1:
2: \documentclass{emulateapj}
3: \usepackage{apjfonts}
4: \usepackage{graphicx}
5: \usepackage{amsmath}
6:
7: \graphicspath{{./}}
8:
9: \bibliographystyle{apj}
10:
11: \newcommand{\Msun}{\rm M_{\odot}}
12: \newcommand{\MB}{M_{\rm B}}
13: \newcommand{\Lbsun}{\rm L_{\odot, \rm B}}
14: \newcommand{\Lsun}{\rm L_{\odot}}
15: \newcommand{\Lbol}{\rm L_{\rm {bol}}}
16: \newcommand{\LB}{\rm L_{\rm B}}
17: \newcommand{\Lb}{\rm L_{\rm B}}
18: \newcommand{\Lfir}{\rm L_{\rm {FIR}}}
19: \newcommand{\Lir}{\rm L_{\rm {IR}}}
20: \newcommand{\Lnir}{\rm L_{\rm {NIR}}}
21: \newcommand{\Lmir}{\rm L_{\rm {MIR}}}
22: \newcommand{\Lradio}{\rm L_{\rm {radio}}}
23: \newcommand{\Lx}{\rm L_{\rm {x-ray}}}
24:
25: \newcommand{\Ledd}{\rm L_{\rm {Edd}}}
26: \newcommand{\BH}{\rm{BH}}
27: \newcommand{\MBH}{\rm M_{\rm{BH}}}
28: \newcommand{\Mbulge}{\rm M_{\rm{bulge}}}
29: \newcommand{\Mstar}{\rm M_{\rm{star}}}
30: \newcommand{\msigma}{\MBH\rm{-}\sigma}
31: \newcommand{\mbulge}{\rm{\MBH\rm{-}\Mbulge}}
32: \newcommand{\zquasar}{J1148+5251}
33: \newcommand{\mdm}{\rm m_{\rm{dm}}}
34: \newcommand{\tQ}{t_{Q}}
35: \newcommand{\Gpc}{\rm {Gpc}}
36: \newcommand{\Mpc}{\rm {Mpc}}
37: \newcommand{\kpc}{\rm {kpc}}
38: \newcommand{\yr}{\rm {yr}}
39:
40: \newcommand{\vir}{\rm{vir}}
41: \newcommand{\Mvir}{M_{\rm{vir}}}
42: \newcommand{\Vvir}{V_{\rm{vir}}}
43: \newcommand{\Rvir}{R_{\rm{vir}}}
44: \newcommand{\Cvir}{C_{\rm{vir}}}
45: \newcommand{\rhoc}{\rho_{\rm{crit}}}
46:
47: \def\Omm{{\Omega_m}}
48: \def\Ommz{{\Omega_m^{\,z}}}
49: \def\Omr{{\Omega_r}}
50: \def\Omk{{\Omega_k}}
51: \def\Oml{{\Omega_{\Lambda}}}
52:
53: \def\beq{\begin{equation}}
54: \def\eeq{\end{equation}}
55:
56: \shorttitle{Modeling Dust in $z \sim 6$ Quasars}
57: \shortauthors{Li et al.}
58:
59: \begin{document}
60:
61: \title{Modeling the Dust Properties of $z \sim 6$ Quasars with ART$^2$ ---
62: All-wavelength Radiative Transfer with Adaptive Refinement Tree}
63:
64: \author
65: {
66: Yuexing Li\altaffilmark{1},
67: Philip F. Hopkins\altaffilmark{1},
68: Lars Hernquist\altaffilmark{1},
69: Douglas P. Finkbeiner\altaffilmark{1},
70: Thomas J. Cox\altaffilmark{1}, \\
71: Volker Springel\altaffilmark{2},
72: Linhua Jiang\altaffilmark{3},
73: Xiaohui Fan\altaffilmark{3},
74: Naoki Yoshida\altaffilmark{4}
75: }
76:
77: \affil{$^{1}$Harvard-Smithsonian Center for Astrophysics, Harvard
78: University, 60 Garden Street, Cambridge, MA 02138, USA}
79: %
80: \affil{$^{2}$Max-Planck-Institute for Astrophysics,
81: Karl-Schwarzschild-Str. 1, 85740 Garching, Germany}
82: %
83: \affil{$^{3}$Department of Astronomy, University of Arizona, 933 N. Cherry
84: Ave, Tucson, AZ 85721}
85: %
86: \affil{$^{4}$Nagoya University, Dept. of Physics, Nagoya, Aichi
87: 464-8602, Japan}
88: %
89: %
90: \email{yxli@cfa.harvard.edu}
91: %
92: \keywords{quasar: formation --- quasar: evolution --- quasar: high redshift
93: --- galaxies: starbursts --- infrared: galaxies --- radiative transfer ---
94: interstellar medium --- dust, extinction --- individual: SDSS J1148+5251}
95:
96: \begin{abstract}
97:
98: The detection of large quantities of dust in $z \sim 6$ quasars by
99: infrared and radio surveys presents puzzles for the formation and
100: evolution of dust in these early systems. Previously \citep{Li2007},
101: we showed that luminous quasars at $z \gtrsim 6$ can form through
102: hierarchical mergers of gas-rich galaxies, and that these systems are
103: expected to evolve from starburst through quasar phases. Here, we
104: calculate the dust properties of simulated quasars and their
105: progenitors using a three-dimensional Monte Carlo radiative transfer
106: code, ART$^2$ -- All-wavelength Radiative Transfer with Adaptive
107: Refinement Tree. ART$^2$ incorporates the radiative equilibrium
108: algorithm developed by \cite{Bjorkman2001} which treats dust emission
109: self-consistently, an adaptive grid method which can efficiently cover
110: a large dynamic range in spatial scales and can capture inhomogeneous
111: density distributions, a multiphase model of the interstellar medium
112: which accounts for the observed scaling relations of molecular clouds,
113: and a supernova-origin model for dust which can explain the existence
114: of dust in cosmologically young, starbursting quasars. By applying
115: ART$^2$ to the hydrodynamic simulations of \cite{Li2007}, we reproduce
116: the observed spectral energy distribution (SED) and inferred dust
117: properties of SDSS J1148+5251, the most distant quasar detected in the Sloan
118: survey. We find that the dust and infrared emission are closely associated
119: with the formation and evolution of the quasar host. As the system evolves
120: from a starburst to a quasar, the SED changes from being dominated by a
121: cold dust bump (peaking at $\sim 50\, \mu$m) to one that includes a
122: prominent hot dust component (peaking at $\sim 3\, \mu$m), and the
123: galaxy evolves from a cold to a warm ultraluminous infrared galaxy
124: (ULIRG) owing to heating and feedback from star formation and the
125: active galactic nucleus (AGN). Furthermore, the AGN activity has
126: significant implications for the interpretation of observable aspects
127: of the hosts. The hottest dust ($T \gtrsim 10^3$~K) is most
128: noticeable only during the peak quasar activity, and correlates with
129: the near-IR flux. However, we find no correlation between the star
130: formation rate and far-IR luminosity during this phase owing to strong
131: AGN contamination. Our results suggest that vigorous star formation in
132: merging progenitors is necessary to reproduce the observed dust
133: properties of $z \sim 6$ quasars, supporting a merger-driven origin
134: for luminous quasars at high redshifts and the starburst-to-quasar
135: evolutionary hypothesis.
136:
137: \end{abstract}
138:
139: \section{INTRODUCTION}
140:
141: High-redshift quasars are important for understanding the formation
142: and evolution of galaxies and supermassive black holes (SMBHs) in the
143: early Universe. In the past few years, nearly two dozen luminous
144: quasars have been discovered by the Sloan Digital Sky Survey (SDSS;
145: \citealt{York2000}) and the Canada-France High-z Quasar Survey
146: (CFHQS; \citealt{Willott2007}) at $z \sim 6$,
147: corresponding to a time when the Universe was less than a billion years old
148: (\citealt{Fan2003, Fan2004, Fan2006B, Fan2006A}). As summarized by
149: \cite{Fan2006C}, these quasars are rare ($\sim 10^{-9}\, \Mpc^{-3}$ comoving);
150: believed to be powered by SMBHs with masses $\sim 10^9\, \Msun$ (e.g.,
151: \citealt{Willott2003, Barth2003}); reside near the end of the epoch of
152: reionization, as indicated by Gunn-Peterson absorption troughs
153: \citep{Gunn1965} in their spectra; and have similar spectral energy
154: distributions (SEDs) and comparable metallicity to lower-redshift counterparts
155: (e.g., \citealt{Elvis1994, Glikman2006, Richards2006, Hopkins2007A}),
156: implying the early presence of ``mature'' quasars and the formation of
157: their hosts in rapid starbursts at even higher redshifts ($z \gtrsim
158: 10$).
159:
160: Follow-up, multi-wavelength observations have been carried out for
161: these $z \sim 6$ quasars, from X-ray (e.g., \citealt{Brandt2002,
162: Strateva2005, Vignali2005, Steffen2006, Shemmer2005, Shemmer2006}), to
163: optical /infrared (e.g., \citealt{Barth2003, Pentericci2003,
164: Freudling2003, White2005, Willott2005, Hines2006}), and radio
165: wavelengths (e.g., \citealt{Carilli2001, Bertoldi2003B, Walter2003,
166: Carilli2004, Wang2007}). A noteworthy result of these studies is their
167: detection of dust in these high-redshift objects. In particular, deep
168: infrared and radio surveys (e.g., \citealt{Robson2004, Bertoldi2003A,
169: Carilli2004, Charmandaris2004, Beelen2006, Hines2006}) have revealed a
170: large amount of cold dust in SDSS J1148+5251 (hereafter J1148+5251),
171: the most distant Sloan quasar detected at redshift $z = 6.42$
172: \citep{Fan2003}. The dust mass is estimated to be $\sim 1-7\times 10^8\,
173: \Msun$. The detection of dust is also reported in the first four CFHQS quasars
174: at $z > 6$, including the new record holder CFHQS J2329-0301 at $z = 6.43$
175: \citep{Willott2007}. \cite{Maiolino2004} argue, from the observed dust
176: extinction curve of SDSS J1048+4637 at $z = 6.2$, that the dust in these
177: high-z systems may be produced by supernovae.
178:
179: \cite{Jiang2006} have performed a comprehensive study of thirteen $z
180: \sim 6$ quasars by combining new {\em Spitzer} observations with
181: existing multi-band data. It appears that nearly all 13 of these
182: quasars exhibit prominent infrared bumps around both $\lambda \sim 3
183: \mu$m and $\lambda \sim 50 \mu$m. In a dusty galaxy, most of the
184: radiation from newly formed stars or an AGN is absorbed by dust and
185: re-emitted at infrared wavelengths. The SEDs of star-forming galaxies
186: typically peak at 50 -- 80 $\mu$m (rest frame) and can be approximated
187: by a modified blackbody spectrum with dust temperatures below 100 K
188: (e.g., \citealt{Dunne2001}), while in quasars, some dust can be
189: heated directly by the AGN to temperatures up to $1200$ K
190: (e.g., \citealt{Glikman2006}) and dominate the near- to mid-IR
191: emission. Therefore, the observations by \cite{Jiang2006} indicate
192: the presence of large amounts of both hot and cold dust in these
193: high-z systems.
194:
195: However, the nature of the dust, and its formation and evolution in
196: the context of the quasar host, are unclear. For example, the dust
197: distribution is unknown. It has been suggested that the hot dust lies
198: in the central regions and is heated by an AGN to produce near-IR
199: emission \citep{Rieke1981, Polletta2000, Haas2003}, while the warm and
200: cold dust can extend to a few kpc and dominate at mid-IR and far-IR
201: wavelengths \citep{Polletta2000, Nenkova2002, Siebenmorgen2005}.
202: However, the relative importance of heating by stars and AGN activity
203: is uncertain. This is essential to an accurate determination of the
204: star formation rate (SFR), but cannot be inferred simply from observed
205: SEDs. Currently, the most common method used to derive a star
206: formation rate is to assume that most of the FIR luminosity comes from
207: young stars. However, if the FIR luminosity is mainly contributed by
208: an AGN, then the SFR would be substantially reduced. Finally, it is
209: not known how the SED and dust content of a quasar and its host evolve
210: with time.
211:
212: In order to address these questions, we must combine a model for the
213: formation of a quasar with radiative transfer calculations that treat
214: dust emission self-consistently, to follow the physical properties,
215: environment, and evolution of quasar hosts and their dust content.
216:
217: Earlier \citep{Li2007}, we developed a quasar formation model which
218: self-consistently accounts for black hole growth, star formation,
219: quasar activity, and host spheroid formation in the context of
220: hierarchical structure formation. We employed a set of multi-scale
221: simulations that included large-scale cosmological simulations and
222: galaxy mergers on galactic scales, together with a self-regulated
223: model for black hole growth, to produce a luminous quasar at $z \sim
224: 6.5$, which has a black hole mass ($\sim 2 \times 10^9\, \Msun$)
225: and a number of properties similar to \zquasar\ \citep{Fan2003}. In our
226: scenario, luminous, high-redshift quasars form in massive halos that originate
227: from rare density peaks in the standard $\Lambda$CDM cosmology, and
228: they grow through hierarchical mergers of gas-rich galaxies. Gravitational
229: torques excited in these mergers trigger large inflows of gas and produce
230: strong shocks that result in intense starbursts \citep{Hernquist1989B,
231: Barnes1991, Barnes1992, Barnes1996, Hernquist1995, Mihos1996} and fuel rapid
232: black hole accretion \citep{DiMatteo2005, Springel2005B}. Moreover, feedback
233: from the accreting black hole disperses the obscuring material, briefly
234: yielding an optically visible quasar \citep{Hopkins2005A, Hopkins2005B,
235: Hopkins2005C, Hopkins2006A}, and regulating the $\msigma$ correlation
236: between the SMBHs and host galaxies \citep{DiMatteo2005, Robertson2006A}. In
237: this picture, quasars are, therefore, descendents of starburst galaxies
238: \citep{Sanders1988, Norman1988, Scoville2003}.
239:
240: To make detailed contact with observations of $z \sim 6$ quasars, it
241: is necessary to theoretically predict the SEDs of these systems and
242: how they evolve with time. Solving radiation transport in dusty,
243: starbursting quasars is, however, difficult owing to the non-locality
244: of the sources, the opacity and multiphase character of the
245: interstellar medium, and the complex morphology of these systems.
246: Over the last several decades, a variety of numerical techniques have
247: been developed to solve the radiative transfer problem with different
248: levels of approximation and in multi-dimensions. Based on the
249: specific algorithms employed, the approaches can be classified into
250: two general categories: {\em finite-difference} and {\em Monte Carlo}
251: methods (e.g., \citealt{Jonsson2006}; see also \citealt{Pascucci2004}
252: who describe the codes as ``grid-based'' or ``particle-based'' by
253: analogy to hydrodynamic solvers).
254:
255: {\em Finite-difference} codes solve the equations of radiative
256: transfer (RT) iteratively using finite convergence criteria. They
257: either solve the moment equations for RT, originally formulated by
258: \cite{Hummer1971} for spherical geometry with a central point source
259: (e.g., \citealt{Scoville1976, Leung1976, Yorke1980, Wolfire1986,
260: Menshchikov1997, Dullemond2000}), or employ ray-tracing methods on a
261: discrete spatial grid for complex density configurations (e.g.,
262: \citealt{Rowan-Robinson1980, Efstathiou1990, Efstathiou1991,
263: Steinacker2003, Folini2003, Steinacker2006}). This technique provides
264: full error control but can be time-consuming.
265:
266: {\em Monte Carlo} methods sample and propagate photons
267: probabilistically (e.g., \citealt{Witt1977, Lefevre1982, Lefevre1983,
268: Whitney1992, Witt1992, Code1995, Lopez1995, Lucy1999, Wolf1999,
269: Bianchi2000, Harries2000, Bjorkman2001, Whitney2003A, Jonsson2006,
270: Pinte2006}). The Monte Carlo technique is more flexible than the
271: finite-difference one because it tracks the scattering, absorption and
272: re-emission of photons (or photon packets) in detail, and can handle
273: arbitrary geometries, but at the cost of computational expense to
274: reduce Poisson noise. However, advances in computing technology and
275: algorithms have made high-accuracy Monte Carlo RT calculations feasible
276: and popular.
277:
278: In particular, \cite{Bjorkman2001} have developed a Monte Carlo code
279: to handle radiative equilibrium and temperature corrections, which
280: calculates dust emission self-consistently. It conserves the total
281: photon energy, corrects the frequency distribution of re-emitted
282: photons, and requires no iteration as the dust opacity is assumed to be
283: independent of temperature. This code has been used in a number of
284: applications, including protostars that have a disk and an envelope
285: with a single heating source in the center (e.g.,
286: \citealt{Whitney1992, Whitney1993, Whitney2003B, Whitney2003A,
287: Whitney2004}); circumstellar envelopes (e.g., \citealt{Wood1996B,
288: Wood1996A, Wood1998}); protoplanetary systems (e.g., \citealt{Wood2002,
289: Rice2003}); and galaxies that include a bulge and a disk with multiple
290: heating sources from these two populations (e.g., \citealt{Wood1997,
291: Wood2000}). This code has also been able to generate optical--far-IR
292: SEDs that reproduce those of a sample of 21 X-ray selected AGNs
293: \citep{Kuraszkiewicz2003}, and the version of \cite{Whitney2003A} has been
294: applied to simulations of galaxy mergers with black hole feedback to study the
295: local ULIRGs \citep{Chakrabarti2007A} and
296: submillimeter galaxies \citep{Chakrabarti2007B}.
297:
298: In order to produce accurate SEDs of quasars and their hosts formed by
299: galaxy mergers, as in \cite{Li2007}, the RT code must satisfy the
300: following requirements: (1) to be able to handle arbitrary geometries
301: and distributed heating sources; (2) resolve the large dynamic ranges
302: in spatial scales and densities in the merger simulations; (3)
303: incorporate a multiphase description of the interstellar medium
304: (ISM); and (4) employ a dust model appropriate for a young starburst
305: system.
306:
307: In many earlier applications, radial logarithmic or uniformly spaced
308: meshes were used to generate density grids. However, in merging
309: galaxies, the ISM is clumpy and irregular owing to shocks and tidal
310: features produced during the interactions, and has multiple density
311: centers, making logarithmic algorithms inefficient. In such a
312: situation, an adaptive grid approach appears ideal for resolving
313: localized high-density regions, while still covering the large volumes
314: of merging systems \citep{Wolf1999, Kurosawa2001, Harries2004,
315: Jonsson2006}.
316:
317: In addition, dust models commonly used in previous works are based on
318: observed extinction curves for the Milky Way (e.g.,
319: \citealt{Weingartner2001, Calzetti1994, Kim1994, Mathis1977,
320: Calzetti2000}), in which the dust is assumed to be produced mainly by
321: old, low-mass stars with ages $> 1$ Gyr \citep{Mathis1990,
322: Whittet2003, Dwek2005}. However, a large amount of dust ($\sim 1-7
323: \times 10^8\, \Msun$) is detected in the $z \simeq 6.42$ quasar host
324: \citep{Bertoldi2003A} at a time when the Universe was only $\sim 850$
325: Myr old. It has been suggested by observations (e.g.,
326: \citealt{Maiolino2004, Maiolino2006, Moseley1989, Dunne2003,
327: Morgan2003B, Sugerman2006}) and theoretical studies (e.g.,
328: \citealt{Todini2001, Nozawa2003, Schneider2004, Nozawa2007, Bianchi2007}) that
329: supernovae can provide fast and efficient dust formation in the early
330: Universe, motivating other choices for the dust model in the
331: RT calculations.
332:
333: Finally, in a multiphase description of the ISM \citep{McKee1977}, as
334: adopted in our simulations \citep{Springel2003A, Springel2003B, Hopkins2006A},
335: the ``hot-phase'' (diffuse) and ``cold-phase'' (dense molecular and
336: ${\rm HI}$ core) components co-exist under pressure equilibrium but
337: have different mass fractions and volume filling factors (i.e., the
338: hot-phase gas is $\le 10\%$ in mass but $\gtrsim 99\%$ in volume),
339: so both phases contribute to the dust extinction and should therefore be
340: included in the RT calculations.
341:
342: We have refined the Monte Carlo RT code developed by \cite{Bjorkman2001} and
343: \cite{Whitney2003A} by implementing: an adaptive grid algorithm similar to
344: that of \cite{Jonsson2006} for the density field and the arbitrarily
345: distributed sources; a multiphase ISM model \citep{Springel2003A} which
346: accounts for observed scaling relations of molecular clouds for the dust
347: distribution; and a supernova-origin dust model for the opacity using the
348: grain size distribution of \cite{Todini2001}. We refer to our new
349: code as ART$^2$ (All-wavelength Radiative Transfer with Adaptive
350: Refinement Tree). ART$^2$ is capable of producing SEDs and images in a wide
351: range of wavelengths from X-ray to millimeter. In the present paper we focus
352: on the dust properties from optical to submillimeter bands. As we show in what
353: follows, ART$^2$ reproduces the spectrum of a single galaxy with bulge and
354: disk calculated with the original code. More important, it captures the
355: inhomogeneous density distribution in galaxy mergers and reproduces the
356: observed SED and dust properties of \zquasar\ based on the simulations of
357: \cite{Li2007}. Therefore, ART$^2$ can be used to predict multi-wavelength
358: properties of quasar systems and their galaxy progenitors.
359:
360: This paper is organized as follows. In \S~2, we describe our
361: computational methods and models, including the multi-scale
362: simulations of \cite{Li2007} for $z \sim 6$ quasar formation, and our
363: implementation of ART$^2$ which incorporates radiative equilibrium as
364: in \cite{Bjorkman2001}, an adaptive grid, a multiphase ISM, and a
365: supernova-origin dust model. In \S~3, we present the multi-wavelength
366: SED from optical to submillimeter of a simulated quasar at $z \sim
367: 6.5$. The dust distribution and properties of the quasar are
368: discussed in \S~4 and we describe their evolution in \S~5. We consider
369: the robustness of our results in \S~6 and summarize in \S~7.
370:
371:
372: \section{Methodology}
373:
374: In order to capture the physical processes underlying the formation
375: and evolution of quasars in the early Universe, and to compare their
376: multi-wavelength properties to observations, we combine quasar
377: formation and radiative transfer calculations. We first perform a
378: set of novel multi-scale simulations that yield a luminous quasar at
379: $z \sim 6.5$ \citep{Li2007}. We then apply the 3-D, Monte Carlo
380: radiative transfer code, ART$^2$, to the outputs of the hydrodynamic
381: galaxy mergers simulations to calculate the SED of the system. The
382: models and simulations of quasar formation are described in detail in
383: \cite{Li2007}. Here, we briefly summarize the modeling of quasar
384: formation, and the specifications of ART$^2$.
385:
386: \subsection{Formation Model of $z \sim 6$ Quasars}
387:
388: \subsubsection{Multi-scale Simulations}
389:
390: The $z \sim 6$ quasars are rare (space density $\sim 1\, \Gpc^{-3}$
391: comoving), and appear to be powered by supermassive black holes of
392: mass $\sim 10^9\, \Msun$. Therefore, simulations of high-redshift
393: quasar formation must consider a large cosmological volume to
394: accommodate the low abundance of this population, have a large dynamic
395: range to follow the hierarchical build-up of the quasar hosts, and
396: include realistic prescriptions for star formation, black hole growth,
397: and associated feedback mechanisms. The multi-scale simulations in
398: \cite{Li2007} include both N-body cosmological calculations in a
399: volume of $3\, \Gpc^3$ to account for the low number density of
400: quasars at $z \sim 6$, and hydrodynamical simulations of individual
401: galaxy mergers on galactic scales to resolve gas-dynamics, star
402: formation, and black hole growth.
403:
404: First, we perform a coarse dark matter-only simulation in a volume of
405: $3\, {\rm Gpc^{3}}$. The largest halo at $z=0$, within which early,
406: luminous quasars are thought to reside \citep{Springel2005A}, is then
407: selected for resimulation at higher resolution. The evolution of this
408: halo and its environment is re-simulated using a multi-grid zoom-in
409: technique \citep{Gao2005} that provides much higher mass and spatial
410: resolution for the halo of interest, while following the surrounding
411: large-scale structure at lower resolution. The merging history of the
412: largest halo at $z \sim 6$, which has then reached a mass of $\sim 7.7
413: \times10^{12}\, \Msun$ through seven major (mass ratio $<$ 5:1) mergers
414: between redshifts 14.4 and 6.5, is extracted. These major mergers are
415: again re-simulated hydrodynamically using galaxy models
416: which include a \cite{Hernquist1990} halo and an exponential disk
417: scaled appropriately for redshift \citep{Robertson2006A}, and adjusted to
418: account for mass accretion through minor mergers. Each of these eight
419: galaxy progenitors has a black hole seed assumed to originate from the
420: remnants of the first stars (\citealt{Abel2002, Bromm2004, Tan2004,
421: Yoshida2003, Yoshida2006, Gao2007}).
422:
423: The simulations were performed using the parallel, N-body/Smoothed
424: Particle Hydrodynamics (SPH) code GADGET2 \citep{Springel2005D}, which
425: conserves energy and entropy using the variational principle
426: formulation of SPH \citep{Springel2002}, and which incorporates a
427: sub-resolution model of a multiphase interstellar medium to describe
428: star formation and supernova feedback \citep{Springel2003A}. Star
429: formation is modeled following the Schmidt-Kennicutt Law
430: (\citealt{Schmidt1959, Kennicutt1998}). Feedback from supernovae is
431: captured by an effective equation of state for star-forming gas
432: \citep{Springel2003A}. A prescription for supermassive black hole
433: growth and feedback is also included, where black holes are
434: represented by collisionless ``sink'' particles that interact
435: gravitationally with other components and accrete gas from their
436: surroundings. The accretion rate is estimated from the local gas
437: density and sound speed using a spherical Bondi-Hoyle
438: \citep{BondiHoyle1944, Bondi1952} model that is limited by the
439: Eddington rate. Feedback from black hole accretion is modeled as
440: thermal energy injected into the surrounding gas \citep{Springel2005B,
441: DiMatteo2005}. We note that implementations of our model for black
442: hole growth and feedback that do not explicitly account for
443: Eddington-limited accretion achieve similar results to the
444: method employed by us (e.g. compare the works of \citealt{DiMatteo2007}
445: and \citealt{Sijacki2007}).
446:
447: These hydrodynamic simulations adopted the $\Lambda$CDM model with
448: cosmological parameters from the first year Wilkinson Microwave Anisotropy
449: Probe data (WMAP1, \citealt{Spergel2003}), ($\Omega_{\rm{m}}$,
450: $\Omega_{\rm{b}}$, $\Omega_{\Lambda}$, $h$, $n_s$, $\sigma_8$)= (0.3, 0.04,
451: 0.7, 0.7, 1, 0.9). In this paper, we use the same parameters.
452:
453:
454: \subsubsection{Hierarchical Assembly of the Quasar System}
455:
456:
457: \begin{figure}
458: \begin{center}
459: \vspace{0.5cm}
460: \includegraphics[width=3.4in]{f1.ps}
461: \vspace{0.5cm}
462: \caption{Evolution of the star formation rate, black hole accretion rate, and
463: masses of the black holes and stars, respectively, of the simulated quasar
464: system at $z \sim 6.5$, adopted from \cite{Li2007}. }
465: \label{Fig_quasar}
466: \end{center}
467: \end{figure}
468:
469: In the simulation analyzed here, the quasar host galaxy builds up
470: hierarchically through seven major mergers of gas-rich progenitors
471: between $z = 14.4 - 6.5$. Gravitational interactions between the
472: merging galaxies form tidal tails, strong shocks and efficient gas
473: inflows that trigger intense starbursts. The highly concentrated gas
474: fuels rapid black hole accretion. Figure~\ref{Fig_quasar} shows the
475: evolution of some aspects of the system. Between $z \sim$~14--9, the
476: merging galaxies are physically small and the interactions occur on
477: scales of tens of kiloparsecs. By $z \sim$~9-7.5, when the last major
478: mergers take place, the interactions have increased dramatically in
479: strength. Galaxies are largely disrupted in close encounters, tidal
480: tails of gas and stars extend over hundreds of kiloparsecs, and
481: powerful bursts of star formation are triggered, resulting in an
482: average star formation rate of $\sim 10^3\, \Msun\, \yr^{-1}$ that
483: peaks at $z\sim 8.5$. During this phase, the black holes are heavily
484: obscured by circumnuclear gas. The luminosity from the starbursts
485: outshines that from the accreting black holes. So, we refer to this
486: period ($z \sim 14-7.5$) as the ``starburst phase.''
487:
488: Once the progenitors have coalesced, the multiple SMBHs from the
489: galaxies merge and grow exponentially in mass and feedback energy via
490: gas accretion. During this period ($z \sim 7.5-6$, hereafter referred
491: to as ``quasar phase''), the black hole luminosity outshines that of
492: the stars. At redshift $z \approx 6.5$, when the galaxies coalesce,
493: the induced high central gas densities bring the SMBH accretion and
494: feedback to a climax. The black hole reaches a mass of
495: $\sim 2\times 10^{9}\, \Msun$, and has a peak bolometric luminosity
496: close to that of \zquasar. Black hole feedback then drives a powerful
497: galactic wind that clears the obscuring material from the center of
498: the system. The SMBH becomes visible briefly as an optically-luminous
499: quasar similar to \zquasar. Once the system relaxes, the SMBH and the host
500: satisfy the relation, $\MBH \approx 0.002\, \Mstar$, similar to that
501: measured in nearby galaxies (\citealt{Magorrian1998, Marconi2003}), as a
502: result of co-eval evolution of both components \citep{Li2007,
503: Robertson2006A, Hopkins2007D}.
504:
505: After $z < 6$ (the ``post-quasar phase''), feedback from star
506: formation and the quasar quenches star formation and self-regulates
507: SMBH accretion. Consequently, both star formation and quasar activity
508: decay, leaving behind a remnant which rapidly reddens. The object will
509: eventually evolve into a cD-like galaxy by the present day. (For an
510: overview of this scenario, see e.g., \citealt{Hopkins2007C,
511: Hopkins2007B}.)
512:
513: The photometric calculations by \cite{Robertson2007} show that this quasar
514: system satisfies a variety of photometric selection criteria based on
515: Lyman-break techniques. The massive stellar spheroid descended from these $z
516: \sim 6$ quasars could be detected at $z \sim 4$ by existing surveys,
517: while the galaxy progenitors at higher redshifts will likely require future
518: surveys of large portions of the sky ($ \gtrsim$ 0.5\%) at wavelengths
519: $\lambda \gtrsim 1\, \mu$m owing to their low number densities.
520:
521: \subsection{ART$^2$: All-wavelength Radiative Transfer with Adaptive
522: Refinement Tree}
523:
524: ART$^2$ is based on a unification of the 3-D Monte Carlo radiative
525: equilibrium code developed by \cite{Bjorkman2001} and
526: \cite{Whitney2003A}, an adaptive grid, a multiphase ISM model, and a
527: supernova-origin dust model. Below, we describe our implementation of
528: ART$^2$.
529:
530:
531: \subsubsection{Monte Carlo Radiative Transfer for Dust in Radiative Equilibrium}
532:
533:
534: \begin{figure}
535: \begin{center}
536: \includegraphics[width=3.5in]{f2.ps}
537: \vspace{0.5cm}
538: \caption{Test run of the radiative equilibrium algorithm adopted from
539: \cite{Bjorkman2001} on 3-D Cartesian coordinates for a simple galaxy with
540: a bulge and a disk. The input spectrum of the stars is a blackbody
541: with a temperature of 15000 K (black curve). The red curve is our output,
542: the blue curve is simulation data from Kenny Wood, while the crosses are
543: observations of a lenticular galaxy NGC 5866, which has an unusual
544: extended dust disk seen exactly edge-on. Both Wood's data and the
545: observations come with the code release package from Kenny
546: Wood (http://www-star.st-and.ac.uk/$\sim$kw25/research/montecarlo/gals/gals.html).}
547: \label{Fig_sed_jaffe}
548: \end{center}
549: \end{figure}
550:
551:
552: The Monte Carlo RT method follows the propagation, scattering,
553: absorption, and reemission of "photon packets" (groups of
554: equal-energy, monochromatic photons that travel in the same
555: direction), by randomly sampling the various probability distribution
556: functions that determine the optical depth, scattering angle, and
557: absorption rates. For the dusty environments we are concerned with,
558: the re-emitted spectrum depends on the temperature of the dust, which
559: is assumed to be in thermal equilibrium with the radiation field. In
560: the traditional scheme, the frequencies of the re-emitted photons are
561: sampled from local reemission spectra with fixed temperatures. The
562: dust radiative equilibrium is ensured by performing the Monte Carlo
563: transfer iteratively until the dust temperature distribution
564: converges. Such methods often require a large number of iterations
565: and are therefore computationally expensive \citep{Bjorkman2001,
566: Pascucci2004}.
567:
568: \cite{Bjorkman2001} proposed a solution to this problem by formulating
569: an ``immediate reemission'' algorithm, in which the dust temperature
570: is immediately updated upon absorption of a photon packet, and the
571: frequencies of re-emitted photons are sampled from a spectrum that
572: takes into account the modified temperature. The advantage of this
573: approach is that dust radiative equilibrium and the radiative transfer
574: solutions are obtained simultaneously without iteration. This
575: algorithm is described in detail in \cite{Bjorkman2001}; here we
576: briefly outline the steps.
577:
578: Assuming that each photon packet carries an energy $E_\gamma$, and after
579: absorption of $N_i$ packets in the $i-th$ grid cell, the total energy absorbed
580: in the cell is
581: \begin{equation}
582: E_i^{\rm abs} = N_i E_\gamma \; .
583: \label{eq:Eabs}
584: \end{equation}
585:
586: In radiative equilibrium, this energy must be re-radiated, with a thermal
587: emissivity of $j_\nu = \kappa_\nu \rho B_\nu(T)$, where $\kappa_\nu$ is the
588: absorptive opacity, $\rho$ is the dust density, and $B_\nu(T)$ is the Planck
589: function at temperature $T$,
590:
591: \begin{equation}
592: B_\nu(T)=\frac{2h_{\rm P}}{c^2}\frac{\nu^3}{e^{h_{\rm p}\nu/(kT)}-1} \; ,
593: \label{eq:planck}
594: \end{equation}
595: where $h_{\rm P}$ is Planck's constant, $c$ the speed of light, and $k$ is
596: Boltzmann's constant. The emitted energy in the time interval $\Delta t$ is
597: \begin{eqnarray}
598: \label{eq:Eem}
599: E_i^{\rm em} &=& 4\pi\Delta t \int dV_i \int \rho \kappa_\nu B_\nu(T) \,d\nu \cr
600: &=& 4\pi\Delta t \kappa_{\rm P}(T_i) B(T_i) m_i \;,
601: \end{eqnarray}
602: where $\kappa_{\rm P}=\int\kappa_\nu B_\nu\,d\nu / B$ is the Planck mean
603: opacity, $B = \sigma T^4 / \pi$ is the frequency integrated Planck
604: function, and $m_i$ is the dust mass in the cell.
605:
606: Equating the absorbed (\ref{eq:Eabs}) and emitted (\ref{eq:Eem})
607: energies, we obtain the dust temperature as follows after absorbing $N_i$ packets:
608: \begin{equation}
609: \sigma T_i^4= { {N_i L }\over {4 N_\gamma \kappa_{\rm P}(T_i) m_i} } \;,
610: \label{eq:REtemp}
611: \end{equation}
612: where $N_\gamma$ is the total number of photon packets in the simulation, and $L$
613: is the total source luminosity. Note that
614: because the dust opacity is temperature-independent, the product
615: $\kappa_{\rm P}(T_i) \sigma T_i^4$ increases monotonically with temperature.
616: Consequently, $T_i$ always increases when the cell absorbs an additional
617: packet.
618:
619: The added energy to be radiated owing to the temperature increase $\Delta T$
620: is determined by a temperature-corrected emissivity $\Delta j_\nu$ in the
621: following approximation when the temperature increase, $\Delta T$, is small:
622: \begin{equation}
623: \Delta j_\nu \approx \kappa_\nu \rho \Delta T {{dB_\nu(T)} \over {dT}} \;.
624: \end{equation}
625: The re-emitted packets, which comprise the diffuse radiation field, then
626: continue to be scattered, absorbed, and re-emitted until they finally escape
627: from the system. This method conserves the total energy exactly, and does not
628: require any iteration as the emergent SED, $\nu L_\nu = \kappa_\nu B_\nu(T)$,
629: corresponds to the equilibrium temperature distribution \citep{Bjorkman2001}.
630:
631: This Monte Carlo radiative equilibrium code works as follows: first, the
632: photon packets are followed to random interaction locations, determined by the
633: optical depth. Then they are either scattered or absorbed with a probability
634: given by the albedo ($a = n_s \sigma_s/(n_s \sigma_s + n_a
635: \sigma_a)$, where $n$ and $\sigma$ are number density and cross section for
636: either scattering or absorption, respectively). If the packet is scattered, a
637: random scattering angle is obtained from the scattering phase function. If
638: instead the packet is absorbed, it is reemitted immediately at a new frequency
639: determined by the envelope temperature, using the algorithm described
640: above. After either scattering, or absorption plus reemission, the photon
641: packet continues to a new interaction location. This process is repeated until
642: all the packets escape the dusty environment. Upon completion of the Monte
643: Carlo transfer, the code produces emergent SEDs and images at any given
644: inclination for a wide range of broadband filters, including those of {\em
645: Hubble Space Telescope} (NICMOS bands), {\em Spitzer} (IRAC and MIPS bands),
646: and {\em SCUBA} submillimeter bands.
647:
648:
649: \cite{Bjorkman2001} tested this algorithm extensively by comparing a
650: set of benchmark calculations to those of \cite{Ivezic1997} for
651: spherical geometry. \cite{Baes2005} critically study the frequency
652: distribution adjustment used by \cite{Bjorkman2001}, and give a firm
653: theoretical basis for their method, although it may fail for small dust
654: grains.
655:
656: One drawback of this code, however, is its fixed geometry and limited
657: dynamic range. It can handle only 2-D or 3-D spherical-polar grids,
658: which are not suitable for capturing the arbitrary geometry and large
659: dynamic ranges characteristic of merging galaxies, which
660: have multiple density centers, and where gas and stars extend hundreds
661: of kpc and shocks produce highly condensed gas on scales of
662: pcs. We have developed an improved version of this code by
663: implementing an adaptive Cartesian grid on top of the code released by
664: Wood~\footnote{http://www-star.st-and.ac.uk/$\sim$kw25/research/montecarlo/gals/gals.html}.
665:
666:
667: To ensure that our implementation of the algorithm is correct, we
668: rerun the test problem that comes with the code release package of
669: Wood using a uniform Cartesian grid. The test problem consists of a
670: simple galaxy with bulge and disk components. As is shown in
671: Figure~\ref{Fig_sed_jaffe}, our code reproduces the result of Wood
672: very well.
673:
674:
675: \subsubsection{Adaptive-mesh Refinement Grid}
676: \label{subsec_grid}
677:
678: \begin{figure*}
679: \begin{center}
680: \includegraphics[width=3.0in]{f3a.ps}
681: \hspace{0.5cm}
682: \includegraphics[width=3.0in]{f3b.ps} \\
683: \includegraphics[width=3.0in]{f3c.ps}
684: \hspace{0.5cm}
685: \includegraphics[width=3.0in]{f3d.ps} \\
686: \includegraphics[width=3.0in]{f3e.ps}
687: \hspace{0.5cm}
688: \includegraphics[width=3.0in]{f3f.ps}
689: \vspace{0.5cm}
690: \caption{Example of the adaptive grid applied to a snapshot of the galaxy
691: merger from \cite{Li2007}. From top to bottom is: uniform grid, and adaptive
692: grids with maximum refinement levels of RL=5, and RL=12, respectively. The
693: label GN indicates the total grid number covering the entire volume (in case
694: of adaptive grids, it provides the number of the finest grid and that of the
695: total grid). This plot demonstrates that adaptive grids with
696: sufficient refinement capture the density distribution well, while the
697: uniform grid fails to do so unless an unreasonably large number of cells is
698: employed. Throughout this paper, we use a maximum refinement level
699: of 12, which has a grid resolution comparable to that of the hydrodynamic
700: simulations in \cite{Li2007}.}
701: \label{Fig_grid}
702: \end{center}
703: \end{figure*}
704:
705:
706: The SPH simulations using GADGET2 \citep{Springel2005D} output
707: hydrodynamic information as particle data. However, the ray tracing in
708: the RT calculation is done on a grid. Therefore, it is necessary to
709: interpolate the particle-based density field onto a
710: grid. \cite{Jonsson2006} performed radiative transfer calculations on
711: SPH simulations of galaxy mergers using an adaptive grid as
712: implemented in his Monte Carlo RT code, SUNRISE. Unfortunately,
713: self-consistent calculations of dust radiative equilibrium and emission
714: are not yet included in SUNRISE.
715:
716: Our algorithm for constructing adaptive grids is similar to that of
717: \cite{Jonsson2006}. We typically start with a base grid of a $4^3$ box
718: covering the entire simulation volume. Each cell is then adaptively
719: refined by dividing it into $2^3$ sub-cells. The refinement is stopped
720: if a predefined maximum refinement level, RL, is reached, or if the
721: total number of particles in the cell becomes less than a certain
722: threshold, whichever criterion is satisfied first. The maximum refinement
723: level used in the present work is 12, and the maximum particle number allowed
724: in the cell is 32, half the number of the SPH smoothing kernel neighbors used
725: in the GADGET2 simulations. The resolution of the finest level is therefore
726: $L_{\rm min}=L_{\rm box}/2^{(\rm RL+1)}$, where $L_{\rm box}$ is the box length,
727: and ${\rm RL}$ is again the maximum refinement level. For example, for the
728: parameters used in this simulation, $L_{\rm box}=200$ kpc, ${\rm RL}=12$, we
729: have $L_{\rm min}= 24.4$ pc.
730:
731: The adaptive-mesh refinement grid serves as an efficient tree for mass
732: assignment owing to fast neighbor finding within the grids for SPH
733: smoothing. After the grid is constructed, the gas properties at the
734: center of each grid cell, such as density, temperature, and
735: metallicity, are calculated using the SPH smoothing kernel
736: \citep{Hernquist1989A} of the original simulation. All physical quantities
737: are assumed to be uniform across a single cell.
738:
739: Figure~\ref{Fig_grid} gives an example of the adaptive grid applied to a
740: snapshot from the galaxy merger simulations in \cite{Li2007}, and compared
741: with a uniform grid. This particular snapshot represents the time when the
742: system reaches the peak quasar phase, when the galaxies are in the final
743: stages of coalescence. The system is highly dynamical as much gas is falling
744: into the center, while feedback from the central massive black hole drives an
745: outflow. The gas distribution is thus inhomogeneous.
746:
747: As is apparent in Figure~\ref{Fig_grid}, a uniform grid of $50^3$ (top
748: panel) barely captures the density distribution with a spatial
749: resolution of $4\, \kpc$. The resolution is worse than an adaptive grid with a
750: moderate maximum refinement level (i.e., RL=5, middle panel), which has a
751: finest cell length of $\sim 3.1\, \kpc$. The grid with a maximum refinement
752: level of 12 shown in the bottom panel fully
753: captures the large dynamic range of the gas density distribution in
754: three dimensions. It has a minimum cell length of $\sim 24.4$ pc for
755: the finest cells, which is comparable to the spatial resolution of the
756: original hydrodynamic simulation of \cite{Li2007}. This resolution is
757: equivalent to a ${\sim 10000 }^3$ uniform grid, which is impractical
758: with existing computational facilities.
759:
760: \subsubsection{A Multiphase Model for the Interstellar Medium}
761: \label{subsec_ism}
762:
763: In determining the dust distribution, we adopt the multiphase model of
764: \cite{Springel2003A} for the ISM. The ISM is then comprised of
765: condensed clouds in pressure equilibrium with an ambient hot gas, as
766: in the picture of \cite{McKee1977}. In the hydrodynamic simulations,
767: individual SPH particles represent regions of gas that contain cold,
768: dense cores embedded in a hot, diffuse medium. The hot and cold phases
769: of the ISM co-exist in pressure equilibrium but have different mass
770: fractions and volume filling factors (i.e., the hot-phase gas is $\le
771: 10\%$ in mass but $\gtrsim 99\%$ in volume). In our previous studies,
772: \cite{Li2007} and \cite{Hopkins2006A} used only the hot-phase density
773: to determine e.g. the obscuration of the central AGN, as the majority
774: of sight lines will pass through only this component owing to its
775: large volume filling factor. However, this method gives only an
776: effective lower limit on the column density.
777:
778: Here, we consider two components of the dust distribution, having
779: different dust-to-gas ratios and being associated with the two phases
780: of the ISM. Within each grid cell in the RT calculation, the hot gas
781: is uniformly distributed, while the cold, dense cores are randomly
782: embedded. Because of the much higher density and smaller volume of
783: the cold phase, it is impractical to resolve these clouds either in
784: hydrodynamic simulations or in our radiative transfer calculations.
785: Consequently, we implement an observationally-motivated,
786: sub-resolution prescription to treat the cold clouds, constrained by
787: the observed mass spectrum and size distribution of molecular clouds
788: in galaxies.
789:
790: Observations of giant molecular clouds (GMCs) in galaxies show that
791: the GMCs follow simple scaling relations \citep{Larson1981},
792: namely a power-law mass distribution, ${\rm d}N/{\rm d}M \propto
793: M^{-2}$ (e.g., \citealt{Fuller1992, Ward-Thompson1994, Andre1996,
794: Blitz2006}), as well as a power-law mass-radius relation $M \propto
795: R^2$ (.e.g, \citealt{Sanders1985, Dame1986, Scoville1987, Solomon1987,
796: Rosolowsky2005, Rosolowsky2007}). It has been suggested by theoretical
797: modeling that the mass function is produced by turbulence in
798: self-gravitating clouds (e.g., \citealt{Elmegreen1996, Elmegreen2002,
799: Ballesteros-Paredes2002, Li2003}), while the mass-radius relation is
800: attributed to virial equilibrium \citep{Larson1981}.
801:
802: Here we incorporate these two empirical relations into our ISM model
803: for the RT calculations. For a given cell, the two-phase break-down in
804: \cite{Springel2003A} determines the hot and cold phase gas density
805: according to pressure equilibrium. The cold clouds are assumed to
806: follow the Larson scaling relations:
807:
808: \begin{eqnarray}
809: \label{eq:ms}
810: \frac{dn}{dM} &=& A M^{-\alpha}, \\
811: M &=& BR^{\beta},
812: \end{eqnarray}
813: where $dn/dM$ is the number density of the clouds differential in cloud
814: mass $M$, and $R$ is the cloud radius. From these equations, one
815: obtains the cloud size distribution
816: \begin{eqnarray}
817: \frac{dn}{dR} &=& \beta A B^{1-\alpha} R^{-(\alpha\beta+1-\beta)} \nonumber\\
818: &=& C R^{-\gamma},
819: \end{eqnarray}
820: where $C = \beta A B^{1-\alpha}$, and $\gamma = \alpha\beta+1-\beta$.
821:
822: Assume that the minimum and maximum values of the cloud mass are $M_0$ and $M_1$,
823: then the normalization constant of the mass spectrum for each cell can be
824: determined using
825: \begin{equation}
826: \int \frac{dn}{dM}MdM = x_c\rho_c,
827: \end{equation}
828: or
829: \begin{equation}
830: A = x_c\rho_c\frac{2-\alpha}{M_1^{2-\alpha} - M_0^{2-\alpha}} \, ,
831: \end{equation}
832: where $\rho_c$ and $x_c$ are the cold gas density and volume filling factor,
833: respectively. The normalization constant of the $M$-$R$ relation may be
834: determined using
835: \begin{equation}
836: x_c = \int \frac{4\pi}{3}R^3\frac{dn}{dM}dM,
837: \end{equation}
838: or
839: \begin{equation}
840: B = \left[\frac{4\pi A}{3\eta x_c}\left(M_1^\eta -
841: M_0^\eta\right)\right]^{\beta/3},
842: \end{equation}
843: where
844: \begin{equation}
845: \eta = 1 + \frac{3}{\beta} - \alpha.
846: \end{equation}
847: The minimum and maximum cloud radii in the cell are therefore
848: \begin{eqnarray}
849: R_0 &=& \left(\frac{M_0}{B}\right)^{1/\beta} \nonumber\\
850: R_1 &=& \left(\frac{M_1}{B}\right)^{1/\beta}.
851: \end{eqnarray}
852:
853: For a photon traveling a distance $L$ in the cell, the average number of
854: cold clouds of radius $R$ the photon will intersect is given by
855: \begin{eqnarray}
856: \label{eq:NR}
857: \frac{dN}{dR} &=& \pi R^2 L \frac{dn}{dR} \nonumber\\
858: &=& \pi C R^{2-\gamma}.
859: \end{eqnarray}
860: Integrating over the cloud radius, one obtains
861: \begin{equation}
862: N = \pi LC \frac{R_1^{3-\gamma} - R_0^{3-\gamma}}{3-\gamma}.
863: \end{equation}
864: Therefore, the average distance the photon must travel to hit a cold cloud
865: (the mean free path) is
866: given by
867: \begin{equation}
868: L_{m} = \frac{3-\gamma}{\pi
869: C\left(R_1^{3-\gamma}-R_0^{3-\gamma}\right)}.
870: \end{equation}
871:
872: In the dust RT calculation, we assume that dust is associated
873: with both the cold and hot phase gases through certain dust-to-gas ratios. When a
874: photon enters a cell, we first determine the distance $L_h$ it travels in the
875: hot phase gas before hitting a cold cloud, which is an exponential
876: distribution function
877: \begin{equation}
878: p(L) = \frac{1}{L_{m}}\exp{\left(-\frac{L}{L_{m}}\right)}.
879: \end{equation}
880: Therefore, $L_h=-L_{m}\ln \xi$, where $\xi$ is a random number
881: uniformly distributed between 0 and 1. The radius of the cloud the photon just
882: hits is also determined randomly assuming the distribution function of
883: Equation~(\ref{eq:NR}); i.e.,
884: \begin{equation}
885: R = \left[R_0^{3-\gamma} +
886: (R_1^{3-\gamma}-R_0^{3-\gamma})\xi\right]^{1/(3-\gamma)}.
887: \end{equation}
888:
889: The distance $L_c$ the photon travels in this cold cloud is again a random
890: variable given by $L_c=2R\sqrt{\xi}$, because clouds are assumed to be
891: uniformly distributed. These equations completely define the statistical
892: procedure for determining the dust column densities associated
893: with $L_h$ and $L_c$ as:
894: \begin{eqnarray}
895: \label{eq:NhNc}
896: N_h &=& \rho_h L_h \nonumber\\
897: N_c &=& \frac{3BR^{\beta-3}L_c}{4\pi}.
898: \end{eqnarray}
899:
900: With a given dust opacity curve, these equations allow one to calculate the
901: optical depths, $\tau_h$ and $\tau_c$ for the hot and cold dust,
902: respectively. They relate the photon path lengths in the multiphase ISM to the
903: total optical depth $\tau_{\rm tot}$, which is then compared with a randomly
904: drawn number $\tau_i$, to determine whether the photon should be stopped for
905: scattering or absorption. In detail, the Monte Carlo ray-tracing
906: procedure for the radiative transfer therefore involves the following steps:
907:
908: \begin{enumerate}
909:
910: \item
911: A photon packet is emitted from either a blackhole or a stellar source with
912: random frequencies consistent with the source spectra. The photon is emitted
913: with a uniformly distributed random direction. The probability of a photon
914: being emitted by any given source is determined by its luminosity relative to
915: the total.
916:
917: \item
918: A random optical depth over which the photon must travel before an interaction
919: with the dust occurs, $\tau_i=-\ln\xi$, is drawn to determine the interaction
920: location. The interaction includes scattering and absorption. In our method,
921: the photon energies are not weighted, only one event is allowed. That
922: is, at any given interaction site, the photon is either scattered or absorbed,
923: but not both.
924:
925: \item
926: Starting from the location of the photon emission, the cumulative optical
927: depth of the photon, $\tau_{\rm tot}$, is calculated stochastically using
928: Equation \ref{eq:NhNc} for both hot and cold dusts. If the photon is stopped
929: for interaction within a single cell, then $\tau_{\rm tot}$ is the sum of
930: contributions from possibly multiple segments of both hot and cold dusts
931: within this cell. If the photon passes through multiple cells before an
932: interaction occurs, then $\tau_{\rm tot}$ is the sum of all contributions from
933: relevant segments in these cells.
934:
935: \item
936: At each boundary between the hot and cold phase gas clouds, or at the boundary
937: of the grid cell, the next interaction point is determined by the comparison
938: between $\tau_i$ and $\tau_{\rm tot}$. If $\tau_i \le \tau_{\rm tot}$, then
939: the photon is either scattered or absorbed, with a probability given by the
940: scattering albedo. The exact interaction location is then determined inside
941: either hot or cold phase gas, such that $\tau_{\rm tot}$ becomes exactly
942: $\tau_i$. If the photon is scattered, its direction and polarization state are
943: altered using the Henyey-Greenstein phase function, and the ray-tracing of the
944: new photon is repeated from step 2. If the photon is absorbed, depending on
945: whether the absorption site occurs in the hot or cold phase dust, the
946: temperature of the appropriate dust cell is raised, and a new photon is
947: reemitted according to the scheme of \citet{Bjorkman2001}. The ray-tracing of
948: the newly emitted photon again restarts from step 2.
949:
950: \item
951: If the photon escapes from the system without reaching the optical depth
952: $\tau_i$, it is then collected in the output spectrum and image. The next
953: photon will be picked up from the source, and the whole Monte Carlo procedure
954: from step 1 will be restarted.
955: \end{enumerate}
956:
957: After all $N_{\gamma}$ photons have been traced, one obtains the
958: dust temperature distribution for both hot and cold dusts in radiative
959: equilibrium, as well as the output spectra and images. Note that such a
960: stochastic procedure outlined here may not be as accurate as physically
961: tracking the locations and sizes of the cold phase gas clouds, which can be
962: rather difficult, if not impossible. However, this method works
963: efficiently for a large number of cold clouds uniformly distributed in the
964: cells, and it gives correct average extinction properties for a multiphase
965: ISM model we are interested here.
966: %
967:
968: Hereafter, we refer to ``HPG-dust'' as the dust that originates from
969: the hot-phase gas, and ``CPG-dust'' as that originating from the
970: cold-phase gas. Note that the gas only determines the distribution and mass
971: of the dust through a given dust-to-gas ratio. The dust temperature is
972: not associated with the gas temperature, and is calculated
973: self-consistently according to radiative equilibrium. We further
974: refer to ``cold dust'', ``warm dust'' and ``hot dust'' as dust with
975: temperatures $T \lesssim 100$~K, $100 \lesssim T \lesssim 1000$~K, and $ 1000
976: \lesssim T \lesssim 1200$~K, respectively.
977:
978: In the RT calculations, we adopt a mass spectrum with $\alpha=1.8$, as
979: suggested by the observations of \cite{Blitz2006}, and an observed
980: mass-radius relation with $\beta=2.0$, which is also a result of the
981: virial theorem. The resulting mass-radius relation in our simulations
982: has a normalization in the range $\sim 10 - 10^4\, \Msun/{\rm
983: pc^2}$. For a cloud size of $\sim 1\, {\rm pc}$, the normalization is
984: $\sim 300\, \Msun/{\rm pc^2}$, similar to observations of the Milky Way
985: \citep{Scoville1987, Rosolowsky2007}. It has been shown that the
986: normalization of the mass-radius relationship depends on galactic
987: environment (e.g., \citealt{Elmegreen1989, Rosolowsky2005,
988: Blitz2007}). For example, it is about $\sim 50\, \Msun/{\rm pc^2}$ in the
989: Large Magellanic Cloud but could be two orders of magnitude higher in
990: ULIRGs. In extreme starburst galaxies, the normalization could go up
991: to $10^4\, \Msun/{\rm pc^2}$, as we find in our simulations.
992:
993: We assume the cold clouds to be in the mass range of $M_0=10^3\,
994: M_\odot$ and $M_1=10^7\, M_\odot$, which is similar to that of
995: protoclusters clouds in star forming galaxies, as shown in
996: \cite{Li2004, Li2005A, Li2005B, Li2006}. For the cold clouds, we have
997: enhanced the pressure by a factor of 10 over the thermal pressure to
998: account for the effects of turbulence (e.g., \citealt{Blitz2006,
999: Chakrabarti2007A}). This enhancement factor is referred to as ``CP''
1000: hereafter. The dust-to-gas ratio of the cold clouds is chosen
1001: to be the same as the Milky Way value (1:124; \citealt{Weingartner2001}), as
1002: found in a large sample of ULIRGs \citep{Dunne2001, Klaas2001}, while that of
1003: the hot, diffuse gas is chosen to be 1\% of the Milky Way value, consistent
1004: with the dust survival rate of sputtering in a hot, diffuse ISM
1005: \citep{Burke1974, Reynolds1997}. Observations of some obscurred or red AGNs
1006: also suggest a ratio significantly lower than that of Milky Way (e.g.,
1007: \citealt{Maiolino2001, Kuraszkiewicz2003, Hall2006}). We will perform a
1008: systematic parameter study of these choices in \S~\ref{sec_param}. However, we
1009: note that these values are found to best reproduce the observed quasar SED at
1010: $z\sim 6.5$. Because the dust opacity is proportional to the gas metallicity,
1011: therefore the dust opacity is weighted by the metallicity of the gas for both
1012: hot and cold gas phases.
1013:
1014:
1015:
1016:
1017: \subsubsection{Supernova-origin Dust Model}
1018: \label{subsec_sndust}
1019:
1020:
1021: \begin{figure}
1022: \begin{center}
1023: \includegraphics[width=3.5in]{f4.ps}
1024: \vspace{0.5cm}
1025: \caption{
1026: Comparison of the dust absorption opacity curves from a supernova-origin dust
1027: model (SN model, with silicate fraction $f_{\rm s}=10\%$, 20\% in red and
1028: black, respectively) and those of \cite{Weingartner2001} (WD01 model,
1029: with $R_{\rm V}=3.1$, 4.0, 5.5, in blue, cyan, and green, respectively). Note
1030: there are two differences between these two models: the silicate feature at $\sim
1031: 9.7\, \mu$m, and a higher opacity in the optical band.}
1032: \label{Fig_opac}
1033: \end{center}
1034: \end{figure}
1035:
1036:
1037: \begin{figure}
1038: \begin{center}
1039: \includegraphics[width=3.5in]{f5.ps}
1040: \vspace{0.5cm}
1041: \caption{Distribution of the optical depths of both HPG- and CPG-dust,
1042: respectively, at $\lambda=0.1\, \mu$m where the dust opacity peaks as shown
1043: in Figure~\ref{Fig_opac}. The density grid corresponds to the simulation
1044: snapshot at $z=6.5$ shown in Figure~\ref{Fig_grid} (bottom panel). The
1045: ``HPG-dust'' and ``CPG-dust'' refer to the dust associated with hot-phase and
1046: cold-phase gas, respectively, as described in \S~\ref{subsec_ism}. }
1047: \label{Fig_tau}
1048: \end{center}
1049: \end{figure}
1050:
1051:
1052: Our understanding of the formation and distribution of dust has benefited
1053: from dust maps of the Milky Way (e.g., \citealt{Gehrz1989, Schlegel1998}). In
1054: particular, \cite{Gehrz1989} concludes from a detailed Galactic
1055: survey of dust-producing stars that dust in the Milky Way originates in three
1056: principle ways: (1) by condensation in winds of evolved,
1057: post-main-sequence objects, which accounts of $\sim 90\%$ of the stellar dust;
1058: (2) by condensation in ejecta from massive novae, supernovae and Wolf-Rayet
1059: stars, which amounts to $< 10\%$; and (3) by slow accretion in molecular
1060: clouds (see also \citealt{Marchenko2006} for a review).
1061:
1062: Over the past several decades, various dust models have been developed
1063: based on the observed extinction curves of the Large Magellanic Cloud
1064: (LMC), the Small Magellanic Cloud (SMC) and the Milky Way (e.g., see
1065: reviews by \citealt{Savage1979, Mathis1990, Calzetti1994,
1066: Dorschner1995, Calzetti2000, Whittet2003, Draine2003}). In these
1067: classical models, dust is assumed to form in the envelopes of
1068: old, low-mass stars such as asymptotic giant branch (AGB) stars
1069: with ages $\gtrsim 1$ Gyrs \citep{Mathis1990, Morgan2003A, Dwek2005,
1070: Marchenko2006}.
1071:
1072: However, this picture may be different for young, high-redshift
1073: objects. Recent deep millimeter and submillimeter observations of
1074: several $z \sim 6$ quasars, which trace the far-infrared thermal dust
1075: emission from these systems, show large masses of dust in these quasar
1076: hosts when the Universe was less than 1 Gyr old (e.g.,
1077: \citealt{Bertoldi2003A, Charmandaris2004, Robson2004, Carilli2004,
1078: Maiolino2004, Beelen2006, Hines2006, Jiang2006, Willott2007}).
1079: In particular,
1080: \cite{Maiolino2004} find that the extinction curve of the reddened
1081: quasar SDSS J1048+46 at $z = 6.2$ is different from those observed at
1082: $z < 4$ (similar to that of the Small Magellanic Cloud,
1083: \citealt{Hopkins2004}), but matches the extinction curve expected for
1084: dust produced by supernovae.
1085:
1086: It has been suggested that core-collapse supernovae (Type-II SNe) may
1087: provide a fast and efficient mechanism for dust production. The
1088: observational evidence for dust formation in SNe comes from
1089: observations of SN1987A (e.g., \citealt{Gehrz1987, Moseley1989,
1090: Roche1993, Spyromilio1993, Colgan1994}), Cassiopeia A
1091: (\citealt{Dunne2003, Dwek2004}; see however \citealt{Krause2004} who
1092: argue that the dust emission is not associated with the remnant), Kepler's
1093: supernova remnant \citep{Morgan2003B}, and SNe 2003gd
1094: \citep{Sugerman2006}. Theoretically, several groups have calculated
1095: dust formation in the ejecta of Type-II SNe \citep{Todini2001,
1096: Nozawa2003, Schneider2004, Hirashita2005, Dwek2007}. In particular,
1097: \cite{Todini2001} have developed a dust model based on standard nucleation
1098: theory and tested it on the well-studied case of SN1987A. They find that SNe
1099: with masses in the range of $12 - 35\, \Msun $ produce about 1\% of the mass
1100: in dust per supernova for primordial metallicity, and $\sim 3\%$ for
1101: solar metallicity.
1102:
1103:
1104:
1105: In the present work, we adopt the dust size distribution of
1106: \cite{Todini2001} for solar metallicity and a $M=22\, \Msun$ SN model, as in
1107: Figure~5 in their paper. This size distribution is then combined with the dust
1108: absorption and scattering cross sections of \cite{Weingartner2001}, to
1109: calculate dust absorption opacity curves \citep{Finkbeiner1999A, Finkbeiner1999B}.
1110: We note that \cite{Bianchi2007} re-analysize the model of \cite{Todini2001} and
1111: follow the evolution of newly condensed grains from the time of formation to
1112: their survival. This new feature shows better agreement with
1113: observations and further supports our motivation to use a SN dust model.
1114: Figure~\ref{Fig_opac} shows dust absorption
1115: opacity curves, in the range $10^{-2} - 10^4\, \mu$m, from both
1116: the supernova-origin dust model (hereafter SN model) and
1117: \cite{Weingartner2001} (hereafter WD model), respectively. The SN
1118: models include silicate fractions $f_{\rm s}=10\%$ (in red) and $f_{\rm
1119: s}=20\%$ (in black), which show slight differences in the silicate
1120: feature at $\sim 9.7\, \mu$m ($f_{\rm s}=20\%$ model has a slightly
1121: higher opacity). The WD models include all the curves commonly used
1122: with extinction $R_{\rm V}=\frac{A(V)}{A(B)-A(V)}=3.1$, 4.0, 5.5,
1123: which differ in the opacity in the UV-NIR ($0.01\, \mu \rm{m} - 1\,
1124: \mu \rm{m}$) bands ($R_{\rm V}=3.1$ model has the highest
1125: opacity).
1126:
1127: Between the SN and WD models, there are two noticeable differences; one is
1128: the silicate feature at $\sim 9.7\, \mu$m, and the second is the UV-NIR
1129: band. On one hand, the WD model has strong peaks around $\sim 9.7\, \mu$m
1130: (silicate feature), while the SN model produces an opacity lower by
1131: nearly one order of magnitude, because the ejecta of SNe with mass
1132: above $15\, \Msun$ predominantly form amorphous carbon grains, which
1133: decreases the silicate fraction in the dust grains. On the other hand,
1134: the SN model increases the opacity in the optical band by a factor of
1135: a few owing to the smaller grain size. We note that the
1136: small grain sizes may cause
1137: quantum fluctuations
1138: in the temperature and high-temperature transients (e.g.,
1139: \citealt{Draine2001}). However, the dust produced by the SN model is dominated
1140: by graphite grains, which have a size distribution from
1141: $\sim 100 - 2000$~\AA, comparable to the size range of the dust grains in the
1142: WD model, so the temperature fluctuation is expected to be insignificant.
1143: Note we do not include polycyclic aromatic hydrocarbon (PAH) features at $\sim
1144: 7.7\, \mu$m, which was modeled in detail by \cite{Li2002} and
1145: \cite{Draine2007}. The PAH feature is a good diagnostic of starbursts
1146: at low redshifts \citep{Houck2005}, however it is very difficult to
1147: detect in $z \gtrsim 6$ systems. Also dust spin is not included in our
1148: modeling, which could be significant around $10$~mm
1149: \citep{Draine1998, Finkbeiner2002}.
1150:
1151: The differences between these two models, in particular the silicate
1152: feature, may be diminished by choosing different grain compositions
1153: and size distributions \citep{Laor1993}. For example, the WD model
1154: would converge to the SN model if the silicate fraction is reduced to
1155: $\sim 5\%$. We note that \cite{Elvis2002} propose an alternative
1156: mechanism for dust production in quasars. They suggest that the
1157: physical conditions (i.e., low temperature and high density) in the
1158: quasar outflow are similar to those in the envelopes of AGB stars, and
1159: hence may produce dust similar to that made by AGB stars. This
1160: scenario does not have the same timescale issue as does the AGB model, but
1161: only requires heavy metals such as carbon and silicate, which should
1162: be available even in high-redshift quasars, as noted in the
1163: introduction. Therefore, quasar winds may also serve as efficient
1164: factories for producing dust with extinction curves similar to those
1165: of the WD model. Currently there are no observations available to
1166: distinguish between these models at high redshifts. We will return to
1167: this in \S~\ref{subsec_dustcomp}.
1168:
1169: Figure~\ref{Fig_tau} shows an example of the resulting optical depths
1170: of both HPG- and CPG-dust at wavelength $\lambda=0.1\, \mu$m where the
1171: dust opacity reaches its maximum, as shown in Figure~\ref{Fig_opac}. This plot
1172: demonstrates that the optical depth is resolved very well in the hot, diffuse
1173: gas which occupies $\sim 99\%$ of the volume, while the optical depth of the
1174: cold, dense cores is usually much larger than unity, in particular in optical
1175: bands with high frequencies. The energy absorbed by the cold dust in these
1176: wavelengths is then re-emitted at infrared or longer wavelengths.
1177:
1178:
1179: \section{The Spectral Energy Distribution of A Quasar System at $z \simeq 6.5$}
1180: \label{sec_sed}
1181:
1182:
1183: \begin{figure*}
1184: \begin{center}
1185: \includegraphics[width=5.0in]{f6.ps}
1186: \vspace{1.5cm}
1187: \caption{
1188: {\small SEDs of a quasar system at $z \simeq 6.5$ determined by applying ART$^2$ to
1189: the SPH simulations of \cite{Li2007}. The input spectra (top
1190: panel) include the stellar spectrum (in blue) calculated with STARBURST99
1191: \citep{Leitherer1999, Vazquez2005}, and the composite spectrum of the AGN (in
1192: black, \citealt{Hopkins2007A}). Note this snapshot corresponds to the peak
1193: quasar phase of the system, the stellar radiation is insignificant compared to
1194: that from the AGN at this evolution stage.
1195: The red curve is the total input spectrum. The
1196: output spectra are shown in the bottom panel. The red curve is
1197: the total SED assuming an isotropic distribution of photon energies in all
1198: directions, while the grey region indicates the range from two orthogonal
1199: angles corresponding to the z-axis and xy-plane of the Cartesian grid,
1200: respectively. The filled black circles with error bars are observations from
1201: \cite{Jiang2006}. } }
1202: \label{Fig_sed_z6}
1203: \end{center}
1204: \end{figure*}
1205:
1206:
1207: We now calculate the spectral energy distribution from UV/optical to
1208: submillimeter of the modeled quasar in \cite{Li2007} by applying
1209: ART$^2$ to our SPH simulations, which provide the gas density field
1210: and heating sources for the RT calculations. The input spectrum
1211: includes that from stars and black holes, as shown in
1212: Figure~\ref{Fig_sed_z6} (top panel). The stellar spectrum is
1213: calculated using the stellar population synthesis code STARBURST99
1214: \citep{Leitherer1999, Vazquez2005}. The age, mass and metallicity of
1215: the stars are taken from the SPH simulations of quasar formation in
1216: \cite{Li2007}, while the stellar initial mass function (IMF) is
1217: assumed to be a top-heavy Kroupa IMF \citep{Kroupa2002}, which
1218: characterizes a starburst better than the classical Salpeter IMF
1219: \citep{Salpeter1955}.
1220:
1221: The input black hole spectrum is a composite template from
1222: \cite{Hopkins2007A}, which consists of a broken power-law (e.g.,
1223: \citealt{Laor1993, Marconi2004}) and an IR component thought
1224: to come from the hottest dust around the AGN, which cannot
1225: be fully resolved in our hydrodynamic simulations. The normalization
1226: of this spectrum is the total bolometric luminosity of the black
1227: holes. Compared to the luminosity calculation in \cite{Li2007}, the
1228: black hole spectrum here is normalized to
1229: the bolometric luminosity of \zquasar. Heating by cosmic microwave
1230: background radiation at different redshifts $z$ is also taken into
1231: account by including a uniform radiation field with temperature of
1232: $T_{\rm cmb}=2.73\times(1+z)$ K.
1233:
1234: The emergent spectrum has a wavelength range of $10^2 - 2\times 10^7
1235: \AA\ $ in the rest frame, and 50 viewing angles (10 in polar angle evenly
1236: divided in ${\rm cos}\theta$ and 5 in azimuthal $\phi$).
1237: We use $10^7$ photon packets isotropically emitted from sources, and the
1238: maximum refinement level of the adaptive grid is RL=12, which is
1239: above the requirement for convergence (see \S~\ref{subsec_res} for
1240: resolution studies).
1241:
1242: The calculated rest-frame SEDs of the system during the peak quasar phase at
1243: $z=6.5$ are shown in Figure~\ref{Fig_sed_z6} (bottom panel). The output SED
1244: depends on the viewing angle. The red curve is the isotropically averaged SED
1245: (e.g., it can be understood as an average SED multiplied by the solid angle
1246: $4\pi$). This averaged SED agrees very well with observations of \zquasar\
1247: \citep{Jiang2006}. A prominent feature of this SED is the two infrared bumps
1248: peaking around $3\, \mu$m and $50\, \mu$m. The $50\, \mu$m bump is produced by
1249: cold dust ($T \sim 50$ K) heated by strong star formation, as commonly seen in
1250: the SEDs of starburst galaxies \citep{Sanders1996, Siebenmorgen2007}. The $3\,
1251: \mu$m bump is produced by the hot dust ($T \sim 1000$ K) heated by the central
1252: AGN. This unique feature appears to be ubiquitous in most of the quasar SEDs
1253: over a wide range of redshifts (e.g., \citealt{Elvis1994, VandenBerk2001,
1254: Telfer2002, Vignali2003, Richards2006, Jiang2006}).
1255:
1256: To demonstrate the line-of-sight dependence, Figure~\ref{Fig_sed_z6} also
1257: shows the range of SEDs viewed from two orthogonal angles of the Cartesian
1258: grid: along the z-axis (upper range) and the xy-plane (lower range),
1259: respectively. This range indicates difference in column density along
1260: the sight lines, and shows that dust extinction in the UV/optical
1261: bands differs by a factor of $\sim 3$, but no difference at wavelengths
1262: longward of $1\, \mu$m. This suggests that the dust is close to optically thin
1263: along these two viewing angles during this time, and that infrared or
1264: submillimeter observations of luminous quasars may not be diagnostics for the
1265: orientation of the host. We have also checked the three major
1266: components of the output SED, namely the scatter, escape and reemission, and
1267: found that the photon energy is conserved. This confirms the photon
1268: conservation algorithm used in the radiative transfer calculation. Moreover,
1269: we find that the emergent SED in the UV/optical bands ($0.01 - 1\, \mu$m) is
1270: dominated by scattering.
1271:
1272: It is interesting to comment on the spectral feature at wavelength
1273: $\lambda_{\rm {rest}} = 9.7\mu$m where the silicate cross section
1274: peaks \citep{Draine2003}. This feature can produce either emission or
1275: absorption in the spectrum depending on the optical depth of the medium at
1276: this wavelength \citep{Bjorkman2001}. In order to have absorption at this
1277: wavelength, the medium has to be optically thick (i.e., $\tau_{\rm
1278: {9.7\mu m}} >> 1$). Generically, this would also result in deep
1279: absorption in the optical/NIR bands ($\sim 0.01 - 1\, \mu$m) owing to
1280: their much higher dust opacity, which is almost two orders of
1281: magnitude higher than that at $9.7\mu$m according to the dust opacity
1282: curve in Figure~\ref{Fig_opac}. Our spectra exhibit absorption when
1283: the system is in the starburst phase (e.g., $z \gtrsim 10$) when the object
1284: is highly obscured. As the system proceeds to the quasar phase (e.g.,
1285: $z < 8$), the emission feature becomes increasingly prominent as dust
1286: becomes more and more transparent to the radiation.
1287:
1288: In observations, both absorption and emission features at
1289: $\lambda_{\rm {rest}} = 9.7\mu$m have been detected for a wide range
1290: of objects up to $z \sim 3$ \citep{Armus2004, Papovich2006}. An
1291: absorption feature is frequently seen in dusty star-forming
1292: galaxies such as M82 (e.g., \citealt{Sturm2000}), Arp 220 (e.g.,
1293: \citealt{Spoon2004}), and ULIRG IRAS08572+3915 which exhibits the most
1294: extreme absorption \citep{Spoon2006}. A comprehensive collection of
1295: the SEDs of starburst nuclei and ULIRGs is reviewed by
1296: \cite{Siebenmorgen2007}, who also provide a library of 7000 SEDs for
1297: dusty galaxies.
1298:
1299: On the other hand, an emission feature is also reported in many
1300: observations \citep{Hao2005, Siebenmorgen2005, Sturm2005}. In
1301: particular, \cite{Hao2007} analyze a sample of 196 local AGNs and
1302: ULIRGs observed by the Infrared Spectrograph (IRS; \citealt{Houck2004}) on
1303: board the Spitzer Space Telescope \citep{Werner2004} to study the distribution
1304: of strengths of the $9.7\mu$m silicate feature. These authors find a wide
1305: range of silicate strengths: quasars are characterized by silicate emission
1306: and Seyfert 1s equally by emission or weak absorption. Seyfert 2s are
1307: dominated by weak silicate absorption, and ULIRGs are characterized by
1308: strong silicate absorption (mean apparent optical depth of about
1309: 1.5). \cite{Spoon2007} find that the same sample of galaxies is
1310: systematically distributed along two distinct branches: one with
1311: AGN-dominated spectra and one with deeply obscured nuclei and
1312: starburst-dominated spectra. These authors suggest that the separation
1313: may reflect a fundamental difference between the dust geometries in
1314: the sources: clumpy for AGNs versus non-clumpy obscuration for
1315: starbursts. For example, \cite{Levenson2007} suggest that the
1316: extremely deep absorption in IRAS08572+3915 requires a source to be
1317: embedded in a smooth distribution of material that is both
1318: geometrically and optically thick. Our simulations show that the dust
1319: distribution around the quasar is quite clumpy, and that the SEDs
1320: changes from starburst to AGN-dominated as the system evolves from the
1321: starburst to quasar phases (see Figure~\ref{Fig_sed_z}). These are
1322: consistent with the observations.
1323:
1324:
1325: \section{Parameter Studies}
1326: \label{sec_param}
1327:
1328: In this Section, we explore the large parameter space involved in this
1329: radiative transfer calculation, and systematically study the resolution
1330: convergence, parameters in the two-ISM model, input spectra for
1331: the BHs, the dust models, and the dust-to-gas ratios for both the cold- and
1332: hot-phase gas.
1333:
1334: \subsection{Resolution Studies}
1335: \label{subsec_res}
1336:
1337: \begin{figure}
1338: \begin{center}
1339: \includegraphics[width=3.2in]{f7a.ps}\\
1340: \vspace{0.5cm}
1341: \includegraphics[width=3.2in]{f7b.ps}
1342: \vspace{1cm}
1343: \caption{
1344: Resolution studies of the number of photons (top panel), and the grid
1345: refinement level (bottom panel), respectively. The top panel shows the SEDs
1346: produced with photon number $N_{\rm ph}=10^4 - 10^9$. These SEDs have similar
1347: shapes, and they converge when $N_{\rm ph} \gtrsim 10^6$. A smaller photon
1348: number results in larger fluctuations in the SED owing to greater Poisson
1349: error. The bottom panel compares the SEDs produced with a uniform grid and those
1350: with adaptive grids of different refinement level. Compared to an adaptive grid
1351: method, SEDs using uniform grids with grid number ${\rm GN}=30^3$ and $50^3$
1352: do not have sufficient dynamic range to resolve both the cold and hot dust, which
1353: result in an underestimate of the dust emission longward of $10\, \mu$m. As
1354: the grid refinement level increases, hot dust is
1355: better resolved,
1356: contributing to the hot dust bump at $1 - 10\, \mu$m. The SEDs converge when the
1357: refinement level goes above 10, which has a minimum cell size approaching
1358: that of the spatial resolution of the original hydrodynamic simulations. Level
1359: 12 is the standard refinement level used in this paper.}
1360: \label{Fig_res}
1361: \end{center}
1362: \end{figure}
1363:
1364:
1365: Figure~\ref{Fig_res} shows resolution studies for photon number (top panel)
1366: and grid size (bottom panel). The SEDs converge when the photon number is
1367: larger than $10^6$. If the photon number is low, then Poisson noise is
1368: significant, which results in large fluctuations in the SEDs. Therefore,
1369: throughout the paper, we use a photon number of $10^7$ for SED production, and
1370: $10^8$ for images in order to have higher signal-to-noise.
1371:
1372: For the adaptive grids, as the refinement level increases, the hot
1373: dust in the central region around the AGN is better resolved,
1374: contributing to the hot dust bump in $1 - 10\, \mu$m. The SEDs converge when
1375: the refinement level goes above 10, which has a minimum cell size close to
1376: the spatial resolution ($\sim$30 pc) of the original hydrodynamic
1377: simulations. Compared to the adaptive grid
1378: method, SEDs using uniform grids with reasonable computing expense
1379: have poor resolution. As demonstrated in Figure~\ref{Fig_grid}, a
1380: $50^3$ uniform grid has a resolution of $4$ kpc, which can only resolve the
1381: dust in the diffuse gas in the outskirts of the system, but not the clumpy,
1382: dense regions around the central AGN. Consequently, the resulting SEDs do not
1383: entirely resolve dust emission in $1 - 10\, \mu$m which comes
1384: partly from hot dust near the AGN, and the cold dust bumps are lower by up to
1385: one order of magnitude.
1386:
1387:
1388: \subsection{ISM Model Parameters}
1389:
1390:
1391: \begin{figure}
1392: \begin{center}
1393: \includegraphics[width=3.5in]{f8.ps}
1394: \vspace{0.5cm}
1395: \caption{Parameter study of the two-phase break down of the ISM with
1396: hot and cold phases. In the legend, HP=1 indicates existence of
1397: hot-phase gas, while CP=10 indicates that the cold pressure is enhanced by a
1398: factor of 10 (see \S~\ref{subsec_ism} for more details). The purple curve
1399: represents the ``no phase break-down'' case in which the cold gas is not
1400: considered; there is only hot-phase gas whose density is the same as that
1401: given directly by the SPH simulations. The blue and cyan curves represent cases
1402: in which no hot-phase gas is present (only cold-phase gas is considered), but the
1403: pressure enhancement factor for cold-phase gas varies from 10 to 100,
1404: respectively. The rest of the colored curves represent cases in which both
1405: hot and cold phases co-exist, with cold gas pressure varying from 1 to
1406: 100. In the RT calculations, we use standard values CP=10 and HP=1.}
1407: \label{Fig_phase}
1408: \end{center}
1409: \end{figure}
1410:
1411:
1412: \begin{figure}
1413: \begin{center}
1414: \includegraphics[width=3.5in]{f9.ps}
1415: \vspace{0.5cm}
1416: \caption{Parameter study of the
1417: power-law indices of the mass spectrum $\alpha$ and
1418: mass-radius relation $\beta$ of the cold clouds (see \S~\ref{subsec_ism} for
1419: more details). The output SED is not sensitive to the ranges of
1420: $\alpha$ and $\beta$ considered here. In the RT simulations, we adopt
1421: $\alpha=1.8$ and $\beta=2.0$ as suggested by observations. }
1422: \label{Fig_MR}
1423: \end{center}
1424: \end{figure}
1425:
1426:
1427: Figure~\ref{Fig_phase} shows the SEDs with various parameters for the
1428: two-phase breakdown of the ISM. When there is no phase break-down
1429: (purple curve), only hot-phase gas is considered (no cold gas). In
1430: this case, the hot gas has both large volume filling factor and
1431: mass fraction, so the extinction is extremely high, which leads
1432: to significant emission in the NIR but little emission in the $20 -
1433: 1000\, \mu$m range (the cold dust is sparse). In cases with cold gas
1434: only (no hot-phase gas), the SEDs show big cold dust bumps but no hot
1435: dust emission, as indicated by the blue and cyan curves. A comparison
1436: of these two curves shows that a larger cold phase pressure leads to a
1437: higher volume filling factor for the cold dust, resulting in stronger
1438: cold dust emission.
1439:
1440: In the cases where both hot and cold phases coexist, the amount of
1441: cold dust emission depends on the cold gas pressure, as indicated by
1442: the other colored curves. We find that cold pressure in the range of
1443: 10 -- 100 is able to produce cold dust bumps that fit the
1444: submillimeter observations of \zquasar. However, because the SED with
1445: CP=10 has both the hot and cold dust bumps that agree better with the
1446: mean SEDs of luminous quasars in the Sloan samples of
1447: \cite{Richards2006}, we choose CP=10 as a standard value in our
1448: calculations.
1449:
1450: Figure~\ref{Fig_MR} shows a study of the output SEDs obtained by
1451: varying the power-law indices of both the mass spectrum and the
1452: mass-radius relation of the cold clouds. Increasing $\alpha$ would
1453: steepen the cloud mass function, leading to more smaller cold clouds
1454: that boost the cold dust bump. A similar effect is seen by increasing
1455: $\beta$. However, the SEDs are not very sensitive to the change of
1456: either $\alpha$ or $\beta$ in the ranges of 1.5 -- 2.5 and 1.5 -- 3.0,
1457: respectively. Changing $\alpha$ from 1.5 to 2.5, or $\beta$ from 1.5
1458: to 3.0 only results in a change in the SED by a factor of about 2. In
1459: the RT calculations, we adopt the standard values of $\alpha=1.8$ and
1460: $\beta=2.0$, as suggested by observations \citep{Blitz2006,
1461: Rosolowsky2007}.
1462:
1463: \subsection{Input Spectrum}
1464:
1465:
1466: \begin{figure}
1467: \begin{center}
1468: \includegraphics[width=3.5in]{f10.ps}
1469: \vspace{0.5cm}
1470: \caption{A parameter study of the input spectrum for the black hole. The solid
1471: curves represent the output SEDs, while the dotted lines represent the input
1472: black hole spectra. The blue and red curves indicate the use of an
1473: intrinsic, broken power-law as in \cite{Marconi2004}, and a composite
1474: spectrum from \cite{Hopkins2007A}, respectively. Both input spectra have the
1475: same bolometric luminosity. A power-law input spectrum would require
1476: pc-scale resolution to resolve the dust near the AGN in order to produce the
1477: near-IR emission, which is below the resolution of our hydrodynamic
1478: simulations. A composite spectrum therefore serves as a sub-resolution
1479: recipe to resolve the dust emission within parsecs of the AGN.}
1480: \label{Fig_inspec}
1481: \end{center}
1482: \end{figure}
1483:
1484:
1485: Figure~\ref{Fig_inspec} shows a comparison of the emergent SEDs using
1486: different input black hole spectrum, namely a broken power-law as in
1487: \cite{Marconi2004} (blue curve), and a composite spectrum from
1488: \cite{Hopkins2007A} (red curve). All other parameters being equal, the SED
1489: using the power-law spectrum shows more emission in the optical bands and
1490: lower emission in the IR than the SED with the composite input spectrum. These
1491: differences owe their origin to differences in the input spectrum. The
1492: composite spectrum includes emission from hot dust residing in the
1493: vicinity of the AGN that is below the resolution limit of our hydrodynamic
1494: simulations. This figure emphasizes the care that must be taken when including
1495: an AGN and performing radiative transfer on scales that are not well-resolved,
1496: and demonstrates that a composite black hole spectrum as in
1497: \cite{Hopkins2007A} provides a viable sub-resolution prescription for the dust
1498: emission contributed within parsecs of the AGN.
1499:
1500: \subsection{Dust Models}
1501: \label{subsec_dustcomp}
1502:
1503: \begin{figure}
1504: \begin{center}
1505: \includegraphics[width=3.5in]{f11.ps}
1506: \vspace{0.5cm}
1507: \caption{
1508: Comparison of the SEDs from using the dust extinction curves of
1509: \cite{Weingartner2001} with $R_{\rm V}=3.1$ (WD model) and that from
1510: Type-II supernovae (SN model). The arrow at $\sim$ 9.43 $\mu$m indicates a
1511: 2$\sigma$ upper limit. \zquasar\ was observed with MIPS at 70 $\mu$m but with
1512: no detection. The arrow indicates a $2\sigma$ upper limit. The current
1513: observations cannot distinguish between these two dust models.}
1514: \label{Fig_dustcomp}
1515: \end{center}
1516: \end{figure}
1517:
1518:
1519: Figure~\ref{Fig_dustcomp} shows the emergent SEDs using dust extinction curves
1520: from the WD model \citep{Weingartner2001} with $R_{\rm V}=3.1$ and our
1521: SN model, respectively. The SEDs agree well at wavelengths $\lambda \gtrsim
1522: 10\, \mu$m, but differ by a factor of a couple in the optical /NIR bands $0.1
1523: - 10\, \mu \rm{m}$. It appears that WD model would require a higher dust-to-gas
1524: ratio than that of the Milky Way by a factor of a few in order to produce the
1525: observed SED of \zquasar. In such case, the WD model would produce a stronger
1526: peak at the $9.7\, \mu$m silicate feature by a similar factor relative to the
1527: SN model. Observation of \zquasar\ at $\sim 9.7\, \mu$m would be helpful in
1528: distinguishing between these two models. However, the current data-point at
1529: that wavelength (observed by MIPS at 70 $\mu$m) is only a $2\sigma$ upper
1530: limit, insufficient to constrain the model.
1531:
1532: As discussed in \S~\ref{subsec_sndust}, \cite{Elvis2002} suggest that
1533: quasars may also be copious producers of dust, as condensation in
1534: quasar outflows is similar to dust formation in the envelopes of AGB
1535: stars. In this case, the dust extinction curve would be similar to the
1536: WD model. However, the contribution of dust from quasar winds in the
1537: early Universe remains unknown. The recent report by
1538: \cite{Stratta2007} of dust extinction in the host galaxy of GRB 050904
1539: at $z = 6.3$ independently supports a supernova-origin dust model.
1540: If both supernovae from starbursts and quasar outflows play important
1541: roles in dust production in these young quasar systems at $z \sim 6$,
1542: then the resulting dust extinction curve would be intermediate to the
1543: WD and SN models shown in Figure~\ref{Fig_dustcomp}. More observations
1544: and deeper surveys for dust in high-z galaxies and quasars will be
1545: necessary to constrain dust formation mechanisms at early
1546: cosmic times.
1547:
1548:
1549: \subsection{Dust-to-gas Ratios}
1550:
1551: \begin{figure}
1552: \begin{center}
1553: \includegraphics[width=3.5in]{f12a.ps}\\
1554: \vspace{0.5cm}
1555: \includegraphics[width=3.5in]{f12b.ps}
1556: \vspace{0.5cm}
1557: \caption{A parameter study of the dust-to-gas ratio relative to Milky Way
1558: value, for the Cold Phase Gas (CPG, top panel) and Hot Phase Gas (HPG,
1559: bottom panel), respectively. This figure shows that the emergent SED is more
1560: sensitive to the dust-to-gas ratio of the HPG than to the CPG owing to the
1561: large covering factor of the HPG. In the regular RT calculations, the
1562: default values are 1 and 1\% Milky Way value for CPG and HPG, respectively.}
1563: \label{Fig_dtg}
1564: \end{center}
1565: \end{figure}
1566:
1567: It has been suggested that ULIRGs have dust-to-gas ratio (mostly cold gas as
1568: in our modeling) close to that of the Milky Way (e.g., \citealt{Dunne2001,
1569: Klaas2001}). While some observations of obscurred, X-ray selected AGNs seem to
1570: suggest that the dust-to-gas ratio in these objects has a wide range, from $\sim
1571: 10^{-3}$ to a few of MW value (e.g., \citealt{Maiolino2001, Kuraszkiewicz2003,
1572: Hall2006}). Although this range may apply mainly to the hot, fully or
1573: partially ionized gas in the circumnuclear regions of the AGNs, it would be
1574: interesting to see how such a range affects the output SED in our calculations.
1575:
1576: Figure~\ref{Fig_dtg} shows the comparsion of the emergent SEDs with different
1577: dust-to-gas ratios in a wide range, for both cold (top panel) and
1578: hot-phase gases (bottom panel), respectively. The values in the plot are
1579: relative to the MW value. A change of two orders of magnitude in the
1580: dust-to-gas ratio of the cold phase gas results in a difference in the cold
1581: dust bump only by a factor of a few. However, a similar change in the
1582: dust-to-gas ratio in the hot phase gas would result in substantial difference
1583: in the output SED. This study shows the extinction is dominated by the hot
1584: dust, which has a much larger covering factor that the cold dust. We find that
1585: using 1 and 1\% of MW value for the cold and hot phase gas, respectively,
1586: reproduce the observation of SDSS J1148 reasonably well. Therefore, we elect
1587: to use them as default values for our standard RT calculations. Note this
1588: choice of 1\% of MW value for the hot, diffuse gas has little effect for the
1589: starburst or ULIRG phase, because a substantial fraction of gas during that
1590: stage is in the cold clouds. However, it may affect the SEDs of major quasar
1591: phase, as the gas is heated by the AGN. We should point out that in our ISM
1592: model, we do not have the warm phase as in the picture of
1593: \cite{McKee1977}. Such a warm phase medium may have a dust-to-gas ratio
1594: similar to that of MW. We will explore such a treatment in future work.
1595:
1596:
1597: \section{The Dust Distribution}
1598: \label{sec_dust}
1599:
1600: \begin{figure}
1601: \begin{center}
1602: \includegraphics[width=3.1in]{f13a.ps} \\
1603: \includegraphics[width=3.1in]{f13b.ps}
1604: \vspace{1cm}
1605: \caption{
1606: Maps of the projected density (top panel) and temperature (bottom
1607: panel) of the HPG-dust (dust associated with hot phase gas). The
1608: origin of the map is the location of the central quasar, and the coordinates
1609: are comoving. Note that the gas only determines the distribution and mass of
1610: the dust; the dust temperature is not associated with the gas temperature but
1611: is calculated self-consistently from the radiation field (see
1612: \S~\ref{subsec_ism} for more details).}
1613: \label{Fig_dustmap}
1614: \end{center}
1615: \end{figure}
1616:
1617: The distribution of dust is essential to investigating dust formation
1618: mechanisms and heating sources. Figure~\ref{Fig_dustmap} shows the projected
1619: spatial distribution of the dust density and temperature in the system at
1620: $z=6.5$. The projected quantity is calculated with $\int f(l){\rm d}l/L$, where
1621: $f(l)$ is density or temperature distribution along the line of sight, while
1622: $L$ is the total length of the sight line. Gravitational torques in the
1623: interaction produce strong shocks and tidal features, making the dust
1624: distribution highly inhomogeneous and clumpy. The dynamic range in density can
1625: be up to six orders of magnitude. Both the dust density and temperature peak
1626: in the central regions near the AGN.
1627:
1628: \begin{figure}
1629: \begin{center}
1630: \vspace{1cm}
1631: \includegraphics[width=3.2in]{f14.ps}
1632: \vspace{1cm}
1633: \caption{
1634: Radial distribution of dust mass (top panel) and temperature (bottom
1635: panel) as a function of the distance from the central AGN. The blue and red
1636: curves represent the CPG- and HPG-dust, respectively, while the black curve in
1637: the bottom panel is a power-law temperature profile, $T \propto
1638: R^{-1/2}$ as expected from heating by the central AGN.}
1639: \label{Fig_rdust}
1640: \end{center}
1641: \end{figure}
1642:
1643: The detailed radial distributions of the dust mass and temperature are
1644: quantified in Figure~\ref{Fig_rdust}. The blue and red curves in this figure
1645: represent CPG- and HPG-dust, dust associated with cold- and hot-phase gas,
1646: respectively, as defined in \S~\ref{subsec_ism}. The CPG-dust has high
1647: density but a small volume filling factor; it is generally optically thick to
1648: radiation. On the other hand, the HPG-dust has a lower density but fills $\gtrsim
1649: 99\%$ of the volume, and is generally optically thin. Both the dust mass and
1650: temperature vary with distance from the central AGN. The cold dust is
1651: typically surrounded by hot dust. In the inner hundred parsecs, the cold cores
1652: are highly condensed and are optically thick even to the hard radiation from
1653: the black hole. However, the HPG-dust at the edges surrounding these cold
1654: cores is heated directly by the central AGN. This is indicated by the
1655: power-law temperature profile in the lower panel of Figure~\ref{Fig_rdust}, $T
1656: \propto R^{-1/2}$ as expected from Equation (\ref{eq:REtemp}): $T^4(R) \propto
1657: E(R) \propto R^{-2}$. The hottest dust is heated up to $\sim 1200$~K, which is
1658: below the dust sublimation temperature $\sim 1600$~K
1659: \citep{Hoenig2006}. Cooler dust ($T \sim$ tens to hundreds K) is distributed
1660: in an extended region from $\sim$ 100 pc to several kiloparsecs. The
1661: heating sources come both from stars and the central AGN. Beyond 10 kpc, the
1662: dust temperature drops to below 100 K. Note that the minimum temperature in
1663: Figure~\ref{Fig_rdust} is $\sim 20$~K, which is the CMB temperature at z=6.5.
1664:
1665: \begin{figure}
1666: \begin{center}
1667: \includegraphics[width=3.2in]{f15.ps}
1668: \vspace{1cm}
1669: \caption{
1670: Mass-temperature distribution of both the CPG- (in blue) and HPG-dust (in
1671: red). In the bottom panel, $M_{\rm cold}$, $M_{\rm warm}$, and $M_{\rm hot}$
1672: represent the total mass of cold-, warm- and hot-dust with $T \lesssim 100$~K,
1673: $100 < T < 1000$~K, and $1000 \lesssim T \lesssim 1200$~K, respectively.}
1674: \label{Fig_mdust}
1675: \end{center}
1676: \end{figure}
1677:
1678: Figure~\ref{Fig_mdust} shows histograms of the dust mass and
1679: temperature. The dust temperature ranges from $\sim 10 - 10^3$ K. The
1680: amount of the hottest dust ($\sim 10^3$ K) is $\sim 1.4\times 10^2\, \Msun$,
1681: while that of the cold dust ($T \lesssim 10^2$ K) is $\sim 1.4\times 10^8\,
1682: \Msun$. This is in agreement with the estimates of the dust detected in the
1683: host of \zquasar\ \citep{Bertoldi2003A, Beelen2006, Jiang2006}.
1684:
1685: The large amount of cold dust ($\sim 1.4\times 10^8\, \Msun$) located
1686: within 3 kpc from the AGN provides efficient cooling for the formation of
1687: molecular gas. In \cite{Narayanan2007}, we calculate carbon monoxide emission
1688: using a non-local thermodynamic equilibrium radiative transfer code
1689: \citep{Narayanan2006b, Narayanan2006c}, and find that CO gas forms in this
1690: region, with a total mass of $\sim 10^{10}\, \Msun$, similar to observations
1691: by \cite{Walter2004}. This suggests that cold dust and molecular gas are
1692: closely associated, and both depend on the star formation history. Our results
1693: show that significant metal enrichment takes place early in the quasar
1694: host, as a result of strong star formation in the progenitors, and that
1695: intense starbursts ($\rm {SFR} \gtrsim 10^3\, \Msun, \yr^{-1}$)
1696: within $\lesssim 10^8$ yr are necessary to produce the observed
1697: properties of dust and molecular gas in the host of \zquasar, a conclusion
1698: which is also supported by the analytical models of \cite{Dwek2007}.
1699:
1700: \section{Evolution of the Quasar System}
1701:
1702: \subsection{Transition from cold to warm ULIRG}
1703:
1704: \begin{figure}
1705: \begin{center}
1706: \includegraphics[width=3.5in]{f16.ps}
1707: \vspace{1cm}
1708: \caption{Evolution of the SEDs of the quasar system and its
1709: galaxy progenitors in the observed frame. The colored curves represent SEDs
1710: from $z \sim 14$ to $z \sim 5.2$, while the black dots are again
1711: observations from \cite{Jiang2006}, as described in the legend. Absorption
1712: of the Lyman line series and continuum by the intergalactic medium
1713: \citep{Madau1995} is taken into account, which results in a sharp drop at
1714: short wavelengths.}
1715: \label{Fig_sed_z}
1716: \end{center}
1717: \end{figure}
1718:
1719:
1720: \begin{figure}
1721: \begin{center}
1722: \includegraphics[width=3.5in]{f17.ps}
1723: \vspace{0.8cm}
1724: \caption{Evolution of rest-frame infrared luminosities of the quasar
1725: system (top panel) and the rest-frame color $f_{25\mu\rm m}/f_{60 \mu\rm m}$
1726: (bottom panel). In the top panel, the solid curves represent the luminosity in
1727: near-IR ($1 - 10\, \mu$m), mid-IR ($10 - 40\, \mu$m), far-IR ($40 - 120\,
1728: \mu$m), and IR ($8 - 1000\, \mu$m), respectively. The grey shades indicate
1729: regimes for LIRG ($\Lir > 10^{11}\, \Lsun$), ULIRG ($\Lir > 10^{12}\,
1730: \Lsun$), and HLIRG ($\Lir > 10^{13}\, \Lsun$),
1731: respectively, as classified by \cite{Sanders1996}. The ratio $f_{25\mu\rm
1732: m}/f_{60 \mu\rm m}$ in the bottom panel is an indicator of the coldness of
1733: the SED with a critical value of 0.3 \citep{Sanders1996}. The quasar system
1734: in our model evolves from ``cold'' ULIRG ($f_{25\mu\rm m}/f_{60 \mu\rm
1735: m}<0.3$) to ``warm'' ULIRG (including HLIRG, $f_{25\mu\rm m}/f_{60 \mu\rm m}
1736: \ge 0.3$) as it transforms from starburst to quasar phases.}
1737: \label{Fig_LIR_color}
1738: \end{center}
1739: \end{figure}
1740:
1741:
1742: In \cite{Li2007}, we show that the host of the $z\sim 6$ quasar
1743: undergoes hierarchical mergers starting from $z\sim 14$, and the
1744: quasar descends from starburst galaxies. The evolution of the SEDs of
1745: the system in the observed frame is shown in
1746: Figure~\ref{Fig_sed_z}. Note that absorption of the Lyman lines and
1747: continuum by the intergalactic medium \citep{Madau1995} is taken into
1748: account. During early stages at $z \sim 14$, the SED of the quasar
1749: progenitor shows only a cold dust bump that peaks around $\sim
1750: 60\mu$m, which is characteristic of starburst galaxies
1751: \citep{Sanders1996}. As the system evolves from starburst to quasar
1752: phases, dust extinction in the UV-optical bands increases, boosting
1753: the emission reprocessed by dust at wavelengths longward of $1\,
1754: \mu$m. The gradually increasing radiation from the accreting black
1755: holes heats the nearby dust to high temperatures, contributing to
1756: the hot dust bump which peaks around $\sim 3\mu$m (rest frame). At the maximum
1757: quasar phase at $z \simeq 6.5$ indicated by the red curve, the hot
1758: dust bump SED reaches its peak with a temperature of $\sim 1200$~K, as
1759: we have seen in the previous section. Such an SED represents
1760: luminous, blue quasars in the samples of \cite{Jiang2006} and
1761: \cite{Richards2006}. As the system ages and reddens in the post-quasar
1762: phase (indicated by the black curve), the total luminosity of the
1763: system drops. However, there is still some residual hot dust,
1764: and the infrared luminosity is dominated by the NIR and MIR. This
1765: resembles the class of infrared-bright, optically-red quasars found in
1766: recent surveys (e.g., \citealt{Brand2006}). Overall, the evolution of
1767: the SED from a starburst to a quasar can be characterized by the slope of
1768: the infrared SED ($3 - 50\, \mu$m), as it decreases from the starburst
1769: to quasar phases owing to the increase of NIR emission from the hot
1770: dust heated by the AGN.
1771:
1772: Infrared luminosities are powerful tools to study starburst galaxies and
1773: quasars. From the results in the previous sections, we see that the emission
1774: in near-IR ($1 - 10\, \mu$m), mid-IR ($10 - 40\, \mu$m) and far-IR ($40 -
1775: 120\, \mu$m) comes from re-emission by hot, warm, and cold dust,
1776: respectively. Note that the meaning of FIR varies in the literature. Here we
1777: use the definition of FIR given by \cite{Condon1992} in order to compare our
1778: results with the observations by \cite{Carilli2004} who used the same FIR
1779: range.
1780:
1781: Using $\Lir$ ($8 - 1000\, \mu$m), \cite{Sanders1996}
1782: classified infrared luminous galaxies into three categories, namely luminous
1783: infrared galaxy (LIRG, $\Lir > 10^{11}\, \Lsun$), ultra-LIRG (ULIRG,
1784: $\Lir > 10^{12}\, \Lsun$), and hyper-LIRG (HLIRG, $\Lir > 10^{13}\,
1785: \Lsun$). Figure~\ref{Fig_LIR_color} (top) shows the evolution of the infrared
1786: luminosities of the quasar system in our simulations. The luminosities
1787: increase with the star formation rate and black hole accretion rate. The
1788: system is ultraluminous most of the time, with periods in HLIRG phases
1789: associated with bursts of star formation or quasar activity. When the last
1790: major mergers take place between $z \sim 9 - 7.5$, the strong shocks and highly
1791: concentrated gas fuel rapid star formation and black hole growth. This
1792: also produces a large amount of dust heated by the central AGN and
1793: stars. As a result, the infrared luminosities increase
1794: dramatically, pushing the system to HLIRG class.
1795:
1796: The infrared luminosities strongly peak at $z\sim 8.7$ when the
1797: star formation rate reaches $\gtrsim 10^4\, \Msun\, \yr^{-1}$,
1798: suggesting a significant contribution from stars in heating the dust
1799: to emit in the range $1 - 1000\, \mu$m. During the major quasar phase
1800: ($z \sim 7.5 - 6$) the star formation declines to $\sim 10^2\,
1801: \Msun\, \yr^{-1}$, and the emission in NIR and MIR outshines that in FIR,
1802: demonstrating that AGN can play a dominant role in heating the dust and
1803: producing NIR and MIR emission. Furthermore, the AGN can also contribute
1804: to FIR emission. For example, while the star formation rate drops by a
1805: factor of $\sim 500$ from $z\sim 8.7$ (when star formation rate peaks)
1806: to $z\sim 6.5$ (when black hole accretion rate peaks, see
1807: Figure~\ref{Fig_quasar}), the $\Lfir$ declines by only a factor of $\sim
1808: 50$, indicating substantial contribution to the $\Lfir$ by the AGN.
1809: This significant AGN contribution has important implications for estimating
1810: the star formation rate. We will discuss this issue at length in the next
1811: section. During the ``post-quasar'' phase at $z<6$, as a result of
1812: feedback which suppresses both star formation and black hole
1813: accretion, and gas depletion which reduces the amount of dust, the
1814: infrared luminosities drop rapidly.
1815:
1816: The flux ratio $f_{25\mu\rm m}/f_{60 \mu\rm m}$ is a color indicator
1817: for the coldness of the SED, and can be used as a diagnostic for AGN
1818: activity, as suggested by \cite{Sanders1996}. For example, starburst
1819: galaxies usually have cold colors $f_{25\mu\rm m}/f_{60 \mu\rm
1820: m}<0.3$, while quasars are warm with $f_{25\mu\rm m}/f_{60 \mu\rm
1821: m}\gtrsim0.3$. The evolution of the color $f_{25\mu\rm m}/f_{60
1822: \mu\rm m}$ of the simulated quasar system is shown in
1823: Figure~\ref{Fig_LIR_color} (bottom panel). The color index is below
1824: 0.3 most of the time, i.e. during both starburst- and post-quasar
1825: phases. However, it rises above 0.3 during the short quasar
1826: phase. This figure clearly demonstrates that the system evolves from a
1827: cold to warm ULIRG (including HLIRG) as it transforms from starburst
1828: to quasar phases. Similar trend is also reported by
1829: \cite{Chakrabarti2007A}. Our results provide further theoretical support for
1830: the starburst-to-quasar conjecture, as suggested by observations
1831: \citep{Sanders1996, Scoville2003}.
1832:
1833:
1834: \subsection{AGN Contamination and the {\rm SFR} -- $\Lfir$ Relation}
1835:
1836: \begin{figure}
1837: \begin{center}
1838: \includegraphics[width=3in]{f18.ps}
1839: \vspace{1cm}
1840: \caption{Contribution to the infrared luminosities, $\Lnir$, $\Lmir$, and
1841: $\Lfir$, from both AGN and stars, respectively. During the
1842: starburst phase, stars are the main heating source for the infrared
1843: emission. However, during the peak quasar phase, AGN heating dominates.
1844: }
1845: \label{Fig_lir_stars_bhs}
1846: \end{center}
1847: \end{figure}
1848:
1849: \begin{figure}
1850: \begin{center}
1851: \includegraphics[width=3.2in]{f19a.ps} \\
1852: \vspace{0.5cm}
1853: \includegraphics[width=3.2in]{f19b.ps} \\
1854: \vspace{0.5cm}
1855: \includegraphics[width=3.2in]{f19c.ps} \\
1856: \vspace{1cm}
1857: \caption{Evolution of the SFR -- $\Lfir$ relation in our
1858: simulations. Here we consider the relation in three cases where $\Lfir$ is
1859: contributed by AGNs only (case 1, top panel), stars only (case 2, middle
1860: panel), and both AGNs and stars (case 3, bottom panel). The colored filled
1861: symbols indicate the system at different redshifts, while the black open diamond
1862: represents the model quasar at $z=6.5$. The black curve in the middle panel
1863: is the least-squares fit to all the data, while that in the bottom panel is the
1864: fit to data points from $z \sim 14 - 7.5$ only (starburst phase). Case 1 has no
1865: SFR -- $\Lfir$ correlation; case 2 has a tight, linear correlation similar to
1866: that used in observations; and case 3 has a non-linear correlation.}
1867: \label{Fig_sfr_lir}
1868: \end{center}
1869: \end{figure}
1870:
1871: AGN contamination in infrared observations of dusty, star-forming
1872: quasar systems has been a long-standing problem. In our simulations, we
1873: see that the contributions from AGNs and stars both vary according to
1874: the activity of these two populations. As shown in
1875: Figure~\ref{Fig_lir_stars_bhs}, during the starburst phase, not
1876: surprisingly nearly all the infrared emission comes from dust heated
1877: by stars. However, during the peak quasar phase, the contribution from
1878: the AGN dominates the infrared light production in the system. This result
1879: has significant implications for the interpretation of observational
1880: data, such as estimates of the star formation rate.
1881:
1882: In particular, owing to the lack of other indicators such as UV flux
1883: or $\rm H{\alpha}$ emission, in observations of high-redshift objects
1884: the far-infrared luminosity is commonly used to estimate the star
1885: formation rate, assuming most or all of $\Lfir$ is contributed by
1886: young stars. For example, \zquasar\ has $\Lfir \sim 2\times 10^{13}\, \Lsun$,
1887: and the star formation rate is estimated to be $\sim 3\times 10^3\,
1888: \Msun\, \yr^{-1}$ \citep{Bertoldi2003A, Carilli2004}. However, these
1889: studies cannot rule out a substantial contribution (e.g. $\sim 50\%$) to
1890: $\Lfir$ by the AGN. If $\Lfir$ is indeed heavily contaminated by AGN
1891: activity, then the SFR -- $\Lfir$ relation would be invalid, and the actual
1892: star formation rate would be much lower.
1893:
1894: To demonstrate this, we show in Figure~\ref{Fig_sfr_lir} the SFR -- $\Lfir$
1895: relation in three cases: (1) only black holes are included as a
1896: heating source of the dust (top panel); (2) only stars are included (middle
1897: panel); and (3) both black holes and stars are included (bottom panel). The
1898: top panel shows that there is no correlation between the SFR and the $\Lfir$
1899: which is produced by AGNs. However, if all the $\Lfir$ is
1900: produced solely from stars, then there is a tight, linear correlation. If the
1901: $\Lfir$ is contributed by both AGN and stars, then the correlation would
1902: change. The bottom panel in Figure~\ref{Fig_sfr_lir} shows that during the
1903: starburst phase where stars dominate dust heating, there is
1904: still a correlation between SFR and $\Lfir$ with modified slope and
1905: normalization. However, during the peak quasar phase where the
1906: AGN dominates dust heating, there is no correlation at all. To summarize,
1907: Figure~\ref{Fig_sfr_lir} gives the following SFR -- $\Lfir$ relations:
1908:
1909: \begin{eqnarray}
1910: \label{eq:SFR}
1911: \frac{\rm SFR}{\Msun\, \yr^{-1}} & = & 2.3\times 10^{-10}\left(\frac{L_{\rm
1912: FIR}}{\Lsun}\right)^{1.01}\; (\rm {stars-only}) \, ,\\
1913: \frac{\rm SFR}{\Msun\, \yr^{-1}} & = & 4.9\times 10^{-7}\left(\frac{L_{\rm
1914: FIR}}{\Lsun}\right)^{0.76}\; (\rm {AGN + stars}) \, .
1915: \end{eqnarray}
1916:
1917: Here, the FIR is in the range of $40 - 120\mu$~m (rest frame). This
1918: correlation in the stars-only case is very close to that found by
1919: \cite{Kennicutt1998B}, $\frac{\rm SFR}{\Msun\, \yr^{-1}} \simeq 1.7\times
1920: 10^{-10}\frac{\Lfir}{\Lsun}$ for starburst galaxies. Note the slight
1921: difference in the normalization might owe to different stellar IMFs used --
1922: we use the top heavy Kroupa IMF \citep{Kroupa2002}, while \cite{Kennicutt1998B}
1923: adopts a Salpeter IMF \citep{Salpeter1955}. Also, the $\Lfir$ in that work
1924: refers the infrared luminosity integrated over the range of $8 -
1925: 1000\mu$~m.
1926:
1927:
1928: In our simulation, the model quasar at $z=6.5$ has $\Lfir \simeq
1929: 1.8\times10^{13}\, \Lsun$, which is close to that of \zquasar\ derived
1930: from radio observation \citep{Bertoldi2003A, Carilli2004}. If we assume a
1931: perfect SFR -- $\Lfir$ correlation as in the middle panel of
1932: Figure~\ref{Fig_sfr_lir} (Equation~[\ref{eq:SFR}]), then the SFR would be $\sim
1933: 4\times 10^3\, \Msun\, \yr^{-1}$, similar to the estimate of \cite{Bertoldi2003A,
1934: Carilli2004}. However, the SFR of the model quasar from our
1935: simulations is about $\sim 112\, \Msun\, \yr^{-1}$, which is significantly
1936: lower than that derived from an ideal SFR -- $\Lfir$ correlation. This
1937: discrepancy is largely due to significant contribution to
1938: the $\Lfir$ by AGN activity. Several studies by \cite{Bertoldi2003A,
1939: Carilli2004} and \cite{Wang2007} find that most $z \sim 6$ quasars
1940: do not seem to show a $\Lfir$ -- $\LB$ correlation, but instead follow a
1941: strong $\Lfir$ -- $\Lradio$ correlation as measured in local star-forming
1942: galaxies \citep{Condon1992}. These authors therefore suggest that these $z
1943: \sim 6$ quasars are strong starbursts, and that most of the $\Lfir$ may come
1944: from stars. However, the $\Lfir$ -- $\Lradio$ correlation may not guarantee the
1945: stellar origin of the $\Lfir$ as the physical basis for this correlation is
1946: unknown. Our results suggest that in quasar systems, $\Lfir$ alone may no
1947: longer be a reliable estimator for star formation rates; other diagnostics
1948: should also be considered. We will study this topic with more detail in a
1949: future paper by investigating the correlations in a multiple luminosity plane,
1950: $\Lb$ -- $\Lfir$ -- $\Lx$ -- $\Lradio$, of starburst galaxies and quasars.
1951:
1952: \subsection{Evolution of the Dust Properties}
1953:
1954: \begin{figure}
1955: \begin{center}
1956: \includegraphics[width=3.5in]{f20.ps}
1957: \vspace{0.5cm}
1958: \caption{Evolution of the dust mass and temperature as a function of
1959: redshift. The dust is heated by both stars and AGN. As the system evolves
1960: from starburst to quasar phases, the amount of hot dust increases. The dust
1961: reaches the highest temperature during the peak quasar phase. After that, the
1962: dust cools owing to the decline of AGN radiation.}
1963: \label{Fig_mdust_z}
1964: \end{center}
1965: \end{figure}
1966:
1967:
1968: Figure~\ref{Fig_mdust_z} shows the evolution of the dust properties,
1969: including the dust mass and temperature, as a function of redshift. As
1970: the system evolves from starburst to quasar phases, the amount of hot
1971: dust increases accordingly owing to the enhanced heating from the
1972: central AGN, as well as replenishment of gas /dust from incoming new
1973: galaxy progenitors during mergers. The amount of cold dust
1974: ($T \lesssim 100$ K) reaches a maximum at $z \sim 8.8$ when the last major
1975: merger takes place, then decreases steadily as the dust is heated to
1976: higher temperatures, or as the gas is consumed by star formation and black
1977: hole accretion. Strong star formation is able to heat the dust nearly
1978: to $T \sim 1000$~K. However, the hottest dust at $T \sim 1200$~K
1979: is associated only with the peak quasar phase at $z \sim 6.5$ when the hot
1980: dust is directly heated by the AGN. After that, the amount of hot dust drops
1981: dramatically as a result of the rapidly declining radiation from the
1982: central AGN.
1983:
1984: \begin{figure}
1985: \begin{center}
1986: \includegraphics[width=3.0in]{f21a.ps}\\
1987: \vspace{0.5cm}
1988: \includegraphics[width=3.0in]{f21b.ps}
1989: \vspace{0.8cm}
1990: \caption{Relations between rest-frame flux $f_{3 \mu\rm m}$ and hot dust
1991: mass $M_{\rm hot}$ (top panel), and $f_{50 \mu\rm m}$) and cold dust mass
1992: $M_{\rm cold}$ (bottom panel). The black curve is the least-squares fit to
1993: all the data in top panel, and data during starburst phase at redshift
1994: $z\sim 14 - 7.5$ in bottom panel.}
1995: \label{Fig_f3_f50_mdust}
1996: \end{center}
1997: \end{figure}
1998:
1999: In observations, the dust mass is usually determined by assuming that the dust
2000: radiates as a black- or grey-body \citep{Hughes1997, Jiang2006}. In the
2001: emergent SEDs from our radiative transfer calculations, the cold dust
2002: bumps at different redshifts appear to peak around $\sim 50 \mu\rm m$
2003: (rest-frame), while the hot dust bumps peak around $\sim 3 \mu\rm
2004: m$. Figure~\ref{Fig_f3_f50_mdust} shows the relations between the dust mass
2005: and fluxes at these two wavelengths. The top panel shows that $M_{\rm hot}$
2006: ($T \gtrsim 10^3$~K) increases with $f_{3\mu \rm m}$ flux, while $M_{\rm
2007: cold}$ ($T \lesssim 10^2$~K) correlates with $f_{50\mu \rm m}$
2008: flux linearly during the starburst phase:
2009:
2010: \begin{eqnarray}
2011: \frac{M_{\rm hot}}{\Msun} & = & 4.2\times 10^{2}\left(\frac{f_{3\mu \rm m}}{\rm
2012: mJy}\right)^{1.22} \,, \\
2013: \frac{M_{\rm cold}}{\Msun}& = & 6.0\times 10^{8}\left(\frac{f_{50\mu \rm
2014: m}}{\rm mJy}\right)^{0.98}\,.
2015: \end{eqnarray}
2016:
2017: However, during the quasar phase, there appears to be an excess in
2018: $f_{50\mu \rm m}$, so the $M_{\rm cold}$ -- $f_{50\mu \rm m}$
2019: correlation does not apply. These results suggest that rest-frame
2020: $3\mu \rm m$ and $50\mu \rm m$ might serve as good diagnostics for
2021: dust, the former for hot dust in quasar systems, and the latter for
2022: cold dust in starburst galaxies.
2023:
2024: \begin{figure}
2025: \begin{center}
2026: \includegraphics[width=3.0in]{f22.ps}
2027: \vspace{1cm}
2028: \caption{Evolution of the relation between rest-frame 3.5 $\mu$m
2029: luminosity and B-Band luminosity (4400 \AA) for the quasar system in our
2030: simulations. The filled circles represent starburst phases, squares
2031: represents quasar phases, and triangles represent post-quasar phases. Colors
2032: indicates redshift of the object. }
2033: \label{Fig_NIR_B}
2034: \end{center}
2035: \end{figure}
2036:
2037: It has been suggested by \cite{Jiang2006} that the NIR-to-optical flux
2038: ratio may be used to probe dust properties. From a sample of
2039: thirteen $z\sim 6$ quasars observed with {\em Spitzer}, these authors
2040: find that two quasars have a remarkably low flux ratio (rest-frame
2041: $3.5\mu \rm m$ to B-band) compared to other quasars at different
2042: redshifts, and they are also weak in FIR. Furthermore, such a low flux
2043: ratio was not seen in low-redshift quasar counterparts. These findings lead to
2044: the suggestion that these two quasars may have different dust properties from
2045: others \citep{Jiang2006}.
2046:
2047: From our model, we find that the dust properties are associated with the
2048: evolutionary stages of the host. Figure~\ref{Fig_NIR_B} shows the evolution of
2049: the ratio $\Lnir/\Lb$ from the quasar system at different stages of its
2050: life. Early on, during the starburst phase, the ratio $\Lnir/\Lb$ is low. As
2051: the system proceeds to its peak quasar activity, the heating from
2052: the central AGN increases the hot dust emission around $3\, \mu$m,
2053: boosting the $\Lnir/\Lb$ ratio. After that, in the post-quasar
2054: phase, as the radiation from the AGN declines, the hot dust emission
2055: drops rapidly, resulting in a lower $\Lnir/\Lb$.
2056: According to our model, there might be two possible explanations for the two
2057: outliers in the \cite{Jiang2006} sample. One possibility is that they may be
2058: young quasars that are still in the starburst phase but have not yet reached
2059: peak quasar activity, so the light from star formation may be dominant or
2060: comparable to that from the accreting SMBH still buried in dense gas. This may
2061: explain the low NIR flux, as well as the B-band luminosity and the narrow
2062: ${\rm Ly}{\alpha}$ emission line, which are primarily produced by the
2063: starburst. From Figure~\ref{Fig_LIR_color}, we note that there are ``valleys''
2064: between major starbursts (or mergers) where the infrared luminosities are
2065: low. This may explain the weak FIR in these two quasars. Another possibility
2066: is that these two objects may be in the post-quasar phase, which would also
2067: have low $\Lnir/\Lb$ and FIR (see also Figure~\ref{Fig_LIR_color} and
2068: \ref{Fig_sfr_lir}).
2069:
2070: However, our model predicts that the same evolution should apply to every
2071: luminous quasar formed in gas-rich mergers. It cannot explain statistically
2072: why such a low $\Lnir/\Lb$ ratio is only found in the $z \sim 6$ quasars, but
2073: not at low redshifts. Perhaps these two quasars indeed have unusual dust
2074: properties that cannot be explained by our current model. More observations of
2075: dust in high-z objects are needed to resolve the issues surrounding
2076: the IR-weak quasars in \cite{Jiang2006}, and to test our model.
2077:
2078:
2079: \section{Discussion}
2080:
2081: Our three-dimensional Monte Carlo radiative transfer code ART$^2$
2082: makes it possible to calculate self-consistently the radiative transfer and
2083: dust emission from galaxies and quasars which have large dynamic ranges and
2084: irregular geometries. It works efficiently on hydrodynamics simulations of
2085: galaxies and mergers, and has flexible grid resolution that can be
2086: adjusted to any SPH resolution in hydrodynamic simulations.
2087:
2088: To date, only a limited number of RT calculations have been performed on
2089: galaxy mergers \citep{Jonsson2006, Chakrabarti2007A, Chakrabarti2007B}. In
2090: particular, the parallel SUNRISE code developed by \cite{Jonsson2006} marked a
2091: milestone in this direction. It has an adaptive grid similar to ours
2092: and works efficiently on the arbitrary geometry in galaxy mergers. However,
2093: radiative equilibrium is not yet included in this code, so SUNRISE can not
2094: currently calculate dust emission self-consistently.
2095:
2096: On the other hand, \cite{Chakrabarti2007A, Chakrabarti2007B} had a similar
2097: approach to ours. They apply the code of \cite{Whitney2003A}, which employs
2098: the same radiative equilibrium algorithm as in \cite{Bjorkman2001}, to
2099: hydrodynamic simulations of galaxy mergers with black holes. However, the
2100: spherical, logarithmically-spaced grid used by these authors is not optimal
2101: for describing an inhomogeneous distribution of gas and dust with multiple
2102: density centers in galaxies and mergers, as exemplified by
2103: Figure~\ref{Fig_grid}. Moreover, our methodology is based on a treatment of
2104: the multiphase ISM that includes extinction from dust in the diffuse phase,
2105: which was ignored in the work of \cite{Chakrabarti2007A, Chakrabarti2007B}.
2106: Furthermore, we adopt an AGN input spectrum that includes emission from dust
2107: on scales that the simulations do not resolve, unlike \cite{Chakrabarti2007A,
2108: Chakrabarti2007B} who employ a power-law AGN spectrum. The tests
2109: described above demonstrate that the near- and mid-IR emission is sensitive to
2110: the resolution of the RT calculation (Figure~\ref{Fig_res}),
2111: extinction by dust in the diffuse phase of the ISM (Figure~\ref{Fig_phase}),
2112: and the template spectrum of the AGN (Figure~\ref{Fig_inspec}). The SEDs in
2113: the work of \cite{Chakrabarti2007A, Chakrabarti2007B} show only cold dust
2114: emission characteristic of a starburst during all evolutionary phases, even
2115: when the quasar peaks (e.g., Figure~20 in \citealt{Chakrabarti2007B}), and,
2116: in particular, do not at any time exhibit hot dust emission. We
2117: attribute the differences between our computed SEDs and those of
2118: \cite{Chakrabarti2007A, Chakrabarti2007B} mainly to the different grid method
2119: employed, the handling of extinction by the diffuse ISM, and our choice of an
2120: AGN input spectrum that includes emission on scales that cannot be resolved in
2121: our hydrodynamic simulations.
2122:
2123: In our radiative transfer calculations, we do not explicitly include a dust
2124: torus near the AGN. The resolution of our RT grid is constrained by
2125: that of the hydrodynamic simulations, which is of $\sim 30$~pc. As a result,
2126: we can not resolve the near vicinity of the AGN. In the Unification Scheme
2127: for AGN \citep{Miller1983}, a parsec-scale dust torus, both optically and
2128: geometrically thick, has been used to successfully explain the different
2129: appearances of Type 1 and Type 2 AGNs. For example, Type 1 AGNs are typically
2130: viewed face-on and show blue UV bump in their spectra, while Type 2 AGNs are
2131: viewed edge-on without the blue UV bump. It is believed that the NIR and MIR
2132: emission of an AGN comes from the hot dust in the torus structure, as
2133: suggested by various observations, for examples, the MIR and NIR detection of a
2134: Seyfert 2 galaxy NGC 1068 \citep{Jaffe2004, Wittkowski2004}, and the MIR
2135: emission of a lensed quasar Q2237+0305 by \cite{Wyithe2002} in the CfA
2136: Redshift Survey \citep{Huchra1985}. The radiative transfer modeling by
2137: \cite{Hoenig2006}, which concentrates on the inner parsecs of the AGN, has
2138: successfully reproduced the SED of NGC 1068 in the NIR and MIR with a
2139: three-dimensional, clumpy dust torus. Such a scale is, however, below the
2140: resolution of our modeling, which follows the dynamical evolution of the
2141: quasar system on a galactic scale. Moreover, it is not clear whether such a
2142: well-defined torus structure would exist in luminous quasars at high
2143: redshifts. Nevertheless, we have adopted a composite AGN spectrum
2144: \citep{Hopkins2007A} as a sub-resolution recipe for the IR emission from the
2145: hottest dust near the AGN. Such an approach is supported by the fact
2146: that the composite spectrum represents the average near- to mid-IR emission in
2147: thousands of quasars \citep{Richards2006}, and that the torus does not
2148: contribute much to the far-IR emission, which comes mainly from cold dust on
2149: kiloparsec scales. Therefore, for spatial scales of interest, the
2150: results should be reliable, and they are expected be similar to those from
2151: calculations with a torus included. We defer such a calculation to the future
2152: when sub-parsec resolution is feasible in hydrodynamic simulations of quasar
2153: formation.
2154:
2155: We should point out that dust destruction is not explicitly included in our
2156: modeling. It has been suggested (e.g., \citealt{McKee1989, Draine1993,
2157: Draine2003, Nozawa2006}) that dust destruction might be efficient in
2158: non-radiative shocks by non-thermal and thermal sputtering owing to a high
2159: shock velocity ($> 100$ km s$^{-1}$) and a high gas temperature ($> 10^6$
2160: K). However, the sputtering timescale for this temperature is of $\sim 3\times
2161: 10^6$ yr for a gas density $n=1\, {\rm cm}^{-1}$, and even shorter if the gas
2162: density is higher \citep{Burke1974, Reynolds1997}. It has also been suggested
2163: that dust grains may be destroyed behind radiative shocks
2164: \citep{Todini2001, Bianchi2007}. However, it is expected that the efficiency cannot
2165: be significant at high redshifts as magnetic fields are likely weak
2166: \citep{Gnedin2000}. In our dust treatment in the RT analysis, the
2167: dust is re-calculated based on the gas content in every snapshot
2168: from the hydrodynamic simulations, which has a time interval of 2
2169: Myr. This is equivalent to dust destruction within this
2170: timescale. Moreover, we assume different dust-to-gas ratios for the
2171: cold and hot gas, which accounts for dust survival in
2172: these different phases of the interstellar medium.
2173:
2174: One limitation of our ART$^2$ code is that it does not include the transient
2175: heating of small dust grains (size $<200${\AA}), which may cause temperature
2176: fluctuation and produce enhanced near-infrared emission \citep{Li2001,
2177: Misselt2001}, in particular for the PAH feature. Although PAH is not
2178: studied in this present work, the code can be improved in the future by
2179: implementing a temperature distribution for those small grains. Moreover,
2180: one potential caveat of our RT work is that it is a post-processing
2181: procedure. It would be ideal to couple the radiation with
2182: hydrodynamics; i.e., having the radiative transfer done simultaneously
2183: with the hydrodynamic simulations (as in e.g., \citealt{Yoshida2007})
2184: as the dynamical evolution of the gas might affect the formation and
2185: properties of the dust. However, this effect is currently uncertain because
2186: such a technique is not yet available for dust emission in galaxy
2187: simulations. Although ART$^2$ is not parallelized, it works fairly efficiently
2188: owing to the adaptive refinement tree.
2189:
2190: We have done extensive resolution studies which include both the number of
2191: photon packets and the refinement level of the grid, and parameters for the
2192: ISM model such as the cloud mass spectrum and size distribution, as well as
2193: different dust models. We find convergence of the modeled SEDs with parameters
2194: within observational ranges (see \S~2 and \S~3). Therefore, we conclude that
2195: the RT calculations we present here with ART$^2$ are robust, and that our
2196: model is more self-consistent, more realistic, and more versatile than
2197: previous approaches.
2198:
2199:
2200: \section{Summary}
2201:
2202: We have implemented a three-dimensional Monte Carlo radiative transfer
2203: code, ART$^2$ -- All-wavelength Radiative Transfer with Adaptive Refinement
2204: Tree, and use it to calculate the dust emission and multi-wavelength
2205: properties of quasars and galaxies. ART$^2$ includes the following essential
2206: implementations:
2207:
2208: \begin{enumerate}
2209:
2210: \item A radiative equilibrium algorithm developed by \cite{Bjorkman2001}. It
2211: conserves the total photon energy, and corrects the dust temperature
2212: without iteration. This algorithm calculates dust emission efficiently and
2213: self-consistently.
2214:
2215: \item An adaptive grid scheme in 3-D Cartesian coordinates similar to that of
2216: \cite{Jonsson2006}, which handles an
2217: arbitrary geometry and covers a large dynamic range over several orders of
2218: magnitude. This is indispensable for capturing the inhomogeneous and clumpy
2219: density distribution in galaxies and galaxy mergers. It easily achieves a
2220: $\sim$10 pc-scale grid resolution equivalent to a $\sim 10000^3$ uniform grid,
2221: which is prohibitive with current computation schemes.
2222:
2223: \item A two-phase ISM model in which the cold, dense clouds are embedded in
2224: a hot, diffuse medium in pressure equilibrium
2225: \citep{Springel2003A}. Moreover, the cold clouds follow a mass spectrum and a
2226: size distribution similar to the \cite{Larson1981} scaling relations
2227: inferred for giant molecular clouds. This model ensures an appropriate
2228: sub-grid recipe for the ISM physics, which is important for studying dust
2229: properties in galaxies.
2230:
2231: \item A supernova-origin dust model in which the dust is produced by Type-II
2232: supernovae, and the size distribution of grains follows that derived by
2233: \cite{Todini2001}. This model may be especially relevant for dust in
2234: high-redshift, young objects. In traditional dust models, the dust is
2235: produced by old, low mass stars such as AGB stars, which are typically over
2236: 1 Gyr old. However, quasar systems at $\sim 6.4$ are only a few hundred
2237: Myr old, and there would be insufficient AGB stars to produce the abundant dust
2238: as observed in \zquasar.
2239:
2240: \item The input spectra include those from both stars, calculated
2241: using STARBURST99 \citep{Leitherer1999, Vazquez2005}, and black holes
2242: represented by a composite AGN template developed by
2243: \cite{Hopkins2007A}. The composite spectrum includes a broken power-law
2244: as the intrinsic black hole spectrum, and infrared components that come from
2245: the torus near the AGN. The torus structure is unresolved in our
2246: hydrodynamic simulations. This template is a sub-grid recipe to include the
2247: hot dust emission within pc-scales which is below the resolution
2248: of our hydrodynamic simulations.
2249:
2250: \end{enumerate}
2251:
2252: ART$^2$ works efficiently on hydrodynamic simulations of galaxies and
2253: quasars performed with GADGET2 \citep{Springel2005D}. By applying
2254: ART$^2$ to the quasar simulations of \cite{Li2007}, in which luminous
2255: quasars at $z \gtrsim 6$ form rapidly through hierarchical mergers of
2256: gas-rich galaxies, we are able to calculate the multi-wavelength SEDs (from
2257: optical to millimeter) and dust properties of the model quasar at $z
2258: \sim 6.5$ and its galaxy progenitors at even higher redshifts.
2259:
2260: We find that a supernova-origin dust model may be able to explain the
2261: dust properties as observed in the high-redshift quasars. Our
2262: calculations reproduce the observed SED and properties such as the
2263: dust mass and temperature of \zquasar, the most distant Sloan quasar.
2264: The dust and infrared emission in quasar hosts are closely associated
2265: with the formation and evolution of the system. As the system
2266: transforms from starburst to quasar phases, the evolution of the SEDs
2267: is characterized by a transition from a cold to warm ULIRG. During the
2268: starburst phase at $z \gtrsim 7.5$, the SEDs of the quasar progenitors
2269: exhibit cold dust bumps (peak at $\sim 50\, \mu$m rest frame) that are
2270: characteristic of starburst galaxies. During the peak quasar phase ($z
2271: \sim 7.5 - 6$), the SEDs show a prominent hot dust bump (peaks at $\sim
2272: 3\, \mu$m rest frame) as observed in luminous quasars at both low and high
2273: redshifts (e.g, \citealt{Richards2006, Jiang2006}).
2274:
2275: Furthermore, we find that during the quasar phase, AGN activity
2276: dominates the heating of dust and the resulting infrared
2277: luminosities. This has significant implications for the interpretation
2278: of observables from the quasar host. The hottest dust ($T \gtrsim
2279: 10^3$~K) is present only during the peak quasar activity, and
2280: correlates strongly with the near-IR flux. The SFR -- $\Lfir$
2281: correlation depends sensitively on the relative heating contributed from AGN
2282: and stars. If the dust heating is dominated by stars, as in the
2283: starburst phase, then there is a tight, linear SFR -- $\Lfir$
2284: correlation similar to the one widely used to interpret observations
2285: \citep{Kennicutt1998B}. However, if both AGN and stars contribute to
2286: the dust emission in the far-IR, then the correlation becomes
2287: non-linear with a modified normalization. If the AGN dominates the
2288: $\Lfir$, as in the peak quasar phase, we find no correlation at all.
2289: Finally, we find correlations between dust masses and rest-frame
2290: fluxes at $3\, \mu$m and $50\, \mu$m. The $f_{3\, \mu \rm m}$ flux may
2291: serve as a good diagnostic for hot dust in quasars, while $f_{50\,
2292: \mu \rm m}$ may be used to estimate the amount of cold dust in
2293: starburst galaxies, as $f_{50\, \mu \rm m}$ correlates linearly with
2294: the cold dust mass.
2295:
2296: Our model demonstrates that massive star formation at higher
2297: redshifts ($ z \gtrsim 10$) is necessary in order to produce the observed
2298: dust properties in these quasars. This suggests that the quasar
2299: hosts should have already built up a large stellar population by $z
2300: \sim 6$. Our results support a merger-driven origin for luminous
2301: quasars at high redshifts, and provide further evidence for the
2302: hypothesis of starburst-to-quasar evolution.
2303:
2304:
2305: \acknowledgments{Special thanks to Kenny Wood\footnotemark[2] and Barbara
2306: Whitney\footnotemark[3] for generously making their Monte Carlo radiative
2307: equilibrium codes publicly available;
2308: \footnotetext[2]{http://www-star.st-and.ac.uk/$\sim$kw25/research/montecarlo/montecarlo.html}
2309: \footnotetext[3]{http://gemelli.colorado.edu/$\sim$bwhitney/codes/codes.html}
2310: and to Patrik Jonsson for sharing publicly his fantastic
2311: radiative transfer code SUNRISE\footnotemark[4]
2312: \footnotetext[4]{http://www.ucolick.org/$\sim$patrik/sunrise/}.
2313: We thank Gurtina Besla, Stephanie Bush, Sukanya Chakrabarti,
2314: Suvendra Dutta, Giovanni Fazio, Jiasheng Huang, Dusan Keres, Desika
2315: Narayanan, Erik Rosolowsky and Josh Younger for many stimulating discussions.
2316: YL thanks Lee Armus, Frank Bertoldi, Bruce Draine, Patrik Jonsson,
2317: Carol Lonsdale, Mike Nolta, Casey Papovich, George Rybicki, Nick
2318: Scoville, Rodger Thompson and Lin Yan for inspiration and
2319: discussions. Finally, we thank the referee for a thoughtful report that has
2320: helped improve this manuscript. YL gratefully acknowledges a Keck
2321: Fellowship sponsored by the Keck Foundation, as well as an Institute for
2322: Theory and Computation Fellowship, under which support most of this work was
2323: done. The computations reported here were performed at the Center for
2324: Parallel Astrophysical Computing at Harvard-Smithsonian Center for
2325: Astrophysics. This work was supported in part by NSF grant 03-07690 and NASA
2326: ATP grant NAG5-13381.}
2327:
2328: \bibliography{ms}
2329:
2330:
2331: \end{document}
2332: