0712.0824/ms.tex
1: %\documentclass[12pt,preprint]{aastex}
2: \documentclass[numberedappendix]{emulateapj}
3: \usepackage{apjfonts}
4: \bibliographystyle{apj}
5: 
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: 
8: \newcommand{\be}{\begin{equation}}
9: \newcommand{\ee}{\end{equation}}
10: \newcommand{\bea}{\begin{eqnarray}}
11: \newcommand{\eea}{\end{eqnarray}}
12: 
13: % Units
14: \newcommand{\hpc}{h^{-1}\mathrm{pc}}
15: \newcommand{\hMsun}{\ h^{-1}\mathrm{M}_{\odot}}
16: \newcommand{\hMpc}{\ h^{-1}\mathrm{Mpc}}
17: \newcommand{\Msun}{M_{\odot}}
18: \newcommand{\kms}{{\,{\mathrm{km}}\,{\mathrm{s}}^{-1}}}
19: \newcommand{\kpc}{{\,{\mathrm{kpc}}}}
20: \newcommand{\Gyr}{{\,{\mathrm{Gyr}}}}
21: 
22: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
23: 
24: \shortauthors{CONROY AND OSTRIKER}
25: \shorttitle{Thermal Balance in the ICM}
26: 
27: \begin{document}
28: 
29: %----------------------------------------------------------------
30: \title{Thermal Balance in the Intracluster Medium: Is AGN Feedback
31:   Necessary?}
32: %----------------------------------------------------------------
33: 
34: \author{Charlie Conroy \& Jeremiah P. Ostriker} 
35: \affil{Department of Astrophysical Sciences, Princeton University,
36:   Princeton, NJ 08544}
37: 
38: \slugcomment{Submitted to ApJ, 6 Dec 2007}
39: 
40: \begin{abstract}
41: 
42:   A variety of physical heating mechanisms are combined with radiative
43:   cooling to explore, via one dimensional hydrodynamic simulations,
44:   the expected thermal properties of the intracluster medium (ICM) in
45:   the context of the cooling flow problem.  Energy injection from type
46:   Ia supernovae, thermal conduction, and dynamical friction (DF) from
47:   orbiting satellite galaxies are considered as potential heating
48:   mechanisms of the central regions of the ICM, both separately and in
49:   conjunction.  The novel feature of this work is the exploration of a
50:   wide range of efficiencies of each heating process.  While DF and
51:   conduction can provide a substantial amount of energy, neither
52:   mechanism operating alone can produce nor maintain an ICM in thermal
53:   balance over cosmological timescales, in stark contrast with
54:   observations.  For simulated clusters with initially isothermal
55:   temperature profiles, both mechanisms acting \emph{in combination}
56:   result in long-term thermal balance for a range of ICM temperatures
57:   and for central electron densities less than $n_e\sim 0.02$
58:   cm$^{-3}$; at greater densities catastrophic cooling invariably
59:   occurs.  Furthermore, these heating mechanisms can neither produce
60:   nor maintain clusters with a declining temperature profile in the
61:   central regions, implying that the observed ``cooling-core''
62:   clusters, which have such declining temperature profiles, cannot be
63:   maintained with these mechanisms alone.  Supernovae heating also
64:   fails to maintain clusters in thermal balance for cosmological
65:   timescales since such heating is largely unresponsive to the
66:   properties of the ICM.  Thus, while there appears to be an abundant
67:   supply of energy capable of heating the ICM in clusters, it is
68:   extremely difficult for the energy deposition to occur in such a way
69:   that the ICM remains in thermal {\it balance} over cosmological
70:   time-scales.  For intracluster media that are not in thermal
71:   balance, the addition of a small amount of relativistic pressure
72:   (provided by e.g. cosmic rays) forestalls neither catastrophic
73:   heating nor cooling.  These conclusions are driven largely by the
74:   fact that 1) DF heating scales approximately as the gas density,
75:   while cooling scales as gas density squared, and thus DF heating
76:   cannot generically balance cooling without fine-tuning; 2)
77:   conduction acts to erase temperature gradients, while most observed
78:   clusters in fact show strong gradients in the inner regions.  These
79:   results strongly suggest that a more dynamic heating process such as
80:   feedback from a central black hole is required to generate the
81:   properties of observed intracluster media.
82: 
83: 
84: \end{abstract}
85: 
86: \keywords{cooling flows --- galaxies: clusters --- hydrodynamics ---
87:   conduction}
88: 
89: %---------------------------------------------------
90: \section{Introduction}
91: \label{section:intro}
92: 
93: Groups and clusters of galaxies are filled with hot plasma in at least
94: approximate pressure equilibrium with the gravitational potential of
95: their dark matter halo.  Observations have demonstrated for decades
96: that the cooling time of central regions of this intracluster medium
97: (ICM) in most ($>70$\%) clusters is shorter than a Hubble time
98: \citep[e.g.][]{Edge92, Peres98, Sanderson06, Vikhlinin06a}; indeed, it
99: is often as short as $\sim 0.1-1$ Gyr.  In the absence of heating, the
100: ICM will thus cool and flow into the central galaxy at the prodigious
101: rate of hundreds of solar masses per year \citep{Cowie77, Fabian77a,
102:   Mathews78}; see \citet{Fabian94} for a review.
103: 
104: Since this (catastrophically) cooling gas has never been observed
105: \citep{Peterson01, Peterson03, Tamura01}, it is now generally supposed
106: that there exists one or more heating mechanisms that maintains the
107: ICM in overall thermal balance.  The tension created by the fact that
108: the ICM is not observed to be significantly cooling, despite the short
109: cooling times near the cluster center, has become known as the cooling
110: flow problem.
111: 
112: Many possible heating mechanisms have been investigated, including
113: thermal conduction \citep[e.g.][]{Binney81, Tucker83, Voigt02,
114:   Fabian02, Zakamska03, Kim03a, Voigt04, Dolag04, Pope06} which
115: carries heat from the abundant thermal reservoir of the cluster gas to
116: the cooling inner parts, gravitational heating, including the heating
117: due to the orbital motions of galaxies, i.e. dynamical friction
118: \citep[DF; e.g.][]{Miller86, Just90, Fabian03, ElZant04b, Kim05,
119:   Dekel07}, turbulent mixing \citep[e.g.][]{Deiss96, Kim03b, Voigt04,
120:   Dennis05}, energy injection from active galactic nuclei (AGN) via
121: bubbles, sound waves, etc. \citep{Binney95, Nusser06, Binney07,
122:   Ruszkowski04a, Ruszkowski04b, Fujita05, Mathews06, Ciotti01, Voit05,
123:   Ciotti07}, and combinations of conduction and AGN
124: \citep{Ruszkowski02, Brighenti03, Fujita05} or conduction and
125: cosmic-rays \citep[CRs;][]{Guo07}, or AGN and preheating
126: \citep{McCarthy07c}.  One of the most comprehensive studies to date
127: was undertaken by \citet{Brighenti02} who included many of the heating
128: mechanisms mentioned above and found that no combination generically
129: reproduced observations.  A stringent constraint on the potential
130: heating mechanism is that it must not only supply of order the
131: necessary energy but must also supply it in a way that approximately
132: balances cooling (locally) so that the ICM does not either heat up or
133: cool on cosmological time-scales.
134: 
135: Many of these mechanisms have been idealized, and it is unclear how
136: effective they would be in observed clusters.  For example, though
137: difficult to constrain observationally, it has been suggested that the
138: conductivity in clusters may be much smaller than is required to stem
139: the cooling flows \citep{Markevitch03, Xiang07}.  In addition, while
140: DF heating may be attractive, since there are a plethora of satellite
141: galaxies in clusters, it has yet to be identified as important in
142: hydrodynamical simulations \citep{Faltenbacher05}, although this may
143: be due to insufficient resolution \citep{Naab07}.  Finally, bubbles
144: from activity in central galaxies may be important
145: \citep[e.g.][]{Churazov01, Ruszkowski02, Bruggen02}, but the mechanism
146: by which they transfer their energy to the cluster gas remains
147: obscure.
148: 
149: Even if one or more of these mechanisms could supply the requisite
150: energy (i.e. result in thermal balance), the mechanism must, in
151: addition, not allow the gas to be thermally unstable \citep{Field65,
152:   Balbus86}.  In what follows we make extensive use of this
153: distinction between thermal {\it balance}, where the net cooling of a
154: given parcel of gas is zero, and thermal {\it stability}, which
155: pertains to the ability of a gas to remain in thermal balance in the
156: presence of isobaric perturbations.  While attention in the literature
157: has focused on investigating the stability of an ICM in thermal
158: balance \citep[e.g.][]{Kim03b}, herein we address the more fundamental
159: challenge of maintaining a cluster in thermal balance with one or more
160: of the heating mechanisms described above.
161: 
162: The physical mechanism causing thermal instability is easy to
163: understand.  Small regions which are slightly over-dense will radiate
164: more than their surroundings (since cooling per unit volume scales as
165: $\rho^2$) and, isobarically contracting, find a new equilibrium, which
166: is at a higher density and lower temperature.  For the temperature
167: domain in question, this leads to further cooling and a thermal
168: runaway results.  If a heating process such as conduction or DF
169: heating exists, there will be a stabilizing influence.  But if the
170: process scales as $\rho$ (per unit volume) then the equilibrium will
171: not be stable, so high temperature under-dense regions will heat
172: exponentially and lower temperature over-dense regions will still cool
173: exponentially.  However, a non-thermal component (such as cosmic rays
174: or tangled magnetic fields) can in principle suppress this instability
175: \citep[][see also the appendix herein]{Cen05, Guo07} because the
176: non-thermal pressure partially decouples the hydrodynamic balance from
177: the thermodynamics of the cluster gas.  In other words, an ICM with a
178: non-thermal component that radiatively cools will lose less total
179: pressure support than an ICM with no non-thermal component.  An ICM
180: with a non-thermal component will thus have to contract less to
181: compensate for the lost thermal pressure support.
182: 
183: Observed clusters can be rather cleanly divided into two categories
184: based on the properties of their ICM.  ``Cooling core'' (CC) clusters
185: are those which have steep temperature drops in their central regions
186: \citep{Sanderson06} and metallicity gradients \citep{DeGrandi01}; many
187: CC clusters have identified radio emission at their centers
188: \citep{Best05} and show signs of AGN activity, including observed
189: sound waves \citep[e.g.][]{Fabian06} apparently emanating from their
190: center, and bubbles at an average distance of $\sim20$ kpc from the
191: cluster center \citep{Birzan04}.  In contrast, non-cooling core (NCC)
192: clusters are approximately isothermal \citep{Sanderson06} within
193: $\sim100$ kpc and show little metallicity gradient \citep{DeGrandi01}.
194: CC clusters generally have central cooling times of $0.1-1.0$ Gyr
195: while the cooling times in NCC clusters are generally somewhat higher
196: at $\sim1$ Gyr \citep{Sanderson06}.  The present work investigates the
197: expected thermal properties of both CC and NCC clusters.
198: 
199: The aim of the present study is not to find one or more heating
200: mechanisms that can offset radiative cooling but rather to understand
201: the general conditions that lead to thermal balance in the ICM for a
202: variety of possible mechanisms.  This requires both a heating
203: mechanism (or mechanisms) that can energetically offset radiation, and
204: also maintain long term thermal balance.  
205: 
206: The rest of this paper proceeds as follows.  In $\S$\ref{sec:methods}
207: the methods are discussed, including the implementation of cooling,
208: supernovae heating, conduction, DF heating, and relativistic pressure,
209: the initial equilibria, and the numerical setup.  $\S$\ref{sec:nexp}
210: contains the results of a series of numerical experiments where the
211: thermal balance of the ICM is investigated.  A discussion and summary
212: of these results can be found in $\S\S$\ref{sec:disc} and
213: \ref{sec:sum}, respectively.  Throughout we assume $h=0.7$ where $h$
214: is the Hubble parameter in units of $100$ km s$^{-1}$ Mpc$^{-1}$.
215: 
216: 
217: \section{Methods \& Physical Processes}
218: \label{sec:methods}
219: 
220: This section reviews the relevant fluid equations, cooling function,
221: and heating mechanisms that will be explored in depth in the following
222: sections.  This section also discusses our implementation of a
223: relativistic fluid component, initial equilibria, and the numerical
224: setup.
225: 
226: \subsection{General Equations}
227: \label{sec:gen}
228: 
229: The hydrodynamical equations are:
230: \be
231: \frac{d \rho_g}{d t} = - \rho_g {\bf \nabla \cdot \bf v},
232: \label{e:hydro1}
233: \ee
234: \be
235: \frac{d{\bf v}}{d t} = -\frac{1}{\rho_g}{\bf \nabla}P_{tot} + \, {\bf g},
236: \label{e:hydro2}
237: \ee
238: \be
239: \frac{de_g}{d t} = - (e_g+P_g) {\bf \nabla \cdot v} + \Gamma - \Lambda,
240: \label{e:hydro3}
241: \ee
242: \noindent
243: where $\rho_g$, $e_g$, $P_g$, $P_{\rm tot}$ and ${\bf v}$ are the gas
244: density, internal energy density, gas pressure, total pressure, and
245: velocity, ${\bf g}$ is the total gravitational acceleration, and
246: $\Gamma$ and $\Lambda$ are the heating and cooling functions per unit
247: volume.  We have explicitly distinguished between the total and gas
248: pressures because below we will allow for the addition of a
249: relativistic fluid that can provide additional pressure support. The
250: equation of state for the gas is:
251: \noindent
252: \be
253: e_g = \frac{1}{\gamma_g-1} P_g,
254: \ee
255: where $\gamma_g=5/3$ is the ratio of specific heats.
256: 
257: The gravitational acceleration is the combination of a (passive) dark
258: matter halo, a central cD galaxy, and the self-gravity of the gas:
259: \be
260: g(r) = g(r)_{\rm{DM}} + g(r)_{\rm{cD}} - \frac{G\,M_{\rm{gas}}(<r)}{r^2},
261: \ee
262: where $g_{\rm{DM}}$ is
263: \be
264: g(r)_{\rm{DM}} = -\frac{2GM_0}{r_s^2}\bigg[\frac{\rm{ln}(1+x)}{x^2} - \frac{1}{x(1+x)}\bigg],
265: \ee
266: and is derived from the NFW \citep{NFW97} density profile:
267: \be
268: \rho_{\rm{DM}} = \frac{M_0/2\pi}{r(r+r_s)^2},
269: \ee
270: \noindent
271: where $G$ is Newton's constant, $x\equiv r/r_s$, $M_0$ is the
272: normalization and $r_s$ is the scale radius, i.e. the radius at which
273: the density profile scales as $r^{-2}$.  For reference, the mass
274: within $2r_s$ is equal to $0.86 M_0$ for the above density profile.
275: We fix $r_s=460$ kpc for comparison to \citet{Kim05}.  This value for
276: $r_s$ is bracketed by the observational range \citep{Vikhlinin06a}.
277: $M_0$ is allowed to vary as discussed in $\S$\ref{s:ie}.
278: 
279: The central cD galaxy mass density profile is taken to be a King
280: profile:
281: \noindent
282: \be
283: \rho_{\rm{cD}}(r) = \frac{\rho_{\rm{cD},0}}{[1+(r/r_{\rm{cD}})^2]^{3/2}},
284: \label{e:cd}
285: \ee
286: where
287: \be
288: \rho_{\rm{cD},0} = \frac{9 \sigma_{\rm{cD}}^2}{4\pi G r_{\rm{cD}}^2}.
289: \ee
290: \noindent
291: In the above equations, $\sigma_{\rm{cD}}$ is the central 1D velocity
292: dispersion of the central galaxy and $r_{\rm{cD}}$ is the core radius
293: for the King profile; these parameters are taken to be 200 $\kms$ and
294: 2.83 kpc, respectively, from a fit to the cD galaxy NGC 6166
295: \citep{Kelson02} which is representative of cD stellar density profiles.
296: Note that the King profile is for all practical purposes quite similar
297: to the more conventional de Vaucouleurs profile, but is more
298: analytically tractable.  The gravitational acceleration associated with
299: this density distribution is
300: \noindent
301: \be
302: g_{\rm{cD}}(r) = -\frac{9\sigma_{\rm{cD}}^2}{r_{\rm{cD}}} \bigg(\frac{1+r'(1+r'^2)^{-0.5}}{r'(r'+\sqrt{1+r'^2})} - \frac{{\rm ln}(r'+\sqrt{1+r'^2})}{r'^2} \bigg),
303: \ee
304: where $r'\equiv r/r_{\rm{cD}}$.
305: 
306: Mass that flows through the inner boundary (1 kpc; see below), and is
307: hence no longer followed directly in the simulation, is added to the
308: cD galaxy.  Our results are unchanged if the mass is added to a
309: central black hole instead.
310: 
311: The gas is assumed to be ideal:
312: \be
313: P = \frac{\rho_g k_B T}{\mu m_\mu} = \frac{\mu_e}{\mu}n_e k_B T,
314: \ee
315: \noindent
316: where $\mu$ is the mean molecular weight, $m_\mu$ is the atomic mass
317: unit, $k_B$ is Boltzmann's constant, $n_e$ is the electron number
318: density and $T$ is the gas temperature.  Throughout, the gas
319: metallicity is taken to be $(1/3) \,Z_\Sun$ (except in $\S$\ref{s:ie}
320: where we discuss the sensitivity of our results to different
321: metallicities).
322: 
323: \subsection{Heating \& Cooling}
324: \label{sec:phy}
325: 
326: This section describes the various heating and cooling mechanisms that
327: are potentially relevant for the thermodynamics of the ICM.  Each
328: mechanism includes an adjustable free parameter that is meant to
329: encapsulate both our ignorance regarding the applicability of the
330: mechanism to real clusters and the uncertain values of the particular
331: parameters relevant for each mechanism; a summary of these parameters
332: is provided in Table \ref{t:models}.
333: 
334: \subsubsection{Radiation}
335: 
336: \begin{figure}[t]
337: \plotone{f1.eps}
338: \vspace{0.5cm}
339: \caption{Cooling function used in the simulations (\emph{solid lines})
340:   compared to the more accurate cooling functions of
341:   \citet{Sutherland93}.  The solid lines are, from bottom to top, for
342:   primordial metallicity, $[Fe/H]=-1.5$ and $[Fe/H]=0.0$.  Note that
343:   even for low metallicity the cooling function deviates from pure
344:   free-free cooling (where $\Lambda \propto T^{1/2}$) at
345:   $T\gtrsim10^7$ K due to recombination cooling.}
346: \vspace{0.5cm}
347: \label{fig:coolfn}
348: \end{figure}
349: 
350: 
351: A simplified cooling function is adopted that broadly captures the
352: metallicity dependent features in the detailed cooling function of
353: \citet{Sutherland93} via the following function:
354: \noindent
355: \be
356: \label{eqn:rad}
357: \Lambda = 2.1 \times 10^{-27} \,n_e^2\, T^{1/2}\bigg[1+\bigg(\frac{1.3\times 10^7\, \zeta(Z)}{T}\bigg)^{1.5}\bigg],
358: \ee
359: \noindent
360: in units of $\rm{erg} \,\rm{s}^{-1} \rm{cm}^{-3}$.  The variable
361: $\zeta$ is a simple function of the gas metallicity, $Z$.  This
362: metallicity-dependent cooling function is shown in Figure
363: \ref{fig:coolfn} (solid lines) along with the detailed cooling
364: functions from \citet{Sutherland93}.  The cooling function is
365: truncated at $T=0.01$ keV because our analytic approximation becomes
366: inaccurate at lower temperature, and because more complex physical
367: phenomena not included herein (such as star formation) become relevant
368: at such temperatures.  Most of our results focus on a single
369: metallicity of $Z=\frac{1}{3} \,Z_\Sun$; in $\S$\ref{s:ie} we briefly
370: discuss how the results change when different metallicities are adopted.
371: 
372: 
373: \subsubsection{Supernovae}
374: 
375: The stellar populations of central galaxies are extremely old, with
376: typical formation epochs at $z>3$ \citep{Thomas05}; there are thus no
377: type II supernovae events in these old systems.  However, type Ia
378: supernovae events are common, even for the old stellar populations
379: comprising cD galaxies.  At low redshift, roughly one type Ia
380: supernovae event occurs every 100 years per $10^{12} M_\Sun$ of old
381: stars \citep{Mannucci05}.  Since each supernova releases $\sim10^{51}$
382: ergs, which we assume is transfered entirely to the ICM, this
383: corresponds to a time-averaged energy injection rate of $10^{49}$ erg
384: yr$^{-1}$ per galaxy of mass $10^{12} M_\Sun$.  The type Ia rate for a
385: co-evolving stellar population may be a decreasing function of time,
386: indicating that the rate at earlier epochs would be higher than the
387: value we adopt here \citep{Greggio05, Mannucci06}.  For our purposes
388: we are primarily interested in type Ia rates at $>1$ Gyr after the
389: formation epoch of the stars, precisely where the time-dependence of
390: the rates are least certain \citep{Mannucci06}.  However, even a
391: modest increase of the rate with redshift would not qualitatively
392: change our conclusions.  We therefore do not include any
393: time-dependence in the type Ia rate in what follows.
394: 
395: Since there are hundreds of galaxies within clusters,
396: supernovae-related energy injection should be treated throughout the
397: cluster.  However, since radiation is capable of substantially cooling
398: only the inner cluster region (because cooling scales as $\rho^2$ at
399: fixed temperature), and since the thermal reservoir in the outer
400: regions is large, we only distribute supernovae within the central
401: galaxy for simplicity.  The supernovae energy injection rate is
402: distributed with the same space density as the cD galaxy (cf. Equation
403: \ref{e:cd}).
404: 
405: Supernovae energy injection is non-negligible in the central regions
406: at the beginning of our simulations.  For a initially isothermal
407: cluster with central electron density $0.020$ cm$^{-3}$ and $T_i=6$
408: keV, the volume-averaged supernovae heating initially dominates
409: cooling within only 2 kpc, while for a central density of $0.005$
410: cm$^{-3}$ supernovae heating dominates within $\sim10$ kpc.  However,
411: since the injection rate is independent of the properties of the ICM,
412: it has little impact on the long-term thermal balance of the ICM.  In
413: particular, while for very low density intracluster media supernovae
414: heating is comparable to radiative cooling on small scales, it is
415: completely ineffective at larger scales (i.e. several tens of kpc)
416: because there are so few stars, and hence so few supernovae, at these
417: larger scales.
418: 
419: We have run many simulations for a variety of initial conditions and
420: additional heating sources (discussed below) with supernovae feedback
421: and indeed find that while at the central regions ($\sim1-10$ kpc) it
422: can be important, it has little effect on the long-term thermal
423: balance of the ICM.  These conclusions are qualitatively similar to
424: those of \citet{Kravtsov00} who used the observed metallicity of the
425: ICM to demonstrate that supernovae heating is insufficient to offset
426: radiative losses.  In order to simplify the discussion in the
427: following sections, we neglect this energy injection process for the
428: remainder of this work.
429: 
430: \subsubsection{Conduction}
431: 
432: The heat flux due to electron conduction may be described by:
433: \be
434: \Gamma_{\rm cond} = {\bf \nabla \cdot}(\kappa {\bf \nabla}T),
435: \ee
436: \noindent
437: where $\kappa$ is the conductivity.  For a fully ionized plasma, the
438: conductivity is:
439: \noindent
440: \be
441: \label{eqn:cond}
442: \kappa = f\,\kappa_{Sp} = f\, \frac{1.84\times 10^{-5}\, T^{5/2}}{\rm{ln}\Lambda_C} \,\,\,\, \rm{erg}\,\rm{s}^{-1}\, \rm{K}^{-1} \, \rm{cm}^{-1},
443: \ee
444: \noindent
445: where $\kappa_{Sp}$ is the classical \citet{Spitzer62} conductivity,
446: $\rm{ln}\Lambda_C\sim37$ is the Coulomb logarithm, and $f$ is a free
447: parameter describing the suppression of conductivity relative to the
448: full Spitzer value.  While it is difficult to constrain the globally
449: averaged value of $f$ in observed clusters, the strength of observed
450: temperature gradients in several clusters suggests that $f\ll 1$
451: \citep{Markevitch03, Xiang07}.
452: 
453: There are many theoretical reasons to expect that $f<1$.  The primary
454: uncertainty on the importance of conduction is the strength and order
455: of magnetic fields in the ICM.  Magnetic fields act to suppress
456: conduction across magnetic field lines (due to the small gyro-radius
457: of electrons) while permitting conduction along the field lines.
458: Observations suggest that the magnetic fields in clusters are tangled
459: with length scales $\sim1-10$ kpc \citep[e.g.][]{Taylor02}.  Tangled
460: magnetic fields or other magnetic phenomena tend to suppress the
461: thermal conductivity to a value roughly $10-30$\% of the fully Spitzer
462: value \citep[e.g.][]{Malyshkin01, Narayan01, Chandran04}, although
463: turbulence may boost the effective conductivity \citep{Cho04}.  The
464: amount of suppression depends in detail on the poorly constrained
465: properties of the cluster magnetic field, and should thus be
466: considered rather uncertain.  Finally, it should be kept in mind that
467: three-dimensional simulations of conduction in the presence of
468: magnetic fields display new instabilities \citep[e.g.][]{Parrish07a}
469: that cannot be captured in our one-dimensional treatment where the
470: conduction is by necessity isotropic.  These considerations lead us to
471: adopt $0.0<f<0.5$ as a plausible range for suppression of
472: isotropic conduction (see Table \ref{t:models}).
473: 
474: 
475: \subsubsection{DF Heating}\label{s:df}
476: 
477: A wealth of energy is stored in the orbital motions of galaxies within
478: clusters.  DF provides a way to transfer that energy to the background
479: matter, including both the dark matter and the ICM, and can be a
480: potential mechanism that balances the cooling flow, as first pointed
481: out by \citet{Miller86}.  Recently, DF heating has experienced a
482: resurgence in popularity \citep{ElZant04b, Kim05, Kim07} thanks in
483: part to detailed calculations of the efficiency of DF in a collisional
484: medium \citep{Ostriker99}, which have been verified by controlled
485: numerical experiments \citep{Sanchez01, Kim07a}. These calculations
486: showed that DF heating is stronger in collisional media (such as the
487: ICM) when galaxies are moving slightly supersonically, as appears to
488: be the case in clusters \citep{Faltenbacher05}, where the average Mach
489: number of galaxies is $\sim 1.3$.  DF will also heat the background
490: dark matter halo, thereby producing a core-like inner density profile,
491: as opposed to the cusp implied by the NFW distribution
492: \citep{ElZant04a}, although this appears to depend on the detailed
493: properties of the accreted objects \citep{Boylan-Kolchin07}.
494: 
495: A simple calculation suggesting that DF heating may be important in
496: real clusters will be presented first, followed by a more detailed
497: discussion of our implementation of DF heating.
498: 
499: \bigskip
500: 
501: $X$-ray luminosities of clusters vary from $\sim10^{43}$ to
502: $\sim10^{45}$ ergs s$^{-1}$, with of order $10$\% emitted within the
503: rapidly cooling inner region \citep{Fabian94}.  Assuming an average
504: $X$-ray luminosity of $10^{44}$ ergs s$^{-1}$, this implies a loss of
505: energy from $X$-ray radiation over a Hubble time of
506: $\sim5\times10^{60}$ ergs within the cooling inner region
507: \citep[e.g.][]{Fabian94}.  The energy liberated during the build-up of
508: the massive central cD galaxy is a plausible source of energy that may
509: balance these radiative losses.  The following is a simple calculation
510: demonstrating that the binding energy of the cD absorbed by the ICM is
511: of the same order as the energy radiated away within the cooling
512: region.
513: 
514: The energy liberated during the build-up of the cD galaxy is
515: \be
516: \Delta E = \int \phi \,{\rm d}m_\ast = 4\pi \, \int \phi \rho_\ast\,r^2 {\rm d}r,
517: \ee
518: \noindent
519: where $\Delta E$ is the binding energy, $\phi$ is the total
520: gravitational potential felt by the cD, $m_\ast$ is a unit of stellar
521: mass, and $\rho_\ast$ is the stellar density profile.  We assume that
522: $\phi$ is static for simplicity, though of course the dark matter halo
523: is being built-up at the same time as the cD.
524: 
525: The cD galaxy projected density profile is approximated as a Sersic
526: profile \citep{Sersic68}, in agreement with a variety of observations.
527: The projected Sersic profile is not analytically invertible, but
528: sufficiently accurate fitting functions exist in the literature.  The
529: de-projected stellar mass density can be approximated by
530: \citep{LimaNeto99}:
531: \noindent
532: \be 
533: \rho_\ast(r) = A \, s^{-\alpha}\,{\rm exp}\big(-b_ns^{1/n}\big),
534: \ee
535: \noindent
536: where $s \equiv r/R_e$, $b_n \approx 2n-0.327$, and $\alpha\approx
537: 1-1.188/(2n) + 0.22/(4n^2)$.  $A$ is an integration constant chosen
538: such that the integral over $\rho_\ast$ equals the total stellar mass
539: $M_\ast$.
540: 
541: For simplicity the total gravitational potential within the inner
542: regions is assumed to be isothermal:
543: \noindent
544: \be
545: \phi(r) = {\rm V}_c^2 \, {\rm ln}(r/r_{\rm cool}),
546: \ee
547: where ${\rm V}_c$ is the circular velocity.  We are only interested in
548: energy liberated within the cooling region ($r<r_{\rm cool}$), and so
549: set the potential to be zero at $r=r_{\rm cool}$.  Define
550: $\gamma\equiv r_{\rm{cool}} / R_e$.
551: 
552: Combining these results yields:
553: \be
554: \Delta E =  V_c^2\, M_\ast \, \, \frac{\int_s^0 {\rm ln}\big(\frac{s'}{\gamma}\big) \,s'^{2-\alpha} \,{\rm exp}\big[-b_n s'^{1/n}\big]{\rm d}s'}{\int_0^s \,s'^{2-\alpha} \,{\rm exp}\big[-b_n s'^{1/n}\big]{\rm d}s'}.
555: \ee
556: \noindent
557: The two integrals represent an average of the logarithmic run of the
558: potential over the stellar density profile.  The integrals are
559: functions only of the Sersic index $n$, the ratio between the cooling
560: radius and the effective radius $\gamma$, and the upper limit of
561: integration $s$ (taken to be the same in both integrals, which is
562: reasonable but not necessary).  For $1<s<3$ and $4<n<10$ the ratio of
563: these two integrals varies from $0.4$ to $1.8$, for $\gamma=1$.  The
564: integral scales roughly as $\sqrt{\gamma}$.
565: 
566: For a typical cD galaxy $M_\ast\sim 10^{12} \,M_\sun$ and for a typical
567: cluster ${\rm V}_c\sim 700\kms$ \citep{Gonzalez07} implying that $\Delta
568: E\sim 10^{61}$ ergs.  This energy can be transferred into both the
569: background ICM and dark matter halo.  If the gas were collisionless,
570: then the fraction of energy shared between the gas and the dark matter
571: would be simply proportional to their relative mass fractions
572: ($\sim1/6$ and $\sim5/6$ for the gas and dark matter respectively).
573: However, the collisional nature of the gas makes DF more efficient at
574: transonic speeds \citep{Ostriker99}; at Mach numbers near unity the
575: efficiency of DF heating on the gas is roughly twice that of the
576: collisionless dark matter.  This implies that $\sim30$\% of the energy
577: released in transferred to the ICM and the rest to the dark matter.
578: The cD binding energy thus absorbed by the ICM is comparable to the
579: energy radiated away within the cooling inner region over a Hubble
580: time.
581: 
582: There were many simplifications made in this calculation.  One
583: complication is that observations indicate that a large fraction of
584: massive ($M_\ast>10^{11.5} M_\Sun$) galaxies were already in place by
585: $z\sim1$ \citep[e.g.][]{Bundy05, Borch06, Fontana06, Brown07}, and
586: hence the energy liberated from the build-up of cDs may not provide
587: much energy to the ICM at $z<1$.  However, the outer envelope of the
588: cD may be growing substantially at $z<1$ from the shredding of
589: satellite galaxies \citep[see e.g.][for a discussion]{Ostriker77,
590:   Conroy07b, Purcell07}, suggesting that energy released from the
591: build-up of the cD is available at late times.  Moreover, the strong
592: small-scale clustering of luminous red galaxies indicates that there
593: are many massive galaxies orbiting near the cluster center, plausibly
594: transferring their orbital energy to the ICM and background halo via
595: DF \citep{Masjedi06}.
596: 
597: These calculations suggest that DF heating may be an important heating
598: mechanism.  While the simulations below address DF heating in more
599: detail, detailed high-resolution three-dimensional simulations are
600: required to fully address the importance of DF heating in the ICM of
601: clusters \citep[see e.g.][]{Kim07}.  Moreover, detailed resolution
602: studies are required to address whether or not current cosmological
603: hydrodynamic simulations are adequately resolving DF \citep[see
604: discussion in][]{Faltenbacher05, Naab07}.
605: 
606: \bigskip
607: 
608: Our implementation of DF heating closely follows that of
609: \citet{Kim05}; the reader is referred to that work for more details.
610: Note that we do not track orbits explicitly nor do we attempt to
611: resolve the actual DF wake.  We take a more approximate,
612: parameterized approach to the spherically-averaged heating rate due
613: to DF.
614: 
615: The heat flux due to DF can be described as:
616: \noindent
617: \be
618: \label{eqn:df}
619: \Gamma_{DF} = n_{\rm{gal}} \langle -{\bf F_{\rm{DF}} \cdot v} \rangle =
620: d \frac{4\pi \rho_g G^2  \overline{M_{\rm{gal}}^2} \langle I/\mathcal{M} \rangle}{c_s} \, n_{\rm{gal}}(r),
621: \ee
622: \noindent
623: where $\mathcal{M}\equiv {\bf v}/c_s$ is the Mach number,
624: $n_{\rm{gal}}$ is the number density of galaxies,
625: $\overline{M_{\rm{gal}}^2}$ is the average squared total mass of the
626: galaxies, and $c_s$ is the isothermal sound speed.  In what follows we
627: will argue for reasonable values for each of the parameters in
628: Equation \ref{eqn:df}, and will then incorporate the uncertainties and
629: possible ranges in all of these parameters into the single free
630: parameter $d$.  The angular brackets indicate an average over the
631: velocity distribution function.
632: 
633: Note that $M_{\rm gal}$ includes both the stellar and dark matter of
634: the satellite galaxies, and that since DF is proportional to the
635: average of the galaxy mass {\it squared}, more massive galaxies are
636: given greater weight.  The quantity $\overline{M_{\rm{gal}}^2}$ is
637: estimated as follows.  We assume that the dark matter-to-stellar mass
638: ratio is constant for satellites, and estimate it by taking the total
639: stellar mass of satellites in clusters to be 1\% of the total cluster
640: mass \citep{Gonzalez07} and the total amount of dark matter mass in
641: substructures to be 10\% \citep{Gao04b}, where the substructures are
642: assumed to be the likely locations of satellite galaxies
643: \citep{Conroy06a}.  These numbers imply an average dark
644: matter-to-stellar mass ratio of 10 for satellite galaxies.  If the
645: total mass is $M_{\rm gal}$, then $\overline{M_{\rm{gal}}^2} =
646: \overline{M_{\rm{star}}^2} + \overline{M_{\rm{dm}}^2} =
647: 101\,\overline{M_{\rm{star}}^2}$.  The stellar mass function provides
648: an estimate of $\overline{M_{\rm{star}}^2}$ and thus of
649: $\overline{M_{\rm{gal}}^2}$.  We adopt the global mass function of
650: \citet{Bell03} where $\alpha=1.1$ and ${\rm log}(M^\ast)=10.9\,M_\Sun$
651: are the best-fit Schechter parameters.  The Schechter parameters for
652: the luminosity function within clusters does not differ strongly from
653: the global value \citep{Hansen07}, and so we adopt the global values
654: for this calculation.  Integrating the stellar mass function to
655: $M_{\rm star}=10^8\,M_\Sun$ leads us to adopt
656: $\overline{M_{\rm{gal}}^2} = (10^{11}\,M_\Sun)^2$ as the fiducial
657: value.  The uncertainty on this quantity is explored via the tunable
658: parameter $d$ discussed above.
659: 
660: 
661: By definition we have:
662: \noindent
663: \be
664: \langle I/\mathcal{M} \rangle = \frac{\int I/\mathcal{M} f({\bf v}) d{\bf v}}{\int f({\bf v}) d{\bf v}},
665: \ee
666: where $f$ is assumed to be Maxwellian \citep{Faltenbacher05}:
667: \be
668: f({\bf v}) = \frac{4\pi N_{\rm{gal}}}{(2\pi\sigma_r^2)^{3/2}}\,{\bf v}^2\, e^{-{\bf v}^2/(2\sigma_r^2)},
669: \ee
670: \noindent
671: where $\sigma_r$ is the radial velocity dispersion of galaxies
672: (assumed for simplicity to be independent of radius), and
673: $N_{\rm{gal}}$ is the total number of galaxies within the cluster.  In
674: the following we take $\sigma_r =1000$ km s$^{-1}$ and
675: $N_{\rm{gal}}=500$.  These parameters are fixed throughout; any
676: variation of these physical quantities is incorporated into the $d$
677: parameter.  Note that DF heating depends on the state variables only
678: through $c_s$, $I/\mathcal{M}$, and $\rho_g$, and that
679: $\Gamma_{DF}\propto \rho_g$ at fixed temperature.  $I$ is the
680: efficiency factor for DF in the collisional case \citep{Ostriker99}:
681: \noindent
682: \begin{displaymath}
683: I \equiv \left\{ \begin{array}{ll}
684:  \frac{1}{2}\,\rm{ln}(1-\mathcal{M}^{-2}) + \rm{ln}(vt/r_{\rm{min}}) & \mathcal{M} >1 \\
685: \frac{1}{2}\,\rm{ln}\bigg(\frac{1+\mathcal{M}}{1-\mathcal{M}}\bigg) -\mathcal{M} & \mathcal{M} <1
686: \end{array} \right.
687: \end{displaymath}
688: \noindent
689: As is common practice, we set $vt=r_{\rm{max}}$, where $r_{\rm max}$
690: and $r_{\rm min}$ are in this problem taken to denote the size of the
691: cluster and the satellite galaxies, respectively.  The factor ${\rm
692:   ln}(vt/r_{\rm{min}})$ thus plausibly varies from $\sim4-10$.  In
693: what follows we set this factor to 6 and note that our results are
694: insensitive to this particular value.  Note that DF heating is more
695: efficient at subsonic speeds than one might naively expect because the
696: relevant efficiency is $\langle I/\mathcal{M} \rangle$, which, though
697: maximal at $\mathcal{M}=1$, decreases by only $50$\% at
698: $\mathcal{M}\approx0.5$ and decreases only weakly at $\mathcal{M}> 1$.
699: This implies that the feedback provided by the Mach number (in the
700: sense that colder/hotter systems will have more/less efficient DF
701: heating) is weaker than one might have expected.
702: 
703: The number density of galaxies, $n_{\rm{gal}}(r)$, is taken to
704: be a modified King profile with the parameters adopted from \citet{Girardi98}:
705: \noindent
706: \be
707: n_{\rm{gal}}(r) = n_{\rm{gal}}(0)[1+(r/r_c)^2]^{-1.2},
708: \ee
709: \noindent
710: where the core radius $r_c=50$ kpc.  The central density is set by
711: requiring that $n_{\rm{gal}}(r)$ integrate to the total number of
712: galaxies, $N_{\rm{gal}}$.  Note that $n_{\rm{gal}}(r)$ does not
713: include the central cD galaxy.  While most cD galaxies are near the
714: center of the halo as defined by $X$-ray imaging \citep{Lin04b}, the
715: cD may be oscillating about the gravitational center with an amplitude
716: of several kpc \citep{VDB05b}.  Such oscillations may provide
717: additional heating in the central regions, but we do not include them
718: herein.
719: 
720: 
721: \begin{deluxetable}{lll}
722: \tablecaption{Summary of Free Parameters}
723: \tablehead{ \colhead{} & \colhead{} &\colhead{Plausible} \\ 
724:  \colhead{Parameter} & \colhead{Comment} &\colhead{Range}}
725: \startdata\\
726: $Z/Z_\Sun$ & Metallicity of the gas (Eqn. \ref{eqn:rad}) & $0.1-0.6$\\
727: $f$ & Fraction of Spitzer conductivity  (Eqn. \ref{eqn:cond}) & $0.0-0.5$ \\
728: $d$ & Normalization of DF heating  (Eqn. \ref{eqn:df}) & $0.1-10.0$\\
729: $\alpha$ & Initial fraction of relativistic pressure  (Eqn. \ref{eqn:nth}) & $0.0-0.3$\\
730: \enddata
731: \label{t:models}
732: \end{deluxetable}
733: \vspace{0.5cm}
734: 
735: \subsection{Relativistic Pressure}
736: \label{s:prel}
737: 
738: Relativistic pressure in the form of cosmic rays could be an important
739: dynamical component of the ICM.  It has long been known that
740: astrophysical shocks efficiently accelerate cosmic rays
741: \citep{Blandford78}.  Numerical simulations suggest that cosmic rays
742: accelerated in shocks could account for $\sim10$\% of the total
743: cluster pressure \citep{Miniati01, Ryu03}.  Cosmic rays may also be
744: generated in buoyantly rising bubbles generated by AGN activity
745: \citep[e.g.][]{Ensslin03}.  Observations suggest that a relativistic
746: component may constitute several tens of percent of the total energy
747: density of clusters \citep{Pfrommer04, Sanders05, Dunn06, Sanders07,
748:   Werner07, Nakar07}.  In addition, it has been shown that a
749: relativistic pressure component can suppress thermal instabilities
750: \citep[][see also the appendix herein]{Cen05}.
751: 
752: There have been several recent attempts to study the effects of
753: cosmic-rays in the ICM numerically \citep{Pfrommer07a, Pfrommer07b,
754:   Pfrommer07c, Guo07}.  The model of \citet{Guo07}, which includes
755: both conduction and cosmic rays injected into the ICM from AGN-induced
756: bubbles, is able to reproduce the temperature and density profiles of
757: observed clusters.  Perhaps the most attractive feature of their model
758: is that their results do not require fine tuning of the various
759: adjustable parameters, including the amount of thermal conductivity.
760: It would be interesting to know whether and to what extent their
761: results rely on the relativistic pressure provided by the CRs, or
762: whether it is due primarily to the energetics associated with the
763: bubbles.  In the present work we test the former hypothesis.
764: 
765: In the following we take a simplified approach when exploring the
766: effects of a relativistic component.  The relativistic pressure
767: ($P_r$) is assumed to be a fixed fraction of the total pressure
768: initially:
769: \noindent
770: \be
771: \label{eqn:nth}
772: P_r = \alpha P_{tot},
773: \ee
774: \noindent
775: and we further assume that the relativistic component is perfectly
776: dynamically coupled to the gas (i.e. there is no motion or diffusion
777: of one component relative to another), and evolves adiabatically
778: (i.e. $P_r\propto \rho_g^{\gamma_r}$).  These requirements lead to a
779: fourth fluid equation:
780: \noindent
781: \be
782: \frac{\partial P_r}{\partial t} = - {\bf \nabla \cdot}({\bf v}P_r) + (1-\gamma_r)\,P_r{\bf \nabla \cdot} {\bf v},
783: \ee
784: \noindent
785: which is simply a statement of energy conservation. In the above
786: equation $\gamma_r=4/3$ and we have made use of the following equation of state:
787: \be
788: e_r = \frac{1}{\gamma_r-1} P_r,
789: \ee
790: \noindent
791: where $e_r$ is the energy density associated with the relativistic
792: fluid.  Note that relativistic pressure also enters into the
793: hydrodynamical equations (Equations \ref{e:hydro1}---\ref{e:hydro3})
794: by contributing to the total pressure.
795: 
796: These simple assumptions are motivated by CR creation and evolution in
797: real clusters.  It may be the case that CRs are created predominately
798: in the cluster center via AGN activity, or it may be that the
799: accretion shock at the cluster outskirts are the predominate source of
800: CRs.  CRs are capable of diffusing out of the cluster, but CR loss via
801: diffusion may be approximately balanced by the creation of new CRs
802: throughout the cluster.  Our simplified treatment is meant to
803: demonstrate the potential importance of CRs generally (insofar as they
804: provide relativistic pressure support); it will be the task of more
805: sophisticated modeling efforts \citep[see e.g.][]{Pfrommer07a,
806:   Pfrommer07b, Pfrommer07c, Guo07} and, ultimately, observations to
807: refine our knowledge of CR production and evolution.  Note that the
808: energetics associated with CR production (whether in e.g. shocks or
809: AGN-related bubbles) is not explored herein.
810: 
811: 
812: \subsection{Initial Equilibria}
813: \label{s:ie}
814: 
815: Every simulated cluster is set up initially in hydrostatic
816: equilibrium.  However, in general the clusters are {\it not} set up in
817: thermal balance.
818: 
819: Observations indicate that there are two classes of intracluster
820: media, the cooling-core (CC) and non-cooling-core (NCC) clusters
821: \citep{Sanderson06}.  Clusters in the former class show a rapidly
822: declining temperature profile with decreasing radius at radii less
823: than $\sim 100$ kpc, while the latter are approximately isothermal
824: within the same physical region.  Furthermore, CC clusters have factors
825: of $\sim2-3$ metallicity gradients within $\sim 100$ kpc while NCC
826: clusters do not \citep{DeGrandi01}.  For our analysis we assume that
827: neither NCC nor CC clusters have an appreciable metallicity gradient,
828: and that the metallicity is $(1/3)Z_\Sun$.  We have selected several
829: cluster runs at random and re-simulated them with metallicity varying
830: from primordial to solar composition and find no qualitative change in
831: our results.
832: 
833: Both of these classes are discussed in the sections that follow.  NCC
834: clusters as approximated as initially isothermal while the CC clusters
835: are assumed to have an initial temperature profile that is described
836: by \citep{Vikhlinin06a}:
837: \noindent
838: \be T(r)/T_{\rm CC} =  \frac{1.15\,(x/0.045)^2 + 0.5}{(x/0.045)^2+1}  \frac{1}{1+(x/0.75)^2},
839: \ee
840: \noindent
841: where $x\equiv r/r_{500}$.  This temperature profile peaks at $T_{\rm
842:   CC}$ where $r\sim0.1 r_{500}$ and drops by a factor of $\sim2-3$
843: toward the center.  Figure \ref{fig:rad_run} (\emph{bottom panel})
844: provides an example of a CC cluster temperature profile.
845: 
846: In each run the cumulative gas fraction at $500$ kpc is fixed at
847: $f_g=0.1$, in agreement with recent observations \citep{Sanderson03,
848:   Vikhlinin06a}.  The gas fraction is less than the universal baryon
849: fraction of $f_b=0.16$ for several reasons \citep[see e.g.][for a
850: discussion]{Kravtsov05}.  The conversion of gas into stars results in
851: a lower gas fraction by several tens of percent \citep{Fukugita98,
852:   Lin03}.  Both adiabatic effects and heating processes further
853: decrease the gas fraction within $\sim500$ kpc \citep{Bode07}.  Fixing
854: the gas fraction implies that simulated clusters with larger central
855: electron densities are embedded in larger dark matter halos, as
856: observed \citep{Vikhlinin06a}.  
857: 
858: With the above simplifications, each simulation is fully specified by
859: the initial central electron density and the temperature profile.
860: These two parameters, along with the assumption of hydrostatic
861: equilibrium, then determines the full density, temperature, and
862: pressure profiles, and the adopted gas mass fraction then determines
863: the mass of the static dark matter halo.  The contribution of the cD
864: galaxy to the gravitational potential is fixed initially and increases
865: in proportion to the amount of mass that flows through the inner
866: boundary condition (see $\S$\ref{sec:gen}).  In practice we
867: iteratively solve for $M_0$ such that hydrostatic balance is achieved.
868: For the central densities explored below, the NFW normalization $M_0$
869: ranges from $\sim(2-7)\times10^{14} M_\Sun$.  Allowing $M_0$ to vary
870: by fixing $f_g$ as opposed to allowing $f_g$ to vary by fixing $M_0$
871: has a negligible effect on our conclusions.
872: 
873: Observationally the temperature of clusters varies from $\sim1-10$
874: keV, and is strongly correlated with the total cluster mass
875: \citep[e.g.][]{Vikhlinin06a}.  The data also cover a wide range in
876: central electron densities, roughly $10^{-2.5}<n_{e,0}<10^{-0.5}$
877: cm$^{-3}$.  We pay more attention to how cooling and heating depends
878: on density because the cooling and heating functions are more
879: sensitive to density than temperature.  Below we include ``observed''
880: central electron densities for comparison with our initial conditions,
881: which are estimated from two sources.  \citet{Zakamska03} provide
882: central densities from hydrostatic fits to observed temperature and
883: density profiles of 10 clusters.  \citet{Vikhlinin06a} provide
884: detailed parametric fits to the electron density profiles of 13
885: clusters (two clusters are in both samples and are only counted once
886: herein).  For the latter sample we quote as the central density the
887: density at 10 kpc because in many cases the density profile at smaller
888: scales is enhanced due to cooling.  Since we are interested in
889: ``initial'' central electron densities, we essentially mask out this
890: inner cooling region when comparing to the initial electron densities
891: in our simulations.  In the simulations the initial central electron
892: densities are approximately constant within the inner 10 kpc; we thus
893: consider this simplification appropriate.
894: 
895: 
896: \begin{figure}[t]
897: \plotone{f2.eps}
898: %\vspace{0.5cm}
899: \caption{Evolution of the temperature profile of clusters with
900:   radiative cooling and no heating for $n_{e,0}=0.02$ cm$^{-3}$.  As
901:   expected, a runaway cooling flow develops and the core collapses
902:   within a cooling time.  \emph{Top Panel:} An initially NCC cluster.
903:   \emph{Bottom Panel:} An initially CC cluster.  The lines are spaced
904:   equally in time.}
905: \vspace{0.5cm}
906: \label{fig:rad_run}
907: \end{figure}
908: 
909: \subsection{Numerical Setup}
910: 
911: We utilize the Lax-Wendroff method \citep{Press92} to integrate the
912: fluid equations in their one-dimensional form.  The spatial grid is
913: logarithmic and has $N=400$ elements with range $1<r<500$ kpc.  We
914: have doubled the number of grid elements for several simulations and
915: extended the spatial grid to 1000 kpc and find no change in the
916: results.  The time-step is determined by the Courant condition:
917: \noindent
918: \be
919: \Delta t_{CFL} = 0.5\, min\bigg( \frac{\Delta r_i}{|{\bf v}_i+c_{s,i}|} \bigg),
920: \ee
921: where $i=[0,N]$ and $c_{s,i}$ is the local isothermal sound speed.
922: 
923: Conduction is implemented with a fully implicit algorithm
924: \citep{Press92}.  The pressure, density, and temperature in the ghost
925: cells are linearly extrapolated from the active zones, both at the
926: inner and outer boundaries.  At the inner boundary, the velocity is
927: also a linear extrapolation from the active zones unless the velocity
928: in the ghost cell is positive (indicating outflow) in which case the
929: velocity is set to zero.  The velocity in the ghost cell at the outer
930: boundary is always zero.
931: 
932: 
933: \section{Numerical Experiments}
934: \label{sec:nexp}
935: 
936: This section presents the results of a series of numerical
937: experiments.  The various physical processes discussed in
938: $\S$\ref{sec:phy} are discussed in a variety of combinations in order
939: to elucidate their various effects.  For each process we discuss the
940: sensitivity of the final state to the free parameters summarized in
941: Table \ref{t:models}.
942: 
943: \subsection{Radiation Only}
944: 
945: \begin{figure}[!t]
946: \plotone{f3.eps}
947: \vspace{0.5cm}
948: \caption{Evolution of the central temperature for simulations that
949:   include both radiation and DF heating.  Each panel shows the
950:   evolution of clusters with a variety of initial central densities
951:   (labeled in the figure in units of cm$^{-3}$).  The normalization of
952:   DF heating here is fixed at $d=1.0$ (cf. Equation \ref{eqn:df}).
953:   \emph{Top Panel:} Initially NCC clusters.  \emph{Bottom Panel:}
954:   Initially CC clusters.  The evolution of CC clusters is
955:   qualitatively similar to NCC clusters in the sense that thermal
956:   balance can only be maintained for a narrow range of central
957:   electron densities.}
958: \vspace{0.5cm}
959: \label{fig:rad_df}
960: \end{figure}
961: 
962: 
963: \begin{figure*}
964: \plottwo{f4a.eps}{f4b.eps}
965: \vspace{0.5cm}
966: \caption{Long-term thermal evolution of intracluster media subjected
967:   to both DF heating and radiative cooling, as a function of initial
968:   central density, $n_{e,0}$ and normalization of DF heating, $d$
969:   (cf. Equation \ref{eqn:df}).  Both NCC clusters with $T_{i,NCC}=6$
970:   keV (\emph{left panel}) and CC clusters with $T_{i,CC}=6$ keV
971:   (\emph{right panel}) are shown.  The color indicates the slope of
972:   central temperature vs. time.  Red indicates the region of parameter
973:   space where the cluster gas catastrophically heats and blue where
974:   the gas catastrophically cools (i.e. where the slope is positive or
975:   negative).  The solid black lines indicate the region of parameter
976:   space where the slope is equal to $\pm0.1$.  Note that while there
977:   is a domain where heating stably balances cooling, sensitive
978:   fine-tuning is required for a real cluster to remain in this domain.
979:   For comparison, the distribution of observed central electron
980:   densities is included at the top (see $\S$\ref{s:ie}) and the
981:   estimated location of several Abell clusters are included in the
982:   $d-n_{e,0}$ plane (\emph{circles}).}
983: \vspace{0.5cm}
984: \label{fig:stabdf}
985: \end{figure*}
986: 
987: As discussed in the Introduction, in the absence of any heating
988: mechanisms the ICM will cool catastrophically. The time required to
989: radiate away all of the internal energy of the gas is defined as the
990: cooling time\footnote{Note that this cooling time differs from the
991:   conventional cooling time by a factor of
992:   $\frac{\gamma_g}{\gamma_g-1}$ because the conventional cooling time
993:   is defined as the time required to cool \emph{isobarically} whereas
994:   the cooling time in Equation \ref{eqn:tcool} is simply the time
995:   required to radiate away all of the thermal energy.}:
996: \noindent
997: \be
998: \label{eqn:tcool}
999: t_{\rm{cool}} \equiv \frac{\rho_g e}{\Lambda} = 5.0\, \bigg(\frac{0.01\, {\rm cm}^{-3}}{n_e}\bigg)  \bigg(\frac{T}{6\, {\rm keV}}\bigg)^{0.5} {\rm Gyr}.
1000: \ee
1001: \noindent
1002: As can be seen from Equation \ref{eqn:tcool}, the central cooling time
1003: for typical clusters is less than a Hubble time.
1004: 
1005: Figure \ref{fig:rad_run} shows the evolution of the ICM for two
1006: clusters with radiative cooling and no heating.  It is clear that the
1007: clusters cool catastrophically in the absence of any heating.  The top
1008: panel shows the evolution of an initially isothermal cluster with
1009: $T=6$ keV and an initial central electron density $n_{e,0}=0.02$
1010: cm$^{-3}$; the bottom panel shows the evolution of an initially CC
1011: cluster with the same central electron density.  The cooling times for
1012: the NCC and CC clusters shown in the figure are $2.4$ and $1.7$ Gyr,
1013: respectively, based on Equation \ref{eqn:tcool}, and adequately
1014: captures the actual time for catastrophic collapse.
1015: 
1016: While there are clusters with temperature profiles similar to that
1017: shown in the top panel of Figure \ref{fig:rad_run}, it is important to
1018: note that not only are these temperature profiles transient (in the
1019: sense that the cluster continues to cool on rapid timescales), but the
1020: density profile of this radiation-only cluster (not shown) is nowhere
1021: observed in nature.  Hence the $X$-ray luminosities, which scale as
1022: the gas density squared, would be out of the range of those observed
1023: in nature.  The conclusion that observed clusters must be periodically
1024: heated is in accord with a large body of previous work.  In
1025: particular, $X$-ray spectroscopy has firmly established that the ICM
1026: is not cooling much below $\sim1/3$ of the ambient temperature
1027: \citep{Peterson01, Peterson03, Tamura01}.
1028: 
1029: The following sections discuss possible heating mechanisms that can
1030: forestall the cooling catastrophe visible in Figure \ref{fig:rad_run}.
1031: 
1032: \subsection{Radiation \& DF Heating}
1033: 
1034: \begin{figure}[!t]
1035: \plotone{f5.eps}
1036: \vspace{0.5cm}
1037: \caption{Evolution of the central temperature for simulations that
1038:   include both radiation and conduction.  The conduction
1039:   normalization is $f=0.1$ (cf. Equation \ref{eqn:cond}).  \emph{Top
1040:     Panel:} Initially NCC clusters with $T_{i,NCC}=6$ keV.
1041:   \emph{Bottom Panel:} CC clusters with $T_{i,CC}=6$ keV.}
1042: \vspace{0.5cm}
1043: \label{fig:rad_cond}
1044: \end{figure}
1045: 
1046: \begin{figure}
1047: \plotone{f6a.eps}
1048: \plotone{f6b.eps}
1049: \vspace{0.5cm}
1050: \caption{Same as Figure \ref{fig:stabdf} except now only thermal
1051:   conduction and radiative cooling are considered, and the clusters
1052:   are initially all NCC.  The normalization of the conductivity, $f$
1053:   (cf. Equation \ref{eqn:cond}), is varied as a function of the
1054:   initial central electron density $n_{e,0}$.  The initial gas
1055:   temperatures are 5 keV (\emph{top panel}) and 8 keV (\emph{bottom
1056:     panel}).}
1057: \vspace{0.5cm}
1058: \label{fig:stabcond}
1059: \end{figure}
1060: 
1061: As discussed in $\S$\ref{s:df}, there is an enormous source of energy
1062: in the orbital motions of satellite galaxies.  While satellite
1063: galaxies can transfer their orbital energy to the ICM through a
1064: variety of mechanisms, we will focus only on DF heating for
1065: simplicity.  Note however, that other mechanisms capable of
1066: transferring the orbital energy of satellites to the background gas
1067: are expected to scale in a similar way as DF heating (i.e. as $\rho_g$
1068: at fixed temperature).  The discussion that follows thus roughly
1069: encompasses a variety of heating mechanisms related to the motions of
1070: satellite galaxies.
1071: 
1072: Figure \ref{fig:rad_df} shows the evolution of the ICM central
1073: temperature, $T_c$, as a function of time for initially NCC
1074: (i.e. isothermal; \emph{top panel}) and CC clusters (\emph{bottom
1075:   panel}).  Each panel shows the central temperature evolution for a
1076: range of initial central gas densities.  The normalization of DF
1077: heating shown in the figure is set to $d=1.0$.
1078: 
1079: Several trends are apparent.  First, it is clear that one can find a
1080: particular set of parameters that leads to approximate thermal balance
1081: (neither a heating nor cooling catastrophe) over a Hubble time.  While
1082: not shown, we note that in the runs where the central temperature does
1083: not change appreciably, neither do the density nor temperature
1084: profiles.  This was first demonstrated by \citet{Kim05}.  In fact, our
1085: fiducial values for DF heating are the same as those in \citet{Kim05}.
1086: The equilibrium NCC model has $T_{i,NCC}=6$ keV and $n_{e,0}=6\times
1087: 10^{-3}$ cm$^{-3}$ which is exactly the set of equilibrium parameters
1088: found in \citet{Kim05}, thereby confirming the results of that work.
1089: 
1090: It is also apparent that DF heating as the sole heating mechanism
1091: leads to serious thermal imbalance for all but a very narrow range of
1092: electron densities, as was also demonstrated by \citet{Kim05}.  This
1093: imbalance arises because at fixed temperature the cooling scales as
1094: $\rho_g^2$ while DF heating scales as $\rho_g$.  Thus at fixed
1095: temperature there can be only one initial density that is in thermal
1096: balance, and that solution will be unstable.
1097: 
1098: In order to more fully elucidate the thermal properties of the ICM,
1099: Figure \ref{fig:stabdf} shows the long-term thermal evolution of the
1100: ICM as a function of initial central density, $n_{e,0}$ and
1101: normalization of DF heating, $d$, for NCC (\emph{left panel}) and CC
1102: (\emph{right panel}) clusters for a fixed initial temperature profile.
1103: In this figure the slope of the central temperature as a function of
1104: time is color-coded according to the sign and steepness of the slope.
1105: From Figure \ref{fig:rad_df} it is clear that $T_c$ for most clusters
1106: evolves at least quasi-linearly, and hence the slope is a reasonable
1107: metric for the evolution of the ICM.
1108: 
1109: Figure \ref{fig:stabdf} demonstrates that there is only a narrow range
1110: of thermal balance in the parameter space of $d$ and $n_{e,0}$.  This
1111: highlights the inability of DF heating operating alone to balance
1112: radiative losses in the ICM.  The normalization of DF heating that is
1113: required to balance radiative losses increases in proportion to the
1114: central electron density ($d\propto n_{e,0}$), which makes the DF
1115: heating term (Equation \ref{eqn:df}) effectively scale as $\rho_g^2$.
1116: This is not surprising because the cooling function scales as
1117: $\rho_g^2$, and so thermal balance can be achieved for a range of
1118: densities if the effective DF heating scales in the same way as
1119: cooling.
1120: 
1121: Nature may in fact provide this required scaling ($d\propto n_{e,0}$).
1122: For example, the number of galaxies per cluster is correlated with the
1123: cluster mass as $N_{\rm gal}\propto M^{0.8}$ \citep{Lin04a}, with
1124: apparently little change at higher redshift \citep{Lin06}, and the
1125: central electron density scales roughly as the cluster mass:
1126: $n_{e,0}\propto M$ \citep[though with large scatter;][]{Vikhlinin06a,
1127:   Zakamska03}.  Since we have incorporated the variation of all free
1128: parameters in Equation \ref{eqn:df} into the single parameter $d$, it
1129: follows that $d\propto N_{\rm gal}$. These additional scalings may
1130: thus in fact produce $d\propto n_{e,0}^\beta$ with $\beta\sim 1$.
1131: 
1132: In order to investigate this explicitly, we have plotted the
1133: approximate locations of several Abell clusters in the $d-n_{e,0}$
1134: plane in Figure \ref{fig:stabdf}, where the central electron density
1135: is estimated from \citet{Vikhlinin06a} and the number of galaxies,
1136: $N_{\rm gal}$, is provided by the luminosity functions of
1137: \citet{Lin04a}.  We have only included the dependence of $d$ on
1138: $N_{\rm gal}$ when estimating the different DF efficiencies for these
1139: clusters since this scaling is the dominant one.  In other words, for
1140: these clusters we adopt our fiducial values relevant for DF heating
1141: except for $N_{\rm gal}$, which we take from the literature.
1142: Moreover, the intracluster media of these clusters span a range in
1143: temperature \citep[from $\sim2-9$ keV;][]{Lin04b}.  Since the region
1144: of thermal balance in Figure \ref{fig:stabdf} is only a weak function
1145: of temperature, we include all the clusters on the same plot for
1146: simplicity.  A trend of increasing $d$ with increasing $n_{e,0}$ is
1147: evident, but the normalization is a factor of $\sim3$ lower than the
1148: region where DF heating balances cooling.  While a factor of three is
1149: probably within the uncertainty of our estimates of $d$ for these
1150: clusters, what is more important is the narrowness of the region.  Our
1151: estimates of $d$ would not only have to be systematically off by a
1152: factor of $\sim3$, but there would have to be rather extreme
1153: fine-tuning, especially at high $n_{e,0}$, for DF heating to be
1154: generically be able to balance radiative losses.  Moreover, even if
1155: clusters initially fell in this narrow range, DF heating would still
1156: respond to perturbations in $n_{e,0}$ only linearly, while cooling
1157: would respond quadratically, and thus it is unlikely that clusters
1158: would remain in that narrow region for long.
1159: 
1160: The conclusion drawn from this section is straightforward and echoes
1161: the conclusion in \citet{Kim05}: DF heating, though an important
1162: reservoir of energy, cannot be the sole heating mechanism operating to
1163: offset radiative losses in the ICM.  This is not because the energy
1164: available is insufficient but rather because there is only a narrow
1165: range of parameter space where DF heating can stably balance radiative
1166: cooling.
1167: 
1168: \subsection{Radiation \& Conduction}\label{s:cond}
1169: 
1170: Thermal conduction can in principle transport heat from the abundant
1171: thermal reservoir of the outer cluster gas to the inner cooling parts.
1172: Invoking conduction as a means to balance radiative losses in the ICM
1173: appears however to require a fair degree of fine tuning of the
1174: conductivity \citep{Zakamska03, Kim03a, Pope06, Guo07}.  As
1175: demonstrated below, this problem is particularly acute in CC clusters
1176: where the {\it observed} temperature profile drops precipitously in
1177: the inner parts --- too little conduction leads to a cooling
1178: catastrophe, while too much conduction tends to produce an isothermal
1179: core.
1180: 
1181: These issues are demonstrated graphically in Figure \ref{fig:rad_cond}
1182: where the evolution of the central temperature is plotted for both NCC
1183: (\emph{top panel}) and CC (\emph{bottom panel}) clusters, for a
1184: conductivity normalization $f=0.1$.  Notice that the evolution is
1185: qualitatively different for NCC and CC clusters.
1186: 
1187: For NCC clusters, which are initially isothermal, only clusters with
1188: very low gas densities are marginally stable, while progressively more
1189: dense clusters quickly runaway.  These trends can be understood with
1190: the aid of the conduction time:
1191: \noindent
1192: \be
1193: t_{\rm cond} \equiv \frac{0.4}{f}\, \bigg(\frac{n_e}{0.01\, {\rm cm}^{-3}}\bigg) \bigg(\frac{\lambda}{100\,{\rm kpc}}\bigg)^2 \bigg(\frac{6\, {\rm keV}}{T}\bigg)^{2.5} {\rm Gyr},
1194: \label{eqn:tcond}
1195: \ee
1196: \noindent
1197: where $f$ is the conductivity normalization and $\lambda$ is the
1198: length scale over which temperature gradients are washed out within a
1199: conduction time \citep[the inner region where the cooling time is
1200: shorter than a Hubble time is generally of order 100
1201: kpc;][]{Sanderson06}.  It is clear that larger densities lead to
1202: longer conduction times\footnote{The conduction time depends on
1203:   density because, while conduction transports energy at fixed
1204:   density, energy scales as $n_e T$, so that the conduction time
1205:   scales as $n_e$.}, and that for $f=0.1$, densities much greater than
1206: $n_{e,0}\sim 10^{-2} $ cm$^{-3}$ result in conduction times
1207: approaching the Hubble time, and much longer than the cooling time.
1208: For these high densities, conduction cannot forestall runaway cooling,
1209: as seen in Figure \ref{fig:rad_cond}.  Thermal conduction alone cannot
1210: therefore stably offset radiative losses in the intracluster media of
1211: NCC clusters.
1212: 
1213: The situation is somewhat more complex in CC clusters because there
1214: the initial temperature profile drops by a factor of $\sim2-3$ within
1215: the inner $\sim 100$ kpc.  The bottom panel of Figure
1216: \ref{fig:rad_cond} shows the central temperature evolution for these
1217: clusters.  For clusters with short conduction times, the inner
1218: (cooler) region isothermalizes to the temperature of the outer cluster
1219: region ($\sim6$ keV for these runs) before radiative effects become
1220: important.  Once isothermal, the clusters then evolve more like the
1221: NCC clusters (\emph{top panel}, Figure \ref{fig:rad_cond}).  CC
1222: clusters that are too dense have conduction times that are longer than
1223: the cooling times, and they thus cool catastrophically.  These trends
1224: persist for the full range of conductivities explored herein
1225: ($0<f<1$).
1226: 
1227: The trends evident for NCC clusters in the top panel of Figure
1228: \ref{fig:rad_cond} are shown for a wide range in conductivity
1229: normalizations and central electron densities in Figure
1230: \ref{fig:stabcond}.  We have simulated NCC clusters with an initial
1231: temperature of 5 keV ({\it top panel}) and 8 keV ({\it bottom panel})
1232: in order to demonstrate explicitly that our results are relatively
1233: insensitive to the ICM temperature.  Unlike Figure \ref{fig:stabdf},
1234: NCC clusters can only cool or remain in thermal balance, since
1235: conduction acting on an isothermal gas cannot produce a steadily
1236: increasing central temperature with time.  It is clear that increasing
1237: the conductivity results in stable intracluster media over a wider
1238: range of initial central electron densities, but the range where
1239: cooling balances heating increases only mildly as the conduction
1240: normalization increases from $f\sim0.2$ to $f\sim1.0$.
1241: 
1242: The DF heating and conduction fine tuning problems are thus somewhat
1243: different. In the former case, too much DF heating will result in
1244: runaway heating, while in the latter, too much conduction simply
1245: isothermalizes the core.  This is not acceptable because the majority
1246: of clusters show a factor of $\sim2-3$ temperature drop within their
1247: cores \citep[e.g.][]{Sanderson06, Vikhlinin06a}.  The implications are
1248: however the same: neither mechanism is generically able to reproduce
1249: observed temperature profiles.
1250: 
1251: It thus appears that previously proposed models for the ICM that
1252: construct clusters initially in thermal balance with conduction as the
1253: only heating source \citep{Zakamska03, Kim03a, Guo07} constitute a
1254: very narrow regime of parameter space and should thus not be
1255: considered realistic solutions to the cooling flow problem, unless a
1256: physical explanation for the fine-tuning is provided \citep[see also
1257: the discussion in][]{Guo07}.
1258: 
1259: 
1260: \subsection{Radiation, DF Heating, \& Conduction}
1261: 
1262: In this section the thermal balance of the ICM is explored when both
1263: DF heating and conduction are operating as heating mechanisms.
1264: 
1265: \begin{figure}[!t]
1266: \plotone{f7.eps}
1267: \vspace{0.5cm}
1268: \caption{Evolution of the central temperature for simulations that
1269:   include radiation, DF heating, and conduction.  The DF heating and
1270:   conductivity normalizations are $d=1.0$ and $f=0.1$, respectively.
1271:   \emph{Top Panel:} Initially NCC clusters.  \emph{Bottom Panel:}
1272:   Initially CC clusters.}
1273: \vspace{0.5cm}
1274: \label{fig:rcd}
1275: \end{figure}
1276: 
1277: Figure \ref{fig:rcd} plots the evolution of the central temperature
1278: for NCC and CC clusters with normalization of DF and conduction set to
1279: $d=1.0$ and $f=0.1$.  Comparing this figure to Figures
1280: \ref{fig:rad_cond} and \ref{fig:rad_df} highlights the stabilizing
1281: effects of the combination of DF heating and conduction.  Note however
1282: that the evolved CC clusters are plagued by the same issues discussed
1283: in $\S$\ref{s:cond}, namely that while the combination of DF and
1284: conduction can maintain CC clusters in thermal balance, the evolved
1285: clusters do not preserve the factor of $\sim2-3$ drop in the
1286: temperature profile characteristic of observed CC clusters.  In other
1287: words, these CC clusters that do not catastrophically cool instead
1288: turn into NCC clusters, and thus these runs cannot explain the
1289: existence of observed CC clusters.
1290: 
1291: \begin{figure}[t]
1292: \plotone{f8a.eps}
1293: \plotone{f8b.eps}
1294: \vspace{0.5cm}
1295: \caption{Same as Figure \ref{fig:stabdf}, now with DF heating, thermal
1296:   conduction, and radiative cooling, for NCC clusters with
1297:   $T_{i,NCC}=6$ keV.  \emph{Top Panel:} NCC clusters with the
1298:   conduction normalization set to $f=0.1$.  \emph{Bottom Panel:} NCC
1299:   clusters with $f=0.5$.  Note that even a small amount of conduction
1300:   makes the clusters much more thermally stable than DF heating alone
1301:   over a much wider range of central gas densities.  As in Figure
1302:   \ref{fig:stabdf} the approximate locations of several Abell clusters
1303:   are included (\emph{circles}).}
1304: \vspace{0.5cm}
1305: \label{fig:rad_cond_df}
1306: \end{figure}
1307: 
1308: 
1309: The increased thermal balance when combining DF heating with
1310: conduction is manifest in Figure \ref{fig:rad_cond_df}, which shows
1311: the evolution of the central temperature of the ICM for initially NCC
1312: clusters with $T_{i,NCC}=6$ keV in the parameter space of initial
1313: central electron density and normalization of DF heating.  This figure
1314: shows the effect of including conduction with normalization $f=0.1$
1315: (\emph{top panel}) and $f=0.5$ (\emph{bottom panel}).
1316: 
1317: Upon comparing this figure to Figures \ref{fig:stabdf} and
1318: \ref{fig:stabcond}, it is clear that the combination of DF heating and
1319: conduction makes the ICM of NCC clusters considerably more thermally
1320: balanced than either mechanism acting alone.  Indeed, for initial
1321: central electron densities $n_{e,0}\lesssim 0.02 $ cm$^{-3}$ the
1322: ICM is thermally balanced for any plausible value of the DF heating
1323: normalization, and for reasonable, not fine-tuned, choices of
1324: conductivity.
1325: 
1326: However, it is still the case that for $n_{e,0}\gtrsim 0.02 $
1327: cm$^{-3}$, no plausible amount of DF heating can offset the cooling
1328: catastrophe.  Increasing the amount of conduction to $f=1.0$ expands
1329: the zone of thermal balance to the right in Figure
1330: \ref{fig:rad_cond_df} only marginally.  While some observed NCC
1331: clusters have a central density lying within the region of thermal
1332: balance in Figure \ref{fig:rad_cond_df}, it is clear that for many
1333: observed clusters processes in addition to DF heating and conduction
1334: are at work in order to effectively balance radiative cooling.
1335: 
1336: These conclusions are rather insensitive to the initial temperature of
1337: the NCC and CC clusters.  This might be expected because the cooling
1338: function scales as $T^{1/2}$ while DF heating scales at $T^{-1/2+p}$
1339: with $p\sim0$ \citep{Kim05}.  For example, decreasing the initial
1340: temperature in Figure \ref{fig:rad_cond_df} by a factor of two results
1341: in somewhat more thermally balanced clusters in the high density
1342: regime.  However, the increased zone of thermal balance is narrow at
1343: high density and hence rather fine tuning would be required for an
1344: observed cluster to remain there.  Moreover, observed clusters with
1345: larger central densities have higher rather than lower temperatures,
1346: and so the problem of generating equilibrium clusters is only
1347: exacerbated if the observed temperature-density relation is
1348: considered.
1349: 
1350: In sum, while it appears that the combination of DF and conduction can
1351: produce intracluster media in thermal balance for a modest range of
1352: central electron densities for NCC clusters, neither NCC clusters with
1353: moderately high central electron densities nor CC clusters of any
1354: density can be kept in thermal balance for cosmological timescales
1355: with these mechanisms alone.
1356: 
1357: \subsection{The Impact of a Relativistic Fluid}
1358: 
1359: A relativistic fluid may be present in observed clusters, perhaps in
1360: the form of cosmic rays \citep{Pfrommer04, Sanders05, Dunn06,
1361:   Sanders07, Werner07, Nakar07}.  We show in the appendix that the
1362: addition of a small amount of relativistic pressure increases the
1363: thermal stability of a gas already in thermal balance (i.e. the growth
1364: of thermal perturbations is damped for intracluster media where the
1365: net heating of a parcel of gas is zero).  In this section we explore
1366: the impact of a relativistic fluid on the thermal {\it balance} of the
1367: ICM.
1368: 
1369: As discussed in $\S$\ref{s:prel}, a relativistic fluid is included in
1370: our simulations by requiring that its pressure be a fixed fraction of
1371: the thermal pressure initially.  The relativistic fluid is assumed to
1372: be massless and perfectly dynamically coupled to the gas.  Its
1373: subsequent adiabatic evolution is then governed by energy
1374: conservation.  Recall that in our treatment the relativistic fluid
1375: provides no additional thermal energy to the system.  The purpose of
1376: this experiment is thus to ascertain the potential importance of a
1377: relativistic fluid on the hydrodynamics of the ICM, not the
1378: thermodynamics.
1379: 
1380: We have run a series of simulations that span the range of parameter
1381: space shown in Figure \ref{fig:stabdf}, i.e. for a range in initial
1382: central electron densities and DF heating normalizations.  In addition
1383: to varying these two variables, we have also varied the initial
1384: fraction of relativistic pressure, $\alpha$, from 10\% to 40\% of the
1385: total pressure, which is a fraction consistent with observations
1386: \citep{Pfrommer04, Sanders05, Dunn06, Sanders07, Werner07, Nakar07}.
1387: We find that the addition of relativistic pressure has a completely
1388: negligible effect on the thermal balance of the gas for the entire
1389: range of parameter space we have explored.
1390: 
1391: At first glance this is somewhat surprising since a thermal stability
1392: analysis indicates that a small amount of relativistic pressure should
1393: increase markedly the stability of a gas already in thermal balance
1394: \citep[][see also $\S$\ref{s:append}]{Cen05}.  The key difference is
1395: that stability analyses assume the gas to be in thermal balance, while
1396: the simulations we have run are almost never set initially in thermal
1397: balance.  In the former case the classic \citet{Field65} instability
1398: (described in the Introduction) is suppressed because slightly
1399: over-dense fluid elements, which thus cool rapidly and hence lose
1400: thermal pressure support, contract less than they would otherwise
1401: thanks to the additional pressure support provided by the relativistic
1402: fluid.  However, in the latter case, where the gas in not in thermal
1403: balance, the additional pressure support provided by the relativistic
1404: fluid is immaterial because the thermal runaway is caused by a more
1405: serious imbalance between heating and cooling, which the relativistic
1406: fluid, providing only pressure support and not energy, cannot remedy.
1407: 
1408: \section{Discussion}
1409: \label{sec:disc}
1410: 
1411: The results of the preceding section imply that both thermal
1412: conduction and DF heating are potentially significant heating
1413: processes in the ICM.  In fact, for large regions of parameter space,
1414: these processes provide more than enough energy to offset radiative
1415: losses.  However, we have found that they can neither produce nor
1416: maintain intracluster media in thermal balance over cosmological
1417: time-scales.  In particular, these processes tend to generate either
1418: runaway cooling or runaway heating, in both cases producing
1419: temperature and density profiles of the ICM nowhere near those
1420: observed in nature \citep{Sanderson06, Vikhlinin06a}.  These
1421: conclusions expand upon those of \citet{Brighenti02} who found that
1422: none of a large range of possible steady-state heating mechanisms
1423: could generically reproduce the observed properties of the ICM.  For
1424: the case of DF heating, these results are driven largely by the fact
1425: that radiation losses scale as the gas density squared, while DF
1426: heating scales linearly with the density.  Since observed intracluster
1427: media exhibit a wide range of densities, fine-tuning is required to
1428: prevent thermal runaways when DF is the dominant heating mechanism.
1429: 
1430: These conclusions have several implications, two of which we now
1431: discuss in detail.  First, in light of our results from one
1432: dimensional simulations, we discuss the continued necessity of
1433: high-resolution, fully three-dimensional, cosmologically embedded
1434: hydrodynamic simulations for understanding the thermodynamics of the
1435: ICM.  Second, we discuss a heating mechanism that may plausibly
1436: generate and maintain the properties of observed intracluster media,
1437: since none of the mechanisms explored herein appear capable of doing
1438: so.
1439: 
1440: \subsection{The Cosmological Context}
1441: 
1442: Our approach has been to simulate in one dimension the evolution of
1443: the ICM when subjected to various heating mechanisms and radiative
1444: cooling in static, isolated clusters, where only the ICM was evolved.
1445: In the real Universe, however, clusters do not appear to evolve in
1446: isolation.  Indeed, cosmological cold dark matter $N$-body simulations
1447: demonstrate that cluster-sized dark matter halos ($M\gtrsim10^{14}\,
1448: M_\Sun$) continually accrete additional halos with a range of masses,
1449: many of which likely host galaxies.  The mass growth rate is
1450: time-dependent, slowing at late times \citep[e.g.][]{Wechsler02}.  The
1451: cosmological context within which clusters evolve may in fact play a
1452: significant role in their thermal history, as suggested by recent
1453: hydrodynamic simulations \citep[e.g.][]{Motl04, Burns07, Nagai07,
1454:   McCarthy07a}.
1455: 
1456: There are a variety of ways in which the cosmological setting can
1457: provide additional heating mechanisms.  A generic class of such
1458: mechanisms may be called ``gravitational'' in the sense that the
1459: energy available for heating ultimately comes from the gravitational
1460: potential energy of infalling material.  DF heating is one example of
1461: this class, but there are others.  In particular, if small gaseous
1462: clumps are accreted, they may transfer their gravitational energy to
1463: the ICM via ram pressure drag and local shocks
1464: \citep[e.g.][]{Murray04, McCarthy07a, Khochfar07, Dekel07}.  Major
1465: mergers can significantly mix the ICM, effectively transporting the
1466: abundant reservoir of energy in the cluster outskirts to the cooler
1467: inner regions \citep[e.g.][]{Burns07}.
1468: 
1469: Recent observations have suggested that the fraction of CC clusters is
1470: increasing with time since at least $z\sim1$ \citep{Vikhlinin06b,
1471:   Ohara07}.  If gravitational heating plays a significant role in the
1472: thermal history of the ICM, then one may expect that as the accretion
1473: rate slows at late times \citep{Wechsler02}, radiative losses would
1474: become increasingly dominant, in at least qualitative agreement with
1475: these recent observations.  More generally, it is clear that the
1476: time-evolution of the abundance of observed CC clusters will provide
1477: strong, unique constraints on the importance of various heating
1478: mechanisms of the ICM.
1479: 
1480: While it may be appealing to invoke major mergers or other strong
1481: mixing processes as a means to transfer energy into the inner cooling
1482: regions of clusters, such processes have potential draw-backs.  The
1483: most serious is that observed CC clusters have strong metallicity
1484: gradients within the cooling region \citep{DeGrandi01}.  If the metals
1485: are produced by type Ia SNe within the central galaxy then the
1486: timescale for generating the observed gradient is $\sim5$ Gyr
1487: \citep{Bohringer04}, which is much longer than the cooling time of the
1488: ICM in most observed clusters \citep{Sanderson06}.  Sedimentation may
1489: also contribute to the metallicity gradient on relevant timescales,
1490: but its relevance is much more uncertain because it can be highly
1491: suppressed in the presence of even modest magnetic fields
1492: \citep[e.g.][]{Fabian77b}.  If mergers are capable of mixing the ICM
1493: such that the inner cooling region is effectively heated, it seems
1494: plausible that they may also destroy the metallicity gradient.
1495: Detailed numerical simulations are required to address these issues.
1496: 
1497: It is clear that 3D cosmologically embedded simulations are preferable
1498: to 1D simulations of the type presented herein because there are
1499: potentially relevant physical processes that cannot be adequately
1500: captured in the latter approach.  While we have highlighted the
1501: potential importance of the cosmological context, it is also true that
1502: certain physical processes cannot be captured simply because our
1503: simulations are in 1D.  For example, processes such as convection and
1504: the Rayleigh-Taylor instability cannot be captured in a 1D simulation.
1505: Additional insight into the cooling flow problem can thus be gained by
1506: simply extending the types of simulations we have run into 3D.
1507: 
1508: In spite of these fact, the motivations for focusing on 1D simulations
1509: are several-fold.  First, cosmologically embedded simulations are very
1510: time-consuming --- a parameter space study similar to what we have
1511: presented in this work is not currently possible with such
1512: simulations.  Second, DF is a resonant process \citep{Tremaine84}, and
1513: so the particle requirement to adequately resolve this phenomenon may
1514: be considerably higher than what is achievable in the current
1515: generation of cosmologically embedded hydrodynamic simulations.  It is
1516: not sufficient to resolve well the dark matter halo and subhalos.
1517: Since much of the DF heating is caused by the inspiralling of the much
1518: smaller stellar systems, a resolution small compared to the $\sim10$
1519: kpc half-light radii of giant ellipticals is required.  Thus, while
1520: our approach to DF heating has been idealized, it was necessary to
1521: explore its possible effects in this parameterized manner because it
1522: is not clear that current 3D simulations are adequately resolving this
1523: phenomenon \citep[see discussion in][]{Faltenbacher05, Naab07}.
1524: Finally, even with sufficient resolution, there are a number of
1525: physical effects that we have demonstrated are potentially relevant
1526: for the thermodynamics of the ICM that are not included in the
1527: majority of cosmologically embedded hydrodynamic simulations.  Such
1528: phenomena include thermal conduction, which ultimately requires the
1529: inclusion of magnetic fields, type Ia supernovae energy injection,
1530: which is included in some but not all current simulations, and energy
1531: injection from accretion onto a central black-hole, which we now
1532: discuss in detail.
1533: 
1534: \subsection{A Self-Regulating Heating Mechanism}
1535: 
1536: In the present work we have focused on heating processes that are not
1537: generally thought of as ``feedback'' mechanisms.  While DF heating is
1538: in principal a feedback mechanism since the heating becomes
1539: inefficient at very and low high Mach numbers, in practice this
1540: feedback is too weak to prevent runaway heating or cooling.  The
1541: preceding discussion of the importance of the cosmological context
1542: suggests that additional gravitational energy is available that can
1543: contribute to the heating of the ICM.  However, the incredibly short
1544: cooling times of many CC clusters ($\ll 1$ Gyr) indicates that heating
1545: from without, via gravitational processes, must still be fine-tuned in
1546: order to prevent runaway cooling.  Heating from within, via the
1547: release of energy from accretion onto a black hole, is in many ways
1548: more appealing because such a process is explicitly self-regulating.
1549: 
1550: Black hole accretion-mediated feedback mechanisms
1551: \citep[e.g][]{Ciotti01, Ruszkowski02, Kaiser03, Guo07, Ciotti07} are
1552: qualitatively different from the mechanisms we have explored.  These
1553: self-regulating feedback processes are thought to work schematically
1554: as follows.  As the gas begins to cool in the inner regions, its
1555: thermal pressure no longer provides sufficient support to the
1556: overlying material, and thus matter flows inward and eventually
1557: accretes onto the central black hole.  The mechanical energy provided
1558: by AGN activity, which is proportional to the rate of mass infall,
1559: then heats up the surrounding gas until the thermal pressure, which is
1560: now increasing due to the energy injection, halts the inward flow of
1561: matter, thereby diminishing the accretion-driven energy injection.
1562: The cycle then begins again in a regulatory fashion, and the ICM thus
1563: neither heats nor cools catastrophically.  It is this regulatory
1564: feedback processes which seems most promising in explaining the
1565: properties of observed intracluster media because the relevant
1566: physical parameters need not be tuned to any particular values for
1567: heating to balance cooling generically \citep[see e.g.][]{Guo07}.
1568: 
1569: A serious constraint on AGN-related feedback is that it may destroy
1570: the observed metallicity gradient in CC clusters if the energy
1571: deposition at the cluster center is sufficient to drive convection.
1572: Recently, \citet{Voit05} have shown that AGN feedback can be effective
1573: at balancing radiative cooling \emph{and} maintaining the observed
1574: metallicity gradient if the AGN outbursts are rather gentle, occur
1575: every $\sim10^8$ years and last in duration for $\sim10^7$ years.  It
1576: remains to be seen if this proposal can be confirmed with both
1577: numerical simulations and, ultimately, observations.
1578: 
1579: While AGN heating may be a promising candidate at generating and
1580: maintaining intracluster media in thermal balance over cosmological
1581: time-scales, simulations of the ICM which include it do so only in a
1582: rather simplistic fashion, in part because a detailed understanding of
1583: the physical processes involved in black hole accretion currently
1584: eludes us.  The way in which energy is transfered from the black hole
1585: to the surrounding ICM is also obscure.  Until these and related
1586: issues are better understood, we should continue to seek out other
1587: potential self-regulating heating processes that could potentially be
1588: relevant for the thermodynamics of the ICM.  Nonetheless, varied
1589: observations of sound waves, bubbles, and strong radio emission
1590: \citep{Birzan04, Best05, Fabian06}, predominantly in CC clusters, in
1591: addition to their very short cooling times \citep{Sanderson06},
1592: strongly suggests that AGN-related activity plays an important role in
1593: the thermodynamics of the ICM.
1594: 
1595: 
1596: 
1597: \section{Summary}
1598: \label{sec:sum}
1599: 
1600: We have presented the results from a series of 1D simulations aimed at
1601: understanding the importance of type Ia supernovae heating, thermal
1602: conduction and DF heating (due to the orbital motions of satellite
1603: galaxies) on the thermal properties of the intracluster media of
1604: clusters.  Both initially non-cooling core (NCC; i.e. isothermal) and
1605: cooling core (CC) clusters were simulated, for a wide range of initial
1606: central electron densities and gas temperatures.
1607: 
1608: Marginalizing over the uncertain efficiencies of DF heating and
1609: thermal conduction, it is clear that only NCC clusters with central
1610: electron densities $n_{e,0}\lesssim 0.02$ cm$^{-3}$ can be maintained
1611: in thermal balance over a Hubble time if both DF and conduction
1612: operate; neither mechanism alone can generate generically stable ICM
1613: at these densities.  At higher densities no reasonable amount of
1614: conduction or DF heating can prevent runaway cooling in NCC clusters.
1615: 
1616: The failure of the combination of conduction and DF heating at either
1617: generating or maintaining observed intracluster media is more
1618: pronounced for CC clusters, which have observed temperature profiles
1619: that decline by a factor of $\sim2-3$ from the outer to inner regions.
1620: This temperature drop is extremely difficult to maintain in the face
1621: of thermal conduction in our simulations, because thermal conduction
1622: acts to erase temperature gradients.  In fact, of the
1623: $\mathcal{O}(10^3)$ simulations run with a wide range of DF heating
1624: and conduction efficiencies and initial gas densities, none generated
1625: stable CC cluster profiles.  Since CC clusters constitute $\sim70$\%
1626: of observed clusters \citep[e.g.][]{Peres98}, we regard this failure
1627: as strong evidence that other heating processes in the ICM must be at
1628: work besides conduction and DF heating.
1629: 
1630: Our results demonstrate that there are numerous energy reservoirs
1631: capable of supplying enough energy to offset radiative losses.  The
1632: crux of the cooling flow problem therefore lies not in finding one or
1633: more mechanisms capable of providing enough energy to the ICM, but
1634: rather in finding one or more mechanisms that can supply the energy in
1635: a way that maintains thermal equilibrium in the ICM over cosmological
1636: time-scales.  Only low-density NCC clusters can be maintained with DF
1637: heating and conduction.  The processes explored herein are not
1638: manifestly self-regulating, and they thus require fine-tuning in order
1639: to generate the observed properties of CC and high-density NCC
1640: intracluster media generically. For these types of observed
1641: intracluster media, it seems likely that an explicitly self-regulatory
1642: feedback process such as black hole accretion-powered energy injection
1643: (i.e. AGN feedback) is required.
1644: 
1645: \acknowledgments 
1646: 
1647: CC gratefully acknowledges Jim Stone, Anatoly Spitkovsky, and Ian
1648: Parrish for extensive help and support with extensive numerical
1649: issues, and Andrey Kravtsov for many fruitful discussions.  We thank
1650: Paul Bode, Woong-Tae Kim, Andrey Kravtsov, and Jim Stone for helpful
1651: comments on an earlier draft.
1652: 
1653: %\bibliography{../master_refs.bib}
1654: \input{ms.bbl}
1655: 
1656: %\begin{thebibliography}{128}
1657: %\expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1658: %\end{thebibliography}
1659: 
1660: 
1661: 
1662: \begin{appendix}
1663: 
1664: \section{Thermal Stability for Generic Heating Mechanisms In the Presence of a Relativistic Fluid}
1665: \label{s:append}
1666: 
1667: The generalized Field criterion \citep{Field65, Balbus86} states that
1668: a gas is thermally unstable to isobaric perturbations if:
1669: \noindent
1670: \be
1671: \frac{\partial( \mathcal{L}/T)}{\partial T}\bigg|_P < 0,
1672: \ee
1673: \noindent
1674: where $\mathcal{L}$ is the net loss function per unit mass defined
1675: such that $\rho_g \mathcal{L} = \Lambda-\Gamma$ where $\Lambda$ and
1676: $\Gamma$ are the cooling and heating rates per unit volume.  In what
1677: follows the subscript ``$g$'', denoting gaseous quantities, will be
1678: omitted for brevity.
1679: 
1680: The following identity holds for a general gas in thermal equilibrium ($\rho \mathcal{L}=0$):
1681: \be
1682: \frac{\partial( \mathcal{L}/T)}{\partial T}\bigg|_P = \frac{1}{\rho T}\bigg[ \frac{\partial(\rho \mathcal{L})}{\partial T}\bigg|_\rho + \frac{\partial \rho}{\partial T}\bigg|_P \frac{\partial(\rho \mathcal{L})}{\partial \rho}\bigg|_T\bigg].
1683: \label{eqn:field}
1684: \ee
1685: The first and third derivatives on the right-hand side are determined by the relevant heating and cooling mechanisms while the second is set by the various sources of pressure support.  The following identity will be useful:
1686: \be
1687: \frac{\partial \rho}{\partial T}\bigg|_P = -\frac{\partial P}{\partial T}\bigg|_\rho \,\bigg/\, \frac{\partial P}{\partial \rho}\bigg|_T.
1688: \ee
1689: \noindent
1690: Assume that the pressure support is provided by both the thermal
1691: pressure of the gas and a relativistic fluid:
1692: \be
1693: P = P_{\rm th} + P_{\rm rel} = K_1 \rho T + P_{\rm rel}(\rho),
1694: \ee
1695: where $K_1$ is a constant and $P_{\rm rel}$ is a function only of density.  It then follows that
1696: \be
1697: \frac{\partial P}{\partial T}\bigg|_\rho = K_1\rho,
1698: \ee
1699: and
1700: \be
1701: \frac{\partial P}{\partial \rho}\bigg|_T = K_1 T + \frac{\partial P_{\rm rel}}{\partial \rho}\bigg|_T,
1702: \ee
1703: and hence:
1704: \be
1705: \frac{\partial \rho}{\partial T}\bigg|_P =\frac{ K_1\rho}{K_1 T + \frac{\partial P_{\rm rel}}{\partial \rho}\big|_T} = \frac{\rho}{T}\frac{P_{\rm th}}{P_{\rm th}+\rho \frac{\partial P_{\rm rel}}{\partial \rho}\big|_T } =  \frac{\rho}{T}\frac{\alpha'}{\alpha' + \frac{\partial{\rm ln}P_{\rm rel}}{\partial {\rm ln} \rho}\big|_T},
1706: \label{eqn:ptot}
1707: \ee
1708: \noindent
1709: where in the last equality we have defined $\alpha'\equiv P_{\rm th} / P_{\rm
1710:   rel}$.
1711: 
1712: Now assume that cooling is dominated by thermal Bremsstrahlung radiation and that the heating has a simply power-law dependence on $\rho$ and $T$.  Then:
1713: \be
1714: \rho \mathcal{L} = \Lambda_0 \rho^2 T^{1/2} - \Gamma_0 \rho^\beta T^{-1/2+\delta},
1715: \ee
1716: \noindent
1717: where $\Lambda_0$ and $\Gamma_0$ are constants and $\beta$ and
1718: $\delta$ are the power-law indices for heating, defined in a way that
1719: will prove useful below.  After some algebra, and using the fact that
1720: $\rho \mathcal{L}=0$, we have
1721: \noindent
1722: \be
1723: \frac{\partial(\rho \mathcal{L})}{\partial T}\bigg|_\rho = (1-\delta)\Lambda_0\rho^2 T^{-1/2},
1724: \ee
1725: and
1726: \be
1727: \frac{\partial(\rho \mathcal{L})}{\partial \rho}\bigg|_T = (2-\beta) \Lambda_0 \rho T^{1/2}.
1728: \ee
1729: Combining these with Equations \ref{eqn:field} and \ref{eqn:ptot} we have, after more algebra:
1730: \be
1731: \frac{\partial( \mathcal{L}/T)}{\partial T}\bigg|_P = \Lambda_0 \rho T^{-3/2}\bigg[ (1-\delta) - (2-\beta)\frac{\alpha'}{\alpha' + \frac{\partial{\rm ln}P_{\rm rel}}{\partial {\rm ln} \rho}\big|_T}\bigg].
1732: \ee
1733: \noindent
1734: The gas is thus thermally unstable if:
1735: \be
1736: \delta > \frac{(\beta-1)\alpha'+\frac{\partial{\rm ln}P_{\rm rel}}{\partial {\rm ln} \rho}\big|_T}{\alpha'+\frac{\partial{\rm ln}P_{\rm rel}}{\partial {\rm ln} \rho}\big|_T} = \frac{(\beta-1)\alpha'+\gamma}{\alpha'+\gamma} = \frac{(\beta-1)+ \frac{\alpha}{1-\alpha}\gamma}{1+\frac{\alpha}{1-\alpha}\gamma},
1737: \ee
1738: \noindent
1739: where we have assumed that $P_{\rm rel}\propto \rho^\gamma$ and
1740: defined $\alpha\equiv P_{\rm rel} / P_{\rm tot}$.
1741: 
1742: This equation recovers the result found in \citet{Kim05}.  There they
1743: explored the stability of DF heating, where $\beta=1$, in the absence
1744: of relativistic pressure ($\alpha=0$).  The instability criterion in
1745: this case is $\delta>0$.  As shown in \citet{Kim05}, DF heating
1746: generally yields a temperature dependence such that $\delta>0$, and
1747: thus DF heating alone is thermally unstable.
1748: 
1749: This equation also implies that heating sources which scale as
1750: $\rho^2$ are thermally stable if the temperature dependence scales as
1751: a power less than $1/2$, which is physically intuitive since $\rho^2
1752: T^{1/2}$ is the scaling of the cooling function.  Moreover, if the
1753: heating source scales as $\rho^2$ then the relativistic fluid has no
1754: influence on the thermal stability.
1755: 
1756: \end{appendix}
1757: 
1758: 
1759: \end{document}