<<

arXiv:1808.02136v2 [astro-ph.CO] 17 Jul 2019 rpitsbitdt Elsevier to submitted Preprint a on norerirwr.Teinrsoeo h MdniyfrteCL pro Navarro–Frenk–White the the for density than DM flatter the be of to slope found inner is The baryonic clusters by work. influenced earlier strongly our in are found clusters th the that of find We corr profiles several density clusters. the and of distribution quantities DM characteristic modified between resu with This clusters distribution. of DM dissipationles formation inner a the by flattens followed physics phase, baryonic dissipative where first the brighte in the forms context, this (BCG) In formation. cluster bas of interpreted scenario are phase lensing from derived profiles density total The and Subaru combined focu s from and analysis measurements weak- Our high-quality have al. that et clusters as Newman X-ray-selected CLASH well from extend clusters as by several clusters profiles on work density high-mass vious 15 (DM) of matter sample a dark for and correlations total the study We Abstract [email protected],[email protected] iatmnod iiaeAtooi,Uiest fCatania Of University Astronomia, e Fisica di Dipartimento e nvriy o22 ot inhiRa,Lnhu as,7 Gansu, Lanzhou, Road, Tianshui South No.222, University, nttt fMdr hsc,CieeAaeyo cecs P Sciences, of Academy Chinese Physics, Modern of Institute ∗ 1 c d hsppri eiae otemmr forclegePo.L Prof. colleague our of memory the to dedicated is paper This nttt fTertclPyis colo hsclScien Physical of School Physics, Theoretical of Institute mi addresses: Email author Corresponding nttt eAto´iaeCˆnisd Espa¸co, Universid Ciˆencias do e Astrof´sica de Instituto orltosi h atrdsrbto nCLASH in distribution matter the in Correlations noioDlPopolo Del Antonino b NNszoed aai,VaS oa6,I913Ctna Ita Catania, I-95123 64, Sofia S. Via Catania, di sezione INFN inis d 8 ap rne 7906Lso,Portugal Lisboa, 1769-016 Grande, Campo C8, Ed. Ciencias, [email protected] 300 epe eulco China of Republic Peoples 730000, aayclusters galaxy a,b ognL Delliou Le Morgan , aai,Italy Catania, China Atnn e Popolo), Del (Antonino ubeSaeTelescope Space Hubble Mra eDelliou) Le (Morgan d eLso,Fclaede Faculdade Lisboa, de ade eadTcnlg,Lanzhou Technology, and ce c,d, il nraDra6 95125 6, Doria Andrea Viale , 00,PolsRpbi of Republic Peoples 30000, s ffieBx1 Lanzhou Box31, Office ost ∗ iu Lee Xiguo , e Xiguo ee, tcutrgalaxy cluster st do h two- the on ed eta DM central e trong-lensing observations. ly l,ranging file, n u pre- our ing uy1,2019 18, July e,1 hsc as physics e n15 on ses t nthe in lts elations phase s ASH their from α = 0.30 to 0.79. We examine correlations of the DM density slope α with the effective radius Re and stellar mass Me of the BCG, finding that these quantities are anti-correlated with a Spearman correlation coefficient of ∼−0.6.

We also study the correlation between Re and the cluster halo mass M500, and the correlation between the total masses inside 5kpc and 100kpc. We find that these quantities are correlated with Spearman coefficients of 0.68 and 0.64, re- spectively. These observed correlations are in support of the physical picture proposed by Newman et al. Keywords: Galaxy Clusters, Galaxy Formation, Weak Gravitational Lensing

1. Introduction

The Λ cold dark matter (ΛCDM) paradigm gives a plethora of correct pre- dictions [1, 2, 3, 4, 5]. However, some of its predictions are at odds with ob- servations. N-body simulations in ΛCDM predict that the spherically averaged density profiles of self-gravitating structures, ranging from dwarf to galaxy clusters, are cuspy and well approximated by the Navarro–Frenk–White (NFW) profile [6, 7]. However, observations [8, 9, 10, 11] and theoretical studies [12, 13, 14, 15, 16, 17] have shown that the inner slopes of the density profile in dwarf galaxies and low-surface-brightness galaxies (LSBs) are usually flatter than simulations, and there is a strong diversity of the dark-matter (DM) distri- bution in these low-mass systems [the so-called “diversity problem”, 18, 17, 19].2 On the observational side, the small dynamic range of observations can cause a degeneracy in the mass profile determination [see 27], and this degeneracy cannot be fully broken due to the lack of HI observations in dwarf spheroidals (dSPhs) and elliptical galaxies. Determinations of their DM structure are thus much more complicated. In the case of dSPhs, there are discrepant results on the cusp-core nature of the density profile [28, 29, 30, 31], sometimes even in the

2In addition to this problem, the ΛCDM paradigm suffers from the cosmological constant problem [20, 21], the unknown nature of dark energy [22, 23, 24], and from several problems at small scales [25, 26]

2 case of the same object studied with different techniques. Similar uncertainties are present in cluster of galaxies, but X-ray observations, lensing and galaxies dynamics overcome them in an easier manner than for the cases of dSPhs or ellipticals. While the NFW profile [12, 6] describes well the observed total density pro- files in galaxy clusters as found in several studies [32, 33, 34, 35, 36, 37, 38, 39, 40, 41], it was also found that the inner DM structure is characterized by a flatter slope within typical scales of the (BCG; from some kpcs to some tens of kpcs). Hence, the cusp-core problem [17, 19] appears to be present in galaxy clusters as well. This discrepancy can be alleviated when the effects of baryonic physics are properly accounted for in N-body simulations [see 42, 43, 13, 44, 15, 45, 46, 14, 47, 16, 48, 49, 50]. In order to study how baryons modify the formation and evolution of clusters, we consider in [51] baryonic clumps interacting with the DM model introduced in [13]. In addition to finding that the central baryonic concentration within 10 kpc plays an important role in shaping the cluster density profile, we reproduced the observed cluster profiles for several massive systems [33, 34, 36], namely A611, A383, MACSJ1423.8+2404, and RXJ1133. In [52], we reproduced the correlations found by [38, 39],3 for MS2137, A963, A383, A611, A2537, A2667, and A2390. For these clusters, the total mass den- sity profiles are well fitted by an NFW profile, while the central DM distribution is shallower than the total mass distribution. The formation picture proposed by [38, 39] is characterized by a dissipa- tional formation of BCGs, followed by a dissipationless phase. In this phase, as described by [42, 53, 43, 54, 55, 13, 56, 17, 51, 47, 50], baryon clumps interact

3The quoted authors found correlations of the inner slope of the DM profile with the size of the BCG, the core radius, namely the constant density core of the cored NFW density profile [see Eq. 2 of 39], and the BCG mass, and finally the correlation between the masses contained inside 5 kpc and 100 kpc. In the present paper, with BCG mass we refer to the stellar mass only.

3 Table 1: Parameters derived for the CLASH sample. First column: cluster name; second:

M500 as given in [41]; third and fourth: innermost 2D density slopes inferred directly from the observed [41] profiles and obtained from our semi-analytical model; fifth: inner 3D density slope from our model; sixth and seventh: stellar and baryonic fractions from our model.

Name M500 α2D α2D,T α3D fstar Fb 14 [10 M⊙] A383 5.88 ±1.73 0.71 ±0.26 0.70 ±0.09 0.37 ±0.09 0.0201 ±0.002 0.1355 ±0.008 A209 9.64 ±1.97 0.67 ±0.29 0.68 ±0.09 0.60 ±0.1 0.0167 ±0.002 0.1417 ±0.01 A2261 15.65 ±3.05 0.77 ±0.26 0.79 ±0.09 0.63 ±0.09 0.0140 ±0.002 0.1480 ±0.012 RXJ2129 4.48 ±1.16 0.49 ±0.26 0.49 ±0.09 0.55 ±0.09 0.0222 ±0.002 0.1323 ±0.006 A611 10.73 ±2.65 0.59 ±0.27 0.58 ±0.09 0.79 ±0.09 0.0161 ±0.002 0.1431 ±0.01 MS2137 8.28 ±2.57 0.86 ±0.25 0.85 ±0.09 0.65 ±0.08 0.0177 ±0.002 0.1398 ±0.009 RXJ2248 12.45 ±3.62 0.45 ±0.28 0.44 ±0.09 0.55 ±0.09 0.0152 ±0.002 0.1450 ±0.011 MACSJ1115 10.67 ±2.22 0.33 ±0.30 0.34 ±0.09 0.39 ±0.1 0.0161 ±0.002 0.1430 ±0.01 MACSJ1931 10.51 ±4.05 0.69 ±0.28 0.70 ±0.09 0.65 ±0.09 0.0162 ±0.002 0.1428 ±0.01 MACSJ1720 9.96 ±2.53 0.59 ±0.26 0.60 ±0.09 0.56 ±0.09 0.0165 ±0.002 0.1421 ±0.01 MACSJ0429 6.85 ±2.1 0.45 ±0.28 0.44 ±0.09 0.48 ±0.09 0.0190 ±0.002 0.1374 ±0.008 MACSJ1206 12.24 ±2.49 0.56 ±0.26 0.57 ±0.09 0.50 ±0.09 0.0153 ±0.002 0.1448 ±0.011 MACSJ0329 6.51 ±1.37 0.65 ±0.27 0.64 ±0.09 0.70 ±0.09 0.0193 ±0.002 0.1368 ±0.008 RXJ1347 22.33 ±4.89 0.39 ±0.30 0.40 ±0.09 0.30 ±0.1 0.0123 ±0.002 0.1528 ±0.014 MACSJ0744 11.94 ±2.81 0.54 ±0.27 0.53 ±0.09 0.55 ±0.09 0.0155 ±0.002 0.1445 ±0.011

4 Table 2: Physical parameters derived for the CLASH sample. First column: cluster name; second: BCG mass derived from our model; third: BCG effective radius; fourth and fifth: spherical total masses inside 5 kpc and 100 kpc.

Name Me Re M5kpc M100kpc 11 11 13 [10 M⊙] [kpc] [10 M⊙] [10 M⊙] A383 9.16 ±0.29 28.7 ±1.5 0.98 ±0.15 1.96 ±0.3 A209 7.84 ±0.29 25 ±1.5 1.31 ±0.15 3.21 ±0.3 A2261 10.5 ±0.29 40 ±1.5 1.75 ±0.15 4.32 ±0.3 RXJ2129 13.73 ±0.29 33 ±1.5 2.43 ±0.15 3.49 ±0.3 A611 12.25 ±0.29 34.6 ±1.5 1.63 ±0.15 3 ±0.3 MS2137 6.55 ±0.29 14 ±1.5 1.95 ±0.15 3.5 ±0.3 RXJ2248 12.45 ±0.29 38.5 ±1.5 2.17 ±0.15 4.3 ±0.3 MACSJ1115 11.44 ±0.29 44.5 ±1.5 2 ±0.15 3.56 ±0.3 MACSJ1931 8.19 ±0.29 31 ±1.5 1.37 ±0.15 3.5 ±0.3 MACSJ1720 8.99 ±0.29 35.8 ±1.5 1.5 ±0.15 3.32 ±0.3 MACSJ0429 13.37 ±0.29 41 ±1.5 2.32 ±0.15 3.28 ±0.3 MACSJ1206 11.96 ±0.29 43 ±1.5 2.08 ±0.15 4.08 ±0.3 MACSJ0329 7.41 ±0.29 20 ±1.5 1.24 ±0.15 2.17 ±0.3 RXJ1347 13.8 ±0.29 46.9 ±1.5 2.45 ±0.15 4.5 ±0.3 MACSJ0744 10 ±0.29 37.1 ±1.5 1.67 ±0.15 3.98 ±0.3

5 with DM through dynamical friction, “heating” DM and reducing the central cusp. Our aims here are to use high-quality gravitational lensing observations from the CLASH survey [57] and investigate if CLASH clusters exhibit correlations that are similar to those observed in the [38, 39] clusters, to characterize the mass distributions of CLASH clusters, and to test the physical picture that was proposed by [38, 39] and confirmed by [52]. To this end, we perform an improved analysis on a sample of 15 X-ray-selected CLASH clusters compared to our previous work [51, 52]. In the present work, we will characterize the total mass density profiles of 15 CLASH clusters by means of a modified version of the semi-analytical model developed by [51, 52]. Here we take into account the following effects:

1. adiabatic contraction (AC) responsible for the steepening of the inner density profiles in the early stage of cluster formation, 2. interaction between baryonic clumps and DM through dynamical friction, which is responsible for “heating” the DM component and flattening the density profile, 3. supernovae (SN) feedback, 4. AGN feedback and other baryonic effects described in detail in Appendix.

The paper is organized as follows. In section 2, we describe the data used and provide a brief summary of our model. In section 3, we discuss the results, and section 4 is devoted to conclusions. Throughout this paper, we adopt a concordance ΛCDM cosmology with −1 −1 Ωm =0.27, ΩΛ =0.73, and h =0.7 with H0 = 100h kms Mpc .

2. Data used and summary of the model

In this study, we use lensing data obtained from the CLASH survey [57], which studied the mass distributions of 25 high-mass clusters using high-quality gravitational lensing observations. Here we focus on a subsample of 15 X-ray- regular CLASH clusters for which strong and weak lensing data are available

6 from both 16-band (HST) and wide-field weak-lensing observations [41]. The wide-field weak-lensing data were taken primarily with Suprime-Cam on the Subaru Telescope [40, 58, 59]. We exclude high-magnification CLASH clusters from our analysis because they are found to be highly disturbed merging systems [57]. We also exclude one X-ray-regular CLASH cluster (RXJ1532) for which no secure identification of multiple images has been made [60] and hence no central strong-lensing infor- mation is available. These selection criteria result in our sample of 15 CLASH clusters (Table 1). The data we use are taken from [41] and given in the form of binned surface mass density profiles, spanning the radial range from 10arcsec to 16 arcmin, and their bin-to–bin covariance matrices. It was found in [41] that the ensemble- averaged surface mass density profile of these X-ray-regular clusters can be well described by cuspy, sharply steepening density profiles, such as the NFW and Einasto profiles. Assuming the spherical NFW profile for each cluster, [41] also found that the concentration–mass relation for the CLASH X-ray-selected subsample is in agreement with ΛCDM predictions, when the CLASH selection function is taken into account. In [41], the binned surface mass density profiles were derived for a sample of 16 X-ray-regular and 4 high-magnification CLASH clusters using the weak- and strong-lensing data of [40, 60]. The mass profile solution for each cluster, N Σ = {Σi}i=1 with N = 15 bins, was obtained from a joint likelihood analy- sis of strong-lensing, weak-lensing shear and magnification data [41, see their

Fig. 11]. The total covariance matrix Cij accounts for the observational errors, the cosmic-noise contribution due to projected uncorrelated large scale struc- ture, the systematic errors due to the residual mass-sheet degeneracy, and the intrinsic variations of the projected cluster mass profile due to halo triaxiality and correlated substructures. In this paper, we generate 3D cluster density profiles whose surface density profiles match those obtained by [41], Σ. For a given 3D density profile ρ(r), we compute the surface mass density Σ(R) by integrating ρ(r) along the line of

7 sight, ∞ Σ(R)=2 ρ(R,l)dl (1) Z0 with R the projected cluster-centric radius [see also Secs. 5.2 and 5.2.2 of 41]. A set of the parameters p that specify the model can be inferred by mini- mizing the χ2 function [see 41],

N 2 ˆ −1 ˆ χ (p)= [Σi − Σi(p)]Cij [Σj − Σj (p)], (2) i,jX=1 where Σˆ i is the surface mass density predicted by the model. Estimates of cluster mass and its radial distribution can be obtained in different ways. The standard approach, as adopted by [41], is to use the NFW profile, which gives a good approximation to the projected total density profile for cluster-size halos out to their virial radius [61]. The NFW density profile is given by ρcδc ρ(r)= 2 , (3) (r/rs)(1 + r/rs) where rs = r∆/c∆ is the scale radius, c∆ the concentration parameter, ρc the critical density of the , r∆ the radius inside which the mean density is

∆ × ρc, and 3 ∆ c∆ δc = . (4) 3 ln(1 + c∆) − c∆/(1 + c∆)

The total mass enclosed within a sphere of radius r∆ is denoted as M∆ = 3 (4π/3)∆ρcr∆. The NFW mass and concentration parameters for the CLASH sample are reported in Tables 2 and 3 of [41]. The typical mass and concentra- 15 −1 tion for the X-ray-selected CLASH sample are M200 ≃ 1.0 × 10 M⊙ h and c200 ≃ 3.8 at z ∼ 0.35 [41, 62]. In our model, described in Appendix A, we chose not to include the 2- halo clustering term, that takes into account the contribution on the projected surface mass density coming from the large-scale clustering. As seen in Figure 3 of [41], the NFW approximation to the surface mass density Σ(R) gives a good estimate to the virial or to the r200m radius. When the fitting is restricted to the radial range ≤ 2Mpc/h, the contribution from the 2-halo term is not important

8 ([40, 41]). In other words, since the results remain the same when our fitting range is limited to < 2Mpc/h, our results are not sensitive to this effect. Although the matter distributions in CLASH clusters is not spherical but triaxial [see 59, 63], as expected for dark-matter dominated cluster-scale halos [e.g. 64], we assumed spherical symmetry in the constraints we obtained. In our spherical modeling of CLASH clusters, we account for the uncertainty arising from projection effects of aspherical cluster shapes in the covariance matrix Cint [see 41]. In order to characterize the 3D density profile of the clusters, we use a mod- ified version of the physical cluster model described in [51, 52]. In those papers

[i.e. 51, 52], the DM density profile was expressed as ρDM = F (M500, Fb, j), with

M500 being the cluster halo mass, fb = Mb/M500 the cluster baryon fraction, and j the random angular momentum The density profile of the clusters studied were reproduced by a) assuming that the cluster final mass in the model is the same as the observed clusters, b) assuming that the cluster baryon fraction is equal to that calculated with the [65] data, and c) adjusting the value of the random angular momentum to reproduce the observed clusters profiles4. The data in the present study does not provide the DM density profile, while we have the total density profile from lensing, and the profile of gas mass recovered from literature. Being the last the major contributor to the baryonic component, it is needed, and used to constrain the baryon fraction. Hence, unlike [51, 52], the DM content for the CLASH clusters cannot be extracted from fitting the model to the total and baryon density profiles separately. The present paper reproduces the CLASH density profiles in a different manner. In the same way as in [51, 52], we assume that the final mass of the protostructure generating each of the CLASH clusters is equal to the ob-

4We recall that clusters of galaxies are not supported by rotation, and that their “ordered” angular momentum, coming from tidal torques, has very similar and low values [some km s−1 66, 67], for all clusters, and in terms of the spin parameter λ can be fixed to the typical value λ = 0.03 [68].

9 served mass of that cluster, namely M500 from [41], and that the surface density given from observation is reproduced by the model. However and contrary to

[51, 52], the cluster baryon fraction Fb is not fixed. Instead, we assume that the system initially starts with a baryon fraction equal to the “universal” baryon fraction fb = 0.17 ± 0.01 [69] while the final baryon fraction is calculated tak- ing into account the major baryonic components: Intra-cluster Medium (ICM) (mainly primordial), intracluster stars, and stars in galaxies, (the latter) deter- mined from the star-formation processes (see Appendix A). In the same way, the “random” angular momentum is no longer a parameter modified to im- prove the agreement between the model and data, but has been fixed following [70, 71]. Given the cluster mass, and the total mass profile, the model repro- duces the DM, stars, and gas profiles. A qualitative summary of this model (see also Appendix A for details) goes as follows. In our model, the protostructure contains baryons and DM. After growing in a linear way, the density contrast becomes large enough to stop the protostructure expansion with the Hubble flow, making it recollapse. DM collapses first and baryons follow. Clumps are formed because of radiative processes, and collapse to the protostructure cen- ter to form stars. At high (z ≃ 5 in the case of a protostructure of 9 10 M⊙), the collapsing DM compresses baryons (adiabatic contraction). The formed clumps transfer energy and angular momentum to DM through dynam- ical friction. Then the amplitude of DM random motions increases, and DM moves toward the outskirts of the protostructure, resulting in a reduction of the central DM density of the forming structure and erasing or flattening of the initial cuspy profile. Protostructures giving rise to rotation supported galaxies suffer from a further flattening due to the acquisition of angular momentum in the collapse phase. SN explosions at a later epoch (z ≃ 2) produce expulsion of gas, and the smallest clumps remaining after star formation are disrupted [50]. AGN feedback has a similar effect on larger scales to that of SN feedback [48].

10 3. Results and discussion

In [13, 17, 51, 52], we showed how the environment, angular momentum, and baryon content influence the characteristics of cluster structure. The inner density profiles of clusters are flatter in clusters with larger angular momentum [70, 72, 73, 74, 75, 71, 76, 77] and larger baryon fraction (especially in the central region). In [52], we reproduced the total and DM density profiles of [38, 39] and those correlations found by these authors. The results obtained by [38, 39] are based on strong-lensing, weak-lensing, and improved with respect to their previous work [33]. The [39] data reduced the degeneracy between the stellar and DM masses, thanks to the determination of the stellar mass scale and by accounting for the BCGs homogeneity [39, see their Sec. 4]. This resulted in a more physically consistent analysis. They showed that the total density profile is well approximated by a cuspy NFW profile, while the DM profile is found to be flatter. In Fig. 1 (top panels), we compare the total surface mass density profiles Σ obtained by [41] (diamond symbols with error bars, its NFW profile fit and the 68% CL – Confidence Level) with our model predictions, shown with dashed lines, for the cases of A383 (left panel) and A611 (right panel). As shown here, our model reproduces well the [41] surface mass density profiles and shows that, similarly to what was found in [38, 39, 52], the total mass profile is well approx- imated by a cuspy NFW-like profile for both clusters. In the bottom parts of the same panels, the solid lines represent the gas surface density, (average value and the 68% CL), obtained by projecting the 3-D density [78] obtained from the ACCEPT project given in https://web.pa.msu.edu/astro/MC2/accept/. The dashed lines are the prediction of our model. In the bottom panels, we show the 3D density profiles of A383 and A611 obtained with our semi-analytical model. The cyan band represents the stellar content, the magenta one is the DM content, the brown one is the total mass, and the blue band the gas. The dashed lines represent the total mass from the [41] data.

11 Total 2-D mass density: Umetsu Total 2-D mass density: Umetsu ____ Total 2-D mass density: NFW fit ____ Total 2-D mass density: NFW fit _ _ _Total 2-D mass density: Our model _ _ _Total 2-D mass density: Our model __ __2-D gass mass density: Our model __ __2-D gass mass density: Our model __ 2-D gass mass density: ACCEPT __ 2-D gass mass density: ACCEPT

o.

Dark Matter (gNFW): Our model Dark Matter (gNFW): Our model Stars (BCG): Our model Stars (BCG): Our model Total mass: Our model Total mass: Our model Total mass: Umetsu Total mass: Umetsu Gas: Our model Gas: Our model

A383

Figure 1: Total, and gas surface mass density profiles for A383 (top-left) and A611 (top-right). The diamond symbols with error bars show the data from [41], while the central and external lines represent the best-fit NFW profile and the 68% CL, respectively. The dashed lines are the result of our model. In the bottom parts of the same panels is plotted the gas surface density obtained projecting the 3-D surface density [78] given in the webpage of the ACCEPT project https://web.pa.msu.edu/astro/MC2/accept/, solid lines, and the prediction of our model, dashed lines. The bottom-left and bottom-right panels show the 3D density profiles of A383 and A611 obtained with our model, respectively. The cyan, magenta, brown, and blue bands represent the stellar, DM, total matter, and gas density profiles, respectively. The dashed line represents the [41] total density profile. The width of the band indicates the 1σ 12 uncertainty [see Sec. 4.3 of 39]. Dark Matter: Newman Stars (BCG): Newman

Gas (ACCEPT) Dark Matter, BCG, gas: Our model A383

Figure 2: Comparison of the [39] DM, stars, and gas density profiles of A383 and A611 with our model’s predictions. The cyan, magenta, and blue bands represent the stellar, DM, and gas density profiles, respectively. The width of the band indicates the 1σ uncertainty [see Sec. 4.3 of 39].

13 18

] 20 −2

22

24

26

_____ Fit Newman 28 A611(Data) Surface brightness [mag arcsec A383 ...... Our Model 30 1 10 Elliptical radius [arcsec]

Figure 3: Comparison of the [39] profiles of surface brightness for the BCGs of A611, and A383 with our model’s predictions. The symbols and the solid lines are the data, and pseudo- isothermal elliptical mass distribution [dPIE, see Eq. 1 of 39] fits to the data, respectively, obtained by [39]. The dotted lines are the dPIE fits done in our model.

The arrows at the bottom represent the three-dimensional BCG half-light radius. In each case, the width of the band indicates the 1σ uncertainty [see Sec. 4.3 of 39]. Tables 1 and 2 summarize all the CLASH parameters found in our analysis.

The 2D density profile slope, α2D, was obtained using 3 adjacent radial bins fitted with a 2-parameter power-law profile, accounting for the averaging effect [41].5 The 3D DM density slope α is obtained from fitting the spherical DM

5The surface mass density profile is described by 15 radial bins. Thus the local slopes can be measured at 13 sets of 3 adjacent radial bins, yielding the logarithmic density slope as a

14 ______MACS1931 MS2137 ______RXJ2248 MACS1720 ______

o. o.

MACS1115 ______A2261 ______RXJ2129 ______A209

o. o.

______MACSJ1206 ______RXJ1347 MACS1931 ______MACS0429 ______MACSJ0744

. o o.

Figure 4: Comparison of the gas mass profiles inferred by JACO [Joint Analysis of Cluster Observations, 79] from Chandra, solid thick lines, with the predictions of our model, dashed lines, for all the clusters except A383, and A611, since for those clusters the comparison was already done in Figs. 1,2.

15 Figure 5: Inner 3D slope of the DM density α versus cluster baryon fraction Fb (left), and α versus stellar mass fraction fstar (right), obtained with our model (square symbols with error bars). In each panel, the solid line represents the orthogonal distance regression (ODR) fit to the data. The Spearman correlation coefficients for the respective relations are 0.43 and -0.44.

profile with a generalized NFW profile (gNFW).6 In the inner ≃ 5 − 10kpc region [see 39, Fig. 3] of A383 and A611, where the BCG mass becomes comparable or larger than the DM mass, Fig. 1 (bottom panels) shows that the density profiles flatten (similarly for the other clusters,

function of the cluster-centric radius. 6The gNFW profile is written as

ρs ρDM(r) = α 3−α , (5) (r/rs) (1 + r/rs) with a central cusp slope given by d log ρDM/d log r →−α for r → 0

16 whose plots were not shown). At this point, we have to notice that while the flattening of the DM distribu- tion occurs within ≃ 10 kpc, region in which the mass of the BCG is dominating, the top row of Fig. 1 shows that the surface density does not extend to radii < 10 kpc. In that region, we reproduced the total density profile given by [41] with our model, as the model produces the three dimensional distribution of total mass, DM, and baryonic density profiles. In the same way as the model reproduces the total mass density profile of [41], and reconstructs the DM, and baryonic distribution, it also reconstructs the density profiles a radii < 10 kpc. In other terms the model gives a ”physical extrapolation” in the region for which there is no data. A legitimate question one can ask is whether the model is robust enough to allow this kind of extrapolation. The model, similarly to hydro-dynamic simulations [see 48, 80, 49], takes into account a large number of physical effects (e.g., gas cooling, star formation, supernovae feedback, AGN feedback, etc.). It has been checked in several different situations. In some studies, it has predicted several important effects in advance of hydro-dynamical simulations, and has also been compared with hydro-simulations. An extensive discussion on the robustness of the model can be found in the ending part of the Appendix A. In addition, the correlations we find in the paper are related to properties of the BCG (Figs. 6-8). In Figs. 6, 7, we show a correlation between the inner slope of DM and parameters of the BCG which, as discussed by [38, 39], is very ∆dlogρ little influenced by the gas ( dlogr ≤ 0.05). Similarly in Fig. 9, the correlation is found between the mass content within 100 kpc, mainly composed of DM, and the mass inside 5 kpc, dominated by the BCG mass. However, as previously discussed, in the CLASH sample there is no data inside 10 kpc. Moreover, as we discuss in the following, there exists only one study on CLASH BCGs [81] in the literature. Nevertheless, that study gives discrepant results compared with [38, 39] for the characteristics of the BCGs of the clusters common to the two studies [i.e. 38, 39, 81], namely A383, and A611, while [38, 39] are in agreement with our BCGs characteristics. Therefore, we

17 prefer not to compare the correlations we found from the model predictions with [81] data because of the apparent tension with other studies. Our results are then based on the model robustness (see the following) and should be checked against future observations. The increased role of the baryon mass at these radii (mainly the BCG mass) steepens the total density profile compared with the DM density profile, whereas for radii ≥ 5−10 kpc, in all clusters, DM dominates over the baryon component. As a result, the outer total and DM density profiles are very similar, and their slopes outside the inner region are comparable in the different clusters, in agree- ment with the NFW profile. Since the total density profile is consistent with the NFW profile at r ≃ 5 − 30 kpc, this implies a “tight coordination” between the stellar distribution and the inner DM profile, as found by [39]. Since the total mass (DM and baryonic matter) in the inner cluster region follows the NFW form, while the baryonic component is dominant in the 5 − 10kpc central region, the DM central density profile is flatter than the NFW profile. These trends are more clearly visible in Fig. 5 and better in Figs. 6 and 7. Although the model, summarized in points a-e of the Appendix A, was validated in previous papers, we compare, in Fig. 2, the model’s prediction for the DM, and baryons (stars) density profile with the observations of [39], as an illustration of how the model equally correctly predicts the DM and baryons distribution. For consistency with Fig. 1, we only display the results for clusters A383 and A611. However, all [39] clusters were checked. As seen in Fig. 2, our model (dashed lines) gives a very good approximation to the DM density profile (magenta band), baryons (i.e. stars, cyan band), and gas density profile. We provide a further check on our model’s baryon distribution predictions in Fig. 3, with a comparison, for the two illustrative clusters, of the model with the observed [39] surface brightness profiles. They both present good agreement between observed and model surface brightness.

For those clusters, their stellar mass-to-light ratio, M∗/L, is equal to: 2.3 (A383), and 2.2 (A611), in agreement with [39, Table 4]. With the same aim to validate the model, in Fig. 4 we compare the gas

18 mass profile, Mgas. This is one of the best proxy to check the SAM predic- tions, because it can be directly recovered from X-ray observations, without any assumption. We present profiles for all clusters (except A611, and A383 for which we already checked the density profile in Fig. 1) obtained from [82, Chandra/JACO dataset]7 with the profiles from our model. As shown, the plots display a good agreement between the observations (average and 68% CL) and the average values of the SAM’s prediction. A comparison between the stellar content of the clusters and our model is not possible because data are lacking. Using the profiles of [83], or calculating Mstar, assuming that all clus- ters have the same baryon fraction, fbar, as in Mstar = fbarMTotal − Mgas, gives average profiles of Mstar. To correctly compare with our predictions, we need measurements of the stellar component which are not yet available. A similar problem is present for the Brightest Cluster Galaxy (BCG). We can only com- pare clusters in which the stellar component and the BCG characteristics have been measured. Such comparison was done for clusters with the corresponding measurements: A611, and A383 plotted in Fig. 28. Recall that in previous pa- pers we showed how our model gives good predictions for the gas, stars, and DM content in structures [e.g. 84], as well as the baryonic Tully-Fisher (TF) re- lation (i.e., baryonic mass, rotation velocity relation), the Mstar −Vrot, Mstar −σ relations (for σ the dispersion velocity and Vrot the circular rotation velocity) and the Mstar − Mhalo relation [see 85]. In Fig. 5 we show the correlation between the 3D DM density slope α and the baryon fraction Fb (left panel) and the correlation between α and the stellar mass fraction fstar (right panel). The Spearman correlation coefficients are ≃ 0.43 and ≃−0.44, respectively. We note that these correlations are stronger than those of the 2D slope of the total density (α2D) with Fb and fstar (not shown). This is because, unlike the 2D total mass profile, the DM density profile varies significantly from cluster to cluster

7 JACO [79] is a tool to derive gas and hydrostatic equilibrium density profiles. 8Although not presented, we get a similar result for MS2137

19 Figure 6 compares the inner 3D DM slope (α) and the BCG stellar mass

(Me). The DM slope parameter α was obtained by parameterizing each cluster with a gNFW profile. From the figure, it is evident that the larger the cluster’s BCG mass, the flatter its inner density profile, in agreement with [51, 39, 52]. This is expected for the following reason: Since the total (DM + baryons) den- sity profile is well represented by a cuspy NFW profile and is dominated by stellar baryons inside 5 − 10 kpc, this implies that the steeper the baryon inner profile, the flatter the DM profile. The best-fit line is derived from orthogo- nal distance regression (ODR), which takes into account the uncertainties in the variables different from the least-square method. We calculated again the Spearman correlation coefficient, which in this case is −0.6, with a p-value of 9 0.02, testifying for the correlation between α and Me. Here, we cannot make a statistical comparison between the parameters de- rived from our modeling and those measured directly from observations, because there is just one systematic study of CLASH BCGs [81] in the literature, and it does not investigate the α–Me correlation. The α–Me relation was studied for two of the CLASH clusters, namely A383, and A611 by [39, 52]. Our results for the α–Me relation are in agreement with [39]. Here, we would like to point out that a comparison of the BCG masses common to [81] and [39, 52] (A383, A611) shows that while those in [52] are in agreement with [39], only A383 agree with [81].

In Fig. 7, we show the α–Re relation, with Re the BCG effective radius (radius containing half of the total light). The figure shows that a cluster with larger Re have a shallower inner slope, which is in line with what was discussed above. That is, clusters with larger BCGs contain more stellar baryons and thus less DM (MDM = Mtotal − Mb) in their central region, resulting in a flatter DM slope. The solid line in the figure represents the ODR fit, with Spearman correlation coefficient of −0.63 and a p-value of 0.0012.

In Fig. 8, we show the correlation between Re and the halo mass M500 (at

9The probability that the “null” hypothesis (the true correlation is zero) is true.

20 ..

Figure 6: Correlation between the inner DM slope α and the BCG mass Me. The BCG stellar mass Me was obtained using our model. For each cluster, the uncertainty of the BCG mass is assumed to be 0.07dex [39, 52]. The solid line is the ODR fit to the data. The Spearman correlation coefficient is -0.6.

21 Figure 7: Correlation between the inner DM slope α and the BCG effective radius Re. The line represents the ODR fit to the data, with a Spearman correlation coefficient of −0.63 and a p-value of 0.0012.

22 r500 ∼ 1.4 Mpc for our sample). The solid line represents again the ODR fit. This comparison exhibits a positive correlation with a Spearman correlation coefficient of −0.68 and a p-value of 0.005. This is in agreement with previous studies [e.g., 86], but in contradiction again with [81]. The latter found no correlation between the BCG mass and the cluster halo mass. They ascribed this lack of correlation to a selection bias in the CLASH sample (whereby clusters with BCGs in a narrow mass range have been selected), but without a clear justification. However, as previously noticed, the BCG mass estimates between [81] and [39] in their overlapping clusters are in striking conflict with each other. Moreover, the inclusion of the high-magnification CLASH clusters in [81] may imply a bias in their results, because they are highly disturbed merging clusters. Our model [13, 17] can explain the resulting density profiles (Fig. 1) and correlations (Figs. 6 and 7) as follows: The DM protostructure starts to collapse at high z (linear phase), forming potential wells for baryons to fall in. In their collapse, baryons radiate away part of their energy and form clumps, condensing into stars [see 87, Secs. 2.2.2 and 2.2.3], while compressing DM [“adiabatic contraction”, 88, 89]. At around z ≥ 5 [13, see Figs. 3 and 5 therein], this process dissipationally produces a steep density profile, the main structure of the BCG [see also 90, 91] with scale radius Re ≃ 30 kpc, similar to size-scales of high- massive galaxies [92, 38, 39]. Extra stars are added in the outer regions by satellite mergers onto the proto-BCG [e.g., 93, 94] Moreover, dynamical friction from DM particles induces orbital decay of the baryon clumps. As a result, DM particles move outward as the baryon clumps move toward the cluster center, thus reducing the central DM density [95, 42, 43, 13, 47, 56, 50]. AGN feedback has been proposed as an alternate mechanism to flatten the DM profile [e.g., 48]. However, it appears to be “too effective” because the 10 kpc core produced is much larger than what is observed [57]. Our physical model thus predicts, at the same time, a flattening of the inner DM density profile and an anti-correlation between the inner DM slope α and the cluster’s central baryon content [43, 13, 17]. This is because the density

23 Figure 8: Correlation between the BCG effective radius Re and the cluster halo mass M500. The solid line shows the ODR fit, with a Spearman correlation coefficient of −0.68 and a p-value of 0.005.

24 . .

Figure 9: Total cluster mass contained within 100 kpc (dominated by DM) versus that con- tained within 5 kpc (dominated by stellar baryons). The solid line is the ODR fit with a Spearman correlation coefficient of 0.64 and a p-value of 0.011.

profile shape is primarily the result of interaction between DM and baryons clumps through dynamical friction and subdominantly of the action of SN/AGN feedback. Dissipative baryonic formation of the proto-BCG (z ≥ 2) follows in our model, where stars merge onto the BCG [e.g., 93, 94] and satellites infall toward the cluster center, thus kinematically “heating” DM and flattening the inner DM slope [95, 42, 13, 47]. A further correlation is expected between the star-dominated, innermost cluster mass and the cluster mass within a region that is already established before the BCG formation and thus almost unchanged subsequently [96]. This has been examined in Fig. 9 by comparing the the cluster mass within 5 kpc

25 (composed mainly of stars) and that within 100 kpc (mainly of DM) for our CLASH sample. We find a Spearman correlation coefficient of 0.64 and a p- value of 0.011. Such a strong correlation implies (see also Sec. 4) that the cluster’s progenitor halo and the innermost BCG region were formed at higher redshifts, and they have been subject to little evolution subsequently, whereas the cluster outskirts have grown via the secondary mass accretion. This “inside- out” growth scenario of cluster formation has been proposed in the framework of ΛCDM [e.g., 97, 98] and recently confirmed from a joint lensing and X-ray analysis of the cluster-halo fundamental plane for the CLASH sample [99, 100]. These correlations confirm the central role of baryons in shaping the DM density profile in galaxy clusters, in particular for the central . 10 kpc region. The BCG characteristics correlate with baryonic and cluster masses. The total density profile agrees well with the NFW form, while the DM density profile exhibits a shallower inner slope with α< 1. Our findings are consistent with [33, 36] on the cluster A383. Our results are also in line with [101], who found that the innermost region of the (non-CLASH) cluster A2589 has a shallow DM profile assuming reasonable values of mass-to- light ratios. Our cored DM profiles have a mean slope of α = 0.54 ± 0.05, in agreement with [48, 49] and [39]. DM-only simulations produce halos with cuspy, steeper density profiles. In the Phoenix project, [102] analyzed zoomed-in re-simulations of cluster-size ha- los drawn from a cosmologically representative volume in ΛCDM, finding that −3 the central density cusp (at their innermost resolved radius of r ∼ 2 × 10 r200) has an average logarithmic slope of α ≃ 1.05 with a halo-to-halo scatter of ∼ 20%. This level of “diversity” is apparently at odds with the observed shal- low slopes of the inner DM density. Including the dominant central baryonic physics on top of such DM simulations can reconcile the discrepancy and re- produce the cored DM distribution, by accounting for dynamical friction from baryonic clumps [95, 42, 54, 13, 47] and density flattening driven by SN/AGN feedback [103, 48]. These mechanisms are described in Appendix Appendix A. On the other hand, more radical solutions have also been proposed: alter-

26 native DM models [e.g., 104, 105, 106, 107], modifications of the matter power spectrum at small scales [e.g. 108], and even modified gravity.

We find that the total density profile ρtot(r) of the X-ray-selected CLASH sample is NFW-like, as found in previous studies [40, 41, 59]. The average slope of the total density for our sample is hγtoti = 1.05 ± 0.02, where γtot =

−d log ρtot/d log r, and the average slope is obtained by linear fitting in the 10 plane of log r–log ρtot in the radial range r = (0.003 − 0.03) × r200. This is in agreement with collisionless DM simulations (γtot ≃ 1), and in line with +0.05 the finding of [38, Sec. 9], hγtoti = 1.16 ± 0.05−0.07 defined in the same range r = (0.003 − 0.03) × r200. Observational results of the asymptotic total density profiles also agree with +0.31 our results (i.e., NFW-like at r & 5 − 10 kpc): αtot =0.96−0.49 for MACSJ1206 11 −1 [109] and αtot =1.08±0.07 for A383 [110]. Excluding the innermost 40 kpch region, the stacked strong+weak lensing analysis of four “superlens” clusters +0.27 (A1689, A1703, A370, Cl0024+17) gives αtot =0.89−0.39 [35].

4. Conclusions

In this paper, we have studied and characterized the total density profiles for a sample of 15 X-ray-selected CLASH clusters by improving our earlier analysis [52] based on several clusters from [38, 39]. The primary goal of this study was to test the physical picture of cluster formation proposed by [38, 39] in the frame work of a modified version of the physical model developed by [51, 52]. To this end, we analyzed binned surface mass density profiles of [41] derived from their strong-lensing, weak-lensing shear and magnification analysis of high-quality HST and Subaru data. For each cluster, we extracted the radial

10 The BCG and the the DM halo are distinct components of the cluster model, while γtot is a derived, composite parameter. 11 It is important to stress that, similarly to [38], hγtoti is the average slope of the total density measured in the range r = (0.003 − 0.03) × r200 and is different from αtot, which is the asymptotic inner slope of the total density assuming the gNFW profile.

27 profile of the total 3D density assuming spherical symmetry. We have used our semi-analytical model to interpret the total 3D density profile, which allows us to compute the baryon density profile and thus the DM density profile for the cluster. The total 3D mass density profile for our sample is characterized by a loga- rithmic slope of hγtoti =1.05±0.02 in the radial range r = (0.003−0.03)×r200, in agreement with several previous studies (e.g. [38, 39]). Stellar mass dominates the total mass at r . 5 − 10 kpc, while the cluster outskirts are dominated by DM. Such segregation reveals a “tight coordination” between the inner DM and stellar distributions, as also implied by interplay of DM and baryons that generates the NFW-like total density profile. The correlation between the mass inside 5 kpc and that inside 100 kpc (Fig. 9) further supports such a tight coordination and points to similar formation time-scales of the BCG and the inner cluster region. Thus, the cluster’s final configuration depends on the baryonic content and their formation process [17]. Therefore, in the context of hierarchical structure formation models, we should expect tight correlations between the final inner baryonic content and the BCG mass, as well as between the total baryonic and cluster masses [see 111]. Since the DM and baryon contents sum to the total mass and the baryons dominate the inner 5 − 10 kpc region, the inner slope of the DM density must be shallower than that of the total density, that is, α < 1, as shown in Fig. 1. The observed inner DM slopes α span the range [0.30, 0.79]. Correlations were also examined between several of the characteristic quan- tities of clusters (e.g., Re, Me, M500). Our findings are summarized as follows:

a. The inner 3D slope of the DM density, α, is anti-correlated with the BCG

effective radius Re. The anti-correlation reflects the balance between DM and the BCG in the cluster center. For an NFW-like total mass profile, clusters with more massive BCGs contain less central DM, implying a flatter DM slope.

b. Similarly, the inner DM slope α and the BCG mass Me are anti-correlated

28 with each other. This indicates again that a larger content of the central baryons gives rise to flatter DM profiles.

c. The cluster halo mass M500 and the BCG effective radius Re are correlated with each other, as found in previous studies [86]. d. The cluster mass inside 5 kpc, dominated by the stellar baryons, and the cluster mass inside 100 kpc, dominated by DM, are correlated with each other. This hints at early formation of the BCG and the inner cluster region, while subsequent, continuous mass accretion played a fundamental role in the growth of cluster outskirts [e.g., 99, 100].

These observed correlations are in support of the physical picture proposed by [38, 39], that clusters form from a dissipative phase that leads to steepening the central stellar density, followed by a second dissipationless phase in which interactions between baryonic clumps and DM through dynamical friction kine- matically heat the latter, leading to flat DM density profiles [95, 42, 54, 17, 51, 47, 50].

Acknowledgments

To our colleague Lee Xiguo, in memoriam. A.D.P. was supported by the Chinese Academy of Sciences and by the Presi- dent’s International Fellowship Initiative, grant no. 2017 VMA0044. M.LeD. acknowledges the financial support by Lanzhou University starting fund. A.D.P. and M.LeD. thank Keiichi Umetsu, who contributed to calculations and im- proving the paper, and the referee Stephano Ettori for discussions which helped improving our paper.

Appendix A. Appendix: the model

In this paper, we used, as in several previous studies [15, 16, 17, 51, 52], a semi-analytical model (SAM) introduced in [13] and extended in [112, 113]. This model incorporates a secondary infall model (SIM) [114, 115, 71, 76] that takes

29 into account the effect of DM adiabatic contraction [88, 89, 116] of ordered and random angular momentum [117, 76], as well as of baryon-DM angular momen- tum transfer through dynamical friction [42, 43, 13, 44, 15, 47, 50], in contrast to previous SIMs. It also accounts for cooling, reionization, star formation, and SN/AGN feedback (see the following). It starts from the Hubble flow expansion of a perturbation starting in the linear phase, following its evolution until the maximum expansion (turn-around) and recollapse to a final density, given by [74]

−1 ρta(xm) d ln f(xi) ρ(x)= 3 1+ (A.1) f(xi)  d ln xm(xi) with f(xi)= x/xi(xi), the so-called collapse factor [see Eq. A18, 13], and

1+ δi xm = g(xi)= xi −1 . (A.2) δi − (Ωi − 1)

Here xm gives the turn-around radius xm(xi), where the initial density parameter is Ωi, and a given shell’s average overdensity reads δi.

The collapse factor f(xi) of a shell with initial radius xi and apapsis (turn- around) xm(xi) can be also written as:

mp(xm) f(xi)= (A.3) mp(xm)+ madd(xm) where mp (permanent component) is the mass contained in shells with apapsis smaller than xm, and madd (additional mass) the contribution of the outer shells passing momentarily through the shell xm. The total mass is thus given by: R dm(x) mT (xm)= mp(xm)+ madd(xm)= mp(xm)+ Prm (x) dx (A.4) Zxm dx where R is the radius of the system (the apapsis of the outer shell) and the distribution of mass m(x)= m(xm), while Pxm (x) is the probability to find the shell with apapsis x inside radius xm, calculated as the ratio of the time the outer shell (with apapsis x) spends inside radius xm to its period. This last quantity can be expressed as

xm dη xp vx(η) Pxm (x)= R x dη (A.5) xp vx(η) R 30 where xp is the pericenter of the shell with apsis x and vx(η) is the radial velocity of the shell with apapsis x as it passes radius η. This radial velocity can be obtained by integrating the equation of motion of the shell: dv2 h2 + j2 m Λ +2µv2 =2 − G T + r v (A.6) dt  r3 r2 3  which can be solved numerically for v. The final density (Eq. A.1) is then a complex aggregate depending on several quantities like angular momentum, dynamical friction, etc. Summarizing, in order to obtain the density profile for a given shell, knowing initial conditions, angular momentum and coefficient of dynamical friction, one needs to integrate the equation of motions (Eq. A.6), then to calculate the probability Pxi , yielding the contribution of the shell to madd (see Eq. A.4), the collapse factor (Eq. A.3) and finally the density profile through Eq. (A.1) (see also [118, Sect. 4]; [71, Sect. 2.1]). The perturbation contains DM and baryons, the latter initially in the gas phase, with the amount set by the cosmic baryon fraction fb =0.17 ± 0.01 [69] [while it is 0.167 in 1]. The tidal torques of larger scales on smaller-scale structures produce the “ordered” angular momentum h through the tidal torque theory [119, 120, 121, 117, 122], while random velocities generate “random” angular momentum [123], j, expressed in terms of the eccentricity e = rmin/rmax [70]. Here rmax denotes the apocentric radius, and rmin the pericentric radius. The dynamical state of the system induces eccentricity to be corrected as in simulations of [71], 0.1 rmax e(rmax) ≃ 0.8 (A.7)  rta  for rmax < 0.1rta. Dynamical friction was introduced in the equation of motion with a decelera- tion term [Eq. A14, 13], whose coefficient proceeds as in [124] and [13, Appendix D]. Baryon collapse induces adiabatic contraction (AC) of DM, according to the following mechanism. From a protostructure of fb = Mb/M500 ≪ 1 baryons

31 12 and 1 − fb DM, the baryons cool and collapse toward the halo center, forming the distribution Mb(r). This compresses DM, relocating particles from ri to

r [Mb(r)+ MDM(r)] = riMi(ri) (A.8)

[88], with Mi(ri) the initial total mass and MDM the final DM distribution. Assuming equal initial baryon and DM distributions [125, 126, 127, 128] and the final Hernquist baryon distribution [129, 128, 127], Mi(ri) and Mb(r) are given, and the DM mass distribution in the absence of shell crossing is obtained by solving Eqs. (A.8) and (A.9),

MDM(r) =(1 − Fb)Mi(ri) (A.9) to find the final halo distribution. Conservation of angular momentum, repre- sented by the product of the orbit-averaged radiusr ¯ with its inner mass [89],

M(¯r)¯r = const., (A.10) using 2 rmax dr r¯ = r , (A.11) Tr Zrmin vr and Tr as the radial period, improves the model. Our treatment of star formation, gas cooling, reionization, and supernovae feedback (SNF) follows [130, 87, Secs. 2.2.2 and 2.2.3] With reionization as in [87], the baryon fraction is reduced in the redshift range z = [11.5, 15] and modified as

fb fb,halo(z,MVir)= 3 , (A.12) [1 + 0.26MF(z)/MVir] calculated with MF the “filtering” mass [see 131] and the virial mass, MVir (∆ =

200) was converted to the cluster halo mass M500 following [132]. Although the [117] treatment yields similar results, a classical cooling flow [e.g., 133] [see Sec. 2.2.2 of 87] is used here for gas cooling.

12 The gas mass, Mgas, and mass in stars, M∗, combine into the total baryonic mass, Mb.

32 We follow [130] to account for star-formation processes. SNF follows the treatment of [134]. The blast-wave SNF [135] was used in [136]. Although our choice of formalism is similar, it is not very fundamental, and our model differs from their SNF model [e.g., 136] in the occurrence of the flattening process before star formation and in our gravitational source of energy, whereas the SNF flattening process occurs after star formation with stellar feedback, which is in act as the energy source after core formation disrupting the gas clouds. The stellar formation occurs once a disc is formed from gas with the star formation rate

ψ =0.03Msf/tdyn, (A.13) which is computed with the disc dynamical time, tdyn, and the gas mass Msf contained in regions where its density is above a given density threshold, n > 9.3 cm−3, in the same way as in [136]. We use the Chabrier [137] initial mass function (IMF), forming stars with

∆M∗ = ψ∆t (A.14) per units of time ∆t. SNF injects energy in the (ISM) at a rate of

2 ∆ESN =0.5ǫhalo ∆M∗ VSN, (A.15)

2 obtained with the energy injected per supernova and per unit solar mass, VSN =

ηSNESN. The fixed efficiency ǫhalo = 0.35 [87] of the disc gas reheating by this energy injection is computed with the supernova number per solar mass, −3 ηSN = 8 × 10 /M⊙, assuming a Chabrier IMF [137], and with the typical 51 energy released in a SN explosion, ESN = 10 erg. This energy injection reheats the gas in proportion to the star-formation number,

∆Mreheat =3.5∆M∗, (A.16) inducing a thermal energy change

2 ∆Ehot =0.5∆MreheatVV ir, (A.17)

33 1/3 where VV ir = (10GH(z)MV ir) with H(z) being the Hubble constant at red- shift z. The last will eject a quantity of the hot gas equal to

∆ESN − ∆Ehot ∆Meject = 2 (A.18) 0.5VV ir from the halo with the condition ∆ESN > ∆Ehot.

All the quantitities evaluated at virial radius (e.g., MVir, VVir) were converted to corresponding quantities at ∆ = 500, by following [132]. Subsequently, the hot gas can be accreted by the halo into its hot component in link with the central galaxy [138, 134]. 11 Accounting for AGN quenching is required for masses M ≃ 6 × 10 M⊙ [139]. Here, we use the prescription of [48, 80] to implement AGN feedback: A 6 −3 conjunction of the star density above 2.4 × 10 M⊙ kpc , the gas density at 10 times the stellar density, and the 3D velocity dispersion exceeding 100kms−1 5 creates an initial (seed) super-massive black hole (SMBH) with 10 M⊙, which grows with accretion and produces AGN feedback following a variant of the [140] model modified according to [48]. Our model is a semi-analytic simulation (SAM), such as, e.g., GALFORM, and GALACTICUS. Its parameters are the same found in (e.g., SPH) simula- tions (e.g., baryon fraction fixed as the cosmic baryon fraction value, a parame- ter fixing the random angular momentum, set as in simulations, star formation rate, density treshold of star formation, parameters related to supernovae and AGN feedback and other parameters discussed in the model section). A more complete description of the model can be found in [113] , and [84]. Finally, the robustness of the model has been demonstrated in various man- ners as summarized below.

a. DM heating by collapsing baryonic clumps that induce cusp flattening was predicted in 2009 for galaxies, and 2012 for clusters. The predictions are in agreement with the studies of [95, 42, 54, 55, 47, 56, 50]. Our model is compared with the smoothed particle hydrodynamics (SPH) simulations of [14] in Fig. 4 of [16].

34 b. Correct predictions for the galaxy density profiles [13, 44] and the cluster density profiles [51] were made before SPH simulations of [14, 141] and [49], respectively. c. The dependence of the inner DM density slope on the halo mass [15, Fig. 2a, solid line] was predicted before the quasi similar results of [136,

Fig. 6, solid lines] presented in terms of the rotation velocity, Vc [Vc = −2 0.316 2.8 × 10 Mvir , 142]. d. The dependence of the cluster baryon fraction on the inner DM slope was also predicted in [51], which was later studied and confirmed in SPH simulations [136]. e. Figure 1 of [112, 113] compares the dependence of the inner DM slope on the halo mass with the [136] simulations. Predictions for the Tully–Fisher

relation, the Faber–Jackson relation, and the M∗ − Mhalo, relation are compared with simulations in Figs. 4 and 5 of [112, 113].

References

[1] E. Komatsu, K. M. Smith, J. Dunkley, et al., Seven-year Wilkin- son Microwave Anisotropy Probe (WMAP) Observations: Cosmo- logical Interpretation, ApJS192 (2011) 18–+. arXiv:1001.4538, doi:10.1088/0067-0049/192/2/18.

[2] Planck Collaboration, P. A. R. Ade, N. Aghanim, C. Armitage- Caplan, M. Arnaud, M. Ashdown, F. Atrio-Barandela, J. Aumont, C. Baccigalupi, A. J. Banday, et al., Planck 2013 results. XVI. Cosmological parameters, A& A571 (2014) A16. arXiv:1303.5076, doi:10.1051/0004-6361/201321591.

[3] A. Del Popolo, Non-baryonic dark matter in cosmology, in: AIP Conf. Proc., Vol. 1548, 2013, pp. 2–63. doi:10.1063/1.4817029.

[4] A. Del Popolo, Dark matter, density perturbations, and structure for-

35 mation, Astronomy Reports 51 (2007) 169–196. arXiv:0801.1091, doi:10.1134/S1063772907030018.

[5] A. Del Popolo, Nonbaryonic Dark Matter in Cosmology, International Journal of Modern Physics D 23 (2014) 30005. arXiv:1305.0456, doi:10.1142/S0218271814300055.

[6] J. F. Navarro, C. S. Frenk, S. D. M. White, A Universal Den- sity Profile from Hierarchical Clustering, ApJ490 (1997) 493–+. arXiv:astro-ph/9611107, doi:10.1086/304888.

[7] J. F. Navarro, A. Ludlow, V. Springel, J. Wang, M. Vogelsberger, S. D. M. White, A. Jenkins, C. S. Frenk, A. Helmi, The diversity and simi- larity of simulated cold dark matter haloes, MNRAS402 (2010) 21–34. arXiv:0810.1522, doi:10.1111/j.1365-2966.2009.15878.x.

[8] B. Moore, Evidence against dissipation-less dark matter from observations of galaxy haloes, Nature370 (1994) 629–631. doi:10.1038/370629a0.

[9] R. A. Flores, J. R. Primack, Observational and theoretical con- straints on singular dark matter halos, ApJL427 (1994) L1–L4. arXiv:astro-ph/9402004, doi:10.1086/187350.

[10] A. Agnello, N. W. Evans, A Virial Core in the Sculptor , ApJL754 (2012) L39. arXiv:1205.6673, doi:10.1088/2041-8205/754/2/L39.

[11] J. J. Adams, J. D. Simon, M. H. Fabricius, R. C. E. van den Bosch, J. C. Barentine, R. Bender, K. Gebhardt, G. J. Hill, J. D. Murphy, R. A. Swaters, J. Thomas, G. van de Ven, Dark Matter Density Profiles Inferred from Stellar and Gas Kinematics, ApJ789 (2014) 63. arXiv:1405.4854, doi:10.1088/0004-637X/789/1/63.

[12] J. F. Navarro, C. S. Frenk, S. D. M. White, The Structure of Cold Dark Matter Halos, ApJ462 (1996) 563. arXiv:astro-ph/9508025, doi:10.1086/177173.

36 [13] A. Del Popolo, The Cusp/Core Problem and the Secondary Infall Model, ApJ698 (2009) 2093–2113. arXiv:0906.4447, doi:10.1088/0004-637X/698/2/2093.

[14] F. Governato, C. Brook, L. Mayer, A. Brooks, G. Rhee, J. Wadsley, P. Jonsson, B. Willman, G. Stinson, T. Quinn, P. Madau, Bulgeless dwarf galaxies and dark matter cores from supernova-driven outflows, Nature463 (2010) 203–206. arXiv:0911.2237, doi:10.1038/nature08640.

[15] A. Del Popolo, On the universality of density pro- files, MNRAS408 (2010) 1808–1817. arXiv:1012.4322, doi:10.1111/j.1365-2966.2010.17288.x.

[16] A. Del Popolo, Non-power law behavior of the radial profile of phase-space density of halos, JCAP7 (2011) 14. arXiv:1112.4185, doi:10.1088/1475-7516/2011/07/014.

[17] A. Del Popolo, Density profile slopes of dwarf galaxies and their environment, MNRAS419 (2012) 971–984. arXiv:1105.0090, doi:10.1111/j.1365-2966.2011.19754.x.

[18] J. D. Simon, A. D. Bolatto, A. Leroy, L. Blitz, E. L. Gates, High- Resolution Measurements of the Halos of Four Dark Matter-Dominated Galaxies: Deviations from a Universal Density Profile, ApJ621 (2005) 757–776. arXiv:astro-ph/0412035, doi:10.1086/427684.

[19] K. A. Oman, J. F. Navarro, A. Fattahi, C. S. Frenk, T. Sawala, S. D. M. White, R. Bower, R. A. Crain, M. Furlong, M. Schaller, J. Schaye, T. Theuns, The unexpected diversity of dwarf galaxy rotation curves, MNRAS452 (2015) 3650–3665. arXiv:1504.01437, doi:10.1093/mnras/stv1504.

[20] S. Weinberg, The cosmological constant problem, Reviews of Modern Physics 61 (1989) 1–23. doi:10.1103/RevModPhys.61.1.

37 [21] A. V. Astashenok, A. del Popolo, Cosmological measure with volume averaging and the vacuum energy problem, Classical and Quantum Gravity 29 (8) (2012) 085014. arXiv:1203.2290, doi:10.1088/0264-9381/29/8/085014.

[22] A. Del Popolo, F. Pace, J. A. S. Lima, Extended Spherical Collapse and the Accelerating Universe, International Journal of Modern Physics D 22 (2013) 50038. arXiv:1207.5789, doi:10.1142/S0218271813500387.

[23] A. Del Popolo, F. Pace, J. A. S. Lima, Spherical collapse model with shear and angular momentum in dark energy cosmologies, MNRAS430 (2013) 628–637. arXiv:1212.5092, doi:10.1093/mnras/sts669.

[24] A. Del Popolo, F. Pace, S. P. Maydanyuk, J. A. S. Lima, J. F. Jesus, Shear and rotation in Chaplygin cosmology, PRD87 (4) (2013) 043527. arXiv:1303.3628, doi:10.1103/PhysRevD.87.043527.

[25] A. Del Popolo, F. Pace, M. Le Delliou, A high precision semi- analytic mass function, JCAP3 (2017) 032. arXiv:1703.06918, doi:10.1088/1475-7516/2017/03/032.

[26] A. Del Popolo, M. Le Delliou, Small Scale Problems of the ΛCDM Model: A Short Review, Galaxies 5 (2017) 17. arXiv:1606.07790, doi:10.3390/galaxies5010017.

[27] A. Del Popolo, On the evolution of aspherical perturbations in the universe: An analytical model, A& A387 (2002) 759–777. arXiv:astro-ph/0202436, doi:10.1051/0004-6361:20020399.

[28] N. C. Amorisco, N. W. Evans, Dark matter cores and cusps: the case of multiple stellar populations in dwarf spheroidals, MNRAS419 (2012) 184–196. arXiv:1106.1062, doi:10.1111/j.1365-2966.2011.19684.x.

[29] J. R. Jardel, K. Gebhardt, The Dark Matter Density Profile of the Fornax Dwarf, ApJ746 (2012) 89. arXiv:1112.0319, doi:10.1088/0004-637X/746/1/89.

38 [30] J. R. Jardel, K. Gebhardt, M. H. Fabricius, N. Drory, M. J. Williams, Measuring Dark Matter Profiles Non-Parametrically in Dwarf Spheroidals: An Application to Draco, ApJ763 (2013) 91. arXiv:1211.5376, doi:10.1088/0004-637X/763/2/91.

[31] J. R. Jardel, K. Gebhardt, Variations in a Universal Dark Matter Profile for Dwarf Spheroidals, ApJL775 (2013) L30. doi:10.1088/2041-8205/775/1/L30.

[32] D. J. Sand, T. Treu, R. S. Ellis, The Dark Matter Density Profile of the Lensing Cluster MS 2137-23: A Test of the Cold Dark Mat- ter Paradigm, ApJL574 (2002) L129–L133. arXiv:astro-ph/0207048, doi:10.1086/342530.

[33] D. J. Sand, T. Treu, G. P. Smith, R. S. Ellis, The Dark Matter Dis- tribution in the Central Regions of Galaxy Clusters: Implications for Cold Dark Matter, ApJ604 (2004) 88–107. arXiv:astro-ph/0309465, doi:10.1086/382146.

[34] A. B. Newman, T. Treu, R. S. Ellis, D. J. Sand, J. Richard, P. J. Mar- shall, P. Capak, S. Miyazaki, The Distribution of Dark Matter Over Three Decades in Radius in the Lensing Cluster Abell 611, ApJ706 (2009) 1078– 1094. arXiv:0909.3527, doi:10.1088/0004-637X/706/2/1078.

[35] K. Umetsu, T. Broadhurst, A. Zitrin, E. Medezinski, D. Coe, M. Postman, A Precise Cluster Mass Profile Averaged from the Highest-quality Lensing Data, ApJ738 (2011) 41. arXiv:1105.0444, doi:10.1088/0004-637X/738/1/41. URL http://adsabs.harvard.edu/abs/2011ApJ...738...41U

[36] A. B. Newman, T. Treu, R. S. Ellis, D. J. Sand, The Dark Matter Distribution in A383: Evidence for a Shallow Density Cusp from Im- proved Lensing, Stellar Kinematic, and X-ray Data, ApJL728 (2011) L39. arXiv:1101.3553, doi:10.1088/2041-8205/728/2/L39.

39 [37] N. Okabe, G. P. Smith, K. Umetsu, M. Takada, T. Futamase, LoCuSS: The Mass Density Profile of Massive Galaxy Clusters at z = 0.2, ApJL769 (2013) L35. arXiv:1302.2728, doi:10.1088/2041-8205/769/2/L35.

[38] A. B. Newman, T. Treu, R. S. Ellis, D. J. Sand, C. Nipoti, J. Richard, E. Jullo, The Density Profiles of Massive, Relaxed Galaxy Clusters. I. The Total Density Over Three Decades in Radius, ApJ765 (2013) 24. arXiv:1209.1391, doi:10.1088/0004-637X/765/1/24.

[39] A. B. Newman, T. Treu, R. S. Ellis, D. J. Sand, The Density Pro- files of Massive, Relaxed Galaxy Clusters. II. Separating Luminous and Dark Matter in Cluster Cores, ApJ765 (2013) 25. arXiv:1209.1392, doi:10.1088/0004-637X/765/1/25.

[40] K. Umetsu, E. Medezinski, M. Nonino, J. Merten, M. Postman, M. Meneghetti, M. Donahue, N. Czakon, A. Molino, S. Seitz, D. Gruen, D. Lemze, I. Balestra, N. Ben´ıtez, A. Biviano, T. Broadhurst, H. Ford, C. Grillo, A. Koekemoer, P. Melchior, A. Mercurio, J. Moustakas, P. Rosati, A. Zitrin, CLASH: Weak-lensing Shear-and-magnification Anal- ysis of 20 Galaxy Clusters, ApJ795 (2014) 163. arXiv:1404.1375, doi:10.1088/0004-637X/795/2/163.

[41] K. Umetsu, A. Zitrin, D. Gruen, J. Merten, M. Donahue, M. Post- man, CLASH: Joint Analysis of Strong-lensing, Weak-lensing Shear, and Magnification Data for 20 Galaxy Clusters, ApJ821 (2016) 116. arXiv:1507.04385, doi:10.3847/0004-637X/821/2/116.

[42] A. A. El-Zant, Y. Hoffman, J. Primack, F. Combes, I. Shlosman, Flat- cored Dark Matter in Cuspy Clusters of Galaxies, ApJL607 (2004) L75– L78. arXiv:astro-ph/0309412, doi:10.1086/421938.

[43] C. Nipoti, T. Treu, L. Ciotti, M. Stiavelli, Galactic cannibalism and cold dark matter density profiles, MNRAS355 (2004) 1119–1124. arXiv:astro-ph/0404127, doi:10.1111/j.1365-2966.2004.08385.x.

40 [44] A. Del Popolo, P. Kroupa, Density profiles of dark matter haloes on galactic and cluster scales, A& A502 (2009) 733–747. arXiv:0906.1146, doi:10.1051/0004-6361/200811404.

[45] V. F. Cardone, M. P. Leubner, A. Del Popolo, Spherical galaxy models as equilibrium configurations in non-extensive statistics, MNRAS414 (2011) 2265–2274. arXiv:1102.3319, doi:10.1111/j.1365-2966.2011.18543.x.

[46] V. F. Cardone, A. Del Popolo, C. Tortora, N. R. Napolitano, Secondary infall model and dark matter scaling relations in intermediate-redshift early-type galaxies, MNRAS416 (2011) 1822–1835. arXiv:1106.0364, doi:10.1111/j.1365-2966.2011.19162.x.

[47] D. R. Cole, W. Dehnen, M. I. Wilkinson, Weakening dark matter cusps by clumpy baryonic infall, MNRAS416 (2011) 1118–1134. arXiv:1105.4050, doi:10.1111/j.1365-2966.2011.19110.x.

[48] D. Martizzi, R. Teyssier, B. Moore, T. Wentz, The effects of baryon physics, black holes and feedback on the mass distribution in clusters of galaxies, MNRAS422 (2012) 3081–3091. arXiv:1112.2752, doi:10.1111/j.1365-2966.2012.20879.x.

[49] D. Martizzi, R. Teyssier, B. Moore, Cusp-core transformations induced by AGN feedback in the progenitors of cluster galaxies, MNRAS432 (2013) 1947–1954. arXiv:1211.2648, doi:10.1093/mnras/stt297.

[50] C. Nipoti, J. Binney, Early flattening of dark matter cusps in dwarf spheroidal galaxies, MNRAS446 (2015) 1820–1828. arXiv:1410.6169, doi:10.1093/mnras/stu2217.

[51] A. Del Popolo, On the density-profile slope of clusters of galaxies, MNRAS424 (2012) 38–51. arXiv:1204.4439, doi:10.1111/j.1365-2966.2012.21141.x.

41 [52] A. Del Popolo, The flat density profiles of massive, and re- laxed galaxy clusters, JCAP7 (2014) 019. arXiv:1407.4347, doi:10.1088/1475-7516/2014/07/019.

[53] C.-P. Ma, M. Boylan-Kolchin, Are Halos of Collisionless Cold Dark Matter Collisionless?, Physical Review Letters 93 (2) (2004) 021301. arXiv:astro-ph/0403102, doi:10.1103/PhysRevLett.93.021301.

[54] E. Romano-D´ıaz, I. Shlosman, Y. Hoffman, C. Heller, Erasing Dark Mat- ter Cusps in Cosmological Galactic Halos with Baryons, ApJL685 (2008) L105–L108. arXiv:0808.0195, doi:10.1086/592687.

[55] E. Romano-D´ıaz, I. Shlosman, C. Heller, Y. Hoffman, Dissect- ing Galaxy Formation. I. Comparison Between Pure Dark Matter and Baryonic Models, ApJ702 (2009) 1250–1267. arXiv:0901.1317, doi:10.1088/0004-637X/702/2/1250.

[56] S. Inoue, T. R. Saitoh, Cores and revived cusps of dark matter haloes in formation through clump clusters, MNRAS418 (2011) 2527– 2531. arXiv:1108.0906, doi:10.1111/j.1365-2966.2011.19873.x.

[57] M. Postman, T. R. Lauer, M. Donahue, G. Graves, D. Coe, J. Moustakas, A. Koekemoer, L. Bradley, H. C. Ford, C. Grillo, A. Zitrin, D. Lemze, T. Broadhurst, L. Moustakas, B. Ascaso, E. Medezinski, D. Kelson, A Brightest Cluster Galaxy with an Extremely Large Flat Core, ApJ756 (2012) 159. arXiv:1205.3839, doi:10.1088/0004-637X/756/2/159.

[58] J. Merten, M. Meneghetti, M. Postman, K. Umetsu, A. Zitrin, E. Medezin- ski, M. Nonino, A. Koekemoer, P. Melchior, D. Gruen, L. A. Mous- takas, M. Bartelmann, O. Host, M. Donahue, D. Coe, A. Molino, S. Jou- vel, A. Monna, S. Seitz, N. Czakon, D. Lemze, J. Sayers, I. Balestra, P. Rosati, N. Ben´ıtez, A. Biviano, R. Bouwens, L. Bradley, T. Broad- hurst, M. Carrasco, H. Ford, C. Grillo, L. Infante, D. Kelson, O. La- hav, R. Massey, J. Moustakas, E. Rasia, J. Rhodes, J. Vega, W. Zheng,

42 CLASH: The Concentration-Mass Relation of Galaxy Clusters, ApJ806 (2015) 4. doi:10.1088/0004-637X/806/1/4.

[59] K. Umetsu, M. Sereno, S.-I. Tam, I.-N. Chiu, Z. Fan, S. Ettori, D. Gruen, T. Okumura, E. Medezinski, M. Donahue, M. Meneghetti, B. Frye, A. Koekemoer, T. Broadhurst, A. Zitrin, I. Balestra, N. Ben´ıtez, Y. Higuchi, P. Melchior, A. Mercurio, J. Merten, A. Molino, M. Nonino, M. Postman, P. Rosati, J. Sayers, S. Seitz, The Projected Dark and Bary- onic Ellipsoidal Structure of 20 CLASH Galaxy Clusters, ApJ860 (2018) 104. arXiv:1804.00664, doi:10.3847/1538-4357/aac3d9.

[60] A. Zitrin, A. Fabris, J. Merten, P. Melchior, M. Meneghetti, A. Koekemoer, D. Coe, M. Maturi, M. Bartelmann, M. Postman, K. Umetsu, G. Seidel, I. Sendra, T. Broadhurst, I. Balestra, A. Bi- viano, C. Grillo, A. Mercurio, M. Nonino, P. Rosati, L. Bradley, M. Carrasco, M. Donahue, H. Ford, B. L. Frye, J. Moustakas, Hubble space telescope combined strong and weak lensing analysis of the clash sample: Mass and magnification models and systematic uncertainties, The Astrophysical Journal 801 (1) (2015) 44. URL http://stacks.iop.org/0004-637X/801/i=1/a=44

[61] M. Oguri, T. Hamana, Detailed cluster lensing profiles at large radii and the impact on cluster weak lensing studies, MNRAS414 (2011) 1851–1861. arXiv:1101.0650, doi:10.1111/j.1365-2966.2011.18481.x.

[62] K. Umetsu, B. Diemer, Lensing Constraints on the Mass Profile Shape and the Splashback Radius of Galaxy Clusters, ApJ836 (2017) 231. arXiv:1611.09366, doi:10.3847/1538-4357/aa5c90.

[63] I.-N. Chiu, K. Umetsu, M. Sereno, S. Ettori, M. Meneghetti, J. Merten, J. Sayers, A. Zitrin, CLUMP-3D: Three-dimensional Shape and Structure of 20 CLASH Galaxy Clusters from Combined Weak and Strong Lensing, ApJ860 (2018) 126. arXiv:1804.00676, doi:10.3847/1538-4357/aac4a0.

43 [64] M. Meneghetti, E. Rasia, J. Vega, J. Merten, M. Postman, G. Yepes, F. Sembolini, M. Donahue, S. Ettori, K. Umetsu, I. Balestra, M. Bartel- mann, N. Ben´ıtez, A. Biviano, R. Bouwens, L. Bradley, T. Broadhurst, D. Coe, N. Czakon, M. De Petris, H. Ford, C. Giocoli, S. Gottl¨ober, C. Grillo, L. Infante, S. Jouvel, D. Kelson, A. Koekemoer, O. Lahav, D. Lemze, E. Medezinski, P. Melchior, A. Mercurio, A. Molino, L. Moscar- dini, A. Monna, J. Moustakas, L. A. Moustakas, M. Nonino, J. Rhodes, P. Rosati, J. Sayers, S. Seitz, W. Zheng, A. Zitrin, The MUSIC of CLASH: Predictions on the Concentration-Mass Relation, ApJ797 (2014) 34. arXiv:1404.1384, doi:10.1088/0004-637X/797/1/34.

[65] S. Giodini, D. Pierini, A. Finoguenov, G. W. Pratt, H. Boehringer, A. Leauthaud, L. Guzzo, H. Aussel, M. Bolzonella, P. Capak, M. Elvis, G. Hasinger, O. Ilbert, J. S. Kartaltepe, A. M. Koekemoer, S. J. Lilly, R. Massey, H. J. McCracken, J. Rhodes, M. Salvato, D. B. Sanders, N. Z. Scoville, S. Sasaki, V. Smolcic, Y. Taniguchi, D. Thompson, COSMOS Collaboration, Stellar and Total Baryon Mass Fractions in Groups and Clusters Since Redshift 1, ApJ703 (2009) 982–993. arXiv:0904.0448, doi:10.1088/0004-637X/703/1/982.

[66] P. Catelan, T. Theuns, Evolution of the angular momentum of from tidal torques: Zel’dovich approximation, MNRAS282 (1996) 436–454. arXiv:arXiv:astro-ph/9604077. URL http://articles.adsabs.harvard.edu/cgi-bin/nph-iarticle_query?1996MNRAS.282..436C&data_type=PDF_HIGH&whole_paper=YES&type=PRINTER&filetype=.pdf

[67] P. Catelan, T. Theuns, Non-linear evolution of the angular momentum of protostructures from tidal torques, MNRAS282 (1996) 455–469. arXiv:arXiv:astro-ph/9604078. URL http://articles.adsabs.harvard.edu/cgi-bin/nph-iarticle_query?1996MNRAS.282..455C&data_type=PDF_HIGH&whole_paper=YES&type=PRINTER&filetype=.pdf

[68] S. Gottl¨ober, G. Yepes, Shape, Spin, and Baryon Fraction of Clusters in the MareNostrum Universe, ApJ664 (2007) 117–122. arXiv:astro-ph/0703164, doi:10.1086/517907.

[69] E. Komatsu, J. Dunkley, M. R. Nolta, et al., Five-Year Wilkin- son Microwave Anisotropy Probe Observations: Cosmological

44 Interpretation, ApJS180 (2009) 330–376. arXiv:0803.0547, doi:10.1088/0067-0049/180/2/330.

[70] V. Avila-Reese, C. Firmani, X. Hern´andez, On the Formation and Evo- lution of Disk Galaxies: Cosmological Initial Conditions and the Grav- itational Collapse, ApJ505 (1998) 37–49. arXiv:astro-ph/9710201, doi:10.1086/306136.

[71] Y. Ascasibar, G. Yepes, S. Gottl¨ober, V. M¨uller, On the physical origin of dark matter density profiles, MN- RAS352 (2004) 1109–1120. arXiv:arXiv:astro-ph/0312221, doi:10.1111/j.1365-2966.2004.08005.x.

[72] K. Subramanian, R. Cen, J. P. Ostriker, The Structure of Dark Mat- ter Halos in Hierarchical Clustering Theories, ApJ538 (2000) 528–542. arXiv:arXiv:astro-ph/9909279, doi:10.1086/309152.

[73] A. Nusser, Self-similar spherical collapse with non-radial mo- tions, MNRAS325 (2001) 1397–1401. arXiv:astro-ph/0008217, doi:10.1046/j.1365-8711.2001.04527.x.

[74] N. Hiotelis, Density profiles in a spherical infall model with non-radial motions, A& A382 (2002) 84–91. arXiv:arXiv:astro-ph/0111324, doi:10.1051/0004-6361:20011620.

[75] M. Le Delliou, R. N. Henriksen, Non-radial motion and the NFW profile, A& A408 (2003) 27–38. arXiv:arXiv:astro-ph/0307046, doi:10.1051/0004-6361:20030922.

[76] L. L. R. Williams, A. Babul, J. J. Dalcanton, Investigating the Ori- gins of Density Profiles, ApJ604 (2004) 18–39. arXiv:astro-ph/0312002, doi:10.1086/381722.

[77] Y. Ascasibar, P. Jean, C. Bœhm, J. Kn¨odlseder, Constraints on dark matter and the shape of the Milky Way dark halo from the 511-

45 keV line, MNRAS368 (2006) 1695–1705. arXiv:astro-ph/0507142, doi:10.1111/j.1365-2966.2006.10226.x.

[78] E. A. Baltz, P. Marshall, M. Oguri, Analytic models of plausible gravitational lens potentials, JCAP1 (2009) 15. arXiv:0705.0682, doi:10.1088/1475-7516/2009/01/015.

[79] A. Mahdavi, H. Hoekstra, A. Babul, J. Sievers, S. T. Myers, J. P. Henry, Joint analysis of cluster observations. i. mass profile of from combined x-ray, sunyaev-zel’dovich, and weak-lensing data, The Astrophysical Journal 664 (1) (2007) 162–180. doi:10.1086/517958. URL https://doi.org/10.1086%2F517958

[80] D. Martizzi, R. Teyssier, B. Moore, The formation of the brightest cluster galaxies in cosmological simulations: the case for active galac- tic nucleus feedback, MNRAS420 (2012) 2859–2873. arXiv:1106.5371, doi:10.1111/j.1365-2966.2011.19950.x.

[81] C. Burke, M. Hilton, C. Collins, Coevolution of brightest cluster galax- ies and intracluster light using CLASH, MNRAS449 (2015) 2353–2367. arXiv:1503.04321, doi:10.1093/mnras/stv450.

[82] M. Donahue, G. M. Voit, A. Mahdavi, K. Umetsu, S. Ettori, J. Merten, M. Postman, A. Hoffer, A. Baldi, D. Coe, N. Czakon, M. Bartel- mann, N. Benitez, R. Bouwens, L. Bradley, T. Broadhurst, H. Ford, F. Gastaldello, C. Grillo, L. Infante, S. Jouvel, A. Koekemoer, D. Kel- son, O. Lahav, D. Lemze, E. Medezinski, P. Melchior, M. Meneghetti, A. Molino, J. Moustakas, L. A. Moustakas, M. Nonino, P. Rosati, J. Sayers, S. Seitz, A. Van der Wel, W. Zheng, A. Zitrin, CLASH- X: A Comparison of Lensing and X-Ray Techniques for Measuring the Mass Profiles of Galaxy Clusters, ApJ794 (2014) 136. arXiv:1405.7876, doi:10.1088/0004-637X/794/2/136.

[83] Y.-T. Lin, J. J. Mohr, S. A. Stanford, K-Band Properties of Galaxy Clus- ters and Groups: Luminosity Function, Radial Distribution, and Halo

46 Occupation Number, ApJ610 (2004) 745–761. arXiv:astro-ph/0402308, doi:10.1086/421714.

[84] A. Del Popolo, F. Pace, M. Le Delliou, X. Lee, Energy transfer from baryons to dark matter as a unified solution to small-scale struc- ture issues of the CDM model, Phys. Rev. D98 (6) (2018) 063517. arXiv:1809.10609, doi:10.1103/PhysRevD.98.063517.

[85] A. Del Popolo, On the dark matter haloes inner structure and galaxy morphology, Ap&SS361 (2016) 222. arXiv:1607.07408, doi:10.1007/s10509-016-2810-4.

[86] A. V. Kravtsov, The Size-Virial Radius Relation of Galaxies, ApJL764 (2013) L31. arXiv:1212.2980, doi:10.1088/2041-8205/764/2/L31.

[87] Y.-S. Li, G. De Lucia, A. Helmi, On the nature of the Milky Way satellites, MNRAS401 (2010) 2036–2052. arXiv:0909.1291, doi:10.1111/j.1365-2966.2009.15803.x.

[88] G. R. Blumenthal, S. M. Faber, R. Flores, J. R. Primack, Contraction of dark matter galactic halos due to baryonic infall, ApJ301 (1986) 27–34. doi:10.1086/163867.

[89] O. Y. Gnedin, A. V. Kravtsov, A. A. Klypin, D. Nagai, Response of Dark Matter Halos to Condensation of Baryons: Cosmological Simula- tions and Improved Adiabatic Contraction Model, ApJ616 (2004) 16–26. arXiv:astro-ph/0406247, doi:10.1086/424914.

[90] A. Immeli, M. Samland, O. Gerhard, P. Westera, Gas physics, disk frag- mentation, and bulge formation in young galaxies, A& A413 (2004) 547– 561. arXiv:astro-ph/0312139, doi:10.1051/0004-6361:20034282.

[91] C. N. Lackner, J. P. Ostriker, Dissipational Versus Dissipationless Galaxy Formation and the Dark Matter Content of Galaxies, ApJ712 (2010) 88– 100. arXiv:1002.0585, doi:10.1088/0004-637X/712/1/88.

47 [92] I. Trujillo, N. M. F¨orster Schreiber, G. Rudnick, M. Barden, M. Franx, H.- W. Rix, J. A. R. Caldwell, D. H. McIntosh, S. Toft, B. H¨aussler, A. Zirm, P. G. van Dokkum, I. Labb´e, A. Moorwood, H. R¨ottgering, A. van der Wel, P. van der Werf, L. van Starkenburg, The Size Evolution of Galaxies since z˜3: Combining SDSS, GEMS, and FIRES, ApJ650 (2006) 18–41. arXiv:astro-ph/0504225, doi:10.1086/506464.

[93] T. Naab, P. H. Johansson, J. P. Ostriker, Minor Mergers and the Size Evolution of Elliptical Galaxies, ApJL699 (2009) L178–L182. arXiv:0903.1636, doi:10.1088/0004-637X/699/2/L178.

[94] C. F. P. Laporte, S. D. M. White, T. Naab, M. Ruszkowski, V. Springel, Shallow dark matter cusps in galaxy clusters, MNRAS424 (2012) 747–753. arXiv:1202.2357, doi:10.1111/j.1365-2966.2012.21262.x.

[95] A. El-Zant, I. Shlosman, Y. Hoffman, Dark Halos: The Flattening of the Density Cusp by Dynamical Friction, ApJ560 (2001) 636–643. arXiv:astro-ph/0103386, doi:10.1086/322516.

[96] L. Gao, A. Loeb, P. J. E. Peebles, S. D. M. White, A. Jenkins, Early Formation and Late Merging of the Giant Galaxies, ApJ614 (2004) 17– 25. arXiv:astro-ph/0312499, doi:10.1086/423444.

[97] R. H. Wechsler, J. S. Bullock, J. R. Primack, A. V. Kravtsov, A. Dekel, Concentrations of Dark Halos from Their Assembly Histories, ApJ568 (2002) 52–70. arXiv:astro-ph/0108151, doi:10.1086/338765.

[98]D. H. Zhao, Y. P. Jing, H. J. Mo, G. B¨orner, Mass and Redshift Dependence of Dark Halo Structure, ApJL597 (2003) L9–L12. arXiv:arXiv:astro-ph/0309375, doi:10.1086/379734. URL http://adsabs.harvard.edu/abs/2003ApJ...597L...9Z

[99] Y. Fujita, K. Umetsu, E. Rasia, M. Meneghetti, M. Donahue, E. Medezin- ski, N. Okabe, M. Postman, Discovery of a New Fundamental Plane

48 Dictating Evolution from Gravitational Lensing, ApJ857 (2018) 118. arXiv:1801.03943, doi:10.3847/1538-4357/aab8fd.

[100] Y. Fujita, K. Umetsu, S. Ettori, E. Rasia, N. Okabe, M. Meneghetti, New Interpretation of the mass-temperature relation and mass cali- bration of galaxy clusters based on the fundamental plane, ArXiv e- printsarXiv:1804.05070.

[101] L. Zappacosta, D. A. Buote, F. Gastaldello, P. J. Humphrey, J. Bullock, F. Brighenti, W. Mathews, The Absence of Adiabatic Contraction of the Radial Dark Matter Profile in the Galaxy Cluster A2589, ApJ650 (2006) 777–790. arXiv:astro-ph/0602613, doi:10.1086/505739.

[102] L. Gao, J. F. Navarro, C. S. Frenk, A. Jenkins, V. Springel, S. D. M. White, The Phoenix Project: the dark side of rich Galaxy clusters, MNRAS425 (2012) 2169–2186. arXiv:1201.1940, doi:10.1111/j.1365-2966.2012.21564.x.

[103] A. Pontzen, F. Governato, Cold dark matter heats up, Nature506 (2014) 171–178. arXiv:1402.1764, doi:10.1038/nature12953.

[104] P. Col´ın, V. Avila-Reese, O. Valenzuela, Substructure and Halo Den- sity Profiles in a Warm Dark Matter Cosmology, ApJ542 (2000) 622–630. arXiv:astro-ph/0004115, doi:10.1086/317057.

[105] J. Sommer-Larsen, A. Dolgov, Formation of Disk Galaxies: Warm Dark Matter and the Angular Momentum Problem, ApJ551 (2001) 608–623. arXiv:astro-ph/9912166, doi:10.1086/320211.

[106] P. J. E. Peebles, Fluid Dark Matter, ApJL534 (2000) L127–L129. arXiv:astro-ph/0002495, doi:10.1086/312677.

[107] M. Kaplinghat, L. Knox, M. S. Turner, Annihilating Cold Dark Mat- ter, Physical Review Letters 85 (2000) 3335. arXiv:astro-ph/0005210, doi:10.1103/PhysRevLett.85.3335.

49 [108] A. R. Zentner, J. S. Bullock, Halo Substructure and the Power Spectrum, ApJ598 (2003) 49–72. arXiv:astro-ph/0304292, doi:10.1086/378797.

[109] K. Umetsu, E. Medezinski, M. Nonino, J. Merten, A. Zitrin, A. Molino, C. Grillo, M. Carrasco, et al., CLASH: Mass Distribution in and around MACS J1206.2-0847 from a Full Cluster Lensing Analysis, ApJ755 (2012) 56. arXiv:1204.3630, doi:10.1088/0004-637X/755/1/56. URL http://adsabs.harvard.edu/abs/2012ApJ...755...56U

[110] A. Zitrin, T. Broadhurst, D. Coe, K. Umetsu, M. Postman, N. Ben´ıtez, M. Meneghetti, E. Medezinski, S. Jouvel, L. Bradley, A. Koekemoer, W. Zheng, H. Ford, J. Merten, D. Kelson, O. Lahav, D. Lemze, A. Molino, M. Nonino, M. Donahue, P. Rosati, A. Van der Wel, M. Bartelmann, R. Bouwens, O. Graur, G. Graves, O. Host, L. Infante, S. Jha, Y. Jimenez- Teja, R. Lazkoz, D. Maoz, C. McCully, P. Melchior, L. A. Moustakas, S. Ogaz, B. Patel, E. Regoes, A. Riess, S. Rodney, S. Seitz, The Clus- ter Lensing and Supernova Survey with Hubble (CLASH): Strong-lensing Analysis of A383 from 16-band HST/WFC3/ACS Imaging, ApJ742 (2011) 117. arXiv:1103.5618, doi:10.1088/0004-637X/742/2/117.

[111] I. M. Whiley, A. Arag´on-Salamanca, G. De Lucia, A. von der Linden, S. P. Bamford, P. Best, M. N. Bremer, P. Jablonka, O. Johnson, B. Milvang- Jensen, S. Noll, B. M. Poggianti, G. Rudnick, R. Saglia, S. White, D. Zarit- sky, The evolution of the brightest cluster galaxies since z ˜ 1 from the ESO Distant Cluster Survey (EDisCS), MNRAS387 (2008) 1253–1263. arXiv:0804.2152, doi:10.1111/j.1365-2966.2008.13324.x.

[112] A. Del Popolo, F. Pace, The Cusp/Core problem: supernovae feedback versus the baryonic clumps and dynamical friction model, Ap&SS361 (2016) 162. arXiv:1502.01947, doi:10.1007/s10509-016-2742-z.

[113] A. Del Popolo, On the dark matter haloes inner structure and

50 galaxy morphology, Ap&SS361 (2016) 222. arXiv:1607.07408, doi:10.1007/s10509-016-2810-4.

[114] J. E. Gunn, J. R. Gott, III, On the Infall of Matter Into Clusters of Galaxies and Some Effects on Their Evolution, ApJ176 (1972) 1. doi:10.1086/151605.

[115] Y. Hoffman, J. Shaham, Local density maxima - Progenitors of structure, ApJ297 (1985) 16–22. doi:10.1086/163498.

[116] M. Gustafsson, M. Fairbairn, J. Sommer-Larsen, Baryonic pinch- ing of galactic dark matter halos, PRD74 (12) (2006) 123522. arXiv:astro-ph/0608634, doi:10.1103/PhysRevD.74.123522.

[117] B. S. Ryden, Galaxy formation - The role of tidal torques and dissipational infall, ApJ329 (1988) 589–611. doi:10.1086/166406.

[118] E. L.Lokas, Universal profile of dark matter haloes and the spherical infall model, MNRAS311 (2000) 423–432. arXiv:astro-ph/9901185, doi:10.1046/j.1365-8711.2000.03082.x.

[119] F. Hoyle, On the Fragmentation of Gas Clouds Into Galaxies and Stars., ApJ118 (1953) 513. doi:10.1086/145780.

[120] P. J. E. Peebles, Origin of the Angular Momentum of Galaxies, ApJ155 (1969) 393. doi:10.1086/149876.

[121] S. D. M. White, Angular momentum growth in protogalaxies, ApJ286 (1984) 38–41. doi:10.1086/162573.

[122] D. J. Eisenstein, A. Loeb, An analytical model for the triax- ial collapse of cosmological perturbations, ApJ439 (1995) 520–541. arXiv:arXiv:astro-ph/9405012, doi:10.1086/175193.

[123] B. S. Ryden, J. E. Gunn, Galaxy formation by gravitational collapse, ApJ318 (1987) 15–31. doi:10.1086/165349.

51 [124] V. Antonuccio-Delogu, S. Colafrancesco, Dynamical friction, secondary infall, and the evolution of clusters of galaxies, ApJ427 (1994) 72–85. doi:10.1086/174122.

[125] H. J. Mo, S. Mao, S. D. M. White, The formation of galac- tic discs, MNRAS295 (1998) 319–336. arXiv:astro-ph/9707093, doi:10.1046/j.1365-8711.1998.01227.x.

[126] V. F. Cardone, M. Sereno, Modelling the Milky Way through adia- batic compression of cold dark matter haloes, A& A438 (2005) 545–556. arXiv:astro-ph/0501567, doi:10.1051/0004-6361:20042035.

[127] T. Treu, L. V. E. Koopmans, The Internal Structure and Forma- tion of Early-Type Galaxies: The Gravitational Lens System MG 2016+112 at z = 1.004, ApJ575 (2002) 87–94. arXiv:astro-ph/0202342, doi:10.1086/341216.

[128] C. R. Keeton, Cold Dark Matter and Strong Gravitational Lensing: Concord or Conflict?, ApJ561 (2001) 46–60. arXiv:astro-ph/0105200, doi:10.1086/323237.

[129] H.-W. Rix, P. T. de Zeeuw, N. Cretton, R. P. van der Marel, C. M. Carollo, Dynamical Modeling of Velocity Profiles: The Dark Halo around the NGC 2434, ApJ488 (1997) 702–719. arXiv:astro-ph/9702126.

[130] G. De Lucia, A. Helmi, The Galaxy and its stellar halo: insights on their formation from a hybrid cosmological approach, MNRAS391 (2008) 14– 31. arXiv:0804.2465, doi:10.1111/j.1365-2966.2008.13862.x.

[131] A. V. Kravtsov, O. Y. Gnedin, A. A. Klypin, The Tumultuous Lives of Galactic Dwarfs and the Missing Satellites Problem, ApJ609 (2004) 482– 497. arXiv:astro-ph/0401088, doi:10.1086/421322.

52 [132] W. Hu, A. V. Kravtsov, Sample Variance Considerations for Clus- ter Surveys, ApJ584 (2003) 702–715. arXiv:astro-ph/0203169, doi:10.1086/345846.

[133] S. D. M. White, C. S. Frenk, Galaxy formation through hierarchical clus- tering, ApJ379 (1991) 52–79. doi:10.1086/170483.

[134] D. J. Croton, V. Springel, S. D. M. White, G. De Lucia, C. S. Frenk, L. Gao, A. Jenkins, G. Kauffmann, J. F. Navarro, N. Yoshida, The many lives of active galactic nuclei: cooling flows, black holes and the luminosities and colours of galaxies, MNRAS365 (2006) 11–28. arXiv:astro-ph/0508046, doi:10.1111/j.1365-2966.2005.09675.x.

[135] G. Stinson, A. Seth, N. Katz, J. Wadsley, F. Governato, T. Quinn, Star formation and feedback in smoothed particle hydrodynamic simulations - I. Isolated galaxies, MNRAS373 (2006) 1074–1090. arXiv:astro-ph/0602350, doi:10.1111/j.1365-2966.2006.11097.x.

[136] A. Di Cintio, C. B. Brook, A. V. Macci`o, G. S. Stinson, A. Knebe, A. A. Dutton, J. Wadsley, The dependence of dark matter profiles on the stellar- to-halo mass ratio: a prediction for cusps versus cores, MNRAS437 (2014) 415–423. arXiv:1306.0898, doi:10.1093/mnras/stt1891.

[137] G. Chabrier, Galactic Stellar and Substellar Initial Mass Func- tion, PASP115 (2003) 763–795. arXiv:astro-ph/0304382, doi:10.1086/376392.

[138] G. De Lucia, G. Kauffmann, S. D. M. White, Chemical enrichment of the intracluster and intergalactic medium in a hierarchical galaxy forma- tion model, MNRAS349 (2004) 1101–1116. arXiv:astro-ph/0310268, doi:10.1111/j.1365-2966.2004.07584.x.

[139] A. Cattaneo, A. Dekel, J. Devriendt, B. Guiderdoni, J. Blaizot, Modelling the galaxy bimodality: shutdown above a critical halo

53 mass, MNRAS370 (2006) 1651–1665. arXiv:astro-ph/0601295, doi:10.1111/j.1365-2966.2006.10608.x.

[140] C. M. Booth, J. Schaye, Cosmological simulations of the growth of supermassive black holes and feedback from active galactic nu- clei: method and tests, MNRAS398 (2009) 53–74. arXiv:0904.2572, doi:10.1111/j.1365-2966.2009.15043.x.

[141] F. Governato, A. Zolotov, A. Pontzen, C. Christensen, S. H. Oh, A. M. Brooks, T. Quinn, S. Shen, J. Wadsley, Cuspy no more: how out- flows affect the central dark matter and baryon distribution in Λ cold dark matter galaxies, MNRAS422 (2012) 1231–1240. arXiv:1202.0554, doi:10.1111/j.1365-2966.2012.20696.x.

[142] A. A. Klypin, S. Trujillo-Gomez, J. Primack, Dark Matter Ha- los in the Standard Cosmological Model: Results from the Bolshoi Simulation, ApJ740 (2011) 102. arXiv:1002.3660, doi:10.1088/0004-637X/740/2/102.

54