Cosmology With Clusters of Galaxies
Cosmology With Clusters of Galaxies
Cosmology With Clusters of Galaxies
S. Borgani
Department of Astronomy, University of Trieste, via Tiepolo 11, I-34131
Trieste, Italy
borgani@ts.astro.it
1 Introduction
Clusters of galaxies occupy a special place in the hierarchy of cosmic struc-
tures. They arise from the collapse of initial perturbations having a typical
comoving scale of about 10 h
1
Mpc
1
. According to the standard model of cos-
mic structure formation, the Universe is dominated by gravitational dynamics
in the linear or weakly nonlinear regime and on scales larger than this. In this
case, the description of cosmic structure formation is relatively simple since
gas dynamical eects are thought to play a minor role, while the dominating
gravitational dynamics still preserves memory of initial conditions. On smaller
scales, instead, the complex astrophysical processes, related to galaxy forma-
tion and evolution, become relevant. Gas cooling, star formation, feedback
from supernovae (SN) and active galactic nuclei (AGN) signicantly change
the evolution of cosmic baryons and, therefore, the observational properties
of the structures. Since clusters of galaxies mark the transition between these
two regimes, they have been studied for decades both as cosmological tools
and as astrophysical laboratories.
In this Chapter I concentrate on the role that clusters play in cosmology.
I will highlight that, in order for them to be calibrated as cosmological tools,
one needs to understand in detail the astrophysical processes which determine
their observational characteristics, i.e. the properties of the cluster galaxy
population and those of the diuse intracluster medium (ICM).
Constraints of cosmological parameters using galaxy clusters have been
placed so far by applying a variety of methods. For example:
1. The mass function of nearby galaxy clusters provides constraints on the
amplitude of the power spectrum at the cluster scale (e.g., [138, 164] and
references therein). At the same time, its evolution provides constraints
on the linear growth rate of density perturbations, which translate into
1
Here h is the Hubble constant in units of 100 kms
1
Mpc
1
S. Borgani: Cosmology with Clusters of Galaxies, Lect. Notes Phys. 740, 287334 (2008)
DOI 10.1007/978-1-4020-6941-3 9 c Springer Science+Business Media B.V. 2008
288 S. Borgani
dynamical constraints on the matter and Dark Energy (DE) density pa-
rameters.
2. The clustering properties (i.e., correlation function and power spectrum)
of the largescale distribution of galaxy clusters provide direct information
on the shape and amplitude of the underlying DM distribution power spec-
trum. Furthermore, the evolution of these clustering properties is again
sensitive to the value of the density parameters through the linear growth
rate of perturbations (e.g., [26, 114] and references therein).
3. The mass-to-light ratio in the optical band can be used to estimate the
matter density parameter,
m
, once the mean luminosity density of the
Universe is known and under the assumption that mass traces light with
the same eciency both inside and outside clusters (see [9, 35, 70], as
examples of the application of this method).
4. The baryon fraction in nearby clusters provides constraints on the matter
density parameter, once the cosmic baryon density parameter is known,
under the assumption that clusters are fair containers of baryons (e.g.,
[60, 168]). Furthermore, the baryon fraction of distant clusters provide a
geometrical constraint on the DE content and equation of state, under the
additional assumption that the baryon fraction within clusters does not
evolve (e.g., [6, 58]).
An extensive presentation of all these methods would probably require a
dedicated book. For this reason, in this contribution I will mostly concentrate
on the method based on the evolution of the cluster mass function. A substan-
tial part of my Lecture will concentrate on the dierent methods that have
been applied so far to weight galaxy clusters. Since all the above cosmological
applications rely on precise measurements of cluster masses, this part of my
contribution will be of general relevance for cluster cosmology.
Also, since most of the cosmological applications of galaxy clusters have
been based so far on Xray surveys, my discussion will be denitely Xray
biased, although I will discuss in some detail methods based on optical obser-
vations and what present and future optical surveys are expected to provide.
I refer to the lecture by Roy Gal in this volume for more details regarding
the properties of galaxy clusters in the optical band. Also, I will refer to the
Lectures by M. Birkinshaw for cosmological studies of clusters based on the
SunyaevZeldovich (SZ) eect, to the Lectures by J.P. Kneib for cluster
studies and mass measurement through gravitational lensing, and to the Lec-
tures by C. Jones and by C. Sarazin for more details about the cosmological
application of the baryon fraction method.
The structure of this Chapter will be as follows. I provide in Sect. 2 a
short introduction to the basics of cosmic structure formation. I will shortly
review the linear theory for the evolution of density perturbations and the
spherical collapse model. In Sect. 3 I will describe the PressSchechter (PS)
formalism to derive the cosmological mass function. I will then introduce ex-
tensions of the PS approach and present the most recent calibrations of the
Clusters and Cosmology 289
mass function from Nbody simulations. In Sect. 4 I will review the meth-
ods to build samples of galaxy clusters, based on optical and Xray observa-
tions, while I will only briey discuss the SZ methodology for cluster surveys.
Section 5 is devoted to the discussion of dierent methods to derive cluster
masses and to review the results of the application of these methods. In Sect. 6
I will describe the cosmological constraints, which have been obtained so far
by tracing the cluster mass function with a variety of methods: distribution
of velocity dispersions, Xray temperature and luminosity functions, and gas
mass function. In this Section I will also critically discuss the reasons for the
dierent, sometimes discrepant, results that have been obtained in the liter-
ature and I will highlight the relevance of properly including the analysis of
the cluster mass function all the statistical and systematic uncertainties in
the relation between mass and observables. Finally, I will describe in Sect. 7
the future perspectives for cosmology with galaxy clusters and which are the
challenges for clusters to keep playing an important role in the era of precision
cosmology.
2 A Concise Handbook of Cosmic Structure Formation
In this section I will briey review the basic concepts of cosmic structure
formation, which are relevant for the study of galaxy clusters as tools for
precision cosmology through the evolution of their mass function. A complete
treatment of models of structure formation can be found in classical cosmology
textbooks (e.g., [42, 123, 124]).
2.1 The Statistics of Cosmic Density Fields
Let (x) be the matter density eld, which is a continuous function of the
position vector x, = its average value computed over a suciently large
(representative) volume of the Universe and
(x) =
(x)
(1)
the corresponding relative density contrast. By denition, it is
= 0 and
(x) 1. If the density eld is traced by a discrete distribution of points
(i.e., galaxies or galaxy clusters) all having the same weight (mass), then
(x) =
i
D
(x x
i
), where
D
(x) is the Dirac deltafunction. The Fourier
representation of the density contrast is given by
(k) =
1
(2)
3/2
_
dx(x)e
ikx
, (2)
with the corresponding dual relation for the inverse Fourier transform.
290 S. Borgani
The 2point correlation function for the density contrast is dened as
(r) = (x
1
)(x
2
) , (3)
which only depends on the modulus of the separation vector, r = |x
1
x
2
|,
under the assumption of statistical isotropy of the density eld. Therefore,
it can be shown that the power spectrum of the density uctuations is the
Fourier transform of the correlation function, so that
P(k) = |
(k)|
2
=
1
2
2
_
dr r
2
(r)
sinkr
kr
, (4)
which, again, depends only on the modulus of the wave-vector k.
In case we are interested in the study of a class of observable structures
of mass M, which arise from the collapse of initial perturbations having size
R (M/ )
1/3
, then it is common to introduce the smoothed density eld,
which is dened as
R
(x) =
M
(x) =
_
(y)W
R
(|x y|) dy . (5)
As such, it is given by the convolution of the density uctuation eld with a
window function, which lters out the uctuation modes having wavelength
<
R. Equation (5) allows us to introduce the variance of the uctuation eld
computed at the scale R, dened as
2
R
=
2
M
=
2
R
=
1
2
2
_
dk k
2
P(k)
W
2
R
(k) , (6)
where
W
R
(k) is the Fourier transform of the window function.
The shape of the window function denes the exact relation between mass
and smoothing scale. For instance, for the tophat window it is
W
R
(k) =
3[sin(kR) kRcos(kR)]
(kR)
3
(7)
while the Gaussian window gives
W
R
(k) = exp
_
(kR)
2
2
_
, . (8)
The corresponding relations between mass scale and smoothing scale are
M = (4/3)R
3
and M = (2R
2
)
3/2
for the tophat and Gaussian lters,
respectively.
The shape of the power spectrum is (essentially) xed once the matter
density parameter,
m
, that associated to the baryonic component,
bar
, and
the Hubble parameter, H
0
, are specied (e.g., [53]). However, its normalization
Clusters and Cosmology 291
can only be xed through a comparison with observational data of the large
scale structure of the Universe or of the anisotropies of the Cosmic Microwave
Background (CMB). A common way of parametrizing this normalization is
through the quantity
8
, which is dened as the variance, computed for a
tophat window having comoving radius R = 8 h
1
Mpc (given in (6)). The
historical reason for this choice of the normalization scale is that the variance
of the galaxy number counts, within the rst redshift surveys, was observed
to be about unity inside spheres of that radius (e.g., [46]). In this way, the
value of
8
for a given cosmology directly provides a measure of the biasing
parameter relating the galaxy and mass distribution, expected for that model.
Furthermore, a tophat sphere of 8 h
1
Mpc radius contains a mass M
5.910
14
m
M
t
+ [(1 +)u] = 0 , (9)
which gives the mass conservation, the Euler equation
u
t
+ 2H(t)u + (u )u =
a
2
, (10)
which gives the relation between the acceleration of the uid element and the
gravitational force, and the Poisson equation
2
= 4G a
2
(11)
which species the Newtonian nature of the gravitational force. In the above
equations, is the gradient computed with respect to the comoving coor-
dinate x, (x) describes the uctuations of the gravitational potential and
292 S. Borgani
H(t) = a/a is the Hubble parameter at the time t. Its timedependence is
given by H(t) = E(t)H
0
, where
E(z) = [(1 +z)
3
m
+ (1 +z)
2
(1
m
DE
) + (1 +z)
3(1+w)
DE
]
1/2
(12)
is related to the density parameter contributed by nonrelativistic matter,
m
, and by Dark Energy (DE),
DE
, with equation of state p = wc
2
(if the
DE term is provided by cosmological constant then w = 1).
In the case of small perturbations, these equations can be linearized by
neglecting all the terms which are of second order in the elds and u. In this
case, after further dierentiating (9) with respect to time, using the Euler
equation to eliminate the term u/t, and using the Poisson equation to
eliminate
2
, one ends up with:
t
2
+ 2H(t)
t
4G = 0 . (13)
This equation describes the Jeans instability of a pressurless uid, with the ad-
ditional Hubble drag term 2H(t)/t, which describes the counteraction
of the expanding background on the perturbation growth. Its eect is to pre-
vent the exponential growth of the gravitational instability taking place in a
nonexpanding background [14]. The solution of the above equation can be
casted in the form:
(x, t) =
+
(x, t
i
)D
+
(t) +
(x, t
i
)D
(t) , (14)
where D
+
and D
(t) = (t/t
i
)
1
. The fact that D
+
(t) a(t) for an EdS Universe should not
be surprising. Indeed, the dynamical timescale for the collapse of a perturba-
tion of uniform density is t
dyn
(G)
1//2
, while the expansion time scale
for the EdS model is t
exp
(G )
1//2
, where is the mean cosmic density.
Since for a linear (small) perturbation it is , then t
dyn
t
exp
, thus show-
ing that the cosmic expansion and the perturbation evolution take place at
the same pace. This argument also leads to understanding the behaviour for
a
m
< 1 model. In this case, the expansion time scale becomes shorter than
the above one at the redshift at which the Universe recognizes that
m
< 1.
This happens at 1 +z
1/3
m
or at 1 +z
1
m
in the presence or absence
of a cosmological constant term, respectively. Therefore, after this redshift,
cosmic expansion takes place at a quicker pace than gravitational instability,
with the result of freezing the perturbation growth.
The exact expression for the growing model of perturbations is given by
D
+
(z) =
5
2
m
E(z)
_
z
1 +z
E(z
)
3
dz
(15)
Clusters and Cosmology 293
0 2 4 6 8 10
Redshift
m
= 1
m
= 0.3
= 0
= 0.7
m
= 0.3
0.1
1
G
r
o
w
t
h
f
a
c
t
o
r
Fig. 1. The redshift dependence of the linear growth mode of perturbations for a
at model with
m
= 1 (solid curve), for a at
m
= 0.3 model with a cosmo-
logical constant (dashed curve) and for an
m
= 0.3 open model with vanishing
cosmological constant (dotted curve)
(e.g., [124]). I show in Fig. 1 the redshift dependence of the linear growth factor
for an Eds model and for two models with
m
= 0.3 both with and without
a cosmological constant term to restore spatial atness. Quite apparently,
the EdS has the faster evolution, while the slowing down of the perturbation
growth is more apparent for the open lowdensity model, the presence of
cosmological constant providing an intermediate degree of evolution. A more
pictorial view is provided in Fig. 2, where we show the dark matter density
elds for two dierent cosmologies and at dierent epochs, as obtained from
Nbody simulations. The two models, an EdS one and a at lowdensity
one with
m
= 0.3, have been tuned so as to have a similar appearance at
z = 0. This gure clearly shows that any observational probe of the degree
of evolution of density perturbations would correspond to a sensitive probe
of cosmological parameters. Such a cosmological test is conceptually dierent
to those provided by the standard geometrical tests based on luminosity and
angularsize distances.
As we shall discuss in the following, clusters of galaxies provide such a
probe, since the evolution of their number density is directly related to the
growth rate of perturbations.
294 S. Borgani
Fig. 2. The evolution of the cluster population from Nbody simulations in two
dierent cosmologies [26]. Left panels describe a at, lowdensity model with
m
=
0.3 and
= 0.7 (L03); right panels are for an Einsteinde-Sitter model (EdS) with
m
= 1. Superimposed on the dark matter distribution, the yellow circles mark
the positions of galaxy clusters with virial temperature T > 3 keV, the size of the
circles is proportional to temperature. Model parameters have been chosen to yield
a comparable space density of clusters at the present time. Each snapshot is 250 h
1
Mpc across and 75 h
1
Mpc thick (comoving with the cosmic expansion)
Clusters and Cosmology 295
2.3 The Spherical Top-Hat Collapse
A spherical perturbation at constant density represents the only case in which
the evolution can be exactly computed. Although the assumptions on which
this model is based are quite restrictive, nevertheless it serves as a very use-
ful guideline to characterize the process of evolution and formation of viri-
alized DM halos. This approach is based on treating the perturbation as a
separate FriedmannLemaitreRobertsonWalker (FLRW) universe, with the
constraint of null velocity at the boundary of the perturbation. Here we will
sketch the derivation in the case of
m
= 1 (e.g., [42], while an extension of
this derivation to more general cosmologies can be found in [54] and [93], with
useful tting functions provided in [31].
Assuming null velocities at an initial time t
i
provides the relation D
+
(t
i
) =
(3/5)(t
i
), between the linear growth mode of the perturbation and the initial
overdensity. The initial density parameter, which characterizes this separate
Universe, is then
p
(t
i
) = (t
i
)(1 +
i
). Therefore, the condition for the
perturbation to re-collapse will be
p
(t
i
) > 1. If this condition is satised,
then we can derive the density within the perturbation at the time t
m
of its
maximum expansion (turnaround) as
p
(t
m
) =
c
(t
i
)
p
(t
i
)
_
p
(t
i
) 1
p
(t
i
)
_
3
. (16)
The time t
m
is given by the solution of the Friedmann equations for a closed
Universe:
t
m
=
2H
i
p
(t
i
)
[
p
(t
i
) 1]
3/2
=
_
3
32G
p
(t
m
)
_
1/2
, (17)
where H
i
is the Hubble parameter within the perturbation. At the same epoch
t
m
, the density of the general cosmic background is (t
m
) = (6Gt
2
m
)
1
.
Therefore, the exact value for the perturbation overdensity at the turn
around is
+
(t
m
) =
p
(t
m
)
(t
m
)
1 =
_
3
4
_
2
1 4.6 . (18)
On the other hand, the lineartheory extrapolation to t
m
would give
+
(t
m
) =
+
(t
i
)
_
t
m
t
i
_
2/3
=
3
5
_
3
4
_
2/3
1.07 . (19)
This demonstrates that the lineartheory extrapolation signicantly underes-
timates overdensities at the turn-around.
After reaching the maximum expansion, the perturbation then evolves by
detaching from the general Hubble expansion and then re-collapses, reaching
virial equilibrium supported by the velocity dispersion of DM particles. This
happens at the virialization time t
vir
, at which the perturbation meets by
296 S. Borgani
denition the virial condition E = K+U = K, being E, K and U the total,
the kinetic and the potential energy, respectively.
At the turnaround point, the perturbation has no kinetic energy, so that
the total energy is
E
m
= U =
3
5
GM
2
R
m
, (20)
where we have used the expression for the potential energy of a uniform spher-
ical density eld of radius R
m
and total mass M. In a similar manner, the
total energy at the virialization is
E
vir
=
U
2
=
1
2
3
5
GM
2
R
vir
. (21)
Therefore, the condition of energy conservation in a dissipationless collapse
gives R
m
= 2R
vir
for the relation between the radii at turn-around and at
virial equilibrium. This allows us to compute the overdensity at t
vir
as
p
(t
vir
)
(t
vir
)
=
_
t
vir
t
m
_
2
_
R
m
R
vir
_
3
p
(t
m
)
(t
m
)
= 2
2
2
3
_
3
4
_
2
= 18
2
178 , (22)
where we have accounted for both the compression of the perturbation density,
due to its shrinking, and of the dilution of the background density as the
Universe expands from t
m
to t
vir
. Equation (22) shows why an overdensity of
about 200 is usually considered as typical for a DM halo which has reached
the condition of virial equilibrium. As for the extrapolation of lineartheory
prediction, it would have given
+
(t
vir
) =
_
t
vir
t
m
_
2/3
+
(t
m
) 1.69 . (23)
The above equation shows the derivation of another fundamental number that
will be used in what follows in order to characterize the mass function of viri-
alized halos. It gives the overdensity that a perturbation in the initial density
eld must have for it to end up in a virialized structure. While the above
derivation holds for an EdS Universe, it can be generalized to any generic
cosmology. For
m
< 1 the increased expansion rate of the Universe causes a
faster dilution of the cosmic density from t
m
to t
vir
and, as a consequence, a
larger value of the overdensity at virialization.
In the following, we will indicate with
vir
the overdensity at virial equi-
librium, computed with respect to the background density, and with
c
the
same quantity expressed in units of the critical density
c
. As a reference, a
at lowdensity model with
m
= 0.3 has
c
100 and
vir
330. Also,
we will use in the following the notation R
N
to indicate the radius of a halo
encompassing an average overdensity equal to N
c
, so that M
N
will denote
the halo mass contained within that radius. As we shall see in the following,
values often used in the literature are N = 200, 500 and 2500.
Clusters and Cosmology 297
3 The Mass Function
The mass function (MF) at redshift z, n(M, z), is dened as the number
density of virialized halos found at that redshift with mass in the range
[M, M + dM]. In this section I will derive the MF expression following the
approach originally devised by Press and Schechter [132] (PS hereafter). After
commenting on the limitations of this approach, I will discuss the accuracy
with which improved derivations of the MF reproduce the exact predictions
from N-body simulations.
3.1 The Press-Schechter Mass Function
The PS derivation of the MF is based on the assumption that the fraction of
matter ending up in objects of a given mass M can be found by looking at
the portion of the initial (Lagrangian) density eld, smoothed on the mass
scale M, lying at an overdensity exceeding a given critical threshold value,
c
. Under the assumption of Gaussian perturbations, the probability for the
linearly-evolved smoothed eld
M
to exceed at redshift z the critical density
contrast
c
reads
p
>
c
(M, z) =
1
2
M
(z)
_
c
exp
_
2
M
2
M
(z)
2
_
d
M
=
1
2
erfc
_
c
2
M
(z)
_
,
(24)
where erfc(x) is the complement error function and
M
(z) =
+
(z)
M
is the
variance at the mass scale M linearly extrapolated at redshift z. Under the
assumption of spherical collapse, the critical overdensity
c
is given by the
linear extrapolation of the overdensity at virial equilibrium, as derived in the
previous section. In this case, it will be
c
=
c
(z) with a weak dependence
upon redshift and cosmological parameters, with
c
1.69 independent of
z only in the case of an EdS cosmology. By denition, the above equation
provides the fraction of unity volume, which ends up by redshift z in objects
with mass above M. Therefore, the fraction of Lagrangian volume in objects
with mass in the range [M, M + dM] is
dp
>
c
(M, z) =
p
>
c
(M, z)
M
dM . (25)
Since the probability of (24) is a decreasing function of mass, the absolute
value is required in order to have a positivedened dierential probability.
Equation (25) shows a fundamental limitation of the PS derivation of the MF.
Indeed, we expect that, as we take the limit of arbitrarily small limiting mass,
we should recover the whole mass content of the Universe. This is to say that,
in the hierarchical clustering picture, all the mass is contained within halos of
arbitrarily small mass. However, integrating (25) over the whole mass range
gives
_
0
dp
>
c
(M, z) = 1/2. This implies that the PS derivation of the mass
298 S. Borgani
function only accounts for half of the total mass at disposition. The basic
reason for this is that, in this derivation, we give zero probability for a point
with
M
<
c
, for a given ltering mass scale M, to have
M
>
c
for some
larger ltering scale M
M
2
M
(z)
d log
M
(z)
d log M
exp
_
2
c
2
M
(z)
2
_
. (26)
This is the expression for the PS mass function. Although we will present
below a more accurate expressions for the MF, this equation already demon-
strates the reason for which the mass function of galaxy clusters is a powerful
probe of cosmological models. Cosmological parameters enter in (26) through
the mass variance
M
, which depends on the power spectrum and on the cos-
mological density parameters, through the linear perturbation growth factor,
and, to a lesser degree, through the critical density contrast
c
. Taking this
expression in the limit of massive objects (i.e., rich galaxy clusters), the MF
shape is dominated by the exponential tail. This implies that the MF becomes
exponentially sensitive to the choice of the cosmological parameters. In other
words, a reliable observational determination of the MF of rich clusters would
allow us to place tight constraints on cosmological parameters.
3.2 Extensions of the PS Approach and N-body Tests
Following [89], an alternative way of recasting the mass function is
f(
M
, z) =
M
dn(M, z)
d ln
1
M
. (27)
In this way, the PS expression is recovered by setting
f(
M
, z) =
_
2
M
exp
_
2
c
2
2
M
_
(28)
Despite its subtle simplicity (e.g., [112]), the PS MF has served for more
than a decade as a guide to constrain cosmological parameters from the mass
Clusters and Cosmology 299
distribution of galaxy clusters. Only with the advent of a new generation of
Nbody simulations, which are able to cover a very large dynamical range,
have signicant deviations of the PS expression from the exact numerical
description been noticed (e.g., [59, 76, 77, 89, 150, 165]). Such deviations have
been usually interpreted in terms of corrections to the PS approach.
Incorporating the eect of nonspherical collapse, the PS expression has
been generalized [146] to
f(
M
, z) =
_
2a
C
_
1 +
_
2
M
a
2
c
_
q
_
c
M
exp
_
a
2
c
2
2
M
_
. (29)
These authors also compared this expression with results from Nbody sim-
ulations, in which the mass of the clusters were estimated with a spherical
overdensity (SO) algorithm, by computing the mass within the radius encom-
passing a mean overdensity equal to the virial one. As a result, they found
the best-tting values a = 0.707, q = 0.3, with the normalization constant
C = 0.3222 obtained from the normalization requirement
_
0
f(
M
)d = 1
(note that the PS expression is recovered for a = 1, q = 0 and C = 1/2; see
also [147]).
Jenkins et al. [89] proposed an alternative expression for the mass function:
f(
M
, z) = 0.315 exp(| ln
1
M
+ 0.61|
3.8
) , (30)
which has been obtained as the best t to the results of a combination of
dierent simulations, covering a wide dynamical range. More recently, Springel
et al. [150] used the largest available single Nbody simulation to verify in
detail the accuracy of (30). The result of this comparison, which is reported
in Fig. 3, demonstrates that this mass function reproduces remarkably well
numerical results over a wide range of sampled halo masses and redshifts,
thereby representing a substantial improvement with respect to the PS mass
function. The accuracy of (30) in reproducing results of numerical experiments
has been also discussed in [59], where it is also pointed out the role of dierent
algorithms to identify clusters and to estimate their mass in simulations, in
[166], where the universality of this expression for a generic cosmology is
discussed, and in [165], where the widest dynamical range to date has been
samples by combining a series of N-body simulations.
In practical applications, the observational mass function of clusters is usu-
ally determined over about one decade in mass. Therefore, it probes the power
spectrum over a relatively narrow dynamical range, and does not provide
strong constraints on the shape of the power spectrum. Using only the number
density of nearby clusters of a given mass M, one can constrain the amplitude
of the density perturbation at the physical scale R (M/
m
crit
)
1/3
which
contains this mass. Since such a scale depends both on M and on
m
, the
mass function of nearby (z
<
0.1) clusters is only able to constrain a relation
between
8
and
m
. In the left panel of Fig. 4 we show that, for a xed value
300 S. Borgani
10
10
10
11
10
12
10
13
10
14
10
16
10
15
10
5
10
4
10
3
10
2
10
1
M
2
/
d
n
/
d
M
z = 10.07
z = 5.72
z = 3.06
z = 1.50
z = 0.00
M [ h
1
M
O
]
0.02
0
100 150 200
R (Mpc/h)
M > 3 10
14
M
O
/h
8
/
8
250 300
0.04
0.06
M > 1 10
14
M
O
/h
Fig. 3. Left panel: the mass function of DM halos (dots with errorbars) identied
at dierent redshifts in the Millenniun Run [150], compared to the predictions of the
mass function by [89] and by [132]. The two model mass functions are plotted with
solid and dotted curves, respectively. Right panel: the relative standard deviation
of
8
as a function of the sample size [166] from the cumulative mass function above
two dierent mass limits. The solid lines indicate the width of the distribution when
including the clustering of clusters
of the observed cluster mass function, the implied value of
8
from (29) in-
creases as the density parameter decreases. Determinations of the cluster mass
function in the local Universe using a variety of samples and methods indi-
cate that
8
m
= 0.4 0.6, where 0.4 0.6, almost independent of the
presence of a cosmological constant term providing spatial atness. As for the
0 0.2 0.8 0.6 0.4 1
0.001
0.0001
0.01
0.1
1
10
n
(
>
M
,
z
)
/
n
(
>
M
,
0
)
m
= 1
m
= 0.3
m
= 0.3
= 0.7
= 0
Redshift
M > 5 10
14
h
1
M
O
0.8 0.6 0.4 1 1.2 1.4
10
8
10
9
10
7
10
6
10
5
0.0001
n
(
>
M
,
z
=
0
)
(
h
1
M
p
c
)
Fig. 4. The sensitivity of the cluster mass function to cosmological models [138].
Left panel: The cumulative mass function at z = 0 for M > 510
14
h
1
M
for three
cosmologies, as a function of
8
, with shape parameter = 0.2; solid line:
m
= 1;
shortdashed line:
m
= 0.3,
= 0. The
shaded area indicates the observational uncertainty in the determination of the local
cluster space density. Right panel: Evolution of n(>M, z) for the same cosmologies
and the same masslimit, with
8
= 0.5 for the
m
= 1 case and
8
= 0.8 for the
lowdensity models
Clusters and Cosmology 301
evolution with redshift, the growth rate of the density perturbations depends
primarily on
m
and, to a lesser extent, on
<
0.2, and assessed the degree of
completeness by resorting to a comparison with mock SDSS surveys extracted
from large Nbody simulations. Once completed, the search of clusters over
the entire SDSS sample will provide about 2500 nearby and mediumdistant
objects. At the same time the next generation of wide eld (> 100 deg
2
)
deep multicolor surveys in the optical and especially the near-infrared will
powerfully enhance the search for distant clusters, out to z
>
1.
4.2 Identication in the X-ray Band
Already from the rst pioneering attempts to map the X-ray sky ( [66],
see [138] for a historical review), clusters were associated with extended
sources, whose dominant emission mechanism was recognized to be ther-
mal bremsstrahlung from optically thin plasma at a temperature of several
keV [40, 61]. The all-sky survey conducted by the the HEAO-1 X-ray Observa-
tory was the rst to provide a uxlimited sample of X-ray identied clusters,
for which both the ux number counts and the X-ray luminosity function have
been computed for the rst time [126]. However, it is only thanks to the much
improved sensitivity of the Einstein Observatory [65] that X-ray surveys were
recognized as an ecient means of constructing samples of galaxy clusters out
to cosmologically interesting redshifts.
First, the X-ray selection has the advantage of revealing physically-bound
systems, because diuse emission from a hot ICM is the direct manifestation of
Clusters and Cosmology 303
the existence of a potential-well within which the gas is in dynamical equilib-
rium with the cool baryonic matter (galaxies) and the dark matter. Second,
the X-ray luminosity is well correlated with the cluster mass (see Fig. 11).
Third, the X-ray emissivity is proportional to the square of the gas density,
hence cluster emission is more concentrated than the optical bidimensional
galaxy distribution. In combination with the relatively low surface density
of X-ray sources, this property makes clusters high contrast objects in the
X-ray sky, and alleviates problems due to projection eects that aect opti-
cal selection. Finally, an inherent fundamental advantage of X-ray selection is
the ability to dene ux-limited samples with well-understood selection func-
tions. This leads to a simple evaluation of the survey volume and therefore
to a straightforward computation of space densities. Nonetheless, there are
some important caveats described below. Pioneering work in this eld [67, 84]
was based on the Einstein Observatory Extended Medium Sensitivity Survey
(EMSS). The EMSS survey covered over 700 square degrees and lead to the
construction of a ux-limited sample of 93 clusters out to z = 0.58, allowing
the cosmological evolution of clusters to be investigated.
The ROSAT satellite, launched in 1990, allowed a signicant step for-
ward in X-ray surveys of clusters. The ROSAT All-Sky Survey (RASS, [155])
was the rst X-ray imaging mission to cover the entire sky, thus paving the
way to large contiguous-area surveys of X-ray selected nearby clusters. In the
northern hemisphere, the largest compilations with virtually complete optical
identication include, the Bright Cluster Sample (BCS, [51]), and the North-
ern ROSAT All Sky Survey (NORAS, [22]). In the southern hemisphere, the
ROSAT-ESO ux limited X-ray (REFLEX) cluster survey [21] has completed
the identication of 452 clusters, the largest, homogeneous compilation to
date. The Massive Cluster Survey (MACS, [52]) is aimed at targeting the
most luminous systems at z > 0.3 which can be identied in the RASS at the
faintest ux levels. The deepest area in the RASS, the North Ecliptic Pole
(NEP, [85]) which ROSAT scanned repeatedly during its All-Sky survey, was
used to carry out a complete optical identication of X-ray sources over a 81
deg
2
region. This study yielded 64 clusters out to redshift z = 0.81.
In total, surveys covering more than 10
4
deg
2
have yielded over 1000 clus-
ters, out to redshift z 0.5. A large fraction of these are new discoveries,
whereas approximately one third are identied as clusters in the Abell or
Zwicky catalogs. For the homogeneity of their selection and the high degree of
completeness of their spectroscopic identications, these samples are now the
basis for a large number of follow-up investigations and cosmological studies.
Besides the all-sky surveys, the ROSAT-PSPC archival pointed obser-
vations were intensively used for serendipitous searches of distant clusters.
These projects, which are now completed, include: the RIXOS survey [38],
the ROSAT Deep Cluster Survey (RDCS, [138, 139]), the Serendipitous High-
Redshift Archival ROSAT Cluster survey (SHARC, [32], the Wide Angle
ROSAT Pointed X-ray Survey of clusters (WARPS, [125]), the 160 deg
2
large area survey [117], the ROSAT Optical X-ray Survey (ROXS, [49]).
304 S. Borgani
ROSAT-HRI pointed observations have also been used to search for distant
clusters in the Brera Multi-scale Wavelet catalog (BMW, [113]).
In Fig. 5, we give an overview of the ux limits and surveyed areas of
all major cluster surveys carried out over the last two decades. RASS-based
surveys have the advantage of covering contiguous regions of the sky so that
the clustering properties of clusters (e.g., [143]) can be investigated. They
also have the ability to unveil rare, massive systems albeit over a limited
redshift and X-ray luminosity range. Serendipitous surveys which are at least
a factor of ten deeper but cover only a few hundreds square degrees, provide
complementary information on lower luminosities, more common systems and
are well suited for studying cluster evolution on a larger redshift baseline.
A number of systematic studies have been carried out to compare the
nature of clusters identied with the optical and the Xray technique (e.g.,
[13, 48, 129]). The general conclusion of these studies is that optically selected
10
15
10
16
0.1 1 10
Area (deg
2
)
CDF
XMM/LH
CHANDRA/XMM
Surveys
RDCS
NEP
ROXS
RIXOS
SHARCS
WARPS
160 deg
2
BMW
BrightSHARC
EMSS
MACS
BCS
REFLEX
NORAS
HEAO1
10
2
10
3
10
4
10
5
A
L
L
S
K
Y
10
14
10
13
10
12
10
11
10
10
F
l
u
x
l
i
m
i
t
[
0
.
5
2
.
0
k
e
v
]
e
r
g
c
m
2
s
1
Fig. 5. Solid angles and ux limits of X-ray cluster surveys carried out over the last
two decades. Dark lled circles represent serendipitous surveys constructed from a
collection of pointed observations. Light shaded circles represent surveys covering
contiguous areas. The hatched region is a predicted locus of current serendipitous
surveys with Chandra and Newton-XMM. From [138]
Clusters and Cosmology 305
clusters are on average underluminous in the X-ray band. This suggests that
optical selection tends to pick up objects which have not yet reached a high
enough density to make the ICM lighting up in Xrays.
In order for a survey to be used for cosmological applications, one needs
to know not only how many clusters it contains, but also the volume within
which each of them is found. In other words, one needs to dene the selection
function of the survey, which depends on the survey strategy and on the details
of the adopted cluster nding algorithm (see [138], for a review). An essential
ingredient for the evaluation of the selection function of X-ray surveys is the
computation of the sky coverage: the eective area covered by the survey as a
function of ux. In general, the exposure time, as well as the background and
the PSF are not uniform across the eld of view of X-ray telescopes, which
introduces vignetting and a degradation of the PSF at increasing o-axis
angles. As a result, the sensitivity to source detection varies signicantly across
the survey area so that only bright sources can be detected over the entire solid
angle of the survey, whereas at faint uxes the eective area decreases. An
example of survey sky coverage is given in the left panel of Fig. 6. By covering
dierent solid angles at varying uxes, these surveys probe dierent volumes
at increasing redshift and therefore dierent ranges in X-ray luminosities at
varying redshifts.
Once the survey ux-limit and the sky coverage are dened one can com-
pute the maximum search volume, V
max
, within which a cluster of a given
luminosity is found in that survey:
V
max
=
_
z
max
0
S[f(L, z)]
_
d
L
(z)
1 +z
_
2
c dz
H(z)
. (31)
10
7
10
8
10
9
0.0 0.5 1.0 1.5
1
10
L
x
= 3 10
44
erg s
1
V
(
>
z
)
m
= 0.3,
= 0.7
100
Redshift
V
o
l
u
m
e
(
>
z
)
(
h
1
M
p
c
)
3
Flux (0.52 keV) erg cm
2
s
1
10
12
10
13
10
14
1
10
A
r
e
a
(
d
e
g
2
)
100
EMSS
160 sqdeg
NEP
RDCS
EMSS
160 sqdeg
NEP
RDCS
1000
5
0
Fig. 6. Left panel: sky coverage as a function of X-ray ux of several serendipitous
surveys. Right panel: corresponding search volumes, V (> z), for a cluster of given
X-ray luminosity (L
X
= 3 10
44
[0.52 keV] L
X
). From [138]
306 S. Borgani
Here S(f) is the survey sky coverage, which depends on the ux f = L/(4d
2
L
),
d
L
(z) is the luminosity distance, and H(z) is the Hubble constant at z. We
dene z
max
as the maximum redshift out to which the ux of an object of
luminosity L lies above the ux limit. The corresponding survey volumes are
shown in the right panel of Fig. 6.
Once again, I emphasize that one of the main advantages of the Xray
selection lies in the fact that the survey selection function can be precisely
computed, thus allowing reliable comparisons between the observed and the
predicted evolution of the cluster population.
4.3 Identication Through the SZ Eect
The SunyaevZeldovich (SZ) eect [154] allows to observe galaxy clusters
by measuring the distortion of the CMB spectrum owing to the hot ICM.
This method does not depend on redshift and provides in principle a reliable
estimate of cluster masses. For these reasons, it is now considered as one of
the most powerful means to nd distant clusters in the years to come. For a
detailed discussion of the SZ technique for cluster identication and for the
ongoing and future surveys, I refer to the lectures by Mark Birkinshaw and to
the reviews in [15, 37]. For the purpose of the present discussion, I show in the
left panel of Fig. 7 a comparison between the limiting mass as a function of
redshift, expected for a Xray and for a SZ cluster survey (from [80]). While
the standard ux dimming with the luminosity distance, f
X
d
2
L
(z), causes
3 10
14
Y
m
a
p
Y
i
n
t
(
M
p
c
/
h
)
2
10
12
10
11
10
10
10
7
10
6
M
200
(M /h)
10
15
2
SZE
Xray
w=0.5
w=0.5
CDM
CDM
z
M
l
i
m
/
M
1 0
10
14
10
15
10
16
5 10
14
5 10
14
5 10
15
Fig. 7. Left panel: limiting cluster virial mass for detection in an Xray and in
a SZ survey (from [80]). Each pair of curves show the results for two
m
= 0.3
cosmologies, having w = 1 and w = 0.5 for the DE equation of state. Right
panel: the relation between the Comptonization parameter and M
200
, from [167].
The upper panel shows the decrement contributed from the gas within 0.5R
200
. The
lower panel indicates the signal from noisefree maps projected on the light cone
Clusters and Cosmology 307
the limiting mass to quickly increase with distance for the Xray selection,
this limiting mass has a much less sensitive dependence on redshift for the
SZ selection. This is the reason why SZ surveys are generally considered as
essentially providing masslimited cluster samples.
It has been recently pointed out [115] that the integrated SZ ux
decrement has a very tight correlation with the total cluster mass (see also
[47]). This fact, joined with the redshiftindependence of the SZ selection,
makes the SZ identication a promising route toward precision cosmology
with galaxy clusters.
A potential problem with the SZ identication of clusters resides in the
possible contamination of the signal from foreground/background structures.
Diuse gas, residing in largescale laments, are likely to provide a negligible
contamination, as a consequence of the comparatively low density and tem-
perature which characterize such structures. However, small halos, which are
expected to be present in large number, contain gas at the virial overdensity.
Since they are not resolved in current SZ observations, their integrated con-
tribution may provide a signicant contamination. Using cosmological hydro-
dynamical simulations, White et al. [167] have created SZ sky maps with the
aim of correlating the SZ signal seen in projection with the actual mass of
clusters.The result of this test is shown in the right panel of Fig. 7. The upper
panel shows the relation between the integrated SZ signal contributed only
from the gas within 0.5R
200
and M
200
, while the lower panel is when using the
actual Compton-y parameter measured from the projected maps. Quite appar-
ently, the scatter in the relation is signicantly increased in projection. Part of
the scatter is due to the dierent redshifts at which clusters seen in projection
are placed. This contribution to the scatter can be removed once redshifts of
clusters are known from followup optical observations. However, a signicant
contribution to the overall scatter is contributed by cluster asphericity and by
contamination from fore/background structures. This highlights the relevance
of keeping this scatter under control for a full exploitation of the SZ signal as
a tracer of the cluster mass.
5 Methods to Estimate Cluster Masses
5.1 The Hydrostatic Equilibrium
The condition of hydrostatic equilibrium determines the balance between the
pressure force and the gravitational force: P
gas
=
gas
, where P
gas
and
gas
are the gas pressure and density, respectively, while is the underlying
gravitational potential. Under the assumption of a spherically symmetric gas
distribution, the above equations read:
dP
gas
dr
=
gas
d
dr
=
gas
GM(< r)
r
2
, (32)
308 S. Borgani
where r is the radial coordinate (cluster-centric distance) and M(< r) is the
total mass contained within r. Using the equation of state of ideal gas to relate
pressure to gas density and temperature, the mass is then given by
M(< r) =
r
G
k
B
T
m
p
_
d ln
gas
d ln r
+
d lnT
d ln r
_
, (33)
where is the mean molecular weight of the gas ( 0.59 for primordial
composition) and m
p
is the proton mass. An often used mass estimator is
based on assuming the model for the gas density prole,
gas
(r) =
0
[1 + (r/r
c
)
2
]
3/2
(34)
[39]. In the above equation, r
c
is the core radius, while is the ratio between
the kinetic energy of any tracer of the gravitational potential (e.g. galaxies)
and the thermal energy of the gas, = m
p
2
v
/(k
B
T) (
v
: onedimensional
velocity dispersion). By further assuming a polytropic equation of state,
gas
P
gas
(: polytropic index), (33) becomes
M(< r) 1.11 10
14
T(r)
keV
r
h
1
Mpc
(r/r
c
)
2
1 + (r/r
c
)
2
M
, (35)
where T(r) is the temperature at the radius r. In its original derivation, the
model was aimed at representing the distribution of isothermal gas sitting
in hydrostatic equilibrium within a Kinglike potential. The corresponding
mass estimator is recovered from (35) by setting = 1 and replacing T(r)
with the global ICM temperature, T
0
. In the absence of accurately resolved
temperature proles from Xray observations, (35) has been used to estimate
cluster masses both in its isothermal (e.g., [136]) and in its polytropic form
(e.g., [56, 62, 120]).
Thanks to the much improved sensitivity of the Chandra and XMM
Newton Xray observatories, temperature proles are now resolved with high
enough accuracy to allow the application of more general methods of mass
estimation (providing tight ML
x
relations; see for example Fig. 8), not nec-
essarily bound to the assumptions of model and of an overall polytropic
form for the equation of state (e.g., [5, 8, 56, 160]).
An alternative way of recasting the isothermal version of (35) between
temperature and mass is based on expressing the mass according to the virial
theorem as M
vir
=
2
v
R
vir
/G, so that
k
B
T =
1.38
_
M
vir
10
15
M
_
3/2
[
m
vir
(z)]
1/3
(1 +z) keV. (36)
This expression, originally introduced in [54], has been sometimes used to
express the MT relation as obtained from hydrodynamical simulations of
galaxy clusters (e.g., [25, 31]).
Clusters and Cosmology 309
10
10
10
13
10
14
10
15
10
14
10
15
kT (keV)
T (keV)
M
h
(
z
)
,
M
h(z) M
500
T
spec
0.1 h(z) M
500
T
mg
h
(
z
)
M
(
M
)
= 2500
Fig. 8. The mass-temperature relation for nearby clusters (from [8]) and for distant
clusters (from [95]), based on a combination of Chandra and XMMNewton data
It is clear that the two crucial assumptions underlying any mass measure-
ments based on the ICM temperature concerns the existence of hydrostatic
equilibrium and of spherical symmetry. While eects of nonspherical geom-
etry can be averaged out by performing the analysis over a large enough
number of clusters, the former can lead to systematic biases in the mass
estimates (e.g., [133]) and references therein). So far, ICM temperature mea-
surements have been based on ts of the observed Xray spectra of clusters
to plasma models, which are dominated at high temperatures by thermal
bremsstrahlung. However, local deviations from isothermality, e.g. due to the
presence of merging cold gas clumps, can bias the spectroscopic temperature
with respect to the actual electron temperature (e.g., [108, 110, 159]). This
bias directly translates into a comparable bias in the mass estimate through
hydrostatic equilibrium (see Sect. 7, below).
5.2 The Dynamics of Member Galaxies
From a historical point of view, the dynamics traced by member galaxies, has
been the rst method applied to measure masses of galaxy clusters [148, 171].
Under the assumption of virial equilibrium, the mass of the cluster can be
estimated by knowing position and redshift for a high enough number of
member galaxies:
M =
2
3
2
v
R
V
G
(37)
(e.g., [99]), where the rst factor accounts for the geometry of projection,
r
is the line-of-sight velocity dispersion and R
V
is the virialization radius,
310 S. Borgani
which depends on the positions of the galaxies with measured redshifts and
recognized as true cluster members:
R
V
= N
2
_
_
i>j
r
1
ij
_
_
1
, (38)
where N is the total number of galaxies, and r
ij
the projected separation
between the i-th and j-th galaxies. This method has been extensively applied
to measure masses for statistical samples of both nearby (e.g., [16, 17, 71, 130,
137]) and distant (e.g., [36, 73]) clusters.
Besides the assumption of virial equilibrium, which may be fullled to dif-
ferent degrees by dierent populations of galaxies (e.g., late vs. early type),
a crucial aspect in the application of the dynamical mass estimator concerns
the rejection of interlopers, i.e. of back/foreground galaxies which lie along
the line-of-sight of the cluster without belonging to it. A spurious inclusion
of nonmember galaxies in the analysis leads in general to an overestimate
of the velocity dispersion and, therefore, of the resulting mass. A number
of algorithms have been developed for interlopers rejection, whose reliability
must be judged on a case-by-case basis (e.g., [68, 156]). A further poten-
tial problem of this analysis concerns the possibility of realizing a uniform
sampling of the cluster potential using galaxies with measured redshifts. For
instance, the technical diculty of packing slits or bers in optical spectro-
scopic observations may lead to an undersampling of the cluster central re-
gions. In turn, this leads to an overestimate of R
V
and, again, of the collapsed
mass.
Tests of the accuracy of mess estimates based on the dynamical virial
method have been performed by using hydrodynamical simulations of galaxy
clusters, in which galaxies are identied from gas cooling and star forma-
tion [18, 63]. For instance, [18] have shown that galaxies identied in the
simulations are fair tracers of the underlying dynamics, with no systematic
bias in the estimate of cluster masses, although a rather large scatter between
true and recovered masses is induced mostly by projection eects.
Quite reassuringly, despite all the assumptions and possible systematics
aecting both dynamical optical and X-ray mass estimates, these two methods
provide in general fairly consistent results for both nearby (e.g., [71, 130]) and
distant (e.g., [96]) clusters. Two examples of such comparisons are shown in
Fig. 9. In the left panel, we report the comparison between Xray and optical
dynamical masses [71]. This plot shows a reasonable agreement among the
two mass estimates, although with some scatter. The right panel reports the
comparison presented in [130]. In this plot, the triangles indicates the cluster
with clear evidences of complex dynamics. Quite interestingly, the agreement
between the two mass estimates is acceptable, with a few outliers which are
generally identied with nonrelaxed clusters.
Clusters and Cosmology 311
0
1
2
3
5
optical M
500
/(10
14
M )
o
p
t
i
c
a
l
M
5
0
0
/
M
5
0
0
(
f
r
o
m
T
x
)
10 15
16 15 14 13 12
12
13
14
15
16
log (M
X
/M )
l
o
g
(
M
o
p
t
/
M
)
Fig. 9. The relation between dynamical optical masses and masses derived from
the X-ray temperature by assuming hydrostatic equilibrium (from [71], left panel,
and from [130], right panel, based on SDSS spectroscopic data)
5.3 The SelfSimilar Scaling
The simplest model to explain the physics of the ICM is based on the assump-
tion that gravity only determines the thermodynamical properties of the hot
diuse gas [90]. Since gravity does not have a preferred scale, we expect clus-
ters of dierent sizes to be the scaled version of each other as long as gravity
only determines the ICM evolution and there are no preferred scales in the
underlying cosmological model. This is the reason why the ICM model based
on the eect of gravity only is said to be self-similar.
If we dene M
c
as the mass contained within the radius R
c
, encompass-
ing a mean density
c
times the critical density, then M
c
c
(z)
c
R
3
c
.
Here
c
(z) is the critical density of the universe which scales with redshift as
c
(z) =
c,0
E
2
(z), where E(z) is given by (12). On the other hand, the clus-
ter size R scales with z and M
c
as R M
1/3
E
2/3
(z). Therefore, assuming
hydrostatic equilibrium, the cluster mass scales with the temperature T as
M
c
T
3/2
E
1
(z) . (39)
If
gas
is the gas density, the corresponding X-ray luminosity for pure thermal
bremsstrahlung emission is
L
X
=
_
V
_
gas
m
p
_
2
(T) dV , (40)
where (T) T
1/2
. Further assuming that the gas distribution traces the
dark matter distribution,
gas
(r)
DM
(r), then
L
X
M
c
T
1/2
T
2
E(z) . (41)
312 S. Borgani
As for the CMB intensity decrement due to the thermal SZ eect we have
S
_
y()d d
2
A
_
Tn
e
d
3
r d
2
A
T
5/2
E
1
(z) , (42)
where y is the Comptonization parameter, d
A
is the angular size distance and
n
e
is the electron number density. We can also write S in a dierent way to
get the explicit dependence on y
0
:
S y
0
d
2
A
_
d y
0
d
2
A
M
2/3
E
4/3
(z) y
0
d
2
A
TE
2
(z) . (43)
In this way, we obtain the following scalings for the central value of the Comp-
tonization parameter:
y
0
T
3/2
E(z) L
3/4
X
E
1/4
(z) . (44)
Equations (39), (41) and (44) are unique predictions for the scaling rela-
tions among ICM physical quantities and, in principle, they provide a way
to relate the cluster masses to observables at dierent redshifts. As we shall
discuss in the following, deviations with respect to these relations witness the
presence of more complex physical processes, beyond gravitational dynamics
only, which aect the thermodynamical properties of the diuse baryons and,
therefore, the relation between observables and cluster masses.
5.4 Phenomenological Scaling Relations
Using the X-ray Luminosity
The relation between X-ray luminosity and temperature of nearby clusters
is considered as one of the most robust observational facts against the self
similar model of the ICM. A number of observational determinations now ex-
ist, pointing toward a relation L
X
T
>
10 keV [3]. While in general the scatter around the besttting relation
is non negligible, it has been shown to be signicantly reduced after excising
the contribution to the luminosity from the cluster cooling regions [106] or by
removing from the sample clusters with evidence of cooling ows [7]. As for
the behaviour of this relation at the scale of groups, T
<
1 keV, the emerging
picture now is that it lies on the extension of the L
X
T relation of clusters,
with no evidence for a steepening [116], although with a signicant increase
of the scatter [121], possibly caused by a larger diversity of the groups popu-
lation when compared to the cluster population. This result is reported in the
left panel of Fig. 10 (from [121]), which shows the L
X
T relation for a set of
clusters with measured ASCA temperatures and for a set of groups.
As for the evolution of the L
X
T relation, a number of analyses have been
performed, using Chandra [57, 86, 109, 161] and XMMNewton [95, 101] data.
Clusters and Cosmology 313
42
44
46
0.5
1
1.5
2
2.5
3
3.5
0 0.2 0.4 0.6 0.8 1 1.2 1.4
L
/
T
B
z
(1+2)
1.2
E(z)(
v
(z)/
v
(0))
1/2
t(0)
2
/(E(z)
3
t(z)
2
) altered similarity
E(z)
t(0)/(E(z)t(z)) cooling threshold
binned data compared with AE99
binned data compared with Markevitch (1998)
1 0.5
log T
X
(keV)
l
o
g
L
X
(
r
5
0
0
)
(
e
r
g
s
1
)
0 0.5
Fig. 10. Left panel: the L
X
T relation for nearby clusters and groups (from [121]).
The star symbols are the for the sample of clusters, with temperature measured from
ASCA, while the lled squares and open circles are for a sample of groups, also with
ASCA temperatures. Right panel: the evolution of the L
X
T relation, normalized
to the local relation (from [109]), using Chandra temperatures of clusters at z > 0.4
Although some dierences exist between the results obtained from dierent
authors, such dierences are most likely due to the convention adopted for
the radii within which luminosity and temperature are estimated. In general,
the emerging picture is that clusters at high redshift are relatively brighter,
at xed temperature. The resulting evolution for a cosmology with
m
= 0.3
and
2
.
4
k
e
V
)
)
[
h
2
1
0
4
0
e
r
g
s
1
]
Fig. 11. The L
X
M relation for nearby clusters (from [136]). Xray luminosities are
from the RASS, while masses are estimated using ASCA temperatures and assuming
hydrostatic equilibrium for isothermal gas
hand, the slope of the relation is found to be steeper than the selfsimilar
scaling, thus consistent with the observed L
X
T relation.
Using the Optical Luminosity
The classical denition of optical richness of clusters is known to be a poor
tracer of the cluster mass (e.g., [26]). However, the increasing quality of pho-
tometric data for the cluster galaxy population and the ever improving ca-
pability of removing fore/background galaxies thanks to larger spectroscopic
galaxy samples have recently allowed dierent authors to demonstrate the
optical/near-IR luminosities to be as reliable tracers of the cluster mass as
the X-ray luminosity.
Two examples of recent calibrations between optical/nearIR luminosity
and mass are shown in Fig. 12. In the left panel we report the result presented
in [100], based on Kband luminosites from the 2MASS and masses obtained
by applying the MT relation by [62]. Although the data points are rather
scattered, they dene a clear correlation. Quite interestingly, the besttting
relation has a slope shallower than unity, thus indicating the Kband mass-
to-light ratio is a (slightly) increasing function of the cluster mass. This result
is in line with previous results using optical luminosities (e.g., [72]) who found
an increasing M/L when passing from galaxy groups to clusters of increasing
richness.
Popesso et al. [130] analysed SDSS data for a set of clusters which have
been identied in the RASS. Their mass estimates come from both Xray
temperature [136] and from the velocity dispersions as estimated from the
SDSS spectroscopic data. The results of their analysis for the i band are shown
Clusters and Cosmology 315
0.01
0.1
1
10
100
10
2 4 6
kT
X
(keV)
M
500
/(10
14
M )
L
o
p
t
/
(
1
0
1
2
L
)
L
5
0
0
(
h
2
1
0
1
2
L
)
8 10
5
M
500
(h
1
10
14
M )
70
7
0
1
1
5
10
100 10 1 0.1 0.01
Fig. 12. The relation between cluster masses and optical/near-IR luminosities. Left
panel: M
500
L
500
relation from [100], using Kband data from the 2MASS survey
and masses from Xray data. The solid line is the bestt power law, while the
dashed line marks the unity slope. Right panel: the same relation, but in the
iSloan band. Open and lled points corresponds to mass estimates based on the
SDSS spectroscopic data and on the MT
X
relation, respectively
in the right panel of Fig. 12. Again, the optical luminosity correlates quite
tightly with the cluster mass, with an intrinsic scatter which is comparable
to, or even smaller than that of the correlation between Xray luminosity and
mass.
These results highlight how cluster samples with precisely measured optical
luminosities can in principle be usefully employed to constrain cosmological
parameters. However, while X-ray luminosity provides at the same time a
tracer of cluster mass and a criterion to precisely determine the sample selec-
tion function, the latter quantity can be extracted from an optically selected
sample only in a rather indirect way.
6 Constraints on Cosmological Parameters
In this section we will review critically results on cosmological constraints
derived from dierent ways of tracing the cosmological mass function of galaxy
clusters.
6.1 The Distribution of Velocity Dispersions
A rst determination of the mass function from velocity dispersions,
v
, of
member galaxies has been attempted in [17]. Girardi et al. [69] used a much
larger sample of nearby clusters with measured velocity dispersions to compare
316 S. Borgani
the resulting mass function with predictions from cosmological models. The
resulting relation between
8
and
m
was such that
8
1 for a ducial value
of the density parameter
m
= 0.3. More recently, data for nearby clusters,
identied in the SDSS, have been used to calibrate a relation between richness
and velocity dispersion [10]. They compared the resulting
v
distribution to
the prediction of cosmological models and found a signicantly lower normal-
ization of the power spectrum,
8
0.7 for
m
= 0.3. Such dierences from
dierent analyses highlight the presence of systematic uncertainties in the
relation between mass and observables (i.e., velocity dispersion and richness).
The application of this method to distant clusters has been applied so far
only to the CNOC sample [34], which comprises 17 clusters selected from the
EMSS out to z 0.6. Still to date, this is the only sample of distant clusters,
with calibrated selection function, for which velocity dispersions have been
reliably measured. Bahcall et al. [11] pointed out that the resulting evolution
of the mass function is consistent with a lowdensity Universe. Borgani et
al. [24] reanalysed this same sample and emphasised that the uncertainties in
the local normalization of the mass function are large enough to make any
constraints on
m
not signicant.
6.2 The Temperature Function
The X-ray Temperature Function (XTF) is dened as the number density of
clusters with given temperature, n(T). As long as a one-to-one relation exist
between temperature and mass, the XTF can be related to the mass function,
n(M), by the relation
n(T) = n[M(T)]
dM
dT
. (45)
In this equation, the ratio dM/dT is provided by the relation between ICM
temperature and cluster mass.
Measurements of cluster temperatures for ux-limited samples of nearby
clusters were rst presented in [83]. These results have been subsequently
rened and extended to larger samples with the advent of ROSAT, Beppo
SAX and, especially, ASCA. XTFs have been computed for both nearby (e.g.,
[88, 106, 127, 128]) and distant (e.g., [50, 55, 81, 82]) clusters, and used to
constrain cosmological models. The starting point in the computation of the
XTF is inevitably a ux-limited sample for which the searching volume of
each cluster can be computed. Then the L
X
T
X
relation and its scatter is
used to derive a temperature limit from the sample ux limit.
Once the XTF is measured from observations, (45) is used to infer the mass
function and, therefore, to constrain cosmological models. A slightly dierent
but conceptually identical approach, has been followed in [136], where masses
for a uxlimited sample of nearby bright RASS clusters have been computed
by applying the assumption of hydrostatic equilibrium, thereby expressing
their results directly in terms of mass function, rather than of XTF.
Clusters and Cosmology 317
Oukbir and Blanchard [122] rst suggested to use the evolution of the XTF
as a way to constrain the value of
m
. Several independent analyses converge
now towards a mild evolution of the XTF, which is interpreted as a case for
a lowdensity Universe, with 0.2
<
m
<
0.6. An example is reported in the
right panel of Fig. 13 (from [82]), which shows the comparison between the
XTFs of the sample of nearby clusters [83] and a sample of EMSS clusters
with ASCA temperatures (see also Fig. 14).
A limitation of the XTFs presented so far is the limited sample size (with
only a few z
>
0.5 measurements), as well as the lack of a homogeneous sample
selection for local and distant clusters. By combining samples with dierent
selection criteria one runs the risk of altering the inferred evolutionary pattern
of the cluster population. This can even give results consistent with a critical
density Universe [19, 41, 158].
Besides the determination of the matter density parameter, the obser-
vational determination of the XTF also allows one to measure the normal-
ization of the power spectrum,
8
. Assuming a ducial value of
m
= 0.3,
dierent (sometimes discrepant) determinations of
8
have been reported by
dierent authors, ranging from
8
0.70.8 (e.g., [55, 82, 136]) to
8
1
(e.g., [106, 128]). Ikebe et al. [88] compared dierent observational determina-
tions of the XTF for nearby clusters (see left panel of Fig. 13) and established
that they all agree with each other reasonably well. Although quite comfort-
able, this result highlights that the discrepant results on the normalization
of the
8
m
relation comes from the cosmological interpretation of the ob-
served XTF, and not from observational uncertainties in its calibration. While
the dierent model mass functions (i.e., whether PressSchechter, Jenkins et
al. or ShethTormen) can in some cases account for part of the dierence,
more in general the dierent results are interpreted in terms of the dierent
10
2
10
9
10
8
10
7
10
6
10
5
10
T (keV)
kT (keV)
n
(
>
k
T
)
(
h
3
M
p
c
3
)
n
(
>
T
)
(
h
3
M
p
c
3
)
5
1
0
9
1
0
8
1
0
7
1
0
6
1
0
5
1
0
4
1 2
Fig. 13. Left panel: a comparison between the XTF for nearby clusters from [88]
(shaded area), [106] (solid line)and [81] (dotted line). Right panel: the evolution of
the XTF from [82]. Open and lled circles are for the local and the distant cluster
sample, respectively
318 S. Borgani
160SD (0.3 < z < 0.6)
160SD (0.6 < z < 0.8)
[
h
5
0
5
M
p
c
3
(
1
0
4
4
e
r
g
s
s
1
)
1
]
L
X
[0.52.0 keV) [h
50
2
ergs s
1
]
10
5
10
6
10
7
10
8
10
9
10
10
10
11
10
43
10
44
10
45
RDCS (0.25 < z < 0.50)
RDCS (0.50 < z < 0.85)
EMSS (0.30 < z < 0.60)
NEP (0.30 < z < 0.80)
WARPS (0.30 < z < 0.50)
WARPS (0.50 < z < 0.95)
SSHARC (0.30 < z < 0.70)
BSHARC (0.30 < z < 0.70)
MACS (0.50 < z < 0.60)
Fig. 14. A compilation of XLF within dierent redshift intervals for independent
X-ray ux-limited surveys (from [119]). The shaded area shows the range of deter-
minations of the local XLF, while the solid curve is the besttting evolving XLF
from [119]
normalization of the MT relation to be used in (45) or to the way in which
the intrinsic scatter and the statistical uncertainties in this relation are in-
cluded in the analysis. We shall critically discuss these issues in Sect. 6.5
below.
Substantially improved observational determinations of the XTF, and cor-
respondingly tighter cosmological constraints, are expected to emerge with the
accumulations of data on the ICM temperature from the Chandra and XMM
Newton satellites. Thanks to the much improved sensitivity of these Xray
telescopes with respect to ASCA, temperature gradients can be measured for
fairly large sets of nearby and mediumdistant (z
<
0.4) clusters, thus allow-
ing more precise determinations of cluster masses. At the same time, reliable
measurements of global temperatures are now emerging for clusters out to the
highest redshifts where they have been secured (e.g., [140]). At the time of
writing, several years after the advent of the new generation of Xray tele-
scopes, no determinations of the XTF from Chandra and XMMNewton data
have been presented, a situation that is expected to change quite soon.
6.3 The Luminosity Function
Another method to trace the evolution of the cluster number density is based
on the Xray luminosity function (XLF), (L
X
), which is dened as the num-
ber density of galaxy clusters having a given Xray luminosity. Similarly to
(45), the XLF can be related to the cosmological mass function of collapsed
halos as
Clusters and Cosmology 319
(L
X
) = n[M(L
X
)]
dM
dL
X
, (46)
where M(L
X
) provides the relation between the observable L
X
and the cluster
mass. The above relation needs to be suitably modied in case an intrinsic
scatter exists in the relation between mass and temperature (see Sect. 6.5,
here below).
A useful observational quantity, that is related to the XLF, is given by
the ux numbercounts, n(S), which is dened as the number of clusters per
steradian, having measured ux S:
n(S) =
_
c
H
0
_
3
_
0
dz
r
2
(z)
E(z)
n[M(S, z); z]
dM
dS
(47)
(e.g., [94]) where r(z) is the radial coordinate appearing in the Friedmann
RobertsonWalker metric:
r(z) =
_
z
0
dz E
1
(z) ;
= 1
m
r(z) =
2
_
m
z + (2
m
) (1
1 +
m
z)
2
m
(1 +z)
;
= 0 . (48)
The ux S is related to the luminosity according to
S =
L
X
4d
2
L
(z)
, (49)
where d
L
(z) = r(z)(1 +z) is the luminosity distance at redshift z.
This quantity can be measured for a uxlimited samples without having
information on cluster redshift and provides useful cosmological information in
the absence of any spectroscopic optical followup. A comparison between dif-
ferent observational determinations of the ux number counts for both nearby
and distant cluster samples (e.g., [138]) show indeed a quite good agreement.
Another quantity, which has been used to derive cosmological constraints
from uxlimited surveys, is the redshift distribution, n(z), which is dened
as the number of clusters found in a survey at a given redshift z:
n(z) =
_
c
H
0
_
3
r
2
(z)
E(z)
_
S
lim
dS f
sky
(S) n[M(S, z); z]
dM
dS
. (50)
In the above expression S
lim
is the limiting completeness ux of the survey,
while f
sky
(S) is the eective uxdependent skycoverage appropriate for the
considered survey. Convolving the mass function with the sky coverage inside
the integral in the above equation is essential to properly account for the dif-
ferent eective area covered at dierent uxes, an aspect which is apparently
overlooked in some analyses (e.g., [157]).
320 S. Borgani
In general, the advantage of using X-ray luminosity as a tracer of the mass
is that L
X
is measured for a much larger number of clusters in samples with
well-dened selection properties. As discussed in Sect. 4.2, the most recent
uxlimited cluster samples contain now a fairly large ( 100) number of
objects, which are homogeneously identied over a broad redshift baseline,
out to z 1.3. This allows nearby and distant clusters to be compared within
the same sample, i.e. with a single selection function. However, since the X-ray
emissivity depends on the square of the gas density, the relation between L
X
and M
vir
, which is based on additional physical assumptions, is more uncertain
than the M
vir
v
or the M
vir
T relations.
A useful parametrization for the relation between temperature and bolo-
metric luminosity can be casted in the form
L
bol
= L
6
_
T
X
6keV
_
(1 +z)
A
_
d
L
(z)
d
L,EdS
(z)
_
2
10
44
h
2
erg s
1
, (51)
with L
6
dening the normalization of the relation and d
L
(z) the luminosity
distance at redshift z for a given cosmology.
Analyses of the number counts from dierent X-ray uxlimited cluster
surveys showed that the resulting constraints on
m
are rather sensitive to
the evolution of the massluminosity relation [24, 94, 107]. On the other hand,
other authors [135, 141, 157] analysed dierent uxlimited surveys and found
results consistent with
m
= 1. Quite intriguingly, this conclusion is common
to analyses which combine a normalization of the local mass function, using
nearby clusters, and to the evolution of the mass function using deep surveys.
Clearly, any uncertainty in the calibration of the selection functions when
combining dierent surveys may induce a spurious signal of evolution of the
cluster population, possibly misinterpreted as an indication for high
m
.
In order to overcome this potential problem, Borgani et al. [29] ana-
lyzed the RDCS sample to trace the cluster evolution over the entire redshift
range, 0.05
<
z
<
1.3, probed by this survey, without resorting to any exter-
nal normalization from a dierent survey of nearby clusters [28]. They found
0.1
<
m
<
0.6 at the 3 condence level, by allowing the ML
X
relation to
change within both the observational and the theoretical uncertainties. In
Fig. 15 we show the resulting constraints on the
8
m
plane (from [138])
and how they vary by changing the parameters dening the ML
X
relation:
the slope and the evolution A of the L
X
T relation (see Equation 51),
the normalization of the MT relation (see (36)), and the overall scatter
ML
X
. Flat geometry is assumed here, i.e.
m
+
= 1.
Similar results have been obtained by combining information on clustering
properties and the redshift distribution from the the REFLEX cluster survey
[144], thus providing
8
0.7 and
m
0.35. One should however notice that,
since these constraints are derived from nearby clusters, the corresponding
estimate of
m
comes from the shape of the CDM power spectrum, rather than
from the growth rate of perturbations. It is rather reassuring that dynamical
and geometrical constraints on
m
are in fact consistent with each other.
Clusters and Cosmology 321
A = 1
= 3
A = 0
= 1.15
ML
X
= 45%
= 2.5
= 1.5 = 1.15 = 0.8
ML
X
= 20%
ML
X
= 20%
1.2
1
0.8
0.6
0.4
1
0.8
0.6
0.4
0.2 0.4 0.6 0.8
m
m
m
0.2 0.4 0.6 0.8 0.2 0.4 0.6 0.8 1
8
Fig. 15. Probability contours in the
8
m
plane from the evolution of the X-ray
luminosity distribution of RDCS clusters. The shape of the power spectrum is xed
to = 0.2 [138]. Dierent panels refer to dierent ways of changing the relation
between cluster virial mass, M, and X-ray luminosity, L
X
, within theoretical and
observational uncertainties (see also [29]). The upper left panel shows the analysis
corresponding to the choice of a reference parameter set. In each panel, we indi-
cate the parameters which are varied, with the dotted contours always showing the
reference analysis
Constraints on
m
from the cluster Xray luminosity and temperature dis-
tribution are thus in line with the completely independent constraints derived
from the baryon fraction in clusters, f
b
(e.g., [4, 58, 168]).
6.4 The Gas Mass Function
An alternative way of tracing the mass function of galaxy clusters is based
on using as its proxy the mass function of the cluster gas content [162]. This
method is based on the assumption that galaxy clusters are fair containers
of cosmic baryons. Similarly to the method based on the baryon fraction,
it relies on the knowledge of the cosmic baryon fraction, either provided by
data on the deuterium abundance in highredshift absorption systems (e.g.,
[92]) combined with predictions of primordial nucleosynthesis, or from the
spectrum of CMB anisotropies (e.g., [103, 149]). This method has the potential
advantage that cluster gas mass is an easier quantity to measure than the total
collapsed mass, since it is essentially related to the total cluster emissivity.
322 S. Borgani
If we dene n
b
(M
b
) to be the baryonic mass function and n(M) the total
mass function, by denition we have n
b
(M
b
) = n(
m
M
b
/
b
). Therefore, once
n
b
(M
b
) and
b
are known from observations, the total mass function can be
computed as a function of
m
, thereby treated as a tting parameter.
While this method has the remarkable advantage of avoiding the uncer-
tainties related to direct estimates of the total collapsed mass, it is aected
by possible violations of the assumption of universality of the baryon content
of clusters. Indeed, while this assumption should be valid for suitably relaxed
and massive clusters, it may be less so when considering all objects belong-
ing to a uxlimited sample, thus including also relatively small clusters and
structures with a complex non-relaxed dynamics.
This method was applied [163] to a set of bright clusters selected from the
RASS [136] and found
8
= 0.72 0.04 with
m
h = 0.13 0.07. Chandra
observations were included in the analysis for a set of clusters extracted from
the 160 deg
2
survey [162]. They found an evolution of the gas mass function,
which is consistent with a at cosmological model with
m
= 0.3.
We emphasize here that the above dierent methods used to reconstruct
the mass function of galaxy clusters consistently prefer relatively low values
of
8
, in the range 0.70.8. Quite remarkably, such values have been shown
now to be required by the 3years WMAP data release [149].
6.5 Including Uncertainties in the Analysis
As we have discussed in the previous sections, most of the analyses of the
cluster populations converge toward a low-density model, with
m
0.3.
However, signicant dierences exist between dierent determinations of the
normalization of the power spectrum,
8
, which amount to up to 20 per cent.
These dierences are much larger that the statistical uncertainties associated
to the nite number of clusters included in the samples, thus indicating that
they arise from unaccounted sources of error, which aects the analyses.
For instance, the role of the uncertain normalization of the mass
temperature relation in the determination of
8
(at xed
m
) from the XTF
analysis has been emphasized by dierent authors (e.g., [88, 127, 145]). In-
creasing the normalization of the MT relation implies that a larger mass
corresponds to a xed temperature value. As a consequence, an observed XTF
translates into a larger mass function, therefore implying a larger
8
(at xed
m
). Since results from hydrodynamical simulations generally imply a larger
MT normalization, a larger
8
is expected when using in the analysis the
simulation predictions. The left panel of Fig. 16 (from [88]) shows how the best
tting values of
8
and
m
from the local XTF change as one uses dierent
masstemperature relations, taken from both observations and simulations.
The right panel of Fig. 16 (from [127]) show the dependence of
8
on the
normalization of the masstemperature relation. Note that this normalization
Clusters and Cosmology 323
0.90
0.85
0.80
0.75
0.70
0.65
1.4 1.6 1.8
2.0 2.2
T
*
Hydro (max)
Hydro (mean) Nevalainen
Finoguenov
Hydro (min)
Horner
8
1.4
1.2
1
0.8
0.6
0
0 0.1 0.2 0.3 0.4 0.5
Fig. 16. Left panel: the dependence of the best-tting values of
m
and
8
,
from the XTF of [88], upon dierent determinations of the masstemperature rela-
tion, from both observational data and from hydrodynamical simulations of galaxy
clusters. Right panel: the dependence of
8
on the normalization of the mass
temperature relation, in the XTF analysis by [127]
is allowed here to vary over the range encompassed by observational and sim-
ulation results. The range of variation of
8
induced by the uncertainty in the
MT relation is at least comparable to the purely statistical uncertainty, as
indicated by the errorbars.
One may wonder why not relying only on the observational determination
of the MT relation, instead of considering also simulation results. As we shall
discuss in Sect. 7, observational results can not necessarily provide the most
reliable determination of the MT scaling.
The eect of changing the normalization of the MT relation on the anal-
ysis of uxlimited samples through the XLF evolution is shown in the lower
left panel of Fig. 15. If this normalization is reduced by 30%, the resulting
8
decreases by 20%.
It is clear that, any uncertainty, both statistical and systematic, in the
tting parameters describing the scaling relations between mass and observ-
able, must be included in the analysis by marginalizing over the probability
distribution function of these parameters. Let be the set of cosmological
parameters that we want to constrain, and W the set of parameters which
dene a scaling relation between mass M and an observable X (i.e.,
v
, L
X
or T). Let us also call P(W) the prior distribution for the uncertainties in the
MX relation. If
2
(, W) gives the goodness of t provided by the choice
of the cosmological parameters, for a given MX relation, then the goodness
of t after marginalizing over the uncertainties in MX reads
324 S. Borgani
2
() =
_
2
(, W) P(W) dW
_
P(W) dW
. (52)
Of course, the marginalization generally induces an increase of the uncertain-
ties in the cosmological parameters. Furthermore, one needs to have a reliable
modeling of both size and distribution of the errors (i.e. whether they have a
uniform, a Gaussian, or a more peculiar distribution).
Besides the errors in the parameters dening the scaling relations, a dif-
ferent source of uncertainty is provided by the intrinsic scatter in these rela-
tions. Intrinsic scatter has the eect of widening the range of possible masses
which correspond to a given value of the observable quantity. This eect can
be included in the analysis by convolving the theoretical mass function with
the distribution of the scatter itself (e.g., [29, 98]). Let be an observable
quantity and () its distribution (i.e., XLF or XTF), to be compared with
observations. Also, let P(M
n(M
, z) P(M
|M; z)
dM
d
d (53)
where n(M, z) is the cosmological mass function at redshift z. If one makes
the standard assumption of Gaussian scatter in the loglog plane, then
P(M
|M) =
_
2
2
ln M
_
1/2
exp
_
x
2
(M
, (54)
where x(M
) = (ln M
ln M)/(
2
ln M
) and
ln M
is the r.m.s. intrinsic
scatter. The eect of this convolution is that of increasing (), for a xed
n(M), as the scatter increases. Therefore, assuming a progressively larger
scatter in the M relation implies a progressively lower
8
(at xed
m
).
An illustrative example of the eect of intrinsic scatter on the determina-
tion of
8
is reported in Fig. 17. In the left panel we show the REFLEX
XLF [20] along with the prediction of the best tting cosmological model
for a given choice of the ML
X
relation, after assuming vanishing intrin-
sic scatter in this relation. In the left panel, we show the same compari-
son, but assuming an Gaussian-distributed intrinsic scatter of 40 per cent in
the ML
X
scaling. As expected, adding the scatter has the eect of increas-
ing the predicted luminosity function, so that
8
has to be lowered from 0.8
to 0.65 to recover the agreement with observations. This example highlights
that a good calibration of the intrinsic scatter in the scaling relations can
be as important as determining the best-tting amplitude and slope of these
relations.
Clusters and Cosmology 325
d
n
/
d
L
[
1
0
4
4
e
r
g
s
1
(
h
1
M
p
c
)
3
]
1
d
n
/
d
L
[
1
0
4
4
e
r
g
s
1
(
h
1
M
p
c
)
3
]
1
0.1
0.01
0.001
0.0001
10
5
10
6
10
7
10
8
10
9
10
10
10
11
10
12
0.1
0.01
0.001
0.0001
10
5
10
6
10
7
10
8
10
9
10
10
10
11
10
12
REFLEX XLF
Boehringer et al. 02
REFLEX XLF
Boehringer et al. 02
m
=
0.3
m
=
0.3
=
0.7; h
=
0.7
=
0.7; h
=
0.7
8
=
0.8
8
=
0.65
ML
=
0%
ML
=
0%
log
10
[L
0.12.4
(erg s
1
)]
log
10
[L
0.12.4
(erg s
1
)]
41 42 43 44 45 46
41 42 43 44 45 46
Fig. 17. The dependence of
8
on the intrinsic scatter assumed in the relation
between Xray luminosity and mass. The two panels show the comparison between
model predictions and the observed XLF from the REFLEX sample [20]. The left
panel assumes vanishing scatter while the right panel assumes a 40 per cent intrinsic
scatter. The best tting values of
8
are reported in both cases
7 The Future
A new era for cosmology with galaxy clusters is now starting. High sensitivity
surveys for blind SZ identication over fairly large contiguous area, 100
deg
2
, have have already started or are planned in the coming years (see the
lectures by M. Birkinshaw in this volume). Also, the Planck satellite will
survey the whole sky, although at a much lower sensitivity, and provide a
large set of clusters identied through the SZ eect. These surveys promise
to identify several thousands clusters, with a fair number of objects expected
to be found at z > 1. In the optical/near-IR bands, imaging with dedicated
telescopes with large eld of view will also allow to secure a large number of
distant clusters. At the same time, Xray observations over contiguous area
(e.g., [169] and serendipitous searches from XMMNewton (e.g., [118]) and
Chandra (e.g., [30]) archives will ultimately cover several hundreds deg
2
down
to ux limits fainter than those reached by the deepest ROSAT pointings.
Preliminary results suggest that identication of z > 1 clusters may eventually
become routine [118]. Ultimately, they will lead to the identication of several
thousands clusters.
Optimized optics for wide-eld X-ray imaging have been originally de-
scribed in a far-reaching paper by Burrows et al. [33] and proposed for the rst
time to be implemented in a dedicated satellite mission in the mid 90s. There
is no doubt that this would be the right time to plan a dedicated wide-eld
X-ray telescope, which should survey the sky over an area of several thousands
deg
2
, with a relatively good, XMMlike or better, point spread function and
a low background. This instrument would be invaluable for studies of galaxy
326 S. Borgani
clusters, thanks to its ability of both identifying extended sources with low
surface brightness. Several missions with a similar prole have been proposed,
although none has been approved so far. Still, the community working on
galaxy clusters is regularly proposing this idea of satellite to dierent Space
Agencies (e.g., [79]).
The samples of galaxy clusters obtainable from large-area SZ and X-ray
surveys contain in principle so much information to allow one to constrain not
only
m
and
8
, but also the Dark Energy (DE) content of the Universe (see,
e.g., [142, 153] for introductory reviews on Dark Energy). The equation of
state of DE is written in the form p = w, where p and are the pressure and
density terms, respectively. The parameter w must take values in the range
1/3 > w 1 for the DE to provide an accelerated cosmic expansion.
Constraining the value of w and its redshift evolution is currently considered
one of the most ambitious targets of modern cosmology. Ones hope is to unveil
the nature of the energy term which dominates the overall dynamics of the
Universe at the present time.
As we have mentioned, the limited statistics prevents current cluster sur-
veys to place signicant constraints on
m
planes, after
marginalizing over the other tting parameters, from dierent SZ and Xray
surveys. In each panel dierent contours indicate the constraints that one can
place by progressively adding information in the analysis. The main message
here is that combining information on the evolution of the cluster population
and on its clustering can place precision constraints on cosmological parame-
ters. These constraints can be further tightened if follow up observations are
available to precisely measure masses for 100 clusters.
These forecasts nicely illustrates the potentiality of the selfcalibration ap-
proach for precision cosmology with future surveys of galaxy clusters. Clearly,
the robustness of these predictions is inextricably linked to the possibility of
accurately modeling the relations between mass and observables.
A line of attack to this problem is based on using detailed hydrodynamical
simulations of galaxy clusters. The great advantages of using simulations is
Clusters and Cosmology 327
8
0.9
0.85
0.8
0.9
0.85
0.8
0.85
0.8
0.9
SPT SPT
Planck Planck
DUET DUET
W
W
W
0.8
0.9
1
0.8
0.9
1
1
0.8
0.9
0.24 0.28 0.32 0.24 0.27 0.3
Fig. 18. Constraints on cosmological parameters from the SZ SPT survey, from
the SZ Planck survey and from a wide-area deep X-ray survey (from top to bottom
panels; from [105]). Dotted lines: no uncertainty in the relation between cluster
mass and observables, using only the dn/dz information; long-dashed line: using
self-calibration on the dn/dz; dotdashed line: as before, but also including the
information on the cluster power spectrum; shortdashed line: using dn/dz and
a calibration of the massobservable relation for 100 clusters; solid line: all the
information combined together
that both cluster mass and observable quantities can be exactly computed.
Furthermore, the eect of observational setups (e.g., response functions of
detectors, etc.) can be included in the analysis and their eect on the scaling
relations quantied. This approach has been applied by dierent groups in the
case of Xray observations [64] to understand the relation between the ICM
temperature, as measured from the tting of the observed spectrum, and
the true massweighted temperature (e.g., [108, 110, 159]). Furthermore,
simulations can also be used to verify in detail the validity of assumptions on
which the mass estimators, applied to observations, are based. The typical ex-
ample is represented by the assumption of hydrostatic equilibrium, discussed
in Sect. 5.1, for which violations in simulations at the 1020 per cent level
have been found (e.g., [12, 27, 91, 134]).
An example of calibration of observational biases in the masstemperature
relation, using hydrodynamical simulations, is shown in Fig. 19 (from [133]),
which provides a comparison between the observed and the simulated MT
relation. Simulations here include radiative cooling, star formation and the
328 S. Borgani
1
0.1
kT
500,sl
(keV) kT
500,sl
(keV)
2 4 6 8 10 2 4 6 8 10
M
5
0
0
(
1
0
1
5
h
1
M
)
.
Fig. 19. The masstemperature relation at /
cr
= 500, in simulations (lled circles
and triangles) and for the observational data (squares with errorbars, [62]). The left
panel is for the true masses of simulated clusters; the right panel is for masses of
simulated clusters estimated by adopting the same procedure applied by Finoguenov
et al. to observational data (from [133])
eect of galactic winds powered by supernovae, and, as such, provide a real-
istic description of the relevant physical processes. The observational results,
which are taken from [62], corresponds to mass estimates based on the hy-
drostatic equilibrium for a polytropic model of the gas distribution [(33)].
In both panels, the temperature has been computed by using a proxy to the
actual spectroscopic temperature, the socalled spectroscopiclike tempera-
ture [110]. The left panel shows the results when exact masses of simulated
clusters are used for the comparison. Based on this result only, the conclusion
would be that simulations do indeed produce too high a MT relation, even in
the presence of a realistic description of gas physics. In the left panel, masses
of simulated clusters are computed instead by using the same procedure as
for observed clusters, i.e. by applying (34) for the hydrostatic equilibrium of a
polytropic model. Quite remarkably, the eect of applying the observational
mass estimator has two eects. First, the overall normalization of the MT
relation is decreased by the amount required to attain a reasonable agreement
with observations. Second, the scatter in the simulated MT relation is sub-
stantially suppressed. This is the consequence of the fact that (34) provides
a one-to-one correspondence between mass and temperature, while only the
cluster-by-cluster variations of and account for the intrinsic diversity of
the cluster thermal structure.
This example illustrates how simulations can be usefully employed as
guidelines to study possible biases on observational mass estimates. However,
it is worth reminding that the reliability of simulation results depends on
our capability to correctly provide a numerical description of all the relevant
physical process. In this sense, understanding in detail the (astro)physics of
Clusters and Cosmology 329
clusters is mandatory in order to calibrate them as tools in the era of precision
cosmology.
As a concluding remark, we emphasize once more that a number of inde-
pendent analyses of the cluster mass function, which have been realized so far,
favor a relatively low normalization of the power spectrum, with
8
0.70.8
for
m
0.3, thus in agreement with the most recent WMAP results [149].
This agreement must be considered as a success for cluster cosmology and
a strong encourgement for future applications to large cluster surveys of the
next generation.
Acknowledgments
I would like to thank the organizers of the GH2005 School, David Hughes,
Omar L opezCruz and Manolis Plionis for having provided an enjoyable
and stimulating environment. I also warmly thank Manolis Plionis and Piero
Rosati for a careful reading of the manuscript and useful suggestions to im-
prove it.
References
1. Abell, G.O.: ApJS 3, 211 (1958)
2. Abell, G.O., Corwin, H.G., Olowin, R.P.: ApJS 70, 1 (1989)
3. Allen, S.W., Fabian, A.C.: MNRAS 297, L57 (1998)
4. Allen, S.W., Schmidt, R.W., Ebeling, H., Fabian, A.C., van Speybroeck,
L.:MNRAS 353, 457 (2004)
5. Allen, S.W., Schmidt, R.W., Fabian, A.C.: MNRAS 328, L37 (2001)
6. Allen, S.W., Schmidt, R.W., Fabian, A.C.: MNRAS 334, L11 (2002)
7. Arnaud, M., Evrard, A.E.: MNRAS 305, 631 (1999)
8. Arnaud, M., Pointecouteau, E., Pratt, G.W.: A&A 441, 893 (2005)
9. Bahcall, N.A., Cen, R., Dave, R., Ostriker, J.P., Yu, Q.: ApJ 541, 1 (2000)
10. Bahcall, N.A., Dong, F., Bode, P., Kim, R., Annis, J., McKay, T.A., Hansen, S.,
Schroeder, J., Gunn, J., Ostriker, J.P., Postman, M., Nichol, R.C., Miller, C.,
Goto, T., Brinkmann, J., Knapp, G.R., Lamb, D.O., Schneider, D.P., Vogeley,
M.S., York, D.G.: ApJ 585, 182 (2003)
11. Bahcall, N.A., Fan, X., Cen, R.: ApJ 485, L53+ (1997)
12. Bartelmann, M., Steinmetz, M.: MNRAS 283, 431 (1996)
13. Basilakos, S., Plionis, M., Georgakakis, A., Georgantopoulos, I., Gaga, T.,
Kolokotronis, V., Stewart, G.: MNRAS 351, 989 (2004)
14. Binney, J., Tremaine, S.: Galactic Dynamics, p. 747. Princeton University
Press, Princeton, NJ (1987)
15. Birkinshaw, M.: Phys. Rep. 310, 97 (1999)
16. Biviano, A., Girardi, M.: ApJ 585, 205 (2003)
17. Biviano, A., Girardi, M., Giuricin, G., Mardirossian, F., Mezzetti, M.: ApJ
411, L13 (1993)
330 S. Borgani
18. Biviano, A., Murante, G., Borgani, S., Diaferio, A., Dolag, K., Girardi, M.:
ArXiv Astrophysics e-prints (2006)
19. Blanchard, A., Sadat, R., Bartlett, J.G., Le Dour, M.: A&A 362, 809 (2000)
20. Bohringer, H., Collins, C.A., Guzzo, L., Schuecker, P., Voges, W., Neumann,
D.M., Schindler, S., Chincarini, G., De Grandi, S., Cruddace, R.G., Edge, A.C.,
Reiprich, T.H., Shaver, P.: ApJ 566, 93 (2002)
21. Bohringer, H., Schuecker, P., Guzzo, L., Collins, C.A., Voges, W., Cruddace,
R.G., Ortiz-Gil, A., Chincarini, G., De Grandi, S., Edge, A.C., MacGillivray,
H.T., Neumann, D.M., Schindler, S., Shaver, P.:A&A 425, 367 (2004)
22. Bohringer, H., Voges, W., Huchra, J.P., McLean, B., Giacconi, R., Rosati, P.,
Burg, R., Mader, J., Schuecker, P., Simic D., Komossa, S., Reiprich, T.H.,
Retzla, J.,Tr umper, J.: ApJS 129, 435 (2000)
23. Bond, J.R., Cole, S., Efstathiou, G., Kaiser, N.: ApJ 379, 440 (1991)
24. Borgani, S., Girardi, M., Carlberg, R.G., Yee, H.K.C., Ellingson, E.: ApJ 527,
561 (1999)
25. Borgani, S., Governato, F., Wadsley, J., Menci, N., Tozzi, P., Quinn, T., Stadel,
J., Lake, G.: MNRAS 336, 409 (2002)
26. Borgani, S., Guzzo, L.: Nature 409, 39 (2001)
27. Borgani, S., Murante, G., Springel, V., Diaferio, A., Dolag, K., Moscardini, L.,
Tormen, G., Tornatore, L., Tozzi, P.: MNRAS 348, 1078 (2004)
28. Borgani, S., Rosati, P., Tozzi, P., Norman, C.: ApJ 517, 40 (1999)
29. Borgani, S., Rosati, P., Tozzi, P., Stanford, S.A., Eisenhardt, P.R., Lidman,
C., Holden, B., Della Ceca, R., Norman, C., Squires, G.: ApJ 561, 13 (2001)
30. Boschin, W.: A&A 396, 397 (2002)
31. Bryan, G.L., Norman, M.L.: ApJ 495, 80 (1998)
32. Burke, D.J., Collins, C.A., Sharples, R.M., Romer, A.K., Holden, B.P., Nichol,
R.C.: ApJ 488, L83+ (1997)
33. Burrows, C.J., Burg, R., Giacconi, R.: ApJ 392, 760 (1992)
34. Carlberg, R.G., Morris, S.L., Yee, H.K.C., Ellingson, E.: ApJ 479, L19+ (1997)
35. Carlberg, R.G., Yee, H.K.C., Ellingson, E., Abraham, R., Gravel, P., Morris,
S., Pritchet, C.J.: ApJ 462, 32 (1996)
36. Carlberg, R.G., Yee, H.K.C., Ellingson, E., Morris, S.L., Abraham, R., Gravel,
P., Pritchet, C.J., Smecker-Hane, T., Hartwick, F.D.A., Hesser, J.E., Hutch-
ings, J.B., Oke, J.B.: ApJ 476, L7+ (1997)
37. Carlstrom, J.E., Holder, G.P., Reese, E.D.: ARAA 40, 643 (2002)
38. Castander, F.J., Bower, R.G., Ellis, R.S., Aragon-Salamanca, A., Mason, K.O.,
Hasinger, G., McMahon, R.G., Carrera, F.J., Mittaz, J.P.D., Perez-Fournon,
I., Lehto, H.J.: Nature 377, 39 (1995)
39. Cavaliere, A., Fusco-Femiano, R.: A&A 49, 137 (1976)
40. Cavaliere, A., Gursky, H., Tucker, W.: Nature 231, 437 (1971)
41. Colafrancesco, S., Mazzotta, P., Vittorio, N.: ApJ 488, 566 (1997)
42. Coles, P., Lucchin, F.: Cosmology: The Origin and Evolution of Cosmic Struc-
ture, Second Edition. Cosmology: The Origin and Evolution of Cosmic Struc-
ture, Second Edition, by Peter Coles, Francesco Lucchin, pp. 512. ISBN 0-471-
48909-3. Wiley-VCH, July 2002.
43. Couch, W.J., Ellis, R.S., MacLaren, I., Malin, D.F.: MNRAS 249, 606 (1991)
44. Dalcanton, J.J.: ApJ 466, 92 (1996)
45. Dalton, G.B., Maddox, S.J., Sutherland, W.J., Efstathiou, G.: MNRAS 289,
263 (1997)
Clusters and Cosmology 331
46. Davis, M., Peebles, P.J.E.: ApJ 267, 465 (1983)
47. Diaferio, A., Borgani, S., Moscardini, L., Murante, G., Dolag, K., Springel, V.,
Tormen, G., Tornatore, L., Tozzi, P.: MNRAS 356, 1477 (2005)
48. Donahue, Megan, Scharf, Caleb, A., Mack, Jennifer, Lee, Y. Paul, Postman,
Marc, Rosati, Piero, Dickinson, Mark, Voit, G. Mark, Stocke, John, T.: ApJ
569, 689 (2002)
49. Donahue, M., Mack, J., Scharf, C., Lee, P., Postman, M., Rosati, P., Dickinson,
M., Voit, G.M., Stocke, J.T.: ApJ 552, L93 (2001)
50. Donahue, M., Voit, G.M.: ApJ 523, L137 (1999)
51. Ebeling, H., Edge, A.C., Allen, S.W., Crawford, C.S., Fabian, A.C., Huchra
J.P.: MNRAS 318, 333 (2000)
52. Ebeling, H., Edge, A.C., Henry, J.P.: ApJ 553, 668 (2001)
53. Eisenstein, D.J., Hu, W.: ApJ 511, 5 (1999)
54. Eke, V.R., Cole, S., Frenk, C.S.: MNRAS 282, 263 (1996)
55. Eke, V.R., Cole, S., Frenk, C.S., Patrick Henry, J.: MNRAS 298, 1145 (1998)
56. Ettori, S., De Grandi, S., Molendi, S.: A&A 391, 841 (2002)
57. Ettori, S., Tozzi, P., Borgani, S., Rosati, P.: A&A 417, 13 (2004)
58. Ettori, S., Tozzi, P., Rosati, P.: A&A 398, 879 (2003)
59. Evrard, A.E., MacFarland, T.J., Couchman, H.M.P., Colberg, J.M., Yoshida,
N., White, S.D.M., Jenkins, A., Frenk, C.S., Pearce, F.R., Peacock, J.A.,
Thomas, P.A.: ApJ 573, 7 (2002)
60. Fabian, A.C.: MNRAS 253, 29P (1991)
61. Felten, J.E., Gould, R.J., Stein, W.A., Woolf, N.J.: ApJ 146, 955 (1966)
62. Finoguenov, A., Reiprich, T.H., Bohringer, H.: A&A 368, 749 (2001)
63. Frenk, C.S., Evrard, A.E., White, S.D.M., Summers, F.J.: ApJ 472, 460 (1996)
64. Gardini, A., Rasia, E., Mazzotta, P., Tormen, G., De Grandi, S., Moscardini,
L.: MNRAS 351, 505 (2004)
65. Giacconi, R., Branduardi, G., Briel, U., et al.: ApJ 230, 540 (1979)
66. Giacconi, R., Murray, S., Gursky, H., Kellogg, E., Schreier, E., Tananbaum,
H.: ApJ 178, 281 (1972)
67. Gioia, I.M., Henry, J.P., Maccacaro, T., Morris, S.L., Stocke, J.T., Wolter, A.:
ApJ 356, L35 (1990)
68. Girardi, M., Biviano, A., Giuricin, G., Mardirossian, F., Mezzetti, M.: ApJ
404, 38 (1993)
69. Girardi, M., Borgani, S., Giuricin, G., Mardirossian, F., Mezzetti, M.: ApJ
506, 45 (1998)
70. Girardi, M., Borgani, S., Giuricin, G., Mardirossian, F., Mezzetti, M.: ApJ
530, 62 (2000)
71. Girardi, M., Giuricin, G., Mardirossian, F., Mezzetti, M., Boschin, W.: ApJ
505, 74 (1998)
72. Girardi, M., Manzato, P., Mezzetti, M., Giuricin, G., Limboz, F.: ApJ 569,
720 (2002)
73. Girardi, M., Mezzetti, M.: ApJ 548, 79 (2001)
74. Gladders, M.D., Yee, H.K.C.: ApJS 157, 1 (2005)
75. Gonzalez, A.H., Zaritsky, D., Dalcanton, J.J., Nelson, A.: ApJS 137, 117 (2001)
76. Governato, F., Babul, A., Quinn, T., Tozzi, P., Baugh, C.M., Katz, N., Lake,
G.: MNRAS 307, 949 (1999)
77. Gross, M.A.K., Somerville, R.S., Primack, J.R., Holtzman, J., Klypin, A.: MN-
RAS 301, 81 (1998)
332 S. Borgani
78. Gunn, J.E., Hoessel, J.G., Oke, J.B.: ApJ 306, 30 (1986)
79. Haiman, Z., Allen, S., Bahcall, N., et al.: ArXiv Astrophysics e-prints (2005)
80. Haiman, Z., Mohr, J.J., Holder, G.P.: ApJ 553, 545 (2001)
81. Henry, J.P.: ApJ 534, 565 (2000)
82. Henry, J.P.: ApJ 609, 603 (2004)
83. Henry, J.P., Arnaud, K.A.: ApJ 372, 410 (1991)
84. Henry, J.P., Gioia, I.M., Maccacaro, T., Morris, S.L., Stocke, J.T., Wolter, A.:
ApJ 386, 408 (1992)
85. Henry, J.P., Gioia, I.M., Mullis, C.R., Voges, W., Briel, U.G., Bohringer, H.,
Huchra, J.P.: ApJ 553, L109 (2001)
86. Holden, B.P., Stanford, S.A., Squires, G.K., Rosati, P., Tozzi, P., Eisenhardt,
P., Spinrad, H.: AJ 124, 33 (2002)
87. Hu, W., Kravtsov, A.V.: ApJ 584, 702 (2003)
88. Ikebe, Y., Reiprich, T.H., Bohringer, H., Tanaka, Y., Kitayama, T.: A&A 383,
773 (2002)
89. Jenkins, A., Frenk, C., White, S., Colberg, J., Cole, S., Evrard, A., Couchman,
H., Yoshida, N.: MNRAS 321, 372 (2001)
90. Kaiser, N.: MNRAS 222, 323 (1986)
91. Kay, S.T., Thomas, P.A., Jenkins, A., Pearce, F.R.: MNRAS 355, 1091 (2004)
92. Kirkman, D., Tytler, D., Suzuki, N., OMeara, J.M., Lubin, D.: ApJS 149, 1
(2003)
93. Kitayama, T., Suto, Y.: ApJ 469, 480 (1996)
94. Kitayama, T., Suto, Y.: ApJ 490, 557 (1997)
95. Kotov, O., Vikhlinin, A.: ApJ 633, 781 (2005)
96. Lewis, A.D., Ellingson, E., Morris, S.L., Carlberg, R.G.: ApJ 517, 587 (1999)
97. Lima, M., Hu, W.: Phys. Rev. D 70, 043504 (2004)
98. Lima, M., Hu, W.: Phys. Rev. D 72, 043006 (2005)
99. Limber, D.N., Mathews, W.G.: ApJ 132, 286 (1960)
100. Lin, Y.-T., Mohr, J.J., Stanford, S.A.: ApJ 591, 749 (2003)
101. Lumb, D.H., Bartlett, J.G., Romer, A.K., Blanchard, A., Burke, D.J., Collins,
C.A., Nichol, R.C., Giard, M., Marty, P.B., Nevalainen, J., Sadat, R., Vauclair,
S.C.: A&A 420, 853 (2004)
102. Lumsden, S.L., Nichol, R.C., Collins, C.A., Guzzo, L.: MNRAS, 258, 1 (1992)
103. MacTavish, C., Ade, P.A.R., Bock, J.J., et al.: ArXiv Astrophysics e-prints
(2005)
104. Majumdar, S., Mohr, J.J.: ApJ 585, 603 (2003)
105. Majumdar, S., Mohr, J.J.: ApJ 613, 41 (2004)
106. Markevitch, M.: ApJ 504, 27 (1998)
107. Mathiesen, B., Evrard, A.E.: MNRAS 295, 769 (1998)
108. Mathiesen, B.F., Evrard, A.E.: ApJ 546, 100 (2001)
109. Maughan, B.J., Jones, L.R., Ebeling, H., Scharf, C.: MNRAS 365, 509 (2006)
110. Mazzotta, P., Rasia, E., Moscardini, L., Tormen, G.: MNRAS 354, 10 (2004)
111. Miller, C.J., Nichol, R.C., Reichart, D., et al.: AJ 130, 968 (2005)
112. Monaco, P.: FCPh 19, 157 (1998)
113. Moretti, A., Guzzo, L., Campana, S., Lazzati, D., Panzera, M.R., Tagliaferri,
G., Arena, S., Braglia, F., DellAntonio, I., Longhetti, M.: A&A 428, 21 (2004)
114. Moscardini, L., Matarrese, S., Mo, H.J.: MNRAS 327, 422 (2001)
115. Motl, P.M., Hallman, E.J., Burns, J.O., Norman, M.L.: ApJ 623, L63 (2005)
116. Mulchaey, J.S., Zabludo, A.I.: ApJ 496, 73 (1998)
Clusters and Cosmology 333
117. Mullis, C.R., McNamara, B.R., Quintana, H., Vikhlinin, A., Henry, J.P., Gioia,
I.M., Hornstrup, A., Forman, W., Jones, C.: ApJ 594, 154 (2003)
118. Mullis, C.R., Rosati, P., Lamer, G., Bohringer, H., Schwope, A., Schuecker, P.,
Fassbender, R.: ApJ 623, L85 (2005)
119. Mullis, C.R., Vikhlinin, A., Henry, J.P., Forman, W., Gioia, I.M., Hornstrup,
A., Jones, C., McNamara, B.R., Quintana, H.: ApJ 607, 175 (2004)
120. Nevalainen, J., Markevitch, M., Forman, W.: ApJ 532, 694 (2000)
121. Osmond, J.P.F., Ponman, T.J.: MNRAS 350, 1511 (2004)
122. Oukbir, J., Blanchard, A.: A&A 262, L21 (1992)
123. Peacock, J.A.: Cosmological Physics. Cosmological Physics, by John A. Pea-
cock, pp. 704. ISBN 052141072X. Cambridge, UK: Cambridge University Press,
January 1999.
124. Peebles, P.J.E.: 1993, Principles of physical cosmology. Princeton Series in
Physics, Princeton, NJ: Princeton University Press, c1993
125. Perlman, E.S., Horner, D.J., Jones, L.R., Scharf, C.A., Ebeling, H., Wegner,
G., Malkan, M.: ApJS 140, 265 (2002)
126. Piccinotti, G., Mushotzky, R.F., Boldt, E.A., Holt, S.S., Marshall, F.E., Ser-
lemitsos, P.J., Shafer, R.A.: ApJ 253, 485 (1982)
127. Pierpaoli, E., Borgani, S., Scott, D., White, M.: MNRAS 342, 163 (2003)
128. Pierpaoli, E., Scott, D., White, M.: MNRAS 325, 77 (2001)
129. Plionis, M., Basilakos, S., Georgantopoulos, I., Georgakakis, A.: ApJ 622, L17
(2005)
130. Popesso, P., Biviano, A., Bohringer, H., Romaniello, M., Voges, W.: A&A 433,
431 (2005)
131. Postman, M., Lubin, L.M., Gunn, J.E., Oke, J.B., Hoessel, J.G., Schneider,
D.P., Christensen, J.A.: AJ 111, 615 (1996)
132. Press, W., Schechter, P.: ApJ 187, 425 (1974)
133. Rasia, E., Mazzotta, P., Borgani, S., Moscardini, L., Dolag, K., Tormen, G.,
Diaferio, A., Murante, G.: ApJ 618, L1 (2005)
134. Rasia, E., Tormen, G., Moscardini, L.: MNRAS 351, 237 (2004)
135. Reichart, D.E., Nichol, R.C., Castander, F.J., Burke, D.J., Romer, A.K.,
Holden, B.P., Collins, C.A., Ulmer, M.P.: ApJ 518, 521 (1999)
136. Reiprich, T., Bohringer, H.: ApJ 567, 716 (2002)
137. Rines, K., Geller, M.J., Kurtz, M.J., Diaferio, A.: AJ 126, 2152 (2003)
138. Rosati, P., Borgani, S., Norman, C.: ARAA 40, 539 (2002)
139. Rosati, P., della Ceca, R., Norman, C., Giacconi, R.: ApJ 492, L21+ (1998)
140. Rosati, P., Tozzi, P., Ettori, S., Mainieri, V., Demarco, R., Stanford, S.A.,
Lidman, C., Nonino, M., Borgani, S., Della Ceca, R., Eisenhardt, P., Holden,
B.P., Norman, C.: AJ 127, 230 (2004)
141. Sadat, R., Blanchard, A., Oukbir, J.: A&A 329, 21 (1998)
142. Sahni, V.: LNP Vol. 653: The Physics of the Early Universe 653, 141 (2005)
143. Schuecker, P., Bohringer, H., Guzzo, L., Collins, C.A., Neumann, D.M.,
Schindler, S., Voges, W., De Grandi, S., Chincarini, G., Cruddace, R., Muller,
V., Reiprich, T.H., Retzla, J., Shaver, P.: A&A 368, 86 (2001)
144. Schuecker, P., Guzzo, L., Collins, C.A., Bohringer, H.: MNRAS 335, 807 (2002)
145. Seljak, U.: MNRAS 337, 769 (2002)
146. Sheth, R., Tormen, G.: MNRAS 308, 119 (1999)
147. Sheth, R.K., Tormen, G.: MNRAS 329, 61 (2002)
148. Smith, S.: ApJ 83, 23 (1936)
334 S. Borgani
149. Spergel, D.N., Bean, R., Dore, O., Nolta, M.R., Bennett, C.L., Hinshaw, G.,
Jarosik, N., Komatsu, E., Page, L., Peiris, H.V., Verde, L., Barnes, C., Halpern,
M., Hill, R.S., Kogut, A., Limon, M., Meyer, S.S., Odegard, N., Tucker, G.S.,
Weiland, J.L., Wollack, E., Wright, E.L.: ArXiv Astrophysics e-prints (2006)
150. Springel, V., White, S.D.M., Jenkins, A., Frenk, C.S., Yoshida, N., Gao, L.,
Navarro, J., Thacker, R., Croton, D., Helly, J., Peacock, J.A., Cole, S., Thomas,
P., Couchman, H., Evrard, A., Colberg, J., Pearce, F.: Nature 435, 629 (2005)
151. Stanford, S.A., Holden, B., Rosati, P., Eisenhardt, P.R., Stern, D., Squires, G.,
Spinrad, H.: AJ 123, 619 (2002)
152. Stanford, S.A., et al.: ApJ 634, L129 (2005)
153. Steinhardt, P.J.: RSPTA 361, 2497 (2003)
154. Sunyaev, R.A., Zeldovich, Y.B.: Comments on Astrophysics and Space Physics
4, 173 (1972)
155. Truemper, J.: Science 260, 1769 (1993)
156. van Haarlem, M.P., Frenk, C.S., White, S.D.M.: MNRAS 287, 817 (1997)
157. Vauclair, S.C., Blanchard, A., Sadat, R., Bartlett, J.G., Bernard, J.-P., Boer,
M., Giard, M., Lumb, D.H., Marty, P., Nevalainen, J.: A&A 412, L37 (2003)
158. Viana, P.T.P., Liddle, A.R.: MNRAS 303, 535 (1999)
159. Vikhlinin, A.: ApJ 640, 710 (2006)
160. Vikhlinin, A., Kravtsov, A., Forman, W., Jones, C., Markevitch, M., Murray,
S.S., Van Speybroeck, L.: ApJ 640, 691 (2005)
161. Vikhlinin, A., VanSpeybroeck, L., Markevitch, M., Forman, W.R., Grego, L.:
ApJ 578, L107 (2002)
162. Vikhlinin, A., Voevodkin, A., Mullis, C.R., VanSpeybroeck, L., Quintana, H.,
McNamara, B.R., Gioia, I., Hornstrup, A., Henry, J.P., Forman, W.R., Jones,
C.: ApJ 590, 15 (2003)
163. Voevodkin, A., Vikhlinin, A.: ApJ 601, 610 (2004)
164. Voit, G.M.: Reviews of Modern Physics 77, 207 (2005)
165. Warren, M.S., Abazajian, K., Holz, D.E., Teodoro, L.: ArXiv Astrophysics
e-prints (2005)
166. White, M.: ApJS 143, 241 (2002)
167. White, M., Hernquist, L., Springel, V.: ApJ 579, 16 (2002)
168. White, S.D.M., Navarro, J.F., Evrard, A.E., Frenk, C.S.: Nature 366, 429
(1993)
169. Willis, J.P., Pacaud, F., Valtchanov, I., Pierre, M., Ponman, T., Read, A.,
Andreon, S., Altieri, B., Quintana, H., Dos Santos, S., Birkinshaw, M., Bremer,
M., Duc, P.-A., Galaz, G., Gosset, E., Jones, L., Surdej, J.: MNRAS 363, 675
(2005)
170. Xue, Y.-J., Wu, X.-P.: ApJ 538, 65 (2000)
171. Zwicky, F.: ApJ 86, 217 (1937)