1: \documentclass{article}
2: \setlength\textheight{10.0in}
3: \setlength\textwidth{6.5in}
4: \oddsidemargin -0.2in
5: \topmargin -1in
6: \input epsf.sty
7:
8: \def\lya{Ly$\alpha$~}
9: \def\Lya{Ly$\alpha$~}
10: \def\HI{\rm H\,I~}
11: \def\zr{z_{\rm reion}}
12: \def\beq{\begin{equation}}
13: \def\beqa{\begin{eqnarray}}
14: \def\eeq{\end{equation}}
15: \def\eeqa{\end{eqnarray}}
16: \def\Omm{{\Omega_m}}
17: \def\gtrsim{>}
18: \def\ga{>}
19: \def\lesssim{<}
20: \def\la{{\lesssim}}
21: \def\Ommz{{\Omega_m^{\,z}}}
22: \def\Omr{{\Omega_r}}
23: \def\Omk{{\Omega_k}}
24: \def\Oml{{\Omega_{\Lambda}}}
25: \def\sun{\odot}
26: \def\xb{{\bf x}}
27: \def\rb{{\bf r}}
28: \def\vb{{\bf v}}
29: \def\ub{{\bf u}}
30: \def\kb{{\bf k}}
31: \def\kB{k}
32: \def\cN{c_{\rm N}}
33:
34: \begin{document}
35: \hfill {\it UNESCO EOLSS ENCYCLOPEDIA}\bigskip\bigskip
36:
37: \centerline
38: {\it{\Large ``Let
39: there be Light'': the Emergence of Structure}}
40: \centerline{\it{\Large out of the Dark Ages in the
41: Early Universe}}\bigskip\bigskip
42:
43:
44: \def\Ref#1{\noindent\hangindent=30pt\hangafter=1\nobreak
45: \frenchspacing #1 \par}
46: \def\sm{\smallskip}
47:
48: {\large
49: % author 1
50: {\bf Abraham Loeb}
51:
52: %affiliation 1
53: Department of Astronomy,
54: Harvard University, 60 Garden St., Cambridge MA, 02138
55: }
56:
57: {\bf Key Words: dark ages, first stars, high-redshift galaxies,
58: reionization, intergalactic medium}
59:
60: \section*{Contents}
61:
62: \noindent
63: 1. Introduction
64:
65: 1.1 Observing our past
66:
67: 1.2 The expanding universe
68:
69: \noindent
70: 2. Galaxy Formation
71:
72: 2.1. Growth of linear perturbations
73:
74: 2.2. Halo properties
75:
76: 2.3. Formation of the first stars
77:
78: 2.4. Gamma-ray Bursts: probing the first stars one star at a time
79:
80: 2.5. Supermassive black holes
81:
82: 2.6. The epoch of reionization
83:
84: 2.7. Post-reionization suppression of low-mass galaxies
85:
86: \noindent
87: 3. Probing the Diffuse Intergalactic Hydrogen
88:
89: 3.1. Lyman-alpha absorption
90:
91: 3.2. 21-cm absorption or emission
92:
93: \noindent
94: 4. Conclusions
95:
96:
97: \section*{Summary}
98: Cosmology is by now a mature experimental science. We are privileged
99: to live at a time when the story of genesis (how the Universe started and
100: developed) can be critically explored by direct observations. Looking deep
101: into the Universe through powerful telescopes we can see images of the
102: Universe when it was younger, because of the finite time it takes light to
103: travel to us from distant sources.
104:
105: Existing data sets include an image of the Universe when it was 0.4 million
106: years old (in the form of the cosmic microwave background), as well as
107: images of individual galaxies when the Universe was older than a billion
108: years. But there is a serious challenge: in between these two epochs was a
109: period when the Universe was dark, stars had not yet formed, and the cosmic
110: microwave background no longer traced the distribution of matter. And this
111: is precisely the most interesting period, when the primordial soup evolved
112: into the rich zoo of objects we now see.
113: %How can astronomers see this dark yet crucial time?
114:
115: %The situation is similar to having a photo album of a person that contains
116: %the first ultra-sound image of him/her as an unborn baby and some
117: %additional photos as a teenager and an adult. We are currently searching
118: %for the missing pages in between that will tell us how the Universe evolved
119: %during its infancy to eventually make galaxies like our own Milky-Way.
120:
121: The observers are moving ahead along several fronts. The first involves
122: the construction of large infrared telescopes on the ground and in space,
123: that will provide us with new photos of the first galaxies. Current plans
124: include ground-based telescopes which are 24-42 meter in diameter, and
125: NASA's successor to the Hubble Space Telescope, called the James Webb Space
126: Telescope. In addition, several observational groups around the globe are
127: constructing radio arrays that will be capable of mapping the
128: three-dimensional distribution of cosmic hydrogen in the infant
129: Universe. These arrays are aiming to detect the long-wavelength (redshifted
130: 21-cm) radio emission from hydrogen atoms.
131: %Coincidentally, this long wavelength (or low frequency) overlaps
132: %with the band used for radio and television broadcasting, and so these
133: %telescopes include arrays of regular radio antennas that one can find in
134: %electronics stores.
135: The images from these antenna arrays will reveal how the non-uniform
136: distribution of neutral hydrogen evolved with cosmic time and eventually
137: was extinguished by the ultra-violet radiation from the first galaxies.
138: Theoretical research has focused in recent years on predicting the expected
139: signals for the above instruments and motivating these ambitious
140: observational projects.
141:
142:
143:
144: \section{Introduction}
145:
146: \subsection{Observing our past}
147:
148: When we look at our image reflected off a mirror at a distance of 1
149: meter, we see the way we looked 6.7 nanoseconds ago, the light travel
150: time to the mirror and back. If the mirror is spaced $10^{19}~{\rm cm}
151: \simeq 3~$pc away, we will see the way we looked twenty one years
152: ago. Light propagates at a finite speed, and so by observing distant
153: regions, we are able to see what the Universe looked like in the past,
154: a light travel time ago (Figure~\ref{fig:z}). The statistical
155: homogeneity of the Universe on large scales guarantees that what we
156: see far away is a fair statistical representation of the conditions
157: that were present in in our region of the Universe a long time ago.
158:
159: \begin{figure}
160: \epsfxsize=15cm \epsfbox{history.eps}
161: \caption{Cosmology is like archeology. The deeper one looks, the older is
162: the layer that is revealed, owing to the finite propagation speed of light.
163: %(illustration from Loeb 2008).
164: }
165: \label{fig:z}
166: \end{figure}
167:
168: This fortunate situation makes cosmology an empirical science. We do
169: not need to guess how the Universe evolved. Using telescopes we can
170: simply see how it appeared at earlier cosmic times. In principle, this
171: allows the entire 13.7 billion year cosmic history of our universe to
172: be reconstructed by surveying the galaxies and other sources of light
173: to large distances (Figure~\ref{fig:history}). Since a greater
174: distance means a fainter flux from a source of a fixed luminosity, the
175: observation of the earliest sources of light requires the development
176: of sensitive instruments and poses challenges to observers.
177:
178: \begin{figure}
179: \epsfxsize=10cm \epsfbox{SciRev2bnd.eps}
180: \caption{Overview of cosmic history, with the age of the universe shown on
181: the top axis and the corresponding redshift on the bottom axis. Yellow
182: represents regions where the hydrogen is ionized, and gray, neutral
183: regions. Stars form in galaxies located within dark matter concentrations
184: whose typical mass grows with time, starting with $\sim 10^5 M_{\odot}$
185: (red circles) for the host of the first star, rising to $10^7$--$10^9
186: M_{\odot}$ (blue circles) for the sources of reionization, and reaching
187: $\sim 10^{12} M_{\odot}$ (green circles) for present-day galaxies like our
188: own Milky Way. Astronomers probe the evolution of the cosmic gas using the
189: absorption of background light (dotted lines) by atomic hydrogen along the
190: line of sight. The classical technique uses absorption by the
191: Lyman-$\alpha$ resonance of hydrogen of the light from bright quasars
192: located within massive galaxies, while a new type of astronomical
193: observation will use the 21-cm line of hydrogen with the cosmic microwave
194: background as the background source.}
195: \label{fig:history}
196: \end{figure}
197:
198: As the universe expands, photon wavelengths get stretched as well.
199: %, so that the spectrum of light we observe today is shifted from the
200: The factor by which the observed wavelength is increased (i.e. shifted
201: towards the red) relative to the emitted one is denoted by $(1+z)$, where
202: $z$ is the cosmological redshift. Astronomers use the known emission
203: patterns of hydrogen and other chemical elements in the spectrum of each
204: galaxy to measure $z$. This then implies that the universe has expanded by
205: a factor of $(1+z)$ in linear dimension since the galaxy emitted the
206: observed light, and cosmologists can calculate the corresponding distance
207: and cosmic age for the source galaxy. Large telescopes have allowed
208: astronomers to observe faint galaxies that are so far away that we see them
209: more than twelve billion years back in time. Thus, we know directly that
210: galaxies were in existence as early as 850 million years after the Big
211: Bang, at a redshift of $z \sim 6.5$ or higher.
212:
213: We can in principle image the Universe only if it is transparent. Earlier
214: than $400\,000$ years after the big bang, the cosmic hydrogen was broken
215: into its constituent electrons and protons (i.e. ``ionized'') and the
216: Universe was opaque to scattering by the free electrons in the dense
217: plasma. Thus, telescopes cannot be used to electromagnetically image the
218: infant Universe at earlier times (or redshifts $\ga 10^3$). The earliest
219: possible image of the Universe was recorded by the COBE and WMAP
220: satellites, which measured the temperature distribution of the cosmic
221: microwave background (CMB) on the sky (Figure~\ref{fig:CMB}).
222:
223: \begin{figure}
224: \epsfxsize=10cm \epsfbox{WMAP.eps}
225: \caption{Images of the Universe shortly after it became transparent, taken
226: by the {\it COBE}\/ and {\it WMAP}\/ satellites (see
227: http://map.gsfc.nasa.gov/ for details). The slight density inhomogeneities
228: in the otherwise uniform Universe imprinted a map of hot and cold spots
229: (shown here as different colors) in the CMB that is observed today. The
230: existence of these anisotropies was predicted three decades before the
231: technology for taking these images became available, in a number of
232: theoretical papers including Sachs \& Wolfe (1967), Rees \& Sciama (1968),
233: Silk (1968), Sunyaev \& Zeldovich (1970), and Peebles \& Yu (1970).}
234: \label{fig:CMB}
235: \end{figure}
236:
237: The CMB, the relic radiation from the hot, dense beginning of the universe,
238: is indeed another major probe of observational cosmology. The universe
239: cools as it expands, so it was initially far denser and hotter than it is
240: today. For hundreds of thousands of years the cosmic gas consisted of a
241: plasma of free protons and electrons, and a slight mix of light nuclei,
242: sustained by the intense thermal motion of these particles. Just like the
243: plasma in our own Sun, the ancient cosmic plasma emitted and scattered a
244: strong field of visible and ultraviolet photons. As mentioned above, about
245: $400\,000$ years after the Big Bang the temperature of the universe dipped
246: for the first time below a few thousand degrees Kelvin. The protons and
247: electrons were now moving slowly enough that they could attract each other
248: and form hydrogen atoms, in a process known as cosmic recombination. With
249: the scattering of the energetic photons now much reduced, the photons
250: continued traveling in straight lines, mostly undisturbed except that
251: cosmic expansion has redshifted their wavelength into the microwave regime
252: today. The emission temperature of the observed spectrum of these CMB
253: photons is the same in all directions to one part in $100\,000$
254: (Figure~\ref{fig:CMB}), which reveals that conditions were nearly uniform
255: in the early universe.
256:
257: It was just before the moment of cosmic recombination (when matter started
258: to dominate in energy density over radiation) that gravity started to
259: amplify the tiny fluctuations in temperature and density observed in the
260: CMB data. Regions that started out slightly denser than average began to
261: contract because the gravitational forces were also slightly stronger than
262: average in these regions. Eventually, after hundreds of millions of years
263: of contraction, the overdense regions stopped expanding, turned around, and
264: eventually collapsed to make bound objects such as galaxies. The gas within
265: these collapsed objects cooled and fragmented into stars. This process,
266: however, would have taken too long to explain the abundance of galaxies
267: today, if it involved only the observed cosmic gas. Instead, gravity is
268: strongly enhanced by the presence of dark matter -- an unknown substance
269: that makes up the vast majority (83\%) of the cosmic density of matter. The
270: motion of stars and gas around the centers of nearby galaxies indicates
271: that each is surrounded by an extended mass of dark matter, and so
272: dynamically-relaxed dark matter concentrations are generally referred to as
273: ``halos''.
274:
275: According to the standard cosmological model, the dark matter is cold
276: (abbreviated as CDM), i.e., it behaves as a collection of collisionless
277: particles that started out at matter domination with negligible thermal
278: velocities and have evolved exclusively under gravitational forces. The
279: model explains how both individual galaxies and the large-scale patterns in
280: their distribution originated from the small initial density
281: fluctuations. On the largest scales, observations of the present galaxy
282: distribution have indeed found the same statistical patterns as seen in the
283: CMB, enhanced as expected by billions of years of gravitational
284: evolution. On smaller scales, the model describes how regions that were
285: denser than average collapsed due to their enhanced gravity and eventually
286: formed gravitationally-bound halos, first on small spatial scales and later
287: on larger ones. In this hierarchical model of galaxy formation, the small
288: galaxies formed first and then merged or accreted gas to form larger
289: galaxies. At each snapshot of this cosmic evolution, the abundance of
290: collapsed halos, whose masses are dominated by dark matter, can be computed
291: from the initial conditions using numerical simulations. The common
292: understanding of galaxy formation is based on the notion that stars formed
293: out of the gas that cooled and subsequently condensed to high densities in
294: the cores of some of these halos.
295:
296: Gravity thus explains how some gas is pulled into the deep potential wells
297: within dark matter halos and forms the galaxies. One might naively expect
298: that the gas outside halos would remain mostly undisturbed. However,
299: observations show that it has not remained neutral (i.e., in atomic form)
300: but was largely ionized by the UV radiation emitted by the galaxies. The
301: diffuse gas pervading the space outside and between galaxies is referred to
302: as the intergalactic medium (IGM). For the first hundreds of millions of
303: years after cosmological recombination, the so-called cosmic ``dark ages'',
304: the universe was filled with diffuse atomic hydrogen. As soon as galaxies
305: formed, they started to ionize diffuse hydrogen in their vicinity. Within
306: less than a billion years, most of the IGM was re-ionized. We have not yet
307: imaged the cosmic dark ages before the first galaxies had formed. One of
308: the frontiers in current cosmological studies aims to study the cosmic
309: epoch of reionization and the first generation of galaxies that triggered
310: it.
311:
312: \subsection{The expanding universe}
313:
314: The modern physical description of the Universe as a whole can be traced
315: back to Einstein, who assumed for simplicity the so-called ``cosmological
316: principle'': that the distribution of matter and energy is homogeneous and
317: isotropic on the largest scales. Today isotropy is well established for the
318: distribution of faint radio sources, optically-selected galaxies, the X-ray
319: background, and most importantly the cosmic microwave background
320: (hereafter, CMB). The constraints on homogeneity are less strict, but a
321: cosmological model in which the Universe is isotropic but significantly
322: inhomogeneous in spherical shells around our special location, is also
323: excluded.
324:
325: In General Relativity, the metric for a space which is spatially
326: homogeneous and isotropic is the Friedman-Robertson-Walker metric, which
327: can be written in the form \beq \label{RW}
328: ds^2=c^2dt^2-a^2(t)\left[\frac{dR^2}{1-k\,R^2}+R^2
329: \left(d\theta^2+\sin^2\theta\,d\phi^2\right)\right]\ , \eeq where $c$ is
330: the speed of light, $a(t)$ is the cosmic scale factor which describes
331: expansion in time $t$, and $(R,\theta,\phi)$ are spherical comoving
332: coordinates. The constant $k$ determines the geometry of the metric; it is
333: positive in a closed Universe, zero in a flat Universe, and negative in an
334: open Universe. Observers at rest remain at rest, at fixed
335: $(R,\theta,\phi)$, with their physical separation increasing with time in
336: proportion to $a(t)$. A given observer sees a nearby observer at physical
337: distance $D$ receding at the Hubble velocity $H(t)D$, where the Hubble
338: constant at time $t$ is $H(t)=d\,a(t)/dt$. Light emitted by a source at
339: time $t$ is observed at $t=0$ with a redshift $z=1/a(t)-1$, where we set
340: $a(t=0) \equiv 1$ for convenience.
341:
342: The Einstein field equations of General Relativity yield the Friedmann
343: equation \beq H^2(t)=\frac{8 \pi G}{3}\rho-\frac{k}{a^2}\ , \eeq which
344: relates the expansion of the Universe to its matter-energy content. The
345: constant $k$ determines the geometry of the universe; it is positive in a
346: closed universe, zero in a flat universe, and negative in an open
347: universe. For each component of the energy density $\rho$, with an equation
348: of state $p=p(\rho)$, the density $\rho$ varies with $a(t)$ according to
349: the thermodynamic relation \beq d (\rho c^2 R^3)=-p d(R^3)\ . \eeq With the
350: critical density \beq \rho_C(t) \equiv \frac{3 H^2(t)}{8 \pi G} \eeq
351: defined as the density needed for $k=0$, we define the ratio of the total
352: density to the critical density as \beq \Omega \equiv \frac{\rho}{\rho_C}\
353: . \eeq With $\Omm$, $\Oml$, and $\Omr$ denoting the present contributions
354: to $\Omega$ from matter (including cold dark matter as well as a
355: contribution $\Omega_b$ from ordinary matter [``baryons''] made of protons
356: and neutrons), vacuum density (cosmological constant), and radiation,
357: respectively, the Friedmann equation becomes \beq \frac{H(t)}{H_0}= \left[
358: \frac{\Omm} {a^3}+ \Oml+ \frac{\Omr}{a^4}+ \frac{\Omk}{a^2}\right]\ , \eeq
359: where we define $H_0$ and $\Omega_0=\Omm+\Oml+\Omr$ to be the present
360: values of $H$ and $\Omega$, respectively, and we let \beq \Omk \equiv
361: -\frac{k}{H_0^2}=1-\Omega_m. \eeq In the particularly simple Einstein-de
362: Sitter model ($\Omm=1$, $\Oml=\Omr=\Omk=0$), the scale factor varies as
363: $a(t) \propto t^{2/3}$. Even models with non-zero $\Oml$ or $\Omk$ approach
364: the Einstein-de Sitter scaling-law at high redshift, i.e.\ when $(1+z) \gg
365: |\Omm^{-1}-1|$ (as long as $\Omr$ can be neglected). In this high-$z$
366: regime the age of the Universe is
367: \begin{equation}
368: t\approx {2\over 3 H_0 \sqrt{\Omega_m}} \left(1+z\right)^{-3/2}\
369: \approx 10^9 {\rm yr} \left({1+z\over 7}\right)^{-3/2} .
370: \end{equation}
371:
372: Recent observations confine the standard set of cosmological parameters to
373: a relatively narrow range. In particular, we seem to live in a universe
374: dominated by a cosmological constant ($\Lambda$) and cold dark matter, or
375: in short a $\Lambda$CDM cosmology (with $\Omk$ so small that it is usually
376: assumed to equal zero) with an approximately scale-invariant primordial
377: power spectrum of density fluctuations, i.e., $n \approx 1$ where the
378: initial power spectrum is $P(k)=\vert \delta_{\bf k} \vert^2\propto k^n$ in
379: terms of the wavenumber $k$ of the Fourier modes $\delta_{\bf k}$ (see \S
380: \ref{sec:lin} below). Also, the Hubble constant today is written as
381: $H_0=100h \mbox{ km s}^{-1}\mbox{Mpc}^{-1}$ in terms of $h$, and the
382: overall normalization of the power spectrum is specified in terms of
383: $\sigma_8$, the root-mean-square amplitude of mass fluctuations in spheres
384: of radius $8\ h^{-1}$ Mpc. For example, the best-fit cosmological
385: parameters matching the WMAP data together with large-scale gravitational
386: lensing observations are $\sigma_8=0.826$, $n=0.953$, $h=0.687$,
387: $\Omega_m=0.299$, $\Omega_\Lambda=0.701$ and $\Omega_b=0.0478$.
388: %CHECK AND UPDATE: WMAP5
389: %
390: %A different cosmological parameter set, also based on the CMB data together
391: %with other large-scale structure measurements is: $\sigma_8=0.785$,
392: %$n=0.957$, $h=0.723$ $\Omega_m=0.253$, $\Omega_\Lambda=0.747$, and
393: %$\Omega_b=0.0425$. The difference between these two parameter sets roughly
394: %represents current $1-\sigma$ parameter uncertainties.
395:
396:
397: \section{Galaxy Formation}
398:
399: \subsection{Growth of linear perturbations}
400:
401: \label{sec:lin}
402:
403: As noted in the Introduction, observations of the CMB show that the
404: universe at cosmic recombination (redshift $z\sim 10^3$) was remarkably
405: uniform apart from spatial fluctuations in the energy density and in the
406: gravitational potential of roughly one part in $\sim 10^5$. The primordial
407: inhomogeneities in the density distribution grew over time and eventually
408: led to the formation of galaxies as well as galaxy clusters and large-scale
409: structure. In the early stages of this growth, as long as the density
410: fluctuations on the relevant scales were much smaller than unity, their
411: evolution can be understood with a linear perturbation analysis.
412:
413: As before, we distinguish between fixed and comoving coordinates. Using
414: vector notation, the fixed coordinate ${\bf r}$ corresponds to a comoving
415: position $\xb=\rb/a$. In a homogeneous Universe with density $\rho$, we
416: describe the cosmological expansion in terms of an ideal pressureless fluid
417: of particles each of which is at fixed $\xb$, expanding with the Hubble
418: flow $\vb=H(t) \rb$ where $\vb=d\rb/dt$. Onto this uniform expansion we
419: impose small perturbations, given by a relative density perturbation \beq
420: \delta(\xb)=\frac{\rho(\rb)}{\bar{\rho}}-1\ , \eeq where the mean fluid
421: density is $\bar{\rho}$, with a corresponding peculiar velocity $\ub \equiv
422: \vb - H \rb$. Then the fluid is described by the continuity and Euler
423: equations in comoving coordinates: \beqa \frac{\partial
424: \delta}{\partial t}+\frac{1}{a}{\bf \nabla} \cdot \left[(1+\delta)
425: \ub\right] &=& 0 \\ \frac{\partial \ub}{\partial t}+H\ub+\frac{1}{a}(\ub
426: \cdot {\bf \nabla}) \ub&=&-\frac{1}{a}{\bf \nabla}\phi\ . \eeqa The
427: potential $\phi$ is given by the Poisson equation, in terms of the density
428: perturbation: \beq \nabla^2\phi=4 \pi G \bar{\rho} a^2 \delta\ . \eeq This
429: fluid description is valid for describing the evolution of collisionless
430: cold dark matter particles until different particle streams cross. This
431: ``shell-crossing'' typically occurs only after perturbations have grown to
432: become non-linear, and at that point the individual particle trajectories
433: must in general be followed. Similarly, baryons can be described as a
434: pressureless fluid as long as their temperature is negligibly small, but
435: non-linear collapse leads to the formation of shocks in the gas.
436:
437: For small perturbations $\delta \ll 1$, the fluid equations can be
438: linearized and combined to yield \beq \frac{\partial^2\delta}{\partial
439: t^2}+2 H \frac{\partial\delta}{\partial t}=4 \pi G \bar{\rho} \delta\
440: . \eeq This linear equation has in general two independent solutions, only
441: one of which grows with time. Starting with random initial conditions, this
442: ``growing mode'' comes to dominate the density evolution. Thus, until it
443: becomes non-linear, the density perturbation maintains its shape in
444: comoving coordinates and grows in proportion to a growth factor $D(t)$. The
445: growth factor in the matter-dominated era is given by \beq D(t)
446: \propto \frac{\left(\Oml a^3+\Omk a+\Omm\right)^{1/2}}{a^{3/2}}\int_0^a
447: \frac{a'^{3/2}\, da'}{\left(\Oml a'^3+\Omk a'+\Omm\right)^{3/2}}\ , \eeq
448: where we neglect $\Omr$ when considering halos forming in the
449: matter-dominated regime at $z \ll 10^4$. In the Einstein-de Sitter model
450: (or, at high redshift, in other models as well) the growth factor is simply
451: proportional to $a(t)$.
452:
453: The spatial form of the initial density fluctuations can be described in
454: Fourier space, in terms of Fourier components \beq \delta_\kb = \int d^3x\,
455: \delta(x) e^{-i \kb \cdot \xb}\ .\eeq Here we use the comoving wave-vector
456: $\kb$, whose magnitude $k$ is the comoving wavenumber which is equal to
457: $2\pi$ divided by the wavelength. The Fourier description is particularly
458: simple for fluctuations generated by inflation. Inflation generates
459: perturbations given by a Gaussian random field, in which different
460: $\kb$-modes are statistically independent, each with a random phase. The
461: statistical properties of the fluctuations are determined by the variance
462: of the different $\kb$-modes, and the variance is described in terms of the
463: power spectrum $P(k)$ as follows: \beq \left<\delta_{\kb} \delta_{{\bf
464: k'}}^{*}\right>=\left(2 \pi\right)^3 P(k)\, \delta^{(3)} \left(\kb-{\bf
465: k'}\right)\ , \eeq where $\delta^{(3)}$ is the three-dimensional Dirac
466: delta function. The gravitational potential fluctuations are sourced by
467: the density fluctuations through Poisson's equation.
468:
469: In standard models, inflation produces a primordial power-law spectrum
470: $P(k) \propto k^n$ with $n \sim 1$. Perturbation growth in the
471: radiation-dominated and then matter-dominated Universe results in a
472: modified final power spectrum, characterized by a turnover at a scale
473: of order the horizon $cH^{-1}$ at matter-radiation equality, and a
474: small-scale asymptotic shape of $P(k) \propto k^{n-4}$. The overall
475: amplitude of the power spectrum is not specified by current models of
476: inflation, and it is usually set by comparing to the observed CMB
477: temperature fluctuations or to local measures of large-scale
478: structure.
479:
480: Since density fluctuations may exist on all scales, in order to determine
481: the formation of objects of a given size or mass it is useful to consider
482: the statistical distribution of the smoothed density field. Using a window
483: function $W(\rb)$ normalized so that $\int d^3r\, W(\rb)=1$, the smoothed
484: density perturbation field, $\int d^3r \delta(\xb) W(\rb)$, itself follows
485: a Gaussian distribution with zero mean. For the particular choice of a
486: spherical top-hat, in which $W=1$ in a sphere of radius $R$ and is zero
487: outside, the smoothed perturbation field measures the fluctuations in the
488: mass in spheres of radius $R$. The normalization of the present power
489: spectrum is often specified by the value of $\sigma_8 \equiv \sigma(R=8
490: h^{-1} {\rm Mpc})$. For the top-hat, the smoothed perturbation field is
491: denoted $\delta_R$ or $\delta_M$, where the mass $M$ is related to the
492: comoving radius $R$ by $M=4 \pi \rho_m R^3/3$, in terms of the current mean
493: density of matter $\rho_m$. The variance $\langle \delta_M \rangle^2$ is
494: \beq \sigma^2(M)= \sigma^2(R)= \int_0^{\infty}\frac{dk}{2 \pi^2} \,k^2 P(k)
495: \left[\frac{3 j_1(kR)}{kR} \right]^2\ ,\label{eqsigM}\eeq where
496: $j_1(x)=(\sin x-x \cos x)/x^2$. The function $\sigma(M)$ plays a crucial
497: role in estimates of the abundance of collapsed objects, as we describe
498: later.
499:
500: Different physical processes contributed to the perturbation growth. In the
501: absence of other influences, gravitational forces due to density
502: perturbations imprinted by inflation would have driven parallel
503: perturbation growth in the dark matter, baryons and photons. However, since
504: the photon sound speed is of order the speed of light, the radiation
505: pressure produced sound waves on a scale of order the cosmic horizon and
506: suppressed sub-horizon perturbations in the photon density. The baryonic
507: pressure similarly suppressed perturbations in the gas below the (much
508: smaller) so-called baryonic {\it Jeans} scale. Since the formation of
509: hydrogen at recombination had decoupled the cosmic gas from its mechanical
510: drag on the CMB, the baryons subsequently began to fall into the
511: pre-existing gravitational potential wells of the dark matter.
512:
513: Spatial fluctuations developed in the gas temperature as well as in the gas
514: density. Both the baryons and the dark matter were affected on small scales
515: by the temperature fluctuations through the gas pressure. Compton heating
516: due to scattering of the residual free electrons (constituting a fraction
517: $\sim10^{-4}$) with the CMB photons remained effective, keeping the gas
518: temperature fluctuations tied to the photon temperature fluctuations, even
519: for a time after recombination. The growth of linear perturbations can be
520: calculated with the standard CMBFAST code (http://www.cmbfast.org), after a
521: modification to account for the fact that the speed of sound of the gas
522: also fluctuates spatially.
523:
524: \begin{figure}
525: \epsfxsize=10cm \epsfbox{binfall9.eps}
526: \caption{Power spectra of density and temperature fluctuations vs.\
527: comoving wavenumber, at redshifts 1200, 800, 400, and 200 (from Barkana \&
528: Loeb 2007). We consider fluctuations in the CDM density (solid curves),
529: baryon density (dotted curves), baryon temperature (short-dashed curves),
530: and photon temperature (long-dashed curves).}
531: \label{fig:photons}
532: \end{figure}
533:
534: The magnitude of the fluctuations in the CDM and baryon densities, and in
535: the baryon and photon temperatures, is shown in Figure~\ref{fig:photons},
536: in terms of the dimensionless combination $[k^3 P(k)/(2 \pi^2)]^{1/2}$,
537: where $P(k)$ is the corresponding power spectrum of fluctuations in terms
538: of the comoving wavenumber $k$ of each Fourier mode. After recombination,
539: two main drivers affect the baryon density and temperature fluctuations,
540: namely, the thermalization with the CMB and the gravitational force that
541: attracts the baryons to the dark matter potential wells. As shown in the
542: figure, the density perturbations in all species grow together on scales
543: where gravity is unopposed, outside the horizon (i.e., at $k \lesssim 0.01$
544: Mpc$^{-1}$ at $z \sim 1000$). At $z=1200$ the perturbations in the
545: baryon-photon fluid oscillate as acoustic waves on scales of order the
546: sound horizon ($k \sim 0.01~{\rm Mpc^{-1}}$), while smaller-scale
547: perturbations in both the photons and baryons are damped by photon
548: diffusion and the drag of the diffusing photons on the baryons. On
549: sufficiently small scales the power spectra of baryon density and
550: temperature roughly assume the shape of the dark matter fluctuations
551: (except for the gas-pressure cutoff at the very smallest scales), due to
552: the effect of gravitational attraction on the baryon density and of the
553: resulting adiabatic expansion on the gas temperature. After the mechanical
554: coupling of the baryons to the photons ends at $z \sim 1000$, the baryon
555: density perturbations gradually grow towards the dark matter perturbations
556: because of gravity. Similarly, after the thermal coupling ends at $z \sim
557: 200$, the baryon temperature fluctuations are driven by adiabatic expansion
558: towards a value of 2/3 of the density fluctuations. As the figure shows, by
559: $z=200$ the baryon infall into the dark matter potentials is well advanced
560: and adiabatic expansion is becoming increasingly important in setting the
561: baryon temperature.
562:
563:
564: \subsection{Halo properties}
565:
566: The small density fluctuations evidenced in the CMB grow over time as
567: described in the previous subsection, until the perturbation $\delta$
568: becomes of order unity, and the full non-linear gravitational problem
569: must be considered. The dynamical collapse of a dark matter halo can
570: be solved analytically only in cases of particular symmetry. If we
571: consider a region which is much smaller than the horizon $cH^{-1}$,
572: then the formation of a halo can be formulated as a problem in
573: Newtonian gravity, in some cases with minor corrections coming from
574: General Relativity. The simplest case is that of spherical symmetry,
575: with an initial ($t=t_i\ll t_0$) top-hat of uniform overdensity
576: $\delta_i$ inside a sphere of radius $R$. Although this model is
577: restricted in its direct applicability, the results of spherical
578: collapse have turned out to be surprisingly useful in understanding
579: the properties and distribution of halos in models based on cold dark
580: matter.
581:
582: The collapse of a spherical top-hat perturbation is described by the
583: Newtonian equation (with a correction for the cosmological constant) \beq
584: \frac{d^2r}{dt^2}=H_0^2 \Oml\, r-\frac{GM}{r^2}\ , \eeq where $r$ is the
585: radius in a fixed (not comoving) coordinate frame, $H_0$ is the present-day
586: Hubble constant, $M$ is the total mass enclosed within radius $r$, and the
587: initial velocity field is given by the Hubble flow $dr/dt=H(t) r$. The
588: enclosed $\delta$ grows initially as $\delta_L=\delta_i D(t)/D(t_i)$, in
589: accordance with linear theory, but eventually $\delta$ grows above
590: $\delta_L$. If the mass shell at radius $r$ is bound (i.e., if its total
591: Newtonian energy is negative) then it reaches a radius of maximum expansion
592: and subsequently collapses. As demonstrated in the previous section, at the
593: moment when the top-hat collapses to a point, the overdensity predicted by
594: linear theory is $\delta_L\, = 1.686$ in the Einstein-de Sitter model, with
595: only a weak dependence on $\Omm$ and $\Oml$. Thus a tophat collapses at
596: redshift $z$ if its linear overdensity extrapolated to the present day
597: (also termed the critical density of collapse) is \beq \delta_{\rm
598: crit}(z)=\frac{1.686}{D(z)}\ ,
599: \label{deltac} \eeq where we set $D(z=0)=1$.
600:
601: Even a slight violation of the exact symmetry of the initial perturbation
602: can prevent the tophat from collapsing to a point. Instead, the halo
603: reaches a state of virial equilibrium by violent relaxation (phase
604: mixing). Using the virial theorem $U=-2K$ to relate the potential energy
605: $U$ to the kinetic energy $K$ in the final state (implying that the virial
606: radius is half the turnaround radius - where the kinetic energy vanishes),
607: the final overdensity relative to the critical density at the collapse
608: redshift is $\Delta_c=18\pi^2 \simeq 178$ in the Einstein-de Sitter model,
609: modified in a Universe with $\Omm+\Oml=1$ to the fitting formula \beq
610: \Delta_c=18\pi^2+82 d-39 d^2\ , \eeq where $d\equiv \Ommz-1$ is evaluated
611: at the collapse redshift, so that \beq \Ommz=\frac{\Omm (1+z)^3}{\Omm
612: (1+z)^3+\Oml+\Omk (1+z)^2}\ .
613: \label{Ommz} \eeq
614:
615: A halo of mass $M$ collapsing at redshift $z$ thus has a virial radius \beq
616: r_{\rm vir}=0.784 \left(\frac{M}{10^8\ h^{-1} \ M_{\sun} }\right)^{1/3}
617: \left[\frac{\Omm} {\Ommz}\ \frac{\Delta_c} {18\pi^2}\right]^{-1/3} \left
618: (\frac{1+z}{10} \right)^{-1}\ h^{-1}\ {\rm kpc}\ , \label{rvir}\eeq and a
619: corresponding circular velocity, \beq V_c=\left(\frac{G M}{r_{\rm
620: vir}}\right)^{1/2}= 23.4 \left( \frac{M}{10^8\ h^{-1} \ M_{\sun}
621: }\right)^{1/3} \left[\frac {\Omm} {\Ommz}\ \frac{\Delta_c}
622: {18\pi^2}\right]^{1/6} \left( \frac{1+z} {10} \right)^{1/2}\ {\rm km\
623: s}^{-1}\ . \label{Vceqn} \eeq In these expressions we have assumed a
624: present Hubble constant written in the form $H_0=100\, h\mbox{ km
625: s}^{-1}\mbox{Mpc}^{-1}$. We may also define a virial temperature \beq
626: \label{tvir} T_{\rm vir}=\frac{\mu m_p V_c^2}{2 \kB}=1.98\times 10^4\
627: \left(\frac{\mu}{0.6}\right) \left(\frac{M}{10^8\ h^{-1} \ M_{\sun}
628: }\right)^{2/3} \left[ \frac {\Omm} {\Ommz}\ \frac{\Delta_c}
629: {18\pi^2}\right]^{1/3} \left(\frac{1+z}{10}\right)\ {\rm K} \ , \eeq where
630: $\mu$ is the mean molecular weight and $m_p$ is the proton mass. Note that
631: the value of $\mu$ depends on the ionization fraction of the gas; for a
632: fully ionized primordial gas $\mu=0.59$, while a gas with ionized hydrogen
633: but only singly-ionized helium has $\mu=0.61$. The binding energy of the
634: halo is approximately\footnote{The coefficient of $1/2$ in
635: equation~(\ref{Ebind}) would be exact for a singular isothermal sphere with
636: $\rho(r)\propto 1/r^2$.} \beq \label{Ebind} E_b= {1\over 2}
637: \frac{GM^2}{r_{\rm vir}} = 5.45\times 10^{53} \left(\frac{M}{10^8\ h^{-1} \
638: M_{\sun} }\right)^{5/3} \left[ \frac {\Omm} {\Ommz}\ \frac{\Delta_c}
639: {18\pi^2}\right]^{1/3} \left(\frac{1+z}{10}\right) h^{-1}\ {\rm erg}\
640: . \eeq Note that the binding energy of the baryons is smaller by a factor
641: equal to the baryon fraction $\Omega_b/\Omm$.
642:
643: Although spherical collapse captures some of the physics governing the
644: formation of halos, structure formation in cold dark matter models proceeds
645: hierarchically. At early times, most of the dark matter is in low-mass
646: halos, and these halos continuously accrete and merge to form high-mass
647: halos. Numerical simulations of hierarchical halo formation indicate a
648: roughly universal spherically-averaged density profile for the resulting
649: halos, though with considerable scatter among different halos. The typical
650: profile has the form \beq \rho(r)=\frac{3 H_0^2} {8 \pi G} (1+z)^3
651: \frac{\Omm}{\Ommz} \frac{\delta_c} {\cN x (1+\cN x)^2}\ ,
652: \label{NFW} \eeq where $x=r/r_{\rm vir}$, and the characteristic density
653: $\delta_c$ is related to the concentration parameter $\cN$ by \beq
654: \delta_c=\frac{\Delta_c}{3} \frac{\cN^3} {\ln(1+\cN)-\cN/(1+\cN)} \ . \eeq
655: The concentration parameter itself depends on the halo mass $M$, at a given
656: redshift $z$.
657:
658:
659: \subsection{Formation of the first stars}
660:
661: Theoretical expectations for the properties of the first galaxies are based
662: on the standard cosmological model outlined in the Introduction. The
663: formation of the first bound objects marked the central milestone in the
664: transition from the initial simplicity (discussed in the previous
665: subsection) to the present-day complexity. Stars and accreting black holes
666: output copious radiation and also produced explosions and outflows that
667: brought into the IGM chemical products from stellar nucleosynthesis and
668: enhanced magnetic fields. However, the formation of the very first stars,
669: in a universe that had not yet suffered such feedback, remains a
670: well-specified problem for theorists.
671:
672: \begin{figure}
673: %\epsfxsize=5cm \epsfbox{Fig2a.eps,Fig2b.eps}
674: \epsfxsize=7cm \epsfbox{Fig2b.eps}
675: \caption{Collapse and fragmentation of a primordial cloud of gas (from
676: Bromm \& Loeb 2004). Shown is the projected gas density at a redshift
677: $z\simeq 21.5$, briefly after gravitational runaway collapse has commenced
678: in the center of the cloud.
679: %{\it Top:} The coarse-grained morphology in a
680: %box with linear physical size of 23.5~pc. At this time in the unrefined
681: %simulation, a high-density clump has formed with an initial mass of $\sim
682: %10^{3}M_{\odot}$. {\it Right:}
683: The refined morphology is plotted in a simulation box with linear
684: physical size of 0.5~pc. The central density peak, vigorously gaining mass
685: by accretion, defines the seed of a metal-free (Population III) star.
686: Searches for metal-poor stars are underway in the halo of the Milky Way
687: galaxy, an environment less crowded by metal-rich (Population I,II) stars
688: than the core of our galaxy. The goal of these searches is to constrain
689: theoretical calculations (such as the one shown here) for the formation of
690: the first stars. }
691: \label{2ab}
692: \end{figure}
693:
694:
695: Stars form when huge amounts of matter collapse to enormous
696: densities. However, the process can be stopped if the pressure exerted by
697: the hot intergalactic gas prevents outlying gas from falling into dark
698: matter concentrations. As the gas falls into a dark matter halo, it forms
699: shocks due to converging supersonic flows and in the process heats up and
700: can only collapse further by first radiating its energy away. This
701: restricts this process of collapse to very large clumps of dark matter that
702: are around $100\,000$ times the mass of the Sun. Inside these clumps, the
703: shocked gas loses energy by emitting radiation from excited molecular
704: hydrogen that formed naturally within the primordial gas mixture of
705: hydrogen and helium.
706:
707: The first stars are expected to have been quite different from the stars
708: that form today in the Milky Way. The higher pressure within the primordial
709: gas due to the presence of fewer cooling agents suggests that fragmentation
710: only occurred into relatively large units, in which gravity could overcome
711: the pressure. Due to the lack of carbon, nitrogen, and oxygen -- elements
712: that would normally dominate the nuclear energy production in modern
713: massive stars -- the first stars must have condensed to extremely high
714: densities and temperatures before nuclear reactions were able to heat the
715: gas and balance gravity. These unusually massive stars produced high
716: luminosities of UV photons, but their nuclear fuel was exhausted after 2--3
717: million years, resulting in a huge supernova or in collapse to a black
718: hole. The heavy elements which were dispersed by the first supernovae in
719: the surrounding gas, enabled the enriched gas to cool more effectively and
720: fragment into lower mass stars. Simple calculations indicate that a carbon
721: or oxygen enrichment of merely $\la 10^{-3}$ of the solar abundance is
722: sufficient to allow solar mass stars to form. These second-generation
723: ``low-metallicity'' stars are long-lived and could in principle be
724: discovered in the halo of the Milky Way galaxy, providing fossil record of
725: the earliest star formation episode in our cosmic environment.
726:
727: Advances in computing power have made possible detailed numerical
728: simulations of how the first stars formed. These simulations begin in
729: the early universe, in which dark matter and gas are distributed uniformly,
730: apart from tiny variations in density and temperature that are
731: statistically distributed according to the patterns observed in the CMB. In
732: order to span the vast range of scales needed to simulate an individual
733: star within a cosmological context, the latest code
734: follows a box 0.3 Mpc in length and zooms in repeatedly on the densest part
735: of the first collapsing cloud that is found within the simulated
736: volume. The simulation follows gravity, hydrodynamics, and chemical
737: processes in the primordial gas, and resolves a scale 10 orders of
738: magnitudes smaller than that of the simulated box. While the resolved scale
739: is still three orders of magnitudes larger than the size of the Sun, these
740: simulations have established that the first stars formed within halos
741: containing $\sim 10^5 M_{\odot}$ in total mass, and indicate that the first
742: stars most likely weighed $\sim 100 M_{\odot}$ each.
743:
744: To estimate {\it when}\/ the first stars formed we must remember that the
745: first $100\,000$ solar mass halos collapsed in regions that happened to
746: have a particularly high density enhancement very early on. There was
747: initially only a small abundance of such regions in the entire universe, so
748: a simulation that is limited to a small volume is unlikely to find such
749: halos until much later. Simulating the entire universe is well beyond the
750: capabilities of current simulations, but analytical models predict that the
751: first observable star in the universe probably formed 30 million years
752: after the Big Bang, less than a quarter of one percent of the Universe's
753: total age of 13.7 billion years.
754:
755: Although stars were extremely rare at first, gravitational collapse
756: increased the abundance of galactic halos and star formation sites with
757: time (Figure~\ref{fig:history}). Radiation from the first stars is expected
758: to have eventually dissociated all the molecular hydrogen in the
759: intergalactic medium, leading to the domination of a second generation of
760: larger galaxies where the gas cooled via radiative transitions in atomic
761: hydrogen and helium. Atomic cooling occurred in halos of mass above
762: $\sim10^8 M_{\odot}$, in which the infalling gas was heated above 10,000 K
763: and became ionized. The first galaxies to form through atomic cooling are
764: expected to have formed around redshift 45, and such
765: galaxies were likely the main sites of star formation by the time
766: reionization began in earnest. As the IGM was heated above 10,000 K by
767: reionization, its pressure jumped and prevented the gas from accreting into
768: newly forming halos below $\sim10^9 M_{\odot}$. The first Milky-Way-sized
769: halo $M = 10^{12} M_{\odot}$ is predicted to have formed 400 million years
770: after the Big Bang, but such halos have become typical
771: galactic hosts only in the last five billion years.
772:
773: Hydrogen is the most abundant element in the Universe, The prominent
774: Lyman-$\alpha$ spectral line of hydrogen (corresponding to a transition
775: from its first excited level to its ground state) provides an important
776: probe of the condensation of primordial gas into the first galaxies.
777: Existing searches for Lyman-$\alpha$ emission have discovered galaxies
778: robustly out to a redshift $z\sim 7$ with some unconfirmed candidate
779: galaxies out to $z\sim 10$. The spectral break owing to Lyman-$\alpha$
780: absorption by the IGM allows to identify high-redshifts galaxies
781: photometrically. Existing observations provide only a preliminary glimpse
782: into the formation of the first galaxies.
783:
784: Within the next decade, NASA plans to launch an infrared space telescope
785: ({\it JWST}\/; Figure~\ref{jwst}) that will image some of the earliest
786: sources of light (stars and black holes) in the Universe. In parallel,
787: there are several initiatives to construct large-aperture infrared
788: telescopes on the ground with the same goal in mind (see
789: http://www.eso.org/public/astronomy/projects/e-elt.html;
790: http://www.tmt.org/;
791: http://www.gmto.org/).
792:
793: \begin{figure}
794: \epsfxsize=10cm \epsfbox{JWST.eps}
795: \caption{A sketch of the current design for the {\it James Webb Space
796: Telescope}, the successor to the {\it Hubble Space Telescope}\/ to be
797: launched in 2013 (see http://www.jwst.nasa.gov/). The current design
798: includes a primary mirror made of beryllium which is 6.5 meters in
799: diameter as well as an instrument sensitivity that spans the full
800: range of infrared wavelengths of 0.6--28$\mu$m and will allow
801: detection of some of the first galaxies in the infant Universe. The
802: telescope will orbit 1.5 million km from Earth at the Lagrange L2
803: point. Note that the sun shield (the large flat screen in the image)
804: is 22m$\times$10m in size.}
805: \label{jwst}
806: \end{figure}
807:
808:
809: The next generation of ground-based telescopes will have a diameter of
810: twenty to thirty meters (Figure~\ref{gmt}). Together with {\it JWST}\/
811: (which will not be affected by the atmospheric background) they will be
812: able to image and make spectral studies of the early galaxies. Given that
813: these galaxies also create the ionized bubbles around them by their UV
814: emission, during reionization the locations of galaxies should correlate
815: with bubbles within the neutral hydrogen. Within a decade it should be
816: possible to explore the environmental influence of individual galaxies by
817: using these telescopes in combination with 21-cm probes of reionization.
818:
819: \begin{figure}
820: \epsfxsize=10cm \epsfbox{gmt.ps}
821: \caption{Artist's conception of the design for one of the future giant
822: telescopes that might detect the first generation of galaxies from the
823: ground. The {\it Giant Magellan Telescope}\/ ({\it GMT}\/) is designed
824: to contain seven mirrors (each 8.4 meter in diameter) and to have a
825: resolving power equivalent to a 24.5 meter (80 foot) primary
826: mirror. For more details see http://www.gmto.org/ .
827: Two other teams are designing
828: competing large telescopes, namely the {\it Thirty Meter Telescope}
829: (see http://www.tmt.org/)
830: and the {\it European Extremely Large Telescope} (see
831: http://www.eso.org/public/astronomy/projects/e-elt.html).
832: }
833: \label{gmt}
834: \end{figure}
835:
836:
837: \subsection{Gamma-ray Bursts: probing the first stars one
838: star at a time}
839:
840: So far, to learn about diffuse IGM gas pervading the space outside and
841: between galaxies, astronomers routinely study its absorption signatures in
842: the spectra of distant quasars, the brightest long-lived astronomical
843: objects. Quasars' great luminosities are believed to be powered by
844: accretion of gas onto black holes weighing up to a few billion times the
845: mass of the Sun that are situated in the centers of massive galaxies. As
846: the surrounding gas spirals in toward the black hole sink, the viscous
847: dissipation of heat makes the gas glow brightly into space, creating a
848: luminous source visible from afar.
849:
850: Over the past decade, an alternative population of bright sources at
851: cosmological distances was discovered, the so-called afterglows of {\it
852: Gamma-Ray Bursts} (GRBs). These events are characterized by a flash of
853: high-energy ($>0.1$ MeV) photons, typically lasting 0.1--100 seconds, which
854: is followed by an afterglow of lower-energy photons over much longer
855: timescales. The afterglow peaks at X-ray, UV, optical and eventually radio
856: wavelengths on time scales of minutes, hours, days, and months,
857: respectively. The central engines of GRBs are believed to be associated
858: with the compact remnants (neutron stars or stellar-mass black holes) of
859: massive stars. Their high luminosities make them detectable out to the edge
860: of the visible Universe. GRBs offer the opportunity to detect the most
861: distant (and hence earliest) population of massive stars, the so-called
862: Population~III (or Pop~III), one star at a time (Figure~\ref{grb}). In the
863: hierarchical assembly process of halos that are dominated by cold dark
864: matter (CDM), the first galaxies should have had lower masses (and lower
865: stellar luminosities) than their more recent counterparts. Consequently,
866: the characteristic luminosity of galaxies or quasars is expected to decline
867: with increasing redshift. GRB afterglows, which already produce a peak flux
868: comparable to that of quasars or starburst galaxies at $z\sim 1-2$, are
869: therefore expected to outshine any competing source at the highest
870: redshifts, when the first dwarf galaxies formed in the Universe.
871:
872: \begin{figure}
873: \epsfxsize=10cm \epsfbox{GRB.ps}
874: \caption{Illustration of a long-duration gamma-ray burst in the popular
875: ``collapsar'' model.
876: %(from NASA E/PO)
877: The collapse of the core of a massive star (which lost its hydrogen
878: envelope) to a black hole generates two opposite jets moving out at a speed
879: close to the speed of light. The jets drill a hole in the star and shine
880: brightly towards an observer who happens to be located within with the
881: collimation cones of the jets. The jets emanating from a single massive
882: star are so bright that they can be seen across the Universe out to the
883: epoch when the first stars formed (for more information see
884: http://swift.gsfc.nasa.gov/). }
885: \label{grb}
886: \end{figure}
887:
888: GRBs, the electromagnetically-brightest explosions in the Universe, should
889: be detectable out to redshifts $z>10$. High-redshift GRBs can be identified
890: through infrared photometry, based on the Lyman-$\alpha$ break induced by
891: absorption of their spectrum at wavelengths below $1.216\, \mu {\rm m}\,
892: [(1+z)/10]$. Follow-up spectroscopy of high-redshift candidates can then be
893: performed on a 10-meter-class telescope. GRB afterglows offer the
894: opportunity to detect stars as well as to probe the metal enrichment level
895: of the intervening IGM. Recently, the {\it Swift} satellite has
896: detected a GRB originating at $z\simeq 6.3$, thus demonstrating the
897: viability of GRBs as probes of the early Universe.
898:
899: Another advantage of GRBs is that the GRB afterglow flux at a given
900: observed time lag after the $\gamma$-ray trigger is not expected to fade
901: significantly with increasing redshift, since higher redshifts translate to
902: earlier times in the source frame, during which the afterglow is
903: intrinsically brighter. For standard afterglow
904: lightcurves and spectra, the increase in the luminosity distance with
905: redshift is compensated by this {\it cosmological time-stretching} effect
906: as shown in Figure~\ref{fig:GRB}.
907:
908: \begin{figure}
909: \epsfxsize=8cm \epsfbox{fig1val.eps}
910: \caption{GRB afterglow flux as a function of time since the $\gamma$-ray
911: trigger in the observer frame (from Barkana \& Loeb 2004a). The flux
912: (solid curves) is calculated at the redshifted Lyman-$\alpha$
913: wavelength. The dotted curves show the planned detection threshold for the
914: {\it James Webb Space Telescope} ({\it JWST}), assuming a spectral
915: resolution $R=5000$ with the near infrared spectrometer, a signal to noise
916: ratio of 5 per spectral resolution element, and an exposure time equal to
917: $20\%$ of the time since the GRB explosion (see
918: http://www.ngst.stsci.edu/nms/main/~). Each set of curves shows a sequence
919: of redshifts, namely $z=5$, 7, 9, 11, 13, and 15, respectively, from top to
920: bottom.}
921: \label{fig:GRB}
922: \end{figure}
923:
924: GRB afterglows have smooth (broken power-law) continuum spectra unlike
925: quasars which show strong spectral features (such as broad emission lines
926: or the so-called ``blue bump'') that complicate the extraction of IGM
927: absorption features. In particular, the extrapolation into the spectral
928: regime marked by the IGM Lyman-$\alpha$ absorption during the epoch of
929: reionization is much more straightforward for the smooth UV spectra of GRB
930: afterglows than for quasars with an underlying broad Lyman-$\alpha$
931: emission line. However, the interpretation may be complicated by the
932: presence of damped Lyman-$\alpha$ absorption by dense neutral hydrogen in
933: the immediate environment of the GRB within its host galaxy. Since GRBs
934: originate from the dense environment of active star formation, such damped
935: absorption is expected and indeed has been seen, including in the most
936: distant GRB at $z=6.3$.
937:
938:
939: \subsection{Supermassive black holes}
940:
941: The fossil record in the present-day Universe indicates that every bulged
942: galaxy hosts a supermassive black hole (BH) at its center. This conclusion
943: is derived from a variety of techniques which probe the dynamics of stars
944: and gas in galactic nuclei. The inferred BHs are dormant or faint most of
945: the time, but occasionally flash in a short burst of radiation that lasts
946: for a small fraction of the age of the Universe. The short duty cycle
947: accounts for the fact that bright quasars are much less abundant than their
948: host galaxies, but it begs the more fundamental question: {\it why is the
949: quasar activity so brief?} A natural explanation is that quasars are
950: suicidal, namely the energy output from the BHs regulates their own growth.
951:
952: Supermassive BHs make up a small fraction, $< 10^{-3}$, of the total mass
953: in their host galaxies, and so their direct dynamical impact is limited to
954: the central star distribution where their gravitational influence
955: dominates. Dynamical friction on the background stars keeps the BH close to
956: the center. Random fluctuations in the distribution of stars induces a
957: Brownian motion of the BH. This motion can be described by the same Langevin
958: equation that captures the motion of a massive dust particle as it responds
959: to random kicks from the much lighter molecules of air around it. The
960: characteristic speed by which the BH wanders around the center is small,
961: $\sim (m_\star/M_{\rm BH})^{1/2}\sigma_\star$, where $m_\star$ and $M_{\rm
962: BH}$ are the masses of a single star and the BH, respectively, and
963: $\sigma_\star$ is the stellar velocity dispersion. Since the random force
964: fluctuates on a dynamical time, the BH wanders across a region that is
965: smaller by a factor of $\sim (m_\star/M_{\rm BH})^{1/2}$ than the region
966: traversed by the stars inducing the fluctuating force on it.
967:
968: The dynamical insignificance of the BH on the global galactic scale is
969: misleading. The gravitational binding energy per rest-mass energy of
970: galaxies is of order $\sim (\sigma_\star/c)^2< 10^{-6}$. Since BH are
971: relativistic objects, the gravitational binding energy of material that
972: feeds them amounts to a substantial fraction its rest mass energy. Even if
973: the BH mass amounts to a fraction as small as $\sim 10^{-4}$ of the
974: baryonic mass in a galaxy, and only a percent of the accreted rest-mass
975: energy is deposited into the gaseous environment of the BH, this slight
976: deposition can unbind the entire gas reservoir of the host galaxy. This
977: order-of-magnitude estimate explains why quasars may be short lived. As
978: soon as the central BH accretes large quantities of gas so as to
979: significantly increase its mass, it releases large amounts of energy and
980: momentum that could suppress further accretion onto it. In short, the BH
981: growth might be {\it self-regulated}.
982:
983: The principle of {\it self-regulation} naturally leads to a correlation
984: between the final BH mass, $M_{\rm bh}$, and the depth of the gravitational
985: potential well to which the surrounding gas is confined. The latter can be
986: characterized by the velocity dispersion of the associated stars, $\sim
987: \sigma_\star^2$. Indeed a correlation between $M_{\rm bh}$ and
988: $\sigma_\star^4$ is observed in the present-day Universe.
989: %The observed power-law relation between $M_{\rm bh}$ and $\sigma_\star$ can
990: %be generalized to a correlation between the BH mass and the circular
991: %velocity of the host halo, $v_c$, which in turn can be related to the halo
992: %mass, $M_{\rm halo}$, and redshift, $z$
993: %\begin{eqnarray}
994: %\label{eq:1}
995: %\nonumber M_{\rm bh}(M_{\rm halo},z) &=&\mbox{const} \times v_c^5\\
996: %&&\hspace{-25mm}= \epsilon_{\rm o} M_{\rm halo} \left(\frac{M_{\rm
997: %halo}}{10^{12}M_{\odot}}\right)^{\frac{2}{3}}
998: %[\zeta(z)]^\frac{5}{6}(1+z)^\frac{5}{2},
999: %\end{eqnarray}
1000: %where $\epsilon_{\rm o}\approx 10^{-5.7}$ is a constant, and as before
1001: %$\zeta\equiv [(\Omega_m/\Omega_m^z)(\Delta_c/18\pi^2)]$, $\Omega_m^z \equiv
1002: %[1+(\Omega_\Lambda/\Omega_m)(1+z)^{-3}]^{-1}$,
1003: %$\Delta_c=18\pi^2+82d-39d^2$, and $d=\Omega_m^z-1$.
1004: If quasars shine near
1005: their Eddington limit as suggested by observations of low and high-redshift
1006: quasars, then
1007: %the above value of $\epsilon_{\rm o}$ implies that
1008: a fraction of $\sim 5$--$10\%$ of the energy released by the quasar over a
1009: galactic dynamical time needs to be captured in the surrounding galactic
1010: gas in order for the BH growth to be self-regulated (see Figure
1011: \ref{merger}). With this interpretation, the $M_{\rm bh}$--$\sigma_\star$
1012: relation reflects the limit introduced to the BH mass by self-regulation;
1013: deviations from this relation are inevitable during episodes of BH growth
1014: or as a result of mergers of galaxies that have no cold gas in them. A
1015: physical scatter around this upper envelope could also result from
1016: variations in the efficiency by which the released BH energy couples to the
1017: surrounding gas.
1018:
1019: \begin{figure}
1020: \epsfxsize=12.5cm \epsfbox{ti.eps}
1021: \caption{Simulation images of a merger of galaxies resulting in quasar
1022: activity that eventually shuts-off the accretion of gas onto the black hole
1023: (from Di Matteo et al. 2005). The upper (lower) panels show a sequence of
1024: snapshots of the gas distribution during a merger with (without) feedback
1025: from a central black hole. The temperature of the gas is color coded.}
1026: \label{merger}
1027: \end{figure}
1028:
1029: Various prescriptions for self-regulation were sketched in the
1030: literature. These involve either energy or momentum-driven winds, with the
1031: latter type being a factor of $\sim v_c/c$ ~less efficient. The quasar
1032: remains active during the dynamical time of the initial gas reservoir,
1033: $\sim 10^7$ years, and fades afterwards due to the dilution of this
1034: reservoir. The BH growth may resume if the cold gas reservoir is
1035: replenished through a new merger. Following early analytic work, extensive
1036: numerical simulations demonstrated that galaxy mergers do produce the
1037: observed correlations between black hole mass and spheroid
1038: properties. Because of the limited resolution near the galaxy nucleus,
1039: these simulations adopt a simple prescription for the accretion flow that
1040: feeds the black hole. The actual feedback in reality may depend crucially
1041: on the geometry of this flow and the physical mechanism that couples the
1042: energy or momentum output of the quasar to the surrounding gas.
1043:
1044: The inflow of cold gas towards galaxy centers during the growth phase of
1045: the BH would naturally be accompanied by a burst of star formation. The
1046: fraction of gas that is not consumed by stars or ejected by
1047: supernova-driven winds, will continue to feed the BH. It is therefore not
1048: surprising that quasar and starburst activities co-exist in Ultra Luminous
1049: Infrared Galaxies, and that all quasars show broad metal lines indicating
1050: pre-enrichment of the surrounding gas with heavy elements.
1051: %Applying a similar self-regulation principle to the stars, leads to the
1052: %expectation that the ratio between the mass of the BH and the mass in stars
1053: %is independent of halo mass (as observed locally but increases with
1054: %redshift as $\propto \xi(z)^{1/2}(1+z)^{3/2}$. A consistent trend has
1055: %indeed been inferred in an observed sample of gravitationally-lensed
1056: %quasars \cite{Rix99}.
1057:
1058: The upper mass of galaxies may also be regulated by the energy output from
1059: quasar activity. This would account for the fact that cooling flows are
1060: suppressed in present-day X-ray clusters, and that massive BHs and stars in
1061: galactic bulges were already formed at $z\sim 2$.
1062: %At some epoch, the quasar energy output may have led to the
1063: %extinction of cold gas in these galaxies and the suppression of further
1064: %star formation in them, leading to an apparent ``anti-hierarchical'' mode
1065: %of galaxy formation where massive spheroids formed early and did not make
1066: %new stars at late times. In the course of subsequent merger events, the
1067: %cores of the most massive spheroids acquired an envelope of collisionless
1068: %matter in the form of already-formed stars or dark matter , without the
1069: %proportional accretion of cold gas into the central BH. The upper limit on
1070: %the mass of the central BH and the mass of the spheroid is caused by the
1071: %lack of cold gas and cooling flows in their X-ray halos.
1072: In the cores of cooling X-ray clusters, there is often an active central BH
1073: that supplies sufficient energy to compensate for the cooling of the
1074: gas. The primary physical process by which this energy couples to the gas
1075: is still unknown.
1076:
1077: The quasars discovered so far at $z\sim 6$ mark the early growth of the
1078: most massive BHs and galactic spheroids. The BHs powering these bright
1079: quasars possess a mass of a few billion solar masses. A quasar radiating
1080: at its Eddington limiting luminosity, $L_E=1.4\times 10^{47}~{\rm
1081: erg~s^{-1}}(M_{\rm bh}/10^9M_\odot)$, with a radiative efficiency,
1082: $\epsilon_{\rm rad}=L_{E}/{\dot M}c^2$, for converting accreted mass into
1083: radiation, would grow exponentially in mass as a function of time $t$,
1084: $M_{\rm bh} =M_{\rm seed}\exp\{t/t_E\}$ from its initial seed mass $M_{\rm
1085: seed}$, on a time scale, $t_E=4.1\times 10^7~{\rm yr} (\epsilon_{\rm
1086: rad}/0.1)$. Thus, the required growth time in units of the Hubble time
1087: $t_{\rm hubble}= 10^9~{\rm yr}[(1+z)/7]^{-3/2}$ is
1088: \begin{equation}
1089: {t_{\rm growth}\over t_{\rm hubble}}=0.7 \left({\epsilon_{\rm rad} \over
1090: 10\%}\right) \left({1+z\over 7}\right)^{3/2}\ln \left({{M_{\rm
1091: bh}/10^9M_\odot} \over M_{\rm seed}/100M_\odot}\right) ~.
1092: \end{equation}
1093: The age of the Universe at $z\sim 6$ provides just sufficient time to grow
1094: a BH with $M_{\rm bh}\sim 10^9M_\odot$ out of a stellar mass seed with
1095: $\epsilon_{\rm rad}=10 \%$. The growth time is shorter for smaller
1096: radiative efficiencies or a higher seed mass.
1097:
1098: \subsection{The epoch of reionization}
1099:
1100: Given the understanding described above of how many galaxies formed at
1101: various times, the course of reionization can be determined universe-wide
1102: by counting photons from all sources of light. Both stars and black holes
1103: contribute ionizing photons, but the early universe is dominated by small
1104: galaxies which in the local universe have central black holes that are
1105: disproportionately small, and indeed quasars are rare above redshift
1106: 6. Thus, stars most likely dominated the production of ionizing UV photons
1107: during the reionization epoch [although high-redshift galaxies should have
1108: also emitted X-rays from accreting black holes and accelerated particles in
1109: collisionless shocks]. Since most stellar ionizing photons are only
1110: slightly more energetic than the 13.6 eV ionization threshold of hydrogen,
1111: they are absorbed efficiently once they reach a region with substantial
1112: neutral hydrogen). This makes the IGM during reionization a two-phase
1113: medium characterized by highly ionized regions separated from neutral
1114: regions by sharp ionization fronts (see Figure~\ref{fig:rei}).
1115:
1116: \begin{figure}
1117: \epsfxsize=8cm \epsfbox{SciRev1.eps}
1118: %\hfill
1119: %\epsfxsize=5cm \epsfbox{image_yz_13.892_enhLoRes.ps}
1120: \caption{The spatial structure of cosmic reionization. The illustration
1121: shows how regions with large-scale overdensities form large concentrations
1122: of galaxies (dots) whose ionizing photons produce enormous joint ionized
1123: bubbles (upper left). At the same time, galaxies are rare within
1124: large-scale voids, in which the IGM is still mostly neutral (lower
1125: right).
1126: %A numerical simulation of reionization (right panel) indeed
1127: %displays such variation in the sizes of ionized bubbles (orange), shown
1128: %overlayed on the density distribution (green).
1129: }
1130: \label{fig:rei}
1131: \end{figure}
1132:
1133: We can obtain a first estimate of the requirements of reionization by
1134: demanding one stellar ionizing photon for each hydrogen atom in the
1135: IGM. If we conservatively assume that stars within the early galaxies
1136: were similar to those observed locally, then each star produced $\sim
1137: 4000$ ionizing photons per baryon. Star formation is observed today to
1138: be an inefficient process, but even if stars in galaxies formed out of
1139: only $\sim10\%$ of the available gas, it was still sufficient to
1140: accumulate a small fraction (of order $0.1\%$) of the total baryonic
1141: mass in the universe into galaxies in order to ionize the entire
1142: IGM. More accurate estimates of the actual required fraction account
1143: for the formation of some primordial stars (which were massive,
1144: efficient ionizers, as discussed above), and for recombinations of
1145: hydrogen atoms at high redshifts and in dense regions.
1146:
1147: From studies of quasar absorption lines at $z\sim 6$ we know that the IGM
1148: is highly ionized a billion years after the big bang. There are hints,
1149: however, that some large neutral hydrogen regions persist at these early
1150: times and so this suggests that we may not need to go to much higher
1151: redshifts to begin to see the epoch of reionization. We now know that the
1152: universe could not have fully reionized earlier than an age of 300 million
1153: years, since WMAP observed the effect of the freshly created plasma at
1154: reionization on the large-scale polarization anisotropies of the CMB and
1155: this limits the reionization redshift; an earlier reionization, when the
1156: universe was denser, would have created a stronger scattering signature
1157: that would be inconsistent with the WMAP observations. In any case, the
1158: redshift at which reionization ended only constrains the overall cosmic
1159: efficiency of ionizing photon production. In comparison, a detailed picture
1160: of reionization as it happens will teach us a great deal about the
1161: population of young galaxies that produced this cosmic phase transition.
1162: A key point is that the spatial distribution of ionized bubbles is
1163: determined by clustered groups of galaxies and not by individual
1164: galaxies. At such early times galaxies were strongly clustered even on very
1165: large scales (up to tens of Mpc), and these scales therefore dominate the
1166: structure of reionization. The basic idea is simple. At high redshift,
1167: galactic halos are rare and correspond to rare, high density peaks. As an
1168: analogy, imagine searching on Earth for mountain peaks above 5000
1169: meters. The 200 such peaks are not at all distributed uniformly but instead
1170: are found in a few distinct clusters on top of large mountain ranges. Given
1171: the large-scale boost provided by a mountain range, a small-scale crest
1172: need only provide a small additional rise in order to become a 5000 meter
1173: peak. The same crest, if it formed within a valley, would not come anywhere
1174: near 5000 meters in total height. Similarly, in order to find the early
1175: galaxies, one must first locate a region with a large-scale density
1176: enhancement, and then galaxies will be found there in abundance.
1177:
1178: The ionizing radiation emitted from the stars in each galaxy initially
1179: produces an isolated ionized bubble. However, in a region dense with
1180: galaxies the bubbles quickly overlap into one large bubble, completing
1181: reionization in this region while the rest of the universe is still mostly
1182: neutral (Figure~\ref{fig:rei}). Most importantly, since the abundance of
1183: rare density peaks is very sensitive to small changes in the density
1184: threshold, even a large-scale region with a small enhanced density (say,
1185: 10\% above the mean density of the universe) can have a much larger
1186: concentration of galaxies than in other regions (e.g., a 50\%
1187: enhancement). On the other hand, reionization is harder to achieve in dense
1188: regions, since the protons and electrons collide and recombine more often
1189: in such regions, and newly-formed hydrogen atoms need to be reionized again
1190: by additional ionizing photons. However, the overdense regions end up
1191: reionizing first since the number of ionizing sources in these regions is
1192: increased so strongly. The large-scale topology of reionization is
1193: therefore inside out, with underdense voids reionizing only at the very end
1194: of reionization, with the help of extra ionizing photons coming in from
1195: their surroundings (which have a higher density of galaxies than the voids
1196: themselves). This is a key prediction awaiting observational testing.
1197:
1198: Detailed analytical models that account for large-scale variations in the
1199: abundance of galaxies confirm that the typical bubble size starts well
1200: below a Mpc early in reionization, as expected for an individual galaxy,
1201: rises to 5--10 Mpc during the central phase (i.e., when the universe is
1202: half ionized), and then by another factor of $\sim$5 towards the end of
1203: reionization. These scales are given in comoving units that scale with the
1204: expansion of the universe, so that the actual sizes at a redshift $z$ were
1205: smaller than these numbers by a factor of $1+z$. Numerical simulations have
1206: only recently begun to reach the enormous scales needed to capture this
1207: evolution. Accounting precisely for gravitational evolution on a wide range
1208: of scales but still crudely for gas dynamics, star formation, and the
1209: radiative transfer of ionizing photons, the simulations confirm that the
1210: large-scale topology of reionization is inside out, and that this topology
1211: can be used to study the abundance and clustering of the ionizing sources
1212: (Figures~\ref{fig:rei}).
1213:
1214: The characteristic observable size of the ionized bubbles at the end
1215: reionization can be calculated based on simple considerations that only
1216: depend on the power-spectrum of density fluctuations and the redshift. As
1217: the size of an ionized bubble increases, the time it takes a 21-cm photon
1218: emitted by hydrogen to traverse it gets longer. At the same time, the
1219: variation in the time at which different regions reionize becomes smaller
1220: as the regions grow larger. Thus, there is a maximum size above which the
1221: photon crossing time is longer than the cosmic variance in ionization
1222: time. Regions bigger than this size will be ionized at their near side by
1223: the time a 21-cm photon will cross them towards the observer from their far
1224: side. They would appear to the observer as one-sided, and hence signal the
1225: end of reionization. These considerations imply a characteristic size for
1226: the ionized bubbles of $\sim 10$ physical Mpc at $z\sim 6$ (equivalent to
1227: 70 Mpc today). This result implies that future radio experiments should be
1228: tuned to a characteristic angular scale of tens of arcminutes for an
1229: optimal detection of 21-cm brightness fluctuations near the end of
1230: reionization (see \S \ref{sec:21-cmAtomic}).
1231:
1232: \subsection{Post-reionization suppression of low-mass galaxies}
1233:
1234: After the ionized bubbles overlapped in each region, the ionizing
1235: background increased sharply, and the IGM was heated by the ionizing
1236: radiation to a temperature $T_{\rm IGM}\gtrsim 10^4$ K. Due to the
1237: substantial increase in the IGM pressure, the smallest mass scale into
1238: which the cosmic gas could fragment, the so-called Jeans mass,
1239: increased dramatically, changing the minimum mass of forming galaxies.
1240:
1241: Gas infall depends sensitively on the Jeans mass. When a halo more massive
1242: than the Jeans mass begins to form, the gravity of its dark matter
1243: overcomes the gas pressure. Even in halos below the Jeans mass, although
1244: the gas is initially held up by pressure, once the dark matter collapses
1245: its increased gravity pulls in some gas. Thus, the Jeans mass is generally
1246: higher than the actual limiting mass for accretion. Before reionization,
1247: the IGM is cold and neutral, and the Jeans mass plays a secondary role in
1248: limiting galaxy formation compared to cooling. After reionization, the
1249: Jeans mass is increased by several orders of magnitude due to the
1250: photoionization heating of the IGM, and hence begins to play a dominant
1251: role in limiting the formation of stars. Gas infall in a reionized and
1252: heated Universe has been investigated in a number of numerical simulations.
1253: Three dimensional numerical simulations found a significant suppression of
1254: gas infall in even larger halos ($V_c \sim 75\ {\rm km\ s}^{-1}$), but this
1255: was mostly due to a suppression of late infall at $z\la 2$.
1256:
1257: When a volume of the IGM is ionized by stars, the gas is heated to a
1258: temperature $T_{\rm IGM}\sim 10^4$ K. If quasars dominate the UV background
1259: at reionization, their harder photon spectrum leads to $T_{\rm IGM}>
1260: 2\times 10^4$ K. Including the effects of dark matter, a given temperature
1261: results in a linear Jeans mass corresponding to a halo circular velocity of
1262: \beq V_J\approx 80 \left(\frac{T_{\rm IGM}}{1.5\times 10^4 {\rm
1263: K}}\right)^{1/2}\ {\rm km\ s}^{-1}. \eeq In halos with a circular
1264: velocity well above $V_J$, the gas fraction in infalling gas equals
1265: the universal mean of $\Omega_b/\Omega_m$, but gas infall is
1266: suppressed in smaller halos. A simple estimate of the limiting
1267: circular velocity, below which halos have essentially no gas infall,
1268: is obtained by substituting the virial overdensity for the mean
1269: density in the definition of the Jeans mass. The resulting estimate is
1270: \beq V_{\rm lim}=34 \left(\frac{T_{\rm IGM}}{1.5\times 10^4 {\rm
1271: K}}\right)^{1/2}\ {\rm km\ s}^{-1}. \eeq This value is in rough
1272: agreement with the numerical simulations mentioned before.
1273:
1274: Although the Jeans mass is closely related to the rate of gas infall at a
1275: given time, it does not directly yield the total gas residing in halos at a
1276: given time. The latter quantity depends on the entire history of gas
1277: accretion onto halos, as well as on the merger histories of halos, and an
1278: accurate description must involve a time-averaged Jeans mass.
1279: The gas content of halos in simulations is well fit by an expression which
1280: depends on the filtering mass, a particular time-averaged Jeans mass.
1281:
1282: The reionization process was not perfectly synchronized throughout the
1283: Universe. Large-scale regions with a higher density than the mean tended to
1284: form galaxies first and reionized earlier than underdense regions. The
1285: suppression of low-mass galaxies by reionization is therefore modulated by
1286: the fluctuations in the timing of reionization. Inhomogeneous reionization
1287: imprint a signature on the power-spectrum of low-mass galaxies. Future
1288: high-redshift galaxy surveys hoping to constrain inflationary parameters
1289: must properly model the effects of reionization; conversely, they will also
1290: place new constraints on the thermal history of the IGM during
1291: reionization.
1292:
1293: \section{Probing the Diffuse Intergalactic Hydrogen}
1294:
1295: \label{sec:absorb}
1296:
1297: \subsection{Lyman-alpha absorption}
1298:
1299: Resonant Lyman-$\alpha$ absorption has thus far proved to be the best probe
1300: of the state of the IGM. The optical depth to absorption by a uniform
1301: intergalactic medium is \beqa \tau_{s}&=&{\pi e^2 f_\alpha \lambda_\alpha
1302: n_{\HI}(z) \over m_e cH(z)} \label{G-P} \\ \nonumber &\approx& 6.45\times
1303: 10^5 x_{\HI} \left({\Omega_bh\over 0.0315}\right)\left({\Omega_m\over
1304: 0.3}\right)^{-1/2} \left({1+z\over 10}\right)^{3/2}\ , \eeqa where
1305: $H\approx 100h~{\rm km~s^{-1}~Mpc^{-1}} \Omega_m^{1/2} (1+z)^{3/2}$ is the
1306: Hubble parameter at redshift $z$; $f_\alpha=0.4162$ and
1307: $\lambda_\alpha=1216$\AA\, are the oscillator strength and the wavelength
1308: of the Lyman-$\alpha$ transition; $n_{\HI}(z)$ is the neutral hydrogen
1309: density at $z$ (assuming primordial abundances); $\Omega_m$ and $\Omega_b$
1310: are the present-day density parameters of all matter and of baryons,
1311: respectively; and $x_{\HI}$ is the average fraction of neutral hydrogen. In
1312: the second equality we have implicitly considered high redshifts.
1313:
1314: \begin{figure}
1315: \epsfxsize=12cm \epsfbox{White.eps}
1316: \caption{Using Lyman-$\alpha$ absorption in quasar spectra to probe the
1317: ionization state of the IGM. This figure from White et al. (2003) shows the
1318: observed spectrum of a $z = 6.28$ quasar (solid curve), and the expected
1319: unabsorbed emission (dashed curve), based on an average over many quasars
1320: seen at lower redshifts. The unabsorbed emission is a sum of smooth
1321: emission (the ``continuum'', dotted curve) plus emission features from
1322: atomic resonances (``emission lines''). }
1323: \label{fig:white}
1324: \end{figure}
1325:
1326: Lyman-$\alpha$ absorption is thus highly sensitive to the presence of even
1327: trace amounts of neutral hydrogen. The lack of full absorption in quasar
1328: spectra then implies that the IGM has been very highly ionized during much
1329: of the history of the universe, from at most a billion years after the big
1330: bang to the present time. At redshifts approaching six, however, the
1331: optical depth increases, and the observed absorption becomes very strong.
1332: An example of this is shown in Figure~\ref{fig:white}, where an observed
1333: quasar spectrum is compared to the unabsorbed expectation for the same
1334: quasar. The prominent Lyman-$\alpha$ emission line, which is produced by
1335: radiating hot gas near the quasar itself, is centered at a wavelength of
1336: 8850\AA, which for the redshift (6.28) of this quasar corresponds to a
1337: rest-frame 1216\AA. Above this wavelength, the original emitted quasar
1338: spectrum is seen, since photons emitted with wavelengths higher than
1339: 1216\AA\ redshift to higher wavelengths during their journey toward us and
1340: never encounter resonance lines of hydrogen atoms. Shorter-wavelength
1341: photons, however, redshift until they hit the local 1216\AA\ and are then
1342: absorbed by any existing hydrogen atoms. The difference between the
1343: unabsorbed expectation and the actual observed spectrum can be used to
1344: measure the amount of absorption, and thus to infer the atomic hydrogen
1345: density. For this particular quasar, this difference is very large (i.e.,
1346: the observed flux is near zero) just to the blue of the Lyman-$\alpha$
1347: emission line.
1348:
1349: \begin{figure}
1350: \epsfxsize=10cm \epsfbox{Fan.eps}
1351: \caption{Spectra of 19 quasars with redshifts $5.74<z<6.42$ from the {\it
1352: Sloan Digital Sky Survey}, taken from Fan et al. (2005). For some of the
1353: highest-redshift quasars, the spectrum shows no transmitted flux shortward
1354: of the Lyman-$\alpha$ wavelength at the quasar redshift (the so-called
1355: ``Gunn-Peterson trough''), indicating a non-negligible neutral fraction in
1356: the IGM.}
1357: \label{fig:qsos}
1358: \end{figure}
1359:
1360: Several quasars beyond $z\sim6.1$ show in their spectra such a (so-called
1361: ``Gunn-Peterson'') trough, a blank spectral region at wavelengths shorter
1362: than \lya at the quasar redshift (Figure~\ref{fig:qsos}). The detection
1363: of Gunn-Peterson troughs indicates a rapid change in the neutral content of
1364: the IGM at $z\sim6$, and hence a rapid change in the intensity of the
1365: background ionizing flux. However, even a small atomic hydrogen fraction of
1366: $\sim 10^{-3}$ would still produce nearly complete \lya absorption.
1367:
1368: While only resonant \lya absorption is important at moderate
1369: redshifts, the damping wing of the \lya line plays a significant role
1370: when neutral fractions of order unity are considered at $z \gtrsim 6$.
1371: The scattering cross-section of the \lya resonance line by neutral
1372: hydrogen is given by
1373: \begin{equation}
1374: \sigma_\alpha(\nu) = {3 \lambda_\alpha^2 \Lambda_\alpha^2 \over 8\pi}
1375: {(\nu/\nu_\alpha)^4\over
1376: 4\pi^2(\nu-\nu_\alpha)^2+(\Lambda_\alpha^2/4)(\nu/\nu_\alpha)^6}\ ,
1377: \label{eq:sig}
1378: \end{equation}
1379: where $\Lambda_\alpha=(8\pi^2 e^2
1380: f_\alpha/3m_ec\lambda_\alpha^2)=6.25\times 10^8~{\rm s^{-1}}$ is the
1381: \lya ($2p\rightarrow 1s$) decay rate, $f_\alpha=0.4162$ is the
1382: oscillator strength, and $\lambda_\alpha=1216$\AA\, and
1383: $\nu_\alpha=(c/\lambda_\alpha)=2.47\times 10^{15}~{\rm Hz}$ are the
1384: wavelength and frequency of the \lya line. The term in the numerator
1385: is responsible for the classical Rayleigh scattering.
1386:
1387: Although reionization is an inhomogeneous process, we consider here a
1388: simple illustrative case of instantaneous reionization. Consider a source
1389: at a redshift $z_s$ beyond the redshift of reionization, $\zr$, and the
1390: corresponding scattering optical depth of a uniform, neutral IGM of
1391: hydrogen density $n_{\rm H,0}(1+z)^3$ between the source and the
1392: reionization redshift. The optical depth is a function of the observed
1393: wavelength $\lambda_{\rm obs}$,
1394: %Defining,
1395: %\begin{equation}
1396: %\Delta \lambda= \lambda- \lambda_\alpha(1+z_s),
1397: %\end{equation}
1398: %we get
1399: \begin{equation}
1400: \tau(\lambda_{\rm obs})=\int_{\zr}^{z_s} dz\, {cdt\over dz}\, n_{\rm
1401: H,0} (1+z)^3 \sigma_\alpha\left[\nu_{\rm obs}(1+z)\right]\ ,
1402: \end{equation}
1403: where $\nu_{\rm obs}=c/\lambda_{\rm obs}$ and
1404: \begin{equation}
1405: {dt\over dz} = \left[(1+z)H(z)\right]^{-1}=H_0^{-1} \times
1406: \left[\Omega_m(1+z)^5+\Omega_\Lambda(1+z)^2+
1407: (1-\Omega_m-\Omega_\Lambda)(1+z)^4\right]^{-1/2}\ .
1408: \end{equation}
1409:
1410: At wavelengths longer than \lya at the source, the optical depth
1411: obtains a small value; these photons redshift away from the line
1412: center along its red wing and never resonate with the line core on
1413: their way to the observer. Considering only the regime in which
1414: $\vert\nu-\nu_\alpha\vert \gg \Lambda_\alpha$, we may ignore the
1415: second term in the denominator of equation~(\ref{eq:sig}). This leads
1416: to an analytical result for the red damping wing of the Gunn-Peterson
1417: trough,
1418: \begin{equation}
1419: \tau(\lambda_{\rm obs})=\tau_s \left(\Lambda\over
1420: 4\pi^2\nu_\alpha\right) {\tilde \lambda}_{\rm obs}^{3/2}\left[
1421: I({\tilde\lambda}_{\rm obs}^{-1}) -
1422: I([(1+\zr)/(1+z_s)]{\tilde\lambda}_{\rm obs}^{-1})\right]\ ,
1423: \label{eq:shift}
1424: \end{equation}
1425: an expression valid for ${\tilde\lambda}_{\rm obs}\geq 1$, where
1426: $\tau_s$ is given in equation~(\ref{G-P}), and we also define
1427: \begin{equation}
1428: {\tilde \lambda}_{\rm obs}\equiv {\lambda_{\rm obs}\over
1429: (1+z_s)\lambda_\alpha}
1430: \end{equation}
1431: and
1432: \begin{equation}
1433: I(x)\equiv {x^{9/2}\over 1-x}+{9\over 7}x^{7/2}+{9\over 5}x^{5/2}+ 3
1434: x^{3/2}+9 x^{1/2}-{9\over 2} \ln\left[ {1+x^{1/2}\over 1-x^{1/2}}
1435: \right]\ .
1436: \end{equation}
1437:
1438: At wavelengths shorter than 912\AA, the photons are absorbed when
1439: they photoionize atoms of hydrogen or helium. The bound-free
1440: absorption cross-section from the ground state of a hydrogenic ion
1441: with nuclear charge $Z$ and an ionization threshold $h\nu_0$, is given
1442: by
1443: \begin{equation}
1444: \sigma_{bf}(\nu)= {6.30\times 10^{-18}\over Z^2}~{\rm cm^2}\times
1445: \left({\nu_0\over
1446: \nu}\right)^4{e^{4-(4\tan^{-1}\epsilon)/\epsilon}\over 1 -
1447: e^{-2\pi/\epsilon}}~~~~{\rm for}~~\nu\geq\nu_0\ ,
1448: \end{equation}
1449: where $\epsilon\equiv \sqrt{{(\nu/\nu_0)}-1}$. For neutral hydrogen, $Z=1$
1450: and $\nu_{{\rm H},0}= (c/\lambda_c) =3.29\times 10^{15}$ Hz ($h\nu_{\rm
1451: H,0}=13.60$ eV); for singly-ionized helium, $Z=2$ and $\nu_{\rm He~II, 0}=
1452: 1.31\times 10^{16}~{\rm Hz}$ ($h\nu_{\rm He~II, 0}=54.42$ eV). The
1453: cross-section for neutral helium is more complicated; when averaged over
1454: its narrow resonances it can be fitted to an accuracy of a few percent up
1455: to $h\nu=50$ keV by the fitting function
1456: \begin{equation}
1457: \sigma_{bf,{\rm
1458: He~I}}(\nu)=9.492\times 10^{-16}~{\rm cm^2}\, \times
1459: \left[(x-1)^2+4.158\right]y^{-1.953}\left(1+ 0.825
1460: y^{1/4}\right)^{-3.188}\ ,
1461: \end{equation}
1462: where $x\equiv[(\nu/3.286\times 10^{15}~{\rm Hz}) -0.4434]$, $y\equiv
1463: x^2+4.563$, and the threshold for ionization is $\nu_{\rm He~I,
1464: 0}=5.938\times 10^{15}~{\rm Hz}$ ($h\nu_{\rm He~I,0}=24.59$ eV).
1465:
1466: For rough estimates, the average photoionization cross-section for a
1467: mixture of hydrogen and helium with cosmic abundances can be
1468: approximated in the range of $54<h\nu \lesssim 10^3$ eV as
1469: $\sigma_{bf}\approx \sigma_0 (\nu/\nu_{\rm H,0})^{-3}$, where
1470: $\sigma_0\approx 6\times 10^{-17}~{\rm cm^2}$.
1471: The redshift factor in the cross-section then cancels exactly the
1472: redshift evolution of the gas density and the resulting optical depth
1473: depends only on the elapsed cosmic time, $t(\zr)-t(z_s)$. At high
1474: redshifts this yields
1475: \begin{eqnarray}
1476: \tau_{bf}(\lambda_{\rm obs})&=&\int_{\zr}^{z_s} dz {cdt\over dz}
1477: n_{0} (1+z)^3 \sigma_{\rm bf}\left[\nu_{\rm obs}(1+z)\right]
1478: %={2cn_{H,0}\sigma_0\over
1479: %3H_0\sqrt{\Omega_m}}\left[{1\over (1+\zr)^{3/2}}-{1\over
1480: %(1+z_s)^{3/2}}\right]
1481: \nonumber \\ &\approx&
1482: 1.5\times 10^2 \left({\Omega_bh\over 0.03}\right) \left({\Omega_m\over
1483: 0.3}\right)^{-1/2}\left({\lambda\over 100{\mbox{\AA}}}\right)^{3}\times
1484: \left[{1\over (1+\zr)^{3/2}}
1485: -{1\over (1+z_s)^{3/2}}\right]\ . \label{eq:bf}
1486: \end{eqnarray}
1487: The bound-free optical depth only becomes of order unity in the
1488: extreme ultraviolet (UV) to soft X-rays, around $h\nu \sim 0.1$ keV, a
1489: regime which is unfortunately difficult to observe due to Galactic
1490: absorption.
1491:
1492:
1493: \subsection{21-cm absorption or emission}
1494: \label{sec:21-cmAtomic}
1495:
1496:
1497: \subsubsection{The spin temperature of the 21-cm transition of hydrogen}
1498:
1499:
1500: The ground state of hydrogen exhibits hyperfine splitting owing to the
1501: possibility of two relative alignments of the spins of the proton and the
1502: electron. The state with parallel spins (the triplet state) has a slightly
1503: higher energy than the state with anti-parallel spins (the singlet
1504: state). The 21-cm line associated with the spin-flip transition from the
1505: triplet to the singlet state is often used to detect neutral hydrogen in
1506: the local universe. At high redshift, the occurrence of a neutral
1507: pre-reionization IGM offers the prospect of detecting the first sources of
1508: radiation and probing the reionization era by mapping the 21-cm emission
1509: from neutral regions. While its energy density is estimated to be only a
1510: $1\%$ correction to that of the CMB, the redshifted 21-cm emission should
1511: display angular structure as well as frequency structure due to
1512: inhomogeneities in the gas density field, hydrogen ionized fraction, and
1513: spin temperature. Indeed, a full mapping of the distribution of H~I as a
1514: function of redshift is possible in principle.
1515:
1516: The basic physics of the hydrogen spin transition is determined as
1517: follows. The ground-state hyperfine levels of hydrogen tend to thermalize
1518: with the CMB background, making the IGM unobservable. If other processes
1519: shift the hyperfine level populations away from thermal equilibrium, then
1520: the gas becomes observable against the CMB in emission or in
1521: absorption. The relative occupancy of the spin levels is usually described
1522: in terms of the hydrogen spin temperature $T_S$, defined by \beq
1523: \frac{n_1}{n_0}=3\, \exp\left\{-\frac{T_*}{T_S}\right\}\ , \eeq where $n_0$
1524: and $n_1$ refer respectively to the singlet and triplet hyperfine levels in
1525: the atomic ground state ($n=1$), and $T_*=0.068$ K is defined by $k_B
1526: T_*=E_{21}$, where the energy of the 21 cm transition is $E_{21}=5.9 \times
1527: 10^{-6}$ eV, corresponding to a frequency of 1420 MHz. In the presence of
1528: the CMB alone, the spin states reach thermal equilibrium with $T_S=T_{\rm
1529: CMB}=2.725 (1+z)$ K on a time-scale of $T_*/(T_{\rm CMB} A_{10}) \simeq 3
1530: \times 10^5 (1+z)^{-1}$ yr, where $A_{10}=2.87 \times 10^{-15}$ s$^{-1}$ is
1531: the spontaneous decay rate of the hyperfine transition. This time-scale is
1532: much shorter than the age of the universe at all redshifts after
1533: cosmological recombination.
1534:
1535: The IGM is observable when the kinetic temperature $T_k$ of the gas differs
1536: from $T_{\rm CMB}$ and an effective mechanism couples $T_S$ to
1537: $T_k$. Collisional de-excitation of the triplet level dominates at very
1538: high redshift, when the gas density (and thus the collision rate) is still
1539: high, but once a significant galaxy population forms in the universe, the
1540: spin temperature is affected also by an indirect mechanism that acts
1541: through the scattering of Lyman-$\alpha$ photons. Continuum UV photons
1542: produced by early radiation sources redshift by the Hubble expansion into
1543: the local Lyman-$\alpha$ line at a lower redshift. These photons mix the
1544: spin states via the Wouthuysen-Field process whereby an atom initially in
1545: the $n=1$ state absorbs a Lyman-$\alpha$ photon, and the spontaneous decay
1546: which returns it from $n=2$ to $n=1$ can result in a final spin state which
1547: is different from the initial one. Since the neutral IGM is highly opaque
1548: to resonant scattering, and the Lyman-$\alpha$ photons receive Doppler
1549: kicks in each scattering, the shape of the radiation spectrum near
1550: Lyman-$\alpha$ is determined by $T_k$ (Field 1959), and the resulting spin
1551: temperature (assuming $T_S \gg T_*$) is then a weighted average of $T_k$
1552: and $T_{\rm CMB}$: \beq T_S=\frac{T_{\rm CMB} T_k (1+x_{\rm tot}) }{T_k +
1553: T_{\rm CMB} x_{\rm tot}}\ , \eeq where $x_{\rm tot} = x_{\alpha} + x_c$ is
1554: the sum of the radiative and collisional threshold parameters. These
1555: parameters are
1556: \begin{equation}
1557: x_{\alpha} = {{P_{10} T_\star}\over {A_{10} T_{\rm CMB}}}\ ,
1558: \end{equation}
1559: and
1560: \begin{equation}
1561: x_c = {{4 \kappa_{1-0}(T_k)\, n_H T_\star}\over {3 A_{10} T_{\rm CMB}}}\ ,\
1562: \end{equation} where $P_{10}$ is the indirect de-excitation rate of the
1563: triplet $n=1$ state via the Wouthuysen-Field process, related to the total
1564: scattering rate $P_{\alpha}$ of Lyman-$\alpha$ photons by $P_{10}=4
1565: P_{\alpha}/27$. Also, the atomic coefficient $\kappa_{1-0}(T_k)$ is
1566: tabulated as a function of $T_k$. The coupling of the spin temperature to
1567: the gas temperature becomes substantial when $x_{\rm tot} \gtrsim 1$; in
1568: particular, $x_{\alpha} = 1$ defines the thermalization rate of
1569: $P_{\alpha}$: \beq P_{\rm th} \equiv \frac{27 A_{10} T_{\rm CMB}}{4 T_*}
1570: \simeq 7.6 \times 10^{-12}\, \left(\frac{1+z}{10}\right)\ {\rm s}^{-1}\
1571: . \eeq
1572:
1573: A patch of neutral hydrogen at the mean density and with a uniform
1574: $T_S$ produces (after correcting for stimulated emission) an optical
1575: depth at a present-day (observed) wavelength of $21 (1+z)$ cm,
1576: \beq \tau(z) = 9.0
1577: \times 10^{-3} \left(\frac{T_{\rm CMB}} {T_S} \right) \left (
1578: \frac{\Omega_b h} {0.03} \right) \left(\frac{\Omm}{0.3}\right)^ {-1/2}
1579: \left(\frac{1+z}{10}\right)^{1/2}\ , \eeq assuming a high redshift
1580: $z\gg1$. The observed spectral intensity $I_{\nu}$ relative to the CMB at a
1581: frequency $\nu$ is measured by radio astronomers as an effective
1582: brightness temperature $T_b$ of blackbody emission at this frequency,
1583: defined using the Rayleigh-Jeans limit of the Planck radiation
1584: formula: $I_{\nu} \equiv 2 k_B T_b \nu^2 / c^2 $.
1585:
1586: The brightness temperature through the IGM is $T_b=T_{\rm CMB}
1587: e^{-\tau}+T_S (1-e^{-\tau})$, so the observed differential antenna
1588: temperature of this region relative to the CMB is \beqa T_b&=&(1+z)^{-1}
1589: (T_S-T_{\rm CMB}) (1-e^{-\tau}) \nonumber \\ &\simeq& 28\, {\rm mK}\,
1590: \left( \frac{\Omega_b h} {0.033} \right) \left(\frac{\Omm}{0.27}\right)^
1591: {-1/2} \left( \frac{1+z} {10} \right)^{1/2} \left( \frac{T_S-T_{\rm CMB}}
1592: {T_S} \right)\ , \eeqa where $\tau \ll 1$ is assumed and $T_b$ has been
1593: redshifted to redshift zero. Note that the combination that appears in
1594: $T_b$ is
1595: \begin{equation}
1596: {T_S - T_{\rm CMB} \over T_S} = {x_{\rm tot}\over 1+ x_{\rm tot}}
1597: \left(1 - {T_{\rm CMB}\over T_k} \right)\ .
1598: \end{equation}
1599: In overdense regions, the observed $T_b$ is proportional to the
1600: overdensity, and in partially ionized regions $T_b$ is proportional to the
1601: neutral fraction. Also, if $T_S \gg T_{\rm CMB}$ then the IGM is observed
1602: in emission at a level that is independent of $T_S$. On the other hand, if
1603: $T_S \ll T_{\rm CMB}$ then the IGM is observed in absorption at a level
1604: that is enhanced by a factor of $T_{\rm CMB} / T_S$. As a result, a number
1605: of cosmic events are expected to leave observable signatures in the
1606: redshifted 21-cm line, as discussed below in further detail.
1607:
1608: \begin{figure}
1609: \epsfxsize=8cm \epsfbox{tempZlong.eps}
1610: \caption{{\em Top panel: }Evolution with redshift $z$ of the CMB
1611: temperature $T_{\rm CMB}$ (dotted curve),the gas kinetic temperature $T_k$
1612: (dashed curve), and the spin temperature $T_S$ (solid curve), taken from
1613: Pritchard \& Loeb (2008). {\em Middle panel: }Evolution of the gas
1614: fraction in ionized regions $x_i$ (solid curve) and the ionized fraction
1615: outside these regions (due to diffuse X-rays) $x_e$ (dotted curve). {\em
1616: Bottom panel: } Evolution of mean 21 cm brightness temperature $T_b$. The
1617: horizontal axis at the top provides the observed photon frequency at the
1618: different redshifts shown at the bottom. Each panel shows curves for three
1619: models in which reionization is completed at different redshifts, namely
1620: $z=6.47$ (thin curves), $z=9.76$ (medium curves), and $z=11.76$ (thick
1621: curves). }
1622: \label{fig:tempZlong}
1623: \end{figure}
1624:
1625: Figure \ref{fig:tempZlong} illustrates the mean IGM evolution for three
1626: examples in which reionization is completed at different redshifts, namely
1627: $z=6.47$ (thin curves), $z=9.76$ (medium curves), and $z=11.76$ (thick
1628: curves). The top panel shows the global evolution of the CMB temperature
1629: $T_{\rm CMB}$ (dotted curve), the gas kinetic temperature $T_k$ (dashed
1630: curve), and the spin temperature $T_S$ (solid curve). The middle panel
1631: shows the evolution of the ionized gas fraction and the bottom panel
1632: presents the mean 21 cm brightness temperature, $T_b$.
1633:
1634: \subsubsection{A handy tool for studying cosmic reionization}
1635:
1636: The prospect of studying reionization by mapping the distribution of atomic
1637: hydrogen across the universe using its prominent 21-cm spectral line has
1638: motivated several teams to design and construct arrays of low-frequency
1639: radio telescopes; the Low Frequency Array (http://www.lofar.org/), the
1640: Mileura Wide-Field Array ({\it
1641: http://www.haystack.mit.edu/ast/arrays/mwa/site/index.html}), the Primeval
1642: Structure Telescope ({\it http://arxiv.org/abs/astro-ph/0502029}), and
1643: ultimately the Square Kilometer Array ({\it http://www.skatelescope.org})
1644: will search over the next decade for 21-cm emission or absorption from
1645: $z\sim 6.5$--15, redshifted and observed today at relatively low
1646: frequencies which correspond to wavelengths of 1.5 to 4 meters.
1647:
1648: The idea is to use the resonance associated with the hyperfine splitting in
1649: the ground state of hydrogen. While the CMB spectrum peaks at a wavelength
1650: of 2 mm, it provides a still-measurable intensity at meter wavelengths that
1651: can be used as the bright background source against which we can see the
1652: expected 1\% absorption by neutral hydrogen along the line of sight. The
1653: hydrogen gas produces 21-cm absorption if its spin temperature is colder
1654: than the CMB and excess emission if it is hotter. Since the CMB covers the
1655: entire sky, a complete three-dimensional map of neutral hydrogen can in
1656: principle be made from the sky position of each absorbing gas cloud
1657: together with its redshift $z$. Different observed wavelengths slice the
1658: Universe at different redshifts, and ionized regions are expected to appear
1659: as cavities in the hydrogen distribution, similar to holes in swiss cheese.
1660: Because the smallest angular size resolvable by a telescope is proportional
1661: to the observed wavelength, radio astronomy at wavelengths as large as a
1662: meter has remained relatively undeveloped. Producing resolved images even
1663: of large sources such as cosmological ionized bubbles requires telescopes
1664: which have a kilometer scale. It is much more cost-effective to use a large
1665: array of thousands of simple antennas distributed over several kilometers,
1666: and to use computers to cross-correlate the measurements of the individual
1667: antennas and combine them effectively into a single large telescope. The
1668: new experiments are being placed mostly in remote sites, because the cosmic
1669: wavelength region overlaps with more mundane terrestrial
1670: telecommunications.
1671:
1672: In approaching redshifted 21-cm observations, although the first inkling
1673: might be to consider the mean emission signal in the bottom panel of
1674: Figure~\ref{fig:tempZlong}, the signal is orders of magnitude fainter than
1675: foreground synchrotron emission from relativistic electrons in the magnetic
1676: field of our own Milky Way as well as other galaxies (see
1677: Figure~\ref{fig:Analytic}). Thus cosmologists have focused on the expected
1678: characteristic variations in $T_b$, both with position on the sky and
1679: especially with frequency, which signifies redshift for the cosmic
1680: signal. The synchrotron foreground is expected to have a smooth frequency
1681: spectrum, and so it is possible to isolate the cosmological signal by
1682: taking the difference in the sky brightness fluctuations at slightly
1683: different frequencies (as long as the frequency separation corresponds to
1684: the characteristic size of ionized bubbles). The 21-cm brightness
1685: temperature depends on the density of neutral hydrogen. As explained in the
1686: previous subsection, large-scale patterns in the reionization are driven by
1687: spatial variations in the abundance of galaxies; the 21-cm fluctuations
1688: reach $\sim$5 mK (root mean square) in brightness temperature
1689: (Figure~\ref{fig:Mellema}) on a scale of 10 Mpc (comoving). While detailed
1690: maps will be difficult to extract due to the foreground emission, a
1691: statistical detection of these fluctuations (through the power spectrum) is
1692: expected to be well within the capabilities of the first-generation
1693: experiments now being built. Current work suggests that the key information
1694: on the topology and timing of reionization can be extracted statistically.
1695:
1696: \begin{figure}
1697: \epsfxsize=16cm \epsfbox{slice_f250_kz75_notitleLoRes.eps}
1698: \caption{Close-up of cosmic evolution during the epoch of reionization, as
1699: revealed in a predicted 21-cm map of the IGM based on a numerical
1700: simulation (from Mellema et al. 2006). This map is constructed from slices
1701: of the simulated cubic box of side 150 Mpc (in comoving units), taken at
1702: various times during reionization, which for the parameters of this
1703: particular simulation spans a period of 250 million years from redshift 15
1704: down to 9.3. The vertical axis shows position $\chi$ in units of Mpc/h
1705: (where $h=0.7$). This two-dimensional slice of the sky (one linear
1706: direction on the sky versus the line-of-sight or redshift direction) shows
1707: $\log_{10}(T_b)$, where $T_b$ (in mK) is the 21-cm brightness temperature
1708: relative to the CMB. Since neutral regions correspond to strong emission
1709: (i.e., a high $T_b$), this slice illustrates the global progress of
1710: reionization and the substantial large-scale spatial fluctuations in
1711: reionization history. Observationally it corresponds to a narrow strip half
1712: a degree in length on the sky observed with radio telescopes over a
1713: wavelength range of 2.2 to 3.4 m (with each wavelength corresponding to
1714: 21-cm emission at a specific redshift slice).}
1715: \label{fig:Mellema}
1716: \end{figure}
1717:
1718: While numerical simulations of reionization are now reaching the
1719: cosmological box sizes needed to predict the large-scale topology of the
1720: ionized bubbles, they do this at the price of limited small-scale
1721: resolution (see Figure \ref{fig:Zahn}). These simulations cannot yet follow
1722: in any detail the formation of individual stars within galaxies, or the
1723: feedback that stars produce on the surrounding gas, such as photo-heating
1724: or the hydrodynamic and chemical impact of supernovae, which blow hot
1725: bubbles of gas enriched with the chemical products of stellar
1726: nucleosynthesis. Thus, the simulations cannot directly predict whether the
1727: stars that form during reionization are similar to the stars in the Milky
1728: Way and nearby galaxies or to the primordial $100 M_{\odot}$ stars. They
1729: also cannot determine whether feedback prevents low-mass dark matter halos
1730: from forming stars. Thus, models are needed that make it possible to vary
1731: all these astrophysical parameters of the ionizing sources and to study the
1732: effect on the 21-cm observations.
1733:
1734: \begin{figure}
1735: \epsfxsize=12cm \epsfbox{ZahnFig10.eps}
1736: \caption{Maps of the 21-cm brightness temperature comparing results of a
1737: numerical simulation and of two simpler numerical schemes, at three
1738: different redshifts (from Zahn et al. 2006). Each map is 65.6 Mpc/$h$ on a
1739: side, with a depth (0.25 Mpc/$h$) that is comparable to the frequency
1740: resolution of planned experiments. The ionized fractions are $x_{\rm
1741: i}=0.13$, 0.35 and 0.55 for $z=8.16$, 7.26 and 6.89 (top to bottom),
1742: respectively. All three maps show a very similar large-scale ionization
1743: topology. \emph{Left column:} Numerical simulation, showing the ionized
1744: bubbles (black) produced by the ionizing sources (blue dots) that form in
1745: the simulation. \emph{Middle column:} Numerical scheme that applies an
1746: analytical model to the final distribution of ionizing sources that form in
1747: the simulation. \emph{Right column:} Numerical scheme that applies the
1748: analytical model to the linear density fluctuations that are the initial
1749: conditions of the simulation.}
1750: \label{fig:Zahn}
1751: \end{figure}
1752:
1753: The theoretical expectations presented here for reionization and for the
1754: 21-cm signal are based on rather large extrapolations from observed
1755: galaxies to deduce the properties of much smaller galaxies that formed at
1756: an earlier cosmic epoch. Considerable surprises are thus possible, such as
1757: an early population of quasars or even unstable exotic particles that
1758: emitted ionizing radiation as they decayed. In any case, the forthcoming
1759: observational data in 21-cm cosmology should make the next few years a very
1760: exciting time.
1761:
1762: At high redshifts prior to reionization, spatial perturbations in the
1763: thermodynamic gas properties are linear and can be predicted precisely (see
1764: section~\ref{sec:lin}). Thus, if the gas is probed with the 21-cm technique
1765: then it becomes a promising tool of fundamental, precision cosmology, able
1766: to probe the primordial power spectrum of density fluctuations imprinted in
1767: the very early universe, perhaps in an era of cosmic inflation. The 21-cm
1768: fluctuations can be measured down to the smallest scales where the baryon
1769: pressure suppresses gas fluctuations, while the CMB anisotropies are damped
1770: on small scales (through the so-called Silk damping). This difference in
1771: damping scales can be seen by comparing the baryon-density and
1772: photon-temperature power spectra in Figure~\ref{fig:photons}. Since the
1773: 21-cm technique is also three-dimensional (while the CMB yields a single
1774: sky map), there is a much large potential number of independent modes
1775: probed by the 21-cm signal: $N_{\rm 21-cm}\sim 3 \times 10^{16}$ compared
1776: to $N_{\rm cmb} \sim 2\times 10^7$. This larger number should provide a
1777: measure of non-Gaussian deviations to a level of $\sim N_{\rm 21
1778: cm}^{-1/2}$, constituting a test of the inflationary origin of the
1779: primordial inhomogeneities which are expected to possess non-Gaussian
1780: deviations $\gtrsim 10^{-6}$.
1781:
1782: The 21cm fluctuations are expected to simply trace the primordial
1783: power-spectrum of matter density perturbations (which is shaped by the
1784: initial conditions from inflation and the dark matter) either before the
1785: first population of galaxies had formed (at redshifts $z>25$) or after
1786: reionization ($z<6$) -- when only dense pockets of self-shielded hydrogen
1787: (such as damped Ly$\alpha$ systems) survive. During the epoch of
1788: reionization, the fluctuations are mainly shaped by the topology of ionized
1789: regions, and thus depend on uncertain astrophysical details involving star
1790: formation. However, even during this epoch, the imprint of peculiar
1791: velocities (which are induced gravitationally by density fluctuations), can
1792: in principle be used to separate the implications for fundamental physics
1793: from the astrophysics.
1794:
1795: \begin{figure}
1796: \epsfxsize=10cm \epsfbox{powerZlong_mid.eps}
1797: \caption{Predicted redshift evolution of the angle-averaged amplitude of
1798: the 21-cm power spectrum ($|\bar{\Delta}_{T_b}|=[k^3P_{\rm
1799: 21-cm}(k)/2\pi^2]^{1/2}$) at comoving wavenumbers $k=0.01$ (solid curve),
1800: 0.1 (dotted curve), 1.0 (short dashed curve), 10.0 (long dashed curve), and
1801: 100.0${\rm Mpc}^{-1}$ (dot-dashed curve). In the model shown, reionization
1802: is completed at $z=9.76$. The horizontal axis at the top shows the
1803: observed photon frequency at the different redshifts. The diagonal
1804: straight (red) lines show various factors of suppression for the
1805: synchrotron Galactic foreground, necessary to reveal the 21-cm signal (from
1806: Pritchard \& Loeb 2008).}
1807: \label{fig:Analytic}
1808: \end{figure}
1809:
1810: Peculiar velocities imprint a particular form of anisotropy in
1811: the 21-cm fluctuations that is caused by gas motions along the line of
1812: sight. This anisotropy, expected in any measurement of density that is
1813: based on a spectral resonance or on redshift measurements, results from
1814: velocity compression. Consider a photon traveling along the line of sight
1815: that resonates with absorbing atoms at a particular point. In a uniform,
1816: expanding universe, the absorption optical depth encountered by this photon
1817: probes only a narrow strip of atoms, since the expansion of the universe
1818: makes all other atoms move with a relative velocity that takes them outside
1819: the narrow frequency width of the resonance line. If there is a density
1820: peak, however, near the resonating position, the increased gravity will
1821: reduce the expansion velocities around this point and bring more gas into
1822: the resonating velocity width. This effect is sensitive only to the
1823: line-of-sight component of the velocity gradient of the gas, and thus
1824: causes an observed anisotropy in the power spectrum even when all physical
1825: causes of the fluctuations are statistically isotropic. This anisotropy is
1826: particularly important in the case of 21-cm fluctuations. When all
1827: fluctuations are linear, the 21-cm power spectrum takes the form \beq
1828: P_{\rm 21-cm}({\bf k}) = \mu^4 P_{\rho}(k) + 2 \mu^2 P_{\rho - {\rm iso}}
1829: (k) + P_{\rm iso}\ , \eeq where $\mu = \cos \theta$ in terms of the angle
1830: $\theta$ between the wave-vector ${\bf k}$ of a given Fourier mode and the
1831: line of sight, $P_{\rm iso}$ is the isotropic power spectrum that would
1832: result from all sources of 21-cm fluctuations without velocity compression,
1833: $P_{\rho}(k)$ is the 21-cm power spectrum from gas density fluctuations
1834: alone, and $P_{\rho - {\rm iso}} (k)$ is the Fourier transform of the
1835: cross-correlation between the density and all sources of 21-cm
1836: fluctuations. The three power spectra can also be denoted $P_{\mu^4}(k)$,
1837: $P_{\mu^2}(k)$, and $P_{\mu^0}(k)$, according to the power of $\mu$ that
1838: multiplies each term. The prediction for these power spectra at high
1839: redshift ($z > 20$), neglecting the effects of any stellar radiation, are
1840: shown in Figure~\ref{21-cmP}. At these redshifts, the 21-cm fluctuations
1841: probe the infall of the baryons into the dark matter potential wells. The
1842: power spectrum shows remnants of the photon-baryon acoustic oscillations on
1843: large scales, and of the baryon pressure suppression on small scales.
1844:
1845: \begin{figure}
1846: \epsfxsize=10cm \epsfbox{binfallA8a.eps}
1847: \caption{Power spectra of 21-cm brightness fluctuations versus comoving
1848: wavenumber (from Barkana \& Loeb 2005c). Show are the three power spectra
1849: that are separately observable, $P_{\mu^4}$ (upper panel), $P_{\mu^2}$
1850: (middle panel), and $P_{\mu^0}$ (lower panel). Each case shows redshifts
1851: 200, 150, 100, 50 (solid curves, from bottom to top), 35, 25, and 20
1852: (dashed curves, from top to bottom).}
1853: \label{21-cmP}
1854: \end{figure}
1855:
1856: Once stellar radiation becomes significant, many processes can contribute
1857: to the 21-cm fluctuations. The contributions include fluctuations in gas
1858: density, temperature, ionized fraction, and \Lya flux. These processes can
1859: be divided into two broad categories: The first, related to {\it
1860: ``physics''}, consists of probes of fundamental, precision cosmology, and
1861: the second, related to {\it ``astrophysics''}, consists of probes of
1862: stars. Both categories are interesting -- the first for precision measures
1863: of cosmological parameters and studies of processes in the early universe,
1864: and the second for studies of the properties of the first
1865: galaxies. However, the astrophysics depends on complex non-linear processes
1866: (collapse of dark matter halos, star formation, supernova feedback), and
1867: must be cleanly separated from the physics contribution, in order to allow
1868: precision measurements of the latter. As long as all the fluctuations are
1869: linear, the anisotropy noted above allows precisely this separation of the
1870: {\it fundamental physics} from the {\it astrophysics} of the 21-cm
1871: fluctuations. In particular, the $P_{\mu^4}(k)$ is independent of the
1872: effects of stellar radiation, and is a clean probe of the gas density
1873: fluctuations. Once non-linear terms become important, there arises a
1874: significant mixing of the different terms; in particular, this occurs on
1875: the scale of the ionizing bubbles during reionization.
1876:
1877: The 21-cm fluctuations are affected by fluctuations in the Lyman-$\alpha$
1878: flux from stars, a result that yields an indirect method to detect and
1879: study the early population of galaxies at $z \sim 20$. The fluctuations are
1880: caused by biased inhomogeneities in the density of galaxies, along with
1881: Poisson fluctuations in the number of galaxies. Observing the power-spectra
1882: of these two sources would probe the number density of the earliest
1883: galaxies and the typical mass of their host dark matter halos. Furthermore,
1884: the enhanced amplitude of the 21-cm fluctuations from the era of \Lya
1885: coupling improves considerably the practical prospects for their
1886: detection. Precise predictions account for the detailed properties of all
1887: possible cascades of a hydrogen atom after it absorbs a photon. Around the
1888: same time, X-rays may also start to heat the cosmic gas, producing strong
1889: 21-cm fluctuations due to fluctuations in the X-ray flux.
1890:
1891: \section{Conclusions}
1892:
1893: The initial conditions of our Universe can be summarized on a single sheet
1894: of paper. Yet the Universe is full of complex structures today, such as
1895: stars, galaxies and groups of galaxies. This review discussed the standard
1896: theoretical model for how complexity emerged from the simple initial state
1897: of the Universe through the action of gravity. In order to test and inform
1898: the related theoretical calculations, large-aperture telescopes and arrays
1899: of radio antennae are currently being designed and constructed.
1900:
1901: The actual transition from simplicity to complexity has not been observed
1902: as of yet. The simple initial conditions were already traced in maps of
1903: the microwave background radiation, but the challenge of detecting the
1904: first generation of galaxies defines one of the exciting frontiers in the
1905: future of cosmology. Once at hand, the missing images of the infant
1906: Universe might potentially surprise us and revise our current ideas.
1907:
1908: \subsection*{Acknowledgements}
1909:
1910: I thank my collaborators on the topics covered by this review: Dan Babich,
1911: Rennan Barkana, Volker Bromm, Steve Furlanetto, Zoltan Haiman, Jonathan
1912: Pritchard, Stuart Wyithe, and Matias Zaldarriaga.
1913:
1914: \noindent
1915:
1916: \subsection*{Bibliography}
1917:
1918: \Ref {Abel T.L., Bryan G.L., Norman M.L. (2002). The Formation of the First Star in the Universe. {\it Science} \textbf{295} 93. [Results from
1919: simulations of the formation of the first stars]. }
1920:
1921: \Ref {Allison A.C. and Dalgarno A. (1969). Spin Change in Collisions of Hydrogen Atoms. {\it Astrophys. J.} \textbf{158} 423. [An early paper
1922: about the effect of collisions on the spin temperature of hydrogen].}
1923:
1924: \Ref {Arons J. and Wingert D.W. (1972). Theoretical Models of Photoionized Intergalactic Hydrogen. {\it Astrophys. J.} \textbf{177} 1. [An early discussion
1925: on the re-ionization of the intergalactic hydrogen, more
1926: than a quarter of a century before th topic gained popularity]. }
1927:
1928: \Ref {Babich D. and Loeb A. (2006). Imprint of Inhomogeneous Reionization on the Power Spectrum of Galaxy Surveys at High Redshifts. {\it Astrophys. J.} \textbf{640} 1. [A discussion on the imprint of cosmic reionization
1929: on the distribution of low-mass galaxies].}
1930:
1931: \Ref {Barkana R. and Loeb A. (2001). In the beginning: the first sources of light and the reionization of the universe. {\it Phys. Rep.} \textbf{349} 125.
1932: [An overview on the physics of the first galaxies and reionization].}
1933:
1934: \Ref { Barkana R. and Loeb A. (2004a). Gamma-Ray Bursts versus Quasars: Ly$\alpha$ Signatures of Reionization versus Cosmological Infall. {\it Astrophys. J.} \textbf{601} 64. [A comparison between quasars and gamma-ray bursts
1935: as probes of the intergalactic medium at high redshifts].}
1936:
1937: \Ref { Barkana R. and Loeb A. (2004b). Unusually Large Fluctuations in the
1938: Statistics of Galaxy Formation at High Redshift. {\it Astrophys. J.}
1939: \textbf{609} 474. [A discussion on the limitations of computer simulations
1940: of reionization owing to their finite size of the simulated region]. }
1941:
1942: \Ref { Barkana R. and Loeb A. (2005a). A Method for Separating the Physics
1943: from the Astrophysics of High-Redshift 21 Centimeter Fluctuations. {\it
1944: Astrophys. J. Lett.} \textbf{624} L65. [The imprint of peculiar velocities
1945: on the 21cm brightness fluctuations can be used to extract cosmological
1946: information during the epoch of reionization]. }
1947:
1948: \Ref { Barkana R. and Loeb A. (2005b). Detecting the Earliest Galaxies through Two New Sources of 21 Centimeter Fluctuations. {\it Astrophys. J.}
1949: \textbf{626} 1. [The imprint of the first galaxies on the 21cm
1950: fluctuations. }
1951:
1952: \Ref { Barkana R. and Loeb A. (2005c). Probing the epoch of early baryonic infall through 21-cm fluctuations. {\it Mon. Not. Roy. Astron. Soc.
1953: Lett.} \textbf{363} L36. [A discussion on the signature of acoustic
1954: oscillations on the 21cm brightness fluctuations]. }
1955:
1956: \Ref { Barkana R. and Loeb A. (2007). The physics and early history of the
1957: intergalactic medium. {\it Rep. Prog. Phys.} \textbf{70} 627.
1958: [A review on the history of the intergalactic medium].}
1959:
1960: \Ref { Bennett C.L. et al. (1996). Four-Year COBE DMR Cosmic Microwave
1961: Background Observations: Maps and Basic Results. {\it Astrophys. J. Lett.}
1962: \textbf{464} L1. [A description of the first robust detection of microwave
1963: background fluctuations by the COBE satellite (for which the Nobel prize
1964: was awarded in 2006].}
1965:
1966: \Ref { Bharadwaj S. and Ali S.S. (2004). The cosmic microwave background radiation fluctuations from HI perturbations prior to reionization. {\it Mon. Not. Roy. Astron. Soc.} \textbf{352} 142. [A discussion on the imprint of peculiar
1967: velocities on 21cm fluctuations].}
1968:
1969: \Ref { Bowman J.D., Morales M.F. and Hewitt J.N. (2006). The Sensitivity of First-Generation Epoch of Reionization Observatories and Their Potential for Differentiating Theoretical Power Spectra. {\it Astrophys. J.} \textbf{638} 20.
1970: [An early discussion on the feasibility of modern measurements of the 21cm fluctuations from the epoch of reionization].}
1971:
1972: \Ref { Bromm V., Coppi P.S., Larson R.B. (2002). The Formation of the First Stars. I. The Primordial Star-forming Cloud. {\it Astrophys. J.} \textbf{564} 23.
1973: [Results from numerical simulations of the formation of the first stars].}
1974:
1975: \Ref { Bromm V. and Larson R.B. (2004). The First Stars. {\it Ann. Rev. Astron. \& Astrophys.} \textbf{42} 79. [An overview on the formation of the
1976: the first stars].}
1977:
1978: \Ref { Bromm V. and Loeb A. (2003). Formation of the First Supermassive
1979: Black Holes. {\it Astrophys. J.} \textbf{596} 34. [An early model for the
1980: production of massive seeds for quasar black holes at early cosmic times].}
1981:
1982: \Ref { Bromm V. and Loeb A. (2004). Accretion onto a primordial
1983: protostar. {\it New Astronomy} \textbf{9} 353. [Results from high-resultion
1984: simulations of the first stars, including an estimate of their final
1985: mass].}
1986:
1987: \Ref { Bromm V., Kudritzki R.P. and Loeb A. (2001). Generic Spectrum and
1988: Ionization Efficiency of a Heavy Initial Mass Function for the First Stars.
1989: {\it Astrophys. J.} \textbf{552} 464. [A pioneering derivation of the
1990: spectrum of the first stars, and the number of ionizing photons they
1991: produce per stellar mass].}
1992:
1993:
1994:
1995: \Ref { Couchman, H.M.P. and Rees M.J. (1986). Pregalactic evolution in
1996: cosmologies with cold dark matter. {\it Mon. Not. Roy. Astron. Soc.}
1997: \textbf{221} 53. [A pioneering paper on the formation of the first
1998: galaxies in a CDM cosmology].}
1999:
2000: \Ref { Chen X. and Miralda-Escud\'e J. (2004). The Spin-Kinetic Temperature
2001: Coupling and the Heating Rate due to Ly$\alpha$ Scattering before
2002: Reionization: Predictions for 21 Centimeter Emission and Absorption. {\it
2003: Astrophys. J.} \textbf{602} 1. [A detailed calculation of the effect of
2004: Lyman-$\alpha$ photons on the spin temperature of intergalactic hydrogen].}
2005:
2006: \Ref { Ciardi B., Ferrara A. and White S.D.M. (2003). Early reionization by
2007: the first galaxies. {\it Mon. Not. Roy. Astron. Soc.} \textbf{344} L7.
2008: [Results from simulations of reionization].}
2009:
2010: \Ref { Ciardi B. and Loeb A. (2000). Expected Number and Flux Distribution
2011: of Gamma-Ray Burst Afterglows with High Redshifts. {\it Astrophys. J.}
2012: \textbf{540} 687. [An early calculation of the rate of gamma-ray bursts
2013: eith high-redshitfs].}
2014:
2015: \Ref { Cole S. et al. (2005). The 2dF Galaxy Redshift Survey:
2016: power-spectrum analysis of the final data set and cosmological
2017: implications. {\it Mon. Not. R. Astron. Soc.} \textbf{362} 505. [Recent
2018: data on the distribution of galaxies on large spatial scales].}
2019:
2020: \Ref { Dijkstra M., Haiman Z., Rees M.J. and Weinberg
2021: D.H. (2004). Photoionization Feedback in Low-Mass Galaxies at High
2022: Redshift. {\it Astrophys. J.} \textbf{601} 666. [A recent discussion on the
2023: suppression of low-mass galaxies after reionization].}
2024:
2025: \Ref { Di Matteo T., Perna R., Abel T. and Rees M.J. (2002). Radio
2026: Foregrounds for the 21 Centimeter Tomography of the Neutral Intergalactic
2027: Medium at High Redshifts. {\it Astrophys. J.} \textbf{564} 576. [A
2028: discussion on the contaminating noise for future 21cm observations].}
2029:
2030: \Ref {Di Matteo T., Springel V., \& Hernquist, L. (2005). Energy input from
2031: quasars regulates the growth and activity of black holes and their host
2032: galaxies. {\it Nature}, \textbf{433}, 604. [Simulations of quasar feedback
2033: on galaxy formation and evolution].}
2034:
2035: \Ref { Eisenstein D.J. et al. (2005). Detection of the Baryon Acoustic Peak
2036: in the Large-Scale Correlation Function of SDSS Luminous Red Galaxies. {\it
2037: Astrophys. J.} \textbf{633} 560. [The first detection of baryonic
2038: oscillations in the distribution of galaxies].}
2039:
2040: \Ref { Efstathiou G. (1992). Suppressing the formation of dwarf galaxies via photoionization. {\it Mon. Not. Roy Astron. Soc.} \textbf{256} 43.
2041: [An early discussion on the suppression of dwarf galaxies
2042: by ionizing radiation].}
2043:
2044: \Ref { Ellis R. (2008). Observations of the High Redshift Universe. {\it
2045: SAAS-Fee Advanced Course 36}, Springer Verlag, Berlin
2046: 2008. http://arxiv.org/abs/astro-ph/0701024. [A recent overview of the
2047: status of observations of high-redshift galaxies].}
2048:
2049: \Ref { Fan X. et al. (2002). Evolution of the Ionizing Background and the
2050: Epoch of Reionization from the Spectra of $ z \sim 6$ Quasars. {\it
2051: Astron. J.} \textbf{123} 1247. [Observational constraints on intergalactic
2052: hydrogen from the spectra of quasars which formed a billion years after the
2053: big bang].}
2054:
2055: \Ref { Fan X. et al. (2003). A Survey of $z>5.7$ Quasars in the Sloan
2056: Digital Sky Survey. II. Discovery of Three Additional Quasars at $
2057: z>6$. {\it Astron. J.} \textbf{125} 1649. [Observations of high redshift
2058: quasars].}
2059:
2060: \Ref { Fan X. et al. (2005). Constraining the Evolution of the Ionizing
2061: Background and the Epoch of Reionization with $ z \sim 6$ Quasars II: A
2062: Sample of 19 Quasars. {\it Astron. J.} \textbf{132} (2006) 117. [A
2063: description od the sample of the highest-redshift quasars].}
2064:
2065: \Ref { Fan X., Carilli C.L. and Keating B. (2006). Observational Constraints on Cosmic Reionization. {\it Ann. Rev. Astron. \& Astrophys.} \textbf{44} 415.
2066: [A review on the constraints drawn from quasar spectra about reionization].}
2067:
2068: \Ref { Field G.B. (1958). Excitation of the Hydrogen 21 cm Line. {\it
2069: Proc. IRE} \textbf{46} 240. [A classic paper on the physics of the
2070: 21 cm line of intergalactic hydrogen].}
2071:
2072: \Ref { Field G.B. (1959). The Time Relaxation of a Resonance-Line
2073: Profile. {\it Astrophys. J.} \textbf{129} 551. [A pioneering paper on the
2074: physics of the 21 cm line from intergalactic hydrogen].}
2075:
2076: \Ref { Fukugita M. and Kawasaki M. (1994). Reionization during Hierarchical
2077: Clustering in a Universe Dominated by Cold Dark Matter. {\it
2078: Mon. Not. Roy. Astron. Soc.} \textbf{269} 563. [An early discussion on
2079: cosmic reionization, about a decade before the topic gained popularity].}
2080:
2081: \Ref { Furlanetto S.R. and Loeb A. (2003). Metal Absorption Lines as Probes of the Intergalactic Medium Prior to the Reionization Epoch. {\it Astrophys. J.}\textbf{588} 18. [A discussion on the detectability of absorption
2082: lines from heavy elements which were produced in stellar interiors
2083: and then dispersed into intergalactic space]. }
2084:
2085: \Ref { Furlanetto S.R., Zaldarriaga M. and Hernquist L. (2004). The Growth
2086: of H II Regions During Reionization. {\it Astrophys. J.} \textbf{613} 1.
2087: [A calculation of the size distribution of ionized bubbles during the epoch
2088: of reionization].}
2089:
2090: \Ref { Furlanetto S.R., Oh S.P. and Briggs F. (2006). Cosmology at low frequencies: The 21 cm transition and the high-redshift Universe. {\it Phys. Rep.} \textbf{433} 181. [An overview on 21cm cosmology].}
2091:
2092: \Ref { Gehrels N. et al. (2004). The Swift Gamma-Ray Burst Mission. {\it
2093: Astrophys. J.} \textbf{611} 1005. [A description of the SWIFT satellite
2094: that is currently detecting gamma-ray bursts and their afterglows].}
2095:
2096: \Ref { Gnedin N.Y. and Ostriker J.P. (1997). Reionization of the Universe
2097: and the Early Production of Metals. {\it Astrophys. J.} \textbf{486} 581.
2098: [An early numerical simulation of the production of heavy elements during
2099: the epoch of reionization]. }
2100:
2101: \Ref { Gnedin N.Y. and Hui L. (1998). Probing the Universe with the
2102: Lyman-alpha forest - I. Hydrodynamics of the low-density intergalactic
2103: medium. {\it Mon. Not. Roy Astron. Soc.} \textbf{296} 44. [A simple model
2104: for the Lyman-$\alpha$ forest in quasar spectra].}
2105:
2106: \Ref { Gnedin N.Y. (2000). Effect of Reionization on Structure Formation in
2107: the Universe. {\it Astrophys. J.} \textbf{542} 535. [Results from early
2108: simulations of the effect of reionization on the assembly
2109: of gas into low-mass galaxies].}
2110:
2111: \Ref { Goodman J. (1995). Geocentrism reexamined. {\it Phys. Rev. D}
2112: \textbf{52} 1821. [A discussion on existing evidence for the homogeneity of
2113: the Universe].}
2114:
2115: \Ref { Gunn J.E. and Peterson B.A. (1965). On the Density of Neutral
2116: Hydrogen in Intergalactic Space. {\it Astrophys. J.} \textbf{142} 1633.
2117: [A seminal paper on the Lyman-$\alpha$ absorption feature of intergalactic
2118: hydrogen]. }
2119:
2120: \Ref { Haiman Z., Thoul A.A. and Loeb A. (1996). Cosmological Formation of
2121: Low-Mass Objects. {\it Astrophys. J.} \textbf{464} 52. [The first
2122: (spherically-symmetric) simulation of the formation of the first gas-rich
2123: galaxies].}
2124:
2125: \Ref { Haiman Z. and Loeb A. (1997). Signatures of Stellar Reionization of
2126: the Universe. {\it Astrophys. J.} \textbf{483} 21. [An early detailed
2127: calculation of reionization by stars in the modern context of cosmological
2128: structure formation].}
2129:
2130: \Ref { Haiman Z., Rees M.J., Loeb A. (1997). Destruction of Molecular
2131: Hydrogen during Cosmological Reionization. {\it Astrophys. J.} \textbf{476}
2132: 458; erratum -- {\it Astrophys. J.} \textbf{484} 985. [Negative feedback of
2133: UV photons on the production of molecular hydrogen in the first
2134: galaxies]. }
2135:
2136: \Ref { Haislip J. et al. (2006). A photometric redshift of $z = 6.39 \pm
2137: 0.12$ for GRB 050904. {\it Nature} \textbf{440} 181. [The discovery of a
2138: gamma-ray burst with the highest redshift known].}
2139:
2140: \Ref { Hirata C.M. (2006). Wouthuysen-Field coupling strength and
2141: application to high-redshift 21-cm radiation. {\it
2142: Mon. Not. Roy. Astron. Soc.} \textbf{367} 259. [A detailed discussion on
2143: the coupling between the spin temperature and the kinetic temperature of
2144: hydrogen through its interaction with Lyman-$\alpha$ photons]. }
2145:
2146: \Ref { Hogan C.J. and Rees M.J. (1979). Spectral appearance of non-uniform gas at high Z. {\it Mon. Not. Roy. Astron. Soc.} \textbf{188} 791.
2147: [A pioneering discussion on the use of resonant lines to probe the
2148: intergalactic gas and study cosmology].}
2149:
2150: \Ref { Hu E.M., Cowie L.L., McMahon R.G., Capak P., Iwamuro F., Kneib J.P.,
2151: Maihara T. and Motohara K. (2002). A Redshift z=6.56 Galaxy behind the
2152: Cluster Abell 370. {\it Astrophys. J. Lett.} {\textbf 568} L75. [A
2153: spectroscopic detection of one of the earliest galaxies known]. }
2154:
2155: \Ref { Iye M. et al. (2006). A galaxy at a redshift z = 6.96. {\it Nature}
2156: \textbf{443} 186. [A spectroscopic detection of one of the earliest
2157: galaxies known]. }
2158:
2159: \Ref { Kaiser N. (1984). On the spatial correlations of Abell
2160: clusters. {\it Astrophys. J. Lett.} \textbf{284} L9. [A pioneering
2161: discussion on the concept of bias in the clustering statistics of
2162: cosmological objects].}
2163:
2164: \Ref { Kaiser N. (1987). Clustering in real space and in redshift
2165: space. {\it Mon. Not. Roy. Astron. Soc.} \textbf{227} 1. [A pioneering
2166: discussion on the effect of peculiar velocities on the clustering of
2167: sources in redshift surveys].}
2168:
2169: \Ref { Kamionkowski M., Spergel D.N. and Sugiyama N. (1994). Small-scale
2170: cosmic microwave background anisotropies as probe of the geometry of the
2171: universe. {\it Astrophys. J. Lett.} \textbf{426} L57. [An early discussion
2172: on the use of microwave background data to constrain the underlying
2173: geometry of the Universe].}
2174:
2175: \Ref { Kitayama T. and Ikeuchi S. (2000). Formation of Subgalactic Clouds
2176: under Ultraviolet Background Radiation. {\it Astrophys. J.} \textbf{529}
2177: 615. [Spherically-symmetric simulations of the suppressing effect of UV
2178: radiation on the collapse low-mass gas coulds].}
2179:
2180: \Ref { Kolb E.W. and Turner M.S. (1990). The early universe. (Redwood City,
2181: CA: Addison-Wesley). [A textbook on the interface
2182: between modern cosmology and particle physics].}
2183:
2184: \Ref { Komatsu E. et al. (2008). Five-Year Wilkinson Microwave Anisotropy
2185: Probe (WMAP) Observations: Cosmological Interpretation, ArXiv e-prints,
2186: 803, arXiv:0803.0547. [The latest cosmological constraints based
2187: on five-years of data gathering by the WMAP satellite].}
2188:
2189: \Ref { Lamb D.Q and Reichart D.E. (2000). Gamma-Ray Bursts as a Probe of
2190: the Very High Redshift Universe. {\it Astrophys. J.} \textbf{536} 1. [An
2191: early discussion on the detectability of gamma-ray bursts out to very high
2192: redshifts].}
2193:
2194: \Ref { Lidz A., Oh S.P. and Furlanetto S.R. (2006). Have We Detected Patchy
2195: Reionization in Quasar Spectra? {\it Astrophys. J. Lett.} \textbf{639}
2196: L47. [An analysis of the implications from absorption spectra of
2197: high-redshift quasar for models of reionization]. }
2198:
2199: \Ref { Loeb A. (2008). First Light. {\it SAAS-Fee Advanced Course 36},
2200: Springer Verlag, Berlin 2008. http://arxiv.org/abs/astro-ph/0701024. [A
2201: recent overview of the underlying physics in studies of the first
2202: galaxies]. }
2203:
2204: \Ref { Loeb A. (2006). The dark ages of the Universe. {\it Scientific American}, \textbf{295}, 46 (http://www.cfa.harvard.edu/~loeb/sciam.pdf).
2205: [A popular level review on the first galaxies and 21-cm cosmology].}
2206:
2207: \Ref { Loeb A. and Rybicki G. (1999). Scattered Lyman-$\alpha$ Radiation
2208: around Sources before Cosmological Reionization {\it Astrophys. J.}
2209: \textbf{524}, 527. [A derivation of the halo of scattered
2210: Lyman-$\alpha$ photons around a source embedded in an expanding
2211: intergalactic medium].}
2212:
2213: \Ref { Loeb A. and Zaldarriaga M. (2004). Measuring the Small-Scale Power
2214: Spectrum of Cosmic Density Fluctuations through 21cm Tomography Prior to
2215: the Epoch of Structure Formation. {\it Phys. Rev. Lett.} \textbf{92}
2216: 211301. [The first calculation of the power-spectrum of 21-cm brightness
2217: fluctuations during the dark ages [prior to the appearance of the first
2218: galaxies)].}
2219:
2220: \Ref { Loeb A. and Wyithe S. (2008). Precise Measurement of the
2221: Cosmological Power Spectrum With a Dedicated 21cm Survey After
2222: Reionization. {\it Phys. Rev. Lett.} \textbf{in press}, ArXiv e-prints,
2223: 801, arXiv:0801.1677. [A study demonstrating that future 21cm data
2224: after reionization can map the matter distribution through most of the
2225: observable volume of the Universe].}
2226:
2227: \Ref { Ma C. and Bertschinger E. (1995). Cosmological Perturbation Theory in the Synchronous and Conformal Newtonian Gauges. {\it Astrophys. J.} \textbf{455} 7. [A comprehensive discussion on the growth of structure
2228: in the Universe].}
2229:
2230: \Ref { Madau P., Meiksin A. and Rees M.J. (1997). 21 Centimeter Tomography
2231: of the Intergalactic Medium at High Redshift. {\it Astrophys. J.}
2232: \textbf{475} 429. [An early discussion on the use of the 21cm line for
2233: three-dimensional mapping of intergalactic hydrogen].}
2234:
2235: \Ref { McQuinn M., Zahn O., Zaldarriaga M., Hernquist L. and Furlanetto
2236: S.R. (2006). Cosmological Parameter Estimation Using 21 cm Radiation from
2237: the Epoch of Reionization. {\it Astrophys. J.} \textbf{653} 815. [A
2238: demonstration of the power of statistical analysis of future 21cm data for
2239: constraining cosmological parameters].}
2240:
2241: \Ref { Mellema G., Iliev I.T., Pen U.L. and Shapiro P.R. (2006). Simulating
2242: Cosmic Reionization at Large Scales II: the 21-cm Emission Features and
2243: Statistical Signals. {\it Mon. Not. Roy. Astron. Soc.} \textbf{372} 679.
2244: [Results from a numerical simulation of reionization by the
2245: first galaxies].}
2246:
2247: %\Ref { Mesinger A. and Haiman Z. (2004). Evidence of a Cosmological
2248: %Str\"{o}mgren Surface and of Significant Neutral Hydrogen Surrounding the
2249: %Quasar SDSS J1030+0524. {\it Astrophys. J. Lett.} \textbf{611} 69. }
2250:
2251: \Ref { Miralda-Escud\'e J. (1998). Reionization of the Intergalactic Medium
2252: and the Damping Wing of the Gunn-Peterson Trough. {\it Astrophys. J.}
2253: \textbf{501} 15. [A derivation of the spectral profile of Lyman-$\alpha$
2254: absorption by a neutral intergalactic medium around a high-redshift
2255: source].}
2256:
2257: \Ref { Miralda-Escud\'e J. and Rees M J (1998). Searching for the Earliest
2258: Galaxies Using the Gunn-Peterson Trough and the Lyman-alpha Emission
2259: Line. {\it Astrophys. J.} \textbf{497} 21. [An early discussion on the
2260: spectral signatures of high-redshift galaxies].}
2261:
2262: \Ref { Miralda-Escud\'e J. (2000). Soft X-Ray Absorption by High-Redshift
2263: Intergalactic Helium. {\it Astrophys. J. Lett.} \textbf{528} L1. [A
2264: discussion on the absorption signature of a neutral intergalactic medium
2265: around an X-ray source]. }
2266:
2267: \Ref { Murray N., Quataert E. and Thompson T.A. (2005). On the Maximum
2268: Luminosity of Galaxies and Their Central Black Holes: Feedback from
2269: Momentum-driven Winds. {\it Astrophys. J.} \textbf{618} 569. [A model for
2270: momentum-regulated growth of supermassive black holes in galaxies].}
2271:
2272: \Ref { Naoz S. and Barkana R. (2005). Growth of linear perturbations before
2273: the era of the first galaxies. {\it Mon. Not. Roy. Astron. Soc.}
2274: \textbf{362} 1047. [A precise calculation of the linear evolution of
2275: density and temperature fluctuations in the cosmic gas during the dark
2276: ages].}
2277:
2278: \Ref{ Navarro J. F., Frenk C. S., White S. D. M. (1997). A Universal
2279: Density Profile from Hierarchical Clustering. {\it Astrophysical Journal}
2280: 490, 493. [Results from numerical simulations that demonstrated the
2281: existence of a universal form for the density profile in dark matter
2282: halos]. }
2283:
2284: \Ref { Navarro J.F. and Steinmetz M. (1997). The Effects of a Photoionizing
2285: Ultraviolet Background on the Formation of Disk Galaxies. {\it
2286: Astrophys. J.} \textbf{478} 13. [Results from three-dimensional
2287: simulations on the effect of UV radiation on the
2288: assembly of gas in low-mass galaxies].}
2289:
2290: \Ref { Oh S. P. (2001). Reionization By Hard Photons. I. X-Rays From the
2291: First Star Clusters. {\it Astrophys. J.} \textbf{553} 499.
2292: [A discussion on reionization by X-ray photons].}
2293:
2294: \Ref { Osterbrock D.E. (1974). Astrophysics of gaseous nebulae. (San Francisco: W. H. Freeman and Company) p.14. [A textbook describing the physics
2295: of ionized regions around a UV source].}
2296:
2297: \Ref { Peebles P.J.E. (1980). The large-scale structure of the universe.
2298: (Princeton: Princeton University Press). [A textbook describing basic
2299: concepts related to the growth of structure in the Universe].}
2300:
2301: \Ref { Peebles P.J.E. (1984). Dark matter and the origin of galaxies and globular star clusters. {\it Astrophys. J.} \textbf{277} 470.
2302: [A pioneering discussion on cold dark matter in galaxies].}
2303:
2304: \Ref { Peebles P.J.E. (1993). Principles of physical
2305: cosmology. (Princeton: Princeton University Press). [A textbook on basic
2306: concepts in the physics of cosmology].}
2307:
2308: \Ref { Peebles P.J.E. and Yu J.T. (1970). Primeval Adiabatic Perturbation
2309: in an Expanding Universe. {\it Astrophys. J.} \textbf{162} 815. [A
2310: pioneering calculation of the temperature anisotropies of the microwave
2311: background].}
2312:
2313: \Ref { Pritchard J.R. and Furlanetto S.R. (2006) Descending from on high:
2314: Lyman-series cascades and spin-kinetic temperature coupling in the 21-cm
2315: line. {\it Mon. Not. Roy. Astron. Soc.} \textbf{367} 1057. [A
2316: comprehensive discussion on the effect of Lyman-series photons on the spin
2317: temperature of hydrogen and the corresponding 21-cm fluctuations].}
2318:
2319: \Ref { Pritchard J.R. and Furlanetto S.R. (2007). 21 cm fluctuations from
2320: inhomogeneous X-ray heating before reionization. {\it
2321: Mon. Not. Roy. Astron. Soc.} \textbf{376} 1680. [A study of the effect of
2322: inhomogeneous X-ray heating by the first galaxies on 21-cm fluctuations]. }
2323:
2324: \Ref {Pritchard J.R. and Loeb A. (2008). Evolution of the 21 cm signal
2325: throughout cosmic history. ArXiv e-prints, 802, arXiv:0802.2102. [A
2326: comprehensive summary of the various sources of 21-cm fluctuations at all
2327: redshifts].}
2328:
2329: \Ref { Purcell E.M. and Field G.B. (1956). Influence of Collisions upon
2330: Population of Hyperfine States in Hydrogen. {\it Astrophys. J.}
2331: \textbf{124} 542. [A pioneering paper on the effect of atomic collisions on
2332: the spin temperature of hydrogen].}
2333:
2334: \Ref { Quinn T., Katz N. and Efstathiou G. (1996). Photoionization and the
2335: formation of dwarf galaxies. {\it Mon. Not. Roy Astron. Soc. Lett.}
2336: \textbf{278} 49. [Results from simulations concerning the suppressing
2337: effect of ionizing radiation on the assembly of gas in low-mass galaxies].}
2338:
2339: %\Ref { Rees M.J. (1986). Baryon concentration in string wakes at Z greater
2340: %than about 200 - Implications for galaxy formation and large-scale
2341: %structure. {\it Mon. Not. Roy. Astron. Soc.} \textbf{222} 27. }
2342:
2343: \Ref { Rees M.J. and Sciama D.W. (1968). Larger scale Density
2344: Inhomogeneities in the Universe. {\it Nature} \textbf{217} 511. [A
2345: pioneering discussion on the imprint of large scale inhomogeneities on
2346: temperature anisotropies of the microwave background through the
2347: time-dependence of the gravitational potential].}
2348:
2349: \Ref { Sachs R.K. and Wolfe A.M. (1967). Perturbations of a Cosmological
2350: Model and Angular Variations of the Microwave Background. {\it
2351: Astrophys. J.} \textbf{147} 73. [A pioneering formal derivation of the
2352: temperature anisotropies in the microwave background owing to density
2353: fluctuations].}
2354:
2355: \Ref { Scott D. and Rees M.J. (1990). The 21-cm line at high redshift: a
2356: diagnostic for the origin of large scale structure. {\it
2357: Mon. Not. Roy. Astron. Soc.} \textbf{247} 510. [An early discussion on the
2358: potential use of the 21-cm line for cosmological studies].}
2359:
2360: \Ref { Seljak U. and Zaldarriaga M. (1996). A Line-of-Sight Integration
2361: Approach to Cosmic Microwave Background Anisotropies. {\it Astrophys. J.}
2362: \textbf{469} 437. [An efficient simplified solution to the equations that
2363: provide the microwave background anisotropies].}
2364:
2365: \Ref { Shapiro P.R. and Giroux M.L. (1987). Cosmological H II regions and
2366: the photoionization of the intergalactic medium. {\it Astrophys. J. Lett.}
2367: \textbf{321} L107. [An early discussion on the evolution of ionized regions
2368: around a UV source embedded within an expanding medium of cosmic gas].}
2369:
2370: \Ref { Shapiro P.R., Giroux M.L. and Babul A. (1994). Reionization in a
2371: cold dark matter universe: The feedback of galaxy formation on the
2372: intergalactic medium. {\it Astrophys. J.} \textbf{427} 25. [An early
2373: discussion on reionization by galaxies].}
2374:
2375: \Ref { Silk J. (1968). Cosmic Black-Body Radiation and Galaxy
2376: Formation. {\it Astrophys. J.} \textbf{151} 459. [A pioneering
2377: derivation of the effect of photon diffusion on the damping
2378: of microwave background anisotropies on small scales].}
2379:
2380: \Ref { Silk J. and Rees M.J. (1998). Quasars and Galaxy Formation. {\it
2381: Astron. \& Astrophys.} \textbf{331} L1. [A schematic discussion on the
2382: expected scaling relations between central black hole mass and
2383: velocity dispersion in galaxies, based on self-regulated growth by
2384: momentum or energy feedback].}
2385:
2386: %\Ref { Spergel D.N. et al. (2006). Wilkinson Microwave Anisotropy Probe (WMAP) Three Year Results: Implications for Cosmology. {\it Astrophys. J. Supp.} \textbf{170} 377. }
2387:
2388: \Ref { Stark D.P., Loeb A. and Ellis R. (2007). An Empirically Calibrated
2389: Model for Interpreting the Evolution of Galaxies during the Reionization
2390: Era. {\it Astrophys. J.} \textbf{668} 627. [A simple theoretical model
2391: for observations of high redshift galaxies].}
2392:
2393: \Ref { Sunyaev R.A. and Zeldovich Y.B. (1970). Small-Scale Fluctuations of
2394: Relic Radiation. {\it APSS} \textbf{7} 3. [A pioneering derivation of the
2395: acoustic oscillation signature in the anisotropies of the microwave
2396: background].}
2397:
2398: \Ref { Tegmark M. et al. (1997). How Small Were the First Cosmological
2399: Objects? {\it Astrophys. J.} \textbf{474} 1. [An early discussion
2400: on the conditions for the formation of the first galaxies]. }
2401:
2402: %\Ref { Tegmark M., Silk J. and Blanchard A. (1994). On the inevitability of
2403: %reionization: Implications for cosmic microwave background
2404: %fluctuations. {\it Astrophys. J.} \textbf{420} 484. [A discussion on the
2405: %imprint of the free electrons produced during reionization on the microwave
2406: %background].}
2407:
2408: \Ref { Thoul A.A. and Weinberg D.H. (1996). Hydrodynamic Simulations of
2409: Galaxy Formation. II. Photoionization and the Formation of Low-Mass
2410: Galaxies. {\it Astrophys. J.} \textbf{465} 608. [Results
2411: from a spherically-symmetric simulation of the suppressed collapse
2412: of gas clouds under the influence of a UV radiation background].}
2413:
2414: \Ref { Totani T., Kawai N., Kosugi G., Aoki K., Yamada T., Iye M., Ohta
2415: K. and Hattori T. (2006). Implications for Cosmic Reionization from the
2416: Optical Afterglow Spectrum of the Gamma-Ray Burst 050904 at $z = 6.3$. {\it
2417: Pub. Astron. Soc. Japan} \textbf{58} 485. [A discussion on the implication
2418: from the spectral data of the gamma-ray burst with the highest known
2419: redshift].}
2420:
2421: \Ref { Trac H., and Cen R. (2007). Radiative Transfer Simulations of Cosmic
2422: Reionization. I. Methodology and Initial Results. {\it Astrophys. J.}
2423: \textbf{671} 1. [Results from a state-of-the-art computer simulation of
2424: cosmic reionization].}
2425:
2426: \Ref { Verner D.A., Ferland G.J., Korista T. and Yakovlev
2427: D.G. (1996). Atomic Data for Astrophysics. II. New Analytic FITS for
2428: Photoionization Cross Sections of Atoms and Ions. {\it Astrophys. J.}
2429: \textbf{465} 487. [A compilation of photo-ionization cross sections for a
2430: variety of atomic and ionic species].}
2431:
2432: %\Ref { Viel M., Haehnelt M.G. and Lewis A. (2006). The Lyman $\alpha$
2433: %forest and WMAP year three. {\it Mon. Not. Roy. Astron. Soc. Lett.}
2434: %\textbf{370} L51.}
2435:
2436: \Ref {Weinberg S. (1972). Gravitation and Cosmology: Principles and
2437: Applications of the General Theory of Relativity. {\it Gravitation and
2438: Cosmology} (New York: Wiley). [A pioneering textbook that established the
2439: currently popular link between cosmology and particle physics].}
2440:
2441: \Ref { Weinberg D.H., Hernquist L. and Katz N. (1997). Photoionization,
2442: Numerical Resolution, and Galaxy Formation. {\it Astrophys. J.}
2443: \textbf{477} 8. [Results from numerical simulations on the effect of
2444: ionizing radiation on galaxy formation].}
2445:
2446: \Ref { White R.L., Becker R.H., Fan X., Strauss M.A. (2003). Probing the
2447: Ionization State of the Universe at $z>6$. {\it Astron. J.} \textbf{126}
2448: 1. [A study of the inference from quasar spectra concerning the ionization
2449: state of the intergalactic medium at redshift $z>6$].}
2450:
2451: \Ref { Wouthuysen S.A. (1952). On the excitation mechanism of the 21-cm
2452: (radio-frequency) interstellar hydrogen emission line. {\it Astron. J.}
2453: \textbf{57} 31. [A pioneering study of the effect of Lyman-$\alpha$ photons
2454: in couplin the spin temperature of hydrogen to its kinetic temperature].}
2455:
2456: \Ref { Wu K.K.S., Lahav O. and Rees M.J. (1999). The large-scale smoothness
2457: of the Universe. {\it Nature} \textbf{397} 225. [A summary of the evidence
2458: for the isotropy and homogeneity of the Universe].}
2459:
2460: \Ref { Wyithe J.S.B. and Loeb A. (2003). Self-regulated Growth of
2461: Supermassive Black Holes in Galaxies. {\it Astrophys. J.} \textbf{595}
2462: 614. [An early model for self-regulated growth of quasars and their
2463: resulting luminosity function].}
2464:
2465: \Ref { Wyithe J.S.B. and Loeb A. (2004a). A large neutral fraction of
2466: cosmic hydrogen a billion years after the Big Bang. {\it Nature}
2467: \textbf{427} 815. [A study linking the size of the ionized regions around
2468: high-redshift quasars with the neutral fraction of the intergalactic
2469: medium].}
2470:
2471: \Ref { Wyithe J.S.B. and Loeb A (2004b). A characteristic size of $\sim$
2472: 10Mpc for the ionized bubbles at the end of cosmic reionization. {\it
2473: Nature} \textbf{432} 194. [A model-independent derivation of the
2474: characteristic size of ionized bubbles at the end of the reionization
2475: epoch].}
2476:
2477: \Ref { Wyithe J.S.B., Loeb A. and Barnes D.G. (2005). Prospects for
2478: Redshifted 21 cm Observations of Quasar H II Regions. {\it Astrophys. J.}
2479: \textbf{634} 715. [A study of the feasibility of detecting ionized regions
2480: around high-redshift quasars as cavities in 21-cm surveys].}
2481:
2482: \Ref { Wyithe J.S.B., and Loeb A. (2008). Fluctuations in 21-cm emission
2483: after reionization. {\it Mon. Not. Roy. Astron. Soc.}, \textbf{383}, 606.
2484: [A derivation of 21-cm fluctuations at low redshifts owing to dense
2485: (Galactic) pockets of hydrogen that are self-shielded from the UV
2486: background after reionization].}
2487:
2488: \Ref { Wyithe J.S.B., Loeb A. and Geil P.M. (2008). Baryonic acoustic
2489: oscillations in 21-cm emission: a probe of dark energy out to high
2490: redshifts. {\it Mon. Not. Roy. Astron. Soc.} \textbf{383}, 1195. [A study
2491: of the detectability of baryonic acoustic oscillations in the 21-cm
2492: brightness fluctuations after reionization].}
2493:
2494: \Ref { Yamamoto K., Sugiyama N. and Sato H. (1997). Cosmological baryon
2495: sound waves coupled with the primeval radiation. {\it Phys. Rev. D}
2496: \textbf{56} 7566. [A comprehensive discussion on the underlying physics of
2497: acoustic oscillations]. }
2498:
2499: %\Ref { Yamamoto K., Sugiyama N. and Sato H. (1998). Evolution of
2500: %Small-Scale Cosmological Baryon Perturbations and Matter Transfer
2501: %Functions. {\it Astrophys. J.} \textbf{501} 442. [A comprehensive
2502: %discussion on the underlying physics of acoustic oscillations].}
2503:
2504: \Ref { Yoshida N., Omukai K., Hernquist L., and Abel T. (2006). Formation
2505: of Primordial Stars in a LCDM Universe. {\it Astrophys. J.} \textbf{652}
2506: 6. [Simulations of the formation of the first stars over a large
2507: cosmological region]. }
2508:
2509: \Ref { Zahn O., Lidz A., McQuinn M., Dutta S., Hernquist L., Zaldarriaga
2510: M. and Furlanetto S.R. (2006). Simulations and Analytic Calculations of
2511: Bubble Growth During Hydrogen Reionization. {\it Astrophys. J.}
2512: \textbf{654} 12. [Results from simulations of
2513: the evolution of cosmic reionization].}
2514:
2515: \Ref { Zhang W., Woosley S. and MacFadyen A.I. (2003). Relativistic Jets in
2516: Collapsars. {\it Astrophys. J.} \textbf{586} 356. [Simulations of the
2517: popular collapsar model for long-duration gamma-ray bursts, in which
2518: relativistic jets are produced by the collapse of a stellar core to a black
2519: hole]. }
2520:
2521: \Ref { Zygelman B. (2005). Hyperfine Level-changing Collisions of Hydrogen
2522: Atoms and Tomography of the Dark Age Universe. {\it Astrophys. J.}
2523: \textbf{622} 1356. [A detailed discussion on the effect of atomic collisions
2524: on the spin temperature of cosmic hydrogen].}
2525:
2526:
2527:
2528: \vfill\eject
2529:
2530: \end{document}
2531:
2532: