A publishing partnership

Articles

OFFSETS BETWEEN THE X-RAY AND THE SUNYAEV–ZEL'DOVICH-EFFECT PEAKS IN MERGING GALAXY CLUSTERS AND THEIR COSMOLOGICAL IMPLICATIONS

, , and

Published 2014 November 14 © 2014. The American Astronomical Society. All rights reserved.
, , Citation Congyao Zhang et al 2014 ApJ 796 138 DOI 10.1088/0004-637X/796/2/138

0004-637X/796/2/138

ABSTRACT

Observations reveal that the peaks of the X-ray map and the Sunyaev–Zel'dovich (SZ) effect map of some galaxy clusters are offset from each other. In this paper, we perform a set of hydrodynamical simulations of mergers of two galaxy clusters to investigate the spatial offset between the maxima of the X-ray and the SZ surface brightness of the merging clusters. We find that significantly large SZ–X-ray offsets (>100 kpc) can be produced during the major mergers of galaxy clusters (with mass > 1 × 1014M). The significantly large offsets are mainly caused by a "jump effect" that occurrs between the primary and secondary pericentric passages of the two merging clusters, during which the X-ray peak may jump to the densest gas region located near the center of the small cluster, but the SZ peak remains near the center of the large one. Our simulations show that merging systems with higher masses and larger initial relative velocities may result in larger offset sizes and longer offset time durations; and only nearly head-on mergers are likely to produce significantly large offsets. We further investigate the statistical distribution of the SZ–X-ray offset sizes and find that (1) the number distribution of the offset sizes is bimodal with one peak located at low offsets ∼0 and the other at large offsets ∼350–450 h−1 kpc, but the objects with intermediate offsets are scarce; and (2) the probabilities of the clusters in the mass range higher than 2 × 1014h−1M that have offsets larger than 20, 50, 200, 300, and 500 h−1 kpc are 34.0%, 11.1%, 8.0%, 6.5%, and 2.0%, respectively, at z = 0.7. The probability is sensitive to the underlying pairwise velocity distribution and the merger rate of clusters. We suggest that the SZ–X-ray offsets provide a probe to the cosmic velocity fields on the cluster scale and the cluster merger rate, and future observations on the SZ–X-ray offsets for a large number of clusters may put strong constraints on them. Our simulation results suggest that the SZ–X-ray offset in the Bullet Cluster, together with the mass ratio of the two merging clusters, requires a relative velocity larger than 3000 km s−1 at an initial separation 5 Mpc. The cosmic velocity distribution at the high-velocity end is expected to be crucial in determining whether there exists an incompatibility between the existence of the Bullet Cluster and the prediction of a ΛCDM model.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Clusters of galaxies, the largest virialized systems known in the universe, are formed from mergers of small structures in the hierarchical structure formation and evolution model (see a review in Kravtsov & Borgani 2012). Collisions of galaxy clusters, with relative velocities up to several hundred or several thousand  km s−1, are probably the most energetic events since the big bang, which re-distribute both the dark matter (DM) and the baryonic matter in the colliding clusters at a megaparsec scale and gas can be shocked and heated. A number of observational features have indicated that some cluster systems are undergoing mergers or are the remnants of recent mergers. For example, the spatial separation of the X-ray emitting gas and the DM clump in the Bullet Cluster is explained by a collision of two clusters, in which gas interacts electromagnetically and moves slower than DM (Clowe et al. 2004, 2006); and the sharp edges shown in some cluster X-ray images are interpreted as either the "shock fronts" or "cold fronts" formed in the intracluster medium (ICM) during cluster collisions (Markevitch et al. 1999, 2000; Markevitch & Vikhlinin 2007). In addition, recent Sunyaev–Zel'dovich (SZ) cluster surveys, including the South Pole Telescope (SPT), the Atacama Cosmology Telescope (ACT), and the Planck SZ surveys, show that the positions of the maxima of the X-ray surface brightness and the SZ effect differ significantly in some clusters (Andersson et al. 2011; Planck Collaboration et al. 2011c; Menanteau et al. 2012; see also an early observation in Komatsu et al. 2001); and also collisions of clusters can lead to the position displacement (Planck Collaboration et al. 2011b). In this paper, we study how the spatial displacement of the X-ray and the SZ signals from a merging cluster is affected by the merging process and explore how the statistical distribution of the displacements is connected to the cluster pairwise velocity field and the cluster merger rate. Study of the cluster merging processes and the possible observational signatures would improve our understanding of the baryonic physics involved in cluster collisions, the cosmic velocity fields, and further the structure formation and evolution model and the underlying cosmological model.

Combination of the X-ray properties and the SZ effects of clusters has been used before to constrain the cosmological parameters (Carlstrom et al. 2002) and investigate baryonic physics in clusters of galaxies (e.g., Andersson et al. 2011; Planck Collaboration et al. 2011a, 2012). The X-ray emission of the ICM gas comes mainly from the thermal bremsstrahlung radiation of the hot electrons, and the X-ray luminosity is proportional to $n_{\rm e}^2T^{1/2}_{\rm e}$, where ne is the number density of electrons and Te is the electron temperature. The SZ effect is the result of energy boost of low-energy cosmic microwave background photons due to the inverse Compton scattering off high-energy electrons in the ICM. Depending on whether the high energy of the electrons is due to their temperature (thermal) or bulk motion (kinematic), the SZ effect can be divided into the thermal effect and the kinematic effect. The thermal SZ effect has been detected in many clusters, and its magnitude is proportional to the integral of the electron pressure along the line of sight (∝neTe) and independent of redshift. The kinematic effect is expected to be a potential probe to the motion of clusters, yet it is much weaker than the thermal effect in high-mass clusters. As the thermal SZ effect and the X-ray brightness of a cluster have a different dependence on gas temperature and density distribution, and the location of the maximum X-ray brightness is likely to be offset from the location of the maximum of the SZ effect because of the re-distribution of the gas in the cluster merging process.

In this paper, we use numerical simulations to study the merging processes of individual clusters and then obtain the distribution maps of the X-ray brightness and the SZ effect in the merging clusters. Many simulations on mergers of two individual clusters have been performed previously to produce the observed configuration/morphology of a specific unrelaxed galaxy cluster (e.g., Springel & Farrar 2007; Mastropietro & Burkert 2008; ZuHone et al. 2009) or investigate various physical effects and features caused by the merging processes (see, e.g., Roettiger et al. 1997; Ritchie & Thomas 2002; Poole et al. 2006; ZuHone 2011). Molnar et al. (2012) simulated mergers of two clusters to particularly reconstruct the morphology of the galaxy cluster CL 0152-1357 and the offset between its SZ effect and X-ray peaks; and in their study, only the high relative velocity (>3000 km s−1) case of two clusters was explored, which is extremely rare in the universe. In our study, we explore the processes in a larger range of the parameter space (e.g., in the initial kinematic distribution of colliding clusters, their total masses and mass ratios), so that the distribution of the offsets can be investigated statistically. We find that the distribution of the SZ–X-ray offset is mainly affected by the merger rate of clusters and the pairwise velocity distribution of those merging clusters. Therefore, observational estimates of the SZ–X-ray offset distribution by SZ and X-ray surveys can be used to put constraints on both the merger rate of clusters and the pairwise velocity distribution of clusters, and further on the ΛCDM model at the cluster scale.

This paper is organized as follows. In Section 2, we describe our numerical methods of the cluster mergers and the initial conditions of the simulations. In Section 3, we present the general results of the simulations and discuss the factors to affect the offsets between the SZ effect and the X-ray peaks. In Section 4, we estimate the probability of the large offsets and compare it with observations. Finally, conclusions are summarized in Section 5.

Throughout the paper, we use a flat ΛCDM cosmology model with Ωm = 0.30, Ωb = 0.05, ΩΛ = 0.70, H = 100 h km s−1 Mpc−1 with h = 0.7.

2. METHOD

We perform numerical simulations of the mergers of two galaxy clusters. In this work, each of the clusters is simplified as a halo with a mixture of only DM and gas. The DM is approximated as collisionless, undergoing only gravity; and the gaseous component is collisional and adiabatic, undergoing both gravity and fluid pressure. Both the particle-based Lagrangian (such as smoothed particle hydrodynamics, SPH) method (Ritchie & Thomas 2002; Poole et al. 2006) and the mesh-based Eulerian such as adaptive mesh refinement, AMR) method (ZuHone 2011; Molnar et al. 2012) have been used to carry out such simulations. Mitchell et al. (2009) investigated the discrepancy occurred in the merger simulations between these two different methods by using GADGET-2 (SPH; Springel et al. 2001; Springel 2005) and FLASH (AMR) codes. They showed that SPH has the advantages in computational speed, effective resolution, and Galilean invariance, but is not good at the treatment of eddies and fluid instabilities, while AMR is on the contrary. Considering the purpose of our work, we choose the SPH code to do the simulations for the following reasons. (1) We need to explore a large parameter space of the merging processes, which demands an efficient calculation speed; (2) we focus on the peak positions of the X-ray and the SZ signals, and the disadvantages of the SPH code do not have significant effects on these; and (3) by applying both the AMR code and the SPH code to simulate the mergers for some cases in the Appendix, and then comparing both the simulation results concerned in this study, we find that different numerical codes do not lead to significant changes.

We have made some simplifications in simulating the physical processes occurred in the clusters. (1) We do not include radiative cooling and various heating mechanisms in our simulations. A pure cooling model has to face the overcooling problem, which is inconsistent with observations (Suginohara & Ostriker 1998); and thus an effective heating form is necessitated to counterbalance the cooling effect, such as AGN feedback (Sijacki et al. 2007). However, the physics of those mechanisms is not well understood yet. As argued in ZuHone (2011), the simplified model for gas physics can serve as a baseline to characterize the effect that we are interested in this work. In addition, Mastropietro & Burkert (2008) showed little evidence that the radiative cooling can change the positions of X-ray peaks, though the surface brightness is obviously modified. (2) The magnetic field is omitted in our simulations, and we assume that it has little influence on the SZ–X-ray offsets. In our work SZ–X-ray offset is formed from the breaking of hydrostatic equilibrium of the ICM gas during cluster energetic collisions. The magnetic energy is usually smaller than 1% of the mechanical energy involved in cluster mergers (Carilli & Taylor 2002), and on average the magnetic pressure in clusters is much smaller than the thermal pressure (Laganà et al. 2010). Moreover, the tangling scale of the magnetic field is about 10 kpc (Carilli & Taylor 2002), which is an order of magnitude smaller than the typical offsets discussed in this work. Nevertheless, the quantitative analysis on the effects after including the ignored physical processes here needs to be investigated in the future.

2.1. Initial Distributions of the Dark Matter and the Gaseous Halo of a Cluster

We assume a spherical symmetric density profile to model the initial mass distributions of the DM and the gas in a cluster. The Navarro–Frenk–White (NFW) profile (Navarro et al. 1997) is used for the DM mass density distribution within the virial radius rvir,

Equation (1)

where ρs and rsrvir/cvir are the scale density and the scale radius, respectively, and cvir is the concentration parameter. The rvir is calculated from the spherical collapse model (Bryan & Norman 1998), and cvir is given by a fitting formula obtained from N-body simulations (Duffy et al. 2008; see also Bullock et al. 2001; Wechsler et al. 2002; Zhao et al. 2003; Macciò et al. 2007). The effect of the evolution of rvir and cvir with redshift will be discussed in Section 4.2. The DM density distribution outside the virial radius is given by

Equation (2)

where we follow Kazantzidis et al. (2004) and implement an exponential cutoff that suppresses the profile on a truncation scale rdecay to avoid a divergent total mass and the parameter δ is set by keeping the first derivative of the DM density profiles continuous at r = rvir. The parameter ρs is set so that the total mass obtained by integrating Equations (1) and (2) over the space is the total DM mass of the cluster. Given the mass density distribution, the distribution function of the DM in the six-dimensional space f(r, v) is assumed to be ergodic and solved via the Eddington's formula (Equation (4.46) in Binney & Tremaine 2008). In both the SPH and the AMR simulations, DM is described by Lagrangian particles. To keep the stability of the DM distribution within the virial radius over cosmological relevant timescales, the truncation scale is set to 0.3rvir, as done in Zemp et al. (2008), and we have tested that a single cluster with the truncated model is stable within the Hubble time in our simulations.

We choose the Burkert profile (Burkert 1995) to represent the initial gas density profile as follows,

Equation (3)

as done in ZuHone et al. (2009), where rc = 0.5rs is the core radius and the normalization density ρc is set so that the baryonic mass fraction within rvir is consistent with the cosmological average value Ωbm = 0.17. For the region outside the virial radius, we assume that the gas density profile traces the DM density profile as follows,

Equation (4)

The specific internal energy of the gas at radius r is deter-mined by

Equation (5)

where the gas is assumed to be in hydrostatic equilibrium and ideal monatomic gas state with mean molecular weight per ion μ = 0.592, γ = 5/3 is the ratio of the heat capacity at constant pressure to that at constant volume, and M(r) is the total mass within radius r. The temperature distribution of the gas can be obtained from its internal energy distribution.

We also test different gas density distribution models in our work, e.g., the β-model (Cavaliere & Fusco-Femiano 1978) or a gas distribution tracing the DM density profile at all radii. We find that the different gas models do not affect our results significantly, as we focus on the positions of the maxima of projected X-ray and SZ maps of merging systems.

2.2. Simulation Settings

In our simulations, the centers of the two clusters with masses M1 and M2 (M1M2) are initially separated by a distance dini set to twice the sum of their virial radii, with initial relative velocity V and impact parameter P. The center of mass of the two clusters is initially put at rest at the origin of the coordinate system in the simulations. The initial position of each cluster center is put at the xy-plane with coordinate z = 0, and their (x, y, z) coordinates are (M2P/(M1 + M2), M2dini/(M1 + M2), 0) and (− M1P/(M1 + M2), − M1dini/(M1 + M2), 0), respectively. The initial relative velocity is along the y-axis, which are (0, − M2V/(M1 + M2), 0) and (0, M1V/(M1 + M2), 0) for the two clusters, respectively.

We explore a large range of the parameter space of the initial conditions. We perform a series of simulations with M1 = 1 × 1014, 2 × 1014, 5 × 1014, 1 × 1015M. The mass ratio ξ(≡ M1/M2) is generally set to be 1, 2, 3, or 5. For some special cases, the values of the mass ratio are set more intensively from 1 to 5. Different from previous work in which a fixed initial relative velocity (1.1 times circular velocity at the virial velocity of the larger cluster) is adopted (McCarthy et al. 2007; ZuHone 2011), our simulations are done for a large range of V from 250 to 4000 km s−1 and show that the SZ–X-ray offset has a strong dependence on the relative velocities (see Section 3.1.1). The impact parameter P spans from 0 to 600 kpc, as head-on or nearly head-on mergers are relevant here (see Section 3.1.2).

We use GADGET-2, an efficient parallel TreeSPH code, to carry out our simulations (Springel et al. 2001). The mass of each gas particle is mgas = 1.25 × 108M, and the mass of each dark matter particle is mDM = 6.25 × 108M. The gravitational softening length is set to 4 kpc, for which other choices (e.g., 1.5 or 15 kpc) are also tested and our main results are not affected.

As mentioned above, we also use FLASH, a mesh-based Eulerian code, to test whether the discrepancy between Eulerian and Lagrangian methods affects our results significantly (Fryxell et al. 2000). FLASH uses adaptive mesh refinement (AMR), and solves the equations of hydrodynamics by piecewise-parabolic method (PPM) of Colella & Woodward (1984). The gravitational potential is computed by the multigrid solver (Ricker 2008). We simulate the cases with ξ = 2, V = 500 km s−1, P = 0 kpc for M1 = 2 × 1014, 5 × 1014, 1 × 1015M. For the box size of 13.0 Mpc of all the FLASH mergers, the finest resolution achieved in our simulations is 12.7 kpc.

2.3. Projection Analysis: The X-ray and the SZ Signals of Merging Clusters

Given any time of the merging processes, we can obtain the two-dimensional maps of the mass surface density, the X-ray surface brightness, and the SZ emission of the simulated merging clusters by the following equations.

  • 1.  
    The mass surface density Σ is given by an integral of the mass density along the line of sight (LOS), i.e.,
    Equation (6)
  • 2.  
    The X-ray surface brightness is obtained by using an approximate expression for the relativistic X-ray thermal bremsstrahlung as follows (see Equation (5.25) in Rybicki & Lightman 1979), considering that the gas temperatures Tgas of some regions heated by shocks can be up to >30 keV,
    Equation (7)
    where $g_{\rm B}=(2\sqrt{3}/\pi)[1+0.79(4.95\times 10^{5}{\rm \,K}/T_{\rm gas})^{1/2}]$ is the frequency average of the velocity averaged Gaunt factor (Rephaeli & Yankovitch 1997).
  • 3.  
    The SZ surface brightness is given by
    Equation (8)
    where σT is the Thomson cross section, me is the electron mass, c is the speed of light, Θ ≡ kBTgas/(mec2), kB is the Boltzmann constant, and Y0, Y1, Y2, Y3, Y4 are the factors given by Equations (2.26)–(2.30) in Itoh et al. (1998).

In practice, we obtain all the above quantities at a given position calculated from the SPH method by smoothing N neighbor particles output from each snapshot of the simulations, where N = 50 ± 2 is a neighbor particle number typically used for smoothing.

The values and the spatial positions at the peaks of the projected mass surface density, X-ray and SZ images can then be obtained. Note that all the maps have been smoothed by a Gaussian distribution function with dispersion σ = 70 kpc. We also try other values of σ = 30, 40 kpc, and find that the variations of the positions of the peaks are smaller than 10%. The SZ peak position is not sensitive to the σ value even when σ ⩾ 140 kpc. A similar result is reported in Molnar et al. (2012) although they use a lower SZ resolution. Thus, not only are the obtained peak positions not affected by the smoothing, but also they can be straightforwardly compared with the current SZ observations, though the current observational resolutions are at the sub-arcminute level.

Finally, we clarify the definition of the "SZ–X-ray offset" to be obtained from our simulations, which is not exactly the same as that obtained from observations in the following two points. (1) The offset that we obtain from our simulations is the distance between the positions of the maxima of the whole X-ray and SZ surface brightness maps, which covers a sufficiently large region. However, the offsets obtained from observations are sometimes the displacements between the X-ray and SZ peaks located in a small local region, hence they can be smaller than the offsets obtained in this work. (2) The centroids of the galaxy clusters obtained in the X-ray survey are always obtained by computing the mean emission-weighted positions after removing the extended secondary X-ray maxima, instead of finding the maximum in the surface brightness maps as done here (e.g., Vikhlinin et al. 2009; Andersson et al. 2011). The above points may cause the underestimate of the probability of large SZ–X-ray offsets in the observations to be discussed in Section 4.

3. SIMULATION RESULTS

In this section, we present the following aspects of our simulation results. (1) We show the projected X-ray and SZ surface brightness maps obtained from Equations (7) and (8), viewed along the z-axis unless otherwise stated. We qualitatively describe the formation and evolution of the SZ–X-ray offsets during the merging process and discuss the impacts of different relative velocities, impact parameters, primary masses, and mass ratios on the maximum and time duration of the offset. (2) We quantitatively measure the time duration of the SZ–X-ray offset, which is used for estimating the probability of the offsets expected in observations in Section 4.

3.1. Mapping the X-Ray and the SZ Emissions from Merging Clusters

Generally, the merging process of two clusters involves the five distinct stages (Poole et al. 2006): pre-interaction stage, primary core–core interaction, apocentric passage, secondary core accretion, and relaxation (see an example illustrated in Figure 2), if the initial relative velocity is not high enough to detach the two clusters from each other at a later time. As mentioned above, the DM and the gas have different physical behavior during the mergers. The DM is collisionless so that the DM halos of the two clusters can go through each other, undergoing only gravity, while gas experiences gas pressure and shocks can be created during collisions of the gas halos. The different behavior can be revealed from the images of the mass surface density, and the X-ray and the SZ emissions of the merging clusters, as DM dominates the entire mass of the merging system and the X-ray and the SZ emissions are merely sensitive to the baryon distribution.

3.1.1. Dependence on Initial Relative Velocities

Figure 1 shows a time sequence of the snapshots of the mass surface density (red contours), the X-ray surface brightness (blue contours), and the SZ emission (green contours) of a head-on merger of two clusters (M1 = 2 × 1014M, ξ = 2) with different initial relative velocities. For simplicity, we show the results of the first three stages (the entire five evolution stages can be viewed in Figure 2). The first column shows a snapshot in the pre-interaction stage for each merging system. The second column represents the time of the first pericentric passage of the two clusters. For simplicity, we set the evolution time t = 0.0 Gyr at the first pericentric passage. As seen from the second column, the blue and the green contours are flattened, i.e., gas is squeezed outward in the direction normal to the collision axis; and the green contours are relatively more flattened, as the outer region of the merging cluster heated by shocks contributes more to the SZ surface brightness than to the X-ray surface brightness. The SZ and the X-ray peaks (labeled by "×" and "+" in the figure) start to separate after the first pericentric passage, when the gas at the inner region of the merging clusters is strongly disturbed. Since the central mass density of the large cluster is lower than that of the small cluster, the larger cluster core is penetrated by the small one, and the offset is gradually stretched (see the third column). The fourth column presents the snapshots with the largest peak offsets occurred during the merging processes. The offset returns to a small value or disappears at a later time as the surrounding gas gradually falls into the gravitational potential of the larger cluster (see the fifth column). After the secondary core–core interaction (though not shown in Figure 1; see the primary core–core interaction in the second column), gas is usually partially relaxed. The SZ–X-ray offsets are then not larger than 100 kpc except for some special massive systems (e.g., bottom panel in Figure 5 below).

Figure 1.

Figure 1. Snapshots of the map of the X-ray and the SZ emissions from two merging clusters at different merging time t. The two clusters have M1 = 2 × 1014M, ξ = 2, P = 0 kpc, with relative velocity V = 500 km s−1 (top row), 1000 km s−1 (middle row), and 2000 km s−1 (bottom row), respectively. The red, green, and blue contours represent the logarithm of the projected mass surface density, SZ, and X-ray surface brightness, respectively (see Equations (6)–(8)); and the ratio in the surface brightness between successive contour curves is 3.16. The "+" and the "×" symbols represent the positions of the maxima of the X-ray and SZ maps, respectively. The second column shows the snapshots at the time of the primary pericentric passage, where the time is set to t = 0.0 Gyr. The fourth column represents the snapshots with the largest SZ–X-ray offset during the merger processes.

Standard image High-resolution image
Figure 2.

Figure 2. Evolution of the SZ–X-ray offsets (solid lines) with the same initial relative velocities as those in Figure 1. The dashed lines are the distances between the mass density centers of the two clusters as a function of time. The characteristic time points of different merging stages are marked in the top panel (M1 = 2 × 1014M, ξ = 2, P = 0 kpc): t1st (primary pericentric passage), tapo (first apocentric passage), t2nd (secondary pericentric passage), and trelax (relaxation). The duration of the offset larger than 200 h−1 kpc is marked by Γ(d > dc) in the middle panel. The symbol "D" represents the time duration between the primary and secondary pericentric passages. This figure shows that with higher initial relative velocity, the maximum of the SZ–X-ray offset appearing after the primary pericentric passages is larger.

Standard image High-resolution image

By comparing the first two rows in Figure 1, one can see that there is no obvious qualitative difference between the evolutionary behavior of the offset for the case with relative velocity V = 500 km s−1 and that with V = 1000 km s−1, except that the largest offset produced during the merging process is larger for the case with a higher relative velocity. For the case with V = 2000 km s−1 (the third row of Figure 1), however, the result is distinctly different as the small cluster escapes away and cannot get back to collide with the large cluster again after its first pericentric passage because of the high relative velocity. The center of the small cluster becomes the maximum of the X-ray brightness map, when it passes through the larger one. The offset can be up to 3 Mpc after their interaction.

Evolution of the SZ–X-ray offsets for different relative velocities is shown in Figure 2. The solid lines give the SZ–X-ray offset as a function of time. The dashed lines represent the evolution of the distance between the mass density centers of the two clusters; and a zero distance is used if the two objects cannot be identified as they overlap or merge together. The merging processes are divided into the five different stages as illustrated in the top panel. Initially, the two clusters are in the pre-interaction stage, and then the primary core–core interaction starts around the time of the first pericentric passage at t1st(= 0). Before t1st, it is the pre-interaction stage. The time of the first apocentric passage is marked as tapo. After tapo, the second core accretion stage starts; and t2nd marks the secondary pericentric passage and the end of the second core accretion stage. As the small cluster in the bound orbit is dragged back and forth during the gradual relaxation process, the apocentric distances damp with time. The whole system appears visually relaxed around the time trelax as marked in the top panel.

As seen from Figure 2, the offset is initially zero. A small jump appears before the first pericentric passage, and it then decreases and disappears as the two clusters get closer. The appearance of the small jump in the pre-interaction stage can be understood as follows: as the two clusters approach each other, their outer layers start to be compressed and heated, during which the position of the maximum SZ signal is affected, but the peak of the X-ray remains located at the center of the larger cluster. After the first pericentric passage, the collision of the two clusters destroys the hydrostatic equilibrium and spherical symmetry of the gas halos. As shown in Figure 2, a significantly large SZ–X-ray offset always occurs between the primary and secondary pericentric passages. The occurrence of the large offset can be understood as follows: the gas core of the larger cluster is disrupted, and the X-ray peak jumps to the densest region around the center of the small cluster. However, the SZ peak does not jump (see the peak positions shown in the fourth column of Figure 1), as the temperature at the center of the small cluster is relatively low (see the top right panel in Figure 6 below), which strongly reduces the SZ surface brightness whose emissivity is proportional to the temperature. The SZ–X-ray offset is therefore nearly boosted to the mass density displacement between the two clusters (hereafter we refer to this as the "jump effect"). As seen from the figure, the maximum offset is positively correlated with the initial relative velocity of the two clusters. The X-ray peaks drop back to the center of the larger cluster mostly after tapo, when the gas falls back to the trough of the gravitational potential well. The end of the large SZ–X-ray offset is indicated by the sharp discontinuity in the solid line. Our results suggest that the jump effect should be the dominant reason to lead to the large offset. Our calculations show that adopting different image smoothing scales σ may result in different time durations of the jump effect if σ > 70 kpc. The duration of the X-ray peaks locating in the center of the small cluster is shorter if σ is larger, as more substructures in the X-ray map is smoothed. For example, for the offset larger than 300 kpc, the duration obtained with σ = 140 kpc is only half of the value obtained with σ = 70 kpc.

As mentioned above, the evolution of the offset shows different behavior for low (e.g., V < 1000 km s−1) and high (e.g., V > 2000 km s−1) relative velocity mergers. Here we define a critical velocity Vcrit to distinguish these bound and unbound collision cases, i.e., the two clusters will merge and relax within 13 ± 0.5 Gyr if V < Vcrit. The merger is referred to as the "merger mode" if V < Vcrit, otherwise as the "flyby mode." We perform a series of low-resolution simulations (with fixed DM and gas particle numbers, 40,000 and 10,000, respectively, for computing efficiency) to identify Vcrit as a function of (M1, P, ξ). The result is shown in Figure 3, and the uncertainty in the obtained Vcrit is ±50 km s−1. We approximate the relation between the Vcrit and the initial condition parameters by the following fitting form:

Equation (9)

where the best-fit parameters are AV = 0.93, αV = 0.30, βV = 0.46, and γV = 0.069. The critical velocity is strongly related with the primary cluster mass and the mass ratio, but not the impact parameter. The fitting form and the best-fit power-law factors are roughly consistent with the expectation from the escape velocity criterion in a simple two-body gravitational interacting system, that is, $V_{\rm crit}^2/2-GM_1(1+\xi ^{-1})/d_{\rm ini}=0$, and thus $V_{\rm crit}\propto M_1^{1/3}(1+\xi ^{-1})^{1/2}(1+f_VP^2/M_1^{2/3})^{-1/4}$, where fV is a factor.

Figure 3.

Figure 3. Critical velocity Vcrit as a function of mass M1 for simulations with different mass ratio ξ (top panel) and impact parameter P (bottom panel). The points are the simulation results, and the dotted lines show the fitting results of Equation (9) for different parameters. See Section 3.1.1.

Standard image High-resolution image

3.1.2. Dependence on Impact Parameters

Figure 4 shows the evolution of the SZ–X-ray offsets for different impact parameters, with V = 500 km s−1. For the low initial relative velocities, the large offsets (>100 kpc) occur in both head-on and nearly head-on impacts (i.e., P < 400 kpc). They can reach up to 600–700 kpc. A larger impact parameter results in a smaller offset, which implies that the size of the SZ–X-ray displacement is strongly related with the intensity of the primary core–core interaction. Molnar et al. (2012) studied the high-initial relative velocity case with V = 4800 km s−1 and find that significant displacements (∼300 kpc) between the SZ and X-ray peaks can be produced for nonzero impact parameters (about 100–250 kpc) and they decrease with increasing impact parameters. We also perform simulations with V = 4000 km s−1 similar to those done in Molnar et al. (2012). For the nonzero impact parameter cases, we find that the patterns of the mass surface density, the X-ray, and the SZ emission are all very similar with those shown in Molnar et al. (2012); for the head-on merger, a displacement up to 150 kpc can also be produced, while the distance between the mass centers of the two clusters is 1.5 Mpc.

Figure 4.

Figure 4. Same as Figure 2, but for different impact parameters with fixed M1 = 2 × 1014M, ξ = 2 and V = 500 km s−1. This figure shows that the mergers with small impact parameters (top and middle panels) can produce a large offset (e.g., >100 kpc), while the mergers with large impact parameters (bottom panels) not.

Standard image High-resolution image

As seen from the bottom panel of Figure 4, there is no significant SZ–X-ray offset for a large impact parameter P = 400 kpc. This can be understood through the density and temperature maps of the colliding clusters shown in Figure 6. Figure 6 compares the density and temperature slices of the gas at coordinate z = 0 plane between the head-on (P = 0 kpc) and the off-axis (P = 400 kpc) mergers. For the head-on merger shown in the top panels, as discussed above, the core of the large cluster is disrupted at the collision, and the denser region is near the center of the small cluster; however, the temperature at the center of the small cluster and its surrounding region is relatively low. However, for the off-axis merger with a large impact parameter shown in the bottom panels, the gas cores of the two clusters sideswipe each other at the primary collision, and the center of the large cluster remains dense; thus the X-ray and SZ peaks are both near the center of the larger cluster. It is worthy to note that though shocks can heat the gas at shock fronts to a relatively high temperature (e.g., a few tens of keV), the X-ray and the SZ peaks still locate near the centers of the clusters, as the observed X-ray and the SZ emission is integrated along the line of sight (see Equations (7) and (8)).

Our simulation results demonstrate that only the head-on or nearly head-on mergers are possible to produce offsets larger than 100 kpc, e.g., P < 400 kpc for the simulation with M1 = 2 × 1014M or P < 600 kpc for M1 = 5 × 1014M(ξ = 2, V = 500 km s−1). In addition, a smaller impact parameter induces a longer time duration of the nonzero offset.

3.1.3. Dependence on Masses

We also investigate the simulations for different masses of the primary cluster. Figure 5 presents the dependence of the SZ–X-ray offset evolution on M1(= 1014, 5 × 1014, 1015M). As seen from the figure, a significantly large SZ–X-ray offset (≳ 400 kpc) can occur in the whole mass range of the galaxy clusters. For the more massive systems, the maximum of the spatial separation between the SZ and the X-ray peaks is larger, and the time duration of a nonzero offset is longer. Our simulation results show that the maxima of the SZ–X-ray offsets (denoted by dmax; see Equation (10) below) for different masses are approximately as large as the first apocentric distances and it also does for different velocities (V < Vcrit). The fitting form to the dependence of dmax on the initial parameters can be obtained first through the following analysis and then from our numerical simulation results. If approximating dynamics of the merging system by the dynamics of a two-body system, dmax can be approximately obtained through the energy conservation law V2/2 − GM1(1 + ξ−1)/dini ≃ −GM1(1 + ξ−1)/dmax, where G is the gravitational constant. By setting ξ = 2 and considering the scaling relation between the virial radius and the mass of galaxy clusters, we have $d_{\rm ini}=f_d M_1^{1/3}$ and $d_{\rm max}\simeq d_{\rm ini}(1-V^2d_{\rm ini}/3GM_1)^{-1} \propto M_1^{1/3}+f_d V^2/3GM_1^{1/3}$, where fd is a factor connecting dini and $M_1^{1/3}$. Thus we fit our numerical simulation results of dmax in the following form,

Equation (10)

and obtain the best-fit parameters Ad = 366 kpc, Bd = 653 kpc, αd = 0.42. Equation (10) is used in the integration of Equation (15) to obtain the offset rate. Here we do not consider the dependence of dmax on impact parameters because (1) dmax is not sensitive to the impact parameter (e.g., P < 400 kpc in Figure 4) unless P is too large to suppress the large offset, and (2) the cases with large impact parameters do not contribute much to the integration of Equation (15) below, as their time durations of the nonzero offsets are shorter (or the 〈Sp〉 term in Equation (15) is zero when P is too large).

Figure 5.

Figure 5. Same as Figure 2, but for different masses of the primary cluster with fixed ξ = 2, V = 500 km s−1 and P = 0 kpc. The SZ–X-ray offset can be significantly large and appear in all the cluster mass range (>1 × 1014M).

Standard image High-resolution image
Figure 6.

Figure 6. Snapshots of the mass density and the temperature slices of the gas at the coordinate z = 0 plane at t = 0.75 Gyr. Top panels are for a head-on merger with P = 0 kpc and bottom panels for an off-axis merger with P = 400 kpc (M1 = 2 × 1014M, ξ = 2, V = 500 km s−1). The red curves are the equi-mass surface density contours of the merging systems. The "+" and the "×" symbols represent the positions of the maxima of the X-ray and the SZ maps, respectively. The X-ray peak locates near the center of the small cluster in the head-on merger, but not in the off-axis one; while the SZ peaks are close to the center of the large cluster in both cases. The figure illustrates the reason of the "jump effect" of the SZ–X-ray peak offset for the head-on merger (see Sections 3.1.1 and 3.1.2).

Standard image High-resolution image

As seen from the middle and the bottom panels of Figure 5, for the mergers of high-mass clusters (e.g., M1 = 5 × 1014, 1 × 1015M), a significantly large offset can still appear after the third pericentric passage (where the SZ peak deviates from the center of the large cluster due to an offset between the positions of the maxima of the gas density and the temperature distributions). However, when doing the statistical analysis in Section 4 below, we consider the offsets (>50 h−1 kpc) triggered merely between the primary and secondary pericentric passages for the following reasons. (1) Massive mergers with cluster masses larger than 5 × 1014M are rare events in the universe, which approximately occupy 5% among all major mergers of galaxy clusters; and (2) the time duration of the offset after the third pericentric passage is nearly five times smaller than that of the offset triggered by the first core–core collision.

3.1.4. Dependence on Mass Ratios

The mass ratio ξ is also an important parameter to affect the values of the offsets. According to our simulations, if the initial relative velocity is below 2000 km s−1 for M1 = 2 × 1014M, we find that only when 1 < ξ < 3 (i.e., major mergers) can the offsets be larger than 100 kpc. The evolution of the SZ–X-ray offsets for different mass ratios with V = 500 km s−1 is shown in Figure 7. As seen from the figure, if the mass ratio is larger, the strength of the collision is weaker and has less power to disturb the large cluster, and thus the offset becomes less significant. For producing the offset larger than 100 kpc with higher ξ, more massive merging systems (e.g., M1 > 5 × 1014M) or higher relative velocities are required. Note that when the mass ratio approaches unity, the offset also turns to be insignificant, as the merging configuration is symmetric.

Figure 7.

Figure 7. Same as Figure 2, but for different mass ratios with fixed M1 = 2 × 1014M, V = 500 km s−1 and P = 0 kpc. The SZ–X-ray offset can be larger than 100 kpc only when 1 < ξ < 3.

Standard image High-resolution image

The dependence of the SZ–X-ray offset on the mass ratio provides a complementary method to constrain the kinematics of an individual merging cluster (see Section 5 below). For example, the Bullet Cluster is a system with two merging clusters with quite different masses; and by doing the simulations with M1 = 5 × 1014M, ξ = 8, P = 0, we find that only when V > 3000 km s−1 can the maximum of the offset be larger than 150 kpc.

3.2. Duration of the SZ–X-ray Offset and the Offset Ratio

The time duration of the SZ–X-ray offset is important in this work, as it is directly associated with the probability of the offset appearing in the observation. As studied in Section 3.1, the duration has a strong relation with the initial parameters V, P, M1, and ξ. To quantitatively describe the duration, we define the "offset ratio" R(d > dc) by

Equation (11)

where Dt2ndt1st is the time duration between the primary and the secondary pericentric passages. Γ(d > dc) is the time duration of the SZ–X-ray offset larger than dc (e.g., see Γ and D marked in the middle panel of Figure 2 where dc = 200 h−1 kpc)3.

The time duration D obtained from our simulations can be fit as a function of the pairwise velocity, cluster mass, and redshift by the following form,

Equation (12)

where AD = 1.53 Gyr, BD = 0.62, αD = −0.77. The relaxation timescale of the merging clusters in our simulations (i.e., trelaxt1st) is about several Gyr, which is typically two to three times longer than D (e.g., see Figures 24, and 5). The time duration of the mergers listed in Poole et al. (2006) also gives a similar result. In this work, we assume trelaxt1st = κD with the factor κ = 2. We will present our detailed studies of the relaxation timescale in a separate paper.

In Figure 8, we present the offset ratio R(d > dc) measured from mergers with different initial conditions. We select dc = 50, 100, 150, 200, 300 h−1 kpc, which fall into the typical offset range in observations. We only consider the "merger mode" whose D is shorter than 6 Gyr in our estimation for its dominant contribution to the large offset in the observation (see Section 4.1.3). Except for the massive mergers (see Section 3.1.3), the offsets larger than 50 h−1 kpc appear only between the primary and secondary collisions, consequently we have 0 ⩽ R(d > dc) < 1.

Figure 8.

Figure 8. Offset ratios R(d > dc) as a function of the initial relative velocity for different initial conditions (see the points with different colors and the connecting solid lines). The dotted lines represent the corresponding best-fit results from Equation (13).

Standard image High-resolution image

The top left panel of Figure 8 shows the offset ratio as a function of the relative velocity with different M1 for dc = 200 h−1 kpc, which indicates that mergers of more massive clusters produce higher offset ratios. For each mass, the ratios reveal an anti-correlation with the relative velocity, and the correlation slopes for different masses are close. The top right panel presents the dependence of the ratio on dc for M1 = 5 × 1014M, where the duration of the larger offset is shorter. However, the difference is not significant at the high-velocity end. The bottom left panel presents the dependence on different impact parameters for M1 = 5 × 1014M. As seen from the panel, the slopes of the curves decrease as the impact parameters increase, and consequently the effective velocity range for a positive offset ratio decreases strongly, e.g., V < 1000 km s−1 for P = 400 kpc. The bottom right panel shows the impact of the different mass ratios on the offset ratio for M1 = 5 × 1014M and P = 0 kpc. We find that if 1.5 < ξ < 2.5, the amplitudes of the offset ratios are close; while if ξ is larger than 3 or approaches to 1, the offset ratio decreases significantly. The non-monotonic feature of the ξ = 1.5 curve that R becomes smaller with decreasing V at the low-V end can be understood as follows: the merging configuration is close to symmetric and thus the offset becomes insignificant; the effect becomes more significant as the relative velocity is lower due to the weaker collision and the longer interaction time.

According to the dependence behavior of the offset ratio on the initial conditions shown in Figure 8, we find that the following functional form fits the data well,

Equation (13)

where the offset ratio is linearly correlated with the initial relative velocity. The mass-scaled terms Dscale and Pscale are included to indicate the dependence on dc and the impact parameter. We fit simultaneously to all of the simulation results but fix ξ = 2. Considering the limited simulation test explored in this study, the mass ratio is not taken as an argument in the above fitting formula. The possible effects caused by different ξ will be discussed at the end of Section 4.2. The best-fit parameters (Al, αl, βl, γl, δl) and their standard errors are given in Table 1. The average deviation between Equation (13) and the simulated data is 10%. Note that Equation (13) is merely suitable for M1 > 1014M and dc = 50–300 h−1 kpc and the ratio R(d > dc) is constrained to be non-negative. Therefore, the effective velocity range for the positive offset ratio is $V<\beta _{\rm l}(M_1/ 10^{14}{\,M_\odot })^{\gamma _{\rm l}}\times 10^3 {\rm \,km\,s^{-1}}$ for P = 0, and the upper limit of the effective velocity range is a few tens of percent larger than the critical velocity Vcrit given in Equation (9). However, when P > 0, the effective velocity could be much smaller than Vcrit.

Table 1. Best-fit Parameters for Equations (13) and (14)

Equation (13) Al αl βl γl δl
−0.267 ± 0.007 −0.150 ± 0.013 1.847 ± 0.041 0.193 ± 0.008 0.430 ± 0.042
Equation (14) As αs βs γs δs
  −2.10 ± 0.50  0.76 ± 0.13 0.390 ± 0.076 −1.04 ± 0.12  −0.672 ± 0.067 

Download table as:  ASCIITypeset image

We investigate those cases with offsets smaller than 50 h−1 kpc below, separately, which is necessary especially when we estimate the observational expectation of the SZ–X-ray offset distribution over all ranges of the offset size in Section 4.4. There are at least the following several reasons to separately investigate the large offsets (i.e., dc ⩾ 50 h−1 kpc) and the small ones (i.e., 0 < dc < 50 h−1 kpc), respectively. (1) As discussed above, the two ranges of the offsets are caused by different reasons. The large offsets are strongly affected by the "jump effect," mostly appearing between the primary and secondary pericentric passages. On the contrary, the small offsets can be viewed during the whole merging processes. (2) As the uncertainty in our offset estimation is about a few kiloparsecs the time duration of the small offsets (≲ 50 h−1 kpc) has a relatively larger error (see Table 1). (3) Furthermore, the statistics of the offsets smaller than 50 h−1 kpc performs relatively irregular behavior, and is more complex to be described. Note that the criterion to separate the large and the small offsets is at 50 h−1 kpc or a few dozens of kiloparsecs, which is reasonable because the core radius of the initial gas distribution is ∼100 h−1 kpc. While the size of the SZ–X-ray offset is comparable with or larger than the core radius, the SZ and the X-ray peaks actually locate near the centers of the big and the small clusters, respectively. Only the disturbed core region of the big cluster itself could not generate such a large offset.

We count the ratio of the small offsets by using our simulations and setting dc = 10, 20, 30, 40 h−1 kpc, respectively. We find that the offset ratios also show a linear correlation with the relative velocity as those of dc ⩾ 50 h−1 kpc, though the correlation has a larger scatter than that shown in Figure 8. In addition, for dc < 50 h−1 kpc, the offset ratios are not strongly related with the impact parameter, and thus we use a mild dependence on the impact parameter in the fitting function of R(d > dc) given by

Equation (14)

where the best-fit parameters (As, αs, βs, γs, δs) and their standard errors are given in Table 1. The average deviation between Equation (14) and the simulated data is 20%.

4. PROBABILITY OF THE SZ–X-RAY OFFSET

By exploring the parameter space of the initial conditions in Section 3, we find that the cluster mergers with M1 > 1014M, P < 400 kpc and ξ ⩽ 3 may form the SZ–X-ray offset larger than 100 kpc. The parameter space almost covers the whole mass range of the galaxy clusters which experienced major mergers. In this section, we investigate the probability of a significant SZ–X-ray offset occurring in observations.

4.1. Model of the Offset Rate

The number of the mergers with observational offset4 dphy > dc per unit redshift (or per unit cosmic time after multiplying by dz/dt) per unit comoving volume at redshift z, which is referred to as the offset rate hereafter, over a given range of mass M0 ∈ [Mmin, Mmax] (M0M1(1 + ξ−1), is the total mass of the merging system), mass ratio ξ ∈ [ξmin, ξmax], impact parameter P ∈ [Pmin, Pmax], and initial relative velocity V ∈ [Vmin, Vmax] can be given by

Equation (15)

where fP(P) and fV(V) are the initial distribution functions of the impact parameters and the relative velocities of merging cluster pairs with ∫fP(P)dP = 1 and ∫fV(V)dV = 1, respectively. In Equation (15), Sp(dphy > dc|M0, ξ, P, V) is defined as the specific probability of a merging system with a given initial condition (M1, ξ, P, V) showing the observed offset dphy > dc in one observational direction, which can be obtained through the ratio of the time duration of the observed offset (>dc) over the total merging time (from the primary pericentric passage to the complete relaxation; see Equation (16) below), and 〈Sp(dphy > dc|M0, ξ, P, V)〉 is the average specific probability of the observed SZ–X-ray offset dphy > dc over all random observational directions. The H(dmaxdc) is the Heaviside step function, where dmax is the maximum of the SZ–X-ray offset during the merger (i.e., Equation (10)). The merger rate B(M0, ξ, z) is defined so that B(M0, ξ, z)dM0dξdz represents the comoving number density of galaxy cluster mergers completed at redshift zz + dz, with primary mass in the range M0M0 + dM0 and mass ratio in the range ξ → ξ + dξ. The Δz is obtained by the cosmic time difference t(z − Δz) − t(z) = (trelaxt1st) − (1/2)(t2ndt1st) for large offsets (e.g., ≳ 50h−1 kpc) and t(z − Δz) − t(z) = (1/2)(trelaxt1st) for small offsets (e.g., ≲ 50h−1 kpc), where the large offsets occur mainly at the primary core–core interaction stage as mentioned in Section 3.1.3.

Below we present the detailed forms of the functions 〈Sp〉, B, fV, and fP in Sections 4.1.14.1.4, respectively.

4.1.1. The Average Specific Probability of the Offset 〈Sp

We define the average specific probability of offset in Equation (15) by

Equation (16)

where 〈Γ(dphy > dc)〉 is the average of the time duration of the observed offsets larger than dc over all possible observational directions. To connect the average duration and the duration observed along the z-axis that was discussed in Section 3, we introduce the following projection factor

Equation (17)

In principle, the projection factor should depend on the initial parameters and dc. We investigate the dependence by using the Monte Carlo method and randomly selecting the observational directions for the given snapshots, and the result is shown in Figure 9.

Figure 9.

Figure 9. Projection factor defined by Equation (17) as a function of dc. Different point colors and types represent different initial parameters. P and ξ are fixed to be 0 kpc and 2 in the top panel; M1 and V are fixed to be 5 × 1014M and 500 km s−1, respectively, in the bottom panel. As seen from the figure, the factor does not significantly depend on the initial conditions.

Standard image High-resolution image

As seen from Figure 9, the values of the points at a given dc are quite close. The only outliers are the red points in the lower panel for the case with M1 = 5 × 1014M, ξ = 1.5, V = 500 km s−1, and P = 0 kpc, which is due to the near symmetry of the ξ = 1.5 merging system. The factors for the mergers with V = 500 km s−1 is around 10% smaller than that for the corresponding cases with V = 1000 km s−1. That is, the projection factor does not significantly depend on the cluster mass, the mass ratio, the impact parameter, and the initial relative velocity. Thus, we assume a universal factor 〈Ap(dc)〉, which is only a function of dc. For simplicity, we set 〈Ap(dc)〉 to be the mean value of the data shown in the top panel of Figure 9 at each dc. For the offsets smaller than 50 h−1 kpc, we also find a similar result that the projection factor can be assumed to be only a function of dc. The results of 〈Ap(dc)〉 are listed in Table 2.

Table 2. The Projection Factor 〈Ap(dc)〉 (Section 4.1.1)

dc(h−1 kpc) 10 20 30 40  
Ap(dc)〉 0.92 0.79 0.79 0.86  
dc(h−1 kpc) 50 100 150 200 300
Ap(dc)〉 0.95 0.90 0.87 0.84 0.73

Download table as:  ASCIITypeset image

By combining Equations (11), (16), and (17), we have

Equation (18)

where κ ≡ (trelaxt1st)/D as defined after Equation (12).

4.1.2. The Merger Rate B

We approximate the merger rate of cluster pairs by using the following universal fitting form of the halo merger rate obtained in Fakhouri et al. (2010, see also Genel et al. 2009):

Equation (19)

where AB = 0.0104, $\tilde{M}=1.2\times 10^{12}{\,M_\odot }$, α = 0.133, β = −1.995, $\tilde{\xi }=0.00972$, γ = 0.263, η = 0.0993. Note that the mass ratio defined in this work is the reciprocal of that in Fakhouri et al. (2010). For the dark matter mass function n(M0, z), we use the universal form derived from the N-body simulations (Tinker et al. 2008; see also Sheth & Tormen 1999; Jenkins et al. 2001),

Equation (20)

where $\overline{\rho }_m$ is the present mean mass density of the universe, σ is the square root of the variance of mass, and f(σ) gives the fraction of the mass associated with halos in a unit range of ln σ−1 (see Equation (3) in Tinker et al. 2008).

4.1.3. Initial Relative Velocity Distribution fV(V)

We approximate the probability distribution function of the pairwise velocity obtained from cosmological simulations with halo masses above 1014M (Thompson & Nagamine 2012) as the distribution of initial relative velocity fV(V) in this work. The 2016 h−1 Mpc box size employed in the simulation of Thompson & Nagamine (2012) guarantees our requirement for the statistic analysis of massive major merging systems. The best-fit of a skewed normal distribution to the simulation results obtained in Thompson & Nagamine (2012) is

Equation (21)

where x = log10(V/ km s−1), ∫fx(x)dx = 1, and the best-fit parameters are (a, e, w) = (− 2.19, 2.90, 0.295). The peak of this distribution locates at 500–600  km s−1. However, the velocity distribution could depend on cluster masses, which was not considered in Thompson & Nagamine (2012). Some other works in the literature also discussed the relative motion of cluster pairs. For example, Dolag & Sunyaev (2013) presented the dependence of the relative motions of cluster pairs on their distance, where the median relative velocity can be described by a simple functional form of 〈V〉 = 270 km s−1 + 1000 km s−1 × d−1 (here d represents the separation of the two clusters and is measured in units of the sum of their virial radii). In our simulations, the initial separation of two merging clusters is set to d = 2, and thus correspondingly the median relative velocity is 〈V〉 = 770 km s−1 if adopting the results from Dolag & Sunyaev (2013), which is about 25% larger than the median value of Equation (21). In addition, Wetzel (2011) suggested that the average value of the infalling velocities of the satellite halos shown in cosmological simulations is approximately the circular velocity of the primary halo at the virial radius (vc), which is smaller than that adopted in our study.

Based on the relative velocity distribution obtained from Thompson & Nagamine (2012), we estimate the probability of the flyby mode defined in Section 3.1.1. It is smaller than 1% of the whole merging events, which is the reason why we only use the merger rate B(M1, ξ, z) in Equation (15) but ignore the contribution from the rare flyby mode. We ignore the redshift evolution of the pairwise velocity distribution for clusters, which seems insignificant at redshift z ∼ 0–0.5 (Thompson & Nagamine 2012). The peak positions of the pairwise velocity distribution in Bouillot et al. (2014) show a slightly increasing trend from z = 0 to 0.5, which however has no significant effect on the result (more discussions seen in Section 4.3).

4.1.4. The Impact Parameter Distribution fP(P)

The distribution of impact parameters for major mergers of massive clusters was investigated in the literature, and it is not easy to give a universal quantitative description for this distribution. We list some works below. Sarazin (2001) suggested that most mergers are expected to involve fairly small impact parameters comparable to the sizes of the gas cores in clusters, which may be also biased to a lower value if most mergers occur along large-scale structure filaments. Wetzel (2011; see also Vitvitska et al. 2002; Benson 2005) presented the distributions of the radial (vr) and the tangential (vt) velocities of DM substructures at the time of crossing within the virial radius of a larger host halo, where the peak of the distribution is centered on vr ≈ 0.89vc and vt ≈ 0.64vc and the impact parameters are implied to be a few hundred kiloparsecs. However, Vitvitska et al. (2002) reported that the tangential velocity decreases with the increase of the secondary mass, and major mergers are significantly more radial than minor mergers. In addition, Benson (2005) found the evidence for a mass dependence of the distributions of orbital parameters, i.e., the orbits of more massive merging systems are more radial and less tangential, though their small sample size limits an accurate determination of this dependence. According to the argument in Poole et al. (2006), the statistics including the secondary or tertiary encounters with the primary one tends to overestimate the tangential velocity, and the average impact parameter is likely to be smaller than what vt/vr implies in Benson (2005). Khochfar & Burkert (2006) also reported a distribution of the impact parameters, which shows a peak around 350 h−1 kpc; however, the application of that result to the analysis of major mergers of galaxy clusters would be limited as the simulation box size is not sufficiently large enough.

In this work, we construct the distribution of the impact parameters by using the following form,

Equation (22)

where λ and μ are the free parameters to control the position and the width of the distribution peak, and AP is the normalization. When the parameters are (AP, λ, μ) = (7.620 × 10−3h kpc−1, 327.1 h−1 kpc, 1.636), Equation (22) reduces to the best-fit distribution (Equation (11)) obtained in Khochfar & Burkert (2006). In this work, we use (AP, λ, μ) = (5.0 × 10−3h kpc−1, 200.0 h−1 kpc, 1.0) as the fiducial model, which gives the distribution peak at 200 h−1 kpc. In Section 4.2 below, we also try various choices of the parameters to test the dependence of the probability of the SZ–X-ray offsets on the distribution of the impact parameters.

4.2. Results of the Expected Offset Probability

We use Equation (15) to estimate the offset rate. In our calculation, the integration limits in Equation (15) are set as follows. Current SZ surveys have discovered clusters in a large mass range from a few 1014 to a few 1015M (Williamson et al. 2011; Reichardt et al. 2013). We set Mmin = 2  ×  1014h−1M and Mmax = 3  ×  1015h−1M in our calculation. We also investigate the results with different Mmin  in Figure 11. We set ξmin  = 1 and ξmax  = 3, as only major mergers could produce the obvious offsets in the simulations. We set Vmax  = Vcrit (see Equation (9)), as we ignore the flyby mode in this work; and Vmin is 10 km s−1. We set Pmin  = 0 and Pmax  = 600 h−1 kpc. We test for a larger value of Pmax  and find no much difference in the result, as mergers with P > 400 kpc only induce small offset ratio.

We define the cumulative offset probability of the SZ–X-ray offset as follows,

Equation (23)

In Figure 10, we show the cumulative probability as a function of dc at different redshifts. As seen from the figure, the cumulative probability is smaller while dc > 50 h−1 kpc at lower redshift. The probability for significant offsets observed at higher redshift is higher because the factor of |dz/dt|(trelaxt1st) in Equation (23) is larger at higher redshift. For all redshift cases, the cumulative probability is flat over the range of 50 < dc < 300 h−1 kpc. This is consistent with the "jump effect" shown in Section 3 that the offset rapidly increases to a few hundred kiloparsecs after the small cluster passes through the larger one, which roughly equals to the displacement between the mass density centers of the two clusters. Forero-Romero et al. (2010) investigated the distribution of displacements between the peaks of the DM and the gas mass densities in clusters from a large nonradiative SPH ΛCDM cosmological simulation. They found that about 10% of the massive clusters (with masses ranging from 2.0  ×  1014h−1M to 2.5  ×  1015h−1M) at redshift z = 0.5 have displacements larger than 100 kpc, and about 3% are larger than 200 kpc. Their results are roughly consistent with ours for dc ∼ 50–100 h−1 kpc; however, when dc > 100 h−1 kpc, their probability is several times smaller than our results. The difference is caused by the jump effect of the X-ray position as discussed above. The period when the X-ray position locates around the center of the small cluster is commonly 0.5 Gyr or longer. This significantly enhances the probability of the offset larger than 100 h−1 kpc, which however might be omitted by the used hierarchical friends-of-friends algorithm.

Figure 10.

Figure 10. Cumulative probabilities of the clusters with offsets larger than dc at different redshift, from 0.1 to 1.0. The merger rate of cluster pairs and their initial relative velocity distribution and impact parameter distribution used here come from the fiducial model based on the currently available numerical simulations discussed in Sections 4.1.24.1.4 (similarly for Figure 11, the right panel of Figures 15 and 16(a) below). The best fits to the points are shown by the solid (z = 0.7, i.e., Equations (24) and (25)) and dotted lines. As seen from the figure, the higher redshift case shows a larger cumulative probability of dc > 50 h−1 kpc. See details in Section 4.2.

Standard image High-resolution image

We calculate the cumulative probability for dc > 300 h−1 kpc by extrapolating Equation (13). As shown in Figure 10, the probabilities reveal an obvious decay when dc ≳ 350 h−1 kpc, mostly because Mmin in Equation (15) is larger when V is close to the peak of the velocity distribution (see Equation (10)).

Specifically, we discuss the probability at z = 0.7 in detail for two reasons, (1) the median redshift of the SZ sample in observations is nearly 0.5 (Marriage et al. 2011; Reichardt et al. 2013); (2) the average redshift of the observed SZ clusters to be compared with our estimation is 0.7 (see Section 4.4). As seen from Figure 10, for the significant offsets with dc = 50, 100, 150, 200, 300 h−1 kpc, the cumulative probabilities of the unrelaxed clusters with SZ–X-ray offset larger than dc are 11.1%, 9.5%, 8.6%, 8.0%, and 6.5%, respectively. The best fit to these probabilities by a power-law form ($\propto d_{\rm c}^{b_{\rm l}}$) is shown as the dotted line, with bl = −0.27. As discussed above that significant offsets might be omitted in Forero-Romero et al. (2010), our best-fit power index (bl = −0.27) is flatter than their result (bl = −1.0).

In addition, the probability of dc = 500 h−1 kpc is approximately one third of that of dc = 300 h−1 kpc at z = 0.7. We assume an exponential decay to describe this behavior and to extend the best-fit power-law form. Consequently, we fit the cumulative probability for dc > 50 h−1 kpc as follows,

Equation (24)

where al = 0.085, bl = −0.27 are the best-fit parameters, and d* is set to 140 h−1 kpc. The best fits of xd and xs are 369 and 115 h−1 kpc, respectively.

For the offsets with dc < 50 h−1 kpc, the jump effect has little influence on the cumulative probability. Our results show that the probabilities at z = 0.7 are 62.9%, 34.0%, 21.4%, and 14.6%, when dc = 10, 20, 30, and 40 h−1 kpc, respectively. We follow Equation (1) in Forero-Romero et al. (2010) to fit the results as

Equation (25)

The best-fit gives as = 0.072, bs = −0.85, approximately consistent with those in Forero-Romero et al. (2010). Since the fitting form is divergent when dc approaches zero, we set an cutoff at the offset where the cumulative probability is 1.0 in Equation (25).

In Figure 11, we show the cumulative probabilities of the SZ–X-ray offset with different Mmin  (Equation (23)) in panel (a) and the contribution of mergers in different cluster mass ranges to the cumulative probability of those clusters with Mmin  = 2 × 1014h−1M in panel (b), respectively. Mergers of the massive systems (>5 × 1014h−1M) tend to produce a larger offset. The cumulative probability of dc = 100 h−1 kpc is 5% for the Mmin  = 1 × 1014h−1M case, while it is about 30% for the Mmin  = 5 × 1015h−1M case (see panel (a)). The results in panel (b) shows that the clusters in the low mass range (2 × 1014–5 × 1014h−1M) dominate the contribution of the offset. In Table 3, we list the best-fit parameters of the cumulative probabilities for Equations (24) and (25) for different redshift and mass range (see Figures 10 and 11(a)). We find that the parameters of as = 0.015(1 + z)3.1, bs = −0.85, al = 0.018(1 + z)3.1, bl = −0.27, xd = 369 h−1 kpc, and xs = 151(1 + z)−0.5h−1 kpc may fit the redshift dependence of the cumulative probability shown in Figure 10 well (Mmin  = 2 × 1014h−1M).

Figure 11.

Figure 11. Left: cumulative probabilities for the clusters with SZ–X-ray offsets larger than dc at redshift z = 0.7. The lines from top to bottom show the results obtained by choosing different Mmin  in Equation (23). The cumulative probability of dc > 50 h−1 kpc increases with increasing Mmin . Right: contribution of mergers in different cluster mass ranges to the cumulative probability obtained with Mmin  = 2 × 1014h−1M. As seen from the panel, the mergers of the clusters in the mass range 2 × 1014–5 × 1014h−1M dominates the contribution of the offsets.

Standard image High-resolution image

Table 3. Best-fit Parameters of the Cumulative Probabilities in Equations (24) and (25) for Different Mass Ranges at z = 0.7

Mass range as bs al bl xd xs
(h−1M) (h−1 kpc) (h−1 kpc)
Mmin  = 1 × 1014 0.026 −1.1 0.040 −0.29 326 109
Mmin  = 2 × 1014 0.072 −0.85 0.085 −0.27 369 115
Mmin  = 3 × 1014 0.13 −0.71 0.14 −0.28 420 114
Mmin  = 5 × 1014 0.27 −0.56 0.26 −0.28 470 151

Download table as:  ASCIITypeset image

The SZ surveys with large sky coverage may be able to determine both the probabilities of the clusters with offsets larger than a given dc and then the offset rate dN(dphy > dc|z)/dz, which depends on the cluster merger rate B in Equation (15) and thus help to constrain it.

Finally, we discuss some possible uncertainties in the above estimation of the cumulative offset probability.

  • 1.  
    In the fitting form of the offset ratio in Equations (13) and (14), we do not consider the mass ratio as an argument. The bottom right panel of Figure 8 reveals that the offset ratio with ξ = 3 is about 30%–50% lower than that with ξ = 2. Thus, we might overestimate the probability by a factor of ≲1.5–2. If we select a tighter constraint of the mass ratio range in Equation (15) with (ξmin , ξmax ) = (1.5, 2.5), the cumulative probability shown in Figure 10 (solid line) becomes 5.4%, 4.6%, 4.2%, 3.9%, and 3.2% for dc =50, 100, 150, 200, and 300 h−1 kpc, respectively, which are nearly half of the values reported above.
  • 2.  
    In this work, the offset ratio R(d > dc) is assumed to be redshift independent. The redshift evolution shown in Figure 10 is purely introduced by the halo merger rate B(M0, ξ, z − Δz) in Equation (15). However, the virial radius of the galaxy cluster is proportional to (1 + z)−1; the concentration parameter is anti-correlated with redshift; and the physical mass densities of the dark matter and gas halos also depend on redshift. We use Figure 12 to indicate the effect on the maximum and the duration of the SZ–X-ray offset from the redshift-dependent physical size of the galaxy cluster. In Figure 12, we show the results of the ratio of R(d > dc) at redshift z = 0.5, 1.0, and  2.0 to its value at redshift z = 0, by performing a series of cluster merger simulations with M1 = 2 × 1014, 5 × 1014, 1 × 1015M, P = 0 kpc, V = 500 km s−1, ξ = 2. As seen from the figure, the deviation of the ratio from unity is smaller than 10% (or 20%) at redshift z = 0.5 (or z = 1) for most of the cases. For dc ⩽ 100 h−1 kpc, the ratios are possibly larger than 1 at higher redshift due to the higher central density of the cluster and thus a strong collision intensity. For larger dc, the ratios shift to a lower level at high redshift. At redshift z = 2, the redshift evolution of the offset ratio becomes more significant, especially for dc ⩾ 200 h−1 kpc. The maximum of the SZ–X-ray offset in the M1 = 2  ×  1014M case is smaller than 200 h−1 kpc at z = 2, which implies that the redshift-dependence of the cluster size (∝(1 + z)−1) plays a more important role on the SZ–X-ray offset size with increasing redshift, although a higher central density enhances the strength of the collision. We conclude that the redshift evolution effect is an important factor for estimating the offset probability only at z ≳ 1. Considering both the smaller offset ratio R(d > dc) and the higher mass limit of Mmin in Equation (15) at high redshift, the cumulative probability of the offset for dc ⩾ 50 h−1 kpc may be suppressed.
Figure 12.

Figure 12. Ratios of R(d > dc) at redshift z to its value at z = 0, obtained from a series of cluster merger simulations with ξ = 2, P = 0 kpc and V = 500 km s−1 at redshift z = 0.5, 1.0, 2.0. Different colors represent the simulation results with different dc, and different types of the points represent different cluster masses, as labeled in the figure. The horizontal dotted line at Rz/Rz = 0 = 1 is a reference line. As seen from the figure, the ratios for a large dc (e.g., ≳ 200 h−1 kpc) decrease at higher redshift.

Standard image High-resolution image

4.3. Impacts of Different Velocity and Impact Parameter Distributions on the Probability

In Section 4.2, we estimate the cumulative probability of the SZ–X-ray offset by using the fiducial model of the distributions of relative velocities and impact parameters presented in Sections 4.1.3 and 4.1.4. The forms and the parameters of the fiducial model are motivated or obtained from cosmological simulations, but there is no observational constraint on them so far. In this section, we discuss the impacts of different distribution parameters on the cumulative offset probability, and the realistic observational results may put constraints on them.

In Figure 13, we explore the influence of the pairwise velocity distribution on the probability. In the left panel, in addition to e = 2.9 in the fiducial model fitted from the cosmological simulation, we also obtain the results with different values of e(=2.5, 3.2, 3.5) in Equation (21) to test the effects of different peak positions of the velocity distribution, where w is fixed to be 0.295. The peaks locate in the range of ∼200–2200 km s−1. In the right panel, we test the effect of different FWHMs in the velocity distribution, where we set different values of w(=0.10, 0.15, 0.29, 0.50) and the value of e is selected to keep the peak positions of the distributions the same as that of the fiducial model. According to the results shown in Figure 13, we find that the amplitude of the probability is sensitive to the peak position of the velocity distribution, but not to the width. The probability obtained in the fiducial model is approximately larger than that obtained with e = 3.2 and 3.5 by a factor of 1.6 and 5.0, respectively. As seen from the figure, the shapes of the cumulative probabilities of dc ⩽ 300 h−1 kpc with different velocity distributions are approximately similar. That is because the shape of the probability is mainly determined by the dependence of the offset ratios and the projection factor on dc (see Equation (18)), and the offset ratios are proportional to $d_{\rm c}^{\alpha _{\rm l}}$ in Equation (13), where the best-fit index αl is approximately a constant and not dependent on the parameters V and P. The same feature also appears in Figure 14 for the effects of different impact parameter distributions below. However, the cumulative probabilities of dc > 300 h−1 kpc are sensitive to the peak position of the velocity distribution. The reason is that the mergers with high relative velocities produce large offsets. In the e = 2.5 case, the upper boundary of the velocity distribution is about 600 km s−1, and only the massive clusters could produce offsets larger than 400 h−1 kpc (see Equation (10)). Due to the relatively small number of the massive clusters, the cumulative probability decays rapidly with dc > 400 h−1 kpc. When the peak position of the relative velocity distribution is higher (e.g., e = 3.2), the cumulative probability becomes flatter at dc > 300 h−1 kpc, as the relatively less massive cluster mergers can also produce relatively large offsets.

Figure 13.

Figure 13. Dependence of the cumulative probability of the SZ–X-ray offset on the pairwise velocity distribution. Left: dependence on the peak position of the distribution. Four velocity distributions with different e = 2.5, 2.9, 3.2, 3.5 (see Equation (21)) are shown in the inset. Different curve types represent different values of e, as labeled in the inset. The curves in each panel are the best-fit of the probability as a function of dc, and the points on the curves represent corresponding simulation results. The solid line represents the fiducial model used in this paper (see Section 4.1.3), similarly for the right panel. Right: dependence on the FWHM of the distribution. We set w = 0.10, 0.15, 0.29, 0.50 in Equation (21), and the value of e is set by keeping the same peak positions as that of the fiducial model. As seen from the figure, the cumulative probability is sensitive to the peak position but not the width of the velocity distribution. See details in Section 4.3.

Standard image High-resolution image
Figure 14.

Figure 14. Dependence of the cumulative probability of the SZ–X-ray offset on the impact parameter distribution. Left: dependence on the peak position of the distribution with the same FWHM, where λ = 50, 200, 500 h−1 kpc in Equation (22). The probability obtained from the distribution reported in Khochfar & Burkert (2006) is also presented (dot–dashed curve). The curves have the similar meanings as those in Figure 13, though for different sets of the parameters in the impact parameter distribution. The solid curves represent the fiducial model used in this work (see Section 4.1.4). Right: dependence on the width of the distribution with the same peak position λ = 200 h−1 kpc. The figure shows that the amplitudes of the cumulative probabilities depend on both the peak position and the width of the impact parameter distributions. See details in Section 4.3.

Standard image High-resolution image

Note that while the velocity distributions shift to the high-velocity end, the flyby mode should not be ignored any more. From our simulations, we find that the time durations of the significant SZ–X-ray offset of the flyby mode are usually 1–1.5 Gyr when V ∼ 2000 km s−1 for different cluster masses (e.g., see the bottom panel of Figure 2). The durations are also inversely proportional to the relative velocity and very sensitive to the impact parameter (because off-axis mergers with high relative velocity, e.g., V > 2000 km s−1, are relatively ineffective in destroying the gas cores in the large clusters, the offsets are strongly suppressed when the impact parameter gets larger). We roughly estimate the effects of the flyby mode on our calculation. We find that when e = 3.2, the effect is smaller than 10%; and when e = 3.5, the probability will increase by ∼50% after including the flyby case. The flyby mode, however, does not significantly weaken the tendency that a higher peak position of the relative velocity distribution results in a smaller probability. Here we stress that the sensitivity of the probability to the velocity is the crux of the Bullet Cluster problem discussed in Lee & Komatsu (2010) and Thompson & Nagamine (2012). The parameter space search of the Bullet Cluster (1E0657-56) in the literature suggested that such a system requires a high relative velocity during the merger (e.g., Mastropietro & Burkert 2008), which possibly challenges the standard ΛCDM model. We suggest that the cumulative probability of the observed SZ–X-ray offset could provide an opportunity to examine the incompatibility existing between observed bullet clusters and cosmological simulations.

In Figure 14, we show the cumulative probabilities obtained from different distributions of the impact parameters. The left panel displays the results of the distributions with different peak positions (i.e., 50, 200, 500 h−1 kpc), but with the same FWHM. The distribution reported in Khochfar & Burkert (2006) is also tested. We find that if the peak positions are smaller than 200 h−1 kpc, the results show little difference. However, if the peak of the distribution shifts to a larger value, the probability is obviously suppressed as shown by the λ = 500 h−1 kpc case in the panel. Compared with our fiducial model, the cumulative probability obtained from Khochfar & Burkert's (2006) distribution is smaller by a factor of 1.7. In the right panel, we change the width of the distribution, but keep the same peak position. We find that a wider distribution gives smaller probability, which is reasonable as the larger impact parameter components contribute more to the distribution in the wider case but less to the large offsets. For the two extreme cases shown in the panel, the probability obtained with μ = 3.0 is approximately two times larger than that obtained with μ = 0.5. As a result, we find that the amplitude of the cumulative probability (dc > 50 h−1 kpc) depends both on the peak position and the width of the impact parameter distributions. However, the shape of the cumulative probability does not, which is different from the dependence on the velocity distributions.

4.4. The SZ–X-Ray Offset in Observations

In this section, we compare our estimation of the cumulative probability of the SZ–X-ray offset with observations in the real universe. This comparison is motivated mainly by the following reasons: (1) in Section 4.2, we estimate the probability of the offset based on the fiducial model, and find that galaxy clusters with significant offsets are not rare, for example, approximately 10% of the clusters have offsets larger than 50 h−1 kpc; (2) the past several years have seen rapid progress in the SZ cluster observation, in terms of the total numbers and the parameter ranges (precision and redshift); and (3) the comparison would potentially provide constraints to the model used in the estimation of the offset probability, e.g., the pairwise velocity distribution. The majority of the observed SZ clusters to be compared with are at redshift z ∼ 0.7 (Andersson et al. 2011) and 1'' corresponds to 5 h−1 kpc at redshift z = 0.7. In the following comparison, we assume that 1'' corresponds to 5 h−1 kpc, for simplicity.

The spatial offset between the X-ray and the SZ peaks shown in observations can be modeled to comprise the two following components,

Equation (26)

where dphy is the physical one produced by the energetic merger defined in Equation (15) and derr is the observational error. The (differential) distribution of the physical offset dphy follows the derivative of the offset cumulative probability (i.e., Equations (24) and (25)) with respect to dc. In this work, we assume that both of the observational errors in the spatial positions of the X-ray and the SZ peaks follow a Gaussian distribution. The standard deviation of derr for the X-ray peak (σX-ray) is set to 1'', according to the current capability of the X-ray instrument (e.g., Chandra X-ray Observatory). For the SZ effect, the typical position uncertainty of the SPT SZ cluster centroid is approximately σSZ = 1'2/SNR ∼ 15'', which dominates the uncertainty derr. In addition, we also test the results by assuming two higher resolutions of the SZ effect (σSZ = 2'', and 8'') in this work.

Figure 15 shows the statistical distribution of the SZ–X-ray offsets dobs by using the Monte-Carlo method to simulate both dphy and derr in Equation (26). The bin size of the offsets is 5'' in both panels of Figure 15, and the histogram represents the percentage of the number fraction of the simulated sample in each bin. To illustrate the effect of different distributions of dphy, we show the observational expectations obtained by assuming that dphy = 0 and dphy follows the derivatives of Equations (24) and (25) in the left and the right panels, respectively; for example, the values of the differential distribution of dphy at 2'', 4'', 6'', 8'', 10'', 20'', 40'', 60'', 80'', and 100'' in the right panel are 2.91 × 10−1, 8.1 × 10−2, 3.8 × 10−2, 2.2 × 10−2, 1.5 × 10−2, 1.3 × 10−3, 5.2 × 10−4, 3.1 × 10−4, 2.3 × 10−3, and 8.9 × 10−4 arcsec−1, respectively. In the left panel of Figure 15, the offsets are contributed purely by the observational errors; while those SZ–X-ray offsets in the right panel are contributed by both the observational errors and the underlying physical ones. The distribution of dphy is displaced in the inset of the right panel of Figure 15 as the solid line, and the discontinuities appear at dc = 10farcs7 and 73farcs8, corresponding to the transitions of the different behavior of the cumulative probability in three different regions of dc as discussed in Section 4.2 (see Equations (24) and 25). As seen from the right panel, the distribution of dobs is bimodal once the underlying physical distribution dphy is considered: some clusters peak around dobs ∼ 0 and the others peak at dobs ∼ 70''–90''. The lack of clusters with the SZ–X-ray offsets around 50''–60'' is mainly due to the "jump effect" as shown in Figures 24, and 5, respectively. The location of the right peak (∼ 70''–90'') is determined by the combination of the following factors: (1) the maximum offset caused by a merger increases with the cluster mass and the initial relative velocity, as seen from Equation (10); (2) the cluster mass function decreases with increasing mass; and (3) the underlying velocity distribution. A much larger offset is more likely to be contributed by mergers of clusters with relatively large masses and high relative velocities (e.g., see Figure 16).

Figure 15.

Figure 15. Distributions of the offsets between X-ray and SZ peaks of galaxy clusters. The histograms represent the results obtained from our model and Monte-Carlo realizations, i.e., the percentage of the cluster number in each offset bin. Left: the distributions obtained by assuming that the offsets are purely contributed by observational errors (i.e., dphy = 0 in Equation (26)). Different histograms represent different resolutions of SZ effects, as labeled in the panel. Right: the distributions obtained by assuming that the offsets are contributed by both the physical offsets and the observational errors. The "+" points are the distributions of the observational sample of 15 SPT SZ clusters obtained in Andersson et al. (2011). The inset in the right panel displays the probability distribution function of the SZ–X-ray physical offsets (green) and their cumulative distribution (red). See Section 4.4.

Standard image High-resolution image
Figure 16.

Figure 16. Distributions of the offsets between X-ray and SZ peaks of galaxy clusters obtained with different cluster mass ranges and pairwise velocity distributions. Top: the distributions of the offsets with different mass ranges Mmin  = 1 × 1014 and 5 × 1014h−1M (see Equation (23)). Bottom: the distributions with different velocity distributions e = 2.5 and 3.2 (see Equation (21)). The figure shows that the bimodal distribution of the SZ–X-ray offsets depends on the cluster mass range and the relative velocity distribution.

Standard image High-resolution image

The distribution of dobs may be significantly biased away from the distribution of dphy, especially when the accuracy in determining the SZ centroid (σSZ) is not sufficient. For example, the current SPT SZ survey (σSZ = 15'') may obtain mis-estimates of the SZ–X-ray offsets ∼10''–20'' for many clusters simply because of the observational errors. However, these mis-estimates may be modeled and the underlying physical SZ–X-ray offset distribution can still be extracted even if σSZ is large. As seen from the right panel of Figure 15, the right peak of the SZ–X-ray offset distribution is hard to be affected by the uncertainties in determining the SZ centroids. If the current SZ surveys can detect many more clusters, the right peak should be able to be revealed. If future SZ surveys can achieve to higher resolution and higher accuracy in determining the SZ centroids (e.g., σSZ = 8'' or 2''), the underlying physical distribution of the SZ–X-ray offset may be better determined, and thus can be used to constrain the physics involved in the mergers of clusters.

We show the observational offset distribution obtained from the sample of Andersson et al. (2011) in Figure 15 (plus signs). The observational sample in Andersson et al. (2011) consists of 15 clusters, obtained from observations of 178 deg2 of the sky surveyed by the SPT. The average redshift of the sample is z = 0.68. The observational data shown in Figure 15 are the values after transferring the observed displacement between SPT detection and X-ray centroid to the distance at redshift z = 0.7 where 1'' ∼ 5 h−1 kpc. As seen from Figure 15, the uncertainties in determining the SZ cluster centroids probably play a dominant role for those clusters with the SZ–X-ray offsets ≲ 20''. It appears there are a few clusters with the SZ–X-ray offsets ≳ 30'', which cannot be due to the observational errors and must be due to large physical offsets. Our model shows that the cumulative probability of existing clusters with the SZ–X-ray offsets ≳ 30''–40'' is roughly 10%, which is roughly consistent with observations. Note that the current observational sample is still small (15), which may lead to large uncertainties in estimating the distribution of the SZ–X-ray offsets; for example, there is only one of the 15 clusters locating in the offset bin 45''–50'', where the Poissonian error is significant and comparison of that bin point with our model results is not reliable. If SZ surveys can detect many more SZ clusters, it would be possible to accurately estimate the observational SZ–X-ray offset distribution and use it to constrain the underlying cluster merging model and the related physics involved in. In addition, we show that the jump effect plays a dominant role in generating significant offsets, but here we do not consider the secondary X-ray maxima in the data analysis, which would suppress the probability to discover the clusters with offsets larger than 10''.

Figure 16(a) presents the model results on the SZ–X-ray offset distribution of clusters within different mass ranges (i.e., different Mmin  in Equation (23)). As seen from the figure, more massive clusters contribute larger offsets in the observation, as the right peak of the distribution resulted from the high mass case locates at a larger spatial scale in the high mass range (e.g., 100''–110'' when Mmin  = 5 × 1014h−1M). In Figure 16(b), we show the distributions obtained with different pairwise velocity distributions. As seen from the figure, the bimodal distribution of the offsets still exist for different velocity distributions. As the mean value of the relative velocity increases, the "jump effect" as shown in Figure 2 is more significant in causing the scarce of the clusters with intermediate offsets. The higher relative velocity case results in more significant offsets (>120''). The dependence of the distribution of the SZ–X-ray offsets on the pairwise relative velocity distribution of clusters suggests that the observed SZ–X-ray offset distribution can be used to probe the cosmological velocity field at the cluster scale.

5. CONCLUSION AND DISCUSSION

In this paper, we perform a series of numerical simulations for mergers of two galaxy clusters to understand the displacements between the spatial positions of the maxima of X-ray and SZ maps of galaxy clusters. The merger of two clusters destroys their initial thermal state, and the SZ–X-ray offset is produced due to the different dependence of the X-ray and SZ emissions on the density and the temperature distributions of the gas. We find significant offsets (≳ 100 kpc) mostly occur between the primary and secondary pericentric passages of the two clusters, due to the "jump effect." After the primary core–core interaction, the densest gas region locates near the center of the small cluster; and the X-ray peak may jump there from the center of the larger cluster, but the SZ peak does not.

Our simulations explore the parameter space over the primary cluster masses, cluster mass ratios, initial relative velocities, and impact parameters of two merging clusters, and we investigate the relation of the maximum and the time duration of the SZ–X-ray offset with the simulation parameter space. Our findings are summarized as follows. (1) A higher initial relative velocity triggers a larger offset. If the initial velocity is high enough (≳ Vcrit), the two colliding clusters cannot be completely relaxed within the Hubble time (i.e., "flyby mode"), different from the behavior of the "merger mode." In the flyby mode, we find that the offset can be even up to 3 Mpc. However, since the pairwise velocity distribution obtained from cosmological simulation reveals that 99% of relative velocities lower than Vcrit, the merger mode dominates the probability of the significantly large offsets appearing in the universe. For the merger mode, the qualitative features of the offsets caused by the mergers with initial velocity V = 500 and 1000 km s−1 show no significant difference. (2) The existence and the sizes of the offsets are sensitive to the impact parameter, since they are strongly related with the intensity of the core–core interactions of the two clusters. Only head-on or nearly head-on mergers can form displacements larger than 100 kpc, e.g., P < 400 kpc for the simulation with M1 = 2 × 1014M and P < 600 kpc for the one with M1 = 5 × 1014M (ξ = 2, V = 500 km s−1). Mergers with smaller impact parameters result in larger sizes and longer durations of the offsets. (3) The SZ–X-ray offsets are strongly related with the masses of the merging clusters. The masses of merging galaxy clusters that possibly result in significant offsets cover the whole mass range of galaxy clusters, i.e., M1 > 1014M. (4) The mass ratio of the two merging clusters is also an important parameter. The significant offsets are mostly formed by major mergers (1 < ξ < 3). For the mergers with ξ > 4, a large primary cluster mass M1 > 5 × 1014M and a high relative velocity V > 1000 km s−1 are required to produce offsets larger than 100 kpc.

By applying the above results summarized from the simulations to the individual cluster "Bullet Cluster" (e.g., see Figure 10 in Hincks et al. 2010), we can give a constraint on the initial relative velocity of the two merging clusters by the SZ–X-ray offset, an observational feature different from that used in previous work (e.g., in Mastropietro & Burkert 2008). In the Bullet Cluster, the maximum of the X-ray image is close to the "bullet," but that of the SZ effect locates near the centroid of the main cluster; and the displacement of the two peaks is around 300 h−1 kpc. Considering the high mass ratio of this merging system (Clowe et al. 2006; Bradač et al. 2006), a relative velocity larger than 3000 km s−1 at the initial separation 5 Mpc is required to reproduce such a significant offset in our simulations. Note that this lower limit of the initial velocity cannot be decreased, even if the projection effect and the relative uncertainty in other initial parameters are considered. The constraint on the initial velocity is in agreement with the conclusion obtained in Mastropietro & Burkert (2008), and it suggests that the SZ–X-ray offset is a good complement to the methods used in Springel & Farrar (2007) and Mastropietro & Burkert (2008), where the velocity is estimated through reproducing the morphology of the bow shock, the brightness, and the projected temperature profile across the shock discontinuity, etc. Constraints on the relative velocity of the merging cluster is important, as Lee & Komatsu (2010; see also Thompson & Nagamine 2012) conclude that the existence of the Bullet Cluster is incompatible with the prediction of a ΛCDM model (see also a contrary result in Lage & Farrar 2014), unless a lower infall velocity solution for 1E0657-56 with ≲ 1800 km s−1 at 2R200 is found. In our study (see Section 3.1.4), we show that considering the SZ–X-ray offset and the mass ratio of the two merging clusters, there is little possibility to find such a low velocity solution.

A high relative velocity for merging clusters like 1E0657-56 was also revealed, for example, Molnar et al. (2012) used simulations to reproduce the SZ–X-ray offset of the merging galaxy cluster CL J0152-1357 and found that a large relative velocity of 4800 km s−1 is necessary to explain the observations. As an indicator of the relative velocity, the SZ–X-ray offset owns the advantages of easy identification, simplicity in the relation with the projection effect, and relatively less sensitivity to the detailed gastrophysics in clusters.

To understand the statistic behavior of the SZ–X-ray offsets, we estimate the cumulative probability of the offset, which is related with the merger rate of galaxy clusters and the duration of the offsets in individual merging events. We find that the cumulative probability shows different behavior depending on whether the offset is smaller than ∼50 h−1 kpc or not. This is caused by the "jump effect," which could significantly enhance the probability of the offset (⩾50 h−1 kpc). We also find that the mergers of the low-mass clusters (i.e., 2 × 1014 < M1 < 5 × 1014h−1M) dominate the contribution of the offsets for clusters larger than 2 × 1014h−1M. The amplitude of the cumulative probability decreases with increasing redshift. For clusters with mass larger than 2 × 1014h−1M at z = 0.7 (the average redshift of the observed SZ clusters compared with our model results), the cumulative probabilities predicted from our fiducial model of the initial relative velocity and the impact parameter distribution of cluster pairs are 62.9%, 34.0%, 21.4%, 14.6%, 11.1%, 9.5%, 8.6%, 8.0%, 6.5%, and 2.0% for SZ–X-ray offsets larger than 10, 20, 30, 40, 50, 100, 150, 200, 300, and 500h−1 kpc, respectively.

We further discuss some possible uncertainties in our estimation of the probability. (1) We do not consider the mass ratio as an argument in the fitting form of the offset ratio (see Equations (13) and (14)), which might cause an overestimation of the probability by a factor smaller than two. (2) In the simulations, we build the initial conditions of the cluster structure at redshift z = 0 but not consider that the size of the cluster is redshift-dependent. We find that the redshift dependence has little effect on the probability when z < 1, though the redshift effect becomes more significant at z > 1.

We investigate the effects of different distributions of relative velocities V and impact parameters P on the probability of offset. (1) We find that the amplitude of the cumulative probability within 50 ⩽ dc ⩽ 300 h−1 kpc is partly controlled by the relative velocity distribution of merging clusters. For example, if the peak position of the relative velocity distribution shifts from 550 km s−1 to 1100 km s−1 (or 2200 km s−1), the obtained probability decreases by a factor of about 1.6 (or 5.0). Regarding the challenge of the Bullet Cluster to the standard ΛCDM model for which a high relative velocity is necessary, the crux of solving this problem is the probability of high relative velocities (e.g., >3000 km s−1), which has been widely discussed through cosmological simulations (Hayashi & White 2006; Lee & Komatsu 2010; Thompson & Nagamine 2012). However, currently there are few constraints from observations. We suggest that the SZ–X-ray offset provides a tool as the probability of the significant offsets is sensitive to the peak position of the relative velocity distribution. (2) We find that both of the peak position and the width of the impact parameter distribution affect the offset probability. If the peak position shifts from 50 h−1 kpc to 500 h−1 kpc with the same distribution width, the amplitude of the probability decreases by a factor of 4.7. As the dependence of the shape of the probability on the impact parameter distribution is weaker than the dependence on the relative velocity distribution at dc > 300 h−1 kpc, it is robust to use the cumulative probability of the SZ–X-ray offset obtained from observations to explore the distribution of relative velocities at the high-velocity end.

We compare the model distribution of the SZ–X-ray offset with observations and they are roughly consistent. However, the current sample of the SZ clusters with the SZ–X-ray offset estimates are still small, which prevents a comprehensive study from comparing the model distribution with the observations and thus putting constraint on the physics involved in the merging processes of clusters. SZ surveys, such as SPT and ACT, will detect hundreds to several thousands of clusters. Many of them are expected to be followed up by X-ray observations and thus have the SZ–X-ray offset measurements. With a substantial increase of the sample size in the future, the underlying physical distribution of the SZ–X-ray offsets can be extracted, and the unique feature of the second peak around the 70''–90'' and the scarce of clusters with dobs at 50''–60'' may be revealed. The physical offset distribution can be used to constrain not only the physics involved in cluster merging processes but also the velocity field at the cluster scale.

In this paper, we compare the cluster merging processes simulated by using the GADGET-2 and FLASH codes. In general, the density and temperature distributions of the merging structure obtained from the two codes are consistent, except that (1) the discontinuity produced by shocks are sharper in the FLASH merger; and (2) the density of the gas core in the inner region of the system produced by the SPH code tends to be higher (by <5%). These deviations in the strength of the surface brightness, however, have little influence on the positions of the X-ray and the SZ peaks, which guarantees the robustness of our simulation results obtained by merely using the SPH code.

Finally, we discuss some other possible uncertainties or assumptions in this work. First, we do not consider the radiative cooling and various heating mechanisms in the simulation. Among the heating mechanisms, AGN feedback is widely proposed for the required energy in solving the overcooling problem in clusters. However, it should not be a key factor in affecting the significant SZ–X-ray offset because its energy is at least one order of magnitude smaller than the gravitational binding energy released from the cluster merger. However, in the inner region of the cluster (e.g., ∼100 h−1 kpc), AGN feedback has the power to alter the baryon distribution, which reduces the central gas density and increases the temperature in massive clusters (Sijacki & Springel 2006; Sijacki et al. 2007). This effect should be relatively significant in the X-ray emission since its emission is proportional to $\rho _{\rm gas}^2$, but not in the SZ effect. The AGN feedback might slightly increase the SZ–X-ray offset duration Γ(d > dc), because the lower central gas density in the large cluster may delay the time of X-ray peak jumping back from the center of the small cluster. Consequently, the probability of the significant SZ–X-ray offset will be enhanced. On the other hand, the lower gas density also mildly reduces the collision strength of the mergers, which could decrease the sizes of the offsets. Though the gastrophysical processes may not have significant influence on the size of the SZ–X-ray offset as they are apt to alter the brightness but not the position of emission peaks, more quantitative explorations on this issue are required in the future. Second, we do not include magnetic field in the simulation. Unless the magnetic field is unusually strong, it should not play an important role in resulting in a significantly large offset, because of the small energy of the normal magnetic field relative to the mechanical energy involving in cluster mergers and its short tangling scale relative to typical size of the SZ–X-ray offset (Carilli & Taylor 2002). If the magnetic field is quite strong as simulated in Lage & Farrar (2014), the offset probability distribution around the high-SZ–X-ray offset peak shown in this work (the right peak in Figure 15) would be enhanced. Third, our simulations show that the relaxation of the two cluster mergers always lasts several gigayears. In such a long relaxation time, multiple mergers may have an influence on the SZ–X-ray offset. We run several typical triple merger cases (the whole parameter space is too large to explore) to examine the significance of this effect. We find that there is no apparent difference between the size and the duration of the offsets formed by binary and triple mergers. Though the behavior of the offset arisen by the triple system is complicated, the possibility of the triple mergers especially both of the two subclusters are nearly as massive as the main cluster is extremely low. Even if multiple mergers are taken into account, the correction for the probability of the offsets should be insignificant.

This research was supported in part by the National Natural Science Foundation of China under nos. 10973001, 11273004, and 11373031. Y.L. is supported in part by the BaiRen program from the National Astronomical Observatories, Chinese Academy of Sciences, and by the Strategic Research Program "The Emergence of Cosmological Structures" of the Chinese Academy of Sciences, grant No. XDB09000000. The software used in this work was developed in part by the DOE NASA ASC and NSF supported Flash Center for Computational Science at the University of Chicago.

APPENDIX: COMPARISON BETWEEN THE SPH AND AMR SIMULATIONS

In the Appendix, we show an example of simulating the merging process of two clusters by using both the SPH and the AMR methods. The parameter settings and the results of the simulation are shown in Figure 17. Figure 17 shows the slices of the simulated gas mass density and temperature distributions at the merger plane, i.e., ρgas(x, y, z = 0) and T(x, y, z = 0), obtained at different merging time. The top and the bottom panels represent the results obtained by the GADGET-2 and FLASH codes, respectively. The overlaid contours show the projected X-ray (blue) and SZ (green) surface brightness maps. As seen from the figure, the density and the temperature distributions of the merging structure obtained from the two codes are consistent in general, except for the two main different points below. (1) The discontinuity produced by shocks are sharper in the FLASH merger, which reveals the advantage of the AMR method in capturing sharp gas features. (2) In the inner region of the system, the density of the gas core produced by the SPH code tends to be higher (by <5%). This minor discrepancy is suggested to be due to the suppression of the turbulence by the SPH method, as reported in Mitchell et al. (2009). These deviations in the strength of the surface brightness, however, affect the positions of the X-ray and SZ peaks little, which guarantees the robustness of our work results obtained by only using the SPH code.

Figure 17.

Figure 17. Gas mass density distribution (panel a) and the temperature distribution (panel b) of two merging clusters at the merger plane at different merging time t = −0.4, 0.0, 0.4 Gyr. The values of the density and the temperature distributions are represented by the color scales. The two merging clusters have M1 = 2 × 1014M and ξ = 2, with initial relative velocity V = 500 km s−1 and zero impact parameter. In each panel, the top row shows the results of GADGET-2, and the bottom row is for those of FLASH. The overlaid blue and the green contours represent the X-ray and the SZ surface brightness viewed along the z-axis, respectively. For clarity, different contour levels are applied at different time. This figure shows that the positions of the X-ray and SZ peaks are affected little by using different simulation methods, which supports the use of the SPH code in a large number of the simulations in this work.

Standard image High-resolution image

Footnotes

  • Hereafter we use the unit of h−1 kpc for dc for comparison with the observation (see Section 4).

  • To distinguish from the SZ–X-ray offset d viewed along the z-axis, we use dphy to represent the offset viewed along from any one randomly given direction.

Please wait… references are loading.
10.1088/0004-637X/796/2/138