E T. L, D N, C A, K N, A V: Preprint Typeset Using L TEX Style Emulateapj v. 05/12/14

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

T HE A STROPHYSICAL J OURNAL , SUBMITTED

Preprint typeset using LATEX style emulateapj v. 05/12/14

MASS ACCRETION AND ITS EFFECTS ON THE SELF-SIMILARITY OF GAS PROFILES


IN THE OUTSKIRTS OF GALAXY CLUSTERS
E RWIN T. L AU 1,2 , DAISUKE NAGAI 1,2,3 , C AMILLE AVESTRUZ 1,2 , K AYLEA N ELSON 2,3 , AND A LEXEY V IKHLININ 4
1 Department

of Physics, Yale University, New Haven, CT 06520, USA; [email protected]


Center for Astronomy and Astrophysics, Yale University, New Haven, CT 06520, USA
3 Department of Astronomy, Yale University, New Haven, CT 06520, USA
4 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
The Astrophysical Journal, submitted

arXiv:1411.5361v1 [astro-ph.CO] 19 Nov 2014

2 Yale

ABSTRACT
Galaxy clusters exhibit remarkable self-similar behavior which allows us to establish simple scaling relationships between observable quantities and cluster masses, making galaxy clusters useful cosmological probes.
Recent X-ray observations suggest that self-similarity may be broken in the outskirts of galaxy clusters. In
this work, we analyze a mass-limited sample of massive galaxy clusters from the Omega500 cosmological
hydrodynamic simulation to investigate the self-similarity of the diffuse X-ray emitting intracluster medium
(ICM) in the outskirts of galaxy clusters. We find that the self-similarity of the outer ICM profiles is better
preserved if they are normalized with respect to the mean density of the universe, while the inner profiles are
more self-similar when normalized using the critical density. However, the outer ICM profiles as well as the
location of accretion shock around clusters are sensitive to their mass accretion rate, which causes the apparent
breaking of self-similarity in cluster outskirts. We also find that the collisional gas does not follow the distribution of collisionless dark matter perfectly in the infall regions of galaxy clusters, leading to 10% departures
in the gas-to-dark matter density ratio from the cosmic mean value. Our results have a number implications
for interpreting observations of galaxy clusters in X-ray and through the Sunyaev-Zeldovich effect and their
application to cluster cosmology.
Subject headings: cosmology: theory galaxies: clusters: general galaxies: clusters: intracluster medium
1. INTRODUCTION

Galaxy clusters are the most massive gravitationally bound


objects in the universe. Most of the baryons are in the form
of hot X-ray emitting gas and reside in the deep gravitational potential wells of galaxy clusters. The hot gas is detectable in both the X-ray and the microwave, through the
Sunyaev-Zeldovich (SZ) effect. Observations of the intracluster medium (ICM) show remarkable regularity over a
wide range of mass and redshift, making galaxy clusters powerful probes for cosmology (e.g. Allen et al. 2011, for review).
Self-similarity is a generic prediction of gravitational structure formation. It features simple scaling relations between
the observable properties and mass of galaxy clusters (Kaiser
1986; Voit 2005; Kravtsov & Borgani 2012, for a recent review). When scaled by cluster mass, the radial profiles of
thermodynamic properties of the ICM display remarkable resemblance (Vikhlinin et al. 2006; Nagai et al. 2007; Pratt et al.
2009; Arnaud et al. 2010) outside of cluster cores where the
effects of non-gravitational physics are small. When integrated, the self-similar ICM quantities can serve as robust observational proxies for cluster mass, such as its thermal energy
content, YX (Kravtsov et al. 2006), or its X-ray luminosity,
LX (Maughan 2007). The outer regions of clusters are ideal
for robust ICM measurements and inferences of cluster mass
since these measurements at large radii are less susceptible
to complex astrophysical processes that affect measurements
in the cluster core, such as gas cooling, star formation, and
energy injections from supernovae and active galactic nuclei.
Both X-ray and SZ observations have recently measured
properties of the ICM out to the virial radius (see, e.g.,
Reiprich et al. 2013, for review). However, these observations
revealed several unexpected results that deviate from theoret-

ical predictions. First, hydrodynamical simulations predict


that the entropy of the ICM should have a power law scaling
with cluster-centric radius, as the entropy is set by the accretion shock of the cluster (Tozzi & Norman 2001; Voit et al.
2003). But, Suzaku X-ray measurements of the outer regions
of clusters showed lower and flatter entropy profiles than the
theoretically predicted power law, breaking the self-similar
scaling (e.g., George et al. 2009; Bautz et al. 2009; Reiprich
et al. 2009; Hoshino et al. 2010; Kawaharada et al. 2010;
Walker et al. 2013; Urban et al. 2014; Okabe et al. 2014). Additionally, Suzaku measured an enclosed gas mass fraction of
the Perseus cluster that curiously exceeded the cosmic baryon
fraction at large radii (Simionescu et al. 2011). With stacked
X-ray data of clusters selected from SZ surveys, McDonald
et al. (2014) found a redshift evolution of the pressure profile in the outer regions of clusters, also indicating departures
from the theoretical expectation.
Galaxy clusters are dynamically young objects that are still
growing via mergers and accretion. The outskirts of galaxy
clusters can be regarded as cosmic melting pots, where infalling materials are being virialized. The accreting gas dissipates heat through shocks and turbulence, establishing the
overall thermodynamical structures in galaxy clusters. An
improved understanding of the cosmic melting pot will advance our ability to use clusters for precision cosmology, particularly in light of ongoing and upcoming multi-wavelength
cluster surveys, including Planck microwave and eROSITA Xray missions.
A variety of astrophysical processes in cluster outskirts can
contribute to the observed deviations from self-similar predictions. Previous theoretical works indicated that processes
such as gas inhomogeneities (Nagai & Lau 2011; Zhuravleva et al. 2013; Vazza et al. 2013; Roncarelli et al. 2013; Ra-

Lau et al

sia et al. 2014) and non-thermal pressure support (e.g, Nelson


et al. 2014a; Shi & Komatsu 2014; Shi et al. 2014) may play
significant roles. These non-equilibrium processes are likely
to be driven by mass accretion in the outskirts of clusters.
Recent N-body simulations suggested that dark matter
(DM) halos accrete mass at different rates depending on their
mass and redshift (e.g., Cuesta et al. 2008; McBride et al.
2009). Variations in the mass accretion rate (MAR) of halos introduce differences in the DM density profiles in halo
outskirts, leading to apparent deviations from self-similarity
(Diemer & Kravtsov 2014). If gas traces DM, we would expect mass accretion to similarly affect the gas distribution in
cluster outskirts. Understanding the effects of MAR on gas
properties in cluster outskirts requires cosmological hydrodynamic simulations that self-consistently follow the dynamics
of DM and gas.
In this work, we analyze a mass-limited sample of galaxy
clusters extracted from the Omega500 cosmological hydrodynamic simulation (Nelson et al. 2014b). We find that the
ICM profiles in cluster outskirts are affected by the mass accretion in two ways. First, the outer ICM profiles are more
self-similar with redshift when they are normalized using the
mean mass density of the universe, because the mean density
traces the redshift evolution of MAR better than the critical
density. Second, at any given redshift, the outer ICM profiles
and the location of accretion shock around clusters are sensitive to the MAR. We also find the collisional gas does not
trace accretion of the collisionless DM perfectly, leading to
departures in the density ratio between gas and dark matter
from the cosmic mean value. We discuss implications of our
results on X-ray and SZ observations of the ICM profiles in
cluster outskirts, mass-observable scaling relations, and the
use of galaxy clusters as cosmological probes.
The paper is organized as follows. In Section 2 we describe
the concepts of cluster mass definitions, self-similar model,
and MAR. In Section 3 we describe our simulated cluster
sample. In Section 4 we examine the dependence of the gas
profiles on the definition of cluster mass and their MAR. We
provide our discussion and conclusions in Section 5.
2. THEORETICAL CONSIDERATIONS
2.1. Cluster Mass Definitions
Galaxy cluster forms at the intersections of large-scale filamentary structures in the universe and they do not have welldefined physical edge. To calculate cluster masses, the common approach is to define the boundary of a cluster as a sphere
enclosing an average matter density equal to a reference overdensity times a reference background density, ref . The
mass of the cluster is then given as,

4
ref (z)R3
(1)
3
where R is the radius within which we compute the enclosed mass. Two common choices of the background density ref are the critical density, c (z) 3H02 E 2 (z)/(8G),
and the mean matter density, m (z) = c (z)m (z), in the standard CDM spatially flat cosmological model, where E 2 (z)
m (1 + z)3 + , m (z) = m (1 + z)3 /E 2 (z), and m (without
the explicit z-dependence) refers to the present-day mass density fraction of the universe. The reference overdensity is
usually chosen to be a value close to 18 2 178, which corresponds to the virial overdensity in the flat matter dominated,
Einstein-de Sitter universe (m = 1 = 1). In the more reM

alistic flat CDM model, the virial overdensity varies with


redshift (e.g., Bryan & Norman 1998).
In the literature, c (z) has been widely used to define cluster masses (ref = c (z), with = c ). In particular, c = 500
has been used in the analyses of Chandra and XMM-Newton
X-ray observations of galaxy clusters, since it corresponds to
the radius out to which gas density and temperature profiles of
the ICM can be reliably measured. c = 200 is also adopted
in recent Suzaku X-ray and Planck SZ observations which extended the measurements of the ICM profiles to larger clustercentric radii. On the other hand, m (z) (1 + z)3 is independent of other cosmological parameters, e.g., the Hubble
parameter. Using m (z) to define DM halos also leads to a
more universal mass function (e.g., White 2002) and has
been used in calibrating mass functions in N-body simulations
(e.g., Jenkins et al. 2001; Tinker et al. 2008).
2.2. Self-Similarity
In the current hierarchical structure formation model,
galaxy cluster of mass M at redshift z forms from gravitational
collapse of the primordial cosmological density perturbation,
when its linear density fluctuation (M, z) reaches the collapse
threshold c = 1.686. Since the primordial density perturbations are well-characterized by the Gaussian distribution,
properties of galaxy clusters are uniquely characterized by
its density peak height, (M, z) c /(M, z), where (M, z)
is the characteristic linear density fluctuation smoothed over
mass scale M at redshift z.
Strictly speaking, self-similarity only holds when the initial
linear density fluctuations and their subsequent gravitational
collapse into cluster halos are scale-free, and there is no physical scale associated with non-gravitational processes operating during cluster formation. These conditions are approximately true for cluster-size halos, where linear density fluctuations follow a power law. During the subsequent collapse
until z & 0.5, the effects of dark energy are also assumed to
be small, keeping cluster growth almost scale-free. Baryonic
physics, such as radiative cooling, star formation and feedback, break self-similarity, but their effects are mostly confined to within the cluster core.
On large scales, the majority of the baryonic component
is in the form of X-ray emitting ICM and is expected to
follow the distribution of the gravitationally dominant DM.
The self-similar model predicts that cluster gas profiles for
a given mass (or peak height) appear universal when they
are scaled with respect to the reference background density
of the universe (see, e.g., Voit 2005). For example, the gas
density is scaled using the mean cosmic baryon overdensity,
defined as gas, fb ref (z), where is the redshift independent chosen overdensity, ref (z) is the reference mass
density of the universe at redshift z, and fb b /m is the
cosmic baryon fraction. Similarly, other quantities, such as
temperature, pressure, entropy, and velocity, can be normalized with appropriate scaling that depends on mass and redshift: T GM m p /(kB 2R ), P gas, kB T /(m p ),
p
2/3
K kB T /(m p gas, ), and Vcirc, GM /R . We
adopt fb = 0.1737 from the WMAP5 cosmology used in this
work, = 0.59 is the mean molecular weight of the fully ionized ICM, m p is the proton mass and kB is the Boltzmann
constant. For the cases where the reference background density is set to the critical or the mean density, we set = c or
= m c /m (z), respectively. In this paper, we consider
two cases: c = 200 and m = 200.

Gas Profiles in Cluster Outskirts

F IG . 1. Projected gas density maps of one of the z = 0 clusters selected from the sample. From left to right: map of the total gas, clumps, and filaments. The
dimension for each panel is 15.6 h1 Mpc 15.6 h1 Mpc, with depth of 1.9 h1 Mpc. The circle in dashed line shows R200m = 3.3 h1 Mpc of the cluster.

3. COSMOLOGICAL HYDRODYNAMIC SIMULATIONS

h = 0.7 and 8 = 0.82, where the Hubble constant is defined as


100 h km s1 Mpc1 and 8 is the mass variance within spheres
of radius 8 h1 Mpc. The simulation is performed on a uniform
5123 grid with 8 levels of mesh refinement, implying a maximum comoving spatial resolution of 3.8 h1 kpc. Our simulations are based on simple non-radiative hydrodynamics, allowing us to isolate the effects of MAR on self-similarity from
the effects of complicated galaxy formation physics, which
are expected to be small in the outskirts of clusters. The current work also serves as a baseline for future studies of the
effects of such physics.
Galaxy clusters are identified in the simulation using a
spherical overdensity halo finder described in Nelson et al.
(2014b). We select clusters with M500c 3 1014 h1 M at
z = 0 and re-simulate the box with higher resolution DM particles in regions of the selected clusters, resulting in an effective
mass resolution of 20483 , which corresponds to a mass resolution of 1.09 109 h1 M . To study the redshift evolution
for the ICM profiles, we extract halos from four redshift outputs: z = {0.0, 0.5, 1.0, 1.5}. At each redshift we apply a mass
cut to ensure a mass-limited sample by comparing our mass
function to that of Tinker et al. (2008) and setting the mass cut
to ensure that the sample is complete above the chosen mass
threshold. The mass-cuts and resulting sample sizes are as
follows: 65 clusters with M200m 6 1014 h1 M at z = 0, 48
clusters with M200m 2.5 1014 h1 M at z = 0.5, 42 clusters
with M200m 1.3 1014 h1 M at z = 1.0, and 42 clusters with
M200m 7 1013 h1 M at z = 1.5.

3.1. Data and Halo Selection

3.2. Substructure Removal and Profile Making

In this work we analyze a sample of simulated galaxy clusters from the Omega500 Simulation Project (Nelson et al.
2014b), which is a high-resolution hydrodynamical simulation of a large cosmological volume with the comoving box
length of 500 h1 Mpc. The simulation is performed using
the Adaptive Refinement Tree (ART) N-body+gas-dynamics
code (Kravtsov 1999; Kravtsov et al. 2002; Rudd et al. 2008),
which is an Eulerian code that uses adaptive refinement in
space and time, and non-adaptive refinement in mass (Klypin
et al. 2001) to achieve the dynamic ranges to resolve the
cores of halos formed in self-consistent cosmological simulations in a flat CDM model with WMAP five-year (WMAP5)
cosmological parameters: m = 1 = 0.27, b = 0.0469,

In this work we are primarily interested in the thermodynamic properties of the diffuse ICM. Gas in dense clumps and
filaments is expected to have different thermodynamical properties from the diffuse ICM. In this work we minimize their
effects by identifying and removing these gaseous substructures directly in simulations. Using the method proposed by
Zhuravleva et al. (2013), we remove gas clumps by excluding
gas cells that have logarithmic density 3.5 above the median
for a given radial bin. Similarly we remove gaseous filamentary structures by excluding radially infalling gas cells with
logarithmic density between 1 and 3.5 above the median
for a given radial bin. As shown in Figure 1, this method can
identify gas associated with clumps and filaments in the clus-

2.3. Mass Accretion Rate


In this work, we propose a proxy for the instantaneous
MAR , defined as the ratio of radial velocity to the circular
velocity measured at some chosen radius R ,

VrDM (r = R )
,
Vcirc,

(2)

where R is a free parameter denoting a certain radius located


inside the infall region where VrDM < 0. Physically, a halo that
is actively accreting in its outskirts will have a large negative
value of .
This MAR proxy has the advantage over :

log10 M (a0 )/M (a1 )
=
(3)
log10 (a0 /a1 )
introduced by Diemer & Kravtsov (2014), because can be
measured for halos at any given redshift, while is an integrated quantity defined between z = 0 and z = 0.5. Note that a
higher means that the halo has experienced a greater physical mass accretion between z = 0 and z = 0.5, where M (a0 )
and M (a1 ) are the mass of the halo at a0 = 1 (z = 0) and its
progenitor at a1 = 0.67 (z = 0.5) respectively. Radial velocity is
also in principle an observable quantity that can be traced by
infalling galaxies or gas in cluster outskirts, while is not a
measurable quantity. We compare these two MAR proxies in
Section 4.3.

Lau et al
0.4

0.2

0.2

Vr /Vcirc,200m

0.4

0.0

0.2

0.4

z = 0.0
z = 0.5
z = 1.0
z = 1.5
gas
DM

0.1

K(r)/K200c

10

0.2

0.4
1
r/R200c

z
z
z
z

= 0.0
= 0.5
= 1.0
= 1.5

0.1

1
r/R200m

10

0.1
0.4
0.2
0.0
0.2
0.4
0.6
0.1

0.0

K(r)/K200m

Vr /Vcirc,200c

0.1

1
r/R200c

0.4
0.2
0.0
0.2
0.4
0.6

0.1

1
r/R200m

F IG . 2. Profiles of average radial velocity (top panels) and gas entropy (bottom panels) at z = 0, 0.5, 1.0, and 1.5. The lower subplot in each panel shows the
deviations of the profiles from different z outputs relative to that of the z = 0.5 clusters. The left panels show the profiles of cluster halos defined using the critical
density, while the right panels show the profiles of the same cluster halos defined using the mean density. In the upper panel, we show the mean radial velocity
profiles for gas and DM in solid and dashed lines respectively. The shaded regions indicate the 1 scatter around the mean gas profiles for the z = 0 clusters.

ter outskirts quite well.


After removing clumps and filaments, we compute the
spherically averaged profiles by dividing the analysis region
for each cluster into 99 spherical logarithmically spaced bins
from 10 h1 kpc to 10 h1 Mpc (comoving) in the radial direction from the cluster center, which is defined as the position
with the maximum binding energy. Our results are insensitive
to the exact choice of binning. We then compute volumeweighted mean profiles of density, pressure, and entropy, and
mass-weighted mean profiles of temperature and velocities
for the gas in each cluster halo, and present the mean profiles averaged the over our cluster sample. For the rest of the
paper, all of the gas profiles are presented with substructures
and filaments removed, unless noted otherwise.

4. RESULTS
4.1. Self-similarity of Gas Profiles

We begin by comparing radial velocity profiles of gas and


DM in clusters defined with respect to the critical density and
the mean density of the universe. The velocity
p profiles are
normalized by the circular velocity Vcirc, GM /R . In
Figure 2 we show the evolution of radial velocities for gas between z = 1.5 and z = 0.0. The top left panel shows the profiles
for clusters normalized by using c = 200, and the top right
panel shows the profiles of the same clusters normalized by
using m = 200. For c = 200, there is significant redshift
evolution in the radial gas velocity profile, where Vr (r = R200c )
varies from 0.3V200c to 0 from z = 1.5 to z = 0. Outside

Gas Profiles in Cluster Outskirts

0.01
0.3
0.2
0.1
0.0
0.1
0.2
0.3
0.1

= 0.0
= 0.5
= 1.0
= 1.5

103
0.4
0.2
0.0
0.2
0.4
0.1

1
r/R200c

gas(r)/200m (r/R200m )2

T (r)/T200m

0.1

0.1

0.01

0.1

0.01

103
0.6
0.4
0.2
0.0
0.2
0.4

1
r/R200c

0.1

1
r/R200c

0.1
P (r)/P200m (r/R200m )3

z
z
z
z

gas(r)/200c (r/R200c )2

T (r)/T200c

0.1

P (r)/P200c (r/R200c )3

0.1

0.01
0.3
0.2
0.1
0.0
0.1
0.2
0.3

0.01

103
0.4
0.2
0.0
0.2
0.4

0.1

r/R200m

1
r/R200m

0.1

0.01

103
0.6
0.4
0.2
0.0
0.2
0.4

0.1

1
r/R200m

F IG . 3. From left to right, we show the profiles of gas temperature, density, and pressure for the cluster halos at z = {0.0, 0.5, 1.0, 1.5}. The upper panels show
the profiles normalized using the critical density 200 c (z), while the lower panels show the same profiles normalized using the mean density 200 m (z). In
each figure, the bottom sub-panel shows the fractional deviations of the profiles with respect to the z = 0.5 clusters. The shaded regions indicate the 1 scatter
around the mean for the z = 0 clusters.

R200c , the radial velocities of DM and gas are mostly negative, indicating infall of DM and gas onto the cluster halos.
The magnitude of the velocities in these infall regions evolves
with redshift: higher-z clusters show more negative radial velocities, indicating their higher MAR. The location of the velocity minimum, where the gas is infalling most strongly, also
evolves with z, and it is located at larger fraction of R200c at
lower z. In the same figures, we also show the average radial
velocities of DM (indicated by the dashed lines). Although
the radial velocity profiles of both gas and DM show similar
qualitative trend with z, gas is generally accreting at a slower
rate than DM; the collisional gas1 experiences ram pressure
from the surrounding ICM, while the collisionless DM does
not.
At large radii r > 3R200c , the gas traces DM, where both
velocities follow the Hubble flow in the expanding universe.
Their normalized radial velocity profiles are independent of
redshift, because the normalization R200c and Vcirc,200c natu1/3
rally account for the Hubble flow: R200c c H 2/3 , and
1/3
Vcirc,200c H , so Vr /r = H Vcirc,200c /R200c , leading to red1 Note that the mean free path of electrons in the ICM plasma could be
large (& 100kpc), which may break down the hydrodynamic approximation
in cluster outskirts. However, the presence of magnetic field may reduce the
effective mean free path below the numerical resolution such that the ICM
can be treated effectively as collisional. This uncertainty must be kept in
mind when interpreting our results.

shift independent Vr /Vcirc,200c r/R200c .


An interesting radius is the turnaround radius Rta , where the
radial velocity becomes zero as the accreting mass detaches
from the Hubble flow. The top left panel of Figure 2 shows
that Rta normalized by R200c is independent of redshift, and
it is located at Rta /R200c 5. We expect Rta to follow the
evolution of the Hubble parameter, as it is the radius where
the dynamics of both dark matter and gas detach from the
Hubble flow. Therefore Rta is well-traced by R200c which is
defined in terms of the Hubble parameter.
On the other hand, Rta does not scale well with R200m . The
radial velocity profiles normalized with Vcirc,200m also show
strong evolution with z, especially at late times (z . 1.0). This
is because R200m m (z)1/3 only accounts for the evolution
of the matter component, but does not account for the effects
of dark energy which drive the accelerated expansion of the
universe at low z.
However, at smaller radii, we find that choosing m = 200
makes the radial velocity profile more universal with z. The
locations of radial velocity minima in both DM and gas for
m = 200 show little redshift evolution compared to the case
of c = 200. This behavior is expected, as the accreting matter undergoes free-fall once decoupled from the Hubble flow,
and its evolution is governed primarily by the gravitational
dynamics governed by the cluster mass, whose mass density
is characterized by the cosmic mean mass density.

Lau et al
1.2

1.2

1.1

1.1

1.0
0.9
0.8
0.7
0.6
0.5

0.15
0.10
0.05
0.00
0.05
0.10
0.15
0.1

z
z
z
z

= 0.0
= 0.5
= 1.0
= 1.5

1
r/R200c

Vcirc (r)/Vcirc,200m

Vcirc (r)/Vcirc,200c

1.0
0.9
0.8
0.7
0.6
0.5

0.15
0.10
0.05
0.00
0.05
0.10
0.15

0.1

1
r/R200m

p
F IG . 4. Profiles of circular velocity Vcirc GM(< r)/r at z = {0.0, 0.5, 1.0, 1.5}. The left panel shows the profiles of the cluster halos defined using the
critical density, while the right panel shows the same profiles with cluster halos defined using the mean density. In each figure, the bottom sub-panel shows the
fractional deviations of the profiles with respect to the profile of the z = 0.5 clusters. The shaded regions indicate the 1 scatter around the mean for the z = 0
clusters.

Similar trends are observed in the entropy profiles of gas.


The profiles are normalized by the self-similar values described in the Section 2.2. The bottom panels of Figure 2
show the redshift evolution of the entropy profiles. Each line
represents the entropy profile averaged over the cluster sample
at z = 0, 0.5, 1.0, and 1.5. Each profile shows a well-defined
entropy peak, which corresponds to the location of the accretion shock2 , Rsh , where the gas radial velocity is minimum.
Some care is needed when interpreting the accretion shock
radius Rsh defined using the entropy peak or the minimum
of the radial infall velocity of gas. When the accreting collisional gas is shocked, it stops infalling and its radial velocity should jumps abruptly to zero. Likewise, the gas entropy profile should jump sharply behind the accretion shock
when the gas is heated through the shock. However, these
sharp jumps are not seen in our radial velocity nor entropy
profiles, which increase more smoothly in the post-shock region. This is partly because gas accretion is aspherical, and
the actual topology of the accretion shocks is rather complex
(e.g., Ryu et al. 2003; Pfrommer et al. 2006; Skillman et al.
2008; Vazza et al. 2009a; Planelles & Quilis 2013; Schaal &
Springel 2014). Spherically averaging the velocity and entropy profiles will smooth out these jumps. Moreover, cold
gas accreting along filaments can penetrate deeper into the
cluster and are shocked at smaller cluster-centric radii (e.g.,
Molnar et al. 2009). This creates a series of shocks with varying velocity and entropy jumps in the clusters infall region
and smoothes out the profiles further.
For c = 200, the accretion shock systematically increases
toward larger cluster-centric radii at lower z. The entropy
measured at r = R200c decreases from 3 K200c at z = 1.5 to
1.5 K200c at z = 0 due to the combination of the shift in the
2 We define accretion shocks as regions where pristine gas from voids falls
into the cluster and gets shock-heated for the first time, which are commonly
referred to as external shocks, in contrast to internal shocks driven by
mergers or accretions through filaments (e.g., Ryu et al. 2003).

location of the entropy peak and the evolution in the entropy


normalization K200c . On the other hand, for m = 200, the
entropy profiles in the radial range of 0.3 r/R200m 1.0
remain roughly constant with z. The accretion shock radius
Rsh /R200m does not evolve much with z, and it is located at
Rsh 1.6R200m at all z.
In Figure 3, we show profiles of other thermodynamic quantities: gas density, temperature and thermal pressure profiles
normalized using c = 200 and m = 200 at different z. We
find that these ICM profiles exhibit similar z-dependence to
that of the entropy profile. For c = 200, there are systematic deviations of the profiles with z outside the cluster core
(r 0.2R200c ). For example at r = R200c , high-z clusters have
lower normalized temperature because they have higher physical mass accretion rate, so that their accretion shocks can
penetrate deeper by pushing the low-temperature pre-shock
regions toward the inner regions of clusters. Similarly, the
evolution in the normalized density and pressure profiles is
also due to the evolution in Rsh /R200c . On the other hand,
switching the halo definition to m = 200 captures the redshift evolution of the Rsh much better and results in thermodynamic profiles that are more universal with z in the radial
range of 0.3 r/R200m 1.0.
The self-similar secondary infall model predicts that the location of shock is located at a fixed fraction of the current
turnaround radius, with Rsh = 0.347Rta , in the Einstein-de Sitter universe (Bertschinger 1985). However, our simulation
shows that Rsh is not proportional to Rta . Rsh evolves as R200m ,
whereas Rta evolves as R200c . This deviation from the selfsimilar secondary infall model is likely due to the increasing
effects of dark energy at low z, which breaks the self-similar
evolution of infalling matter that are currently turning around.
We will investigate the origin of this deviation in a future paper.
We find little dependence of the gas profiles on cluster mass
or density peak height c /(M, z) defined in Section 2.2,

Gas Profiles in Cluster Outskirts

4.2. Evolution of the Circular Velocity Profile

The differences in the self-similarity between the inner and


outer regions can be understood in terms of the evolution of
the gravitational potential well, which determines the thermodynamical properties of the cluster gas. In Figure 4, we plot
the circular velocity profile Vcirc (r) which we use as a proxy
for the gravitational potential, normalized using c = 200 and
m = 200 at z = {0.0, 0.5, 1.0, 1.5}. For c = 200, the circular velocity profile is more universal with z at r R200c ,
indicating that the evolution of the cluster potential in the inner region is well captured by c (z). On the other hand for
m = 200, Vcirc evolves significantly at r R200m , but exhibits
an enhanced level of self-similarity at r R200m .
The reason behind the dependence is that the gravitational
potential well of the cluster is already set during its early stage
of formation (Li et al. 2007; van den Bosch et al. 2014), while
the outer region is more sensitive to the recent mass growth
of the halo. N-body simulations have shown that halo grows
in two phases: an early fast growth phase when the halo is
formed via violent relaxation and phase mixing, followed by
slow growth phase via smooth accretion (Wechsler et al. 2002;
Zhao et al. 2003). The initial fast growth phase determines
the inner density and hence the potential well of the halo.
For a massive cluster-size halo at z = 0, its fast growth phase
still occurs at high redshift (z & 1) when the universe is still
Einstein-de Sitter (i.e., flat, m = 1 universe), with its gravitational potential forming and scaling with the background density where m (z) = c (z). During the subsequent slow growth
phase in the epoch of dark energy domination (m (z) < c (z)),
accretion only adds mass in the halo outskirts, leaving the
inner mass distribution unchanged. For the constant mass,
the radius defined with respect to the critical density scales as
Rc c (z)1/3 E(z)2/3 , which evolves much slower than
Rm (1 + z)1 at late times when the universe is no longer
matter-dominated. Therefore, Rc tracks the slowly changing
interior better, while Rm tracks the outer gas profiles determined by the mass accretion at late times. Further works are
needed to understand the tight self-similar scaling of the inner
profiles with c (z). However, we note that in real clusters the
inner profiles are modified by baryonic physics which breaks
the self-similarity (e.g., McDonald et al. 2014).

0.6
R = 1.25 R200m
R = 1.0 R200m
R = 1.5 R200m

0.4

200m

0.2
0.0
0.2
0.4
0.6
0.8

200m
F IG . 5. Comparison between two MAR proxies 200m versus 200m for
the z = 0 clusters. Here, we consider three values of 200m based on the DM
radial velocity at r = 1.25R200m (black circles), 1.0R200m (red squares), and
1.5R200m (blue triangles), normalized by the circular velocity at r = R200m .

0.7 200m 0.4 (22)


0.4 200m 0.3 (21)

0.4

0.3 200m (22)


gas
DM

0.2
Vr /Vcirc,200m

because our sample contains only massive clusters. On average, halos with higher mass or peak height are on average
accreting more rapidly compared to those with low mass or
peak height (Cuesta et al. 2008; McBride et al. 2009; Diemer
& Kravtsov 2014), which can introduce mass or peak height
dependence in the outskirt gas profiles.
Baryonic processes such as radiative cooling, star formation, and energy feedback from supernovae or active galactic nuclei can significantly influence the gas profiles in the
inner regions and break self-similar behavior. These physical processes can also change the thermodynamical structure
at larger radii and influence gas dynamics in low mass halos, such as galaxies and galaxy groups (e.g., Faucher-Gigure
et al. 2011; van de Voort et al. 2011). We expect the effects of
baryonic physics to be considerably smaller in the outskirts of
massive halos, where gravitational physics dominates the gas
dynamics. However, further work is necessary to fully quantify how baryons affect the MAR and the self-similarity of gas
profiles in all halos. A study of gas flows in group and galaxysize halos will help address these issues (Wetzel & Nagai, in
prep.).

0.0
0.2
0.4
0.6

0.1

1
r/R200m

F IG . 6. Profiles of radial velocity of DM (dashed lines) and gas (solid


lines) in different bins of MAR 200m for the z = 0 clusters. The shaded
regions represent 1 scatter in the profiles with the most negative 200m .
The scatter for the other 200m bins are similar in size. The numbers in the
parentheses indicate the number of clusters in each 200m bin.

4.3. Effects of Mass Accretion Rate on Gas Profiles


Next we study the dependence of the gas profiles on the
MAR in the present-day universe at z = 0. We use the MAR
proxy defined in Equation (2) as a probe the mass accretion
rate of the cluster. We choose m = 200 to define since it
scales out the redshift dependence of the mass accretion rate,
as shown in Section 4.1. We first compare our new MAR

Lau et al
0.7 200m 0.4 (22)

10

0.4 200m 0.3 (21)

0.1

P/P200m (r/R200m )3

T /T200m

K/K200m

0.3 200m (22)

0.1

0.01

103
0.1
0.1

0.01

0.1

r/R200m

0.1

r/R200m

r/R200m

F IG . 7. Profiles of gas entropy (left), temperature (middle), and pressure (right) scaled with R200m as a function of MAR 200m for the z = 0 clusters. We have
multiplied the pressure profile by the radius cubed to show its dependence on 200m more clearly. The shaded regions represent 1 scatter in the profiles with the
most negative 200m . The scatter for the other 200m bins are similar in size. The numbers in the parentheses indicate the number of clusters in each 200m bin.

0
0.1

gas = d log gas /d log r

gas(r)/200m (r/R200m )

0.01

0.7 200m 0.4 (22)

0.4 200m 0.3 (21)

103

3
4
5

0.3 200m (22)

0.1

1
r/R200m

0.1

1
r/R200m

F IG . 8. Profiles of gas density (left) and its logarithmic slope (right) as a function of the cluster-centric radius r/R200m for different MAR 200m at z = 0. The
pressure profiles are multiplied by the square of the normalized radius to demonstrate its dependence on 200m more clearly. The shaded regions represent 1
scatter in the profile with the most negative 200m . The scatter for the other 200m bins are similar in size. The numbers in the parentheses indicate the number
of clusters in each 200m bin.

proxy 200m , based on DM infall velocities defined in Equation (2), with 200m given by Equation (3). In Figure 5, we
plot 200m as a function of 200m , with R /R200m = 1.0, 1.25,
and 1.5. There is an anti-correlation between 200m and 200m
for the above values of R /R200m , suggesting that the instantaneous MAR proxy 200m is a good alternative probe of 200m .
In particular, we find that setting R = 1.25R200m minimizes
the scatter in 200m 200m , elucidating the clearest dependence of the thermodynamic profiles on the MAR.
Figure 6 shows the radial velocity profiles of gas and DM
for clusters in three bins of 200m , with each bin containing
clusters that lie in the top, middle, and bottom third of the
distribution in 200m . Rapidly accreting clusters (with more
negative 200m ) show more negative gas radial velocity in the
infall region (0.6 . r/R200m . 3) where Vr < 0. The radial gas

velocity becomes progressively less negative, with decreasing size in the infall regions for systems with a less negative
200m . Note that the radial velocity of gas in the infall region is
generally less negative than that of DM because the collisional
gas experiences shocks and ram-pressure of the surrounding
ICM. Note, although we treat the DM and gas flows as spherical mass shells moving radially, in reality they are likely to
occur anisotropically through mergers and accretions along
filaments (Zinger et al., in prep).
In Figure 7 we plot the profiles of entropy, temperature, and
pressure for the three different bins of 200m . The left panel
of the figure shows the entropy profile. The peak of the entropy profile indicates the accretion shock radius Rsh shifts
towards the inner regions of clusters with higher MAR. As
more rapidly accreting halos accumulate gas with higher mo-

1.2

1.2

1.1

1.1

1.0
0.9
0.8
z
z
z
z

0.7
0.6
0.1

= 0.0
= 0.5
= 1.0
= 1.5

1
r/R200m

gas(r)/DM (r) [b /DM]

gas(r)/DM (r) [b /DM]

Gas Profiles in Cluster Outskirts

1.0
0.9
0.8
0.7

0.7 200m 0.4 (22)

0.4 200m 0.3 (21)

0.3 200m (22)

0.6
0.1

1
r/R200m

F IG . 9. The ratio of mean gas density to mean DM density profile normalized by the cosmic baryon to DM density for different redshifts z = 0, 0.5, 1.0, and
1.5 (left panel) and for different MAR 200m at z = 0 (right panel). The shaded regions represent 1 scatter in the profile for clusters at z = 0 (left panel) and the
profile with the most negative 200m (right panel). The scatter for the other z and 200m bins are similar in size.

mentum flux, the pre-shocked gas penetrates deeper into the


interior region (e.g., Voit et al. 2003; McCourt et al. 2013)
and shifting the entropy peak inward. The height of the entropy peak also decreases from 10 K200m to 8 K200m from
the least to most rapidly accreting clusters. The small drop
in the entropy for the highly accreting cluster is likely due to
the deeper penetration of the accretion shock into the hotter
ICM leading to smaller Mach number thus smaller entropy
increase. Since the entropy profile inside the accretion shock
is monotonically increasing, the decrease in both the accretion shock radius and entropy peak for higher MAR leads to
relatively small increase in the outer entropy profiles: e.g.,
the entropy at R200m differs by 20% between the least and the
most rapidly accreting systems. Here we do not see the level
of entropy flattening seen in observations (Simionescu et al.
2011; Walker et al. 2013) due to changes in MAR, which is
suggested as a possible explanation for the flattening (e.g.,
Cavaliere et al. 2011).
Other processes may be responsible for the observed flattening of the entropy profiles in the outskirts of relaxed
clusters. For example, the radially-dependent ICM inhomogeneities can lead to overestimates in the gas density profiles
derived from X-ray observations, causing the observed entropy profile to flatten at the large cluster-centric radii (Nagai & Lau 2011). Gas motions induced by gas accretion
and mergers (e.g., Vazza et al. 2009b; Nelson et al. 2012) as
well as plasma effects (e.g., magnetothermal instability, Parrish et al. 2012) can also bias the gas temperature and entropy
profiles low at large cluster-centric radii by keeping some of
the energy of the infalling gas non-thermal.
The average gas temperature in more rapidly accreting clusters is lower than the value of the whole cluster sample, especially in their outskirts where gas is actively accreting, but
the impact of MAR on the temperature profiles is smaller
within R200m than the entropy profile. At r = R200m , the most
rapidly accreting clusters have temperature that is about 9%
lower than the least accreting systems. The differences be-

come larger at r > R200m , reaching 15% at r = 1.5R200m , this is


because accretion shocks of high MAR clusters are located at
smaller radii where the gas is still pre-shocked and has lower
temperature.
Note that our results on the temperature profile and its
dependence on MAR are qualitatively different from results
based on idealized simulations by McCourt et al. (2013), who
reported that clusters with higher accretion rates have higher
temperatures than slowly accreting systems. The discrepancy
could be due to their assumption of instantaneous thermalization of accreting gas at the accretion shocks, which could
cause the gas temperature to be overestimated. Taking into
account the residual kinetic energy from incomplete thermalization in the form of non-thermal pressure could account
for this problem. High MAR clusters are expected to have
a higher non-thermal pressure fraction due to the increased
level of merger- and accretion-induced gas motions (Nelson
et al. 2014a; Shi & Komatsu 2014; Shi et al. 2014). Therefore, the over-predicted gas temperature in the McCourt et al.
model can be lowered by the correspondingly larger amount
of non-thermal pressure present in the high MAR clusters,
which could bring their results in better agreement to the results of cosmological simulations.
In the right panel of Figure 7 we show the thermal pressure
profile and its dependence on MAR. We multiply the pressure
profile by the radius cubed to show more clearly the dependence on 200m in the outskirts. Rapidly accreting clusters
exhibit lower thermal pressure than slowly accreting systems
by 40%, as recently accreted gas is less thermalized and
less dense (see Figure 8).
The left panel of Figure 8 shows the effect of the MAR
200m on gas density. We have multiplied the density profile by the radius squared to show the dependence on 200m
more clearly. Similar to the pressure profile, there is a
factor of 2 difference in the gas density between the least
and most rapidly accreting clusters at R200m . The larger inflow in rapidly accreting clusters is responsible for the de-

10

Lau et al

crease in density at 0.5 r/R200m 1. The right panel


of Figure 8 shows the logarithmic gas density slope gas
d log gas /d log r. More rapidly accreting clusters have shallower density slopes in both DM and gas densities at 0.1
r/R200m 0.3 and slightly steeper slope at 0.3 r/R200m 1.
We note, however, that there is a large scatter in the gas density slopes for a given 200m bin.
In this section we examine how well gas density traces DM
density in cluster outskirts and its dependence on redshift and
MAR. The relation between gas and DM densities can be useful for prescribing gas distribution in DM-only simulations.
Note that in this section, we do not exclude substructures and
filaments in either gas or DM components.
We find that gas traces DM in a uniform manner with redshift when we define clusters with m = 200. The left panel of
Figure 9 shows the ratio of gas-to-DM density (gas /DM ) as a
function of the cluster-centric radius at z = {0.0, 0.5, 1.0, 1.5}.
Here, the gas and DM densities shown are the mean values in each spherical bin, and the ratio is normalized to the
cosmic baryon-to-DM ratio (b /DM ). Similar to the thermodynamic profiles discussed in Section 4.1, the profile of
the gas-to-DM ratio is universal with z in the radial range
0.3 r/R200m 2. While the gas density roughly traces the
DM density in this radial range, there is 10% 20% deviation
in the gas-to-DM density ratio from the cosmic value around
the accretion shock. The value of gas /DM is below the cosmic value in the intermediate region (0.6 r/R200m 1), but
exceeds the cosmic value in the inner (0.3 r/R200m 0.6)
and outer (1.0 r/R200m 3) regions of clusters, and asymptotically approaches the cosmic value beyond 3R200m . The deviation from the cosmic value is on average about 10% around
the accretion shock, but could reach to more than 20% for individual clusters. This pattern in the gas-to-DM density ratio
profile originates from the difference in the dynamics between
gas and DM discussed in Section 4.1. Shock heating and ram
pressure cause gas to lag behind DM during infall, creating an
overdensity of gas around the accretion shock Rsh , indicated
by the peak in the gas /DM profile. The collisionless DM, on
the other hand, penetrates further into the inner region of the
cluster, undergoes core passage, and accumulates at the first
apocenter passage, leading to slightly overdense DM density
at r . 0.8R200m . This feature corresponds to the outermost
caustic of the DM (Diemer & Kravtsov 2014; Adhikari et al.
2014). In the right panel of Figure 9, we also show the dependence of MAR in the profiles of the gas-to-DM density ratio
for the z = 0 clusters. The profiles follow the same pattern as
in the left panel. The peak of the profiles, which corresponds
to the location of the accretion shock, occurs at smaller radii
for clusters with the highest MAR.
Recently, Patej & Loeb (2014) proposed an analytical
model of gas distribution in galaxy clusters that depends on
the ratio of the gas density jump to DM density jump at the
accretion shock. They find that fitting their model to observed
gas density profiles infers a similar gas density jump to that
of DM density around the accretion shock, consistent with
our findings that the gas density traces DM density to within
. 20% in the cluster outskirts. Their model can be further
improved by considering the effect of differential dynamics
between the collisionless DM and collisional gas that leads to
the deviation of the gas-to-DM density ratio from the cosmic
mean. This will provide an unique approach for constraining
the physics of cluster accretion shocks based on observations

fit
200m < 3.25
200m 3.25

2.4
2.2

Rsh /R200m

4.4. How Does Gas Trace Dark Matter in Cluster Outskirts?

2.6

2.0
1.8
1.6
1.4
1.2
1.0
1.0

0.8

0.6

0.4 0.2
200m

0.0

0.2

F IG . 10. Relation between the shock radius Rsh /R200m and the MAR
200m of clusters at z = {0.0, 0.5, 1.0, 1.5} binned by their peak heights 200m .
The filled and empty points represent clusters with 200m greater than or less
than 3.25, respectively. The line represents the best-fit Rsh /R200m -200m relation.

of the inner ICM profiles.


4.5. Location of the Accretion Shock

The dependence of outskirt gas profiles on 200m is similar to that of z for c = 200 (see Figures 2 and 3). The accretion shock occurs closer to the cluster center for systems
with more negative values of 200m . Similarly, for c = 200,
high-z clusters on average have their accretion shock closer
to the cluster center. This suggests that the apparent evolution of the profiles for c = 200 originates from the evolution
in the physical MAR, where high-z clusters experience more
rapid mass accretion than low-z counterparts. While normalizing clusters with respect to m = 200 will account for the
redshift dependence of the average MAR, the residual differences in MAR between clusters at a given redshift contribute
to the scatter in the outskirt gas profiles. We further investigate how the location of the accretion shock depends on redshift, mass, and accretion rate. In Figure 10, we characterize
the relationship between the location of the accretion shock
Rsh in units of R200m and the MAR proxy 200m . We divide the
cluster sample into a high and low peak height bin, splitting
at 200m = c /(M200m , z) = 3.25. The accretion shock radius
Rsh /R200m has a linear correlation that is independent of peak
height. Halos with different peak height occupy different regions along the relation, where low peak height halos tend to
have slightly larger Rsh /R200m and low MAR (more positive
200m ), although this trend is fairly weak.
We quantify the best-fit relation between Rsh /R200m and
200m by performing linear least square fit:
Rsh /R200m = A + B200m

(4)

where A = 1.990 0.030 and B = 0.782 0.067. Note that


this accretion shock radius is considerably larger than R200c
( 0.58 R200m at z = 0), or the virial radius Rvir ( 0.78 R200m
at z = 0). This redshift independent location of the accretion
shocks should be useful for modeling how accretion shocks
generate non-thermal pressure (Shi & Komatsu 2014), cos-

Gas Profiles in Cluster Outskirts

accretion shocks, and steepens the gas density profile


relative to the DM profile at large cluster-centric radii.
Recent ultra-deep (& 2 Msec) Chandra observation of
Abell 133 and Abell 1795 (Vikhlinin et al., in prep.)
may be able to detect the steepening in X-ray emissions
in the diffuse ICM component, after properly removing
point sources, clumps, and filaments.

mic rays (see Brunetti & Jones 2014, for review), and nonequilibrium electrons (e.g., Avestruz et al. 2014) as well as
assessing their effects on the hydrostatic mass bias (e.g., Lagan et al. 2010). Note further that we defined the shock
radius using the peak of the azimuthally averaged entropy
profile, while the actual topology of the accretion shock is
quite complicated, which contributes to the large scatter in
the Rsh /R200m 200m relation.
5. CONCLUSIONS AND DISCUSSION

In this work we investigated the self-similarity of the diffuse


X-ray emitting gas profiles in the outskirts of galaxy clusters
using a mass-limited sample of simulated clusters extracted
from the Omega500 cosmological hydrodynamic simulation.
Our main results are summarized below:
1. The radial profiles of the diffuse ICM in the outskirts of
galaxy clusters at r & R200c 0.6R200m exhibit remarkable self-similarity with redshift when they are normalized with respect to the mean density of the universe,
while in the inner regions of clusters they are more
self-similar when normalized with respect to the critical density. This difference in the scaling property of
the ICM radial profiles originates from the fact that the
outer gas profiles are determined by late time accretion
governed by the mean density of the universe, while
the inner profiles are determined by the gravitational
potential that is set when the universe is still matterdominated and stays roughly constant afterward.
2. The diffuse ICM profiles in cluster outskirts depend on
the mass accretion rate (MAR) of the cluster. For example, we find that the pressure and temperature profiles of
low MAR clusters are systematically higher than those
of high MAR clusters, because a significant fraction of
kinetic energy associated with accreting materials has
not yet thermalized by the accretion shocks in clusters
with high MAR. This means that the selection functions
of SZ surveys depend on the MAR of clusters by being
more sensitive to low MAR clusters which have higher
thermal pressure in their outskirts. Our work suggests
that MAR must be taken into account when interpreting SZ observations of cluster outskirts, including the
recent Plancks stacked SZ measurements which detected thermal pressure profiles around massive clusters
out to 3 R500c 1.2 R200m (Planck Collaboration Int.
V 2013).
3. Gas does not trace DM perfectly in the infall regions
of galaxy clusters, because the infall velocities of the
collisional gas are less than those of the collisionless
DM. This causes the gas-to-DM density ratio to deviate from the cosmic mean value by about 10% near the

11

4. The mean accretion shock Rsh is located at the fixed


fraction of R200m (Rsh /R200m 1.6) of galaxy clusters
for a wide range of redshift (0 z 1.5). However,
there is also a large scatter in the accretion shock radius (Rsh /R200m 1.0 2.4), depending on the MAR
of clusters. Higher MAR clusters have smaller the accretion shock radius. These results can be useful for
modeling physical processes (such as generation of turbulence and cosmic-rays) related to accretion and shock
heating at outer boundaries of galaxy clusters.
5. Our results suggest that the critical density is still preferred in defining cluster mass and radius, for calibrations of observable-mass relations (e.g., M YX and
M YSZ ) based on the current generation of X-ray and
SZ profile measurements, which mostly probe gas out
to r . R500c . Since the outer profiles are more selfsimilar when they are normalized with respect to the
mean mass density, the exploitation of cluster outskirts
for cosmology requires some care. For example, using the critical density in normalizing the outer ICM
profiles can introduce redshift-dependent systematic biases in cluster scaling relations. Furthermore, scaling relations of cluster outskirts are expected to show
larger scatter due to variations in MAR. Detailed understanding of physical processes and observational biases will be critical for interpreting data from the nextgeneration of X-ray and SZ missions, such as SMARTX3 and Athena+4 .

We thank Benedikt Diemer, Oleg Gnedin, Eiichiro Komatsu, Andrey Kravtsov, Avi Loeb, Xun Shi, and Andrew Wetzel for useful discussion and/or comments on the
manuscript. This work was supported in part by NSF grants
AST-1412768 & 1009811, NASA ATP grant NNX11AE07G,
NASA Chandra grants GO213004B and TM4-15007X, the
Research Corporation, and by the facilities and staff of the
Yale University Faculty of Arts and Sciences High Performance Computing Center. CA acknowledges support from
the NSF Graduate Student Research Fellowship and Alan D.
Bromley Fellowship from Yale University.

REFERENCES
Adhikari, S., Dalal, N., & Chamberlain, R. T. 2014, ArXiv e-prints,
arXiv:1409.4482
Allen, S. W., Evrard, A. E., & Mantz, A. B. 2011, ARA&A, 49, 409
Arnaud, M., Pratt, G. W., Piffaretti, R., et al. 2010, A&A, 517, A92
Avestruz, C., Nagai, D., Lau, E. T., & Nelson, K. 2014, ApJ, submitted,
arXiv:1410.8142
Bautz, M. W., Miller, E. D., Sanders, J. S., et al. 2009, PASJ, 61, 1117
3
4

http://smart-x.cfa.harvard.edu/
http://athena2.irap.omp.eu/

Bertschinger, E. 1985, ApJS, 58, 39


Brunetti, G., & Jones, T. W. 2014, International Journal of Modern Physics
D, 23, 30007
Bryan, G. L., & Norman, M. L. 1998, ApJ, 495, 80
Cavaliere, A., Lapi, A., & Fusco-Femiano, R. 2011, ApJ, 742, 19
Cuesta, A. J., Prada, F., Klypin, A., & Moles, M. 2008, MNRAS, 389, 385
Diemer, B., & Kravtsov, A. V. 2014, ApJ, 789, 1
Faucher-Gigure, C.-A., Kere, D., & Ma, C.-P. 2011, MNRAS, 417, 2982
George, M. R., Fabian, A. C., Sanders, J. S., Young, A. J., & Russell, H. R.
2009, MNRAS, 395, 657

12
Hoshino, A., Henry, J. P., Sato, K., et al. 2010, PASJ, 62, 371
Jenkins, A., Frenk, C. S., White, S. D. M., et al. 2001, MNRAS, 321, 372
Kaiser, N. 1986, MNRAS, 222, 323
Kawaharada, M., Okabe, N., Umetsu, K., et al. 2010, ApJ, 714, 423
Klypin, A., Kravtsov, A. V., Bullock, J. S., & Primack, J. R. 2001, ApJ, 554,
903
Kravtsov, A. V. 1999, PhD thesis, New Mexico State Univ.
Kravtsov, A. V., & Borgani, S. 2012, ARA&A, 50, 353
Kravtsov, A. V., Klypin, A., & Hoffman, Y. 2002, ApJ, 571, 563
Kravtsov, A. V., Vikhlinin, A., & Nagai, D. 2006, ApJ, 650, 128
Lagan, T. F., de Souza, R. S., & Keller, G. R. 2010, A&A, 510, A76
Li, Y., Mo, H. J., van den Bosch, F. C., & Lin, W. P. 2007, MNRAS, 379, 689
Maughan, B. J. 2007, ApJ, 668, 772
McBride, J., Fakhouri, O., & Ma, C.-P. 2009, MNRAS, 398, 1858
McCourt, M., Quataert, E., & Parrish, I. J. 2013, MNRAS, 432, 404
McDonald, M., Benson, B. A., Vikhlinin, A., et al. 2014, ApJ, 794, 67
Molnar, S. M., Hearn, N., Haiman, Z., et al. 2009, ApJ, 696, 1640
Nagai, D., Kravtsov, A. V., & Vikhlinin, A. 2007, ApJ, 668, 1
Nagai, D., & Lau, E. T. 2011, ApJ, 731, L10
Nelson, K., Lau, E. T., & Nagai, D. 2014a, ApJ, 792, 25
Nelson, K., Lau, E. T., Nagai, D., Rudd, D. H., & Yu, L. 2014b, ApJ, 782,
107
Nelson, K., Rudd, D. H., Shaw, L., & Nagai, D. 2012, ApJ, 751, 121
Okabe, N., Umetsu, K., Tamura, T., et al. 2014, PASJ, in press,
arXiv:1406.3451
Parrish, I. J., McCourt, M., Quataert, E., & Sharma, P. 2012, MNRAS, 419,
L29
Patej, A., & Loeb, A. 2014, ArXiv e-prints, arXiv:1411.2971
Pfrommer, C., Springel, V., Enlin, T. A., & Jubelgas, M. 2006, MNRAS,
367, 113
Planck Collaboration Int. V. 2013, A&A, 550, A131
Planelles, S., & Quilis, V. 2013, MNRAS, 428, 1643
Pratt, G. W., Croston, J. H., Arnaud, M., & Bhringer, H. 2009, A&A, 498,
361
Rasia, E., Lau, E. T., Borgani, S., et al. 2014, ApJ, 791, 96
Reiprich, T. H., Basu, K., Ettori, S., et al. 2013, Space Sci. Rev., 177, 195

Lau et al
Reiprich, T. H., Hudson, D. S., Zhang, Y.-Y., et al. 2009, A&A, 501, 899
Roncarelli, M., Ettori, S., Borgani, S., et al. 2013, MNRAS, 432, 3030
Rudd, D. H., Zentner, A. R., & Kravtsov, A. V. 2008, ApJ, 672, 19
Ryu, D., Kang, H., Hallman, E., & Jones, T. W. 2003, ApJ, 593, 599
Schaal, K., & Springel, V. 2014, ArXiv e-prints, arXiv:1407.4117
Shi, X., & Komatsu, E. 2014, MNRAS, 442, 521
Shi, X., Komatsu, E., Nelson, K., & Nagai, D. 2014, MNRAS, submitted,
arXiv:1408.3832
Simionescu, A., Allen, S. W., Mantz, A., et al. 2011, Science, 331, 1576
Skillman, S. W., OShea, B. W., Hallman, E. J., Burns, J. O., & Norman,
M. L. 2008, ApJ, 689, 1063
Tinker, J., Kravtsov, A. V., Klypin, A., et al. 2008, ApJ, 688, 709
Tozzi, P., & Norman, C. 2001, ApJ, 546, 63
Urban, O., Simionescu, A., Werner, N., et al. 2014, MNRAS, 437, 3939
van de Voort, F., Schaye, J., Booth, C. M., Haas, M. R., & Dalla Vecchia, C.
2011, MNRAS, 414, 2458
van den Bosch, F. C., Jiang, F., Hearin, A., et al. 2014, MNRAS, in press,
arXiv:1409.2750
Vazza, F., Brunetti, G., & Gheller, C. 2009a, MNRAS, 395, 1333
Vazza, F., Brunetti, G., Kritsuk, A., et al. 2009b, A&A, 504, 33
Vazza, F., Eckert, D., Simionescu, A., Brggen, M., & Ettori, S. 2013,
MNRAS, 429, 799
Vikhlinin, A., Kravtsov, A., Forman, W., et al. 2006, ApJ, 640, 691
Voit, G. M. 2005, Rev. Mod. Phys., 77, 207
Voit, G. M., Balogh, M. L., Bower, R. G., Lacey, C. G., & Bryan, G. L.
2003, ApJ, 593, 272
Walker, S. A., Fabian, A. C., Sanders, J. S., Simionescu, A., & Tawara, Y.
2013, MNRAS, 432, 554
Wechsler, R. H., Bullock, J. S., Primack, J. R., Kravtsov, A. V., & Dekel, A.
2002, ApJ, 568, 52
White, M. 2002, ApJS, 143, 241
Zhao, D. H., Mo, H. J., Jing, Y. P., & Brner, G. 2003, MNRAS, 339, 12
Zhuravleva, I., Churazov, E., Kravtsov, A., et al. 2013, MNRAS, 428, 3274

You might also like