This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

ADAPTIVELY REFINED LARGE EDDY SIMULATIONS OF A GALAXY CLUSTER: TURBULENCE MODELING AND THE PHYSICS OF THE INTRACLUSTER MEDIUM

, , , and

Published 2009 November 17 © 2009. The American Astronomical Society. All rights reserved.
, , Citation A. Maier et al 2009 ApJ 707 40 DOI 10.1088/0004-637X/707/1/40

0004-637X/707/1/40

ABSTRACT

We present a numerical scheme for modeling unresolved turbulence in cosmological adaptive mesh refinement codes. As a first application, we study the evolution of turbulence in the intracluster medium (ICM) and in the core of a galaxy cluster. Simulations with and without subgrid scale (SGS) model are compared in detail. Since the flow in the ICM is subsonic, the global turbulent energy contribution at the unresolved length scales is smaller than 1% of the internal energy. We find that the production of turbulence is closely correlated with merger events occurring in the cluster environment, and its dissipation locally affects the cluster energy budget. Because of this additional source of dissipation, the core temperature is larger and the density is smaller in the presence of SGS turbulence than in the standard adiabatic run, resulting in a higher entropy core value.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Simulations of cosmological structure formation often share two important attributes. First, the ubiquitous presence of spatially localized features such as shocks, clumps, or composition discontinuities that need to be numerically resolved or at least adequately modeled; and second, moderate or large Reynolds numbers of the baryonic component indicating that fully developed, i.e., space-filling turbulence is responsible for the mixing and dissipation properties of the gas. Despite great advances in computational fluid dynamics, an accurate handling of both aspects has so far proven to be very difficult, because dedicated numerical techniques seem to be mutually incompatible.

The most powerful technique for grid-based solvers to resolve localized and anisotropic structures in a flow is adaptive mesh refinement (AMR; Berger & Oliger 1984; Berger & Colella 1989). This technique has proven to be very well suited for several astrophysical problems (Norman 2005). However, in the case of astrophysically relevant Reynolds numbers even with AMR, we cannot resolve all the relevant length scales down to the dissipative one (Schmidt et al. 2006). Even if this condition may be achieved in regions of maximum refinement, as is possibly the case in the core of galaxy clusters (where the effective viscosity is still the subject of debate), turbulence from coarser areas of the grid continuously flows into these regions without being properly accounted for.

In engineering applications as well as other fields of computational fluid dynamics, subgrid scale (SGS) models have been developed in order to mimic the influence of unresolved turbulence on the resolved scales. This technique is often referred to as Large Eddy Simulations (LES; Lesieur & Metais 1996). In astrophysics, SGS models have already been extensively used in simulations of Type Ia supernova explosions (Niemeyer & Hillebrandt 1995; Reinecke et al. 2002; Röpke & Hillebrandt 2005; Röpke et al. 2007). In this framework, Schmidt et al. (2006) presented a formulation of SGS models based on the filtering approach of Germano (1992). Other applications of SGS models in astrophysical problems have been proposed by Pope et al. (2008) and, in an approach specially designed for Rayleigh–Taylor-driven turbulence, by Scannapieco & Brüggen (2008).

In this paper, a numerical method that combines LES and AMR for the study of astrophysical turbulent flows will be presented. We will refer to this new tool as fearless (Fluid mEchanics with Adaptively Refined Large Eddy SimulationS). With the combined use of grid refinement and SGS model, fearless is very suitable for simulations of intermittent turbulent flows in clumped media.

The formation and evolution of the cosmological large-scale structure is a typical case of turbulence generation in a strongly clumped medium. The concordance model of cosmological structure formation explains the formation of clusters through a hierarchical sequence of mergers of lower mass systems (e.g., Ostriker 1993). In particular, mergers of subhalos play a fundamental role in determining the structure and dynamics of massive clusters of galaxies. Furthermore, it is known that major mergers induce temperature inhomogeneities and bulk motions with velocities of the order of 1000 km s-1 in the intracluster medium (ICM; Norman & Bryan 1999). This results in complex hydrodynamic flows where most of the kinetic energy is quickly dissipated to heat by shocks, but some part may, in principle, also excite long-lasting turbulent gas motions. Besides merger processes, it is also known that galactic motions (Bregman & David 1989; Kim 2007) and active galactic nucleus (AGN) outflows (Heinz et al. 2006; Sijacki & Springel 2006) can stir the ICM.

The problem of the turbulent state of the ICM is still controversial, both from the theoretical point of view of constraining the kinematic viscosity of the fluid (Reynolds et al. 2005; Narayan & Medvedev 2001; Jones 2008), and from the observational side, since a direct observation of turbulent emission-line broadening is beyond the reach of current X-ray observatories (Sunyaev et al. 2003; Inogamov & Sunyaev 2003; Brüggen et al. 2005; Dolag et al. 2005; Rebusco et al. 2008). Nonetheless, some indirect ways of investigating turbulence in clusters gave encouraging results (see Iapichino & Niemeyer 2008 for an overview) and call for a better theoretical understanding of the problem. A clear example for the relevance of cluster turbulence for precision cosmology is provided by recent results that demonstrate the sensitivity of hydrostatic mass estimates on assumptions about the level of turbulence (Lau et al. 2009).

In numerical simulations of merging clusters (Schindler & Mueller 1993; Roettiger et al. 1997; Ricker & Sarazin 2001; Fujita et al. 2004a, 2004b; Takizawa 2005a; Iapichino et al. 2008), it has been shown that infalling subclusters generate a laminar bulk flow but inject turbulent motions via Kelvin–Helmholtz instabilities at the interfaces between the bulk flows and the primary cluster gas. Such eddies redistribute the energy of the merger through the cluster volume and decay into turbulent velocity fields, eventually developing a turbulent cascade with a spectrum of fluctuations expected to be close to a Kolmogorov spectrum (Dolag et al. 2005). Numerical simulations focused on the role of turbulence in astrophysical flows in general, and especially for clusters, have been restricted to measuring passively statistical quantities like velocity dispersion from simulation data (e.g., Norman & Bryan 1999; Dolag et al. 2005). The active role of small-scale velocity fluctuations on the large-scale flow has not been taken into account so far.

A previous attempt of modeling turbulence in hydrodynamic simulation of cluster formation has been performed by Iapichino & Niemeyer (2008). In that work, the authors focused on better definitions of the AMR criteria for refining the computational grid where and when the flow in the ICM was turbulent (Schmidt et al. 2009; Iapichino et al. 2008). Though useful, this numerical strategy can follow only a narrow range of large length scales along the turbulent cascade, being the Kolmogorov length scale for turbulent dissipation much lower than the spatial resolution. Besides this theoretical shortcoming, also numerically it is questionable whether the mixing forced at the mesh length scale correctly represents the physics of turbulence (Mitchell et al. 2009).

These arguments motivate the application of fearless to cluster simulations as a more consistent approach. We show below that the additional degree of freedom given by the local turbulence intensity on unresolved scales has a measurable impact on the features of the ICM. In addition to the direct dynamical coupling to the resolved fluid equations, the ability to separate unresolved kinetic energy from thermal energy allows a more accurate computation of the local temperature and entropy than without the SGS model.

This work is structured as follows: in Section 2, the formalism of the SGS model and of fearless is introduced. Some numerical tests and consistency checks are presented in Section 3, and the setup of the galaxy cluster simulations is described in Section 4. The results are presented in Section 5 and discussed in Section 6, where our conclusions are drawn.

2. SGS MODEL AND fearless

2.1. Germano Decomposition

The dynamics of a compressible, viscous, self-gravitating fluid with density ρ(ri, t), momentum density ρvi(ri, t), and total energy density ρe(ri, t) at spatial position (r1, r2, r3) is given by the following set of equations:

Equation (1)
Equation (2)
Equation (3)

where p is the pressure, gi the gravitational acceleration, and σ'ij the viscous stress tensor. Note that the Einstein sum convention applies to repeated indices.

As shown by Schmidt et al. (2006), these equations can be decomposed into large-scale (resolved) and small-scale (unresolved) parts using the filter formalism proposed by Germano (1992) in terms of density-weighted quantities.4 By means of filtering, any field quantity a can be split into a smoothed part 〈a〉 and a fluctuating part a', where 〈a〉 varies only at scales greater than the prescribed filter length. We define density-weighted filtered quantities according to Favre (1969) by

Equation (4)

Following Schmidt et al. (2006), filtered equations for compressible fluid dynamics can be derived:

Equation (5)

Equation (6)

Equation (7)

where we introduced the total resolved energy $e_{\mathrm{res}}=\hat{e}_{\mathrm{int}}+\frac{1}{2}\hat{v_i}\hat{v_i}$, $\hat{e}_{\mathrm{int}}$ being the filtered internal energy, and the generalized moments that are generically defined by

Equation (8)

Equation (9)

Equation (10)

for Favre-filtered quantities a, b, c, etc. Germano interpreted the trace of $\hat{\tau }(v_i,v_j)\langle \rho \rangle$as the squared velocity fluctuation, $q^{2}:=\hat{\tau }(v_i,v_i)/\langle \rho \rangle$. The evolution of the corresponding turbulent energy, $e_{\mathrm{t}}=\frac{1}{2}q^{2}$, is given by

Equation (11)

where

Equation (12)
Equation (13)
Equation (14)
Equation (15)
Equation (16)
Equation (17)

and

Equation (18)

The explicit forms of the quantities $\mathbb {D},\lambda,\epsilon,\Gamma$, and $\hat{\tau }(v_i,v_j)$ are unknown and have to be modeled in terms of closure relations, i.e., functions of the filtered flow quantities (or their derivatives) and the turbulent energy, et. The closures for all these terms represent the SGS model.

2.2. SGS Closures

In the following, we consider a simplified set of equations to model the influence of the turbulent small-scale (SGS) motions on the numerically resolved scales ℓ ⩾ ℓΔ, neglecting the influence of the viscous stress tensor 〈σ'ij〉 (which is a very good approximation for high Reynolds numbers) and the turbulent transport of heat given by the divergence of $\hat{\tau }(v_j,e_{\mathrm{int}})$ in Equations (6) and (7). Moreover, gravitational effects on unresolved scales are neglected, i.e., we set Γ = 0 in Equation (11) for the turbulent energy. For the terms (12), (13), and (16), we adopt SGS closures that have been applied in LES of incompressible turbulence. The numerical study by Schmidt et al. (2006) demonstrated that these closures can be carried over to transonic turbulence, for which the unresolved turbulent velocity fluctuations are small compared to the speed of sound. Additionally, we utilize the pressure-dilatation model of Sarkar (1992) in order to account for moderate compressibility effects. Since we concentrate on the dynamics of the gas in the ICM, where the Mach numbers may locally approach unity, the SGS closures outlined subsequently serve as a reasonable approximation. In supersonic flow regions, on the other hand, the SGS model is deactivated in order to maintain stability (see Section 2.4).

The flux of kinetic energy from resolved scales toward SGSs, i.e., the rate of turbulent energy production, is given by the contraction of the turbulent stress tensor and the Jacobian of the resolved velocity field. Since $\hat{\tau }_{kk}=\langle \rho \rangle q^2$, we split the tensor in a symmetric trace-free part $\hat{\tau }^*_{ij}$ and a diagonal part:

Equation (19)

The model for $\hat{\tau }^*_{ij}$ is based on the turbulent viscosity hypothesis (Boussinesq 1877), which means that $\hat{\tau }^*_{ij}$ is assumed to be of the same form as the stress tensor σ'ij of a Newtonian fluid. Hence,

Equation (20)

with a turbulent dynamic viscosity ηt = 〈ρ〉νt = 〈ρ〉CνlΔq and

Equation (21)

The turbulence production term is therefore modeled as

Equation (22)

We set Cν = 0.05 (Sagaut 2006).

The SGS transport of turbulent energy (Equation (12)) is modeled by a gradient-diffusion hypothesis, stating that the non-linear term is proportional to the turbulent velocity q2 gradient (Sagaut 2006)

Equation (23)

The diffusion coefficient has been calibrated to $C_{\mathbb {D}} \approx 0.4$ by numerical experiments (Schmidt et al. 2006).

For sufficiently high Reynolds numbers, viscous energy dissipation (Equation (16)) becomes entirely an SGS effect. The simplest expression that can be built from the characteristic turbulent velocity and length scale for dissipation is

Equation (24)

For our simulations, we set Cepsilon = 0.5 (Sagaut 2006).

The effect of unresolved pressure fluctuations in compressible turbulence is described by the λ term (Equation (15)). A simple closure for subsonic turbulent flow is (Deardorff 1973)

Equation (25)

where $C_{\lambda }=-\frac{1}{5}$ (Fureby et al. 1997). Sarkar (1992) performed simulations of simple compressible flows and investigated the influence of the mean Mach number of the flow on the turbulent dissipation, epsilon, and the pressure dilatation, λ. Based on this analysis, he suggested different models for these terms, which we will describe in the following sections. These modifications have been proven to yield good results for transonic turbulence (Shyy & Krishnamurty 1997).

As a major effect of compressibility from direct numerical simulation (DNS), Sarkar (1992) identified that the growth rate of kinetic energy decreases if the initial turbulent Mach number increases. This means that the dissipation of kinetic energy (and, therefore, of the turbulent energy) increases with the turbulent Mach number Mt = q/cs, where cs is the speed of sound. Sarkar (1992) suggested to account for this effect by using

Equation (26)

with α1 = 0.5 as a model for the dissipation of turbulent energy.

Based on a decomposition of all variables of the equation for instantaneous pressure,

Equation (27)

into a mean and a fluctuating part and comparisons with DNSs of simple compressible flows, Sarkar (1992) proposed a different model for the pressure dilatation

Equation (28)

with α2 = 0.15 and α3 = 0.2 obtained from a curve fit of the model with DNS. Unfortunately, Sarkar (1992) does not specify a value for α4, so there is some confusion in the literature about it. For example, Shyy & Krishnamurty (1997) set α4 = 0 and still found the Sarkar model in good agreement with their DNS simulation. In this work, we adopt α4 = α22/2. With this choice, the effective production of turbulent energy vanishes for a turbulent Mach number, Mt = 1/α2. In the following, we will sometimes refer to the "Sarkar SGS" when the Equations (26) and (28) are used in the model.

2.3. Filtered Equations in Comoving Coordinates

Simulations with a comoving cosmological background require a formulation of the filtered fluid dynamical equations in comoving coordinates. Applying the Germano decomposition (Section 2.1) in a comoving coordinate system with spatially homogeneous scale factor a(t), we obtain5

Equation (29)

Equation (30)

Equation (31)

Equation (32)

With respect to the non-comoving equations listed in Section 2.1, the only term to additionally implement is the last one on the right-hand side of Equation (32). Furthermore, SGS closures have to be expressed in terms of the Jacobian of the velocity in comoving coordinates,

Equation (33)

In particular, the trace-free rate of strain tensor in comoving coordinates is given by

Equation (34)

2.4. Limits of the SGS Model

Numerical difficulties result from constraints on the validity of SGS closures. In particular, the turbulent viscosity hypothesis expressed by Equation (22) was devised to account for the production of turbulence by shear in moderately compressible flow. This is typically encountered in the dense, central regions of galaxy clusters. However, the surrounding low-density gas can be accelerated very quickly in the gravitational field of the cluster. Moreover, high-velocity gradients are encountered in the vicinity of shocks, which are produced by gas accretion onto filaments, sheets, and halos as well as by the merging of substructures. In order to inhibit unphysical production of turbulent energy by these mechanisms, which are not accommodated in the present formulation of the SGS model, several numerical safeguarding mechanisms have been introduced.

First of all, we implemented a simple shock detector, which identifies strong negative divergence. A cell is marked if the velocity jump corresponding to the negative divergence becomes greater than the sound speed across the cell width lΔ:

Equation (35)

In cells satisfying the above criterion, the source and transport terms of the SGS model (Equations (22) and (23)) are disabled. The turbulent energy is only advected in these cells, and no coupling to the velocity and the resolved energy is applied.

Besides the previous check, an additional constraint is imposed on the magnitude of the turbulent energy, via the turbulent Mach number. Basically, the SGS model breaks down once Mt becomes large compared to unity, therefore a threshold for this quantity is set to $M_\mathrm{{t,max}}=\sqrt{2}$. This value for the maximal turbulent Mach number is motivated by the theory of isothermal turbulence, where the effective gas pressure can be expressed as

Equation (36)

and, consequently, peff is limited to the adiabatic value $\frac{5}{3}\rho c_{\mathrm{s}}^2$. We verified that this threshold does not harm our cluster simulations, because in the hotter gas phases (T > 105 K) turbulence is largely subsonic, and the threshold is rarely reached.

A supplementary low-temperature cutoff ensures that the sound speed does not drop to excessively low values, which occur in cosmological simulations especially in the low-density voids. We set the lower limit of the temperature to Tmin = 10 K. This threshold ensures numerical stability and does not affect the baryon physics appreciably, apart from possibly making the shocks on accreting structures weaker.

2.5. Combining AMR and LES

In LES, the filtered Equations (29)–(32) are solved using an SGS model as outlined in the previous section. However, the closure relations we use and, in fact, the very concept of SGS turbulence energy only applies if the velocity fluctuations on SGSs are nearly isotropic. This limits the LES methodology to flows where all anisotropies stemming from large-scale features, like boundary conditions or external forces, can be resolved. In the fearless method, the grid resolution ℓΔ is locally adjusted by AMR in order to ensure that the anisotropic, energy-containing scales are resolved everywhere. On the other hand, it is assumed that turbulence is asymptotically isotropic on length scales comparable to or less than ℓΔ. It is very difficult to justify the latter assumption a priori, because there are no refinement criteria that would guarantee asymptotic isotropy on the smallest resolved length scales. By careful analysis of simulation results, however, one can gain confidence that AMR resolves turbulent regions appropriately.

As an infrastructure for the implementation of fearless, we chose Enzo v. 1.0 (O'Shea et al. 2005), an AMR, grid-based hybrid (hydrodynamics plus N-body) code based on the PPM solver (Colella & Woodward 1984) and especially designed for simulations of cosmological structure formation. When a grid location is flagged for refinement in Enzo, a new finer grid is created, and the cell values on the finer grid are generated by interpolating them from the coarser grid using a conservative interpolation scheme. At each time step of the coarse grid, the values from the fine grid are averaged and the values computed on the coarse grid (in the region where fine and coarse grid overlap) are replaced. However, this approach does not account for the inherent scale dependence of the turbulent energy. Assuming Kolmogorov scaling (Kolmogorov 1941; Frisch 1995), the turbulent energies at two different levels of refinement with cell size lΔ,1 and lΔ,2, respectively, are statistically related by

Equation (37)

The assumption of Kolmogorov scaling has a caveat: since grid regions that are flagged for refinement cannot be preconditioned to ensure local isotropy, the above scaling relation is an estimate that applies to regions in which turbulence is fully developed. Near pronounced inhomogeneities or in regions undergoing rapid temporal evolution deviations from Kolmogorov scaling are inevitable. For this reason, there is still room for improvement of the method.

Using the above scaling relation, we implemented a simple algorithm to adjust the turbulent energy budget when grids are refined or derefined. The following procedure is used once a grid is refined.

  • 1.  
    Interpolate the values from the coarse to the fine grid using the standard interpolation scheme from Enzo.
  • 2.  
    On the finer grid, correct the values of the velocity components, $\hat{v}_i$, and the turbulent energy, $e_{\mathrm{t}}=\frac{1}{2}q^2$, as follows
    Equation (38)
    Equation (39)
    where $\hat{e}_{\mathrm{kin}}$ is the resolved kinetic energy, rΔ is the refinement factor of the mesh, and the primed quantities are the final values on the fine grid. The resolved energy is adjusted such that the sum of resolved energy and turbulent energy remains conserved.

Apart from adjusting the energy budget, the resolved flow should feature velocity fluctuations on length scales smaller than the cutoff length of the parent grid if a refined grid is generated. To address this problem, we observe that the smallest pre-existing eddies that are inherited from the parent grid will produce new eddies of smaller size within a turnover time. Although this implies a small delay because of the higher time resolution of the refined grid, the flow will rapidly adjust itself to the new grid resolution.

For grid derefinement we reverse this procedure.

  • 1.  
    Average the values from the fine grid and replace the corresponding values on the coarse grid.
  • 2.  
    In the regions of the coarse grid covered by finer grids, correct the velocity components and the turbulent energy:
    Equation (40)
    Equation (41)
    Here, primed quantities denote the final values on the coarse grid. The resolved energy is adjusted to maintain energy conservation and a positive kinetic energy.

3. NUMERICAL TESTS

We applied two consistency tests of the SGS model in simulations of forced isotropic turbulence in a periodic box. First, energy conservation was checked in adiabatic turbulence simulations; and, second, the scaling of the turbulent energy over several levels of resolution was investigated for isothermal turbulence. To simulate driven turbulent flow, a random forcing mechanism based on the Ornstein–Uhlenbeck process was applied (Schmidt 2004). This process generates a solenoidal (i.e., divergence-free) stochastic force field, which accelerates the fluid at large length scales ll0, where l0 is the size of the computational box. The strength of the force field is characterized by a forcing Mach number, Mf. The Mach numbers of the flow becomes comparable to Mf once the forcing has been applied over a period of time that is defined by the integral time, tint = l0/Mfcs.

3.1. Energy Conservation

For global energy conservation, it turned out to be important to compute the turbulent stress term in Equation (6) indirectly as

Equation (42)

The reason behind it is that only by using this rearrangement of the terms we can ensure that we do not introduce small numerical errors which would violate the local sum,

Equation (43)

leading to a big error in global energy conservation.

As a testing case, we run a LES of driven turbulence as outlined above on a static grid of 2563 grid points and periodic boundary conditions. The adiabatic index $\gamma =\frac{5}{3}$ and Mf = 0.68.

Figure 1 shows the typical time development of the mass-weighted mean energies in our simulation including the energy ef injected into the system by random forcing. It is evident from the curve of the turbulent energy that after one integral timescale, our simulation reaches an equilibrium between production and dissipation of turbulent energy.

Figure 1.

Figure 1. Time evolution of mass-weighted energy averages in the driven turbulence simulation. The different energy components are indicated by colors.

Standard image High-resolution image

In Figure 2, we plot the time development of the relative error $\frac{\Delta e(t)}{e(0)}$ of the mean total energy, defined as the sum of the mass-weighted means of internal energy, kinetic energy, turbulent energy minus the injected energy by the forcing,

Equation (44)

where $\hat{e} = \frac{\langle \rho e \rangle }{\langle \rho \rangle }$. It demonstrates that with our basic model, the relative error in energy is comparable to the error without SGS model, and is around 1%. The energy conservation of the model using the Sarkar modifications is equally good.

Figure 2.

Figure 2. Relative error of the total energy in the driven turbulence simulation. The different lines indicate simulations run with different versions of the SGS model, or without it, as shown in the legend.

Standard image High-resolution image

It is also instructive to plot the difference between the energy contributions (internal and kinetic energy) in the simulations with and without the SGS model. These differences are shown in Figure 3. One can conclude from this figure that, at the beginning of the turbulent driving, the turbulent energy produced in our simulation with SGS model is found in the kinetic energy of the simulation without SGS model. In contrast, from t = 1.2 tint on, most of the turbulent energy can be found in the internal energy of the simulation without SGS model. Turbulent energy can therefore be interpreted as a kind of buffer which prevents the kinetic energy in our simulation to be converted instantly into thermal energy.

Figure 3.

Figure 3. Time evolution of the energy differences (red line: kinetic energy; green line: internal energy) between simulations with and without the SGS model (basic version), compared to the evolution of the turbulent energy (blue line).

Standard image High-resolution image

3.2. Scaling of Turbulent Energy

A necessary condition for the validity of the turbulent energy transfer algorithms explained in Section 2.5 is that an AMR simulation should approximately reproduce the results of static grid simulations corresponding to the different levels of refinement. To test this, we compared an AMR simulation of driven turbulence with a 323 root grid resolution and three additional levels (with a refinement factor of 2 between each level) to three static grid simulations with resolutions of 323, 643, 1283, and 2563. In order to allow for the comparison of averaged quantities, refinement of the entire domain was enforced at all levels of the AMR simulation. In order to reach a statistically stationary root-mean-square (rms) Mach number, we ran the simulations for nearly isothermal gas with γ = 1.01 (for adiabatic turbulence, after an initial rise, the rms Mach number gradually decreases with time because of the dissipative heating of the gas). We used a supersonic forcing Mach number, Mf = 2.7, to check whether a consistent turbulent energy budget could be achieved for highly compressible turbulence, albeit the scaling relations (Equation (37)) of incompressible turbulence were utilized.

The results of this consistency check can be seen in Figure 4. We observe that the time development of the mean turbulent energy is very similar on the different levels of the AMR simulation compared to the static grid simulations, except for some deviations at the root level. On the other hand, comparing these results to a simulation without correcting turbulent energy at grid refinement/derefinement (Figure 5), it is evident that the scaling of the turbulent energy in the latter case is inconsistent.

Figure 4.

Figure 4. Thick lines: mean mass-weighted turbulent energy for each level of the AMR simulation, using our procedure of transferring turbulent energy at grid refinement/derefinement. Thin lines: the corresponding evolution of turbulent energy of the static grid simulations. The colors indicate the AMR level or the static grid resolution.

Standard image High-resolution image
Figure 5.

Figure 5. Thick lines: mean mass-weighted turbulent energy for each level of the AMR simulation without transferring turbulent energy at grid refinement/derefinement. Thin lines: the corresponding evolution of turbulent energy of the static grid simulations. The colors indicate the AMR level or the static grid resolution.

Standard image High-resolution image

4. DETAILS OF THE CLUSTER SIMULATIONS

4.1. Simulation Setup

We performed simulations of cluster formation with Enzo, following Iapichino & Niemeyer (2008). We will compare two runs: one of them was done with the public version of Enzo (without SGS model), and the other with fearless implemented, in its version including the Sarkar correction (Equations (26) and (28)). The simulations were done using a flat ΛCDM background cosmology with a dark energy density ΩΛ = 0.7, a total (including baryonic and dark matter) matter density Ωm = 0.3, a baryonic matter density Ωb = 0.04, the Hubble parameter set to h = 0.7, the mass fluctuation amplitude σ8 = 0.9, and the scalar spectral index n = 1. Both simulations were started with the same initial conditions at redshift zini = 60, using the Eisenstein & Hu (1999) transfer function, and evolved to z = 0. The simulations are adiabatic with a heat capacity ratio γ = 5/3 assuming a fully ionized gas with a mean molecular weight mμ = 0.6 u. Cooling physics, magnetic fields, feedback, and transport processes are neglected.

The simulation box has a comoving size of 128 Mpc h−1. It is resolved with a root grid (level l = 0) of 1283 cells and 1283 N-body particles. A static child grid (l = 1) is nested inside the root grid with a size of 64 Mpc h−1, 1283 cells and 1283 N-body particles. The mass of each particle in this grid is 9 × 109 M h−1. Inside this grid, in a volume of 38.4 Mpc h−1, adaptive grid refinement from level l = 2 to l = 7 is enabled using an overdensity refinement criterion as described in Iapichino & Niemeyer (2008) with an overdensity factor f = 4.0. The refinement factor between two levels was set to rΔ = 2, allowing for an effective resolution of 7.8 kpc h−1.

The static and dynamically refined grids were nested around the place of formation of a galaxy cluster, identified using the HOP algorithm (Eisenstein & Hut 1998). Since the realization of the initial conditions was chosen identical to Iapichino & Niemeyer (2008), this study is based on the same cluster analyzed in that work. The cluster has a virial mass of Mvir = 5.95 × 1014 M h−1 and a virial radius of Rvir = 1.37 Mpc h−1.

4.2. Local Kolmogorov Scaling

In static grid simulations, one often chooses to use the grid resolution lΔ as characteristic length scale to compute a characteristic velocity or eddy turnover time for this scale. However, in an AMR code it is not trivial to compute the turbulent velocity ql associated with a characteristic length scale l = lΔ, since lΔ varies in time and space. To circumvent this difficulty, we assume that below the grid resolution, turbulent velocity locally scales according to Kolmogorov

Equation (45)

We thereby assume that locally a Kolmogorov-like energy cascade sets in, at a length scale given by the resolution of the grid at that position. This local hypothesis holds here only for the analysis of our simulations, and is similar to the assumption done in Section 2.5 for managing the grid refinement and derefinement.

As a characteristic scale of our analysis, we choose the length scale of our highest resolved regions, which is lmin = 7.8 kpc h−1. The turbulent velocity in the most finely resolved regions can be computed directly from the values of the turbulent energy $q(l) = \sqrt{2 e_{\mathrm{t}}}$ on the grid; the turbulent velocity in less finely refined regions is scaled down according to our local Kolmogorov hypothesis as

Equation (46)

5. RESULTS

5.1. Turbulent Energy Scaling in the Cluster Simulation

In Section 3.2, we studied the temporal evolution of the turbulent energy at different resolutions in a simulation of driven turbulence. In this section, we repeat this analysis for our fearless cluster simulation. Figure 6 shows the evolution of the mass-weighted mean turbulent energy for every level of our AMR simulation. We see from the plot that the turbulent energy on the higher AMR levels l (meaning at smaller scales) is higher at early times (2 Gyr < t < 6 Gyr). Later this picture changes, but not completely. For example, the turbulent energy at l = 4 stays above the turbulent energy at l = 3 throughout the simulation. The magnitude of et along the AMR levels suggests to locate the turbulence injection length scale between 125 and 250 kpc h−1, corresponding to the effective resolutions of levels 3 and 4. This is only a rough qualitative estimate, but nevertheless in agreement with theoretical expectations (e.g., Subramanian et al. 2006).

Figure 6.

Figure 6. Time evolution of the mean turbulent energy for each level of refinement. This analysis has been performed on a test run identical to that described in Section 4.1 with only six AMR levels.

Standard image High-resolution image

Particularly noteworthy are the turbulent energy fluctuations on smaller scales at the time 2 Gyr < t < 6 Gyr, corresponding to a redshift z = 3–1. We can interpret these large fluctuations as evidence for violent major mergers that happen at that time, producing turbulent energy which is then dissipated into internal energy, heating up the cluster gas. However, at t > 12 Gyr, the simulation seems to globally reach some kind of stable state, comparable to what has been found in the driven turbulence simulations.

5.2. Spatial Distribution of Turbulent Energy

Before performing a quantitative analysis of the cluster properties in the fearless run, it is useful to visually inspect the generation and the spatial distribution of the turbulent energy in the ICM and around the cluster, in order to compare the simulations with the theoretical expectations of cluster mergers. Figures 7 and 8 present a time series of density and turbulent velocity slices, where several merger events in the cluster outskirts can be identified.

Figure 7.

Figure 7. Slices of baryon density (left-hand panels, a and c) and turbulent velocity $q = \sqrt{2 e_{\mathrm{t}}}$ scaled to 7.8 kpc h−1 (right-hand panels, b and d) at different redshifts z, for the fearless run. The density is logarithmically color coded as overdensity with respect to the average baryon density in the colorbar on the left of panel a, whereas q is linearly coded in km s-1, according to the colorbar on the left of panel b. The overlayed contours show density. The slices show a region of 6.4 × 6.4 Mpc h−1 around the center of the main cluster followed in the simulation. Panels a and b refer to z = 0.15, and panels c and d to z = 0.1.

Standard image High-resolution image
Figure 8.

Figure 8. Same as Figure 7, but in this case panels a and b refer to z = 0.05, and panels c and d to z = 0. The black rectangle in panel a denotes the projection on the slice of a small volume, including a subclump and its wake, analyzed in Table 3.

Standard image High-resolution image

In the density slice at redshift z = 0.15 (Figure 7(a)), we can see a filament extending from the lower left to the upper right corner of the figure; material is falling onto the cluster along this structure. On both sides, the inflow of relatively cold gas from the filament onto the ICM produces a moderate increase of turbulent energy (cf. Nagai & Kravtsov 2003). From the upper right side there is not only a smooth inflow of matter, but two small clumps are approaching the cluster. During the simulation, these two clumps merge with the main cluster (Figures 7(c)–8(c)) and one of them (on the left) is assimilated completely at redshift z = 0. A substructure approaches the cluster from the lower left corner along the filament (Figure 8(a)) and another one is visible only at z = 0.05 just to the right of the cluster core, when it crosses the slicing plane.

The merging process can be followed much more easily in terms of production of turbulent energy, visualized by the turbulent velocity $q = \sqrt{2 e_{\mathrm{t}}}$. In Figure 7(b) at z = 0.15, a marked peak of turbulent energy in the center of our cluster, resulting from a former massive merger, can be seen. The turbulent energy produced by this merger declines (Figures 7(d)–8(d)), and at z = 0 it is dissipated into internal energy nearly completely, as confirmed by our further analysis in Section 5.4.

The two approaching clumps described above continue to drive turbulence in the cluster. Thereby, the left clump can be identified in the turbulent velocity slice at z = 0.15 (Figure 7(b)) as a ring-like structure, showing that turbulence is not produced at the center of the infalling clump but at the front (behind a bow shock) and in the wake of the infalling material. The right clump only shows some turbulence production in its wake, which might be due to its smaller size and smaller velocity. On their way toward the main cluster and through its ICM, both clumps show a relevant production of turbulent energy (Figures 7(d)–8(b)). The considerable amount of turbulent energy can even be identified after the two clumps have merged with the main cluster (Figure 8(d)) and the left one is not easily visible in the corresponding density slice (Figure 8(c)). From this point of view, the distribution of turbulent energy traces the local merging history of a galaxy cluster until it is dissipated into internal energy completely.

The morphological evolution of the cluster gives a clear sense of the markedly local behavior of the production and dissipation of turbulence, which is confirmed to be an intermittent process in the ICM.

5.3. Cluster Energy Budget

It is extremely difficult to apply an energy analysis similar to that performed in Section 3.1 to a galaxy cluster. Different than a periodic box, a galaxy cluster is an open system, with a growth over time of negative gravitational potential energy. Nevertheless, a comparison of the energy contributions of the two simulations at z = 0 is useful to understand the role of the SGS model.

The results are summarized in Table 1, where the energies in a sphere centered at the cluster center and with r = Rvir are reported. Different than elsewhere in this work, in Table 1 the energies are not specific, i.e., Etot = ρetot, etc. This choice allows a better evaluation of the energy budget but it does not differ appreciably from the analysis of specific energies, since the baryon masses in the two runs agree within 1%.

Table 1. Energy Contributions in a Sphere of r = Rvir, Centered at the Cluster Core, at z = 0

Quantity Adiabatic Run fearless Run
Etot(1063 erg) 2.6458 2.6426 (−0.1%)
Eint(1063 erg) 2.1982 2.2082 (+0.5%)
Ekin(1062 erg) 4.476 4.168 (−6.9%)
Et(1061 erg)  ⋅⋅⋅ 1.762

Notes. The total energy Etot is defined as the sum of Eint, Ekin and, in the fearless run, Et. The turbulent energy reported here is not scaled as described in Section 5.1.

Download table as:  ASCIITypeset image

The total energy remains basically unaltered in the two simulations, whereas the most important change is the decrease of Ekin in the fearless run. The missing kinetic energy is transferred mostly to Et, which acts as a buffer between the resolved kinetic energy and Eint. The SGS model transfers energy from Et to Eint either adiabatically (via the pressure dilatation term, Equation (15)) or irreversibly (via the dissipative term, Equation (16)). Turbulent dissipation is thus added to the dominant numerical dissipation, resulting in a moderate increase of Eint, though it is less relevant than the variation of Ekin.

In both runs, the kinetic energy contribution in the cluster Ekin is smaller than Eint. The mass-weighted average of the Mach number in the cluster is about 0.6, in agreement with the known fact that the flow in the ICM is, on average, mildly subsonic. In this regime, it is not surprising that the subgrid energy, Et, is about 2 orders of magnitude smaller than Eint on the global level. The energy contribution from unresolved scales is globally negligible, though locally turbulence can play a more significant role, as will be discussed in Section 5.4.

Finally, we note that the ratio of the turbulent production term Σ (Equation (13)) to the turbulent dissipation term epsilon (Equation (16)) in the cluster core is

Equation (47)

suggesting that the turbulent flow in the ICM is globally in a regime of near equilibrium of production and dissipation of turbulent energy.

5.4. Radial Profiles and Local Analyses

The results of Section 5.3, referring to global features of the galaxy cluster, will be complemented by a local comparison in terms of radial profiles of selected physical quantities and an analysis of the cluster core in this section.

As a first interesting result of the comparison between radial profiles in the fearless simulation and in the standard adiabatic run, one can notice (Figure 9) that the temperature profile of the fearless run deviates slightly from that of the adiabatic run. This is especially apparent at the center, where T is larger in the cluster core for r ≲ 0.07 Rvir, with respect to the standard run. Consequently (Figure 10), the core in the fearless run is less dense, so that the ICM remains in hydrostatic equilibrium. The local energy budget in the cluster core is therefore modified by the SGS model.

Figure 9.

Figure 9. Radial profiles of mass-weighted temperature at z = 0. The dotted line refers to the simulation without SGS model, whereas the solid line is for the fearless simulation (in its version including the Sarkar corrections).

Standard image High-resolution image
Figure 10.

Figure 10. Same as Figure 9, but showing the baryon density.

Standard image High-resolution image

In Table 2, we explore this feature in detail, reporting the mass-weighted averages of selected variables in a sphere within 0.07 Rvir from the cluster center. Because of the adjustment of the cluster hydrostatic equilibrium, the mass enclosed in this sphere is significantly different in the two runs (it decreases by 10% in the fearless run), thus it is more convenient to present specific energies in Table 2.

Table 2. Mass-weighted Averages in a Sphere of r = 0.07 Rvir, Centered at the Cluster Core, at z = 0

Quantity Adiabatic Run fearless Run
Σ/epsilon  ⋅⋅⋅ 0.59
etot(1016 cm2 s-2) 1.3781 1.4189
eint(1016 cm2 s-2) 1.2529 1.3030
ekin(1015 cm2 s-2) 1.252 1.138
et(1013 cm2 s-2)  ⋅⋅⋅ 2.078
pres/(pres + ptherm)(%) 1.46 1.52
vrms(km s-1) 196 204

Download table as:  ASCIITypeset image

First, the low value of the Σ/epsilon ratio indicates that, at z = 0, in the cluster core region the dissipation of turbulence is dominant with respect to its production. This confirms the result of the morphological analysis in Section 5.2 that turbulence is not produced locally in the core by mergers at z < 0.15, but that it decays in this region. The impact of this turbulent dissipation on the local energy budget of the cluster core can seen from the comparison of the energy contributions in Table 2. Similar to the global analysis in Table 1, there is a clear decrease of ekin, transferred both to et and eint. Both etot and eint are higher in the fearless run, pointing to the existence of an energy flux from the resolved scales to the thermal reservoir through the turbulent buffer, leading to the increase of the internal energy. We interpret this additional energy contribution as caused by the turbulent dissipation introduced by the SGS model.

In the cluster core, the energy content at the SGSs is marginal. Apparently the relative contribution to the total energy is even smaller than in Table 1, but one should notice that, for consistency, in that table both Ekin and Et are reported according to the original scale separation introduced by the AMR resolution, and without rescaling Et as described in Section 5.1. In the cluster core, the refinement level is maximum, therefore the unresolved part of the turbulent cascade is relatively smaller than elsewhere, and so is et. In Table 2, we use the scaled definition of et; but in the core, it differs from the unscaled one only marginally, because in this region the resolution is lmin almost everywhere.

To further quantitatively appreciate the contribution of et to the energy budget, Figure 11 reports the profile of the turbulent pressure support pt/(pt + ptherm) in the cluster, where the turbulent pressure is defined as pt = 1/3 ρq2, and ptherm is the usual thermodynamical pressure. This ratio is also equal to the ratio of the corresponding energies (et/(et + eint)). At the length scale of the effective spatial resolution of the simulations lmin = 7.8 kpc h−1, the contribution of the turbulent pressure (or energy) is well below 1%, although it increases at larger central distances.

Figure 11.

Figure 11. Radial profile of the turbulent contribution to the pressure support pt/(pt + ptherm), as defined in the text, for the fearless simulation at z = 0.

Standard image High-resolution image

In analogy with the turbulent pressure, we define a "resolved pressure"

Equation (48)

where the rms velocity (Iapichino & Niemeyer 2008) is defined as

Equation (49)

Here, 〈v〉 is the mass-weighted average of the velocity in the analysis volume. This quantity essentially probes the contribution of turbulent motions at length scales of the order of 0.07 Rvir ∼ 90 kpc h−1. As shown in Table 2, the pressure contribution at these length scales is at the percent level, and is slightly higher in the fearless simulation. Interestingly, the rms velocity in the fearless run is also somewhat larger than in the adiabatic case.

The changes in the temperature and density profiles are also reflected on the entropy which is defined, as is customary in astrophysics, as

Equation (50)

with γ = 5/3. The entropy in the cluster core is higher in the fearless run as compared to the standard run (Figure 12). This result is consistent both with the locally increased dissipation of turbulent to internal energy provided by the SGS model and with the higher degree of mixing induced in the cluster core, shown by vrms in Table 2.

Figure 12.

Figure 12. Same as Figure 9, but showing the mass-weighted entropy (as defined in the text).

Standard image High-resolution image

The radial distribution of turbulent energy is displayed in Figure 13 in terms of the turbulent velocity scaled to lmin. The turbulent velocity at this length scale is below 100 km s-1. There is a pronounced peak at r = 0.6 Rvir, which is correlated with analogous trends in the turbulent pressure (Figure 11) and in the radial velocity profile (Figure 14). This structure is clearly linked with the most prominent merging clump shown in Figure 8(d), and analyzed below in Table 3.

Figure 13.

Figure 13. Radial profile of the turbulent energy scaled at the length scale lmin, as described in Section 4.2, for the fearless simulation at z = 0.

Standard image High-resolution image
Figure 14.

Figure 14. Same as Figure 9, but showing the mass-weighted baryon radial velocity.

Standard image High-resolution image

Table 3. Mass-weighted Averages in a Volume of (512 × 768 × 1280) kpc h−1, Containing a Subclump and its Wake, at z = 0.05

Quantity Adiabatic Run fearless Run
Σ/epsilon  ⋅⋅⋅ 1.13
etot(1016 cm2 s-2) 1.7447 1.5746
eint(1015 cm2 s-2) 5.4281 5.9607
ekin(1016 cm2 s-2) 1.2032 0.9786
et(1014 cm2 s-2)  ⋅⋅⋅ 2.290
pt/(pt + ptherm)(%)  ⋅⋅⋅ 3.70

Download table as:  ASCIITypeset image

There is an appealing similarity between the intervals of radii where q and the pressure ratio are larger (r < 0.1 Rvir and 0.4 Rvir < r < 0.8 Rvir), and the corresponding intervals where the temperature and entropy of the fearless run are slightly larger than those computed for the adiabatic run. The opposite trend occurs in the interval in between, where vturb is comparatively smaller. The effects are very small, but suggest that the SGS model plays the same role in the ICM that was shown above for the cluster core, and in Section 5.3 for the global quantities. In case of radial profiles, the spherical averaging combined with the intermittent behavior of turbulence tends to mask the turbulent effects. This can be better understood with a comparison of the values of q in Figure 13 and in the right-hand panels of Figures 7 and 8: the peak values in the slices are much larger than the spherical averages in the profile.

The idea that, locally, the turbulence and its modeling can play a sizeable role is further corroborated by the data in Table 3, reporting the analysis at z = 0.05 of a small volume (512 × 768 × 1280 kpc h−1) that contain one of the clumps presented in Section 5.2 and its wake (cf. Figure 8(a)). The morphology of this accreting subcluster in the fearless run is not substantially different from the adiabatic one. The energy content, however, is rather different from that in the cluster core: in the region under consideration ekin is dominant with respect to eint. The importance of the turbulence, injected by the hydrodynamic instabilities in the wake of the moving clump, is testified by the large ratio Σ/epsilon and by the turbulent pressure support, which is at the level of some percent, about 1 order of magnitude larger than the spherical averages in Figure 11. Despite of the slightly smaller average of etot in the fearless run with respect to the adiabatic one, one can see an increase of eint, mostly at the expenses of ekin, resulting from the turbulent dissipation. The decrease of ekin is rather large (19%) but can be partially ascribed to the difficulty of comparing energy budgets in such open volumes.

It is important to stress the deep difference between the turbulent velocity profile in Figure 13 and the profile of vrms, defined by Equation (49) (see also Norman & Bryan 1999; Iapichino & Niemeyer 2008). In the former case, the mass-weighted average of a local quantity (i.e., defined in every cell) is computed for each spherical shell, whereas in the latter case vrms is interpreted shell-wise as the standard deviation with respect to the average 〈v〉. Clearly, the latter definition does not retain any information related to a length scale, and can be interpreted as turbulent velocity only in a loose sense. From this point of view, the turbulent velocity provided by the SGS model is a more powerful probe of the features of a turbulent flow. On the other hand, spherically averaged velocity dispersions (and the derived turbulent pressure) are meaningful in comparison with observations, for example, in the procedure for estimating the cluster mass (cf. Rasia et al. 2006; Piffaretti & Valdarnini 2008). According to this different definition, the spherically averaged turbulent pressure of the simulated cluster (in a run similar to the adiabatic one presented here) is reported in Iapichino & Niemeyer (2008). It reaches values around 10%, in agreement with the values found recently in simulations by Lau et al. (2009) for relaxed clusters. The turbulent pressure is somewhat smaller in the fearless run because of its slightly reduced content in kinetic energy, but the difference is small, and is not expected to significantly affect the estimates of the cluster mass.

6. DISCUSSION AND CONCLUSIONS

LES are based on the notion of filtering the fluid dynamic equations at a specific length scale, thus performing a scale separation between the resolved and the unresolved flow. The latter is treated by means of a SGS model, which in turn is coupled to the hydrodynamic equations governing the former. In principle, a single scale separation is incompatible with AMR codes, often used to study astrophysical phenomena.

One of the aims of the present work was to address this numerical problem by means of developing, implementing, and applying a new numerical scheme that uses AMR and LES in combination, which we called fearless. This novel tool is suitable for modeling turbulent flows over a wide range of length scales, a key feature in the treatment of many astrophysical flows including the ICM.

We showed that the idea of our approach to correct the velocity and the kinetic energy at grid refinement/derefinement, according to local Kolmogorov scaling, produces consistent results in simulations of driven turbulence. We demonstrated that energy conservation and the scaling of turbulent energy in our adaptive simulations is consistent with static grid simulations.

To our knowledge, this work shows the first application of an SGS model to simulations of the formation and evolution of a galaxy cluster. The results give rise to several interesting implications with regard to the physics of galaxy clusters and to the numerical methods employed for their exploration in computational cosmology.

The production of turbulence induced by minor mergers, analytically studied by Subramanian et al. (2006) and addressed by several numerical investigations (Heinz et al. 2003; Takizawa 2005a, 2005b; Asai et al. 2007; Dursi & Pfrommer 2008; Iapichino et al. 2008), is accurately tracked by the newly defined turbulent subgrid energy (Figures 7 and 8), although the level of resolution of the idealized setups cannot be reached by cosmological simulations. The visualization and subsequent analysis and postprocessing of turbulence and related quantities is therefore easier and more consistent. Turbulence in the ICM appears to be subsonic, in agreement with previous results. The average ratio between the dissipation and the production term Σ/epsilon in the SGS model is close to unity, namely typical of a system where the turbulence is roughly stationary. On the other hand, this ratio is locally variable (cf. Table 2) and, together with the intermittent nature of turbulence in the ICM (Section 5.2; see also Iapichino & Niemeyer 2008), delivers the picture of a flow where turbulent motions are randomly initiated by merger events and then gradually decay (Frisch 1995; Subramanian et al. 2006). We also notice that, in simulations of driven turbulence in a periodic box (Figure 3), the decrease of ekin (noticed in our cluster simulation) is linked to an increase of et only in the early driving phase, not in the later equilibrium stage.

The morphological evolution of the minor merger events and the subsequent injection of turbulence in the ICM (Figures 7 and 8) appear to be rather localized and intermittent, confirming the feature of turbulent flows as being not very volume-filling (Subramanian et al. 2006; Iapichino & Niemeyer 2008). The dissipation of turbulent to internal energy is thus modeled as a markedly local process, consistent with the theoretical expectations.

The effect of the SGS model on the cluster energy budget is well exemplified by the comparison of our simulations at z = 0 (Table 1). Although the value of et is small compared to eint, this energy buffer is locally effective in transferring the kinetic energy to the thermal component. The dissipative effects are therefore more relevant in those locations where et is relatively large, like the cluster core (Figure 13). In general, the main contribution of the fearless approach is to add a more physically motivated contribution to the energy dissipation, which in Eulerian codes is otherwise purely numerical. In fearless, part of the energy flux from resolved scales to the thermal reservoir is retained in the buffer turbulent energy, et, and is further dissipated (turbulent dissipation) according to a local and temporal evolution determined by the SGS model.

Besides local effects, the importance of the SGS model for the overall cluster structure appears small, because of the modest subgrid energy contribution (Figure 11). One remark about the simulated cluster is important at this point: as also verified in Iapichino & Niemeyer (2008), this structure is very relaxed (see also Figure 14). Simulations of more perturbed structures with recent or ongoing major mergers are in preparation (S. Paul et al. 2009, in preparation), because they will help to clarify the role of the turbulent energy (and of its modeling) in the cluster energy budget in cases where its magnitude is larger. From this viewpoint, the radial increase of turbulent pressure support in the cluster outskirt (Figure 11) is interesting for physical mechanisms (like the acceleration of cosmic rays and magnetic field amplification) where the knowledge of the turbulent state of the flow is needed.

More turbulence in the cluster core is required, for example, to reproduce the iron abundance profile in cool core clusters. Following Dennis & Chandran (2005), a turbulent diffusion coefficient can be defined as Dturb ∼ 0.1 q l, where q is the turbulent velocity at the length scale l. Using l = lmin and q ∼ 60 km s-1, we find Dturb ≃ 2 × 1028 cm2 s-1. We notice that this value is smaller than the estimates of the effective diffusion coefficient in the cluster models of Rebusco et al. (2005) and Rebusco et al. (2006), which aim to reproduce the turbulent diffusion of metals in the cores of selected clusters. In particular, the cited models require much larger turbulent velocities. In the framework of our cluster simulation, these velocities could be injected into the ICM by a vigorous merger event. Another possibility, explicitly suggested by the authors cited above, is to invoke the action of an AGN outflow as an additional stirring agent in the cluster core.

The enhanced temperature profile in the fearless run is somehow reminiscent of the theoretical predictions about the role of turbulent heating in cluster cores (Dennis & Chandran 2005). We notice an apparent misunderstanding in the literature regarding this point. In our model (and in the theory of turbulent flows in general), the dissipation of turbulent energy does not act as an additional energy source but simply releases the energy arising from the virialization process on a longer timescale than the quick shock heating. Nevertheless, we showed that turbulence, and the turbulent dissipation as well, can be rather localized. Naively, one could think that an effective turbulent heating in cool cores would require a peak of turbulent energy in the cluster core, whose existence and magnitude should be justified theoretically. Again, the stirring induced by AGN activity is an open possibility which deserves further investigation. However, the model of Dennis & Chandran (2005) includes radiative cooling and thermal conduction, and a detailed comparison is beyond the scope of the present work. We observe that additional physics which is here not addressed (thermal conduction and magnetic fields) could bring further interesting implications for the energy budget in the ICM and the turbulent mixing (Sharma et al. 2009, and references therein).

Consequent to both the enhanced dissipation and fluid mixing is the larger value of entropy in the fearless cluster core. A long-standing problem in cluster simulations is the shape of the entropy profile, which smoothly decreases in the center in smoothed particle hydrodynamics (SPH) simulations whereas it flattens inside the core in runs with grid-based codes (Frenk et al. 1999). This issue has been debated recently in several works (among others, Dolag et al. 2005; Wadsley et al. 2008; Kawata et al. 2009), because it is controversial which difference of SPH and mesh-based codes it results from. It has been claimed that the source of discrepancy probably lies in the treatment of fluid mixing (Mitchell et al. 2009): the weaknesses of SPH in this regard are known, but the ability of mesh codes to model the turbulent cascade on length scales comparable with the grid resolution has not been addressed in a satisfactory way. It is therefore unclear whether the flat core entropy in grid codes correctly represents the physics of the ICM, or perhaps numerical effects harm the robustness of this feature. Recently, Springel (2009) pointed out that the core temperature and entropy in grid-based codes are affected by a spurious increase, caused by the N-body noise in the gravitational force field. In our opinion, the higher entropy core value in the fearless run suggests that the typical flat entropy core is a hydrodynamic feature which requires a better understanding of the numerics in mesh codes, and is at least not primarily caused by N-body noise.

The SGS model applied in this work has to be considered as an intermediate solution to address some basic questions related to dynamics of the turbulent ICM. A more elaborate model that is able to handle the complexity of the flow (wide range of Mach numbers and large density gradients as well as pronounced inhomogeneities) in simulations of large-scale structure evolution is under development (W. Schmidt & C. Federrath 2009, in preparation). This first application shows the promising perspectives for the use of an SGS model in combination with AMR and its potential impact on many branches of numerical astrophysics.

L.I. acknowledges useful discussions with V. Antonuccio-Delogu, M. Bartelmann, S. Borgani, G. Murante, and M. Viel. The simulations described in this work were performed using the Enzo code, developed by the Laboratory for Computational Astrophysics at the University of California in San Diego (http://lca.ucsd.edu). We thank the Enzo development team, in particular D. Collins, for invaluable support in improving our modifications to the code. The numerical simulations were carried out on the SGI Altix 4700 HLRB2 of the Leibniz Computing Centre in Munich (Germany). The research of J.C.N. and L.I. was partly supported by the Alfried Krupp Prize for Young University Teachers of the Alfried Krupp von Bohlen und Halbach Foundation.

Footnotes

  • For a review, see Röpke & Schmidt (2009).

  • The comoving density $\tilde{\rho } = a^3 \rho$ and the comoving pressure $\tilde{p} = a^3 p$ are introduced to shorten the equations.

Please wait… references are loading.
10.1088/0004-637X/707/1/40