scispace - formally typeset
Search or ask a question
Journal ArticleDOI

The cluster lensing and supernova survey with hubble: an overview

TL;DR: The Cluster Lensing And Supernova Survey with Hubble (CLASH) as mentioned in this paper is a 524-orbit Multi-Cycle Treasury Program to use the gravitational lensing properties of 25 galaxy clusters to accurately constrain their mass distributions.
Abstract: The Cluster Lensing And Supernova survey with Hubble (CLASH) is a 524-orbit Multi-Cycle Treasury Program to use the gravitational lensing properties of 25 galaxy clusters to accurately constrain their mass distributions. The survey, described in detail in this paper, will definitively establish the degree of concentration of dark matter in the cluster cores, a key prediction of structure formation models. The CLASH cluster sample is larger and less biased than current samples of space-based imaging studies of clusters to similar depth, as we have minimized lensing-based selection that favors systems with overly dense cores. Specifically, 20 CLASH clusters are solely X-ray selected. The X-ray-selected clusters are massive (kT > 5 keV) and, in most cases, dynamically relaxed. Five additional clusters are included for their lensing strength (θ_Ein > 35" at z_s = 2) to optimize the likelihood of finding highly magnified high-z (z > 7) galaxies. A total of 16 broadband filters, spanning the near-UV to near-IR, are employed for each 20-orbit campaign on each cluster. These data are used to measure precise (σ_z ~ 0.02(1 + z)) photometric redshifts for newly discovered arcs. Observations of each cluster are spread over eight epochs to enable a search for Type Ia supernovae at z > 1 to improve constraints on the time dependence of the dark energy equation of state and the evolution of supernovae. We present newly re-derived X-ray luminosities, temperatures, and Fe abundances for the CLASH clusters as well as a representative source list for MACS1149.6+2223 (z = 0.544).

Summary (5 min read)

1. INTRODUCTION

  • It is a “dark” universe where ∼23% of its mass-energy density is made up of weakly interacting (and, as yet, undetected) nonbaryonic particles (a.k.a. dark matter, DM) and ∼73% is as yet unknown physics (a.k.a. dark energy) that is driving an accelerated expansion of the metric (e.g., Komatsu et al.
  • Measure the profiles and substructures of DM in galaxy clusters with unprecedented precision and resolution.
  • Based on a current census of the HST data archive, CLASH will produce a six-fold increase in the number of lensing clusters observed to a depth of 20 orbits and, more importantly, will vastly increase the number of lensing clusters with extensive multiband HST imaging.
  • Subsequent sections describe the cluster sample (Section 3), survey design (Section 4), data pipeline (Section 5), and supporting observations using other facilities (Section 6).

2.1. Galaxy Cluster Dark Matter Profiles and Formation Times

  • Recent observations suggest that galaxy clusters formed earlier in their universe than in simulated ΛCDM universes.
  • Similarly, clusters have also been found to have somewhat larger than expected Einstein radii, a direct and particularly accurate measure of the projected mass in a halo’s core (Broadhurst & Barkana 2008; Richard et al.
  • Importantly, cluster elongation along the line of sight (a potential bias in concentration measurements) can be measured by the combination of lensing and X-ray analysis (Morandi et al. 2011; Newman et al. 2011).
  • While the robustness of this new result is still being assessed, it raises the possibility that the combination of new observations of an unbiased sample of clusters and new simulations may be able to bridge the concentration gap.

2.2. Improved Constraints on the Dark Energy Equation of State and SNe Evolution

  • The biggest cosmological surprise in decades came from observations of high-redshift SNe Ia, providing the first evidence that the expansion of the universe now appears to be accelerating (Riess et al.
  • The goal for cosmologists now is to measure the equation of state of dark energy, w = P/(ρc2), and its time variation in the hope of discriminating between viable explanations.
  • This is the distribution of times that elapse between a brief burst of star formation and the subsequent SN Ia explosions.
  • By z > 1.5, the measurements are most sensitive to evolution if present (e.g., the changing C/O ratio of the donor star), providing the means to diagnose and calibrate the degree of SN Ia evolution in dark energy measurements.

2.3. Detection and Characterization of z > 7 Galaxies

  • The majority of galaxies at z = 6–7 have been discovered via two methods: (1) deep pencil-beam surveys, such as the Hubble Ultra Deep Field (HUDF; Beckwith et al. 2006) and the Great Observatories Origins Deep Survey (Giavalisco et al. 2004) and (2) degree-size surveys with 10 m class groundbased telescopes, such as the Subaru Deep Field.
  • Both use a color selection to search for “dropout” candidates—galaxies with deep IGM absorption at the wavelengths shortward of the redshifted Lyα break.
  • Gravitational lensing by clusters amplifies the flux of background sources considerably.

2.4. Galaxy Evolution

  • ΛCDM provides a robust theoretical framework for the evolution of DM halos.
  • ΛCDM predicts that structures form hierarchically, with small halos forming early and later assembling into larger halos.
  • Several open questions remain about the origin and nature of this stellar mass growth.
  • Thus, massive galaxies are predicted to have substantial gradients in the origin of their stars, with the innermost stars having formed in situ and the outer stars largely accreted through merging.
  • There are several observational challenges to measuring and characterizing these two modes of growth.

3. CLASH CLUSTER SAMPLE

  • The CLASH program is robustly measuring galaxy cluster DM profiles and concentrations for a systematic comparison with those realized in cosmological simulations.
  • Specifically, their cluster sample size and selection criteria were chosen to allow the robust measurement of deviations from the predicted cluster concentration distribution of ∼15% or more at high statistical confidence (∼99% C.L.) given a relatively unbiased ensemble of clusters (Section 2.1).

3.1. Cluster Sample Selection

  • To date, robust joint SL+WL analyses have only been performed on a small, highly biased sample of five to ten clusters .
  • A handful of the clusters in the CLASH X-ray-selected subset have some evidence for departures from symmetric X-ray surface brightness distributions.
  • These systems are briefly discussed in Section 3.3.
  • These clusters were also selected to cover a wide redshift range (0.18 < z < 0.90 with a median zmed = 0.40) allowing us to probe the full c(M, z) relations expected from simulations.

3.2. Cluster Sample Size Requirements

  • The required size of their “relaxed” cluster sample is derived from the goal to measure “average” cluster concentrations to ∼10% (after accounting for variations in mass and redshift) and to detect a ∼15% deviation from the concentrations of simulated clusters at 99% confidence.
  • Being more conservative and assuming a factor of two (100%) lensing bias (Meneghetti et al. 2010), the observed concentrations would still be ∼20% greater than expectations.

3.3. Notes on Clusters with Possible Substructure

  • While their X-ray selection criteria favor the inclusion of highly relaxed clusters in the CLASH sample, for eight of their clusters the dynamical state is somewhat ambiguous.
  • Clusters with mass ratio values below 0.95 are considered unrelaxed.
  • These two clusters are included in the A08 compilation (as well as in Schmidt & Allen 2007, hereafter SA07), and were classified in these works as dynamically relaxed.
  • This cluster shows evidence for substructure (SA07; M08).
  • Some evidence of merger activity along the line of sight may be suggested by the very high velocity dispersion of 1580 km s−1.

4. SURVEY DESIGN AND IMPLEMENTATION

  • The CLASH program consists of 524 HST orbits, including 50 for SN follow-up.
  • The multiband observations span the near-ultraviolet to near-infrared (2000–17000 Å).
  • Indeed, there are often overlaps in time when the “A” and “B” orientations are both being executed.
  • When the entire sequence of exposures for a cluster is completed, the region covered by all 16 filters subtends an area of 4.08 arcmin2.
  • CLASH clusters will be observed over the course of three annual HST observing cycles, with 10, 10, and 5 clusters to be observed in cycles 18, 19, and 20, respectively.

4.1. Filter Selection and Exposure Times

  • Redshift estimates for multiply lensed images are crucial for breaking lensing degeneracies and tightening constraints on mass profiles (e.g., Broadhurst et al.
  • With continuous sampling of the broad wavelength range from the NUV to NIR (∼ 2000–17000 Å) that is enabled with WFC3 and ACS the authors can now obtain very accurate photometric redshifts (photo-z’s) for most of the lensed objects down to an apparent F775W AB magnitude limit of 26.
  • The authors performed simulations to inform their filter selection and estimate their eventual photo-z precision.
  • The limiting magnitudes in Table 5 are for a 0.′′4 diameter circular aperture and a point source with a flat Fν spectrum.
  • The five NIR filters provide the ability to identify z > 7 galaxies with high confidence.

4.2. Dither Pattern

  • In each orbit the authors use a compact four-point dither pattern that provides half-pixel sampling along both detector axes.
  • The dither pattern serves to both improve the spatial sampling of the point spread function (PSF), especially for the WFC3/IR detector with its large pixel scale of 0.128 arcsec pixel−1, and to help remove hot pixels and other detector imperfections that may be unaccounted for in the calibration reference files.
  • In subsequent epochs involving WFC3/IR observations, either in prime or parallel, the authors use a slightly larger dither pattern to help identify and remove persistence artifacts from compact sources, which, if uncorrected, could possibly be misidentified as SN candidates.
  • While their small-scale dither patterns are much smaller than is needed to cover the WFC3/UVIS and ACS/WFC detector gaps, their cluster observations are obtained at two orientations, leaving only two small diamond-shaped regions (∼4.4 arcsec2 each) in the central cluster area without data in all 16 filters.

4.3. Observation Cadence and Supernova Follow-up

  • Parallel observations are being obtained with a primary science goal of detecting SNe Ia.
  • The ability to reprogram the later of the two orientations to follow-up an SN detected early in a cluster observing sequence sets the upper limit on the angular offset between the two orientations—both orientations must be accessible during the entire cluster sequence.
  • This constraint means that the two orientations cannot be more than ∼30◦ apart.
  • The CLASH and CANDELS (Grogin et al. 2011; Koekemoer et al. 2011) SN programs are tightly coordinated and, in fact, share a common pool of reserve orbits from which both programs can draw upon for follow-up.

4.4. ACS Failure Options

  • While ACS functionality was restored in SM4, it is now only a “single-string” instrument, meaning there is no redundant path if the main CCD electronics box experiences a failure.
  • This was a constraint imposed by the nature of the ACS repair.
  • The use of SNe Ia for cosmology in strongly lensed regions is fraught with difficulty (see Section 2.2).
  • If ACS fails permanently, the authors would abandon the dual orientation strategy and would, most likely, abandon the multiple epoch exposures, allowing the observations of each cluster to be completed on a much shorter time frame.

5. THE CLASH DATA PIPELINE

  • The bulk of their HST data analyses make use of mosaics of globally aligned and co-added images.
  • To accomplish this, the authors use the MosaicDrizzle pipeline (Koekemoer et al. 2002).
  • The authors find that this uncorrected CTE can most significantly affect their UVIS photometry as follows.
  • Trails from cosmic rays can leak into photometric apertures of non-detections, artificially boosting their observed fluxes.
  • This enables a mask to be generated that can flag any suspect pixels in the initial exposure of a CLASH visit.

5.1. Object Detection and Characterization

  • SExtractor (Bertin & Arnouts 1996) is used to detect objects and measure their photometry.
  • The authors prune these detections from their catalog by rejecting any object with only a single 5σ detection in one UVIS/ACS filter, as measured by SExtractor.
  • It also performs slightly better at deblending these fainter objects, including those at high-z as well as arcs (strongly lensed galaxies), from brighter nearby cluster galaxies.
  • For arcs that are either missing from the detection list or that are only partially detected, the authors construct manual photometric apertures.
  • The authors force SExtractor to adopt these apertures using the software package SExSeg (Coe et al. 2006).

5.2. Photometric Redshifts

  • Both software packages use χ2 minimization and template fitting but differ in their specific templates and their assumed priors.
  • LPZ uses the SED library optimized for the COSMOS survey (Ilbert et al. 2009) without template interpolation.
  • BPZ currently uses PEGASE SED templates (Fioc & Rocca-Volmerange 1997) which have been heavily recalibrated based on the FIREWORKS spectroscopic and photometric catalog (Wuyts et al. 2008).
  • BPZ allows for interpolation between adjacent templates and uses an empirically derived prior on redshift and type based on observed magnitude.

6. SUPPORTING OBSERVATIONS

  • As discussed above, having both weak and strong-lensing information as well as information about the cluster baryonic mass distribution are critical for deriving robust mass profiles and concentrations.
  • All CLASH clusters have X-ray imaging as well as wide-field multiband ground-based optical imaging.
  • The X-ray imaging is from Chandra/ACIS (Garmire et al. 2003) and some of the clusters have XMM/EPIC (Strüder et al.

7. SUMMARY

  • The precision to which these measurements are being made will provide an unprecedented foil against which the authors will challenge and ultimately expand their current ideas about structure formation and the nature of dark energy.
  • The 16-band HST imaging yields precise (2%(1 + z)) photometric redshifts for all galaxies brighter than F775W AB mag 26, including hundreds of strongly lensed galaxies.
  • CLASH data will also provide the mass calibrators for the next generation of big cosmological surveys such as the Dark Energy Survey (DES), Sunyaev-Zel’dovich surveys (e.g., the South Pole Telescope), and next generation X-ray cluster surveys.
  • The authors thank Jay Anderson and Norman Grogin for providing the ACS CTE and bias-striping correction algorithms used in their data pipeline.
  • The CLASH Multi-Cycle Treasury Program (GO-12065) is based on observations made with the NASA/ESA Hubble Space Telescope.

Did you find this useful? Give us your feedback

Figures (24)

Content maybe subject to copyright    Report

The Astrophysical Journal Supplement Series, 199:25 (23pp), 2012 April doi:10.1088/0067-0049/199/2/25
C
2012. The American Astronomical Society. All rights reserved. Printed in the U.S.A.
THE CLUSTER LENSING AND SUPERNOVA SURVEY WITH HUBBLE: AN OVERVIEW
Marc Postman
1
,DanCoe
1
, Narciso Ben
´
ıtez
2
, Larry Bradley
1
, Tom Broadhurst
3
, Megan Donahue
4
, Holland Ford
5
,
Or Graur
6
, Genevieve Graves
7
, Stephanie Jouvel
8
, Anton Koekemoer
1
, Doron Lemze
5
, Elinor Medezinski
5
,
Alberto Molino
2
, Leonidas Moustakas
9
, Sara Ogaz
1
, Adam Riess
1,5
, Steve Rodney
5
, Piero Rosati
10
, Keiichi Umetsu
11
,
Wei Zheng
5
, Adi Zitrin
6
, Matthias Bartelmann
12
, Rychard Bouwens
13
, Nicole Czakon
8
, Sunil Golwala
8
,OleHost
14
,
Leopoldo Infante
15
, Saurabh Jha
16
, Yolanda Jimenez-Teja
2
, Daniel Kelson
17
, Ofer Lahav
14
, Ruth Lazkoz
3
,
Dani Maoz
6
, Curtis McCully
16
, Peter Melchior
18
, Massimo Meneghetti
19
, Julian Merten
12
, John Moustakas
20
,
Mario Nonino
21
, Brandon Patel
16
, Enik
¨
oReg
¨
os
22
, J ack Sayers
8
, Stella Seitz
23
, and Arjen Van der Wel
24
1
Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21208, USA; postman@stsci.edu
2
Instituto de Astrof
´
ısica de Andaluc
´
ıa (CSIC), C/Camino Bajo de Hu
´
etor 24, Granada 18008, Spain
3
Department of Theoretical Physics, University of the Basque Country, P. O. Box 644, 48080 Bilbao, Spain
4
Department of Physics and Astronomy, Michigan State University, East Lansing, MI 48824, USA
5
Department of Physics and Astronomy, The Johns Hopkins University, 3400 North Charles Street, Baltimore, MD 21218, USA
6
School of Physics and Astronomy, Tel Aviv University, Tel-Aviv 69978, Israel
7
Department of Astronomy, University of California, 601 Campbell Hall, Berkeley, CA 94720, USA
8
Jet Propulsion Laboratory, California Institute of Technology, MS 169-327, Pasadena, CA 91109, USA
9
ESO-European Southern Observatory, D-85748 Garching bei M
¨
unchen, Germany
10
Institute of Astronomy and Astrophysics, Academia Sinica, P. O. Box 23-141, Taipei 10617, Taiwan
11
Institut f
¨
ur Theoretische Astrophysik, ZAH, Albert-Ueberle-Straß e 2, 69120 Heidelberg, Germany
12
Leiden Observatory, Leiden University, P. O. Box 9513,2300 RA Leiden, The Netherlands
13
Department of Physics, Mathematics, and Astronomy, California Institute of Technology, Pasadena, CA91125, USA
14
Department of Physics & Astronomy. University College London, Gower Street, London WCIE 6 BT, UK
15
Departamento de Astrono
´
ıa y Astrof
´
ısica, Pontificia Universidad Cat
´
olica de Chile, V. Mackenna 4860, Santiago 22, Chile
16
Department of Physics and Astronomy, Rutgers University, 136 Frelinghuysen Rd., Piscataway, NJ 08854, USA
17
Observatories of the Carnegie Institution of Washington, Pasadena, CA 91 101, USA
18
Center for Cosmology and Astro-Particle Physics, & Department of Physics; The Ohio State University, 191 W. Woodruff Ave., Columbus, Ohio 43210, USA
19
INAF, Osservatorio Astronomico di Bologna, & INFN, Sezione di Bologna; Via Ranzani 1, I-40127 Bologna, Italy
20
Center for Astrophysics and Space Sciences, University of California at San Diego. 9500 Gilman Dr., MC 0424, La Jolla, CA 92093, USA
21
INAF-Osservatorio Astronomico di Trieste, via G.B. Tiepolo 11, 40131 Trieste, Italy
22
European Laboratory for Particle Physics (CERN). CH-1211, Geneva 23, Switzerland
23
Universit
¨
ats-Sternwarte, M
¨
unchen, Scheinerstr. 1, D-81679 M
¨
unchen. Germany
24
Max-Planck Institute for Astronomy, K
¨
onigstuhl 17, D-69117, Heidelberg, Germany
Received 2011 June 16; accepted 2011 December 5; published 2012 March 14
ABSTRACT
The Cluster Lensing And Supernova survey with Hubble (CLASH) is a 524-orbit Multi-Cycle Treasury Program
to use the gravitational lensing properties of 25 galaxy clusters to accurately constrain their mass distributions.
The survey, described in detail in this paper, will definitively establish the degree of concentration of dark matter
in the cluster cores, a key prediction of structure formation models. The CLASH cluster sample is larger and less
biased than current samples of space-based imaging studies of clusters to similar depth, as we have minimized
lensing-based selection that favors systems with overly dense cores. Specifically, 20 CLASH clusters are solely
X-ray selected. The X-ray-selected clusters are massive (kT > 5 keV) and, in most cases, dynamically relaxed.
Five additional clusters are included for their lensing strength (θ
Ein
> 35

at z
s
= 2) to optimize the likelihood of
finding highly magnified high-z (z>7) galaxies. A total of 16 broadband filters, spanning the near-UV to near-IR,
are employed for each 20-orbit campaign on each cluster. These data are used to measure precise (σ
z
0.02(1 + z))
photometric redshifts for newly discovered arcs. Observations of each cluster are spread over eight epochs to enable
a search for Type Ia supernovae at z>1 to improve constraints on the time dependence of the dark energy equation
of state and the evolution of supernovae. We present newly re-derived X-ray luminosities, temperatures, and Fe
abundances for the CLASH clusters as well as a representative source list for MACS1149.6 + 2223 (z = 0.544).
Key words: dark energy dark matter Galaxy: evolution Galaxy: formation gravitational lensing: strong
gravitational lensing: weak
Online-only material: color figures, machine-readable table
1. INTRODUCTION
The universe has proven to be far more intriguing in its
composition than we knew it to be even just 14 years ago. It
is a “dark” universe where 23% of its mass-energy density is
made up of weakly interacting (and, as yet, undetected) non-
baryonic particles (a.k.a. dark matter, DM) and 73% is as
yet unknown physics (a.k.a. dark energy) that is driving an
accelerated expansion of the metric (e.g., Komatsu et al. 2011;
Riess et al. 2011). TheHubble Space Telescope (HST) has played
a key role in providing evidence for and constraining the nature
of both of these mysterious dark components (e.g., Riess et al.
1998; Perlmutter et al. 1999; Clowe et al. 2006).
Clusters of galaxies, by virtue of their position at the high end
of the cosmic mass power spectrum, provide a powerful way to
constrain the frequency of high amplitude perturbations in the
primordial density field. As such, they play a direct and funda-
mental role in testing cosmological models and in constraining
1

The Astrophysical Journal Supplement Series, 199:25 (23pp), 2012 April Postman et al.
the properties of DM, providing unique and independent tests
of any viable cosmology and structure formation scenario, and
possible modifications of the laws of gravity. A key ingredient
of such cluster-based cosmological tests is the mass distribu-
tion of clusters, both on (sub) Mpc scales and across the range
of populations. The best and highest resolution maps of DM
distribution in massive galaxy clusters come from observations
of strong gravitational lensing made by the Advanced Camera
for Surveys (ACS; Ford et al. 2003) on board HST. As part
of a Guaranteed Time Observation (GTO) program, deep (20-
orbit) multiband (4–6 filter) observations were obtained for five
galaxy clusters: A1689 (Broadhurst et al. 2005; Limousin et al.
2007; Coe et al. 2010), A1703 (Limousin et al. 2008; Saha &
Read 2009; Zitrin et al. 2010), A2218 (El
´
ıasd
´
ottir et al. 2007),
CL0024+1654 (Jee et al. 2007; Zitrin et al. 2009; Umetsu et al.
2010), and MS1358+6245 (Zitrin et al. 2011c). These results
have contributed to a reported tension between observed and
simulated galaxy cluster DM halos. Observed clusters appear to
have denser cores than simulated clusters of similar total (virial)
mass (e.g., Broadhurst et al. 2008;Ogurietal.2009; Sereno et al.
2010). Understanding the true constraints from observed con-
centration and mass profile measurements on ΛCDM structure
formation models is one of the important problems that can be
tackled with a deep, high angular resolution imaging survey of
a significantly larger and more homogeneously selected sample
of galaxy clusters.
In 2009 May, NASA successfully executed the final planned
Hubble Servicing mission, SM4, with the installation of the
Wide Field Camera 3 (WFC3; Kimble et al. 2008) and the repair
of ACS. Shortly thereafter, the Hubble Multi-Cycle Treasury
(MCT) Program was conceived to permit ambitious programs
(>500 orbits) with broad scientific potential that could not be
accomplished within the constraints of a single HST observing
cycle and that would take full advantage of the final era of a
newly refurbished HST.
The Cluster Lensing and Supernova survey with Hubble
(CLASH) was one of three MCT programs selected. CLASH
has four main science goals.
1. Measure the profiles and substructures of DM in galaxy
clusters with unprecedented precision and resolution.
2. Detect Type Ia supernovae (SNe Ia) out to redshift z 2.5
to measure the time dependence of the dark energy equation
of state and potential evolutionary effects in the SNe
themselves.
3. Detect and characterize some of the most distant galaxies
yet discovered at z>7.
4. Study the internal structure and evolution of the galaxies in
and behind these clusters.
To accomplish these objectives, the CLASH program targets 25
massive galaxy clusters and will image each in 16 passbands
using WFC3/UVIS, WFC3/IR, and ACS
/WFC. CLASH has
been allocated 524 HST orbits, spread out over cycles 18, 19, and
20. The majority of these orbits (474) are for cluster imaging
and, simultaneously, for the parallel SN search program. An
additional 50 orbits were allocated as a reserve for SN follow-
up observations. Based on a current census of the HST data
archive, CLASH will produce a six-fold increase in the number
of lensing clusters observed to a depth of 20 orbits and, more
importantly, will vastly increase the number of lensing clusters
with extensive multiband HST imaging.
Motivations for each of the main CLASH science goals are
provided in Sections 2.12.4. Subsequent sections describe the
cluster sample (Section 3), survey design (Section 4), data
pipeline (Section 5), and supporting observations using other
facilities (Section 6). Data products intended for public distri-
bution to the community are briefly described in Section 7.AB
magnitudes are used throughout (Oke 1974). The cosmologi-
cal parameters H
o
= 100 h km s
1
Mpc
1
,h= 0.7, Ω
m
=
0.30, and Λ = 0.70 are assumed in this paper.
2. SCIENTIFIC MOTIVATION
2.1. Galaxy Cluster Dark Matter Profiles and Formation Times
Recent observations suggest that galaxy clusters formed
earlier in our universe than in simulated ΛCDM universes. These
observations include the detection of perhaps unexpectedly
massive galaxy clusters at z>1 (Stanford et al. 2006;
Eisenhardt et al. 2008; Jee et al. 2009, 2011; Huang et al.
2009; Rosati et al. 2009; Papovich et al. 2010; Schwope et al.
2010; Gobat et al. 2011; Foley et al. 2011) and the finding that
some clusters at intermediate redshift (z 0.3) have denser
cores than clusters of similar mass produced in simulations
(Broadhurst et al. 2008; Broadhurst & Barkana 2008; Oguri
et al. 2009; Richard et al. 2010; Sereno et al. 2010; Zitrin et al.
2011a). While the evidence to date for early cluster growth is
suggestive, possible explanations include departures from the
Gaussian initial density fluctuation spectrum or higher levels
of dark energy in the past, so-called Early Dark Energy (EDE;
Fedeli & Bartelmann 2007; Sadeh & Rephaeli 2008; Francis
et al. 2009; Grossi & Springel 2009). If a significant quantity of
EDE (for example, Ω
DE
0.1atz = 6) suppressed structure
growth in the early universe, then clusters would have had to
start forming sooner to yield the numbers we observe today.
These scenarios remain allowable within current observational
constraints as described in the above papers, although some non-
Gaussian models can be ruled out by using the cosmic X-ray
background measurements (Lemze et al. 2009). The CLASH
data permit significant advances to be made toward supporting
or rejecting observational evidence for early cluster growth by
measuring core densities for a larger, less biased sample of
clusters.
In cosmological simulations, cold-dark-matter-(CDM)-
dominated halos of all masses consistently evolve to have a
roughly “universal” density profile that steepens with radius.
Functional forms that fit such a profile well include the “NFW”
profile (Navarro et al. 1996, 1997) and the Einasto/S
´
ersic pro-
file (S
´
ersic 1963; Einasto 1965; Navarro et al. 2004, 2010).
Furthermore, each simulated halo’s core density is related to the
background density of the universe at the halo’s formation time.
Halos that form later, including the most massive galaxy clus-
ters, are found to have the least dense cores in a relative sense.
Determining the relationship between the shape and depth of a
halo’s gravitational potential and its total mass as a function of
time thus provides fundamental constraints on structure forma-
tion.
In practice, the relative core densities, or “concentrations,
are measured (both in simulated and observed halos) as c
vir
=
r
vir
/r
2
, a ratio between the virial radius and the inner radius
at which the density slope of the fitted profile is isothermal
(ρ r
2
). Analyses of gravitational lensing spanning such a
large range of radii allow one to map the (primarily dark) matter
profiles of observed halos and measure their concentrations.
DM profiles are best mapped in cluster cores using strong-
lensing analysis of multiband HST imaging. The lensing-
based mass profile mapping is extended to the virial radii of
2

The Astrophysical Journal Supplement Series, 199:25 (23pp), 2012 April Postman et al.
Figure 1. Mass profiles measured for four well-studied, strongly lensing galaxy
clusters that are not included in the CLASH sample. All have similar mass
profiles as measured from Hubble observations of strong-lensing and Subaru
observations of weak-lensing distortion and magnification (Umetsu et al. 2011a,
their Figure 6). The averaged mass profile is in remarkably good agreement
with the standard NFW form (Umetsu et al. 2011b, their Figure 1) as shown
by the gray area (2σ confidence interval of the NFW fit), though with a higher
concentration than predicted from cosmological simulations. Both strong- and
weak-lensing probes are required to map the continuously steepening mass
profile from the inner core (10 kpc h
1
) out to beyond the virial radius
(2Mpch
1
).
(A color version of this figure is available in the online journal.)
Figure 2. Joint strong- and weak-lensing analyses are required to obtain tight
constraints on cluster concentrations as shown here for CL0024+17, a non-
CLASH cluster (Umetsu et al. 2010, from their Figures 15 and 17). Confidence
levels of 68.3%, 95.4%, and 99.7% are plotted in the c
vir
M
vir
plane.
(A color version of this figure is available in the online journal.)
clusters using weak-lensing analysis of wider field ground-based
multiband imaging, such as from Subaru. Results from Umetsu
et al. (2011a, 2011b) for four of the currently best-studied (non-
CLASH) clusters are shown in Figure 1.
Joint modeling of strong and weak lensing (SL+WL) yields
significantly better constraints on concentrations than either
probe alone. Quantitatively, Meneghetti et al. (2010) found their
joint SL+WL analyses of simulated clusters yield concentra-
tions to 11% accuracy, while WL-only and SL-only analy-
ses yielded 33% and 59% scatters, respectively. Figure 2
Figure 3. Mass profiles of the best-studied clusters to date are revealed to have
higher central density concentrations than simulated clusters of similar mass
and redshift. Reconciliation may be within reach given results from the latest
simulations and an estimated lensing bias which CLASH will avoid. The plotted
lines are mean concentrations at three different redshifts for clusters in these
simulations as calculated from the fitting formulae provided in those papers.
However, note that halos of this great mass are rare (or even non-existent at
these redshifts) in these simulations, so these results are mainly extrapolations,
as designatedby the dashed lines. The thinner dashed lines aboveillustrate a 50%
observational bias applied to the Prada et al. (2011) results. This bias has been
roughly estimated for non-CLASH clusters such as these which were selected
for study based on exceptional lensing strength (Hennawi et al. 2007; Oguri &
Blandford 2009; Meneghetti et al. 2010, 2011). Results from the currently best-
studied clusters are plotted here as squares (Umetsu et al. 2011a) and circles
(Oguri et al. 2009) labeled with abbreviated names and described further in
Table 1.
(A color version of this figure is available in the online journal.)
demonstrates how SL and WL analyses combine to yield ro-
bust constraints on the mass and concentration of CL0024+17
(Umetsu et al. 2010).
The best-studied (SL+WL) galaxy clusters to date have
been found to have overly high concentrations (dense cores)
compared to halos in N-body simulations with similar masses,
as shown in Figure 3 (Broadhurst et al. 2008;Ogurietal.2009;
Sereno et al. 2010) and detailed in Table 1. Recent simulations
(Prada et al. 2011) yield cluster concentrations that are over 50%
higher than previous simulations (e.g., Duffy et al. 2008). This
is a product of upturns as a function of both mass and redshift
found in these newer simulations, which are not yet understood.
Similarly, clusters have also been found to have somewhat larger
than expected Einstein radii, a direct and particularly accurate
measure of the projected mass in a halo’s core (Broadhurst &
Barkana 2008; Richard et al. 2010; Zitrin et al. 2011a, 2011b).
Unfortunately, the best-studied clusters to date have also
been among the strongest gravitational lenses known. Such
lensing-selected clusters are highly biased toward halos with
high concentrations, both intrinsically and as projected on the
sky due to halo elongation along the line of sight (Hennawi et al.
2007; Oguri & Blandford 2009; Meneghetti et al. 2010, 2011).
These biases are estimated to lead to systematically higher
concentrations by as much as 50% or more. However, such
a bias is insufficient to account for the discrepancy between
the observations and the predictions, that is until a very recent
analysis of new simulations was performed (see Figure 3 and
discussion below).
More robust conclusions require analysis of a larger, unbiased
cluster sample. Progress toward this goal has been made by
3

The Astrophysical Journal Supplement Series, 199:25 (23pp), 2012 April Postman et al.
Tab l e 1
Concentration Measurements for Previously Well-studied Lensing-selected Clusters
Constraints
a
Publication Cluster zM
vir
c
vir
χ
2
/dof
(10
15
M
h
1
)
SL+WL+mag Umetsu et al. (2011a)
b
A1689 0.187 1.34
+0.20
0.16
13.82
+1.72
1.62
4.73/17
SL+WL+mag Umetsu et al. (2011a)
b
A1703 0.281 1.29
+0.22
0.19
6.89
+1.04
0.91
7.14/19
SL+WL+mag Umetsu et al. (2011a)
b
A370 0.375 2.26
+0.26
0.23
4.56 ± 0.33 14.07/24
SL+WL+mag Umetsu et al. (2011a)
b
CL0024+17 0.395 1.37
+0.20
0.18
7.77
+0.97
0.87
11.47/20
SL+WL+mag Umetsu et al. (2011a)
b
RXJ1347.51145 0.451 1.73
+0.12
0.11
5.96
+0.37
0.35
45.06/25
RE+WL Oguri et al. (2009) SDSS J1531+3414 0.335 0.7
+0.29
0.24
7.9
+3.0
1.5
8.1/6
RE+WL Oguri et al. (2009) SDSS J1446+3032 0.464 0.8
+0.3
0.22
8.3
+3.9
3.1
6.4/6
RE+WL Oguri et al. (2009) SDSS J21110115 0.637 0.9
+0.41
0.32
14.1
+25.9
9.3
7.5/6
Notes. Of these eight clusters, only one—RXJ1347.51145—is in the CLASH sample.
a
SL = strong lensing; WL = weak lensing; mag = magnification bias (number counts); RE = Einstein radius.
b
The parameters here are from standard NFW fits to profiles in Umetsu et al. (2011a), and not from the generalized NFW (gNFW) fits
giveninthatpaper.
LoCuSS, the Local Cluster Substructure Survey (Smith et al.
2005). A large sample of 165 clusters between 0.15 <z<0.30
was selected based on X-ray brightness. Strong-lensing analyses
of 20 of these based on HST imaging (mostly single-band
“snapshots”) were presented by Richard et al. (2010). Okabe
et al. (2010a) published weak-lensing analyses of 30 LoCuSS
clusters, 22 of these being more “secure” based on multiband
Subaru imaging, and 9 of these 22 overlapping with the Richard
et al. (2010) subset. Subaru images are especially desirable for
WL studies as they enable excellent galaxy shape measurements
to be performed over a wide area. Multiband imaging is
also critical to properly select background galaxies and avoid
significantly diluting the weak-lensing signal (and thus the virial
mass and concentration) with unlensed foreground galaxies
(Medezinski et al. 2007, 2010). Stacked WL-only analyses have
been performed on large cluster samples (Johnston et al. 2007;
Mandelbaum et al. 2008), but we re-emphasize the need for
combining strong + weak lensing analyses in addition to cross-
comparisons with mass profile estimates from other techniques.
For example, mass concentrations can be measured from X-
ray profiles (Buote et al. 2007; Ettori et al. 2010), although
these are subject to uncertainties due to assumptions about
hydrostatic equilibrium. Importantly, cluster elongation along
the line of sight (a potential bias in concentration measurements)
can be measured by the combination of lensing and X-ray
analysis (Morandi et al. 2011; Newman et al. 2011). The caustic
technique (Diaferio & Geller 1997; Rines & Diaferio 2006)is,
like lensing-based methods, independent of the dynamical state
of the cluster and also provides an important cross check on mass
estimates from lensing and gas kinematics. Further discussion
of some of the previous results from various methods is given in
Comerford & Natarajan (2007), Coe (2010), Rines et al. (2010),
and King & Mead (2011).
Reconciliation of high observed concentrations with results
from simulations may ultimately come from significantly reduc-
ing the observational sample bias along with finding higher con-
centrations in simulated clusters. Baryons, for example, are cur-
rently absent from those cosmological simulations large enough
to produce massive clusters. Baryons can result in significant
“adiabatic contraction” on galaxy scales, however they consti-
tute a much smaller fraction of the mass on cluster scales. Simu-
lations including baryons show that cluster halo concentrations
are likely only varied by 10% or so relative to DM-only ha-
los, and that the direction of variation (increase or decrease) is
not even clear, depending on the gas physics assumed (Duffy
et al. 2010; Mead et al. 2010; King & Mead 2011). Nonetheless,
measuring the velocity dispersion of the brightest cluster galaxy
(BCG) as an additional constraint on the inner (80 kpc) mass
profile, where its stellar mass is a non-negligible component of
the matter distribution, provides for a more thorough mapping
of the total mass profile. Some clusters have been shown to have
inner mass profiles that are shallower than NFW (Sand et al.
2004, 2008; Newman et al. 2009, 2011
). Deviations of cluster
mass profiles from NFW or Einasto forms at small and/or large
radii may slightly bias concentration measurements (Oguri &
Hamana 2011). The use of multiple probes of the matter distri-
bution, as CLASH is designed to do, will enable such biases to
be measured and the significance of deviations determined.
More recently, an analysis by Prada et al. (2011) found that
clusters in the Bolshoi and MultiDark simulations have con-
centrations 50% higher than clusters in previous simulations.
While the robustness of this new result is still being assessed, it
raises the possibility that the combination of new observations of
an unbiased sample of clusters and new simulations may be able
to bridge the concentration gap. We stress here that estimates of
the observational bias from previous cluster studies have large
uncertainties and likely vary for each cluster. CLASH will deter-
mine mass profiles and concentrations for a new cluster sample
free of lensing selection bias. As we demonstrate in Section 3.2,
CLASH is designed to detect (or rule out) with 99% statistical
confidence average deviations of 15% or more from predicted
concentrations. However, given the outstanding uncertainties in
the expected concentrations, we prefer to recast the problem as
follows: CLASH will deliver robust observational concentration
measurements for a sample of clusters that simulations will be
tasked to reproduce. Ultimately, this will lead to a better calibra-
tion of many mass estimation techniques and, consequently, to
a better understanding of structure formation on cluster scales
and perhaps of our cosmological model as well.
2.2. Improved Constraints on the Dark Energy Equation of
State and SNe Evolution
The biggest cosmological surprise in decades came from
observations of high-redshift SNe Ia, providing the first evidence
that the expansion of the universe now appears to be accelerating
(Riess et al. 1998; Perlmutter et al. 1999), and indicating that the
universe is dominated by “dark energy. The presence of dark
4

The Astrophysical Journal Supplement Series, 199:25 (23pp), 2012 April Postman et al.
Figure 4. Dark energy and evolution sensitivity. Three models consistent with
current data are shown: w
0
= 0.9, w
a
= 0 (black); w
0
= 1, w
a
= 0.8 (red);
and w
0
= 1,w
a
= 0 (dashed). Also plotted blue is the SNe Ia evolution model
of Dom
´
ınguez et al. (2001), where the peak luminosity changes 3% per solar
mass change in the donor star. The error bars show the constraints at present, as
projected after CLASH, and if HST continues to collect SNe Ia at the present
rate for seven years.
(A color version of this figure is available in the online journal.)
energy has galvanized cosmologists as they seek to understand
it. Observations of high-redshift SNe Ia have continued to lead
the way in measuring the properties of dark energy (e.g., Riess
et al. 2011). The goal for cosmologists now is to measure
the equation of state of dark energy, w = P/(ρc
2
), and its
time variation in the hope of discriminating between viable
explanations. A departure of the present equation of state, w
0
,
from 1 or a detection of its variation, ∂w/∂z, would invalidate
an innate vacuum energy (i.e., the cosmological constant) as the
source of dark energy and would point toward a present epoch
of “weak inflation. A difference between the expansion history
and the growth history of structure expected for w(z) would
point toward a breakdown in general relativity as the cosmic
scale factor approaches unity.
HST paired with ACS is a unique tool in this investigation,
providing the only means to collect SNe Ia at 1 <z<1.5,
which, in turn, provide the only constraints we have to date
on the time variation of w. From the 23 SNe Ia at z>1
with HST data (Riess et al. 2004, 2007) we have learned: (1)
that cosmic expansion was once decelerating before it recently
began accelerating, (2) that dark energy, i.e., an energy density
with w<0, was already present during this prior decelerating
phase, (3) that SNe Ia at a look-back time of 10 Gyr appear
both spectroscopically and photometrically similar to those seen
locally, and (4) no rapid change is seen in w(z) and thus no
departure is yet seen from the cosmological constant, though the
constraint on the time variation remains an order of magnitude
worse than on the w
0
.
SNe Ia play a central role, not only as distance indicators for
cosmography, but also as major contributors to cosmic metal
production and distribution. The measurement of high-redshift
SN Ia rates is therefore integral to understanding the history
of chemical enrichment. The high-redshift rate cannot easily
be predicted from the star formation rate because the nature
and, hence, the timescales of the process behind the growth
of the white dwarf toward the Chandrasekhar mass are not
known. The two leading, competing scenarios are accretion
from a close binary companion—the single-degenerate scenario
(Whelan & Iben 1973; Nomoto 1982), or merger with another
white dwarf, following loss of orbital energy and angular
momentum by emission of gravitational waves—the double-
Tab l e 2
Estimated Number of z 1 SNe Ia to be Discovered During
the CLASH Program
Redshift Estimated Number
Range of SNe Ia Found
1.0 z<1.5 7–11
1.5 z<2.0 4–13
2.0 z 2.704
1.0 z 2.7 11–28
Notes. These estimates are the 68% confidence intervals, accounting
for both statistical and systematic uncertainties. The assumed SN Ia
rate is from Graur et al. (2011).
degenerate scenario (Iben & Tutukov 1984; Webbink 1984).
One way to constrain the different progenitor scenarios is to
measure the delay-time distribution (DTD) of SNe Ia. This is
the distribution of times that elapse between a brief burst of star
formation and the subsequent SN Ia explosions. Observations
have suggested variousdifferent forms for the DTD (e.g., Dahlen
et al. 2004, 2008; Mannucci et al. 2006; Pritchet et al. 2008).
However, a number of more recent measurements and analyses
point to a DTD that is a power law of index ≈−1 (Totani
et al. 2008; Maoz et al. 2010, 2011; Maoz & Badenes 2010;
Brandt et al. 2010; Horiuchi & Beacom 2010; Graur et al. 2011).
Specifically, Graur et al. (2011) have shown that such a DTD fits
well the measured SN Ia rate out to z 2. Small sample sizes
are the major limiting factor at high redshifts. Large high-z SN
samples are therefore needed to resolve the issue, and control
possible biases in cosmological studies (due to the evolving
SN Ia channel mix).
The CLASH survey uses ACS in parallel with the cluster
program to continue the discovery of SNe Ia at 1 <z<1.5, the
objects which tell us about the variation in w. With WFC3
in parallel, CLASH will yield SNe Ia at 1.5 <z<2.5.
Observations in this fully matter-dominated epoch provide the
unique chance to test SN Ia distance measurements for the
deleterious effects of evolution independent of our ignorance
of dark energy (Riess & Livio 2006). Because the SNe Ia
are detected when these cameras are in parallel, they are far
from the cluster core (2 Mpc at the median cluster redshift
of z = 0.4) and, hence, the effects of lensing are small (and
correctable), making the SNe usable for improving the limits
on the redshift variation of the dark energy equation of state.
At z<1, SN Ia distance measurements are most sensitive to
the static component of dark energy, w
0
.At1<z<1.5, the
measurements are most sensitive to the dynamic component, w
a
.
By z>1.5, the measurements are most sensitive to evolution
if present (e.g., the changing C/O ratio of the donor star),
providing the means to diagnose and calibrate the degree of
SN Ia evolution in dark energy measurements. Figure 4 shows
how variations in the DE equation of state or an evolution in the
white dwarf C/O ratio can change the SN Ia distance modulus
as a function of redshift.
Accounting for the systematic uncertainty introduced by the
cosmic star formation history, Graur et al. (2011) predicted the
SN Ia rate out to higher redshifts (their Figure 13), which we use
here, in conjunction with the CLASH observational parameters,
to estimate the number of z 1 SNe Ia that will be discovered
over the course of this program. These estimates are presented
in Table 2 and are the 68% confidence intervals, accounting for
both statistical and systematic uncertainties.
5

Citations
More filters
Journal ArticleDOI
TL;DR: The Illustris Project as mentioned in this paper is a series of large-scale hydrodynamical simulations of galaxy formation, which includes primordial and metal-line cooling with self-shielding corrections, stellar evolution, stellar feedback, gas recycling, chemical enrichment, supermassive black hole growth, and feedback from active galactic nuclei.
Abstract: We introduce the Illustris Project, a series of large-scale hydrodynamical simulations of galaxy formation. The highest resolution simulation, Illustris-1, covers a volume of (106.5 Mpc)^3, has a dark mass resolution of 6.26 × 10^6 M_⊙, and an initial baryonic matter mass resolution of 1.26 × 10^6 M_⊙. At z = 0 gravitational forces are softened on scales of 710 pc, and the smallest hydrodynamical gas cells have an extent of 48 pc. We follow the dynamical evolution of 2 × 1820^3 resolution elements and in addition passively evolve 1820^3 Monte Carlo tracer particles reaching a total particle count of more than 18 billion. The galaxy formation model includes: primordial and metal-line cooling with self-shielding corrections, stellar evolution, stellar feedback, gas recycling, chemical enrichment, supermassive black hole growth, and feedback from active galactic nuclei. Here we describe the simulation suite, and contrast basic predictions of our model for the present-day galaxy population with observations of the local universe. At z = 0 our simulation volume contains about 40 000 well-resolved galaxies covering a diverse range of morphologies and colours including early-type, late-type and irregular galaxies. The simulation reproduces reasonably well the cosmic star formation rate density, the galaxy luminosity function, and baryon conversion efficiency at z = 0. It also qualitatively captures the impact of galaxy environment on the red fractions of galaxies. The internal velocity structure of selected well-resolved disc galaxies obeys the stellar and baryonic Tully–Fisher relation together with flat circular velocity curves. In the well-resolved regime, the simulation reproduces the observed mix of early-type and late-type galaxies. Our model predicts a halo mass dependent impact of baryonic effects on the halo mass function and the masses of haloes caused by feedback from supernova and active galactic nuclei.

2,012 citations

Posted Content
TL;DR: In 2014, the Science Definition Team (SDT) of the Wide Field Infrared Survey Telescope (WFIRST) mission presented a design reference mission (DRM) for an implementation of WFIRST using one of the 2.4m, Hubble-quality telescopes recently made available to NASA as discussed by the authors.
Abstract: This report describes the 2014 study by the Science Definition Team (SDT) of the Wide-Field Infrared Survey Telescope (WFIRST) mission. It is a space observatory that will address the most compelling scientific problems in dark energy, exoplanets and general astrophysics using a 2.4-m telescope with a wide-field infrared instrument and an optical coronagraph. The Astro2010 Decadal Survey recommended a Wide Field Infrared Survey Telescope as its top priority for a new large space mission. As conceived by the decadal survey, WFIRST would carry out a dark energy science program, a microlensing program to determine the demographics of exoplanets, and a general observing program utilizing its ultra wide field. In October 2012, NASA chartered a Science Definition Team (SDT) to produce, in collaboration with the WFIRST Study Office at GSFC and the Program Office at JPL, a Design Reference Mission (DRM) for an implementation of WFIRST using one of the 2.4-m, Hubble-quality telescope assemblies recently made available to NASA. This DRM builds on the work of the earlier WFIRST SDT, reported by Green et al. (2012) and the previous WFIRST-2.4 DRM, reported by Spergel et. (2013). The 2.4-m primary mirror enables a mission with greater sensitivity and higher angular resolution than the 1.3-m and 1.1-m designs considered previously, increasing both the science return of the primary surveys and the capabilities of WFIRST as a Guest Observer facility. The addition of an on-axis coronagraphic instrument to the baseline design enables imaging and spectroscopic studies of planets around nearby stars.

1,009 citations

Journal ArticleDOI
Peter A. R. Ade, Nabila Aghanim, Monique Arnaud, M. Ashdown, J. Aumont, Carlo Baccigalupi, A. J. Banday, R. B. Barreiro, J. G. Bartlett, N. Bartolo, E. Battaner, Richard A. Battye, K. Benabed, Alain Benoit, A. Benoit-Lévy, J.-P. Bernard, Marco Bersanelli, P. Bielewicz, J. J. Bock, Anna Bonaldi, Laura Bonavera, J. R. Bond, Julian Borrill, François R. Bouchet, M. Bucher, Carlo Burigana, R. C. Butler, Erminia Calabrese, Jean-François Cardoso, A. Catalano, Anthony Challinor, A. Chamballu, R.-R. Chary, H. C. Chiang, P. R. Christensen, Sarah E. Church, David L. Clements, S. Colombi, L. P. L. Colombo, C. Combet, B. Comis, F. Couchot, A. Coulais, B. P. Crill, A. Curto, F. Cuttaia, Luigi Danese, R. D. Davies, R. J. Davis, P. de Bernardis, A. de Rosa, G. de Zotti, Jacques Delabrouille, F.-X. Désert, Jose M. Diego, Klaus Dolag, H. Dole, S. Donzelli, Olivier Doré, Marian Douspis, A. Ducout, X. Dupac, George Efstathiou, F. Elsner, Torsten A. Enßlin, H. K. Eriksen, E. Falgarone, James R. Fergusson, Fabio Finelli, Olivier Forni, M. Frailis, A. A. Fraisse, E. Franceschi, A. Frejsel, S. Galeotta, S. Galli, K. Ganga, M. Giard, Y. Giraud-Héraud, E. Gjerløw, J. González-Nuevo, Krzysztof M. Gorski, Serge Gratton, A. Gregorio, Alessandro Gruppuso, Jon E. Gudmundsson, F. K. Hansen, Duncan Hanson, D. L. Harrison, Sophie Henrot-Versille, C. Hernández-Monteagudo, D. Herranz, S. R. Hildebrandt, E. Hivon, Michael P. Hobson, W. A. Holmes, Allan Hornstrup, W. Hovest, Kevin M. Huffenberger 
TL;DR: In this article, the authors present constraints on cosmological parameters using number counts as a function of redshift for a sub-sample of 189 galaxy clusters from the Planck SZ (PSZ) catalogue.
Abstract: We present constraints on cosmological parameters using number counts as a function of redshift for a sub-sample of 189 galaxy clusters from the Planck SZ (PSZ) catalogue. The PSZ is selected through the signature of the Sunyaev--Zeldovich (SZ) effect, and the sub-sample used here has a signal-to-noise threshold of seven, with each object confirmed as a cluster and all but one with a redshift estimate. We discuss the completeness of the sample and our construction of a likelihood analysis. Using a relation between mass $M$ and SZ signal $Y$ calibrated to X-ray measurements, we derive constraints on the power spectrum amplitude $\sigma_8$ and matter density parameter $\Omega_{\mathrm{m}}$ in a flat $\Lambda$CDM model. We test the robustness of our estimates and find that possible biases in the $Y$--$M$ relation and the halo mass function are larger than the statistical uncertainties from the cluster sample. Assuming the X-ray determined mass to be biased low relative to the true mass by between zero and 30%, motivated by comparison of the observed mass scaling relations to those from a set of numerical simulations, we find that $\sigma_8=0.75\pm 0.03$, $\Omega_{\mathrm{m}}=0.29\pm 0.02$, and $\sigma_8(\Omega_{\mathrm{m}}/0.27)^{0.3} = 0.764 \pm 0.025$. The value of $\sigma_8$ is degenerate with the mass bias; if the latter is fixed to a value of 20% we find $\sigma_8(\Omega_{\mathrm{m}}/0.27)^{0.3}=0.78\pm 0.01$ and a tighter one-dimensional range $\sigma_8=0.77\pm 0.02$. We find that the larger values of $\sigma_8$ and $\Omega_{\mathrm{m}}$ preferred by Planck's measurements of the primary CMB anisotropies can be accommodated by a mass bias of about 40%. Alternatively, consistency with the primary CMB constraints can be achieved by inclusion of processes that suppress power on small scales relative to the $\Lambda$CDM model, such as a component of massive neutrinos (abridged).

738 citations

Journal ArticleDOI
TL;DR: In this article, the effects of self-interacting dark matter (SIDM) on the density profiles and substructure counts of dark matte r halos from the scales of spiral galaxies to galaxy clusters are studied.
Abstract: We use cosmological simulations to study the effects of self-interacting dark matter (SIDM) on the density profiles and substructure counts of dark matte r halos from the scales of spiral galaxies to galaxy clusters, focusing explicitly on mod els with cross sections over dark matter particle mass σ/m = 1 and 0.1 cm 2 /g. Our simulations rely on a new SIDM N-body algorithm that is derived self-consistently from the Boltz mann equation and that reproduces analytic expectations in controlled numerical experiments. We find that well-resolved SIDM halos have constant-density cores, with significantly lowe r central densities than their CDM counterparts. In contrast, the subhalo content of SIDM halos is only modestly reduced compared to CDM, with the suppression greatest for large hosts and small halo-centric distances. Moreover, the large-scale clustering and halo circular vel ocity functions in SIDM are effectively identical to CDM, meaning that all of the large-scale successes of CDM are equally well matched by SIDM. From our largest cross section runs we are able to extract scaling relations for core sizes and central densities over a range o f halo sizes and find a strong correlation between the core radius of an SIDM halo and the NFW scale radius of its CDM counterpart. We construct a simple analytic model, based on CDM scaling relations, that captures all aspects of the scaling relations for SIDM halos. Our results show that halo core densities in σ/m = 1 cm 2 /g models are too low to match observations of galaxy clusters, low surface brightness spirals (LSBs), and dwarf spheroidal galaxies. However, SIDM with σ/m ≃ 0.1 cm 2 /g appears capable of reproducing reported core sizes and central densities of dwarfs, LSBs, and galaxy clusters without the need for velocity dependence. Higher resolution simulations over a wider range of masses will be required to confirm this expectation. We discuss constraints arising from the Bullet cluster observ ations, measurements of dark matter density on small-scales and subhalo survival requirements, and show that SIDM models with σ/m ≃ 0.1 cm 2 /g ≃ 0.2 barn/GeV are consistent with all observational constraints.

736 citations


Cites background from "The cluster lensing and supernova s..."

  • ...The CLASH multi-cycle treasury p ogram on theHubble Space Telecope is finding many new strongly lensed galaxies in about a set of 25 massive clusters (Postman et al. 2012)....

    [...]

Journal ArticleDOI
Peter A. R. Ade1, Nabila Aghanim2, Monique Arnaud3, M. Ashdown  +282 moreInstitutions (70)
TL;DR: In this article, the authors presented cluster counts and corresponding cosmological constraints from the Planck full mission data set and extended their analysis to the two-dimensional distribution in redshift and signal-to-noise.
Abstract: We present cluster counts and corresponding cosmological constraints from the Planck full mission data set. Our catalogue consists of 439 clusters detected via their Sunyaev-Zeldovich (SZ) signal down to a signal-to-noise ratio of 6, and is more than a factor of 2 larger than the 2013 Planck cluster cosmology sample. The counts are consistent with those from 2013 and yield compatible constraints under the same modelling assumptions. Taking advantage of the larger catalogue, we extend our analysis to the two-dimensional distribution in redshift and signal-to-noise. We use mass estimates from two recent studies of gravitational lensing of background galaxies by Planck clusters to provide priors on the hydrostatic bias parameter, (1−b). In addition, we use lensing of cosmic microwave background (CMB) temperature fluctuations by Planck clusters as an independent constraint on this parameter. These various calibrations imply constraints on the present-day amplitude of matter fluctuations in varying degrees of tension with those from the Planck analysis of primary fluctuations in the CMB; for the lowest estimated values of (1−b) the tension is mild, only a little over one standard deviation, while it remains substantial (3.7σ) for the largest estimated value. We also examine constraints on extensions to the base flat ΛCDM model by combining the cluster and CMB constraints. The combination appears to favour non-minimal neutrino masses, but this possibility does little to relieve the overall tension because it simultaneously lowers the implied value of the Hubble parameter, thereby exacerbating the discrepancy with most current astrophysical estimates. Improving the precision of cluster mass calibrations from the current 10%-level to 1% would significantly strengthen these combined analyses and provide a stringent test of the base ΛCDM model.

606 citations


Cites background from "The cluster lensing and supernova s..."

  • ...Among these, the Weighing the Giants (WtG, von der Linden et al. 2014a), CLASH (Postman et al. 2012; Merten et al. 2015; Umetsu et al. 2014), and Canadian Cluster Comparison Project (CCCP, Hoekstra et al. 2015) programmes offer constraints on our mass bias factor, 1−b, through direct comparison of…...

    [...]

  • ...Among these, the Weighing the Giants (WtG, von der Linden et al. 2014a), CLASH (Postman et al. 2012; Merten et al. 2015; Umetsu et al. 2014), and Canadian Cluster Comparison Project (CCCP, Hoekstra et al. 2015) programmes offer constraints on our mass bias factor, 1−b, through direct comparison of the lensing masses to the Planck mass proxy, MYz....

    [...]

References
More filters
Journal ArticleDOI
TL;DR: In this paper, the mass density, Omega_M, and cosmological-constant energy density of the universe were measured using the analysis of 42 Type Ia supernovae discovered by the Supernova Cosmology project.
Abstract: We report measurements of the mass density, Omega_M, and cosmological-constant energy density, Omega_Lambda, of the universe based on the analysis of 42 Type Ia supernovae discovered by the Supernova Cosmology Project. The magnitude-redshift data for these SNe, at redshifts between 0.18 and 0.83, are fit jointly with a set of SNe from the Calan/Tololo Supernova Survey, at redshifts below 0.1, to yield values for the cosmological parameters. All SN peak magnitudes are standardized using a SN Ia lightcurve width-luminosity relation. The measurement yields a joint probability distribution of the cosmological parameters that is approximated by the relation 0.8 Omega_M - 0.6 Omega_Lambda ~= -0.2 +/- 0.1 in the region of interest (Omega_M <~ 1.5). For a flat (Omega_M + Omega_Lambda = 1) cosmology we find Omega_M = 0.28{+0.09,-0.08} (1 sigma statistical) {+0.05,-0.04} (identified systematics). The data are strongly inconsistent with a Lambda = 0 flat cosmology, the simplest inflationary universe model. An open, Lambda = 0 cosmology also does not fit the data well: the data indicate that the cosmological constant is non-zero and positive, with a confidence of P(Lambda > 0) = 99%, including the identified systematic uncertainties. The best-fit age of the universe relative to the Hubble time is t_0 = 14.9{+1.4,-1.1} (0.63/h) Gyr for a flat cosmology. The size of our sample allows us to perform a variety of statistical tests to check for possible systematic errors and biases. We find no significant differences in either the host reddening distribution or Malmquist bias between the low-redshift Calan/Tololo sample and our high-redshift sample. The conclusions are robust whether or not a width-luminosity relation is used to standardize the SN peak magnitudes.

16,838 citations

Journal ArticleDOI
TL;DR: In this article, the authors used spectral and photometric observations of 10 Type Ia supernovae (SNe Ia) in the redshift range 0.16 " z " 0.62.
Abstract: We present spectral and photometric observations of 10 Type Ia supernovae (SNe Ia) in the redshift range 0.16 " z " 0.62. The luminosity distances of these objects are determined by methods that employ relations between SN Ia luminosity and light curve shape. Combined with previous data from our High-z Supernova Search Team and recent results by Riess et al., this expanded set of 16 high-redshift supernovae and a set of 34 nearby supernovae are used to place constraints on the following cosmo- logical parameters: the Hubble constant the mass density the cosmological constant (i.e., the (H 0 ), () M ), vacuum energy density, the deceleration parameter and the dynamical age of the universe ) " ), (q 0 ), ) M \ 1) methods. We estimate the dynamical age of the universe to be 14.2 ^ 1.7 Gyr including systematic uncer- tainties in the current Cepheid distance scale. We estimate the likely e†ect of several sources of system- atic error, including progenitor and metallicity evolution, extinction, sample selection bias, local perturbations in the expansion rate, gravitational lensing, and sample contamination. Presently, none of these e†ects appear to reconcile the data with and ) " \ 0 q 0 " 0.

16,674 citations


"The cluster lensing and supernova s..." refers background in this paper

  • ...The Hubble Space Telescope (HST) has played a key role in providing evidence for and constraining the nature of both of these mysterious dark components (e.g., Riess et al. 1998; Perlmutter et al. 1999; Clowe et al. 2006)....

    [...]

  • ...…surprise in decades came from observations of high-redshift type-Ia supernovae (SNe Ia), providing the first evidence that the expansion of the Universe now appears to be accelerating (Riess et al. 1998; Perlmutter et al. 1999), and indicating the Universe is dominated by “dark energy.”...

    [...]

Journal ArticleDOI
TL;DR: In this article, a reprocessed composite of the COBE/DIRBE and IRAS/ISSA maps, with the zodiacal foreground and confirmed point sources removed, is presented.
Abstract: We present a full-sky 100 μm map that is a reprocessed composite of the COBE/DIRBE and IRAS/ISSA maps, with the zodiacal foreground and confirmed point sources removed. Before using the ISSA maps, we remove the remaining artifacts from the IRAS scan pattern. Using the DIRBE 100 and 240 μm data, we have constructed a map of the dust temperature so that the 100 μm map may be converted to a map proportional to dust column density. The dust temperature varies from 17 to 21 K, which is modest but does modify the estimate of the dust column by a factor of 5. The result of these manipulations is a map with DIRBE quality calibration and IRAS resolution. A wealth of filamentary detail is apparent on many different scales at all Galactic latitudes. In high-latitude regions, the dust map correlates well with maps of H I emission, but deviations are coherent in the sky and are especially conspicuous in regions of saturation of H I emission toward denser clouds and of formation of H2 in molecular clouds. In contrast, high-velocity H I clouds are deficient in dust emission, as expected. To generate the full-sky dust maps, we must first remove zodiacal light contamination, as well as a possible cosmic infrared background (CIB). This is done via a regression analysis of the 100 μm DIRBE map against the Leiden-Dwingeloo map of H I emission, with corrections for the zodiacal light via a suitable expansion of the DIRBE 25 μm flux. This procedure removes virtually all traces of the zodiacal foreground. For the 100 μm map no significant CIB is detected. At longer wavelengths, where the zodiacal contamination is weaker, we detect the CIB at surprisingly high flux levels of 32 ± 13 nW m-2 sr-1 at 140 μm and of 17 ± 4 nW m-2 sr-1 at 240 μm (95% confidence). This integrated flux ~2 times that extrapolated from optical galaxies in the Hubble Deep Field. The primary use of these maps is likely to be as a new estimator of Galactic extinction. To calibrate our maps, we assume a standard reddening law and use the colors of elliptical galaxies to measure the reddening per unit flux density of 100 μm emission. We find consistent calibration using the B-R color distribution of a sample of the 106 brightest cluster ellipticals, as well as a sample of 384 ellipticals with B-V and Mg line strength measurements. For the latter sample, we use the correlation of intrinsic B-V versus Mg2 index to tighten the power of the test greatly. We demonstrate that the new maps are twice as accurate as the older Burstein-Heiles reddening estimates in regions of low and moderate reddening. The maps are expected to be significantly more accurate in regions of high reddening. These dust maps will also be useful for estimating millimeter emission that contaminates cosmic microwave background radiation experiments and for estimating soft X-ray absorption. We describe how to access our maps readily for general use.

15,988 citations


"The cluster lensing and supernova s..." refers methods in this paper

  • ...Extinction corrections are derived from the Schlegel et al. (1998) IR dust emission maps, and the resulting Aλ coefficients, in AB mag per unit E(B-V), are given in Table 5....

    [...]

Journal ArticleDOI
TL;DR: In this paper, the authors presented a reprocessed composite of the COBE/DIRBE and IRAS/ISSA maps, with the zodiacal foreground and confirmed point sources removed.
Abstract: We present a full sky 100 micron map that is a reprocessed composite of the COBE/DIRBE and IRAS/ISSA maps, with the zodiacal foreground and confirmed point sources removed. Before using the ISSA maps, we remove the remaining artifacts from the IRAS scan pattern. Using the DIRBE 100 micron and 240 micron data, we have constructed a map of the dust temperature, so that the 100 micron map can be converted to a map proportional to dust column density. The result of these manipulations is a map with DIRBE-quality calibration and IRAS resolution. To generate the full sky dust maps, we must first remove zodiacal light contamination as well as a possible cosmic infrared background (CIB). This is done via a regression analysis of the 100 micron DIRBE map against the Leiden- Dwingeloo map of H_I emission, with corrections for the zodiacal light via a suitable expansion of the DIRBE 25 micron flux. For the 100 micron map, no significant CIB is detected. In the 140 micron and 240 micron maps, where the zodiacal contamination is weaker, we detect the CIB at surprisingly high flux levels of 32 \pm 13 nW/m^2/sr at 140 micron, and 17 \pm 4 nW/m^2/sr at 240 micron (95% confidence). This integrated flux is ~2 times that extrapolated from optical galaxies in the Hubble Deep Field. The primary use of these maps is likely to be as a new estimator of Galactic extinction. We demonstrate that the new maps are twice as accurate as the older Burstein-Heiles estimates in regions of low and moderate reddening. These dust maps will also be useful for estimating millimeter emission that contaminates CMBR experiments and for estimating soft X-ray absorption.

14,295 citations

Journal ArticleDOI
TL;DR: In this paper, the authors present observations of 10 type Ia supernovae (SNe Ia) between 0.16 0 and 4.0 sigma confidence levels, for two fitting methods respectively.
Abstract: We present observations of 10 type Ia supernovae (SNe Ia) between 0.16 0) and a current acceleration of the expansion (i.e., q_0 0, the spectroscopically confirmed SNe Ia are consistent with q_0 0 at the 3.0 sigma and 4.0 sigma confidence levels, for two fitting methods respectively. Fixing a ``minimal'' mass density, Omega_M=0.2, results in the weakest detection, Omega_Lambda>0 at the 3.0 sigma confidence level. For a flat-Universe prior (Omega_M+Omega_Lambda=1), the spectroscopically confirmed SNe Ia require Omega_Lambda >0 at 7 sigma and 9 sigma level for the two fitting methods. A Universe closed by ordinary matter (i.e., Omega_M=1) is ruled out at the 7 sigma to 8 sigma level. We estimate the size of systematic errors, including evolution, extinction, sample selection bias, local flows, gravitational lensing, and sample contamination. Presently, none of these effects reconciles the data with Omega_Lambda=0 and q_0 > 0.

14,295 citations

Related Papers (5)
Anton M. Koekemoer, Sandra M. Faber, Henry C. Ferguson, Norman A. Grogin, Dale D. Kocevski, David C. Koo, Kamson Lai, Jennifer M. Lotz, Ray A. Lucas, Elizabeth J. McGrath, Sara Ogaz, Abhijith Rajan, Adam G. Riess, S. Rodney, L. G. Strolger, Stefano Casertano, Marco Castellano, Tomas Dahlen, Mark Dickinson, Timothy Dolch, Adriano Fontana, Mauro Giavalisco, Andrea Grazian, Yicheng Guo, Nimish P. Hathi, Kuang-Han Huang, Kuang-Han Huang, Arjen van der Wel, Hao Jing Yan, Viviana Acquaviva, David M. Alexander, Omar Almaini, Matthew L. N. Ashby, Marco Barden, Eric F. Bell, Frédéric Bournaud, Thomas M. Brown, Karina Caputi, Paolo Cassata, Peter Challis, Ranga-Ram Chary, Edmond Cheung, Michele Cirasuolo, Christopher J. Conselice, Asantha Cooray, Darren J. Croton, Emanuele Daddi, Romeel Davé, Duilia F. de Mello, Loic de Ravel, Avishai Dekel, Jennifer L. Donley, James Dunlop, Aaron A. Dutton, David Elbaz, Giovanni Fazio, Alexei V. Filippenko, Steven L. Finkelstein, Chris Frazer, Jonathan P. Gardner, Peter M. Garnavich, Eric Gawiser, Ruth Gruetzbauch, Will G. Hartley, B. Haussler, Jessica Herrington, Philip F. Hopkins, J.-S. Huang, Saurabh Jha, Andrew Johnson, Jeyhan S. Kartaltepe, Ali Ahmad Khostovan, Robert P. Kirshner, Caterina Lani, Kyoung-Soo Lee, Weidong Li, Piero Madau, Patrick J. McCarthy, Daniel H. McIntosh, Ross J. McLure, Conor McPartland, Bahram Mobasher, Heidi Moreira, Alice Mortlock, Leonidas A. Moustakas, Mark Mozena, Kirpal Nandra, Jeffrey A. Newman, Jennifer L. Nielsen, Sami Niemi, Kai G. Noeske, Casey Papovich, Laura Pentericci, Alexandra Pope, Joel R. Primack, Swara Ravindranath, Naveen A. Reddy, Alvio Renzini, Hans Walter Rix, Aday R. Robaina, David J. Rosario, Piero Rosati, S. Salimbeni, Claudia Scarlata, Brian Siana, Luc Simard, Joseph Smidt, D. Snyder, Rachel S. Somerville, Hyron Spinrad, Amber N. Straughn, Olivia Telford, Harry I. Teplitz, Jonathan R. Trump, Carlos J. Vargas, Carolin Villforth, C. Wagner, P. Wandro, Risa H. Wechsler, Benjamin J. Weiner, Tommy Wiklind, Vivienne Wild, Grant W. Wilson, Stijn Wuyts, Min S. Yun 
Frequently Asked Questions (16)
Q1. What can be used to help distinguish between highly reddened objects and truly distant objects?

lens models can be used to help discriminate between highly reddened objects and truly distant, high-redshift objects, as the projected positions of the lensed images are strong functions of the source redshifts. 

cluster elongation along the line of sight (a potential bias in concentration measurements) can be measured by the combination of lensing and X-ray analysis (Morandi et al. 2011; Newman et al. 2011). 

Clusters of galaxies, by virtue of their position at the high end of the cosmic mass power spectrum, provide a powerful way to constrain the frequency of high amplitude perturbations in the primordial density field. 

In order to continue to achieve the same depths, the authors would require exposure times in the redder filters to be increased by 80%, adding about four orbits of integration time to each cluster. 

The ability to reprogram the later of the two orientations to follow-up an SN detected early in a cluster observing sequence sets the upper limit on the angular offset between the two orientations—both orientations must be accessible during the entire cluster sequence. 

A key ingredient of such cluster-based cosmological tests is the mass distribution of clusters, both on (sub) Mpc scales and across the range of populations. 

CLASH may detect dozens of relatively bright (magnified to m < 26.7 AB) z > 7 galaxies, including some bright enough for spectroscopic follow-up. 

The reason, in part, is a lack of photons: at redshift z = 8 and 10, an L∗ galaxy would have an apparent magnitude in the first NIR detection band of 28.2 and 29.6, respectively. 

Subaru images are especially desirable for WL studies as they enable excellent galaxy shape measurements to be performed over a wide area. 

The authors prune these detections from their catalog by rejecting any object with only a single 5σ detection in one UVIS/ACS filter, as measured by SExtractor. 

Figure 4 shows how variations in the DE equation of state or an evolution in the white dwarf C/O ratio can change the SN Ia distance modulus as a function of redshift. 

Strong-lensing analyses of 20 of these based on HST imaging (mostly single-band “snapshots”) were presented by Richard et al. (2010). 

The inclusion of NUV photometry, for example, resolves one of the most common photo-z degeneracies between the Balmer break in z ∼ 0.2 galaxies and the Lyman break in z ∼ 3 galaxies (Rafelski et al. 2009). 

The combination of the X-ray and millimeter-wave observations allows the mass scaling relations to be accurately calibrated for use in cosmological surveys (e.g., Okabe et al. 2010b). 

This process likely happens through “dry” (dissipationless) merging, since such massive galaxies are observed to have old stellar populations locally. 

These cluster counts tell us much about cosmological parameters through their impact on both the volume and the growth of perturbations.