Probing of the Interactions Between the Hot Plasmas and Galaxies in Clusters over a Cosmological Timescale

Similar documents
Clusters of Galaxies Groups: Clusters poor rich Superclusters:

This document is provided by JAXA.

Clusters of Galaxies " High Energy Objects - most of the baryons are in a hot (kt~ k) gas." The x-ray luminosity is ergs/sec"

Clusters of Galaxies Groups: Clusters poor rich Superclusters:

Clusters: Observations

Hubble s Law and the Cosmic Distance Scale

Chandra Analysis of a Possible Cooling Core Galaxy Cluster at z = 1.03

Infrared Mass-to-Light Profile Throughout the Infall Region of the Coma Cluster

Large-Scale Structure

X- ray surface brightness fluctuations and turbulence in galaxy clusters. Jeremy Sanders. Andy Fabian. Sanders & Fabian 2011, MNRAS, submitted

Clusters and Groups of Galaxies

arxiv:astro-ph/ v1 6 May 2004

Galaxy formation and evolution. Astro 850

Publ. Astron. Obs. Belgrade No. 86 (2009), TURKISH NATIONAL OBSERVATORY (TUG) VIEW OF CLUSTERS OF GALAXIES

Luke Leisman Calvin College. Photo Credit: SLOAN Digital Sky Survey. Great Lakes Cosmology Workshop June 15, 2010

Discovery of eight lensing clusters of galaxies

Exam 4 Review EXAM COVERS LECTURES 22-29

Clusters of Galaxies

arxiv:astro-ph/ v1 1 Nov 2006

Observational Cosmology

Clusters of galaxies and the large scale structure of the universe. Gastão B. Lima Neto IAG/USP

Contents. List of Participants

Galaxy clusters. Dept. of Physics of Complex Systems April 6, 2018

Chapter 25: Galaxy Clusters and the Structure of the Universe

Modern Image Processing Techniques in Astronomical Sky Surveys

Where are oxygen synthesized in stars?

Dust properties of galaxies at redshift z 5-6

Chandra Observation of Point Sources in the X-Ray Elliptical Galaxy NGC 1407

Coma Cluster Matthew Colless. Encyclopedia of Astronomy & Astrophysics P. Murdin

9.1 Large Scale Structure: Basic Observations and Redshift Surveys

Chapter 10: Unresolved Stellar Populations

Clusters and Groups of Galaxies

Unveiling the Structure of Galaxy Clusters with Combined ESO-VLT, WFI, and XMM-Newton Observations

arxiv: v1 [astro-ph.co] 15 Nov 2010

Astronomy 422. Lecture 15: Expansion and Large Scale Structure of the Universe

Inclination-Dependent Extinction Effects in Disk Galaxies in the Sloan Digital Sky Survey aa aa. A Senior Honors Thesis

Physics of Galaxies 2016 Exercises with solutions batch I

MASS PROFILES OF X-RAY BRIGHT RELAXED GROUPS: METHODS AND SYSTEMATICS

Upcoming class schedule

Techniques for measuring astronomical distances generally come in two variates, absolute and relative.

Course of Galaxies course organizer: Goeran Ostlin ESSAY. X-ray physics of Galaxy Clusters

Hubble s Law. Tully-Fisher relation. The redshift. λ λ0. Are there other ways to estimate distances? Yes.

arxiv:astro-ph/ v1 6 Dec 1999

The shapes of faint galaxies: A window unto mass in the universe

Cosmic Distance Determinations

The HII Regions of Sextans A

Lecture 32: The Expanding Universe Readings: Sections 26-5 and 28-2

Chapter 8: Simple Stellar Populations

The cosmic distance scale

Survey of Astrophysics A110

TOTAL TO CENTRAL LUMINOSITY RATIOS OF MASSIVE

Metals in clusters of galaxies observed with Suzaku and XMM- Newton

arxiv:astro-ph/ v1 13 Jan 2004

Lecture 11: SDSS Sources at Other Wavelengths: From X rays to radio. Astr 598: Astronomy with SDSS

Clusters: Observations

Simulations Applied to the Bright SHARC XCLF: Results and Implications

New Results on the AGN Content of Galaxy Clusters

High Redshift Universe

Stellar Mass Estimates of Galaxies in Superclusters at z~1

Galaxy Clusters: Cosmic High-Energy Laboratories to Study the Structure of Our Universe

Galaxy Luminosity Function

A Cluster of Galaxies, Abell 496

PoS(extremesky2009)103

SUPPLEMENTARY INFORMATION

Redshifted Broad Absorption Troughs in Quasars

The King's University College Astronomy 201 Mid-Term Exam Solutions

Clusters of Galaxies Ch 7 S&G! Coma Cluster-the nearest massive cluster! The apparent nature of clusters depends on the wavelength one looks at!

Direct empirical proof of dark matter?

Moment of beginning of space-time about 13.7 billion years ago. The time at which all the material and energy in the expanding Universe was coincident

Gamma-Ray Astronomy. Astro 129: Chapter 1a

X-ray and Sunyaev-Zel dovich Effect cluster scaling relations: numerical simulations vs. observations

Star formation in XMMU J : a massive galaxy cluster at z=1.4

Weak lensing calibrated scaling relations for galaxy groups and clusters in the COSMOS and CFHTLS fields

Quasars and Active Galactic Nuclei (AGN)

Brief update (3 mins/2 slides) on astrophysics behind final project

The effect of large scale environment on galaxy properties

Probing the Outskirts of Strongly Merging Double Clusters

3 The lives of galaxies

The Turmoil in IC1262

DETECTION OF A TEMPERATURE STRUCTURE IN THE COMA CLUSTER OF GALAXIES WITH ASCA

Low Surface Brightness Galaxies Erwin de Blok. Encyclopedia of Astronomy & Astrophysics P. Murdin

Clocks and Scales to understand the physics of BSS

Active Galaxies and Galactic Structure Lecture 22 April 18th

EVOLUTION OF DUST EXTINCTION AND SUPERNOVA COSMOLOGY

Lecture 15: Galaxy morphology and environment

arxiv:astro-ph/ v1 16 Apr 2004

The SINFONI Nearby Elliptical Lens Locator Survey (SNELLS)

Extended Chandra Multi-Wavelength Project (ChaMPx): Source Catalog and Applications

Normal Galaxies (Ch. 24) + Galaxies and Dark Matter (Ch. 25) Symbolically: E0.E7.. S0..Sa..Sb..Sc..Sd..Irr

Galaxy classification

The Radio/X-ray Interaction in Abell 2029

A New Analysis in the Field of the Open Cluster Collinder 223

Dark Matter. Galaxy Counts Redshift Surveys Galaxy Rotation Curves Cluster Dynamics Gravitational Lenses ~ 0.3 Ω M Ω b.

A100 Exploring the Universe: The Milky Way as a Galaxy. Martin D. Weinberg UMass Astronomy

The Caustic Technique An overview

The origin of lopsidedness in galaxies

HW#6 is due; please pass it in. HW#7 is posted; it requires you to write out answers clearly and completely explaining your logic.

The Formation of Galaxies: connecting theory to data

Clusters, lensing and CFHT reprocessing

AS1001:Extra-Galactic Astronomy

Transcription:

Probing of the Interactions Between the Hot Plasmas and Galaxies in Clusters over a Cosmological Timescale Liyi Gu 1, Poshak Gandhi 2, JinLin Han 3, Naohisa Inada 4, Madoka Kawaharada 5, Tadayuki Kodama 6, Saori Konami 7, Kazuhiro Nakawaza 8, Kazuhiro Shimasaku 9, Zhonglue Wen 3,Haiguang Xu 10, Kazuo Makishima 1,8,11 1 Research Center for the Early Universe, School of Science, University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033 2 Group of Astronomy and Astrophysics, University of Durham, United Kingdom 3 National Astronomical Observatories, Chinese Academy of Sciences 4 Department of Physics, Nara National College of Technology, Japan 5 Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency 6 National Astronomical Observatory of Japan 7 Department of Physics, Tokyo Metropolitan University 8 Department of Physics, School of Science, University of Tokyo 9 Department of Astronomy, School of Science, University of Tokyo 10 Department of Physics and Astronomy, Shanghai JiaoTong University 11 MAXI team, Institute of Physical and Chemical Research, Japan E-mail(LG): lygu@juno.phys.s.u-tokyo.ac.jp Abstract Based on optical and X-ray data for two samples (34 clusters in z = 0.1 0.9 and 340 clusters in z = 0.0 0.5), we have detected, for the first time, a significant evolution spanning a time interval of 6 Gyr in relative spatial distribution of member galaxy versus the ICM mass; while the galaxy component was as spatially extended as the ICM at z 0.9, towards the lower redshifts, it has indeed become more centrally-concentrated relative to the ICM sphere. We further discovered that this galaxy infall exhibits no dependence on the galaxy/cluster mass, implying a non-gravitational origin. The profound evolution is likely related to strong electromagnetic interaction (Makishima et al. 2001) between galaxies and ICM which continues on cosmological timescale. Key words: galaxies:clusters:general intergalactic medium X-rays: galaxies: clusters 1. Introduction Nearby clusters of galaxies are generally characterized by galaxies distributions which are much more concentrated, in terms of mass, than their intracluster medium (ICM). This is a well-known but barely explored issue, which cannot be simply explained by physics of gravity: since the ICM is hydrostatistically confined, while galaxies are dynamically confined by the same gravitational potential of cluster, the two baryonic components should share approximately the same specific energy, and thus similar distribution scale height. It is found that the stellar light distributions in nearby galaxy clusters are even more concentrated than the X- ray measured metal mass in the ICM (e.g., Kawaharada et al. 2009). Werner et al. (2013) reported that the iron abundance of ICM shows a flat distribution out to virial radius. Because the heavy elements in the ICM are all created in galaxies which are now distributing near the bottom of cluster potential, it takes a huge amount of energy ( 10 45 erg s 1 ) to transport the metals out to the observed radius. To overcome the above difficulties, let us recall a simple, but long-forgotten scenario proposed in Makishima et al. (2001) that galaxies moving through the cluster will interact strongly with the ICM, transfer their free energies and metals to the surrounding ICM, and will gradually fall to the cluster center (Fig. 1). This model requires the electromagnetism between ICM and galaxies are effective even on large scale. Yet it has little been investigated, and most of astrophysicists consider that the ICM would freely flow through interstellar space of moving galaxies. Surprisingly, the simple model can not only explain the baryon/heavy element distributions of nearby clusters, but also potentially solve many long-lasting puzzles as follows. 1. It means a discovery of a totally new large-scale energy flow between the two major cosmic baryons. - 382 -

where ICM density is much lower than those in the latter sample, the difference in two results might actually hint for an ICM effect on the galaxy distribution. To address this issue, we jointly analyze high quality optical and X-ray data for two samples of clusters, and study in details the relative spatial distributions of stellar component vs. ICM components. 2. Construction of Two Samples Fig. 1. Schematic illustration of the model in Makishima et al. (2001). Left panels are young (distant) clusters while right ones are old (nearby). Upper and lower panels show the evolution of matter and metallicity, respectively. 2. It gives a solution to the famous enigma: why the hot ICM in each cluster is kept thermo-dynamically stable, even through the observed X-ray emission, 10 44 erg s 1 per cluster, would significantly cool off the ICM over the Hubble time? Simply, the ICM is likely to be heated by the moving galaxies. 3. It provides answer to a widely recognized but never answered problem: why galaxies evolve differently in- and out-side clusters? This environmental effect might be due to that the spiral galaxies falling to cluster center have a higher chance to merge with others and evolve to be ellipticals. One important observational expectation of this model is that member galaxies should distribute out to cluster peripheral regions in the early stage, and gradually fall towards center relative to the ICM. So far no systematic study has been made to compare directly the spatial extents of galaxies and ICM for different redshifts. Several works studied the galaxy distribution alone: by stacking the SDSS DR7 data of over 20,000 optically-selected groups and clusters, Budzynski et al. (2012) reported that the galaxies density profile does not vary significantly in shape over z = 0.15 0.4. On the other hand, by analyzing the spectroscopic and photometric data of 15 X-ray bright rich clusters at z = 0.18 0.55, Ellingson et al. (2001) showed that the galaxies gets more centrally-concentrated towards lower redshifts. Since the former sample is dominated by groups and poor clusters, To cover large range in redshift, we selected 34 clusters with z = 0.1 0.9 for a pencil-beam survey (sample I). Main results of sample I were reported in Gu et al. (2013). To enhance statistics of the results, a much broader survey (sample II) was constructed with z = 0.0 0.5. The selection procedures of both sample are described as follows. Sample I was primarily constructed based on the previous flux-limited samples presented in Croston et al. (2008; 31 clusters), Leccardi & Molendi (2008; 48 clusters), and Ettori et al. (2004; 28 clusters). First, to study clusters with sufficient member galaxies and similar extent, we selected M 500 > 2 10 14 M systems with R 500 withintherangeof700 1400kpc. Second, toavoid strong merging clusters, the positions of X-ray peak and the central dominant galaxy are required to coincide with each other within 50 kpc, and the X-ray morphology should be approximately symmetric. Finally we selected 34 clusters that survive both criteria. The sample was further categorized into three subsamples by their redshifts, i.e., low-redshift subsample (z 0.11 0.22, hereafter subsample I-L), intermediate-redshift subsample (z 0.22 0.45, hereafter subsample I-M), and highredshift subsample (z 0.45 0.89, hereafter subsample I-H). Sample II includes all clusters available in the Chandra and XMM-Newton archives, and covered by the Sloan Digital Sky Survey (SDSS) up to DR9. The neighboring Virgo, Coma, and Perseus clusters are not selected. The sample includes 340 clusters. By comparing this sample with existing catalogs of the NORAS and REFLEX surveys (Bohringer et al. 2000), we found it covers a 80% of sky density of NORAS/REFLEX at X-ray flux of> 5 10 12 ergs 1 cm 2. Thismeansthatthesample II includes most of the X-ray bright objects in the SDSS region. Similar to Sample I, Sample II was further categorized into three subsamples by their redshift, i.e., lowredshift subsample (z 0.0 0.08, hereafter subsample II-L), intermediate-redshift subsample (z 0.08 0.22, hereafter subsample II-M), and high-redshift subsample (z 0.22 0.45, hereafter subsample II-H). As an example, we plot in Figure 2 the optical and X-ray images of six Sample II clusters. - 383 -

observed with similar depth and seeing as the cluster central regions. We fit the combined central + offset I- band surface brightness profiles with a King model plus constant, S I (r) = S 0 [1+(r/r c ) 2 ] 3/2 +S bkg (King 1962), where S 0 is a normalization, r c is core radius, and S bkg represents the background component. The backgroundsubtracted galaxy fluxes enclosed in radius r were combined to calculate the integral light profile L opt,bs (r). Color-Magnitude Filtering Fig. 3. Subsample-averaged galaxy number density profiles of sample I, compared with the mean profile reported in Budzynski et al. (2012). Fig. 2. RGB composite images (red: SDSS-i, green: SDSS-g, blue: Chandra/XMM-Newton 0.3 8.0 kev) of six sample II clusters. Detected member galaxies are marked by green regions. R 500 is shown in white circle. 3. Data Analysis and Results 3.1. Sample I 3.1.1. Optical Light Profiles We analyzed the B-, R-, and I- band photometric data of the 34 clusters observed with the UH88 telescope. To calculate the radial profiles of the stellar light, we first need to define the galaxy membership of each cluster. Here we carried out two independent approaches, i.e., background subtraction (hereafter BS), and color magnitude filtering (hereafter CMF) methods. Background Subtraction First, we divided the I-band images of central pointings into three or four radial bins, each with a width of 1. To exclude apparent foreground galaxies, galaxies with flux (FLUX AUTO) higher than the cluster central galaxy were discarded from the surface brightness calculation. Then we estimated the background contribution based on data of local field regions about 1.5 5.1 Mpc ( 1 3 R 200 ) offset from the cluster center, which were To cross-check the background subtraction result, we applied a member galaxy selection based on the standard color-magnitude relation, that the old passively evolving cluster galaxies form a tight horizontal relation in the color-magnitude space (e.g. Kodama et al. 1998). The colors m R m I were acquired from the Kron-type I- and R- band magnitudes calculated with the SExtractor, and corrected for photometric zero points. A Gaussian fitting to the number-color histogram gave the approximate central color (C 0 ), as well as the associated error color (σ C ). To determine the slope of the relation, we performed linear fitting in magnitude range 14 < m I < 24. Then the red-sequence members were selected as data within the fitting uncertainty range of the derived relations, typically δ(m R m I ) 0.1 0.2 mag. The galaxy number density profiles obtained with thecmfmethodisshowninfigure3. Theintegral light profiles, L opt,cmf (r), were obtained based on the fluxes of the selected galaxies. 3.1.2. X-ray Mass Profiles To derive cluster gas mass profiles, the gas density profiles were calculated based on deprojection spectral analysis with the XMM-Newton EPIC (for z < 0.5 clusters )and Chandra ACIS (for z > 0.5 clusters) data after point source removal and background correction. First we extracted spectra from several concentric annulus regions for each cluster. The extracted spectra were fit in 0.7 8.0 kev with an absorbed, single-temperature - 384 -

APEC model in XSPEC, when the projection effect was corrected by using the PROJCT model. Then the 3- D gas density profiles were calculated from the best-fit model normalization of each annulus. To obtain 2-D ICM mass profiles, M X,2D (r), in the same manner as the integral optical light profiles, we projected along line-ofsight the best-fit density profiles, and integrated for each radius the enclosed ICM mass. Fig.4. (a) Average GLIMR profiles of sample I. Points with filled circles and boxes are results with the BS and CMF methods, respectively. (b) Subsample-average GLIMR profiles obtained in approximated same rest-frame bands. (c) Subsample-average GLNMR profiles with the two methods. 3.1.3. Galaxy Light vs ICM Mass Ratio Profiles To enable direct comparison between the slopes of L opt (r) and M X,2D (r) profiles obtained in 3.1 and 3.2, respectively, we normalized each of them to its central values within r = 0.25R 500. The galaxy light vs. ICM mass ratio profiles (hereafter GLIMR profiles) were then calculated as the ratios between the normalized profiles. For each subsample (i.e., I-L, I-M, and I-H), the mean GLIMR profile was calculated at four radii, i.e., r = 0.35R 500, 0.45R 500, 0.55R 500, and 0.65R 500. As shown in Figure 4a, the average GLIMR profiles of subsample I-L clusters drop steeply towards outer regions, in contrast, the GLIMR profiles of subsample I-H clusters show apparently flatter distributions than the low redshift counterparts, and the profiles of subsample I-M clusters appear to have intermediate gradients. To quantify the significance of the possible evolution, we carried out Kolmogorov-Smirnov (K-S) test among the three subsamples. Given the obtained D-statistic, the hypothesis that the GLIMRs follow the same distribution at 0.65 R 500 can be rejected on 94% and 98% confidence levels between subsamples I-L and I-M, and between subsamples I-M and I-H, respectively. 3.1.4. Systematics and Biases A major concern of the GLIMR profiles is possible redshift-dependent deviation biased by radial color gradient of galaxies, and possible galaxy luminosity uncertainties, coupled with the obvious differences in the restframe color among the three subsamples. As shown in Boselli & Gavazzi (2006), the distribution of blue galaxies has been observed to be considerably flatter than that of the red ones. Since higher-redshift clusters are observed in bluer rest-frame light, the apparent evolution in Figure 4a could simply because the subsample I-H is contributed more by bluer galaxies which have flatter spatial distributions. To examine for such effects on stellar distribution, we carried out investigations in two independent approaches. First, to achieve a redshiftindependent response to different color galaxies, the integral light profiles were re-calculated based on the same rest-frame bands. The B-band, R-band, and I-band data were used for z = 0.119 0.235, z = 0.235 0.600, and z = 0.600 0.890 clusters, respectively, so that the fluxes are all measured in approximately rest-frame B- band. As shown in Figure 4b, the average CMF-GLIMR profiles derived in the same rest-frame band are consistent within 5% scatter with the one derived with I-band data alone. As another approach to more thoroughly eliminate the possible redshift-dependence of color gradient and galaxy luminosity, we adopted galaxy number counts instead of the optical luminosity, and derived the galaxy number vs. ICM mass ratio (hereafter GNIMR) profile for each cluster. Both the BS and the CMF methods were employed to derive the galaxy number density profile for each clusters. The GNIMR profiles shown in Figure 4c were calculated in a similar way to the GLIMR profiles. The BS and CMF methods yield consistent profiles within 1σ error ranges. The shapes of the subsample-averaged GN- IMR profiles exhibit significant dependence on redshift, and the evolution is consistent with the GLIMR profiles. Hence, from the above two examinations, i.e., the GLIMR profiles obtained in the same rest-frame band and the GNIMR profiles, we can exclude the possibility that the detect GLIMR evolution is caused by the radius- /redshift- dependent galaxy color/luminosity variation. 3.2. Sample II 3.2.1. Optical Light Profiles To select member galaxies for the sample clusters, we used the optical photometric data of the Sloan Digital Sky Survey (SDSS-III). For each galaxy in the sky field (R < 2.5 Mpc) of sample clusters, we obtained photometric redshift (hereafter z phot ), K-correction, and absolute magnitude M r from a table Photoz which is based on the method of Csabai et al. (2007). Saturated objects and those with blending problems were removed using flags. Following Wen et al. (2009), for a cluster at - 385 -

z cl, themembergalaxiesareselectedwitharedshiftfilter of z cl 0.04 < z phot < z cl +0.04, and a limiting magnitude of M e r 20.5. M e r is r band absolute magnitude corrected for passive evolution, M e r(z) = M r (z) + Qz, where Q = 1.62 (Blanton et al. 2003). As shown in Wen et al. (2012), z phot of member galaxies are consistent with the spectroscopic redshifts (hereafter z spec ) within an average scatter of 0.02 0.03. To exclude the z phot with apparent bad photometry, we discarded objects with z phot errors > 0.08(1 + z phot ). This removed about 20% of the total galaxies. As shown in Wen et al. (2009), the selected member galaxies are complete > 90% for rich clusters, and 80% for groups. To further enhance the accuracy of member galaxy selection, we applied the z spec measured with the SDSS DR9 (Ahn et al. 2012). For a cluster with recession velocity of v cl, we selected objects with velocity differences v v cl < 2500 km s 1. By further applying the limiting magnitude M e r 20.5, a new member galaxy catalog was then created. The results of spectroscopic and photometric selections were combined to obtain the final catalog: all galaxies selected by z spec were considered as members, while the z phot selection was used in case when z spec is not available. profiles. First we examine how the member galaxy distributions change with cluster mass. The sample was separated into three mass bins, i.e., 5 10 13 2 10 14 M, 2 10 14 6 10 14 M, and 6 10 14 2 10 15 M. As shown in Figure 5, the mean density profiles of the more massive clusters are systematically larger than the low mass ones, which is known as the mass-richness relation. To prove the galaxy distribution evolution indicated in Figure 2, we then calculated mean density profiles for the three redshift bins (II-L, II-M, and II-H; 2). As shown in Figure 5, the high redshift bins have slightly larger galaxy number than the low redshift bins. This is probably due to the selection effect that high redshift bins have more massive clusters. Another apparent feature is that the radial profiles of low redshift clusters are more centrally-peaked compared to the high redshift ones. Because the gradient difference among the three redshiftdependent bins are apparently stronger than that of the mass-dependent bins, it cannot be explained merely by the selection effect alone. To compensate the intrinsic difference in cluster scale, we have further scaled the radial profile of each cluster by dividing both the radius and number density by R 500. As shown in Figure 5, the radial profiles of three mass-dependent bins exhibit similar gradients in general, except for the central 0.2R 500 region of low-mass bin in which the radial profile is slightly steeper. For the redshift-dependent bins, however, the gradient difference can still be significantly detected. This means that the member galaxies in nearby sample II objects are indeed more centrally-concentrated than those in the distant counterparts. 3.2.2. X-ray Mass Profiles The X-ray mass profiles of sample II were obtained in the same way as sample I ( 3.1.2). The subsample-mean profiles are presented in Figure 5. The clusters in lower redshift bins exhibit relatively lower densities in the center, while the three subsamples agree at the outer region ( R 500 ). This feature can be seen even in the R 500 - scaled profiles, though the vertical differences between three subsamples are much smaller. Fig. 5. (upper) Mean galaxy number density profiles of three mass bins of sample II. The raw and R 500 -scaled profiles are in left and right panels, respectively. (middle) Mean galaxy number density profiles of three redshift-subsamples. (lower) Mean ICM number density profiles of three redshift-subsamples. For each cluster, galaxies selected with this method were combined to calculate the surface number density 3.2.3. Galaxy Number vs ICM Mass Ratio Profiles To compare the radial gradients of galaxy and ICM distributions in sample II, we carried out the same analysis as was done for sample I ( 3.1.3). The galaxy number profiles were normalized to r = 0.25R 500 and divided by the ICM mass profiles normalized in the same way. As showninfig. 6,weplottedtheGLNMRprofilesinacontinuous form for the sample II, and compared them with those of sample I. The two sets of profiles are in rough consistent with each other. Both samples exhibit clear evolution that galaxies get more centrally-concentrated - 386 -

gravitationally along the galaxy orbit (e.g., Rephaeli & Salpeter 1980). To examine whether such a process may reproduce the observed GLIMR/GLNMR evolution or not, we investigated quantitatively the orbital evolution of a model galaxy that is affected by gravitational drag from a model cluster, which is given by Fig. 6. Subsample-average GLIMR profiles of sample I (points with errors) and sample II (curves). relative to the ICM in lower redshifts. As shown in Figure 5, the GLNMR evolution is caused by a combined effect in galaxy density and ICM mass distributions, when the evolution in galaxy distribution contributes a main role. As proved in Gu et al. (2013), this evolution means that galaxy are infalling from cluster outer region. 4. Origin of the Galaxy Infall Fig. 7. Predicted evolution of galaxy orbit from the initial position of 300 kpc (solid line), 500 kpc (dash line), and 700 kpc (dotted line) for a galaxy with mass of 1 10 11 M (left) and 1 10 12 M (right). Dynamical friction is considered. Fig. 8. Subsample-average GLIMR profiles of sample I, for galaxies of mass < 1 10 11 M (cross), plotted against the profiles with all galaxies (filled circle). One candidate process to enhance galaxy infall is dynamical friction, which occurs when the potential of a moving galaxy interacts with a wake that is created F DF (R) 4πρ(R)(Gm)2 v 2 (1) (e.g., Ostriker 1999), where ρ(r) is the total mass density, G is gravitational constant, m is galaxy mass, and v is galaxy velocity. As shown in Figure 7, the dynamical friction is expected to induce a spatial mass segregation in galaxy clusters that massive galaxies become concentrated towards center more quickly, while the less massive ones are much less affected. To assess the effect of dynamical friction, it is hence crucial to examine whether or not the GLIMR evolution depends strongly on the galaxy mass. Here we set a mass limit of 1 10 11 M to sample I, which was converted to the optical flux threshold by using the Bruzual & Charlot stellar model. As shown in Figure 7, gravitational drag on galaxies with mass below this limit is expected to be negligible. By applying the same threshold to the galaxies in the offset region, we calculate the GLIMR profiles for m < 1 10 11 M galaxies with the BS method. As shown in Figure 8, the GLIMR profiles for low-mass galaxies still exhibit significant evolution among subsample I-H, I-M, and I-L. Since this feature cannot be explained by dynamical friction, we conclude that the gravitational effect alone is insufficient to account for the observed GLIMR evolution. Other physics, e.g., ICM drag, should also be responsible for the sink of the cluster galaxies towards cluster center, especially for the low- mass ones. References Ahn, C. et al. 2012, ApJS, 203, 21 Blanton, M. et al. 2003, ApJ, 592, 819 Bohringer, H. et al. 2000, ApJS, 129, 435 Budzynski, J. M. et al. 2012, MNRAS, 423, 104 Croston, J. H. et al. 2008, A&A, 487, 431 Csabai, I. et al. 2003, AJ, 125, 580 Ellingson, E. et al. 2001, ApJ, 547, 609 Ettori, S. et al. 2004, A&A, 417, 13 Gu, L. et al. 2013 ApJ, 767, 157 Kawaharada, M. et al. 2009 ApJ, 691, 971 Kodama, T. et al. 1998, A&A, 334, 99 Makishima, K. et al. 2001 PASJ, 53, 401 Wen, Z. et al. 2009, ApJS, 183, 197 Wen, Z. et al. 2013, MNRAS, 436, 275 Werner, N. et al. 2013 Nature, 502, 656-387 -