astro-ph0603360/p.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%% author.tex %%%%%%%%%%%%%%%%%%%%%%%%%
2: %
3: % sample root file for your contribution to a "contributed book"
4: %
5: % "contributed book"
6: %
7: % Use this file as a template for your own input.
8: %
9: %%%%%%%%%%%%%%%%%%%%%%%% Springer-Verlag %%%%%%%%%%%%%%%%%%%%%%%%%%
10: 
11: 
12: % RECOMMENDED %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
13: \documentclass[vecphys]{svmult}
14: % choose options for [] as required from the list
15: % in the Reference Guide, Sect. 2.2
16: \usepackage{amsmath}
17: 
18: \usepackage{makeidx}         % allows index generation
19: \usepackage{graphicx}        % standard LaTeX graphics tool
20:                              % when including figure files
21: \usepackage{multicol}        % used for the two-column index
22: \usepackage[bottom]{footmisc}% places footnotes at page bottom
23: % etc.
24: % see the list of further useful packages
25: % in the Reference Guide, Sects. 2.3, 3.1-3.3
26: 
27: %***
28: %***Definitions
29: %***
30: 
31: \def\lya{Ly$\alpha$~}
32: \def\ljeans{\lambda_{\rm J}}
33: \def\centering{ }
34: \def\ion{ }
35: \def\citep{\cite}
36: \def\citet{\cite}
37: \def\HII{H II~}
38: \def\sun{\odot}
39: \def\ie{i.e.~}
40: \def\ga{>}
41: \def\gtrsim{>}
42: \def\la{<}
43: \def\del{\delta}
44: \def\lesssim{<}
45: \def\kjeans{k_{\rm J}}
46: \def\mjeans{M_{\rm J}}
47: \def\beq{\begin{equation}}
48: \def\eeq{\end{equation}}
49: \def\ba{\begin{eqnarray}}
50: \def\ee{\end{equation}}
51: \def\lapprox{\la}
52: \def\Msun{$M_\odot$}
53: \def\lyasource{{\dot N_\alpha}}
54: \def\beqa{\begin{eqnarray}}
55: \def\eeqa{\end{eqnarray}}
56: \def\xb{{\bf x}}
57: \def\rb{{\bf r}}
58: \def\vb{{\bf v}}
59: \def\ub{{\bf u}}
60: \def\kb{{\bf k}}
61: \def\Omm{{\Omega_m}}
62: \def\Ommz{{\Omega_m^{\,z}}}
63: \def\Omr{{\Omega_r}}
64: \def\Omk{{\Omega_k}}
65: \def\Oml{{\Omega_{\Lambda}}}
66: \def\nb{\bar{n}}
67: \def\etal{et al.\ }
68: \def\Ng{N_\gamma}
69: \def\zr{z_{\rm reion}}
70: \def\HI{\rm H\,I~}
71: \def\cN{c_{\rm N}}
72: \def\kB{k}
73: \def\Ni{N_{\rm ion}}
74: \def\fg{f_{\rm gas}}
75: \def\fe{f_{\rm eject}}
76: \def\fin{f_{\rm int}}
77: \def\fw{f_{\rm wind}}
78: \def\NGST{{\it JWST}\,}
79: \def\dh{{\delta_{H}}}
80: \def\n{{\bf n}}
81: \def\k{{\bf k}}
82: \newcommand{\delh}{{\delta_{H}}}
83: \newcommand{\bx}{{\bf x}}
84: \newcommand{\bn}{{\bf n}}
85: \newcommand{\br}{{\bf r}}
86: \newcommand{\bk}{{\bf k}}
87: \newcommand{\Lya}{Ly$\alpha$~}
88: \newcommand{\Lyb}{Ly$\beta$~}
89: \newcommand{\td}{{\tilde{\delta}}}
90: 
91: \makeindex             % used for the subject index
92:                        % please use the style sprmidx.sty with
93:                        % your makeindex program
94: 
95: 
96: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
97: 
98: \begin{document}
99: 
100: \title*{First Light}
101: % Use \titlerunning{Short Title} for an abbreviated version of
102: % your contribution title if the original one is too long
103: \author{Abraham Loeb\inst{1}}
104: % Use \authorrunning{Short Title} for an abbreviated version of
105: % your contribution title if the original one is too long
106: \institute{Department of Astronomy, Harvard University, 60 Garden St.,
107: Cambridge, MA 02138 \texttt{aloeb@cfa.harvard.edu}}
108: %
109: % Use the package "url.sty" to avoid
110: % problems with special characters
111: % used in your e-mail or web address
112: %
113: \maketitle
114: 
115: \begin{abstract}
116: 
117: The first dwarf galaxies, which constitute the building blocks of the
118: collapsed objects we find today in the Universe, had formed hundreds of
119: millions of years after the big bang. This pedagogical review describes the
120: early growth of their small-amplitude seed fluctuations from the epoch of
121: inflation through dark matter decoupling and matter-radiation equality, to
122: the final collapse and fragmentation of the dark matter on all mass scales
123: above $\sim 10^{-4}M_\odot$. The condensation of baryons into halos in the
124: mass range of $\sim 10^5$--$10^{10}M_\odot$ led to the formation of the
125: first stars and the re-ionization of the cold hydrogen gas, left over from
126: the big bang.  The production of heavy elements by the first stars started
127: the metal enrichment process that eventually led to the formation of rocky
128: planets and life.
129: 
130: A wide variety of instruments currently under design [including
131: large-aperture infrared telescopes on the ground or in space ({\it JWST}),
132: and low-frequency arrays for the detection of redshifted 21cm radiation],
133: will establish better understanding of the first sources of light during an
134: epoch in cosmic history that was largely unexplored so far. Numerical
135: simulations of reionization are computationally challenging, as they
136: require radiative transfer across large cosmological volumes as well as
137: sufficently high resolution to identify the sources of the ionizing
138: radiation. The technological challenges for observations and the
139: computational challenges for numerical simulations, will motivate intense
140: work in this field over the coming decade.
141: %Better understanding of the properties and environmental influence
142: %of first sources of light will establish the scientific substitute for
143: %traditional texts (such as the biblical story of ``genesis''), that
144: %dominated notions on this subject in past centuries.
145: 
146: \noindent
147: {\bf Disclaimer:} {\it This review was written as an introductory text for
148: a series of lectures at the SAAS-FEE 2006 winter school, and so it includes
149: a limited sample of references on each subject. It does not intend to
150: provide a comprehensive list of all up-to-date references on the topics
151: under discussion, but rather to raise the interest of beginning graduate
152: students in the related literature.}
153: 
154: 
155: 
156: \end{abstract}
157: 
158: %***
159: %***Table of Contents
160: %***
161: %
162: %\newpage
163: %\tableofcontents
164: %\newpage
165: 
166: %<>**************************************************************************
167: 
168: \section{\bf Opening Remarks}
169: \label{sec1}
170: 
171: When I open the daily newspaper as part of my morning routine, I often see
172: lengthy descriptions of conflicts between people on borders, properties, or
173: liberties. Today's news is often forgotten a few days later.  But when one
174: opens ancient texts that have appealed to a broad audience over a longer
175: period of time, such as the Bible, what does one often find in the opening
176: chapter?...  a discussion of how the constituents of the Universe
177: (including light, stars and life) were created.  Although humans are often
178: occupied with mundane problems, they are curious about the big picture. As
179: citizens of the Universe, we cannot help but wonder how the first sources
180: of light formed, how life came to existence, and whether we are alone as
181: intelligent beings in this vast space.  As astronomers in the twenty first
182: century, we are uniquely positioned to answer these big questions with
183: scientific instruments and a quantitative methodology. In this pedagogical
184: review, intended for students preparing to specialize in cosmology, I will
185: describe current ideas about one of these topics: {\it the appearance of
186: the first sources of light and their influence on the surrounding
187: Universe}. This topic is one of the most active frontiers in present-day
188: cosmology. As such it is an excellent area for a PhD thesis of a graduate
189: student interested in cosmology.  I will therefore highlight the unsolved
190: questions in this field as much as the bits we understand.
191: 
192: \section{\bf Excavating the Universe for Clues About Its History}
193: 
194: When we look at our image reflected off a mirror at a distance of 1 meter,
195: we see the way we looked 6 nano-seconds ago, the light travel time to the
196: mirror and back. If the mirror is spaced $10^{19}~{\rm cm}=3$pc away, we
197: will see the way we looked twenty one years ago. Light propagates at a
198: finite speed, and so by observing distant regions, we are able to see how
199: the Universe looked like in the past, a light travel time ago. The
200: statistical homogeneity of the Universe on large scales guarantees that
201: what we see far away is a fair statistical representation of the conditions
202: that were present in in our region of the Universe a long time ago.
203: 
204: \begin{figure}
205: \centering
206: %\picplace{5cm}{2cm}{history.eps}
207: %??rotate image by 90 degrees??
208: \includegraphics[height=6cm]{history.eps}
209: \caption{Cosmology is like archeology. The deeper one looks, the older is
210: the layer that one is revealing, owing to the finite propagation speed of
211: light.}
212: \label{fig:1}       % Give a unique label
213: \end{figure}
214: 
215: This fortunate situation makes cosmology an empirical science. We do not
216: need to guess how the Universe evolved. Using telescopes we can simply see
217: the way it appeared at earlier cosmic times. Since a greater distance means
218: a fainter flux from a source of a fixed luminosity, the observation of the
219: earliest sources of light requires the development of sensitive instruments
220: and poses challenges to observers.
221: 
222: We can in principle image the Universe only if it is transparent. Earlier
223: than 0.4 million years after the big bang, the cosmic plasma was ionized
224: and the Universe was opaque to Thomson scattering by the dense gas of free
225: electrons that filled it. Thus, telescopes cannot be used to image the
226: infant Universe at earlier times (or redshifts $\ga 10^3$). The earliest
227: possible image of the Universe was recorded by COBE and WMAP (see Fig. 2).
228: 
229: \begin{figure}
230: \centering
231: \includegraphics[height=6cm]{WMAP.eps}
232: \caption{Images of the Universe shortly after it became transparent, taken
233: by the {\it COBE} and {\it WMAP} satellites (see http://map.gsfc.nasa.gov/
234: for details).  The slight density inhomogeneties in the otherwise uniform
235: Universe, imprinted hot and cold brightness map of the cosmic microwave
236: background.  The existence of these anisotropies was predicted three
237: decades before the technology for taking this image 
238: became available in a number of theoretical papers, 
239: including \cite{zel,sachs,sciama,silk,peeb}.}
240: \end{figure}
241: 
242: \begin{figure}
243: \centering
244: \includegraphics[height=6cm]{xe.ps}
245: \caption{The optical depth of the Universe to electron scattering (upper
246: panel) and the ionization fraction (lower panel) as a function of redshift
247: before reionization.  Observatories of electromagnetic radiation cannot
248: image the opaque Universe beyond a redshift of $z\sim 1100$.}
249: \end{figure}
250: 
251: \section{Bakground Cosmological Model}
252: 
253: %<>
254: \subsection{The Expanding Universe}
255: \label{sec2.1}
256: 
257: The modern physical description of the Universe as a whole can be traced
258: back to Einstein, who argued theoretically for the so-called ``cosmological
259: principle'': that the distribution of matter and energy must be homogeneous
260: and isotropic on the largest scales. Today isotropy is well established
261: (see the review by Wu, Lahav, \& Rees 1999 \cite{Wu99}) for the distribution of faint
262: radio sources, optically-selected galaxies, the X-ray background, and most
263: importantly the cosmic microwave background (hereafter, CMB; see, e.g.,
264: Bennett et al.\ 1996 \cite{Be96}). The constraints on homogeneity are less strict, but
265: a cosmological model in which the Universe is isotropic but significantly
266: inhomogeneous in spherical shells around our special location, is also
267: excluded \cite{Goodman95}.
268: 
269: In General Relativity, the metric for a space which is spatially
270: homogeneous and isotropic is the Friedman-Robertson-Walker metric, which
271: can be written in the form \beq \label{RW}
272: ds^2=dt^2-a^2(t)\left[\frac{dR^2}{1-k\,R^2}+R^2
273: \left(d\theta^2+\sin^2\theta\,d\phi^2\right)\right]\ , \eeq where $a(t)$ is
274: the cosmic scale factor which describes expansion in time, and
275: $(R,\theta,\phi)$ are spherical comoving coordinates. The constant $k$
276: determines the geometry of the metric; it is positive in a closed Universe,
277: zero in a flat Universe, and negative in an open Universe. Observers at
278: rest remain at rest, at fixed $(R,\theta,\phi)$, with their physical
279: separation increasing with time in proportion to $a(t)$. A given observer
280: sees a nearby observer at physical distance $D$ receding at the Hubble
281: velocity $H(t)D$, where the Hubble constant at time $t$ is
282: $H(t)=d\,a(t)/dt$. Light emitted by a source at time $t$ is observed at
283: $t=0$ with a redshift $z=1/a(t)-1$, where we set $a(t=0) \equiv 1$
284: for convenience (but note that old textbooks may use a different 
285: convention).
286: 
287: The Einstein field equations of General Relativity yield the Friedmann
288: equation (e.g., Weinberg 1972 \cite{We72}; Kolb \& Turner 1990
289: \cite{Kolb90}) \beq H^2(t)=\frac{8 \pi G}{3}\rho-\frac{k}{a^2}\ ,\eeq which
290: relates the expansion of the Universe to its matter-energy content. For
291: each component of the energy density $\rho$, with an equation of state
292: $p=p(\rho)$, the density $\rho$ varies with $a(t)$ according to the
293: equation of energy conservation \beq d (\rho R^3)=-p d(R^3)\ . \eeq With
294: the critical density \beq \rho_C(t) \equiv \frac{3 H^2(t)}{8 \pi G} \eeq
295: defined as the density needed for $k=0$, we define the ratio of the total
296: density to the critical density as \beq \Omega \equiv \frac{\rho}{\rho_C}\
297: . \eeq With $\Omm$, $\Oml$, and $\Omr$ denoting the present contributions
298: to $\Omega$ from matter (including cold dark matter as well as a
299: contribution $\Omega_b$ from baryons), vacuum density (cosmological
300: constant), and radiation, respectively, the Friedmann equation becomes \beq
301: \frac{H(t)}{H_0}= \left[ \frac{\Omm} {a^3}+ \Oml+ \frac{\Omr}{a^4}+
302: \frac{\Omk}{a^2}\right]\ , \eeq where we define $H_0$ and
303: $\Omega_0=\Omm+\Oml+\Omr$ to be the present values of $H$ and $\Omega$,
304: respectively, and we let \beq \Omk \equiv -\frac{k}{H_0^2}=1-\Omega_m. \eeq
305: In the particularly simple Einstein-de Sitter model ($\Omm=1$,
306: $\Oml=\Omr=\Omk=0$), the scale factor varies as $a(t) \propto
307: t^{2/3}$. Even models with non-zero $\Oml$ or $\Omk$ approach the
308: Einstein-de Sitter behavior at high redshift, i.e.\ when $(1+z) \gg
309: |\Omm^{-1}-1|$ (as long as $\Omr$ can be neglected). In this high-$z$
310: regime the age of the Universe is,
311: \begin{equation}
312: t\approx {2\over 3 H_0 \sqrt{\Omega_m}} \left(1+z\right)^{-3/2}~.
313: \end{equation}
314: The Friedmann equation implies that models with $\Omk=0$ converge to the
315: Einstein-de Sitter limit faster than do open models.
316: 
317: In the standard hot Big Bang model, the Universe is initially hot and the
318: energy density is dominated by radiation. The transition to matter
319: domination occurs at $z \sim 3500$, but the Universe remains hot enough
320: that the gas is ionized, and electron-photon scattering effectively couples
321: the matter and radiation. At $z \sim 1100$ the temperature drops below
322: $\sim 3000$K and protons and electrons recombine to form neutral
323: hydrogen. The photons then decouple and travel freely until the present,
324: when they are observed as the CMB \cite{WMAP}.
325: 
326: \subsection{Composition of the Universe}
327: 
328: According to the standard cosmological model, the Universe started at the
329: big bang about 14 billion years ago.  During an early epoch of accelerated
330: superluminal expansion, called inflation, a region of microscopic size was
331: stretched to a scale much bigger than the visible Universe and our local
332: geometry became flat. At the same time, primordial density fluctuations
333: were generated out of quantum mechanical fluctuations of the vacuum.  These
334: inhomogeneities seeded the formation of present-day structure through the
335: process of gravitational instability. The mass density of ordinary
336: (baryonic) matter makes up only a fifth of the matter that led to the
337: emergence of structure and the rest is the form of an unknown dark matter
338: component.  Recently, the Universe entered a new phase of accelerated
339: expansion due to the dominance of some dark vacuum energy density over the
340: ever rarefying matter density.
341: 
342: \noindent
343: The basic question that cosmology attempts to answer is: 
344: 
345: \noindent
346: {\bf What are the ingredients (composition and initial conditions) of the
347: Universe and what processes generated the observed structures in it?}
348: 
349: \noindent
350: In detail, we would like to know: 
351: 
352: \noindent
353: {\it (a)} Did inflation occur and when? If so, what drove it and how did it
354: end?
355: 
356: \noindent
357: {\it (b)} What is the nature of of the dark energy and how does it change
358: over time and space?
359: 
360: \noindent
361: {\it (c)} What is the nature of the dark matter and how did it regulate the
362: evolution of structure in the Universe?
363: 
364: Before hydrogen recombined, the Universe was opaque to electromagnetic
365: radiation, precluding any possibility for direct imaging of its evolution.
366: The only way to probe inflation is through the fossil record that it left
367: behind in the form of density perturbations and gravitational waves.
368: Following inflation, the Universe went through several other milestones
369: which left a detectable record. These include: baryogenesis (which resulted
370: in the observed asymmetry between matter and anti-matter), the electroweak
371: phase transition (during which the symmetry between electromagnetic and
372: weak interactions was broken), the QCD phase transition (during which
373: protons and neutrons were assembled out of quarks and gluons), the dark
374: matter freeze-out epoch (during which the dark matter decoupled from the
375: cosmic plasma), neutrino decoupling, electron-positron annihilation, and
376: light-element nucleosynthesis (during which helium, deuterium and lithium
377: were synthesized). The signatures that these processes left in the Universe
378: can be used to constrain its parameters and answer the above questions.
379: 
380: Half a million years after the big bang, hydrogen recombined and the
381: Universe became transparent. The ultimate goal of observational cosmology
382: is to image the entire history of the Universe since then. Currently, we
383: have a snapshot of the Universe at recombination from the CMB, and detailed
384: images of its evolution starting from an age of a billion years until the
385: present time. The evolution between a million and a billion years has not
386: been imaged as of yet.
387: 
388: Within the next decade, NASA plans to launch an infrared space telescope
389: (JWST) that will image the very first sources of light (stars and black
390: holes) in the Universe, which are predicted theoretically to have formed in
391: the first hundreds of millions of years. In parallel, there are several
392: initiatives to construct large-aperture infrared telescopes on the ground
393: with the same goal in
394: mind\footnote{http://www.eso.org/projects/owl/}$^{,}$
395: \footnote{http://celt.ucolick.org/}$^{,}$\footnote{http://www.gmto.org/}.
396: The neutral hydrogen, relic from cosmological recombination, can be mapped
397: in three-dimensions through its 21cm line even before the first galaxies
398: formed \cite{Loeb04}. Several groups are currently constructing
399: low-frequency radio arrays in an attempt to map the initial inhomogeneities
400: as well as the process by which the hydrogen was re-ionized by the first
401: galaxies.
402: 
403: \begin{figure}
404: \centering
405: \includegraphics[height=6cm]{JWST.eps}
406: \caption{A sketch of the current design for the {\it James Webb Space
407: Telescope}, the successor to the {\it Hubble Space Telescope} to be
408: launched in 2011 (http://www.jwst.nasa.gov/). The current design includes a
409: primary mirror made of beryllium which is 6.5 meter in diameter as well as
410: an instrument sensitivity that spans the full range of infrared wavelengths
411: of 0.6--28$\mu$m and will allow detection of the first galaxies in the
412: infant Universe.
413: %The size of the sun shield (the large flat screen
414: %in the image) is 22m$\times$10m. 
415: The telescope will orbit 1.5 million km from Earth at the Lagrange L2
416: point.}
417: \end{figure}
418: 
419: The next generation of ground-based telescopes will have a diameter of
420: twenty to thirty meter. Together with JWST (that will not be affected by
421: the atmospheric backgound) they will be able to image the first
422: galaxies. Given that these galaxies also created the ionized bubbles around
423: them, the same galaxy locations should correlate with bubbles in the
424: neutral hydrogen (created by their UV emission). Within a decade it would
425: be possible to explore the environmental influence of individual galaxies
426: by using the two sets of instruments in concert \cite{WyBar}.
427: 
428: \begin{figure}
429: \centering
430: \includegraphics[height=6cm]{gmt.ps}
431: \caption{Artist conception of the design for one of the future giant
432: telescopes that could probe the first generation of galaxies from the
433: ground. The {\it Giant Magellan Telescope} (GMT) will contain seven mirrors
434: (each 8.4 meter in diameter) and will have the resolving power equivalent
435: to a 24.5 meter (80 foot) primary mirror. For more details see
436: http://www.gmto.org/}
437: \label{gmt}
438: \end{figure}
439: 
440: 
441: The dark ingredients of the Universe can only be probed indirectly through
442: a variety of luminous tracers.  The distribution and nature of the dark
443: matter are constrained by detailed X-ray and optical observations of
444: galaxies and galaxy clusters. The evolution of the dark energy with cosmic
445: time will be constrained over the coming decade by surveys of Type Ia
446: supernovae, as well as surveys of X-ray clusters, up to a redshift of two.
447: 
448: On large scales ($\ga 10$Mpc) the power-spectrum of primordial density
449: perturbations is already known from the measured microwave background
450: anisotropies, galaxy surveys, weak lensing, and the Ly$\alpha$
451: forest. Future programs will refine current knowledge, and will search for
452: additional trademarks of inflation, such as gravitational waves (through
453: CMB polarization), small-scale structure (through high-redshift galaxy
454: surveys and 21cm studies), or the Gaussian statistics of the initial
455: perturbations.
456: 
457: %ADD/summarize in a Table all cosmological parameters based on Seljak/WMAP 
458: %(zero curvature, running, pure inflationary LCDM). In the following we
459: %assume the dark matter is cold, ignore neutrinos and adopt $\Omega_b=...$.
460: 
461: %\section{Linear Perturbations}
462: %
463: %\subsection{Growth of Dark Matter Inhomogeneities}
464: %
465: %\subsubsection{The Seeds Planted by Inflation}
466: %
467: %ADD Quantum fluctuation freeze and become classical after horizon crossing,
468: %Decoherence guarantees classical behaviour and deterministic evolution at
469: %later times. (borrow discussion from Liddle \& Lyth, p. 182-183).  Each
470: %${\bf k}$-mode is uncorrelated (meaning gaussian statistics) because of the
471: %nature of quantum fluctuations.
472: %
473: The big bang is the only known event where particles with energies
474: approaching the Planck scale [$(\hbar c^5/G)^{1/2}\sim 10^{19}~{\rm GeV}$]
475: interacted.  It therefore offers prospects for probing the unification
476: physics between quantum mechanics and general relativity (to which string
477: theory is the most-popular candidate).  Unfortunately, the exponential
478: expansion of the Universe during inflation erases memory of earlier cosmic
479: epochs, such as the Planck time.
480: 
481: %\subsubsection{Growth During the Radiation-Dominated Era
482: %through Dark Matter Decoupling}
483: %
484: %Acoustic oscillations: inflation produce synchronized perturbations at some
485: %initial time slice. 
486: %
487: %Derive formalism from Dodelson (including super-horizon)
488: %
489: 
490: 
491: %<>
492: \subsection{Linear Gravitational Growth}
493: \label{sec2.2}
494: 
495: Observations of the CMB (e.g., Bennett et al.\ 1996 \cite{Be96}) show that the
496: Universe at recombination was extremely uniform, but with spatial
497: fluctuations in the energy density and gravitational potential of
498: roughly one part in $10^5$. Such small fluctuations, generated in the
499: early Universe, grow over time due to gravitational instability, and
500: eventually lead to the formation of galaxies and the large-scale
501: structure observed in the present Universe.
502: 
503: As before, we distinguish between fixed and comoving
504: coordinates. Using vector notation, the fixed coordinate ${\bf r}$
505: corresponds to a comoving position $\xb=\rb/a$. In a homogeneous
506: Universe with density $\rho$, we describe the cosmological expansion
507: in terms of an ideal pressureless fluid of particles each of which is
508: at fixed $\xb$, expanding with the Hubble flow $\vb=H(t) \rb$ where
509: $\vb=d\rb/dt$. Onto this uniform expansion we impose small
510: perturbations, given by a relative density perturbation \beq
511: \delta(\xb)=\frac{\rho(\rb)}{\bar{\rho}}-1\ , \eeq where the mean
512: fluid density is $\bar{\rho}$, with a corresponding peculiar velocity
513: $\ub \equiv \vb - H \rb$. Then the fluid is described by the
514: continuity and Euler equations in comoving coordinates \cite{p80,Peebles}:
515: \beqa \frac{\partial \delta}{\partial t}+\frac{1}{a}{\bf
516: \nabla} \cdot \left[(1+\delta) \ub\right] &=& 0 \\ \frac{\partial
517: \ub}{\partial t}+H\ub+\frac{1}{a}(\ub \cdot {\bf \nabla})
518: \ub&=&-\frac{1}{a}{\bf \nabla}\phi\ .  \eeqa The potential $\phi$ is
519: given by the Poisson equation, in terms of the density perturbation:
520: \beq \nabla^2\phi=4 \pi G \bar{\rho} a^2 \delta\ . \eeq This fluid
521: description is valid for describing the evolution of collisionless
522: cold dark matter particles until different particle streams cross.
523: This
524: ``shell-crossing'' typically occurs only after perturbations have
525: grown to become non-linear, and at that point the individual particle
526: trajectories must in general be followed. Similarly, baryons can be
527: described as a pressureless fluid as long as their temperature is
528: negligibly small, but non-linear collapse leads to the formation of
529: shocks in the gas.
530: 
531: For small perturbations $\delta \ll 1$, the fluid equations can be
532: linearized and combined to yield \beq \frac{\partial^2\delta}{\partial
533: t^2}+2 H \frac{\partial\delta}{\partial t}=4 \pi G \bar{\rho} \delta\
534: . \eeq This linear equation has in general two independent solutions, only
535: one of which grows with time. Starting with random initial conditions, this
536: ``growing mode'' comes to dominate the density evolution. Thus, until it
537: becomes non-linear, the density perturbation maintains its shape in
538: comoving coordinates and grows in proportion to a growth factor $D(t)$. The
539: growth factor in the matter-dominated era is given by \cite{p80} \beq
540: D(t) \propto \frac{\left(\Oml a^3+\Omk a+\Omm\right)^{1/2}}{a^{3/2}}\int_0^a
541: \frac{a'^{3/2}\, da'}{\left(\Oml a'^3+\Omk a'+\Omm\right)^{3/2}}\ , \eeq where
542: we neglect $\Omr$ when considering halos forming in the matter-dominated
543: regime at $z \ll 10^4$. In the Einstein-de Sitter model (or, at high redshift,
544: in other models as well) the growth factor is simply proportional to
545: $a(t)$.
546: 
547: The spatial form of the initial density fluctuations can be described in
548: Fourier space, in terms of Fourier components \beq \delta_\kb = \int d^3x\,
549: \delta(x) e^{-i \kb \cdot \xb}\ .\eeq Here we use the comoving wavevector
550: $\kb$, whose magnitude $k$ is the comoving wavenumber which is equal to
551: $2\pi$ divided by the wavelength. The Fourier description is particularly
552: simple for fluctuations generated by inflation (e.g., Kolb \& Turner 1990
553: \cite{Kolb90}). Inflation generates perturbations given by a Gaussian
554: random field, in which different $\kb$-modes are statistically independent,
555: each with a random phase. The statistical properties of the fluctuations
556: are determined by the variance of the different $\kb$-modes, and the
557: variance is described in terms of the power spectrum $P(k)$ as follows:
558: \beq \left<\delta_{\kb} \delta_{{\bf k'}}^{*}\right>=\left(2 \pi\right)^3
559: P(k)\, \delta^{(3)} \left(\kb-{\bf k'}\right)\ , \eeq where $\delta^{(3)}$
560: is the three-dimensional Dirac delta function.  The gravitational potential
561: fluctuations are sourced by the density fluctuations through Poisson's
562: equation.
563: 
564: In standard models, inflation produces a primordial power-law spectrum
565: $P(k) \propto k^n$ with $n \sim 1$. Perturbation growth in the
566: radiation-dominated and then matter-dominated Universe results in a
567: modified final power spectrum, characterized by a turnover at a scale
568: of order the horizon $cH^{-1}$ at matter-radiation equality, and a
569: small-scale asymptotic shape of $P(k) \propto k^{n-4}$. The overall
570: amplitude of the power spectrum is not specified by current models of
571: inflation, and it is usually set by comparing to the observed CMB
572: temperature fluctuations or to local measures of large-scale
573: structure.
574: 
575: Since density fluctuations may exist on all scales, in order to determine
576: the formation of objects of a given size or mass it is useful to consider
577: the statistical distribution of the smoothed density field.  Using a window
578: function $W(\rb)$ normalized so that $\int d^3r\, W(\rb)=1$, the smoothed
579: density perturbation field, $\int d^3r \delta(\xb) W(\rb)$, itself follows
580: a Gaussian distribution with zero mean. For the particular choice of a
581: spherical top-hat, in which $W=1$ in a sphere of radius $R$ and is zero
582: outside, the smoothed perturbation field measures the fluctuations in the
583: mass in spheres of radius $R$. The normalization of the present power
584: spectrum is often specified by the value of $\sigma_8 \equiv \sigma(R=8
585: h^{-1} {\rm Mpc})$. For the top-hat, the smoothed perturbation field is
586: denoted $\delta_R$ or $\delta_M$, where the mass $M$ is related to the
587: comoving radius $R$ by $M=4 \pi \rho_m R^3/3$, in terms of the current mean
588: density of matter $\rho_m$. The variance $\langle \delta_M \rangle^2$ is
589: \beq \sigma^2(M)= \sigma^2(R)= \int_0^{\infty}\frac{dk}{2 \pi^2} \,k^2 P(k)
590: \left[\frac{3 j_1(kR)}{kR} \right]^2\ ,\label{eqsigM}\eeq where
591: $j_1(x)=(\sin x-x \cos x)/x^2$. The function $\sigma(M)$ plays a crucial
592: role in estimates of the abundance of collapsed objects, as we describe
593: later.
594: 
595: Species that decouple from the cosmic plasma (like the dark matter or the
596: baryons) would show fossil evidence for acoustic oscillations in their
597: power spectrum of inhomogeneities due to sound waves in the radiation fluid
598: to which they were coupled at early times.  This phenomenon can be
599: understood as follows. Imagine a localized point-like perturbation from
600: inflation at $t=0$. The small perturbation in density or pressure will send
601: out a sound wave that will reach the sound horizon $c_s t$ at any later
602: time $t$.  The perturbation will therefore correlate with its surroundings
603: up to the sound horizon and all $k$-modes with wavelengths equal to this
604: scale or its harmonics will be correlated.
605: %ADD: this point can be captured and quantified
606: %through the Green's function approach to the evolution of perturbations in
607: %real space (Bashinsky \& Bertschinger).
608: The scales of the perturbations that grow to become the first collapsed
609: objects at $z< 100$ cross the horizon in the radiation dominated era
610: after the dark matter decouples from the cosmic plasma. Next we consider
611: the imprint of this decoupling on the smallest-scale structure of
612: the dark matter.
613: 
614: \subsection{The Smallest-Scale Power Spectrum of Cold Dark Matter}
615: 
616: %ADD simplify and cut-down discussion to include only
617: %the essential ingredients for an introduction on the subject.
618: %
619: A broad range of observational data involving the dynamics of galaxies, the
620: growth of large-scale structure, and the dynamics and nucleosynthesis of
621: the Universe as a whole, indicate the existence of dark matter with a mean
622: cosmic mass density that is $\sim 5$ times larger than the density of the
623: baryonic matter \cite{Jungman,WMAP}.  The data is consistent with a dark
624: matter composed of weakly-interacting, massive particles, that decoupled
625: early and adiabatically cooled to an extremely low temperature by the
626: present time \cite{Jungman}.  The Cold Dark Matter (CDM) has not been
627: observed directly as of yet, although laboratory searches for particles
628: from the dark halo of our own Milky-Way galaxy have been able to restrict
629: the allowed parameter space for these particles. Since an alternative
630: more-radical interpretation of the dark matter phenomenology involves a
631: modification of gravity \cite{Beken}, it is of prime importance to find
632: direct fingerprints of the CDM particles. One such fingerprint involves the
633: small-scale structure in the Universe \cite{Green}, on which we focus in
634: this section.
635: 
636: The most popular candidate for the CDM particle is a Weakly Interacting
637: Massive Particle (WIMP).  The lightest supersymmetric particle (LSP) could
638: be a WIMP (for a review see \cite{Jungman}).  The CDM particle mass depends
639: on free parameters in the particle physics model but typical values cover a
640: range around $M\sim 100~{\rm GeV}$ (up to values close to a TeV). In many
641: cases the LSP hypothesis will be tested at the Large Hadron Collider
642: (e.g. \cite{battaglia}) or in direct detection experiments
643: (e.g. \cite{baltz}).
644: 
645: The properties of the CDM particles affect their response to the
646: small-scale primordial inhomogeneities produced during cosmic
647: inflation. The particle cross-section for scattering off standard model
648: fermions sets the epoch of their thermal and kinematic decoupling from the
649: cosmic plasma (which is significantly later than the time when their
650: abundance freezes-out at a temperature $T\sim M$). Thermal decoupling is
651: defined as the time when the temperature of the CDM stops following that of
652: the cosmic plasma while kinematic decoupling is defined as the time when
653: the bulk motion of the two species start to differ.  For CDM the epochs of
654: thermal and kinetic decoupling coincide.  They occur when the time it takes
655: for collisions to change the momentum of the CDM particles equals the
656: Hubble time.  The particle mass determines the thermal spread in the speeds
657: of CDM particles, which tends to smooth-out fluctuations on very small
658: scales due to the free-streaming of particles after kinematic decoupling
659: \cite{Green,Green2}.  Viscosity has a similar effect before the CDM fluid
660: decouples from the cosmic radiation fluid \cite{Hofmann}.  An important
661: effect involves the memory the CDM fluid has of the acoustic oscillations
662: of the cosmic radiation fluid out of which it decoupled.  Here we consider
663: the imprint of these acoustic oscillations on the small-scale power
664: spectrum of density fluctuations in the Universe. Analogous imprints of
665: acoustic oscillations of the baryons were identified recently in maps of
666: the CMB \cite{WMAP}, and the distribution of nearby galaxies
667: \cite{Eisenste}; these signatures appear on much larger scales, since the
668: baryons decouple much later when the scale of the horizon is larger.  The
669: discussion in this section follows Loeb \& Zaldarriaga (2005) \cite{Loe05}.
670: 
671: \noindent{\bf Formalism}
672: 
673: Kinematic decoupling of CDM occurs during the radiation-dominated era.  For
674: example, if the CDM is made of neutralinos with a particle mass of $\sim
675: 100~{\rm GeV}$, then kinematic decoupling occurs at a cosmic temperature of
676: $T_{\rm d}\sim 10~{\rm MeV}$ \cite{Hofmann,Chen}.  As long as $T_d \ll
677: 100~{\rm MeV}$, we may ignore the imprint of the QCD phase transition
678: (which transformed the cosmic quark-gluon soup into protons and neutrons)
679: on the CDM power spectrum \cite{Schmid}.  Over a short period of time
680: during this transition, the pressure does not depend on density and the
681: sound speed of the plasma vanishes, resulting in a significant growth for
682: perturbations with periods shorter than the length of time over which the
683: sound speed vanishes. The transition occurs when the temperature of the
684: cosmic plasma is $\sim 100-200~{\rm MeV}$ and lasts for a small fraction of
685: the Hubble time. As a result, the induced modifications are on scales
686: smaller than those we are considering here and the imprint of the QCD phase
687: transition is washed-out by the effects we calculate.
688: 
689: At early times the contribution of the dark matter to the energy density is
690: negligible. Only at relatively late times when the cosmic temperature drops
691: to values as low as $\sim 1$ eV, matter and radiation have comparable
692: energy densities.  As a result, the dynamics of the plasma at earlier times
693: is virtually unaffected by the presence of the dark matter particles.  In
694: this limit, the dynamics of the radiation determines the gravitational
695: potential and the dark matter just responds to that potential.  We will use
696: this simplification to obtain analytic estimates for the behavior of the
697: dark matter transfer function.
698: 
699: The primordial inflationary fluctuations lead to acoustic modes in the
700: radiation fluid during this era. The interaction rate of the particles in
701: the plasma is so high that we can consider the plasma as a perfect fluid
702: down to a comoving scale,
703: \begin{equation}
704: \lambda_f \sim \eta_{d}/\sqrt{N} \ \ \ \ ;  \ \ \ N \sim n \sigma t_d  ,
705: \end{equation}
706: where $\eta_d=\int_{0}^{t_d} dt/a(t)$ is the conformal time (i.e.  the
707: comoving size of the horizon) at the time of CDM decoupling, $t_d$;
708: $\sigma$ is the scattering cross section and $n$ is the relevant particle
709: density.  (Throughout this section we set the speed of light and Planck's
710: constant to unity for brevity.) The damping scale depends on the species
711: being considered and its contribution to the energy density, and is the
712: largest for neutrinos which are only coupled through weak interactions. In
713: that case $N\sim (T/T_{d}^\nu)^3$ where $T_{d}^\nu\sim 1\ {\rm MeV}$ is the
714: temperature of neutrino decoupling. At the time of CDM decoupling $N\sim
715: M/T_d \sim 10^4$ for the rest of the plasma, where $M$ is the mass of the
716: CDM particle. Here we will consider modes of wavelength larger than
717: $\lambda_f$, and so we neglect the effect of radiation diffusion damping
718: and treat the plasma (without the CDM) as a perfect fluid.
719: 
720: The equations of motion for a perfect fluid during the radiation era can be
721: solved analytically. We will use that solution here, following the notation
722: of Dodelson \cite{Dodelson}.  As usual we Fourier decompose fluctuations and
723: study the behavior of each Fourier component separately. For a mode of
724: comoving wavenumber $k$ in Newtonian gauge, the gravitational potential
725: fluctuations are given by:
726: \begin{equation}
727: \Phi= 3\Phi_p\left[{\sin(\omega\eta) -\omega\eta\cos(\omega\eta)
728: \over (\omega\eta)^3}\right],
729: \label{eq:phi}
730: \end{equation}
731: where $\omega=k/\sqrt{3}$ is the frequency of a mode and $\Phi_p$ is its
732: primordial amplitude in the limit $\eta \rightarrow 0$. In this section we
733: use conformal time $\eta=\int dt/a(t)$ with $a(t)\propto t^{1/2}$ during
734: the radiation-dominated era.  Expanding the temperature anisotropy in
735: multipole moments and using the Boltzmann equation to
736: describe their evolution, the monopole $\Theta_0$ and dipole $\Theta_1$ of
737: the photon distribution can be written in terms of the gravitational
738: potential as \cite{Dodelson}:
739: \begin{eqnarray}
740: \Theta_0&=&\Phi\left({x^2\over 6}+{1\over 2}\right)+{x\over 2}\Phi^\prime
741: \nonumber \\ \Theta_1&=&-{x^2\over 6}\left(\Phi^\prime +{1\over x}
742: \Phi\right)
743: \label{eq:theta01}
744: \end{eqnarray}
745: where $x\equiv k\eta$ and a prime denotes a derivative with respect to $x$.
746: 
747: The solutions in equations (\ref{eq:phi}) and (\ref{eq:theta01}) assume
748: that both the sound speed and the number of relativistic degrees of freedom
749: are constant over time. As a result of the QCD phase transition and of
750: various particles becoming non-relativistic, both of these assumptions are
751: not strictly correct. The vanishing sound speed during the QCD phase
752: transition provides the most dramatic effect, but its imprint is on scales
753: smaller than the ones we consider here because the transition occurs at a
754: significantly higher temperature and only lasts for a fraction of the
755: Hubble time \cite{Schmid}.
756: 
757: Before the dark matter decouples kinematically, we will treat it as a fluid
758: which can exchange momentum with the plasma through particle collisions. At
759: early times, the CDM fluid follows the motion of the plasma and is involved
760: in its acoustic oscillations. The continuity and momentum equations for the
761: CDM can be written as:
762: \begin{eqnarray}
763: \dot \delta_c+\theta_c &=& 3 \dot \Phi \nonumber \\ \dot \theta_c + {\dot a
764: \over a} \theta_c &=& k^2 c_s^2 \delta_c - k^2 \sigma_c - k^2 \Phi +
765: \tau_c^{-1} (\Theta_1 - \theta_c)
766: \label{eq:dmcontmom}
767: \end{eqnarray}
768: where a dot denotes an $\eta$-derivative, $\delta_c$ is the dark matter density
769: perturbation, $\theta_c$ is the divergence of the dark matter velocity
770: field and $\sigma_c$ denotes the anisotropic stress. In writing these
771: equations we have followed Ref. \cite{Ma}. The term $\tau_c^{-1} (\Theta_1 -
772: \theta_c)$ encodes the transfer of momentun between the radiation and CDM
773: fluids and $\tau_c^{-1}$ provides the collisional rate of momentum transfer,
774: \begin{equation}
775: \tau_c^{-1}= n \sigma {T \over M} a, 
776: \end{equation}  
777: with $n$ being the number density of particles with which the dark matter
778: is interacting, $\sigma(T)$ the average cross section for interaction and
779: $M$ the mass of the dark matter particle. The relevant scattering partners
780: are the standard model leptons which have thermal abundances.  For detailed
781: expressions of the cross section in the case of supersymmetric (SUSY) dark
782: matter, see Refs. \cite{Chen,Green2}. For our purpose, it is sufficient to
783: specify the typical size of the cross section and its scaling with cosmic
784: time,
785: \begin{equation}
786: \sigma\approx {T^2 \over M_\sigma^4}  ,
787: \end{equation}
788: where the coupling mass $M_\sigma$ is of the order of the weak-interaction
789: scale ($\sim 100$ GeV) for SUSY dark matter. This equation should be taken
790: as the definition of $M_\sigma$, as it encodes all the uncertainties in the
791: details of the particle physics model into a single parameter. The
792: temperature dependance of the averaged cross section is a result of the
793: available phase space. Our results are quite insensitive to the details
794: other than through the decoupling time. Equating $\tau_c^{-1}/a$ to the
795: Hubble expansion rate gives the temperature of kinematic decoupling:
796: \begin{equation}
797: T_d= \left({M_\sigma^4 M \over M_{pl}}\right)^{1/4}\approx 10 \ {\rm MeV}
798: \left({M_\sigma \over 100 \ {\rm GeV}}\right) \left( M \over 100\ {\rm GeV}
799: \right)^{1/4}  .
800: \end{equation}
801: 
802: The term $k^2 c_s^2 \delta_c$ in Eq. (\ref{eq:dmcontmom}) results from the
803: pressure gradient force and $c_s$ is the dark matter sound speed.  In the
804: tight coupling limit, $\tau_c \ll H^{-1}$ we find that $c_s^2\approx f_c
805: T/M$ and that the shear term is $k^2 \sigma_c \approx f_v c_s^2 \tau_c
806: \theta_c$.  Here $f_{v}$ and $f_{c}$ are constant factors of order unity.
807: We will find that both these terms make a small difference on the scales of
808: interest, so their precise value is unimportant.
809: 
810: By combining both equations in (\ref{eq:dmcontmom}) into a single equation
811: for $\delta_c$ we get
812: \begin{eqnarray}
813: \delta_c^{\prime\prime}&+& {1\over x}\left[1+ F_{\rm
814: v}(x)\right]\delta_c^\prime +c_s^2(x) \delta_c \nonumber \\ = &S(x)&
815: -3F_{\rm v}(x)\Phi^\prime+{x_d^4 \over x^5} \left(3\theta_0^\prime
816: -\delta_c^\prime\right),
817: \label{eq:delta}
818: \end{eqnarray}
819: where $x_d=k\eta_d$ and $\eta_d$ denotes the time of kinematic decoupling
820: which can be expressed in terms of the decoupling temperature as,
821: \begin{eqnarray}
822: \eta_d= {2 t_d (1+z_d) }\approx {M_{pl} \over T_0 T_d} &\approx& 10 \ {\rm
823: pc} \ \left({T_d \over 10\ {\rm MeV}}\right)^{-1} \nonumber \\ &\propto&
824: M_\sigma^{-1} M^{-1/4} ,
825: \end{eqnarray}
826: with $T_0=2.7$K being the present-day CMB temperature and $z_d$ being the
827: redshift at kinematic decoupling.  We have also introduced the source
828: function,
829: \begin{equation}
830: S(x)\equiv -3\Phi^{\prime\prime}+\Phi-{3\over x}\Phi^\prime.
831: \label{eq:s}
832: \end{equation}
833: For $x\ll x_d$, the dark matter sound speed is given by 
834: \begin{equation}
835: c_s^2(x)=c_s^2(x_d) {x_d\over x}, 
836: \label{eq:sound}
837: \end{equation}
838: where $c_s^2(x_d)$ is the dark matter sound speed at kinematic decoupling
839: (in units of the speed of light),
840: \begin{equation}
841: c_s(x_d) \approx 10^{-2} f_c^{1/2} \left({T_d \over 10\ {\rm MeV}}
842: \right)^{1/2} \left({M \over 100\ {\rm GeV}} \right)^{-1/2}.
843: \end{equation}
844: In writing (\ref{eq:sound}) we have assumed that prior to decoupling the
845: temperature of the dark matter follows that of the plasma. For the
846: viscosity term we have,
847: \begin{equation}
848: F_{v}(x)= f_{v} c_s^2(x_d) x_d^2  \left({x_d\over x}\right)^5.
849: \label{eq:F_v_before}
850: \end{equation}
851: 
852: \noindent{\bf Free streaming after kinematic decoupling}
853: 
854: In the limit of the collision rate being much slower than the Hubble
855: expansion, the CDM is decoupled and the evolution of its perturbations is
856: obtained by solving a Boltzman equation: \begin{equation} {\partial f \over
857: \partial \eta} +{d r_i \over d\eta} {\partial f \over \partial r_i} + {d
858: q_i \over d\eta} {\partial f \over \partial q_i} =0,
859: \end{equation}
860: where $f(\vec r,\vec q,\eta)$ is the distribution function which depends on
861: position, comoving momentum $\vec q$, and time. The comoving momentum
862: 3-components are ${d x_i / d\eta} = q_i/a$. We use the Boltzman equation to
863: find the evolution of modes that are well inside the horizon with $x\gg
864: 1$. In the radiation era, the gravitational potential decays after horizon
865: crossing (see Eq. \ref{eq:phi}). In this limit the comoving momentum
866: remains constant, ${d q_i /d\eta} =0$ and the Boltzman equation becomes,
867: \begin{equation}
868: {\partial f \over \partial \eta} +{q_i \over a} {\partial f \over \partial
869: r_i} =0.
870: \end{equation}
871: We consider a single Fourier mode and write $f$ as, 
872: \begin{equation}
873: f(\vec r,\vec q,\eta)=f_0(q) [1+\delta_F(\vec q,\eta) e^{i \vec k \cdot
874: \vec r}],
875: \end{equation}
876: where $f_0(q)$ is the unperturbed distribution,
877: \begin{equation}
878: f_0(q) = n_{\rm CDM} \left( {M \over 2\pi T_{\rm CDM} } \right)^{3/2}
879: \exp\left[-{1\over 2} {M q^2 \over T_{\rm CDM}}\right]
880: \end{equation} 
881: where $n_{\rm CDM}$ and $T_{\rm CDM}$ are the present-day density and
882: temperature of the dark matter.
883: 
884: Our approach is to solve the Boltzman equation with initial conditions
885: given by the fluid solution at a time $\eta_*$ (which will depend on
886: $k$). The simplified Boltzman equation can be easily solved to give
887: $\delta_F(\vec q,\eta)$ as a function of the initial conditions
888: $\delta_F(\vec q,\eta_*)$,
889: \begin{equation}
890: \delta_F(\vec q,\eta)=\delta_F(\vec q,\eta_*) \exp[-i \vec q \cdot \vec k
891: {\eta_* \over a(\eta_*)} \ln (\eta/\eta_*)] .
892: \label{solbol}
893: \end{equation}   
894: 
895: The CDM overdensity $\delta_c$ can then be expressed in terms of the
896: perturbation in the distribution function as,
897: \begin{equation}
898: \delta_c(\eta)={1\over n_{\rm CDM}} \int d^3q \ f_0(q) \ \delta_F(\vec q,\eta).
899: \end{equation} 
900: We can use (\ref{solbol}) to obtain the evolution of $\delta_c$ in terms of
901: its value at $\eta_*$,
902: \begin{equation}
903: \delta_c(\eta)= \exp\left[-{1\over 2} {k^2 \over k_f^2} \ln^2({\eta\over
904: \eta_*})\right] \ \left[\delta|_{\eta_*} + {d \delta \over d
905: \eta}|_{\eta_*} \eta_* \ln({\eta\over \eta_*})\right] ,
906: \label{delcbol}
907: \end{equation}
908: where $k_f^{-2}= \sqrt{(T_d/ M)} \eta_d$. The exponential term is
909: responsible for the damping of perturbations as a result of free streaming and
910: the dispersion of the CDM particles after they decouple from the plasma. The
911: above expression is only valid during the radiation era. The free streaming
912: scale is simply given by $\int dt (v/a) \propto \int dt a^{-2}$ which grows
913: logarithmically during the radiation era as in equation (\ref{delcbol}) but
914: stops growing in the matter era when $a\propto t^{2/3}$.
915: 
916: Equation (\ref{delcbol}) can be used to show that even during the free
917: streaming epoch, $\delta_c$ satisfies equation (\ref{eq:delta}) but with a
918: modified sound speed and viscous term. For $x \gg x_d$ one should use,
919: \begin{eqnarray}
920: c_s^2(x) &=& c_s^2(x_d) \left({x_d\over x}\right)^2 \left[1+ x_d^2
921: c_s^2(x_d) \ln^2({x\over x_d})\right] \nonumber \\ F_{v}(x)&=&2 c_s^2(x_d)
922: x_d^2 \ln\left({x_d\over x}\right)
923: \end{eqnarray}
924: The differences between the above scalings and those during the tight
925: coupling regime are a result of the fact that the dark matter temperature
926: stops following the plasma temperature but rather scales as $a^{-2}$ after
927: thermal decoupling, which coincides with the kinematic decoupling.
928: We ignore the effects of heat transfer during the
929: fluid stage of the CDM because its temperature is controlled by the much
930: larger heat reservoir of the radiation-dominated plasma at that stage.
931: 
932: To obtain the transfer function we solve the dark matter fluid equation
933: until decoupling and then evolve the overdensity using equation
934: (\ref{delcbol}) up to the time of matter--radiation equality. In practice,
935: we use the fluid equations up to $x_*=10\  {\rm max}(x_d, 10)$ so as to switch
936: into the free streaming solution well after the gravitational potential has
937: decayed. In the fluid equations, we smoothly match the sound speed and
938: viscosity terms at $x = x_d$. As mentioned earlier, because $c_s(x_d)$ is
939: so small and we are interested in modes that are comparable to the size of
940: the horizon at decoupling, i.e. $x_d \sim {\rm few}$, both the dark
941: matter sound speed and the associated viscosity play only a minor
942: role, and our simplified treatment is adequate.
943: 
944: \begin{figure}
945: \centering
946: \includegraphics[height=6cm]{time.ps}
947: \caption{The normalized amplitude of CDM fluctuations $\delta/\Phi_P$ for a
948: variety of modes with comoving wavenumbers
949: $\log(k\eta_d)=(0,1/3,2/3,1,4/3,5/3,2)$ as a function of $x\equiv k\eta$,
950: where $\eta=\int_0^t dt/a(t)$ is the conformal time coordinate. The dashed
951: line shows the temperature monopole $3\theta_0$ and the uppermost (dotted)
952: curve shows the evolution of a mode that is uncoupled to the cosmic
953: plasma.}
954: \label{figtime}
955: \end{figure}
956: 
957: In Figure \ref{figtime} we illustrate the time evolution of modes during
958: decoupling for a variety of $k$ values. The situation is clear. Modes that
959: enter the horizon before kinematic decoupling oscillate with the
960: radiation fluid. This behavior has two important effects. In the absence of
961: the coupling, modes receive a ``kick" by the source term $S(x)$ as they
962: cross the horizon. After that they grow logarithmically. In our case, modes
963: that entered the horizon before kinematic decoupling follow the plasma
964: oscillations and thus miss out on both the horizon ``kick" and the
965: beginning of the logarithmic growth. Second, the decoupling from the
966: radiation fluid is not instantaneous and this acts to further damp the
967: amplitude of modes with $x_d \gg 1$. This effect can be understood as
968: follows.  Once the oscillation frequency of the mode becomes high compared
969: to the scattering rate, the coupling to the plasma effectively damps the
970: mode. In that limit one can replace the forcing term $\Theta_0^\prime$ by
971: its average value, which is close to zero. Thus in this regime, the
972: scattering is forcing the amplitude of the dark matter oscillations to
973: zero.  After kinematic decoupling the modes again grow logarithmically but
974: from a very reduced amplitude. {\it The coupling with the plasma induces
975: both oscillations and damping of modes that entered the horizon before
976: kinematic decoupling. This damping is different from the free streaming
977: damping that occurs after kinematic decoupling}.
978: 
979: 
980: \begin{figure}
981: \centering
982: \includegraphics[height=6cm]{tf.ps}
983: \caption{Transfer function of the CDM density perturbation amplitude
984: (normalized by the primordial amplitude from inflation).  We show two
985: cases: {\it (i)} $T_d/M=10^{-4}$ and $T_d/T_{\rm eq}=10^7$; {\it (ii)}
986: $T_d/M=10^{-5}$ and $T_d/T_{\rm eq}=10^7$. In each case the oscillatory
987: curve is our result and the other curve is the free-streaming only result
988: that was derived previously in the literature [4,7,8].  }
989: \label{figtransfer}
990: \end{figure}
991: 
992: In Figure \ref{figtransfer} we show the resulting transfer function of the
993: CDM overdensity. The transfer function is defined as the ratio between the
994: CDM density perturbation amplitude $\delta_c$ when the effect of the
995: coupling to the plasma is included and the same quantity in a model where
996: the CDM is a perfect fluid down to arbitrarily small scales (thus, the
997: power spectrum is obtained by multiplying the standard result by the square
998: of the transfer function).  This function shows both the oscillations and
999: the damping signature mentioned above. The peaks occur at multipoles of the
1000: horizon scale at decoupling,
1001: \begin{equation} k_{peak}= (8,15.7,24.7,..) \eta_d^{-1} \propto {M_{pl} \over
1002: T_0 T_d}.
1003: \end{equation} This same scale determines the ``oscillation" damping. The
1004: free streaming damping scale is,
1005: \begin{equation}
1006: \eta_d c_d(\eta_d) \ln(\eta_{eq}/\eta_d) \propto {M_{pl} M^{1/2} \over T_0
1007: T_d^{3/2}} \ln(T_d/T_{\rm eq}),
1008: \end{equation}
1009: where $T_{\rm eq}$ is the temperature at matter radiation equality,
1010: $T_{\rm eq}\approx 1\ {\rm eV}$. The free streaming scale is parametrically
1011: different from the ``oscillation" damping scale. However for our fiducial
1012: choice of parameters for the CDM particle they roughly coincide.
1013: 
1014: The CDM damping scale is significantly smaller than the scales observed
1015: directly in the Cosmic Microwave Background or through large scale
1016: structure surveys. For example, the ratio of the damping scale to the scale
1017: that entered the horizon at Matter-radiation equality is
1018: $\eta_{d}/\eta_{eq}\sim T_{eq}/T_d \sim 10^{-7}$ and to our present horizon
1019: $\eta_{d}/\eta_{0}\sim (T_{eq} T_0)^{1/2}/T_{d} \sim 10^{-9}$. In the
1020: context of inflation, these scales were created 16 and 20 {\it e}--folds
1021: apart. Given the large extrapolation, one could certainly imagine that a
1022: change in the spectrum could alter the shape of the power spectrum around
1023: the damping scale.  However, for smooth inflaton potentials with small
1024: departures from scale invariance this is not likely to be the case. On
1025: scales much smaller than the horizon at matter radiation equality, the
1026: spectrum of perturbations density before the effects of the damping are
1027: included is approximately,
1028: \begin{eqnarray}
1029: \Delta^2(k)&\propto& \exp\left[(n-1)\ln(k\eta_{eq})+ {1\over 2} \alpha^2
1030: \ln(k\eta_{eq})^2 + \cdots\right]\nonumber \\ && \times \ln^2(k\eta_{eq}/8)
1031: \end{eqnarray}
1032: where the first term encodes the shape of the primordial spectrum and the
1033: second the transfer function. Primordial departures from scale invariance
1034: are encoded in the slope $n$ and its running $\alpha$. The effective slope
1035: at scale $k$ is then,
1036: \begin{equation}
1037: {\partial \ln \Delta^2 \over \partial \ln k}= (n-1)+\alpha \ln(k\eta_{eq})
1038: + {2 \over \ln(k\eta_{eq}/8)}.
1039: \end{equation}
1040: For typical values of $(n-1) \sim 1/60$ and $\alpha\sim 1/60^2$ the slope
1041: is still positive at $k\sim \eta_d^{-1}$, so the cut-off in the power will
1042: come from the effects we calculate rather than from the shape of the
1043: primordial spectrum. However given the large extrapolation in scale, one
1044: should keep in mind the possibility of significant effects resulting from
1045: the mechanisms that generates the density perturbations.
1046: 
1047: \noindent{\bf Implications}
1048: We have found that acoustic oscillations, a relic from the epoch when the
1049: dark matter coupled to the cosmic radiation fluid, truncate the CDM power
1050: spectrum on a comoving scale larger than effects considered before, such as
1051: free-streaming and viscosity \cite{Green,Green2,Hofmann}. For SUSY dark
1052: matter, the minimum mass of dark matter clumps that form in the Universe is
1053: therefore increased by more than an order of magnitude to a value of
1054: \footnote{Our definition of the cut-off mass follows the convention of the
1055: Jeans mass, which is defined as the mass enclosed within a sphere of radius
1056: $\lambda_{\rm J}/2$ where $\lambda_{\rm J}\equiv 2\pi/k_{\rm J}$ is the
1057: Jeans wavelength \cite{Haiman}.}
1058: \begin{eqnarray}
1059: M_{\rm cut}&=& {4\pi\over 3} \left({\pi \over k_{\rm cut}}\right)^{3}
1060: \Omega_M \rho_{\rm crit} \nonumber \\ &\simeq& 10^{-4} \left({T_d\over
1061: 10~{\rm MeV}}\right)^{-3} M_\odot,
1062: \label{mcut}
1063: \end{eqnarray}
1064: where $\rho_{\rm crit}=(H_0^2/8\pi G)=9\times 10^{-30}~{\rm g~cm^{-3}}$ is
1065: the critical density today, and $\Omega_M$ is the matter density for the
1066: concordance cosmological model \cite{WMAP}. We define the cut-off
1067: wavenumber $k_{\rm cut}$ as the point where the transfer function first
1068: drops to a fraction $1/e$ of its value at $k\rightarrow 0$. This
1069: corresponds to $k_{\rm cut}\approx 3.3 \ \eta_d^{-1}$.
1070: 
1071: \begin{figure}
1072: \centering
1073: %\picplace{5cm}{2cm}{history.ps}
1074: \includegraphics[height=6cm]{moo.eps}
1075: \label{figMoore}
1076: \caption{A slice through a numerical simulation of the first 
1077: dark matter condensations to form in the Universe. Colors represent
1078: the dark matter density at $z=26$. The simulated volume is 60 comoving pc
1079: on a side, simulated with 64 million particles each weighing
1080: $1.2\times 10^{-10}M_\odot$ (!). (from Diemand, Moore, \& Stadel 2005 \cite{Moore}). }
1081: \end{figure} 
1082: 
1083: Recent numerical simulations \cite{Moore,Gao} of the earliest and smallest
1084: objects to have formed in the Universe (see Fig. \ref{figMoore}), need to
1085: be redone for the modified power spectrum that we calculated in this
1086: section.  Although it is difficult to forecast the effects of the acoustic
1087: oscillations through the standard Press-Schechter formalism \cite{Press},
1088: it is likely that the results of such simulations will be qualitatively the
1089: same as before except that the smallest clumps would have a mass larger
1090: than before (as given by Eq. \ref{mcut}).
1091: 
1092: Potentially, there are several observational signatures of the smallest CDM
1093: clumps.  As pointed out in the literature \cite{Moore,Stoehr}, the smallest
1094: CDM clumps could produce $\gamma$-rays through dark-matter annihilation in
1095: their inner density cusps, with a flux in excess of that from nearby dwarf
1096: galaxies.  If a substantial fraction of the Milky Way halo is composed of
1097: CDM clumps with a mass $\sim 10^{-4} M_\odot$, the nearest clump is
1098: expected to be at a distance of $\sim 4\times 10^{17}$ cm.  Given that the
1099: characteristic speed of such clumps is a few hundred ${\rm km~s^{-1}}$, the
1100: $\gamma$-ray flux would therefore show temporal variations on the
1101: relatively long timescale of a thousand years.  Passage of clumps through
1102: the solar system should also induce fluctuations in the detection rate of
1103: CDM particles in direct search experiments.
1104: Other observational effects have rather limited prospects for
1105: detectability.  Because of their relatively low-mass and large size ($\sim
1106: 10^{17}~{\rm cm}$), the CDM clumps are too diffuse to produce any
1107: gravitational lensing signatures (including {\it femto-}lensing
1108: \cite{Gould}), even at cosmological distances. 
1109: 
1110: 
1111: The smallest CDM clumps should not affect the intergalactic baryons which
1112: have a much larger Jeans mass. However, once objects above
1113: $\sim 10^6M_\odot$ start to collapse at redshifts $z<30$, the baryons would
1114: be able to cool inside of them via molecular hydrogen transitions and the
1115: interior baryonic Jeans mass would drop.  The existence of dark matter
1116: clumps could then seed the formation of the first stars inside these
1117: objects \cite{Bromm}.
1118: %, but it would be difficult to infer $M_{\rm cut}$
1119: %indirectly from the mass function of the first stars.
1120: 
1121: %\subsubsection{Growth through Matter-Radiation Equality and 
1122: %Matter Domination}
1123: 
1124: \subsection{Structure of the Baryons}
1125: 
1126: \noindent{\bf Early Evolution of Baryonic Perturbations on Large Scales}
1127: 
1128: The baryons are coupled through Thomson scattering to the radiation fluid
1129: until they become neutral and decouple. After cosmic recombination, they
1130: start to fall into the potential wells of the dark matter and their early
1131: evolution was derived by Barkana \& Loeb (2005) \cite{BLinf}.
1132: 
1133: On large scales, the dark matter (dm) and the baryons (b) are affected only
1134: by their combined gravity and gas pressure can be ignored. The evolution of
1135: sub-horizon linear perturbations is described in the matter-dominated
1136: regime by two coupled second-order differential equations \cite{Peebles}:
1137: \beqa \ddot{\delta}_{\rm dm} + 2H \dot {\delta}_{\rm dm} & = & 4 \pi G
1138: \bar{\rho}_m \left(f_{\rm b} \delta_{\rm b} + f_{\rm dm} \delta_{\rm
1139: dm}\right)\ , \nonumber \\ \ddot{\delta}_{\rm b}+ 2H \dot {\delta}_{\rm b}
1140: & = & 4 \pi G \bar{\rho}_m \left(f_{\rm b} \delta_{\rm b} + f_{\rm dm}
1141: \delta_{\rm dm}\right)\ , \label{coupled}\eeqa where $\delta_{\rm dm}(t)$
1142: and $\delta_{\rm b}(t)$ are the perturbations in the dark matter and
1143: baryons, respectively, the derivatives are with respect to cosmic time $t$,
1144: $H(t)=\dot{a}/a$ is the Hubble constant with $a=(1+z)^{-1}$, and we assume
1145: that the mean mass density $\bar{\rho}_m(t)$ is made up of respective mass
1146: fractions $f_{\rm dm}$ and $f_{\rm b}=1-f_{\rm dm}$. Since these linear
1147: equations contain no spatial gradients, they can be solved spatially for
1148: $\delta_{\rm dm}(\bx,t)$ and $\delta_{\rm b}(\bx,t)$ or in Fourier space
1149: for $\td_{\rm dm}(\bk,t)$ and $\td_{\rm b}(\bk,t)$.
1150: 
1151: Defining $\delta_{\rm tot}
1152: \equiv f_{\rm b} \delta_{\rm b} + f_{\rm dm} \delta_{\rm dm}$ and
1153: $\delta_{\rm b-} \equiv \delta_{\rm b} - \delta_{\rm tot}$ , we find
1154: \beqa \ddot{\delta}_{\rm tot} + 2H \dot {\delta}_{\rm tot} & = & 4 \pi
1155: G \bar{\rho}_m \delta_{\rm tot}\ , \nonumber \\ \ddot{\delta}_{\rm
1156: b-}+ 2H \dot{\delta}_{\rm b-} & = & 0\ .  \eeqa Each of these
1157: equations has two independent solutions. The equation for $\delta_{\rm
1158: tot}$ has the usual growing and decaying solutions, which we denote
1159: $D_1(t)$ and $D_4(t)$, respectively, while the $\delta_{\rm b-}$
1160: equation has solutions $D_2(t)$ and $D_3(t)$; we number the solutions
1161: in order of declining growth rate (or increasing decay rate). We
1162: assume an Einstein-de Sitter, matter-dominated Universe in the
1163: redshift range $z=20$--150, since the radiation contributes less than
1164: a few percent at $z < 150$, while the cosmological constant and the
1165: curvature contribute to the energy density less than a few percent at
1166: $z > 3$. In this regime $a \propto t^{2/3}$ and the solutions are
1167: $D_1(t)=a/a_i$ and $D_4(t)=(a/a_i)^{-3/2}$ for $\delta_{\rm tot}$, and
1168: $D_2(t)=1$ and $D_3(t)=(a/a_i)^{-1/2}$ for $\delta_{\rm b-}$, where we
1169: have normalized each solution to unity at the starting scale factor
1170: $a_i$, which we set at a redshift $z_i=150$. The observable baryon
1171: perturbation can then be written as \beq \td_{\rm b}(\bk,t) =
1172: \td_{\rm b-} + \td_{\rm tot} = \sum_{m=1}^4 \td_{m}(\bk)\, D_m(t)\
1173: , \label{eq:delb} \eeq and similarly for the dark matter perturbation,
1174: \beq \td_{\rm dm}={1\over f_{\rm dm}}\left(\td_{\rm
1175: tot}-f_b\td_b\right)= \sum_{m=1}^4 \td_{m}(\bk)\, C_m(t)\ ,
1176: \label{eq:deldm} 
1177: \eeq where $C_i=D_i$ for $i=1,4$ and $C_i=-(f_{\rm b}/f_{\rm dm})D_i$ for
1178: $i=2,3$. We may establish the values of $\td_{m}(\bk)$ by inverting the
1179: $4\times 4$ matrix ${\bf A}$ that relates the 4-vector
1180: $(\td_1,\td_2,\td_3,\td_4)$ to the 4-vector that represents the initial
1181: conditions $(\td_{\rm b},\td_{\rm dm},\dot{\td}_{\rm b},\dot{\td}_{\rm
1182: dm})$ at the initial time.
1183: 
1184: \begin{figure}
1185: \includegraphics[width=84mm]{binfallA1.eps}
1186: \caption{Redshift evolution of the amplitudes of the independent modes of
1187: the density perturbations (upper panel) and the temperature perturbations
1188: (lower panel), starting at redshift 150 (from Barkana \& Loeb 2005
1189: \cite{BLinf}). We show $m=1$ (solid curves), $m=2$ (short-dashed curves),
1190: $m=3$ (long-dashed curves), $m=4$ (dotted curves), and $m=0$ (dot-dashed
1191: curve). Note that the lower panel shows one plus the mode amplitude.}
1192: \label{fig:Tevol}
1193: \end{figure}
1194: 
1195: \begin{figure}
1196: \includegraphics[width=84mm]{binfallA2.eps}
1197: \caption{Power spectra and initial perturbation amplitudes versus
1198: wavenumber (from \cite{BLinf}). The upper panel shows $\td_{\rm b}$ (solid
1199: curves) and $\td_{\rm dm}$ (dashed curves) at $z=150$ and 20 (from bottom
1200: to top).  The lower panel shows the initial ($z=150$) amplitudes of $\td_1$
1201: (solid curve), $\td_2$ (short-dashed curve), $\td_3$ (long-dashed curve),
1202: $\td_4$ (dotted curve), and $\td_T(t_i)$ (dot-dashed curve). Note that if
1203: $\td_1$ is positive then so are $\td_3$ and $\td_T(t_i)$, while $\td_2$ is
1204: negative at all $k$, and $\td_4$ is negative at the lowest $k$ but is
1205: positive at $k > 0.017$ Mpc$^{-1}$.}
1206: \label{fig:del}
1207: \end{figure}
1208: 
1209: 
1210: Next we describe the fluctuations in the sound speed of the cosmic gas
1211: caused by Compton heating of the gas, which is due to scattering of the
1212: residual electrons with the CMB photons. The evolution of the temperature
1213: $T$ of a gas element of density $\rho_b$ is given by the first law of
1214: thermodynamics: \beq dQ = \frac{3} {2} \kB dT - \kB T d \log \rho_{\rm b}\
1215: , \eeq where $dQ$ is the heating rate per particle. Before the first
1216: galaxies formed, \beq \frac{dQ} {d t} = 4 \frac{\sigma_{\rm T}\, c} {m_e}
1217: \, \kB (T_\gamma - T) \rho_\gamma x_e(t) \ , \eeq where $\sigma_T$ is the
1218: Thomson cross-section, $x_e(t)$ is the electron fraction out of the total
1219: number density of gas particles, and $\rho_\gamma$ is the CMB energy
1220: density at a temperature $T_\gamma$. In the redshift range of interest, we
1221: assume that the photon temperature ($T_\gamma = T_\gamma^0/a$) is spatially
1222: uniform, since the high sound speed of the photons (i.e., $c/\sqrt{3}$)
1223: suppresses fluctuations on the sub-horizon scales that we consider, and the
1224: horizon-scale $\sim 10^{-5}$ fluctuations imprinted at cosmic recombination
1225: are also negligible compared to the smallWe establish the values of
1226: $\td_{m}(\bk)$ by inverting the $4\times 4$ matrix ${\bf A}$ that relates
1227: the 4-vector $(\td_1,\td_2,\td_3,\td_4)$ to the 4-vector that represents
1228: the initial conditions $(\td_{\rm b},\td_{\rm dm},\dot{\td}_{\rm
1229: b},\dot{\td}_{\rm dm})$ at the initial time.  er-scale fluctuations in the
1230: gas density and temperature. Fluctuations in the residual electron fraction
1231: $x_e(t)$ are even smaller.  Thus, \beq \frac{dT} {dt} = \frac{2} {3} T
1232: \frac{d \log \rho_{\rm b}} {dt} + \frac{x_e(t)}{t_\gamma}\, (T_\gamma -
1233: T)\, a^{-4}\ , \eeq where $t_\gamma^{-1} \equiv \bar{\rho}_\gamma^0
1234: ({8\sigma_{\rm T}\, c}/{3 m_e}) = 8.55 \times 10^{-13} {\rm yr}^{-1}$.
1235: After cosmic recombination, $x_e(t)$ changes due to the slow recombination
1236: rate of the residual ions: \beq {d x_e(t)\over dt} = - \alpha_B(T) x_e^2(t)
1237: \bar{n}_H (1+y)\ , \eeq where $\alpha_B(T)$ is the case-B recombination
1238: coefficient of hydrogen, $\bar{n}_H$ is the mean number density of hydrogen
1239: at time $t$, and $y=0.079$ is the helium to hydrogen number density
1240: ratio. This yields the evolution of the mean temperature, ${d \bar{T}}/{dt}
1241: = - 2 H \bar{T} + {x_e(t)}{t_\gamma^{-1}}\, (T_\gamma - \bar{T})\,
1242: a^{-4}$. In prior analyses \cite{Peebles, Ma} a spatially uniform
1243: speed of sound was assumed for the gas at each redshift. Note that we refer
1244: to $\delta p/ \delta \rho$ as the square of the sound speed of the fluid,
1245: where $\delta p$ is the pressure perturbation, although we are analyzing
1246: perturbations driven by gravity rather than sound waves driven by pressure
1247: gradients.
1248: 
1249: Instead of assuming a uniform sound speed, we find the first-order
1250: perturbation equation, \beq \frac{d \delta_T} {d t} = \frac{2}{3}
1251: \frac{d \delta_b} {dt} - \frac{x_e(t)} {t_\gamma} \frac{T_\gamma}
1252: {\bar{T}} a^{-4} \delta_T\ , \label{eq:order1} \eeq
1253: where we defined the fractional temperature perturbation $\delta_T$.  Like
1254: the density perturbation equations, this equation can be solved separately
1255: at each $\bx$ or at each $\bk$. Furthermore, the solution $\delta_T (t)$ is
1256: a linear functional of $\delta_b(t)$ [for a fixed function $x_e(t)$].
1257: Thus, if we choose an initial time $t_i$ then using Eq.~(\ref{eq:delb}) we
1258: can write the solution in Fourier space as \beq \td_T (\bk,t) =
1259: \sum_{m=1}^4 \td_{m}(\bk)\, D^T_m(t) + \td_T (\bk,t_i)\, D^T_0(t)\ ,
1260: \label{eq:delT} \eeq where $D^T_m(t)$ is the solution of
1261: Eq.~(\ref{eq:order1}) with $\delta_T = 0$ at $t_i$ and with the
1262: perturbation mode $D_m(t)$ substituted for $\delta_b(t)$, while $D^T_0(t)$
1263: is the solution with no perturbation $\delta_b(t)$ and with $\delta_T = 1$
1264: at $t_i$. By modifying the CMBFAST code (http://www.cmbfast.org/), we can
1265: numerically solve Eq.~(\ref{eq:order1}) along with the density perturbation
1266: equations for each $\bk$ down to $z_i=150$, and then match the solution to
1267: the form of Eq.~(\ref{eq:delT}).
1268: 
1269: \begin{figure}
1270: \includegraphics[width=84mm]{binfallA4.eps}
1271: \caption{Relative sensitivity of perturbation amplitudes at $z=150$ to
1272: cosmological parameters (from \cite{BLinf}). For variations in a parameter
1273: $x$, we show $d{\rm log}\sqrt{P(k)}/d{\rm log}(x)$. We consider variations
1274: in $\Omega_{\rm dm} h^2$ (upper panel), in $\Omega_b h^2$ (middle panel),
1275: and in the Hubble constant $h$ (lower panel). When we vary each parameter
1276: we fix the other two, and the variations are all carried out in a flat
1277: $\Omega_{\rm total}=1$ universe. We show the sensitivity of $\td_1$ (solid
1278: curves), $\td_2$ (short-dashed curves), $\td_3$ (long-dashed curves), and
1279: $\td_T(t_i)$ (dot-dashed curves).}
1280: \label{fig:sens}
1281: \end{figure}
1282: 
1283: Figure~\ref{fig:Tevol} shows the time evolution of the various independent
1284: modes that make up the perturbations of density and temperature, starting
1285: at the time $t_i$ corresponding to $z_i=150$. $D^T_2(t)$ is identically
1286: zero since $D_2(t)=1$ is constant, while $D^T_3(t)$ and $D^T_4(t)$ are
1287: negative. Figure~\ref{fig:del} shows the amplitudes of the various
1288: components of the initial perturbations. We consider comoving wavevectors
1289: $k$ in the range 0.01 -- 40 Mpc$^{-1}$, where the lower limit is set by
1290: considering sub-horizon scales at $z=150$ for which photon perturbations
1291: are negligible compared to $\delta_{\rm dm}$ and $\delta_{\rm b}$, and the
1292: upper limit is set by requiring baryonic pressure to be negligible compared
1293: to gravity. $\td_2$ and $\td_3$ clearly show a strong signature of the
1294: large-scale baryonic oscillations, left over from the era of the
1295: photon-baryon fluid before recombination, while $\td_1$, $\td_4$, and
1296: $\td_T$ carry only a weak sign of the oscillations. For each quantity, the
1297: plot shows $[k^3 P(k)/(2 \pi^2)]^{1/2}$, where $P(k)$ is the corresponding
1298: power spectrum of fluctuations.  $\td_4$ is already a very small correction
1299: at $z=150$ and declines quickly at lower redshift, but the other three
1300: modes all contribute significantly to $\td_{\rm b}$, and the $\td_T(t_i)$
1301: term remains significant in $\td_T(t)$ even at $z \la 100$. Note that at
1302: $z=150$ the temperature perturbation $\td_T$ has a different shape with
1303: respect to $k$ than the baryon perturbation $\td_{\rm b}$, showing that
1304: their ratio cannot be described by a scale-independent speed of sound.
1305: 
1306: The power spectra of the various perturbation modes and of
1307: $\td_T(t_i)$ depend on the initial power spectrum of density
1308: fluctuations from inflation and on the values of the fundamental
1309: cosmological parameters ($\Omega_{\rm dm}$, $\Omega_b$,
1310: $\Omega_{\Lambda}$, and $h$). If these independent power spectra can
1311: be measured through 21cm fluctuations, this will probe the basic
1312: cosmological parameters through multiple combinations, allowing
1313: consistency checks that can be used to verify the adiabatic nature and
1314: the expected history of the perturbations. Figure~\ref{fig:sens}
1315: illustrates the relative sensitivity of $\sqrt{P(k)}$ to variations in
1316: $\Omega_{\rm dm} h^2$, $\Omega_b h^2$, and $h$, for the quantities
1317: $\td_1$, $\td_2$, $\td_3$, and $\td_T(t_i)$. Not shown is $\td_4$,
1318: which although it is more sensitive (changing by order unity due to
1319: $10\%$ variations in the parameters), its magnitude always remains
1320: much smaller than the other modes, making it much harder to
1321: detect. Note that although the angular scale of the baryon
1322: oscillations constrains also the history of dark energy through the
1323: angular diameter distance, we have focused here on other cosmological
1324: parameters, since the contribution of dark energy relative to matter
1325: becomes negligible at high redshift.
1326: 
1327: 
1328: \noindent
1329: {\bf Cosmological Jeans Mass}
1330: 
1331: The Jeans length $\ljeans$ was originally defined (Jeans 1928 \cite{Jeans28}) in Newtonian
1332: gravity as the critical wavelength that separates oscillatory and
1333: exponentially-growing density perturbations in an infinite, uniform, and
1334: stationary distribution of gas. On scales $\ell$ smaller than $\ljeans$,
1335: the sound crossing time, $\ell/c_s$ is shorter than the gravitational
1336: free-fall time, $(G\rho)^{-1/2}$, allowing the build-up of a pressure force
1337: that counteracts gravity. On larger scales, the pressure gradient force is
1338: too slow to react to a build-up of the attractive gravitational force.  The
1339: Jeans mass is defined as the mass within a sphere of radius $\ljeans/2$,
1340: $\mjeans=(4\pi/3)\rho(\ljeans/2)^3$.  In a perturbation with a mass greater
1341: than $\mjeans$, the self-gravity cannot be supported by the pressure
1342: gradient, and so the gas is unstable to gravitational collapse. The
1343: Newtonian derivation of the Jeans instability suffers from a conceptual
1344: inconsistency, as the unperturbed gravitational force of the uniform
1345: background must induce bulk motions (compare to Binney \& Tremaine 1987
1346: \cite{Bi87}).  However, this inconsistency is remedied when the analysis is
1347: done in an expanding Universe.
1348: 
1349: The perturbative derivation of the Jeans instability criterion can be
1350: carried out in a cosmological setting by considering a sinusoidal
1351: perturbation superposed on a uniformly expanding background.  Here, as
1352: in the Newtonian limit, there is a critical wavelength $\ljeans$ that
1353: separates oscillatory and growing modes.  Although the expansion of
1354: the background slows down the exponential growth of the amplitude to a
1355: power-law growth, the fundamental concept of a minimum mass that can
1356: collapse at any given time remains the same (see, e.g. Kolb \& Turner
1357: 1990 \cite{Kolb90}; Peebles 1993 \cite{Peebles}).
1358: 
1359: We consider a mixture of dark matter and baryons with density parameters
1360: $\Omega_{\rm dm}^{\,z}=\bar{\rho}_{\rm dm}/\rho_{\rm c}$ and $\Omega_{\rm
1361: b}^{\,z}=\bar{\rho}_{\rm b}/\rho_{\rm c}$, where $\bar{\rho}_{\rm dm}$ is
1362: the average dark matter density, $\bar{\rho}_{\rm b}$ is the average
1363: baryonic density, $\rho_{\rm c}$ is the critical density, and $\Omega_{\rm
1364: dm}^{\,z} + \Omega_{\rm b}^{\,z} = \Ommz$ is given by
1365: equation(\ref{Ommz}). We also assume spatial fluctuations in the gas and
1366: dark matter densities with the form of a single spherical Fourier mode
1367: on a scale much smaller than the horizon,
1368: \begin{eqnarray}
1369: \frac{\rho_{\rm dm}(r,t)-\bar{\rho}_{\rm dm}(t)}
1370: {\bar{\rho}_{\rm dm}(t)} & = &
1371: \delta_{\rm dm}(t) \frac{\sin(kr)}{kr}\ , \\
1372: \frac{\rho_{\rm b}(r,t)-\bar{\rho}_{\rm b}(t)}
1373: {\bar{\rho}_{\rm b}(t)} & = &
1374: \delta_{\rm b}(t) \frac{\sin(kr)}{kr}\ ,
1375: \end{eqnarray}
1376: where $\bar{\rho}_{\rm dm}(t)$ and $\bar{\rho}_{\rm b}(t)$ are the
1377: background densities of the dark matter and baryons, $\delta_{\rm
1378: dm}(t)$ and $\delta_{\rm b}(t)$ are the dark matter and baryon
1379: overdensity amplitudes, $r$ is the comoving radial coordinate, and $k$
1380: is the comoving perturbation wavenumber.  We adopt an ideal gas
1381: equation-of-state for the baryons with a specific heat ratio
1382: $\gamma$=$5/3$.  Initially, at time $ t=t_{\rm i}$, the gas
1383: temperature is uniform $ T_{\rm b}(r,t_{\rm i})$=$ T_{\rm i}$, and the
1384: perturbation amplitudes are small $\delta_{\rm dm,i},\delta_{\rm
1385: b,i}\ll 1$.  We define the region inside the first zero of
1386: $\sin(kr)/(kr)$, namely $0<kr<\pi$, as the collapsing ``object''.
1387: 
1388: The evolution of the temperature of the baryons $T_{\rm b}(r,t)$ in
1389: the linear regime is determined by the coupling of their free
1390: electrons to the CMB through Compton
1391: scattering, and by the adiabatic expansion of the gas.  Hence, $
1392: T_{\rm b}(r,t)$ is generally somewhere between the CMB temperature,
1393: $T_{\gamma}\propto (1+z)^{-1}$ and the adiabatically-scaled
1394: temperature $T_{\rm ad}\propto (1+z)^{-2}$.  In the limit of tight
1395: coupling to $T_{\gamma}$, the gas temperature remains uniform.  On the
1396: other hand, in the adiabatic limit, the temperature develops a
1397: gradient according to the relation \beq T_{\rm b} \propto
1398: \rho_b^{(\gamma-1)}.  \eeq
1399: 
1400: The evolution of a cold dark matter
1401: overdensity, $\delta_{\rm dm}(t)$, in the linear regime is described by the
1402: equation (\ref{coupled}),
1403: \begin{equation}
1404:        {\ddot{\delta}}_{\rm dm}
1405:  + 2H{\dot \delta}_{\rm dm}
1406:        ={3\over 2}H^2 \left(\Omega_{\rm b} \delta_{\rm b} +
1407: \Omega_{\rm dm}
1408:         \delta_{\rm dm}\right)\label{dm}
1409: \end{equation}
1410: whereas the evolution of the overdensity of the baryons, $\delta_{\rm
1411: b}(t)$, with the inclusion of their pressure force is described by
1412: (see \S 9.3.2 of \cite{Kolb90}),
1413: \begin{equation}
1414:         {\ddot{\delta}}_{\rm b}+
1415:        2H{\dot \delta}_{\rm b} ={3\over 2}H^2
1416: \left(\Omega_{\rm b} \delta_{\rm b} +
1417:        \Omega_{\rm dm} \delta_{\rm dm}\right) -\frac{\kB T_{\rm
1418:        i}}{\mu m_p} \left({k\over a}\right)^2
1419:        \left(\frac{a_{\rm i}}{a} \right)^{(1+\beta)} \left(\delta_{\rm
1420:        b}+{2\over 3}\beta [\delta_{\rm b}-\delta_{\rm
1421:        b,i}]\right)\label{b}.
1422: \end{equation}
1423: Here, $H(t)={\dot a}/a$ is the Hubble parameter at a cosmological time
1424: $t$, and $\mu=1.22$ is the mean molecular weight of the neutral
1425: primordial gas in atomic units. The parameter $\beta$ distinguishes
1426: between the two limits for the evolution of the gas temperature. In
1427: the adiabatic limit $\beta=1$, and when the baryon temperature is
1428: uniform and locked to the background radiation, $\beta=0$. The last
1429: term on the right hand side (in square brackets) takes into account
1430: the extra pressure gradient force in $\nabla(\rho_{\rm b} T)=(T \nabla
1431: \rho_{\rm b}+\rho_{\rm b}\nabla T)$, arising from the temperature
1432: gradient which develops in the adiabatic limit. The Jeans wavelength
1433: $\ljeans=2\pi/\kjeans$ is obtained by setting the right-hand side of
1434: equation~(\ref{b}) to zero, and solving for the critical wavenumber
1435: $\kjeans$.  As can be seen from equation~(\ref{b}), the critical
1436: wavelength $\ljeans$ (and therefore the mass $\mjeans$) is in general
1437: time-dependent.  We infer from equation~(\ref{b}) that as time
1438: proceeds, perturbations with increasingly smaller initial wavelengths
1439: stop oscillating and start to grow.
1440: 
1441: To estimate the Jeans wavelength, we equate the right-hand-side of
1442: equation~(\ref{b}) to zero. We further approximate $\delta_{\rm
1443: b}\sim \delta_{\rm dm}$, and consider sufficiently high redshifts at
1444: which the Universe is matter dominated and flat, $(1+z)\gg {\rm
1445: max}[(1-\Omega_m-\Omega_\Lambda)/\Omega_m,
1446: (\Omega_\Lambda/\Omega_m)^{1/3}]$. In this regime, $\Omega_{\rm
1447: b}\ll\Omm \approx 1$, $H\approx 2/(3t)$, and
1448: $a=(1+z)^{-1}\approx (3H_0{\sqrt{\Omega_m}}/2)^{2/3}t^{2/3}$, where
1449: $\Omega_m=\Omega_{\rm dm}+\Omega_b$ is the total matter
1450: density parameter.  Following cosmological recombination at $z\approx
1451: 10^3$, the residual ionization of the cosmic gas keeps its temperature
1452: locked to the CMB temperature (via Compton scattering) down to a
1453: redshift of \cite{Peebles}
1454: \begin{equation}
1455: 1+z_t\approx 160 (\Omega_b h^2/0.022)^{2/5}\ .
1456: \end{equation}
1457: In the redshift range between recombination and $z_t$, $\beta=0$ and
1458: \beq
1459: \kjeans\equiv (2\pi/\ljeans)=[2\kB T_{\gamma}(0)/3\mu
1460: m_p]^{-1/2} {\sqrt{\Omega_m}} H_0\ ,
1461: \eeq
1462: so that the Jeans mass is therefore redshift independent and obtains the
1463: value (for the total mass of baryons and dark matter)
1464: \begin{equation}
1465: M_{\rm J}\equiv {4\pi\over 3} \left({\ljeans\over 2}\right)^3
1466: {\bar\rho}(0)
1467: = 1.35\times 10^5 \left({\Omega_mh^2\over 0.15}\right)^{-1/2} M_\odot\ .
1468: \end{equation}
1469: 
1470: Based on the similarity of $\mjeans$ to the mass of a globular
1471: cluster, Peebles \& Dicke (1968) \cite{Pe68} suggested that globular clusters form
1472: as the first generation of baryonic objects shortly after cosmological
1473: recombination. Peebles \& Dicke assumed a baryonic Universe, with a
1474: nonlinear fluctuation amplitude on small scales at $z\sim 10^3$, a
1475: model which has by now been ruled out. The lack of a dominant mass of dark
1476: matter inside globular clusters makes it unlikely that they formed
1477: through direct cosmological collapse, and more likely that they
1478: resulted from fragmentation during the process of galaxy formation.
1479: 
1480: 
1481: \begin{figure}
1482: \centering
1483: \includegraphics[height=6cm]{temp.eps}
1484: \caption{Thermal history of the baryons, left over from the big bang,
1485: before the first galaxies formed.  The
1486: residual fraction of free electrons couple the gas temperture $T_{\rm gas}$
1487: to the cosmic microwave background temperature [$T_\gamma\propto (1+z)$]
1488: until a redshift $z\sim 200$. Subsequently the gas temperature cools
1489: adiabatically at a faster rate [$T_{\rm gas}\propto (1+z)^2$].  Also
1490: shown is the spin temperature of the 21cm transition of hydrogen $T_{\rm
1491: s}$ which interpolates between the gas and radiation temperature
1492: and will be discussed in detail later in this review.}
1493: \end{figure}
1494: 
1495: At $z\la z_t$, the gas temperature declines adiabatically as
1496: $[(1+z)/(1+z_t)]^2$ (i.e., $\beta=1$) and the total Jeans mass obtains
1497: the value,
1498: \begin{equation}
1499: \mjeans= 4.54\times 10^3\left({\Omega_mh^2\over 0.15}\right)^{-1/2}
1500: \left({\Omega_b h^2\over 0.022}\right)^{-3/5}
1501: \left({1+z\over 10}\right)^{3/2}~M_\odot.
1502: \label{eq:m_j}
1503: \end{equation}
1504: 
1505: It is not clear how the value of the Jeans mass derived above relates
1506: to the mass of collapsed, bound objects. The above analysis is
1507: perturbative (Eqs.~\ref{dm} and \ref{b} are valid only as long as
1508: $\delta_{\rm b}$ and $\delta_{\rm dm}$ are much smaller than unity),
1509: and thus can only describe the initial phase of the collapse.  As
1510: $\delta_{\rm b}$ and $\delta_{\rm dm}$ grow and become larger than
1511: unity, the density profiles start to evolve and dark matter shells may
1512: cross baryonic shells \cite{Haiman94} due to their
1513: different dynamics. Hence the amount of mass enclosed within a given
1514: baryonic shell may increase with time, until eventually the dark
1515: matter pulls the baryons with it and causes their collapse even
1516: for objects below the Jeans mass.
1517: 
1518: %since material at various radii collapses at different times, there is
1519: %no reason to assume that the collapsed object includes the entire mass
1520: %within the original density peak.  Indeed, it has been realized
1521: %(Larson 1969) that in the non-linear regime the density and
1522: %temperature profiles steepen considerably and the central part of the
1523: %sphere becomes exceedingly dense and hot.  An object may form in the
1524: %center, and subsequently accrete material from the original peak.  In
1525: %the limit of high masses ($ \lambda\gg\lambda_{\rm Jeans}$) and in the
1526: %absence of radiative cooling, the gas particles virialize through a
1527: %shock that propagates outwards from the center of the sphere.  The
1528: %shock structure and the evolution of the density profile is
1529: %self-similar, and near the center the density approaches the power-law
1530: %profile $\rho\propto r^{-2.25}$ (this was first predicted by Gott
1531: %1975, and later proved via a semi-analytic approach by Bertschinger
1532: %1985).  Although the initial perturbation has a definite mass, it is
1533: %not clear what mass to associate with the end-product of the collapse.
1534: %The relation between the mass of the bound object $M_{\rm obj}$ and
1535: %the parameters of the initial perturbation can be expressed as $M_{\rm
1536: %obj}=M_{\rm obj}(k,\delta_{\rm dm,i},\delta_{\rm b,i},t_{\rm i})$.
1537: %This relation cannot be inferred directly from linear theory.
1538: 
1539: Even within linear theory, the Jeans mass is related only to the evolution
1540: of perturbations at a given time. When the Jeans mass itself varies with
1541: time, the overall suppression of the growth of perturbations depends on a
1542: time-weighted Jeans mass. Gnedin \& Hui (1998) \cite{Gnedin98} showed that
1543: the correct time-weighted mass is the filtering mass $M_F=(4 \pi/3)\,
1544: \bar{\rho}\, (2 \pi a/k_F)^3$, in terms of the comoving wavenumber $k_F$
1545: associated with the ``filtering scale'' (note the change in convention from
1546: $\pi/k_J$ to $2\pi/k_F$). The wavenumber $k_F$ is related to the Jeans
1547: wavenumber $\kjeans$ by \beq \frac{1}{k^2_F (t)}=\frac{1}{D(t)} \int_0^t
1548: dt' \, a^2(t') \frac{\ddot{D} (t')+ 2 H(t') \dot{D}(t')} {k_J^2 (t')} \,
1549: \int_{t'}^t \frac{dt''} {a^2(t'')}\ , \eeq where $D(t)$ is the linear
1550: growth factor.  At high redshift (where $\Ommz \rightarrow 1$), this
1551: relation simplifies to \cite{Gnedin2000b} \beq \frac{1}{k^2_F (t)}=
1552: \frac{3}{a} \int_0^a \frac{d a'}{k_J^2(a')} \left( 1-\sqrt{ \frac{a'} {a}}~
1553: \right)\ . \eeq Then the relationship between the linear overdensity of the
1554: dark matter $\delta_{\rm dm}$ and the linear overdensity of the baryons
1555: $\delta_b$, in the limit of small $k$, can be written as \cite{Gnedin98}
1556: \beq \frac{\delta_b} {\delta_{\rm dm}} = 1-\frac{k^2}{k_F^2}+O (k^4)\
1557: . \eeq
1558: 
1559: Linear theory specifies whether an initial perturbation, characterized
1560: by the parameters $ k$, $\delta_{\rm dm,i}$, $\delta_{\rm b,i}$ and
1561: $t_{\rm i}$, begins to grow.  To determine the minimum mass of
1562: nonlinear baryonic objects resulting from the shell-crossing and
1563: virialization of the dark matter, we must use a different model which
1564: examines the response of the gas to the gravitational potential of a
1565: virialized dark matter halo.
1566: 
1567: 
1568: %<>
1569: \subsection{Formation of Nonlinear Objects}
1570: \label{sec2.3}
1571: 
1572: \subsection{Spherical Collapse}
1573: 
1574: Let us consider a spherically symmetric density or velocity perturbation of
1575: the smooth cosmological background, and examine the dynamics of a test
1576: particle at a radius $r$ relative to the center of symmetry.  Birkhoff's
1577: (1923) \cite{Birk23} theorem implies that we may ignore the mass outside this radius in
1578: computing the motion of our particle.  We further find that the
1579: relativistic equations of motion describing the system reduce to the usual
1580: Friedmann equation for the evolution of the scale factor of a homogeneous
1581: Universe, but with a density parameter $\Omega$ that now takes account of
1582: the additional mass or peculiar velocity.  In particular, despite the
1583: arbitrary density and velocity profiles given to the perturbation, only the
1584: total mass interior to the particle's radius and the peculiar velocity at
1585: the particle's radius contribute to the effective value of $\Omega$.  We
1586: thus find a solution to the particle's motion which describes its departure
1587: from the background Hubble flow and its subsequent collapse or expansion.
1588: This solution holds until our particle crosses paths with one from a
1589: different radius, which happens rather late for most initial profiles.
1590: 
1591: As with the Friedmann equation for a smooth Universe, it is possible to
1592: reinterpret the problem into a Newtonian form.  Here we work in an inertial
1593: (i.e.  non-comoving) coordinate system and consider the force on the
1594: particle as that resulting from a point mass at the origin (ignoring
1595: the possible presence of a vacuum energy density):
1596: %1-38
1597: \begin{equation}\label{eq:1a}
1598: \frac {d^2r} {dt^2} = - \frac {GM}{r^2}, 
1599: \end{equation}
1600: where $G$ is Newton's constant, $r$ is the distance of the particle from
1601: the center of the spherical perturbation, and $M$ is the total mass within
1602: that radius.  As long as the radial shells do not cross each other, the
1603: mass $M$ is constant in time.  The initial density profile determines $M$,
1604: while the initial velocity profile determines $dr/dt$ at the initial time.
1605: As is well-known, there are three branches of solutions: one in which the
1606: particle turns around and collapses, another in which it reaches an
1607: infinite radius with some asymptotically positive velocity, and a third
1608: intermediate case in which it reachs an infinite radius but with a velocity
1609: that approaches zero.  These cases may be written as \cite{Gu72}:
1610: %1-39
1611: \begin{align}
1612: \label{eq:2} 
1613: \left.
1614: \begin{array}{r}
1615: r = A(\cos \eta -1)  \\  
1616: t = B(\eta - \sin \: \eta) 
1617: \end{array} 
1618: \right\}
1619: &&{\rm Closed} &&(0 \leq \eta \leq 2 \pi )
1620: \end{align}
1621: %\end{equation}
1622: \begin{align}\label{eq:3}
1623: \left.
1624: \begin{array}{r}
1625: r = A{\eta ^2} /2 \\ 
1626: t = {B\eta^3} / 6 
1627: \end{array}
1628: \right\}
1629: &&{\rm Flat} &&(0 \leq \eta \leq \infty ) 
1630: \end{align}
1631: \begin{align}\label{eq:4}
1632: \left.
1633: \begin{array}{r}
1634: r = A({\rm cosh} \: \eta -1)  \\  
1635: t = B({\rm sinh} \: \eta - \eta) 
1636: \end{array} 
1637: \right\}
1638: &&{\rm Open} &&(0 \leq \eta \leq \infty )
1639: \end{align}
1640: where ${A^3} = GM{B^2}$ applies in all cases.  All three solutions have
1641: $r^3 = 9GM{t^2}/2$ as $t$ goes to zero, which matches the linear theory
1642: expectation that the perturbation amplitude get smaller as one goes back
1643: in time.  In the closed case, the shell turns around at time $\pi B$ and
1644: radius $2A$ and collapses to zero radius at time $2\pi B$.
1645: 
1646: We are now faced with the problem of relating the spherical collapse
1647: parameters $A, B,$ and $M$ to the linear theory density perturbation
1648: $\delta$ \cite{p80}.  We do this by returning to the equation of
1649: motion.  Consider that at an early epoch (i.e. scale factor $a_i \ll$ 1),
1650: we are given a spherical patch of uniform overdensity $\delta_i$ (the
1651: so-called `top-hat' perturbation).  If $\Omega$ is essentially unity at
1652: this time and if the perturbation is pure growing mode, then the initial
1653: velocity is radially inward with magnitude $\delta_i H( t_i )r/3$, where
1654: $H(t_i)$ is the Hubble constant at the initial time and $r$ is the radius
1655: from the center of the sphere. This can be easily seen from the continuity
1656: equation in spherical coordinates.  The equation of motion (in noncomoving
1657: coordinates) for a particle beginning at radius $r_i$ is simply
1658: %1-42
1659: \begin{equation}\label{eq:5} 
1660: \frac {d^2r}{dt^2} = - \frac{GM}{r^2} + \frac {\Lambda r}{3}, 
1661: \end{equation}
1662: where $M = (4 \pi /3) r^3_i \rho_i (1 + \delta_i) $ and $\rho_i$
1663: is the background density of the Universe at time $t_i$.  We next define
1664: the dimensionless radius $x = ra_i/r_i$ and rewrite equation (\ref{eq:5})
1665: as
1666: \begin{equation}\label{eq:6}
1667: \frac {l}{H^2_0} \frac {d^2x}{dt^2}=- \frac {\Omega_m}{2x^2}(1+ \delta_i
1668: )+\Omega_\Lambda x.
1669: \end{equation}
1670: Our initial conditions for the integration of this orbit are 
1671: \begin{equation} \label{eq:7}
1672: x(t_i) = a_i
1673: \end{equation}
1674: \begin{equation}\label{eq:8}
1675: \frac {dx}{dt}(t_i) = H(t_1)x \left( 1- \frac {\delta_i}{3} \right) =H_0 a_i \left( 1- \frac 
1676: {\delta_i}{3} 
1677: \right) 
1678: \sqrt { \frac {\Omega_m} {a^3_i} + \frac {\Omega_k} {a^2_i} + \Omega_\Lambda },
1679: \end{equation}
1680: where $H(t_1)=H_0[\Omega_m/a^3(t_1)+(1-\Omega_m)]^{1/2}$ is the Hubble
1681: parameter for a flat Universe at a a cosmic time $t_1$.  Integrating
1682: equation (\ref{eq:6}) yields
1683: \begin{equation}\label{eq:9}
1684: \frac {1}{H^2_0} {\left( \frac {dx}{dt} \right)}^2 =\frac {\Omega_m}{x}(1+ \delta_i )+\Omega_\Lambda x^2 
1685: + 
1686: K,
1687: \end{equation}
1688: where $K$ is a constant of integration.  Evaluating this at the initial
1689: time and dropping terms of $O(a_i)$ (but $\delta_i \sim a_i$, so we keep
1690: ratios of order unity), we find
1691: \begin{equation}\label{eq:10}
1692: K = - \frac {5 \delta_i }{3a_i} \Omega_m + \Omega_k.	
1693: \end{equation}
1694: If $K$ is sufficiently negative, the particle will turn-around and the
1695: sphere will collapse at a time
1696: \begin{equation}\label{eq:11}
1697: H_0 t_{coll} = 2 \int_0^{a_{\rm max}} da \left( \Omega_m/a + K + \Omega_\Lambda a^2 \right)^{-1/2},
1698: \end{equation}
1699: where $a_{\rm max}$ is the value of $a$ which sets the denominator of the
1700: integral to zero.
1701: 
1702: For the case of $\Lambda$ = 0, we can determine the spherical collapse
1703: parameters $A$ and $B.$ $K > 0 \: (K < 0)$ produces an open (closed) model.
1704: Comparing coefficients in the energy equations [eq. (\ref{eq:9}) and the
1705: integration of (\ref{eq:1a})], one finds
1706: %1-49
1707: \begin{equation}\label{eq:12}
1708: A = \frac {\Omega_mr_i}{2a_i} \left| \frac {5 \delta_i }{3a_i} \Omega_m - \Omega_k\right|^{-1}	
1709: \end{equation}
1710: \begin{equation}\label{eq:13} 
1711: B = \frac {\Omega_m}{2H_0} \left| \frac {5 \delta_i }{3a_i} \Omega_m - \Omega_k \right|^{-3/2},		
1712: \end{equation}
1713: where $\Omega_k = 1 - \Omega_m$.  In particular, in an $\Omega = 1$
1714: Universe, where $1 + z = (3H_0t/2)^{-2/3}$, we find that a shell collapses
1715: at redshift $1 + z_c = 0.5929\delta_i/a_i$, or in other words a shell
1716: collapsing at redshift $z_c$ had a linear overdensity extrapolated to the
1717: present day of $\delta_0 = 1.686(1 + z_c).$
1718: 
1719: While this derivation has been for spheres of constant density, we may treat 
1720: a general spherical density profile $\delta_i(r)$ up until shell crossing \cite{Gu72}.  
1721: A particular radial shell evolves according to the mass interior to it; therefore, we
1722: define the average overdensity $\overline{\delta_i}$
1723: \begin{equation}\label{eq:14}
1724: \overline{\delta_i}(R) = \frac {3}{4 \pi R^3} \int_0^R d^3r\delta_i(r), 
1725: \end{equation}
1726: so that we may use $\overline{\delta_i}$ in place of ${\delta_i}$ in the
1727: above formulae.  If $\overline{\delta_i}$ is not monotonically decreasing
1728: with $R$, then the spherical top-hat evolution of two different radii will
1729: predict that they cross each other at some late time; this is known as
1730: shell crossing and signals the breakdown of the solution.  Even
1731: well-behaved $\overline{\delta_i}$ profiles will produce shell crossing if
1732: shells are allowed to collapse to $r = 0$ and then reexpand, since these
1733: expanding shells will cross infalling shells.  In such a case, first-time
1734: infalling shells will never be affected prior to their turn-around; the
1735: more complicated behavior after turn-around is a manifestation of
1736: virialization.  While the end state for general initial conditions cannot
1737: be predicted, various results are known for a self-similar collapse, in which
1738: $\delta (r)$ is a power-law \cite{Fi84,Be85}, as well as for the case of 
1739: secondary infall models \cite{Go75,Gu77,Ho85}.
1740: 
1741: \subsection{Halo Properties}
1742: 
1743: The small density fluctuations evidenced in the CMB grow over time as
1744: described in the previous subsection, until the perturbation $\delta$
1745: becomes of order unity, and the full non-linear gravitational problem
1746: must be considered. The dynamical collapse of a dark matter halo can
1747: be solved analytically only in cases of particular symmetry. If we
1748: consider a region which is much smaller than the horizon $cH^{-1}$,
1749: then the formation of a halo can be formulated as a problem in
1750: Newtonian gravity, in some cases with minor corrections coming from
1751: General Relativity. The simplest case is that of spherical symmetry,
1752: with an initial ($t=t_i\ll t_0$) top-hat of uniform overdensity
1753: $\delta_i$ inside a sphere of radius $R$. Although this model is
1754: restricted in its direct applicability, the results of spherical
1755: collapse have turned out to be surprisingly useful in understanding
1756: the properties and distribution of halos in models based on cold dark
1757: matter.
1758: 
1759: The collapse of a spherical top-hat perturbation is described by the
1760: Newtonian equation (with a correction for the cosmological constant) \beq
1761: \frac{d^2r}{dt^2}=H_0^2 \Oml\, r-\frac{GM}{r^2}\ , \eeq where $r$ is the
1762: radius in a fixed (not comoving) coordinate frame, $H_0$ is the present-day
1763: Hubble constant, $M$ is the total mass enclosed within radius $r$, and the
1764: initial velocity field is given by the Hubble flow $dr/dt=H(t) r$. The
1765: enclosed $\delta$ grows initially as $\delta_L=\delta_i D(t)/D(t_i)$, in
1766: accordance with linear theory, but eventually $\delta$ grows above
1767: $\delta_L$. If the mass shell at radius $r$ is bound (i.e., if its total
1768: Newtonian energy is negative) then it reaches a radius of maximum expansion
1769: and subsequently collapses. As demonstrated in the previous section, at the
1770: moment when the top-hat collapses to a point, the overdensity predicted by
1771: linear theory is $\delta_L\, = 1.686$ in the Einstein-de Sitter model, with
1772: only a weak dependence on $\Omm$ and $\Oml$. Thus a tophat collapses at
1773: redshift $z$ if its linear overdensity extrapolated to the present day
1774: (also termed the critical density of collapse) is \beq \delta_{\rm
1775: crit}(z)=\frac{1.686}{D(z)}\ ,
1776: \label{deltac} \eeq where we set $D(z=0)=1$.
1777: 
1778: Even a slight violation of the exact symmetry of the initial perturbation
1779: can prevent the tophat from collapsing to a point. Instead, the halo
1780: reaches a state of virial equilibrium by violent relaxation (phase
1781: mixing). Using the virial theorem $U=-2K$ to relate the potential energy
1782: $U$ to the kinetic energy $K$ in the final state (implying that the virial
1783: radius is half the turnaround radius - where the kinetic energy vanishes),
1784: the final overdensity relative to the critical density at the collapse
1785: redshift is $\Delta_c=18\pi^2 \simeq 178$ in the Einstein-de Sitter model,
1786: modified in a Universe with $\Omm+\Oml=1$ to the fitting formula (Bryan \&
1787: Norman 1998 \cite{BM98}) \beq \Delta_c=18\pi^2+82 d-39 d^2\ , \eeq where $d\equiv
1788: \Ommz-1$ is evaluated at the collapse redshift, so that \beq
1789: \Ommz=\frac{\Omm (1+z)^3}{\Omm (1+z)^3+\Oml+\Omk (1+z)^2}\ .
1790: \label{Ommz} \eeq
1791: 
1792: A halo of mass $M$ collapsing at redshift $z$ thus has a virial radius
1793: \beq r_{\rm vir}=0.784 \left(\frac{M}{10^8\ h^{-1} \ M_{\sun}
1794: }\right)^{1/3} \left[\frac{\Omm} {\Ommz}\ \frac{\Delta_c}
1795: {18\pi^2}\right]^{-1/3} \left (\frac{1+z}{10} \right)^{-1}\ h^{-1}\
1796: {\rm kpc}\ , \label{rvir}\eeq and a corresponding circular velocity,
1797: \beq V_c=\left(\frac{G M}{r_{\rm vir}}\right)^{1/2}= 23.4 \left(
1798: \frac{M}{10^8\ h^{-1} \ M_{\sun} }\right)^{1/3} \left[\frac {\Omm}
1799: {\Ommz}\ \frac{\Delta_c} {18\pi^2}\right]^{1/6} \left( \frac{1+z} {10}
1800: \right)^{1/2}\ {\rm km\ s}^{-1}\ . \label{Vceqn} \eeq In these
1801: expressions we have assumed a present Hubble constant written in the
1802: form $H_0=100\, h\mbox{ km s}^{-1}\mbox{Mpc}^{-1}$. We may also define
1803: a virial temperature \beq \label{tvir} T_{\rm vir}=\frac{\mu m_p
1804: V_c^2}{2 \kB}=1.98\times 10^4\ \left(\frac{\mu}{0.6}\right)
1805: \left(\frac{M}{10^8\ h^{-1} \ M_{\sun} }\right)^{2/3} \left[ \frac
1806: {\Omm} {\Ommz}\ \frac{\Delta_c} {18\pi^2}\right]^{1/3}
1807: \left(\frac{1+z}{10}\right)\ {\rm K} \ , \eeq where $\mu$ is the mean
1808: molecular weight and $m_p$ is the proton mass. Note that the value of
1809: $\mu$ depends on the ionization fraction of the gas; for a fully
1810: ionized primordial gas $\mu=0.59$, while a gas with ionized hydrogen
1811: but only singly-ionized helium has $\mu=0.61$. The binding energy of
1812: the halo is approximately\footnote{The coefficient of $1/2$ in
1813: equation~(\ref{Ebind}) would be exact for a singular isothermal
1814: sphere, $\rho(r)\propto 1/r^2$.} \beq \label{Ebind} E_b= {1\over 2}
1815: \frac{GM^2}{r_{\rm vir}} = 5.45\times 10^{53} \left(\frac{M}{10^8\
1816: h^{-1} \ M_{\sun} }\right)^{5/3} \left[ \frac {\Omm} {\Ommz}\
1817: \frac{\Delta_c} {18\pi^2}\right]^{1/3} \left(\frac{1+z}{10}\right)
1818: h^{-1}\ {\rm erg}\ . \eeq Note that the binding energy of the baryons
1819: is smaller by a factor equal to the baryon fraction $\Omega_b/\Omm$.
1820: 
1821: Although spherical collapse captures some of the physics governing the
1822: formation of halos, structure formation in cold dark matter models procedes
1823: hierarchically. At early times, most of the dark matter is in low-mass
1824: halos, and these halos continuously accrete and merge to form high-mass
1825: halos. Numerical simulations of hierarchical halo formation indicate a
1826: roughly universal spherically-averaged density profile for the resulting
1827: halos (Navarro, Frenk, \& White 1997, hereafter NFW \cite{Na97}), though
1828: with considerable scatter among different halos (e.g., \cite{Bu00}). The NFW profile has the form \beq \rho(r)=\frac{3 H_0^2} {8 \pi G}
1829: (1+z)^3 \frac{\Omm}{\Ommz} \frac{\delta_c} {\cN x (1+\cN x)^2}\ ,
1830: \label{NFW} \eeq where $x=r/r_{\rm vir}$, and the characteristic density
1831: $\delta_c$ is related to the concentration parameter $\cN$ by \beq
1832: \delta_c=\frac{\Delta_c}{3} \frac{\cN^3} {\ln(1+\cN)-\cN/(1+\cN)} \ . \eeq
1833: The concentration parameter itself depends on the halo mass $M$, at a given
1834: redshift $z$ \cite{WBPKD02}.
1835: 
1836: More recent N-body simulations indicate deviations from the original NFW
1837: profile; for details and refined fitting formula see \cite{nav04}.
1838: 
1839: 
1840: %\subsection{Universal Mass Profile of Dark Matter Halos}
1841: 
1842: 
1843: %<>
1844: \section{Nonlinear Growth}
1845: 
1846: %\subsection{The Press-Schechter Formalism for the Abundance of Dark Matter 
1847: %Halos}
1848: 
1849: 
1850: \subsection{The Abundance of Dark Matter Halos}
1851: \label{sec2.4}
1852: 
1853: In addition to characterizing the properties of individual halos, a
1854: critical prediction of any theory of structure formation is the
1855: abundance of halos, i.e.\ the number density of halos as a function of
1856: mass, at any redshift. This prediction is an important step toward
1857: inferring the abundances of galaxies and galaxy clusters. While the
1858: number density of halos can be measured for particular cosmologies in
1859: numerical simulations, an analytic model helps us gain physical
1860: understanding and can be used to explore the dependence of abundances
1861: on all the cosmological parameters. 
1862: 
1863: A simple analytic model which successfully matches most of the numerical
1864: simulations was developed by Press \& Schechter (1974) \cite{Press}. The model is based
1865: on the ideas of a Gaussian random field of density perturbations, linear
1866: gravitational growth, and spherical collapse. To determine the abundance of
1867: halos at a redshift $z$, we use $\delta_M$, the density field smoothed on a
1868: mass scale $M$, as defined in \S \ref{sec2.2}. Since $\delta_M$ is
1869: distributed as a Gaussian variable with zero mean and standard deviation
1870: $\sigma(M)$ [which depends only on the present linear power spectrum, see
1871: equation~(\ref{eqsigM})], the probability that $\delta_M$ is greater than
1872: some $\delta$ equals \beq \int_{\delta}^{\infty}d\delta_M \frac{1}{\sqrt{2
1873: \pi}\, \sigma(M)} \exp \left[- \frac{\delta_M^2} {2
1874: \,\sigma^2(M)}\right]={1\over 2} {\rm erfc}\left(\frac{\delta} {\sqrt{2}
1875: \,\sigma(M) } \right)\ . \label{PS1} \eeq The fundamental ansatz is to
1876: identify this probability with the fraction of dark matter particles which
1877: are part of collapsed halos of mass greater than $M$, at redshift
1878: $z$. There are two additional ingredients: First, the value used for
1879: $\delta$ is $\delta_{\rm crit}(z)$ given in equation~(\ref{deltac}), which
1880: is the critical density of collapse found for a spherical top-hat
1881: (extrapolated to the present since $\sigma(M)$ is calculated using the
1882: present power spectrum); and second, the fraction of dark matter in halos
1883: above $M$ is multiplied by an additional factor of 2 in order to ensure
1884: that every particle ends up as part of some halo with $M>0$. Thus, the
1885: final formula for the mass fraction in halos above $M$ at redshift $z$ is
1886: \beq
1887: \label{PSerfc} F(>M | z)={\rm erfc}\left(\frac{\delta_{\rm crit}(z)} 
1888: {\sqrt{2}\,\sigma(M) } \right)\ . \eeq
1889: 
1890: This ad-hoc factor of 2 is necessary, since otherwise only positive
1891: fluctuations of $\delta_M$ would be included. Bond et al.\ (1991) \cite{bond91}
1892: found an alternate derivation of this correction factor, using a
1893: different ansatz. In their derivation, the factor of 2 has a more
1894: satisfactory origin, namely the so-called ``cloud-in-cloud'' problem:
1895: For a given mass $M$, even if $\delta_M$ is smaller than $\delta_{\rm
1896: crit}(z)$, it is possible that the corresponding region lies inside a
1897: region of some larger mass $M_L>M$, with $\delta_{M_L}>\delta_{\rm
1898: crit}(z)$. In this case the original region should be counted as
1899: belonging to a halo of mass $M_L$. Thus, the fraction of particles
1900: which are part of collapsed halos of mass greater than $M$ is larger
1901: than the expression given in equation~(\ref{PS1}). Bond et al.\ showed
1902: that, under certain assumptions, the additional contribution results
1903: precisely in a factor of 2 correction.
1904: 
1905: \begin{figure}
1906: \centering
1907: \includegraphics[height=6cm]{col.ps}
1908: \caption{Fraction of baryons that assembled into dark matter halos with a
1909: virial temperature of $T_{\rm vir}>10^4$K as a function of redshift. These
1910: baryons are above the temperature threshold for gas cooling and
1911: fragmentation via atomic transitions.  After reionization the temperature
1912: barrier for star formation in galaxies is raised because the photo-ionized
1913: intergalactic medium is already heated to $\sim 10^{4}$K and it can
1914: condense only into halos with $T_{\rm vir}>10^5$K.}
1915: \label{collapsed}
1916: \end{figure}
1917: 
1918: Differentiating the fraction of dark matter in halos above $M$ yields
1919: the mass distribution. Letting $dn$ be the comoving number density of
1920: halos of mass between $M$ and $M+dM$, we have \beq\frac{dn}{dM}=
1921: \sqrt{\frac{2}{\pi}}\, \frac{\rho_m}{M}\, \frac{-d(\ln \sigma)}{dM}
1922: \,\nu_c\, e^{-\nu_c^2/2}\ , \eeq where $\nu_c=\delta_{\rm crit}(z)/
1923: \sigma(M)$ is the number of standard deviations which the critical
1924: collapse overdensity represents on mass scale $M$. Thus, the abundance
1925: of halos depends on the two functions $\sigma(M)$ and $\delta_{\rm
1926: crit} (z)$, each of which depends on the energy content of the
1927: Universe and the values of the other cosmological parameters. Since
1928: recent observations confine the standard set of parameters to a
1929: relatively narrow range, we illustrate the abundance of halos and
1930: other results for a single set of parameters: $\Omm=0.3$, $\Oml=0.7$,
1931: $\Omega_b=0.045$, $\sigma_8=0.9$, a primordial power spectrum index
1932: $n=1$ and a Hubble constant $h=0.7$. 
1933: %These parameter values are based primarily on the following observational
1934: %results: CMB temperature anisotropy measurements on large scales (Bennett
1935: %et al.\ 1996) and on the scale of $\sim 1^\circ$ (Lange et al.\ 2000; Balbi
1936: %et al.\ 2000); the abundance of galaxy clusters locally (Viana \& Liddle
1937: %1999; Pen 1998; Eke, Cole, \& Frenk 1996) and as a function of redshift
1938: %(Bahcall \& Fan 1998; Eke, Cole, Frenk, \& Henry 1998); the baryon density
1939: %inferred from big bang nucleosynthesis (see the review by Tytler et al.\
1940: %2000); distance measurements used to derive the Hubble constant (Mould et
1941: %al.\ 2000; Jha et al.\ 1999; Tonry et al.\ 1997); and indications of cosmic
1942: %acceleration from distances based on type Ia supernovae (Perlmutter et al.\
1943: %1999; Riess et al.\ 1998).
1944: 
1945: Figure \ref{fig2a} shows $\sigma(M)$ and $\delta_{\rm crit}(z)$, with
1946: the input power spectrum computed from Eisenstein \& Hu (1999) \cite{Eis99}. The
1947: solid line is $\sigma(M)$ for the cold dark matter model with the
1948: parameters specified above. The horizontal dotted lines show the value
1949: of $\delta_{\rm crit}(z)$ at $z=0, 2, 5, 10, 20$ and 30, as indicated
1950: in the figure. From the intersection of these horizontal lines with
1951: the solid line we infer, e.g., that at $z=5$ a $1-\sigma$ fluctuation
1952: on a mass scale of $2\times 10^7 M_{\sun}$ will collapse. On the other
1953: hand, at $z=5$ collapsing halos require a $2-\sigma$ fluctuation on a
1954: mass scale of $3\times 10^{10} M_{\sun}$, since $\sigma(M)$ on this
1955: mass scale equals about half of $\delta_{\rm crit}(z=5)$. Since at
1956: each redshift a fixed fraction ($31.7\%$) of the total dark matter
1957: mass lies in halos above the $1-\sigma$ mass, Figure \ref{fig2a} shows
1958: that most of the mass is in small halos at high redshift, but it
1959: continuously shifts toward higher characteristic halo masses at lower
1960: redshift. Note also that $\sigma(M)$ flattens at low masses because of
1961: the changing shape of the power spectrum. Since $\sigma \rightarrow
1962: \infty$ as $M \rightarrow 0$, in the cold dark matter model all the
1963: dark matter is tied up in halos at all redshifts, if sufficiently
1964: low-mass halos are considered.
1965:   
1966: %%%%%%%% Figure 2.1
1967: \begin{figure}%
1968: \centering
1969: \includegraphics[height=6cm]{IIfig1.ps}
1970: %voffset=-500 hscale=65 vscale=65}
1971: %\vspace{6.2in}
1972: %Used IIfig1.mon
1973: \caption{Mass fluctuations and collapse thresholds in cold dark matter
1974: models. The horizontal dotted lines show the value of the extrapolated
1975: collapse overdensity $\delta_{\rm crit}(z)$ at the indicated
1976: redshifts. Also shown is the value of $\sigma(M)$ for the cosmological
1977: parameters given in the text (solid curve), as well as $\sigma(M)$ for
1978: a power spectrum with a cutoff below a mass $M=1.7\times 10^8
1979: M_{\sun}$ (short-dashed curve), or $M=1.7\times 10^{11} M_{\sun}$
1980: (long-dashed curve). The intersection of the horizontal lines with the
1981: other curves indicate, at each redshift $z$, the mass scale (for each
1982: model) at which a $1-\sigma$ fluctuation is just collapsing at $z$
1983: (see the discussion in the text).}
1984: \label{fig2a}
1985: \end{figure}
1986: %%%%%%%%
1987:   
1988: Also shown in Figure \ref{fig2a} is the effect of cutting off the
1989: power spectrum on small scales. The short-dashed curve corresponds to
1990: the case where the power spectrum is set to zero above a comoving
1991: wavenumber $k=10\, {\rm Mpc}^{-1}$, which corresponds to a mass
1992: $M=1.7\times 10^8 M_{\sun}$. The long-dashed curve corresponds to a
1993: more radical cutoff above $k=1\, {\rm Mpc}^{-1}$, or below
1994: $M=1.7\times 10^{11} M_{\sun}$. A cutoff severely reduces the
1995: abundance of low-mass halos, and the finite value of $\sigma(M=0)$
1996: implies that at all redshifts some fraction of the dark matter does
1997: not fall into halos. At high redshifts where $\delta_{\rm crit}(z) \gg
1998: \sigma(M=0)$, all halos are rare and only a small fraction of the dark
1999: matter lies in halos. In particular, this can affect the abundance of
2000: halos at the time of reionization, and thus the observed limits on
2001: reionization constrain scenarios which include a small-scale cutoff in
2002: the power spectrum \cite{BHO00}.
2003: 
2004: In figures \ref{fig2b} -- \ref{fig2e} we show explicitly the properties of
2005: collapsing halos which represent $1-\sigma$, $2-\sigma$, and $3-\sigma$
2006: fluctuations (corresponding in all cases to the curves in order from bottom
2007: to top), as a function of redshift. No cutoff is applied to the power
2008: spectrum. Figure \ref{fig2b} shows the halo mass, Figure \ref{fig2c} the
2009: virial radius, Figure \ref{fig2d} the virial temperature (with $\mu$ in
2010: equation~(\ref{tvir}) set equal to $0.6$, although low temperature halos
2011: contain neutral gas) as well as circular velocity, and Figure \ref{fig2e}
2012: shows the total binding energy of these halos. In figures \ref{fig2b} and
2013: \ref{fig2d}, the dotted curves indicate the minimum virial temperature
2014: required for efficient cooling with primordial atomic species only (upper
2015: curve) or with the addition of molecular hydrogen (lower curve). Figure
2016: \ref{fig2e} shows the binding energy of dark matter halos. The binding
2017: energy of the baryons is a factor $\sim \Omega_b/\Omega_m\sim 15\%$
2018: smaller, if they follow the dark matter. Except for this constant factor,
2019: the figure shows the minimum amount of energy that needs to be deposited
2020: into the gas in order to unbind it from the potential well of the dark
2021: matter. For example, the hydrodynamic energy released by a single
2022: supernovae, $\sim 10^{51}~{\rm erg}$, is sufficient to unbind the gas in
2023: all $1-\sigma$ halos at $z\ga 5$ and in all $2-\sigma$ halos at $z\ga 12$.
2024: 
2025: %%%%%%%% Figure 2b
2026: \begin{figure}
2027: \centering
2028: \includegraphics[height=6cm]{IIfig2.ps}
2029: %Used IIfig2.mon
2030: \caption{Characteristic properties of collapsing halos: Halo mass.
2031: The solid curves show the mass of collapsing halos which correspond to
2032: $1-\sigma$, $2-\sigma$, and $3-\sigma$ fluctuations (in order from
2033: bottom to top). The dotted curves show the mass corresponding to the
2034: minimum temperature required for efficient cooling with primordial
2035: atomic species only (upper curve) or with the addition of molecular
2036: hydrogen (lower curve).}
2037: \label{fig2b}
2038: \end{figure}
2039: %%%%%%%%
2040: 
2041: %%%%%%%% Figure 2c
2042: \begin{figure}
2043: \centering
2044: \includegraphics[height=6cm]{IIfig3.ps}
2045: %Used IIfig3.mon
2046: \caption{Characteristic properties of collapsing halos: Halo virial
2047: radius. The curves show the virial radius of collapsing halos which
2048: correspond to $1-\sigma$, $2-\sigma$, and $3-\sigma$ fluctuations (in
2049: order from bottom to top).}
2050: \label{fig2c}
2051: \end{figure}
2052: %%%%%%%%
2053:    
2054: %%%%%%%% Figure 2d
2055: \begin{figure}
2056: \centering
2057: \includegraphics[height=6cm]{IIfig4.ps}
2058: %Used IIfig4.mon
2059: \caption{Characteristic properties of collapsing halos: Halo virial
2060: temperature and circular velocity. The solid curves show the virial
2061: temperature (or, equivalently, the circular velocity) of collapsing
2062: halos which correspond to $1-\sigma$, $2-\sigma$, and $3-\sigma$
2063: fluctuations (in order from bottom to top). The dotted curves show the
2064: minimum temperature required for efficient cooling with primordial
2065: atomic species only (upper curve) or with the addition of molecular
2066: hydrogen (lower curve).}
2067: \label{fig2d}
2068: \end{figure}
2069: %%%%%%%%
2070: 
2071: %%%%%%%% Figure 2e
2072: \begin{figure}
2073: \centering
2074: \includegraphics[height=6cm]{IIfig5.ps}
2075: %Used IIfig5.mon
2076: \caption{Characteristic properties of collapsing halos: Halo binding
2077: energy. The curves show the total binding energy of collapsing halos
2078: which correspond to $1-\sigma$, $2-\sigma$, and $3-\sigma$
2079: fluctuations (in order from bottom to top).}
2080: \label{fig2e}
2081: \end{figure}
2082: %%%%%%%%
2083:  
2084: At $z=5$, the halo masses which correspond to $1-\sigma$, $2-\sigma$,
2085: and $3-\sigma$ fluctuations are $1.8\times 10^7 M_{\sun}$, $3.0\times
2086: 10^{10} M_{\sun}$, and $7.0\times 10^{11} M_{\sun}$, respectively. The
2087: corresponding virial temperatures are $2.0 \times 10^3$ K, $2.8 \times
2088: 10^5$ K, and $2.3 \times 10^6$ K. The equivalent circular velocities
2089: are 7.5 ${\rm km\ s}^{-1}$, 88 ${\rm km\ s}^{-1}$, and 250 ${\rm km\
2090: s}^{-1}$. At $z=10$, the $1-\sigma$, $2-\sigma$, and $3-\sigma$
2091: fluctuations correspond to halo masses of $1.3\times 10^3 M_{\sun}$,
2092: $5.7\times 10^7 M_{\sun}$, and $4.8\times 10^9 M_{\sun}$,
2093: respectively. The corresponding virial temperatures are 6.2 K, $7.9
2094: \times 10^3$ K, and $1.5 \times 10^5$ K. The equivalent circular
2095: velocities are 0.41 ${\rm km\ s}^{-1}$, 15 ${\rm km\ s}^{-1}$, and 65
2096: ${\rm km\ s}^{-1}$. Atomic cooling is efficient at $T_{\rm vir} \ga
2097: 10^4$ K, or a circular velocity $V_c \ga 17\ {\rm km\ s}^{-1}$. This
2098: corresponds to a $1.2-\sigma$ fluctuation and a halo mass of
2099: $2.1\times 10^8 M_{\sun}$ at $z=5$, and a $2.1-\sigma$ fluctuation and
2100: a halo mass of $8.3\times 10^7 M_{\sun}$ at $z=10$. Molecular hydrogen
2101: provides efficient cooling down to $T_{\rm vir} \sim 300$ K, or a
2102: circular velocity $V_c \sim 2.9\ {\rm km\ s}^{-1}$. This corresponds
2103: to a $0.81-\sigma$ fluctuation and a halo mass of $1.1\times 10^6
2104: M_{\sun}$ at $z=5$, and a $1.4-\sigma$ fluctuation and a halo mass of
2105: $4.3\times 10^5 M_{\sun}$ at $z=10$.
2106: 
2107: In Figure  \ref{fig2f} we show the halo mass function $dn/d\ln(M)$ at
2108: several different redshifts: $z=0$ (solid curve), $z=5$ (dotted
2109: curve), $z=10$ (short-dashed curve), $z=20$ (long-dashed curve), and
2110: $z=30$ (dot-dashed curve). Note that the mass function does not
2111: decrease monotonically with redshift at all masses. At the lowest
2112: masses, the abundance of halos is higher at $z>0$ than at $z=0$.
2113: 
2114: %%%%%%%% Figure 2f
2115: \begin{figure}
2116: \centering
2117: \includegraphics[height=6cm]{IIfig6.ps}
2118: %Used IIfig6.mon
2119: \caption{Halo mass function at several redshifts: $z=0$ (solid curve),
2120: $z=5$ (dotted curve), $z=10$ (short-dashed curve), $z=20$ (long-dashed
2121: curve), and $z=30$ (dot-dashed curve).}
2122: \label{fig2f}
2123: \end{figure}
2124: %%%%%%%%
2125: 
2126: \subsection {The Excursion-Set (Extended Press-Schechter) Formalism}
2127: 
2128: The usual Press-Schechter formalism makes no attempt to deal with the
2129: correlations between halos or between different mass scales.  In
2130: particular, this means that while it can generate a distribution of halos
2131: at two different epochs, it says nothing about how particular halos in one
2132: epoch are related to those in the second.  We therefore would like some
2133: method to predict, at least statistically, the growth of individual halos
2134: via accretion and mergers.  Even restricting ourselves to spherical
2135: collapse, such a model must utilize the full spherically-averaged density
2136: profile around a particular point.  The potential correlations between the
2137: mean overdensities at different radii make the statistical description
2138: substantially more difficult.
2139: 
2140: The excursion set formalism (Bond et al. 1991 \cite{bond91}) seeks to
2141: describe the statistics of halos by considering the statistical properties
2142: of $\overline{\delta}(R)$, the average overdensity within some spherical
2143: window of characteristic radius $R$, as a function of $R$.  While the
2144: Press-Schechter model depends only on the Gaussian distribution of
2145: $\overline{\delta}$ for one particular $R$, the excursion set considers all
2146: $R$.  Again the connection between a value of the linear regime $\delta$
2147: and the final state is made via the spherical collapse solution, so that
2148: there is a critical value $\delta_c (z)$ of $\overline{\delta}$ which is
2149: required for collapse at a redshift $z$.
2150: 
2151: For most choices of window function, the functions $\overline{\delta}(R)$
2152: are correlated from one $R$ to another such that it is prohibitively
2153: difficult to calculate the desired statistics directly [although Monte
2154: Carlo realizations are possible \cite{bond91}].
2155: However, for one particular choice of a window function, the correlations
2156: between different $R$ greatly simplify and many interesting quantities may
2157: be calculated \cite{bond91,LC93}.  The key is to use a $k$-space top-hat
2158: window function, namely $W_k = 1$ for all $k$ less than some critical $k_c$
2159: and $W_k = 0$ for $k > k_c$.  This filter has a spatial form of $W(r) \:
2160: \propto \: j_1 (k_c r) / k_c r$, which implies a volume $6 \pi^2 /
2161: k^{3}_{c}$ or mass $6\pi^2 \rho_b /k^{3}_{c}$.  The characteristic radius
2162: of the filter is $\sim k^{-1}_{c}$, as expected.  Note that in real space,
2163: this window function converges very slowly, due only to a sinusoidal
2164: oscillation, so the region under study is rather poorly localized.
2165: 
2166: The great advantage of the sharp $k$-space filter is that the difference at
2167: a given point between $\overline{\delta}$ on one mass scale and that on
2168: another mass scale is statistically independent from the value on the
2169: larger mass scale.  With a Gaussian random field, each
2170: $\delta_k$ is Gaussian distributed independently from the others.  For this
2171: filter,
2172: \begin{equation}\label{eq:15}
2173: \overline{\delta}(M) = \int_{k<k_c(M)} \frac {d^3k}{(2 \pi )^3} \delta_k, 
2174: \end{equation}
2175: meaning that the overdensity on a particular scale is simply the sum of the
2176: random variables $\delta_k$ interior to the chosen $k_c$.  Consequently,
2177: the difference between the $\overline{\delta}(M)$ on two mass scales is
2178: just the sum of the $\delta_k$ in the spherical shell between the two
2179: $k_c$, which is independent from the sum of the $\delta_k$ interior to the
2180: smaller $k_c$.  Meanwhile, the distribution of $\overline{\delta}(M)$ given
2181: no prior information is still a Gaussian of mean zero and variance
2182: \begin{equation}\label{eq:16}
2183: \sigma^2(M) = \frac {1}{2 \pi^2} \int_{k<k_c(M)} dk \: k^2P(k). 
2184: \end{equation}
2185: 
2186: If we now consider $\overline{\delta}$ as a function of scale $k_c$, we see
2187: that we begin from $\overline{\delta} = 0$ at $k_c = 0 \: (M = \infty )$
2188: and then add independently random pieces as $k_c$ increases.  This
2189: generates a random walk, albeit one whose stepsize varies with $k_c$.  We
2190: then assume that, at redshift $z$, a given function $\overline{\delta}(k_c
2191: )$ represents a collapsed mass $M$ corresponding to the $k_c$ where the
2192: function first crosses the critical value $\delta_c (z)$.  With this
2193: assumption, we may use the properties of random walks to calculate the
2194: evolution of the mass as a function of redshift.
2195: 
2196: It is now easy to rederive the Press-Schechter mass function, including the
2197: previously unexplained factor of 2 \cite{bond91,LC93,Whi94}.  The fraction
2198: of mass elements included in halos of mass less than $M$ is just the
2199: probability that a random walk remains below $\delta_c (z)$ for all $k_c$
2200: less than $K_c$, the filter cutoff appropriate to $M$.  This probability
2201: must be the complement of the sum of the probabilities that: {\it (a)}
2202: $\overline{\delta} (K_c) > \delta_c(z)$; or that {\it (b)}
2203: $\overline{\delta}(K_c) < \delta_c(z)$ but $\overline{\delta}(k'_c) >
2204: \delta_c(z)$ for some $k'_c < K_c$.  But these two cases in fact have equal
2205: probability; any random walk belonging to class {\it (a)} may be reflected
2206: around its first upcrossing of $\delta_c (z)$ to produce a walk of class
2207: {\it (b)}, and vice versa.  Since the distribution of
2208: $\overline{\delta}(K_c)$ is simply Gaussian with variance $\sigma^2(M)$,
2209: the fraction of random walks falling into class {\it (a)} is simply $(1/
2210: \sqrt{2\pi\sigma^2}) \int_{\delta_c (z)}^{\infty} d \delta \; \exp \{ -
2211: \delta^2 / 2 \sigma^2 (M) \}$.  Hence, the fraction of mass elements
2212: included in halos of mass less than $M$ at redshift $z$ is simply
2213: %1-58
2214: \begin{equation}\label{eq:17} 
2215: F(<M)=1-2 \times \frac {1} { \sqrt{2\pi\sigma^2}} \int_{\delta_c
2216: (z)}^{\infty} d \delta \: \exp \{ - \delta^2 /2 \sigma^2 (M) \}
2217: \end{equation} 
2218: which may be differentiated to yield the Press-Schechter mass function. We
2219: may now go further and consider how halos at one redshift are related to
2220: those at another redshift.  If we are given that a halo of mass $M_2$
2221: exists at redshift $z_2$, then we know that the random function
2222: $\overline{\delta}(k_c)$ for each mass element within the halo first
2223: crosses $\delta(z_2)$ at $k_{c2}$ corresponding to $M_2$.  Given this
2224: constraint, we may study the distribution of $k_c$ where the function
2225: $\overline{\delta}(k_c)$ crosses other thresholds.  It is particularly easy
2226: to construct the probability distribution for when trajectories first cross
2227: some $\delta_c (z_1) \: > \: \delta_c (z_2)$ (implying $z_1 \: > \: z_2$);
2228: clearly this occurs at some $k_{c1} \: > k_{c2}$.  This problem reduces to
2229: the previous one if we translate the origin of the random walks from
2230: $(k_c,\overline{\delta}) = (0,0) \: {\rm to} \: (k_{c2},\delta_c(z_2))$.
2231: We therefore find the distribution of halo masses $M_1$ that a mass element
2232: finds itself in at redshift $z_1$ given that it is part of a larger halo of
2233: mass $M_2$ at a later redshift $z_2$ is \cite{bond91,Bow91})
2234: %1-59
2235: \begin{equation}\label{eq:18}\begin{split} 
2236: & \frac { dP } { dM_1 } (M_1,z_1 | M_2,z_2) = \\ & \sqrt {\frac {2}{\pi}}
2237: \frac {\delta_c (z_1) - \delta_c (z_2)} { [ \sigma^2 (M_1) - \sigma^2
2238: (M_2)]^{3/2} } \left | \frac { d \sigma (M_1) } { d M_1} \right | \exp \left
2239: \{ - \frac { [ \delta_c (z_1) - \delta_c (z_2) ] ^2 } { 2 [ \sigma^2 (M_1)
2240: - \sigma^2 (M_2) ] } \right \}.
2241: \end{split}\end{equation} 
2242: This may be rewritten as saying that the quantity 
2243: \begin{equation}\label{eq:19}
2244: \tilde {v} = \frac {\delta_c (z_1) - \delta_c (z_2)} { \sqrt {\sigma^2
2245: (M_1) - \sigma^2 (M_2)} }
2246: \end{equation} 
2247: is distributed as the positive half of a Gaussian with unit variance;
2248: equation (\ref{eq:19}) may be inverted to find $M_1(\tilde {v})$.
2249: 
2250: We seek to interpret the statistics of these random walks as those of
2251: merging and accreting halos.  For a single halo, we may imagine that as we
2252: look back in time, the object breaks into ever smaller pieces, similar to
2253: the branching of a tree.  Equation (\ref{eq:18}) is the distribution of the
2254: sizes of these branches at some given earlier time.  However, using this
2255: description of the ensemble distribution to generate Monte Carlo
2256: realizations of single merger trees has proven to be difficult.  In all
2257: cases, one recursively steps back in time, at each step breaking the final
2258: object into two or more pieces.  An elaborate scheme (Kauffmann \& White
2259: 1993 \cite{KW93}) picks a large number of progenitors from the ensemble
2260: distribution and then randomly groups them into sets with the correct total
2261: mass.  This generates many (hundreds) possible branching schemes of equal
2262: likelihood.  A simpler scheme (Lacey \& Cole 1993 \cite{LC93}) assumes that
2263: at each time step, the object breaks into two pieces.  One value from the
2264: distribution (\ref{eq:18}) then determines the mass ratio of the two
2265: branchs.
2266: 
2267: One may also use the distribution of the ensemble to derive some additional
2268: analytic results.  A useful example is the distribution of the epoch at
2269: which an object that has mass $M_2$ at redshift $z_2$ has accumulated half
2270: of its mass \cite{LC93}.  The probability that the formation time is
2271: earlier than $z_1$ is equal to the probability that at redshift $z_1$ a
2272: progenitor whose mass exceeds $M_2/2$ exists:
2273: %1-61
2274: \begin{equation}\label{eq:20}  
2275: P(z_f > z_1) = \int_{M_2/2}^{M_2} \frac{M_2}{M} \frac {dP}{dM}
2276: (M,z_1|M_2,z_2)dM,
2277: \end{equation} 
2278: where $dP/dM$ is given in equation (\ref{eq:18}).  The factor of $M_2/M$
2279: corrects the counting from mass weighted to number weighted; each halo of
2280: mass $M_2$ can have only one progenitor of mass greater than $M_2/2$.
2281: Differentiating equation (\ref{eq:20}) with respect to time gives the
2282: distribution of formation times.  This analytic form is an excellent match
2283: to scale-free N-body simulations \cite{LC94}.  On the
2284: other hand, simple Monte Carlo implementations of equation (\ref{eq:18})
2285: produce formation redshifts about 40\% higher \cite{LC93}.  As there may be
2286: correlations between the various branches, there is no unique Monte Carlo
2287: scheme.
2288: 
2289: Numerical tests of the excursion set formalism are quite encouraging.  Its
2290: predictions for merger rates are in very good agreement with those measured
2291: in scale-free N-body simulations for mass scales down to around 10\% of the
2292: nonlinear mass scale (that scale at which $\sigma_M= 1$ ), and
2293: distributions of formation times closely match the analytic predictions
2294: \cite{LC94}.  The model appears to be a promising method for tracking the
2295: merging of halos, with many applications to cluster and galaxy formation
2296: modeling.  In particular, one may use the formalism as the foundation of
2297: semi-analytic galaxy formation models \cite{KWG93}.  The excursion set
2298: formalism may also be used to derive the correlations of halos in the
2299: nonlinear regime \cite{mw96}.
2300: 
2301: \subsection{Response of Baryons to Nonlinear Dark Matter Potentials}
2302: 
2303: The dark matter is assumed to be cold and to dominate gravity, and so
2304: its collapse and virialization proceeds unimpeded by pressure
2305: effects. In order to estimate the minimum mass of baryonic objects, we
2306: must go beyond linear perturbation theory and examine the baryonic
2307: mass that can accrete into the final gravitational potential well
2308: of the dark matter.
2309: 
2310: For this purpose, we assume that the dark matter had already
2311: virialized and produced a gravitational potential $\phi({\bf r})$ at a
2312: redshift $z_{\rm vir}$ (with $\phi\rightarrow0$ at large distances,
2313: and $\phi<0$ inside the object) and calculate the resulting
2314: overdensity in the gas distribution, ignoring cooling (an assumption
2315: justified by spherical collapse simulations which indicate that
2316: cooling becomes important only after virialization; see Haiman et al.\
2317: 1996 \cite{Haiman}).
2318: 
2319: After the gas settles into the dark matter potential well, it
2320: satisfies the hydrostatic equilibrium equation,
2321: \begin{equation}
2322: \nabla p_{\rm b} = -\rho_{\rm b} \nabla \phi
2323: \label{eq:hyd}
2324: \end{equation}
2325: where $p_{\rm b}$ and $\rho_{\rm b}$ are the pressure and mass density
2326: of the gas.  At $z \la 100$ the gas temperature is decoupled from the
2327: CMB, and its pressure evolves adiabatically (ignoring atomic or
2328: molecular cooling),
2329: \begin{equation}
2330: {p_{\rm b}\over {\bar p}_{\rm b}} = \left({\rho_{\rm b}\over {\bar
2331: \rho}_{\rm b}}\right)^{5/3}
2332: \label{eq:adi}
2333: \end{equation}
2334: where a bar denotes the background conditions. We substitute
2335: equation~(\ref{eq:adi}) into~(\ref{eq:hyd}) and get the solution,
2336: \begin{equation}
2337: {\rho_{\rm b}\over {\bar \rho}_{\rm b}}= \left(1- {2\over 5}{\mu
2338: m_p\phi\over {\kB \bar T}}\right)^{3/2}
2339: \end{equation}
2340: where ${\bar T}= {\bar p}_{\rm b} \mu m_p/(\kB {\bar \rho}_{\rm b})$
2341: is the background gas temperature.  If we define $T_{\rm vir}=
2342: -{1\over 3}m_p\phi/\kB$ as the virial temperature for a potential
2343: depth $-\phi$, then the overdensity of the baryons at the
2344: virialization redshift is
2345: \begin{equation}
2346: \delta_{\rm b} = {\rho_{\rm b}\over {\bar \rho}_{\rm b}} - 1 =
2347: \left(1+
2348: {6\over 5}{T_{\rm vir}\over {\bar T}}\right)^{3/2} - 1 .
2349: \label{eq:del}
2350: \end{equation}
2351: This solution is approximate for two reasons: (i) we assumed that the
2352: gas is stationary throughout the entire region and ignored the
2353: transitions to infall and the Hubble expansion at the interface
2354: between the collapsed object and the background intergalactic medium
2355: (henceforth IGM), and (ii) we ignored entropy production at the
2356: virialization shock surrounding the object.  Nevertheless, the result
2357: should provide a better estimate for the minimum mass of collapsed
2358: baryonic objects than the Jeans mass does, since it incorporates the
2359: nonlinear potential of the dark matter.
2360: 
2361: We may define the threshold for the collapse of baryons by the
2362: criterion that their mean overdensity, $\delta_{\rm b}$, exceeds a
2363: value of 100, amounting to $\ga 50\%$ of the baryons that would
2364: assemble in the absence of gas pressure, according to the spherical
2365: top-hat collapse model. Equation~(\ref{eq:del}) then
2366: implies that $T_{\rm vir} > 17.2\, {\bar T}$.
2367: %PUT IN ANOTHER SECTION: This result
2368: %is in good agreement with the
2369: %threshold for the collapse of baryonic objects out of a reheated
2370: %(photo-ionized) IGM with ${\bar T}\approx 10^4~{\rm K}$; in that
2371: %case,
2372: %numerical simulations imply a threshold circular velocity of
2373: %$50$--$75~{\rm
2374: %km~s^{-1}}$ (Thoul \& Weinberg; Steinmetz; Kitayama \& Ikeuchi 1999=
2375: %astro-ph/9908084 *give more details and check that cooling is not
2376: %important
2377: %and the gas evolves adiabatically*) while we predict threshold of
2378: %$v_c\approx 70~{\rm km~s^{-1}}$.
2379: 
2380: As mentioned before, the gas temperature evolves at $z\la 160$ according to
2381: the relation ${\bar T}\approx 170 [(1+z) /100]^2\ {\rm K}$. This implies
2382: that baryons are overdense by $\delta_{\rm b} > 100$ only inside halos with
2383: a virial temperature $T_{\rm vir}\ga 2.9\times 10^3~[(1+z)/100]^2\ {\rm
2384: K}$. Based on the top-hat model, this implies a minimum halo mass for
2385: baryonic objects of
2386: \begin{equation}
2387: M_{\rm min}= 5.0 \times 10^3  \left({\Omega_m h^2\over
2388: 0.15}\right)^{-1/2} \left({\Omega_b h^2\over 0.022}\right)^{-3/5}
2389: \left({1+z\over 10}\right)^{3/2}~M_\odot,
2390: \label{eq:M_min}
2391: \end{equation}
2392: where we consider sufficiently high redshifts so that $\Ommz \approx
2393: 1$. This minimum mass is coincidentally almost identical to the naive
2394: Jeans mass calculation of linear theory in equation~(\ref{eq:m_j})
2395: despite the fact that it incorporates shell crossing by the dark
2396: matter, which is not accounted for by linear theory. Unlike the Jeans
2397: mass, the minimum mass depends on the choice for an overdensity
2398: threshold [taken arbitrarily as $\delta_{\rm b}>100$ in
2399: equation~(\ref{eq:M_min})]. To estimate the minimum halo mass which
2400: produces any significant accretion we set, e.g., $\delta_{\rm b}=5$,
2401: and get a mass which is lower than $M_{\rm min}$ by a factor of 27.
2402: 
2403: Of course, once the first stars and quasars form they heat the
2404: surrounding IGM by either outflows or radiation. As a result, the
2405: Jeans mass which is relevant for the formation of new objects changes
2406: \cite{G097,g00}). The most dramatic change
2407: occurs when the IGM is photo-ionized and is consequently heated to a
2408: temperature of $\sim(1$--$2)\times 10^4$ K. 
2409: 
2410: \section{\bf Fragmentation of the First Gaseous Objects to Stars}
2411: \label{sec4}
2412: 
2413: \subsection{Star Formation}
2414: \label{sec4.1}
2415: 
2416: As mentioned in the preface, the fragmentation of the first gaseous
2417: objects is a well-posed physics problem with well specified initial
2418: conditions, for a given power-spectrum of primordial density
2419: fluctuations.
2420: %Magnetic fields are expected to be dynamically unimportant in the
2421: %primordial cosmic plasma and the the molecular chemistry involves
2422: %only hydrogen (H$_2$ and HD). 
2423: This problem is ideally suited for three-dimensional computer
2424: simulations, since it cannot be reliably addressed in idealized 1D or
2425: 2D geometries.
2426: %(such as, CITE: Lahav
2427: %1986; Omukai \& Nishi ; Haiman, Thoul, \& Loeb... ).
2428: 
2429: Recently, two groups have attempted detailed 3D simulations of the
2430: formation process of the first stars in a halo of $\sim 10^6 M_\odot$ by
2431: following the dynamics of both the dark matter and the gas components,
2432: including H$_2$ chemistry and cooling. Bromm, Coppi, \& Larson (1999)
2433: \cite{BCL99} have used a Smooth Particle Hydrodynamics (SPH) code to
2434: simulate the collapse of a top-hat overdensity with a prescribed solid-body
2435: rotation (corresponding to a spin parameter $\lambda=5\%$) and additional
2436: small perturbations with $P(k)\propto k^{-3}$ added to the top-hat
2437: profile. Abel et al.\ (2002) \cite{ABN02} isolated a high-density filament
2438: out of a larger simulated cosmological volume and followed the evolution of
2439: its density maximum with exceedingly high resolution using an Adaptive Mesh
2440: Refinement (AMR) algorithm.
2441: 
2442: %%%%%%%% Figure: Cooling
2443: \begin{figure} % fig4.3
2444: \centering
2445: \includegraphics[height=6cm]{cooling.ps}
2446: \caption{Cooling rates as a function of temperature for a primordial
2447: gas composed of atomic hydrogen and helium, as well as molecular
2448: hydrogen, in the absence of any external radiation. We assume a
2449: hydrogen number density $n_H=0.045\ {\rm cm}^{-3}$, corresponding to
2450: the mean density of virialized halos at $z=10$. The plotted quantity
2451: $\Lambda/n_H^2$ is roughly independent of density (unless $n_H \ga 10\
2452: {\rm cm}^{-3}$), where $\Lambda$ is the volume cooling rate (in
2453: erg/sec/cm$^3$). The solid line shows the cooling curve for an atomic
2454: gas, with the characteristic peaks due to collisional excitation of
2455: \ion{H}{1} and \ion{He}{2}. The dashed line shows the additional
2456: contribution of molecular cooling, assuming a molecular abundance
2457: equal to $1\%$ of $n_H$.}
2458: \label{cooling}
2459: \end{figure}
2460: %%%%%%%%  
2461: 
2462: The generic results of Bromm et al.\ (1999 \cite{BCL99}; see also Bromm 2000 \cite{Bro00}) are
2463: illustrated in Figure \ref{fig4a}. The collapsing region forms a disk which
2464: fragments into many clumps. The clumps have a typical mass $\sim
2465: 10^2$--$10^3M_\odot$. This mass scale corresponds to the Jeans mass for a
2466: temperature of $\sim 500$K and the density $\sim 10^4~{\rm cm^{-3}}$ where
2467: the gas lingers because its cooling time is longer than its collapse time
2468: at that point (see Fig. \ref{fig4b}). Each clump accretes mass slowly
2469: until it exceeds the Jeans mass and collapses at a roughly constant
2470: temperature (isothermally) due to H$_2$ cooling that brings the gas to a
2471: fixed temperature floor.  The clump formation efficiency is high in this
2472: simulation due to the synchronized collapse of the overall top-hat
2473: perturbation.
2474: 
2475: \begin{figure} % fig4.3
2476: \centering
2477: \includegraphics[height=6cm]{IV_fig1.ps}
2478: %%\special{psfile=IV_fig1.ps
2479: %hoffset=30 voffset=-370 hscale=70 vscale=70}
2480: %\vspace{3.3in}
2481: \caption{Numerical results from Bromm et al.\ (1999) \cite{BCL99}, showing gas
2482: properties at $z=31.2$ for a collapsing slightly inhomogeneous top-hat
2483: region with a prescribed solid-body rotation.  {\bf (a)} Free electron
2484: fraction (by number) vs.\ hydrogen number density (in cm$^{-3}$). At
2485: densities exceeding $n\sim 10^{3}$ cm$^{-3}$, recombination is very
2486: efficient, and the gas becomes almost completely neutral.\ {\bf (b)}
2487: Molecular hydrogen fraction vs.\ number density. After a quick initial
2488: rise, the H$_{2}$ fraction approaches the asymptotic value of $f\sim
2489: 10^{-3}$, due to the H$^{-}$ channel.  {\bf (c)} Gas temperature vs.\ 
2490: number density. At densities below $\sim 1$ cm$^ {-3}$, the gas
2491: temperature rises because of adiabatic compression until it reaches
2492: the virial value of $T_{vir}\simeq 5000$ K.  At higher densities,
2493: cooling due to H$_{2}$ drives the temperature down again, until the
2494: gas settles into a quasi-hydrostatic state at $T\sim 500$ K and $n\sim
2495: 10^{4}$ cm$^{-3}$.  Upon further compression due to accretion and the
2496: onset of gravitational collapse, the gas shows a further modest rise
2497: in temperature. {\bf (d)} Jeans mass (in $M_{\odot}$) vs.\ number
2498: density. The Jeans mass reaches a value of $M_{J}\sim 10^{3}M_{\odot}$
2499: for the quasi-hydrostatic gas in the center of the potential well, and
2500: reaches the resolution limit of the simulation, $M_{\rm res}\simeq 200
2501: M_{\odot}$, for densities close to $n=10^{8}$ cm$^{-3}$.  }
2502: \label{fig4a}
2503: \end{figure}
2504:   
2505: \noindent
2506: \begin{figure} % fig4.4
2507: \centering
2508: \includegraphics[height=6cm]{IV_fig2.ps}
2509: %\special{psfile=IV_fig2.ps
2510: %hoffset=30 voffset=-320 hscale=70 vscale=70}
2511: %\vspace{3.3in}
2512: \caption{Gas and clump morphology at $z=28.9$ in
2513: the simulation of Bromm et al.\ (1999) \cite{BCL99}.
2514: {\it Top row:} The remaining gas in the diffuse phase.
2515: {\it Bottom row:} Distribution of clumps. The numbers next to the dots
2516: denote clump mass in units of $M_{\odot}$. 
2517: {\it Left panels:} Face-on view.
2518: {\it Right panels:} Edge-on view.
2519: The length of the box is 30 pc.
2520: The gas has settled into a flattened
2521: configuration with two dominant clumps of mass close to 
2522: $20,000 M_{\odot}$. During the subsequent evolution, the clumps
2523: survive
2524: without merging, and grow in mass only slightly by accretion of
2525: surrounding
2526: gas. }
2527: \label{fig4b}
2528: \end{figure}
2529:  
2530: Bromm (2000) \cite{Bro00} has simulated the collapse of one of the
2531: above-mentioned clumps with $\sim 1000 M_\odot$ and demonstrated that it
2532: does not tend to fragment into sub-components. Rather, the clump core of
2533: $\sim 100M_\odot$ free-falls towards the center leaving an extended
2534: envelope behind with a roughly isothermal density profile.  At very high
2535: gas densities, three-body reactions become important in the chemistry of
2536: H$_2$.  Omukai \& Nishi (1998) \cite{ON1998} have included these reactions
2537: as well as radiative transfer and followed the collapse in spherical
2538: symmetry up to stellar densities. Radiation pressure from nuclear burning
2539: at the center is unlikely to reverse the infall as the stellar mass builds
2540: up.  These calculations indicate that each clump may end as a single
2541: massive star; however, it is conceivable that angular momentum may
2542: eventually halt the collapsing cloud and lead to the formation of a binary
2543: stellar system instead.
2544: 
2545: The Jeans mass, which is defined based on small fluctuations in a
2546: background of {\it uniform}\/ density, does not strictly apply in the
2547: context of collapsing gas cores. We can instead use a slightly modified
2548: critical mass known as the Bonnor-Ebert mass \cite{Bo56,Eb55}. For baryons
2549: in a background of uniform density $\rho_b$, perturbations are unstable to
2550: gravitational collapse in a region more massive than the Jeans
2551: mass. Instead of a uniform background, we consider a spherical,
2552: non-singular, isothermal, self-gravitating gas in hydrostatic equilibrium,
2553: i.e., a centrally-concentrated object which more closely resembles the gas
2554: cores found in the above-mentioned simulations. In this case, small
2555: fluctuations are unstable and lead to collapse if the sphere is more
2556: massive than the Bonnor-Ebert mass $M_{\rm BE}$, given by the same
2557: expression the Jeans Mass but with a different coefficient (1.2
2558: instead of 2.9) and with $\rho_b$ denoting in this case the gas (volume)
2559: density at the surface of the sphere,
2560:  \beq M_{\rm BE}=1.2\,
2561: \frac{1}{\sqrt{\rho_b}}\,\left( \frac{k T} {G \mu m_p} \right)^{3/2}\
2562: . \label{MJb} \eeq
2563: 
2564: In their simulation, Abel et al.\ (2000)\cite{Abel} adopted the actual
2565: cosmological density perturbations as initial conditions. The simulation
2566: focused on the density peak of a filament within the IGM, and evolved it to
2567: very high densities (Fig. \ref{fig4c}). Following the initial collapse of
2568: the filament, a clump core formed with $\sim 200M_\odot$, amounting to only
2569: $\sim 1\%$ of the virialized mass.  Subsequently due to slow cooling, the
2570: clump collapsed subsonically in a state close to hydrostatic equilibrium
2571: (see Fig. \ref{fig4d}).  Unlike the idealized top-hat simulation of Bromm
2572: et al.\ (2001) \cite{BKL2001}, the collapse of the different clumps within
2573: the filament is not synchronized.  Once the first star forms at the center
2574: of the first collapsing clump, it is likely to affect the formation of
2575: other stars in its vicinity.
2576: 
2577: As soon as nuclear burning sets in the core of the proto-star, the
2578: radiation emitted by the star starts to affect the infall of the
2579: surrounding gas towards it.  The radiative feedback involves
2580: photo-dissociation of H$_2$, Ly$\alpha$ radiation pressure, and
2581: photo-evaporation of the accretion disk. Tan \& McKee \cite{Tan2003}
2582: studied these effects by extrapolating analytically the infall of gas from
2583: the final snapshot of the above resolution-limited simulations to the scale
2584: of a proto-star; they concluded that nuclear burning (and hence the
2585: feedback) starts when the proton-star accretes $\sim 30M_\odot$ and
2586: accretion is likely to be terminated when the star reaches $\sim
2587: 200M_\odot$.
2588: 
2589: \noindent
2590: \begin{figure} % fig4.5
2591: \centering
2592: \includegraphics[height=6cm]{IV_fig3.ps}
2593: %angle=-90
2594: %hoffset=0 voffset=-50 hscale=60 vscale=60}
2595: %\vspace{6.2in}
2596: \caption{Zooming in on the core of a star forming region with the {\it
2597: Adaptive Mesh Refinement} simulation of Abel et al.\ (2000)
2598: \cite{Abel}. The panels show different length scales, decreasing clockwise
2599: by an order of magnitude between adjacent panels. Note the large dynamic
2600: range of scales which are being resolved, from 6 kpc (top left panel) down
2601: to 10,000 AU (bottom left panel).}
2602: \label{fig4c}
2603: \end{figure}
2604:   
2605: \noindent
2606: \begin{figure} % fig4.6
2607: %\epsscale{0.7}
2608: %\plotone{IV_fig4.eps}
2609: \centering
2610: \includegraphics[height=6cm]{IV_fig4.eps}
2611: \caption{Gas profiles from the simulation of Abel et al.\
2612: (2000)\cite{Abel}.  The cell size on the finest grid corresponds to $0.024$
2613: pc, while the simulation box size corresponds to 6.4 kpc.  Shown are
2614: spherically-averaged mass-weighted profiles around the baryon density peak
2615: shortly before a well defined fragment forms ($z=19.1$). Panel (a) shows
2616: the baryonic number density, enclosed gas mass in solar mass, and the local
2617: Bonnor-Ebert mass $M_{\rm BE}$ (see text).
2618: %($\sim 27M_\odot (T/1{\rm K})^{1.5}/\sqrt{n}$).  
2619: Panel (b) plots the molecular hydrogen fraction (by number) $f_{H_2}$
2620: and the free electron fraction $x$.  The H$_2$ cooling time,
2621: $t_{H_2}$, the time it takes a sound wave to travel to the center,
2622: $t_{\rm cross}$, and the free--fall time $t_{\rm ff}=[3\pi/(32G\rho
2623: )]^{1/2}$ are given in panel (c). Panel (d) gives the temperature in K
2624: as a function of radius.  The bottom panel gives the local sound
2625: speed, $c_s$ (solid line with circles), the rms radial velocities of
2626: the dark matter (dashed line) and the gas (dashed line with asterisks)
2627: as well as the rms gas velocity (solid line with square symbols). The
2628: vertical dotted line indicates the radius ($\sim 5$ pc) at which the
2629: gas has reached its minimum temperature allowed by H$_2$ cooling. The
2630: virial radius of the $5.6\times 10^{6}M_\odot$ halo is 106 pc.  }
2631: \label{fig4d}
2632: \end{figure}
2633:   
2634: If the clumps in the above simulations end up forming individual very
2635: massive stars, then these stars will likely radiate copious amounts of
2636: ionizing radiation \cite{CBA84,TS00,BKL2001} and expel strong
2637: winds.  Hence, the stars will have a large effect on their interstellar
2638: environment, and feedback is likely to control the overall star formation
2639: efficiency.  This efficiency is likely to be small in galactic potential
2640: wells which have a virial temperature lower than the temperature of
2641: photoionized gas, $\sim 10^4$K. In such potential wells, the gas may go
2642: through only a single generation of star formation, leading to a
2643: ``suicidal'' population of massive stars.
2644: 
2645: The final state in the evolution of these stars is uncertain; but if their
2646: mass loss is not too extensive, then they are likely to end up as black
2647: holes \cite{CBA84,FWH00}. The remnants may provide the seeds of quasar black
2648: holes \cite{Lar99}.  Some of the massive stars may end their lives by
2649: producing gamma-ray bursts. If so then the broad-band afterglows of these
2650: bursts could provide a powerful tool for probing the epoch of reionization
2651: \cite{LR00,CL00}).
2652: There is no better way to end the dark ages than with $\gamma$-ray burst
2653: fireworks.
2654: 
2655: {\it Where are the first stars or their remnants located today?} The very
2656: first stars formed in rare high-$\sigma$ peaks and hence are likely to
2657: populate the cores of present-day galaxies \cite{WS99}. However, the bulk
2658: of the stars which formed in low-mass systems at later times are expected
2659: to behave similarly to the collisionless dark matter particles and populate
2660: galaxy halos \cite{Loe98}.
2661: 
2662: 
2663: \subsection{The Mass Function of Stars}
2664: 
2665: Currently, we do not have direct observational constraints on how
2666: the first stars, the so-called Population~III stars, formed at
2667: the end of the cosmic dark ages. It is, therefore, instructive to
2668: briefly summarize what we have learned about star formation in the
2669: present-day Universe, where theoretical reasoning is guided by a
2670: wealth of observational data (see \cite{Pud2002} for a recent review).
2671: 
2672: Population~I stars form out of cold, dense molecular gas that is
2673: structured in a complex, highly inhomogeneous way. The molecular
2674: clouds are supported against gravity by turbulent velocity fields and
2675: pervaded on large scales by magnetic fields.  Stars tend to form in
2676: clusters, ranging from a few hundred up to $\sim 10^{6}$ stars. It
2677: appears likely that the clustered nature of star formation leads to
2678: complicated dynamics and tidal interactions that transport angular
2679: momentum, thus allowing the collapsing gas to overcome the classical
2680: centrifugal barrier \cite{Lar2002}.  The initial 
2681: mass function (IMF) of Pop~I stars is
2682: observed to have the approximate Salpeter form (e.g., \cite{Kr02})
2683: \begin{equation}
2684: \frac{{\rm d}N}{{\rm d log}M}\propto M^{x} \mbox{\ ,}
2685: \end{equation}
2686: where
2687: \begin{equation}
2688: x\simeq \left\{
2689: \begin{array}{rl}
2690: -1.35 & \mbox{for \ }M\ge 0.5 M_{\odot}\\
2691: 0.0 & \mbox{for \ }0.007 \le M\le 0.5 M_{\odot}\\
2692: \end{array}
2693: \right. \mbox{\ .}
2694: \end{equation}
2695: The lower cutoff in mass corresponds roughly to the opacity limit for
2696: fragmentation. This limit reflects the minimum fragment mass, set when the
2697: rate at which gravitational energy is released during the collapse exceeds
2698: the rate at which the gas can cool (e.g., \cite{MJR1976}).  The most
2699: important feature of the observed IMF is that $\sim 1 M_{\odot}$ is the
2700: characteristic mass scale of Pop~I star formation, in the sense that most
2701: of the mass goes into stars with masses close to this value. In Figure
2702: \ref{firstfig}, we show the result from a recent hydrodynamical simulation of
2703: the collapse and fragmentation of a molecular cloud core
2704: \cite{BBB2002,BBB2003}.  This simulation illustrates the highly dynamic and
2705: chaotic nature of the star formation process\footnote{See http://
2706: www.ukaff.ac.uk/starcluster for an animation.}.
2707: 
2708: \begin{figure}
2709:   \includegraphics[height=.3\textheight]{Fig1a.eps}
2710:   \includegraphics[height=.3\textheight]{Fig1b.eps}
2711: \caption{A hydrodynamic simulation of the collapse and fragmentation
2712: of a turbulent molecular cloud in the present-day Universe (from
2713: \cite{BBB2003}).  The cloud has a mass of $50 M_{\odot}$.  The panels
2714: show the column density through the cloud, and span a scale of 0.4 pc
2715: across.  {\it Left:} The initial phase of the collapse. The turbulence
2716: organizes the gas into a network of filaments, and decays thereafter
2717: through shocks.  {\it Right:} A snapshot taken near the end of the
2718: simulation, after 1.4 initial free-fall times of $2\times 10^{5}$yr.
2719: Fragmentation has resulted in $\sim 50$ stars and brown dwarfs.
2720: The star formation efficiency is $\sim 10$\% on the scale of the
2721: overall cloud, but can be much larger in the dense sub-condensations.
2722: This result is in good agreement with what is observed in local star-forming
2723: regions.}
2724: \label{firstfig}
2725: \end{figure}
2726: 
2727: The metal-rich chemistry, magnetohydrodynamics, and radiative
2728: transfer involved in present-day star formation is complex, and we
2729: still lack a comprehensive theoretical framework that predicts the IMF
2730: from first principles. Star formation in the high redshift Universe,
2731: on the other hand, poses a theoretically more tractable problem due to
2732: a number of simplifying features, such as: (i) the initial absence of
2733: heavy metals and therefore of dust; and (ii) the absence of
2734: dynamically-significant magnetic fields, in the pristine gas left over
2735: from the big bang. The cooling of the primordial gas does then only
2736: depend on hydrogen in its atomic and molecular form.  Whereas in the
2737: present-day interstellar medium, the initial state of the star forming
2738: cloud is poorly constrained, the corresponding initial conditions for
2739: primordial star formation are simple, given by the popular
2740: $\Lambda$CDM model of cosmological structure formation. We now turn to
2741: a discussion of this theoretically attractive and important problem.
2742: 
2743: {\it How did the first stars form?} A complete answer to this question
2744: would entail a theoretical prediction for the Population~III IMF, which is
2745: rather challenging. Let us start by addressing the simpler problem of
2746: estimating the characteristic mass scale of the first stars. As mentioned
2747: before, this mass scale is observed to be $\sim 1 M_{\odot}$ in the
2748: present-day Universe.  
2749: 
2750: Bromm \& Loeb (2004) \cite{BL04} carried out idealized simulations of the
2751: protostellar accretion problem and estimated the final mass of a
2752: Population~III star.  Using the smoothed particle hydrodynamics (SPH)
2753: method, they included the chemistry and cooling physics relevant for the
2754: evolution of metal-free gas (see \cite{BCL02} for details). Improving on
2755: earlier work \cite{BCL99,BCL02} by initializing the simulations according
2756: to the $\Lambda$CDM model, they focused on an isolated overdense region
2757: that corresponds to a 3$\sigma-$peak \cite{BL04}: a halo containing a total
2758: mass of $10^{6}M_{\odot}$, and collapsing at a redshift $z_{\rm vir}\simeq
2759: 20$. In these runs, one high-density clump has formed at the center of the
2760: minihalo, possessing a gas mass of a few hundred solar masses.  Soon after
2761: its formation, the clump becomes gravitationally unstable and undergoes
2762: runaway collapse. Once the gas clump has exceeded a threshold density of
2763: $10^{7}$ cm$^{-3}$, it is replaced by a sink particle which is a
2764: collisionless point-like particle that is inserted into the simulation.
2765: This choice for the density threshold ensures that the local Jeans mass is
2766: resolved throughout the simulation.  The clump (i.e., sink particle) has an
2767: initial mass of $M_{\rm Cl}\simeq 200M_{\odot}$, and grows subsequently by
2768: ongoing accretion of surrounding gas.  High-density clumps with such masses
2769: result from the chemistry and cooling rate of molecular hydrogen, H$_{2}$,
2770: which imprint characteristic values of temperature, $T\sim 200$~K, and
2771: density, $n\sim 10^{4}$ cm$^{-3}$, into the metal-free gas \cite{BCL02}.
2772: Evaluating the Jeans mass for these characteristic values results in
2773: $M_{J}\sim \mbox{\ a few \ }\times 10^{2}M_{\odot}$, which is close to the
2774: initial clump masses found in the simulations.
2775: 
2776: 
2777: \begin{figure}
2778:   \includegraphics[height=.3\textheight]{Fig2a.eps}
2779:   \includegraphics[height=.3\textheight]{Fig2b.eps}
2780: \caption{Collapse and fragmentation of a primordial cloud (from
2781: \cite{BL04}).  Shown is the projected gas density at a redshift $z\simeq
2782: 21.5$, briefly after gravitational runaway collapse has commenced in the
2783: center of the cloud.  {\it Left:} The coarse-grained morphology in a box
2784: with linear physical size of 23.5~pc.  At this time in the unrefined
2785: simulation, a high-density clump (sink particle) has formed with an initial
2786: mass of $\sim 10^{3}M_{\odot}$.  {\it Right:} The refined morphology in
2787: a box with linear physical size of 0.5~pc.  The central density peak,
2788: vigorously gaining mass by accretion, is accompanied by a secondary clump.}
2789: \label{2ab}
2790: \end{figure}
2791: 
2792: \begin{figure}[ht]
2793:   \includegraphics[height=.3\textheight]{Fig3a.eps}
2794:   \includegraphics[height=.3\textheight]{Fig3b.eps}
2795: \caption{Accretion onto a primordial protostar (from \cite{BL04}).  The
2796: morphology of this accretion flow is shown in Fig.~\ref{2ab}.  {\it Left:}
2797: Accretion rate (in $M_{\odot}$~yr$^{-1}$) vs. time (in yr) since molecular
2798: core formation.  {\it Right:} Mass of the central core (in $M_{\odot}$)
2799: vs. time.  {\it Solid line:} Accretion history approximated as:
2800: $M_{\ast}\propto t^{0.45}$.  Using this analytical approximation, we
2801: extrapolate that the protostellar mass has grown to $\sim 150 M_{\odot}$
2802: after $\sim 10^{5}$~yr, and to $\sim 700 M_{\odot}$ after $\sim 3\times
2803: 10^{6}$~yr, the total lifetime of a very massive star.  }
2804: \label{fig3ab}
2805: \end{figure}
2806: 
2807: The high-density clumps are clearly not stars yet. To probe the subsequent
2808: fate of a clump, Bromm \& Loeb (2004) \cite{BL04} have re-simulated the
2809: evolution of the central clump with sufficient resolution to follow the
2810: collapse to higher densities. Figure~\ref{2ab} ({\it right panel})
2811: shows the gas density on a scale of 0.5~pc, which is two orders of magnitude
2812: smaller than before. Several features are evident in this plot. First, the
2813: central clump does not undergo further sub-fragmentation, and is likely to
2814: form a single Population~III star. Second, a companion clump is visible at
2815: a distance of $\sim 0.25$~pc. If negative feedback from the first-forming
2816: star is ignored, this companion clump would undergo runaway collapse on its
2817: own approximately $\sim 3$~Myr later.  This timescale is comparable to the
2818: lifetime of a very massive star (VMS)\cite{BKL2001}.  If the second clump
2819: was able to survive the intense radiative heating from its neighbor, it
2820: could become a star before the first one explodes as a supernova
2821: (SN). Whether more than one star can form in a low-mass halo thus crucially
2822: depends on the degree of synchronization of clump formation.  Finally, the
2823: non-axisymmetric disturbance induced by the companion clump, as well as the
2824: angular momentum stored in the orbital motion of the binary system, allow
2825: the system to overcome the angular momentum barrier for the collapse of the
2826: central clump (see \cite{Lar2002}).
2827: 
2828: 
2829: The recent discovery of stars like HE0107-5240 with a mass of $0.8
2830: M_{\odot}$ and an iron abundance of ${\rm [Fe/H]} = -5.3$ \cite{Cr02} shows
2831: that at least some low mass stars could have formed out of extremely
2832: low-metallicity gas.  The above simulations show that although the majority
2833: of clumps are very massive, a few of them, like the secondary clump in
2834: Fig.~\ref{2ab}, are significantly less massive. Alternatively, low-mass
2835: fragments could form in the dense, shock-compressed shells that surround
2836: the first hypernovae \cite{MBH03}.
2837: 
2838: {\it How massive were the first stars?} Star formation typically
2839: proceeds from the `inside-out', through the accretion of gas onto a
2840: central hydrostatic core.  Whereas the initial mass of the hydrostatic
2841: core is very similar for primordial and present-day star formation
2842: \cite{ON1998}, the accretion process -- ultimately responsible for
2843: setting the final stellar mass, is expected to be rather different. On
2844: dimensional grounds, the accretion rate is simply related to the sound
2845: speed cubed over Newton's constant (or equivalently given by the ratio
2846: of the Jeans mass and the free-fall time): $\dot{M}_{\rm acc}\sim
2847: c_s^3/G \propto T^{3/2}$. A simple comparison of the temperatures in
2848: present-day star forming regions ($T\sim 10$~K) with those in
2849: primordial ones ($T\sim 200-300$~K) already indicates a difference in
2850: the accretion rate of more than two orders of magnitude.
2851: 
2852: The above refined simulation enables one to study the three-dimensional
2853: accretion flow around the protostar (see also
2854: \cite{OP2001,Rip2002,Tan2003}).  The gas may now reach densities of
2855: $10^{12}$ cm$^{-3}$ before being incorporated into a central sink
2856: particle. At these high densities, three-body reactions \cite{PSS1983} 
2857: convert the gas into a fully molecular form.  Figure~\ref{fig3ab}
2858: shows how the molecular core grows in mass over the first $\sim 10^{4}$~yr
2859: after its formation. The accretion rate ({\it left panel}) is initially
2860: very high, $\dot{M}_{\rm acc}\sim 0.1 M_{\odot}$~yr$^{-1}$, and
2861: subsequently declines according to a power law, with a possible break at
2862: $\sim 5000$~yr. The mass of the molecular core ({\it right panel}), taken
2863: as an estimator of the proto-stellar mass, grows approximately as:
2864: $M_{\ast}\sim \int \dot{M}_{\rm acc}{\rm d}t \propto t^{0.45}$. A rough
2865: upper limit for the final mass of the star is then: $M_{\ast}(t=3\times
2866: 10^{6}{\rm yr})\sim 700 M_{\odot}$. In deriving this upper bound, we have
2867: conservatively assumed that accretion cannot go on for longer than the
2868: total lifetime of a massive star.
2869: 
2870: {\it Can a Population~III star ever reach this asymptotic mass limit?}  The
2871: answer to this question is not yet known with any certainty, and it depends
2872: on whether the accretion from a dust-free envelope is eventually terminated
2873: by feedback from the star (e.g., \cite{OP2001,Rip2002,Tan2003,OI2002}). The
2874: standard mechanism by which accretion may be terminated in metal-rich gas,
2875: namely radiation pressure on dust grains \cite{WC1987}, is evidently not
2876: effective for gas with a primordial composition. Recently, it has been
2877: speculated that accretion could instead be turned off through the formation
2878: of an H~II region \cite{OI2002}, or through the photo-evaporation of the
2879: accretion disk \cite{Tan2003}. The termination of the accretion process
2880: defines the current unsolved frontier in studies of Population~III star
2881: formation. Current simulations indicate that the first stars were
2882: predominantly very massive ($\ga 30M_\odot$), and consequently rather
2883: different from present-day stellar populations. The crucial question then
2884: arises: {\it How and when did the transition take place from the early
2885: formation of massive stars to that of low-mass stars at later times?}  We
2886: address this problem next.
2887: 
2888: The very first stars, marking the cosmic Renaissance of structure
2889: formation, formed under conditions that were much simpler than the
2890: highly complex environment in present-day molecular clouds.
2891: Subsequently, however, the situation rapidly became more complicated
2892: again due to the feedback from the first stars on the IGM.  Supernova
2893: explosions dispersed the nucleosynthetic products from the first
2894: generation of stars into the surrounding gas (e.g., \cite{MFR01,MFM02,
2895: TSD02}), including also dust grains produced in the explosion itself
2896: \cite{LH97,TodF01}.  Atomic and molecular cooling became much more
2897: efficient after the addition of these metals. Moreover, the
2898: presence of ionizing cosmic rays, as well as of UV and X-ray
2899: background photons, modified the thermal and chemical behavior of
2900: the gas in important ways (e.g., \cite{MBA01,MBA03}).
2901: 
2902: Early metal enrichment was likely the dominant effect that brought
2903: about the transition from Population~III to Population~II star
2904: formation.  Recent numerical simulations of collapsing primordial
2905: objects with overall masses of $\sim 10^{6}M_{\odot}$, have shown that
2906: the gas has to be enriched with heavy elements to a minimum level of
2907: $Z_{\rm crit}\simeq 10^{-3.5}Z_{\odot}$, in order to have any effect
2908: on the dynamics and fragmentation properties of the system
2909: \cite{Om00,BFCL01,BrL03}.  Normal, low-mass (Population~II) stars are
2910: hypothesized to only form out of gas with metallicity $Z\ge Z_{\rm
2911: crit}$.  Thus, the characteristic mass scale for star formation is
2912: expected to be a function of metallicity, with a discontinuity at
2913: $Z_{\rm crit}$ where the mass scale changes by $\sim$ two orders of
2914: magnitude. The redshift where this transition occurs has important
2915: implications for the early growth of cosmic structure, and the
2916: resulting observational signature (e.g.,
2917: \cite{Wyithe03,FL03,MBH03,Sch02}) 
2918: include the extended 
2919: nature of reionization \cite{FL05}.
2920: 
2921: %Important caveats, however, remain. The determination of the critical
2922: %metallicity mentioned above \cite{BFCL01,BrL03} implicitly assumes that the
2923: %gas at temperatures below $\sim 8000$~K is maintained in ionization
2924: %equilibrium by cosmic rays, with an ionization rate that is scaled
2925: %from the Galactic value by the factor $Z/Z_{\odot}$. The cosmic-ray
2926: %flux in the early Universe might well have not obeyed this simple
2927: %relation, and it is not clear whether cosmic rays could have
2928: %successfully `activated' the metals.
2929: 
2930: For additional detailes about the properties of the first stars, see the
2931: comprehensive review by Bromm \& Larson (2004) \cite{Bromm}.
2932: 
2933: \subsection{Gamma-ray Bursts: Probing the First Stars One 
2934: Star at a Time}
2935: 
2936: Gamma-Ray Bursts (GRBs) are believed to originate in compact remnants
2937: (neutron stars or black holes) of massive stars. Their high luminosities
2938: make them detectable out to the edge of the visible Universe
2939: \cite{CL00,LR00}.  GRBs offer the opportunity to detect the most distant
2940: (and hence earliest) population of massive stars, the so-called
2941: Population~III (or Pop~III), one star at a time.  In the hierarchical
2942: assembly process of halos which are dominated by cold dark matter (CDM),
2943: the first galaxies should have had lower masses (and lower stellar
2944: luminosities) than their low-redshift counterparts.  Consequently, the
2945: characteristic luminosity of galaxies or quasars is expected to decline
2946: with increasing redshift. GRB afterglows, which already produce a peak flux
2947: comparable to that of quasars or starburst galaxies at $z\sim 1-2$, are
2948: therefore expected to outshine any competing source at the highest
2949: redshifts, when the first dwarf galaxies have formed in the Universe.
2950: 
2951: \begin{figure}
2952: \centering
2953: \includegraphics[height=6cm]{GRB.ps}
2954: \caption{Illustration of a long-duration gamma-ray burst in the popular
2955: ``collapsar'' model.
2956: %(credit: NASA E/PO).
2957: The collapse of the core of a massive star (which lost
2958: its hydrogen envelope) to a black hole generates two opposite jets moving
2959: out at a speed close to the speed of light. The jets drill a hole in the
2960: star and shine brightly towards an observer who happened to be located
2961: within with the collimation cones of the jets. The jets emenating from a
2962: single massive star are so bright that they can be seen across the Universe
2963: out to the epoch when the first stars have formed.  Upcoming observations
2964: by the {\it Swift} satellite will have the sensitivity to reveal whether
2965: the first stars served as progenitors of gamma-ray bursts
2966: (for updates see http://swift.gsfc.nasa.gov/).}
2967: \label{grb}
2968: \end{figure}
2969: 
2970: The first-year polarization data from the {\it Wilkinson Microwave
2971: Anisotropy Probe} ({\it WMAP}) indicates an optical depth to electron
2972: scattering of $\sim 17\pm 4$\% after cosmological recombination
2973: \cite{Kog03,WMAP}.  This implies that the first stars must have formed at
2974: a redshift $z\sim $10--20, and reionized a substantial fraction of the
2975: intergalactic hydrogen around that time \cite{Cen03,CFW03,SL03,WL03,YBH04}.
2976: Early reionization can be achieved with plausible star formation parameters
2977: in the standard $\Lambda$CDM cosmology; in fact, the required optical depth
2978: can be achieved in a variety of very different ionization histories since
2979: {\it WMAP} places only an integral constraint on these histories
2980: \cite{HH03}. One would like to probe the full history of reionization in
2981: order to disentangle the properties and formation history of the stars that
2982: are responsible for it. GRB afterglows offer the opportunity to detect
2983: stars as well as to probe the metal
2984: enrichment level \cite{FL03} of the intervening IGM.
2985: 
2986: GRBs, the electromagnetically-brightest explosions in the Universe, should
2987: be detectable out to redshifts $z>10$ \cite{CL00,LR00}.
2988: High-redshift GRBs can be identified through infrared
2989: photometry, based on the Ly$\alpha$ break induced by absorption of their
2990: spectrum at wavelengths below $1.216\, \mu {\rm m}\, [(1+z)/10]$. Follow-up
2991: spectroscopy of high-redshift candidates can then be performed on a
2992: 10-meter-class telescope. Recently, the ongoing {\it Swift} mission
2993: \cite{Geh04} has detected a GRB originating at $z\simeq 6.3$
2994: (e.g., \cite{Hai05}), thus demonstrating the viability of
2995: GRBs as probes of the early Universe.
2996: 
2997: There are four main advantages of GRBs relative to traditional cosmic
2998: sources such as quasars:
2999: 
3000: \noindent {\it (i)} The GRB afterglow flux at a given observed time lag
3001: after the $\gamma$-ray trigger is not expected to fade significantly with
3002: increasing redshift, since higher redshifts translate to earlier times in
3003: the source frame, during which the afterglow is intrinsically brighter
3004: \cite{CL00}. For standard afterglow lightcurves and spectra, the
3005: increase in the luminosity distance with redshift is compensated by this
3006: {\it cosmological time-stretching} effect.
3007: 
3008: \begin{figure}
3009: \centering
3010: \includegraphics[height=10cm]{fig1val.eps}
3011: \caption{GRB afterglow flux as a function of time since the $\gamma$-ray
3012: trigger in the observer frame (taken from \cite{BL04}). The flux
3013: (solid curves) is calculated at the redshifted Ly$\alpha$ wavelength. The
3014: dotted curves show the planned detection threshold for the {\it James Webb
3015: Space Telescope} ({\it JWST}), assuming a spectral resolution $R=5000$
3016: with the near infrared spectrometer, a signal to noise ratio of 5 per
3017: spectral resolution element, and an exposure time equal to $20\%$ of the
3018: time since the GRB explosion (see  http://www.ngst.stsci.edu/nms/main/~).
3019: Each set of curves shows a sequence of redshifts, namely $z=5$, 7, 9, 11,
3020: 13, and 15, respectively, from top to bottom.}
3021: \label{fig1val}
3022: \end{figure}
3023: 
3024: 
3025: \noindent {\it (ii)} As already mentioned, in the standard $\Lambda$CDM
3026: cosmology, galaxies form hierarchically, starting from small masses and
3027: increasing their average mass with cosmic time. Hence, the characteristic
3028: mass of quasar black holes and the total stellar mass of a galaxy were
3029: smaller at higher redshifts, making these sources intrinsically fainter
3030: \cite{WL02}.  However, GRBs are believed to originate from a stellar mass
3031: progenitor and so the intrinsic luminosity of their engine should not
3032: depend on the mass of their host galaxy. GRB afterglows are therefore
3033: expected to outshine their host galaxies by a factor that gets larger with
3034: increasing redshift.
3035: 
3036: \noindent {\it (iii)} Since the progenitors of GRBs are believed to be
3037: stellar, they likely originate in the most common star-forming galaxies at
3038: a given redshift rather than in the most massive host galaxies, as is the
3039: case for bright quasars \cite{GRBquasar}. Low-mass host galaxies
3040: induce only a weak ionization effect on the surrounding IGM and do not
3041: greatly perturb the Hubble flow around them. Hence, the Ly$\alpha$ damping
3042: wing should be closer to the idealized unperturbed IGM case
3043: and its detailed spectral shape should be easier
3044: to interpret. Note also that unlike the case of a quasar, a GRB afterglow
3045: can itself ionize at most $\sim 4\times 10^4 E_{51} M_\odot$ of hydrogen if
3046: its UV energy is $E_{51}$ in units of $10^{51}$ ergs (based on the
3047: available number of ionizing photons), and so it should have a negligible
3048: cosmic effect on the surrounding IGM. 
3049: 
3050: \noindent
3051: {\it (iv)} GRB afterglows have smooth (broken power-law) continuum spectra
3052: unlike quasars which show strong spectral features (such as broad emission
3053: lines or the so-called ``blue bump'') that complicate the extraction of IGM
3054: absorption features. In particular, the continuum extrapolation into the
3055: Ly$\alpha$ damping wing (the so-called Gunn-Peterson absorption trough)
3056: during the epoch of reionization is much more straightforward for the
3057: smooth UV spectra of GRB afterglows than for quasars with an underlying
3058: broad Ly$\alpha$ emission line \cite{GRBquasar}.
3059: 
3060: The optical depth of the uniform IGM to Ly$\alpha$ absorption is
3061: given by (Gunn \& Peterson 1965 \cite{GP}),
3062: \begin{equation}
3063: \tau_{s}={\pi e^2 f_\alpha \lambda_\alpha n_{\HI}(z_s) \over m_e
3064: cH(z_s)} \approx 6.45\times 10^5 x_{\HI} \left({\Omega_bh\over
3065: 0.03}\right)\left({\Omega_m\over 0.3}\right)^{-1/2} \left({1+z_s\over
3066: 10}\right)^{3/2} \label{G-P}
3067: \end{equation}
3068: where $H\approx 100h~{\rm km~s^{-1}~Mpc^{-1}}\Omega_m^{1/2}(1+z_s)^{3/2}$
3069: is the Hubble parameter at the source redshift $z_s>>1$, $f_\alpha=0.4162$
3070: and $\lambda_\alpha=1216$\AA~ are the oscillator strength and the
3071: wavelength of the Ly$\alpha$ transition; $n_{\HI}(z_s)$ is the neutral
3072: hydrogen density at the source redshift (assuming primordial abundances);
3073: $\Omega_m$ and $\Omega_b$ are the present-day density parameters of all
3074: matter and of baryons, respectively; and $x_{\HI}$ is the average fraction
3075: of neutral hydrogen. In the second equality we have implicitly considered
3076: high-redshifts, $(1+z)\gg {\rm
3077: max}\left[(1-\Omega_m-\Omega_\Lambda)/\Omega_m,
3078: (\Omega_\Lambda/\Omega_m)^{1/3}\right]$, at which the vacuum energy density
3079: is negligible relative to matter ($\Omega_\Lambda\ll \Omega_m$) and the
3080: Universe is nearly flat; for $\Omega_m=0.3, \Omega_\Lambda=0.7$ this
3081: corresponds to the condition $z\gg 1.3$ which is well satisfied by the
3082: reionization redshift.
3083: 
3084: At wavelengths longer than \lya at the source, the optical depth
3085: obtains a small value; these photons redshift away from the line
3086: center along its red wing and never resonate with the line core on
3087: their way to the observer. The red damping wing of the Gunn-Peterson
3088: trough (Miralda-Escud\'e 1998 \cite{Mir98})
3089: \begin{equation}
3090: \tau(\lambda_{\rm obs})=\tau_s \left(\Lambda\over
3091: 4\pi^2\nu_\alpha\right) {\tilde \lambda}_{\rm obs}^{3/2}\left[
3092: I({\tilde\lambda}_{\rm obs}^{-1}) -
3093: I([(1+z_{i})/(1+z_s)]{\tilde\lambda}_{\rm obs}^{-1})\right]~~{\rm
3094: for}~~{\tilde\lambda}_{\rm obs}\geq 1~,
3095: \label{eq:shift}
3096: \end{equation}
3097: where $\tau_s$ is given in equation~(\ref{G-P}), also we define
3098: \begin{equation}
3099: {\tilde \lambda}_{\rm obs}\equiv {\lambda_{\rm obs}\over
3100: (1+z_s)\lambda_\alpha}
3101: \end{equation}
3102: and
3103: \begin{equation}
3104: I(x)\equiv {x^{9/2}\over 1-x}+{9\over 7}x^{7/2}+{9\over 5}x^{5/2}+ 3
3105: x^{3/2}+9 x^{1/2}-{9\over 2} \ln\left[ {1+x^{1/2}\over 1-x^{1/2}}
3106: \right]\ .
3107: \end{equation}
3108: 
3109: \begin{figure}
3110: \centering
3111: \includegraphics[height=10cm]{fig2val.eps}
3112: \caption{Expected spectral shape of the Ly$\alpha$ absorption trough due to
3113: intergalactic absorption in GRB afterglows (taken from
3114: \cite{BL04}). The spectrum is presented in terms of the flux density
3115: $F_{\nu}$ versus relative observed wavelength $\Delta \lambda$, for a
3116: source redshift $z=7$ (assumed to be prior to the final reionization phase)
3117: and the typical halo mass $M=4 \times 10^8 M_{\odot}$ expected for GRB host
3118: galaxies that cool via atomic transitions. {\it Top panel:} Two examples
3119: for the predicted spectrum including IGM HI absorption (both resonant and
3120: damping wing), for host galaxies with (i) an age $t_S=10^7$ yr, a UV escape
3121: fraction $f_{\rm esc}=10\%$ and a Scalo initial mass function (IMF) in
3122: solid curves, or (ii) $t_S=10^8$ yr, $f_{\rm esc}=90\%$ and massive
3123: ($>100M_\odot$) Pop III stars in dashed curves. The observed time after the
3124: $\gamma$-ray trigger is one hour, one day, and ten days, from top to
3125: bottom, respectively. {\it Bottom panel:} Predicted spectra one day after a
3126: GRB for a host galaxy with $t_S=10^7$ yr, $f_{\rm esc}=10\%$ and a Scalo
3127: IMF. Shown is the unabsorbed GRB afterglow (dot-short dashed curve), the
3128: afterglow with resonant IGM absorption only (dot-long dashed curve), and
3129: the afterglow with full (resonant and damping wing) IGM absorption (solid
3130: curve). Also shown, with 1.7 magnitudes of extinction, are the afterglow
3131: with full IGM absorption (dotted curve), and attempts to reproduce this
3132: profile with a damped Ly$\alpha$ absorption system in the host galaxy
3133: (dashed curves).  (Note, however, that damped absorption of this type could
3134: be suppressed by the ionizing effect of the afterglow UV radiation on the
3135: surrounding interstellar medium of its host galaxy\cite{PL98}.) Most
3136: importantly, the overall spectral shape of the Ly$\alpha$ trough carries
3137: precious information about the neutral fraction of the IGM at the source
3138: redshift; averaging over an ensemble of sources with similar redshifts can
3139: reduce ambiguities in the interpretation of each case due to particular
3140: local effects.}
3141: \label{fig2val}
3142: \end{figure}
3143: 
3144: Although the nature of the central engine that powers the relativistic jets
3145: of GRBs is still unknown, recent evidence indicates that long-duration GRBs
3146: trace the formation of massive stars (e.g.,
3147: \cite{T97,Wij98,BN00,Kul00,BKD02,Nat05}) and in particular that long-duration
3148: GRBs are associated with Type Ib/c supernovae \cite{Sta03}. Since the first
3149: stars in the Universe are predicted to be predominantly massive
3150: \cite{ABN02,BCL02,Bromm}, their death might give rise to large numbers of
3151: GRBs at high redshifts.  In contrast to quasars of comparable brightness,
3152: GRB afterglows are short-lived and release $\sim 10$ orders of magnitude
3153: less energy into the surrounding IGM. Beyond the scale of their host
3154: galaxy, they have a negligible effect on their cosmological
3155: environment\footnote{Note, however, that feedback from a single GRB or
3156: supernova on the gas confined within early dwarf galaxies could be
3157: dramatic, since the binding energy of most galaxies at $z>10$ is lower than
3158: $10^{51}~{\rm ergs}$ \cite{BL01}.}. Consequently, they are ideal probes
3159: of the IGM during the reionization epoch.  Their rest-frame UV spectra can
3160: be used to probe the ionization state of the IGM through the spectral shape
3161: of the Gunn-Peterson (Ly$\alpha$) absorption trough, or its metal
3162: enrichment history through the intersection of enriched bubbles of
3163: supernova (SN) ejecta from early galaxies \cite{FL03}.  Afterglows that are
3164: unusually bright ($>10$mJy) at radio frequencies should also show a
3165: detectable forest of 21~cm absorption lines due to enhanced HI column
3166: densities in sheets, filaments, and collapsed minihalos within the IGM
3167: \cite{Carilli04,FL02}.
3168: 
3169: Another advantage of GRB afterglows is that once they fade away, one may
3170: search for their host galaxies. Hence, GRBs may serve as signposts of the
3171: earliest dwarf galaxies that are otherwise too faint or rare on their own
3172: for a dedicated search to find them. Detection of metal absorption lines
3173: from the host galaxy in the afterglow spectrum, offers an unusual
3174: opportunity to study the physical conditions (temperature, metallicity,
3175: ionization state, and kinematics) in the interstellar medium of these
3176: high-redshift galaxies.  As Figure \ref{fig2val} indicates, damped
3177: Ly$\alpha$ absorption within the host galaxy may mask the clear signature
3178: of the Gunn-Peterson trough in some galaxies \cite{BL04}.  A small
3179: fraction ($\sim 10$) of the GRB afterglows are expected to originate at
3180: redshifts $z>5$ \cite{BL02,BL06}.  This subset of afterglows can be
3181: selected photometrically using a small telescope, based on the Ly$\alpha$
3182: break at a wavelength of $1.216\, \mu {\rm m}\, [(1+z)/10]$, caused by
3183: intergalactic HI absorption.  The challenge in the upcoming years will be
3184: to follow-up on these candidates spectroscopically, using a large (10-meter
3185: class) telescope.  GRB afterglows are likely to revolutionize observational
3186: cosmology and replace traditional sources like quasars, as probes of the
3187: IGM at $z>5$.  The near future promises to be exciting for GRB astronomy as
3188: well as for studies of the high-redshift Universe.
3189: 
3190: It is of great importance to constrain the Pop~III star formation mode, and
3191: in particular to determine down to which redshift it continues to be
3192: prominent. The extent of the Pop~III star formation will affect models of
3193: the initial stages of reionization (e.g.,
3194: \cite{WL03,CFW03,Sok04,YBH04,ABS06}) and metal enrichment (e.g.,
3195: \cite{MBH03,FL03,FL05,Sch03,SSR04}), and will determine whether planned
3196: surveys will be able to effectively probe Pop~III stars (e.g.,
3197: \cite{Sca05}).  The constraints on Pop~III star formation will also
3198: determine whether the first stars could have contributed a significant
3199: fraction to the cosmic near-IR background (e.g.,
3200: \cite{SBK02,SF03,Kas05,MS05,DAK05}).  To constrain high-redshift star
3201: formation from GRB observations, one has to address two
3202: major questions:
3203: 
3204: \noindent {\it (1)} {\it What is the signature of GRBs that originate in
3205: metal-free, Pop~III progenitors?} Simply knowing that a given GRB came from
3206: a high redshift is not sufficient to reach a definite conclusion as to the
3207: nature of the progenitor. Pregalactic metal enrichment was likely
3208: inhomogeneous, and we expect normal Pop~I and II stars to exist in galaxies
3209: that were already metal-enriched at these high redshifts
3210: \cite{BL06}. Pop~III and Pop~I/II star formation is thus predicted to have
3211: occurred concurrently at $z > 5$. How is the predicted high mass-scale for
3212: Pop~III stars reflected in the observational signature of the resulting
3213: GRBs? Preliminary results from numerical simulations of Pop III star
3214: formation indicate that circumburst densities are systematically higher in
3215: Pop~III environments. GRB afterglows will then be much brighter than for
3216: conventional GRBs. In addition, due to the systematically increased
3217: progenitor masses, the Pop~III distribution may be biased toward
3218: long-duration events.
3219: 
3220: \noindent {\it (2)}
3221: The modelling of Pop~III cosmic star formation histories has a number of
3222: free parameters, such as the star formation efficiency and the strength of
3223: the chemical feedback. The latter refers to the timescale for, and spatial
3224: extent of, the distribution of the first heavy elements that were produced
3225: inside of Pop~III stars, and subsequently dispersed into the IGM by
3226: supernova blast waves. Comparing with theoretical GRB redshift
3227: distributions one can use the GRB redshift
3228: distribution observed by {\it Swift} to calibrate the free model
3229: parameters. In particular, one can use this strategy to measure the
3230: redshift where Pop~III star formation terminates.
3231: 
3232: \begin{figure}[tp!]
3233: \includegraphics[height=.4\textheight]{fig1b.ps}
3234: \caption{Theoretical prediction for the comoving star formation rate
3235: (SFR) in units of $M_{\odot}$~yr$^{-1}$~Mpc$^{-3}$, as a function of
3236: redshift (from \cite{BL06}).  We assume that cooling in primordial gas is
3237: due to atomic hydrogen only, a star formation efficiency of
3238: $\eta_\ast=10\%$, and reionization beginning at $z_{\rm reion}\approx 17$.
3239: {\it Solid line:} Total comoving SFR.  {\it Dotted lines:} Contribution to
3240: the total SFR from Pop~I/II and Pop~III for the case of weak chemical
3241: feedback.  {\it Dashed lines:} Contribution to the total SFR from Pop~I/II
3242: and Pop~III for the case of strong chemical feedback.  Pop~III star
3243: formation is restricted to high redshifts, but extends over a significant
3244: range, $\Delta z\sim 10-15$.
3245: \label{fig1bl}}
3246: \end{figure}
3247: 
3248: Figures~\ref{fig1bl} and \ref{fig2bl} illustrate these issues (based on
3249: \cite{BL06}).  Figure \ref{fig2bl} leads to the robust expectation that
3250: $\sim 10$\% of all {\it Swift} bursts should originate at $z > 5$. This
3251: prediction is based on the contribution from Population~I/II stars which
3252: are known to exist even at these high redshifts. Additional GRBs could be
3253: triggered by Pop~III stars, with a highly uncertain efficiency. Assuming
3254: that long-duration GRBs are produced by the collapsar mechanism, a Pop~III
3255: star with a close binary companion provides a plausible GRB progenitor.
3256: The Pop~III GRB efficiency, reflecting the probability of forming
3257: sufficiently close and massive binary systems, to lie between zero (if
3258: tight Pop~III binaries do not exist) and $\sim 10$ times the empirically
3259: inferred value for Population~I/II (due to the increased fraction of black
3260: hole forming progenitors among the massive Pop~III stars).
3261: 
3262: A key ingredient in determining the underlying star formation history from
3263: the observed GRB redshift distribution is the GRB luminosity function,
3264: which is only poorly constrained at present.  The improved statistics
3265: provided by {\it Swift} will enable the construction of an empirical
3266: luminosity function. With an improved luminosity function it would be
3267: possible to re-calibrate the theoretical prediction in Figure~2 more
3268: reliably.
3269: 
3270: 
3271: \begin{figure}[t]
3272: \includegraphics[height=.4\textheight]{fig2b.ps}
3273: \caption{Predicted GRB rate to be observed by {\it Swift} (from \cite{BL06}).
3274: Shown is the observed
3275: number of bursts per year, $dN_{\rm GRB}^{\rm obs}/d\ln (1+z)$, as a function
3276: of redshift.  All rates are calculated with a constant GRB efficiency,
3277: $\eta_{\rm GRB}\simeq 2\times 10^{-9}$~bursts $M_{\odot}^{-1}$, using the
3278: cosmic SFRs from Fig.~\ref{fig1bl}.
3279: {\it Dotted lines:} Contribution to the observed GRB
3280: rate from Pop~I/II and Pop~III for the case of weak chemical feedback.
3281: {\it Dashed lines:} Contribution to the GRB rate from Pop~I/II and Pop~III
3282: for the case of strong chemical feedback. 
3283: {\it Filled circle:} GRB rate from Pop~III stars if these were
3284: responsible for reionizing the Universe at $z\sim 17$.  
3285: \label{fig2bl}
3286: }
3287: \end{figure}
3288: 
3289: In order to predict the observational signature of high-redshift GRBs, it
3290: is important to know the properties of the GRB host systems.  Within
3291: variants of the popular CDM model for structure formation, where small
3292: objects form first and subsequently merge to build up more massive ones,
3293: the first stars are predicted to form at $z\sim 20$--$30$ in minihalos of
3294: total mass (dark matter plus gas) $\sim 10^6 M_{\odot}$
3295: \cite{Teg97,BL01,YBH04}.  These objects are the sites for the formation
3296: of the first stars, and thus are the potential hosts of the
3297: highest-redshift GRBs.  {\it What is the environment in which the earliest
3298: GRBs and their afterglows did occur?}  This problem breaks down into two
3299: related questions: (i) what type of stars (in terms of mass, metallicity,
3300: and clustering properties) will form in each minihalo?, and (ii) how will
3301: the ionizing radiation from each star modify the density structure of the
3302: surrounding gas? These two questions are fundamentally intertwined. The
3303: ionizing photon production strongly depends on the stellar mass, which in
3304: turn is determined by how the accretion flow onto the growing protostar
3305: proceeds under the influence of this radiation field. In other words, the
3306: assembly of the Population~III stars and the development of an HII region
3307: around them proceed simultaneously, and affect each other.  As a
3308: preliminary illustration, Figure~\ref{fig3ab} describes the photo-evaporation
3309: as a self-similar champagne flow \cite{Shu02} with parameters appropriate
3310: for the Pop~III case.
3311: 
3312: \begin{figure}[t]
3313: \includegraphics[height=.4\textheight]{fig3b.ps}
3314: \caption{Effect of photoheating from a Population~III star on the density
3315: profile in a high-redshift minihalo (from \cite{BrL06}).  The curves,
3316: labeled by the time after the onset of the central point source, are
3317: calculated according to a self-similar model for the expansion of an HII
3318: region. Numerical simulations closely conform to this analytical
3319: behavior. Notice that the central density is significantly reduced by the
3320: end of the life of a massive star, and that a central core has developed
3321: where the density is constant.}
3322: \label{fig3b}
3323: \end{figure}
3324: 
3325: Notice that the central density is significantly reduced by the end of the
3326: life of a massive star, and that a central core has developed where the
3327: density is nearly constant. Such a flat density profile is markedly
3328: different from that created by stellar winds ($\rho \propto
3329: r^{-2}$). Winds, and consequently mass-loss, may not be important for
3330: massive Population~III stars \cite{BHW01,K02},
3331: and such a flat density profile may be
3332: characteristic of GRBs that originate from metal-free Population~III
3333: progenitors.
3334: 
3335: \begin{figure}[t]
3336: \includegraphics[height=.3\textheight]{fig4b.eps}
3337: \caption{Supernova explosion in the high-redshift Universe (from \cite{BYH03}).
3338: The snapshot is taken $\sim 10^6$~yr after the explosion with
3339: total energy $E_{\rm SN}\simeq 10^{53}$~ergs. We show the projected gas
3340: density within a box of linear size 1~kpc. The SN bubble has expanded to a
3341: radius of $\sim 200$~pc, having evacuated most of the gas in the
3342: minihalo. {\it Inset:} Distribution of metals. The stellar ejecta ({\it gray
3343: dots}) trace the metals and are embedded in pristine metal-poor gas ({\it
3344: black dots}).  }
3345: \label{ffig4b}
3346: \end{figure}
3347: 
3348: The first galaxies may be surrounded by a shell of highly enriched material
3349: that was carried out in a SN-driven wind (see Fig.~\ref{ffig4b}). A GRB in
3350: that galaxy may show strong absorption lines at a velocity separation
3351: associated with the wind velocity.  Simulating these winds and calculating
3352: the absorption profile in the featureless spectrum of a GRB afterglow, will
3353: allow us to use the observed spectra of high-$z$ GRBs and directly probe the
3354: degree of metal enrichment in the vicinity of the first star forming
3355: regions (see \cite{FL03} for a semi-analytic treatment).
3356: 
3357: As the early afterglow radiation propagates through the interstellar
3358: environment of the GRB, it will likely modify the gas properties close to
3359: the source; these changes could in turn be noticed as time-dependent
3360: spectral features in the spectrum of the afterglow and used to derive the
3361: properties of the gas cloud (density, metal abundance, and size).  The UV
3362: afterglow radiation can induce detectable changes to the interstellar
3363: absorption features of the host galaxy \cite{PL98}; dust destruction could
3364: have occurred due to the GRB X-rays \cite{WD00,FKR01}, and molecules could
3365: have been destroyed near the GRB source \cite{DH02}.  Quantitatively, all
3366: of the effects mentioned above strongly depend on the exact properties of
3367: the gas in the host system.
3368: 
3369: Most studies to date have assumed a constant efficiency of forming GRBs per
3370: unit mass of stars. This simplifying assumption could lead, under different
3371: circumstances, to an overestimation or an underestimation of the frequency
3372: of GRBs. Metal-free stars are thought to be massive \cite{ABN02,BCL02} and
3373: their extended envelopes may suppress the emergence of relativistic jets
3374: out of their surface (even if such jets are produced through the collapse
3375: of their core to a spinning black hole). On the other hand, low-metallicity
3376: stars are expected to have weak winds with little angular momentum loss
3377: during their evolution, and so they may preferentially yield rotating
3378: central configurations that make GRB jets after core collapse.
3379: 
3380: {\it What kind of metal-free, Pop~III progenitor stars may lead to GRBs?}
3381: Long-duration GRBs appear to be associated with Type Ib/c supernovae
3382: \cite{Sta03}, namely progenitor massive stars that have lost their hydrogen
3383: envelope. This requirement is explained theoretically in the collapsar
3384: model, in which the relativistic jets produced by core collapse to a black
3385: hole are unable to emerge relativistically out of the stellar surface if
3386: the hydrogen envelope is retained \cite{MWH01}.  The question then arises
3387: as to whether the lack of metal line-opacity that is essential for
3388: radiation-driven winds in metal-rich stars, would make a Pop~III star
3389: retain its hydrogen envelope, thus quenching any relativistic jets and
3390: GRBs.
3391: 
3392: Aside from mass transfer in a binary system, individual Pop~III stars could
3393: lose their hydrogen envelope due to either: (i) violent pulsations,
3394: particularly in the mass range $100$--$140 M_{\odot}$, or (ii) a wind
3395: driven by helium lines.  The outer stellar layers are in a state where
3396: gravity only marginally exceeds radiation pressure due to
3397: electron-scattering (Thomson) opacity. Adding the small, but still
3398: non-negligible contribution from the bound-free opacity provided by
3399: singly-ionized helium, may be able to unbind the atmospheric
3400: gas. Therefore, mass-loss might occur even in the absence of dust or any
3401: heavy elements.
3402: 
3403: %<>
3404: \subsection{Emission Spectrum of Metal-Free Stars}
3405: \label{sec4.1.2}
3406: 
3407: The evolution of metal-free (Population III) stars is qualitatively
3408: different from that of enriched (Population I and II) stars. In the
3409: absence of the catalysts necessary for the operation of the CNO cycle,
3410: nuclear burning does not proceed in the standard way. At first,
3411: hydrogen burning can only occur via the inefficient PP chain. To
3412: provide the necessary luminosity, the star has to reach very high
3413: central temperatures ($T_{c}\simeq 10^{8.1}$ K). These temperatures
3414: are high enough for the spontaneous turn-on of helium burning via the
3415: triple-$\alpha$ process. After a brief initial period of
3416: triple-$\alpha$ burning, a trace amount of heavy elements
3417: forms. Subsequently, the star follows the CNO cycle. In constructing
3418: main-sequence models, it is customary to assume that a trace mass
3419: fraction of metals ($Z\sim 10^{-9}$) is already present in the star
3420: (El Eid 1983 \cite{EFO83}; Castellani et al.\ 1983 \cite{CCT83}).
3421: 
3422: Figures \ref{fig4e} and \ref{fig4f} show the luminosity $L$ vs.\
3423: effective temperature $T$ for zero-age main sequence stars in the mass
3424: ranges of $2$--$90M_\odot$ (Fig. \ref{fig4e}) and $100$--$1000
3425: M_\odot$ (Fig. \ref{fig4f}). Note that above $\sim 100M_\odot$ the
3426: effective temperature is roughly constant, $T_{\rm eff}\sim 10^5$K,
3427: implying that the spectrum is independent of the mass distribution of
3428: the stars in this regime (Bromm, Kudritzky, \& Loeb 2001 \cite{BKL2001}). As is
3429: evident from these figures (see also Tumlinson \& Shull 2000 \cite{TS00}), both
3430: the effective temperature and the ionizing power of metal-free (Pop
3431: III) stars are substantially larger than those of metal-rich (Pop I)
3432: stars.  Metal-free stars with masses $\ga 20M_\odot$ emit between
3433: $10^{47}$ and $10^{48}$ H I and He I ionizing photons
3434: per second per solar mass of stars, where the lower value applies to
3435: stars of $\sim 20M_\odot$ and the upper value applies to stars of $\ga
3436: 100 M_{\odot}$ (see Tumlinson \& Shull 2000 \cite{TS00} and Bromm et al.\ 2001 \cite{BKL2001} for
3437: more details). Over a lifetime of $\sim 3\times 10^6$ years these
3438: massive stars produce $10^4$--$10^5$ ionizing photons per stellar
3439: baryon. However, this powerful UV emission is suppressed as soon as
3440: the interstellar medium out of which new stars form is enriched by
3441: trace amounts of metals. Even though the collapsed fraction of baryons
3442: is small at the epoch of reionization, it is likely that most of the
3443: stars responsible for the reionization of the Universe formed out of
3444: enriched gas.
3445: 
3446: \noindent
3447: \begin{figure} 
3448: \centering
3449: \includegraphics[height=6cm]{IV_fig5.eps}
3450: \caption{Luminosity vs.\ effective temperature for zero-age main
3451: sequences stars in the mass range of $2$--$90M_\odot$ (from Tumlinson
3452: \& Shull 2000 \cite{TS00}). The curves show Pop I ($Z_{\odot}$ = 0.02) and Pop III
3453: stars of mass 2, 5, 8, 10, 15, 20, 25, 30, 35, 40, 50, 60, 70, 80, and
3454: 90 \Msun. The diamonds mark decades in metallicity in the approach to
3455: $Z = 0$ from 10$^{-2}$ down to 10$^{-5}$ at 2 \Msun, down to
3456: 10$^{-10}$ at 15 \Msun, and down to 10$^{-13}$ at 90 \Msun. The
3457: dashed line along the Pop III ZAMS assumes pure H-He composition,
3458: while the solid line (on the left) marks the upper MS with $Z_{\rm C}
3459: = 10^{-10}$ for the $M \geq 15$ \Msun\ models. Squares mark the points
3460: corresponding to pre-enriched evolutionary models from El Eid et al.\
3461: (1983) \cite{EFO83} at 80 \Msun\ and from Castellani et al.\ (1983) \cite{CCT83} for 25 \Msun.  }
3462: \label{fig4e}
3463: \end{figure}
3464:   
3465: \noindent
3466: \begin{figure}
3467: \centering
3468: \includegraphics[height=6cm]{IV_fig6.eps}
3469: \caption{Same as Figure \ref{fig4e} but for very massive stars above
3470: $100M_\odot$ (from Bromm, Kudritzki, \& Loeb 2001 \cite{BKL2001}).  {\it Left solid
3471: line:} Pop III zero-age main sequence (ZAMS).  {\it Right solid line:}
3472: Pop I ZAMS. In each case, stellar luminosity (in $L_{\odot}$) is
3473: plotted vs.\ effective temperature (in K).  {\it Diamond-shaped
3474: symbols:} Stellar masses along the sequence, from $100 M_{\odot}$
3475: (bottom) to $1000 M_{\odot}$ (top) in increments of $100 M_{\odot}$.
3476: The Pop III ZAMS is systematically shifted to higher effective
3477: temperature, with a value of $\sim 10^{5}$ K which is approximately
3478: independent of mass. The luminosities, on the other hand, are almost
3479: identical in the two cases.  }
3480: \label{fig4f}
3481: \end{figure}
3482:   
3483: {\it Will it be possible to infer the initial mass function (IMF) of
3484: the first stars from spectroscopic observations of the first
3485: galaxies?}  Figure \ref{fig4g} compares the observed spectrum from a
3486: Salpeter IMF ($dN_\star/dM\propto M^{-2.35}$) and a heavy IMF (with
3487: all stars more massive than $100M_\odot$) for a galaxy at
3488: $z_s=10$. The latter case follows from the assumption that each of the
3489: dense clumps in the simulations described in the previous section ends
3490: up as a single star with no significant fragmentation or mass
3491: loss. The difference between the plotted spectra cannot be confused
3492: with simple reddening due to normal dust. Another distinguishing
3493: feature of the IMF is the expected flux in the hydrogen and helium
3494: recombination lines, such as Ly$\alpha$ and He II 1640 \AA, from the
3495: interstellar medium surrounding these stars. We discuss this next.
3496: 
3497: \noindent
3498: \begin{figure} 
3499: \centering
3500: \includegraphics[height=6cm]{IV_fig7.eps}
3501: \caption{ Comparison of the predicted flux from a Pop III star cluster
3502: at $z_{s}=10$ for a Salpeter IMF (Tumlinson \& Shull 2000 \cite{TS00}) and a
3503: massive IMF (Bromm et al.\ 2001 \cite{BKL2001}).  Plotted is the observed flux (in
3504: $\mbox{nJy}$ per $10^{6}M_{\odot}$ of stars) vs.\ observed wavelength
3505: (in $\mu$m) for a flat Universe with $\Omega_{\Lambda}=0.7$ and
3506: $h=0.65$.  {\it Solid line:} The case of a heavy IMF.  {\it Dotted
3507: line:} The fiducial case of a standard Salpeter IMF.  The cutoff below
3508: $\lambda_{obs} = 1216\mbox{\,\AA \,} (1+z_{s})=1.34\mu$m is due to
3509: complete Gunn-Peterson absorption (which is artificially assumed to be
3510: sharp). Clearly, for the same total stellar mass, the observable flux
3511: is larger by an order of magnitude for stars which are biased towards
3512: having masses $\ga 100M_\odot\,$.  }
3513: \label{fig4g}
3514: \end{figure}
3515:   
3516: %<>
3517: \subsection{Emission of Recombination Lines from the First Galaxies}
3518: \label{sec4.1.3}
3519: 
3520: The hard UV emission from a star cluster or a quasar at high redshift
3521: is likely reprocessed by the surrounding interstellar medium,
3522: producing very strong recombination lines of hydrogen and helium (Oh
3523: 1999 \cite{Oh99}; Tumlinson \& Shull 2000 \cite{TS00}; see also Baltz, Gnedin \& Silk
3524: 1998 \cite{Baltz98}). We define $\dot{N}_{\rm ion}$ to be the production rate of
3525: ionizing photons by the source. The emitted luminosity $L_{\rm
3526: line}^{\rm em}$ per unit stellar mass in a particular recombination
3527: line is then estimated to be
3528: \begin{equation}
3529: L_{\rm line}^{\rm em} = p_{\rm line}^{\rm em} h\nu \dot{N}_{\rm ion}
3530: (1 - p^{\rm esc}_{\rm cont}) p^{\rm esc}_{\rm line} \mbox{\ \ \ ,}
3531: \end{equation}
3532: where $p_{\rm line}^{\rm em}$ is the probability that a recombination leads
3533: to the emission of a photon in the corresponding line, $\nu$ is the
3534: frequency of the line and $p^{\rm esc}_{\rm cont}$ and $p^{\rm esc}_{\rm
3535: line}$ are the escape probabilities for the ionizing photons and the line
3536: photons, respectively. It is natural to assume that the stellar cluster is
3537: surrounded by a finite H II region, and hence that $p^{\rm esc}_{\rm cont}$
3538: is close to zero \cite{WL00,RS00}.  In addition, $p^{\rm esc}_{\rm line}$
3539: is likely close to unity in the H II region, due to the lack of dust in the
3540: ambient metal-free gas. Although the emitted line photons may be scattered
3541: by neutral gas, they diffuse out to the observer and in the end survive if
3542: the gas is dust free. Thus, for simplicity, we adopt a value of unity for
3543: $p^{\rm esc}_{\rm line}$.
3544: 
3545: As a particular example we consider case B recombination which yields
3546: $p_{\rm line}^{\rm em}$ of about 0.65 and 0.47 for the Ly${\alpha}$ and He
3547: II 1640\,\AA \,lines, respectively. These numbers correspond to an electron
3548: temperature of $\sim 3\times 10^4$K and an electron density of $\sim
3549: 10^{2}-10^{3}$ cm$^{-3}$ inside the H II region \cite{SH95}. For example,
3550: we consider the extreme and most favorable case of metal-free stars all of
3551: which are more massive than $\sim 100M_\odot$. In this case $L_{\rm
3552: line}^{\rm em} = 1.7\times 10^{37}$ and $2.2\times 10^{36}$ erg
3553: s$^{-1}M_{\odot}^{-1}$ for the recombination luminosities of Ly$\alpha$ and
3554: He II 1640\,\AA\,per stellar mass \cite{BKL2001}. A cluster of $10^{6}
3555: M_{\odot}$ in such stars would then produce 4.4 and 0.6 $\times
3556: 10^{9}L_{\odot}$ in the Ly$\alpha$ and He II
3557: 1640\,\AA\,lines. Comparably-high luminosities would be produced in other
3558: recombination lines at longer wavelengths, such as He II 4686\,\AA\,and
3559: H$\alpha$ \cite{Oh99,OHR00}.
3560: 
3561: The rest--frame equivalent width of the above emission lines measured
3562: against the stellar continuum of the embedded star cluster at the line
3563: wavelengths is given by
3564: \begin{equation}
3565: W_{\lambda} =\left(\frac{L_{\rm line}^{\rm em}}{L_{\lambda}}\right)
3566: \mbox{\ \ \ ,}
3567: \end{equation}
3568: where $L_{\lambda}$ is the spectral luminosity per unit wavelength of
3569: the stars at the line resonance.  The extreme case of metal-free stars
3570: which are more massive than $100M_\odot$ yields a spectral luminosity
3571: per unit frequency $L_{\nu} = 2.7\times 10^{21}$ and $1.8\times
3572: 10^{21}$ erg s$^{-1}$ Hz$^{-1}M_{\odot}^{-1}$ at the corresponding
3573: wavelengths \cite{BKL2001}.  Converting to $L_{\lambda}$, this
3574: yields rest-frame equivalent widths of $W_{\lambda}$ = 3100\,\AA\,and
3575: 1100\,\AA\,for Ly$\alpha$ and He II  1640\,\AA\,,
3576: respectively. These extreme emission equivalent widths are more than
3577: an order of magnitude larger than the expectation for a normal cluster
3578: of hot metal-free stars with the same total mass and a Salpeter IMF
3579: under the same assumptions concerning the escape probabilities and
3580: recombination \cite{Kud00}. The equivalent widths are, of
3581: course, larger by a factor of $(1+z_{s})$ in the observer frame.
3582: Extremely strong recombination lines, such as Ly$\alpha$ and
3583: He II  1640\,\AA, are therefore expected to be an additional
3584: spectral signature that is unique to very massive stars in the early
3585: Universe. The strong recombination lines from the first luminous
3586: objects are potentially detectable with \NGST \cite{OHR00}.
3587: 
3588: 
3589: \section{Supermassive Black holes}
3590: 
3591: \subsection{The Principle of Self-Regulation}
3592: 
3593: The fossil record in the present-day Universe indicates that every bulged
3594: galaxy hosts a supermassive black hole (BH) at its center
3595: \cite{Kor03}. This conclusion is derived from a variety of techniques which
3596: probe the dynamics of stars and gas in galactic nuclei.  The inferred BHs
3597: are dormant or faint most of the time, but ocassionally flash in a short
3598: burst of radiation that lasts for a small fraction of the Hubble time. The
3599: short duty cycle acounts for the fact that bright quasars are much less
3600: abundant than their host galaxies, but it begs the more fundamental
3601: question: {\it why is the quasar activity so brief?}  A natural explanation
3602: is that quasars are suicidal, namely the energy output from the BHs
3603: regulates their own growth.
3604: 
3605: Supermassive BHs make up a small fraction, $< 10^{-3}$, of the total mass
3606: in their host galaxies, and so their direct dynamical impact is limited to
3607: the central star distribution where their gravitational influence
3608: dominates. Dynamical friction on the background stars keeps the BH close to
3609: the center. Random fluctuations in the distribution of stars induces a
3610: Brownian motion of the BH. This motion can be decribed by the same Langevin
3611: equation that captures the motion of a massive dust particle as it responds
3612: to random kicks from the much lighter molecules of air around it
3613: \cite{Cha02}.  The characteristic speed by which the BH wanders around the
3614: center is small, $\sim (m_\star/M_{\rm BH})^{1/2}\sigma_\star$, where
3615: $m_\star$ and $M_{\rm BH}$ are the masses of a single star and the BH,
3616: respectively, and $\sigma_\star$ is the stellar velocity dispersion. Since
3617: the random force fluctuates on a dynamical time, the BH wanders across a
3618: region that is smaller by a factor of $\sim (m_\star/M_{\rm BH})^{1/2}$
3619: than the region traversed by the stars inducing the fluctuating force on
3620: it.
3621: 
3622: The dynamical insignificance of the BH on the global galactic scale is
3623: misleading. The gravitational binding energy per rest-mass energy of
3624: galaxies is of order $\sim (\sigma_\star/c)^2< 10^{-6}$.  Since BH are
3625: relativistic objects, the gravitational binding energy of material that
3626: feeds them amounts to a substantial fraction its rest mass energy. Even if
3627: the BH mass occupies a fraction as small as $\sim 10^{-4}$ of the baryonic
3628: mass in a galaxy, and only a percent of the accreted rest-mass energy leaks
3629: into the gaseous environment of the BH, this slight leakage can unbind the
3630: entire gas reservoir of the host galaxy! This order-of-magnitude estimate
3631: explains why quasars are short lived.  As soon as the central BH accretes
3632: large quantities of gas so as to significantly increase its mass, it
3633: releases large amounts of energy that would suppress further accretion onto
3634: it. In short, the BH growth is {\it self-regulated}.
3635: 
3636: The principle of {\it self-regulation} naturally leads to a correlation
3637: between the final BH mass, $M_{\rm bh}$, and the depth of the gravitational
3638: potential well to which the surrounding gas is confined which can be
3639: characterized by the velocity dispersion of the associated stars, $\sim
3640: \sigma_\star^2$. Indeed such a correlation is observed in the present-day
3641: Universe \cite{Tre02}. The observed power-law relation between $M_{\rm bh}$
3642: and $\sigma_\star$ can be generalized to a correlation between the BH mass
3643: and the circular velocity of the host halo, $v_c$ \cite{Fer02}, which in
3644: turn can be related to the halo mass, $M_{\rm halo}$, and redshift, $z$
3645: \cite{WL03}
3646: \begin{eqnarray}
3647: \label{eq:1}
3648: \nonumber M_{\rm bh}(M_{\rm halo},z) &=&\mbox{const} \times v_c^5\\
3649: &&\hspace{-25mm}= \epsilon_{\rm o} M_{\rm halo} \left(\frac{M_{\rm
3650: halo}}{10^{12}M_{\odot}}\right)^{\frac{2}{3}}
3651: [\zeta(z)]^\frac{5}{6}(1+z)^\frac{5}{2},
3652: \end{eqnarray}
3653: where $\epsilon_{\rm o}\approx 10^{-5.7}$ is a constant, and as before
3654: $\zeta\equiv [(\Omega_m/\Omega_m^z)(\Delta_c/18\pi^2)]$, $\Omega_m^z \equiv
3655: [1+(\Omega_\Lambda/\Omega_m)(1+z)^{-3}]^{-1}$,
3656: $\Delta_c=18\pi^2+82d-39d^2$, and $d=\Omega_m^z-1$.  If quasars shine near
3657: their Eddington limit as suggested by observations of low and high-redshift
3658: quasars \cite{Flo03,Wil03}, then the above value of $\epsilon_{\rm o}$
3659: implies that a fraction of $\sim 5$--$10\%$ of the energy released by the
3660: quasar over a galactic dynamical time needs to be captured in the
3661: surrounding galactic gas in order for the BH growth to be self-regulated
3662: \cite{WL03}.
3663: 
3664: \begin{figure}
3665: \centering
3666: \includegraphics[height=6cm]{loebF1.eps}
3667: \caption{Comparison of the observed and model luminosity functions (from
3668: \cite{WL03}). The data points at $z<4$ are summarized in Ref. \cite{Pei95},
3669: while the light lines show the double power-law fit to the {\it 2dF} quasar
3670: luminosity function \cite{Boy00}.  At $z\sim4.3$ and $z\sim6.0$ the data is
3671: from Refs. \cite{Fan01a}. The grey regions show the 1-$\sigma$ range of
3672: logarithmic slope ($[-2.25,-3.75]$ at $z\sim4.3$ and $[-1.6,-3.1]$ at
3673: $z\sim6$), and the vertical bars show the uncertainty in the
3674: normalization. The open circles show data points converted from the X-ray
3675: luminosity function \cite{Bar03} of low luminosity quasars using the median
3676: quasar spectral energy distribution.  In each panel the vertical dashed
3677: lines correspond to the Eddington luminosities of BHs bracketing the
3678: observed range of the $M_{\rm bh}$--$v_{\rm c}$ relation, and the vertical
3679: dotted line corresponds to a BH in a $10^{13.5}M_\odot$ galaxy.}
3680: \label{fig1}
3681: \end{figure}
3682: 
3683: With this interpretation, the $M_{\rm bh}$--$\sigma_\star$ relation
3684: reflects the limit introduced to the BH mass by self-regulation; deviations
3685: from this relation are inevitable during episodes of BH growth or as a
3686: result of mergers of galaxies that have no cold gas in them.  A physical
3687: scatter around this upper envelope could also result from variations in the
3688: efficiency by which the released BH energy couples to the surrounding gas.
3689: 
3690: Various prescriptions for self-regulation were sketched by Silk \& Rees
3691: \cite{Sil98}. These involve either energy or momentum-driven winds, where
3692: the latter type is a factor of $\sim v_c/c$ ~less efficient
3693: \cite{Beg04,Kin03,Mur04}. Wyithe \& Loeb \cite{WL03} demonstrated that a
3694: particularly simple prescription for an energy-driven wind can reproduce
3695: the luminosity function of quasars out to highest measured redshift, $z\sim
3696: 6$ (see Figs. \ref{fig1} and \ref{fig2}), as well as the observed
3697: clustering properties of quasars at $z\sim 3$ \cite{WLcl} (see
3698: Fig. \ref{fig3}). The prescription postulates that: {\it (i)}
3699: self-regulation leads to the growth of $M_{\rm bh}$ up the
3700: redshift-independent limit as a function of $v_c$ in Eq. (\ref{eq:1}), for
3701: all galaxies throughout their evolution; and {\it (ii)} the growth of
3702: $M_{\rm bh}$ to the limiting mass in Eq. (\ref{eq:1}) occurs through halo
3703: merger episodes during which the BH shines at its Eddington luminosity
3704: (with the median quasar spectrum) over the dynamical time of its host
3705: galaxy, $t_{\rm dyn}$.  This model has only one adjustable parameter,
3706: namely the fraction of the released quasar energy that couples to the
3707: surrounding gas in the host galaxy. This parameter can be fixed based on
3708: the $M_{\rm bh}$--$\sigma_\star$ relation in the local Universe
3709: \cite{Fer02}.  It is remarkable that the combination of the above simple
3710: prescription and the standard $\Lambda$CDM cosmology for the evolution and
3711: merger rate of galaxy halos, lead to a satisfactory agreement with the rich
3712: data set on quasar evolution over cosmic history. 
3713: 
3714: The cooling time of the heated gas is typically longer than its dynamical
3715: time and so the gas should expand into the galactic halo and escape the
3716: galaxy if its initial temperature exceeds the virial temperature of the
3717: galaxy \cite{WL03}. The quasar remains active during the dynamical time of
3718: the initial gas reservoir, $\sim 10^7$ years, and fades afterwards due to
3719: the dilution of this reservoir.  Accretion is halted as soon as the quasar
3720: supplies the galactic gas with more than its binding energy. The BH growth
3721: may resume if the cold gas reservoir is replenished through a new merger.
3722: 
3723: Following the early analytic work, extensive numerical simulations by
3724: Springel, Hernquist, \& Di Matteo (2005) \cite{SDH05} (see also Di Matteo et al. 2005 \cite{DSH05})
3725: demonstrated that galaxy mergers do produce the observed correlations
3726: between black hole mass and spheroid properties when a similar energy
3727: feedback is incorporated. Because of the limited resolution near the galaxy
3728: nucleus, these simulations adopt a simple prescription for the accretion
3729: flow that feeds the black hole. The actual feedback in reality may depend
3730: crucially on the geometry of this flow and the physical mechanism that
3731: couples the energy or momentum output of the quasar to the surrounding gas.
3732: 
3733: \begin{figure}
3734: \centering
3735: \includegraphics[height=4.77cm]{ti.eps}
3736: \caption{Simulation images of a merger of galaxies resulting in quasar
3737: activity that eventually shuts-off the accretion of gas onto the black hole
3738: (from Di Matteo et al. 2005 \cite{DSH05}). The upper (lower) panels show a sequence of
3739: snapshots of the gas distribution during a merger with (without) feedback
3740: from a central black hole. The temperature of the gas is color coded.}
3741: \label{merger}
3742: \end{figure}
3743: 
3744: 
3745: \begin{figure}
3746: \centering
3747: \includegraphics[height=6cm]{loebF2.eps}
3748: \caption{The comoving density of supermassive BHs per unit BH mass (from
3749: \cite{WL03}). The grey region shows the estimate based on the observed
3750: velocity distribution function of galaxies in Ref. \cite{She03} and the
3751: $M_{\rm bh}$--$v_c$ relation in Eq. (\ref{eq:1}). The lower bound
3752: corresponds to the lower limit in density for the observed velocity
3753: function while the grey lines show the extrapolation to lower densities. We
3754: also show the mass function computed at $z=1$, 3 and 6 from the
3755: Press-Schechter\cite{ps74} halo mass function and Eq.~(\ref{eq:1}), as well
3756: as the mass function at $z\sim2.35$ and $z\sim3$ implied by the observed
3757: density of quasars and a quasar lifetime of order the dynamical time of the
3758: host galactic disk, $t_{\rm dyn}$ (dot-dashed lines).}
3759: \label{fig2}
3760: \end{figure}
3761: 
3762: Agreement between the predicted and observed correlation function of
3763: quasars (Fig. \ref{fig3}) is obtained only if the BH mass scales with
3764: redshift as in Eq. (\ref{eq:1}) and the quasar lifetime is of the
3765: order of the dynamical time of the host galactic disk \cite{WLcl},
3766: \begin{equation}
3767: t_{\rm dyn}= 10^7~[\xi(z)]^{-1/2}\left({1+z\over 3}\right)^{-3/2}~{\rm yr}.
3768: \label{eq:life}
3769: \end{equation}
3770: This characterizes the timescale it takes low angular momentum gas to
3771: settle inwards and feed the black hole from across the galaxy before
3772: feedback sets in and suppresses additional infall. It also characterizes
3773: the timescale for establishing an outflow at the escape speed from the host
3774: spheroid.
3775: 
3776: The inflow of cold gas towards galaxy centers during the growth phase of
3777: the BH would naturally be accompanied by a burst of star formation.  The
3778: fraction of gas that is not consumed by stars or ejected by supernovae,
3779: will continue to feed the BH. It is therefore not surprising that quasar
3780: and starburst activities co-exist in Ultra Luminous Infrared Galaxies
3781: \cite{Gen02}, and that all quasars show broad metal lines indicating a
3782: super-solar metallicity of the surrounding gas \cite{Ham03}. Applying a
3783: similar self-regulation principle to the stars, leads to the expectation
3784: \cite{WL03,Kau00} that the ratio between the mass of the BH and the mass in
3785: stars is independent of halo mass (as observed locally \cite{Mag98}) but
3786: increases with redshift as $\propto \xi(z)^{1/2}(1+z)^{3/2}$. A
3787: consistent trend has indeed been inferred in an observed sample of
3788: gravitationally-lensed quasars \cite{Rix99}.
3789: 
3790: 
3791: \begin{figure}
3792: \centering
3793: \includegraphics[height=6cm]{loebF3.eps}
3794: \caption{Predicted correlation function of quasars at various redshifts
3795: in comparison to the {\it 2dF} data \cite{Cro01} (from \cite{WLcl}). The
3796: dark lines show the correlation function predictions for quasars of various
3797: apparent B-band magnitudes. The {\it 2dF} limit is $B\sim20.85$. The lower
3798: right panel shows data from entire {\it 2dF} sample in comparison to the
3799: theoretical prediction at the mean quasar redshift of $\langle
3800: z\rangle=1.5$. The $B=20.85$ prediction at this redshift is also shown by
3801: thick gray lines in the other panels to guide the eye. The predictions are
3802: based on the scaling $M_{\rm bh}\propto v_c^5$ in Eq. (\ref{eq:1}).  }
3803: \label{fig3}
3804: \end{figure}
3805: 
3806: 
3807: The upper mass of galaxies may also be regulated by the energy output from
3808: quasar activity. This would account for the fact that cooling flows are
3809: suppressed in present-day X-ray clusters \cite{Fab04,Boe02,OhS04}, and that
3810: massive BHs and stars in galactic bulges were already formed at $z\sim
3811: 2$. The quasars discovered by the {\it Sloan Digital Sky Survey} ({\it
3812: SDSS}) at $z\sim 6$ mark the early growth of the most massive BHs and
3813: galactic spheroids. The present-day abundance of galaxies capable of
3814: hosting BHs of mass $\sim 10^9M_\odot$ (based on Eq. \ref{eq:1}) already
3815: existed at $z\sim 6$ ~\cite{Loe03}. At some epoch, the quasar energy output
3816: may have led to the extinction of cold gas in these galaxies and the
3817: suppression of further star formation in them, leading to an apparent
3818: ``anti-hierarchical'' mode of galaxy formation where massive spheroids
3819: formed early and did not make new stars at late times. In the course of
3820: subsequent merger events, the cores of the most massive spheroids acquired
3821: an envelope of collisionless matter in the form of already-formed stars or
3822: dark matter \cite{Loe03}, without the proportional accretion of cold gas
3823: into the central BH. The upper limit on the mass of the central BH and the
3824: mass of the spheroid is caused by the lack of cold gas and cooling flows in
3825: their X-ray halos. In the cores of cooling X-ray clusters, there is often
3826: an active central BH that supplies sufficient energy to compensate for the
3827: cooling of the gas \cite{Boe02,Fab04,Beg04}. The primary physical process
3828: by which this energy couples to the gas is still unknown.
3829: 
3830: \begin{figure}
3831: \begin{center}
3832: \includegraphics[height=6cm]{loebF4.eps}
3833: %[width=.95\textwidth]{loebF4.eps}
3834: \end{center}
3835: \caption[]{ The global influence of magnetized quasar outflows on the
3836: intergalactic medium (from \cite{Fur01}). {\it Upper Panel:} Predicted
3837: volume filling fraction of magnetized quasar bubbles $F(z)$, as a function
3838: of redshift.  \emph{Lower Panel:} Ratio of normalized magnetic energy
3839: density, $\bar{u}_B/\epsilon_{-1}$, to the fiducial thermal energy density
3840: of the intergalactic medium $u_{fid} = 3 n(z) k T_{IGM}$, where $T_{IGM} =
3841: 10^4 ~{\rm K}$, as a function of redshift (see \cite{Fur01} for more
3842: details).  In each panel, the solid curves assume that the blast wave
3843: created by quasar ouflows is nearly (80\%) adiabatic, and that the minimum
3844: halo mass of galaxies, $M_{h,min}$, is determined by atomic cooling before
3845: reionization and by suppression due to galactic infall afterwards (top
3846: curve), $M_{h,min} = 10^9 M_\odot$ (middle curve), and $M_{h,min} = 10^{10}
3847: M_\odot$ (bottom curve). The dashed curve assumes a fully-radiative blast
3848: wave and fixes $M_{h,min}$ by the thresholds for atomic cooling and infall
3849: suppression.  The vertical dotted line indicates the assumed redshift of
3850: complete reionization, $z_r=7$.  }
3851: \label{fig4}
3852: \end{figure}
3853: 
3854: 
3855: \subsection{Feedback on Large Intergalactic Scales}
3856: 
3857: Aside from affecting their host galaxy, quasars disturb their large-scale
3858: cosmological environment. Powerful quasar outflows are observed in the form
3859: of radio jets \cite{Beg84} or broad-absorption-line winds
3860: \cite{Bran03}. The amount of energy carried by these outflows is largely
3861: unknown, but could be comparable to the radiative output from the same
3862: quasars. Furlanetto \& Loeb \cite{Fur01} have calculated the intergalactic
3863: volume filled by such outflows as a function of cosmic time (see
3864: Fig. \ref{fig4}). This volume is likely to contain magnetic fields and
3865: metals, providing a natural source for the observed magnetization of the
3866: metal-rich gas in X-ray clusters \cite{Kro01} and in galaxies \cite{Dal90}.
3867: The injection of energy by quasar outflows may also explain the deficit of
3868: Ly$\alpha$ absorption in the vicinity of Lyman-break galaxies
3869: \cite{Ade03,Cro02} and the required pre-heating in X-ray clusters
3870: \cite{Bor02,Boe02}.
3871: 
3872: Beyond the reach of their outflows, the brightest {\it SDSS} quasars at
3873: $z>6$ are inferred to have ionized exceedingly large regions of gas (tens
3874: of comoving Mpc) around them prior to global reionization (see
3875: Fig. \ref{fig5} and Refs. \cite{WB1,WL04c}). Thus, quasars must have
3876: suppressed the faint-end of the galaxy luminosity function in these regions
3877: before the same occurred throughout the Universe.  The recombination time
3878: is comparable to the Hubble time for the mean gas density at $z\sim 7$ and
3879: so ionized regions persist \cite{Oh03} on these large scales where
3880: inhomogeneities are small.  The minimum galaxy mass is increased by at
3881: least an order of magnitude to a virial temperature of $\sim 10^5$K in
3882: these ionized regions \cite{BL01}.  It would be particularly interesting
3883: to examine whether the faint end ($\sigma_\star < 30{\rm km~s^{-1}}$) of
3884: the luminosity function of dwarf galaxies shows any moduluation on
3885: large-scales around rare massive BHs, such as M87.
3886: 
3887: To find the volume filling fraction of relic regions from $z\sim 6$, we
3888: consider a BH of mass $M_{\rm bh}\sim3\times10^9M_\odot$. We can estimate
3889: the comoving density of BHs directly from the observed quasar luminosity
3890: function and our estimate of quasar lifetime.  At $z\sim 6$, quasars
3891: powered by $M_{\rm bh}\sim3\times10^9M_\odot$ BHs had a comoving density of
3892: $\sim 0.5\mbox{Gpc}^{-3}$\cite{WL03}.  However, the Hubble time exceeds
3893: $t_{\rm dyn}$ by a factor of $\sim 2\times 10^2$ (reflecting the square
3894: root of the density contrast of cores of galaxies relative to the mean
3895: density of the Universe), so that the comoving density of the bubbles
3896: created by the $z\sim 6$ BHs is $\sim10^2\mbox{Gpc}^{-3}$ (see
3897: Fig. \ref{fig2}). The density implies that the volume filling fraction of
3898: relic $z\sim 6$ regions is small, $<10\%$, and that the nearest BH that had
3899: $M_{\rm bh}\sim3\times10^9M_\odot$ at $z\sim 6$ (and could have been
3900: detected as an {\it SDSS} quasar then) should be at a distance $d_{\rm
3901: bh}\sim \left(4\pi/3\times10^2\right)^{1/3}\mbox{Gpc} \sim140\mbox{Mpc}$
3902: which is almost an order-of-magnitude larger than the distance of M87, a
3903: galaxy known to possess a BH of this mass \cite{For94}.
3904: 
3905: \begin{figure}
3906: \begin{center}
3907: \includegraphics[height=6cm]{loebF5.eps}
3908: %[width=.8\textwidth]{loebF5.eps}
3909: \end{center}
3910: \caption[]{Quasars serve as probes of the end of reionization. The measured
3911: size of the HII regions around {\it SDSS} quasars can be used
3912: \cite{WL04b,Mes04} to demonstrate that a significant fraction of the
3913: intergalactic hydrogen was neutral at $z\sim 6.3$ or else the inferred size
3914: of the quasar HII regions would have been much larger than observed
3915: (assuming typical quasar lifetimes \cite{Mar03}).  Also, quasars can be
3916: used to measure the redshift at which the intergalactic medium started to
3917: transmit Ly$\alpha$ photons\cite{WB1,WL04c}. The upper panel illustrates how
3918: the line-of-sight towards a quasar intersects this transition redshift. The
3919: resulting Ly$\alpha$ transmission of the intrinsic quasar spectrum is shown
3920: schematically in the lower panel.}
3921: \label{fig5}
3922: \end{figure}
3923: 
3924: 
3925: {\it What is the most massive BH that can be detected dynamically in a
3926: local galaxy redshift survey?} {\it SDSS} probes a volume of
3927: $\sim1\mbox{Gpc}^3$ out to a distance $\sim30$ times that of M87.  At the
3928: peak of quasar activity at $z\sim 3$, the density of the brightest quasars
3929: implies that there should be $\sim100$ BHs with masses of
3930: $3\times10^{10}M_\odot$ per $\mbox{Gpc}^{3}$, the nearest of which will be
3931: at a distance $d_{\rm bh}\sim130\mbox{Mpc}$, or $\sim 7$ times the distance
3932: to M87.  The radius of gravitational influence of the BH scales as $M_{\rm
3933: bh}/v_{\rm c}^2\propto M_{\rm bh}^{3/5}$. We find that for the nearest
3934: $3\times10^9M_\odot$ and $3\times10^{10}M_\odot$ BHs, the angular radius of
3935: influence should be similar.  Thus, the dynamical signature of $\sim
3936: 3\times 10^{10}M_\odot$ BHs on their stellar host should be detectable.
3937: 
3938: \subsection{What seeded the growth of the supermassive black holes?}
3939: 
3940: The BHs powering the bright {\it SDSS} quasars possess a mass of a few
3941: $\times 10^9 M_\odot$, and reside in galaxies with a velocity dispersion of
3942: $\sim 500 {\rm km~s^{-1}}$\cite{nature}.  A quasar radiating at its
3943: Eddington limiting luminosity, $L_E=1.4\times 10^{46}~{\rm
3944: erg~s^{-1}}(M_{\rm bh}/10^8M_\odot)$, with a radiative efficiency,
3945: $\epsilon_{\rm rad}=L_{E}/{\dot M}c^2$ would grow exponentially in mass as
3946: a function of time $t$, $M_{\rm bh} =M_{\rm seed}\exp\{t/t_E\}$ on a time
3947: scale, $t_E=4.1\times 10^7~{\rm yr} (\epsilon_{\rm rad}/0.1)$. Thus, the
3948: required growth time in units of the Hubble time $t_{\rm hubble}= 9\times
3949: 10^8~{\rm yr}[(1+z)/7]^{-3/2}$ is
3950: \begin{equation}
3951: {t_{\rm growth}\over t_{\rm hubble}}=0.7 \left({\epsilon_{\rm rad} \over
3952: 10\%}\right) \left({1+z\over 7}\right)^{3/2}\ln \left({{M_{\rm
3953: bh}/10^9M_\odot} \over M_{\rm seed}/100M_\odot}\right) ~.
3954: \end{equation}
3955: The age of the Universe at $z\sim 6$ provides just sufficient time to grow
3956: an {\it SDSS} BH with $M_{\rm bh}\sim 10^9M_\odot$ out of a stellar mass
3957: seed with $\epsilon_{\rm rad}=10 \%$ \cite{Hai01}. The growth time is
3958: shorter for smaller radiative efficiencies, as expected if the seed
3959: originates from the optically-thick collapse of a supermassive star (in
3960: which case $M_{\rm seed}$ in the logarithmic factor is also larger).
3961: 
3962: \begin{figure}
3963: \begin{center}
3964: \includegraphics[height=6cm]{loebF6.eps}
3965: %[width=.8\textwidth]{loebF6.eps}
3966: \end{center}
3967: \caption[]{SPH simulation of the collapse of an early dwarf galaxy with a
3968: virial temperature just above the cooling threshold of atomic hydrogen and
3969: no H$_2$ (from \cite{Bro03}).  The image shows a snapshot of the gas
3970: density distribution at $z\approx 10$, indicating the formation of two
3971: compact objects near the center of the galaxy with masses of $2.2\times
3972: 10^{6}M_{\odot}$ and $3.1\times 10^{6}M_{\odot}$, respectively, and radii
3973: $<1$ pc. Sub-fragmentation into lower mass clumps is inhibited as long as
3974: molecular hydrogen is dissociated by a background UV flux.  These
3975: circumstances lead to the formation of supermassive stars
3976: \cite{Loe94} that inevitably collapse and trigger the birth of supermassive
3977: black holes \cite{Loe94,Bau02}.  The box size is 200 pc.  }
3978: \label{fig6}
3979: \end{figure}
3980: 
3981: {\it What was the mass of the initial BH seeds?  Were they planted in early
3982: dwarf galaxies through the collapse of massive, metal free (Pop-III) stars
3983: (leading to $M_{\rm seed}$ of hundreds of solar masses) or through the
3984: collapse of even more massive, i.e. supermassive, stars \cite{Loe94} ?}
3985: ~Bromm \& Loeb \cite{Bro03} have shown through a hydrodynamical simulation
3986: (see Fig. \ref{fig6}) that supermassive stars were likely to form in early
3987: galaxies at $z\sim 10$ in which the virial temperature was close to the
3988: cooling threshold of atomic hydrogen, $\sim 10^4$K. The gas in these
3989: galaxies condensed into massive $\sim 10^6M_\odot$ clumps (the progenitors
3990: of supermassive stars), rather than fragmenting into many small clumps (the
3991: progenitors of stars), as it does in environments that are much hotter than
3992: the cooling threshold. This formation channel requires that a galaxy be
3993: close to its cooling threshold and immersed in a UV background that
3994: dissociates molecular hydrogen in it. These requirements should make this
3995: channel sufficiently rare, so as not to overproduce the cosmic mass density
3996: of supermassive BH.
3997: 
3998: The minimum seed BH mass can be identified observationally through the
3999: detection of gravitational waves from BH binaries with {\it Advanced LIGO}
4000: \cite{WL04a} or with {\it LISA} \cite{WL03b}.  Most of the mHz binary
4001: coalescence events originate at $z>7$ if the earliest galaxies included BHs
4002: that obey the $M_{\rm bh}$--$v_c$ relation in Eq. (\ref{eq:1}). The number
4003: of {\it LISA} sources per unit redshift per year should drop substantially
4004: after reionization, when the minimum mass of galaxies increased due to
4005: photo-ionization heating of the intergalactic medium.  Studies of the
4006: highest redshift sources among the few hundred detectable events per year,
4007: will provide unique information about the physics and history of BH growth
4008: in galaxies \cite{Vol04}.
4009: 
4010: The early BH progenitors can also be detected as unresolved point sources,
4011: using the future {\it James Webb Space Telescope} ({\it
4012: JWST}). Unfortunately, the spectrum of metal-free massive and supermassive
4013: stars is the same, since their surface temperature $\sim 10^5$K is
4014: independent of mass \cite{BKL2001}. Hence, an unresolved cluster of massive
4015: early stars would show the same spectrum as a supermassive star of the same
4016: total mass.
4017: 
4018: It is difficult to ignore the possible environmental impact of
4019: quasars on {\it anthropic} selection. One may wonder whether it is not a
4020: coincidence that our Milky-Way Galaxy has a relatively modest BH mass of
4021: only a few million solar masses in that the energy output from a much more
4022: massive (e.g. $\sim 10^9M_\odot$) black hole would have disrupted the
4023: evolution of life on our planet. A proper calculation remains to be done
4024: (as in the context of nearby Gamma-Ray Bursts \cite{Sca02}) in order to
4025: demonstrate any such link.
4026: 
4027: 
4028: %\noindent{\bf Impact of the First Stars}
4029: 
4030: \section{\bf Radiative Feedback from the First Sources of Light}
4031: \label{sec6}
4032: 
4033: \subsection{Escape of Ionizing Radiation from Galaxies}
4034: \label{sec6.1}
4035: 
4036: The intergalactic ionizing radiation field, a key ingredient in the
4037: development of reionization, is determined by the amount of ionizing
4038: radiation escaping from the host galaxies of stars and quasars.  The value
4039: of the escape fraction as a function of redshift and galaxy mass remains a
4040: major uncertainty in all current studies, and could affect the cumulative
4041: radiation intensity by orders of magnitude at any given redshift. Gas
4042: within halos is far denser than the typical density of the IGM, and in
4043: general each halo is itself embedded within an overdense region, so the
4044: transfer of the ionizing radiation must be followed in the densest regions
4045: in the Universe. Reionization simulations are limited in resolution and
4046: often treat the sources of ionizing radiation and their immediate
4047: surroundings as unresolved point sources within the large-scale
4048: intergalactic medium (see, e.g., Gnedin 2000 \cite{g00}).  The
4049: escape fraction is highly sensitive to the three-dimensional distribution
4050: of the UV sources relative to the geometry of the absorbing gas within the
4051: host galaxy (which may allow escape routes for photons along particular
4052: directions but not others). 
4053: 
4054: The escape of ionizing radiation ($h\nu > 13.6$eV, $\lambda < 912$ {\AA})
4055: from the disks of present-day galaxies has been studied in recent years in
4056: the context of explaining the extensive diffuse ionized gas layers observed
4057: above the disk in the Milky Way \cite{RTKMH95} and other galaxies
4058: \cite{Rand96, HWR99}. Theoretical models predict that of order 3--14\% of
4059: the ionizing luminosity from O and B stars escapes the Milky Way disk
4060: \cite{DS94,DSF99}.  A similar escape fraction of $f_{\rm esc}=6$\% was
4061: determined by Bland-Hawthorn \& Maloney (1999) \cite{BM99} based on
4062: H$\alpha$ measurements of the Magellanic Stream.  From {\it Hopkins
4063: Ultraviolet Telescope} observations of four nearby starburst galaxies
4064: (Leitherer et al.\ 1995 \cite{LFHL95}; Hurwitz, Jelinsky, \& Dixon 1997
4065: \cite{HJD97}), the escape fraction was estimated to be in the range
4066: 3\%$<f_{\rm esc} < 57$\%.  If similar escape fractions characterize high
4067: redshift galaxies, then stars could have provided a major fraction of the
4068: background radiation that reionized the IGM \cite{MS96,Madau99}.  However,
4069: the escape fraction from high-redshift galaxies, which formed when the
4070: Universe was much denser ($\rho\propto (1+z)^3$), may be significantly
4071: lower than that predicted by models ment to describe present-day galaxies.
4072: Current reionization calculations assume that galaxies are isotropic point
4073: sources of ionizing radiation and adopt escape fractions in the range $5\%
4074: < f_{\rm esc} < 60\%$ \cite{g00}.
4075: 
4076: Clumping is known to have a significant effect on the penetration and
4077: escape of radiation from an inhomogeneous medium
4078: \cite{Boi90,WG96,Neufeld91,HS99,BFDA00}.  The inclusion of clumpiness
4079: introduces several unknown parameters into the calculation, such as the
4080: number and overdensity of the clumps, and the spatial correlation between
4081: the clumps and the ionizing sources.  An additional complication may arise
4082: from hydrodynamic feedback, whereby part of the gas mass is expelled from
4083: the disk by stellar winds and supernovae (\S \ref{sec7}).
4084: 
4085: Wood \& Loeb (2000) \cite{WL00} used a three-dimensional radiation transfer
4086: code to calculate the steady-state escape fraction of ionizing photons from
4087: disk galaxies as a function of redshift and galaxy mass. The gaseous disks
4088: were assumed to be isothermal, with a sound speed $c_s\sim 10~{\rm
4089: km~s^{-1}}$, and radially exponential, with a scale-length based on the
4090: characteristic spin parameter and virial radius of their host halos. The
4091: corresponding temperature of $\sim 10^4$ K is typical for a gas which is
4092: continuousely heated by photo-ionization from stars.  The sources of
4093: radiation were taken to be either stars embedded in the disk, or a central
4094: quasar. For stellar sources, the predicted increase in the disk density
4095: with redshift resulted in a strong decline of the escape fraction with
4096: increasing redshift. The situation is different for a central quasar. Due
4097: to its higher luminosity and central location, the quasar tends to produce
4098: an ionization channel in the surrounding disk through which much of its
4099: ionizing radiation escapes from the host. In a steady state, only
4100: recombinations in this ionization channel must be balanced by ionizations,
4101: while for stars there are many ionization channels produced by individual
4102: star-forming regions and the total recombination rate in these channels is
4103: very high. Escape fractions $\ga 10\%$ were achieved for stars at $z\sim
4104: 10$ only if $\sim 90\%$ of the gas was expelled from the disks or if dense
4105: clumps removed the gas from the vast majority ($\ga 80\%$) of the disk
4106: volume (see Fig. \ref{fig6a}). This analysis applies only to halos with
4107: virial temperatures $\ga 10^4$ K. Ricotti \& Shull (2000) \cite{RS00}
4108: reached similar conclusions but for a quasi-spherical configuration of
4109: stars and gas.  They demonstrated that the escape fraction is substantially
4110: higher in low-mass halos with a virial temperature $\la 10^4$ K.  However,
4111: the formation of stars in such halos depends on their uncertain ability to
4112: cool via the efficient production of molecular hydrogen.
4113: 
4114: \noindent
4115: \begin{figure} 
4116: %\epsscale{0.7}
4117: %\plotone{VI.2_fig1.ps}
4118: \centering
4119: \includegraphics[height=6cm]{VI.2_fig1.ps}
4120: \caption{Escape fractions of stellar ionizing photons from a gaseous
4121: disk embedded within a $10^{10}M_\odot$ halo which have formed at
4122: $z=10$ (from Wood \& Loeb 2000 \cite{WL00}). The curves show three different cases
4123: of clumpiness within the disk. The volume filling factor refers to
4124: either the ionizing emissivity, the gas clumps, or both, depending on
4125: the case. The escape fraction is substantial ($\ga 1\%$) only if the
4126: gas distribution is highly clumped.  }
4127: \label{fig6a}
4128: \end{figure}
4129:   
4130: The main uncertainty in the above predictions involves the distribution of
4131: the gas inside the host galaxy, as the gas is exposed to the radiation
4132: released by stars and the mechanical energy deposited by supernovae.  Given
4133: the fundamental role played by the escape fraction, it is desirable to
4134: calibrate its value observationally.  Steidel, Pettini, \& Adelberger
4135: \cite{Ste00} reported a  detection of significant Lyman
4136: continuum flux in the composite spectrum of 29 Lyman break galaxies (LBG)
4137: with redshifts in the range $z = 3.40\pm 0.09$. They co-added the spectra
4138: of these galaxies in order to be able to measure the low flux. Another
4139: difficulty in the measurement comes from the need to separate the
4140: Lyman-limit break caused by the interstellar medium from that already
4141: produced in the stellar atmospheres. After correcting for intergalactic
4142: absorption, Steidel et al.\ \cite{Ste00} inferred a ratio between
4143: the emergent flux density at 1500\AA~ and 900\AA~ (rest frame) of $4.6 \pm
4144: 1.0$. Taking into account the fact that the stellar spectrum should already
4145: have an intrinsic Lyman discontinuity of a factor of $\sim 3$--5, but that
4146: only $\sim 15$--$20\%$ of the 1500\AA~ photons escape from typical LBGs
4147: without being absorbed by dust (Pettini et al.\ 1998 \cite{PKSDAG98a};
4148: Adelberger et al.\ 2000 \cite{Ade00}), the inferred 900\AA~ escape fraction
4149: is $f_{\rm esc} \sim 10$--$20\%$.  Although the galaxies in this sample
4150: were drawn from the bluest quartile of the LBG spectral energy
4151: distributions, the measurement implies that this quartile may itself
4152: dominate the hydrogen-ionizing background relative to quasars at $z\sim 3$.
4153: 
4154: %<>
4155: \subsection{Propagation of Ionization Fronts in the IGM}
4156: \label{sec6.2}
4157: 
4158: The radiation output from the first stars ionizes hydrogen in a
4159: growing volume, eventually encompassing almost the entire IGM within a
4160: single H II  bubble. In the early stages of this process, each
4161: galaxy produces a distinct H II  region, and only when the
4162: overall H II  filling factor becomes significant do neighboring
4163: bubbles begin to overlap in large numbers, ushering in the ``overlap
4164: phase'' of reionization. Thus, the first goal of a model of
4165: reionization is to describe the initial stage, when each source
4166: produces an isolated expanding H II  region.
4167: 
4168: We assume a spherical ionized volume $V$, separated from the surrounding
4169: neutral gas by a sharp ionization front. Indeed, in the case of a stellar
4170: ionizing spectrum, most ionizing photons are just above the hydrogen
4171: ionization threshold of 13.6 eV, where the absorption cross-section is high
4172: and a very thin layer of neutral hydrogen is sufficient to absorb all the
4173: ionizing photons. On the other hand, an ionizing source such as a quasar
4174: produces significant numbers of higher energy photons and results in a
4175: thicker transition region.
4176: 
4177: In the absence of recombinations, each hydrogen atom in the IGM would
4178: only have to be ionized once, and the ionized proper volume $V_p$
4179: would simply be determined by \beq \nb_H V_p=\Ng\ , \eeq where $\nb_H$
4180: is the mean number density of hydrogen and $\Ng$ is the total number
4181: of ionizing photons produced by the source. However, the increased
4182: density of the IGM at high redshift implies that recombinations cannot
4183: be neglected. Indeed, in the case of a steady ionizing source (and
4184: neglecting the cosmological expansion), a steady-state volume would be
4185: reached corresponding to the Str\"{o}mgren sphere, with recombinations
4186: balancing ionizations: \beq \alpha_B \nb_H^2 V_p=\frac{d\, \Ng}{dt}\ ,
4187: \eeq where the recombination rate depends on the square of the density
4188: and on the case B recombination coefficient $\alpha_B=2.6\times 10^{-13}$
4189: cm$^3$ s$^{-1}$ for hydrogen at $T=10^4$ K. The exact evolution for an
4190: expanding H II  region, including a non-steady ionizing source,
4191: recombinations, and cosmological expansion, is given by (Shapiro \&
4192: Giroux 1987 \cite{SG87}) \beq \nb_H\left( \frac{dV_p}{dt}-3 H V_p\right)=
4193: \frac{d\, \Ng}{dt} - \alpha_B \left<n_H^2\right> V_p\ . \label{front}
4194: \eeq In this equation, the mean density $\nb_H$ varies with time as
4195: $1/a^3(t)$. A critical physical ingredient is the dependence of
4196: recombination on the square of the density. This means that if the IGM
4197: is not uniform, but instead the gas which is being ionized is mostly
4198: distributed in high-density clumps, then the recombination time is
4199: very short. This is often dealt with by introducing a volume-averaged
4200: clumping factor $C$ (in general time-dependent), defined
4201: by\footnote{The recombination rate depends on the number density of
4202: electrons, and in using equation~(\ref{clump}) we are neglecting the
4203: small contribution caused by partially or fully ionized helium.} \beq
4204: C=\left<n_H^2\right>/\nb_H^2 \label{clump}\ . \eeq
4205: 
4206: If the ionized volume is large compared to the typical scale of clumping,
4207: so that many clumps are averaged over, then equation~(\ref{front}) can be
4208: solved by supplementing it with equation~(\ref{clump}) and specifying
4209: $C$. Switching to the comoving volume $V$, the resulting equation is \beq
4210: \frac{dV}{dt}= \frac{1}{\nb_H^0} \frac{d\, \Ng}{dt}- \alpha_B \frac{C}{a^3}
4211: \nb_H^0 V\ , \label{HIIreg} \eeq where the present number density of
4212: hydrogen is \beq \nb_H^0=1.88\times 10^{-7} \left(\frac{\Omega_b
4213: h^2}{0.022}\right)\ {\rm cm}^{-3}\ . \eeq This number density is lower than
4214: the total number density of baryons $\nb_b^0$ by a factor of $\sim 0.76$,
4215: corresponding to the primordial mass fraction of hydrogen. The solution for
4216: $V(t)$ (generalized from Shapiro \& Giroux 1987 \cite{SG87}) around a source which
4217: turns on at $t=t_i$ is \beq V(t)=\int_{t_i}^t \frac{1}{\nb_H^0} \frac{d\,
4218: \Ng}{dt'} \,e^{F(t',t)} dt'\ ,\label{HIIsoln} \eeq where \beq
4219: F(t',t)=-\alpha_B \nb_H^0 \int_{t'}^t \frac{C(t'')} {a^3(t'')}\, dt''\
4220: \label{Fgen}. \eeq At high redshift (when $(1+z) \gg |\Omm^{-1}-1|$), the
4221: scale factor varies as \beq a(t)\simeq \left(\frac{3}{2}\sqrt{\Omm} H_0
4222: t\right)^{2/3}\ , \eeq and with the additional assumption of a constant $C$
4223: the function $F$ simplifies as follows. Defining \beq f(t)=a(t)^{-3/2}
4224: \label{foft}\ , \eeq we derive \beq F(t',t)=-\frac{2}{3}\frac{ \alpha_B
4225: \nb_H^0} {\sqrt{\Omm} H_0}\,C \left[f(t')-f(t)\right]=-0.262 \left[f(t')
4226: -f(t)\right] \ , \eeq where the last equality assumes $C=10$ and our
4227: standard choice of cosmological parameters: $\Omm=0.3$, $\Oml=0.7$, and
4228: $\Omega_b=0.045$. Although this expression for $F(t',t)$ is in general an
4229: accurate approximation at high redshift, in the particular case of the
4230: $\Lambda$CDM model (where $\Omm+\Oml=1$) we get the exact result by
4231: replacing equation~(\ref{foft}) with \beq
4232: f(t)=\sqrt{\frac{1}{a^3}+\frac{1-\Omm}{\Omm}}\ . \label{fLCDM} \eeq
4233: 
4234: The size of the resulting H II  region depends on the halo which
4235: produces it. Consider a halo of total mass $M$ and baryon fraction
4236: $\Omega_b/\Omm$. To derive a rough estimate, we assume that baryons
4237: are incorporated into stars with an efficiency of $f_{\rm star}=10\%$,
4238: and that the escape fraction for the resulting ionizing radiation is
4239: also $f_{\rm esc}=10\%$. If the stellar IMF is similar to the one
4240: measured locally \cite{Sca98}, then $N_{\gamma}
4241: \approx 4000$ ionizing photons are produced per baryon in stars (for a
4242: metallicity equal to $1/20$ of the solar value). We define a parameter
4243: which gives the overall number of ionizations per baryon, \beq \Ni
4244: \equiv N_{\gamma} \, f_{\rm star}\, f_{\rm esc}\ . \eeq If we neglect
4245: recombinations then we obtain the maximum comoving radius of the
4246: region which the halo of mass $M$ can ionize, \beq r_{\rm max}=
4247: \left(\frac{3}{4\pi}\, \frac{\Ng} {\nb_H^0} \right)^{1/3} =
4248: \left(\frac{3}{4\pi}\, \frac{\Ni} {\nb_H^0}\, \frac{\Omega_b}{\Omm}\,
4249: \frac{M}{m_p} \right)^{1/3}= 680\, {\rm kpc} \left( \frac{\Ni}{40}\,
4250: \frac{M} {10^8 M_{\sun}}\right)^{1/3}\ , \label{rmax} \eeq for our
4251: standard set of parameters. The actual radius never reaches this size
4252: if the recombination time is shorter than the lifetime of the ionizing
4253: source. For an instantaneous starburst with the Scalo (1998) 
4254: \cite{Sca98} IMF, the production rate of ionizing photons can be
4255: approximated as \beq \frac{d\,
4256: \Ng}{dt}=\frac{\alpha-1}{\alpha} \frac{\Ng} {t_s}\times \left\{
4257: \begin{array}{ll} \ \ \, 1 & \mbox{if
4258: $t<t_s$,} \\ \left(\frac{t}{t_s}\right)^{-\alpha}\ & \mbox{otherwise,}
4259: \end{array} \right. \label{zoltan} \eeq
4260: where $\Ng=4000$, $\alpha=4.5$, and the most massive stars fade away
4261: with the characteristic timescale $t_s=3\times 10^6$ yr. In figure
4262:  \ref{fig6b} we show the time evolution of the volume ionized by such
4263: a source, with the volume shown in units of the maximum volume $V_{\rm
4264: max}$ which corresponds to $r_{\rm max}$ in equation~(\ref{rmax}). We
4265: consider a source turning on at $z=10$ (solid curves) or $z=15$
4266: (dashed curves), with three cases for each: no recombinations, $C=1$,
4267: and $C=10$, in order from top to bottom (Note that the result is
4268: independent of redshift in the case of no recombinations). When
4269: recombinations are included, the volume rises and reaches close to
4270: $V_{\rm max}$ before dropping after the source turns off. At large $t$
4271: recombinations stop due to the dropping density, and the volume
4272: approaches a constant value (although $V \ll V_{\rm max}$ at large $t$
4273: if $C=10$).
4274: 
4275: %%%%%%%% Figure6.2
4276: \begin{figure}
4277: \centering
4278: \includegraphics[height=6cm]{VI3fig1.ps}
4279: \caption{Expanding H II  region around an isolated ionizing
4280: source. The comoving ionized volume $V$ is expressed in units of the
4281: maximum possible volume, $V_{\rm max}=4\pi r_{\rm max}^3/3$ [with
4282: $r_{\rm max}$ given in equation~(\ref{rmax})], and the time is
4283: measured after an instantaneous starburst which produces ionizing photons
4284: according to equation~(\ref{zoltan}). We consider a source turning on at
4285: $z=10$ (solid curves) or $z=15$ (dashed curves), with three cases for each:
4286: no recombinations, $C=1$, and $C=10$, in order from top to bottom. The
4287: no-recombination curve is identical for the different source redshifts.}
4288: \label{fig6b}
4289: \end{figure}
4290: %%%%%%%%
4291: 
4292: We obtain a similar result for the size of the H II region around a galaxy
4293: if we consider a mini-quasar rather than stars. For the typical quasar
4294: spectrum (Elvis et al.\ 1994 \cite{ELV94}), roughly 11,000 ionizing photons
4295: are produced per baryon incorporated into the black hole, assuming a
4296: radiative efficiency of $\sim 6\%$. The efficiency of incorporating the
4297: baryons in a galaxy into a central black hole is low ($\la 0.6\%$ in the
4298: local Universe, e.g.\ Magorrian et al.\ 1998 \cite{Mag98}), but the escape
4299: fraction for quasars is likely to be close to unity, i.e., an order of
4300: magnitude higher than for stars (see previous sub-section). Thus, for every
4301: baryon in galaxies, up to $\sim 65$ ionizing photons may be produced by a
4302: central black hole and $\sim 40$ by stars, although both of these numbers
4303: for $\Ni$ are highly uncertain. These numbers suggest that in either case
4304: the typical size of H II regions before reionization may be $\la 1$ Mpc or
4305: $\sim 10$ Mpc, depending on whether $10^8 M_{\sun}$ halos or $10^{12}
4306: M_{\sun}$ halos dominate.
4307: 
4308: 
4309: The ionization front around a bright transient source like a quasar expands
4310: at early times at nearly the speed of light. This occurs when the HII
4311: region is sufficiently small so that the production rate of ionizing
4312: photons by the central source exceeds their consumption rate by hydrogen
4313: atoms within this volume. It is straightforward to do the accounting for
4314: these rates (including recombinations) taking the light propagation delay
4315: into account.  This was done by Wyithe \& Loeb \cite{WL04b} [see also White
4316: et al. (2003) \cite{WB1}] who derived the general equation for the
4317: relativistic expansion of the {\em comoving} radius [$r=(1+z)r_{\rm p}$]
4318: of the quasar \HII region in an IGM with a neutral filling fraction $x_{\rm
4319: HI}$ (fixed by other ionizing sources) as,
4320: \begin{equation}
4321: \label{Vev2}
4322: \frac{dr}{dt}=c(1+z)\left[\frac{\dot{N}_{\gamma} - \alpha_{\rm B}C x_{\rm
4323: HI}\left(\bar{n}^0_{\rm H}\right)^2 \left(1+z\right)^3 \left({4\pi\over
4324: 3}r^3\right)} {\dot{N}_{\gamma} + 4\pi r^2 \left(1+z\right) c x_{\rm
4325: HI}\bar{n}^0_{\rm H}}\right],
4326: \end{equation}
4327: where $c$ is the speed of light, $C$ is the clumping factor, $\alpha_{\rm
4328: B}=2.6\times10^{-13}$cm$^3$s$^{-1}$ is the case-B recombination coefficient
4329: at the characteristic temperature of $10^4$K, and $\dot{N}_{\gamma}$ is the
4330: rate of ionizing photons crossing a shell at the radius of the HII region
4331: at time $t$.  Indeed, for $\dot{N}_\gamma\rightarrow \infty$ the
4332: propagation speed of the proper radius of the HII region $r_p=r/(1+z)$
4333: approaches the speed of light in the above expression,
4334: $(dr_p/dt)\rightarrow c$.  The actual size of the HII region along the
4335: line-of-sight to a quasar can be inferred from the extent of the spectral
4336: gap between the quasar's rest-frame Ly$\alpha$ wavelength and the start of
4337: Ly$\alpha$ absorption by the IGM in the observed spectrum.  Existing data
4338: from the SDSS quasars \cite{WL04b,Mes04, Wy05} provide typical values of
4339: $r_p\sim 5$Mpc and indicate for plausible choices of the quasar lifetimes
4340: that $x_{\rm HI}>0.1$ at $z>6$.  These ionized bubbles could be imaged
4341: directly by future 21cm maps of the regions around the highest-redshift
4342: quasars \cite{Tozzi,WyLo,WyBar}.
4343: 
4344: The profile of the Ly$\alpha$ emission line of galaxies has also been
4345: suggested as a probe of the ionization state of the IGM
4346: \cite{Loeb_Rybicki,Santos,Haiman_Cen,Haiman05,Madau_Rees,Loeb_Hernquist,Rhoads}.
4347: If the IGM is neutral, then the damping wing of the Gunn-Peterson trough in
4348: equation (\ref{eq:shift}) is modified since Ly$\alpha$ absorption starts
4349: only from the near edge of the ionized region along the line-of-sight to
4350: the source \cite{Haiman_Cen,Madau_Rees}. Rhoads \& Malhotra \cite{Rhoads}
4351: showed that the observed abundance of galaxies with Ly$\alpha$ emission at
4352: $z\sim 6.5$ indicates that a substantial fraction (tens of percent) of the
4353: IGM must be ionized in order to allow transmission of the observed
4354: Ly$\alpha$ photons.  However, if these galaxies reside in groups, then
4355: galaxies with peculiar velocities away from the observer will
4356: preferentially Doppler-shift the emitted Ly$\alpha$ photons to the red wing
4357: of the Ly$\alpha$ resonance and reduce the depression of the line profile
4358: \cite{Loeb_Hernquist,Haiman_Mesinger}. Additional uncertainties in the
4359: intrinsic line profile based on the geometry and the stellar or gaseous
4360: contents of the source galaxy \cite{Loeb_Hernquist,Santos}, as well as the
4361: clustering of galaxies which ionize their immediate environment in groups
4362: \cite{WL04c,Furlanetto_Hernquist}, limits this method from reaching
4363: robust conclusions. Imaging of the expected halos of scattered Ly$\alpha$
4364: radiation around galaxies embedded in a neutral IGM
4365: \cite{Loeb_Rybicki,Rybicki_Loeb} provide a more definitive test of the
4366: neutrality of the IGM, but is more challenging observationally.
4367: 
4368: 
4369: \subsection{Reionization of Hydrogen}
4370: \label{sec6.3}
4371: 
4372: In this section we summarize recent progress, both analytic and
4373: numerical, made toward elucidating the basic physics of reionization
4374: and the way in which the characteristics of reionization depend on the
4375: nature of the ionizing sources and on other input parameters of
4376: cosmological models.
4377: 
4378: The process of the reionization of hydrogen involves several distinct
4379: stages. The initial, ``pre-overlap'' stage (using the terminology of Gnedin
4380: \cite{g00}) consists of individual ionizing sources turning on and
4381: ionizing their surroundings. The first galaxies form in the most massive
4382: halos at high redshift, and these halos are biased and are preferentially
4383: located in the highest-density regions. Thus the ionizing photons which
4384: escape from the galaxy itself (see \S \ref{sec6.1}) must then make their
4385: way through the surrounding high-density regions, which are characterized
4386: by a high recombination rate. Once they emerge, the ionization fronts
4387: propagate more easily into the low-density voids, leaving behind pockets of
4388: neutral, high-density gas. During this period the IGM is a two-phase medium
4389: characterized by highly ionized regions separated from neutral regions by
4390: ionization fronts.  Furthermore, the ionizing intensity is very
4391: inhomogeneous even within the ionized regions, with the intensity
4392: determined by the distance from the nearest source and by the ionizing
4393: luminosity of this source.
4394: 
4395: The central, relatively rapid ``overlap'' phase of reionization begins when
4396: neighboring H II  regions begin to overlap. Whenever two ionized
4397: bubbles are joined, each point inside their common boundary becomes exposed
4398: to ionizing photons from both sources. Therefore, the ionizing intensity
4399: inside H II  regions rises rapidly, allowing those regions to expand
4400: into high-density gas which had previously recombined fast enough to remain
4401: neutral when the ionizing intensity had been low. Since each bubble
4402: coalescence accelerates the process of reionization, the overlap phase has
4403: the character of a phase transition and is expected to occur rapidly, over
4404: less than a Hubble time at the overlap redshift. By the end of this stage
4405: most regions in the IGM are able to see several unobscured sources, and
4406: therefore the ionizing intensity is much higher than before overlap and it
4407: is also much more homogeneous. An additional ingredient in the rapid
4408: overlap phase results from the fact that hierarchical structure formation
4409: models predict a galaxy formation rate that rises rapidly with time at the
4410: relevant redshift range. This process leads to a state in which the
4411: low-density IGM has been highly ionized and ionizing radiation reaches
4412: everywhere except for gas located inside self-shielded, high-density
4413: clouds. This marks the end of the overlap phase, and this important
4414: landmark is most often referred to as the 'moment of reionization'.
4415: 
4416: Some neutral gas does, however, remain in high-density structures
4417: which correspond to Lyman Limit systems and damped Ly$\alpha$
4418: systems seen in absorption at lower redshifts. The high-density
4419: regions are gradually ionized as galaxy formation proceeds, and the
4420: mean ionizing intensity also grows with time. The ionizing intensity
4421: continues to grow and to become more uniform as an increasing number
4422: of ionizing sources is visible to every point in the IGM. This
4423: ``post-overlap'' phase continues indefinitely, since collapsed objects
4424: retain neutral gas even in the present Universe. The IGM does,
4425: however, reach another milestone at $z \sim 1.6$, the breakthrough
4426: redshift \cite{Mad99}. Below this redshift, all
4427: ionizing sources are visible to each other, while above this redshift
4428: absorption by the Ly$\alpha$ forest implies that only
4429: sources in a small redshift range are visible to a typical point in
4430: the IGM.
4431: 
4432: Semi-analytic models of the pre-overlap stage focus on the evolution
4433: of the H II  filling factor, i.e., the fraction of the volume of
4434: the Universe which is filled by H II  regions. We distinguish
4435: between the naive filling factor $F_{\rm H\ II}$ and the actual
4436: filling factor or porosity $Q_{\rm H\ II}$. The naive filling factor
4437: equals the number density of bubbles times the average volume of each,
4438: and it may exceed unity since when bubbles begin to overlap the
4439: overlapping volume is counted multiple times. However, as explained
4440: below, in the case of reionization the linearity of the physics means
4441: that $F_{\rm H\ II}$ is a very good approximation to $Q_{\rm H\ II}$
4442: up to the end of the overlap phase of reionization.
4443: 
4444: The model of individual H II  regions presented in the previous
4445: section can be used to understand the development of the total filling
4446: factor. Starting with equation~(\ref{HIIreg}), if we assume a common
4447: clumping factor $C$ for all H II  regions then we can sum each
4448: term of the equation over all bubbles in a given large volume of the
4449: Universe, and then divide by this volume. Then $V$ is replaced by the
4450: filling factor and $\Ng$ by the total number of ionizing photons
4451: produced up to some time $t$, per unit volume. The latter quantity
4452: equals the mean number of ionizing photons per baryon times the mean
4453: density of baryons $\nb_b$. Following the arguments leading to
4454: equation~(\ref{rmax}), we find that if we include only stars then \beq
4455: \frac {\nb_\gamma} {\nb_b}= \Ni F_{\rm col}\ , \label{ngnb} \eeq where
4456: the collapse fraction $F_{\rm col}$ is the fraction of all the baryons
4457: in the Universe which are in galaxies, i.e., the fraction of gas which
4458: settles into halos and cools efficiently inside them. In writing
4459: equation~(\ref{ngnb}) we are assuming instantaneous production of
4460: photons, i.e., that the timescale for the formation and evolution of
4461: the massive stars in a galaxy is short compared to the Hubble time at
4462: the formation redshift of the galaxy. In a model based on
4463: equation~(\ref{HIIreg}), the near-equality between $F_{\rm H\ II}$ and
4464: $Q_{\rm H\ II}$ results from the linearity of this equation. First,
4465: the total number of ionizations equals the total number of ionizing
4466: photons produced by stars, i.e., all ionizing photons contribute
4467: regardless of the spatial distribution of sources; and second, the
4468: total recombination rate is proportional to the total ionized volume,
4469: regardless of its topology. Thus, even if two or more bubbles overlap
4470: the model remains an accurate approximation for $Q_{\rm H\ II}$ (at
4471: least until $Q_{\rm H\ II}$ becomes nearly equal to 1). Note, however,
4472: that there still are a number of important simplifications in the
4473: model, including the assumption of a homogeneous (though possibly
4474: time-dependent) clumping factor, and the neglect of feedback whereby
4475: the formation of one galaxy may suppress further galaxy formation in
4476: neighboring regions. These complications are discussed in detail below
4477: and in \S \ref{sec6.5} and \S \ref{sec7}.
4478: 
4479: Under these assumptions we convert equation~(\ref{HIIreg}), which describes
4480: individual H II regions, to an equation which statistically describes the
4481: transition from a neutral Universe to a fully ionized one (compare to Madau
4482: et al.\ 1999 \cite{Mad99} and Haiman \& Loeb 1997 \cite{hl97}): \beq
4483: \frac{dQ_{\rm H\ II}}{dt}=\frac{\Ni}{0.76} \frac{dF_{\rm col}}{dt}-
4484: \alpha_B \frac{C}{a^3} \nb_H^0 Q_{\rm H\ II}
4485: \label{QIIeqn}\ , \eeq where we assumed a primordial mass fraction of 
4486: hydrogen of 0.76. The solution (in analogy with
4487: equation~(\ref{HIIsoln})) is \beq Q_{\rm H\ II}(t) =\int_{0}^t
4488: \frac{\Ni} {0.76} \frac{dF_{\rm col}}{dt'}\,e^{F(t',t)} dt'\ , \eeq
4489: where $F(t',t)$ is determined by equations~(\ref{Fgen})--(
4490: \ref{fLCDM}).
4491:   
4492: A simple estimate of the collapse fraction at high redshift is the mass
4493: fraction (given by equation~(\ref{PSerfc}) in the Press-Schechter model) in
4494: halos above the cooling threshold, which is the minimum mass of halos in
4495: which gas can cool efficiently. Assuming that only atomic cooling is
4496: effective during the redshift range of reionization, the minimum mass
4497: corresponds roughly to a halo of virial temperature $T_{\rm vir}=10^4$ K,
4498: which can be converted to a mass using equation~(\ref{tvir}). With this
4499: prescription we derive (for $\Ni=40$) the reionization history shown in
4500: Fig. \ref{fig6c} for the case of a constant clumping factor $C$. The
4501: solid curves show $Q_{\rm H\ II}$ as a function of redshift for a clumping
4502: factor $C=0$ (no recombinations), $C=1$, $C=10$, and $C=30$, in order from
4503: left to right. Note that if $C \sim 1$ then recombinations are unimportant,
4504: but if $C \ga 10$ then recombinations significantly delay the reionization
4505: redshift (for a fixed star-formation history). The dashed curve shows the
4506: collapse fraction $F_{\rm col}$ in this model. For comparison, the vertical
4507: dotted line shows the $z=5.8$ observational lower limit (Fan et al.\ 2000 \cite{f0})
4508: on the reionization redshift.
4509: 
4510: %%%%%%%% Figure 
4511: \begin{figure}
4512: \centering
4513: \includegraphics[height=6cm]{VI4fig1.ps}
4514: \caption{Semi-analytic calculation of the reionization of the IGM (for
4515: $\Ni=40$), showing the redshift evolution of the filling factor
4516: $Q_{\rm H\ II}$. Solid curves show $Q_{\rm H\ II}$ for a clumping
4517: factor $C=0$ (no recombinations), $C=1$, $C=10$, and $C=30$, in order
4518: from left to right. The dashed curve shows the collapse fraction
4519: $F_{\rm col}$, and the vertical dotted line shows the $z=5.8$
4520: observational lower limit (Fan et al.\ 2000 \cite{f0}) on the reionization
4521: redshift.}
4522: \label{fig6c}
4523: \end{figure}
4524: %%%%%%%%
4525:   
4526: Clearly, star-forming galaxies in CDM hierarchical models are capable of
4527: ionizing the Universe at $z\sim 6$--15 with reasonable parameter
4528: choices. This has been shown by a large number of theoretical,
4529: semi-analytic calculations
4530: \cite{fk94,sgb94,hl97,VS99,CO00,CFA00,Wyithe03,Cen03,Tumlinson04} as well
4531: as numerical simulations
4532: \cite{CO93,G097,g00,ANM99,RS99,Ciardi03,Sokasian03,Kohler05,Iliev05}. Similarly,
4533: if a small fraction ($\la 1\%$) of the gas in each galaxy accretes onto a
4534: central black hole, then the resulting mini-quasars are also able to
4535: reionize the Universe, as has also been shown using semi-analytic models
4536: \cite{fk94,HL98,VS99,Wyithe03}.
4537: %Note that the prescription whereby a constant fraction of the galactic mass
4538: %accretes onto a central black hole is based on local observations which
4539: %indicate that $z=0$ galaxies harbor central black holes of mass equal to
4540: %$\sim 0.2$--$0.6\%$ of their bulge mass. Although the bulge constitutes
4541: %only a fraction of the total baryonic mass of each galaxy, the higher
4542: %gas-to-stellar mass ratio in high redshift galaxies, as well as their high
4543: %merger rates compared to their low redshift counterparts, suggest that a
4544: %fraction of a percent of the total gas mass in high-redshift galaxies may
4545: %have contributed to the formation of quasar black holes.
4546: 
4547: Although many models yield a reionization redshift around 7--12, the exact
4548: value depends on a number of uncertain parameters affecting both the source
4549: term and the recombination term in equation~(\ref{QIIeqn}). The source
4550: parameters include the formation efficiency of stars and quasars and the
4551: escape fraction of ionizing photons produced by these sources. The
4552: formation efficiency of low mass galaxies may also be reduced by feedback
4553: from galactic outflows. These parameters affecting the sources are
4554: discussed elsewhere in this review (see \S \ref{sec6.1}, and
4555: \ref{sec7}). Even when the clumping is inhomogeneous, the recombination
4556: term in equation~(\ref{QIIeqn}) is generally valid if $C$ is defined as in
4557: equation~(\ref{clump}), where we take a global volume average of the square
4558: of the density inside ionized regions (since neutral regions do not
4559: contribute to the recombination rate). The resulting mean clumping factor
4560: depends on the density and clustering of sources, and on the distribution
4561: and topology of density fluctuations in the IGM. Furthermore, the source
4562: halos should tend to form in overdense regions, and the clumping factor is
4563: affected by this cross-correlation between the sources and the IGM density.
4564: 
4565: Miralda-Escud\'e, Haehnelt, \& Rees (2000) \cite{MHR00} presented a simple model
4566: for the distribution of density fluctuations, and more generally they
4567: discussed the implications of inhomogeneous clumping during
4568: reionization. They noted that as ionized regions grow, they more
4569: easily extend into low-density regions, and they tend to leave behind
4570: high-density concentrations, with these neutral islands being ionized
4571: only at a later stage. They therefore argued that, since at
4572: high-redshift the collapse fraction is low, most of the high-density
4573: regions, which would dominate the clumping factor if they were
4574: ionized, will in fact remain neutral and occupy only a tiny fraction
4575: of the total volume. Thus, the development of reionization through the
4576: end of the overlap phase should occur almost exclusively in the
4577: low-density IGM, and the effective clumping factor during this time
4578: should be $\sim 1$, making recombinations relatively unimportant (see
4579: Fig. \ref{fig6c}). Only in the post-reionization phase,
4580: Miralda-Escud\'e et al.\ (2000) \cite{MHR00} argued, do the high density clouds and
4581: filaments become gradually ionized as the mean ionizing intensity
4582: further increases.
4583: 
4584: The complexity of the process of reionization is illustrated by the
4585: numerical simulation of Gnedin \cite {g00} of stellar
4586: reionization (in $\Lambda$CDM with $\Omm=0.3$). This simulation uses a
4587: formulation of radiative transfer which relies on several rough
4588: approximations; although it does not include the effect of shadowing behind
4589: optically-thick clumps, it does include for each point in the IGM the
4590: effects of an estimated local optical depth around that point, plus a local
4591: optical depth around each ionizing source. This simulation helps to
4592: understand the advantages of the various theoretical approaches, while
4593: pointing to the complications which are not included in the simple
4594: models. Figures \ref{fig6d} and \ref{fig6e}, taken from Figure 3 in 
4595: \cite{g00}, show the state of the simulated Universe just
4596: before and just after the overlap phase, respectively. They show a thin (15
4597: $h^{-1}$ comoving kpc) slice through the box, which is 4 $h^{-1}$ Mpc on a
4598: side, achieves a spatial resolution of $1 h^{-1}$ kpc, and uses $128^3$
4599: each of dark matter particles and baryonic particles (with each baryonic
4600: particle having a mass of $5\times 10^5 M_{\sun}$).  The figures show the
4601: redshift evolution of the mean ionizing intensity $J_{21}$ (upper right
4602: panel), and visually the logarithm of the neutral hydrogen fraction (upper
4603: left panel), the gas density (lower left panel), and the gas temperature
4604: (lower right panel). Note the obvious features resulting from the periodic
4605: boundary conditions assumed in the simulation. Also note that the intensity
4606: $J_{21}$ is defined as the intensity at the Lyman limit, expressed in units
4607: of $10^{-21}\, \mbox{ erg cm}^{-2} \mbox{ s}^{-1} \mbox{ sr}
4608: ^{-1}\mbox{Hz}^{-1}$. For a given source emission, the intensity inside H
4609: II regions depends on absorption and radiative transfer through the IGM
4610: (e.g., Haardt \& Madau 1996 \cite{HM96}; Abel \& Haehnelt 1999 \cite{AH99})
4611: 
4612: %%%%%%%% Figure 2
4613: \begin{figure}
4614: \centering
4615: \includegraphics[height=6cm]{VI4fig2.ps}
4616: \caption{Visualization at $z=7.7$ of a numerical simulation of
4617: reionization, adopted from Figure 3c of \cite{g00}. The panels
4618: display the logarithm of the neutral hydrogen fraction (upper left), the
4619: gas density (lower left), and the gas temperature (lower right). Also shown
4620: is the redshift evolution of the logarithm of the mean ionizing intensity
4621: (upper right). Note the periodic boundary conditions.}
4622: \label{fig6d}
4623: \end{figure}
4624:   
4625: %%%%%%%%
4626: 
4627: %%%%%%%% Figure6.5
4628: \begin{figure}
4629: \centering
4630: \includegraphics[height=6cm]{VI4fig3.ps}
4631: \caption{Visualization at $z=6.7$ of a numerical simulation of
4632: reionization, adopted from Figure 3e of \cite{g00}. The panels
4633: display the logarithm of the neutral hydrogen fraction (upper left), the
4634: gas density (lower left), and the gas temperature (lower right). Also shown
4635: is the redshift evolution of the logarithm of the mean ionizing intensity
4636: (upper right). Note the periodic boundary conditions.}
4637: \label{fig6e}
4638: \end{figure}
4639:   
4640: %%%%%%%%
4641: 
4642: Figure \ref{fig6d} shows the two-phase IGM at $z=7.7$, with ionized bubbles
4643: emanating from one main concentration of sources (located at the right edge
4644: of the image, vertically near the center; note the periodic boundary
4645: conditions). The bubbles are shown expanding into low density regions and
4646: beginning to overlap at the center of the image. The topology of ionized
4647: regions is clearly complex: While the ionized regions are analogous to
4648: islands in an ocean of neutral hydrogen, the islands themselves contain
4649: small lakes of dense neutral gas. One aspect which has not been included in
4650: theoretical models of clumping is clear from the figure. The sources
4651: themselves are located in the highest density regions (these being the
4652: sites where the earliest galaxies form) and must therefore ionize the gas
4653: in their immediate vicinity before the radiation can escape into the low
4654: density IGM. For this reason, the effective clumping factor is of order 100
4655: in the simulation and also, by the overlap redshift, roughly ten ionizing
4656: photons have been produced per baryon. Figure \ref{fig6e} shows that by
4657: $z=6.7$ the low density regions have all become highly ionized along with a
4658: rapid increase in the ionizing intensity. The only neutral islands left are
4659: the highest density regions (compare the two panels on the left). However,
4660: we emphasize that the quantitative results of this simulation must be
4661: considered preliminary, since the effects of increased resolution and a
4662: more accurate treatment of radiative transfer are yet to be
4663: explored. Methods are being developed for incorporating a more complete
4664: treatment of radiative transfer into three dimensional cosmological
4665: simulations (e.g., \cite{ANM99,RS99,Ciardi03,Sokasian03,Kohler05,Iliev05}).
4666: 
4667: Gnedin, Ferrara, \& Zweibel (2000) \cite{GFZ00} investigated an additional effect
4668: of reionization. They showed that the Biermann battery in cosmological
4669: ionization fronts inevitably generates coherent magnetic fields of an
4670: amplitude $\sim 10^{-19}$ Gauss. These fields form as a result of the
4671: breakout of the ionization fronts from galaxies and their propagation
4672: through the H I  filaments in the IGM. Although the fields are
4673: too small to directly affect galaxy formation, they could be the seeds
4674: for the magnetic fields observed in galaxies and X-ray clusters today.
4675: 
4676: If quasars contribute substantially to the ionizing intensity during
4677: reionization then several aspects of reionization are modified compared to
4678: the case of pure stellar reionization. First, the ionizing radiation
4679: emanates from a single, bright point-source inside each host galaxy, and
4680: can establish an escape route (H II  funnel) more easily than in the case
4681: of stars which are smoothly distributed throughout the galaxy (\S
4682: \ref{sec6.1}). Second, the hard photons produced by a quasar penetrate
4683: deeper into the surrounding neutral gas, yielding a thicker ionization
4684: front.  Finally, the quasar X-rays catalyze the formation of $H_2$
4685: molecules and allow stars to keep forming in very small halos.
4686: 
4687: Oh (1999) \cite{Oh99} showed that star-forming regions may also produce
4688: significant X-rays at high redshift. The emission is due to inverse Compton
4689: scattering of CMB photons off relativistic electrons in the ejecta, as well
4690: as thermal emission by the hot supernova remnant. The spectrum expected
4691: from this process is even harder than for typical quasars, and the hard
4692: photons photoionize the IGM efficiently by repeated secondary
4693: ionizations. The radiation, characterized by roughly equal energy per
4694: logarithmic frequency interval, would produce a uniform ionizing intensity
4695: and lead to gradual ionization and heating of the entire IGM. Thus, if this
4696: source of emission is indeed effective at high redshift, it may have a
4697: crucial impact in changing the topology of reionization. Even if stars
4698: dominate the emission, the hardness of the ionizing spectrum depends on the
4699: initial mass function. At high redshift it may be biased toward massive,
4700: efficiently ionizing stars, but this remains very much uncertain.
4701: 
4702: Semi-analytic as well as numerical models of reionization depend on an
4703: extrapolation of hierarchical models to higher redshifts and lower-mass
4704: halos than the regime where the models have been compared to observations
4705: (see e.g. \cite{Wyithe03,Cen03,Tumlinson04}). These models have the
4706: advantage that they are based on the current CDM paradigm which is
4707: supported by a variety of observations of large-scale structure, galaxy
4708: clustering, and the CMB. The disadvantage is that the properties of
4709: high-redshift galaxies are derived from those of their host halos by
4710: prescriptions which are based on low redshift observations, and these
4711: prescriptions will only be tested once abundant data is available on
4712: galaxies which formed during the reionization era (see \cite{Wyithe03} for
4713: the sensitivity of the results to model parameters). An alternative
4714: approach to analyzing the possible ionizing sources which brought about
4715: reionization is to extrapolate from the observed populations of galaxies
4716: and quasars at currently accessible redshifts. This has been attempted,
4717: e.g., by Madau et al.\ (1999) \cite{Mad99} and Miralda-Escud\'e et al.\ (2000) \cite{MHR00}. The
4718: general conclusion is that a high-redshift source population similar to the
4719: one observed at $z=3$--4 would produce roughly the needed ionizing
4720: intensity for reionization.  However, Dijkstra, Haiman, \& Loeb (2004)
4721: \cite{Dijkstra04} constrained the role of quasars in reionizing the
4722: Universe based on the unresolved flux of the X-ray background. At any
4723: event, a precise conclusion remains elusive because of the same kinds of
4724: uncertainties as those found in the models based on CDM: The typical escape
4725: fraction, and the faint end of the luminosity function, are both not well
4726: determined even at $z=3$--4, and in addition the clumping factor at high
4727: redshift must be known in order to determine the importance of
4728: recombinations. Future direct observations of the source population at
4729: redshifts approaching reionization may help resolve some of these
4730: questions.
4731: 
4732: 
4733: %<>
4734: \subsection{Photo-evaporation of Gaseous Halos After Reionization}
4735: \label{sec6.4}
4736: 
4737: The end of the reionization phase transition resulted in the emergence
4738: of an intense UV background that filled the Universe and heated the
4739: IGM to temperatures of $\sim 1$--$2\times 10^4$K (see the previous
4740: section). After ionizing the rarefied IGM in the voids and filaments
4741: on large scales, the cosmic UV background penetrated the denser
4742: regions associated with the virialized gaseous halos of the first
4743: generation of objects. A major fraction of the collapsed gas had been
4744: incorporated by that time into halos with a virial temperature $\la
4745: 10^4$K, where the lack of atomic cooling prevented the formation of
4746: galactic disks and stars or quasars. Photoionization heating by the
4747: cosmic UV background could then evaporate much of this gas back into
4748: the IGM. The photo-evaporating halos, as well as those halos which did
4749: retain their gas, may have had a number of important consequences just
4750: after reionization as well as at lower redshifts.
4751: 
4752: In this section we focus on the process by which gas that had already
4753: settled into virialized halos by the time of reionization was
4754: evaporated back into the IGM due to the cosmic UV background. This
4755: process was investigated by Barkana \& Loeb (1999) \cite{BL99} using 
4756: semi-analytic methods and idealized numerical calculations. They first considered an 
4757: isolated spherical, centrally-concentrated dark matter halo containing
4758: gas. Since most of the photo-evaporation occurs at the end of overlap,
4759: when the ionizing intensity builds up almost instantaneously, a sudden
4760: illumination by an external ionizing background may be assumed.
4761: Self-shielding of the gas implies that the halo interior sees a
4762: reduced intensity and a harder spectrum, since the outer gas layers
4763: preferentially block photons with energies just above the Lyman limit.
4764: It is useful to parameterize the external radiation field by a
4765: specific intensity per unit frequency, $\nu$, \beq J_{\nu}=10^{-21}\,
4766: J_{21}\, \left(\frac{\nu}{\nu_L}\right)^ {-\alpha}\mbox{ erg
4767: cm}^{-2}\mbox{ s}^{-1}\mbox{ sr} ^{-1}\mbox{Hz}^{-1}\ , \label{Inu}
4768: \eeq where $\nu_L$ is the Lyman limit frequency, and $J_{21}$ is the
4769: intensity at $\nu_L$ expressed in units of $10^{-21}\, \mbox{ erg
4770: cm}^{-2} \mbox{ s}^{-1} \mbox{ sr} ^{-1}\mbox{Hz}^{-1}$. The intensity
4771: is normalized to an expected post--reionization value of around unity
4772: for the ratio of ionizing photon density to the baryon density.
4773: Different power laws can be used to represent either quasar spectra
4774: ($\alpha \sim 1.8$) or stellar spectra ($\alpha \sim 5$).
4775: 
4776: Once the gas is heated throughout the halo, some fraction of it
4777: acquires a sufficiently high temperature that it becomes unbound. This
4778: gas expands due to the resulting pressure gradient and eventually
4779: evaporates back into the IGM. The pressure gradient force (per unit
4780: volume) 
4781: $\kB \nabla (T \rho/\mu m_p)$ competes with the gravitational force of
4782: $\rho\, G M/r^2$. Due to the density gradient, the ratio between the
4783: pressure force and the gravitational force is roughly equal to the ratio
4784: between the thermal energy $\sim \kB T$ and the gravitational binding
4785: energy $\sim \mu m_p G M/r$ (which is $\sim \kB T_{\rm vir}$ at the virial
4786: radius $r_{\rm vir}$) per particle. Thus, if the kinetic energy exceeds the
4787: potential energy (or roughly if $T>T_{\rm vir}$), the repulsive pressure
4788: gradient force exceeds the attractive gravitational force and expels the
4789: gas on a dynamical time (or faster for halos with $T\gg T_{\rm vir}$).
4790: 
4791:   
4792: The left panel of Figure \ref{fig6.7} (adopted from Fig. 3 of
4793: Barkana \& Loeb 1999 \cite{BL99}) shows the fraction of gas within the virial
4794: radius which becomes unbound after reionization, as a function of the
4795: total halo circular velocity, with halo masses at $z=8$ indicated at
4796: the top. The two pairs of curves correspond to spectral index
4797: $\alpha=5$ (solid) or $\alpha=1.8$ (dashed). In each pair, a
4798: calculation which assumes an optically-thin halo leads to the upper
4799: curve, but including radiative transfer and self-shielding modifies
4800: the result to the one shown by the lower curve. In each case
4801: self-shielding lowers the unbound fraction, but it mostly affects only
4802: a neutral core containing $\sim 30\%$ of the gas. Since high energy
4803: photons above the Lyman limit penetrate deep into the halo and heat
4804: the gas efficiently, a flattening of the spectral slope from
4805: $\alpha=5$ to $\alpha=1.8$ raises the unbound gas fraction. This
4806: figure is essentially independent of redshift if plotted in terms of
4807: circular velocity, but the conversion to a corresponding mass does
4808: vary with redshift. The characteristic circular velocity where most of
4809: the gas is lost is $\sim 10$--$15~{\rm km~s^{-1}}$, but clearly the
4810: effect of photo-evaporation is gradual, going from total gas removal
4811: down to no effect over a range of a factor of $\sim 100$ in halo mass.
4812: 
4813: %%%%%%%% Figure 1
4814: \begin{figure}
4815: \centering
4816: \includegraphics[height=6cm]{VI5fig1.ps}
4817: \includegraphics[height=6cm]{VI5fig2.ps}
4818: \caption{Effect of photo-evaporation on individual halos and on the
4819: overall halo population. The left panel shows the unbound gas fraction
4820: (within the virial radius) versus total halo velocity dispersion or
4821: mass, adopted from Figure 3 of Barkana \& Loeb (1999) \cite{BL99}. The two pairs
4822: of curves correspond to spectral index $\alpha=5$ (solid) or
4823: $\alpha=1.8$ (dashed), in each case at $z=8$. In each pair, assuming
4824: an optically-thin halo leads to the upper curve, while the lower curve
4825: shows the result of including radiative transfer and self
4826: shielding. The right panel shows the total fraction of gas in the
4827: Universe which evaporates from halos at reionization, versus the
4828: reionization redshift, adopted from Figure 7 of Barkana \& Loeb
4829: (1999) \cite{BL99}. The solid line assumes a spectral index $\alpha=1.8$, and the dotted line assumes 
4830: $\alpha=5$.}
4831: \label{fig6.7}
4832: \end{figure}
4833: %%%%%%%%
4834: 
4835: Given the values of the unbound gas fraction in halos of different masses,
4836: the Press-Schechter mass function (\S \ref{sec2.4}) can be used to
4837: calculate the total fraction of the IGM which goes through the process of
4838: accreting onto a halo and then being recycled into the IGM at
4839: reionization. The low-mass cutoff in this sum over halos is given by the
4840: lowest mass halo in which gas has assembled by the reionization
4841: redshift. This mass can be estimated by the linear Jeans mass $\mjeans$ in
4842: equation (\ref{eq:m_j}). The Jeans mass does not in general precisely equal
4843: the limiting mass for accretion (see the discussion in the next
4844: section). Indeed, at a given redshift some gas can continue to fall into
4845: halos of lower mass than the Jeans mass at that redshift. On the other
4846: hand, the larger Jeans mass at higher redshifts means that a time-averaged
4847: Jeans mass may be more appropriate, as indicated by the filtering mass. In
4848: practice, the Jeans mass is sufficiently accurate since at $z\sim 10$--20
4849: it agrees well with the values found in the numerical spherical collapse
4850: calculations of Haiman, Thoul, \& Loeb (1996) \cite{Haiman}.
4851: 
4852: The right panel of Figure  \ref{fig6.7} (adopted from Fig. 7 of
4853: Barkana \& Loeb 1999 \cite{BL99}) shows the total fraction of gas in the 
4854: Universe which evaporates from halos at reionization, versus the reionization
4855: redshift. The solid line assumes a spectral index $\alpha=1.8$, and
4856: the dotted line assumes $\alpha=5$, showing that the result is
4857: insensitive to the spectrum. Even at high redshift, the amount of gas
4858: which participates in photo-evaporation is significant, which suggests
4859: a number of possible implications as discussed below. The gas fraction
4860: shown in the figure represents most ($\sim 60$--$80\%$ depending on
4861: the redshift) of the collapsed fraction before reionization, although
4862: some gas does remain in more massive halos.
4863: 
4864: The photo-evaporation of gas out of large numbers of halos may have
4865: interesting implications. First, gas which falls into halos and is expelled
4866: at reionization attains a different entropy than if it had stayed in the
4867: low-density IGM. The resulting overall reduction in the entropy is expected
4868: to be small -- the same as would be produced by reducing the temperature of
4869: the entire IGM by a factor of $\sim 1.5$ -- but localized effects near
4870: photo-evaporating halos may be more significant. Furthermore, the resulting
4871: $\sim 20~{\rm km~s^{-1}}$ outflows induce small-scale fluctuations in
4872: peculiar velocity and temperature. These outflows are usually well below
4873: the resolution limit of most numerical simulations, but some outflows were
4874: resolved in the simulation of Bryan et al.\ (1998) \cite{BMAN98}. The
4875: evaporating halos may consume a significant number of ionizing photons in
4876: the post-overlap stage of reionization \cite{HAM00,Iliev05}, but a
4877: definitive determination requires detailed simulations which include the
4878: three-dimensional geometry of source halos and sink halos.
4879: 
4880: Although gas is quickly expelled out of the smallest halos,
4881: photo-evaporation occurs more gradually in larger halos which retain some
4882: of their gas. These surviving halos initially expand but they continue to
4883: accrete dark matter and to merge with other halos. These evaporating gas
4884: halos could contribute to the high column density end of the Ly$\alpha$
4885: forest \cite{BSS88}. Abel \& Mo (1998) \cite{AM98} suggested that, based on
4886: the expected number of surviving halos, a large fraction of the Lyman limit
4887: systems at $z\sim 3$ may correspond to mini-halos that survived
4888: reionization. Surviving halos may even have identifiable remnants in the
4889: present Universe. These ideas thus offer the possibility that a population
4890: of halos which originally formed prior to reionization may correspond
4891: almost directly to several populations that are observed much later in the
4892: history of the Universe. However, the detailed dynamics of
4893: photo-evaporating halos are complex, and detailed simulations are required
4894: to confirm these ideas.  Photo-evaporation of a gas cloud has been followed
4895: in a two dimensional simulation with radiative transfer, by Shapiro \& Raga
4896: (2000) \cite{SR00}. They found that an evaporating halo would indeed appear
4897: in absorption as a damped Ly$\alpha$ system initially, and as a weaker
4898: absorption system subsequently. Future simulations \cite{Iliev05} will
4899: clarify the contribution to quasar absorption lines of the entire
4900: population of photo-evaporating halos.
4901: 
4902: \subsection{Suppression of the Formation of Low Mass Galaxies}
4903: \label{sec6.5}
4904: 
4905: At the end of overlap, the cosmic ionizing background increased sharply,
4906: and the IGM was heated by the ionizing radiation to a temperature $\ga
4907: 10^4$ K. Due to the substantial increase in the IGM temperature, the
4908: intergalactic Jeans mass increased dramatically, changing the minimum mass
4909: of forming galaxies \cite{MJR86,EF92,G097,MR98}.
4910: 
4911: Gas infall depends sensitively on the Jeans mass. When a halo more massive
4912: than the Jeans mass begins to form, the gravity of its dark matter
4913: overcomes the gas pressure. Even in halos below the Jeans mass, although
4914: the gas is initially held up by pressure, once the dark matter collapses
4915: its increased gravity pulls in some gas \cite{Haiman}. Thus, the Jeans mass
4916: is generally higher than the actual limiting mass for accretion. Before
4917: reionization, the IGM is cold and neutral, and the Jeans mass plays a
4918: secondary role in limiting galaxy formation compared to cooling. After
4919: reionization, the Jeans mass is increased by several orders of magnitude
4920: due to the photoionization heating of the IGM, and hence begins to play a
4921: dominant role in limiting the formation of stars. Gas infall in a reionized
4922: and heated Universe has been investigated in a number of numerical
4923: simulations. Thoul \& Weinberg (1996) \cite{Th96} inferred, based on a
4924: spherically-symmetric collapse simulation, a reduction of $\sim 50\%$ in
4925: the collapsed gas mass due to heating, for a halo of circular velocity
4926: $V_c\sim 50\ {\rm km\ s}^{-1}$ at $z=2$, and a complete suppression of
4927: infall below $V_c \sim 30\ {\rm km\ s}^{-1}$. Kitayama \& Ikeuchi (2000)
4928: \cite{KI00} also performed spherically-symmetric simulations but included
4929: self-shielding of the gas, and found that it lowers the circular velocity
4930: thresholds by $\sim 5\ {\rm km\ s}^{-1}$. Three dimensional numerical
4931: simulations \cite{QKE96,WHK97,NS97} found a significant suppression of gas
4932: infall in even larger halos ($V_c \sim 75\ {\rm km\ s}^{-1}$), but this was
4933: mostly due to a suppression of late infall at $z\la 2$.
4934: 
4935: When a volume of the IGM is ionized by stars, the gas is heated to a
4936: temperature $T_{\rm IGM}\sim 10^4$ K. If quasars dominate the UV background
4937: at reionization, their harder photon spectrum leads to $T_{\rm IGM}>
4938: 2\times 10^4$ K. Including the effects of dark matter, a given temperature
4939: results in a linear Jeans mass corresponding to a halo circular velocity of
4940: \beq V_J=81 \left(\frac{T_{\rm IGM}}{1.5\times 10^4 {\rm K}}\right)^{1/2}\
4941: \left[\frac{1}{\Ommz}\ \frac{\Delta_c}{18 \pi^2}\right]^{1/6}\ {\rm km\
4942: s}^{-1}, \eeq where we used equation~(\ref{Vceqn}) and assumed
4943: $\mu=0.6$. In halos with $V_c>V_J$, the gas fraction in infalling gas
4944: equals the universal mean of $\Omega_b/\Omega_m$, but gas infall is
4945: suppressed in smaller halos. Even for a small dark matter halo, once it
4946: collapses to a virial overdensity of $\Delta_c/\Ommz$ relative to the mean,
4947: it can pull in additional gas. A simple estimate of the limiting circular
4948: velocity, below which halos have essentially no gas infall, is obtained by
4949: substituting the virial overdensity for the mean density in the definition
4950: of the Jeans mass. The resulting estimate is \beq V_{\rm lim}=34
4951: \left(\frac{T_{\rm IGM}}{1.5\times 10^4 {\rm K}}\right)^{1/2}\ {\rm km\
4952: s}^{-1}. \eeq This value is in rough agreement with the numerical
4953: simulations mentioned before.  A more recent study by Dijkstra et
4954: al. (2004) \cite{Dijkstra04} indicates that at the high redshifts of $z>10$ gas could
4955: nevertheless assemble into halos with circular velocities as low as
4956: $v_c\sim 10~{\rm km~s^{-1}}$, even in the presence of a UV background.
4957: 
4958: Although the Jeans mass is closely related to the rate of gas infall at a
4959: given time, it does not directly yield the total gas residing in halos at a
4960: given time. The latter quantity depends on the entire history of gas
4961: accretion onto halos, as well as on the merger histories of halos, and an
4962: accurate description must involve a time-averaged Jeans mass. Gnedin
4963: \cite{Gnedin2000b} showed that the gas content of halos in simulations is
4964: well fit by an expression which depends on the filtering mass, a particular
4965: time-averaged Jeans mass (Gnedin \& Hui 1998 \cite{Gnedin98}). Gnedin
4966: \cite{Gnedin2000b} calculated the Jeans and filtering masses using the mean
4967: temperature in the simulation to define the sound speed, and found the
4968: following fit to the simulation results: \beq \bar{M_g}=\frac{f_b M}{\left
4969: [1+ \left(2^{1/3}-1\right) M_C/M \right]^3}\ , \eeq where $\bar{M_g}$ is
4970: the average gas mass of all objects with a total mass $M$,
4971: $f_b=\Omega_b/\Omm$ is the universal baryon fraction, and the
4972: characteristic mass $M_C$ is the total mass of objects which on average
4973: retain $50\%$ of their gas mass. The characteristic mass was well fit by
4974: the filtering mass at a range of redshifts from $z=4$ up to $z\sim 15$.
4975: 
4976: The reionization process was not perfectly synchronized throughout the
4977: Universe. Large-scale regions with a higher density than the mean tend to
4978: form galaxies first and reionize earlier than underdense regions (see
4979: detailed discussion in \S \ref{scatter}). The suppression of low-mass
4980: galaxies by reionization will therefore be modulated by the fluctuations in
4981: the timing of reionization.  Babich \& Loeb (2005) \cite{Bab} considered
4982: the effect of inhomogeneous reionization on the power-spectrum of low-mass
4983: galaxies.  They showed that the shape of the high redshift galaxy power
4984: spectrum on small scales in a manner which depends on the details of epoch
4985: of reionization. This effect is significantly larger than changes in the
4986: galaxy power spectrum due to the current uncertainty in the inflationary
4987: parameters, such as the tilt of the scalar power spectrum $n$ and the
4988: running of the tilt $\alpha$.  Therefore, future high redshift galaxies
4989: surveys hoping to constrain inflationary parameters must properly model the
4990: effects of reionization, but conversely they will also be sensitive to the
4991: thermal history of the high redshift intergalactic medium.
4992: 
4993: %<>*************************************************************************
4994: 
4995: \section{\bf Feedback from Galactic Outflows}
4996: \label{sec7}
4997: 
4998: \subsection{Propagation of Supernova Outflows in the IGM}
4999: \label{sec7.1}
5000: 
5001: Star formation is accompanied by the violent death of massive stars in
5002: supernova explosions. In general, if each halo has a fixed baryon fraction
5003: and a fixed fraction of the baryons turns into massive stars, then the
5004: total energy in supernovae outflows is proportional to the halo mass. The
5005: binding energy of the gas in the halo is proportional to the halo mass
5006: squared. Thus, outflows are expected to escape more easily out of low-mass
5007: galaxies, and to expel a greater fraction of the gas from dwarf
5008: galaxies. At high redshifts, most galaxies form in relatively low-mass
5009: halos, and the high halo merger rate leads to vigorous star
5010: formation. Thus, outflows may have had a great impact on the earliest
5011: generations of galaxies, with consequences that may include metal
5012: enrichment of the IGM and the disruption of dwarf galaxies. In this
5013: subsection we present a simple model for the propagation of individual
5014: supernova shock fronts in the IGM. We discuss some implications of this
5015: model, but we defer to the following subsection the brunt of the discussion
5016: of the cosmological consequences of outflows.
5017: 
5018: For a galaxy forming in a given halo, the supernova rate is related to the
5019: star formation rate. In particular, for a Scalo (1998) \cite{Sca98} initial
5020: stellar mass function, if we assume that a supernova is produced by each
5021: $M>8 M_{\odot}$ star, then on average one supernova explodes for every 126
5022: $M_{\odot}$ of star formation, expelling an ejecta mass of $\sim 3\,
5023: M_{\odot}$ including $\sim 1\, M_{\odot}$ of heavy elements. We assume that
5024: the individual supernovae produce expanding hot bubbles which merge into a
5025: single overall region delineated by an outwardly moving shock front. We
5026: assume that most of the baryons in the outflow lie in a thin shell, while
5027: most of the thermal energy is carried by the hot interior. The total
5028: ejected mass equals a fraction $\fg$ of the total halo gas which is lifted
5029: out of the halo by the outflow. This gas mass includes a fraction $\fe$ of
5030: the mass of the supernova ejecta itself (with $\fe \le 1$ since some metals
5031: may be deposited in the disk and not ejected). Since at high redshift most
5032: of the halo gas is likely to have cooled onto a disk, we assume that the
5033: mass carried by the outflow remains constant until the shock front reaches
5034: the halo virial radius. We assume an average supernova energy of
5035: $10^{51}E_{51}$ erg, a fraction $\fw$ of which remains in the outflow after
5036: it escapes from the disk. The outflow must overcome the gravitational
5037: potential of the halo, which we assume to have a Navarro, Frenk, \& White
5038: (1997) \cite{Na97} density profile [NFW; see equation (\ref{NFW})]. Since the entire
5039: shell mass must be lifted out of the halo, we include the total shell mass
5040: as well as the total injected energy at the outset. This assumption is
5041: consistent with the fact that the burst of star formation in a halo is
5042: typically short compared to the total time for which the corresponding
5043: outflow expands.
5044: 
5045: The escape of an outflow from an NFW halo depends on the concentration
5046: parameter $\cN$ of the halo. Simulations by Bullock et al.\ (2000) \cite{Bu00}
5047: indicate that the concentration parameter decreases with redshift, and
5048: their results may be extrapolated to our regime of interest (i.e., to
5049: smaller halo masses and higher redshifts) by assuming that \beq
5050: \cN=\left(\frac{M}{10^9 M_{\sun}}\right) ^{-0.1}\, \frac{25}{(1+z)}\
5051: . \eeq Although we calculate below the dynamics of each outflow in
5052: detail, it is also useful to estimate which halos can generate
5053: large-scale outflows by comparing the kinetic energy of the outflow to
5054: the potential energy needed to completely escape (i.e., to infinite
5055: distance) from an NFW halo. We thus find that the outflow can escape
5056: from its originating halo if the circular velocity is below a critical
5057: value given by \beq V_{\rm crit} = 200 \sqrt{\frac{E_{51}\fw
5058: (\eta/0.1)} {\fg\ g(\cN)}}\ {\rm km\ s}^{-1} \ , \label{Vcrit} \eeq
5059: where the efficiency $\eta$ is the fraction of baryons incorporated in
5060: stars, and \beq g(x)=\frac{x^2}{(1+x)\ln(1+x)-x} \ .  \eeq Note that
5061: the contribution to $\fg$ of the supernova ejecta itself is $0.024
5062: \eta \fe$, so the ejecta mass is usually negligible unless $\fg \la
5063: 1\%$. Equation (\ref{Vcrit}) can also be used to yield the maximum gas
5064: fraction $\fg$ which can be ejected from halos, as a function of their
5065: circular velocity. Although this equation is most general, if we
5066: assume that the parameters $\fg$ and $\fw$ are independent of $M$ and
5067: $z$ then we can normalize them based on low-redshift observations. If
5068: we specify $\cN \sim 10$ (with $g(10)=6.1$) at $z=0$, then setting
5069: $E_{51}=1$ and $\eta=10\%$ yields the required energy efficiency as a
5070: function of the ejected halo gas fraction: \beq \fw = 1.5 \fg
5071: \left[\frac{V_{\rm crit}}{100\ {\rm km\ s}^{-1}} \right]^2\ . \eeq A
5072: value of $V_{\rm crit} \sim 100\ {\rm km\ s}^{-1}$ is suggested by
5073: several theoretical and observational arguments which are discussed in
5074: the next subsection. However, these arguments are not conclusive, and
5075: $V_{\rm crit}$ may differ from this value by a large factor,
5076: especially at high redshift (where outflows are observationally
5077: unconstrained at present). Note the degeneracy between $\fg$ and $\fw$
5078: which remains even if $V_{\rm crit}$ is specified. Thus, if $V_{\rm
5079: crit} \sim 100\ {\rm km\ s}^{-1}$ then a high efficiency $\fw \sim 1$
5080: is required to eject most of the gas from all halos with $V_c < V_{\rm
5081: crit}$, but only $\fw \sim 10\%$ is required to eject 5--10$\%$ of the
5082: gas. The evolution of the outflow does depend on the value of $\fw$
5083: and not just the ratio $\fw/\fg$, since the shell accumulates material
5084: from the IGM which eventually dominates over the initial mass carried
5085: by the outflow.
5086: 
5087: We solve numerically for the spherical expansion of a galactic outflow,
5088: elaborating on the basic approach of Tegmark, Silk, \& Evrard (1993)
5089: \cite{TSE93}. We assume that most of the mass $m$ carried along by the
5090: outflow lies in a thin, dense, relatively cool shell of proper radius
5091: $R$. The interior volume, while containing only a fraction $\fin \ll 1$ of
5092: the mass $m$, carries most of the thermal energy in a hot, isothermal
5093: plasma of pressure $p_{\rm int}$ and temperature $T$. We assume a uniform
5094: exterior gas, at the mean density of the Universe (at each redshift), which
5095: may be neutral or ionized, and may exert a pressure $p_{\rm ext}$ as
5096: indicated below. We also assume that the dark matter distribution follows
5097: the NFW profile out to the virial radius, and is at the mean density of the
5098: Universe outside the halo virial radius. Note that in reality an overdense
5099: distribution of gas as well as dark matter may surround each halo due to
5100: secondary infall.
5101: 
5102: The shell radius $R$ in general evolves as follows: \beq m \frac{d^2R}
5103: {dt^2}= 4 \pi R^2 \delta p-\left (\frac{dR}{dt} - H R\right) \frac{dm}
5104: {dt}- \frac{G m} {R^2} \left(M(R)+ \frac{1} {2}m \right) + \frac{8}{3}
5105: \pi G R m \rho_{\Lambda} \ , \eeq where the right-hand-side includes
5106: forces due to pressure, sweeping up of additional mass, gravity, and a
5107: cosmological constant, respectively.  The shell is accelerated by
5108: internal pressure and decelerated by external pressure, i.e., $\delta
5109: p=p_{\rm int}-p_{\rm ext}$. In the gravitational force, $M(R)$ is the
5110: total enclosed mass, not including matter in the shell, and $\frac{1}
5111: {2}m$ is the effective contribution of the shell mass in the
5112: thin-shell approximation \cite {OM88}. The interior
5113: pressure is determined by energy conservation, and evolves according
5114: to \cite{TSE93}: \beq \frac{d p_{\rm int}} {dt}=\frac{L}{2
5115: \pi R^3}-5\, \frac{p_{\rm int}} {R}\, \frac{d R}{dt} \ , \eeq where
5116: the luminosity $L$ incorporates heating and cooling terms. We include
5117: in $L$ the supernova luminosity $L_{\rm sn}$ (during a brief initial
5118: period of energy injection), cooling terms $L_{\rm cool}$, ionization
5119: $L_{\rm ion}$, and dissipation $L_{\rm diss}$. For simplicity, we
5120: assume ionization equilibrium for the interior plasma, and a
5121: primordial abundance of hydrogen and helium. We include in $L_{\rm
5122: cool}$ all relevant atomic cooling processes in hydrogen and helium,
5123: i.e., collisional processes, Bremsstrahlung emission, and Compton
5124: cooling off the CMB. Compton scattering is the dominant cooling
5125: process for high-redshift outflows. We include in $L_{\rm ion}$ only
5126: the power required to ionize the incoming hydrogen upstream, at the
5127: energy cost of 13.6 eV per hydrogen atom. The interaction between the
5128: expanding shell and the swept-up mass dissipates kinetic energy. The
5129: fraction $f_d$ of this energy which is re-injected into the interior
5130: depends on complex processes occurring near the shock front, including
5131: turbulence, non-equilibrium ionization and cooling, and so (following
5132: Tegmark et al.\ 1993 \cite{TSE93}) we let \beq L_{\rm diss}=\frac{1}{2} f_d
5133: \frac{dm}{dt} \left( \frac{dR}{dt} - H R \right)^2\ , \eeq where we
5134: set $f_d=1$ and compare below to the other extreme of $f_d=0$.
5135: 
5136: In an expanding Universe, it is preferable to describe the propagation of
5137: outflows in terms of comoving coordinates since, e.g., the critical result
5138: is the maximum {\it comoving}\, size of each outflow, since this size
5139: yields directly the total IGM mass which is displaced by the outflow and
5140: injected with metals. Specifically, we apply the following transformation
5141: \cite{Shan80,Voit96}: \beq d\hat{t}=a^{-2} dt,\ \ \hat{R}=a^{-1}R,\ \
5142: \hat{p}=a^5 p,\ \hat{\rho}= a^3 \rho\ . \eeq For $\Oml=0$, Voit (1996)
5143: \cite{Voit96} obtained (with the time origin $\hat{t}=0$ at redshift
5144: $z_1$): \beq \hat{t}=\frac{2}{\Omm H_0} \left[ \sqrt{1+\Omm
5145: z_1}-\sqrt{1+\Omm z}\, \right]\ , \eeq while for $\Omm+\Oml=1$ there is no
5146: simple analytic expression. We set $\beta=\hat{R}/\hat{r}_{\rm vir}$, in
5147: terms of the virial radius $r_{\rm vir}$ [equation (\ref{rvir})] of the
5148: source halo. We define $\alpha_S^1$ as the ratio of the shell mass $m$ to
5149: $\frac{4}{3} \pi \hat{\rho}_b\, \hat{r}_{\rm vir}^3$, where
5150: $\hat{\rho}_b=\rho_b(z=0)$ is the mean baryon density of the Universe at
5151: $z=0$. More generally, we define \beq \alpha_S(\beta) \equiv
5152: \frac{m}{\frac{4}{3} \pi \hat{\rho}_b \, \hat{\rho}^3} = \left\{
5153: \begin{array}{ll} \alpha_S^1 / \beta^3 & \mbox{if $\beta<1$} \\
5154: 1+\left(\alpha_S^1-1 \right)/\beta^3 & \mbox{otherwise.}
5155: \end{array} \right. \eeq Here we assumed, as noted above, that the shell
5156: mass is constant until the halo virial radius is reached, at which point
5157: the outflow begins to sweep up material from the IGM. We thus derive the
5158: following equations: \beq \frac{d^2 \hat{R}} {d \hat{t}^2}= \left\{
5159: \begin{array}{ll} \frac{3}{\alpha_S(\beta)} \frac{\hat{p}}{\hat{\rho}_b\,
5160: \hat{R} }-\frac{a}{2} \hat{R} H_0^2 \Omm \bar{\delta} (\beta) & \mbox{if
5161: $\beta<1$} \vspace{.1in} \\ \frac{3}{\alpha_S(\beta) \hat{R}}
5162: \left[\frac{\hat{p}}{\hat{\rho}_b}-\left( \frac{d\hat{R}} {d\hat{t}}
5163: \right)^2 \right]-\frac{a}{2} \hat{R} H_0^2 \Omm \bar{\delta} (\beta) +
5164: \frac{a}{4} \hat{R} H_0^2 \Omega_b \alpha_S(\beta) & \mbox{otherwise,}
5165: \end{array} \right. \eeq along with \beq \frac{d}{d\hat{t}} \left(
5166: \hat{R}^5 \hat{p}_{\rm int} \right)= \frac{a^4}{2 \pi} L \hat{R}^2\ . \eeq
5167: In the evolution equation for $\hat{R}$, for $\beta < 1$ we assume for
5168: simplicity that the baryons are distributed in the same way as the dark
5169: matter, since in any case the dark matter halo dominates the gravitational
5170: force. For $\beta>1$, however, we correct (via the last term on the
5171: right-hand side) for the presence of mass in the shell, since at $\beta \gg
5172: 1$ this term may become important. The $\beta > 1$ equation also includes
5173: the braking force due to the swept-up IGM mass. The enclosed mean
5174: overdensity for the NFW profile [Eq. (\ref{NFW})] surrounded by matter at
5175: the mean density is \beq \bar\delta(\beta)= \left\{
5176: \begin{array}{ll} \frac{ \Delta_c}{\Ommz \beta^3} \frac{\ln (1+\cN
5177: \beta) - \cN \beta/(1+ \cN \beta)} {\ln (1+c )-c/ (1+c)}& \mbox{if
5178: $\beta<1$} \vspace{.1in} \\ \left( \frac{ \Delta_c}{\Ommz}-1
5179: \right)\frac{1} {\beta^3} & \mbox{otherwise.}
5180: \end{array} \right. \eeq
5181: 
5182: The physics of supernova shells is discussed in Ostriker \& McKee (1988)
5183: \cite{OM88} along with a number of analytical solutions.  The propagation of
5184: cosmological blast waves has also been computed by Ostriker \& Cowie (1981)
5185: \cite{OC81}, Bertschinger (1985) \cite{Be85} and Carr \& Ikeuchi (1985)
5186: \cite{CI85}.  Voit (1996) \cite{Voit96} derived an exact analytic solution to the fluid equations
5187: which, although of limited validity, is nonetheless useful for understanding
5188: roughly how the outflow size depends on several of the parameters.  The solution
5189: requires an idealized case of an outflow which at all times expands into a
5190: homogeneous IGM.  Peculiar gravitational forces, and the energy lost in escaping
5191: from the host halo, are neglected, cooling and ionization losses are also assumed
5192: to be negligible, and the external pressure is not included.  The dissipated energy
5193: is assumed to be retained, i.e., $f_d$ is set equal to unity.  Under these
5194: conditions, the standard Sedov self-similar solution 
5195: \cite{Sed59,Sed93} generalizes to the cosmological case as follows 
5196: \cite{Voit96}:
5197: \beq \hat{R}=\left( \frac{\xi \hat{E}_0}{ \hat{\rho}_b} \right) ^{1/5}
5198: \hat{t}^{\,2/5}\ , \label{Voit} \eeq where $\xi=2.026$ and $\hat{E}_0=
5199: E_0/(1+z_1)^2$ in terms of the initial (i.e., at $t=\hat{t}=0$ and $z=z_1$) energy
5200: $E_0$.  Numerically, the comoving radius is \beq \hat{R}= 280 \left(
5201: \frac{0.022}{\Omega_b h^2}\, \frac{E_0}{10^{56}{\rm erg}} \right) ^{1/5}
5202: \left(\frac{10}{1+z_1} \, \frac{\hat{t}}{10^{10}{\rm yr}} \right) ^{2/5}\ {\rm kpc}\ .  \eeq
5203: 
5204: In solving the equations described above, we assume that the shock
5205: front expands into a pre-ionized region which then recombines after a
5206: time determined by the recombination rate. Thus, the external pressure
5207: is included initially, it is turned off after the pre-ionized region
5208: recombines, and it is then switched back on at a lower redshift when
5209: the Universe is reionized. When the ambient IGM is neutral and the
5210: pressure is off, the shock loses energy to ionization. In practice we
5211: find that the external pressure is unimportant during the initial
5212: expansion, although it {\it is}\, generally important after
5213: reionization. Also, at high redshift ionization losses are much
5214: smaller than losses due to Compton cooling. In the results shown
5215: below, we assume an instantaneous reionization at $z=9$.
5216: 
5217: Figure \ref{figVII1} shows the results for a starting redshift $z=15$, for
5218: a halo of mass $5.4 \times 10^7 M_{\sun}$, stellar mass $8.0 \times 10^5
5219: M_{\sun}$, comoving $\hat{r}_{\rm vir}=12$ kpc, and circular velocity
5220: $V_c=20$ km/s. We show the shell comoving radius in units of the virial
5221: radius of the source halo (top panel), and the physical peculiar velocity
5222: of the shock front (bottom panel). Results are shown (solid curve) for the
5223: standard set of parameters $\fin=0.1$, $f_d=1$, $\fw=75\%$, and
5224: $\fg=50\%$. For comparison, we show several cases which adopt the standard
5225: parameters except for no cooling (dotted curve), no reionization
5226: (short-dashed curve), $f_d=0$ (long-dashed curve), or $\fw=15\%$ and
5227: $\fg=10\%$ (dot-short dashed curve). When reionization is included, the
5228: external pressure halts the expanding bubble. We freeze the radius at the
5229: point of maximum expansion (where $d \hat{R}/d\hat{t}=0$), since in reality
5230: the shell will at that point begin to spread and fill out the interior
5231: volume due to small-scale velocities in the IGM. For the chosen parameters,
5232: the bubble easily escapes from the halo, but when $\fw$ and $\fg$ are
5233: decreased the accumulated IGM mass slows down the outflow more
5234: effectively. In all cases the outflow reaches a size of 10--20 times
5235: $\hat{r}_{\rm vir}$, i.e., 100--200 comoving kpc. If all the metals are
5236: ejected (i.e., $\fe=1$), then this translates to an average metallicity in
5237: the shell of $\sim 1$--5$\times 10^{-3}$ in units of the solar metallicity
5238: (which is $2\%$ by mass). The asymptotic size of the outflow varies roughly
5239: as $\fw^{1/5}$, as predicted by the simple solution in equation
5240: (\ref{Voit}), but the asymptotic size is rather insensitive to $\fg$ (at a
5241: fixed $\fw$) since the outflow mass becomes dominated by the swept-up IGM
5242: mass once $\hat{R} \ga 4 \hat{r}_{\rm vir}$. With the standard parameter
5243: values (i.e., those corresponding to the solid curve), Figure \ref{figVII1}
5244: also shows (dot-long dashed curve) the Voit (1996) \cite{Voit96} solution
5245: of equation (\ref{Voit}). The Voit solution behaves similarly to the
5246: no-reionization curve at low redshift, although it overestimates the shock
5247: radius by $\sim 30\%$, and the overestimate is greater compared to the more
5248: realistic case which does include reionization.
5249: 
5250: %%%%%%%% Figure VII1
5251: \begin{figure}
5252: \centering
5253: \includegraphics[height=6cm]{VIIfig1.ps}
5254: \caption{Evolution of a supernova outflow from a $z=15$ halo of circular
5255: velocity $V_c=20$ km/s. Plotted are the shell comoving radius in units of
5256: the virial radius of the source halo (top panel), and the physical peculiar
5257: velocity of the shock front (bottom panel). Results are shown for the
5258: standard parameters $\fin=0.1$, $f_d=1$, $\fw=75\%$, and $\fg=50\%$ (solid
5259: curve). Also shown for comparison are the cases of no cooling (dotted
5260: curve), no reionization (short-dashed curve), $f_d=0$ (long-dashed curve),
5261: or $\fw=15\%$ and $\fg=10\%$ (dot-short dashed curve), as well as the
5262: simple Voit (1996) \cite{Voit96} solution of equation (\ref{Voit}) for the
5263: standard parameter set (dot-long dashed curve). In cases where the outflow
5264: halts, we freeze the radius at the point of maximum expansion.}
5265: \label{figVII1}
5266: \end{figure}
5267: %%%%%%%%
5268: 
5269: Figure \ref{figVII2} shows different curves than Figure \ref{figVII1} but
5270: on an identical layout. A single curve starting at $z=15$ (solid curve) is
5271: repeated from Figure \ref{figVII1}, and it is compared here to outflows
5272: with the same parameters but starting at $z=20$ (dotted curve), $z=10$
5273: (short-dashed curve), and $z=5$ (long-dashed curve). A $V_c=20$ km/s halo,
5274: with a stellar mass equal to $1.5\%$ of the total halo mass, is chosen at
5275: the three higher redshifts, but at $z=5$ a $V_c=42$ km/s halo is
5276: assumed. Because of the suppression of gas infall after reionization, we
5277: assume that the $z=5$ outflow is produced by supernovae from a stellar mass
5278: equal to only $0.3\%$ of the total halo mass (with a similarly reduced
5279: initial shell mass), thus leading to a relatively small final shell
5280: radius. The main conclusion from both figures is the following: In all
5281: cases, the outflow undergoes a rapid initial expansion over a fractional
5282: redshift interval $\delta z/z \sim 0.2$, at which point the shell has
5283: slowed down to $\sim 10$ km/s from an initial 300 km/s. The rapid
5284: deceleration is due to the accumulating IGM mass. External pressure from
5285: the reionized IGM completely halts all high-redshift outflows, and even
5286: without this effect most outflows would only move at $\sim 10$ km/s after
5287: the brief initial expansion. Thus, it may be possible for high-redshift
5288: outflows to pollute the Lyman alpha forest with metals without affecting
5289: the forest hydrodynamically at $z \la 4$. While the bulk velocities of
5290: these outflows may dissipate quickly, the outflows do sweep away the IGM
5291: and create empty bubbles. The resulting effects on observations of the
5292: Lyman alpha forest should be studied in detail (some observational
5293: signatures of feedback have been suggested recently by Theuns, Mo, \&
5294: Schaye 2000 \cite{TMS00}).
5295: 
5296: %%%%%%%% Figure VII2
5297: \begin{figure}
5298: \centering
5299: \includegraphics[height=6cm]{VIIfig2.ps}
5300: \caption{Evolution of supernova outflows at different redshifts. The
5301: top and bottom panels are arranged similarly to Figure \ref{figVII1}.
5302: The $z=15$ outflow (solid curve) is repeated from Figure
5303: \ref{figVII1}, and it is compared here to outflows with the same
5304: parameters but starting at $z=20$ (dotted curve), $z=10$ (short-dashed
5305: curve), and $z=5$ (long-dashed curve). A $V_c=20$ km/s halo is assumed
5306: except for $z=5$, in which case a $V_c=42$ km/s halo is assumed to
5307: produce the outflow (see text).}
5308: \label{figVII2}
5309: \end{figure}
5310: %%%%%%%%
5311: 
5312: Furlanetto \& Loeb (2003) \cite{FL03} derived the evolution of the
5313: characteristic scale and filling fraction of supernova-driven bubbles based
5314: on a refinement of this formalism (see also their 2001 paper for
5315: quasar-driven outflows).  The role of metal-rich outflows in smearing the
5316: transition epoch between Pop-III (metal-free) and Pop II (metal-enriched)
5317: stars, was also analysed by Furlanetto \& Loeb (2005) \cite{FL05}, who
5318: concluded that a double-reionization history in which the ionization
5319: fraction goes through two (or more) peaks is unlikely.
5320: 
5321: %<>
5322: 
5323: \subsection{Effect of Outflows on Dwarf Galaxies and on the IGM}
5324: \label{sec7.2}
5325: 
5326: Galactic outflows represent a complex feedback process which affects
5327: the evolution of cosmic gas through a variety of phenomena. Outflows
5328: inject hydrodynamic energy into the interstellar medium of their host
5329: galaxy. As shown in the previous subsection, even a small fraction of
5330: this energy suffices to eject most of the gas from a dwarf galaxy,
5331: perhaps quenching further star formation after the initial burst. At
5332: the same time, the enriched gas in outflows can mix with the
5333: interstellar medium and with the surrounding IGM, allowing later
5334: generations of stars to form more easily because of metal-enhanced
5335: cooling. On the other hand, the expanding shock waves may also strip
5336: gas in surrounding galaxies and suppress star formation.
5337: 
5338: Dekel \& Silk (1986) \cite{DS86} attempted to explain the different
5339: properties of diffuse dwarf galaxies in terms of the effect of galactic
5340: outflows. They noted the observed trends whereby lower-mass dwarf galaxies
5341: have a lower surface brightness and metallicity, but a higher mass-to-light
5342: ratio, than higher mass galaxies. They argued that these trends are most
5343: naturally explained by substantial gas removal from an underlying dark
5344: matter potential. Galaxies lying in small halos can eject their remaining
5345: gas after only a tiny fraction of the gas has turned into stars, while
5346: larger galaxies require more substantial star formation before the
5347: resulting outflows can expel the rest of the gas. Assuming a wind
5348: efficiency $\fw \sim 100\%$, Dekel \& Silk showed that outflows in halos
5349: below a circular velocity threshold of $V_{\rm crit} \sim 100$ km/s have
5350: sufficient energy to expel most of the halo gas. Furthermore, cooling is
5351: very efficient for the characteristic gas temperatures associated with
5352: $V_{\rm crit} \la 100$ km/s halos, but it becomes less efficient in more
5353: massive halos. As a result, this critical velocity is expected to signify a
5354: dividing line between bright galaxies and diffuse dwarf galaxies. Although
5355: these simple considerations may explain a number of observed trends, many
5356: details are still not conclusively determined. For instance, even in
5357: galaxies with sufficient energy to expel the gas, it is possible that this
5358: energy gets deposited in only a small fraction of the gas, leaving the rest
5359: almost unaffected.
5360: 
5361: Since supernova explosions in an inhomogeneous interstellar medium lead to
5362: complicated hydrodynamics, in principle the best way to determine the basic
5363: parameters discussed in the previous subsection ($\fw$, $\fg$, and $\fe$)
5364: is through detailed numerical simulations of individual galaxies. Mac Low
5365: \& Ferrara (1999) \cite{Mac99} simulated a gas disk within a $z=0$ dark
5366: matter halo. The disk was assumed to be azimuthally symmetric and initially
5367: smooth. They represented supernovae by a central source of energy and mass,
5368: assuming a constant luminosity which is maintained for 50 million
5369: years. They found that the hot, metal-enriched ejecta can in general escape
5370: from the halo much more easily than the colder gas within the disk, since
5371: the hot gas is ejected in a tube perpendicular to the disk without
5372: displacing most of the gas in the disk. In particular, most of the metals
5373: were expelled except for the case with the most massive halo considered
5374: (with $10^9 M_{\sun}$ in gas) and the lowest luminosity ($10^{37}$ erg/s,
5375: or a total injection of $2 \times 10^{52}$ erg). On the other hand, only a
5376: small fraction of the total gas mass was ejected except for the least
5377: massive halo (with $10^6 M_{\sun}$ in gas), where a luminosity of $10^{38}$
5378: erg/s or more expelled most of the gas. We note that beyond the standard
5379: issues of numerical resolution and convergence, there are several
5380: difficulties in applying these results to high-redshift dwarf
5381: galaxies. Clumping within the expanding shells or the ambient interstellar
5382: medium may strongly affect both the cooling and the hydrodynamics. Also,
5383: the effect of distributing the star formation throughout the disk is
5384: unclear since in that case several characteristics of the problem will
5385: change; many small explosions will distribute the same energy over a larger
5386: gas volume than a single large explosion [as in the Sedov (1959)
5387: \cite{Sed59} solution; see, e.g., equation (\ref{Voit})], and the geometry
5388: will be different as each bubble tries to dig its own escape route through
5389: the disk. Also, high-redshift disks should be denser by orders of magnitude
5390: than $z=0$ disks, due to the higher mean density of the Universe at early
5391: times.  Thus, further numerical simulations of this process are required in
5392: order to assess its significance during the reionization epoch.
5393: 
5394: Some input on these issues also comes from observations. Martin (1999)
5395: \cite{Mart99} showed that the hottest extended X-ray emission in galaxies
5396: is characterized by a temperature of $\sim 10^{6.7}$ K. This hot gas, which
5397: is lifted out of the disk at a rate comparable to the rate at which gas
5398: goes into new stars, could escape from galaxies with rotation speeds of
5399: $\la 130$ km/s. However, these results are based on a small sample which
5400: includes only the most vigorous star-forming local galaxies, and the
5401: mass-loss rate depends on assumptions about the poorly understood transfer
5402: of mass and energy among the various phases of the interstellar medium.
5403: 
5404: Many authors have attempted to estimate the overall cosmological effects of
5405: outflows by combining simple models of individual outflows with the
5406: formation rate of galaxies, obtained via semi-analytic methods
5407: \cite{CR86,TSE93,Voit96,NT97,FPS00,SB00} or numerical simulations
5408: \cite{G097,Gne98,CO99,AHWKG2000a}. The main goal of these calculations is
5409: to explain the characteristic metallicities of different environments as a
5410: function of redshift. For example, the IGM is observed to be enriched with
5411: metals at redshifts $z\la 5$. Identification of C IV, Si IV an O VI
5412: absorption lines which correspond to Ly$\alpha$ absorption lines in the
5413: spectra of high-redshift quasars has revealed that the low-density IGM has
5414: been enriched to a metal abundance (by mass) of $Z_{\rm IGM}\sim 10^{-2.5
5415: (\pm 0.5)}Z_\odot$, where $Z_\odot=0.019$ is the solar metallicity
5416: \cite{MY87,Tyt95,SC96,LSBR98,CS98,Song97,ESSP00}. The metal enrichment has
5417: been clearly identified down to H I column densities of $\sim
5418: 10^{14.5}~{\rm cm^{-2}}$. The detailed comparison of cosmological
5419: hydrodynamic simulations with quasar absorption spectra has established
5420: that the forest of Ly$\alpha$ absorption lines is caused by the
5421: smoothly-fluctuating density of the neutral component of the IGM
5422: \cite{CMOR94,ZAN95,HKWM96}. The simulations show a strong correlation
5423: between the H I column density and the gas overdensity $\delta_{\rm gas}$
5424: \cite{DHKW99}, implying that metals were dispersed into regions with an
5425: overdensity as low as $\delta_{\rm gas}\sim 3$ or possibly even lower.
5426: 
5427: In general, dwarf galaxies are expected to dominate metal enrichment at
5428: high-redshift for several reasons. As noted above and in the previous
5429: subsection, outflows can escape more easily out of the potential wells of
5430: dwarfs. Also, at high redshift, massive halos are rare and dwarf halos are
5431: much more common. Finally, as already noted, the Sedov (1959) \cite{Sed59}
5432: solution [or equation (\ref{Voit})] implies that for a given total energy
5433: and expansion time, multiple small outflows fill large volumes more
5434: effectively than would a smaller number of large outflows. Note, however,
5435: that the strong effect of feedback in dwarf galaxies may also quench star
5436: formation rapidly and reduce the efficiency of star formation in dwarfs
5437: below that found in more massive galaxies.
5438: 
5439: Cen \& Ostriker (1999) \cite{CO99} showed via numerical simulation that
5440: metals produced by supernovae do not mix uniformly over cosmological
5441: volumes.  Instead, at each epoch the highest density regions have much
5442: higher metallicity than the low-density IGM.  They noted that early star
5443: formation occurs in the most overdense regions, which therefore reach a
5444: high metallicity (of order a tenth of the solar value) by $z \sim 3$, when
5445: the IGM metallicity is lower by 1--2 orders of magnitude.  At later times,
5446: the formation of high-temperature clusters in the highest-density regions
5447: suppresses star formation there, while lower-density regions continue to
5448: increase their metallicity.  Note, however, that the spatial resolution of
5449: the hydrodynamic code of Cen \& Ostriker is a few hundred kpc, and anything
5450: occurring on smaller scales is inserted directly via simple parametrized
5451: models.  Scannapieco \& Broadhurst (2000) \cite{SB00} implemented expanding
5452: outflows within a numerical scheme which, while not a full gravitational
5453: simulation, did include spatial correlations among halos.  They showed that
5454: winds from low-mass galaxies may also strip gas from nearby galaxies (see
5455: also Scannapieco, Ferrara, \& Broadhurst 2000 \cite{SFB00}), thus
5456: suppressing star formation in a local neighborhood and substantially
5457: reducing the overall abundance of galaxies in halos below a mass of $\sim
5458: 10^{10} M_{\sun}$.  Although quasars do not produce metals, they may also
5459: affect galaxy formation in their vicinity via energetic outflows
5460: \cite{ER88,BW91,Sil98,NSS98}.
5461: 
5462: Gnedin \& Ostriker (1997) \cite{G097} and Gnedin (1998) \cite{Gne98}
5463: identified another mixing mechanism which, they argued, may be dominant at
5464: high redshift ($z \ga 4$).  In a collision between two protogalaxies, the
5465: gas components collide in a shock and the resulting pressure force can
5466: eject a few percent of the gas out of the merger remnant.  This is the
5467: merger mechanism, which is based on gravity and hydrodynamics rather than
5468: direct stellar feedback.  Even if supernovae inject most of their metals in
5469: a local region, larger-scale mixing can occur via mergers.  Note, however,
5470: that Gnedin's (1998) \cite{Gne98} simulation assumed a comoving star
5471: formation rate at $z \ga 5$ of $\sim 1 M_{\sun}$ per year per comoving
5472: Mpc$^3$, which is 5--10 times larger than the observed rate at redshift
5473: 3--4.  Aguirre et al.\ \cite{AHWKG2000a} used outflows implemented in
5474: simulations to conclude that winds of $\sim 300$ km/s at $z \la 6$ can
5475: produce the mean metallicity observed at $z \sim 3$ in the Ly$\alpha$
5476: forest.  In a separate paper Aguirre et al. \cite{AHKGW2000b} explored
5477: another process, where metals in the form of dust grains are driven to
5478: large distances by radiation pressure, thus producing large-scale mixing
5479: without displacing or heating large volumes of IGM gas.  The success of
5480: this mechanism depends on detailed microphysics such as dust grain
5481: destruction and the effect of magnetic fields.  The scenario, though, may
5482: be directly testable because it leads to significant ejection only of
5483: elements which solidify as grains.
5484: 
5485: Feedback from galactic outflows encompasses a large variety of processes
5486: and influences. The large range of scales involved, from stars or quasars
5487: embedded in the interstellar medium up to the enriched IGM on cosmological
5488: scales, make possible a multitude of different, complementary approaches,
5489: promising to keep galactic feedback an active field of research.
5490: 
5491: 
5492: \section{The Frontier of 21cm Cosmology}
5493: 
5494: \subsection{Mapping Hydrogen Before Reionization}
5495: 
5496: The small residual fraction of free electrons after cosmological
5497: recombination coupled the temperature of the cosmic gas to that of the
5498: cosmic microwave background (CMB) down to a redshift, $z\sim 200$
5499: \cite{Peebles}. Subsequently, the gas temperature dropped adiabatically as
5500: $T_{\rm gas}\propto (1+z)^2$ below the CMB temperature $T_{\gamma}\propto
5501: (1+z)$.  The gas heated up again after being exposed to the photo-ionizing
5502: ultraviolet light emitted by the first stars during the {\it reionization
5503: epoch} at $z\lesssim 20$. Prior to the formation of the first stars, the
5504: cosmic neutral hydrogen must have resonantly absorbed the CMB flux through
5505: its spin-flip 21cm transition \cite{Field,Scott,Tozzi,Zalda04}.  The linear
5506: density fluctuations at that time should have imprinted anisotropies on the
5507: CMB sky at an observed wavelength of $\lambda=21.12[(1+z)/100]$ meters. We
5508: discuss these early 21cm fluctuations mainly for pedagogical purposes.
5509: Detection of the earliest 21cm signal will be particularly challenging
5510: because the foreground sky brightness rises as $\lambda^{2.5}$ at long
5511: wavelengths in addition to the standard $\sqrt{\lambda}$ scaling of the
5512: detector noise temperature for a given integration time and fractional
5513: bandwidth.  The discussion in this section follows Loeb \& Zaldarriaga
5514: (2004) \cite{Loeb04}.
5515: 
5516: 
5517: \begin{figure} 
5518: \centering
5519: \includegraphics[width=0.7\columnwidth,angle=-90]{21cm.ps} 
5520: \caption{The 21cm transition of hydrogen. The higher energy level the spin
5521: of the electron (e-) is aligned with that of the proton (p+).  A spin flip
5522: results in the emission of a photon with a wavelength of 21cm (or a
5523: frequency of 1420MHz).}
5524: \label{21cm}
5525: \end{figure}
5526: 
5527: We start by calculating the history of the spin temperature, $T_s$, defined
5528: through the ratio between the number densities of hydrogen atoms in the
5529: excited and ground state levels, ${n_1/ n_0}=(g_1/
5530: g_0)\exp\left\{-{T_\star/ T_s}\right\},$ 
5531: \begin{equation} 
5532: {n_1\over
5533: n_0}={g_1\over g_0}\exp\left\{-{T_\star\over T_s}\right\}, 
5534: \label{eq:spin}
5535: \end{equation} 
5536: where subscripts $1$ and $0$ correspond to the excited and
5537: ground state levels of the 21cm transition, $(g_1/g_0)=3$ is the ratio of
5538: the spin degeneracy factors of the levels, $n_{\rm H}=(n_0+n_1)\propto
5539: (1+z)^3$ is the total hydrogen density, and $T_\star=0.068$K is the
5540: temperature corresponding to the energy difference between the levels.  The
5541: time evolution of the density of atoms in the ground state is given by,
5542: \begin{eqnarray} \left( {\partial\over \partial t} + 
5543: 3{{\dot a}\over a} \right) & n_0 & =-n_0\left(C_{01}+B_{01}I_\nu\right)
5544: \nonumber \\ + & n_1 & \left(C_{10}+A_{10}+B_{10}I_\nu\right),
5545: \label{eq:evolution} \end{eqnarray} where $a(t)=(1+z)^{-1}$ is the cosmic
5546: scale factor, $A$'s and $B$'s are the Einstein rate coefficients, $C$'s are
5547: the collisional rate coefficients, and $I_\nu$ is the blackbody intensity
5548: in the Rayleigh-Jeans tail of the CMB, namely
5549: $I_\nu=2kT_{\gamma}/\lambda^2$ with $\lambda=21$ cm \cite{RL}.  Here a dot
5550: denotes a time-derivative.  The $0\rightarrow 1$ transition rates can be
5551: related to the $1\rightarrow 0$ transition rates by the requirement that in
5552: thermal equilibrium with $T_s=T_\gamma=T_{\rm gas}$, the right-hand-side of
5553: Eq. (\ref{eq:evolution}) should vanish with the collisional terms balancing
5554: each other separately from the radiative terms. The Einstein coefficients
5555: are $A_{10}=2.85\times 10^{-15}~{\rm s^{-1}}$, $B_{10}=(\lambda^3/2hc)
5556: A_{10}$ and $B_{01}=(g_1/g_0)B_{10}$ \cite{Field,RL}.  The collisional
5557: de-excitation rates can be written as $C_{10}={4\over 3} \kappa(1-0) n_{\rm
5558: H}$, where $\kappa(1-0)$ is tabulated as a function of $T_{\rm gas}$
5559: \cite{AD,Zyg}.
5560: 
5561: Equation (\ref{eq:evolution}) can
5562: be simplified to the form,
5563: \begin{eqnarray}
5564: {d\Upsilon \over dz} & = & -\left[H(1+z)\right]^{-1}
5565: \left[-\Upsilon(C_{01}+B_{01}I_\nu) \right. \nonumber \\ 
5566: && \left. +
5567: (1-\Upsilon)(C_{10}+A_{10}  + B_{10}I_\nu)\right],
5568: \label{eq:upsilon}
5569: \end{eqnarray}
5570: where $\Upsilon\equiv n_0/n_{\rm H}$, $H\approx
5571: H_0\sqrt{\Omega_m}(1+z)^{3/2}$ is the Hubble parameter at high redshifts
5572: (with a present-day value of $H_0$), and $\Omega_m$ is the density
5573: parameter of matter.  The upper panel of Fig. \ref{dtb} shows the results
5574: of integrating Eq.~(\ref{eq:upsilon}). Both the spin temperature and the
5575: kinetic temperature of the gas track the CMB temperature down to $z\sim
5576: 200$. Collisions are efficient at coupling $T_s$ and $T_{gas}$ down to
5577: $z\sim 70$ and so the spin temperature follows the kinetic temperature
5578: around that redshift. At much lower redshifts, the Hubble expansion makes
5579: the collision rate subdominant relative the radiative coupling rate to the
5580: CMB, and so $T_s$ tracks $T_{\gamma}$ again. Consequently, there is a
5581: redshift window between $30\lesssim z \lesssim 200$, during which the
5582: cosmic hydrogen absorbs the CMB flux at its resonant 21cm
5583: transition. Coincidentally, this redshift interval precedes the appearance
5584: of collapsed objects \cite{BL01} and so its signatures are not contaminated by
5585: nonlinear density structures or by radiative or hydrodynamic feedback
5586: effects from stars and quasars, as is the case at lower redshifts
5587: \cite{Zalda04}.
5588: 
5589: 
5590: \begin{figure}[th]
5591: \centering
5592: \includegraphics[height=6cm]{fig21_1.ps}
5593: \caption{{\it Upper panel:} Evolution of the gas, CMB and spin temperatures
5594: with redshift [4].  {\it Lower panel:} ${dT_b/d\dh}$ as function of
5595: redshift.  The separate contributions from fluctuations in the density and
5596: the spin temperature are depicted. We also show ${dT_b/d\dh} a \propto
5597: {dT_b/d\dh} \times \dh$, with an arbitrary normalization.}
5598: \label{dtb}
5599: \end{figure}
5600: 
5601: During the period when the spin temperature is smaller than the CMB
5602: temperature, neutral hydrogen atoms absorb CMB photons. The resonant 21cm
5603: absorption reduces the brightness temperature of the CMB by,
5604: \begin{equation}
5605: T_b =\tau \left( T_s-T_{\gamma}\right)/(1+z) ,
5606: \end{equation}
5607: where the optical depth for resonant 21cm absorption is,
5608: \begin{equation}
5609: \tau= {3c\lambda^2hA_{10}n_{\rm H}\over 32 \pi k T_s H(z)} .
5610: \label{eq:tau} 
5611: \end{equation}
5612: 
5613: Small inhomogeneities in the hydrogen density $\dh\equiv (n_{\rm H}-{\bar
5614: n_{\rm H}})/{\bar n}_{\rm H}$ result in fluctuations of the 21cm absorption
5615: through two separate effects. An excess of neutral hydrogen directly
5616: increases the optical depth and also alters the evolution of the spin
5617: temperature. For now, we ignore the additional effects of peculiar
5618: velocities (Bharadwaj \& Ali 2004 \cite{Indian}; Barkana \& Loeb 2004 \cite{BL04a}) as well as
5619: fluctuations in the gas kinetic temperature due to the adiabatic
5620: compression (rarefaction) in overdense (underdense) regions \cite{BLinf}.
5621: Under these approximations, we can write an equation for the resulting
5622: evolution of $\Upsilon$ fluctuations,
5623: \begin{eqnarray}
5624: {d\delta \Upsilon \over dz} & = & \left[H(1+z)\right]^{-1}
5625: \left\{[C_{10}+C_{01}+(B_{01}+B_{10})I_\nu]\delta\Upsilon \right. \nonumber \\
5626:  && \left. + \left[ C_{01} \Upsilon
5627: -C_{10}(1-\Upsilon)\right]\dh\right\},
5628: \label{eq:dup}
5629: \end{eqnarray}
5630: leading to spin temperature fluctuations,
5631: \begin{equation}
5632: {\delta T_s\over {\bar T}_s}= -{1\over
5633: \ln[3\Upsilon/(1-\Upsilon)]}{\delta\Upsilon\over \Upsilon (1-\Upsilon)}.
5634: \label{eq:dts}
5635: \end{equation}
5636: The resulting brightness temperature fluctuations can be related to the derivative,
5637: \begin{equation}
5638: {\delta T_b\over {\bar T}_b}=\dh +  {T_{\gamma}\over ({\bar T}_s-T_{\gamma})}
5639: {\delta T_s\over {\bar T}_s}.
5640: \label{eq:dtb}
5641: \end{equation}
5642: The spin temperature fluctuations ${\delta T_s/ {T}_s}$ are proportional to
5643: the density fluctuations and so we define, 
5644: \beq
5645: \label{weight} 
5646: {d T_b \over d \dh} \equiv {\bar T}_b + {T_{\gamma} {\bar T}_b
5647: \over ({\bar T}_s-T_{\gamma})} {\delta T_s\over {\bar T}_s \dh}, 
5648: \eeq
5649: through ${\delta T_b}=({d T_b /d \dh}) \dh$.  We ignore fluctuations
5650: in $C_{ij}$ due to
5651: fluctuations in $T_{\rm gas}$ which are very small \cite{AD}.
5652: Figure \ref{dtb} shows ${dT_b/d\dh}$ as a function of redshift, including
5653: the two contributions to ${dT_b/d\dh}$, one originating directly from
5654: density fluctuations and the second from the associated changes in the spin
5655: temperature \cite{Scott}.  Both contributions have the same sign, because
5656: an increase in density raises the collision rate and lowers the spin
5657: temperature and so it allows $T_s$ to better track $T_{\rm gas}$.  Since
5658: $\dh$ grows with time as $\dh \propto a$, the signal peaks at $z\sim 50$, a
5659: slightly lower redshift than the peak of ${dT_b/d\dh}$.
5660: 
5661: Next we calculate the angular power spectrum of the brightness temperature
5662: on the sky, resulting from density perturbations with a power spectrum
5663: $P_{\delta}(k)$, 
5664: \beq
5665: \label{pdel} \langle \dh(\k_1) \dh(\k_2) \rangle = (2\pi)^3
5666: \delta^D(\k_1+\k_2) P_{\delta}(k_1).  \eeq where $\dh(\k)$ is the Fourier
5667: tansform of the hydrogen density field, $\k$ is the comoving wavevector,
5668: and $\langle \cdots \rangle$ denotes an ensemble average (following the
5669: formalism described in \cite{Zalda04}).  The 21cm brightness temperature
5670: observed at a frequency $\nu$ corresponding to a distance $r$ along the
5671: line of sight, is given by \beq
5672: \label{los} \delta T_b(\n,\nu)=\int dr W_{\nu}(r)
5673: \ {dT_b \over d\dh} \dh(\n,r), \eeq where $\n$ denotes the direction of
5674: observation, $W_{\nu}(r)$ is a narrow function of $r$ that peaks at the
5675: distance corresponding to $\nu$. The details of this function depend on the
5676: characteristics of the experiment.  The brightness fluctuations in Eq.
5677: \ref{los}
5678: can be expanded in spherical harmonics with expansion coefficients
5679: $a_{lm}(\nu)$. The angular power spectrum of map $C_{l}(\nu)= \langle
5680: |a_{lm}(\nu)|^2 \rangle$ can be expressed in terms of the 3D power spectrum
5681: of fluctuations in the density $P_{\delta}(k)$, \beqa
5682: \label{fcldef} C_{l}(\nu)&=&4
5683: \pi \int {d^3k \over (2\pi)^3} P_{\delta}(k) \alpha_l^2(k,\nu) \nonumber \\
5684: \alpha_l(k,\nu)&=& \int dr W_{r_0}(r) {dT_b\over d\dh}(r) j_l(kr).  \eeqa
5685: Our calculation ignores inhomogeneities in the hydrogen ionization
5686: fraction, since they freeze at the earlier recombination epoch ($z\sim
5687: 10^3$) and so their amplitude is more than an order of magnitude smaller
5688: than $\dh$ at $z\lesssim 100$.  The gravitational
5689: potential perturbations induce a redshift distortion effect that is of
5690: order $\sim (H/ck)^2$ smaller than $\dh$ for the
5691: high--$l$ modes of interest here.  
5692: 
5693: \begin{figure}
5694: \centering
5695: \includegraphics[height=6cm]{fig21_2.ps}
5696: \caption{Angular power spectrum of 21cm anisotropies on the sky at
5697: various redshifts.  From
5698: top to bottom, $z=55,40,80,30,120,25,170$.}
5699: \label{clfig}
5700: \end{figure}
5701: 
5702: Figure \ref{clfig} shows the angular power spectrum at various
5703: redshifts. 
5704: The signal peaks around $z\sim 50$ but maintains a substantial
5705: amplitude over the full range of $30\lesssim z\lesssim 100$.  
5706: The ability to probe the small scale power of density fluctuations is only
5707: limited by the Jeans scale, below which the dark matter inhomogeneities are
5708: washed out by the finite pressure of the gas. Interestingly, the
5709: cosmological Jeans mass reaches its minimum value, $\sim 3\times 10^4
5710: M_\odot$, within the redshift interval of interest here which 
5711: corresponds to modes of angular scale $\sim$ arcsecond on the sky. During
5712: the epoch of reionization, photoionization heating raises the Jeans mass by
5713: several orders of magnitude and broadens spectral features, thus limiting
5714: the ability of other probes of the intergalactic medium, such as the
5715: Ly$\alpha$ forest, from accessing the same very low mass scales. The 21cm
5716: tomography has the additional advantage of probing the majority of the
5717: cosmic gas, instead of the trace amount ($\sim 10^{-5}$) of neutral
5718: hydrogen probed by the Ly$\alpha$ forest after reionization.  Similarly to
5719: the primary CMB anisotropies, the 21cm signal is simply shaped by gravity,
5720: adiabatic cosmic expansion, and well-known atomic physics, and is not
5721: contaminated by complex astrophysical processes that affect the
5722: intergalactic medium at $z\lesssim 30$.
5723: 
5724: Characterizing the initial fluctuations is one of the
5725: primary goals of observational cosmology, as it offers a window into the
5726: physics of the very early Universe, namely the epoch of inflation during
5727: which the fluctuations are believed to have been produced. 
5728: In most models of inflation, the evolution of the Hubble parameter during
5729: inflation leads to departures from a scale-invariant spectrum that are of
5730: order $1/N_{\rm efold}$ with $N_{\rm efold}\sim 60$ being the number of
5731: $e$--folds between the time when the scale of our horizon was of order the
5732: horizon during inflation and the end of inflation \cite{lidlith}. 
5733: Hints that the standard
5734: $\Lambda$CDM model may have too much power on galactic scales have inspired
5735: several proposals for suppressing the power on small scales. Examples
5736: include the possibility that the dark matter is warm and it decoupled while
5737: being relativistic so that its free streaming erased small-scale power
5738: \cite{Ba01}, or direct modifications of inflation that produce a cut-off in
5739: the power on small scales \cite{kamlidd}. An unavoidable collisionless
5740: component of the cosmic mass budget beyond CDM, is provided by massive
5741: neutrinos (see \cite{neut} for a review). Particle physics experiments
5742: established the mass splittings among different species which translate
5743: into a lower limit on the fraction of the dark matter accounted for by
5744: neutrinos of $f_\nu > 0.3 \%$, while current constraints based on galaxies
5745: as tracers of the small scale power imply $f_\nu < 12 \%$ \cite{tegsdss}.
5746: 
5747: Figure \ref{clcomp} shows the 21cm power spectrum for various
5748: models that differ in their level of small scale power. It is clear that a
5749: precise measurement of the 21cm power spectrum will dramatically improve
5750: current constraints on alternatives to the standard $\Lambda$CDM spectrum.
5751: 
5752: \begin{figure}
5753: \centering
5754: \includegraphics[height=6cm]{fig21_3.ps}
5755: \caption{{\it Upper panel:} Power spectrum of 21cm anisotropies at $z=55$
5756: for a $\Lambda$CDM scale-invariant power spectrum, a model with $n=0.98$, a
5757: model with $n=0.98$ and $\alpha_r\equiv {1\over 2} (d^2\ln P/d\ln
5758: k^2)=-0.07$, a model of warm dark matter particles with a mass of 1 keV,
5759: and a model in which $f_\nu=10\% $ of the matter density is in three
5760: species of massive neutrinos with a mass of $0.4~{\rm eV}$ each. {\it Lower
5761: panel:} Ratios between the different power spectra and the scale-invariant
5762: spectrum. }
5763: \label{clcomp}
5764: \end{figure}
5765: 
5766: The 21cm signal contains a wealth of information about the 
5767: initial fluctuations. A full sky map at a single photon 
5768: frequency measured up to $l_{\rm max}$, can 
5769: probe the power spectrum up to $k_{\rm max}\sim (l_{\rm max}/10^4) {\rm
5770: Mpc}^{-1}$. Such a map contains $l_{\rm max}^2$ independent samples. By
5771: shifting the photon frequency, one may obtain many independent measurements
5772: of the power. When measuring a mode $l$, which corresponds to a wavenumber
5773: $k\sim l/r$, two maps at different photon frequencies will be independent
5774: if they are separated in radial distance by $1/k$. Thus, an experiment that
5775: covers a spatial range $\Delta r$ can probe a total of $k\Delta r\sim l
5776: \Delta r/r$ independent maps. An experiment that detects the 21cm signal
5777: over a range $\Delta\nu$ centered on a frequency $\nu$, is sensitive to
5778: $\Delta r/r\sim 0.5 (\Delta\nu/\nu)(1+z)^{-1/2}$, and so it 
5779: measures a total of $N_{\rm 21cm}\sim 3 \times 10^{16} (l_{\rm max}/10^6)^3
5780: (\Delta\nu/\nu) (z/100)^{-1/2}$ independent samples.
5781: 
5782: This detection capability cannot be reproduced even remotely by other
5783: techniques. For example, the primary CMB anisotropies are damped on small
5784: scales (through the so-called Silk damping), and probe only modes with $l
5785: \leq 3000$ ($k\leq 0.2 \ {\rm Mpc}^{-1}$). The total number of modes
5786: available in the full sky is $N_{\rm cmb} = 2 l_{\rm max}^2\sim 2\times
5787: 10^7 (l_{\rm max}/3000)^2$, including both temperature and polarization
5788: information.
5789: 
5790: The sensitivity of an experiment depends strongly on its particular design,
5791: involving the number and distribution of the antennae for an
5792: interferometer. Crudely speaking, the uncertainty in the measurement of
5793: $[{l(l+1)C_l/2\pi}]^{1/2}$ is dominated by noise, $N_\nu$, which is
5794: controlled by 
5795: the sky brightness $I_\nu$ at the observed frequency $\nu$ \cite{Zalda04}, 
5796: \beqa
5797: \label{err} 
5798: N_\nu && \sim 0.4 {\rm mK }\left({I_\nu [50{\rm MHz}]\over 5\times 10^{5} {
5799: \rm Jy\ sr^{-1}}}\right) \left({l_{\rm min}\over 35}\right) \left( {5000
5800: \over l_{\rm max}}\right) \left( {0.016 \over f_{\rm cover}}\right)
5801: \nonumber \\&& \times \left({1 \ {\rm year} \over t_0}\right)^{1/2}
5802: \left({\Delta \nu \over 50{\rm MHz}
5803: %\over 0.1
5804: }\right)^{-1/2} \ \left({50\ {\rm MHz }\over \nu}\right)^{2.5} , \eeqa
5805: where $l_{\rm min}$ is the minimum observable $l$ as determined by the
5806: field of view of the instruments, $l_{\rm max}$ is the maximum observable
5807: $l$ as determined by the maximum separation of the antennae, $f_{\rm
5808: cover}$ is the fraction of the array area thats is covered by telescopes,
5809: $t_0$ is the observation time and $\Delta \nu$ is the frequency range over
5810: which the signal can be detected.  Note that the assumed sky temperature of
5811: $0.7\times 10^4$K at $\nu=50$MHz (corresponding to $z\sim 30$) is more than
5812: six orders of magnitude larger than the signal.  We have already included
5813: the fact that several independent maps can be produced by varying the
5814: observed frequency.  The numbers adopted above are appropriate for the
5815: inner core of the {\it LOFAR} array ({\it http://www.lofar.org}), planned
5816: for initial operation in 2006. The predicted signal is $\sim 1{\rm mK}$,
5817: and so a year of integration or an increase in the covering fraction are
5818: required to observe it with {\it LOFAR}.  Other experiments whose goal is
5819: to detect 21cm fluctuations from the subsequent epoch of reionization at
5820: $z\sim 6-12$ (when ionized bubbles exist and the fluctuations are larger)
5821: include the Mileura Wide-Field Array (MWA;
5822: http://web.haystack.mit.edu/arrays/MWA/), the Primeval Structure Telescope
5823: ({\it PAST}; {\it http://arxiv.org/abs/astro-ph/0502029}), and in the more
5824: distant future the Square Kilometer Array ({\it SKA}; {\it
5825: http://www.skatelescope.org}).  The main challenge in detecting the
5826: predicted signal from higher redshifts involves its appearance at low
5827: frequencies where the sky noise is high.  Proposed space-based instruments
5828: \cite{RadioAstr} avoid the terrestrial radio noise and the increasing
5829: atmospheric opacity at $\nu< 20 \ {\rm MHz}$ (corresponding to $z> 70$).
5830: 
5831: \begin{figure}
5832: \centering
5833: \includegraphics[height=6cm]{mwa.eps}
5834: \caption{Prototype of the tile design for the {\it Mileura Wide-Field
5835: Array} (MWA) in western Australia, aimed at detecting redshifted 21cm from
5836: the epoch of reionization. Each 4m$\times$4m tile contains 16 dipole
5837: antennas operating in the frequency range of 80--300MHz. Altogether the
5838: initial phase of MWA (the so-called ``Low-Frequency Demostrator'') will
5839: include 500 antenna tiles with a total collecting area of 8000 m$^2$ at
5840: 150MHz, scattered across a 1.5 km region and providing an angular
5841: resolution of a few arcminutes.}
5842: \end{figure}
5843: 
5844: The 21cm absorption is replaced by 21cm emission from neutral hydrogen as
5845: soon as the intergalactic medium is heated above the CMB temperature by
5846: X-ray sources during the epoch of reionization \cite{ChenMi}. This occurs
5847: long before reionization since the required heating requires only a modest
5848: amount of energy, $\sim 10^{-2}~{\rm eV}[(1+z)/30]$, which is three orders
5849: of magnitude smaller than the amount necessary to ionize the Universe.  As
5850: demonstrated by Chen \& Miralda-Escude (2004) \cite{ChenMi}, heating due
5851: the recoil of atoms as they absorb Ly$\alpha$ photons \cite{Madau} is not
5852: effective; the Ly$\alpha$ color temperature reaches equilibrium with the
5853: gas kinetic temperature and suppresses subsequent heating before the level
5854: of heating becomes substantial. Once most of the cosmic hydrogen is
5855: reionized at $z_{\rm reion}$, the 21cm signal is diminished. The optical
5856: depth for free-free absorption after reionization, $\sim 0.1 [(1+z_{\rm
5857: reion})/20]^{5/2}$, modifies only slightly the expected 21cm anisotropies.
5858: Gravitational lensing should modify the power spectrum \cite{Pen} at high
5859: $l$, but can be separated as in standard CMB studies (see \cite{lensing}
5860: and references therein).  The 21cm signal should be simpler to clean as it
5861: includes the same lensing foreground in independent maps obtained at
5862: different frequencies.
5863: 
5864: 
5865: \begin{figure}
5866: \centering
5867: \includegraphics[height=10cm,angle=-90]{spin_temp.ps}
5868: \caption{Schematic sketch of the evolution of the kinetic temperature
5869: ($T_k$) and spin temperature ($T_s$) of cosmic hydrogen. Following
5870: cosmological recombination at $z\sim 10^3$, the gas temperature (orange
5871: curve) tracks the CMB temperature (blue line; $T_\gamma\propto (1+z)$) down
5872: to $z\sim 200$ and then declines below it ($T_k\propto (1+z)^2$) until the
5873: first X-ray sources (accreting black holes or exploding supernovae) heat it
5874: up well above the CMB temperature. The spin temperature of the 21cm
5875: transition (red curve) interpolates between the gas and CMB
5876: temperatures. Initially it tracks the gas temperature through collisional
5877: coupling; then it tracks the CMB through radiative coupling; and eventually
5878: it tracks the gas temperature once again after the production of a cosmic
5879: background of UV photons between the Ly$\alpha$ and the Lyman-limit
5880: frequencies that redshift or cascade into the Ly$\alpha$ resonance (through
5881: the Wouthuysen-Field effect [Wouthuysen 1952 \cite{Wout}; Field 1959
5882: \cite{Field}]). Parts of the curve are exaggerated for pedagogic purposes.
5883: The exact shape depends on astrophysical details about the first galaxies,
5884: such as their production of X-ray binaries, supernovae, nuclear accreting
5885: black holes, and their generation of relativistic electrons in
5886: collisionless shocks which produce UV and X-ray photons through
5887: inverse-Compton scattering of CMB photons. }
5888: \label{spin}
5889: \end{figure}
5890: 
5891: The large number of independent modes probed by the 21cm signal would
5892: provide a measure of non-Gaussian deviations to a level of $\sim N_{\rm 21
5893: cm}^{-1/2}$, constituting a test of the inflationary origin of the
5894: primordial inhomogeneities which are expected to possess deviations
5895: $\gtrsim 10^{-6}$ \cite{malda}.
5896:  
5897: \subsection{The Characteristic Observed Size of Ionized Bubbles at the
5898: End of Reionization} 
5899: \label{size}
5900: 
5901: \begin{figure}
5902: \centering
5903: \includegraphics[height=6cm]{cheeze.eps}
5904: \caption{21cm imaging of ionized bubbles during the epoch of reionization
5905: is analogous to slicing swiss cheese. The technique of slicing at intervals
5906: separated by the typical dimension of a bubble is optimal for revealing
5907: different pattens in each slice.}  \label{swiss}
5908: \end{figure}
5909: 
5910: The first galaxies to appear in the Universe at redshifts $z\ga 20$ created
5911: ionized bubbles in the intergalactic medium (IGM) of neutral hydrogen (\HI)
5912: left over from the Big-Bang. It is thought that the ionized bubbles grew
5913: with time, surrounded clusters of dwarf galaxies\cite{BL04,FZH} and
5914: eventually overlapped quickly throughout the Universe over a narrow
5915: redshift interval near $z\sim 6$. This event signaled the end of the
5916: reionization epoch when the Universe was a billion years old.  Measuring
5917: the unknown size distribution of the bubbles at their final overlap phase
5918: is a focus of forthcoming observational programs aimed at highly redshifted
5919: 21cm emission from atomic hydrogen.  In this sub-section we follow Wyithe
5920: \& Loeb (2004) \cite{Bubble} and show that the combined constraints of cosmic variance
5921: and causality imply an observed bubble size at the end of the overlap epoch
5922: of $\sim 10$ physical Mpc, and a scatter in the observed redshift of
5923: overlap along different lines-of-sight of $\sim 0.15$. This scatter is
5924: consistent with observational constraints from recent spectroscopic data on
5925: the farthest known quasars. This result implies that future radio
5926: experiments should be tuned to a characteristic angular scale of $\sim
5927: 0.5^\circ$ and have a minimum frequency band-width of $\sim 8$ MHz for an
5928: optimal detection of 21cm flux fluctuations near the end of reionization.
5929: 
5930: During the reionization epoch, the characteristic bubble size (defined here
5931: as the spherically averaged mean radius of the \HII regions that contain
5932: most of the ionized volume\cite{FZH}) increased with time as smaller
5933: bubbles combined until their overlap completed and the diffuse IGM was
5934: reionized.  However the largest size of isolated bubbles (fully surrounded
5935: by \HI boundaries) that can be {\em observed} is finite, because of the
5936: combined phenomena of cosmic variance and causality. Figure \ref{ffig1}
5937: presents a schematic illustration of the geometry. There is a surface on
5938: the sky corresponding to the time along different lines-of-sight when the
5939: diffuse (uncollapsed) IGM was {\em most recently neutral}. We refer to it
5940: as the Surface of Bubble Overlap (SBO). There are two competing sources for
5941: fluctuations in the SBO, each of which is dependent on the characteristic
5942: size, $R_{\rm SBO}$, of the ionized regions just before the final
5943: overlap. First, the finite speed of light implies that 21cm photons
5944: observed from different points along the curved boundary of an \HII region
5945: must have been emitted at different times during the history of the
5946: Universe.  Second, bubbles on a comoving scale $R$ achieve reionization
5947: over a spread of redshifts due to cosmic variance in the initial conditions
5948: of the density field smoothed on that scale.  The characteristic scale of
5949: \HII bubbles grows with time, leading to a decline in the spread of their
5950: formation redshifts\cite{BL04} as the cosmic variance is averaged over an
5951: increasing spatial volume.  However the 21cm light-travel time across a
5952: bubble rises concurrently. Suppose a signal 21cm photon which encodes the
5953: presence of neutral gas, is emitted from the far edge of the ionizing
5954: bubble. If the adjacent region along the line-of-sight has not become
5955: ionized by the time this photon reaches the near side of the bubble, then
5956: the photon will encounter diffuse neutral gas. Other photons emitted at
5957: this lower redshift will therefore also encode the presence of diffuse
5958: neutral gas, implying that the first photon was emitted prior to overlap,
5959: and not from the SBO. Hence the largest observable scale of \HII regions
5960: when their overlap completes, corresponds to the first epoch at which the
5961: light crossing time becomes larger than the spread in formation times of
5962: ionized regions.  Only then will the signal photon leaving the far side of
5963: the HII region have the lowest redshift of any signal photon along that
5964: line-of-sight.
5965: 
5966: \begin{figure}
5967: \centering
5968: \includegraphics[height=6cm]{Fan.ps}
5969: \caption{Spectra of 19 quasars with redshifts $5.74<z<6.42$ from the {\it
5970: Sloan Digital Sky Survey} \cite{Fan05}. For some of the highest-redshift
5971: quasars, the spectrum shows no transmitted flux shortward of the Ly$\alpha$
5972: wavelength at the quasar redshift (the so-called ``Gunn-Peterson trough''),
5973: indicating a non-negligible neutral fraction in the IGM (see the analysis
5974: of Fan et al. \cite{Fan05} for details).  }
5975: \end{figure}
5976: 
5977: The observed spectra of some quasars beyond $z\sim6.1$ show a Gunn-Peterson
5978: trough\cite{GP,f2} (Fan et al. 2005 \cite{Fan05}), a blank spectral region at
5979: wavelengths shorter than \lya at the quasar redshift, implying the presence
5980: of \HI in the diffuse IGM.  The detection of Gunn-Peterson troughs
5981: indicates a rapid change\cite{f1,Pe,WB1} in the neutral content of the IGM
5982: at $z\sim6$, and hence a rapid change in the intensity of the background
5983: ionizing flux. This rapid change implies that overlap, and hence the
5984: reionization epoch, concluded near $z\sim6$. The most promising
5985: observational probe\cite{Zalda04,Miguel1} of the reionization epoch is redshifted
5986: 21cm emission from intergalactic \HI.  Future observations using low
5987: frequency radio arrays (e.g. LOFAR, MWA, and PAST) will allow a direct
5988: determination of the topology and duration of the phase of bubble
5989: overlap. In this section we determine the expected angular scale and
5990: redshift width of the 21cm fluctuations at the SBO theoretically, and show
5991: that this determination is consistent with current observational
5992: constraints.
5993: 
5994: We start by quantifying the constraints of causality and cosmic
5995: variance. First suppose we have an \HII region with a physical radius
5996: $R/(1+\langle z\rangle)$. For a 21cm photon, the light crossing time of
5997: this radius is
5998: \begin{equation}
5999: \label{causality}
6000: \langle\Delta z^2\rangle^{1/2} =
6001: \left|\frac{dz}{dt}\right|_{\langle z\rangle}
6002: \frac{R}{c(1+\langle z\rangle)},
6003: \end{equation}
6004: where at the high-redshifts of interest
6005: $(dz/dt)=-(H_0\sqrt{\Omega_m})(1+z)^{5/2}$.  Here, $c$ is the speed of
6006: light, $H_0$ is the present-day Hubble constant, $\Omega_m$ is the present
6007: day matter density parameter, and $\langle z\rangle$ is the mean redshift
6008: of the SBO.  Note that when discussing this crossing time, we are referring
6009: to photons used to probe the ionized bubble (e.g. at 21cm), rather than
6010: photons involved in the dynamics of the bubble evolution.
6011: 
6012: Second, overlap would have occurred at different times in different regions
6013: of the IGM due to the cosmic scatter in the process of structure formation
6014: within finite spatial volumes\cite{BL04}. Reionization should be completed
6015: within a region of comoving radius $R$ when the fraction of mass
6016: incorporated into collapsed objects in this region attains a certain
6017: critical value, corresponding to a threshold number of ionizing photons
6018: emitted per baryon. The ionization state of a region is governed by the
6019: enclosed ionizing luminosity, by its over-density, and by dense pockets of
6020: neutral gas that are self shielding to ionizing radiation.  There is an
6021: offset \cite{BL04} $\delta z$ between the redshift when a region of mean
6022: over-density $\bar{\delta}_{\rm R}$ achieves this critical collapsed
6023: fraction, and the redshift ${\bar z}$ when the Universe achieves the same
6024: collapsed fraction on average.  This offset may be computed\cite{BL04} from
6025: the expression for the collapsed fraction\cite{bond91} $F_{\rm col}$ within a
6026: region of over-density $\bar{\delta}_{\rm R}$ on a comoving scale $R$,
6027: \begin{equation}
6028: \label{scatter}
6029: F_{\rm col}(M_{\rm min})=\mbox{erfc}\left[\frac{\delta_{\rm
6030: c}-\bar{\delta}_{\rm R}}{\sqrt{2[\sigma_{\rm R_{\rm min}}^2-\sigma_{\rm
6031: R}^2]}}\right]
6032: %,\hspace{2mm}\mbox{yielding}\hspace{2mm} 
6033: \rightarrow 
6034: \frac{\delta
6035: z}{(1+\bar{z})}=\frac{\bar{\delta}_{\rm R}}{\delta_{\rm
6036: c}(\bar{z})}-\left[1-\sqrt{1-\frac{\sigma_{\rm R}^2}{\sigma_{\rm R_{\rm
6037: min}}^2}}\right],
6038: \end{equation}
6039: where $\delta_{\rm c}(\bar{z})\propto (1+\bar{z})$ is the collapse
6040: threshold for an over-density at a redshift $\bar{z}$; $\sigma_{\rm R}$ and
6041: $\sigma_{R_{\rm min}}$ are the variances in the power-spectrum linearly
6042: extrapolated to $z=0$ on comoving scales corresponding to the region of
6043: interest and to the minimum galaxy mass $M_{\rm min}$, respectively.  The
6044: offset in the ionization redshift of a region depends on its linear
6045: over-density, $\bar{\delta}_{\rm R}$. As a result, the distribution of
6046: offsets, and therefore the scatter in the SBO may be obtained directly from
6047: the power spectrum of primordial inhomogeneities. As can be seen from
6048: equation~(\ref{scatter}), larger regions have a smaller scatter due to
6049: their smaller cosmic variance.
6050: 
6051: Note that equation~(\ref{scatter}) is independent of the critical
6052: value of the collapsed fraction required for reionization. Moreover,
6053: our numerical constraints are very weakly dependent on the minimum
6054: galaxy mass, which we choose to have a virial temperature of $10^4$K
6055: corresponding to the cooling threshold of primordial atomic gas.  The
6056: growth of an \HII bubble around a cluster of sources requires that the
6057: mean-free-path of ionizing photons be of order the bubble radius or
6058: larger. Since ionizing photons can be absorbed by dense pockets of
6059: neutral gas inside the \HII region, the necessary increase in the
6060: mean-free-path with time implies that the critical collapsed fraction
6061: required to ionize a region of size $R$ increases as well. This larger
6062: collapsed fraction affects the redshift at which the region becomes
6063: ionized, but not the scatter in redshifts from place to place which is
6064: the focus of this sub-section. Our results are therefore independent
6065: of assumptions about unknown quantities such as the star formation
6066: efficiency and the escape fraction of ionizing photons from galaxies,
6067: as well as unknown processes of feedback in galaxies and clumping of
6068: the IGM.
6069: 
6070: Figure~\ref{ffig2} displays the above two fundamental constraints.  The
6071: causality constraint (Eq.~\ref{causality}) is shown as the blue line,
6072: giving a longer crossing time for a larger bubble size. This contrasts with
6073: the constraint of cosmic variance (Eq.~\ref{scatter}), indicated by the red
6074: line, which shows how the scatter in formation times decreases with
6075: increasing bubble size. The scatter in the SBO redshift and the
6076: corresponding fluctuation scale of the SBO are given by the intersection of
6077: these curves. We find that the thickness of the SBO is $\langle\Delta
6078: z^2\rangle^{1/2}\sim0.13$, and that the bubbles which form the SBO have a
6079: characteristic comoving size of $\sim60$Mpc (equivalent to 8.6 physical
6080: Mpc). At $z\sim6$ this size corresponds to angular scales of $\theta_{\rm
6081: SBO}\sim0.4$ degrees on the sky.
6082: 
6083: A scatter of $\sim0.15$ in the SBO is somewhat larger than the value
6084: extracted from existing numerical
6085: simulations\cite{g00,yoshida}. The difference is most likely due
6086: to the limited size of the simulated volumes; while the simulations
6087: appropriately describe the reionization process within limited regions
6088: of the Universe, they are not sufficiently large to describe the
6089: global properties of the overlap phase\cite{BL04}. The scales over
6090: which cosmological radiative transfer has been simulated are smaller
6091: than the characteristic extent of the SBO, which we find to be $R_{\rm
6092: SBO}\sim70$ comoving Mpc.
6093: 
6094: We can constrain the scatter in the SBO redshift observationally using the
6095: spectra of the highest redshift quasars. Since only a trace amount of
6096: neutral hydrogen is needed to absorb Ly$\alpha$ photons, the time where the
6097: IGM becomes \lya transparent need not coincide with bubble
6098: overlap. Following overlap the IGM was exposed to ionizing sources in all
6099: directions and the ionizing intensity rose rapidly. After some time the
6100: ionizing background flux was sufficiently high that the \HI fraction fell
6101: to a level at which the IGM allowed transmission of resonant \lya
6102: photons. This is shown schematically in Figure~\ref{ffig1}.  The lower
6103: wavelength limit of the Gunn-Peterson trough corresponds to the \lya
6104: wavelength at the redshift when the IGM started to allow transmission of
6105: \lya photons {\em along that particular line-of-sight}.  In addition to the
6106: SBO we therefore also define the Surface of \lya Transmission (hereafter
6107: SLT) as the redshift along different lines-of-sight when the diffuse IGM
6108: became transparent to Ly$\alpha$ photons.
6109: 
6110: The scatter in the SLT redshift is an observable which we would like to
6111: compare with the scatter in the SBO redshift.  The variance of the density
6112: field on large scales results in the biased clustering of
6113: sources\cite{BL04}.  \HII regions grow in size around these clusters of
6114: sources. In order for the ionizing photons produced by a cluster to advance
6115: the walls of the ionized bubble around it, the mean-free-path of these
6116: photons must be of order the bubble size or larger.  After bubble overlap,
6117: the ionizing intensity at any point grows until the ionizing photons have
6118: time to travel across the scale of the new mean-free-path, which represents
6119: the horizon out to which ionizing sources are visible.  Since the
6120: mean-free-path is larger than $R_{\rm SBO}$, the ionizing intensity at the
6121: SLT averages the cosmic scatter over a larger volume than at the SBO.  This
6122: constraint implies that the cosmic variance in the SLT redshift must be
6123: smaller than the scatter in the SBO redshift. However, it is possible that
6124: opacity from small-scale structure contributes additional scatter to the
6125: SLT redshift.
6126: 
6127: If cosmic variance dominates the observed scatter in the SLT redshift, then
6128: based on the spectra of the three $z>6.1$ quasars\cite{f2,WB1} we
6129: would expect the scatter in the SBO redshift to satisfy $\langle\Delta
6130: z^2\rangle^{1/2}_{\rm obs}\ga0.05$. In addition, analysis of the {\it
6131: proximity effect} for the size of the \HII regions around the two highest
6132: redshift quasars\cite{WL04b,Mes04} implies a neutral fraction that is
6133: of order unity (i.e. pre-overlap) at $z\sim6.2-6.3$, while the transmission
6134: of Ly$\alpha$ photons at $z\la6$ implies that overlap must have completed
6135: by that time. This restricts the scatter in the SBO to be $\langle\Delta
6136: z^2\rangle^{1/2}_{\rm obs}\la0.25$. The constraints on values for the
6137: scatter in the SBO redshift are shaded gray in Figure~\ref{ffig2}.  It is
6138: reassuring that the theoretical prediction for the SBO scatter of
6139: $\langle\Delta z^2\rangle^{1/2}_{\rm obs}\sim0.15$, with a characteristic
6140: scale of $\sim70$ comoving Mpc, is bounded by these constraints.
6141: 
6142: The possible presence of a significantly neutral IGM just beyond the
6143: redshift of overlap\cite{WL04b,Mes04} is encouraging for upcoming
6144: 21cm studies of the reionization epoch as it results in emission near an
6145: observed frequency of 200 MHz where the signal is most readily
6146: detectable. Future observations of redshifted 21cm line emission at $6\la
6147: z\la 6.5$ with instruments such as LOFAR, MWA, and PAST, will be
6148: able to map the three-dimensional distribution of HI at the end of
6149: reionization. The intergalactic \HII regions will imprint a 'knee' in the
6150: power-spectrum of the 21cm anisotropies on a characteristic angular scale
6151: corresponding to a typical isolated \HII region\cite{Zalda04}. Our results
6152: suggest that this characteristic angular scale is large at the end of
6153: reionization, $\theta_{\rm SBO}\sim 0.5$ degrees, motivating the
6154: construction of compact low frequency arrays. An SBO thickness of $\langle
6155: \Delta z^2\rangle^{1/2}\sim0.15$ suggests a minimum frequency band-width of
6156: $\sim8$ MHz for experiments aiming to detect anisotropies in 21cm emission
6157: just prior to overlap. These results will help guide the design of the next
6158: generation of low-frequency radio observatories in the search for 21cm
6159: emission at the end of the reionization epoch.
6160: \begin{figure}
6161: \centering
6162: \includegraphics[height=8cm]{figS_1.eps}
6163: \caption{The distances to the observed Surface of Bubble
6164: Overlap (SBO) and Surface of Ly$\alpha$ Transmission (SLT) fluctuate
6165: on the sky. The SBO corresponds to the first region of diffuse neutral
6166: IGM {\em observed} along a random line-of-sight. It fluctuates across
6167: a shell with a minimum width dictated by the condition that the light
6168: crossing time across the characteristic radius $R_{\rm SBO}$ of
6169: ionized bubbles equals the cosmic scatter in their formation
6170: times. Thus, {\it causality} and {\it cosmic variance} determine the
6171: characteristic scale of bubbles at the completion of bubble overlap.
6172: After some time delay the IGM becomes transparent to Ly$\alpha$
6173: photons, resulting in a second surface, the SLT.  The upper panel
6174: illustrates how the lines-of-sight towards two quasars (Q1 in red and
6175: Q2 in blue) intersect the SLT with a redshift difference $\delta
6176: z$. The resulting variation in the observed spectrum of the two
6177: quasars is shown in the lower panel.  Observationally, the ensemble of
6178: redshifts down to which the Gunn-Peterson troughs are seen in the
6179: spectra of $z>6.1$ quasars is drawn from the probability distribution
6180: $dP/dz_{\rm SLT}$ for the redshift at which the IGM started to allow
6181: \lya transmission along random lines-of-sight. The observed values of
6182: $z_{\rm SLT}$ show a small scatter\cite{f2} in the SLT redshift around
6183: an average value of $\langle z_{\rm SLT}\rangle\approx 5.95$. Some
6184: regions of the IGM may have also become transparent to Ly$\alpha$
6185: photons prior to overlap, resulting in windows of transmission inside
6186: the Gunn-Peterson trough (one such region may have been seen\cite{WB1}
6187: in SDSS J1148+5251).  In the existing examples, the portions of the
6188: Universe probed by the lower end of the Gunn-Peterson trough are
6189: located several hundred comoving Mpc away from the background quasar,
6190: and are therefore not correlated with the quasar host galaxy. The
6191: distribution $dP/dz_{\rm SLT}$ is also independent of the redshift
6192: distribution of the quasars. Moreover, lines-of-sight to these quasars
6193: are not causally connected at $z\sim 6$ and may be considered
6194: independent. }
6195: \label{ffig1}
6196: \end{figure}
6197: 
6198: 
6199: \begin{figure}
6200: \centering
6201: \includegraphics[height=8cm]{figS_2.eps}
6202: \caption{Constraints on the scatter in the SBO redshift and
6203: the characteristic size of isolated bubbles at the final overlap stage,
6204: $R_{\rm SBO}$ (see Fig. 1). The characteristic size of \HII regions grows
6205: with time. The SBO is observed for the bubble scale at which the light
6206: crossing time (blue line) first becomes smaller than the cosmic scatter in
6207: bubble formation times (red line).  At $z\sim6$, the implied scale $R_{\rm
6208: SBO}\sim60$ comoving Mpc (or $\sim 8.6$ physical Mpc), corresponds to a
6209: characteristic angular radius of $\theta_{\rm SLT}\sim 0.4$ degrees on the
6210: sky.  After bubble overlap, the ionizing intensity grows to a level at
6211: which the IGM becomes transparent to Ly$\alpha$ photons. The collapsed
6212: fraction required for \lya transmission within a region of a certain size
6213: will be larger than required for its ionization. However, the scatter in
6214: equation~(\ref{scatter}) is not sensitive to the collapsed fraction, and so
6215: may be used for both the SBO and SLT.  The scatter in the SLT is smaller
6216: than the cosmic scatter in the structure formation time on the scale of the
6217: mean-free-path for ionizing photons.  This mean-free-path must be longer
6218: than $R_{\rm SBO}\sim60$Mpc, an inference which is supported by analysis of
6219: the Ly$\alpha$ forest at $z\sim4$ where the mean-free-path is
6220: estimated\cite{ME1} to be $\sim 120$ comoving Mpc at the Lyman limit (and
6221: longer at higher frequencies). If it is dominated by cosmic variance, then
6222: the scatter in the SLT redshift provides a lower limit to the SBO
6223: scatter. The three known quasars at $z>6.1$ have \lya transmission
6224: redshifts of\cite{WB1,f2} $z_{\rm SLT}=5.9$, 5.95 and 5.98,
6225: implying that the scatter in the SBO must be $\ga0.05$ (this scatter may
6226: become better known from follow-up spectroscopy of Gamma Ray Burst
6227: afterglows at $z>6$ that might be discovered by the {\it SWIFT}
6228: satellite\cite{GRBquasar,BL02}).  
6229: The observed scatter in the SLT redshift is somewhat smaller than the
6230: predicted SBO scatter, confirming the expectation that cosmic variance is
6231: smaller at the SLT. The scatter in the SBO redshift must also be $\la0.25$
6232: because the lines-of-sight to the two highest redshift quasars have a
6233: redshift of \lya transparency at $z\sim6$, but a neutral fraction that is
6234: known from the {\it proximity effect}\cite{WL04b} to be substantial at
6235: $z\ga6.2-6.3$. The excluded regions of scatter for the SBO are shown in
6236: gray.  }
6237: \label{ffig2}
6238: \end{figure}
6239: 
6240: The full size distribution of ionized bubbles has to be calculated from a
6241: numerical cosmological simulation that includes gas dynamics and radiative
6242: transfer. The simulation box needs to be sufficiently large for it to
6243: sample an unbiased volume of the Universe with little cosmic variance, but
6244: at the same time one must resolve the scale of individual dwarf galaxies
6245: which provide (as well as consume) ionizing photons (see discussion at the
6246: last section of this review). Until a reliable simulation of this magnitude
6247: exists, one must adopt an approximate analytic approach to estimate the
6248: bubble size distribution. Below we describe an example for such a method,
6249: developed by Furlanetto, Zaldarriaga, \& Hernquist (2004) \cite{FZH}.
6250: 
6251: The criterion for a region to be ionized is that galaxies inside of it
6252: produce a sufficient number of ionizing photons per baryon.  This condition
6253: can be translated to the requirement that the collapsed fraction of mass in
6254: halos above some threshold mass $M_{\rm min}$ will exceed some threshold,
6255: namely $F_{\rm col}>\zeta^{-1}$. The minimum halo mass most likely
6256: corresponds to a virial temperature of $10^4$K relating to the threshold
6257: for atomic cooling (assuming that molecular hydrogen cooling is suppressed
6258: by the UV background in the Lyman-Werner band).  We would like to find the
6259: largest region around every point that satisfies the above condition on the
6260: collapse fraction and then calculate the abundance of ionized regions of
6261: this size.  Different regions have different values of $F_{\rm col}$
6262: because their mean density is different. In the extended Press-Schechter
6263: model (Bond et al. 1991 \cite{bond91}; Lacey \& Cole 1993 \cite{LC93}), the collapse fraction in a
6264: region of mean overdensity $\delta_M$ is
6265: \begin{equation}
6266: F_{\rm col}={\rm erfc}\left({\delta_c-\delta_M\over \sqrt{2[\sigma_{\rm
6267: min}^2-\sigma^2(M,z)]}}\right). 
6268: \end{equation}
6269: where $\sigma^2(M,z)$ is the variance of density fluctuations on 
6270: mass scale $M$, $\sigma^2_{\rm min}\equiv\sigma^2(M_{\rm min},z)$,
6271: and $\delta_c$ is the collapse threshold. This equation can be used
6272: to derive the condition on the mean overdensity within
6273: a region of mass $M$ in order for it to be ionized,
6274: \begin{equation}
6275: \delta_M>\delta_B(M,z)\equiv \delta_c-{\sqrt{2}}K(\zeta)[\sigma_{\rm
6276: min}^2-\sigma^2(M,z)]^{1/2},
6277: \label{barrier}
6278: \end{equation}
6279: where $K(\zeta)={\rm erfc}^{-1}(1-\zeta^{-1})$. Furlanetto et al.
6280: \cite{FZH} showed how to construct the mass function of ionized regions
6281: from $\delta_B$ in analogy with the halo mass function (Press \& Schechter
6282: 1974 \cite{Press}; Bond et al. 1991 \cite{bond91}). The barrier in equation (\ref{barrier}) is well
6283: approximated by a linear dependence on $\sigma^2$,
6284: \begin{equation}
6285: \delta_B\approx B(M)=B_0+B_1\sigma^2(M),
6286: \end{equation}
6287: in which case the mass function has an analytic solution (Sheth 1998 \cite{sheth98}),
6288: \begin{equation}
6289: n(M)={\sqrt{2\over \pi}}{{\bar{\rho}}\over M^2}\left | {d\ln \sigma
6290: \over d\ln M}\right | {B_0\over \sigma(M)}\exp\left[-{B^2(M)\over
6291: 2\sigma^2(M)}\right],
6292: \end{equation}
6293: where ${\bar{\rho}}$ is the mean mass density. This solution provides the
6294: comoving number density of ionized bubbles with mass in the range of
6295: $(M,M+dM)$. The main difference of this result from the Press-Schechter
6296: mass function is that the barrier in this case becomes more difficult to
6297: cross on smaller scales because $\delta_B$ is a decreasing function of mass
6298: $M$.  This gives bubbles a characteristic size. The size evolves with
6299: redshift in a way that depends only on $\zeta$ and $M_{\rm min}$.
6300: 
6301: One limitation of the above analytic model is that it ignores the non-local
6302: influence of sources on distant regions (such as voids) as well as the
6303: possible shadowing effect of intervening gas.  Radiative transfer effects
6304: in the real Universe are inherently three-dimensional and cannot be fully
6305: captured by spherical averages as done in this model. Moreover, the value
6306: of $M_{\rm min}$ is expected to increase in regions that were already
6307: ionized, complicating the expectation of whether they will remain ionized
6308: later.  The history of reionization could be complicated and non monotonic
6309: in individual regions, as described by Furlanetto \& Loeb (2005) \cite{FL05}. Finally,
6310: the above analytic formalism does not take the light propagation delay into
6311: account as we have done above in estimating the characteristic bubble size
6312: at the end of reionization. Hence this formalism describes the observed
6313: bubbles only as long as the characteristic bubble size is sufficiently
6314: small, so that the light propagation delay can be neglected compared to
6315: cosmic variance.  The general effect of the light propagation delay on the
6316: power-spectrum of 21cm fluctuations was quantified by Barkana \& Loeb
6317: (2005) \cite{BLinf}.
6318: 
6319: 
6320: 
6321: \subsection{Separating the ``Physics'' from the ``Astrophysics'' of the 
6322: Reionization Epoch with 21cm Fluctuations}
6323: 
6324: The 21cm signal can be seen from epochs during which the cosmic gas was
6325: largely neutral and deviated from thermal equilibrium with the cosmic
6326: microwave background (CMB). The signal vanished at redshifts $z\ga
6327: 200$, when the residual fraction of free electrons after cosmological
6328: recombination kept the gas kinetic temperature, $T_{k}$, close to the
6329: CMB temperature, $T_\gamma$. But during $200\ga z\ga 30$ the gas
6330: cooled adiabatically and atomic collisions kept the spin temperature
6331: of the hyperfine level population below $T_\gamma$, so that the gas
6332: appeared in absorption \cite{Scott,Loeb04}. As the Hubble expansion
6333: continued to rarefy the gas, radiative coupling of $T_s$ to $T_\gamma$
6334: began to dominate and the 21cm signal faded. When the first galaxies
6335: formed, the UV photons they produced between the Ly$\alpha$ and Lyman
6336: limit wavelengths propagated freely through the Universe, redshifted
6337: into the Ly$\alpha$ resonance, and coupled $T_s$ and $T_{k}$ once
6338: again through the Wouthuysen-Field \cite{Wout,Field} effect by which
6339: the two hyperfine states are mixed through the absorption and
6340: re-emission of a Ly$\alpha$ photon \citep{Madau, Ciardi}. Emission
6341: above the Lyman limit by the same galaxies initiated the
6342: process of reionization by creating ionized bubbles in the neutral
6343: cosmic gas, while X-ray photons propagated farther and heated $T_{k}$
6344: above $T_\gamma$ throughout the Universe. Once $T_s$ grew larger than
6345: $T_\gamma$, the gas appeared in 21cm emission. The ionized bubbles
6346: imprinted a knee in the power spectrum of 21cm fluctuations
6347: \citep{Zalda04}, which traced the H I  topology until the
6348: process of reionization was completed \citep{FZH}.
6349: 
6350: The various effects that determine the 21cm fluctuations can be separated
6351: into two classes. The density power spectrum probes basic cosmological
6352: parameters and inflationary initial conditions, and can be calculated
6353: exactly in linear theory. However, the radiation from galaxies, both
6354: Ly$\alpha$ radiation and ionizing photons, involves the complex, non-linear
6355: physics of galaxy formation and star formation. If only the sum of all
6356: fluctuations could be measured, then it would be difficult to extract the
6357: separate sources, and in particular, the extraction of the power spectrum
6358: would be subject to systematic errors involving the properties of
6359: galaxies. Barkana \& Loeb (2005) \cite{BL05a} showed that the unique
6360: three-dimensional properties of 21cm measurements permit a separation of
6361: these distinct effects. Thus, 21cm fluctuations can probe astrophysical
6362: (radiative) sources associated with the first galaxies, while at the same
6363: time separately probing the physical (inflationary) initial conditions of
6364: the Universe. In order to affect this separation most easily, it is
6365: necessary to measure the three-dimensional power spectrum of 21cm
6366: fluctuations. The discussion in this section follows Barkana \& Loeb
6367: (2005) \cite{BL05a}.
6368: 
6369: \noindent{\bf Spin temperature history}
6370: 
6371: %The spin temperature $T_s$ is defined through the ratio between the
6372: %number densities of hydrogen atoms in the excited and ground state
6373: %levels, ${n_1/ n_0}=(g_1/ g_0)\exp\left\{-{T_\star/ T_s}\right\},$
6374: %where subscripts $1$ and $0$ correspond to the excited and ground
6375: %state levels of the 21cm transition, $(g_1/g_0)=3$ is the ratio of the
6376: %spin degeneracy factors of the levels, and $T_\star=0.0682$K
6377: %corresponds to the energy difference between the levels. 
6378: 
6379: As long as the spin-temperature $T_s$ is smaller than the CMB temperature
6380: $T_{\gamma} = 2.725 (1+z)$ K, hydrogen atoms absorb the CMB, whereas if
6381: $T_s > T_{\gamma}$ they emit excess flux. In general, the resonant 21cm
6382: interaction changes the brightness temperature of the CMB by
6383: \citep{Scott,Madau} $T_b =\tau \left( T_s-T_{\gamma}\right)/(1+z)$, where
6384: the optical depth at a wavelength $\lambda=21$cm is \beq \label{tau} \tau=
6385: \frac {3c\lambda^2h A_{10}n_{\rm H}} {32 \pi k T_s\, (1+z)\, (dv_r/dr)}
6386: x_{\rm HI}\ , \eeq where $n_H$ is the number density of hydrogen,
6387: $A_{10}=2.85\times 10^{-15}~{\rm s^{-1}}$ is the spontaneous emission
6388: coefficient, $x_{\rm HI}$ is the neutral hydrogen fraction, and $dv_r/dr$
6389: is the gradient of the radial velocity along the line of sight with $v_r$
6390: being the physical radial velocity and $r$ the comoving distance; on
6391: average $dv_r/dr = H(z)/ (1+z)$ where $H$ is the Hubble parameter.  The
6392: velocity gradient term arises because it dictates the path length over
6393: which a 21cm photon resonates with atoms before it is shifted out of
6394: resonance by the Doppler effect \citep{Sobolev}.
6395: 
6396: For the concordance set of cosmological parameters \citep{WMAP}, the
6397: mean brightness temperature on the sky at redshift $z$ is 
6398: \begin{equation}
6399: T_b = 28\,
6400: {\rm mK}\,
6401: \left( \frac{\Omega_b h}{.033} \right) \left(
6402: \frac{\Omega_m}{.27} \right)^{-\frac{1}{2}} 
6403: \left[{({1+z})\over{10}}\right]^{1/2} \left[{({T_s - T_{\gamma}})\over
6404: {T_s}}\right]
6405: \bar{x}_{\rm HI}, 
6406: \end{equation}
6407: where $\bar{x}_{\rm HI}$ is the mean neutral
6408: fraction of hydrogen.  The spin temperature itself is coupled to $T_k$
6409: through the spin-flip transition, which can be excited by collisions
6410: or by the absorption of \Lya photons.  As a result, the combination
6411: that appears in $T_b$ becomes \citep{Field} $(T_s - T_{\gamma})/T_s =
6412: [x_{\rm tot}/(1+ x_{\rm tot})] \left(1 - T_{\gamma}/T_k \right)$,
6413: where $x_{\rm tot} = x_{\alpha} + x_c$ is the sum of the radiative and
6414: collisional threshold parameters. These parameters are $x_{\alpha} =
6415: {4 P_{\alpha} T_\star}/{27 A_{10} T_{\gamma}}$ and $x_c = {4
6416: \kappa_{1-0}(T_k)\, n_H T_\star}/{3 A_{10} T_{\gamma}}$, where
6417: $P_{\alpha}$ is the \Lya scattering rate which is proportional to the
6418: \Lya intensity, and $\kappa_{1-0}$ is tabulated as a function of $T_k$
6419: \citep{AD, Zyg}. The coupling of the spin temperature
6420: to the gas temperature becomes substantial when $x_{\rm tot} \ga 1$.
6421: 
6422: \noindent{\bf Brightness temperature fluctuations}
6423: 
6424: Although the mean 21cm emission or absorption is difficult to measure due
6425: to bright foregrounds, the unique character of the fluctuations in $T_b$
6426: allows for a much easier extraction of the signal \citep{Shaver, Zalda04,
6427: Miguel1, Miguel2, Santos}. We adopt the notation $\delta_A$ for the
6428: fractional fluctuation in quantity $A$ (with a lone $\delta$ denoting
6429: density perturbations). In general, the fluctuations in $T_b$ can be
6430: sourced by fluctuations in gas density ($\delta$), \Lya flux (through
6431: $\delta_{x_{\alpha}}$) neutral fraction ($\delta_{x_{\rm HI}}$), radial
6432: velocity gradient ($\delta_{d_rv_r}$), and temperature, so we find \beqa
6433: \label{dsignal} \delta_{T_b} & = & \left( 1 + \frac{x_c} {\tilde{x}_{\rm
6434: tot}} \right) \delta + \frac{x_{\alpha}} {\tilde{x}_{\rm tot}}
6435: \delta_{x_{\alpha}} + \delta_{x_{\rm HI}} - \delta_{d_rv_r} \nonumber \\ &
6436: & + (\gamma_a - 1) \left[ \frac{T_{\gamma}} {T_k - T_{\gamma}} + \frac{x_c}
6437: {\tilde{x}_{ \rm tot}}\, \frac{d \log(\kappa_{1-0})} {d \log(T_k)}
6438: \right]\, \delta \ , \eeqa where the adiabatic index is $\gamma_a = 1 +
6439: (\delta_{T_k} / \delta)$, and we define $\tilde{x}_{\rm tot} \equiv (1 +
6440: x_{\rm tot}) x_{\rm tot}$. Taking the Fourier transform, we obtain the
6441: power spectrum of each quantity; e.g., the total power spectrum $P_{T_b}$
6442: is defined by \beq \label{pTb} \langle \td_{T_b} (\bk_1) \td_{T_b} (\bk_2)
6443: \rangle = (2\pi)^3 \delta^D(\bk_1+\bk_2) P_{T_b}(\bk_1)\ , \eeq where
6444: $\td_{T_b} (\bk)$ is the Fourier transform of $\delta_{T_b}$, $\bk$ is the
6445: comoving wavevector, $\delta^D$ is the Dirac delta function, and $\langle
6446: \cdots \rangle$ denotes an ensemble average.  In this analysis, we consider
6447: scales much bigger than the characteristic bubble size and the early phase
6448: of reionization (when ${\bar{\delta_{x_{\rm HI}}}}<<1$), so that the
6449: fluctuations $\delta_{x_{\rm HI}}$ are also much smaller than unity. For a
6450: more general treatment, see McQuinn et al. (2005) \cite{McQuinn}.
6451: 
6452: \bigskip
6453: \medskip
6454: \noindent{\bf The separation of powers}
6455: 
6456: The fluctuation $\delta_{T_b}$ consists of a number of isotropic
6457: sources of fluctuations plus the peculiar velocity term
6458: $-\delta_{d_rv_r}$. Its Fourier transform is simply proportional to
6459: that of the density field \citep{kaiser, Indian}, \beq \label{vgrad}
6460: \td_{d_rv_r} = -\mu^2 \td , \eeq where $\mu = \cos\theta_k$ in terms 
6461: of the angle $\theta_k$ of $\bk$ with respect to the line of
6462: sight. The $\mu^2$ dependence in this equation results from taking the
6463: radial (i.e., line-of-sight) component ($\propto \mu$) of the peculiar
6464: velocity, and then the radial component ($\propto \mu$) of its
6465: gradient. Intuitively, a high-density region possesses a velocity
6466: infall towards the density peak, implying that a photon must travel
6467: further from the peak in order to reach a fixed relative redshift,
6468: compared with the case of pure Hubble expansion. Thus the optical
6469: depth is always increased by this effect in regions with $\delta >
6470: 0$. This phenomenon is most properly termed {\it velocity
6471: compression}.
6472: 
6473: We therefore write the fluctuation in Fourier space as 
6474: \beq \label{Tbk} \td_{T_b} (\bk) = \mu^2 \td(\bk) + \beta \td(\bk) + 
6475: \td_{\rm rad}(\bk)\ , \eeq
6476: where we have defined a coefficient $\beta$ by collecting all terms
6477: $\propto \delta$ in Eq.~(\ref{dsignal}), and have also combined
6478: the terms that depend on the radiation fields of \Lya photons and
6479: ionizing photons, respectively. We assume that these radiation fields
6480: produce isotropic power spectra, since the physical processes that
6481: determine them have no preferred direction in space. The total power
6482: spectrum is
6483: \beqa \label{powTb}
6484: P_{T_b}(\bk) & = & \mu^4 P_{\delta}(k) + 2 \mu^2 [\beta P_{\delta}(k)
6485: + P_{\delta\cdot{\rm rad}}(k)]+ \nonumber \\ & & [\beta^2
6486: P_{\delta}(k) + P_{\rm rad}(k) + 2 \beta P_{\delta\cdot{\rm rad}}(k)]\
6487: ,
6488: \eeqa 
6489: where we have defined the power spectrum $P_{\delta\cdot{\rm
6490: rad}}$ as the Fourier transform of the cross-correlation 
6491: function,
6492: \beq \label{xi} 
6493: \xi_{\delta\cdot{\rm rad}} (r) =
6494: \langle \delta(\br_1)\, \delta_{\rm rad} (\br_1 + \br) \rangle\ .  
6495: \eeq 
6496: 
6497: We note that a similar anisotropy in the power spectrum has been
6498: previously derived in a different context, i.e., where the use of
6499: galaxy redshifts to estimate distances changes the apparent
6500: line-of-sight density of galaxies in redshift surveys
6501: \citep{kaiser,lilje,hamilton,fisher}. However, galaxies are
6502: intrinsically complex tracers of the underlying density field, and in
6503: that case there is no analog to the method that we demonstrate below
6504: for separating in 21cm fluctuations the effect of initial conditions
6505: from that of later astrophysical processes.
6506: 
6507: The velocity gradient term has also been examined for its global effect on
6508: the sky-averaged power and on radio visibilities \citep{t00, Indian}.
6509: The other sources of 21cm perturbations are isotropic and would
6510: produce a power spectrum $P_{T_b}(k)$ that could be measured by averaging
6511: the power over spherical shells in $\bk$ space. In the simple case where
6512: $\beta = 1$ and only the density and velocity terms contribute, the
6513: velocity term increases the total power by a factor of $\langle (1+\mu^2)^2
6514: \rangle = 1.87$ in the spherical average. However, instead of averaging the
6515: signal, we can use the angular structure of the power spectrum to greatly
6516: increase the discriminatory power of 21cm observations. We may break up
6517: each spherical shell in $\bk$ space into rings of constant $\mu$ and
6518: construct the observed $P_{T_b}(k,\mu)$. Considering Eq.~(\ref{powTb}) as a
6519: polynomial in $\mu$, i.e., $\mu^4 P_{\mu^4} + \mu^2 P_{\mu^2} + P_{\mu^0}$,
6520: we see that the power at just three values of $\mu$ is required in order to
6521: separate out the coefficients of 1, $\mu^2$, and $\mu^4$ for each $k$.
6522: 
6523: If the velocity compression were not present, then only the
6524: $\mu$-independent term (times $T_b^2$) would have been observed, and its
6525: separation into the five components ($T_b$, $\beta$, and three power
6526: spectra) would have been difficult and subject to degeneracies. Once the
6527: power has been separated into three parts, however, the $\mu^4$ coefficient
6528: can be used to measure the density power spectrum directly, with no
6529: interference from any other source of fluctuations. Since the overall
6530: amplitude of the power spectrum, and its scaling with redshift, are well
6531: determined from the combination of the CMB temperature fluctuations and
6532: galaxy surveys, the amplitude of $P_{\mu^4}$ directly determines the mean
6533: brightness temperature $T_b$ on the sky, which measures a combination of
6534: $T_s$ and $\bar{x}_{\rm HI}$ at the observed redshift. McQuinn et
6535: al. (2005) \cite{McQuinn} analysed in detail the parameters that can be
6536: constrained by upcoming 21cm experiments in concert with future CMB
6537: experiments such as Planck
6538: (http://www.rssd.esa.int/index.php?project=PLANCK).  Once $P_{\delta}(k)$
6539: has been determined, the coefficients of the $\mu^2$ term and the
6540: $\mu$-independent term must be used to determine the remaining unknowns,
6541: $\beta$, $P_{\delta\cdot{\rm rad}}(k)$, and $P_{\rm rad}(k)$. Since the
6542: coefficient $\beta$ is independent of $k$, determining it and thus breaking
6543: the last remaining degeneracy requires only a weak additional assumption on
6544: the behavior of the power spectra, such as their asymptotic behavior at
6545: large or small scales. If the measurements cover $N_k$ values of wavenumber
6546: $k$, then one wishes to determine $2 N_k + 1$ quantities based on $2 N_k$
6547: measurements, which should not cause significant degeneracies when $N_k \gg
6548: 1$. Even without knowing $\beta$, one can probe whether some sources of
6549: $P_{\rm rad}(k)$ are uncorrelated with $\delta$; the quantity $P_{\rm
6550: un-\delta}(k) \equiv P_{\mu^0}- P_{\mu^2}^2/(4 P_{\mu^4})$ equals $P_{\rm
6551: rad} - P_{\delta\cdot{\rm rad}}^2 / P_{\delta}$, which receives no
6552: contribution from any source that is a linear functional of the density
6553: distribution (see the next subsection for an example).
6554: 
6555: \noindent{\bf Specific epochs}
6556: 
6557: At $z \sim 35$, collisions are effective due to the high gas density,
6558: so one can measure the density power spectrum \citep{Loeb04} and the
6559: redshift evolution of $n_{\rm HI}$, $T_{\gamma}$, and $T_k$. At $z\la
6560: 35$, collisions become ineffective but the first stars produce a
6561: cosmic background of \Lya photons (i.e. photons that redshift into
6562: the \Lya resonance) that couples $T_s$ to $T_k$. During
6563: the period of initial \Lya coupling, fluctuations in the
6564: \Lya flux translate into fluctuations in the 21cm brightness
6565: \citep{BarkL05}. This signal can be observed from $z \sim 25$ until the
6566: \Lya coupling is completed (i.e., $x_{\rm tot} \gg 1$) at $z \sim
6567: 15$. At a given redshift, each atom sees \Lya photons that were
6568: originally emitted at earlier times at rest-frame wavelengths between
6569: \Lya and the Lyman limit. Distant sources are time retarded, and since
6570: there are fewer galaxies in the distant, earlier Universe, each atom
6571: sees sources only out to an apparent source horizon of $\sim 100$
6572: comoving Mpc at $z \sim 20$. A significant portion of the flux comes
6573: from nearby sources, because of the $1/r^2$ decline of flux with
6574: distance, and since higher Lyman series photons, which are degraded to
6575: \Lya photons through scattering, can only be seen from a small
6576: redshift interval that corresponds to the wavelength interval between
6577: two consecutive atomic levels.
6578: 
6579: \begin{figure}
6580: \centering
6581: \includegraphics[height=6cm]{losfig1.ps}
6582: \caption{Observable power spectra during the period of initial \Lya 
6583: coupling. {\it Upper panel:} Assumes adiabatic cooling. {\it Lower
6584: panel:} Assumes pre-heating to 500 K by X-ray sources. Shown are
6585: $P_{\mu^4}=P_{\delta}$ (solid curves), $P_{\mu^2}$ (short-dashed
6586: curves), and $P_{\rm un-\delta}$ (long-dashed curves), as well as
6587: for comparison $2 \beta P_{\delta}$ (dotted curves).}
6588: \label{Pkfig} 
6589: \end{figure}
6590: 
6591: There are two separate sources of fluctuations in the \Lya flux
6592: \citep{BarkL05}. The first is density inhomogeneities.  Since gravitational
6593: instability proceeds faster in overdense regions, the biased
6594: distribution of rare galactic halos fluctuates much more than the
6595: global dark matter density.  When the number of sources seen by each
6596: atom is relatively small, Poisson fluctuations provide a second source
6597: of fluctuations. Unlike typical Poisson noise, these fluctuations are
6598: correlated between gas elements at different places, since two nearby
6599: elements see many of the same sources. Assuming a scale-invariant
6600: spectrum of primordial density fluctuations, and that $x_{\alpha}=1$
6601: is produced at $z=20$ by galaxies in dark matter halos where the gas
6602: cools efficiently via atomic cooling, Figure \ref{Pkfig} shows the
6603: predicted observable power spectra. The figure suggests that $\beta$ can
6604: be measured from the ratio $P_{\mu^2} / P_{\mu^4}$ at $k \ga 1$
6605: Mpc$^{-1}$, allowing the density-induced fluctuations in flux to be
6606: extracted from $P_{\mu^2}$, while only the Poisson fluctuations
6607: contribute to $P_{\rm un-\delta}$. Each of these components probes the
6608: number density of galaxies through its magnitude, and the distribution
6609: of source distances through its shape. Measurements at $k \ga 100$
6610: Mpc$^{-1}$ can independently probe $T_k$ because of the smoothing
6611: effects of the gas pressure and the thermal width of the 21cm line.
6612: 
6613: After \Lya coupling and X-ray heating are both completed, reionization
6614: continues. Since $\beta = 1$ and $T_k \gg T_{\gamma}$, the
6615: normalization of $P_{\mu^4}$ directly measures the mean neutral
6616: hydrogen fraction, and one can separately probe the density
6617: fluctuations, the neutral hydrogen fluctuations, and their
6618: cross-correlation.
6619: 
6620: \noindent{\bf Fluctuations on large angular scales}
6621: 
6622: Full-sky observations must normally be analyzed with an angular and
6623: radial transform \citep{FZH, Santos, Indian}, rather than a
6624: Fourier transform which is simpler and yields more directly the
6625: underlying 3D power spectrum \citep{Miguel1, Miguel2}. The 21cm
6626: brightness fluctuations at a given redshift -- corresponding to a
6627: comoving distance $r_0$ from the observer -- can be expanded in
6628: spherical harmonics with expansion coefficients $a_{lm}(\nu)$, where
6629: the angular power spectrum is
6630: \beqa \label{cldef} C_{l}(r_0) & = & \langle |a_{lm}(\nu)|^2 \rangle 
6631: =4 \pi \int \frac {k^2 dk} {2 \pi^2} \biggl[ G_l^2(k r_0)
6632: P_{\delta}(k) + \nonumber \\ & & 2 P_{\delta\cdot{\rm rad}}(k) G_l(k
6633: r_0) j_l(k r_0) + P_{\rm rad}(k) j_l^2(k r_0) \biggl] \ , \eeqa with
6634: $G_l(x) \equiv J_l(x) + (\beta - 1) j_l(x)$ and $J_l(x)$ being a
6635: linear combination of spherical Bessel functions \citep{Indian}.
6636: 
6637: In an angular transform on the sky, an angle of $\theta$ radians
6638: translates to a spherical multipole $l \sim 3.5/ \theta$. For
6639: measurements on a screen at a comoving distance $r_0$, a multipole $l$
6640: normally measures 3D power on a scale of $k^{-1} \sim \theta r_0 \sim
6641: 35/l$ Gpc for $l\gg 1$, since $r_0 \sim 10$ Gpc at $z \ga 10$. This
6642: estimate fails at $l \la 100$, however, when we consider the sources
6643: of 21cm fluctuations. The angular projection implied in $C_l$ involves
6644: a weighted average (Eq.~\ref{cldef}) that favors large scales
6645: when $l$ is small, but density fluctuations possess little large-scale
6646: power, and the $C_l$ are dominated by power around the peak of $k
6647: P_\delta(k)$, at a few tens of comoving Mpc.
6648: 
6649: \begin{figure}
6650: \centering
6651: \includegraphics[height=6cm]{losfig2.ps}
6652: \caption{Effect of large-scale power on the angular 
6653: power spectrum of 21cm anisotropies on the sky. This example shows the
6654: power from density fluctuations and velocity compression, assuming a
6655: warm IGM at $z=12$ with $T_s=T_k \gg T_{\gamma}$. Shown is the $\%$
6656: change in $C_l$ if we were to cut off the power spectrum above $1/k$
6657: of 200, 180, 160, 140, 120, and 100 Mpc (top to bottom). Also shown
6658: for comparison is the cosmic variance for averaging in bands of
6659: $\Delta l \sim l$ (dashed lines).}
6660: \label{fclfig} 
6661: \end{figure}
6662: 
6663: Figure \ref{fclfig} shows that for density and velocity fluctuations, even
6664: the $l=1$ multipole is affected by power at $k^{-1} > 200$ Mpc only at the
6665: $2\%$ level. Due to the small number of large angular modes available on
6666: the sky, the expectation value of $C_l$ cannot be measured precisely at
6667: small $l$. Figure \ref{fclfig} shows that this precludes new information
6668: from being obtained on scales $k^{-1} \ga 130$ Mpc using angular structure
6669: at any given redshift. Fluctuations on such scales may be measurable using
6670: a range of redshifts, but the required $\Delta z \ga 1$ at $z \sim 10$
6671: implies significant difficulties with foreground subtraction and with the
6672: need to account for time evolution.
6673: 
6674: \section{Major Challenge for Future Theoretical Research: 
6675: {\it radiative transfer during reionization requires a large dynamic range,
6676: challenging the capabilities of existing simulation codes}}
6677: 
6678: Observations of the cosmic microwave background \citep{WMAP}
6679: have confirmed the notion that the present large-scale structure in
6680: the Universe originated from small-amplitude density fluctuations at
6681: early cosmic times. Due to the natural instability of gravity, regions
6682: that were denser than average collapsed and formed bound halos, first
6683: on small spatial scales and later on larger and larger scales. At each
6684: snapshot of this cosmic evolution, the abundance of collapsed halos,
6685: whose masses are dominated by cold dark matter, can be computed from
6686: the initial conditions using numerical simulations and can be
6687: understood using approximate analytic models \citep{ps74, bond91}. The
6688: common understanding of galaxy formation is based on the notion that
6689: the constituent stars formed out of the gas that cooled and
6690: subsequently condensed to high densities in the cores of some of these
6691: halos \citep{wr78}.
6692: 
6693: The standard analytic model for the abundance of halos \citep{ps74,
6694: bond91} considers the small density fluctuations at some early,
6695: initial time, and attempts to predict the number of halos that will
6696: form at some later time corresponding to a redshift $z$. First, the
6697: fluctuations are extrapolated to the present time using the growth
6698: rate of linear fluctuations, and then the average density is computed
6699: in spheres of various sizes. Whenever the overdensity (i.e., the
6700: density perturbation in units of the cosmic mean density) in a sphere
6701: rises above a critical threshold $\delta_c(z)$, the corresponding
6702: region is assumed to have collapsed by redshift $z$, forming a halo
6703: out of all the mass that had been included in the initial spherical
6704: region. In analyzing the statistics of such regions, the model
6705: separates the contribution of large-scale modes from that of
6706: small-scale density fluctuations. It predicts that galactic halos will
6707: form earlier in regions that are overdense on large scales \citep{k84,
6708: b86, ck89, mw96}, since these regions already start out from an
6709: enhanced level of density, and small-scale modes need only supply the
6710: remaining perturbation necessary to reach $\delta_c(z)$. On the other
6711: hand, large-scale voids should contain a reduced number of halos at
6712: high redshift. In this way, the analytic model describes the
6713: clustering of massive halos.
6714: 
6715: As gas falls into a dark matter halo, it can fragment into stars only if
6716: its virial temperature is above $10^4$K for cooling mediated by atomic
6717: transitions [or $\sim 500$ K for molecular ${\rm H}_2$ cooling; see
6718: Fig. \ref{cooling}]. The abundance of dark matter halos with a virial
6719: temperature above this cooling threshold declines sharply with increasing
6720: redshift due to the exponential cutoff in the abundance of massive halos at
6721: early cosmic times. Consequently, a small change in the collapse threshold
6722: of these rare halos, due to mild inhomogeneities on much larger spatial
6723: scales, can change the abundance of such halos dramatically. Barkana \&
6724: Loeb (2004) \cite{BL04a} have shown that the modulation of galaxy formation
6725: by long wavelength modes of density fluctuations is therefore amplified
6726: considerably at high redshift; the discussion in this section follows their
6727: analysis.
6728: 
6729: \noindent{\bf Amplification of Density Fluctuations}
6730: 
6731: Galaxies at high redshift are believed to form in all halos above some
6732: minimum mass $M_{\rm min}$ that depends on the efficiency of atomic and
6733: molecular transitions that cool the gas within each halo. This makes useful
6734: the standard quantity of the collapse fraction $F_{\rm col}(M_{\rm min})$,
6735: which is the fraction of mass in a given volume that is contained in halos
6736: of individual mass $M_{\rm min}$ or greater (see Fig. \ref{collapsed}). If
6737: we set $M_{\rm min}$ to be the minimum halo mass in which efficient cooling
6738: processes are triggered, then $F_{\rm col}(M_{\rm min})$ is the fraction of
6739: all baryons that reside in galaxies. In a large-scale region of comoving
6740: radius $R$ with a mean overdensity $\bar{\delta}_R$, the standard result is
6741: \citep{bond91}
6742: \begin{equation} F_{\rm col}(M_{\rm min})={\rm erfc}\left[
6743: \frac{\delta_c(z)- \bar{\delta}_R} {\sqrt{2 \left[S(R_{\rm min}) - S(R)
6744: \right]}} \right]\ , \label{eq:Fcol} \end{equation} where
6745: $S(R)=\sigma^2(R)$ is the variance of fluctuations in spheres of radius
6746: $R$, and $S(R_{\rm min})$ is the variance in spheres of radius $R_{\rm
6747: min}$ corresponding to the region at the initial time that contained a mass
6748: $M_{\rm min}$. In particular, the cosmic mean value of the collapse
6749: fraction is obtained in the limit of $R \rightarrow \infty$ by setting
6750: $\bar{\delta}_R$ and $S(R)$ to zero in this expression. Throughout this
6751: section we shall adopt this standard model, known as the extended
6752: Press-Schechter model. Whenever we consider a cubic region, we will
6753: estimate its halo abundance by applying the model to a spherical region of
6754: equal volume. Note also that we will consistently quote values of comoving
6755: distance, which equals physical distance times a factor of $(1+z)$.
6756: 
6757: At high redshift, galactic halos are rare and correspond to high peaks in
6758: the Gaussian probability distribution of initial fluctuations. A modest
6759: change in the overall density of a large region modulates the threshold for
6760: high peaks in the Gaussian density field, so that the number of galaxies is
6761: exponentially sensitive to this modulation. This amplification of
6762: large-scale modes is responsible for the large statistical fluctuations
6763: that we find.
6764: 
6765: In numerical simulations, periodic boundary conditions are usually assumed,
6766: and this forces the mean density of the box to equal the cosmic mean
6767: density. The abundance of halos as a function of mass is then biased in
6768: such a box (see Fig.~\ref{bias}), since a similar region in the real
6769: Universe will have a distribution of different overdensities
6770: $\bar{\delta}_R$.  At high redshift, when galaxies correspond to high
6771: peaks, they are mostly found in regions with an enhanced large-scale
6772: density. In a periodic box, therefore, the total number of galaxies is
6773: artificially reduced, and the relative abundance of galactic halos with
6774: different masses is artificially tilted in favor of lower-mass halos. Let
6775: us illustrate these results for two sets of parameters, one corresponding
6776: to the first galaxies and early reionization ($z=20$) and the other to the
6777: current horizon in observations of galaxies and late reionization
6778: ($z=7$). Let us consider a resolution equal to that of state-of-the-art
6779: cosmological simulations that include gravity and gas
6780: hydrodynamics. Specifically, let us assume that the total number of dark
6781: matter particles in the simulation is $N = 324^3$, and that the smallest
6782: halo that can form a galaxy must be resolved into 500 particles;
6783: \citet{converge} showed that this resolution is necessary in order to
6784: determine the star formation rate in an individual halo reliably to within
6785: a factor of two. Therefore, if we assume that halos that cool via molecular
6786: hydrogen must be resolved at $z=20$ (so that $M_{\rm min}=7 \times 10^5
6787: M_{\odot}$), and only those that cool via atomic transitions must be
6788: resolved at $z=7$ (so that $M_{\rm min}=10^8 M_{\odot}$), then the maximum
6789: box sizes that can currently be simulated in hydrodynamic comological
6790: simulations are $l_{\rm box}=1$ Mpc and $l_{\rm box}=6$ Mpc at these two
6791: redshifts, respectively.
6792: 
6793: \begin{figure} 
6794: \centering
6795: \includegraphics[height=6cm]{Nbody1.eps} 
6796: \caption{Bias in the halo mass distribution in simulations. Shown is the
6797: amount of mass contained in all halos of individual mass $M_{\rm min}$ or
6798: greater, expressed as a fraction of the total mass in a given volume. This
6799: cumulative fraction $F_{\rm col}(M_{\rm min})$ is illustrated as a function
6800: of the minimum halo mass $M_{\rm min}$. We consider two cases of redshift
6801: and simulation box size, namely $z=7$, $l_{\rm box}=6$ Mpc (upper curves),
6802: and $z=20$, $l_{\rm box}=1$ Mpc (lower curves). At each redshift, we
6803: compare the true average distribution in the Universe (dotted curve) to the
6804: biased distribution (solid curve) that would be measured in a simulation
6805: box with periodic boundary conditions (for which $\bar{\delta}_R$ is
6806: artificially set to zero).}
6807: \label{bias}
6808: \end{figure}
6809: 
6810: At each redshift we only consider cubic boxes large enough so that the
6811: probability of forming a halo on the scale of the entire box is
6812: negligible. In this case, $\bar{\delta}_R$ is Gaussian distributed with
6813: zero mean and variance $S(R)$, since the no-halo condition $\sqrt{S(R)} \ll
6814: \del_c(z)$ implies that at redshift $z$ the perturbation on the scale $R$
6815: is still in the linear regime. We can then calculate the probability
6816: distribution of collapse fractions in a box of a given size (see
6817: Fig.\ref{prob}). This distribution corresponds to a real variation in the
6818: fraction of gas in galaxies within different regions of the Universe at a
6819: given time. In a numerical simulation, the assumption of periodic boundary
6820: conditions eliminates the large-scale modes that cause this cosmic
6821: scatter. Note that Poisson fluctuations in the number of halos within the
6822: box would only add to the scatter, although the variations we have
6823: calculated are typically the dominant factor. For instance, in our two
6824: standard examples, the mean expected number of halos in the box is 3 at
6825: $z=20$ and 900 at $z=7$, resulting in Poisson fluctuations of a factor of
6826: about 2 and 1.03, respectively, compared to the clustering-induced scatter
6827: of a factor of about 16 and 2 in these two cases.
6828: 
6829: \begin{figure}
6830: \centering
6831: \includegraphics[height=8cm]{Nbody2.eps}
6832: \caption{Probability distribution within a small volume of the total
6833: mass fraction in galactic halos. The normalized distribution of the
6834: logarithm of this fraction $F_{\rm col}(M_{\rm min})$ is shown for two
6835: cases: $z=7$, $l_{\rm box}=6$ Mpc, $M_{\rm min}=10^8 M_{\odot}$ (upper
6836: panel), and $z=20$, $l_{\rm box}=1$ Mpc, $M_{\rm min}=7 \times 10^5
6837: M_{\odot}$ (bottom panel). In each case, the value in a periodic box
6838: ($\bar{\delta}_R=0$) is shown along with the value that would be
6839: expected given a plus or minus $1-\sigma$ fluctuation in the mean
6840: density of the box (dashed vertical lines). Also shown in each case is
6841: the mean value of $F_{\rm col}(M_{\rm min})$ averaged over large
6842: cosmological volumes (solid vertical line).}  
6843: \label{prob}
6844: \end{figure}
6845: 
6846: Within the extended Press-Schechter model, both the numerical bias and the
6847: cosmic scatter can be simply described in terms of a shift in the redshift
6848: (see Fig. \ref{shifti}). In general, a region of radius $R$ with a mean
6849: overdensity $\bar{\delta}_R$ will contain a different collapse fraction
6850: than the cosmic mean value at a given redshift $z$. However, at some wrong
6851: redshift $z + \Delta z$ this small region will contain the cosmic mean
6852: collapse fraction at $z$. At high redshifts ($z > 3$), this shift in
6853: redshift was derived by Barkana \& Loeb \cite{BL04a}
6854: from equation~(\ref{eq:Fcol})
6855: [and was already mentioned in Eq. (\ref{scatter})]
6856: \begin{equation} \Delta z = \frac{\bar{\delta}_R}{\delta_0} - (1+z)
6857: \times \left[ 1 - \sqrt{1 - \frac{S(R)} {S(R_{\rm min})}}\ \right]\ ,
6858: \end{equation} where $\delta_0 \equiv \delta_c(z)/(1+z)$ is
6859: approximately constant at high redshifts \citep{p80}, and equals 1.28 for
6860: the standard cosmological parameters (with its deviation from the
6861: Einstein-de Sitter value of 1.69 resulting from the existence of a
6862: cosmological constant). Thus, in our two examples, the bias is -2.6 at
6863: $z=20$ and -0.4 at $z=7$, and the one-sided $1-\sigma$ scatter is 2.4 at
6864: $z=20$ and 1.2 at $z=7$.
6865: 
6866: \begin{figure}
6867: \centering
6868: \includegraphics[height=8cm]{Nbody3.eps}
6869: \caption{Cosmic scatter and numerical bias, expressed as the change in
6870: redshift needed to get the correct cosmic mean of the collapse
6871: fraction. The plot shows the $1-\sigma$ scatter (about the biased value) in
6872: the redshift of reionization, or any other phenomenon that depends on the
6873: mass fraction in galaxies (bottom panel), as well as the redshift bias
6874: [expressed as a fraction of $(1+z)$] in periodic simulation boxes (upper
6875: panel). The bias is shown for $M_{\rm min}=7 \times 10^5 M_{\odot}$ (solid
6876: curve), $M_{\rm min}=10^8 M_{\odot}$ (dashed curve), and $M_{\rm min}=3
6877: \times 10^{10} M_{\odot}$ (dotted curve). The bias is always negative, and
6878: the plot gives its absolute value. When expressed as a shift in redshift, the
6879: scatter is independent of $M_{\rm min}$.}
6880: \label{shifti}
6881: \end{figure}
6882: 
6883: \noindent{\bf Matching Numerical Simulations}
6884: 
6885: \label{massfn}
6886: 
6887: Next we may develop an improved model that fits the results
6888: of numerical simulations more accurately. The model constructs the
6889: halo mass distribution (or mass function); cumulative quantities such
6890: as the collapse fraction or the total number of galaxies can then be
6891: determined from it via integration. We first define $f(\del_c(z),S)\,
6892: dS$ to be the mass fraction contained at $z$ within halos with mass in
6893: the range corresponding to $S$ to $S+d S$. 
6894: As derived earlier, the Press-Schechter halo abundance is 
6895: \beq \frac{dn}{dM} = \frac{\bar{\rho}_0}{M} \left|\frac{d S}{d M}
6896: \right| f(\del_c(z),S)\ , \label{eq:abundance} \eeq where $dn$ is the
6897: comoving number density of halos with masses in the range $M$ to
6898: $M+dM$, and
6899: \beq f_{\rm PS}(\del_c(z),S) =
6900: \frac{1} {\sqrt{2 \pi}} \frac{\nu }{S} \exp\left[-\frac{\nu^2}{2}
6901: \right]\ , \label{eq:PS} \eeq where $\nu=\del_c(z)/\sqrt{S}$ is the
6902: number of standard deviations that the critical collapse overdensity
6903: represents on the mass scale $M$ corresponding to the variance $S$.
6904: 
6905: However, the Press-Schechter mass function fits numerical simulations only
6906: roughly, and in particular it substantially underestimates the abundance of
6907: the rare halos that host galaxies at high redshift. The halo mass function
6908: of \cite{shetht99} [see also \cite{shethmot}] adds two free parameters that
6909: allow it to fit numerical simulations much more accurately
6910: \citep{jenkins}. These N-body simulations followed very large volumes
6911: at low redshift, so that cosmic scatter did not compromise their
6912: accuracy. The matching mass function is given by \beq f_{\rm
6913: ST}(\del_c(z),S) = A' \frac{\nu }{S} \sqrt{\frac{a'} {2 \pi}} \left[
6914: 1+\frac{1}{(a' \nu^2)^{q'}} \right] \exp\left[-\frac{a' \nu^2}{2} \right]\
6915: , \label{eq:ST} \eeq with best-fit parameters \citep{shetht02} $a'=0.75$
6916: and $q'=0.3$, and where normalization to unity is ensured by taking
6917: $A'=0.322$.
6918: 
6919: In order to calculate cosmic scatter we must determine the biased halo
6920: mass function in a given volume at a given mean density. Within the
6921: extended Press-Schechter model \citep{bond91}, the halo mass
6922: distribution in a region of comoving radius $R$ with a mean
6923: overdensity $\bar{\delta}_R$ is given by \beq f_{\rm
6924: bias-PS}(\del_c(z), \bar{\delta}_R,R,S)=f_{\rm PS}(\del_c(z)-
6925: \bar{\delta}_R,S-S(R))\ \label{eq:ePS}. \eeq The corresponding collapse
6926: fraction in this case is given simply by eq.~(\ref{eq:Fcol}). Despite
6927: the relatively low accuracy of the Press-Schechter mass function, the
6928: {\it relative change} is predicted rather accurately by the extended
6929: Press-Schechter model. In other words, the prediction for the halo
6930: mass function in a given volume compared to the cosmic mean mass
6931: function provides a good fit to numerical simulations over a wide
6932: range of parameters \citep{mw96,casas02}.
6933: 
6934: For the improved model (derived in \cite{BL04a}), we adopt a hybrid
6935: approach that combines various previous models with each applied where it
6936: has been found to closely match numerical simulations. We obtain the halo
6937: mass function within a restricted volume by starting with the Sheth-Torme
6938: formula for the cosmic mean mass function, and then adjusting it with a
6939: relative correction based on the extended Press-Schechter model. In other
6940: words, we set \ba & & f_{\rm bias}(\del_c(z),\bar{\delta}_R,R,S) =
6941: \nonumber \\ & & f_{\rm ST}(\del_c(z),S)\ \times \left[ \frac{f_{\rm PS}
6942: (\del_c(z)-\bar{\delta}_R,S-S(R))} {f_{\rm PS}(\del_c(z),S)} \right]\
6943: . \label{eq:bias} \end{eqnarray} As noted, this model is based on fits to
6944: simulations at low redshifts, but we can check it at high redshifts as
6945: well. Figure~\ref{fig-Lars} shows the number of galactic halos at $z \sim
6946: 15-30$ in two numerical simulations run by \citet{yoshida}, and our
6947: predictions given the cosmological input parameters assumed by each
6948: simulation. The close fit to the simulated data (with no additional free
6949: parameters) suggests that our hybrid model (solid lines) improves on the
6950: extended Press-Schechter model (dashed lines), and can be used to calculate
6951: accurately the cosmic scatter in the number of galaxies at both high and
6952: low redshifts. The simulated data significantly deviate from the expected
6953: cosmic mean [eq.~(\ref{eq:ST}), shown by the dotted line], due to the
6954: artificial suppression of large-scale modes outside the simulated box.
6955: 
6956: \begin{figure}
6957: \centering
6958: \includegraphics[height=8cm]{Nbody4.eps} 
6959: \caption{Halo mass function at high redshift in a 1 Mpc box at the
6960: cosmic mean density. The prediction  (solid lines) of
6961: the hybrid model of Barkana \& Loeb (2004) \cite{BL04a} is
6962: compared with the number of halos above mass $7 \times 10^5 M_{\odot}$
6963: measured in the simulations of \citet{yoshida} [data points are taken from
6964: their Figure~5]. The cosmic mean of the halo mass function
6965: (dotted lines) deviates significantly from the simulated values, since
6966: the periodic boundary conditions within the finite simulation box
6967: artificially set the amplitude of large-scale modes to zero. The
6968: hybrid model starts with the Sheth-Tormen mass function and applies a
6969: correction based on the extended Press-Schechter model; in doing so,
6970: it provides a better fit to numerical simulations than the pure
6971: extended Press-Schechter model (dashed lines) used in the previous
6972: figures. We consider two sets of cosmological parameters, the
6973: scale-invariant $\Lambda$CDM model of \citet{yoshida} (upper curves),
6974: and their running scalar index (RSI) model (lower curves).}
6975: \label{fig-Lars} 
6976: \end{figure}
6977: 
6978: As an additional example, we consider the highest-resolution first
6979: star simulation \citep{ABN02}, which used $l_{\rm box}=128$ kpc and
6980: $M_{\rm min}=7 \times 10^5 M_{\odot}$.  The first star forms within
6981: the simulated volume when the first halo of mass $M_{\rm min}$ or
6982: larger collapses within the box. To compare with the simulation, we
6983: predict the redshift at which the probability of finding at least one
6984: halo within the box equals $50\%$, accounting for Poisson
6985: fluctuations. We find that if the simulation formed a population of
6986: halos corresponding to the correct cosmic average [as given by
6987: eq.~(\ref{eq:ST})], then the first star should have formed already at
6988: $z=24.0$. The first star actually formed in the simulation box only at
6989: $z=18.2$ \citep{ABN02}. Using eq.~(\ref{eq:bias}) we can account for
6990: the loss of large-scale modes beyond the periodic box, and predict a
6991: first star at $z=17.8$, a close match given the large Poisson
6992: fluctuations introduced by considering a single galaxy within the box.
6993: 
6994: The artificial bias in periodic simulation boxes can also be seen in
6995: the results of extensive numerical convergence tests carried out by
6996: \citet{converge}. They presented a large array of numerical
6997: simulations of galaxy formation run in periodic boxes over a wide
6998: range of box size, mass resolution, and redshift. In particular, we
6999: can identify several pairs of simulations where the simulations in
7000: each pair have the same mass resolution but different box sizes; this
7001: allows us to separate the effect of large-scale numerical bias from
7002: the effect of having poorly-resolved individual halos. 
7003: %Specifically,
7004: %their simulations Z1 and R4 [see Table~1 in \citet{converge}] used the
7005: %same particle mass but R4 had a box length larger by a factor of 3.375
7006: %. The simulations Q1 and D4 are similarly related, as are Q2 and
7007: %D5. In each case, the smaller simulation substantially underestimated
7008: %the star formation rate at high redshift [see Figure~10 in
7009: %\citet{converge}], Z1 by a factor of 5 at $z=15$ compared to R4, Q1 by
7010: %a factor of 3 at $z=8$ compared to D4, and Q2 by a factor of 3 at
7011: %$z=10$ compared to D5.
7012: 
7013: %There have been previous attempts to develop a model for
7014: %the halo mass function in different environments, so that the model
7015: %would be consistent with the Sheth-Tormen mass function of
7016: %eq.~(\ref{eq:ST}) which accurately fits the cosmic mean mass function
7017: %measured in numerical simulations. In order to identify the specific
7018: %requirements for such a consistency, we first consider the analogous
7019: %case of the extended Press-Schechter model and its relation to the
7020: %Press-Schechter formula for the cosmic mean mass function. The
7021: %extended Press-Schechter model is consistent with the mean mass
7022: %function in the sense that $f_{\rm bias-PS}$ evaluated for an infinite
7023: %box (i.e., in the limit where $\bar{\delta}_R$ and $S(R)$ both vanish)
7024: %yields the Press-Schechter mass function: \beq f_{\rm bias-PS}
7025: %(\del_c(z), 0,\infty,S)=f_{\rm PS}(\del_c(z),S)\ . \label{eq:cons} \eeq
7026: %This condition does not suffice, however, since there is an additional
7027: %self-consistency test that any viable model must satisfy. Consider any
7028: %fixed scale $R$. Suppose we consider a very large number $N$ of
7029: %spheres of radius $R$ within the Universe. The mean density
7030: %$\bar{\delta}_R$ in each sphere is determined according to a
7031: %probability distribution $p(\bar{\delta}_R)$; we assume that $R$ is
7032: %large enough so that the probability of forming a halo out of all the
7033: %mass on the scale $R$ is negligible, and so the distribution is a
7034: %Gaussian with zero mean and variance $S(R)$ (see also \S 2.1). The
7035: %number of galaxies in each sphere is given in the extended
7036: %Press-Schechter model by eq.~(\ref{eq:ePS}). As $N \rightarrow
7037: %\infty$, the halo mass function averaged over all these spheres must
7038: %approach the cosmic mean value, and it must also approach the
7039: %ensemble-averaged mass function, where the averaging is performed over
7040: %the probability distribution of $\bar{\delta}_R$. This yields the
7041: %following self-consistency requirement, which is indeed satisfied by
7042: %the extended Press-Schechter model: \begin{eqnarray} & & \int f_{\rm
7043: %bias-PS}(\del_c(z), \bar{\delta}_R,R,S)\, p(\bar{\delta}_R)\,
7044: %d\bar{\delta}_R = \nonumber \\ & & f_{\rm bias-PS}(\del_c(z),
7045: %0,\infty,S)\ .  \label{eq:scons} \end{eqnarray}
7046: %
7047: %Now we again consider attempts to construct an improved model that is
7048: %consistent with the Sheth-Tormen mass function. Such a model must
7049: %satisfy eq.~(\ref{eq:cons}) (except with $f_{\rm ST}$ on the
7050: %right-hand side), and it must also satisfy eq.~(\ref{eq:scons}) in
7051: %order to be self-consistent. The latter equation must be satisfied
7052: %separately for every scale $R$ large enough to avoid collapsing [i.e.,
7053: %that satisfies $\sqrt{S(R)} \ll \del_c(z)$]. Previous proposed models
7054: %\citep{shetht02, gottl03} satisfied simple consistency but not the
7055: %self-consistency test. Our hybrid model of eq.~(\ref{eq:bias})
7056: %satisfies both requirements (with respect to the Sheth-Tormen mass
7057: %function), a result that follows immediately from the fact that the
7058: %extended Press-Schechter model also satisfies both requirements (with
7059: %respect to the Press-Schechter mass function). Thus, our hybrid model
7060: %is the first self-consistent model that is also consistent with the
7061: %Sheth-Tormen mass function. As demonstrated in this section, it also
7062: %matches results from a wide array of numerical simulations
7063: %\citep{ABN02, yoshida, mw96, casas02, jenkins}.
7064: 
7065: \bigskip
7066: \noindent{\bf Implications}
7067: 
7068: \noindent
7069: {\it (i) The nature of reionization} 
7070: 
7071: A variety of papers in the literature \citep{aw72, fk94, sgb94,
7072: hl97, g00, BL01, BL01a} maintain that reionization ended with a fast,
7073: simultaneous, overlap stage throughout the Universe. This view has been
7074: based on simple arguments and has been supported by numerical simulations
7075: with small box sizes. The underlying idea was that the ionized hydrogen (H
7076: II ) regions of individual sources began to overlap when the typical size
7077: of each H II bubble became comparable to the distance between nearby
7078: sources. Since these two length scales were comparable at the critical
7079: moment, there is only a single timescale in the problem -- given by the
7080: growth rate of each bubble -- and it determines the transition time between
7081: the initial overlap of two or three nearby bubbles, to the final stage
7082: where dozens or hundreds of individual sources overlap and produce large
7083: ionized regions. Whenever two ionized bubbles were joined, each point
7084: inside their common boundary became exposed to ionizing photons from both
7085: sources, reducing the neutral hydrogen fraction and allowing ionizing
7086: photons to travel farther before being absorbed. Thus, the ionizing
7087: intensity inside H II regions rose rapidly, allowing those regions to
7088: expand into high-density gas that had previously recombined fast enough to
7089: remain neutral when the ionizing intensity had been low. Since each bubble
7090: coalescence accelerates the process, it has been thought that the overlap
7091: phase has the character of a phase transition and occurs rapidly. Indeed,
7092: the simulations of reionization \citep{g00} found that the
7093: average mean free path of ionizing photons in the simulated volume rises by
7094: an order of magnitude over a redshift interval $\Delta z = 0.05$ at $z=7$.
7095: 
7096: These results imply that overlap is still expected to occur rapidly, but
7097: only in localized high-density regions, where the ionizing intensity and
7098: the mean free path rise rapidly even while other distant regions are still
7099: mostly neutral. In other words, the size of the bubble of an individual
7100: source is about the same in different regions (since most halos have masses
7101: just above $M_{\rm min}$), but the typical distance between nearby sources
7102: varies widely across the Universe. The strong clustering of ionizing
7103: sources on length scales as large as 30--100 Mpc introduces long timescales
7104: into the reionization phase transition. The sharpness of overlap is
7105: determined not by the growth rate of bubbles around individual sources, but
7106: by the ability of large groups of sources within overdense regions to
7107: deliver ionizing photons into large underdense regions. 
7108: 
7109: Note that the recombination rate is higher in overdense regions
7110: because of their higher gas density. These regions still reionize
7111: first, though, despite the need to overcome the higher recombination
7112: rate, since the number of ionizing sources in these regions is
7113: increased even more strongly as a result of the dramatic amplification
7114: of large-scale modes discussed earlier.
7115: 
7116: \noindent
7117: {\it (ii) Limitations of current simulations} 
7118: 
7119: The shortcomings of current simulations do not amount simply to a shift of
7120: $\sim 10\%$ in redshift and the elimination of scatter. The effect
7121: mentioned above can be expressed in terms of a shift in redshift only
7122: within the context of the extended Press-Schechter model, and only if the
7123: total mass fraction in galaxies is considered and not its distribution as a
7124: function of galaxy mass. The halo mass distribution should still have the
7125: wrong shape, resulting from the fact that $\Delta z$ depends on $M_{\rm
7126: min}$. A self-contained numerical simulation must directly evolve a very
7127: large volume.
7128: 
7129: Another reason that current simulations are limited is that at high
7130: redshift, when galaxies are still rare, the abundance of galaxies grows
7131: rapidly towards lower redshift. Therefore, a $\sim 10\%$ relative error in
7132: redshift implies that at any given redshift around $z \sim 10$--20, the
7133: simulation predicts a halo mass function that can be off by an order of
7134: magnitude for halos that host galaxies (see Fig. \ref{fig-Lars}). This
7135: large underestimate suggests that the first generation of galaxies formed
7136: significantly earlier than indicated by recent simulations.  Another
7137: element missed by simulations is the large cosmic scatter. This scatter can
7138: fundamentally change the character of any observable process or feedback
7139: mechanism that depends on a radiation background. Simulations in periodic
7140: boxes eliminate any large-scale scatter by assuming that the simulated
7141: volume is surrounded by identical periodic copies of itself. In the case of
7142: reionization, for instance, current simulations neglect the collective
7143: effects described above, whereby groups of sources in overdense regions may
7144: influence large surrounding underdense regions. In the case of the
7145: formation of the first stars due to molecular hydrogen cooling, the effect
7146: of the soft ultraviolet radiation from these stars, which tends to
7147: dissociate the molecular hydrogen around them \citep{hrl97, rgs02, Oh03},
7148: must be reassessed with cosmic scatter included.
7149: 
7150: \noindent{\it (iii) Observational consequences} 
7151: 
7152: The spatial fluctuations that we have calculated also affect current and
7153: future observations that probe reionization or the galaxy population at
7154: high redshift. For example, there are a large number of programs searching
7155: for galaxies at the highest accessible redshifts (6.5 and beyond) using
7156: their strong Ly$\alpha$ emission \citep{h02, r03, m03, k03}. These programs
7157: have previously been justified as a search for the reionization redshift,
7158: since the intrinsic emission should be absorbed more strongly by the
7159: surrounding IGM if this medium is neutral. For any particular source, it
7160: will be hard to clearly recognize this enhanced absorption because of
7161: uncertainties regarding the properties of the source and its radiative and
7162: gravitational effects on its surroundings
7163: \citep{nature,GRBquasar,s03}. However, if the luminosity function of
7164: galaxies that emit Ly$\alpha$ can be observed, then faint sources, which do
7165: not significantly affect their environment, should be very strongly
7166: absorbed in the era before reionization. Reionization can then be detected
7167: statistically through the sudden jump in the number of faint
7168: sources\cite{Rhoads}. The above results alter the expectation for such
7169: observations. Indeed, no sharp ``reionization redshift'' is
7170: expected. Instead, a Ly$\alpha$ luminosity function assembled from a large
7171: area of the sky will average over the cosmic scatter of $\Delta z \sim
7172: 1$--2 between different regions, resulting in a smooth evolution of the
7173: luminosity function over this redshift range. In addition, such a survey
7174: may be biased to give a relatively high redshift, since only the most
7175: massive galaxies can be detected, and as we have shown, these galaxies will
7176: be concentrated in overdense regions that will also get reionized
7177: relatively early.
7178: 
7179: The distribution of ionized patches during reionization will likely be
7180: probed by future observations, including small-scale anisotropies of the
7181: cosmic microwave background photons that are rescattered by the ionized
7182: patches \citep{a96, gh98, san03}, and observations of 21 cm emission by the
7183: spin-flip transition of the hydrogen in neutral regions \citep{t00, cgo02,
7184: fsh03}. Previous analytical and numerical estimates of these signals have
7185: not included the collective effects discussed above, in which rare groups
7186: of massive galaxies may reionize large surrounding areas. The transfer of
7187: photons across large scales will likely smooth out the signal even on
7188: scales significantly larger than the typical size of an H II bubble due to
7189: an individual galaxy. Therefore, even the characteristic angular scales
7190: that are expected to show correlations in such observations must be
7191: reassessed.
7192: 
7193: The cosmic scatter also affects observations in the present-day Universe
7194: that depend on the history of reionization. For instance, photoionization
7195: heating suppresses the formation of dwarf galaxies after reionization,
7196: suggesting that the smallest galaxies seen today may have formed prior to
7197: reionization \citep{bkw01, s02, b02}. Under the popular view that assumed a
7198: sharp end to reionization, it was expected that denser regions would have
7199: formed more galaxies by the time of reionization, possibly explaining the
7200: larger relative abundance of dwarf galaxies observed in galaxy clusters
7201: compared to lower-density regions such as the Local Group of galaxies
7202: \citep{t02, b03}. The above results undercut the basic assumption of this
7203: argument and suggest a different explanation. Reionization
7204: occurs roughly when the number of ionizing photons produced starts to
7205: exceed the number of hydrogen atoms in the surrounding IGM. If the
7206: processes of star formation and the production of ionizing photons are
7207: equally efficient within galaxies that lie in different regions, then
7208: reionization in each region will occur when the collapse fraction reaches
7209: the same critical value, even though this will occur at different times in
7210: different regions. Since the galaxies responsible for reionization have the
7211: same masses as present-day dwarf galaxies, this estimate argues for a
7212: roughly equal abundance of dwarf galaxies in all environments today. This
7213: simple picture is, however, modified by several additional effects. First,
7214: the recombination rate is higher in overdense regions at any given time, as
7215: discussed above. Furthermore, reionization in such regions is accomplished
7216: at an earlier time when the recombination rate was higher even at the mean
7217: cosmic density; therefore, more ionizing photons must be produced in order
7218: to compensate for the enhanced recombination rate. These two effects
7219: combine to make overdense regions reionize at a higher value of $F_{\rm
7220: col}$ than underdense regions. In addition, the overdense regions, which
7221: reionize first, subsequently send their extra ionizing photons into the
7222: surrounding underdense regions, causing the latter to reionize at an even
7223: lower $F_{\rm col}$. Thus, a higher abundance of dwarf galaxies today is
7224: indeed expected in the overdense regions.
7225: 
7226: The same basic effect may be even more critical for understanding the
7227: properties of large-scale voids, 10--30 Mpc regions in the present-day
7228: Universe with an average mass density that is well below the cosmic
7229: mean. In order to predict their properties, the first step is to
7230: consider the abundance of dark matter halos within them. Numerical
7231: simulations show that voids contain a lower relative abundance of rare
7232: halos \citep{mw02,co00,bh03}, as expected from the raising of the
7233: collapse threshold for halos within a void. On the other hand,
7234: simulations show that voids actually place a larger fraction of their
7235: dark matter content in dwarf halos of mass below $10^{10} M_{\odot}$
7236: \citep{gottl03}. This can be understood within the extended
7237: Press-Schechter model. At the present time, a typical region in the
7238: Universe fills halos of mass $10^{12} M_{\odot}$ and higher with most
7239: of the dark matter, and very little is left over for isolated dwarf
7240: halos. Although a large number of dwarf halos may have formed at early
7241: times in such a region, the vast majority later merged with other
7242: halos, and by the present time they survive only as substructure
7243: inside much larger halos. In a void, on the other hand, large halos
7244: are rare even today, implying that most of the dwarf halos that formed
7245: early within a void can remain as isolated dwarf halos till the
7246: present. Thus, most isolated dwarf dark matter halos in the present
7247: Universe should be found within large-scale voids \citep{infall}.
7248: 
7249: However, voids are observed to be rather deficient in dwarf galaxies as
7250: well as in larger galaxies on the scale of the Milky Way mass of $\sim
7251: 10^{12}M_\odot$ \cite{BigVoid,el-ad,pVoids}. A deficit of large galaxies
7252: is naturally expected, since the total mass density in the void is
7253: unusually low, and the fraction of this already low density that assembles
7254: in large halos is further reduced relative to higher-density regions. The
7255: absence of dwarf galaxies is harder to understand, given the higher
7256: relative abundance expected for their host dark matter halos. The standard
7257: model for galaxy formation may be consistent with the observations if some
7258: of the dwarf halos are dark and do not host stars. Large numbers of dark
7259: dwarf halos may be produced by the effect of reionization in suppressing
7260: the infall of gas into these halos. Indeed, exactly the same factors
7261: considered above, in the discussion of dwarf galaxies in clusters compared
7262: to those in small groups, apply also to voids. Thus, the voids should
7263: reionize last, but since they are most strongly affected by ionizing
7264: photons from their surroundings (which have a higher density than the voids
7265: themselves), the voids should reionize when the abundance of galaxies
7266: within them is relatively low.
7267: 
7268: \bigskip
7269: \bigskip
7270: \bigskip
7271: \noindent
7272: {\bf Acknowledgements}
7273: 
7274: I thank my young collaborators with whom my own research in this field was
7275: accomplished: Dan Babich, Rennan Barkana, Volker Bromm, Benedetta Ciardi,
7276: Daniel Eisenstein, Steve Furlanetto, Zoltan Haiman, Rosalba Perna, Stuart
7277: Wyithe, and Matias Zaldarriaga. I thank Donna Adams for her highly
7278: professional assistance with the latex file, and Dan Babich \& 
7279: Matt McQuinn for their helpful comments on the manuscript.
7280: 
7281: %
7282: %
7283: % BibTeX users please use
7284: % \bibliographystyle{}
7285: % \bibliography{}
7286: %
7287: % Non-BibTeX users please follow the syntax
7288: % the syntax of "referenc.tex" for your own citations
7289: \input{referenc}
7290: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%  }
7291: 
7292: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7293: \printindex
7294: \end{document}
7295: 
7296: \bibitem[Di Matteo et al.(2005)]{2005Natur.433..604D} Di Matteo, T., 
7297: Springel, V., \& Hernquist, L.\ 2005, \nat, 433, 604 
7298: 
7299: \bibitem[Springel et al.(2005)]{2005ApJ...620L..79S} Springel, V., Di 
7300: Matteo, T., \& Hernquist, L.\ 2005, \apjl, 620, L79 
7301:  
7302: \bibitem[Loeb \& Zaldarriaga(2004)]{2004PhRvL..92u1301L} Loeb, A., \& 
7303: Zaldarriaga, M.\ 2004, Physical Review Letters, 92, 211301 
7304:  
7305: \bibitem[Wyithe \& Loeb(2004)]{2004Natur.432..194W} Wyithe, J.~S.~B., \& 
7306: Loeb, A.\ 2004, \nat, 432, 194 
7307:  
7308: \bibitem[Barkana \& Loeb(2004)]{2004ApJ...609..474B} Barkana, R., \& Loeb, 
7309: A.\ 2004, \apj, 609, 474 
7310:  
7311:  
7312: \bibitem[Fan et al.(2005)]{Fan} Fan, X. et al. 2005, AJ, submitted, 
7313: astro-ph/0512082 
7314: 
7315: \bibitem[Barkana \& Loeb(2005)]{2005MNRAS.363L..36B} Barkana, R., \& Loeb, 
7316: A.\ 2005a, \mnras, 363, L36 
7317: 
7318: \bibitem[Wouthuysen(1952)]{1952AJ.....57R..31W} Wouthuysen, S.~A.\ 1952, 
7319: \aj, 57, 31 
7320:  
7321: \bibitem[Field(1959)]{1959ApJ...129..536F} Field, G.~B.\ 1959, \apj, 129, 
7322: 536 
7323: 
7324: \bibitem[Dijkstra et al.(2004)]{2004ApJ...601..666D} Dijkstra, M., Haiman, 
7325: Z., Rees, M.~J., \& Weinberg, D.~H.\ 2004, \apj, 601, 666 
7326: 
7327: \bibitem[Malhotra \& Rhoads(2005)]{lala} Malhotra, S., \& Rhoads, J. 2005,
7328: ApJL, submitted, astro-ph/0511196
7329: 
7330: \bibitem[Furlanetto \& Loeb(2001)]{2001ApJ...556..619F} Furlanetto, S.~R., 
7331: \& Loeb, A.\ 2001, \apj, 556, 619 
7332: 
7333: \bibitem[Furlanetto \& Loeb(2003)]{2003ApJ...588...18F} Furlanetto, S.~R., 
7334: \& Loeb, A.\ 2003, \apj, 588, 18 
7335: 
7336: \bibitem[Furlanetto \& Loeb(2005)]{2005ApJ...634....1F} Furlanetto, S.~R., 
7337: \& Loeb, A.\ 2005, \apj, 634, 1 
7338: 
7339: 
7340: %new
7341: 
7342: 
7343: \bibitem[McQuinn et al.(2005)]{McQuinn}
7344: McQuinn, M., Zahn, O., Zaldarriaga, M., Hernquist, L., \& Furlanetto, S. R.
7345: 2005 ApJ, submitted; preprint astro-ph/0512263
7346: 
7347: \bibitem[Barkana \& Loeb(2005)]{BL05}
7348: Barkana, R., \& Loeb, A. 2005, MNRAS, submitted; preprint astro-ph/0512453
7349: 
7350: \bibitem[Furlanetto et al.(2004)]{FZH} Furlanetto, S.~R., 
7351: Zaldarriaga, M., \& Hernquist, L.\ 2004, \apj, 613, 1 
7352: 
7353: \bibitem[Furlanetto \& Loeb(2005)]{FL05} Furlanetto, S.~R., 
7354: \& Loeb, A.\ 2005, \apj, 634, 1 
7355:   
7356: \bibitem[Sheth(1998)]{1998MNRAS.300.1057S} Sheth, R.~K.\ 1998, \mnras, 300, 
7357: 1057 
7358: 
7359: \bibitem[Bromm \& Larson(2004)]{2004ARA&A..42...79B} Bromm, V., \& Larson, 
7360: R.~B.\ 2004, \araa, 42, 79 
7361:  
7362: 
7363: 
7364: \bibitem[1]{WyL03a}Wyithe, J. S. B., \& Loeb, A. 2003a, ApJ, in press
7365: (astro-ph/0209056)
7366: \bibitem[2]{Cen03}Cen, R. 2003, ApJ, submitted
7367: (astro-ph/0210473)
7368: \bibitem[3]{FL03}Furlanetto, S. R., \& Loeb, A. 2003, ApJ, in press
7369: (astro-ph/0211496)
7370: \bibitem[4]{BL2001}Barkana, R., \& Loeb, A. 2001, Physics Reports, 349, 125
7371: \bibitem[5]{Kap02}Kaplinghat, M., Chu, M., Haiman, Z.,
7372: Holder, G., Knox, L., \& Skordis, C. 2002, ApJ, submitted (astro-ph/0207591)
7373: \bibitem[6]{Yos2003}Yoshida, N., Abel, T., Hernquist, L., \& Sugiyama, N.
7374: 2003, ApJ, submitted
7375: \bibitem[7]{HL2001}Haiman, Z., \& Loeb, A. 2001, ApJ, 552, 459
7376: \bibitem[8]{Fan2003}Fan, X., et al. 2003, AJ, in press
7377: (astro-ph/0301135)
7378: \bibitem[9]{BCL1999}Bromm, V., Coppi, P. S., \& Larson, R. B. 1999, ApJ, 527, L5
7379: \bibitem[10]{BCL2002}Bromm, V., Coppi, P. S., \& Larson, R. B. 2002, ApJ, 564, 23
7380: \bibitem[11]{NaU2001}Nakamura, F., \& Umemura, M. 2001, ApJ, 548, 19
7381: \bibitem[12]{ABN2002}Abel, T., Bryan, G. L., \& Norman, M. L. 2002, Science, 295, 93
7382: \bibitem[13]{Pud2002}Pudritz, R. E. 2002, Science, 295, 68
7383: \bibitem[14]{Lar2002}Larson, R. B. 2002, MNRAS, 332, 155
7384: \bibitem[15]{Kr02}Kroupa, P. 2002, Science, 295, 82
7385: \bibitem[16]{MJR1976}Rees, M. J. 1976, MNRAS, 176, 483
7386: \bibitem[17]{BBB2002}Bate, M. R., Bonnell, I. A., \& Bromm, V. 2002, MNARS, 332, L65
7387: \bibitem[18]{BBB2003}Bate, M. R., Bonnell, I. A., \& Bromm, V. 2003, MNRAS, 
7388: in press (astro-ph/0212380)
7389: 
7390: \bibitem[19]{BrL2003}
7391: Bromm, V., \& Loeb, A. 2003a, ApJ, 596, 34 
7392: 
7393: \bibitem{BrL03}
7394: Bromm, V., \& Loeb, A. 2003b, Nature, 425, 812
7395: 
7396: \bibitem{BrL06} Bromm, V., \& Loeb, A. 2006, to appear in Proc.  of ``Gamma
7397: Ray Bursts in the Swift Era''; preprint astro-ph/0601216
7398: 
7399: \bibitem[20]{BKL2001}Bromm, V., Kudritzki, R. P., \& Loeb, A. 2001, ApJ, 552, 464
7400: \bibitem[21]{Cr02}Christlieb, N., et al. 2002, Nature, 419, 904
7401: \bibitem[22]{MBH03}Mackey, J., Bromm, V., \& Hernquist, L. 2003, ApJ, in press
7402: (astro-ph/0208447)
7403: 
7404: 
7405: \bibitem[23]{ON1998}Omukai, K., \& Nishi, R. 1998, ApJ, 508, 141
7406: \bibitem[24]{OP2001}Omukai, K., \& Palla, F. 2001, ApJ, 561, L55
7407: \bibitem[25]{Rip2002}Ripamonti, E., Haardt, F., Ferrara, A., \& Colpi, M.
7408: 2002, MNRAS, 334, 401
7409: \bibitem[26]{Tan2003}Tan, J. C., \& McKee, C. F. 2003, these proceedings
7410: \bibitem[27]{PSS1983}Palla, F., Salpeter, E. E., \& Stahler, S. W. 1983, ApJ, 271, 632
7411: \bibitem[28]{OI2002}Omukai, K., \& Inutsuka, S. 2002, MNRAS, 332, 59
7412: \bibitem[29]{WC1987}Wolfire, M. G., \& Cassinelli, J. P. 1987, ApJ, 319, 850
7413: \bibitem[30]{MFR01}Madau, P., Ferrara, A., \& Rees, M. J. 2001, ApJ, 555, 92
7414: \bibitem[31]{MFM02}Mori, M., Ferrara, A., \& Madau, P. 2002, ApJ, 571, 40
7415: \bibitem[32]{TSD02}Thacker, R.J., Scannapieco, E., \& Davis, M. 2002, ApJ, 581, 836
7416: \bibitem[33]{LH97} Loeb, A.,\& Haiman, Z.\ 
7417: 1997, ApJ, 490, 571
7418: \bibitem[34]{TodF01} Todini, P., \& Ferrara, A. 2001, MNRAS, 325, 726
7419: \bibitem[35]{MBA01}Machacek, M. E., Bryan, G. L., \& Abel, T. 2001, ApJ, 548, 509
7420: \bibitem[36]{MBA03}Machacek, M. E., Bryan, G. L., \& Abel, T. 2003, MNRAS, 338, 27
7421: \bibitem[37]{Om00}Omukai, K. 2000, ApJ, 534, 809
7422: \bibitem[38]{BFCL01}Bromm, V., Ferrara, A., Coppi, P. S., \& Larson, R. B. 2001, MNRAS, 328, 969
7423: \bibitem[39]{Sch02} Schneider, R., Ferrara, A., Natarajan, 
7424: P., \& Omukai, K.\ 2002, ApJ, 571, 30 
7425: \bibitem[40]{LR00}Lamb, D. Q., \& Reichart, D. E. 2000, ApJ, 536, 1
7426: \bibitem[41]{Cia00} Ciardi, B., \& Loeb, A.\ 2000, ApJ, 540, 687 
7427: \bibitem[42]{Bloom01} Bloom, 
7428: J.~S., Kulkarni, S.~R., \& Djorgovski, S.~G.\ 2002, AJ, 123, 1111
7429: \bibitem[43]{kul} Kulkarni, S. R., et 
7430: al.\ 2000, Proc. SPIE, 4005, 9  
7431: \bibitem[44]{Tot97}Totani, T. 1997, ApJ, 486, L71
7432: \bibitem[45]{Wij98}Wijers, R. A. M. J., Bloom, J. S., Bagla, J. S., 
7433: \& Natarajan, P. 1998, MNRAS, 294, L13
7434: \bibitem[46]{BlNat00}Blain, A. W., \& Natarajan, P. 2000, MNRAS, 312, L35
7435: \bibitem[47]{BrL2002}Bromm, V., \& Loeb, A. 2002, ApJ, 575, 111
7436: 
7437: \bibitem[48]{Heg03} Heger, A., Fryer, C. L., Woosley, S. E.,
7438: Langer, N., \& Hartmann, D. H. 2003, ApJ, submitted
7439: (astro-ph/0212469)
7440: \bibitem[49]{BL03} Barkana, R., \& Loeb, A. 2003, Nature, in press
7441: (astro-ph/0209515)
7442: \bibitem[50]{Mi98} Miralda-Escud\'{e}, J.\ 
7443: 1998, ApJ, 501, 15 
7444: \bibitem[51]{Rees1984}Rees, M. J. 1984, ARA\&A, 22, 471
7445: \bibitem[52]{Beck2001}Becker, R. H., et al. 2001, AJ, 122, 2850
7446: \bibitem[53]{Djor2001}Djorgovski, S.G., Castro, S., Stern, D.,
7447: \& Mahabal, A. A. 2001, ApJ, 560, L5
7448: \bibitem[54]{Fan2002}Fan, X., Narayanan, V. K., Strauss, M. A., White, R. L.,
7449: Becker, R. H., Pentericci, L., \& Rix, H.-W.  2002, AJ, 123, 1247
7450: \bibitem[55]{LP02} Loeb, A., \& Peebles, P. J. E. 2002, ApJ, submitted
7451: (astro-ph/0211465)
7452: \bibitem[56]{LR1994}Loeb, A., \& Rasio, F. A. 1994, ApJ, 432, 52
7453: \bibitem[57]{OH02}Oh, S.P., \& Haiman, Z. 2002, ApJ, 569, 558
7454: \bibitem[58]{HRL1997}Haiman, Z., Rees, M. J., \& Loeb, A. 1997, ApJ, 476, 458;
7455: erratum 484, 985
7456: \bibitem[59]{WyL2003}Wyithe, J. S. B., \& Loeb, A. 2003b, ApJ, submitted 
7457: (astro-ph/0211556)
7458: \bibitem[60]{BS1999} Baumgarte, T. W., \& Shapiro, S. L. 1999, ApJ, 526, 941
7459: 
7460: 
7461: \bibitem[1]{CL00} 
7462: Ciardi, B., \& Loeb, A. 2000, ApJ, 540, 687
7463: 
7464: \bibitem[2]{LR00} 
7465: Lamb, D. Q., \& Reichart, D. E. 2000, ApJ, 536, 1
7466: 
7467: \bibitem[3]{Kog03} 
7468: Kogut, A., et al. 2003, ApJS, 148, 161
7469: 
7470: \bibitem[4]{Spe03} 
7471: Spergel, D. N., et al. 2003, ApJS, 148, 175
7472: 
7473: \bibitem[5]{C03} 
7474: Cen, R. 2003, ApJ, 591, L5
7475: 
7476: \bibitem[6]{CFW03} 
7477: Ciardi, B., Ferrara, A., \& White, S.D.M. 2003, MNRAS, 344, L7
7478: 
7479: \bibitem[7]{SL03} 
7480: Somerville, R. S., \& Livio, M. 2003, ApJ, 593, 611
7481: 
7482: \bibitem[8]{WL03} 
7483: Wyithe, J. S. B., \& Loeb, A. 2003, ApJ, 588, L69
7484: 
7485: \bibitem[9]{YBH04} 
7486: Yoshida, N., Bromm, V., \& Hernquist, L. 2004, ApJ, 605, 579
7487: 
7488: \bibitem[10]{HH03} 
7489: Haiman, Z., \& Holder, G. P. 2003, ApJ, 595, 1
7490: 
7491: \bibitem[11]{BarL04} 
7492: Barkana, R., \& Loeb, A. 2004, ApJ, 601, 64
7493: 
7494: \bibitem[12]{FL03} 
7495: Furlanetto, S. R., \& Loeb, A. 2003, ApJ, 588, 18
7496: 
7497: \bibitem[13]{Geh04} 
7498: Gehrels, N., et al. 2004, ApJ, 611, 1005
7499: 
7500: \bibitem[14]{Hai05} 
7501: Haislip, J., et al. 2005, Nature, submitted (astro-ph/0509660)
7502: 
7503: \bibitem[15]{WL02} 
7504: Wyithe, J. S. B., \& Loeb, A. 2002, ApJ, 581, 886
7505: 
7506: \bibitem[16]{T97} 
7507: Totani, T. 1997, ApJ, 486, L71
7508: 
7509: \bibitem[17]{Wij98} 
7510: Wijers, R.A.M.J., Bloom, J.S., Bagla, J.S., \& Totani, T. 1997, ApJ, 486, L71
7511: 1998, MNRAS, 294, L13
7512: 
7513: \bibitem[18]{BN00} 
7514: Blain, A. W., \& Natarajan, P. 2000, MNRAS, 312, L35
7515: 
7516: \bibitem[19]{Kul00} 
7517: Kulkarni, S. R., et al. 2000, Proc. SPIE, 4005, 9
7518: 
7519: \bibitem[20]{BKD02} 
7520: Bloom, J. S., Kulkarni, S. R., \& Djorgovski, S. G. 2002, AJ, 123, 1111
7521: 
7522: \bibitem[21]{Nat05} 
7523: Natarajan, P., Albanna, B., Hjorth, J., Ramirez-Ruiz, E.,
7524: Tanvir, N., \& Wijers, R.A.M.J. 
7525: 2005, MNRAS, 364, L8
7526: 
7527: \bibitem[22]{Sta03} 
7528: Stanek, K. Z., et al. 2003, ApJ, 591, L17
7529: 
7530: \bibitem[23]{ABN02} 
7531: Abel, T., Bryan, G., \& Norman, M. L. 2002, Science, 295, 93
7532: 
7533: \bibitem[24]{BCL02} 
7534: Bromm, V., Coppi, P. S., \& Larson, R. B. 2002, ApJ, 564, 23
7535: 
7536: \bibitem[25]{BLar04} 
7537: Bromm, V., \& Larson, R. B. 2004, \araa, 42, 79
7538: 
7539: \bibitem{BarL01} 
7540: Barkana, R., \& Loeb, A. 2001, Phys. Rep., 349, 125
7541: 
7542: \bibitem[27]{FL02} 
7543: Furlanetto, S. R., \& Loeb, A. 2002, ApJ, 579, 1
7544: 
7545: \bibitem[28]{BL02} 
7546: Bromm, V., \& Loeb, A. 2002, ApJ, 575, 111
7547: 
7548: \bibitem[29]{BL06} 
7549: Bromm, V., \& Loeb, A. 2006, ApJ, in press (astro-ph/0509303)
7550: 
7551: \bibitem[30]{Sok04} 
7552: Sokasian, A., Yoshida, N., Abel, T., Hernquist, L., 
7553: \& Springel, V. 2004, MNRAS, 350, 47
7554: 
7555: \bibitem[31]{ABS06} 
7556: Alvarez, M. A., Bromm, V., \& Shapiro, P. R. 2006, ApJ, in press
7557: (astro-ph/0507684)
7558: 
7559: \bibitem[32]{MBH03} 
7560: Mackey, J., Bromm, V., \& Hernquist, L. 2003, ApJ, 586, 1
7561: 
7562: \bibitem[33]{FL05} 
7563: Furlanetto, S. R., \& Loeb, A. 2005, ApJ, 634, 1
7564: 
7565: \bibitem[34]{Sch03} 
7566: Schaye, J., Aguirre, A., Kim, T.-S., Theuns, T., Rauch, M., 
7567: \& Sargent, W. L. W. 2003, ApJ, 596, 768
7568: 
7569: \bibitem[35]{SSR04} 
7570: Simcoe, R. A., Sargent, W. L. W., \& Rauch, M. 2004, ApJ, 606, 92
7571: 
7572: \bibitem[36]{Sca05} 
7573: Scannapieco, E., Madau, P., Woosley, S., Heger, A., 
7574: \& Ferrara, A. 2005, ApJ, 633, 1031
7575: 
7576: \bibitem[37]{SBK02} 
7577: Santos, M. R., Bromm, V., \& Kamionkowski, M. 2002, MNRAS, 336, 1082
7578: 
7579: \bibitem[38]{SF03} 
7580: Salvaterra, R., \& Ferrara, A. 2003, MNRAS, 339, 973
7581: 
7582: \bibitem[39]{Kas05} 
7583: Kashlinsky, A., Arendt, R. G., Mather, J., \& Moseley, S. H.
7584: 2005, Nature, 438, 45
7585: 
7586: \bibitem[40]{MS05} 
7587: Madau, P., \& Silk, J. 2005, MNRAS, 359, L37
7588: 
7589: \bibitem[41]{DAK05} 
7590: Dwek, E., Arendt, R. G., \& Krennrich, F. 2005, ApJ, 635, 784
7591: 
7592: \bibitem[42]{Teg97} 
7593: Tegmark, M., Silk, J., Rees, M. J., Blanchard, A., Abel, T., \& Palla, F. 
7594: 1997, ApJ, 474, 1
7595: 
7596: \bibitem[43]{Shu02} 
7597: Shu, F. H., Lizano, S., Galli, D., Cant\'{o}, J., 
7598: \& Laughlin, G. 2002, \apj, 580, 969
7599: 
7600: \bibitem[44]{BHW01} 
7601: Baraffe, I., Heger, A., \& Woosley, S. E. 2001, ApJ, 550, 890
7602: 
7603: \bibitem[45]{K02} 
7604: Kudritzki, R. P. 2002, ApJ, 577, 389
7605: 
7606: \bibitem[46]{BYH03} 
7607: Bromm, V., Yoshida, N., \& Hernquist, L. 2003, ApJ, 596, L135
7608: 
7609: \bibitem[47]{BL04} 
7610: Bromm, V., \& Loeb, A. 2004, NewA, 9, 353
7611: 
7612: \bibitem[48]{BCL99} 
7613: Bromm, V., Coppi, P. S., \& Larson, R. B. 1999, ApJ, 527, L5
7614: 
7615: \bibitem[49]{KW02} 
7616: Kitsionas, S., \& Whitworth, A. P. 2002, MNRAS, 330, 129
7617: 
7618: \bibitem[50]{BL03} 
7619: Bromm, V., \& Loeb, A. 2003, ApJ, 596, 34
7620: 
7621: \bibitem[51]{OP01} 
7622: Omukai, K., \& Palla, F. 2001, ApJ, 561, L55
7623: 
7624: \bibitem[52]{OP03} 
7625: Omukai, K., \& Palla, F. 2003, ApJ, 589, 677
7626: 
7627: \bibitem[53]{Rip02} 
7628: Ripamonti, E., Haardt, F., Ferrara, A., \& Colpi, M. 2002, MNRAS, 334, 401
7629: 
7630: \bibitem[54]{TM04} 
7631: Tan, J. C., \& McKee, C. F. 2004, ApJ, 603, 383 
7632: 
7633: \bibitem[55]{PSS83} 
7634: Palla, F., Salpeter, E. E.., \& Stahler, S. W. 1983, ApJ, 271, 632
7635: 
7636: \bibitem[56]{L03} 
7637: Larson, R. B. 2003, Rep. Prog. Phys., 66, 1651
7638: 
7639: \bibitem[57]{ON98} 
7640: Omukai, K., \& Nishi, R. 1998, ApJ, 508, 141
7641: 
7642: \bibitem[58]{OI02} 
7643: Omukai, K., \& Inutsuka, S. 2002, MNRAS, 332, 59
7644: 
7645: \bibitem[59]{WC87} 
7646: Wolfire, M. G., \& Cassinelli, J. P. 1987, ApJ, 319, 850
7647: 
7648: \bibitem[60]{PL98} 
7649: Perna, R., \& Loeb, A. 1998, \apj, 501, 467 
7650: 
7651: \bibitem[61]{WD00} 
7652: Waxman, E., \& Draine, B.~T.\ 2000, \apj, 537, 796 
7653: 
7654: \bibitem[62]{FKR01} 
7655: Fruchter, A., Krolik, J.~H., \& Rhoads, J.~E.\ 2001, \apj, 563, 597 
7656: 
7657: \bibitem[63]{DH02} 
7658: Draine, B.~T., \& Hao, L.\ 2002, \apj, 569, 780 
7659: 
7660: \bibitem[64]{MWH01} 
7661: MacFadyen, A. I., Woosley, S. E., \& Heger, A. 2001, ApJ, 550, 410
7662: 
7663: \bibitem[Carilli et al.(2004)]{Carilli04} Carilli, C.~L., Gnedin, 
7664: N., Furlanetto, S., \& Owen, F.\ 2004, New Astronomy Review, 48, 1053 
7665:  
7666: 
7667: 
7668: \bibitem[Dijkstra et al.(2004)]{Dijkstra04} Dijkstra, M., Haiman, 
7669: Z., \& Loeb, A.\ 2004, \apj, 613, 646 
7670: 
7671: \bibitem[Kohler et al.(2005)]{Kohler05} Kohler, K., Gnedin, 
7672: N.~Y., \& Hamilton, A.~J.~S.\ 2005, ArXiv Astrophysics e-prints, 
7673: arXiv:astro-ph/0511627 
7674:  
7675: \bibitem[Iliev et al.(2005)]{Iliev05} Iliev, I.~T., Mellema, G., Pen,
7676: U.-L., Merz, H., Shapiro, P.~R., \& Alvarez, M.~A.\ 2005, preprint
7677: astro-ph/0512187
7678:  
7679: \bibitem[Sokasian et al.(2003)]{Sokasian03} Sokasian, A., Abel, 
7680: T., Hernquist, L., \& Springel, V.\ 2003, \mnras, 344, 607 
7681:  
7682: \bibitem[Ciardi et al.(2003)]{Ciardi03} Ciardi, B., Stoehr, F., 
7683: \& White, S.~D.~M.\ 2003, \mnras, 343, 1101 
7684: 
7685: 
7686: \bibitem[Wyithe \& Loeb(2003)]{Wyithe03} Wyithe, J.~S.~B., \& 
7687: Loeb, A.\ 2003, \apj, 586, 693 
7688: 
7689: \bibitem[Cen(2003)]{Cen03} Cen, R.\ 2003, \apj, 591, 12  
7690: 
7691: \bibitem[Tumlinson et al.(2004)]{Tumlinson04} Tumlinson, J., 
7692: Venkatesan, A., \& Shull, J.~M.\ 2004, \apj, 612, 602 
7693: 
7694: \bibitem[White et al.(2003)]{White03} White, R.L., Becker, R.H., Fan, X.,
7695: Strauss, M.A. 2003, AJ, 126, 1 
7696:   
7697: \bibitem[Wyithe et al.(2005)]{Wy05} Wyithe, J.~S.~B., Loeb, 
7698: A., \& Carilli, C.\ 2005, \apj, 628, 575 
7699:  
7700: \bibitem[Wyithe \& Loeb(2004)]{WyLo} Wyithe, J.~S.~B., \& 
7701: Loeb, A.\ 2004, \apj, 610, 117 
7702:  
7703: \bibitem[Wyithe et al.(2005)]{WyBar} Wyithe, J.~S.~B., Loeb, 
7704: A., \& Barnes, D.~G.\ 2005, \apj, 634, 715 
7705:  
7706: \bibitem[Mesinger \& Haiman(2004)]{Mesing} Mesinger, A., \& 
7707: Haiman, Z.\ 2004, \apjl, 611, L69 
7708:  
7709: 
7710: \bibitem[Loeb \& Rybicki(1999)]{Loeb_Rybicki} Loeb, A., \& Rybicki, 
7711: G.~B.\ 1999, \apj, 524, 527 
7712: 
7713: \bibitem[Rybicki \& Loeb(1999)]{Rybicki_Loeb} Rybicki, G.~B., \& 
7714: Loeb, A.\ 1999, \apjl, 520, L79 
7715:  
7716: \bibitem[Santos(2004)]{Santos} Santos, M.~R.\ 2004, \mnras,
7717: 349, 1137 
7718: 
7719: \bibitem[Haiman \& Cen(2005)]{Haiman05} Haiman, Z., \& Cen, R.\ 
7720: 2005, \apj, 623, 627 
7721: 
7722: \bibitem[Cen \& Haiman(2000)]{Haiman_Cen} Cen, R., \& Haiman, Z.\ 
7723: 2000, \apjl, 542, L75 
7724:  
7725: \bibitem[Madau \& Rees(2000)]{Madau_Rees} Madau, P., \& Rees, 
7726: M.~J.\ 2000, \apjl, 542, L69 
7727:  
7728: \bibitem[Loeb et al.(2005)]{Loeb_Hernquist} Loeb, A., Barkana, R., \& 
7729: Hernquist, L.\ 2005, \apj, 620, 553 
7730:  
7731: \bibitem[Malhotra \& Rhoads(2005)]{Rhoads} Malhotra, S., \& 
7732: Rhoads, J.\ 2005, ArXiv Astrophysics e-prints, arXiv:astro-ph/0511196 
7733:  
7734: \bibitem[Wyithe \& Loeb(2005)]{Wyithe_Loeb} Wyithe, J.~S.~B., \& 
7735: Loeb, A.\ 2005, \apj, 625, 1 
7736: 
7737: \bibitem[Furlanetto et al.(2006)]{Furlanetto_Hernquist} Furlanetto, S.~R.,
7738: Zaldarriaga, M., \& Hernquist, L.\ 2006, \mnras, 365, 1012
7739: 
7740: \bibitem[Cen et al.(2005)]{Haiman_Mesinger} Cen, R., Haiman, Z., \& 
7741: Mesinger, A.\ 2005, \apj, 621, 89 
7742: 
7743: %new
7744: 
7745: \bibitem[Chen \& Miralda-Escud{\'e}(2004)]{ChenMi} Chen, X., 
7746: \& Miralda-Escud{\'e}, J.\ 2004, \apj, 602, 1 
7747:  
7748: \bibitem[Madau et al.(1997)]{Meiksin} Madau, P., Meiksin, A., 
7749: \& Rees, M.~J.\ 1997, \apj, 475, 429 
7750:  
7751: \bibitem[Wyithe et al.(2005)]{LoWy05} Wyithe, J.~S.~B., Loeb, 
7752: A., \& Barnes, D.~G.\ 2005, \apj, 634, 715 
7753:  
7754:  
7755: 
7756: