Skip to main content

Full text of "Enhanced Detectability of Pre-reionization 21-cm Structure"

See other formats


Submitted to ApJ Letters June 30, 2010 

Preprint typeset using I4TgX style emulateapj v. 5/25/10 



O 
£ 

O 
m 

O 

u 

Oh 
i 

o 

CO 
C3 



O 

o 
p 

o 
o 

> 

X 



ENHANCED DETECTABILITY OF PRE-REIONIZATION 21 -CM STRUCTURE 

Marcelo A. Alvarez 1 , Ue-Li Pen 1 , and Tzu-Ching Chang 1 - 2 

Submitted to ApJ Letters June 30, 2010 

ABSTRACT 

Before the universe was reionized, it was likely that the spin temperature of intergalactic hydrogen was de- 
coupled from the CMB by UV radiation from the first stars through the Wouthuysen-Field effect. If the IGM 
had not yet been heated above the CMB temperature by that time, then the gas would appear in absorption rela- 
tive to the CMB. Large, rare sources of X-rays could inject sufficient heat into the neutral IGM, so that ST/, > 
at comoving distances of tens to hundreds of Mpc, resulting in large 21 -cm fluctuations with ST], ~ 250 mK 
on arcminute to degree angular scales, an order of magnitude larger in amplitude than that caused by ionized 
bubbles during reionization, <57i ~ 25 mK. This signal could therefore be easier to detect and probe higher 
redshifts than that due to patchy reionization. For the case in which the first objects to heat the IGM are QSOs 
hosting 10 7 -M Q black holes with an abundance exceeding ~ 1 Gpc" 3 at z ~ 15, observations with either the 
Arecibo Observatory or the Five Hundred Meter Aperture Spherical Telescope (FAST) could detect and image 
their fluctuations at greater than 5-<r significance in about a month of dedicated survey time. Additionally, 
existing facilities such as MWA and LOFAR could detect the statistical fluctuations arising from a population 
of 10 5 -M© black holes with an abundance of ~ 10 4 Gpc" 3 at z ~ 10- 12. 
Subject headings: cosmology: theory — dark ages, reionization, first stars 



intergalactic medium 



1. INTRODUCTION 

The 2 1 cm transition is among the most promising probes of 
the high-redshift universe. The observed differential bright- 
ness temperature from the fully neutral IGM at the mean den- 
sity with spin temperature T s (z) is given by 



ST b (z) ~ 29 mK 



1+z 
10 



1/2 



1 



Tcmb(z) 
T s (z) 



(1) 



(e.g., Madau et al.|1997] l. In the absence of an external radi- 
ation field, only the hot, dense gas in minihalos would have 
been able to develop a spin temperature di fferent from that 
of the CMB at z ~ 20 ( [Shapiro et aT1[2006| . Howeve r, radi- 
ation emitted by the earliest generations of stars (e.g., Ciardi 
& Madau|[2003 ; Barkana & Loeb 2005) and X-ray sources 



(Chuzhoy et al. 2006, Chen & Miralda-Escude 2008) could 
have coupled the spin temperature to t he IGM temperature 
via th e "Wouthuysen-Field effect" (Wouthuysen 1952; Field 
1 1959} long before reionizing the universe, at redshifts as high 
as z = 20-30, causing the IGM to appear in absorption with 
respect to the CMB, with ST b ~ -250 mK. 

Only a modest amount of heating is necessary to raise the 
gas temperature above the CMB temperature of ~ 30-60 K 
at z ~ 10-20. By the time reionization was well underway, it 
is generally believed that the neutral component of the IGM 
had already been heated to T s 3> 7cmbi an d therefore the ob- 
served brightness temperature will be an order of magnitude 
lower in amplitu de, 8Tb ~ 25 mK, t han when the IGM was in 
absorption (e.g., Furlanet to et al.|2006) . 

Clearly, there should be some transition epoch, during 
which sources of X-ray radiation heated the neutral IGM, 
creating "holes" in the absorption. Natural candidates for 
these early sources of heating are quasars (e.g., Chuzhoy et al. 
2006; Zaroubi et al. 2007; Thomas & Zaroubi 2008; Chen & 



the quasar population at z > 6, with t he only constraints com- 
ing from th e bright end at z ~ 6 (e.g., Fan et al.|2002 Willott 
et al. 2010 1, with inferred black hole masses ^ 10 8 M , lumi- 
nosities <; 10 46 erg s~ l , and a comoving abundance <; 1 Gpc" 3 . 

Quasars are not the only high-redshift objects able to pro- 
duce X-rays. High-redshift supernovae (Oh 2001 ) and X-ray 
binaries could have also been sources. Theoretical models 
typically parametrize X-ray production associated with star 
formation by fx, normalized so that fx = 1 corresponds to that 
observed for local starburst galaxies (e.g., Furlanetto 2006, 
Pritchard & Furlanetto 2007). 

Our aim in this paper is to explore the enhanced 21 -cm sig- 
nature of early quasars. We will therefore assume that the 
first galaxies were efficient enough to couple the spin temper- 
ature to the kinetic temperature by z — 20, but did not heat the 
IGM above the CMB temperature, i.e. /^<1. As shown by 



2C 

M 



Miralda-Escude 2008). Unfortunately, little is known about 



Electronic address: malvarez@cita.utoronto.ca 

1 Canadian Institute for Theoretical Astrophysics, University of 
Toronto, 60 St.George St., Toronto, ON M5S 3H8, Canada 

2 IAA, Academia Sinica, PO Box 23-141, Taipei, 10617, Taiwan 



Pritchard & Loeb (2010), even for values of f x > 1 it is still 
possible that there is a significant period between the time 
when the spin temperature was coupled to the kinetic temper- 
ature and the time when the mean IGM was heated above the 
CMB temperature. In addition, the first galaxies were likely 
clustered around the quasars and may have actually increased 
the sizes of the heated regions. On the other hand, X-ray- 
emitting galaxies could have also formed outside the regions 
that would have been heated by the quasars, making the 21- 
cm features due to early, rare quasars that we predict here less 
prominent. 

This paper is organized as follows. In §2 we present our 
forecasted signal from partially-heated regions and a brief 
discussion of the expected black hole abundance at high- 
redshift. In §3 we assess the detectability of the predicted 
signal under various assumptions about the early quasar 
population. We end in §4 with a discussion of possi- 
ble survey strategies and requirements of detection on the 
high-redshift black hole abundance. All calculations were 
done assuming a flat universe with (Q m h 2 ,nbh 2 ,h,n,,as) = 
(0.1 33, 0. 0225, 0.7 1,0.96, 0.8), consistent with WMAP 7- 
year data (Komatsu et al. 2010). All distances are comoving. 



ALVAREZ, PEN & CHANG 



2. FORECAST 

Our predictions will focus on accreting black holes with 
masses greater than ~ 1O 5 M . It is of course possible that a 
more abundant population of black holes with lower masses 
existed at these early times, left behind as re mnants of early 
generations of massive Pop III stars ( [Heger et al. 2003 ), act- 
ing as "seeds" for the observed z ~ 6 supermassive black hole 
population (e.g., Li et al. 2007), and powering "miniquasars" 
(Madau et al. 2004). However, radiative feedback from the 



10 Mq in th e first halos with virial temperature 
Bromm & Loeb|2003 Begelman 



progenitor stars (Johnson & Bromm 2007) as well as the ac- 
cretion radiation itself (Alvarez et al. 2009), would have likely 
substantially limited their growth and corresponding X-ray 
emission at early times. An alternative scenario for form- 
ing the seeds is gaseous collapse to black holes with masses 
greater than 

greater than ~ 10 4 K (e.g., 

et al. 2006). Formation of black holes by thTs mechanism may 
have been quite a rare occurence (e.g., Dijkstra et al. 2008 ). It 
is this scenario, in which most accreting black holes in the uni- 
verse were relatively rare and more massive than ~ 1O 5 M , 
that is most consistent with the predictions we make here. 

Because the heating is a time-dependent effect, we will 
parametrize the total energy radiated during accretion as E m = 
LtQso = cMbhc 2 , where we take the radiative efficiency e = 
0.1. In principle some of the rest mass energy, i.e. the ini- 
tial seed mass of the black hole, did not contribute to heating 
the surroundings, but in general the seed mass is expected to 
be small compared to the mass of the black hole after it un- 
dergoes its first episode of radiatively-efficient accretion as a 
quasar, so we neglect it. 

2.1. Quasar spectral energy distribution 

We assume the spectral energy distribution of the quasar 
is given by a broken power-law, S„ oc v~°- 5 at v < i>\„ and 



Bolton & 



S v oc v at v > Vb, with hv\, = 11.8 eV (e.g. 
Haehnelt 2007), approximate ly consistent with the template 
spectra of Telfer et al. (2002 1. If the quasar has a total bolo 
metric luminosity L, then the spectral energy distribution is 



S v 



L 






V < Z/b, 

v > v h . 



(2) 



2.2. The 21 -cm profile around individual quasars 

The temperature profile surrounding a quasar can be ob- 
tained by considering the fraction of the radiated energy ab- 
sorbed per atom at comoving distance r and redshift z, 



^m(r,z) = 



(1+z) 2 



, S v a v 
dp—, — (hv- 



foto)Xi,exp[-7Y/(r)], 



4irLr 2 J Vm " hv 

(3) 
where \v is the fraction of photoelectron energy, hv-hvm, 
which goes into heat, with the rest being lost to secondary 
ionizations and excitations. In the limit in which the ionized 
fraction of the gas is low, 0. 1 < \ v < 0.3 for all hv > 25 eV 



( Shull & van Steenberg[J98? i. In what follows we will make 

the approximation that S u = 0.2 for all v 

energy radiated is Lt qso = eMbhc 2 , with f, 

the Hubble time. In this case, the relative brightness tempera 

ture is given by 



and assume that the 
qso short compared to 



ST h (M BH ,r,z) = 29 mK 



1- 



7cmb(z) 



T s (M BH ,r,z) 



1+z 
10 



"I 1/2 



(4) 





50 









-50 


M 

£ 


-100 


-150 


E- 


-200 
-250 




-300 




-350 




-400 



01 

E- 



100 =- 



10 =- 




10 100 

r [Mpc] 



10 3 



FIG. 1. — Top: Profiles of the observed differential brightness temperature 
versus comoving distance at three different redshifts around a quasar that has 
radiated 10 per cent of the rest-mass energy of 10 5 , 10 7 , and 10 9 Mp 
labeled. Bottom: Spin temperature of the IGM, obtained via equation I 



where 



T s (M B H,r,z) = T 1GM (z)+ — r ffl (r,z). 

3k b 



(5) 



In calculating the optical depth, we assume that the IGM is 
completely neutral at the mean density, «h(z), so tnat T v{r,z) = 
rnu{z)(l +z) _1 ov. In reality, the quasar's own H II region will 
reduce the opacity at small radii, but given that we are con- 
cerned with the small amount of heating happening at much 
larger radii, we neglect the H II region when calculating the 
optical depth. 

Shown in Fig. [T] are profiles of the observed differential 
brightness temperature versus comoving distance at three dif- 
ferent redshifts. The total energy radiated by the quasar, ZiQso 
in each curve corresponds to 10 per cent of the rest mass, as 
labeled. Close to the quasar, the gas is heated above the CMB 
temperature, and 5Tb ~ 30-40 mK. Further away, the heating 
from the quasar declines due to spherical dilution and attenu- 
ation of the lowest energy radiation, with STf, finally reaching 
about -220 to -340 mK at large distances. The FWHM size of 
the fluctuations are about 20, 80, and 400 Mpc for black holes 
of mass 10 5 , 10 7 , and 10 9 M , respectively. 



PRE-REIONIZATION 21 -CM STRUCTURE 



2.3. Black hole abundance at high-z 

How do the black hole densities and masses we assume 
here compare to those observed at z ~ 6 and expected at 
higher redshifts? Willott et al. (2010) constructed the mass 
function of black holes in the range 10 8 M Q < M B h < 3 x 
10 9 M Q at z = 6, finding it to be well-fitted by dn/dlnM Bll ~ 
<j>*(M Bii /M*)- 1 exp(-M BH /M*), with 0* = 5.34 Gpc" 3 and 
M* = 2.2 x 10 9 Mq. Integrating the black hole mass function, 
one finds p BH (> M BH ) ~ 7 x 10 9 and 3 x 10 10 M Gpc" 3 , for 
A^bh = 10 9 , and 10 8 Mq, respectively, somewhat larger than 
the black hole mass density we find which maximizes the 21- 
cm power spectrum (§3.2). Matching the abundance of black 
holes great er than a given mass to the dark matter halo mass 
function of Warren et al. (2006) at z = 6 ("abundance match- 
ing" - e.g., Kravtsov et al. 2004 ), we obtain M halo = 2.3 x 10 12 
and 4.7 x 10 1Z M for M BH = 10 8 and 10 9 M , respectively, 
implying a value of M B ^/M\ i . i \ ~ 4 x 10" 5 to 2 x 10 -4 over 
the same range. 

More detailed predictions would require extrapolating the 
MfiH-A^haio relationship to lower masses and higher redshifts, 
or making highly uncertain assumptions about the formation 
mechanism of the high-redshift seeds and their accretion his- 
tory. For example, the ratio M BH /Mh a io ~ 10~ 4 we determine 
here by abundance matching at z = 6 and M B u = 10 8 - 10 9 M Q 
would be a significant underestimate in atomic cooling halos 
where black hole formation by direct collapse took place, in 
which it is possible that M B u/Mh- d i could approach the limit- 
ing value of ilb/ft m ~ 0.17. Clearly much more work is re- 
quired in understanding the high-redshift quasar population, 
and for this reason the constraints provided by either detec- 
tion or non-detection of the signal we predict here would be 
very valuable. 

3. DETECTABILITY 

In this section, we estimate the detectability of individ- 
ual sources as well the statistical detection of their power 
spectrum. In the case of individual objects, we focus on a 
novel approach, using single-dish filled aperture telescopes 
like Arecibcr] and FASlF] while for the power spectrum we 
will simply refer to existing sensitivity estimates for facilities 
such as LOFAR0 MWA0 and SKAQ 

3.1. Individual quasars 

In order to determine the necessary integration time, we 
convolve the profiles shown in Fig. [Tjwith a half -power beam 
width of 



100 



% = 26' 



1+z 
11 



/ ^dish 

\300m 



(6) 



where offish is the effective dish diameter. Converting angle on 
the sky to comoving distance, we obtain the comoving reso- 
lution, 

D ^ 70 Mpcf-^V7^V' 2 . (7) 



This implies that the fluctuation created by a 10 7 M Q -black 
hole would be just resolvable with a single 300-m dish like 



http://www.naic.edu 



http://fast.bao.ac.cn 

5 http://www.lofar. org 

6 http://www.mwatelescope.org 

7 http://www.skatelescope.org 







l^-l 


10 


c 




»<— 1 




e 




o 


1 


■^H 




\ 




*J 




<^ 


0.1 


Z 




\ 




C/2 





^ 



J3 

E- 

T 

X 

(0 

B 



1 1 Mini — i 1 1 Mini — i 1 1 1 mil — i i inini — i i mini — mt 




z=10 % 
z=15 



z=20 

; I I I I I U-LLli 



10 4 10 5 10 6 10 7 10 8 10 9 10 10 
M BH [M ] 

FIG. 2. — Bottom: Peak fluctuation amplitude, measured with respect to 
the mean background absorption (A7b)(j), as a function of black hole mass, 
for three different redshift. Top: Signal-to-noise ratio at the same redshifts 
for a bandwidth corresponding to the beam width and an integration time of 
10 minutes. The detectability declines rapidly in the interval 10 < z < 20. 

Arecibo (see Fig. IT), lower mass black holes would be un- 
resolved point sources, and the profile of higher mass black 
holes could actually be measured. For an integration time of 
At, the sensitivity is given by 



ST err = 22 mK 



At 



-1/2 



/ , \ 1/2 /I , \ 2.35 



60s J V 300m 



where we have used 

At/ = 4 MHz 



ll 



-1/2 



(8) 



(9) 



p \ n+z 

70MpcJ ^ 11 

for the bandwidth corresponding to a comoving distance D, 
STar = T^/y/AvAt, and r SYS ~ 3 x 10 5 mK[(l +z)/ll] 27 
(e.g., |Furlanetto et al.|2006) . 

We calculate the maximum fluctuation amplitude as a func- 
tion of quasar black hole mass and redshift, <57b, max (Mbh,z), 
which is a convolution of the beam with the individual pro- 
files plotted in Fig. [TJ 



ST b , 



2 
fflD 



d69e ie i / dlST b (r w ), (10) 

o Jo 



where we use a Gaussian profile with a FWHM of f?b(z), such 



ALVAREZ, PEN & CHANG 





1 


1 


"i ' 


1 




10 4 




n qso = 2X10 S 


2xl0 4 


Gpc" 3 


- 






M BH = 10? 


10 5 


M e 


: 






/ ■■' / / ■■' / 


\ y 

\y 


\\\ 


~ 


-100 




/ ■'' / / •' / 
1 ■*' / / ■' ^ 




e 


g 10 

w 

CO 


Z ■' / 
.■' / 
/ 
1 

— I 


/ / / .-' / 

r / -'' f 
/ 1:1 
J :' J 
/ 
/ / 
/ 
/ 
/ 

z-10 




\ \\ 
\\ 

\ 1 


— 






z-15 




\ 


- 


0.1 




z-20 




1 11 

til mi 


— 




1 


1 


ill I 


i ^ 



0.001 0.01 0.1 1 

k [1/Mpc] 

FIG . 3 . — Shown is the spherically-averaged three-dimensional power spec- 
trum of 21 -cm fluctuations for a black hole density 1 Mq/Mpc 3 , for three 
different redshifts, as labeled. Large, relatively rare 10 7 Mq black holes have 
a power spectrum which peaks at k ~ 0.03 Mpc~', while for Mbh = 10 5 Mq 
the power spectrum which peaks at k ~ 0.15Mpc~ 3 . Such a signal at z = 10 
should be easily detectable by LOFAR, MWA, or SKA. 



that f?b(z) = 2\/2ln28 g (z), and rj e = r 2 e + I 2 , where rg is the 
projected comoving distance perpendicular to the line of sight 
corresponding to the angle 9. 

Shown in Fig. [2] are the resulting fluctuation ampli- 
tudes with respect to the mean back ground absorption, 
^7b. max (Mbh,z)- (STt,)(z), as well as the signal-to-noise ratio, 
[ST b>m3x (M b b,z)- (ST b )(z)]/5T b , en {z) for an integration time of 
10 minutes. As can be seen from the figure both the signal, 
and to a greater extent the signal-to-noise, decline rapidly with 
increasing redshift. This is due to several reasons. First, the 
intrinsic signal declines with increasing redshift because ab- 
sorption is relatively weaker at higher redshifts when the IGM 
has not had as much time to cool (see Fig.[T|. In addition, the 
beamwidth gets larger, in proportion to l+z, while the angu- 
lar size of the fluctuations at fixed black hole mass actually 
decline with redshift, as seen from Fig. [T] 

3.2. Power spectrum from Poisson fluctuations 

To estimate the fluctuating background from a superposi- 
tion of sources, we assume all black holes at a given red- 
shift have the same mass, Mbh, and comoving number den- 
sity, «bh- We then generate a realization of random black 
hole positions within a periodic box, and calculate the heat- 
ing from each source using equation d5V Because the heating 
rate is proportional to mass and we assume a random spa- 
tial distribution, the two length scales in the system are the 

mean separation of sources, r sep oc n Bt [ , and the distance out 
to which an individual black hole can effectively heat the IGM 

1/3 

above the CMB temperature, r bs .. di oc M B ' H . For this simplified 
case, in which the spatial distribution of black holes is ran- 
dom and all black holes have the same mass, the shape of the 
power spectrum is determined only by the ratio rh eat /r s , 



roughly that density at which the signal is maximized; a lower 
black hole mass density results in a lower overall signal, while 
a higher density leads to overlap of the individual regions, and 
the fluctuations become saturated. 

Shown in Fig. Rlis our predicted power spectrum for Pois- 
son fluctuations of black holes of mass M BH = 10 5 and 10 7 M Q 
at z =10, 15, and 20. Because the black hole mass density 
is the same in both cases, the curves have the same shape, 
but are shifted in wavenumber such that k~ 3 oc Mbh- Be- 
cause the individual regions are only weakly overlapping for 
Pbh = 2 x 10 9 M Q Gpc~~ , the shape of the curve is quite close 
to the Fourier transform of an individual region, so that 



P(k) 



sin kr 



POO 

1/2 oc / r 2 dr5T h (r)- 

Jo kr 



(11) 



with STb(r) given by eqs. Q and (|5j. 

The amplitude of this signal at k ~ 0.1 Mpc -1 
([fc 3 f(A;)/(27r 2 )] 1 / 2 ~ 50 mK) is almost an order of magni- 
tude greater than that expect ed from ionized bubbles when 
Ts ^ Tcmb (— 6 mK, e.g., |Furlanetto et al.||2004|), which 



compensates for the increased foregrounds at the higher 
redshifts corresponding to the signal we predict here. For 



>l/3 = n 1 ! 3 



sep 



(«bh^bh) ' = Pbh ■ Our choice of p B H = 2 x 10 y M Gpc is 



example, McQuinn et al. (2006) estimated the sensitivity of 
various facilities to the spherically-averaged power spectrum, 
finding (at z = 12) ST 2 en .(k - 0. lMpc -1 ) ~ 30mK 2 for LOFAR 
and MWA, and T b 2 en (k - O.lMpc" 1 ) ~ O.lmK 2 for SKA, for 
their adopted array configurat ions and 1 000 hr of integration 
time (Fig. 6 and Table 1 o f [McQuinn et al.|2 006 ). Thus, the 
power spectrum shown in Fig. 4 would be easily detectable 
by either of these th ree experiments at z y 12 for the survey 
parameters used by McQuinn et al."|(|2006|l. 



4. SURVEY STRATEGIES 

Since the signal comes from a large angular scale on the 
sky, a filled aperture maximizes the sensitivity. The two 
largest collecting area telescopes are Arecibo and, under con- 
struction, FAST. They are at latitudes 5=18 (26) degrees, re- 
spectively. To stabilize the baselines, a drift scan at 80 MHz 
maps a strip of width 40 (90) arc minutes by length 360 de- 
grees cos S every day. With Arecibo, a custom feedline with 
pairwise correlations of dipoles would allow a frequency de- 
pendent illumination of the mirror, allowing the frequency in- 
dependent removal of foregrounds. It would also allow op- 
eration as an interferometer, increasing stability and rejection 
of interference, and enabling arbitrary apodization of the sur- 
face. Should grating sidelobes from support structures be- 
come a problem, one could also remove the carriage house, 
and mount the feedline on a pole from the center of the dish. 
This would result in a clean, unblocked aperture with fre- 
quency independent beam. Even with the carriage support 
blocking, the side lobes would still be frequency independent 
for an appropriately scaled illumination pattern. 

For FAST, a focal plane array also enables a frequency de- 
pendent illumination of the primary, resulting in a frequency 
independent beam on the sky. It also increases the survey 
speed by the number of receivers used. Only one receiver is 
needed every half wave length, roughly two meters. Hundred 
pixel surveys seem conceivable at low cost, since only small 
bandwidths would be needed, and the system temperature is 
sky limited, even with cheap TV amplifiers. 



PRE-REIONIZATION 21 -CM STRUCTURE 



1U 




1 1 ' ' ' 


1 1 1 








X~ 






kv = 4 MHz 


xX 


10 5 






x X — 








^x - 








x'X 


10 4 






xy X, 






X'X X<2 








// XX - 






<5Q= 1140 dee 2 


XX x : X - 


10 3 




t obs = 1 month x^/ 


Xy^ 1 






S'X 


x-y : 


Z 100 




SX 


/xX — 






S-X / 








^ XX^ 


- y 6Q= 40 deg 2 \ 


10 


X 


/ X-X 


t obs= 1 da y ! 






XX 




1 




X-X 
X X 


z-10 - 






XX 

S'x 


z-15 : 




??- 




- 


0.1 






z-20 






1 1 < < < 


1 1 ; 



10 



100 10 3 10 4 

n bh [Gpc- 3 ] 



10 5 



FIG. 4. — Number of black holes expected within a 1-month (upper curves) 
and 1-day (lower curves) survey in drift-scan mode that scans each point 
in the sky nine times, as a function of comoving black hole density. The 
integration time per pixel in each case is about 7, 14, and 25 minutes for 
Z = 10, 15, and 20. With a 14-minute integration time, a 10 7 -M© black hole 
would be detectable at greater than 5-<r significance (see Fig.|2} 



one month survey operating at 88 MHz (z ~ 15) with a 300- 
m dish like Arecibo, which scans each point in the sky nine 
times would have a pixel size of about 0.4 deg 2 during which 
each pixel was integrated for about 15 minutes and 1 140 deg 2 
were surveyed. If 10 7 — M® black holes had a number density 
of 1 Gpc -3 at that time, one would expect to discover about 
three or four of them at greater than 5-er significance since 
each pixel would have been integrated for about 15 minutes 
at z = 15 (see dotted lines in Figs. SandHjl. Black holes with 
masses greater than 10 7 M and similar abundances would be 
easily detectable, allowing for followup with longer baseline 
facilities such as LOFAR to determine the detailed shape of 
their 21-cm profile. Detecting a 1O 6 -M black hole at 5-a 
significance would require a significantly longer integration 
time on each pixel, about 4 hours, so that only about 70 deg 2 
could be surveyed in a month. Thus, the minimum spatial 
density of 10 6 -M Q black holes would be about 5 Gpc" 3 at 
z = 15. Smaller black holes would mean even smaller fields of 
view and longer integration times per pixel, so in those cases 
detecting their statistical signature in the spherically-averaged 
power spectrum discussed in §3.2 may be a better approach. 

Regardless of whether high-redshift quasar X-ray sources 
will be detected directly by the means we propose here, it 
is clear that the pre-reionization universe offers a wealth of 
information that can be uniquely probed with the 21-cm tran- 
sition, and it is therefore vital that theoretical models for the 
signal originating from before the epoch of reionization con- 
tinue to be developed. 



We note that the ratio of signal to foreground in this regime 
is comparable to that during reionization. With a filled aper- 
ture, it may be easier to achieve foreground subtraction. This 
has been demonstrated for intensity mapping at z ~ 0.8 using 
the filled aperture of GBT (Chang et al. 2010), where a similar 
foreground ratio exists of galactic synchrotron to 21cm. 

Shown in Fig. [4] is the expected number of black holes 
within the field ofview for a given redshift and survey band- 
width, versus their comoving number density. A dedicated 



We wish to thank J. Peterson and R. M. Thomas for helpful 
discussions on Arecibo and 21-cm signatures of early QSOs, 
resepectively, and J. R. Pritchard for comments on an earlier 
draft of the paper. M. A. A. and T C. are grateful for the 
hospitality of the Aspen Center for Physics, where this work 
was completed. We acknowledge financial support by ClfAR 
andNSERC. 



REFERENCES 



Alvarez, M. A., Wise, J. H., & Abel, T. 2009, ApJ, 701, L133 

Barkana, R. & Loeb, A. 2005, ApJ, 626, 1 

Begelman, M. C, Volonteri, M., & Rees, M. J. 2006, MNRAS, 370, 289 

Bolton, J. S. & Haehnelt, M. G. 2007, MNRAS, 374, 493 

Bromm, V. & Loeb, A. 2003, ApJ, 596, 34 

Chang, T.-C, Pen, U.-L., Bandurak, K., & Peterson, J. 2010, Nature, in press 

Chen, X. & Miralda-Escude, J. 2008, ApJ, 684, 18 

Chuzhoy, L., Alvarez, M. A., & Shapiro, P. R. 2006, ApJ, 648, LI 

Ciardi, B. & Madau, P. 2003, ApJ, 596, 1 

Dijkstra, M.. Haiman, Z., Mesinger, A., & Wyithe. J. S. B. 2008, MNRAS, 

391, 1961 
Fan, X., Narayanan, V. K., Strauss, M. A., White, R. L., Becker, R. H., 

Pentericci, L., & Rix, H.-W. 2002, AJ, 123, 1247 
Field, G. B. 1959, ApJ, 129, 536 
Furlanetto, S. R. 2006, MNRAS, 371, 867 

Furlanetto, S. R., Oh, S. P., & Briggs, F. H. 2006, Phys. Rep., 433, 181 
Furlanetto, S. R., Zaldarriaga, M., & Hernquist, L. 2004, ApJ, 613, 1 
Heger, A., Fryer, C. L., Woosley. S. E., Langer, N., & Hartmann, D. H. 

2003, ApJ, 591, 288 
Johnson, J. L. & Bromm, V. 2007, MNRAS, 374, 1557 
Komatsu, E. et al. 2010, ApJ, submitted, arXiv:1001.4538 
Kravtsov, A. V., Berlind, A. A., Wechsler, R. H., Klypin, A. A., Gottlober, 

S., Allgood. B., & Primack, J. R. 2004, ApJ, 609, 35 
Li, Y., Hernquist, L., Robertson, B., Cox, T. J., Hopkins, P. F, Springel, V., 

Gao, L., Di Matteo, T., Zentner, A. R., Jenkins, A., & Yoshida, N. 2007, 

ApJ, 665, 187 



Madau, P., Meiksin, A., & Rees, M. J. 1997, ApJ, 475, 429 

Madau, P., Rees, M. J., Volonteri, M., Haardt, F, & Oh, S. P. 2004, ApJ, 

604, 484 
McQuinn, M., Zahn, O., Zaldarriaga, M., Hernquist, L., & Furlanetto, S. R. 

2006, ApJ, 653, 815 
Oh, S. P. 2001, ApJ, 553,499 

Pritchard, J. R. & Furlanetto, S. R. 2007, MNRAS, 376, 1680 
Pritchard, J. R. & Loeb, A. 2010, ArXiv e-prints 
Shapiro, P. R., Ahn, K., Alvarez, M. A., Iliev, I. T., Mattel, H., & Ryu, D. 

2006, ApJ, 646, 681 
Stall, J. M. & van Steenberg, M. E. 1985, ApJ, 298, 268 
Telfer, R. C, Zheng, W., Kriss, G. A., & Davidsen, A. F. 2002, ApJ, 565, 

773 
Thomas, R. M. & Zaroubi, S. 2008, MNRAS, 384, 1080 
Wan-en, M. S., Abazajian, K., Holz, D. E., & Teodoro, L. 2006, ApJ, 646, 

881 
Willott, C. J., Albert, L., Arzoumanian, D., Bergeron, J., Crampton, D., 

Delorme, P., Hutchings, J. B., Omont. A., Reyle, C, & Schade, D. 2010, 

ArXiv e-prints 
Wouthuysen, S. A. 1952, AJ, 57, 31 
Zaroubi, S., Thomas, R. M., Sugiyama, N, & Silk, J. 2007, MNRAS, 375, 

1269