} Bernardi, M., Sheth, R. K., Tundo, E., \& Hyde, J. B. 2007, ApJ, 660, 267
2050: %\bibitem{} Blanton, M. R., et al. 2001, AJ, 121, 2358
2051: \bibitem{} Binney, J., \& Merrifield, M. 2000, \emph{Galactic Astronomy},
2052: Princeton Univ. Press
2053: \bibitem{} Bolton, J. S., Haehnelt, M. G., Viel, M., Springel, V.
2054: 2005, MNRAS, 357, 1178
2055: \bibitem{} Bongiorno, A., et al. 2007, A\&A, 472, 443
2056: \bibitem{} Borys, C., Smail, I., Chapman, S. C., Blain, A. W.,
2057:   Alexander, D. M., Ivison, R. J. 2005, ApJ, 635, 853
2058: \bibitem{} Brown, M. J. I., et al. 2006, ApJ, 638, 88
2059: \bibitem{} Bundy, K., et al. ApJ, submitted, arXiv/0710.2105
2060: \bibitem{} Caputi, K. I., McLure, R. J., Dunlop, J. S., Cirasuolo,
2061: M., \& Schael, A. M. 2006, MNRAS, 366, 609
2062: \bibitem{} Cavaliere, A., Morrison, P.,
2063: \& Wood, K. 1971, ApJ, 170, 223
2064: \bibitem{} Cavaliere, A., \& Vittorini, V. 2000, ApJ,
2065: 543, 599
2066: \bibitem{} Churazov, E., et al. 2007, A\&A, 467, 529
2067: \bibitem{} Comastri, A., Setti, G., Zamorani, G., \& Hasinger, G.
2068: 1995, A\&A, 296, 1
2069: \bibitem{} Comastri, A. 2004, Review for
2070: {\it "Supermassive Black Holes in the Distant Universe"}, Ed. A. J.
2071: Barger, Kluwer Academic, astroph/0403693
2072: \bibitem{} Conroy, C., Ho, S., \& White, M. 2007, MNRAS, 379, 1491
2073: \bibitem{} Constantin, A., \& Vogeley, M. S. 2006, 650, 727
2074: \bibitem{} Cool, R. J., Kochanek, C. S., Eisenstein, D. J., Stern, D.,
2075: Brand, K., Brown, M. J. I., Dey, A., Eisenhardt, P. R., Fan, X.,
2076: Gonzalez, A. H., Green, R. F., Jannuzi, B. T., McKenzie, E. H.,
2077: Rieke, G. H., Rieke, M., Soifer, B. T., Spinrad, H., \& Elston, R.
2078: J. 2006, AJ, 132, 823
2079: \bibitem{} Cowie, L. L., et al. 2002, ApJ, 566, L5
2080: \bibitem{} Cowie, L. L., et al. 2003, ApJ, 584, L57
2081: \bibitem{} Croom, S. M., et al. 2004, MNRAS, 349, 1397
2082: \bibitem{} Croton, D. J., Springel, V., White, S. D. M., De Lucia,
2083: G., Frenk, C. S., Gao, L., Jenkins, A., Kauffmann, G., Navarro, J.
2084: F., \& Yoshida, N. 2006, MNRAS, 365, 11
2085: \bibitem{} Croton, D. J. 2006b, MNRAS, 369, 1808
2086: %\bibitem{} Maiolino, R. et al., 1998, \aap, 331, 519
2087: \bibitem{} da Angela, J., et al. 2006, MNRAS, submitted,
2088: astroph/0612401
2089: \bibitem{} Dasyra, K. M., Tacconi, L. J., Davies, R. I., Naab, T., Genzel,
2090: R., Lutz, D., Sturm, E., Baker, A. J., Veilleux, S., Sanders, D. B.,
2091: \& Burkert, A. 2006, ApJ, 651, 835
2092: \bibitem{} Dasyra, K. M., Tacconi, L. J., Davies, R. I., Genzel, R., Lutz,
2093: D., Peterson, B. M., Veilleux, S., Baker, A. J., Schweitzer, M., \&
2094: Sturm, E. 2007, ApJ, 657, 102
2095: \bibitem{} De Zotti, G., Shankar, F., Lapi, A., Granato, G. L.,
2096: Silva, L., Cirasuolo, M., Salucci, P., \& Danese, L. 2006, MmSAIt,
2097: 77, 661
2098: %\bibitem{} Dzanovic, D., Benson, A. J., Frenk, C. S., \& Sharples,
2099: %R. 2006, MNRAS submitted, astroph/0612719
2100: \bibitem{} Elvis, M., Risaliti, G., \& Zamorani, G. 2002, ApJ, 565,
2101: L75
2102: \bibitem{} Fabian, A. C., \& Iwasawa, K. 1999, MNRAS, 303, 34
2103: \bibitem{} Fan, X., et al. 2001, AJ, 121, 54
2104: %\bibitem{} Fan, X., et al. 2001b, AJ, 122, 2833
2105: \bibitem{} Fan, X., et al. 2004, ApJ, 128, 515
2106: \bibitem{} Fardal, M. A., Katz, N., Weinberg, D. H., \& Dav\'e, R. 2007,
2107: MNRAS, 379, 985
2108: \bibitem{} Ferrarese, L., \& Merritt, D. 2000, ApJ, 539, L9
2109: \bibitem{} Ferrarese, L. 2002, Proceedings of the 2nd KIAS Astrophysics Workshop, Seoul, Korea, astroph/0203047
2110: \bibitem{} Fioc, M., \& Rocca-Volmerange, B. 1997, A\&A, 326, 950
2111: \bibitem{} Fiore, F., et al. 2003, A\&A, 409, 79
2112: \bibitem{} Fontana, A., et al. 2006, A\&A, 459, 745
2113: \bibitem{} Fontanot, F., Cristiani, S., Monaco, P., Nonino, M.,
2114: Vanzella, E., Brandt, W. N., Grazian, A., \& Mao, J. 2007, A\&A,
2115: 461, 39
2116: \bibitem{} Fontanot, F., Monaco, P., Cristiani, S., \& Tozzi, P.
2117: 2006, MNRAS, 373, 1173
2118: \bibitem{} Frontera, F., Orlandini, M.,
2119: Landi, R., Comastri, A., Fiore, F., Setti, G., Amati, L., Costa, E.,
2120: Masetti, N., \& Palazzi, E. 2007, ApJ, 666, 86
2121: %\bibitem{} Fukugita, M., Shimasaku, K., \& Ichikawa, T. 1995, PASP, 107, 945
2122: \bibitem{} Fukugita, M., Hogan, C. J., \& Peebles, P. J. E. 1998, ApJ, 503,
2123: 518
2124: \bibitem{} Gammie, C. F., Shapiro, S. L., \& McKinney, J, C. 2004,
2125: ApJ, 602, 312
2126: \bibitem{} Gilli, R., Comastri, A., \& Hasinger, G. 2006, A\&A, 463,
2127: 79
2128: \bibitem{} Graham, A. W., Erwin, P., Caon, N., \& Trujillo, I. 2001, ApJ, 563,
2129: L11
2130: \bibitem{} Graham, A. W., et al. 2007, MNRAS, 378, 198
2131: \bibitem{} Graham, A. W. 2007, MNRAS, 379, 711
2132: \bibitem{} Granato, G. L., De Zotti, G., Silva, L., Bressan, A., \&
2133:  Danese, L. 2004, ApJ, 600, 580
2134: \bibitem{} Granato, G. L., Silva, L., Lapi, A., Shankar, F., De Zotti, G., Danese,
2135: L. 2006, MNRAS, 368L, 72
2136: \bibitem{} Greene, J. E., \& Ho, L. C. 2006, ApJ, 641, 21
2137: \bibitem{} Greene, J. E., \& Ho, L. C. 2007, ApJ, 667, 131
2138: \bibitem{} Greve, T. R., et al. 2005, MNRAS, 359, 1165
2139: \bibitem{} Kennefick, J. D., Djorgovski, S. G., \& de Carvalho, R. R.
2140: 1995, AJ, 110, 2553
2141: \bibitem{} Haiman, Z., Ciotti, L., Ostriker, J. P. 2004, ApJ, 606, 763
2142: \bibitem{} H\"{a}ring, N., \&  Rix, H. W.  2004, ApJ, 604, 89
2143: %\bibitem{} Hao, C. N., Xia, X. Y., Mao, S., Wu, H., \& Deng, Z. G.
2144: %2005, ApJ, 625, 78
2145: \bibitem{} Hasinger, G., et al. 2001, A\&A, 365, 45
2146: \bibitem{} Hasinger, G., Miyaji, T., Schmidt, M. 2005, A\&A, 441,
2147: 417
2148: \bibitem{} Heckman, T. M., Kauffmann, G., Brinchmann, J.,
2149: Charlot, S., Tremonti, C., \& White, S. D. M. 2004, ApJ, 613, 109
2150: \bibitem{} Hickox, R. C. et al.\ 2007, ApJ, in press, astroph/0708.3678
2151: \bibitem{} Hopkins, A. W., \& Beacom, J. F. 2006, 651, 142
2152: \bibitem{} Hopkins, P. F., Hernquist, L., Cox,
2153: T. J., Di Matteo, T., Robertson, B., \& Springel, V. 2006a, ApJS,
2154: 163, 1
2155: \bibitem{} Hopkins, P. F., Robertson, B., Krause, E., Hernquist,
2156: L., \& Cox, T. J. 2006b, ApJ, 652, 107
2157: \bibitem{} Hopkins, P. F., Richards, G. T., Hernquist, L.
2158: 2007a, ApJ, 654, 731 (HRH07)
2159: \bibitem{} Hopkins, P. F., Hernquist, L., Cox, T. J., Robertson, B., \& Krause,
2160: E. 2007b, ApJ, in press, astroph/0701351
2161: \bibitem{} Hopkins, P. F., Lidz, A., Hernquist, L., Coil, A. L., Myers, A. D., Cox, T. J., \& Spergel,
2162: D. 2007c, ApJ, 662, 110
2163: \bibitem{} Hosokawa, T. 2004, ApJ, 606, 139
2164: \bibitem{} Hughes, S. A., \& Blandford, R. D. 2003, ApJ, 585, L10
2165: \bibitem{} Hunt, M. P., Steidel, C. C.,
2166: Adelberger, K. L., \& Shapley, A. E. 2004, ApJ, 605, 625
2167: \bibitem{} Islam, R. R., Taylor, J. E., \& Silk, J. 2004, MNRAS,
2168: 354, 427
2169: \bibitem{} Jiang, L., et al. 2006, AJ, 131, 2788
2170: \bibitem{} Kaspi, S., et al. 2000, ApJ, 533, 631
2171: \bibitem{} Kembhavi, A. K., \& Narlikar, J. V. 1999, in ``Quasars and Active Galactic
2172: Nuclei", Cambridge Univ. Press, Cambridge
2173: \bibitem{} King, A. R., \& Pringle, J. E. 2006, MNRAS, 373, 90
2174: \bibitem{} Kollmeier, J. A., et al. 2006, ApJ, 648, 128
2175: \bibitem{} La Franca, F., et al. 2005, ApJ, 635, 864
2176: \bibitem{} Lapi, A., Shankar, F., Mao, J., Granato, G. L.,
2177: Silva, L., De Zotti, G., \& Danese, L. 2006, ApJ, 650, 42
2178: \bibitem{} Lauer, T. R., Faber, S. M., Richstone, D., Gebhardt, K.,
2179:  Tremaine, S., Postman, M., Dressler, A., Aller, M. C., Filippenko, A. V., Green, R., Ho, L. C., Kormendy, J., Magorrian, J., \& Pinkney,
2180:  J. 2006, ApJ, 670, 249
2181: \bibitem{} Lauer, T. R., et al. 2007a, ApJ, 662, 808
2182: \bibitem{} Lauer, T. R., Tremaine, S., Richstone, D., \& Faber, S.
2183: M., 2007b, ApJ, 670, 249
2184: \bibitem{} Lawrence, J. K. 1991, MNRAS, 252, 586
2185: \bibitem{} Lidz, A., Hopkins, P. F., Cox, T. J., Hernquist, L., \&
2186: Robertson, B. 2006, ApJ, 641, 41
2187: \bibitem{} Mahabal, A., Stern, D., Bogosavljevic, M., Djorgovski,
2188: S.G., \& Thompson, D. 2005, ApJ, 634, L9
2189: \bibitem{} Magorrian, J., et al. 1998, AJ, 115, 2285
2190: \bibitem{} Magdziarz, P., \& Zdziarski, A. A. 1995, MNRAS, 273, 837
2191: \bibitem{} Malbon, R. K., Baugh, C. M., Frenk, C. S., \& Lacey, C.
2192: G. 2006, MNRAS, submitted, astroph/0607424
2193: \bibitem{} Maller, A. H., Katz, N., Kere\v{s}, D., Dav\'{e}, R., \& Weinberg, D. H. 2006, ApJ, 647, 763
2194: \bibitem{} Marconi, A., \& Hunt, L. 2003, ApJL, 589, L21
2195: \bibitem{} Marconi, A., Risaliti, G., Gilli, R., Hunt, L. K.,
2196: Maiolino, R., \& Salvati, M. 2004, MNRAS, 351, 169
2197: \bibitem{} Marshall, F. E., et al. 1980, ApJ, 235, 4
2198: \bibitem{} Martinez-Sansigre, A. et al., 2005, Nature, 436, 666
2199: \bibitem{} Masjedi, M., Hogg, D. W., \& Blanton, M. R. 2008, ApJ,
2200: 679, 260
2201: \bibitem{} Matt, G., Fabian, A. C., Guainazzi, M., Iwasawa, K., Bassani, L., \& Malaguti,
2202: G. 2000, MNRAS, 318, 173
2203: \bibitem{} McLure, R. J., Dunlop., J. S. 2002, MNRAS, 331, 795
2204: \bibitem{} McLure, R. J., Dunlop., J. S. 2004, MNRAS, 352, 1390
2205: \bibitem{} McLure, R. J., Jarvis, M. J.,
2206: Targett, T. A., Dunlop, J. S., \& Best, P. N. 2006, MNRAS, 368, 1359
2207: \bibitem{} Merloni, A. 2004, MNRAS, 353, 1035
2208: \bibitem{} Merloni, A., Rudnick, G., \& Di Matteo, T. 2004, MNRAS,
2209: 354, 37
2210: \bibitem{} Merritt, D., \& Milosavljevi\'{c} M. 2005, LRR, 8, 8
2211: \bibitem{} Miyaji, T., Hasinger, G., \& Schmidt, M. 2000, A\&A, 353, 25
2212: \bibitem{} Myers, A. D., Brunner, R. J., Nichol, R. C., Richards, G. T., Schneider, D. P., \& Bahcall, N.
2213: A. 2007, ApJ, 658, 85
2214: \bibitem{} Miralda-Escud\'{e}, J. 2003, ApJ, 597, 66
2215: \bibitem{} Monaco, P., \& Fontanot, F. 2005, MNRAS, 359, 283
2216: \bibitem{} Murray, N., Quataert, E., Thompson, T. A. 2005, ApJ, 618, 569
2217: \bibitem{} Nakamura, O., Fukugita, M., Yasuda, N., Loveday, J.,
2218: Brinkmann, J., Schneider, D. P., Shimasaku, K., \& SubbaRao, M.
2219: 2003, AJ, 125, 1682
2220: \bibitem{} Nandra, K., Mushotzky, R. F., Arnaud, K., Steidel, C. C.,
2221: Adelberger, K. L., Gardner, J. P., Teplitz, H. I., \& Windhorst, R.
2222: A. 2002, ApJ, 576, 625
2223: \bibitem{} Nandra, K., Laird, E. S., \& Steidel, C. C. 2005, MNRAS,
2224: 360, L39
2225: \bibitem{} Narayan, R., Mahadevan, R., \& Quataert, E. 1998,
2226: in {\it "The Theory of Black Hole Accretion Discs"}, eds. M. A.
2227: Abramowicz, G. Bjornsson, and J. E. Pringle (Cambridge: Cambridge
2228: Univ. Press), p.148, astroph/9803141
2229: \bibitem{} Netzer, H., Trakhtenbrot, B. 2007, ApJ, 654, 754
2230: \bibitem{} Netzer, H., Lira, P., Trakhtenbrot, B., Shemmer, O., \& Cury, I. 2007, ApJ,
2231: 671, 1256
2232: \bibitem{} Osmer, P. S. 2004, in Carnegie Observatories Astrophysics Series,
2233: Vol. 1: Coevolution of Black Holes and Galaxies, ed. L.C. Ho
2234: (Cambridge Univ. Press), in press, astroph/0304150
2235: \bibitem{} Pannella, M., Hopp, U., Saglia, R. P., Bender, R.,
2236: Drory, N., Salvato, M., Gabasch, A., \& Feulner, G. 2006, ApJ, 639,
2237: 1
2238: \bibitem{} Peeples, M. S., \& Martini, P. 2006, ApJ, 652, 1097
2239: \bibitem{} Pei, Y. C. 1995, ApJ, 438, 623
2240: \bibitem{} Peng, C. Y., Impey, C. D., Rix, H. W., Kochanek, C. S.,
2241: Keeton, C. R., Falco, E. E., Leh\'{a}r, J., \& McLeod, B. A. 2006,
2242: ApJ, 649, 616
2243: \bibitem{} Polletta, M., Weedman, D., Hoenig, S., Lonsdale, C. J.,
2244:   Smith, H. E., \& Houck, J. \ 2008, ApJ, 675, 960
2245: \bibitem{} Porciani, C., Magliocchetti, M., \& Norberg, P. 2004
2246: MNRAS, 355, 1010
2247: \bibitem{} Porciani, C., \& Norberg, P. 2006, MNRAS, 371, 1824
2248: \bibitem{} Rees, M. J. 1984, ARA\&A, 22, 471
2249: \bibitem{} Richards, G. T., et al. 2006, AJ, 131, 2766
2250: \bibitem{} Richstone, D., et al. 1998, Nat, 395, A14
2251: \bibitem{} Rigby, J. R., Rieke, G. H., Donley, J. L.,
2252: Alonso-Herrero, A., \& P\'{e}rez-Gonz\'{a}lez, P. G. 2006, ApJ, 645,
2253: 115
2254: \bibitem{} Salucci, P., Szuszkiewicz, E., Monaco, P., \& Danese, L.
2255: 1999, MNRAS, 307, 637
2256: \bibitem{} Salpeter, E. E. 1964, ApJ, 140, 796
2257: \bibitem{} Sanders, D. B., et al. 1988, ApJ, 325, 74
2258: \bibitem{} Schirber, M., \& Bullock, J. S. 2003, ApJ, 584, 110
2259: \bibitem{} Shankar, F., Salucci, P., Granato, G. L., De Zotti, G., \& Danese,
2260: L. 2004, MNRAS, 354, 1020 (S04)
2261: \bibitem{} Shankar, F., Lapi, A., Salucci, P., De Zotti, G., \& Danese,
2262: L. 2006, ApJ, 643, 14
2263: \bibitem{} Shankar, F., \& Mathur, S. 2007, ApJ, 660, 1051
2264: \bibitem{} Shankar, F., \& Ferrarese, L. 2008, ApJ, submitted
2265: \bibitem{} Shankar, F., Bernardi, M., \& Z\'{o}ltan, H. 2008, ApJ,
2266: submitted
2267: \bibitem{} Shapiro, S. L. 2005, ApJ, 620, 59
2268: \bibitem{} Shen, Y., et al. 2007a, AJ, 133, 2222
2269: \bibitem{} Shen, Y., Greene, J. E., Strauss, M., Richards, G. T., \& Schneider, D. P. 2007b, ApJ, submitted, arXiv/0709.3098
2270: \bibitem{} Sheth, R. K., \& Tormen, G. 1999, MNRAS, 308, 119
2271: \bibitem{} Sheth, R. K., Hui, L., Diaferio, A., \& Scoccimarro, R.
2272: 2001, MNRAS, 325, 1288
2273: \bibitem{} Sheth, R. K., et al. 2003, ApJ, 594, 225
2274: \bibitem{} Shields, G. A., Menezes, K. L., Massart, C. A., \& Vanden Bout,
2275: P. 2006, ApJ, 641, 683
2276: \bibitem{} Shinozaki, K., Miyaji, T., Ishisaki, Y., Ueda, Y., \& Ogasaka,
2277: Y. 2006, 131, 2843
2278: \bibitem{} Silverman, J. D., et al. 2008, ApJ, 679, 118
2279: \bibitem{} Small, T. A., \& Blandford, R. D. 1992, MNRAS, 259, 725
2280: \bibitem{} Smith, R. E., Peacock, J. A., Jenkins, A., White, S. D. M.,
2281: Frenk, C. S., Pearce, F. R., Thomas, P. A., Efstathiou, G., \&
2282: Couchman, H. M. P. 2003, MNRAS, 341, 1311
2283: \bibitem{} So{\l}tan, A. 1982, MNRAS, 200, 115
2284: \bibitem{} Steed, A., \& Weinberg, D. H. 2003, astroph/0311312 (SW03)
2285: \bibitem{} Steffen, A. T., Strateva, I., Brandt, W. N., Alexander,
2286: D. M., Koekemoer, A. M., Lehmer, B. D., Schneider, D. P, \& Vignali,
2287: C. 2006, AJ, 131, 2826
2288: \bibitem{} Steidel, C. C., Hunt, M. P., Shapley, A. E.,
2289: Adelberger, K. L., Pettini, M., Dickinson, M., \& Giavalisco, M.
2290: 2002, ApJ, 576, 653
2291: \bibitem{} Stern, D., et al. 2005, ApJ, 631, 163
2292: \bibitem{} Tacconi, L. J., Neri, R., Chapman, S. C., Genzel, R., Smail,
2293: I., Ivison, R. J., Bertoldi, F., Blain, A., Cox, P., Greve, T.,
2294: Omont, A. 2006, 640, 228
2295: \bibitem{} Tamura, N., Ohta, K., \& Ueda, Y. 2006, MNRAS, 365, 134
2296: \bibitem{} Tinker, J. L., Weinger, D. H., Zheng Z., \& Zehavi, I.
2297: 2005, ApJ, 631, 41
2298: \bibitem{} Tonry, J., et al. 2001, ApJ, 546, 681
2299: \bibitem{} Tozzi, P., et al. 2006, A\&A, 451, 457
2300: \bibitem{} Treister, E., et al. 2006, ApJ, 640, 603
2301: \bibitem{} Treu, T., Ellis, R. S., Liao, T. X., van Dokkum, P. G.,
2302: Tozzi, P., Coil, A., Newman, J., Cooper, M. C., \& Davis, Marc 2005,
2303: ApJ, 633, 174
2304: \bibitem{} Tundo, E., Bernardi, M., Hyde, J. B., Sheth, R. K., \& Pizzella,
2305: A. 2007, ApJ, 663, 53
2306: \bibitem{} Ueda, Y., Akiyama, M., Ohta, K., \& Miyaji, T. 2003, ApJ, 598, 886 (U03)
2307: \bibitem{} Vale, A.,\& Ostriker, J. P. 2004, MNRAS, 353, 189
2308: \bibitem{} Vestergaard, M. 2004, ApJ, 601, 676
2309: \bibitem{} Vittorini, V., Shankar, F., \& Cavaliere, A. 2005, MNRAS,
2310: 363, 1376
2311: \bibitem{} Volonteri, M., Madau, P., \& Haardt, F. 2003, ApJ, 593,
2312: 661
2313: \bibitem{} Volonteri, M., Madau, P., Quataert, E., \& Rees, M. J. 2005,
2314: ApJ, 620, 69
2315: \bibitem{} Volonteri, M., Salvaterra, R., \& Haardt, F. 2006, MNRAS,
2316: 373, 121
2317: \bibitem{} Wilman, R. J., \& Fabian, A. C. 1999, MNRAS 309, 862
2318: \bibitem{} Wisotzki, L. 1999, Reviews in Modern Astronomy 12,
2319: Astronomical Instruments and Methods at the turn of the 21st
2320: Century. Edited by Reinhard E. Schielicke. Published by
2321: Astronomische Gesellschaft, Hamburg, p. 231
2322: \bibitem{} Wolf, C., et al. 2003, A\&A, 408, 499
2323: \bibitem{} Wyithe, J. S. B. 2004, MNRAS, 1082, 1098
2324: \bibitem{} Yoo, J., \& Miralda-Escud\'{e}, J. 2004, ApJ, 614, 25
2325: \bibitem{} Yoo, J., Miralda-Escud\'{e}, J., Weinberg, D. H., Zheng, Z., \& Morgan,
2326: C. W. 2007, ApJ, submitted, astroph/0702199
2327: \bibitem{} Yu, Q., \& Lu, Y. 2004, ApJ, 602, 603
2328: \bibitem{} Yu, Q., \& Tremaine, S. 2002, MNRAS, 335, 965
2329: \end{thebibliography}
2330: %;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;;
2331: 
2332: \newpage
2333: 
2334: %%%%%%%%TABLES%%%%%%%%%%%%%%%%%%%%%
2335: 
2336: \begin{deluxetable}{cccccc}
2337: \tablecolumns{6} \tablewidth{0pc} \tablecaption{AGN Luminosity
2338: Function parameters} \tablehead{\colhead{Parameter} &  &  &  & & }
2339: \startdata
2340: $p_1$         &  $4.23$ ($z<4$) & $-0.615z+6.690$ & $3.0$\, ($z=6$)   &  &   \\
2341: $p_2$         &  $-1.5$ at all $z$&   &  & & \\
2342: $\gamma_1$     &  $0.86$ at all $z$&    &  & & \\
2343: $\gamma_2$     &  $2.6$\, ($z<0.5$) & $-0.933z+3.067$ & $2.32$\, ($0.8<z<4$) & $0.24z+1.36$ & $2.8$\, ($z=6$)  \\
2344: $A$           &  $5.04\times 10^{-6}\, {\rm Mpc^{-3}}$    &   &   & & \\
2345: $z_c^*$       &  $1.9$ &     &  & & \\
2346: $\log L_a$    &  $44.6$\, ($z<3$) & 0.0667$z$+44.4 & 44.8\, ($z=6$)    &   &   \\
2347: $\log L_*$    &  $43.94$&     &  & & \\
2348: \enddata
2349: \tablecomments{ \, List of the parameters entering the AGN
2350: luminosity function described in \S~\ref{sec|AGNLF}. Some of the
2351: parameters assume different values in different redshift bins as
2352: quoted. A smooth linear transition in redshift $z$ in the values has
2353: been applied between discontinuous redshift bins. Luminosities are
2354: the inferred values in the $2-10$ keV band, in units of ${\rm erg\,
2355: s^{-1}}$.} \label{Table|LF}
2356: \end{deluxetable}
2357: 
2358: \newpage
2359: 
2360: \begin{deluxetable}{cccccc}
2361: \tablecolumns{6} \tablewidth{0pc} \tablecaption{AGN BOLOMETRIC
2362: LUMINOSITY FUNCTIONS}
2363: \tablehead{$z$ & $\log L$ & $\log \Phi$  & $\log \Phi_{\rm no CT}$ & $\log \Phi_{\rm Double CT}$ & $\log \Phi_{\rm HRH07}$\\
2364: (1) & (2) & (3) & (4)  & (5) & (6)} \startdata
2365: 0.020 & 41.00 & -1.859 & -1.994 & -1.757 & -1.374\\
2366: 0.020 & 41.25 & -2.074 & -2.208 & -1.972 & -1.578\\
2367: ..........\\
2368: 0.020 & 48.00 & -10.08 & -10.16 & -10.01 & -9.539\\
2369: 0.260 & 41.00 & -1.903 & -2.036 & -1.796 & -1.609\\
2370: 0.260 & 41.25 & -2.089 & -2.222 & -1.984 & -1.787\\
2371: 0.260 & 41.50 & -2.267 & -2.399 & -2.162 & -1.967\\
2372: ..........\\
2373: \enddata
2374: \tablecomments{\,(1) redshift; (2) bolometric luminosity in
2375: logarithmic scale in units of ${\rm erg\, s^{-1}}$; (3) reference
2376: bolometric AGN luminosity function in logarithmic scale in units of
2377: ${\rm Mpc^{-3}\, dex^{-1}}$; (4) reference bolometric AGN luminosity
2378: function with no Compton-thick sources (i. e., with $\log N_H/{\rm
2379: cm^{-2}}>24$); (5) reference bolometric AGN luminosity function with
2380: Double contribution from Compton-thick sources (i. e., with $\log
2381: N_H/{\rm cm^{-2}}\le 26$); (6) bolometric AGN luminosity function
2382: from HRH07. The full table is available in electronic form.}
2383: \label{Table|FitsLF}
2384: \end{deluxetable}
2385: 
2386: \newpage
2387: 
2388: \begin{deluxetable}{ccccccc}
2389: \tablecolumns{7} \tablewidth{0pc} \tablecaption{BLACK HOLE MASS
2390: FUNCTIONS} \tablehead{$z$ & $\log M_{\bullet}$ & $\log \Phi$  &
2391: $\log \Phi_{\rm no CT}$ &
2392: $\log \Phi_{\rm Double CT}$ & $\log \Phi_{\rm HRH07}$ & $\log \Phi_{\rm HRH07_{\rm EXTRA}}$\\
2393: (1) & (2) & (3) & (4)  & (5) & (6) & (7)} \startdata
2394: 0.02 & 5.0 & -1.264 & -1.398 & -1.162 & -0.928 & -0.952\\
2395: 0.02 & 5.2 & -1.394 & -1.528 & -1.292 & -1.060 & -1.082\\
2396: ..........\\
2397: 0.02 & 9.6 & -4.683 & -4.779 & -4.604 & -4.436 & -4.964\\
2398: 0.26 & 5.0 & -1.482 & -1.614 & -1.376 & -1.129 & -1.131\\
2399: 0.26 & 5.2 & -1.608 & -1.739 & -1.501 & -1.247 & -1.248\\
2400: 0.26 & 5.4 & -1.723 & -1.855 & -1.616 & -1.365 & -1.364\\
2401: ..........\\
2402: \enddata
2403: \tablecomments{\,(1) redshift; (2) black hole mass in logarithmic
2404: scale in units of ${\rm M_{\odot}}$; (3) black hole mass function in
2405: our reference model in logarithmic scale in units of ${\rm
2406: Mpc^{-3}\, dex^{-1}}$; (4) reference black hole mass function in the
2407: same model but with no Compton-thick sources; (5) reference black
2408: hole mass function in the same model but with Double contribution
2409: from Compton-thick sources; (6) black hole mass function in the same
2410: model but adopting the HRH07 luminosity function; (7) black hole
2411: mass function obtained adopting the HRH07 luminosity function with
2412: $\epsilon=0.09$ and $\dot{m}=1$, as in Figure~\ref{fig|Models}d. The
2413: full table is available in electronic form.} \label{Table|Fits}
2414: \end{deluxetable}
2415: %\end{landscape}
2416: %
2417: \begin{deluxetable}{ccccccc}
2418: \tablecolumns{7} \tablewidth{0pc} \tablecaption{BLACK HOLE DUTY
2419: CYCLES} \tablehead{$z$ & $\log M_{\bullet}$ & $P_0$  & $P_0^{\rm no
2420: CT}$ &
2421: $P_0^{\rm Double CT}$ & $P_0^{\rm HRH07}$ & $P_0^{\rm HRH07_{\rm EXTRA}}$\\
2422: (1) & (2) & (3) & (4)  & (5) & (6) & (7)} \startdata
2423: 0.02 &   5.0 &   0.00995 &  0.00995 &   0.00995 &  0.01401  &  0.00727\\
2424: 0.02 &   5.2 &   0.00941 &  0.00942 &   0.00942 &  0.01299  &  0.00668\\
2425: ..........\\
2426: 0.02 &   9.6 &   0.00008 &  0.00008 &   0.00008 &  0.00011  &  0.00009\\
2427: 
2428: 0.26 &   5.0 &   0.03063 &  0.03057 &   0.03050 &  0.01947  &  0.01050\\
2429: 0.26 &   5.2 &   0.03246 &  0.03231 &   0.03215 &  0.01837  &  0.00987\\
2430: 0.26 &   5.4 &   0.03273 &  0.03243 &   0.03212 &  0.01729  &  0.00924\\
2431: ..........\\
2432: \enddata
2433: \tablecomments{\,(1) redshift; (2) black hole mass in logarithmic
2434: scale in units of ${\rm M_{\odot}}$; (3) duty cycle for our
2435: reference model; (4) duty cycle for the reference model with no
2436: Compton-thick sources; (5) duty cycle for the reference model with
2437: Double contribution from Compton-thick sources; (6) duty cycle for
2438: the model obtained adopting the HRH07 luminosity function; (7) duty
2439: cycle for the model obtained adopting the HRH07 luminosity function
2440: with $\epsilon=0.09$ and $\dot{m}=1$. The full table is available in
2441: electronic form.} \label{Table|DutyCycles}
2442: \end{deluxetable}
2443: 
2444: 
2445: %%%%%%%%FIGURES%%%%%%%%%%%%%%%%%
2446: %
2447: \begin{figure}
2448: \includegraphics[angle=00,scale=0.95]{fig1.eps}
2449: \caption{The bolometric AGN luminosity function. Curves show the
2450: model described in \S~\ref{sec|AGNLF}. The \emph{solid} line is the
2451: total LF including very Compton-thick sources, the
2452: \emph{short-dashed} line includes sources with column densities up
2453: to $\log N_H/{\rm cm^{-2}}=24$, while the \emph{dot-dashed} line
2454: includes only sources with $\log N_H/{\rm cm^{-2}} \le 22$. The data
2455: from optical surveys are from Hunt et al. (2004), Pei (1995),
2456: Wisotzki (1999), Jiang et al. (2006), Cool et al. (2006), Shankar \&
2457: Mathur (2007; derived from data of Stern et al. 2000; Willott et al.
2458: 2004; Mahabal et al. 2005), VIMOS-VLT Deep Survey (Bongiorno et al.
2459: 2007), 2SLAQ (Richards et al. 2005), SDSS (Richards et al. 2006; Fan
2460: et al. 2001, 2004), COMBO-17 (Wolf et al. 2003), Kennefick et al.
2461: (1994). The data from X-ray surveys are from Ueda et al. (2003),
2462: Barger et al. (2003), Barger et al. (2005), Barger \& Cowie (2005),
2463: Nandra et al. (2005) and Silverman et al. (2008). GOODS
2464: (multi-wavelength) data are from Fontanot et al. (2007). }
2465: \label{fig|BolLF}
2466: \end{figure}
2467: %
2468: \begin{figure}
2469: \includegraphics[angle=00,scale=0.95]{fig2.eps}
2470: \caption{Same as Figure~\ref{fig|BolLF} but now all the data have
2471: been corrected for the obscured fraction as given in HRH07.} \label{fig|BolLFcorrected}
2472: \end{figure}
2473: %
2474: \newpage
2475: \begin{figure}
2476: \includegraphics[angle=00,scale=0.9]{fig3.eps}
2477: \caption{The X-ray background. Curves represent the integration of
2478: our AGN bolometric LF converted to hard X-ray bands, as described in
2479: the text. The \emph{dashed} line includes sources with column
2480: densities up to $\log N_H/{\rm cm^{-2}}\le 22$, the
2481: \emph{dot-dashed} line those up to $\log N_H/{\rm cm^{-2}} \le 23$,
2482: and the \emph{triple-dot dashed} line those with $\log N_H/{\rm
2483: cm^{-2}}\le 24$. The \emph{solid} line refers to the total when
2484: including Compton-thick sources up to $\log N_H/{\rm cm^{-2}}\le
2485: 25$, with a dependence on luminosity following that in U03. Data are
2486: taken from a compilation in Frontera et al. (2007).}
2487: \label{fig|XRBGs}
2488: \end{figure}
2489: %
2490: \begin{figure}
2491: \includegraphics[angle=00,scale=1.2]{fig4.eps}
2492: \caption{Comparison between the bolometric AGN luminosity function
2493: described in \S~\ref{sec|AGNLF} (\emph{solid} line) and the
2494: recently derived estimate by HRH07 (\emph{dashed}
2495: line), shown with its 1-$\sigma$ uncertainty in the bolometric
2496: correction (\emph{gray area}). Curves are vertically offset by $-1.5\times z$.} \label{fig|HopkinsLF}
2497: \end{figure}
2498: %
2499: %
2500: \begin{figure}
2501: \includegraphics[angle=00,scale=1.]{fig5.eps}
2502: \caption{Comparison among estimates of the local black hole
2503: mass function. Lines show estimates using various calibrations of the $M_{\bullet}-L_{\rm sph}$,
2504: $M_{\bullet}-\sigma$, or $M_{\bullet}-M_{\rm star}$ relations as described in the text,
2505: assuming a 0.3-dex intrinsic scatter in all cases. The grey band
2506: encompasses the range of these estimates. Filled small circles show the
2507: determination of Hopkins et al. (2007b) using the black hole
2508: ``fundamental plane'' and open circles show the determination of Graham et al. (2007)
2509: using the relation between black hole mass and S\'{e}rsic index.} \label{fig|LocalMFs}
2510: \end{figure}
2511: %
2512: \begin{figure}
2513: \includegraphics[angle=00,scale=1.1]{fig6.eps}
2514: \caption{Black hole growth and stellar mass growth. (\emph{a}) Average black hole accretion rate as
2515: computed via equation~(\ref{eq|soltan}) compared to the SFR as given
2516: by Hopkins \& Beacom (2006) and Fardal et al. (2007),
2517: scaled by the factor $M_{\bullet}/M_{\rm STAR}=0.5\times
2518: 1.6\times 10^{-3}$. The grey area shows the 3-$\sigma$ uncertainty region
2519: from Hopkins \& Beacom (2006). (\emph{b})
2520: Cumulative black hole mass density as a function of redshift. The solid line is the prediction
2521: based on the bolometric AGN luminosity function. The
2522: \emph{light-gray area} is the local value of the black hole mass
2523: density with its systematic uncertainty as given in
2524: Figure~\ref{fig|LocalMFs}. The \emph{dark squares} are estimates of
2525: the black hole mass function at $z=1$ and $z=2$ obtained from the
2526: stellar mass function of Caputi et al. (2006) and Fontana et al.
2527: (2006), scaled by the local ratio $M_{\bullet}/M_{\rm STAR}=1.6\times
2528: 10^{-3}$. The lines are the integrated stellar mass densities based on the SFR histories
2529: in panel (\emph{a}), scaled by $8\times 10^{-4}$.} \label{fig|SFR-Rhoz}
2530: \end{figure}
2531: %
2532: %
2533: \begin{figure}
2534: \includegraphics[angle=00,scale=1.17]{fig7.eps}
2535: \caption{Results for a reference model that agrees well with the
2536: local black hole mass function, with $\epsilon=0.065$, $\dot{m}_0=0.60$,
2537: and an initial duty cycle $P_0(z=6)=0.5$. (\emph{a}) The accreted mass function shown at different
2538: redshifts as labeled, compared with the local mass function
2539: (\emph{grey area}). (\emph{b}) Similar to (\emph{a}) but plotting
2540: $M_{\bullet}\Phi(M_{\bullet})$ instead of $\Phi(M_{\bullet})$.
2541: (\emph{c}) Duty-cycle as a function of black hole mass and redshift as labeled.
2542: (\emph{d}) Average accretion histories for black holes of
2543: different relic mass $M_{\bullet}$ at $z=0$ as a function of
2544: redshift; the \emph{dashed} lines represent the curves when the
2545: black hole has a luminosity below $L_{\rm MIN}(z)$ and therefore
2546: does not grow in mass; the minimum black hole mass corresponding to
2547: $L_{\rm MIN}(z)$ at different redshifts is plotted with a
2548: \emph{dot-dashed} line.} \label{fig|bestFitModel}
2549: \end{figure}
2550: %
2551: %
2552: \begin{figure}
2553: \includegraphics[angle=00,scale=1.]{fig8.eps}
2554: \caption{Effect of model inputs on the predicted black hole
2555: mass function at $z=0$. (\emph{a}) Varying $\epsilon$ at fixed $\dot{m}_0=0.6$.
2556: (\emph{b}) Varying $\dot{m}_0=$ at fixed $\epsilon=0.065$. (\emph{c})
2557: \emph{Solid} and \emph{dotted} curves show the effect of removing or doubling the fraction of Compton-thick
2558: sources, keeping $\epsilon=0.065$ and $\dot{m}_0=0.6$ fixed at their reference values.
2559: \emph{Dashed} and \emph{dot-dashed} curves show these modified models with parameter values chosen to reproduce
2560: the local mass function. (\emph{d}) The {\it solid}
2561: curve shows the effect of adopting the
2562: HRH07 LF, with $\epsilon=0.065$ and $\dot{m}_0=0.6$. The {\it dashed}
2563: curve shows a model
2564: with the HRH07 LF and parameter values $\epsilon=0.09$ and $\dot{m}_0=1$.} \label{fig|Models}
2565: \end{figure}
2566: %
2567: \begin{figure}
2568: \includegraphics[angle=00,scale=1.]{fig9.eps}
2569: \caption{Models incorporating our standard luminosity function estimate compared to
2570: the properties of the local black hole mass function described in \S~\ref{subsec|models}: integrated
2571: mass density ($\rho_{\bullet}$, \emph{upper left}), peak mass ($M_{\rm PEAK}$,
2572: \emph{upper right}), width ($M_{\bullet, 1/2}$,
2573: \emph{lower left}), and asymmetry ($\Delta
2574: \rho_{\bullet}/\rho_{\bullet}$, \emph{lower right}). The \emph{thick}
2575: lines show model predictions as a function of radiative efficiency $\epsilon$, for several values
2576: of $\dot{m}_0$ as labeled. The horizontal \emph{grey band} shows observational estimates of the four
2577: quantities based on the grey band in Figure~\ref{fig|LocalMFs}. Horizontal \emph{long-dashed} and \emph{short-dashed}
2578: lines show the values derived from the local mass functions of Hopkins et al. (2007b) and
2579: Graham et al. (2007), respectively.} \label{fig|multiparam}
2580: \end{figure}
2581: %
2582: \begin{figure}
2583: \includegraphics[angle=00,scale=1.]{fig10.eps}
2584: \caption{Local black hole mass function parameters predicted for different LF inputs.
2585: The accretion rate is
2586: fixed to $\dot{m}_0=0.75$ in all cases. The \emph{thick} lines refer to models
2587: with our standard LF estimate (\emph{solid} line), with no contribution of Compton-thick sources (\emph{dotted} line),
2588: with only Type I AGNs included (\emph{dashed} line), with Type I
2589: AGNs multiplied by a constant factor of 3 at all redshifts and all
2590: luminosities (\emph{dot-dashed} line), and with the HRH07 LF (\emph{long-dashed} line).
2591: Grey band and horizontal thin lines are as
2592: in Figure~\ref{fig|multiparam}.} \label{fig|multiparamLF}
2593: \end{figure}
2594: %
2595: \begin{figure}
2596: \includegraphics[angle=00,scale=1.17]{fig11.eps}
2597: \caption{Model with $\epsilon=0.065$ and decreasing $\dot{m}_0(z)$ as given in
2598: equation~(\ref{eq|mdotz}), compared to our reference model with constant $\dot{m}_0=0.6$
2599: and $\epsilon=0.065$.
2600: Panels (a), (b), (c) show, respectively, the evolution of the mass function, the evolution
2601: of the duty cycle, and mass growth along characteristics, in the same format as Figure~\ref{fig|bestFitModel}.
2602: Reference model results are shown by \emph{squares} in (a), by
2603: \emph{thinner lines} in (b), and by \emph{long-dashed lines} in (c).}
2604: \label{fig|etamdot}
2605: \end{figure}
2606: %
2607: \begin{figure}
2608: \includegraphics[angle=00,scale=1.3]{fig12.eps}
2609: \caption{Model with $\epsilon=0.075$ and a mass-dependent Eddington
2610: ratio $\dot{m}_0=0.445(M_{\bullet}/10^9\, {\rm M_{\odot}})^{0.5}$
2611: (equation~\ref{eq|EddMbh}), compared to our reference model. The
2612: format is similar to panels (a) and (b) of
2613: Figure~\ref{fig|etamdot}.} \label{fig|dmdtMbh}
2614: \end{figure}
2615: %
2616: \begin{figure}
2617: \includegraphics[angle=00,scale=1.]{fig13.eps}
2618: \caption{Evolution of the black hole mass function in a model with black
2619: hole mergers. Accretion-driven growth is computed assuming $\epsilon=0.065$,
2620: $\dot{m}_0=0.60$, and each black hole has a probability $P_{\rm merg}=0.5$
2621: of merging with another black hole of equal mass per Hubble time $t_H(z)$.
2622: Squares show the $z=0$ predictions of the reference model (same accretion
2623: parameters, no merging), and the long-dashed line shows the $z=0$ mass function
2624: for a merger model with the HRH07 LF and accretion parameters $\epsilon=0.09$,
2625: $\dot{m}_0=1$.} \label{fig|merging}
2626: \end{figure}
2627: %
2628: \begin{figure}
2629: \includegraphics[angle=00,scale=1.3]{fig14.eps}
2630: \caption{Panel \emph{a}: cumulative space density of predicted hosts (including those
2631: with inactive black holes of the same  mass) that have a luminosity
2632: higher than $L$ for four different redshifts. Panel \emph{b}:
2633: predicted bias as a function of luminosity and redshift,
2634: computed according to equations~(\ref{eq|nhosts}) and
2635: ~(\ref{eq|bias}); the data are from da Angela et al. (2006; \emph{filled triangles}),
2636: Myers et al. (2007; \emph{filled squares}), Porciani \& Norberg (2007; \emph{filled} and \emph{open circles}), and
2637: Shen et al. (2007a; \emph{open diamonds}), all corrected to $\sigma_8=0.8$. Both panels refer to the predictions of our reference model.}
2638: \label{fig|hosts}
2639: \end{figure}
2640: %
2641: %\begin{figure}
2642: %\includegraphics[angle=00,scale=1.3]{fig12.eps}
2643: %\caption{Cumulative total number of predicted hosts (inclusive of
2644: %the inactive ones with same black hole mass) which have a luminosity
2645: %higher than $L$ for four different models; in each panel the four
2646: %set of different lines correspond to different redshift bins as
2647: %labeled. Panel \emph{a}: our reference model (\emph{thin} lines) is
2648: %compared with a model with the same Eddington limit but different
2649: %radiative efficiency, $\epsilon=0.3$ (\emph{thick} lines); we find
2650: %no difference in the number of hosts, this is because even if the
2651: %number of hosts should be higher in our reference model, this effect
2652: %is compensated by a lower luminosity threshold for a given black
2653: %hole mass, as predicted by our definition of Eddington limit given
2654: %in equation~(\ref{eq|Medd}). Panel \emph{b}: here we run two models
2655: %with different radiative efficiency as labeled but using the
2656: %classical definition of Eddington limit (see
2657: %equation~(\ref{eq|MeddCla})); the number of hosts in the model with
2658: %higher radiative efficiency (\emph{thick} lines) is now lower. Panel
2659: %\emph{c}: here we fix the radiative efficiency to our reference
2660: %value of $\epsilon=0.065$ and vary the Eddington ratio; the number
2661: %of hosts above a given luminosity $L$ increases in the model with
2662: %lower $\dot{m}_0$, as more massive, less numerous black holes are
2663: %associated to the same luminosity, in agreement with
2664: %equation~(\ref{eq|Lbol}). Panel \emph{d}: we compare the reference
2665: %model with the model discussed in \S~\ref{subsec|etamdot}, in
2666: %which we have considered a decreasing $\dot{m}(z)$; it can be seen
2667: %that at high luminosities at low redshifts the $\dot{m}(z)$ model,
2668: %predicts more hosts at high luminosities, as the black hole mass
2669: %function is higher at the high-mass end with respect to our
2670: %reference model (see Figure~\ref{fig|etamdot}\emph{a}). In all
2671: %outputs shown in this Figure the outputs have been obtained by
2672: %extrapolating the luminosity function below $L_{\rm min}$; this is
2673: %done to get a full comparison at all luminosities.}
2674: %\label{fig|hosts}
2675: %\end{figure}
2676: %
2677: %\begin{figure}
2678: %includegraphics[angle=00,scale=1.3]{figBias.eps}
2679: %\caption{Predicted bias as a function of luminosity and redshift,
2680: %computed according to equations~(\ref{eq|bias}) and
2681: %~(\ref{eq|mhalo}); the four panels represent the predicted bias
2682: %corresponding to the models shown in the four panels of
2683: %Figure~\ref{fig|hosts}; all the lines correspond to the same
2684: %redshift bins as in Figure~\ref{fig|hosts}. It can be seen the
2685: %general trend that an higher number density in
2686: %Figure~\ref{fig|hosts}, corresponds to a lower effective halo mass
2687: %and therefore a lower bias in this Figure, and viceversa.}
2688: %\label{fig|bias}
2689: %\end{figure}
2690: %
2691: %
2692: \end{document}
2693: