Astronomy
&
Astrophysics
A&A 525, A79 (2011)
DOI: 10.1051/0004-6361/201014415
c ESO 2010
Merging history of three bimodal clusters⋆,⋆⋆
S. Maurogordato1 , J. L. Sauvageot2 , H. Bourdin3 , A. Cappi1,4 , C. Benoist1 , C. Ferrari1 , G. Mars1 , and K. Houairi5,6
1
2
3
4
5
6
Université de Nice Sophia-Antipolis, CNRS, Observatoire de la Côte d’Azur, UMR 6202 Cassiopée, BP 4229,
06304 Nice Cedex 4, France
e-mail:
[email protected]
CEA, DSM, Irfu, Service d’Astrophysique, C.E. Saclay, 91191 Gif-sur-Yvette Cedex, France
Dipartimento di Fisica, Università degli Studi di Roma “Tor Vergata”, via della Ricerca Scientifica, 1, 00133 Roma, Italy
INAF - Osservatorio Astronomico di Bologna, via Ranzani 1, 40127 Bologna, Italy
ONERA, Optics Department, BP 72, 92322 Chatillon Cedex, France
CNES, 18 avenue Édouard Belin, 31401 Toulouse Cedex 09, France
Received 12 March 2010 / Accepted 20 August 2010
ABSTRACT
We present a combined X-ray and optical analysis of three bimodal galaxy clusters selected as merging candidates at z ∼ 0.1. These
targets are part of MUSIC (MUlti-Wavelength Sample of Interacting Clusters), which is a general project designed to study the
physics of merging clusters by means of multi-wavelength observations. Observations include spectro-imaging with XMM-Newton
EPIC camera, multi-object spectroscopy (260 new redshifts), and wide-field imaging at the ESO 3.6 m and 2.2 m telescopes. We build
a global picture of these clusters using X-ray luminosity and temperature maps together with galaxy density and velocity distributions.
Idealized numerical simulations were used to constrain the merging scenario for each system. We show that A2933 is very likely an
equal-mass advanced pre-merger ∼200 Myr before the core collapse, while A2440 and A2384 are post-merger systems (∼450 Myr
and ∼1.5 Gyr after core collapse, respectively). In the case of A2384, we detect a spectacular filament of galaxies and gas spreading
over more than 1 h−1 Mpc, which we infer to have been stripped during the previous collision. The analysis of the MUSIC sample
allows us to outline some general properties of merging clusters: a strong luminosity segregation of galaxies in recent post-mergers;
the existence of preferential axes – corresponding to the merging directions – along which the BCGs and structures on various scales
are aligned; the concomitance, in most major merger cases, of secondary merging or accretion events, with groups infalling onto the
main cluster, and in some cases the evidence of previous merging episodes in one of the main components. These results are in good
agreement with the hierarchical scenario of structure formation, in which clusters are expected to form by successive merging events,
and matter is accreted along large-scale filaments.
Key words. X-rays: galaxies: clusters – galaxies: clusters: intracluster medium – galaxies: clusters: individual: Abell 2384 –
galaxies: clusters: individual: Abell 2440 – galaxies: clusters: individual: Abell 2933
1. Introduction
In standard Cold Dark Matter cosmological models, including
the concordance ΛCDM, the general growth of structures starts
from the primordial density fluctuations generated by inflation,
and is driven by gravity in a hierarchical way, i.e. smaller structures form first, then merge into progressively more massive
systems; however, if the acceleration of the expansion is due
to a cosmological constant, the process of structure formation
will completely stop in the future (Krauss & Starkman 2000;
Nagamine & Loeb 2003; Busha et al. 2005).
Galaxy clusters are the largest gravitationally bound objects
of the hierarchy; they accrete smaller groups coming from the
filaments of the cosmic web and occasionally merge with other
clusters of comparable mass, releasing an exceptionally high
amount of energy. Merging clusters are therefore ideal laboratories to study the process of structure formation and how its
affects galaxy evolution. In this scenario, one expects to find
⋆
Based on data obtained with the European Southern Observatory,
Chile (programs 072.A-0595, 075.A-0264, and 079.A-0425).
⋆⋆
Tables 5–7 are only available in electronic form at the CDS via
anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via
http://cdsarc.u-strasbg.fr/viz-bin/qcat?J/A+A/525/A79
a large fraction of clusters in their formation process at high
redshift (which is corroborated by the large fraction of irregular morphologies observed). However, observations of clusters
at high redshift are difficult and time consuming, so an alternative choice is to search for rarer but more easily observable
merging candidates at low redshift. In this way, we can probe
in detail the merging signatures and try to shed some light
on the process of cluster formation and how it affects galaxy
evolution.
A crucial progress in the study of merging clusters was
provided by the spectral and imaging capabilities of the last generation of X-ray satellites. Before then, X-ray spectroscopic information had not been available or of very poor quality, and the
main information about mergers was based on morphology. With
the spatially resolved spectroscopy and high resolution imaging offered by Chandra and XMM, the situation has radically
changed, and sophisticated algorithms have been developed to
achieve good and reliable temperature maps (see for instance
Bourdin et al. 2004), since the temperature is the more accurate tracer of the energy transfer from the collision to the X-ray
gas itself. Strong signatures of the merging events have been detected in these maps (Vikhlinin et al. 2000; Markevitch et al.
2000 and 2002), and cold fronts and bow shocks are now well
Article published by EDP Sciences
A79, page 1 of 19
A&A 525, A79 (2011)
established as common merger features. Thanks to this observational progress, major mergers of galaxy clusters now appear far
more complex than previously foreseen.
The full understanding of the complex processes at work
in merging requires dedicated numerical simulations. Much
progress has been made in this field, starting from the pioneering works of Schindler & Böhringer (1993) and Roettiger et al.
(1997). Ricker & Sarazin (2001, hereafter RS01) described the
violent relaxation of gas in a dark matter potential well for a variety of idealized merging systems, paying special attention to
the impact parameter and the mass ratio between units. Poole
et al. (2006) analyzed merging of idealized relaxed clusters with
sophisticated SPH simulations including cooling and star formation, and detected the major transient signatures existing in
observed temperature maps.
Combining optical with X-ray data has been shown to be extremely effective in unveiling the complex history of merging
clusters (to mention a few, Arnaud et al. 2000; Maurogordato
et al. 2000; Donnelly et al. 2001; Barrena et al. 2002, 2007;
Boschin et al. 2004; Owers et al. 2009). These studies have revealed various peculiar properties of the galaxy distribution in
the individual merging clusters, such as strong signatures in the
density and velocity distribution, and strong alignments effects.
However, a larger sample with both X-ray and optical observations is clearly needed to test the generality of these properties,
and their dependence on the merging stage.
Motivated by these reasons, we started an observational
program, MUSIC (MUltiwavelength Sample of Interacting
Clusters) to define and analyze a sample of merging clusters at
different stages of the merging process. The first targets (A3921,
A1750, A2065, and A2255) were selected from the merging
clusters observed during the XMM Guaranteed Time by one of
us (Sauvageot), and we established an optical follow-up program
at ESO.
To cover systems representing various stages of merging, we extended the initial sample by including targets from
Kolokotronis et al. (2001) which compared the density distributions of the gas (from ROSAT/HRI maps) and of the galaxies
(from APM maps). We selected the systems with a clear bimodal
morphology in both the X-ray and the optical maps as probable pre-merger candidates, and those with a distorted morphology and a pronounced segregation between gas (collisional) and
galaxies (non collisional), i.e. with features that are expected of
mergers at a more advanced stage. We selected clusters around
z ∼ 0.1, because at this redshift evolutionary effects on galaxies are negligible and the field of view of X-ray and optical instruments offers a good coverage of the system (30′ with XMM
and wide field imagers such as WFI at the ESO 2.2 m, corresponding to 2.3 h−1 Mpc at z = 0.1). With a 4 m class telescope,
spectroscopy is also easily performed as faint as R < 19, a limit
corresponding to ∼L∗ + 3.5 at this redshift (with a best-fit parameter MR∗ = −22.97 estimated by Popesso et al. (2005), fitting a
Schechter function to the cluster galaxy luminosity function of
the RASS-SDSS cluster survey). The aims of the program are to
characterize the merging scenario (epoch of collision, geometry,
and mass ratios) and to test the impact of the merging process on
the properties of galaxies, in particular the star formation. For
example, the combined X-ray and optical analysis allowed us to
determine the merging scenario of the first two clusters (a premerger for A1750, Belsole et al. 2004; and a recent post-merger
for A3921, Belsole et al. 2005; Ferrari et al. 2005).
Here we present the analysis of three additional merging
candidates: A2933, A2440, and A2384, based on XMM X-ray
spectro-imaging obtained in 2005 and 2006 (A2933 and A2440),
A79, page 2 of 19
or retrieved from the XSA database (A2384) and optical observations carried out in 2003, 2005, and 2007 at ESO, including wide-field multi-band imaging with the Wide Field Imager
on the 2.2 m telescope, and multi-object spectroscopy with
EFOSC2 at the 3.6 m telescope.
The clusters A2933, A2440 and A2384 share common features: an irregular morphology (they have been classified as
Bautz-Morgan type III, II, and II/III, respectively), low Abell
richness classes (1, 0, and 1), and redshifts slightly less than
0.1 (0.0925, 0.0906, and 0.0943 according to Struble & Rood
1999). In the optical, A2933 was observed as a target of the ESO
Nearby Abell Cluster Survey (Katgert et al. 1996) and the southern Abell redshift survey (Muriel et al. 2002; Way et al. 2005;
Coenda et al. 2006). A2440 was identified as a pre-merging cluster from the dynamical and X-ray/optical analysis of Beers et al.
(1991) and Mohr et al. (1996). A2384 is also a classical example
of a bimodal cluster (Ulmer & Cruddace 1982; West et al. 1995).
A weak-lensing analysis of A2384 was performed by Cypriano
et al. (2004). Both A2440 and A2384 belong to the flux-limited
sample of bright clusters of galaxies from the southern part of
the ROSAT All-Sky Survey (de Grandi et al. 1999).
In Sects. 2 and 3, we present the data, the reduction procedure, and the methodology used in X-ray and optical respectively. In Sects. 4−6, we analyze the data for A2933, A2440,
and A2384 respectively and propose a merging scenario for each
cluster. In Sect. 7 we present our numerical simulations of merging clusters that are used in Sect. 8 to refine the scenarios. In the
following, we adopt the cosmological parameters of a ΛCDM
model with ΩM = 0.3, ΩΛ = 0.7, and H0 = 70 km s−1 Mpc−1 .
With these parameters, at z = 0.1 one degree corresponds to a
physical length of 6.6 Mpc.
2. X-ray observations: scientific products
The three systems presented in this paper were observed by
XMM-Newton EPIC cameras for about 30 ks in full frame mode
with the medium filter. The observation were screened for proton flares on a high and low energy light curve basis. A summary
of the effective exposure time remaining after this screening process is provided in Table 1.
The brightness of the X-ray emitting intra-cluster medium
(ICM) was mapped from a multi-scale algorithm using Haar
wavelets, especially suited to denoise images dominated by shot
noise. First proposed by Jammal & Bijaoui (2004), this algorithm was adapted to analyze X-ray astronomical images and
correct signal distorsions related to the spatially variable detector response, and mirror effective area (Bourdin et al. 2008). We
estimated the ICM brightness from the spatial distribution of low
energy X-ray photons (with energy lower than 2.5 keV), since
their spectral distribution is expected to have a weak dependence
on temperature.
We used the X-ray data set in order to map the ICM temperature. To do so, we used the wavelet spectral-imaging algorithm
proposed in Bourdin et al. (2004) and Bourdin et al. (2008).
X-ray spectral-imaging can be performed only on data of sufficiently high count statistics, and it is necessary to gather photons within large detector regions, the shape of which needs to
be optimised to detect thermal structure. To overcome this major
difficulty, we developed a strategy based on wavelet transforms,
where we first estimated the ICM temperature with its associated
confidence range at various angular scales and locations in the
field of view. The detection of the temperature features and the
reconstruction of an ICM temperature map was finally obtained
by applying a threshold to the wavelet coefficients. The wavelet
S. Maurogordato et al.: Merging history of three bimodal clusters
Table 1. XMM-Newton-EPIC observations used in our analysis, with effective exposure time corresponding to each instrument.
Cluster name
A2933
A2440
A2384
XMM-Newton
obs. IDs
0305060101
0401920101
0101902701
Center coordinates
01h 40m 41.2s −54◦ 33′ 26.0′′
22h 23m 52.6s −01◦ 36′ 57.0′′
21h 52m 14.2s −19◦ 42′ 19.8′′
MOS1 effective
exposure time (ks)
21.5 (69.3%)
25.3 (55.4%)
17.2 (65.4%)
MOS2 effective
exposure time (ks)
27.7 (89.0%)
24.9 (54.3%)
16.8 (63.6%)
PN effective
exposure time (ks)
16.3 (68.5 %)
17.8 (50.3 %)
10.2 (51.6 %)
Notes. In brackets: fraction of the useful exposure time after solar-flare “cleaning”.
analysis was performed following a shift-invariant algorithm using B-spline wavelet coefficients, and applying a threshold of 1σ
confidence level above the noise fluctuation. Point sources have
to be masked before building the map, to avoid pollution from
the ICM signal at different spatial frequencies.
3. Optical observations: data description
and reduction methods
All three clusters were allocated time at ESO for wide-field
imaging with the WFI instrument at the 2.2 m telescope (total of 9.75 h), and extensive multi-object spectroscopy with
EFOSC2 at the 3.6 m telescope (total of 6 nights), in 2003, 2005,
and 2007, respectively (programs 072.A-0595, 075.A-0264, and
079.A-0425). As wide-field imaging of A2933 in service mode
was not performed, we used the APM galaxy survey (Maddox
et al. 1990) and the DSS images in the analysis of this cluster.
We retrieved the relevant galaxy catalog in a region of 30′ × 30′
covering A2933 from the Southern sky catalogue based on the
UKST SES R survey. For A2384 and A2440, we obtained imaging of the central 30′ × 30′ field in the R (filter ESO/844) and
B (filter ESO/878) passbands. For each filter, 8 dithered images
were obtained leading to a total exposure time of 40 min. These
sets of images were reduced and combined using the ESO/MVM
package Alambic (Vandamme et al. 2002), and the catalogs were
extracted with SExtractor (Bertin & Arnouts 1996). The typical
seeing is 1 arcsec. The magnitudes used in the present paper are
the total magnitudes “MAGAUTO ” provided by SExtractor. Stars
and galaxies were classified (up to a magnitude of 21 in both
bands) by extracting the stellar sequences within the magnitudehalf light radius diagrams. Fainter objects were all considered as
galaxies. Finally, catalogs with B − R color were built by associating objects with positions in the B and R catalogs differing by
less than 1.5 arcsec.
For A2440 and A2384, the final catalog of galaxies including both B and R passbands and covering a field of 30′ × 30′
was used to analyze the color properties and the spatial distribution of the cluster galaxies. The color-magnitude diagram was
used to identify the red sequence defined by the cluster elliptical
population (Lopez-Cruz et al. 2004). We selected galaxies with
colors within 3σ of this relation as high probability members
of the cluster and from this selection we derived the projected
density maps for different magnitude cuts in the R-band (corresponding typically to L∗ +1, L∗ +2, L∗ +3). In the case of A2933,
R band magnitudes from the APM catalog were used to compute
density maps without color selection.
Spectroscopy was performed with EFOSC2 using the
grism #03, which covers the spectral range 3050−6010 Å, with
a FWHM of 7.5 Å. In general, for each pointing we obtained
2 × 45 min exposures. Wavelength calibration was performed
in real time, taking spectra of arc calibration lamps (heliumargon) after each exposure. Data reduction was performed using
our dedicated IRAF package speXtra for automatic extraction
and wavelength calibration of spectra. Radial velocities were
obtained using the cross-correlation technique (Tonry & Davis
1981) with the rvsao package. Cross-correlation was performed
with velocity standards observed with the same instrumental
configuration during the observing run. The typical velocity error is ∼50 km s−1 , an estimate that we could confirm by performing a cross-check against the data available in the literature (see
Sect. 4.3).
The spectroscopic catalogs resulting from our observations
are listed for the three clusters in Tables 5−7, where columns are
as follows: 1) identification number; 2) and 3) right ascension
and declination (J2000.0); 4) radial velocity; 5) velocity error
for the cross-correlation; 6) RTR parameter (Tonry & Davis 1981;
when RTR > 3, the cross-correlation redshift can be considered
as reliable); and 7) quality flag for the redshift (0: high precision,
1: medium precision).
Objects with a velocity within a ±5000 km s−1 window centered on the mean velocity were selected as cluster member candidates. Possible interlopers were rejected by the gap technique.
The velocity location and scale of the various clusters (and subclusters) were determined with the biweight estimator using the
ROSTAT package (Beers et al. 1990).
Ten normality tests (provided by ROSTAT) were applied to
the data. The Dip test for unimodality (Hartigan & Hartigan
1985) was computed with the diptest package implemented by
Martin Maechler in the R environment. The Pvalue was computed
in each case according to the value of the Dip test and the number of objects according to the table provided in Hartigan &
Hartigan (1985). The results of this analysis are displayed in
Table 2, where columns are as follows: 1) name of the cluster
(or subcluster); 2) and 3) right ascension and declination of the
brightest galaxy (BCG) of the (sub)cluster; 4) angular radius θ of
the circle in which galaxies were selected for the ROSTAT analysis; 5) and 6) estimation of location and scale with the biweight
technique; 7) number of redshifts used in the analysis; 8) radial velocity of the BCG; 9) number of statistical tests excluding
Gaussianity at more than 10% confidence; 10) and 11) value of
the Dip test and of its P-value; and 12) mean temperature computed within radius θ and its error.
To test for multi-modality, we tried in each case to fit the entire velocity distribution with a mixture of Gaussian functions.
For this purpose, we have used the powerful EMMIX algorithm
(McLachlan & Krishnan 1997; McLachlan et al. 1999). This
program is quite flexible and allows a variety of choices in the
fitting. In particular, it does not require to introduce any guess
for the initial partition of velocities and the form of the covariance matrix. We applied EMMIX to the data, trying to fit the
velocity distribution with mixtures of Ng Gaussian distributions,
with Ng varying from 1 to 5. We retained the best fit on the basis
of the P-value, and in the case of identical values of P-value, on
the criterion of the smallest partition number. When a mixture of
Gaussians had been fitted to the data, we recomputed the value
of the location and scale of each partition with ROSTAT.
A79, page 3 of 19
A&A 525, A79 (2011)
Table 2. Characteristics of the clusters and subclusters.
Name
A2933
A2933N
A2933S
A2440
A2440A
A2440B
A2440C
A2384
A2384N
A2384S
RA(BCG)
–
01 40 35.520
01 40 59.520
–
22 24 13.663
22 23 56.941
22 23 47.814
–
21 52 21.962
21 52 09.556
Dec(BCG)
–
–54 30 54.00
–54 37 26.40
–
–01 31 34.34
–01 34 59.78
–01 39 01.14
–
–19 32 48.65
–19 43 23.71
θp
15
3.5
3.5
15
2.5
2.5
2.5
15
5
5
CBI (km s−1 ) S BI (km s−1 ) Nobj VBCG (km s−1 ) Ntest Dip Test Pvalue T X (KeV)
27 281 ± 103 682 ± 65
47
–
0
0.0340 0.02
–
0.31
26 919 ± 131 578 ± 70
24 27 424 ± 100
0
0.056
0.12 2.83−0.26
0.19
27 648 ± 166 568 ± 142
17 28 028 ± 100
0
0.071
0.15 2.80−0.17
27 251 ± 93
940 ± 70
103
–
0
0.022
0.01
–
28 005 ± 75
178 ± 90
10 27 925 ± 100 10
0.067
0.03
−−
0.20
27 212 ± 195 946 ± 95
29
27 032 ±16
1
0.050
0.1 4.07−0.19
0.26
27 234 ± 188 949 ± 150
25
26 816 ±19
0
0.059
0.15 4.57−0.20
28 263 ± 154 1114 ± 106 56
–
0
0.038
0.1
–
0.20
28 139 ± 182 1176 ± 120 44
27 602 ± 55
0
0.043
0.1 4.05−0.19
0.28
28 814 ± 241 877 ± 290
11
28 696 ± 53
0
0.066
0.03 4.02−0.26
Notes. Location, scale, and X-ray mean temperatures of the different subclusters are computed in circles of θp arcmin centered on the position of
the BCG (associated with the corresponding X-ray maximum). The measured velocity of the BCG is also listed.
Fig. 1. Galaxy density maps as a function of galaxy luminosity. From top to bottom: A2933, A2440 and A2384. The field of view displayed is
15′ × 15′ for A2933 and A2440, and 30′ × 30′ for A2384. From left to right, magnitude cuts are: R < 18, R < 19, R < 20 and R < 21. The density
maps of A2440 and A2384 have been computed for red sequence galaxies.
4. A2933
4.1. X-ray gas morphology and thermal structures
In X-rays, A2933 is a well defined bimodal system. In Fig. 2 we
show the temperature map of A2933 with the X-ray brightness
contours superimposed. Looking at the low energy contours,
both components look quite regular and have similar brightnesses. We clearly detect the interaction zone as a hot region located in-between the two maxima. The northern unit (A2933N)
is slightly elongated along the east-west direction, while the
southern unit (A2933S) is orientated along the general NWSE direction connecting the two units. In the temperature map,
the most prominent feature is the hot region between the two
A79, page 4 of 19
units, which is statistically significant, because the exposure time
of our XMM/EPIC observation was specifically adapted to ensure good statistics in the region of interaction. The temperature distribution of A2933N appears to be quite elongated, while
A2933S looks more regular.
4.2. Galaxy density distribution
The morphology of Abell 2933 has been very poorly studied at
optical wavelengths. In the following analysis, projected density
maps were inferred in the field of the cluster from galaxy positions and magnitudes in the APM catalog.
S. Maurogordato et al.: Merging history of three bimodal clusters
Fig. 2. Left: galaxy density map of A2933 (magnitude cut: R < 20) overlaid on the X-ray luminosity contours (EPIC-XMM counts in the
0.5−2.5 keV band corrected for background and vignetting). The black crosses indicate the BCG positions. Right: ICM temperature map of
A2933 overlaid on the X-ray luminosity contours (EPIC-XMM data analyzed by means of wavelet spectral-imaging; see Bourdin et al. 2004,
2008, for details).
The galaxy density distribution shows a bimodal structure,
with two subclusters located on a NW/SE axis at a separation of
∼7 arcmin (Fig. 1). The relative importance of the two components varies with the magnitude cut-off. Selecting bright galaxies (R < 18), the NW subcluster (A2933N) is dominant. When
including fainter objects, the SE (A2933S) subcluster becomes
progressively more prominent (Fig. 1). At faint magnitudes, the
NW subcluster shows an extended tail in the NW direction. In
Fig. 3, we also show the isodensity contours corresponding to
R < 20, superimposed on the DSS image of A2933. A2933N
hosts a couple of very bright galaxies, and is centered on a BCG
with a very close companion, while A2933S hosts a single bright
elliptical galaxy. One can also note a small density clump at the
southern extremity of A2933S. There is a general alignment of
the two BCGs, the two subclusters, and the global cluster structure along the NW/SE direction joining A2933N and A2933S.
4.3. Galaxy velocity distribution
A2933 was included in the Southern Abell Redshift Survey
(SARS, Way et al. 2005), and on the basis of 53 redshifts Muriel
et al. (2002) estimated the cluster global velocity and velocity
dispersion (27 709 ± 105 km s−1 and 759 ± 72 km s−1 , respectively); they also showed that the velocity dispersion as a function of radius appears to be constant to 5 Mpc/h from the cluster
center. Their data, however, sample the cluster and its environment on large scales and at relatively bright magnitudes. Since
we are interested in the merging process, we observed the central
region including the two main clumps at deeper magnitudes.
We measured 71 redshifts in the 30′ × 30′ field of A2933,
listed in Table 5, from which 47 objects are identified as cluster members. We have only 11 redshifts (among 53) in common
with them, because of the different region sampled. Excluding
one discrepant case, the redshifts of common objects are all
in excellent agreement, with a mean velocity difference of
14 km s−1 and a standard deviation of 59 km s−1 . From our
redshift sample, we analyzed the velocity distribution in the
central region of A2933. We measured a location of CBI =
27 281 ± 103 km s−1 and a scale of S BI = 682 ± 65 km s−1
(Table 2). Our estimate of velocity location is significantly lower
(∼400 km s−1 ) than that found by Muriel et al. (2002). These differences may be caused by variations in sampling. Their sample
covers a much larger area, that is not centered on the two subclusters but is shifted to the south, which we show to be a region
of higher mean velocity. We checked that their estimate of mean
velocity and our estimate for southern component (see below)
are consistent.
The velocity histogram (plotted in Fig. 4) clearly shows a
multi–modal distribution, with a sharp peak at ∼26 200 km s−1 ,
and a more dispersed structure, with two wider peaks at
∼26 800 km s−1 and ∼27 500 km s−1 . In Fig. 3, cluster members are indicated by different symbols and colors for galaxies with velocities within each of the three peaks in the histogram. Most of the galaxies with velocities located in the sharp
peak at ∼26 200 km s−1 (purple squares) are located in A2933N.
Moreover, the region of A2933S, including its southern extension, is populated mostly by galaxies in the highest velocity peak
at ∼27 500 km s−1 , while objects with velocities within the peak
at ∼26 800 km s−1 are mainly in the A2933N region.
To quantitatively assess the significance of the multimodality in the velocity distribution suggested by the histogram,
we applied our set of statistical tests (running a set of tests is
useful because they have different sensitivities to subclustering;
see Pinkney et al. 1996). Ten out of 10 of the normality tests
provided by ROSTAT do not find significant deviations from
Gaussianity (see Table 2). However, the Dip-test excludes unimodality at more than 2% significance.
We finally applied EMMIX, where the data were fitted
without a priori constraints on the covariance matrix. A three
Gaussian mixture inferred a P-value of 0.03, indicating that
the Gaussian hypothesis is excluded at a very high significance
level. This mixture fits the main structure with a Gaussian of
A79, page 5 of 19
A&A 525, A79 (2011)
15
10
5
0
cz [km/s]
Fig. 3. DSS region (17′ × 20′ ) centered on A2933 (North is up and
East is to the left). Galaxies identified as cluster members from spectroscopy are marked with different symbols corresponding to the
three major peaks in the velocity histogram (purple squares: [26 000,
26 500] km s−1 , cyan diamonds: [26 500, 27 200] km s−1 , orange circles:
[27 200, 28 500] km s−1 ) which is plotted in Fig. 4. The isocontours corresponding to the galaxy density map with R < 19 are superimposed.
The velocity histograms for the galaxies in the two circles (which have a
3.5 arcmin radius and are centered on the two subclusters) are displayed
in Fig. 5.
Table 3. Results of the mixture of Gaussian with EMMIX.
Name part v1 (km s−1 ) S BI (km s−1 )
A2933 1
27 364 ± 84
501 ± 45
A2933 2
26 145 ± 27
65 ± 25
A2933 3
28 873 ± −
–
A2440 1
27 080 ± 96
889 ± 62
A2440 2
28 058 ± 25
33 ± 25
A2440 3
30 360 ± −
–
A2384 1 29 760 ± 122
365 ± 75
A2384 2
28 546 ± 55
200 ± 27
A2384 3 27 312 ± 168 654 ± 135
Ng Allocation rate
39
0.984
6
0.999
2
1.0
96
0.95
5
0.30
2
1.0
13
0.968
20
0.974
23
0.833
Fig. 4. Velocity histogram of A2933 (binning of 250 km s−1 ). The best
Gaussian fit for the whole distribution (dotted line) is centered on the
vertical solid line which gives the location value. Location and scale
of the Gaussian (i.e. mean velocity and velocity dispersion) were estimated with ROSTAT. We also show the two Gaussian functions corresponding to partition 1 and 2 of the best mixture 3 partitions fit by
EMMIX (dashed lines) and the composite function (solid line).
5
0
cz [km/s]
Notes. Errors have been estimated with ROSTAT.
mean V = 27 364 km s−1 and dispersion σ = 501 km s−1 (partition 1), the sharp peak at lower velocity with a Gaussian having mean V ∼ 26 145 km s−1 and a very low dispersion σ =
65 km s−1 (partition 2), and the small excess at high velocities
(partition 3). Results are listed in Table 3. The sharp low velocity excess consists of six galaxies: all but one are located in the
northern subcluster (purple squares in Fig. 3). These galaxies
are likely to form a low velocity group probably infalling into
A2933N. In the following, we exclude partition 3 (with only 2
objects) from the analysis. No significant results are obtained
with EMMIX when fitting with mixture with a higher partitioning to the data.
A79, page 6 of 19
Fig. 5. Velocity histograms of A2933N (blue) and A2933S (red). For
each subcluster we included galaxies within a radius of 3.5′ from
its X-ray center. Solid lines: velocities of the brightest galaxy in the
Northern (blue) and Southern (red) subclusters.
4.4. X-ray/optical combined analysis
The projected density maps of A2933 for galaxies and for gas
provide similar views of the cluster (Fig. 2): the X-ray and optical isocontours of the NW and SE subclusters, A2933N and
A2933S, closely correspond and are centered on the cold cores
in the temperature map (Fig. 2). The X-ray contours are roughly
centered on the BCGs for both the NW and SE subclusters.
These results suggest that the system is in a pre-merger phase.
S. Maurogordato et al.: Merging history of three bimodal clusters
However, in the case of the NW subcluster, the BCG seems
slightly offset eastwards with respect to the X-ray centroid. The
inner X-ray isocontours and the temperature map are elongated
along an E-W axis while the outer isocontours are orientated
along the general SE-NW axis also connecting the two brightest
galaxies of A2933N.
As the density distribution of galaxies shows a clear bimodal
behavior, we analyzed separately the velocity distributions of
A2933N and A2933S, selecting the galaxies within a radius of
3.5′ from each subcluster center (see Fig. 3). The corresponding
histograms are displayed in Fig. 5. The velocity distributions of
the two subclusters are offset from each other, the NW subcluster lying at lower velocities than the SE subcluster, but there is
significant overlap. Moreover, they are quite different. A2933N
has a multi-peak structure with the sharp excess at low velocities
previously identified with EMMIX, while A2933S has a more
continuous distribution with an extended low velocity tail that
falls in the velocity range of A2933N.
Departure from normality was tested when analyzing both
subclusters individually, and neither the normality tests nor the
Dip test allowed us to significantly exclude the hypothesis of
a Gaussian velocity distribution. The ROSTAT analysis of the
two velocity distributions shows that the scales (velocity dispersions) are comparable for both subclusters. These scales are in
good agreement with those expected from the measured X-ray
temperatures, assuming the typical scaling relation between σ
and kT (Lubin & Bahcall 1993; Girardi et al. 1998; Wu et al.
1998, see Fig. 6), suggesting that gas and galaxies are in equilibrium.
The dynamical analysis also confirms a significant offset for locations (mean velocities) of ∼730 ± 210 km s−1 :
A2933S has a significantly higher mean velocity than A2933N
(∼27 648 km s−1 vs. ∼26 919 km s−1 ). This trend is also followed
by the velocities of the brightest cluster members: the BCG in
A2933N (BCG1) and its close companion have comparable velocities (27 350 km s−1 and 27 424 km s−1 ); the second brightest
galaxy has a lower velocity (26 951 km s−1 ), while the brightest
galaxy in A2933S (BCG2) has a higher velocity (28 028 km s−1 ).
This implies that there is a significant offset of ∼600±140 km s−1
between BCG1 and BCG2.
The velocities of the two BCGs also differ from the mean
velocity of their host subcluster by 505 ± 165 km s−1 and 380 ±
200 km s−1 for A2933N and A2933S, respectively, which is statistically significant for A2933N.
However, they coincide with the locations of the highest peaks in the velocity histograms of their respective subclusters A2933N and A2933S (excluding the sharp peak at
∼26 200 km s−1 ), and these peaks are slightly shifted from the
mean velocity of each subcluster, due to the skewness of the velocity distributions (Fig. 5). All these results suggest that the velocity distributions have started to mix with each other, but that
the structure of the two subclusters is still not strongly affected;
this implies that the clusters have not yet crossed each other.
Therefore the combined X-ray-optical analysis indicates that
A2933N and A2933S are in an advanced pre-merger stage. The
presence of two very bright galaxies along an NE-SW axis, the
elongations of the inner X-ray contours and the temperature map
along the E-W direction of A2933N, and the existence of a distinct low velocity, low dispersion velocity component, suggest
that A2933N has also undergone some previous merging and is
still accreting a small group.
Fig. 6. The σ − T X relation for the subclusters A2933N and A2933S
(red), A2440B and A2440C (green), A2384N and A2384S(blue). The
straight line is the relation σ = 102.47±0.08 T 0.67±0.09 (Wu et al. 1998).
4.5. Dynamical analysis
We now provide mass estimates for the subclusters and the total mass of A2933 (analogous sections are devoted to A2440
and A2384). For each system, we discuss the uncertainties in
the mass estimates. An obvious uncertainty is related to the assumption of dynamical or hydrostatic equilibrium. For example, velocity dispersions might not simply reflect the subcluster
potential and might be overestimated. Another uncertainty is due
to the limited field in which it is possible to estimate the harmonic radius: the circle within which we estimate this radius
cannot intersect the nearby subcluster(s). As a consequence, the
virial radius might be underestimated. The mass may also be underestimated due to the pressure term.
4.5.1. Mass estimates
From the previous analysis, A2933 is very likely in a pre-merger
stage, in which subclusters begin to interact but cores have not
yet crossed each other. We may then assume that the states of the
two subclusters do not strongly deviate from dynamical equilibrium, so that the virial theorem and the hydrostatic equilibrium
can be safely applied to determine optical and X-ray mass estimates. These estimates will be used in the following to constrain
the parameters of the collision by means of a two-body analysis.
In the optical, we estimated the harmonic radius of the two
subclusters from the projected distribution of all galaxies with
R < 20 and within a radius of 3.5 arcmin from the BCG; from
the harmonic radius we derived the virial radius and, in combination with the velocity dispersion, we estimated the virial
mass. Finally, extrapolating the density profile, we estimated the
mass M200 corresponding to a density contrast ρ̄/ρc = 200 (for
more details see e.g. Maurogordato et al. 2008). We find for
A2933N M200 = 2.0 ± 0.4 × 1014 M⊙ and for A2933S M200 =
(2.2 ± 0.4) × 1014 M⊙ , with a ratio MA2933S /MA2933N = 1.1 ± 0.6.
We can also estimate the mass ratio using the X-ray scaling
relations. The mean X-ray temperatures of A2933N and A2933S
A79, page 7 of 19
A&A 525, A79 (2011)
are respectively kT A2933N = 2.11±0.25 keV and kT A2933S = 2.7±
0.2 keV. Assuming M ∝ T X3/2 , we obtain a mass ratio between
the two components around 1.5:1.
On the other hand, when we estimate masses using the
brightness and temperature profiles of each unit, assuming
hydrostatic equilibrium for each subcluster, we find that
MA2933N = 2.6±0.4×1014 M⊙ and MA2933S = 2.76±1.5×1014 M⊙
MA2933S
= 1.2 ± 0.6.
leading to a mass ratio of M
A2933N
While the exact values could be systematically affected by
deviations from equilibrium, there is a good consistency between
the optical and X-ray mass estimates, and we can conclude that
the two subclusters have comparable masses, with a total mass
for A2933 of M200 ∼ 5 × 1014 M⊙ .
BIa
BIb
BO
4.5.2. Two-body analysis
We applied the two-body dynamical formalism (Beers et al.
1982; Gregory & Thomson 1984) to the two subclusters in
A2933. One of the variants of this method establishes a relationship between the total mass M of the bimodal cluster and the
angle α between the plane of the sky and the collision axis of
the two clumps (Barrena et al. 2002; Ferrari et al. 2005). The input parameters derived from the observations are the relative radial velocity between the two subclusters Vr and their projected
spatial separation Rp . For these observables, we chose to take
the values of the projected spatial separation and the radial relative velocity (at rest) between the two BCGs (Rp = 0.75 Mpc
and Vr = 550 km s−1 ), which are more clearly defined than that
of the subclusters. We obtained similar results when using values from the subcluster centroids. In addition, the formalism requires us to define the time t0 that has elapsed since the epoch of
the last encounter between the two subclusters. We tested different scenarios, including a pre-merging scenario (t0 ∼ 12.2 Gyr,
i.e. the age of the Universe at the redshift of A2933, implying
that the two systems have never yet crossed each other), a recent
(t0 = 0.5 Gyr) and an older (t0 = 1.0 Gyr) post-merger event.
Considering the total mass of the cluster derived from optical
observations (5 × 1014 M⊙ ), the post-merger cases provide only
one possible bound outgoing solution (i.e. the two subclusters
are moving apart). In the pre-merger case, Fig. 7 reveals two possible bound incoming solutions, the first with α ∼ 15◦ (BIa) and
the second with α ∼ 75◦ (BIb), and a bound outgoing solution
(BO). For BIa, the “real” separation between the two subclusters
appears to be small (∼0.8 Mpc) and the relative velocity high
(V ∼ 2000 km s−1 ), while for BIb the two components appears
to be more distant (∼3.5 Mpc) and the relative velocity comparable to the projected observed one (V ∼ 560 km s−1 ). The results
are listed in Table 4.
4.6. The scenario for A2933
The combined X-ray optical analysis has shown: a) two subclusters, visible in both X-ray and optical maps, and associated with
cool cores detected in the temperature maps; b) a hot region in
between the two subclusters; c) a perturbed velocity distribution,
showing a velocity offset (at rest) of ∼700 km s−1 between the
two subclusters. This suggests that the interaction between the
two subclusters is in its initial phase. This lead us to exclude
the post-merger cases in the two-body analysis, and to examine
the pre-merger one. Solutions with large values of α: BIb, and
UO are quite improbable since they would imply large spatial
separations between the subclusters (∼3.5 Mpc and 11 Mpc, respectively), which is unlikely taken into account the clear signs
A79, page 8 of 19
UO
0
20
40
60
80
Fig. 7. The sum of the virial masses of A2933N and A2933S as a function of the projection angle α. The horizontal lines show the mass estimate (full line) and errors (dashed lines) (see Sect. 4.5.1). The projected
distance of 0.75 Mpc and radial velocity difference at rest of 550 km s−1
are derived from the two BCG’s location and velocities. The two systems were at zero separation 12.2 Gyr ago. The dotted line represents
the Newtonian criterion for gravitational binding. Two bound incoming
(BIa and BIb) and one bound outgoing (BO) are found compatible with
our mass estimates.
of interaction between the two merging units. The most likely
solution for A2933 is thus (BIa), i.e. a two-body pre-merger
nearly on the plane of the sky (α ∼ 15◦ ). The high relative velocity (∼2000 km s−1 ) between the two components implied by
this solution is a common feature in merging clusters.
5. A2440
Mohr et al. (1996) performed a combined X-ray and optical
analysis of A2440, based on optical imaging and spectroscopy
(48 redshifts), and Einstein X-ray data. Our work significantly
extends this study, being based on a larger spectroscopic sample
(more than double the total number of redshifts) and new X-ray
data, including the gas temperature maps.
5.1. X-ray gas morphology and thermal structures
The X-ray observations detect a gas emission with an elongated
structure connecting two brightness peaks (B) and (C) to a northern and a less luminous component (A) (Fig. 8). The two brightness peaks are associated with cool cores (kT ≃ 2.8−3.5 keV),
and are surrounded by hotter gas (kT ≃ 5 keV); a significant
estimate of the temperature of the (A) component is not possible due to the strong contamination from the innermost cluster
regions.
The southern emission peak (C) is delimited by gas brightness and temperature jumps across the white sector shown in
Fig. 8. As discussed above, these discontinuities are signatures
of a cold front delimiting the edge of the southern cold core.
To investigate the nature of the gas density jump observable
across the white sector in Fig. 8, we model the gas 3D structure
S. Maurogordato et al.: Merging history of three bimodal clusters
Table 4. Two-body model solutions for the A2933N-A2933S, A2440A-A2440(B+C), and A2384N-A2384S systems.
Cluster
A2933N/S
t0
12.2 Gyr
A2933N/S
A2933N/S
A2440A/B+C
0.5 Gyr
1.0 Gyr
12.2 Gyr
A2384N/S
12.2 Gyr
A2384N/S
A2384N/S
A2384N/S
0.5 Gyr
1.0 Gyr
2.0 Gyr
Solution
BIa
BIb
BO
BO
BO
BIa
BIb
BO
BIa
BIb
BO
BO
BO
BIa
BIb
α(deg)
15
78
86
47
65
11.37
82.4
87.5
17
77
46
65
75
25
53
R(Mpc)
0.8
3.5
10.9
1.1
1.8
0.61
4.5
14.3
1.2
5.0
1.6
2.6
4.4
1.2
1.9
Rm (Mpc)
4.1
4.7
47.
1.5
2.5
5.3
6.1
75
6.5
7.3
2.0
3.2
6.0
2.2
2.3
V(km s−1 )
2120
560
551
1055
862
3700
740
730
3440
1030
1390
1110
1030
2354
1240
Notes. Various possibilities are considered for the evolutionary phase: the pre-merger case (t0 = 12.2 Gyr), and post-mergers seen at t0 after
the first passage. For each solution, we derive the angle α between the line connecting the two components and the plane of the sky, the spatial
separation of the subclusters R, their separation at maximum expansion Rm and their relative velocity V.
Fig. 8. A2440: left: galaxy density map (mag. cuts: R < 19) overlaid on the X-ray luminosity contours (EPIC-XMM counts in the .5−2.5 keV
band corrected for background and vignetting). Black crosses indicate the BCG positions. Right: ICM temperature map overlaid on the X-ray
luminosity contours (EPIC-XMM data analyzed through wavelet spectral-imaging, see Bourdin et al. 2004, 2008, for details).
with disrupted density and temperature profiles (see Eqs. (20)
and (21) of Bourdin et al. 2008). As shown in Fig. 10, fitting
the slope, jump position, and amplitude of these functions reveals a discontinuity in gas density, Dn = 1.82 ± 0.06, and 3D
temperature, DT = 1.76+0.08
−0.06 , located at 110 kpc to the south of
the cool core. The pressure continuity that is measurable across
the jump (DP = DT /Dn = 0.97 ± 0.05) identifies the discontinuity as a cold front. As already observed in various interacting
systems, this cold front is likely to delimit a stripped cold core
moving outwards from the hotter embedding ICM.
5.2. Galaxy density distribution
The galaxy density maps computed for RS galaxies for different magnitude cuts (Fig. 1) show that the general structure of
A2440 is strongly elongated along a NE/SW axis and includes
multiple clumps. At bright magnitudes (R < 18), four groups
are detected: the two most significant subclusters B and C are
identified as the main cluster, and two fainter ones are detected
at the NE extent: A1 and A2. When including faint galaxies,
the two components corresponding to the main cluster B and
C progressively merge to appear as a single dominant elongated
structure (R < 21), while the NE groups A1 and A2 are less
visible and a new (but faint) component A3 is detected on the
western side.
The three brightest galaxies are located near the three density
peaks A1, B, and C identified at R < 18 and R < 19 (Fig. 11).
There is a strong alignment between the main axis of the
bright galaxies, the position angle of the subclusters and the
NE/SW axis of the whole cluster.
A79, page 9 of 19
A&A 525, A79 (2011)
5.3. Galaxy velocity distribution
The first analyses of the dynamics of A2440 were performed by
Beers et al. (1991) and Mohr et al. (1996). We now have a much
larger dataset: from our EFOSC2 observations, 97 redshifts have
been obtained in the 30′ by 30′ field centered on A2440. We
merged this catalog with redshifts available in the literature,
which led to the final redshift catalog of 150 objects (with
10 twice observed objects). A comparison with 10 measured
redshifts available in the literature gives a mean difference of
30 km s−1 and a standard deviation of 60 km s−1 . Restricting this
sample to cluster members, we are left with 103 objects. The
velocity histogram is shown in Fig. 12.
In Fig. 11, cluster members are indicated by purple squares
and orange circles when belonging to one of the two major velocity peaks in the histogram at ∼27 000 km s−1 and
28 000 km s−1 respectively. One can see that galaxies belonging
to both velocity peaks populate subclusters B and C, while all
galaxies in subcluster A1 belong to the highest velocity peak.
Measurements of location and scale of the velocity distribution with the biweight estimator give CBI = 27 251 ± 93 km s−1
and S BI = 940 ± 70 km s−1 . While none of the 10 ROSTAT normality tests identify significant deviations from Gaussianity, the
Dip test rejects unimodality at the 1 percent level (Table 2).
EMMIX fits a mixture with two partitions isolating the main
component and an excess of two galaxies at V ∼ 28 058 km s−1
with a very significant P-value (0.01). With three partitions,
EMMIX isolates the main component with mean velocity V ∼
27 080 km s−1 and velocity dispersion σ = 890 km s−1 (partition 1), a very sharp Gaussian (σ ∼ 33 km s−1 ), corresponding to the previously mentioned excess with mean velocity
V ∼ 28 058 km s−1 (partition 2), and a group of two galaxies
at ∼30 337 km s−1 (partition 3), which are not considered in the
following. The best-fit Gaussian functions (partition 1 and 2) are
plotted in Fig. 12. However, the relatively high value of Pvalue
obtained (0.23) indicates that this three partition fit is less significant than the two partition one. No significant results are
obtained with EMMIX when fitting with a larger number of
partitions.
5.4. A2440: X-ray/optical combined analysis
Examining Fig. 8, where we show the X-ray isocontours and
the galaxy projected density distribution, we see that the central
X-ray bimodal structure (B+C) is nearly coincident with the optical bimodal structure (B+C) in the density map at R < 19 (and
coincident with that at R < 18). The X-ray A component is coincident with the NE subcluster A1 identified in the galaxy density maps, while the subclusters A2 and A3 are not detected in
X-rays. The three X-ray maxima A, B, and C are well centered
on the three corresponding BCGs (Fig. 8). We have seen that at
fainter magnitudes the optical bimodal structure merges into one
structure: its density peak is centered between the X-ray maxima B and C. These three subclusters, identified in X-rays and
with optical counterparts, are referred to as A2440A, A2440B,
and A2440C.
To analyze the velocity distribution in A2440A, A2440B,
and A2440C, we defined three subsamples including galaxies
within circular regions of radius 2.5 arcmin (the largest compatible with negligible overlap) and centered on the corresponding
X-ray maxima, which coincide with the BCGs positions. These
circular regions are shown in Fig. 11.
In Fig. 13, we show the corresponding velocity histograms
with different colors (blue, green, and red for A2440A, A2440B,
A79, page 10 of 19
and A2440C, respectively). The histograms indicate that most
galaxies in A are in the velocity bin at ∼28 000 km s−1 (EMMIX
partition 2). Within the limits of the small number of counts per
bin, there is no evidence of another significant segregation in
velocity for galaxies in the regions B and C: the velocity distribution seems spatially mixed.
The results of the ROSTAT analysis for the regions corresponding to the three X-ray subclusters are listed in Table 2.
Both the mean velocity and velocity dispersions of the B and
C regions are comparable, and consistent with the global cluster
values. This suggests two possible alternatives: either A2440B
and A2440C have not yet interacted, and the merging is occurring in the plane of the sky, or they have already crossed, but
are seen just after core passage. The velocity of the BCG in
the B region, BG2, is close to the velocity of its host subcluster, while BG3 in the C region has a (∼2σ) velocity offset of
(418 ± 188 km s−1 ). The velocity dispersions of the two main
components A2440B and A2440C are higher than the values expected from their X-ray temperature (see Fig. 6), but the deviation from the σ − T X relation is not statistically significant.
The mean velocity of A2440A is ≈750 km s−1 higher than
that of the main cluster at a 3σ level. The small velocity dispersion of A is more typical of a group. The velocity of the brightest
galaxy in A, VG1 = 27 925 km s−1 , is consistent with the mean
velocity of its subcluster, implying that it is at rest in the potential well. This dynamical analysis is in very good agreement with
that obtained by Beers et al. (1991) and Mohr et al. (1996).
These results show that A2440A is a group at higher velocity with respect to the main system, which includes the two subclusters A2440B and A2440C: these two components must have
recently crossed each other or are close to merging in the plane
of the sky.
We also applied the normality tests to the three subclusters.
In the case of the A subcluster, the 10 ROSTAT tests and Dip
test exclude normality at more than the 10% level. This result
remains unchanged when excluding a low velocity galaxy that is
3000 km s−1 offset from the major velocity peak. We note however that the analysis of this subcluster relies on only 11 (10) velocities. For subclusters B and C, all the normality tests (except
for one in the case of B) and the Dip test are consistent with a
unimodal Gaussian distribution.
5.5. Dynamical analysis
5.5.1. Mass estimates
We estimated the mass of the three components in A2440 following the same method as for A2933. However, to determine the
harmonic radius we selected the galaxies in the red sequence, defined using our B and R photometric catalogues, with R < 20; in
this way, background contamination was minimized. For each
component, we selected galaxies satisfying the above criteria
within a radius of 3.0 arcmin from the respective BCG.
As a caveat, we emphasize that we consider subclusters potentially in a post-merger phase, in which case one expects important distortions in morphology and velocity field. However,
previous analyses have shown that the velocity distributions in
both components B and C can be assumed to be Gaussian. We
therefore assume that the two main components, A2440B and
A2440C, are now not far from dynamical equilibrium. In contrast, A2440A has a non-Gaussian velocity distribution, but this
was determined using only a few redshifts. However, its characteristics are similar to those of a group, and its contribution to
the whole mass should be a priori negligible.
S. Maurogordato et al.: Merging history of three bimodal clusters
We find for A2440A, A2440B, and A2440C, respectively,
that M200 = 0.1 ± 0.02 × 1014 M⊙ , M200 = 5.6 ± 1.1 × 1014 M⊙ ,
and M200 = 5.4 ± 1.1 × 1014 M⊙ . As previously discussed, B and
C appear to be the main subclusters; the estimated total mass of
A2440 is MA2440 ∼ 1.2 × 1015 M⊙ .
5.5.2. A2440: two-body model
The two-body model was applied to the central system composed of A2440B and A2440C, with input parameters Rp =
0.442 Mpc and Vr = 200 km s−1 derived from the BCGs. As
expected in terms of both projected separation and velocities,
the two-body model is not well constrained by the observations. Bound solutions exist in both the pre-merger case with
t0 = 12.2 Gyr (two incoming and one outgoing), and in the postmerger case (one outgoing for t0 < 0.5 Gyr, two incoming, and
one outgoing for t0 > 0.5 Gyr). Therefore, according to the twobody analysis, A2440B and A2440C may be systems seen either
before or after first encounter.
We also tested the two-body system composed of A2440A
and the whole complex A2440(B+C). To estimate Rp and Vr , we
used the position and velocity of the centroid of the whole system A2440(B+C), obtaining Rp = 0.6 Mpc and Vr = 730 km s−1 .
In the pre-merger hypothesis (t0 = 12.2 Gyr), there are two
bound incoming (BIa, BIb), and one bound outgoing (UO) solutions (Fig. 14). If seen after first passage, one bound outgoing
solution exists (BO). The different solutions are listed in Table 4.
5.6. The scenario for A2440
We can now use the complete set of information collected from
our analysis of A2440 to determine the merging scenario. The
bimodal X-ray emissivity of the main component is evidence of
the interaction of two massive subclusters being likely to undergo or to have undergone a merging event. At first sight, the
existence of two units with cool cores, and a close correspondence between the gas and bright galaxy distributions, which are
well centered on the BCGs, are indicative of a pre-merger event.
This was also the conclusion of Mohr et al. (1996) for A2440,
who observed the similarity between galaxy projected density
maps and X-ray emission maps from Einstein.
However the pre-merger hypothesis is strongly disfavored by
the lack of any strong temperature enhancement between the two
maxima, as we would expect in a pre-merger event (RS01). The
similar velocity distributions of the two subclusters can be understood in terms of both a pre-merger and post-merger, but with
different implications: in the pre-merger case it is indicative of
a merging in the plane of the sky, while in the post-merger case
it is due to a very recent collision. The observed difference of
clustering for faint and bright galaxies, with a good coincidence
between the gas and bright galaxies centroids for A2440-B and
A2440-C subclusters, but a segregation between the centroids of
the gas and faint galaxies distributions also indicate that merging
has already occurred. These results have already been found for
several post-merger clusters (Biviano et al. 1996; Maurogordato
et al. 2008).
The evidence of a cold front delimiting the core of the
Southern subcluster also favors a post-merger scenario, where
the stripped subcluster is now moving away from its companion, as shown for instance in A1201 by Owers et al. (2009).
The velocity offset of BG2 with respect to A2440-C also implies
that there is some dynamical activity in this region. To make
this scenario compatible with the observed bimodal structure,
we argue that the collision must have occurred with a relatively
high impact parameter because a too small impact parameter
would have led to the disruption of the two former subcluster
cores during the violent relaxation of the gas. The similar values obtained for the temperatures and velocity dispersions of
A2440B and A2440C suggests that these components have comparable masses. If one identifies the hot structures surrounding
A2440B and A2440C as remnants of a shock wave, a scenario
with an equal mass collision with a high impact parameter after the maximum core collapse, when the shock wave begins to
propagate towards the outskirts of the new structure, is therefore
very likely to explain the A2440(B+C) complex.
The A2440A subcluster is probably a group, with a significant velocity offset (∼750 km s−1 ) with respect to the main cluster A2440(B+C). The coincidence of the BCG position with the
maximum of the gas and galaxy distribution and the consistency
of the BCG velocity with the subcluster mean velocity, suggest
that the BCG is at the center of the subcluster in dynamical equilibrium. We should expect a Gaussian velocity distribution for
this subcluster, which is not the case: this may be due to the
small number of available redshifts (10) and contamination from
galaxies not belonging to A2440A.
The lack of signatures of gas compression between A2440A
and A2440B excludes a pre-merger hypothesis implying small
physical separations (BIa). The two viable pre-merger solutions
are then incoming (BIb) or outgoing (BO), both with a major
component along the line of sight (α ∼ 80 degrees). The postmerger case remains unlikely, as there is a good correspondence
between the gas and galaxy distributions and the X-ray temperature map does not show any evidence of shocked gas in the
periphery of A2440A. Therefore A2440A is probably a group
infalling for the first time into the main cluster component along
the NE/SW axis.
6. A2384
6.1. X-ray gas morphology and thermal structures
Looking at Fig. 9, it is clear that the X-ray morphology of A2384
is very peculiar. The general shape of the whole cluster is very
elongated, including a northern primary maximum, connected
to a southern secondary maximum by a continuous gas distribution. The temperature of this gas is around 3−4 keV, while its
surroundings are hotter (up to 4−5 keV). Both subclusters exhibit cool cores.
6.2. Galaxy density distribution
A2384 is one of the best examples of colinear distribution of
substructures in a cluster (see West et al. 1995). At all magnitude
limits, the galaxy distribution shows a very elongated structure,
extending over 1.2 Mpc along the N-S axis (Fig. 1); embedded
in this structure, there are two main subclusters roughly centered
on two bright galaxies (which we call BCG1 and BCG2 in the
following; see Fig. 15). The northern subcluster is more densely
populated than the southern one. At bright magnitudes (R < 18),
the galaxy distribution is strongly concentrated in the northern
(A2384N) component. Besides BCG1, in A2384N there are two
other very bright galaxies, while in A2384S there are no bright
galaxies around BCG2. Near A2384N, a clump of galaxies to the
west causes an elongation in the density map. While A2384S is
very elongated along the general N/S axis of the system, the internal contours of A2384N are oriented towards the NW/SE, and
rotate to the W-E and general NE/SW direction on larger scales.
A79, page 11 of 19
A&A 525, A79 (2011)
Fig. 9. Same as Fig. 8 for A2384.
Fig. 10. Top: gas brightness and projected temperature profiles observable across the white sector in Fig. 8 for A2440. Bottom: ICM density and
pressure profile estimated above the Southern gas clump in A2440. These profiles reveal a density discontinuity at constant pressure or “cold front”
feature (see vertical dashed lines on the plots).
When including fainter objects, the isodensity contours of the
two components are elongated towards each other (see Fig. 1).
6.3. Galaxy velocity distribution
Before our observations, very few redshifts were available in the
literature for this cluster. In our 30′ × 30′ field, only 4 galaxies
A79, page 12 of 19
already had a redshift; we reobserved two of them, and our redshifts are in agreement with the literature values within the estimated errors. Our final catalog includes 84 redshifts, and the
number of cluster members is 56.
The mean location of the cluster, CBI = 28 263±154 km s−1 ,
is well defined, but the scale is quite high (SBI = 1114 ±
120 km s−1 ). The velocity distribution (Fig. 16) is very broad
S. Maurogordato et al.: Merging history of three bimodal clusters
5
0
cz [km/s]
Fig. 13. Velocity histograms of A2440A (blue), A2440B (green) and
A2440C (red). For each subcluster we included galaxies within a radius
of 2.5′ from its center. Dotted lines: velocities of the brightest galaxies
in A2440A (blue), A2440B (green) and A2440C (red).
Fig. 11. WFI R-band image of the A2440 field (12′ ×14′ ). Galaxies identified as cluster members from spectroscopy (Flag 0 and 1) are marked
with different symbols according to the velocity range (purple squares:
[24 000, 27 600] km s−1 ; orange circles [27 600, 30 000] km s−1 ). The
isocontours of red sequence galaxy density maps with magnitude limit
R < 19 are superimposed. The velocity histograms of the three regions
with 2.5 arcmin radius delimited by circles (corresponding to the three
subclusters A2440A, A2440B, and A2440C) are displayed in Fig. 13.
North is up and East is to the left.
BIa
BIb
BO
UO
15
0
10
20
40
60
80
Fig. 14. Two-body model for the system A2440A/A2440(B+C). The
horizontal lines show the mass estimate (full line) and errors (dashed
lines) (see Sect. 5.5.1). The projected distance of 0.6 Mpc and the radial velocity difference at rest of 730 km s−1 were derived from the position and velocity of the two centroids. The two systems are assumed
to have been at zero separation 12.2 Gyr ago. The dotted line represents
the Newtonian criterion for gravitational binding. Two bound incoming (BIa and BIb) and one unbound outgoing (UO) solutions are found
compatible with the mass estimates.
5
0
cz [km/s]
Fig. 12. Velocity histogram of A2440 (binning of 250 km s−1 ). The best
Gaussian fit for the whole distribution (dotted line) is centered on the
vertical solid line which gives the location value. We also show the two
Gaussian functions (dashed lines) corresponding to partitions 1 and 2
in the best 3 partition mixture model by EMMIX and the composite
function (solid line).
and has three main peaks. This extended velocity distribution
could be either due to a very deep potential (massive system)
or multi-modality. The hypothesis of a massive system is improbable, given the low cluster optical luminosity and richness.
The battery of normality tests do not indicate significant departures from normality. The Dip test excludes unimodality with a
P-value of 0.1.
In Fig. 15, cluster members in the three velocity peaks are
identified with different symbols and colors. We have very few
A79, page 13 of 19
A&A 525, A79 (2011)
Fig. 15. WFI R-band image (20′ × 25′ ) centered on A2384. Galaxies identified as cluster members from spectroscopy (flags 0 and 1) are marked
with different symbols as in the previous cases (purple squares: [24 000, 28 000] km s−1 , cyan diamonds: [28 000, 29 000] km s−1 and orange circles:
[29 000, 31 000] km s−1 ). Isocontours corresponding to red sequence galaxy density maps at R < 19 are superimposed. The velocity histograms of
the three regions delimited by circles with 3 arcmin radius, which include the two subclusters and the intermediary region, are displayed in Fig. 18.
North is up and East is to the left.
redshifts in A2384S, because it is a poor structure. A striking
feature is the high fraction of objects in the filament joining
A2384N and A2384S, which are associated to the second velocity peak (∼28 500 km s−1 ).
For A2384, the best-fit solution was found by EMMIX for a
partition of three Gaussians roughly centered on the peaks visible in the velocity histogram, with a P-value of 0.16. This indicates that, while the fit quality is greatly enhanced by using 3
Gaussians instead of 1, the null hypothesis of Gaussianity is not
rejected. The three Gaussian functions are visualized in Fig. 16;
the sum of these functions reproduces the data quite well. The results are summarized in Table 3. The whole velocity distribution
can be reproduced by the combination of two Gaussians with intermediate velocity distributions (∼365 km s−1 and ∼665 km s−1
corresponding to partitions 1 and 3, respectively) at mean velocities ∼1200 km s−1 higher and lower than the mean velocity of
the whole cluster, and by a third strongly peaked component of
low dispersion, σ ∼ 250 km s−1 peaked at ∼28 500 km s−1 (corresponding to partition 2). In the following section, we discuss
the spatial distribution of the galaxies assigned to the different
velocity partitions with respect to the X-ray and optical density
distribution.
6.4. A2384: X-ray/optical combined analysis
Looking at Fig. 9, both the gas and the galaxy maps approximately define the same bimodal structure. The northern and
southern overdensities (A2384N and A2384S) correspond to the
cool cores in the X-ray temperature map. The X-ray maximum
of A2384S is coincident with the BCG2 position, while BCG1
A79, page 14 of 19
is slightly offset westwards (15 arcsec) with respect to the X-ray
maximum of A2384N (see Fig. 9). This offset cannot be due to
astrometric errors, as the precision of the optical images is superior to 1′′ , while the precision of X-ray images is superior to
4′′ , and there is a very good correspondence between the optical
and X-ray positions of point sources. As BCG1 also has a significant velocity offset with respect to A2384N (see below), this
indicates that it is not at rest in the subcluster potential. However,
the velocity distribution of A2384N has a large dispersion, and
at the cluster redshift 15 arcsec corresponds to a physical projected separation of about 25 kpc, which means that BCG1 is
close to the cluster center, which is probably included in its extended halo.
As already noted, the velocity dispersion of A2384 is unexpectedly high with respect to its optical luminosity and to the low
values of its X-ray luminosity and temperature. A cluster with
velocity dispersion ∼1000 km s−1 should have an X-ray temperature T ∼ 7 keV, according to the σ − T X relation (Fig. 6), which
is much higher than the observed value. Moreover, from a weak
lensing analysis Cypriano et al. (2004) fit a velocity dispersion of
737 ± 126 km s−1 or 797 ± 108 km s−1 depending on the assumed
profile (SIS and SIE, respectively). These velocity dispersions
corresponds to T ∼ 3.45 keV and T ∼ 4.02 keV, respectively,
which are consistent with the observed values in X-rays. This
suggests that the measured velocity dispersion is overestimated.
We analyzed the velocity distribution in A2384N and
A2384S, selecting galaxies in circles of radius 5 arcmin centered on the X-ray maxima. The velocity histograms in the corresponding regions are shown in Fig. 17. The mean velocity of
A2384S is ∼675 km s−1 higher than that of A2384N. The same
trend is followed by the BCGs of the two subclusters, which
S. Maurogordato et al.: Merging history of three bimodal clusters
15
10
10
5
5
0
0
cz [km/s]
cz [km/s]
−1
Fig. 16. Velocity histogram of A2384 (binning of 250 km s ). The best
Gaussian fit for the whole distribution (dotted line) is centered on the
vertical solid line which gives the location value. Location and scale of
the Gaussian were estimated with ROSTAT. We also show the three
Gaussian functions (dashed lines) corresponding to the best mixture
model by EMMIX and the composite function (solid line).
show an even larger offset of ∼1000 km s−1 . We also found
that the position of the BCG1 has a small offset with respect
to the centroid of A2384N, while the position of BCG2 corresponds to the centroid of A2384S. The velocity of BCG1 is
600 km s−1 lower than the mean velocity of galaxies in A2384N,
while the velocity of BCG2 is consistent with that of A2384S
(see Table 2).
The velocity dispersions of the two subclusters are also quite
large: ∼1200 km s−1 for A2384N and ∼900 km s−1 for A2384S.
Both subclusters are above the σ − T X relation; the deviation is
at the 3σ level for A2384N, but within the errors in the case of
A2384S. ROSTAT normality tests reject the Gaussian hypothesis for neither A2384-N, nor A2384S. The Dip-test however
excludes unimodality for A2384-N (A2384-S) with a P-value of
0.1 (0.01). This suggests that the subclusters are still dynamically perturbed.
We also defined three subsamples by selecting galaxies
within a radius of 3 arcmin from the centers of A2384N,
A2384S, and the intermediate region (see Fig. 18). The galaxies
in the northern component show a large spread in their velocity distribution, which is skewed towards lower values and has a
peak at ∼28 500 km s−1 , while galaxies in the southern and especially in the intermediate region are more strongly concentrated
in the same peak; this peak corresponds to EMMIX partition 2.
6.5. Dynamical analysis
6.5.1. Mass estimates
We applied a method similar to that applied to A2440, determining the harmonic radius of each component by selecting only the
galaxies in the red sequence, with R < 20 and within a radius of
5 arcmin from the respective BCG.
However, as we previously discussed, the velocity distribution of this system is not simple. Our mass estimates could therefore be severely affected. We find for A2384N and A2384S that
Fig. 17. Velocity histograms of A2384N (blue), and A2384S (red). For
each subcluster we included galaxies within a radius of 5′ from its center. Velocity bins of 500 km s−1 are used. Dashed lines: velocities of the
brightest galaxies in A2384N (blue) and A2384S (red).
5
0
cz [km/s]
Fig. 18. Velocity histograms of A2384N (blue), the intermediate region
(green) and A2384S (red). For each subcluster we included galaxies
within a radius of 3′ from its center. Velocity bins of 500 km s−1 are
used. Dashed lines: velocities of the brightest galaxies in A2384N (blue)
and A2384S (red).
M200 = 1.4 ± 0.3 × 1015 M⊙ and M200 = 0.6 ± 0.1 × 1015 M⊙ , respectively. As the velocity dispersion is probably overestimated,
we consider in the following M200 = 2.0 × 1015 M⊙ as an upper
bound to the sum of the mass of the two components.
6.5.2. Two-body model
We applied the two-body model to A2384N and A2384S, assuming as values of Rp and Vr the projected separation and relative
velocity at rest of the BCGs (Rp = 1.14 Mpc, Vr = 1000 km s−1 ).
We tested several scenarios: the pre-merger case (t0 = 12.2 Gyr),
A79, page 15 of 19
A&A 525, A79 (2011)
BO
0
20
40
60
80
Fig. 19. The sum of the virial masses of A2384N and A2384S as a function of the projection angle α. The horizontal lines show the mass value
(full line) and errors (dashed lines) as estimated in Sect. 6.4. The projected distance of 1.14 Mpc and the radial velocity difference at rest of
1000 km s−1 were derived from the two BCGs positions and velocities.
It is assumed that the two systems are 1.0 Gyr after the first passage. The
dotted line represents the Newtonian criterion for gravitational binding. One unbound outgoing (UO) is found compatible with our mass
estimates.
and a range of post-mergers (t0 = 0.2, 0.5, 1., 1.5 and 2 Gyr). In
the pre-merger case, two incoming bound solution and one unbound solutions were found. In the post-merger case, only an
unbound solution exists if the merging is recent (t0 < 0.5 Gyr),
and a bound outgoing solution exists if the merger is older
(t0 > 0.5 Gyr). Figure 19 shows the case of a post-merger seen
1.0 Gyr after the encounter for which a unique bound outgoing
solution exists. At t0 > 1.5 Gyr, two bound incoming solutions
and a bound outgoing solution are possible.
previous collision with A2384S or/and a secondary merger event
within A2384N itself.
At first sight, A2384S is more regular than A2384N. It has
a good spatial coincidence between its gas and galaxy distributions, coincidence between the position and velocity of the BCG
and the X-ray center and mean velocity of the subcluster, respectively. Its velocity distribution is very broad and non-Gaussian,
but this may be due to contamination by objects in the periphery
of A2384N.
This complex structure could be understood in a scenario in
which a low mass cluster of galaxies has crossed the environment of a more massive one and has been stripped of a large
fraction of its gas and galaxies. The “head” of the system would
then correspond to A2384S, and the tail of galaxies belonging to
the original cluster affected by A2384N would lie in the intermediary region.
We conclude that the most likely scenario for A2384 is a
post-merger between two unequal mass clusters. However, if the
merging has already occurred, we have to explain why the two
cool cores have remained unaffected by the merging process.
Numerical simulations (Poole et al. 2006) have shown that the
initial cool cores of the primary and secondary components survive the first core crossing, and disappear after the second pericentric passage.
It is however difficult to constrain the parameters of the scenario, as the mass of A2384 is not well determined. From the
2-body analysis, bound solutions exists only if the merger event
is older than 0.5 Gyr. Older mergers (1−1.5 Gyr) with larger
physical separations between the units are favoured by the lack
of signatures of compression in the X-ray gas (Table 4). For
mergers older than 1.5 Gyr, one can see that incoming solutions,
corresponding to second infall, are again possible. However,
these also predict small values of the real separation between
the two components, which are unlikely from the lack of signature of compression of the gas in the temperature maps, as
noted before. One is then left with the most probable hypothesis
of two bound outgoing subclusters seen more than 1.0 Gyr after
the first passage, with a collision near the line of sight. Figure 17
shows as an example the bound outgoing solution in a 1.0 Gyr
post-merger, which is in good agreement with both our X-ray
and optical observations.
6.5.3. A2384 proposed scenario
The bimodal structure of A2384 detected in both the galaxy
and gas density maps suggests that the two components either
are about to or have undergone a merger event. The presence
of two well-separated components with cold cores, coincident
in the X-ray and the optical, is often observed in the case of
pre-mergers.
However, the trail of gas and the bridge of galaxies extending from A2384N to A2384S are difficult to explain if the subclusters have not yet interacted. If A2384 were a pre-merger, its
temperature map should exhibit a compression region between
the two subclusters, which is not the case. The velocity distribution shows a large dispersion (σ ∼ 1200 km s−1 ), inconsistent
with the value of the temperature if it simply reflects the mass
of the system. This high value of the velocity dispersion and the
velocity distributions of the two components being mixed imply
that the merging has already occurred. A2384N shows several
signatures of previous dynamical activity: a slight segregation
between gas and galaxies, an offset of the position of the BCG
from to the peak of the X-ray emitting gas, and of its velocity
from the mean velocity of the subcluster. This suggests that the
dynamical state of A2384N is still disturbed possibly due to a
A79, page 16 of 19
7. “Idealized” simulation and XMM emulation
In the previous sections, we combined optical and X-ray observations to constrain the merging scenario for each one of our
three bimodal systems. To help define the scenario and test its
consistency, we tried to reproduce the observed X-ray properties with numerical simulations. Adopting an approach similar
to RS01, we placed two clusters in a box and allowed them to
evolve under gravity in an adiabatic framework. The initial parameters of the collision are inferred from the X-ray and optical
observational constraints.
To this aim, we developed a set of programs that allow us
to reproduce the observational process and analyze the simulated clusters in the same way as we analyze real ones,
adopting an approach similar to e.g. Gardini et al. (2004) for
XMM observations.
Using an emission plasma model, where gas pressure and
density are known everywhere, we can compute the emitted
X-ray photons as a function of telescope area and exposure time.
Fixing the direction of observation, we collect photons from the
different cells along the line of sight, taking into account that we
S. Maurogordato et al.: Merging history of three bimodal clusters
observe clusters in projection (as long as we can assume that the
thin plasma approach is valid).
The first step in developing a good “idealized” merger simulation is to check the way we implement the hydrostatic equilibrium in each (sub)cluster (hereafter “unit”) to be sure that,
during collision, what we see is really due to the collision and
not a spurious result of numerical errors in the initial units. Once
this is achieved (by allowing a unit to evolve alone in the middle
of the box for a very long time), we are able to build collision
simulations.
The total mass profile of our “perfect cluster” follows a clas1
sic NFW profile, ρdm (r) ∝ r(1+r/r
2 (Navarro et al. 1996), while
S)
the gas follows a profile described in Suto et al. (1998) adapted to
keep the gas mass fraction constant at large radii. The dark matter profile is derived from the difference between the total mass
profile and the gas. We note that the NFW profile is derived from
pure dark matter simulations, thus this approach is likely to be
safe as long as the gas mass fraction is not too high (thus not in
the central region of cooling-flow clusters). The velocity dispersion of dark matter particles is computed (as in RS01) using the
virial equation drd [ρdm (r)σ2 (r)] = − GM(r)
ρdm (r).
r2
After fixing the gas-to-mass ratio, the concentration parameter c200 = Mr200
, the mass ratio between units and the impact paS
rameter (defined as the minimal distance between the two initial
trajectories in unit of the NFW scaling radius rS ), we reconstruct
a pre-collision scenario by computing two units as described
above. We place these two units in a rather large box to avoid
any border effects. We compute the initial velocity considering
the pre-merger evolution as a free fall of the two units, exactly
in the way described in RS01. We then compute the evolution of
the system with the AMR hydro-NBody simulation RAMSES
(Teyssier et al. 2002). At each time step, for all the cells in the
simulated volume, we store the pressure, velocity, and density of
the gas, and position and velocity of dark matter particles. After
running these simulations, the following steps are the computation of the photons emitted by the simulated volume and their
ray-tracing using the EPIC/XMM instrumental response. These
steps are fully described in Bourdin et al. (2004). We finally have
a dataset that can be reduced in the same way as real XMM/EPIC
observations.
All these previously described steps are important for a correct comparison between observation and simulation. Because
of the complexity of this procedure, it is impossible to explore
the whole parameter space when determining the optimal collision parameters. By looking simultaneously at the brightness,
temperature maps of the X-ray gas, and galaxy density maps,
we first guess a set of possible parameters that we place into the
complete simulation stream and iterate manually to optimize the
set of parameters. Therefore, we do not pretend to fit the data, but
we attempt to obtain a scenario reproducing the main features of
the observed merging system.
8. Comparison with simulations and discussion
Taking as input the scenario inferred from observations for the
main merger event, we obtained density and temperature maps
as they would be observed with XMM, and compared them with
the real observations.
In the case of A2933, the most likely scenario is a pre-merger
of two components of comparable mass for which there is a
small angle between the collision axis and the plane of the sky.
We simulated an equal-mass merger in the plane of the sky with
an impact parameter set to zero, and allowed it to evolve with
Fig. 20. Simulated wavelet reconstructed temperature map with luminosity contours superimposed, derived from the simulation which best
reproduces the observed features of A2933.
time. The resulting temperature map and luminosity contours for
this scenario are presented in Fig. 20. The X-ray properties, in
particular the temperature contrast between the high temperature
region and the mean temperature of each unit and its relative angular distance, are closely reproduced by a pre-merger scenario
∼200 Myr before the core collapse.
For A2440, which is expected to be a post-merger case, we
simulated a variety of equal mass mergers with different impact
parameters and tried to find the best configuration reproducing
simultaneously the observed temperature and luminosity distributions. In Fig. 21, we present our best candidate. This scenario
corresponds to a collision with an impact parameter of 15 rs seen
450 Myr after the maximum core collapse. We managed to reproduce the most striking features revealed by the observations,
in particular the hot region surrounding the X-ray maximum of
each unit, and the temperature contrast between the hottest region and the colder ones. Since our code follows a purely adiabatic approach, we do not expect of course to reconstruct the
two cool cores.
In the case of A2384, we performed different trials allowing the impact parameter value to vary. If the impact parameter
is not zero, it should be large enough to allow the gas of the
smaller unit to follow the influence of its original potential well
after the collision (we note that this is also necessary to explain
the remaining “cool” core of each unit). At the same time, the
impact parameter should be small enough to enable the gas to
be stripped efficiently. In these limits, the gas is stripped along
a rather significant distance as observed. Simulations by Poole
et al. (2006) show a gas surface density similar to that of A2384,
in the case of a 3:1 mass ratio seen at the first apocentric passage of the secondary cluster, at ∼1.5 Gyr after the closest approach (see their Fig. 4 central panel). In these scenarios, it is
A79, page 17 of 19
A&A 525, A79 (2011)
Fig. 21. Simulated wavelet reconstructed temperature map with luminosity contours superimposed, derived from the simulation which best
reproduces the observed features of A2440.
nevertheless very difficult to keep the axial symmetry observed
in luminosity and in temperature maps and impossible to explain
that the BCG of the south unit still lies at the X-ray maximum.
On the other hand, in a null impact parameter collision, if a
small group encounters a heavier unit, its gas can escape significantly after the first maximum core collapse only if its central
density is very high compare to the central density of the massive object. Simulations by Poole et al. (2006) are able to strip
gas even with a null parameter thanks to their initial gas profile (S ∝ r1.1 ) profile, which leads to a high inner gas density
in the core. In contrast, if the small unit has a concentration parameter higher than the massive one (say by a factor 2) then the
gas appears sticked to its own unit even after the first maximum
core collapse. Unfortunately, we cannot simulate such a complex
scenario in the adiabatic scheme since we have no cooling thus
the inner temperature of the small unit increases strongly with
the concentration parameter, preventing the correct description
of the gas luminosity and temperature during the collision.
9. Conclusion
We have performed a combined X-ray/optical analysis of three
bimodal clusters at low redshift, selected as merger candidates
from the sample of Kolokotronis et al. (2001). We have confirmed the disturbed dynamical state of these clusters and reconstructed the general trends of their main merging scenario.
A2933 has been shown to most likely be an equal-mass advanced
pre-merger (∼200 Myr before the core collapse). A2440 is a recent equal-mass merger (∼450 Myr after core collapse) with another subcluster infalling along the main axis of the cluster, and
A2384 is the result of an older collision between two units of
mass ratio 1:3 (∼1.0−1.5 Gyr after core collapse).
A79, page 18 of 19
We emphasize the complementarity of the X-ray and optical
data in determining the parameters of the scenario (angle, massratios, epoch) that were refined by running simulations of idealized cluster collisions. Our analysis confirms the efficiency of
the selection based on the comparison of gas and density maps to
identify merging clusters when (as is frequently the case) information on temperature and velocity distribution is not available.
Moreover, the level of segregation between galaxies and gas provides some indication of the merging stage. This approach is far
more powerful than the density maps taken individually in either X-rays or the optical. From X-ray density maps only, the
three bimodal clusters presented in this paper would have been
expected to be in the very last phase before the first maximum
core collapse, while two of them are shown to be post-mergers
with surviving cores in the two units. The properties of the temperature maps are essential to assess the merging stage: the hot
region in between the two sub-clusters in A2933 is indicative of
a pre-merger, the spectacular hot structures in the periphery of
A2440 and A2384 have been identified as remnants of a postmerger shock wave, and the cold front detected in the S/W region of A2440 is consistent with a a post-merger. On the other
hand, redshift information is needed to understand the dynamics
of the system, determine the mass of the subclusters and the relative motion of the BCGs, and finally date the merger event and
constrain its axis using the two-body analysis.
Numerical simulations have allowed us to refine the collision parameters. Thanks to our XMM simulation tools, we correctly handle all projection and instrumental effects. Thus we can
rely on these “observed simulations” and select the most likely
scenario based on measured criteria such as distances, temperature, and luminosity. We have been able to reproduce the main
features of A2933 and A2440, simulating the collision of two
systems initially in complete hydrostatic equilibrium (with an
approach similar to RS01) and adding a complete set of programs to reproduce X-ray observations for simulations and obtain EPIC/XMM synthetic observations (with an approach similar to Gardini et al. 2004). The case of A2384 is too complex to
be described with our approach, but we have been able to place
constraints on its collision parameters.
The results obtained from the analysis of these clusters can
be combined to our previous work on other MUSIC clusters to
draw a more general picture of merging clusters properties. In
the three clusters studied in the present paper, we have shown
the existence of preferential axes following the general position angle of the cluster, as previously shown in other merging
clusters (Arnaud et al. 2000; Plionis et al. 2003; Maurogordato
et al. 2006). Subclusters generally host one or more BCGs. In
most cases, the different merging subclusters and their brightest BCGs are aligned along this direction. This corroborates the
property of preferential alignment of BCGs with their host cluster (Niederste-Ostholt et al. 2010) on the scale of subclusters. In
A2384, we have detected a spectacular filament of galaxies and
gas probably stripped from the colliding group along the merging axis.
Another interesting aspect is the position and motion of the
BCGs relative to their subclusters. We have found that their
angular coordinates generally correspond (within the errors) to
the X-ray centroids of their host subclusters (A2933S, A2440A,
A2440B, A2440C, A2384S) but in some cases there is a small
offset (A2933N, A2384N). Similarly, in some cases the radial
velocities of the BCGs are consistent with the mean velocity
of their respective subcluster (A2440A, A2440B, A2384S), but
in other cases they are offset by 300−500 km s−1 (A2933N,
A2933S, A2440C, A2384N). These offsets are signatures of
S. Maurogordato et al.: Merging history of three bimodal clusters
ongoing dynamical activity in the subclusters, as found in
A3921B (Ferrari et al. 2005) and in A2163A (Maurogordato
et al. 2008), which were both identified as recent mergers.
In an analogous way, the beginning of interaction between
A2933N and A2933S, the probable recent mergers occuring
within A2933N and A2384N, and the recent merger between
A2440B and A2440C can explain the observed BCG offsets.
We have also found that, in addition to the main merging
event, a large fraction of MUSIC clusters appear to contain secondary merging events (A2933N and A2384N) and infalling
groups, e.g., A2440 (this work), A521 (Ferrari et al. 2003), and
A2163 (Maurogordato et al. 2008).
These results are consistent with a hierarchical scenario of
structure formation where clusters form by successive mergers
and accretion of matter along large-scale filaments, which may
cause alignments with structures on various scales (Basilakos
et al. 2006; Lee & Evrard 2007; Faltenbacher et al. 2008).
Moreover, we have also shown that in recent post-merger
clusters, such as A2440 (this work) and A2163 (Maurogordato
et al. 2008), galaxies exhibit a strong luminosity segregation,
similar to the case of the Coma cluster studied by Biviano et al.
(1996). We have also found a deviation from the σ − T X relation for the subclusters in a post-merger stage, the largest one
(at the 3σ level) being detected for A2384N. At variance, both
pre-merging subclusters A2933N and A2933S follow the σ − T X
relation.
The impact of the merging process on galaxy properties in
these clusters, in particular on star formation, will be addressed
in a forthcoming paper.
To test the validity of these properties on a large sample, we
are currently extending this work to a subsample of the C4 SDSS
cluster sample (Miller et al. 2004) with both optical and X-ray
available data.
Acknowledgements. The authors want to thank Romain Teyssier for the use of
RAMSES Hydro-NBody code and Albert Bijaoui and Eric Slezak for providing their program estimating density maps through a multi-scale approach, and
Monique Arnaud for fruitful discussions. We thank the Programme National de
Cosmologie et Galaxies of CNRS for his constant support on this program, the
Observatoire de la Côte d’Azur and the Laboratoire Cassiopée, CNRS, for specific funding of this project. H.B. acknowledge financial support from contract
ASI-INAF I/088/06/0. We also want to thanks an anonymous referee for his/her
comments and suggestions which helped us to improve the quality of the paper.
References
Arnaud, M., Maurogordato, S., Slezak, E., & Rho, J. 2000, A&A 355, 461
Barrena, R., Biviano, A., Ramella, M., Falco, E., & Seitz, S. 2002, A&A, 386,
816
Barrena, R., Boschin, W., Girardi, M., & Spolaor, M. 2007, A&A, 469, 861
Basilakos, S., Plionis, M., Yepes, G., et al. 2006, MNRAS, 365, 2, 539
Beers, T. C., Geller, M. J., & Huchra, J. P. 1982, ApJ, 257, 23
Beers, T. C., Flynn, K., & Gebhardt, K. 1990, AJ, 100, 32
Beers, T. C., Forman, W., Huchra, J. P., Jones, C., & Gebhardt, K. 1991, AJ, 102,
1581
Belsole, E., Pratt, G. W., Sauvageot, J.-L., & Bourdin, H. 2004, A&A, 415, 821
Belsole, E., Sauvageot, J.-L., Pratt, G. W., & Bourdin, H. 2005, A&A, 430, 385
Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393
Biviano, A., Durret, F., Gerbal, D., et al. 1996, A&A, 311, 95
Boschin, W., Girardi, M., Barrena, R., et al. 2004, A&A, 416, 839
Bourdin, H., & Mazzotta, P. 2008, A&A, 479, 307
Bourdin, H., Sauvageot, J.-L., Slezak, E., Bijaoui, A., & Teyssier, R. 2004, A&A,
414, 429
Busha, M. T., Evrard, A. E., Adams, F. C., & Wechsler, R. H. 2005, MNRAS,
363, L11
Coenda, V., Muriel, H., Donzelli, C. J., et al. 2006, AJ, 131, 1989
Cypriano, E. S., Sodré, L. Jr., Kneib, J.-P., & Campusano, L. E. 2004, ApJ, 613,
95
Donnelly, R. H., Forman, W., Jones, C., et al. 2001, ApJ, 562, 254
Faltenbacher, A., Jing, Y. P., Li, C., et al. 2008, ApJ, 675, 146
Ferrari, C., Maurogordato, S., Cappi, A., & Benoist, C. 2003, A&A, 399, 813
Ferrari, C., Benoist, C., Maurogordato, S., Cappi, A., & Slezak, E. 2005, A&A,
430, 19
Gardini, A., Rasia, E., Mazzotta, P., et al. 2004, MNRAS, 351, 505
Girardi, M., Giuricin, G., Mardirossian, F., Mezzetti, M., & Boschin, W. 1998,
ApJ, 505, 74
Hartigan, J. A., & Hartigan, P. M. 1985, Ann. Stat. 13, 1, 70
Jammal, G., & Bijaoui, A. 2004, Signal Processing, 84, 1049
Katgert, P., Mazure, A., Perea, J., et al. 1996, A&A, 310, 8
Kolokotronis, V., Basilakos, S., Plionis, M., & Georgantopoulos, I. 2001,
MNRAS, 320, 49
Krauss, L. M., & Starkman, G. D. 2000, ApJ, 531, 22
Lee, J., & Evrard, A. E. 2007, ApJ, 657, 30
López-Cruz, O., Barkhouse, W. A., & Yee, H. K. C. 2004, ApJ, 614, 679
Lubin, L. M., & Bahcall, N. 1993, ApJ, 415, L17
Maddox, S. J., Efstathiou, G., Sutherland, W. J., & Loveday, J. 1990, MNRAS,
243, 692
Markevitch, M., Ponman, T. J., Nulsen, P. E. J., et al. 2000, ApJ, 541, 542
Markevitch, M., Gonzalez, A. H., David, L., et al. 2002, ApJ, 567, L27
Maurogordato, S., Proust, D., Beers, T. C., et al. 2000, A&A, 355, 848
Maurogordato, S., Ferrari, C., Benoist, C., et al. 2006, in Proc. XLIst Rencontres
de Moriond, From dark halos to Light, ed. S. Maurogordato, J. Tran Thanh
Van, & L. Tresse (The Gioi Publishers)
Maurogordato, S., Cappi, A., Ferrari, C., et al. 2008, A&A, 481, 593
McLachlan, G. J., & Krishnan, T. 1997, The EM Algorithm and Extensions
(Wiley)
McLachlan, G. J., Peel, D., Basford, K. E., & Adams, P. 1999, J. Statistic
Software, 4, 2
Miller, N. A., Owen, F. N., Hill, J. M., et al. 2004, ApJ, 613, 841
Mohr, J. J., Geller, M. J., & Wegner, G. 1996, AJ, 112, 1816
Muriel, H., Quintana, H., Infante, L., Lambas, D. G., & Way, M. J. 2002, AJ,
124, 1934
Nagamine, K., & Loeb, A. 2003, New Astron., 8, 439
Navarro, J. F., Frenk, C. S., & White, S. D. M. 1996, ApJ, 462, 563
Niederste-Ostholdt, M., Strauss, M., Dong, F., Koester, B. P., & McKay, T. A.
2010, MNRAS, 405, 2023
Owers, M. S., Nulsen, P. E. J., Couch, W. J., Markevitch, M., & Poole, G. B.
2009, ApJ, 692, 702
Pinkney, J., Roettiger, K., Burns, J. O., & Bird, C. M. 1996, ApJS, 104, 1
Plionis, M., Benoist, C., Maurogordato, S., Ferrari, C., & Basilakos, S. 2003,
ApJ, 594, 144
Poole, G. B., Fardal, M. A., Babul, A., et al. 2006, MNRAS, 373, 881
Popesso, P., Bohringer, H., Romaniello, M., et al. 2005, A&A, 433, 415
Richstone, D. O., Loeb, A., & Turner, E. L. 1992, ApJ, 393, 477
Ricker, P. M., & Sarazin, C. L. 2001, ApJ, 561, 621
Roettiger, K., Loken, C., & Burns, J. O. 1997, ApJ, 109, 307
Sauvageot, J. L., Belsole, E., & Pratt, G. W. 2005, A&A, 444, 673
Schindler, S., & Böhringer, H. 1993, A&A, 269, 83
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500, 525
Suto, Y., Sasaki, S., & Makino, N. 1998, ApJ, 509, 544
Tonry, J. L., & Davis, M. 1981, ApJ, 246, 666
Teyssier, R. 2002, A&A, 385, 337
Ulmer, M. P., & Cruddace, R. G. 1982, ApJ, 258, 434
Vandame, B. 2002, SPIE Proc., 4847, 123
Vikhlinin, A., Markevitch, M., & Murray, S. S. 2000, ApJ, 551, 160
Way, M. J., Quintana, H., Infante, L., Lambas, D. G., & Muriel, H. 2005, AJ,
130, 2012
West, M. J., Jones, C., & Forman, W. 1995, ApJ, 451, L5
Wu, X.-P., Fang, L.-Z., & Xu, W. 1998, A&A, 338, 813
A79, page 19 of 19