You are on page 1of 17

The Astrophysical Journal, 680:829Y845, 2008 June 20 A

# 2008. The American Astronomical Society. All rights reserved. Printed in U.S.A.

EFFECT OF PRIMORDIAL BLACK HOLES ON THE COSMIC MICROWAVE BACKGROUND


AND COSMOLOGICAL PARAMETER ESTIMATES
Massimo Ricotti
Department of Astronomy, University of Maryland, College Park, MD 20742; ricotti@astro.umd.edu
and
Jeremiah P. Ostriker and Katherine J. Mack
Department of Astrophysical Sciences, Princeton University, Peyton Hall, Ivy Lane, Princeton, NJ 08544
Received 2007 August 24; accepted 2008 February 21

ABSTRACT
We investigate the effect of nonevaporating primordial black holes (PBHs) on the ionization and thermal history
of the universe. X-rays emitted by gas accretion onto PBHs modify the cosmic recombination history, producing
measurable effects on the spectrum and anisotropies of the cosmic microwave background (CMB). Using the third-year
WMAP data and COBE FIRAS data we improve existing upper limits on the abundance of PBHs with masses >0.1 M
by several orders of magnitude. The new upper limits still allow PBHs to be important for the origin of supermassive
black holes and ultraluminous X-ray sources. Fitting WMAP3 data with cosmological models that do not allow for
nonstandard recombination histories, as produced by PBHs or other early energy sources, may lead to an underestimate
of the best-fit values of the amplitude of linear density fluctuations (8 ) and the scalar spectral index (ns ). Cosmological
parameter estimates are affected because models with PBHs allow for larger values of the Thomson scattering optical
depth, whose correlation with other parameters may not be correctly taken into account when PBHs are ignored. Values
of e  0:2, ns  1, and 8  0:9 are allowed at 95% CF. This result may relieve recent tension between WMAP3 data
and clusters data on the value of 8 . PBHs may increase the primordial molecular hydrogen abundance by up to 2 orders
of magnitude, this promoting cooling and star formation. The suppression of galaxy formation due to X-ray heating
is negligible for models consistent with the CMB data. Thus, the formation rate of the first galaxies and stars would
be enhanced by a population of PBHs.
Subject headinggs: black hole physics — cosmic microwave background — cosmology: observations —
cosmology: theory — cosmological parameters — early universe
Online material: color figures

1. INTRODUCTION evaporating PBHs. In addition, we find that the existence of PBHs


may affect the best-fit estimates of cosmological parameters. In
During the radiation era, before the formation of the first stars particular, the value of e , 8 , and ns can be underestimated in
and galaxies, small perturbations of the energy density of the uni- models which do not account for the existence of PBHs. This
verse on scales comparable to the particle horizon may become result is more general than the case of PBHs discussed in this
gravitationally unstable. The outcome of the collapse can be the paper. Recently we became aware that similar effects on estimates
direct formation of primordial black holes (PBHs; Hawking 1971; of cosmological parameters, and in particular ns, have also been
Carr & Hawking 1974; Musco et al. 2005; Harada & Carr 2005). found in Bean et al. (2003, 2007), although not from the mod-
If these PBHs form at a sufficiently high mass, they do not evap- eling of PBH accretion. Any source of energy injection at early
orate but instead begin to grow by accretion, producing X-rays. times that modifies the recombination history may lead to under-
A population of sufficiently massive and numerous PBHs may estimating e , ns , and 8 if the effect is excluded a priori when
provide an important and observable source of energy injection fitting the WMAP3 data. This scenario may ease recent tensions
into the cosmic gas before the formation of nonlinear large-scale between the WMAP3 analysis that favors low values of 8  0:74
structures and galaxies. Other authors have addressed this prob- (assuming a standard recombination history) versus Ly forest
lem before (Barrow & Silk 1979; Carr 1981; Gnedin et al. 1995; (e.g., Viel et al. 2006) and clusters data that may favor larger
Miller & Ostriker 2001). Here, we improve on previous cal- values of 8  0:9 (e.g., Evrard et al. 2008, but see also Bode
culations by modeling the accumulation of dark matter around et al. 2007). Finally, ionization from the X-rays of accreting PBHs
PBHs, the proper motion of PBHs, and feedback effects including will increase the amount of H2 significantly and the resultant
Compton drag. We simulate the ionization and thermal history of extra cooling will enhance early star formation.
primordial plasma using semianalytical calculations as in Ricotti This paper is organized as follows. In x 2 we review the prop-
& Ostriker (2004). The early energy injection by PBHs may pro- erties of PBHs and the current constraints. In x 3 we introduce
duce observable distortions of the CMB spectrum (Battistelli et al. the basic equations for gas and dark matter accretion and for the
2000) and may affect CMB anisotropies. We fit cosmological accretion luminosity of PBHs. In x 4 we evaluate the effect of
models which include the effect of PBHs on the cosmic recom- local and global feedback processes. In x 5 we present the results
bination history to the third-year WMAP data (Spergel et al. 2007). of calculations of the ionization and thermal history of the IGM,
This is done by modifying the Monte Carlo code CosmoMC and in x 6, data from WMAP3 and COBE are used to constrain the
(Lewis & Bridle 2002) appropriately. Our main goal is to im- mass and abundance of PBHs. A summary and discussion is pre-
prove existing upper limits on the mass and abundance of non- sented in x 7. Throughout the rest of the paper we use the following
829
830 RICOTTI, OSTRIKER, & MACK Vol. 680

cosmological parameters (h ¼ 0:73,


m ¼ 0:238,
b ¼ 0:0418, about 20% of the Galactic halo mass. Recent results seem to dis-
8 ¼ 0:74, and ns ¼ 0:95) from WMAP3 (Spergel et al. 2007). favor this claim (e.g., Tisserand et al. 2007). The goal of the
present work is to better constrain the abundance of relatively
2. PRIMORDIAL BLACK HOLES
massive (M k 0:1 M ) PBHs by modeling their effect on the
In the Newtonian regime, the theory of PBH formation can ionization history of the universe. To further motivate the present
be understood in simple terms. The Jeans length in a static and study we briefly review the current status of theoretical works
homogeneous fluid with sound speed vs is RJ  (c/vs )RSch , where and observational limits on the existence of PBHs. For a more
RSch is the Schwarzschild radius of a black hole with density equal
pffiffiffi comprehensive review see Carr (2006).
to the mean cosmic value. During the radiation era vs  c/ 3,
2.1. Current Observational Limits on PBHs
thus RJ ! RSch . This means that the self-gravitating regime ap-
pears when the perturbation is very close to the black hole regime. Many physical processes may lead to the formation of PBHs.
The critical overdensity needed to trigger PBH collapse is w P1/3, For example, PBHs may form from perturbations after inflation,
where P ¼ wc 2 is the cosmic equation of state (Carr 1975; Green during phase transitions of the cosmic equation of state or from
et al. 2004). The typical mass of a PBH is approximately equal to topological defects (e.g., Carr & Hawking 1974; Carr 2006). The-
or smaller than the mass within the particle horizon at the redshift oretically, it is unclear what is the largest allowed mass of PBHs.
of its formation, zf : Inflationary theories predict an almost scale-invariant (ns  1)
    initial spectrum of perturbations. This spectrum has the remark-
MPBH 1 þ zeq 2  2 able property that all perturbations have the same amplitude when
 f hor  f hor ; ð1Þ
Mhor (zeq ) 1 þ zf
PBH they enter the horizon. Thus, in this case, it may appear that the
probability of PBH formation is almost independent of their mass.
where f hor ¼ MPBH /Mhor (zf )  1 is the ratio of the mass of the Small-mass PBHs would have a larger probability of formation if
PBH to the mass of the particle horizon at z ¼ zf ,  is the density the spectrum were slightly blue (with ns > 1) and vice versa if the
parameter of PBHs at z ¼ zf , Mhor (zeq )  3:1 ; 1016 M is the spectrum were red, as seems to be indicated by available data
mass of the horizon at the redshift of matter-radiation equality (Spergel et al. 2007; Chongchitnan & Efstathiou 2007). Jedamzik
zeq  3000 (for
m h2 ¼ 0:127; Spergel et al. 2007). The rela- (1997) have shown that during a first-order phase transition
tionship (1 þ zf ) ¼
PBH (1 þ zeq ), valid for nonevaporating the cosmic equation of state may become softer (wT1/3). As
PBHs, relates the density parameter of PBHs today,
PBH (M ) ¼ a result, the formation of PBHs may be substantially enhanced
PBH /crit , to the one at the time of formation, (M ). For example, at that particular mass scale. PBHs with masses 1 M may form
if a tiny fraction   109 of the cosmic energy density collapses during the QCD (quark-hadron) phase transition at t  105 s, or
into PBHs during the quark-hadron phase transition at t  105 s, PBHs with mass 105 M may form during the eþ -e annihilation
it follows from equation (1) that about 30% of the present day era. Softening of the equation of state alone is unlikely to produce
dark matter is made of PBHs with mass MPBH  1 M (we as- interesting values for the abundance of PBHs. Relatively large
sumed f hor ¼ 1). Nothing prevents the dark matter from being amplitudes of density perturbations on the relevant scales are nec-
a mixture of weakly interacting particles (WIMPs) and PBHs essary. They can be produced in inflationary models (Yokoyama
with
dm ¼
PBH þ
WIMP . However, equation (1) shows that 1997, 1998a, 1998b) or in bubble collisions following second-
fine tuning of the value of  is required in order to have
PBH  order phase transitions (Steinhardt 1982; Crawford & Schramm

WIMP . Thus, it is often considered more likely that dark matter 1982; Hawking et al. 1982; La & Steinhardt 1989). For example,
is dominated by either WIMPs or PBHs. PBHs may have an ex- PBHs with masses of 100Y1000 M may be produced in two-
tended mass range if they form at different epochs, and, even if stage inflationary models designed produce a large and negative
they all form at one epoch, simulations show that their mass dis- running spectral index and a low WMAP quadrupole (Kawasaki
tribution is broad (e.g., Choptuik 1993; Evans & Coleman 1994). et al. 2006). One interesting feature of this model is a bump in
Thus, present-day dark matter may be composed of a mixture of the power spectrum at kpc scales with overdensity  > 1. Other
relatively massive and tiny PBHs, some of which may be com- possible scenarios for PBH formation involve non-Gaussian
pletely evaporated or have left Planck-mass relics. In practice, perturbations produced by cosmic strings and cosmic string or
using astrophysical tests only, sufficiently small PBHs would domain wall collapse (e.g., Polnarev & Zembowicz 1991; Rubin
be virtually indistinguishable from WIMPs. It is worth point- et al. 2001; Stojkovic et al. 2005; Nozari 2007). A more complete
ing out that no fine tuning of (zf ) is required to produce
dm ¼ discussion on the topic can be found in Carr (2006) and Mack et al.

PBH  1. Any value of   1 and zf is allowed, but, of course, (2007). It is generally considered unlikely that PBH formation
different values of the entropy of the universe and different red- occurred after t ¼ 1 s, when MPBH  Mhor k 105 M , because
shift of equivalence would be produced. Specifically, the radiation the physics in this domain is sufficiently understood and their
density parameter today is related to  and zf by the relationship formation would affect primordial nucleosynthesis.

rad;0 ¼
dm /(1 þ zeq )  1 (1 þ zf )1 . In addition, a scenario Observationally, the abundance of PBHs is well constrained
in which PBHs produced after inflation evaporate leaving only only at very small masses. PBHs with masses smaller than 5 ;
Planck mass relics may explain why
rad;0 h2  4:35 ; 105 and 1014 g should not exist today because they evaporate in less than a
provide a justification for the large value of the entropy of the Hubble time by emission of Hawking radiation (Hawking 1974).
universe (Alexander & Mészáros 2007). The radiation from evaporating PBHs affects nucleosynthesis and
PBHs are a unique probe of the early universe, of high-energy the CMB spectrum and may overproduce the observed gamma-
processes and of quantum gravity. The small-mass scales at which ray background. Hence, upper limits on the abundance of PBHs
PBHs may form are inaccessible to cosmic microwave back- with masses 1 g < MPBH < 1015 g are quite stringent: (M ) 
ground (CMB) experiments. As of today there is no solid evidence 1020 Y1022 (e.g., Carr 2003). The number density of PBHs with
for the existence of PBHs, but their presence would be very dif- masses larger than 1015 g is poorly constrained because they emit
ficult to detect even if they constitute the bulk of the dark matter. negligible amounts of Hawking radiation (their timescale for
Early results from the MACHO collaboration (Alcock et al. 2000) evaporation is longer than the age of the universe today). Dy-
suggested a possible detection of solar mass objects constituting namical constraints (Lacey & Ostriker 1985; Carr 1994; Moore
No. 2, 2008 EFFECT OF PBHs ON CMB 831

1993; Jin et al. 2005), effects on the matter power spectrum where rB  GM v2 2
eA is the Bondi-Hoyle radius and veA  (v þ
2 1=2
(Afshordi et al. 2003), and statistical studies of binary stars in cs ) (Bondi & Hoyle 1944). The numerical values of the Bondi
the Galactic halo (Yoo et al. 2004) have been used to rule out radius and accretion rate are
PBHs more massive than 1000 M as the main constituent of the   2
dark matter, but Mack et al. (2007, hereafter Paper I), summariz- M veA
rB  1:3 ; 10 pc
4
; ð2Þ
ing these constraints, found that a domain remains for which mas- 1 M 5:7 km s1
sive PBHs can make a significant contribution to the dark matter. ṀB  2 ; 1012 g s1 kngas
Searches for microlensing events toward the Large Magellanic  
Cloud by the MACHO and EROS collaborations so far provide M 2 veA 3
the most stringent constraints on the existence of nonevaporating ; 1
: ð3Þ
1 M 5:7 km s
PBHs. They have ruled out PBHs as the bulk of the Galactic dark
matter in the mass range 107 M < M < 30 M (Alcock et al. The mean cosmic gas density is
1998, 2001). Alcock et al. (2000) have claimed a positive detec-  3
tion of massive compact halo objects (MACHOs) with M  1þz
0:1Y1 M constituting about 20% of the Milky Way dark matter ngas ’ 200 cm3 : ð4Þ
1000
halo. More recent works by the EROS collaboration have not con-
firmed this result (Hamadache et al. 2006; Tisserand et al. 2007), Compton heating/cooling of the gas from CMB radiation keeps
but this mass range is of particular interest because it is about the temperature of the gas coupled to the radiation temperature.
the mass of PBHs that may form during quark-hadron phase tran- The gas temperature evolves as T / (1 þ z) even after recombi-
sition as proposed by Jedamzik (1997). The gray curves in Figure 9 nation at z  1000 due to residual ionization fraction xe  104 .
(left), below, summarize the current status of observational upper At redshift zd  100, due to the expansion of the universe, the
limits on the abundance of PBHs with MPBH > 1015 g. The two electron temperature decouples from the CMB and the gas tem-
black curves labeled ‘‘FIRAS’’ and ‘‘WMAP3’’ show the new perature decreases adiabatically [T / (1 þ z)2 ]. The temperature
upper limits derived in the present work. of the IGM (neglecting for the moment the feedback produced
by accreting PBHs) is well approximated by the fitting formula
2.2. PBH Growth and ‘‘Clothing’’ Dark Halo  
zþ1 ad
PBHs more massive than 105 M may only exist if substantial Tgas ¼ (2730 K )  1= ; ð5Þ
accretion takes place after their formation or if many PBHs which 1000
a  þ ad
form in a cluster merge into a larger one (Mészáros 1975a; Khlopov
et al. 2005; Chisholm 2006). In the Newtonian approximation, where a  1/(1 þ z) is the scale parameter, and ad is the scale
PBHs can only grow if their mass is comparable to the horizon parameter at photon-electron temperature decoupling where zd ’
mass. In this case their mass increases at the same rate as the 132(
b h2 /0:022)2 5 and  ¼ 1:72. The gas sound speed is cs ¼
=

horizon mass (Zel’Dovich & Novikov 1967). Relativistic cal- (5:7 km s1 )(Tgas /2730)1 2 . Thus, from equation (5) we have
=

culations by Carr & Hawking (1974) have shown that growth by


accretion is always slower than the growth of the horizon mass.  1=2
Thus, it is generally accepted that the mass of PBHs does not 1þz
cs  (5:7 km s1 ) for z 3 zd  132: ð6Þ
increase significantly after their formation. In models in which the 1000
equation of state becomes stiff w  1 (Lin et al. 1976) or in brane
world cosmologies in which extra dimensions are sufficiently For spherical accretion onto a point mass, assuming a nonviscous
large, growth by accretion could be significant and in principle fluid, the parameter k is of order unity. In the present study, we
may lead to the production of massive PBHs starting from smaller need to take into account the effects of the growth of a dark halo
seeds (Bicknell & Henriksen 1978; Guedens et al. 2002a, 2002b; around PBHs, the Hubble expansion and the coupling of the CMB
Majumdar 2003; Tikhomirov & Tsalkou 2005). radiation to the gas through Compton scattering (Compton drag).
Scenarios in which the bulk of the dark matter is not made of All these effects can be folded into the calculation of the accre-
PBHs are particularly interesting because, after the redshift of tion eigenvalue k. In a companion paper (Ricotti 2007, hereafter
matter-radiation equivalence, each PBH seeds the growth of a Paper II), we have derived analytical relationships for k for the
dark halo that, over time, becomes considerably more massive cases of a point mass and an extended dark halo, including the
than the PBH in its center (Paper I). The halo gravitational po- aforementioned cosmological effects. In the following calcula-
tential may increase the gas accretion rate onto the central PBH by tions we will use the results from Paper II to derive the accretion
several orders of magnitude. If the accretion is near the Eddington rate onto ‘‘naked’’ PBHs (appropriate only if PBHs constitute the
rate, the radiation efficiency of PBHs may be large. At later times, bulk of the dark matter) and for the general case of accretion onto
when PBHs are accreted by the massive halos of galaxies, they PBHs including the growth of a ‘‘clothing’’ dark halo of mass Mh .
likely lose much of their dark matter ‘‘clothing’’ and become To obtain this result, we must first estimate the dark matter ac-
‘‘naked’’ again. cretion (x 3.1) and the effects of PBH velocity (x 3.2), the angular
momentum (x 3.3) of the dark and baryonic matter and PBH clus-
3. BASIC EQUATIONS tering (x 3.4). From the gas accretion rate, we can then estimate
Here we review the theoretical framework for the accretion of the accretion luminosity, which will ultimately be responsible
dark matter and baryons by primordial black holes. A point mass, for altering the evolution of the IGM.
M, immersed in an hydrogen gas with constant number density
ngas and sound speed cs that moves at velocity v with respect to the 3.1. Growth of a Dark Halo around PBHs
gas accretes at a rate
If PBHs do not constitute the bulk of the dark matter, they seed
the accumulation of a dark halo of WIMPs which grows propor-
ṀB ¼ k4m H ngas veA rB2 ; tionally to t 2=3 (Bertschinger 1985; Mack et al. 2007). A similar
832 RICOTTI, OSTRIKER, & MACK Vol. 680

Fig. 1.—Left: Ratio of the baryonic to dark matter power spectrum as a function of wavenumber k. Each curve, from bottom to the top, corresponds to scale factors
a ¼ 0:001, 0.002, 0.005, 0.05, and 0.1, respectively. Right: The solid curve shows the mean relative velocity between the dark matter and baryons in a sphere of comoving
radius r0 . The dashed and dotted curves show the velocity dispersion within a sphere of comoving radius r0 for the baryons and for the dark matter, respectively.

result holds if the dark matter is made only of PBHs but with a of nonlinear perturbations dominates the motion of PBHs. In
very broad distribution of masses. x 3.2.2 we estimate the fraction of PBHs that are captured by
After the redshift of equivalence the mass of the dark halo galactic halos with mass >MPBH as a function of redshift. The
surrounding PBHs grow proportionally to the cosmic scale velocity of PBHs falling into the potential wells of these halos
parameter: is sufficient to suppress the accretion of gas onto those PBHs that,
  due to their large mass, are still accreting efficiently at low redshift.
z þ 1 1 Here, we do not model the contribution to the global emissivity
Mh (z) ¼ i MPBH ; ð7Þ of those massive PBHs that, due to dynamical friction, come to
1000
rest at the center of the galactic halos into which they are accreted.
where i ¼ 3 (see Paper I). The growth of the PBH during the
3.2.1. Linear Regime
radiation era is of order of unity. After z  30, depending on the
environment, the dark halo may stop growing: if a PBH evolves in In linear theory, gas and dark matter perturbations are coupled.
isolation it can continue to grow; otherwise, it will lose mass due Before the redshift of photon-electron temperature ‘‘decoupling’’
to tidal forces as it is incorporated into a larger galactic dark halo. at zd  100, the growth of gas inhomogeneities on small scales
If Mh > MPBH and the accretion is not perfectly spherical, the is suppressed by Silk damping (Silk 1968). At these scales and
dark halo has a self-similar power-law density profile  / r redshifts, the gas flow lags behind the dark matter until it is able
with   2:25 (Bertschinger 1985) truncated at a halo radius to catch up at later times. This process is illustrated quantitatively
in the left panel of Figure 1, which shows the ratio of the gas to
 1=3  1 dark matter power spectra in our fiducial CDM model. The
Mh 1þz
rh ¼ 0:019 pc ; ð8Þ curves, from bottom to the top, show the ratio of the power spectra
1 M 1000
of the baryons to dark matter calculated using the code ‘‘lingers’’
in the Graphic1 package (Bertschinger 1985) at scale parameters
where rh is about one-third of the turnaround radius ( Paper I a ¼ 0:001, 0.002, 0.005, 0.05, and 0.1, respectively.
and II ). From the power spectra of dark matter and baryons we calculate
the ensemble average of the center-of-mass velocity of a patch of
3.2. Proper Motions of PBHs the universe of comoving radius r0 (Ostriker & Suto 1990):
The motion of PBHs relative to the surrounding gas strongly Z
reduces the Bondi accretion rate (see eq. [3]). This proper motion H2 1 2
hVi i2 ¼ 2 f (k; t)Pi (k)w 2 ðk; r0 Þ dk; ð9Þ
is determined by the relative amplitude of inhomogeneities of 2 0 i
dark matter and gas. More precisely, assuming that PBHs behave
like dark matter particles and neglecting binary interactions, we where w is the ‘‘window function’’ (here we use ‘‘top-hat’’ win-
estimate PBHs relative velocity with respect to the gas in the linear dow function), and we have used the growth factor of cold dark
and nonlinear regimes. We show that at redshifts z k 30Y50, due matter perturbations in a flat universe.
to the imperfect coupling of gas and dark matter perturbations The index i ¼ dm; b refers to dark matter and baryons, re-
(i.e., Silk damping), the peculiar velocity of PBHs with respect spectively. For the dark matter the function fdm (k; t) ¼ 1 and for
of the gas is of the order of the gas sound speed. At redshifts the baryons fb (k; t) accounts for the faster rate of growth of the
z P 5Y10, depending somewhat on the mass of PBHs, the growth baryonic perturbations at z < 1000 as they catch up with the dark
No. 2, 2008 EFFECT OF PBHs ON CMB 833

matter. We estimate the importance of fb (k; t) from Figure 1 as-


suming, for simplicity, that Pb (k; t)/Pdm (k; t) ¼ 1 for k < keq and
Pb (k; t)/Pdm (k; t) ¼ const(t) for k  keq. It turns out that the in-
crease of the value of hVb i due to fb (k; t) ¼
6 1 is negligible. This is
because at z  1000, when the correction factor fb (k; t) at k  keq
is large, the integral in equation (9) is dominated by the growth
of the largest scale perturbations for which fb (k; t) ¼ 1. At lower
redshifts the correction factor fb (k; t) becomes close to unity.
The ensemble average of the velocity variance within a patch
of comoving radius r0 is calculated in a similar fashion:
Z
2 H2 1 2
hi i ¼ 2 f (k; t)Pi (k)½1  w 2 (k; r0 ) dk: ð10Þ
2 0 i
This equation will be used later (x 3.3) to estimate the angular
momentum of gas and dark matter accreting onto PBHs.
Since the flow of the gas and the dark matter trace each other,
their mean relative velocity is hVrel i  hVdm i  hVbar i. The right
panel of Figure 1 shows the mean relative velocity hVrel i of PBHs
with respect to the baryons (solid curves), and the velocity dis- Fig. 2.—Luminosity-weighted effective velocity of PBHs (thick curves): hveA iA
persion hdm;b i for the gas (dashed curve) and the dark matter (solid curve) is weighted assuming l / ṁ2 , and hveA iB (dashed curve) is weighted
(dotted curve) within a sphere of comoving radius r0 at z ¼ 1000. assuming l / ṁ. See the text for details. The thin curves show the variance of
Power-law fits for hi as a function of r0 and z are the velocity distribution, hVrel i (dashed line) and the gas sound speed, cs (dotted
line), respectively.
 0:85  
  r0 1 þ z 1
b  0:35 km s1 ; ð11Þ for  ¼ 3. The calculations presented in this section on the effect
1 Mpc 1000
 0:85   of linear perturbations on the accretion luminosity of PBHs are
 1
 r0 1 þ z 1=2 summarized in Figure 2. The thick solid and dashed curves show
dm  1:58 km s : ð12Þ hveA iB from equation (15), and hveA iA from equation (14), respec-
1 Mpc 1000
tively, as a function of redshift. For comparison, the thin dashed
The fits are accurate to about 5% for r0 < 1 Mpc in the redshift and dotted curves show hVrel i and the sound speed of the gas,
range 50 < z < 2000. respectively. Neglecting feedback effects, the PBH velocity is
In x 3.6 we discuss in detail the radiative efficiency of accreting comparable to the gas sound speed at redshifts z P 500. On av-
PBHs. Here we anticipate that the radiative efficiency and the erage, the motion of PBHs in the linear regime reduces the gas
accretion luminosity L depend on the gas accretion rate ṁ: L / accretion rate by a factor of a few with respect to the static case.
ṁ 2 / v6 3
eA if ṁ <1 (spherical accretion) and L / ṁ / veA if ṁ k1 3.2.2. Nonlinear Regime
(thin-disk accretion), where we have used the expression for the
Bondi accretion rate and effective velocity veA defined in x 3 (see The velocity of PBHs falling into the gravitational potential
eq. [3]). The effect of the linear velocity field, which is Gaussian, of nearby galactic halos can be roughly estimated assuming that
on the mean accretion luminosity of PBH is given by the statis- it is comparable to the circular velocity of virialized halos of
tical ensemble average of v eA weighted by the distribution func- 2  density perturbations: vp  vc (2 ; z). We calculate vc (2 ; z)
tion of the PBHs relative velocities with respect to the gas: using the Press-Schechter formalism (Press & Schechter 1974).
Z 1 Adopting a top-hat window function to calculate the variance of
 fM (v; ) dv density perturbations, we find (M )  10:2  0:79 log (M ) for
hveA i ¼ ; ð13Þ
0 (cs2 þ v 2 )=2 masses M < 1010 M . Hence, in the mass range of interest,
 
where  ¼ 3 or 6, depending on the value of the accretion rate. M2  ¼ 8:8 ; 1012 M exp½1:8ð z þ 1Þ; ð16Þ
The linear velocity field is Gaussian; thus, the distribution of the
moduli of the 3D velocity field is a Maxwellian, fM (v; ) with  ¼ is a sufficiently accurate approximation for the mass of 2  per-
hVrel i. Finally, integrating equation (13) and defining MPBH ¼ turbations as a function of redshift. From equation (16) we cal-
hVrel i/cs , we find culate the circular velocity and thus typical proper velocity of
8  1=6 PBHs induced by nonlinear structures, as a function of redshift:
>
<hc i p16 3    
s ffiffiffiffiffiffi MPBH for MPBH > 1; 1 M2 1=3 z þ 1 1=2
hveA iA  2 ð14Þ vp  vc ¼ (17 km s ) : ð17Þ
>
: 108 M 10
2 1=2
cs (1 þ MPBH ) for MPBH < 1;
The fraction of the dark matter and hence of PBHs pffiffiffi in virialized
for  ¼ 6 and halos with mass >Mmin is fvir (z) ¼ 1  erf (
min / 2), where erf is
the error function and
¼ c /(Mmin ; z). We assume that virialized
hveA iB  halos with mass larger than Mmin ¼ maxðMPBH ; M2  Þ can reduce
8 "rffiffiffiffi  #1=3
> the accretion onto PBHs of mass MPBH . Thus, we assume that a
>
<cs MPBH 2 2
fraction fvir (z; Mmin ) of PBHs have velocity vp relative to the gas
ln MPBH for MPBH > 1;
 e ð15Þ and accrete gas with overdensity  ¼ 200. The remaining PBHs
>
>
:  1=2 are in the uniform density intergalactic medium and are subject
cs 1 þ M2PBH for MPBH < 1; to global and local thermal feedback discussed in x 4.
834 RICOTTI, OSTRIKER, & MACK Vol. 680

Dynamical friction may allow the most massive PBHs to spiral tivistic corrections. We conclude that the angular momentum
to the center of the halos in less than a Hubble time (see x 7), of the gas is negligible when the halo mass is
where they meet favorable conditions for accretion. For smaller  2:35
PBHs inside virialized halos, their contribution to the accretion 1:17 4:33 1 þ z
luminosity is important only in the redshift range 5 P z P 30. At Mh < (1746 M )D (z) : ð20Þ
1000
z k 30 the fraction of PBHs in virialized halos of mass >1 M
is negligible. At z P 5, the several massive halos have formed Using MPBH (z) calculated in x 3.2.1 and D ¼ 10, the flow is
and the rate of gas accretion onto PBHs embedded in these halos quasi-spherical for any halo mass Mh P 1700Y17;000 M , al-
declines rapidly due to the increasing relative velocity of the PBHs most independently of redshift. At redshifts z < 100, the dark
with respect to the gas: vrel k 10 km s1 for Mdm k 108 M. halo mass is about 30 times the mass of the PBH in its center. We
Adopting WMAP3 cosmological parameters (8 ¼ 0:74, ns ¼ conclude that an accretion disk forms at redshifts z < 100 around
0:95), at z  10 only a fraction 10%Y20% of PBHs is inside PBHs more massive than MPBH  500 M . Thermal feedback
virialized halos of mass >1 M and a few percent in halos with effects, which are important at z < 100, will increase our simple
mass k106 M . For the purpose of our calculations most PBHs at estimate of the critical mass for disk formation. So far we did
z  10 are accreting from a uniform density intergalactic medium not take into account feedback effects and Compton drag which
and are not affected by collapsed dark matter halos. Nevertheless, are important in reducing the gas angular momentum. Previous
the gas accretion onto the PBHs in the intergalactic medium is studies have found that due to Compton drag the flow is nearly
strongly suppressed at z < 10 due to global thermal feedback spherical at redshifts z > 100 (e.g., Loeb 1993; Umemura et al.
(i.e., X-ray heating; see Fig. 6, below). 1993).
Similar arguments apply to the dark matter component. The
3.3. Angular Momentum of Accreted Material
flow is sufficiently spherical for the dark matter to be directly
The angular momentum of the accreting gas determines whether accreted onto the PBH if dm < cDRSch /rh . This inequality is
a disk forms or the accretion is quasi-spherical. In turn, the ge- satisfied for PBHs with masses
ometry of gas accretion determines the radiative efficiency (ṁ),
i.e., the efficiency of conversion of gravitational energy into ra-  
1 þ z 2:9
diation (see x 3.6). The angular momentum of accreting dark MPBH > (7:18 ; 105 M )D2:6 ;
matter determines the density profile of the dark halo enveloping 1000
a PBH and whether the mass of a PBH can grow substantially  0:9
MPBH 2:6 1 þ z
by accreting a fraction of its enveloping dark halo. The angu- > (7:76 ; 10 )D
13
; ð21Þ
Mhor 1000
lar momentum of the gas and dark matter accreting onto PBHs
can be estimated from equation (12) for the mean values of the
velocity dispersion within a comoving volume of radius r0 (see in terms of the horizon mass Mhor (z). Including relativistic cor-
the right of Fig. 1). For the gas we estimate b (rB;com ) at the rections (D  10), only PBHs with mass k103 M at z  1000
Bondi radius (see eq. [3]), and for the dark matter we estimate (note the steep dependence of the critical mass on redshift) may
dm (rh;com ) at the turnaround radius (see eq. [8]). If we neglect grow substantially from direct accretion of dark matter. The ge-
the proper motion of PBHs, the Bondi radius of a ‘‘naked’’ ometry of the dark halo becomes more spherical with the increas-
PBH is approximately constant at redshifts z < 100: rB;com  ing mass of the PBH, since the ratio of the halo radius to the
(1:3 ; 107 Mpc)(MPBH /1 M ). The proper motion of PBHs 2=3
Schwarzschild radius, rh /RSch / MPBH , decreases with increas-
produces a small reduction of the Bondi radius that we parameter- ing mass. We conclude that direct accretion of dark matter into
ize with the function (z) ¼ max ½MPBH (z); 1, where MPBH ¼ PBHs is typically negligible. In most cases the density profile of
hveA i/cs is the PBH Mach number. Replacing Mh with the mass the dark halo is well described by the Bertschinger self-similar
of the PBH in the previous expression, we obtain an estimate of solution with log slope  ¼ 2:25.
the Bondi radius including the enveloping dark halo. Finally,
from equation (12) we obtain 3.4. PBH Clustering
 1   Recently it has been suggested by several authors (e.g.,
1:7 1 þ z Mh 0:85
b  b;0 (z) ; ð18Þ Dokuchaev et al. 2004; Carr 2006; Chisholm 2006) that PBHs
1000 1 M have high probability of forming clusters or binaries. A precise
    calculation of the effect of clustering is complex and beyond the
1 þ z 1=2 Mh 0:28
dm  dm;0 ; ð19Þ scope of the present paper. Nevertheless, there are simple argu-
1000 1 M
ments that we can use to show that clustering may increase the
radiative efficiency of PBHs.
where b;0 ¼ 3:8 ; 107 km s1 and dm;0 ¼ 1:4 ; 104 km s1 .
We consider two regimes determined by the relationship be-
Let us first consider the accretion geometry of the gas. Ap-
tween the typical size of the cluster (or the binary separation), rcl ,
plying conservation of angular momentum we find that the ro-
and the Bondi radius of the whole system, r B;tot .
tational (i.e., tangential) velocity of the gas at a distance r from
the black hole is v(r)r  b rB , where rB is the Bondi radius. 1. If rcl > r B;tot , the orbital velocities of the PBHs are smaller
Neglecting relativistic effects, if the velocity is much smaller than the effective translational velocity of the system. Hence, we
than the Keplerian velocity in the proximity of the black hole, can neglect the orbital velocities to estimate the accretion rate.
then the accretion is quasi-spherical; if vice versa, a disk can To a first approximation in this case, we can ignore the fact that
form. The Keplerian velocity at radius r is vKep ¼ c(RSch /r)1 2 < c,
=
the PBHs are clustered or in binary systems for the purpose of
2
where RSch ¼ 2GM /c is the Schwarzschild radius of the PBH calculating their accretion luminosity.
(RSch  3 km for a 1 M black hole). Thus, the gas accretion 2. If rcl < rB;tot , the orbital velocities of the PBHs are larger
is quasi-spherical if b < 2D (z)2 cs2 /c. We have assumed cs / than the effective translational velocity of the system. Thus, the
(1 þ z)1 2 and the constant D  1Y10 takes into account rela-
=
material is not accreted directly onto the PBHs but near the center
No. 2, 2008 EFFECT OF PBHs ON CMB 835

Fig. 3.—Dimensionless accretion rate of baryonic matter onto a ‘‘naked’’ PBH (without enveloping dark halo) from the IGM as a function of redshift. The curves from
bottom to top refer to MPBH ¼ 1, 10, 100, 300, 103, 104, and 105 M . For the sake of simplicity, we have assumed a toy model for the thermal and ionization history:
instantaneous recombination at z ¼ 1000, instantaneous reionization at z ¼ 7 (reheating temperature of 20,000 K), and constant ionization fraction, xe , between
recombination and reionization. Left, For a gas with electron fraction xe ¼ 103 ; right, for xe ¼ 1. In these plots, feedback and heating by PBHs and accretion onto PBHs
falling inside virialized halos are neglected. The motion of the PBH due to linear density perturbations is included.

of mass of the system and subsequently onto the PBHs. Because where the dimensionless sonic radius (xcr  rrc /rB ) is xcr ¼
the PBHs are orbiting the center of mass of the cluster, the angular ½1 þ (1 þ ) ˆ The effective gas viscosity, eA ¼  þ H,
ˆ 1=2 /.
momentum of the accreted material is large and so the accretion is the sum of two terms. The first term, due to Compton drag, is
geometry is disklike rather than spherical. The formation of a disk (z) ¼ 2:06 ; 1023 xe (1 þ z)4 s1, and the second, due to the
can increase the radiative efficiency by a factor of 10 (see x 3.6). cosmic expansion, is the Hubble parameter H(z). It is easier to
Further work is needed to better understand the accretion rate of understand the physical meaning of the dimensionless viscosity
PBHs in this regime. due to the Hubble expansion in terms of the recession velocity
of the gas due to the Hubble flow at the Bondi radius: vHub 
3.5. Gas Accretion Rate r B H(z). As expected, if v Hub /cs ¼ ˆHub T1 we can neglect the
We now have all that is needed to estimate the accretion rate onto cosmic expansion. It turns out that in the cases in which the vis-
PBHs from equation (3). We define the dimensionless accretion lu- cous term due to cosmic expansion is important, the assumption of
minosity l  Lbol /LEd, where LEd  1:3 ; 1038 (MPBH /1 M )erg s1. steady state accretion fails and the ‘‘Bondi-type’’ solution tran-
The radiative efficiency determines the accretion luminosity for a sitions to a self-similar solution studied in detail by Bertschinger
given accretion rate: Lbol ¼ Ṁb c 2 . Thus, defining the Eddington (1985). This transition happens for relatively massive PBHs with
accretion rate as ṀEd  LEd /c 2 ¼ 1:44 ; 1017 (MPBH /M ) g s1 M > xx (Paper II). The dimensionless viscosity depends on the
and the dimensionless accretion rate as ṁ  Ṁb /ṀEd , we have redshift, the mass of the PBH, and the temperature and ionization
l ¼ ṁ. fraction of the cosmic gas:
The dimensionless Bondi-Hoyle accretion rate for a PBH with-   
out a dark halo is ˆ MPBH z þ 1 3=2  cs 3

104 M 1000 5:74 km s1
" #
  
1 þ z 3 MPBH  veA 3  x  z þ 1 5=2
e
ṁ ¼ (1:8 ; 103 k) 1
: ð22Þ ; 0:257 þ 1:45 : ð24Þ
1000 1 M 5:74 km s 0:01 1000

This equation is valid only if PBHs constitute all the dark matter: Figure 3 shows the accretion rate ṁ from equations (22)Y(23) as
fPBH ¼
PBH /
dm ¼ 1. If the accreting object is a point mass a function of redshift. The curves from bottom to top refer to
(i.e., a ‘‘naked’’ PBH ), k depends only on the value of the di- MPBH ¼ 1, 10, 100, 103, 104, and 105 M . We assume a gas with
mensionless gas viscosity ˆ ¼  eA rB /cs due to Compton drag and constant electron fraction xe ¼ 103 after recombination in the
ˆ
the gas equation of state: if T1 we find k  1 and if ˆ 31 we left panel and xe ¼ 1 in the right panel. Thermal feedback is ne-
ˆ
have k  1/. The fit to the accretion eigenvalue for an isothermal glected and recombination is instantaneous at z ¼ 1000.
gas is given by the formula (see Paper II for an extensive study on If f PBH < 1 we need to consider the effect of a growing dark
cosmological spherical accretion in a viscous gas) matter halo with mass Mh (z) ¼ i MPBH ½(1 þ z)/10001 . For
i ¼ 3 we find
    
9=2 1þz MPBH  veA 3
k ¼ exp xcr2 ; ð23Þ ṁ ¼ (0:016k) : ð25Þ
3 þ ˆ0:75 1000 1 M 5:74 km s1
836 RICOTTI, OSTRIKER, & MACK Vol. 680

Fig. 4.—Same as in Fig. 3, but including the growth of the dark halo surrounding the PBH ( ¼ 2:25 and i ¼ 3). The curves from bottom to top refer to MPBH ¼ 1,
10, 100, 300, 103, 104, and 105 M .

In the simulations presented in x 5, we impose that the dark halo rates ṁ k 1. If the accretion rate is ṁ P 1 the radiative efficiency
stops growing when all the available dark matter has been ac- is smaller by a factor / ṁ (e.g., Park & Ostriker 2001). De-
creted, i.e., when f PBH i ½(1 þ z)/10001 ¼ 1. The gas accretion pending on the angular momentum of accreted material, a thick
onto an extended dark halo of mass Mh and power-law density disk may form. Otherwise the accretion geometry is spherical.
profile with logarithmic slope 2 <  < 3 is related to the one The later case is the most conservative as the accretion efficiency
for a point mass (with MPBH ¼ Mh ) by the following scaling is minimum. In addition, this case is appropriate for most PBH
relationships ( Paper II): masses which have ṁ < 1 (see x 3.3). We consider the following
accretion regimes:
( ) ( )ˆ ( ) ( )
ˆ h  p= 1p ; k h  p= 1p k; Thin disk.—If ṁ > 1 we assume that the gas accreted by the
 p=(1p) PBH forms a thin accretion disk and has radiative efficiency
( )
rcrh  rcr ; ð26Þ ¼ 0:1. We assume that the bolometric luminosity cannot exceed
2
the Eddington limit and that, due to complex feedback effects
where ¼ r B /rh < 2, p ¼ 3  , and that we do not attempt to model here, the PBHs accrete near the
Eddington limit only for a fraction of time fduty . Observations
   2 of AGN at redshifts z < 3 show that typically fduty is about 3%
( ) 1=10
 ¼ 1 þ 10ˆ h exp (2  ) : ð27Þ (e.g., Shapley et al. 2003). However, for PBHs, the duty cycle
2
can be substantially different as the nature of the feedback mech-
When the Bondi radius rB is larger than twice the radius of the anisms which regulate the cycle are not well understood. We let
dark halo (  2), the dark halo behaves the same as a point fduty be a free parameter and we show that the limits on the PBH
mass in terms of accretion rate, sonic radius, and dimensionless abundance scale linearly with it. In summary,
viscosity. The numerical value of is
l ¼ f PBH min(0:1ṁ; 1) if ṁ > 1: ð29Þ
 2=3   2
Mh 1þz cs
¼ 0:22 : ð28Þ Figures 3Y4 show that this choice of the radiative efficiency is
1 M 1000 1 km s1 appropriate for PBHs with masses MPBH k 100 M . The spec-
trum of the radiation emitted by the accreting material depends on
We adopt the parameters  ¼ 2:25, i ¼ 3, and rh given in the accretion geometry, accretion rate, and the mass of the black
equation (8), valid for a quasi-spherical dark halo in which the hole. For a thin disk, any given radius emits blackbody radiation
dark matter does not fall directly onto the PBH. Figure 4 shows with different emission temperature (multicolor disk). The disk
the accretion rate, ṁ, from equations (25)Y(26) as a function of is hot in the inner parts and colder in the outer parts. The max-
redshift, with the curves representing the same masses as Figure 3 imum temperature of the disk depends on the black hole mass:
with the same simplifying assumptions going into the simulation. 1=4
Tmax / MPBH ṁ. Beyond photon energies kTmax the emission is
dominated by nonthermal radiation produced by a diffuse hot
3.6. Accretion Luminosity corona around the accretion disk. The spectrum of the nonthermal
The radiative efficiency of accreting black holes is known to component is a power law of the form
L
/
 with   1:5.
depend on the accretion rate ṁ and the geometry of the accretion Guided by observations we model the spectrum as a double power
flow. Observations and theoretical models suggest that the radi- law with a break at energies h
¼ kTmax with slope  ¼ 0:3 at
ative efficiency approaches a constant value  0:1 for accretion low energy and  ¼ 1:6 at high energy.
No. 2, 2008 EFFECT OF PBHs ON CMB 837

Spherical accretion or ADAF.—If ṁ < 1 the accretion efficiency radiation background that increases the ionization fraction of the
depends on the flow geometry. If the infall is quasi-spherical the gas and heats the IGM. The UV radiation, on the other hand, is
radiative efficiency is minimal. The dominant emission mech- absorbed not very far from the emitting sources, so UV photons
anism is thermal bremsstrahlung. Most of the radiation will orig- produce spheres of fully ionized hydrogen around each PBH.
inate from the region just outside the event horizon. For the case These processes introduce global and local feedback effects which
of accretion from a neutral H i gas, including relativistic effects, may alter the accretion rate. Here we describe these effects and
the efficiency for conversion of rest-mass energy into radiation comment on their importance for the accretion calculation.
is ¼ 0:011ṁ (Shapiro 1973a, 1973b). Hence,
4.1. Global Feedback
2
l ¼ 0:011ṁ if ṁ < 1 (spherical accretion): ð30Þ We use a semianalytic code, described in detail in Ricotti &
Ostriker (2004) and based on Chiu & Ostriker (2000), to follow
The spectrum is well approximated by a power law of the form the chemical, ionization, and thermal history of the IGM from

L
/
0:5 at
> 13:6 eV and has an exponential cutoff at
 recombination to the redshift of the formation of the first galaxies.
5 ; 105 eV (Shapiro 1973a). This case gives the most conser- We consider a gas of primordial composition and include H2
vative estimate of the effect of PBHs on the cosmic ionization formation /destruction processes. Although the code includes a
history. In addition some authors have found that when the effects stellar reionization and galaxy formation formalism, in the present
of magnetic field and cosmic rays are included in the calculations, paper we focus on the cosmic epochs preceding the formation
the radiative efficiency can be larger than the 0:01ṁ assumed in of the first galaxies. In addition to the emission from PBHs we
our fiducial case (Mészáros 1975b). include the contribution from the CMB radiation background that
If the gas has nonnegligible angular momentum, an advection gives us a redshift of recombination zrec  1000, in good agree-
dominated accretion flow (ADAF) may form (Narayan & Yi ment with observations.
1995). In this case the radiative efficiency is a factor of 10 larger: The thermal feedback by the X-ray background is calculated
¼ 0:1ṁ. Hence, l ¼ 0:1ṁ2 . This case is not particularly rele- self-consistently: the temperature and ionization of the cosmic
vant for our study as we have seen in x 3.3 that only PBHs with gas is determined by the luminosity of PBHs, and the PBH ac-
masses >1000 M can form accretion disks and typically for cretion rate is a function of the gas temperature and ionization.
these masses we find ṁ > 1. The thermal and chemical history is calculated iteratively until
3.6.1. Is the Emerging Radiation Trapped by the Accreting Gas? the solution converges. The total emissivity per unit mass from
accretion onto PBHs is proportional to lf PBH . At z < 30, when
Assuming spherical accretion and negligible pressure with the age of the universe is t  107 yr, accretion onto PBHs stops
respect to the gravitational potential, the gas falling onto PBHs due to their peculiar velocities. Even if we assume that PBHs
acquires a velocity that approaches the free fall velocity near the accrete at the Eddington rate before z ¼ 30, the accretion time-
black hole.1 From mass conservation it follows that the gas den- scale tSalp ¼ 4:4 ; 108 yr is much larger than the age of the uni-
sity profile is g / r1:5 in the inner parts. Hence, the gas column verse at this redshift. Hence, the growth of PBHs by gas accretion
density Ng / r1=2 diverges for radii r ! 0 and so does the is negligible.
Compton scattering optical depth for outgoing radiation.
We checked whether the accreting gas may become opaque 4.2. Local Feedback and Duty Cycle
to Compton scattering. In this case, the infalling gas would trap In addition to tracking the X-ray background, the semianalytic
the emerging X-ray photons (which are mostly emitted near the code simulates the evolution of the Strömgren spheres produced
black hole horizon) or the X-ray photons may be reprocessed by UV radiation. These ionized bubbles have a small volume
into lower energy photons. The optical depth for emerging ra- filling factor and thus a negligible effect on the ionization history.
diation emitted at a distance rmin ¼ RSch from the black hole, Nevertheless, the gas is ionized and heated by UV radiation within
Z the H ii regions, leading to local feedback effects that may reduce
T rmax ṁ
e ¼ g dr ¼ 1=2 ; ð31Þ the gas accretion rate.
mp rmin  Local feedback may reduce the gas accretion rate if : (1) the
radius of the H ii region exceeds the Bondi radius (r H ii > rB ),
where   1 is the distance in units of the black hole Schwarzschild and (2) if the gas temperature inside the H ii region is higher than
radius and where we have assumed rmax 3 rmin . Hence, for any the temperature outside. Although this is always the case at low
value of the accretion rate smaller than ṁ   1=2  1 the gas is
redshifts, at z > 30 the Compton coupling between the free elec-
optically thin to Compton scattering. Since we assume that PBHs trons inside the H ii regions and the CMB radiation is very ef-
with ṁ > 1 form an thin accretion disk, we can safely neglect fective in keeping the gas temperature near the CMB temperature.
trapping of radiation emitted near the black hole in all the models
If the final size of the H ii region is larger than the Bondi radius
we consider. Similar calculations show that the gas accreting onto (r H ii /rB > 1) and the temperature inside the H ii region is much
PBHs is transparent (to Compton scattering) to external back- larger than that outside, then the gas accretion rate may stop or
ground radiation.
decrease. For simplicity, let us consider the most extreme scenario,
4. FEEDBACK PROCESSES in which when r H ii /rB  1, the luminosity becomes negligible
and the H ii region recombines. With this assumption the time-
Gas accretion onto PBHs produces radiation that heats and averaged luminosity can be estimated to be
ionizes the IGM, thus raising its temperature above the value given
in equation (5). X-ray photons, having a mean free path larger than l l
the mean distance between the sources, tend to build up a uniform hlit ¼ ¼ ¼ fduty l; ð32Þ
1 þ toA =ton 1 þ (r H ii =rB )1=3
1
For realistic cases in which the angular momentum of accreting gas is im-
portant we refer to previous well-known studies of hot optically thin accretion where toA ¼ trec;H is the H i recombination timescale and ton ¼
flows (Shapiro et al. 1976; Narayan & Yi 1995). trec;H (rB /r H ii )1=3 is the timescale it takes for the ionization front to
838 RICOTTI, OSTRIKER, & MACK Vol. 680

reach the Bondi radius r B . Thus, it follows toA /ton ¼ (r H ii /r B )1 3 .


=

Equation (32) gives a rough estimate of the duty cycle produced


by local feedback.
4.2.1. Temperature Structure of the H ii Region
We have performed time-dependent 1D radiative transfer sim-
ulations of the H ii regions around PBHs to estimate r H ii and the
temperature TH ii inside the Strömgren radius. The basic results
can be also derived analytically. The radius of the Strömgren
sphere around a PBH of mass MPBH at the center of a spherical
gas cloud with density profile ngas ¼ nH0 (r/r0 ) , where nH0 (z) is
the mean cosmic density of the gas, is
 1=3  2
S0;49 1þz
r H ii ¼ (0:77 pc) ; ð33Þ
f 1000

where f ¼ 1/3 for ¼ 0 (i.e., assuming gas at uniform density)


or f  ln (r0 /rmin )  10 assuming a free falling gas that be-
comes collisionally ionized at rmin (i.e., we adopt ¼ 1:5 and
r0 /rmin  104 ). We have defined the quantities Fig. 5.—Temperature structure of the H ii region around a PBH at z ¼ 500
emitting S0 ¼ 1052 ionizing photons s1. The curves show the temperature
profile after 2 yr (dashed line), 100 yr (dotted line), and 4600 yr (solid line) after
  1 the source turns on. The source spectrum is one appropriate for spherical accretion
MPBH X
S0;49  (6 ; 10 )l3
; ð34Þ onto a black hole, with log slope  ¼ 0:5.
1 M 100

the number of ionizing photons emitted per second in units of only 56% (41%) higher than the temperature outside, which is
1049 photons s1, and roughly equal to the CMB temperature TCMB  1375. We con-
clude that at high redshift, even if the Bondi radius is smaller
 Z 1 than the Strömgren radius, the increase of the IGM temperature
1
h
0 L
due to local feedback is negligible. Hence, we can neglect the
X  d
; ð35Þ
Lbol
0 h
local contribution to thermal feedback (from UV photos) with
respect to the global thermal feedback from X-ray heating, that
is always included self-consistently (see x 4.1)
the mean energy of the emitted photons in Rydberg units (h
0 ¼
More sophisticated radiative transfer simulations confirm the
13:6 eV ). Assuming a power-law spectrum with index  ¼ 0:5
results of the analytical calculations of r H ii and TH ii . In Figure 5
in the frequency interval 13:6 <
< 5 ; 105 eV, appropriate for
we show the time evolution of the temperature profile inside an
spherical accretion, we find X  192. For thin disk accretion we
H ii region around a PBH with S0 ¼ 1052 s1 at z  500. The 1D
find X ¼ /(  1)  3 for  ¼ 1:5. In order to estimate the
radiative transfer code used to produce the temperature plots is
value of the Strömgren radius we assumed an H i recombination
described in detail in Ricotti et al. (2001). In addition, in the
coefficient  ¼ 1:28 ; 1012 cm3 s1, appropriate for a gas at
present simulation we have also included Compton heating from
T  2000 K. This temperature is typical for the gas inside an H ii
X-rays emitted near the PBH. The increase of the gas temper-
region at z  500.
ature at small radii evident in Figure 5, T / r2 , is produced by
We now estimate the temperature inside the H ii region as-
Compton heating. We estimate that Compton heating becomes
suming that the dominant gas heating is H i photoionization and
dominant over photoionization heating at radii
the dominant cooling is Compton cooling. In a uniform density
gas in ionization and thermal equilibrium the temperature inside  1=2  
the H ii region is constant. At equilibrium the ionization rate X kT 1=2
r< (T S0 trec;H )1=2
per hydrogen atom equals the recombination rate: tion ¼ trec;H  hh
i me c 2
(124 yr)½(1 þ z)/10003 , and the heating rate is hh
i/tion , where  3=2
1=2 1 þ z
hh
i  0:365 eV/(  þ 2) is the mean photon energy deposited  (0:01 pc)S0;49 : ð37Þ
into the gas per hydrogen atom. Neglecting Compton heating, 1000
which is only important in a small volume around the black hole,
a fraction C  0:0268/( þ 2)X of the radiation emitted by the After a time-dependent phase during which the H ii region reaches
black hole is deposited as heat inside the H ii region. If we assume its Strömgren radius, the temperature inside the H ii region de-
that the heating rate equals the Compton cooling rate k(TH ii  creases to a constant value TH ii  2100 K, about 56% higher
TCMB )/tComp , where tComp  (0:76 yr)½(1 þ z)/10004 , we find than the temperature outside the H ii regions, in agreement with
the analytical estimate. We have verified that the temperature TH ii
 2 is independent of the source luminosity and depends on  and the
TH ii 0:36 1þz
1þ : ð36Þ redshift according to equation (36).
TCMB þ2 1000
4.2.2. Strömgren versus Bondi Radii
Thus, for spherical accretion with  ¼ 0:5 (disk accretion with Let us now estimate the ratio r H ii /rB , which, along with the
 ¼ 1:5) the temperature inside the H ii regions at z  500 is H ii region gas temperature, determines whether local feedback
No. 2, 2008 EFFECT OF PBHs ON CMB 839

will reduce the gas accretion rate. From equation (3) and equa- feedback, the ionized gas inside the H ii region increases the
tion (33) with f ¼ 10, we find gas viscosity due to Compton drag. Compton drag is negligible
 1=3   at z < 100 (see Figs. 3Y4), but reduces the accretion rate at
r H ii l MPBH 2=3 z > 100 when, instead, thermal feedback is negligible.
 3 ; 103
rB X 1 M 5. SIMULATING THE COSMIC IONIZATION
 2  2
1þz veA We simulate the ionization, chemical, and thermal history of
; 1
; ð38Þ
1000 5:7 km s the universe after recombination using a modification of the semi-
analytic code discussed in detail in Ricotti & Ostriker (2004).
for ‘‘naked’’ PBHs. If we include the growth of an extended dark The original code is tailored to simulate UV and X-ray reioniza-
halo, r H ii /rB decreases by a factor 0:3½(1 þ z)/1000 at z < 1000. tion from high-redshift galaxies at z < 30. Here we do not include
We shall consider two cases for the accretion luminosity: case A, any ionizing source other than PBHs. In order to simulate standard
appropriate for small accretion rates ṁ < 1 and case B, appro- recombination properly we model the redshifted CMB blackbody
priate for ṁ  1Y10 (see x 3.6). radiation including the effect of Thomson opacity. We test the
recombination history calculation in the absence of PBHs
Case A.—l / ṁ2 . The case l / ṁ2 is the most interesting
against a widely used code RECFAST (Seager et al. 1999). We
because it is the most likely to occur for PBHs and because the
can reproduce the recombination history at redshifts z < 900
ratio r H ii /rB is independent of the gas effective velocity veA and
(when xe < 0:1) with accuracy of a few percent. Using the equa-
PBH mass. Indeed, r H ii / l 2=3 / v2 2
eA and rB / veA . Assuming tions derived in this work we model the UV and X-ray emission
¼ 0:01 (spherical accretion) we find that the ratio
from accreting PBHs including feedback effects, the secondary
  ionization due to fast photoelectrons, and Compton heating/
r H ii X 1=3 cooling, and we solve the chemical network for a gas of primordial
2 ð39Þ
rB 200 composition (e.g., molecular hydrogen formation /dissociation).
We run a grid of models with a range of PBH masses spanning
is a constant of order of unity for a ‘‘naked’’ PBH with mass several orders of magnitude, between 103 and 108 M .
MPBH < Mcr  104 M and ṁ < 1. For PBHs more massive than For the sake of simplicity, throughout the rest of the paper we
Mcr , the constant is smaller by a factor (MPBH /Mcr )2 3 ; thus, local
=
consider the effects of a population of equal mass PBHs which
radiative feedback can be safely neglected because r H ii < rB . account for a fraction fPBH of the dark matter. The results can be
Including the growth of the dark halo around PBHs we find used to estimate the effect of an arbitrary PBH mass function by
   integrating the optical depth de (M ) in each mass bin with weight
r H ii 1þz X 1=3 function (dfPBH /dM )1 2 . See below for an explanation of the
=

¼ 0:66 < 1; ð40Þ choice of the weighting function. Special care should be taken in
rB 1000 200
excluding from the integration PBHs which are sufficiently small
valid only at redshifts z k 100. At lower redshifts, r H ii /rB in- to be part of the dark halos of more massive PBHs. In addition we
creases with respect to the estimate in equation (40) but its value need to check that the proper velocities of small PBHs, due to their
remains of order unity or smaller for ṁ < 1. interaction with more massive ones, are negligible (i.e., subsonic).
Case B.—l / ṁ. This case applies to the case of relatively Fitting formulas for e as a function of MPBH and fPBH are pro-
massive PBHs with typical masses >103 M that have 1 < ṁ < vided at the end of this section.
10 or 0:1 < l < 1. It is sufficient to consider the case of a PBH Independently of the radiation emitted by the gas falling di-
clothed in its dark halo. We find rectly onto the PBHs, gas inflow onto the PBH can produce
shocks and collisional ionization. For the idealized case of an
   
r H ii MPBH 2=3 X 1=3 isothermal gas the sonic point is located at 0:5rB , where rB is the
4 : ð41Þ Bondi radius. At this point a shock develops that ionizes the gas,
rB 104 M 3
even if the temperature remains below 10,000 K due to Compton
cooling and molecular hydrogen cooling. The volume filling
Thus, r H ii < rB for MPBH k105 M. Also in this case we can
factor of the collisionally ionized gas is fV  nPBH (rB /2)3 , where
neglect local feedback with the exception of the mass range 103 Y
nPBH is the PBH physical number density. Assuming that a frac-
105 M where local thermal feedback may reduce accretion by a
tion of the IGM fV is fully ionized and the rest of the cosmic gas
factor of a few (from eq. [32] we obtain a reduction fduty  30%
is neutral we have a volume-weighted ionization fraction due to
for MPBH  103 M ).
shocks: hxe iV  1011 fPBH (MPBH /1 M )2 . This effect is negligible
In summary, the local feedback due to the formation of an H ii for any PBH mass assuming values of fPBH allowed by obser-
region around the PBH can be neglected in most cases. At high vational constraints.
redshift the temperature inside the H ii regions is only a few tens Examples of simulations with MPBH ¼ 100 and 1000 M are
of a percent higher than the temperature outside (see eq. [36]); shown in Figure 6. Here, we describe simulations for models
thus, even if the Bondi radius is smaller than the Strömgren radius, that are consistent with COBE FIRAS and WMAP3 data. In these
the reduction of the accretion rate due to thermal feedback is models the contribution of PBHs to the total optical depth to
negligible. At lower redshifts (z < 100) the temperature inside Thomson scattering typically does not exceed e  0:03Y0:04
H ii regions is larger than the temperature outside but we found at 68% confidence level and e  0:07Y0:09 at 95% confidence
that the Strömgren radius is typically smaller than the Bondi level. The contribution to the optical depth is uncorrelated with the
radius. value of e produced by ionization sources in high-redshift gal-
The only exceptions are massive PBHs with 103 M < axies (at z < 30) which is e ¼ 0:09 0:03 (see x 6).
MPBH < 105 M for which we estimate fduty P 30%. For this The mean electron fraction increases approximately as xe /
case, since r H ii > rB , we need to also assume that the gas is (z þ 1)1 from xe  103 at z  900 to values close to xe 
fully ionized at the Bondi radius. Hence, in addition to thermal 101 to 102 at z  10. Hence, the viscous effect of Compton drag
840 RICOTTI, OSTRIKER, & MACK Vol. 680

Fig. 6.—Left: Simulation of early ionization by PBHs with mass MPBH ¼ 100 M and abundance fPBH ¼ 104 . (a) The ionization fraction xe (the dashed line shows
the recombination history without PBHs); (b) the temperature (the dashed line shows the thermal history without PBHs); (c) the He ionization fraction; (d ) the H2
abundance; (e) the accretion rate, ṁ (solid curve) and luminosity l (dashed curve); ( f ) the optical depth to Thomson scattering e . Right: Same as in the left, but for
MPBH ¼ 1000 M and fPBH ¼ 106 (solid curves) and fPBH ¼ 107 (dotted curves).

can be neglected after recombination. The main contribution to the the thermal and ionization history would be somewhat modified
partial ionization of the cosmic gas after recombination is from X- at z < 30, but our main conclusions and the value of e would
ray emission at redshifts z > 100. At smaller redshifts, due to the not be affected because they are dominated by the PBH accre-
rapid decline of the accretion rate onto PBHs, the ionization tion rate at z > 100.
fraction keeps increasing slowly, due to redshifted X-ray back- Depending on the mass of PBHs we encounter different re-
ground photons (Ricotti & Ostriker 2004). The gas temperature gimes for the accretion.
is approximately equal to the CMB temperature at z > 200 and
1. 1 M < MPBH < 30 M .—The dark halo has negligible
ranges between 100 and 1000 K afterward. The He ii ionization
effect on the accretion rate. The accretion is spherical and the
fraction becomes 1%Y10% and the molecular abundance xH2 
luminosity l / ṁ2. Global feedback effects are negligible.
104 Y105 after redshift z  100. The molecular abundance is
2. 30 M < MPBH < 300 M .—The dark halo increases the
10Y100 times larger than the standard value, xH2  106 , obtained
accretion rate onto PBHs, mostly at redshifts z < 100. The ac-
neglecting PBHs. In all models the accretion rate onto PBHs
cretion is spherical and the luminosity l / ṁ2. Global feedback
declines rapidly at low redshift (z  10), mainly because of the
effects are important in reducing the accretion rate at z < 100.
increase of the temperature of the IGM due to X-ray heating from
3. 300 M < MPBH < 3000 M .—The dark halo increases
PBHs. In addition, at z < 6Y10, due to the growth of density per-
the accretion rate onto PBHs at all redshifts. The accretion rate
turbations, an increasing fraction of PBHs are captured or tidally
is large (ṁ k 1); thus, a thin disk may form and the luminosity
stripped by galactic halos (see x 3.2.2). The accretion of gas and
is l / fduty ṁ. Global feedback effects moderately reduce the
dark matter for these halos ceases to be spherical and modeling
accretion rate.
the accretion onto such PBHs becomes nontrivial. However, the
4. MPBH > 3000 M .—The dark halo increases the accre-
contribution of these PBHs to the global radiation emissivity is
tion rate onto PBHs at all redshifts. For masses >5 ; 104 M ,
negligible for two reasons. First, for WMAP3 cosmological pa-
the Hubble expansion becomes important and the Bondi solutions
rameter, only a small fraction of all PBHs are captured by dark
transition to self-similar infall solutions (see Paper II). PBHs ac-
halos of comparable or larger mass, even at z  4. In addition,
crete at the Eddington rate with a duty cycle fduty . Global feed-
those PBHs that are incorporated or infalling into more massive
back effects do not reduce the accretion rate significantly for
dark halos move with respect to the gas at a speed comparable to
values e P 0:05, compatible with WMAP3 (see x 6).
the virial velocity of the halo, substantially quenching the gas
accretion rate. Such PBHs will not emit much radiation, unless, The qualitative dependence of xe (z) and e (z) on the mass and
due to dynamical friction, they are able to settle at the center of abundance of PBHs can be understood assuming ionization equi-
the halo in a Hubble time. librium. For xe (z)T1 and neglecting the temperature dependence
We do not include ionization sources other than PBHs in our of the recombination coefficient, we have xe (z)2 / l(z; MPBH ) fPBH .
simulations. However, we assume that at redshift z < 7, due to the For PBHs with masses smaller than MPBH  100 M , we have
2 1=2
observed reionization and reheating of the IGM, the temperature ṁ < 1 and l / MPBH . Hence, e / xe / MPBH fPBH . If the mass
is 20,000 K. The IGM reheating explains the sharp drop of PBHs of the PBH is k100 M , the accretion rate is typically ṁ > 1,
l / MPBH , and e / xe / (MPBH fduty fPBH )1 2 . Finally, if MPBH k
=
accretion rate at z ¼ 7 evident in the bottom left of Figure 6. If
we had included galactic sources of ionization in addition to PBHs, 1000 M , we have ṁ  10, PBHs accrete nearly at the Eddington
No. 2, 2008 EFFECT OF PBHs ON CMB 841

to the value e ¼ 0:09 quoted by the WMAP3 team, which must


be produced by ionization sources other than PBHs.
In order to constrain models which allow for the existence of
PBHs we include an additional cosmological parameter describ-
ing the deviation from the standard recombination history calcu-
lated using RECFAST (Seager et al. 1999). We have modified the
publicly available codes CAMCMB and CosmoMC (Lewis &
Bridle 2002) to include an additional degree of freedom. The
redshift dependence of the ionization history after recombination
depends weakly on the mass of PBHs for values of e P 0:1.
We modify the recombination history xe;rec (z) given by RECFAST
as follows:
"   #
1 þ z 1
xe (z) ¼ xe;rec (z) þ min xe0 ; 0:1 : ð43Þ
1000

The free parameter xe0 is the ionization fraction at z ¼ 1000.


The ionization fraction increases proportionally to the scale
parameter after recombination and is constant at later times. In
addition, we model the instantaneous reionization at z ¼ zrei
produced by galactic sources. We will show that the effects on
the CMB polarization anisotropies of PBHs and other ionizing
Fig. 7.—Power spectrum of temperature and polarization anisotropies for sources are uncorrelated.
the best-fit WAMP3 model with e ¼ 0:09 and redshift of reionization zrei ¼ 11 Using the Markhov chain Monte Carlo code CosmoMC with
(crosses) and a model with the same e but zrei ¼ 7 and modified recombination
history with constant residual ionization fraction xe0 ¼ 2:5 ; 103 . From top to WMAP3 data we find 1 and 2  marginalized confidence limits
bottom: TT, TE, and EE power spectra. for the new free parameter: xe0 < 2 ; 104 (68% CF) and xe0 <
4:2 ; 104 (95% CF). These upper limits on the fractional ion-
ization correspond to upper limits on e ¼ e  e;rei of 0.05
limit and e / xe / ( fduty fPBH )1 2 is independent of the PBH
=
(68% CF ) and 0.1 (95% CF ).
mass. To first order, the value of e produced by PBHs of mass As illustrated in Figure 8, the new parameter does not correlate
MPBH and abundance fPBH can be parameterized as follows: with the redshift of reionization zrei , but because it increases the
total e it correlates with 8 and ns , increasing their values with
e  respect to the ones quoted by the WMAP3 team (Spergel et al.
8   2007). Invoking a nonstandard recombination history may ease
>
> MPBH 1=2
>
> 0:05 fPBH if MPBH < 100 M ; the tension between the low value of 8  0:74 from the WMAP3
>
<  1 M
> analysis (assuming standard recombination history) and clusters
MPBH  1=2 MPBH ð42Þ data that instead seem to favor larger values of 8  0:9 (Evrard
>
> 0:1 fduty fPBH if 102 < < 103 ;
> 1 M 1 M et al. 2008, but see Bode et al. 2007).
>
>
>
: 5 1=2 Using the results of the simulations presented in x 5 that are
10 fduty fPBH if MPBH > 103 M : approximately summarized by equation (42), we are able to con-
strain the abundance of PBHs with masses 1 M < MPBH <
The fit is derived from a grid of simulations with e < 0:2 and 108 M . The results are shown by the thick solid line in Figure 9
is not accurate for large values of e due to feedback and sat- (left).
uration effects.
6.2. Spectral Distortions
6. EFFECTS OF PBHs ON THE CMB SPECTRUM The injection of some form of energy density, U , in the ex-
AND ANISOTROPIES panding universe at redshifts z < 107 cannot be fully thermalized,
and so it produces observable deviations of the CMB spectrum
6.1. Temperature and Polarization Anisotropies
from a perfect blackbody (e.g., Burigana et al. 1991). There are
In x 5 we have shown that a signature of the existence of PBHs two types of spectral distortions: - and y-distortions. Energy
is the modification of the cosmic recombination history. It is pos- injection in the redshift range 107 < z < 105 produces a Bose-
sible to construct models with e  1 due to the large residual frac- Einstein spectrum (i.e., -distortions) with chemical potential
tional ionization of the cosmic gas after recombination produced U /U ¼ 0:71, where U is the unperturbed energy density of
by X-ray ionization. Of course, such models are ruled out by CMB the CMB. Gas accretion onto PBHs before the redshift of equal-
observations. The WMAP3 constraint on e is e  0:09 0:03. ity is negligible, hence nonevaporating PBHs do not produce
This limit on e assumes that the intergalactic medium becomes -distortions. In the redshift interval 105 < z < 1000 the Comp-
partially or fully ionized by stars and/or black holes starting at tonization distortion dominates and has amplitude U /U ¼ 4y.
redshift z P 30. As illustrated in Figure 7, the partial ionization Using the spectrometer FIRAS on the COBE satellite, Fixsen
from PBHs and the ionization from high-redshift galaxies forming et al. (1996) found the following upper limits for the deviation
at z < 30 produce very different signatures on the CMB polar- of the CMB from a Planck spectrum:   9 ; 105 and y  1:5 ;
ization anisotropies. PBHs affect small angular scales with l k 10, 105 at 95% confidence. It follows that U /U  6 ; 105 at
while radiation emitted by high-redshift galaxies affects larger 95% confidence from 107 < z < 1000. In this section we cal-
scales with l < 10. Hence, PBHs do not contribute significantly culate the value of the y-parameter produced by a population of
842 RICOTTI, OSTRIKER, & MACK Vol. 680

Fig. 8.—Likelihood isocontours of cosmological parameters as a function of xe0 , which describes the modified recombination history produced by PBHs (see the text).
We have used WMAP3 data and a modified version of CosmoMC ( Lewis & Bridle 2002). The equation used for the nonstandard recombination history is given in eq. (43)
and the free parameter xe0 describes the electron fraction at z ¼ 1000. Clockwise from the top-left corner: (1) marginalized likelihood of xe0 (in units of 104 ) with 68%
and 95% confidence limits; (2) optical depth e vs. xe0 ; (3) amplitude of the power spectrum vs. xe0 ; (4) scalar tilt of the initial power spectrum depth ns vs. xe0 . [See the
electronic edition of the Journal for a color version of this figure.]

PBHs of mass MPBH and compare it to observed upper limits sorbed by the gas and exchanged with the CMB radiation. As a
from FIRAS to set upper limits on their abundance. In order to zeroth order approximation we can use equation (44), integrating
calculate the total value of the y-parameter we need to distin- between zd and zrec , to estimate an upper limit for the y-distortion.
guish between three epochs for the energy injection: The integration shows that for PBH masses MPBH < 103 M ,
y2 Ty1 even when we assume that all the energy is absorbed by
1. Before the redshift of last scattering at zrec  1000 the uni-
the gas. In x 4.2.1 we estimated that only a fraction C  1:4 ;
verse is optically thick to Compton scattering; hence, all the energy
104 of the radiation emitted by PBHs is deposited into heat. A
emitted by accreting PBHs is absorbed by the cosmic gas. During
more realistic estimate of U is obtained by multiplying our
this epoch the y-parameter is
upper limit for y2 by C. Hence, we do not need to worry further
Z zrec about getting a more precise estimate of y2 for PBH of any mass
1 dz dU (z)
y1 ¼ ; ð44Þ as y2 Ty1 .
4U (zeq ) zeq aH(z) dt 3. After photon-electron temperature decoupling, Compton
heating/cooling becomes negligible and we can calculate the
where dU (z)/dt is the total energy per unit comoving volume y-parameter using the relationship
per unit time emitted by accretion onto PBHs and U (z) is the Z 10
unperturbed comoving energy density of the CMB at redshift z. kB
y3 ¼ (Te  TCMB ) de : ð45Þ
We will show that this contribution to the total y-parameter me c 2 zd
is dominant. We integrate dU (z)/dt starting at the redshift of
matter-radiation equality because before zeq ptheffiffiffi effective cosmic Let us estimate the upper limit for y3. Assuming Te ¼ constant 3
sound speed approaches the value cs  c/ 3 and the gas ac- TCMB , we have y3 ¼ 1:726 ; 106 (Te /104 K)e . In all the
cretion rate onto PBHs is negligible. models consistent with WMAP3 data the gas temperature is
2. Between last scattering and photon-electron temperature Te < 1000 K and e between redshift z  10 and z  100 is
decoupling the estimate of the y-parameter is complicated by <0.05. It follows that the upper limit on y3 is 8:6 ; 109 , which
the fact that only a fraction of the energy emitted by PBHs is ab- is negligible when compared to the FIRAS upper limit. Finally,
No. 2, 2008 EFFECT OF PBHs ON CMB 843

Fig. 9.—Left: Upper limits on the present abundance of PBHs. The thick lines are the results obtained in the present work. The solid lines show the upper limits using
WMAP3 data (CMB anisotropies) for two values of the black hole duty cycle fduty ¼ 1 and 0.1. The dashed lines show the limits using COBE FIRAS data (CMB
spectral distortions) at 95% and 68% confidence. The other lines refer to previous upper limits from microlensing (EROS and MACHO collaborations) and dynamical
constraints (see introduction). Right: Upper limits on the abundance of PBHs at the epoch of their formation  as a function of their mass. We assume that the mass of
PBHs is a fraction f hor of the mass of the horizon at the epoch of their formation. The thick curves show the upper limits obtained in the present work and the thin dotted
curve are limits from the EROS collaboration (microlensing experiment). [See the electronic edition of the Journal for a color version of this figure.]

the value of the y-parameter is the sum of the y during each epoch. the CMB anisotropies and spectrum. In Paper I we focused on
Thus, in our case, we have y  y1 . studying the formation and growth of the dark matter halo which
envelops PBHs that do not constitute the bulk of the dark matter.
Let us now estimate y1 . The total energy density emitted per
In the second paper (Ricotti 2007), we study in detail the Bondi-
unit time per unit comoving volume is U /dt ¼ lLEd nPBH . Thus,
type accretion solutions onto PBHs including the effects of
we find
Compton drag, Hubble expansion, and the growth of the dark
Z zrec matter halo. Finally, this work focuses on modeling the accretion
LEd crit
dm l(MPBH ; z) luminosity of PBHs including feedback effects and observational
y1 ¼ fPBH dz : ð46Þ
4MPBH aR T04 (1 þ zeq ) zeq aH(z) signatures.
We find that if a fraction fPBH of the dark matter is in PBHs
Using equation (30) for the dimensionless accretion luminosity with mass >0.1 M , the energy released due to gas accretion may
l (see also Fig. 4), we obtain the value of the y-parameter as a produce spectral distortions of the CMB radiation and keep the
function of MPBH and fPBH . Imposing y  1:5 ; 105 , we obtain universe partially ionized after recombination. The limits on the
upper limits for fPBH (MPBH ) at 95% confidence. The results are mass and abundances of PBHs set from observations of the X-ray
summarized in Figure 9 (left). background are much less restrictive than those from the CMB.
In summary, before the redshift of recombination gas accre- The modified recombination history produces observable sig-
tion onto PBHs with mass <100 M is not greatly reduced by natures on the spectrum of polarization anisotropies of the CMB
Compton drag. Although the accretion luminosity during this at angular scales l k 10. Hence, the effect of PBHs cannot be
epoch does not contribute to increase e , the energy injection confused with the effect of ionization by high-redshift galaxies
produce spectral distortions of the CMB, increasing the value of which affect polarization anisotropies on larger angular scales.
the y-parameter. We find that the existence of PBHs with masses We are able to improve the constraints on fPBH for PBHs with
<100 M is best constrained by upper limits on the y-parameter masses >0.1 M by several orders of magnitude using WMAP3
from FIRAS. and COBE FIRAS data. The results are summarized in Figure 9
(left). The upper limits on the abundance of PBHs with masses
0:1 M < MPBH < 108 M at the epoch of their formation, ,
7. DISCUSSION AND SUMMARY are shown in Figure 9 (right). We use equation (1) to derive  as
During the radiation era, mildly nonlinear perturbations with a function of MPBH , fPBH and the ratio f hor ¼ MPBH /Mhor between
/  0:5Y1 can collapse directly into primordial black holes the PBH mass and the mass Mhor of the horizon at the epoch of
(PBHs). Such black holes may have masses ranging from the PBH formation.
Planck mass to a million solar masses, depending on the redshift Fitting WMAP3 data with cosmological models that do not
of their formation and the details of the formation mechanism. allow for nonstandard recombination histories as produced by
The abundance of evaporating PBHs with masses <1015 g is con- PBHs or other early energy sources may lead to an underestimate
strained by observations to be a fraction  P1022 of the mean of the best-fit values of the amplitude of linear density fluctu-
energy density of the universe at the time of their formation, but ations, 8 , and the scalar spectral index, ns . This happens because
the existence of PBHs with masses larger than 1015 g is poorly the contribution of PBHs to the optical depth to Thomson scat-
constrained. It is not ruled out that the bulk of the dark matter may tering, which is uncorrelated with the contribution from galactic
be composed of PBHs with masses in the range between 1015 and ionization sources, can be e P 0:05 (e P 0:1 at 95% CF ).
1026 g or Planck mass relics with M  105 g. In this work, the Since ns and 8 are correlated with e , their best-fit values may
last of a series of three papers, we study the effects of a ( yet increase to ns  1 and 8  0:9 (at 95% CF). This is a general
undetected) population of nonevaporating PBHs on the thermal result (see also Bean et al. 2003, 2007) that may reduce recent
and ionization history of the universe and their signatures on tensions between WMAP3 data and other data, such as Ly forest
844 RICOTTI, OSTRIKER, & MACK Vol. 680

(e.g., Viel et al. 2006) and clusters data (Evrard et al. 2008), on the creases steeply with decreasing mass for MPBH < 1000 M. Thus,
value of 8 . although a smaller fraction of the seed PBHs can grow substan-
A population of intermediate-mass black holes (IMBHs) with tially, the number of seeds available can be much larger. PBHs
masses of about 100Y1000 M is still allowed and may be wide- with masses smaller than 100 M , assuming Bondi type accretion
spread if a fraction of the ultraluminous X-ray sources (ULX ) from the ISM of a typical high-z galaxy, are unlikely to accrete
observed in nearby galaxies host IMBHs (Miller et al. 2003; rapidly enough to grow to SMBH masses in less than 1 Gyr, even
Dewangan et al. 2006). The origin of IMBHs is unknown, but if if they constitute a few percent of the dark matter (Kuranov et al.
they are produced by Population III stars, their number may fall 2007; Pelupessy et al. 2007; M. Ricotti & F. Köckert 2008, in
short in explaining the observed ULXs population (e.g., Kuranov preparation).
et al. 2007; Pelupessy et al. 2007). We find that, if all or a fraction The increased fractional ionization of the cosmic gas produced
of observed ULXs are PBHs with masses MPBH  100Y1000 M by nonstandard recombination also increases the primordial mo-
with fPBH  105 , they would increase the best-fit value of the lecular hydrogen abundance to xH2  104 to 105 after redshift
optical depth to Thomson scattering to e  0:2. Since the scalar z  100. This value is between 10 and 100 times larger than the
spectral index ns and the amplitude of density fluctuations As and standard value, xH2  106 , obtained neglecting PBHs. The in-
8 are correlated to e , their best fits also increase to ns  1 and crease of the cosmic Jeans mass due to X-ray heating is negligible
8  0:9. PBHs in this mass range may be produced in two-stage for models consistent with the CMB data. Therefore, the for-
inflationary models designed to fit the low WMAP quadrupole mation rate of the first galaxies and stars may be enhanced if a
(Kawasaki et al. 2006). We emphasize again that this effect is population of PBHs exists. Several aspects of first-star and galaxy
more general than the specific case of PBHs discussed in this formation physics would be affected by the enhanced molecular
paper. Any mechanism or energy source that modifies the standard fraction: (1) the mass of the first stars may be reduced due to for-
recombination history may affect the estimate of cosmological mation of HD molecules (Nagakura & Omukai 2005); (2) the inter-
parameters in a way similar to that discussed here. galactic medium would be optically thick to H2 photodissociating
Our results are in contradiction with the suggestion that radiation in the Lyman-Werner bands, allowing molecular hy-
MACHOs are PBHs with mass 0.1Y1 M and fPBH  0:2 drogen to survive in the low-density IGM even at relatively low
(Alcock et al. 2000). Such a PBH population would produce redshifts z  10Y15; and (3) the epoch of domination of the first
spectral distortions incompatible with COBE FIRAS data. stars and galaxies would probably start earlier and perhaps last
The luminous QSOs found by Sloan Digital Sky Survey at longer. The number of first galaxies that remain completely dark
z  6 are thought to be powered by 108 Y109 M SMBHs. It is would be reduced. It is not obvious that the star formation effi-
difficult to produce such massive black holes starting from small ciency and other internal properties of the first galaxies would
seeds by gas accretion because the age of the universe at z ¼ 6 is be affected because feedback effects such as photoevaporation
a few tens the Salpeter accretion timescale. A few massive PBHs from internal sources and SN explosions are probably dominant
or numerous less massive PBHs may help explain the origin of (Ricotti et al. 2002a, 2002b). We leave quantitative calculations
SMBHs at high redshift and in present day galaxies by producing on the impact of PBHs on the formation of the first galaxies to
relatively massive ‘‘seeds.’’ Are the upper limits on the number of a future work.
PBHs derived in this work compatible with this scenario? The Planned and ongoing CMB experiments, such as the 5th year
fraction of mass in SMBHs today is approximately
SMBH /
dm  WMAP data release and the European Planck satellite, will allow
2:13 ; 105 (Gebhardt et al. 2000; Ricotti & Ostriker 2004). For us to improve the upper limits on the abundance of PBHs or may
PBHs with mass >1000 M , we found fPBH ¼
PBH /
dm P provide indirect evidence of their existence. The future NASA
106 /fduty . Hence, assuming that only a fraction FAGN  1 of space telescope JWST may be able to observe the formation of the
PBHs is incorporated into SMBHs and grows by gas accretion first stars if their number is much larger than expected in the stan-
by a factor Xacc  1, we have fPBH Xacc FAGN  2 ; 105 or dard cosmological model. Although difficult to model in detail,
Xacc FAGN k 20fduty . The most massive PBHs have FAGN ! 1 qualitatively we expect that the formation of the first stars and
because they spiral in to the centers of galaxies by dynamical galaxies is enhanced in some cosmological models with PBHs.
friction on a shorter timescale [tfric /tHub (z)  0:02Mdm (z)/MPBH ,
where tHub is the Hubble time] and because they may accrete gas
more efficiently. Hence, for fduty  3% and FAGN ¼ 1 we find
Xacc k 1, indicating that even scenarios with negligible mass M. R. acknowledge stimulating discussions with Niayesh
accretion onto PBHs (i.e., only growth through mergers) are Afshordi, Carlo Burigana, Andrea Ferrara, Avi Loeb, Cole
consistent with the observed mass in SMBHs today. Miller, Eve Ostriker, Ruben Salvaterra, David Spergel, and Matias
Less massive PBHs have lower probability for growing to Zaldarriaga. This work was supported in part by NASA grant
masses typical of SMBHs because the Bondi accretion rate is NNX07AH10G (M. R.) and a NSF Graduate Research Fellow-
/ M 2 . However, the upper limit on the abundance of PBHs in- ship ( K. J. M.).

REFERENCES
Afshordi, N., McDonald, P., & Spergel, D. N. 2003, ApJ, 594, L71 Bode, P., Ostriker, J. P., Weller, J., & Shaw, L. 2007, ApJ, 663, 139
Alcock, C., et al. 1998, ApJ, 499, L9 Bondi, H., & Hoyle, F. 1944, MNRAS, 104, 273
———. 2000, ApJ, 542, 281 Burigana, C., Danese, L., & de Zotti, G. 1991, A&A, 246, 49
———. 2001, ApJ, 550, L169 Carr, B. 1994, ARA&A, 32, 531
Alexander, S., & Mészáros, P. 2007, preprint ( hep-th/0703070) Carr, B. J. 1975, ApJ, 201, 1
Barrow, J. D., & Silk, J. 1979, Gen. Relativ. Gravitation, 10, 633 ———. 1981, MNRAS, 194, 639
Battistelli, E. S., Fulcoli, V., & Macculi, C. 2000, NewA, 5, 77 ———. 2003, in Quantum Gravity: From Theory to Experimental Search,
Bean, R., Melchiorri, A., & Silk, J. 2003, Phys. Rev. D, 68, 083501 ed. D. Giulini, C. Kiefer, & C. Lämmerzahl ( Berlin: Springer), 301
———. 2007, Phys. Rev. D, 75, 063505 ———. 2006, in Inflating Horizons of Particle Astrophysics and Cosmology,
Bertschinger, E. 1985, ApJS, 58, 39 ed. H. Suzuki et al. ( Tokyo: Universal Academy Press), 129
Bicknell, G. V., & Henriksen, R. N. 1978, ApJ, 225, 237 Carr, B. J., & Hawking, S. W. 1974, MNRAS, 168, 399
No. 2, 2008 EFFECT OF PBHs ON CMB 845
Chisholm, J. R. 2006, Phys. Rev. D, 73, 083504 Miller, J. M., Fabbiano, G., Miller, M. C., & Fabian, A. C. 2003, ApJ, 585, L37
Chiu, W. A., & Ostriker, J. P. 2000, ApJ, 534, 507 Miller, M. C., & Ostriker, E. C. 2001, ApJ, 561, 496
Chongchitnan, S., & Efstathiou, G. 2007, J. Cosmol. Astropart. Phys., 1, 11 Moore, B. 1993, ApJ, 413, L93
Choptuik, M. W. 1993, Phys. Rev. Lett., 70, 9 Musco, I., Miller, J. C., & Rezzolla, L. 2005, Classical Quantum Gravity, 22, 1405
Crawford, M., & Schramm, D. N. 1982, Nature, 298, 538 Nagakura, T., & Omukai, K. 2005, MNRAS, 364, 1378
Dewangan, G. C., Titarchuk, L., & Griffiths, R. E. 2006, ApJ, 637, L21 Narayan, R., & Yi, I. 1995, ApJ, 452, 710
Dokuchaev, V., Eroshenko, Y., & Rubin, S. 2004, preprint (astro-ph/0412479) Nozari, K. 2007, Astropart. Phys., 27, 169
Evans, C. R., & Coleman, J. S. 1994, Phys. Rev. Lett., 72, 1782 Ostriker, J. P., & Suto, Y. 1990, ApJ, 348, 378
Evrard, A. E., et al. 2008, ApJ, 672, 122 Park, M.-G., & Ostriker, J. P. 2001, ApJ, 549, 100
Fixsen, D. J., Cheng, E. S., Gales, J. M., Mather, J. C., Shafer, R. A., & Wright, Pelupessy, F. I., Di Matteo, T., & Ciardi, B. 2007, ApJ, 665, 107
E. L. 1996, ApJ, 473, 576 Polnarev, A., & Zembowicz, R. 1991, Phys. Rev. D, 43, 1106
Gebhardt, K., et al. 2000, ApJ, 539, L13 Press, W. H., & Schechter, P. 1974, ApJ, 193, 437
Gnedin, N. Y., Ostriker, J. P., & Rees, M. J. 1995, ApJ, 438, 40 Ricotti, M. 2007, ApJ, 662, 53
Green, A. M., Liddle, A. R., Malik, K. A., & Sasaki, M. 2004, Phys. Rev. D, Ricotti, M., Gnedin, N. Y., & Shull, J. M. 2001, ApJ, 560, 580
70, 041502 ———. 2002a, ApJ, 575, 33
Guedens, R., Clancy, D., & Liddle, A. R. 2002a, Phys. Rev. D, 66, 083509 ———. 2002b, ApJ, 575, 49
———. 2002b, Phys. Rev. D, 66, 43513 Ricotti, M., & Ostriker, J. P. 2004, MNRAS, 352, 547
Hamadache, C., et al. 2006, A&A, 454, 185 Rubin, S. G., Sakharov, A. S., & Khlopov, M. Y. 2001, J. Exp. Theor. Phys.,
Harada, T., & Carr, B. J. 2005, Phys. Rev. D, 71, 104009 92, 921
Hawking, S. 1971, MNRAS, 152, 75 Seager, S., Sasselov, D. D., & Scott, D. 1999, ApJ, 523, L1
———. 1974, Nature, 248, 30 Shapiro, S. L. 1973a, ApJ, 180, 531
Hawking, S. W., Moss, I. G., & Stewart, J. M. 1982, Phys. Rev. D, 26, 2681 ———. 1973b, ApJ, 185, 69
Jedamzik, K. 1997, Phys. Rev. D, 55, 5871 Shapiro, S. L., Lightman, A. P., & Eardley, D. M. 1976, ApJ, 204, 187
Jin, S., Ostriker, J. P., & Wilkinson, M. I. 2005, MNRAS, 359, 104 Shapley, A. E., Steidel, C. C., Pettini, M., & Adelberger, K. L. 2003, ApJ, 588, 65
Kawasaki, M., Takayama, T., Yamaguchi, M., & Yokoyama, J. 2006, Phys. Silk, J. 1968, ApJ, 151, 459
Rev. D, 74, 043525 Spergel, D. N., et al. 2007, ApJS, 170, 377
Khlopov, M. Y., Rubin, S. G., & Sakharov, A. S. 2005, Astropart. Phys., 23, 265 Steinhardt, P. J. 1982, Phys. Rev. D, 25, 2074
Kuranov, A. G., Popov, S. B., Postnov, K. A., Volonteri, M., & Perna, R. 2007, Stojkovic, D., Freese, K., & Starkman, G. D. 2005, Phys. Rev. D, 72, 045012
MNRAS, 377, 835 Tikhomirov, V. V., & Tsalkou, Y. A. 2005, Phys. Rev. D, 72, 121301
La, D., & Steinhardt, P. J. 1989, Phys. Lett. B, 220, 375 Tisserand, P., et al. 2007, A&A, 469, 387
Lacey, C. G., & Ostriker, J. P. 1985, ApJ, 299, 633 Umemura, M., Loeb, A., & Turner, E. L. 1993, ApJ, 419, 459
Lewis, A., & Bridle, S. 2002, Phys. Rev. D, 66, 103511 Viel, M., Haehnelt, M. G., & Lewis, A. 2006, MNRAS, 370, L51
Lin, D. N. C., Carr, B. J., & Fall, S. M. 1976, MNRAS, 177, 51 Yokoyama, J. 1997, A&A, 318, 673
Loeb, A. 1993, ApJ, 403, 542 ———. 1998a, Phys. Rev. D, 58, 083510
Mack, K. J., Ostriker, J. P., & Ricotti, M. 2007, ApJ, 665, 1277 ———. 1998b, Phys. Rev. D, 58, 083510
Majumdar, A. S. 2003, Phys. Rev. Lett., 90, 031303 Yoo, J., Chanamé, J., & Gould, A. 2004, ApJ, 601, 311
Mészáros, P. 1975a, A&A, 38, 5 Zel’Dovich, Y. B., & Novikov, I. D. 1967, Soviet Astron., 10, 602
———. 1975b, A&A, 44, 59

You might also like