ArticlePDF Available

Abstract and Figures

In this paper we use deep CFHT and SUBARU $uBVRIz$ archival images of the Abell 383 cluster (z=0.187) to estimate its mass by weak lensing. To this end, we first use simulated images to check the accuracy provided by our KSB pipeline. Such simulations include both the STEP 1 and 2 simulations, and more realistic simulations of the distortion of galaxy shapes by a cluster with a Navarro-Frenk-White (NFW) profile. From such simulations we estimate the effect of noise on shear measurement and derive the correction terms. The R-band image is used to derive the mass by fitting the observed tangential shear profile with a NFW mass profile. Photometric redshifts are computed from the uBVRIz catalogs. Different methods for the foreground/background galaxy selection are implemented, namely selection by magnitude, color and photometric redshifts, and results are compared. In particular, we developed a semi-automatic algorithm to select the foreground galaxies in the color-color diagram, based on observed colors. Using color selection or photometric redshifts improves the correction of dilution from foreground galaxies: this leads to higher signals in the inner parts of the cluster. We obtain a cluster mass that is ~ 20% higher than previous estimates, and is more consistent the mass expected from X--ray data. The R-band luminosity function of the cluster is finally computed.
Content may be subject to copyright.
arXiv:1102.1837v1 [astro-ph.CO] 9 Feb 2011
Astronomy & Astrophysics
manuscript no. a383 c
ESO 2011
February 10, 2011
A weak lensing analysis of the Abell 383 cluster.
Zhuoyi Huang1, Mario Radovich2, Aniello Grado1, Emanuella Puddu1, Anna Romano3,
Luca Limatola1, and Liping Fu4,1
1INAF - Osservatorio Astronomico di Capodimonte, via Moiariello 16, I-80131, Napoli, Italy
2INAF - Osservatorio Astronomico di Padova, vicolo dell’Osservatorio 5, I-35122, Padova, Italy
3INAF - Osservatorio Astronomico di Roma, Monte Porzio, I-00185, Roma, Italy
4Key Lab for Astrophysics, Shanghai Normal University, 100 Guilin Road, 200234, Shanghai, China
received; accepted
ABSTRACT
Aims.
In this paper we use deep CFHT and SUBARU uBVRIz archival images of the Abell 383 cluster (z=0.187) to estimate its mass
by weak lensing.
Methods.
To this end, we first use simulated images to check the accuracy provided by our KSB pipeline. Such simulations include
both the STEP 1 and 2 simulations, and more realistic simulations of the distortion of galaxy shapes by a cluster with a Navarro-
Frenk-White (NFW) profile. From such simulations we estimate the eect of noise on shear measurement and derive the correction
terms. The Rband image is used to derive the mass by fitting the observed tangential shear profile with a NFW mass profile.
Photometric redshifts are computed from the uBV RIz catalogs. Dierent methods for the foreground/background galaxy selection are
implemented, namely selection by magnitude, color and photometric redshifts, andresults are compared. In particular, we developed
a semi-automatic algorithm to select the foreground galaxies in the color-color diagram, based on observed colors.
Results.
Using color selection or photometric redshifts improves the correction of dilution from foreground galaxies: this leads to
higher signals in the inner parts of the cluster. We obtain a cluster mass Mvir =7.5+2.7
1.9×1014 M: such value is 20% higher than
previous estimates, and ismore consistent the mass expected from X–ray data. The R-band luminosity function of the cluster is finally
computed, giving a total luminosity Ltot =(2.14 ±0.5) ×1012 Land a mass to luminosity ratio M/L300M/L.
Key words. Galaxies: clusters: individual: Abell 383 – Galaxies: fundamental parameters – Cosmology: dark matter
1. Introduction
Weak gravitational lensing is a unique technique that allows to
probe the distribution of dark matter in the Universe. It mea-
sures the very small distortions in the shapes of faint back-
ground galaxies, due to foreground mass structures. This tech-
nique requires a very accurate measurement of the shape pa-
rameters as well as the removal of the systematic eects af-
fecting them. In addition, galaxies to be used in the weak lens-
ing analysis must be carefully selected so that they do not in-
clude a significant fraction of unlensed sources, with redshift
smaller than that of the lens. This would introduce a dilution
and therefore an underestimated signal, in particular towards the
cluster center (see Broadhurst et al. 2005): such eect may be
the reason of the under-prediction by weak lensing of the ob-
served Einstein radius, by a factor of 2.5 (Smith et al. 2001;
Bardeau et al. 2005). The optimal case would happen if photo-
metric redshifts were available. Even if for weak lensing high
accuracy in their estimate are not required for individual galax-
Based on: data collected at Subaru Telescope (University of Tokyo)
and obtained from the SMOKA, which is operated by the Astronomy
Data Center, National Astronomical Observatory of Japan; observa-
tions obtained with MegaPrime/MegaCam, a joint project of CFHT and
CEA/DAPNIA, at the Canada-France-Hawaii Telescope (CFHT) which
is operated by the National Research Council (NRC) of Canada, the
Institute National des Sciences de l’Univers of the Centre National de
la Recherche Scientifique of France, and the University of Hawaii. This
work is based in part on data products produced at TERAPIX and the
Canadian Astronomy Data Centre as part of the Canada-France-Hawaii
Telescope Legacy Survey, a collaborative project of NRC and CNRS.
ies, on average we need at least σz/(1 +z)<0.1: this implies
having observations in several bands, spanning a good wave-
length range. If few bands are available, an uncontaminated
background sample can be obtained by selecting only galaxies
redder than the cluster red sequence (Broadhurst et al. 2005).
However, such method often does not allow to get a number
density of background sources high enough to allow an accu-
rate weak lensing measure. Including galaxies bluer than the red
sequence (Okabe et al. 2010) requires a careful selection of the
color oset, as bluer galaxies can be still contaminated by late–
types members of the cluster. Finally, if more than two bands
are available, Medezinski et al. (2010) discussed how to identify
cluster members and the foreground population as overdensities
in the color-color space.
In this paper we exploit deep uBVRIz images of the clus-
ter Abell 383, taken with the MEGACAM and SUPRIME cam-
era mounted on the 3.6m CFHT and 8m SUBARU telescopes
respectively, and publicly available. The mass of the cluster is
derived by weak lensing, and values obtained by dierent selec-
tion methods are compared. The properties of the cluster are re-
viewed in Sect.2. Data reduction is discussed in Sect.3. In Sect. 4
we describe the algorithm used for the shape measurement, and
some improvements for the removal of biases. The accuracy in
the mass estimate is derived by comparison with simulations. In
Sect. 5, we first summarize the dierent methods for the selec-
tion of the background galaxies from which the lensing signal
is measured. Such methods are applied to the case of Abell 383,
and the masses derived in such way are then compared. Finally,
in Sect. 6 we compare the mass derived in this paper with lit-
1
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
erature values, both by X–rays and weak lensing; a comparison
is also done with the mass expected for the R–band luminosity,
derived from the luminosity function of Abell 383.
A standard cosmology was adopted in this paper: Λ=0.7,
M=0.3, H0=70 km s1Mpc1, giving a scale of 2.92
kpc/arcsec at the redshift of Abell 383.
2. Abell 383
Abell 383 is an apparently well relaxed cluster of galaxies of
richness class 2 and of Bautz-Morgan type II-III (Abell et al.
1989), located at z =0.187 (Fetisova et al. 1993). It is domi-
nated by the central cD galaxy, a blue-core emission-line Bright
Cluster Galaxy (BCG), that is aligned with the X-ray peak
Smith et al. (2001). Abell 383 is one of the clusters of the
XBACs sample (X-ray-Brightest Abell-type Clusters), observed
in the ROSAT All-Sky Survey (RASS; Voges 1992): its X-ray
luminosity is 8.03 ×1044 erg sec1in the 0.1-2.4 keV band and
its X-ray temperature is 7.5 keV (Ebeling et al. 1996). A small
core radius, a steep surface brightness profile and an inverted de-
projected temperature profile, show evidence of the presence of
a cooling flow, as supported by the strong emission lines in the
optical spectra of its BCG (Rizza et al. 1998).
An extensive study of this cluster was carried out by
Smith et al. (2001), in which lensing and X-ray properties were
analyzed on deep optical HST images and ROSAT HRI data,
respectively. A complex system of strong lensed features (a gi-
ant arc, two radial arcs in the center and numerous arclets) were
identified in its HST images, some of which are also visible in
the deep SUBARU data used here.
3. Data retrieval and reduction
The cluster Abell 383 was observed with the SUPRIME cam-
era mounted at the 8m SUBARU telescope: SUPRIME is
a ten CCDs mosaic, with a 34 ×27 arcmin2field of view
(Miyazaki et al. 2002). Data are publicly available in the BVRIz
filters, with total exposure times of 7800s (R), 6000s (B,V),
3600s (I) and 1500s (z); they were retrieved using the SMOKA1
Science Archive facility. Data were collected from seven dier-
ent runs, amounting to 55GB. Details about the observation
nights and exposure times for each band are given in Table 1. The
data reduction was done using the VST–Tube imaging pipeline,
specifically developed for the VLT Survey Telescope (VST,
Capaccioli et al. 2005) but adaptable to other existing or future
multi-CCD cameras (Grado et al. 2011) .
The field of Abell 383 was also observed in the uband with
the MEGACAM camera attached on the Canada-France Hawaii
Telescope (CFHT), with a total exposure time of 10541s. The
preprocessed images were retrieved from the CADC archive.
The basic reduction steps were performed for each frame,
namely overscan correction, flat fielding, correction of the ge-
ometric distortion due to the optics, and sky background sub-
traction. It is worth mentioning that to improve the photometric
accuracy, a sky superflat was used.
Geometric distortions were first removed from each expo-
sure, using the Scamp tool 2and taking the USNO-B1 as the as-
trometric reference catalog. The internal accuracy provided by
Scamp, as measured by positions of the same sources in dier-
ent exposures, is 0.05 arcsec. Dierent exposures were then
1http://smoka.nao.ac.jp/
2Stuff, Skymaker, SWarp and SExtractor are part of the
Astromatic software developed by E. Bertin, see www.astromatic.net
17 18 19 20
−0.4 0.0 0.2 0.4
R
zphot zSDSS
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
0.0 0.2 0.4 0.6
zSDSS
zphot
Fig.1. Density plots comparing the photometric redshifts in the
Abell 383 field available from the SDSS, and those here com-
puted from the uBVRIz photometry.
16 17 18 19 20 21 22 23
−0.2 0.0 0.2 0.4 0.6 0.8 1.0
R
V−R
Fig.2. VRvs Rcolor plot: red and blue points are galaxies at
zphot =0.187±0.1 and classified as early and late-type, respec-
tively; dashed lines shows the ±1σlevels.
stacked together using SWarp. The coaddition was done in such
a way that all the images had the same scale and size.
The photometric calibration of the BVRI bands was per-
formed using the standard Stetson fields, observed in the same
nights of the data. In the case of the zband, we used a point-
ing also covered by the Stripe 82 scans in the Sloan Digital Sky
Survey (SDSS); the SDSS photometry of sources identified as
point-like was used to derive the zero point in the SUBARU im-
age.In the case of the MEGACAM-CFHT uband images, re-
duced images and photometric zero points are already available
from the CADC public archive, hence only astrometric calibra-
tion and stacking were required.
Table 2 summarizes the photometric properties of the fi-
nal coadded images (average FWHMs and limiting magnitudes
for point-like sources) for each band. All magnitudes were
converted to the AB system; magnitudes of sources classified
2
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
Table 1. Summary of observations done with the MEGACAM
(u) and SUPRIME (BVRIz) cameras, used in this paper.
Date Band Exp. Time Total
23 Dec 2003 u8381s
21 Jan 2004 u2160s 10541s
09 Sept 2002 B2400s
08 Jan 2008 B2400s
09 Jan 2008 B1200s 6000s
02 Oct 2002 V4320s
01 Oct 2005 V1800s 6120s
12 Nov 2007 Rc2400s
08 Jan 2008 Rc5400s 7800s
02 Oct 2002 Ic3600s 3600s
10 Sept 2002 z1500s 1500s
Table 2. Photometric properties of the coadded images; FWHM
(arcsec) and limiting magnitudes (in the AB system) were com-
puted for point-like sources. The signal to noise (SNR) is defined
as SNR=FLUX AUTO/FLUXERR AUTO.
Band FWHM mag (SNR=5) mag (SNR=10)
u1.3 26.1 25.3
B0.99 27.0 26.2
V0.95 26.6 25.7
Rc0.82 27.1 26.1
Ic0.97 25.0 24.2
z0.74 24.7 23.9
as galaxies were corrected for Galactic extinction using the
Schlegel maps (Schlegel et al. 1998).
The weak lensing analysis was done on the Rband im-
age. The masking of reflection haloes and diraction spikes
near bright stars are performed by ExAM, a code developed
for this purpose. In short, ExAM takes the SExtractor cata-
log as input, locates the stellar locus in the size-magnitude di-
agram (see Sect. 4.1), picks out stars with spike-like features
from the isophotal shape analysis, and outputs mask region file
that may be visualized in the DS9 software, and finally cre-
ates a mask image in FITS format. The reflection haloes are
masked by estimating the background contrast near the bright
stars, whose positions are obtained from the USNO–B1. The ef-
fective area available after removal of regions masked in such
way was 801 arcmin2. Catalogs for the other bands were ex-
tracted using SExtractor in dual–mode, where the Rband im-
age was used as the detection image.
Photometric redshifts were computed from uBVRIz pho-
tometry, using the ZEBRA code (Feldmann et al. 2006). This
software allows to define 6 basic templates (elliptical, Sbc,
Sbd, irregular and two starburst SEDs), and to compute log-
interpolations between each pair of adjacent templates. We first
applied to the BVz magnitudes the oset derived within the
COSMOS survey by Capak et al. (2007), that is +0.19 (B),
+0.04 (V), -0.04 (z). We then convolved the stellar spectra from
the Pickles’ library (Pickles 1985) with the transmission curves
used for each filter, and derived the osets for the other filters,
that is: 0.0 (u), 0.0 (R), +0.05 (I). Fig. 3 shows the comparison
−0.5 0.5 1.5
2.5 2.0 1.5 1.0 0.5 0.0 −0.5
(BV)AB
(uB)AB
*
*
***
**
*
**
*
*
*
*
***
**
**
*
***
*
*
**
*
*
*
*
*
**
*
**
*
**
**
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
**
*
*
*
*
*
*
*
**
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
**
*
*
*
*
*
****
*
*
*
***
**
*
**
*
*
*
*
***
**
**
*
***
*
*
**
*
*
*
*
*
**
*
**
*
**
**
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
**
*
*
*
*
*
*
*
**
*
*
*
*
*
*
*
*
*
*
*
*
**
*
*
*
*
*
*
*
*
*
*
**
*
*
*
*
*
****
*
**
*
*
*
***
**
*
**
*
*
*
***
**
*
−1.0 0.0 1.0 2.0
−0.5 0.0 0.5 1.0 1.5 2.0
(BV)AB
(RI)AB
*
*
*
**
**
*
**
*
*
*
*
***
*
*
**
*
*
**
*
*
**
*
*
*
*
*
**
*
**
*
**
*
*
**
*
**
*
*
**
*
**
*
*
*
**
*
*
**
*
*
**
**
*
***
*
*
*
*
*
*
*
***
**
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
*
**
**
******
*
*
*
**
**
*
**
*
*
*
*
***
*
*
**
*
*
**
*
*
**
*
*
*
*
*
**
*
**
*
**
*
*
**
*
**
*
*
**
*
**
*
*
*
**
*
*
**
*
*
**
**
*
***
*
*
*
*
*
*
*
**
*
**
*
*
*
*
*
*
*
*
*
*
**
*
*
*
*
*
*
**
**
******
*
*
**
***
*
***
*
*
**
***
*
***
−0.5 0.0 0.5 1.0
−0.2 0.0 0.2 0.4 0.6
(VR)AB
(Iz)AB
*
*
*
**
**
*
**
**
*
*
*
**
*
*
**
*
*
**
*
*
*
*
*
*
**
*
*
*
*
**
*
**
*
*
*
*
*
**
*
*
**
*
**
*
*
*
**
*
*
*
*
*
*
**
**
*
*
**
*
*
*
*
*
*
*
**
*
*
*
*
*
*
*
**
*
**
*
*
*
*
*
*
*
**
**
**
****
*
*
*
**
**
*
**
**
*
*
*
**
*
*
**
*
*
**
*
*
*
*
*
*
**
*
*
*
*
**
*
**
*
*
*
*
*
**
*
*
**
*
**
*
*
*
**
*
*
*
*
*
*
**
**
*
*
**
*
*
*
*
*
*
*
**
**
*
*
*
*
*
**
**
**
*
*
*
*
*
*
**
**
**
****
**
*
*
**
*
*
***
**
*
*
**
*
*
***
Fig.3. Observed (red dots) and model colors (black dots) for
stars, after the osets given in the text were applied. Model col-
ors were derived convolving the Pickles’ library of stellar spectra
with the filter transmission curves.
1000 2000 3000 4000
Redshift
N
0.1 0.3 0.5 0.7 0.9 1.1 1.3 1.5 1.7 1.9 2.1 2.3 2.5 2.7 2.9
Fig.4. Distribution of the photometric redshifts computed from
the uBVRIz data for R<25 mag.
of the model colors with those derived for the stars in our cata-
logs, after the above osets were applied. We also verified that
the osets derived in this way are consistent with those obtained
by running ZEBRA in the so-called photometry-check mode,
that allows to compute magnitude osets minimizing the aver-
age residuals of observed versus template magnitudes.
We then removed from the catalog those galaxies with a pho-
tometric redshift zph >3 and σz/(1 +z)>0.1, where σzwas
derived from the 68%–level errors computed in ZEBRA. The
distribution of the so-obtained photometric redshifts is displayed
in Fig. 4. The accuracy of the so obtained photometric redshifts
was derived from the comparison with the SDSS photometric
redshifts of galaxies in the SDSS (DR7), whose rms error is
0.025 for r<20 mag. We derived (Fig. 1) a systematic o-
set z/(1 +z)=0.003 and an rms error σz/(1 +z)=0.07.
As a further check, we extracted from our catalog those galax-
ies classified by ZEBRA as early-type, with R<23 mag and
|zphot 0.187|=0.1. As expected, such galaxies define a red se-
3
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
quence (Fig.2): this was fitted as VR=a+bR, with a=0.5,
b=6×103.
4. Shape measurement
Ellipticities of galaxies were estimated using the KSB approach
(Luppino & Kaiser 1997): even if such algorithm does not allow
to achieve accurate measurements of very low shear signals, γ.
103, it is nevertheless adequate in the case of weak lensing by
clusters, as discussed e.g. by Gill et al. (2009) and Romano et al.
(2010).
In our KSB implementation, the SExtractor software was
modified to compute all the relevant quantities, namely the raw
ellipticity e, the smear polarizability Psm, and the shear polariz-
ability Psh. The centers of detected sources were measured using
the windowed centroids in SExtractor.
The KSB approach assumes that the PSF can be described
as the sum of an isotropic component (simulating the eect of
seeing) and an anisotropic part. The correction of the observed
ellipticity eobs for the anisotropic part is computed as:
eaniso =eobs Psmp,(1)
where (starred terms indicate that they are derived from mea-
surement of stars): p=e
obs/Psm.(2)
The intrinsic ellipticity eof a galaxy, and the reduced shear, g=
γ/(1 κ), are then related by:
eaniso =e+Pγg(3)
The term Pγ, introduced by Luppino & Kaiser (1997) as the pre–
seeing shear polarizability, describes the eect of seeing and is
defined to be:
Pγ=Psh Psm Psh
PsmPsh Psmq.(4)
The final output of the pipeline is the quantity eiso =
eaniso/Pγ, from which the average reduced shear, hgi=heisoi,
provided that the average intrinsic ellipticity vanishes, hei=0.
The ellipticity calculation is done by using a window func-
tion in order to suppress the outer, noisy part of a galaxy: the
function is usually chosen to be Gaussian with size θ. The size
of the window function is commonly taken as the radius con-
taining 50% of the total flux of the galaxy (as given by e.g. the
FLUX RADIUS parameter in SExt ractor). In our case, we pro-
ceed as follows. We define a set of bins with θvarying between 2
and 10 pixels (sources with smaller and larger sizes are rejected
in our analysis), and a step of 0.5 pixel. For each bin we compute
eobs,Psh and Psm, and the ellipticity signal to noise ratio defined
by Eq. 16 in Erben et al. (2001):
SNe(θ)=RI(θ)W(|θ|)d2θ
σsky qRW2(|θ|)d2θ
.(5)
The optimal size of the window function, θmax, is then defined as
the value that maximizes SNe. Fig.5 shows the typical trend of
SNe, which was normalized for display purposes, as a function
of θ. It can be seen (Fig. 6) that, on average, there is a con-
stant oset between θmax and FLUX RADIUS. Below SNe 5,
FLUX RADIUS starts to decrease: this provides an estimate of
limit on SNe, below which the shape measurement is not mean-
ingful any more.
2 3 4 5 6 7
0.0 0.2 0.4 0.6 0.8 1.0
θ [pixels]
Normalized SNe
Fig.5. SNe is displayed as a function of the window function
size, θ, used to measure ellipticities. For display purposes, galax-
ies were selected to have the same value of θmax, and SNe was
normalized so that min(SNe)=0,max(SNe)=1. The vertical lines
indicate the average (solid) and standard deviation (dashed) of
FLUX RADIUS for the same galaxies.
0 10 20 30 40 50
0.0 0.1 0.2 0.3 0.4 0.5
SNe
(FLUX_RADIUSθ) [pixels]
Fig.6. Running median of FLUX RADIUS-θmax as a function
of SNe. The vertical line shows the limit chosen for the selection
of background galaxies.
The terms pand q, derived from stars, must be evaluated
at each galaxy position: this is usually done fitting them by a
polynomial (see e.g. Radovich et al. 2008), whose order must be
chosen to fit the observed trend, without overfitting. The usage
of the window function introduces a calibration factor, which
is compensated by the Pγterm. This implies that stellar terms
must be computed and fitted with the same value of θused for
each galaxy (Hoekstra et al. 1998). An alternative approach, not
based on a constant (and somehow arbitrary) order polynomial,
is given e.g. by Generalized Additive Models: we found that
the implementation in R (function GAM in the mgcv library)
provides good results. Fig. 7 shows fitting and residuals of the
anisotropic PSF component: from the comparison between the
results obtained with polynomial and GAM fitting we see that
4
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
in the latter case we obtain lower residuals, in particular in the
borders of the image. To quantify the improvementcompared to
the usage of the polynomial, we obtain eaniso,1=(1±5)×104,
eaniso,2=(2±9) ×104with a polynomial of order 3,
eaniso,1=(2 ±4) ×104,eaniso,2=(1 ±7) ×104with the
GAM algorithm. The values of the fitted terms pand qat the po-
sitions of the galaxies are predicted by GAM, that also provides
an estimate of the standard errors of the predictions, pand q.
From error propagation, the uncertainty on eiso was computed
as:
e2
iso =(eaniso/Pγ)2+(eaniso(Pγ)2Pγ)2,(6)
where (eaniso)2=(eobs)2+(Psmp)2and (Pγ)2=(Psmq)2;
uncertainties on the measured values of Psm and Psh were not
considered.
For each galaxy, a weight is defined as:
w=1
e02+ ∆e2
iso
,(7)
where e00.3 is the typical intrinsic rms of galaxy elliptici-
ties.
4.1. Star-galaxy classification
Stars and galaxies were separated in the magnitude
(MAG AUTO) vs. size plot. Instead of using e.g.
FLUX RADIUS as the estimator of size, we used the quantity
δ=MU MAX-MAG AUTO, where MU MAX is the peak
surface brightness above background. Saturated stars were
found in the locus of sources with constant MU MAX; in the
δvs. MAG AUTO plot, stars are identified as sources in the
vertical branch. Sources with δlower than stars were classified
as spurious detections. In addition, we rejected those sources for
which δis 2σlarger than the median value. This is to exclude
from the sample of stars used to compute the PSF correction
terms, those sources for which the shape measurement may be
wrong due to close blended sources, noise, etc.
We further excluded those galaxies with w<1 or SNe <5,
for which the ellipticity measurement is not meaningful.
4.2. Error estimate in shear and mass measurement
In the case of faint galaxies used for the weak lensing analy-
sis, the ellipticity is underestimated due to noise. Such eect is
not included in the Pγterm, which can be only computed on
stars with high signal to noise ratio. Schrabback et al. (2010)
proposed the following parametrization for such bias, as a func-
tion of the signal to noise:
m=ekem
em=a(SNe)b,(8)
where emand ekare the ellipticities before and after the correc-
tion respectively.Such parameters were derived using the STEP1
(Heymans et al. 2006) and STEP2 (Massey et al. 2007) simula-
tions, where both PSF and shear are constant for each simulated
image. We obtain a=0.1 and b=0.45, which corresponds to
a bias mchanging from 5% for SNe =5 to <2% for SNe >50.
After such correction was applied, we computed again the aver-
age shear from the STEP1 and STEP2 simulations, and obtained
a typical bias of 3% for SNe =5.
We then estimated the accuracy on the mass that can be ob-
tained from an image with the same noise and depth as in the
0 2000 6000 10000
0 2000 6000
0 2000 6000 10000
0 2000 6000
0 2000 6000 10000
0 2000 6000
−0.10 0.00 0.05 0.10
−0.10 0.00 0.05 0.10
0 2000 6000 10000
0 2000 6000
−0.10 0.00 0.05 0.10
−0.10 0.00 0.05 0.10
Fig.7. PSF anisotropy correction derived with the GAM algo-
rithm: the first three panels show the ellipticity pattern (mea-
sured, fitted and residuals; X and Y are in pixels). The scale is
displayed by the arrows in the upper right part of each panel
(e=0.05). In the next panel, black dots are the measured val-
ues, green dots are after the correction; values rejected during
fitting are marked in red. The last row shows for comparison the
corrected ellipticities obtained using for the fit a polynomial of
order 3.
Rband SUBARU image. To this end we dropped the assump-
tion on constant shear, and produced more realistic simulations:
the eect on galaxy shapes by weak lensing from a galaxy cluster
was produced using the SHUFF code, that will be described in a
separate paper (Huang et al., in preparation). To summarize, the
code takes as input a catalog of galaxies produced by the Stuff
tool; it computes the shear produced by a standard mass profile
(e.g. Navarro-Frenk-White, NFW hereafter) and applies it to the
ellipticities of the galaxies behind the cluster. Such catalog is
then used in the SkyMaker software, configured with the tele-
scope parameters suitable for the SUBARU telescope and with
the exposure time of the Rband image, producing a simulated
image; the background rms of such image was set to be as close
as possible to that of the real image. We considered for the lens
5
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
Table 3. Masses derived by simulations with NFW model fitting
(M3D) and aperture densitometry (M2D). Input masses used for
each simulation are given in the first column (Min).
Min M3D M2D/M3D
1014 M1014 M
0.316 0.31 ±0.13 1.3±0.9
1.00 1.01 ±0.24 1.4±0.4
3.16 3.12 ±0.36 1.4±0.3
10.0 10.1±0.5 1.3±0.1
a range of masses at log Mvir/M=13.5,14,14.5,15.0, and a
NFW mass profile with cvir =6. Each simulation was repeated
50 times for each mass value, randomly changing the morphol-
ogy, position and redshift of the galaxies.
For each of these images, we run our lensing pipeline with
the same configuration used for the real data. The density of
background galaxies used for the lensing analysis was 20
gals arcmin2. The fit of the mass was done as described in
Radovich et al. (2008) and Romano et al. (2010): the expres-
sions for the radial dependence of tangential shear γTderived
by Bartelmann (1996) and Wright & Brainerd (2000) were used,
and the NFW parameters (Mvir,cvir) were derived using a max-
imum likelihood approach. In addition, the 2D projected mass
can be derived in a non-parametric way by aperture densitome-
try, where the mass profile of the cluster is computed by the ζ
statistics (Fahlman et al. 1994; Clowe et al. 1998):
ζ(θ1)=¯κ(θθ1)¯κ(θ2< θ θout)=2Zθ2
θ1
hγTidlnθ(9)
+2
1(θ2out)2Zθout
θ2
hγTidlnθ.
The mass is estimated as Map(θ1)=πθ2
1ζ(θ1)Σcrit, and θout is
chosen so that ¯κ(θ2, θout)0.
The average errors on mass estimate obtained in such way
are displayed in Table 3, showing that masses can be estimated
within an uncertainty of <20% for M1014M. Such accuracy
only includes the contribute due to shape measurement and mass
fitting method, but it does not include the uncertainty due to the
selection of the lensed galaxies.
Finally, the masses derived by aperture densitometry are
1.3 higher than those obtained by mass fitting: this is in agree-
ment with Okabe et al. (2010), who find M2D/M3D =1.34 for
virial overdensity.
5. Results
One of the most critical source of systematic errors, which can
lead to an underestimation of the true WL signal, is dilution of
the distortion due to the contamination of the background galaxy
catalog by unlensed foreground and cluster member galaxies
(see e.g. Broadhurst et al. 2005). The dilution eect increases as
the cluster-centric distance decreases because the number den-
sity of cluster galaxies that contaminate the faint galaxy catalog
is expected to roughly follow the underlying density profile of
the cluster. Thus, correcting for the dilution eect is important
to obtain unbiased, accurate constraints on the cluster parameters
and mass profile.
As discussed by Broadhurst et al. (2005); Okabe et al.
(2010); Oguri et al. (2010), the selection of background galaxies
to be used for the weak lensing analysis can be done taking those
galaxies redder than the cluster red sequence. However, such
selection produces a low number density (10 galaxies/arcmin2
in our case), and correspondingly high uncertainties in the de-
rived parameters. In the following, we compare the results ob-
tained by dierent methods. We first assumed that no informa-
tion on the redshift is available, and that photometry from only
one band is available (magnitude cut), or from more than two
bands (color selection). Finally, we included in our analysis the
photometric redshifts. The density of background galaxies is
25-30 galaxies/arcmin2, see Table 4.
In order to derive the mass, we need to know the critical
surface density:
Σcrit =c2
4πGDs
DlDls =c2
4πG1
Dlβ,(10)
Dls,Ds, and Dlbeing the angular distances between lens and
source, observer and source, and observer and lens respectively.
This quantity should be computed for each lensed galaxy. As
the reliability of photometric redshifts for the faint background
galaxies is not well known, we prefer to adopt the single sheet
approximation, where all background galaxies are assumed to
lie at the same redshift, defined as β(zs)=hβ(z)i. In the case
of the selection based only on magnitude or colors, such value
was derived from the COSMOS catalog of photometric redshifts
(Capak et al. 2007), to which the same cuts used for the Abell
383 catalog are applied. Later on, we computed β(zs) from the
photometric redshifts themselves, and compared such two val-
ues.The mass was computed both by fitting a NFW profile
(M3D =Mvir), and by aperture densitometry (M2D). In the case
of the NFW fit, in addition to a 2-parameters fit (virial mass Mvir
and concentration cvir), we also show (MNFW) the results ob-
tained using the relation proposed by Bullock et al. (2001) be-
tween Mvir,cvir and the cluster redshift zcl:
cvir =K
1+zcl Mvir
M!α
,(11)
with M=1.5×1013/h M,K=9, α=0.13.
The shear profiles obtained from the dierent methods dis-
cussed below are displayed in Fig. 8; also displayed are the av-
erage values of tangential and radial components of shear, com-
puted in bins selected to contain at least 200 galaxies, and cen-
tered on the BCG: this is also where the peak of the X–ray emis-
sion is located (Rizza et al. 1998). In order to check the possible
error introduced by a wrong center, we considered a grid around
the position of the BCG, with a step of 2 arcsec: for each posi-
tion in the grid, we took it as the center, performed the fit and
derived the mass. Within 30 arcsec, we obtain that the rms is
σ(Mvir)<5%. The NFW parameters obtained by model fitting,
and the reduced χ2computed from the binned average tangential
shear, are given in Table 4.
The magnitude cut is the simplest approach as it only re-
quires photometry from the same band in which the lensing mea-
surement is done. Taking galaxies in the range 23 <R<26 mag
(a), produces a sample dominated by faint background galaxies,
but the inner regions of the cluster may still present an unknown
contamination by cluster galaxies.
To improve the selection, we proceeded as follows. The locus
of foreground galaxies was first found, allowing a better separa-
tion of dierent galaxy populations (see Medezinski et al. 2010,
and references therein), compared e.g. to methods based on the
selection of only red galaxies. Here we considered two colors
selections, namely Bzvs. Vz(b) and BVvs. VI(c),
6
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
100 200 300 400 500 600 700
−0.05 0.00 0.05 0.10 0.15
θ (arcsec)
γ
mnfw
nfw
a
500 1000 1500 2000
r (kpc)
100 200 300 400 500 600 700
−0.05 0.00 0.05 0.10 0.15
θ (arcsec)
γ
mnfw
nfw
b
500 1000 1500 2000
r (kpc)
100 200 300 400 500 600 700
−0.05 0.00 0.05 0.10 0.15
θ (arcsec)
γ
mnfw
nfw
c
500 1000 1500 2000
r (kpc)
100 200 300 400 500 600 700
−0.05 0.00 0.05 0.10 0.15
θ (arcsec)
γ
mnfw
nfw
d
500 1000 1500 2000
r (kpc)
Fig.8. Shear profiles obtained with the dierent selection methods (see Table 4, where the parameters derived for each model are
given). The fitting was done using the maximum likelihood approach. Binned points are shown for display purposes only. In panel
a, the curves obtained by the two models (nfw/mnfw) overlap.
with 21 <R<26 mag. For intermediate redshifts (z0.2),
foreground and background objects are well separated in such
two colors diagrams. We explored the possibility to find the
best selection criteria based only on the observed colors, with-
out any information on the redshift of the galaxies. To this end,
we developed a semi-automatic procedure, implemented in the
R language (R Development Core Team 2010). We first selected
bright (R<21 mag) galaxies, which are expected to be mainly at
zph <0.2. A kernel density estimate, obtained by the kde2d pack-
age in R (Venables & Ripley 2002) was then applied to these
points, giving the plots displayed in Fig.9: the normalization was
done so that the value of the maximum density in the binned data
was equal to one. The boundary of the foreground galaxy region
was then defined by the points within the same density level l
(e.g., l=0.2). Such region was converted to a polygon using the
splancs3package in R, that also allows to select for a given cata-
log those sources whose colors lie inside or outside the polygon.
A comparison with the model colors obtained in ZEBRA from
the convolution of the spectral templates with filter transmission
curves, shows that colors inside the area selected in such way
are consistent with those expected for galaxies at redshift <0.2.
Galaxies classified as foreground in such way were therefore ex-
cluded from the weak lensing analysis.
We finally used photometric redshifts (d), both for the selec-
tion of background galaxies, defined as those with 21 <R<26
mag, 0.3<zph <3, and to compute the average value of β: we
3http://www.maths.lancs.ac.uk/˜rowlings/Splancs
obtain in such way β(zs)=0.74, in good agreement with the
value obtained from the COSMOS catalog with the same mag-
nitude and redshift selection, β(zs)=0.73.
The eect of the dierent selections on the residual pres-
ence of cluster galaxies is displayed in Fig. 11, showing the den-
sity of background galaxies computed in dierent annuli around
the cluster: a clear increase of the density in the inner regions
is visible in case a, which indicates that magnitude selection
alone does not allow to completely remove the contamination by
cluster galaxies. Such contamination is greatly reduced by color
selection, and the optimal result is given by photometric red-
shifts, as expected. As a further check, we also found for each
method that the tangential shear signals of the rejected ’fore-
ground/cluster’ galaxies average out. In the following discus-
sion, we take as reference the results from case d, which is very
close to cin terms of uncertainties on fitted parameters, density
of background galaxies and residuals in the radial component of
the shear.
It was pointed out in Hoekstra (2003) and Hoekstra et al.
(2010) that large-scale structures along the line of sight provide
a source of uncertainty on cluster masses derived by weak lens-
ing, which is usually ignored and increases as a larger radius
(θmax) is used in the fitting. The uncertainty introduced by such
component on the mass estimate can be 10-20% for a cluster
with M=1015Mat z0.2, θmax =10 arcmin as in our case
(see Fig. 6 and 7 in Hoekstra (2003)), which is comparable to
the uncertainties derived in the fitting.
7
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
Table 4. Best-fit NFW parameters: for each selection method, in the upper row both Mvir and cvir were taken as free parameters,
in the lower row the (Mvir,cvir,zcl) relation (Bullock et al. 2001) was used. The reduced χ2was computed from the binned values
displayed in Fig. 8. Also given are the values of M200: the larger uncertainties are due to the fact they were derived from the fitted
values (Mvir,cvir). The values of M2D before and after applying the factor 1.34 (Sect.4.2) are displayed in the upper and lower rows,
respectively.
Method Mvir cvir rvir M200 χ2/d.o.f. M2D Density
1014 Marcsec 1014 M1014 Mgal/arcmin2
a 7.783.34
2.21 4.731.99
1.49 69888
73 6.447.94
3.23 1.13 12.3 ±2.98 30
7.801.65
1.52 4.710.13
0.12 69946
49 6.461.98
1.68 1.13 9.20 ±2.22
b 7.612.94
2.04 5.772.33
1.69 69380
69 6.426.85
3.21 0.40 8.14 ±3.15 25
9.001.85
1.69 4.620.13
0.11 73347
49 7.442.20
1.86 0.70 6.07 ±2.35
c 7.512.90
2.00 5.402.13
1.52 69079
68 6.306.15
3.08 0.87 10.5 ±3.03 28
8.381.69
1.57 4.670.13
0.11 71645
48 6.942.01
1.73 0.61 7.82 ±2.26
d 7.542.66
1.91 5.682.11
1.60 691.0573
64 6.356.45
3.02 1.1 10.2 ±2.94 25
8.731.75
1.60 4.640.12
0.11 725.7746
47 7.222.09
1.77 1.90 7.59 ±2.19
Also displayed in Table 4 are the projected masses computed
by aperture densitometry from Eq. 10, at a distance from the
cluster center r=rvir, and θ2=900′′,θout =1000′′ . A good
agreement is found between these masses and the values com-
puted from parametric fits, if we take into account the expected
ratio M2D/M3D =1.34 (see Sect. 4.2).
From the catalog based on photometric redshifts selec-
tion (case d), we finally derived the S-map introduced by
Schirmer et al. (2004), that is: S=MapMap , where
Map =Piet,iwiQ(|θiθ0|)
Piwi(12)
σ2
Map =Pie2
iw2
iQ2(|θiθ0|)
2(Piwi)2,
The map was obtained by defining a grid of points along the
image; the tangential components et,iof the lensed galaxy ellip-
ticities were computed taking as center each point in such grid.
The weight wiwas defined in Eq. 7, and Qis a Gaussian function
as in Radovich et al. (2008):
Q(|θθ0|)=1
πθ2
sexp (θθ0)2
θ2
s!,(13)
where θ0and θsare the center and size of the aperture (θs
1.5 arcmin). The S-map is displayed in Fig. 10, showing a quite
circular mass distribution centered on the BCG.
6. Discussion
Several mass measurements of this cluster are available in liter-
ature, based on dierent data and/or methods. Schmidt & Allen
(2007) used Chandra data and modeled the dark mat-
ter halo by a generalized NFW profile, obtaining a mass
value Mvir =9.16+1.89
1.85 ×1014Mand a concentration value
Table 5. Luminosity function parameters, and uncertainties. Φ
is in units of deg2.
Best-fit Standard error
α-1.06 0.07
m18.32 0.39
Φ/1031.273 0.374
cvir =5.08+0.55
1.03.
A weak lensing analysis of Abell 383 was done by
Bardeau et al. (2007) using CFH12K data in the B,R,Ifilters.
For the shape measurements they used a Bayesian method
implemented into the IM2SHAPE software. To retrieve the
weak lensing signal they selected the background galaxies
as those within 21.6<R<24.9 mag and (RI)&0.7,
obtaining a number density of 10 gal arcmin2. Their fit of
the shear profile by a NFW profile gave as result a mass of
M200 =4.19 ±1.46 ×h1
70 1014Mat r200 =1.32 ±0.17h1
70 Mpc
and a concentration value of c=2.62 ±0.69.
Another weak lensing mass estimate of Abell 383 from
CHF12K data was obtained by Hoekstra (2007) using two bands,
B(7200 s) and R(4800 s). The background sample was se-
lected by a magnitude cut 21 <R<24.5, from which cluster
red sequence galaxies were discarded. The remaining contam-
ination was estimated from the stacking of several clusters by
assuming that fraction of cluster galaxies fgc was a function of
radius r1. This function was used to correct the tangential
shear measurements. As discussed in Okabe et al. (2010), this
kind of calibration method does not allow to perform an unbi-
ased cluster-by-cluster correction. Assuming a NFW profile, the
fitted virial mass of Abell 383 was Mvir =2.81.6
1.5×h11014M.
8
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
0.0 0.5 1.0 1.5
0.5 1.0 1.5 2.0 2.5 3.0
V−z
B−z
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
0.2
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
B−V
V−I
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
0.2
Fig.9. Color selection of foreground galaxies (zphot <0.2). The
contour in red is the density level chosen for the selection; the
points display the model colors computed by ZEBRA for this
redshift range.
Abell 383 belongs to the clusters sample selected for the
Local Cluster Substructure Survey (LoCuSS) project (P.I. G.
Smith). Within this project, a weak lensing analysis of this clus-
ter has been recently performed by Okabe et al. (2010) using
SUBARU data in two filters, i(36 minutes) and V(30 min-
utes). In addition to a magnitude cut 22 <i<26 mag, they used
the color information to select galaxies redder and bluer than the
cluster red sequence. Looking at the trend of the lensing signal
as function of the color oset, they selected the sample where
the dilution were minimized, obtaining a background sample of
34 gal arcmin2for the computation of tangential shear pro-
file of the cluster. The fit of this profile by a NFW model did
not give an acceptable fit: the virial mass computed assuming
a NFW profile was Mvir =3.62+1.15
0.86 ×h11014Mwith a high
concentration parameter cvir =8.87+5.22
3.05. The same authors also
derived the projected mass, obtaining M2D=8.69 ×h11014 M
at the virial radius.
Scaling such mass values to h=0.7, we derive Mvir
(4 5) ×1014 M. This value is still consistent within uncer-
tainties with the value derived in the present analysis (Mvir =
(7.52.7
1.9×1014 M,cvir =5.72.1
1.6, case d). In our case, we obtain
Fig.10. Weak lensing S-map showing the mass distribution de-
rived by weak lensing; overlaid is the central region of the Abell
383 field.
1 2 3 4 5 6 7
15 20 25 30 35
r (arcmin)
ρ
no sel
a
b
c
d
Fig.11. Density (gal/arcmin2) of background galaxies used for
the lensing analysis, as a function of the distance from the clus-
ter, for the dierent selection methods here considered. The case
with no selection is also displayed for comparison.
however a better agreement with both the values of Mvir and cvir
given by the X-ray data, and a better consistency between para-
metric and non-parametric mass estimates, once the projection
factor is taken into account. This may be due either to how the
selection of background/foreground galaxies was done, or to a
higher accuracy in the shape measurement as a combination of
the deeper image and calibration of the bias due to SNR (Sect. 4).
9
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
Fig.12. R-band LF of the galaxies in Abell 383. Data points are
derived by binning the data in magnitude bins of 0.5 mag and
error bars by Poissonian errors; the curve traces the best-fitting
LF and the shaded area marks the model uncertainty, obtained
by a bootstrap technique.
Finally, we computed the luminosity function (LF hereafter)
in the R-band and derived the total luminosity by integrating
the fitted Schechter (Schechter 1976) function as described in
Radovich etal. (2008). In order to obtain the cluster LF, that is
the number of galaxies per unit luminosity and volume belong-
ing to the cluster, we need to remove from our catalog all the
background and foreground galaxies. Usually, this is done by
statistically subtracting the galaxy counts in a control field from
galaxy counts in the cluster direction. Here we take advantage
of the selection on the color-color diagrams described in Sect. 5,
and extract a catalog that includes cluster, foreground and resid-
ual background galaxies. The last two components were further
removed by the statistical subtraction, where we defined as clus-
ter area the circular region around the cluster center of radius
r=10.2 arcmin (1.3Mpc); the control field was instead defined as
the area outside the circle of radius r=15.3 arcmin. For the best-
fitting procedure to the Schechter function, we adopted conven-
tional routines of minimization on the binned distributions. Best-
fitting parameters are listed in Table 5. The R-band total lumi-
nosity, calculated as the Schechter integral, is Ltot =(2.14 ±0.5)
×1012 L. The errors were estimated by the propagation of the
68%-confidence-errors of each parameter.
For comparison, we then used the relation in Popesso et al.
(2007) between M200 and the optical luminosity, Lop, to see
whether the mass obtained in this paper is consistent with the
value expected for that luminosity. According to this relation,
the mass expected for such luminosity is M200 =(4.73 ±1.3)×
1014 M, in good agreement with the value derived by our weak
lensing analysis, M200 6.3×1014 M(case d), corresponding
to M/L300M/L.
7. Conclusions
We have computed the cluster Abell 383 mass by weak lens-
ing, using a deep R-band image taken with the Suprime cam-
era on the Subaru telescope. Catalogs extracted from combined
CFHT+SUBARU uBVRIz images were used to derive photo-
metric redshifts, and improve the weak lensing analysis. The
data were reduced using a pipeline developed in–house, that was
specifically designed for wide-field imaging data. The elliptic-
ities, from which the shear signal was derived, were measured
using a pipeline based on the KSB approach. We discussed some
aspects that may improve the results, namely the size of the win-
dow used to suppress the noise from the outer part of the galax-
ies, the selection of a limit on SNR below which the measure-
ment ellipticity is not accurate enough, and a weighting scheme
where uncertainties on spatial fitting of the PSF correction terms
were taken into account.
The accuracy on the mass estimate by weak lensing available
with our KSB pipeline was first derived on simulated images
which were built in such a way to mimic as closely as possi-
ble the background noise and the depth of the real image. From
these simulations we conclude that the mass can be measured
with an uncertainty 5-10% for log M/M14.5. Such ac-
curacy takes into account the measurement errors of the ellip-
ticity, but not the errors due to e.g. the foreground/background
galaxy separation, that may introduce an underestimate of the
mass. The impact of such selection was evaluated by comparing
three methods for the foreground/background galaxy separation,
namely: magnitude cut in one band, color selection and usage of
photometric redshifts. All methods gave consistent estimates of
the total virial mass, but from the shear profile it can be seen that
a dilution of the signal in the inner regions is still present, in the
case of a simple magnitude cut. Color selection and photometric
redshifts provide better results, even if the accuracy of the pho-
tometric redshifts is not high due to the few available bands. The
virial mass of Abell 383 here obtained by NFW model fitting is
in agreement with the value obtained from the non-parametric
mass estimate, that is Mvir 7×1014 M. Other previous weak
lensing analyses give Mvir (45)×1014 M: the value found in
this paper seems more in agreement with the value found by X-
ray data, and we also have a better agreement between paramet-
ric and non-parametric estimates, compared to e.g. Okabe et al.
(2010).
Finally, we estimated the R-band LF of Abell 383, and de-
rived the total R-band luminosity of the cluster: starting from this
value and using the relation between mass and luminosity found
for clusters by Popesso et al. (2007), we conclude that the mass
derived by weak lensing is consistent with the value expected for
this luminosity.
8. Acknowledgments
L.F., Z.H. and M.R. acknowledge the support of the European
Commission Programme 6-th framework, Marie Curie Training
and Research Network “DUEL”, contract number MRTN-
CT-2006-036133. L.F. was partly supported by the Chinese
National Science Foundation Nos. 10878003 & 10778725, 973
Program No. 2007CB 815402, Shanghai Science Foundations
and Leading Academic Discipline Project of Shanghai Normal
University (DZL805), and Chen Guang project with No.
10CG46 of Shanghai Municipal Education Commission and
Shanghai Education Development Foundation. A.R. acknowl-
edges support from the Italian Space Agency (ASI) contract
Euclid-IC I/031/10/0. We are grateful to the referee for the useful
comments that improved this paper.
References
Abell, G. O., Corwin, Jr., H. G., & Olowin, R. P. 1989, ApJS, 70, 1
10
Zhuoyi Huang et al.: A weak lensing analysis of the Abell 383 cluster.
Bardeau, S., Kneib, J., Czoske, O., et al. 2005, A&A, 434, 433
Bardeau, S., Soucail, G., Kneib, J., et al. 2007, A&A, 470, 449
Bartelmann, M. 1996, A&A, 313, 697
Broadhurst, T.,Takada, M., Umetsu, K., et al. 2005, ApJ, 619, L143
Bullock, J. S., Kolatt, T. S., Sigad, Y.,et al. 2001, MNRAS, 321, 559
Capaccioli, M., Mancini, D., & Sedmak, G. 2005, The Messenger, 120, 10
Capak, P., Aussel, H., Ajiki, M., et al. 2007, ApJS, 172, 99
Clowe, D., Luppino, G. A., Kaiser, N., Henry, J. P., & Gioia, I. M. 1998, ApJ,
497L, 61
Ebeling, H., Voges, W., Bohringer, H., et al. 1996, MNRAS, 281, 799
Erben, T., Waerbeke, L. V., Bertin, E., Mellier, Y., & Schneider, P. 2001, A&A,
366, 717
Fahlman, G., Kaiser, N., Squires, G., & Woods, D. 1994, ApJ, 437, 56
Feldmann, R., Carollo, C. M., Porciani, C., et al. 2006, MNRAS, 372, 565
Fetisova, T. S., Kuznetsov, D. Y., Lipovetskij, V. A., Starobinskij, A. A., &
Olowin, R. P. 1993, Astronomy Letters, 19, 198
Gill, M. S. S., Young, J. C., Draskovic, J. P.,et al. 2009, ArXiv:0909.3856
Grado, A., Capaccioli, M., Limatola, L., & Getman, F. 2011, ArXiv:1102.1588
Heymans, C., Van Waerbeke, L., Bacon, D., et al. 2006, MNRAS, 368, 1323
Hoekstra, H. 2003, MNRAS, 339, 1155
Hoekstra, H. 2007, MNRAS, 379, 317
Hoekstra, H., Franx, M., Kuijken, K., & Squires, G. 1998, ApJ, 504, 636
Hoekstra, H., Hartlap, J.,Hilbert, S., & van Uitert, E.2010, ArXiv:1011.1084
Luppino, G. & Kaiser, N. 1997, ApJ, 475, 20
Massey, R., Heymans, C., Berg´e, J., et al. 2007, MNRAS, 376, 13
Medezinski, E., Broadhurst, T., Umetsu, K., et al. 2010, MNRAS, 405, 257
Miyazaki, S., Komiyama, Y., Sekiguchi, M., et al. 2002, PASJ, 54, 833
Oguri, M., Takada, M., Okabe, N., & Smith, G. P. 2010, MNRAS, 405, 2215
Okabe, N., Takada, M., Umetsu, K., Futamase, T., & Smith, G. P. 2010, PASJ,
62, 811
Pickles, A. J. 1985, ApJS, 59, 33
Popesso, P., Biviano, A., B¨oringher, H., & Romaniello, M. 2007, A&A, 464, 451
R Development Core Team. 2010, R: A Language and Environment for
Statistical Computing, R Foundation for Statistical Computing, Vienna,
Austria, ISBN 3-900051-07-0
Radovich, M., Puddu, E., Romano, A., Grado, A., & Getman, F. 2008, A&A,
487, 55
Rizza, E., Burns, J. O., Ledlow, M. J., et al. 1998, MNRAS, 301, 328
Romano, A., Fu, L., Giordano, F., et al. 2010, A&A, 514, A88+
Schechter, P. 1976, ApJ, 203, 297
Schirmer, M., Erben, T., Schneider, P., Wolf, C., & Meisenheimer, K. 2004,
A&A, 420, 75
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525
Schmidt, R. W. & Allen, S. W. 2007, MNRAS, 379, 209
Schrabback, T., Hartlap, J., Joachimi, B., et al. 2010, A&A, 516, 63
Smith, G. P., Kneib, J., Ebeling, H., Czoske, O., & Smail, I. 2001, ApJ, 552, 493
Venables, W. N. & Ripley, B. D. 2002, Modern Applied Statistics with S, 4th
edn. (New York: Springer), isbn 0-387-95457-0
Voges, W. 1992, in European Southern Observatory Conference and Workshop
Proceedings, Vol. 43, European Southern Observatory Conference and
Workshop Proceedings, ed. A. Heck & F. Murtagh, 139–+
Wright, C. O. & Brainerd, T. G. 2000, ApJ, 534, 34
11
... The science rationale of the survey span a wide range of topics. Multi-epoch observations of well-studied fields allow us to study the rate of SN explosions and the properties of the SN-host population [25,26] as well as the low-luminosity AGN population discovered by means of their optical variability [27,28]. Deep high-fidelity optical images will also allow us to study galaxy groups and clusters at intermediate and high redshift. ...
... Upon reduction of the full BV R dataset, we decided to focus our VST observations toward obtaining deep ugri optical imaging of the CDFS field while obtaining additional u-band imaging of the ES1 field. In order to maximize the efficient use of the INAF Guaranteed Time and strengthen the scientific synergy of the two projects, we also joined forces with the SUDARE project [25], whose main aim was to carry out a supernova search [26] and study AGN variability [27,28] with VST. As part of this collaboration, SUDARE obtained multi-epoch gri data within VOICE-CDFS, while VOICE obtained u-band observations and additional gri data over the same field. ...
... A number of weak-lensing analyses of the cluster have been carried out (Squires et al. 1997;Radovich et al. 2008;Okabe et al. 2011;Soucail 2012). Okabe et al. (2011) andSoucail (2012) both find evidence for a bimodal mass distribution with the peak of the X-ray gas located between the two mass peaks. ...
Article
Full-text available
Golovich et al. present an optical imaging and spectroscopic survey of 29 radio relic merging galaxy clusters. In this paper, we study this survey to identify substructure and quantify the dynamics of the mergers. Using a combined photometric and spectroscopic approach, we identify the minimum number of substructures in each system to describe the galaxy populations and estimate the line-of-sight velocity difference between likely merging subclusters. We find that the line-of-sight velocity components of the mergers are typically small compared with the maximum 3D relative velocity (usually <1000 km s ⁻¹ and often consistent with zero). We also compare our systems to n -body simulation analogs and estimate the viewing angle of the clean mergers in our ensemble. We find that the median system’s separation vector lies within 40° (17°) at a 90% (50%) confidence level. This suggests that the merger axes of these systems are generally in or near the plane of the sky, matching findings in magnetohydrodynamical simulations. In 28 of the 29 systems we identify substructures in the galaxy population aligned with the radio relic(s) and presumed associated merger-induced shock. From this ensemble, we identify eight systems to include in a “gold” sample that is prime for further observation, modeling, and simulation study. Additional papers will present weak-lensing mass maps and dynamical modeling for each merging system, ultimately leading to new insight into a wide range of astrophysical phenomena at some of the largest scales in the universe.
... Upon reduction of the full BV R dataset, we decided to focus our VST observations toward obtaining deep ugri optical imaging of the CDFS field while obtaining additional u-band imaging of the ES1 field. In order to maximize the efficient use of the INAF Guaranteed Time and strengthen the scientific synergy of the two projects, we also joined forces with the SUDARE project [32], whose main aim was to carry out a supernova search [33] and study AGN variability [25,26] with VST. As part of this collaboration, SUDARE obtained multi-epoch gri data within VOICE-CDFS, while VOICE obtained u-band observations and additional gri data over the same field. ...
Article
Full-text available
We present the VST Optical Imaging of the CDFS and ES1 Fields (VOICE) Survey, a VST INAF Guaranteed Time program designed to provide optical coverage of two 4 deg$^2$ cosmic windows in the Southern hemisphere. VOICE provides the first, multi-band deep optical imaging of these sky regions, thus complementing and enhancing the rich legacy of longer-wavelength surveys with VISTA, Spitzer, Herschel and ATCA available in these areas and paving the way for upcoming observations with facilities such as the LSST, MeerKAT and the SKA. VOICE exploits VST's OmegaCAM optical imaging capabilities and completes the reduction of WFI data available within the ES1 fields as part of the ESO-Spitzer Imaging Extragalactic Survey (ESIS) program providing $ugri$ and $uBVR$ coverage of 4 and 4 deg$^2$ areas within the CDFS and ES1 field respectively. We present the survey's science rationale and observing strategy, the data reduction and multi-wavelength data fusion pipeline. Survey data products and their future updates will be released at http://www.mattiavaccari.net/voice/ and on CDS/VizieR.
Article
Context . Galaxy clusters form at the intersections of the filamentary large scale structure in merging events and by the accretion of matter along these filaments. Imprints of these formation processes should be visible in the intracluster medium and can arise in shock fronts, which are detectable via discontinuities in, for example, the gas temperature and density profiles. However, relatively few observational examples of prominent shocks have been detected in X-rays so far. Aims . In this study, we investigate the X-ray properties of the intracluster gas and the radio morphology of the extraordinary cluster A2163. This cluster shows an irregular morphology in various wavelengths and has one of the most luminous and extended radio halos known. Additionally, it is one of the hottest clusters known. We aim to measure the temperature and density profiles in two azimuthal directions to search for the presence of shock fronts. Methods . We performed a spectral analysis of data from two Suzaku observations, one in the north-east (NE) and one in the southwest (SW) direction of A2163, and used archival XMM-Newton data to remove point sources in the field of view. We deprojected the temperature and density profiles and accounted for the Suzaku point spread function. From the detected discontinuities in the density and temperature profiles, we estimated the Mach numbers and velocities of the shock fronts. To compare our findings in the X-ray regime with the radio emission, we obtained radio images of the cluster from an archival Very Large Array (VLA) observation at 20 cm. Results . We identify three shock fronts in A2163 in our spectral X-ray study. A clear shock front lies in the NE direction at a distance of 1.4 Mpc from the center, with a Mach number of M = 1.7 +0.3−0.2 , estimated from the temperature discontinuity. This shock coincides with the position of a known radio relic. We identify two additional shocks in the SW direction, one with M = 1.5 +0.5−0.3 at a distance of 0.7 Mpc, which is likely related to a cool core remnant, and a strong shock with M = 3.2 +0.6−0.7 at a distance of 1.3 Mpc, which also closely matches the radio contours. The complex structure of A2163 as well as the different Mach numbers and shock velocities suggest a merging scenario with two unequal merging constituents, where two shock fronts emerged at an early stage of the merger and traveled outwards while an additional shock front developed in front of the merging cluster cores.
Article
Full-text available
Strong gravitational lensing by galaxy clusters magnifies background galaxies, enhancing our ability to discover statistically-significant samples of galaxies at z>6, in order to constrain the high-redshift galaxy luminosity functions. Here, we present the first five lens models out of the Reionization Lensing Cluster Survey (RELICS) Hubble Treasury Program, based on new HST WFC3/IR and ACS imaging of the clusters RXC J0142.9+4438, Abell 2537, Abell 2163, RXC J2211.7-0349, and ACT-CLJ0102-49151. The derived lensing magnification is essential for estimating the intrinsic properties of high-redshift galaxy candidates, and properly accounting for the survey volume. We report on new spectroscopic redshifts of multiply imaged lensed galaxies behind these clusters, which are used as constraints, and detail our strategy to reduce systematic uncertainties due to lack of spectroscopic information. In addition, we quantify the uncertainty on the lensing magnification due to statistical and systematic errors related to the lens modeling process, and find that in all but one cluster, the magnification is constrained to better than 20% in at least 80% of the field of view, including statistical and systematic uncertainties. The five clusters presented in this paper span the range of masses and redshifts of the clusters in the RELICS program. We find that they exhibit similar strong lensing efficiencies to the clusters targeted by the Hubble Frontier Fields within the WFC3/IR field of view. Outputs of the lens models are made available to the community through the Mikulski Archive for Space Telescopes.
Chapter
Full-text available
The Kilo Degree survey (KiDS) is a large-scale optical imaging survey carried out with the VLT Survey Telescope (VST), which is the ideal tool for galaxy evolution studies. We expect to observe millions of galaxies for which we extract the structural parameters in four wavebands (u, g, r and i). This sample will represent the largest dataset with measured structural parameters up to a redshift z = 0. 5. In this paper we will introduce the sample, and describe the 2D fitting procedure using the 2DPHOT environment and the validation of the parameters with an external catalog.
Article
Full-text available
We present a mass estimate of the Planck-discovered cluster PLCK G100.2-30.4, derived from a weak lensing analysis of deep SUBARU griz images. We perform a careful selection of the background galaxies using the multi-band imaging data, and undertake the weak lensing analysis on the deep (1hr) r-band image. The shape measurement is based on the KSB algorithm; we adopt the PSFex software to model the Point Spread Function (PSF) across the field and correct for this in the shape measurement. The weak lensing analysis is validated through extensive image simulations. We compare the resulting weak lensing mass profile and total mass estimate to those obtained from our re-analysis of XMM-Newton observations, derived under the hypothesis of hydrostatic equilibrium. The total integrated mass profiles are in remarkably good agreement, agreeing within 1$\sigma$ across their common radial range. A mass $M_{500} \sim 7 \times 10^{14} M_\odot$ is derived for the cluster from our weak lensing analysis. Comparing this value to that obtained from our reanalysis of XMM-Newton data, we obtain a bias factor of (1-b) = 0.8 $\pm$ 0.1. This is compatible within 1$\sigma$ with the value of (1-b) obtained by Planck Collaboration XXIV from their calibration of the bias factor using newly-available weak lensing reconstructed masses.
Article
The measurement of the mass of clusters of galaxies is crucial for their use in cosmology and astrophysics. Masses can be efficiently determined with weak lensing (WL) analyses. I compiled from Literature a Catalog of weak Lensing Clusters (LC^2). Cluster identifiers, coordinates, and redshifts have been standardised. WL masses were reported to over-densities of 2500, 500, 200, and to the virial one in the reference Lambda-CDM model. Duplicate entries were carefully handled. I produced three catalogs: LC^2-single, with 485 unique groups and clusters analysed with the single-halo model; LC^2-substructure, listing substructures in complex systems; LC^2-all, listing all the 822 WL masses found in literature. The catalogs are publicly available at https://www.dropbox.com/sh/hukhb24c3ahiun2/AADVuW7yUAA2XjyDrFwofejAa?dl=0
Article
Full-text available
Abell 383 is a famous rich cluster (z = 0.1887) imaged extensively as a basis for intensive strong and weak lensing studies. Nonetheless there are few spectroscopic observations. We enable dynamical analyses by measuring 2360 new redshifts for galaxies with r$_{petro} \leq 20.5$ and within 50$^\prime$ of the BCG (Brightest Cluster Galaxy: R.A.$_{2000} = 42.014125^\circ$, Decl$_{2000} = -03.529228^\circ$). We apply the caustic technique to identify 275 cluster members within 7$h^{-1}$ Mpc of the hierarchical cluster center. The BCG lies within $-11 \pm 110$ km s$^{-1}$ and 21 $\pm 56 h^{-1}$ kpc of the hierarchical cluster center; the velocity dispersion profile of the BCG appears to be an extension of the velocity dispersion profile based on cluster members. The distribution of cluster members on the sky corresponds impressively with the weak lensing contours of Okabe et al. (2010) especially when the impact of foreground and background structure is included. The values of R$_{200}$ = $1.22\pm 0.01 h^{-1}$ Mpc and M$_{200}$ = $(5.07 \pm 0.09)\times 10^{14} h^{-1}$ M$_\odot$ obtained by application of the caustic technique agree well with recent completely independent lensing measures. The caustic estimate extends direct measurement of the cluster mass profile to a radius of $\sim 5 h^{-1}$ Mpc.
Article
We describe the recent improvements made on the techniques for the estimates of the total mass of galaxy clusters using the X-ray data, obtained by the Chandra and XMM satellites, and the weak lensing observations, performed by Subaru telescope. Furthermore, we show the current state of X-ray and weak lensing mass comparison obtained by simulations and observational data. In particular, we use a subsample of 21 galaxy clusters of the LOCUSS High-Lx sample, for which e find that X-ray masses overcome weak lensing mass by about 10%.
Article
Full-text available
We present bright galaxy number counts in five broad bands ($u', g', r', i', z'$) from imaging data taken during the commissioning phase of the Sloan Digital Sky Survey (SDSS). The counts are derived from two independent stripes of imaging scans along the Celestial Equator, one each toward the North and the South Galactic cap, covering about 230 and 210 square degrees, respectively. A careful study is made to verify the reliability of the photometric catalog. For galaxies brighter than $r^* = 16$, the catalog produced by automated software is examined against eye inspection of all objects. Statistically meaningful results on the galaxy counts are obtained in the magnitude range $12 \le r^* \le 21$, using a sample of 900,000 galaxies. The counts from the two stripes differ by about 30% at magnitudes brighter than $r^*= 15.5$, consistent with a local $2\sigma$ fluctuation due to large scale structure in the galaxy distribution. The shape of the number counts-magnitude relation brighter than $r^* = 16$ is well characterized by $N \propto 10^{0.6m}$, the relation expected for a homogeneous galaxy distribution in a ``Euclidean'' universe. In the magnitude range $ 16 < r^* < 21$, the galaxy counts from both stripes agree very well, and follow the prediction of the no-evolution model, although the data do not exclude a small amount of evolution. We use empirically determined color transformations to derive the galaxy number counts in the $B$ and $I_{814}$ bands. We compute the luminosity density of the universe at zero redshift in the five SDSS bands and in the $B$ band. We find ${\cal L}_{B} = 2.4 \pm 0.4 \times 10^8L_\odot h $Mpc$^{-3}$, for a reasonably wide range of parameters of the Schechter luminosity function in the $B$ band. Comment: 48 pages, 15 figures
Article
Full-text available
We explore the dark matter distribution in MS 1224.7+2007 using the gravitational distortion of the images of faint background galaxies. Projected mass image reconstruction reveals a highly significant concentration coincident with the X-ray and optical location. The concentration is seen repeatably in reconstructions from independent subsamples, and the azimuthally averaged tangential shear pattern is also clearly seen in the data. The projected mass within a 2.76 min radius aperture is approximately = 3.5 x 1014 solar mass/h. This is approximately = 3 times larger than that predicted if mass traces light with M/L = 275 as derived form virial analysis. It is very hard to attribute the discrepancy to a statistical fluctuation, and a further indication of a significant difference between the mass and the light comes from a second mass concentration, which is again seen in independent subsamples but which is not seen at all in the cluster light. We find a mass per galaxy visible to I = 22 of approximately 8 x 1022 solar mass/h, which, if representative of the universe, implies a density parameter Omega approximately = 2. We find a null detection of any net shear from large-scale structure with a precision of 0.9% per component. This is much smaller than the possible detection in a recent comparable study, and the precision here is at about the level of the rms shear fluctuations in realistic models.
Article
Full-text available
Context: A 2163 is among the richest and most distant Abell clusters, presenting outstanding properties in different wavelength domains. X-ray observations have revealed a distorted gas morphology and strong features have been detected in the temperature map, suggesting that merging processes are important in this cluster. However, the merging scenario is not yet well-defined. Aims: We have undertaken a complementary optical analysis, aiming to understand the dynamics of the system, to constrain the merging scenario and to test its effect on the properties of galaxies. Methods: We present a detailed optical analysis of A 2163 based on new multicolor wide-field imaging and medium-to-high resolution spectroscopy of several hundred galaxies. Results: The projected galaxy density distribution shows strong subclustering with two dominant structures: a main central component (A), and a northern component (B), visible both in optical and in X-ray, with two other substructures detected at high significance in the optical. At magnitudes fainter than R=19, the galaxy distribution shows a clear elongation approximately with the east-west axis extending over 4~h70-1 Mpc, while a nearly perpendicular bridge of galaxies along the north-south axis appears to connect (B) to (A). The (A) component shows a bimodal morphology, and the positions of its two density peaks depend on galaxy luminosity: at magnitudes fainter than R = 19, the axis joining the peaks shows a counterclockwise rotation (from NE/SW to E-W) centered on the position of the X-ray maximum. Our final spectroscopic catalog of 512 objects includes 476 new galaxy redshifts. We have identified 361 galaxies as cluster members; among them, 326 have high precision redshift measurements, which allow us to perform a detailed dynamical analysis of unprecedented accuracy. The cluster mean redshift and velocity dispersion are respectively z= 0.2005 ± 0.0003 and 1434 ± 60 km s-1. We spectroscopically confirm that the northern and western components (A 2163-B and A 2163-C) belong to the A 2163 complex. The velocity distribution shows multi-modality, with an overall bimodal structure peaking at ~59 200 km s-1 and ˜60 500 km s-1. A significant velocity gradient (~1250 km s-1) is detected along the NE/SW axis of the cluster, which partially explains the detected bimodality. A 2163 appears to be exceptionally massive: the cluster virial mass is M_vir = 3.8 ± 0.4 × 1015~M_&sun;~h70-1. Conclusions: Our analysis of the optical data, combined with the available information from X-ray observations and predictions of numerical simulations, supports a scenario in which A 2163-A has undergone a recent (t ˜ 0.5 Gyr) merger along a NE/SW (or E-W) axis, and A 2163-B is connected to the main complex, and is probably infalling on A 2163-A. Based on data obtained with the European Southern Observatory, Chile (runs 073.A-0672 and 077.A-0813) and with the Canada France Hawaii Telescope. Table 1 is only available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/481/593
Article
Full-text available
Aims:We present a cosmic shear analysis and data validation of 15 square degree high-quality R-band data of the Garching-Bonn Deep Survey obtained with the Wide Field Imager of the MPG/ESO 2.2 m telescope. Methods: We measure the two-point shear correlation functions to calculate the aperture mass dispersion. Both statistics are used to perform the data quality control. Combining the cosmic shear signal with a photometric redshift distribution of a galaxy sub-sample obtained from two square degree of UBVRI-band observations of the Deep Public Survey we determine constraints for the matter density Omega_m, the mass power spectrum normalisation sigma8 and the dark energy density Omega_Lambda in the magnitude interval Rin [21.5,24.5]. In this magnitude interval the effective number density of source galaxies is n=12.5 arcmin-2, and their mean redshift is bar z=0.78. To estimate the posterior likelihood we employ the Monte Carlo Markov Chain method. Results: Using the aperture mass dispersion we obtain for the mass power spectrum normalisation sigma_8=0.80± 0.10 (1sigma statistical error) at a fixed matter density Omega_m=0.30 assuming a flat universe with negligible baryon content and marginalising over the Hubble parameter and the uncertainties in the fitted redshift distribution. Based on observations made with ESO Telescopes at the La Silla Observatory. Appendix A is only available in electronic form at http://www.aanda.org
Article
Full-text available
A study is made of stellar and galaxy colors using a spectrophotometric synethesis technique. We show that use of good color response functions and a modern determination of the spectroscopic energy distribution for a alpha-Lyr gives synthetic colors in a good agreement with photometric observations to about 0.05 mag. The synthetic method then is applied to sutdy galaxy colors using the spectrophotometric atlas of Kennicutt (1992), and a comparison is made with observed galaxy colors. The K-correction is calculated and compared with that of Coleman, Wu and Weedman (1980). We then study colors of galaxies in various photometric band systems and obtain color transformation laws, which enable us to find offsets among different systems. We include 48 photometric bands in our study. (SECTION: Galaxies)
Article
The VLT Survey Telescope (VST) is now a few months from its completion. this paper briefly reviews the project, accounts for its current status, anticipates the calendar of future milestones up to first light, and lists the scientific programmes for the observing time guaranteed to the OAC by ESO for procurement of the telescope.
Article
This is an all-sky catalog of 4073 rich clusters of galaxies, each having at least 30 members within the magnitude range m3 to m3+2 (m3 is the magnitude of the third brightest cluster member) and each with a nominal redshift less than 0.2. The southern data have been collected from a survey of UK 1.2 m Schmidt telescope IIIa-J plates and films and have been reduced to the systems defined by the northern data previously published by Abell (1958). A revised northern catalog, including Bautz-Morgan (1970) types and redshifts where known, is also included.
Article
Continuous spectrophotometry has been obtained for 200 objects at a resolution of 10-17 A over the wavelength region 3600-10,000 A. Kron-Cousins BVRI colors are computed from the spectra and compared with published photoelectric photometry. The (V - R)C color index is used to group the individual observations to form synthesis standard spectra for 48 common spectral types. The standard groups include a solar abundance sequence of most spectral types and luminosity classes, metal-rich and metal weak G - K giant-branch sequences, and horizontal-branch giants. The variations with color, luminosity, and metallicity of several prominent line strengths are discussed. The spectral atlas is available as a FITS magnetic tape.
Article
A new analytic approximation for the luminosity function for galaxies is proposed, which shows good agreement with both a luminosity distribution for bright nearby galaxies and a composite luminosity distribution for cluster galaxies. The analytic expression is proportional to L⁻⁵/⁴e⁻/sup L*/, where L* is a characteristic luminosity corresponding to a characteristic absolute magnitude M*/sub B0/âââ/sub .6/For an individual cluster, the characteristic magnitude may be determined with an accuracy of approx.0.25 mag, suggesting its use as a standard candle. The analytic expression is used to compute an expected richness-absolute magnitude correlation for first ranked cluster galaxies and an expected dispersion, which are compared with the data of Sandage and Hardy. (AIP)
Article
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.