0704.2607/ms.tex
1: %\documentclass[12pt,preprint]{aastex}  % for e-submission to ApJ - one column
2: 
3: %\documentclass[12pt,preprint2]{aastex}  % for e-submission to ApJ - two column
4: 
5: \documentclass{emulateapj}   % to make it look like ApJ
6: 
7: %\usepackage{graphicx,natbib}
8: 
9: \def\etal{{\sl et al.}}
10: \def\kms{{\rm km/s}}
11: \def\Ms{{\rm M_\odot}}
12: 
13: \citestyle{aa}  % correct formatting for ApJ style files
14: 
15: \begin{document}
16: 
17: \title{The Santa Fe Light Cone Simulation Project: I. Confusion and the WHIM in Upcoming Sunyaev-Zel'dovich Effect Surveys}
18: \author{Eric J. Hallman\altaffilmark{1,2}, Brian W. O'Shea\altaffilmark{3}, 
19: Jack O. Burns\altaffilmark{2}, Michael L. Norman\altaffilmark{4},
20: Robert Harkness\altaffilmark{5} \& Rick Wagner\altaffilmark{4}}
21: 
22: \altaffiltext{1}{National Science Foundation Astronomy and
23:   Astrophysics Postdoctoral Fellow}
24: 
25: \altaffiltext{2}{Center for Astrophysics and Space Astronomy,
26:   Department of Astrophysics and Planetary Sciences, 
27:   University of Colorado at Boulder, Boulder, CO 80309; hallman, burns@casa.colorado.edu}
28: 
29: \altaffiltext{3}{Theoretical Astrophysics (T-6), Los Alamos National
30: Laboratory, Los Alamos, NM 87545; bwoshea@lanl.gov}
31: 
32: \altaffiltext{4}{Center for Astrophysics and Space Sciences,
33: University of California at San Diego, La Jolla, CA 
34: 92093; mnorman, rpwagner@cosmos.ucsd.edu}
35: 
36: \altaffiltext{5}{San Diego Supercomputing Center, MC0505, 9500 Gilman
37:   Drive, La Jolla, CA 92093; harkness@sdsc.edu}
38: 
39: 
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
41: \begin{abstract}
42:  We present the first results from a new generation of simulated large sky coverage ($\sim$100 square degrees) Sunyaev-Zeldovich effect (SZE) cluster surveys
43: using the cosmological adaptive mesh refinement N-body/hydro code Enzo.
44: We have simulated a
45: very large (512$^3$h$^{-3}$Mpc$^3$) volume with unprecedented dynamic
46: range. We have generated simulated light cones to match the resolution and
47: sensitivity of current and future SZE instruments. Unlike many
48: previous studies of this type, our
49: simulation includes unbound gas, where an appreciable fraction of the
50: baryons in the universe reside.
51: 
52: We have found that cluster line-of-sight overlap may be a
53: significant issue in upcoming single-dish SZE surveys. Smaller
54: beam surveys ($\sim$1$\arcmin$) have more than one massive
55: cluster within a beam diameter 5-10\% of the time, and a larger beam
56: experiment like Planck has multiple clusters per beam 60\% of the time. 
57: We explore the contribution of unresolved halos and unbound gas to the
58: SZE signature at the maximum decrement. We find that there
59: is a contribution from gas outside clusters of $\sim$16\% per object on average for upcoming surveys. This adds both bias and scatter to the deduced value
60: of the integrated SZE, adding difficulty in
61: accurately calibrating a cluster Y-M relationship. 
62: 
63: Finally, we find that in images where objects with M $>$ 5
64: $\times$10$^{13}$ M$_{\odot}$ have had their SZE signatures removed,
65: roughly a third of the total SZE flux still remains. This gas exists at least partially in the Warm Hot
66: Intergalactic Medium (WHIM), and will possibly be detectable with the
67: upcoming generation of SZE surveys.
68: \end{abstract}
69: 
70: \keywords{cosmology: theory--galaxies:clusters:general--cosmology:observations--hydrodynamics--methods:numerical--cosmology:cosmic microwave background}
71: 
72: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
73: \section{Introduction}\label{sec:Intro}
74: Clusters of galaxies form from the highest peaks in the
75: primordial spectrum of density perturbations generated by inflation in
76: the early universe. They are the most massive virialized structures in
77: the universe, and as such are rare objects. 
78: The number density of galaxy clusters as a function of mass and
79: redshift is strongly dependent on a number of cosmological
80: parameters. In particular, counting the abundance of clusters above
81: some lower mass limit as a function of cluster redshift places
82: constraints on $\Omega_b$, $\Omega_M$, $\sigma_8$, and the dark energy
83: equation of state parameter, $w$ \citep{wang, haiman01}.  
84: 
85: Observational measurement of the cluster abundance over large
86: sky areas and redshift range is required to
87: generate cosmological parameter constraints which are complementary
88: to constraints from the cosmic microwave background (CMB), type Ia
89: supernovae, Big Bang nucleosynthesis (BBN), the Lyman-$\alpha$ forest,
90: and galaxy redshift surveys \citep{tozzi}. Cluster survey yields
91: depend on the value of the minimum cluster luminosity probed as a function of
92: redshift, the scaling between cluster luminosity and mass, the growth function 
93: of structure, and the redshift evolution
94: of the comoving volume element, all of which depend on a complex combination
95: of cosmological parameters \citep{rosati} and intracluster medium
96: (ICM) physics \citep{ev04}. 
97: 
98: \subsection{Sunyaev-Zel'dovich Effect Surveys}
99: The Sunyaev-Zel'dovich Effect \citep{sz} is a process by
100: which the hot electrons trapped in the large dark matter potential
101: wells of clusters inverse Compton scatter CMB photons to higher
102: energy, resulting in a low frequency ($<$ 218 GHz) decrement, and a corresponding
103: high frequency ($>$ 218 GHz) increment in the CMB on the angular scale of the
104: cluster. The strength of the SZE decrement/increment is characterized
105: by the Compton $y$ parameter, which results from the line of sight
106: integral of the thermal pressure
107: \begin{equation}
108: y = \int \sigma_T n_e \frac{k_b T}{m_e c^2} dl,
109: \end{equation}
110: where $n_e$ is the electron density and $T$ is the gas electron
111: temperature. We also define the integrated Compton $y$ parameter as
112: \begin{equation}
113: Y = \int y dA
114: \end{equation}
115: which is the integration over the area subtended by a circle
116: corresponding to some relevant cluster physical radius. 
117: The observed temperature fluctuation corresponding to a
118: give value of  $y$ in a given frequency band is
119: \begin{equation}
120: \frac{\Delta T}{T} = y g(x),
121: \end{equation}
122: where 
123: \begin{equation}
124: g(x) = \left(x \frac{e^x + 1}{e^x -1} - 4\right) [1 + \delta_{sze}(x,T_e)],
125: \end{equation}
126: $x = h\nu/kT_{cmb}$ and $\delta_{SZE}$ is a relativistic correction as
127: described in \citet{itoh}. For our purposes in this paper we have
128: neglected the relativistic correction, but will explore it in future
129: work. This correction is small (less than 1\% at the maximum decrement
130: frequency) for clusters with T $<$ 10 keV. 
131: 
132: The SZE is particularly useful in cosmological studies due to its
133: near redshift independence \citep{reph,birk,carlstrom}.  Therefore, observations of clusters are
134: not limited to low redshift as in the X-ray, but can extend to as high
135: as z $\sim$ 2, where the number of massive clusters becomes small. An additional consequence of the redshift independence
136: of SZE surveys is that a flux-limited survey is also approximately a mass
137: limited survey \citep{reph,haiman01}. These two properties make SZE surveys uniquely
138: valuable for cluster abundance counts and determination of the cluster
139: mass function with redshift, provided one can obtain independent
140: optical redshifts for the galaxies in the identified objects. The near redshift
141: independence of the SZE creates unique complications for large area surveys
142: which do not seriously affect other types of surveys (e.g., optical and X-ray).  In
143: particular, the contribution to the sky signal from both low mass and
144: distant halos, as well as unbound gas, may be a significant source of
145: confusion.  
146: 
147: There are several
148: upcoming millimeter wavelength cluster surveys with new telescopes including
149: the Atacama Pathfinder Experiment Sunyaev-Zeldovich survey (APEX-SZ) \citep{apex}, the South Pole Telescope
150: (SPT) \citep{spt}, and the Atacama Cosmology
151: Telescope (ACT) \citep{act}, in addition to the space-based \textit{Planck} Surveyor \citep{planck}, which will conduct large blind surveys of clusters
152: using the SZE. Table \ref{sz_inst} shows the values
153: of the relevant instrumental characteristics for the single-dish
154: survey telescopes we have taken from the literature and used in the
155: following analysis. For this study, we limit ourselves to results
156: using a single band, $\sim$144 GHz, where the SZE decrement is
157: maximal, though these results will generalize to other bands and
158: to multiwavelength studies, since the SZE from all gas will have roughly the same
159: spectral signature (modified slightly by relativistic effects). However, for removing contaminating signals (e.g. radio
160: point sources, CMB) multiwavelength coverage will be very desirable.
161: 
162: The variation in survey characteristics for these instruments has
163: important consequences for cluster surveys. For example, the distribution of sources detected as
164: a function of cluster redshift should be different for each
165: survey. This is because although the SZE surface brightness does not diminish with distance,
166: the angular size of the objects does vary with redshift, and may be
167: larger or smaller than the instrument beam for any given cluster. This selection effect is
168: modified also by the volume sampled as a function of redshift in a
169: fixed angular field and the growth rate of structure in an $\Lambda$CDM
170: universe. It is important to understand these selection effects
171: in order to constrain cosmology. We must be able to determine the
172: correct distribution of clusters as a function of redshift from the
173: surveys or systematic errors in estimated cosmological parameters will
174: result. 
175: 
176: \begin{table}
177: \caption{Characteristics of Upcoming SZE Surveys}
178: \begin{center}
179: \begin{tabular}{cccc}
180: \hline
181: \hline
182: Survey & Angular Coverage & Beam Size ($\sim$144 GHz) & RMS per beam\\
183: \hline
184: APEX-SZ & TBD & 1.0$\arcmin$ & 10 $\mu$K \\
185: SPT & 4000 deg$^2$ & 1.0$\arcmin$ & 10 $\mu$K \\
186: ACT & 100 deg$^2$ & 1.7$\arcmin$ & 2 $\mu$K \\
187: Planck & All-Sky & 7.1$\arcmin$ & 6.0 $\mu$K \\
188: \hline
189: \end{tabular}
190: \end{center}
191: \label{sz_inst}
192: \end{table}
193: 
194: There are also several centimeter
195: wave ($\approx$30 GHz) interferometers that are performing surveys of the SZE, such as the Arcminute Microkelvin Imager (AMI) \citep{ami} and
196: the Sunyaev-Zel'dovich Array (SZA) \citep{sza}. SZE surveys have the potential to
197: strongly constrain the $w$ parameter for dark energy, since they
198: sample clusters to large redshift.
199: 
200: Determining the abundance and distribution in mass and redshift of
201: massive galaxy clusters from observables is a challenging
202: exercise.  In the realm of cluster
203: abundance counts, one needs to know with high precision the mass range
204: of clusters probed in a flux-limited survey as a function of cluster
205: redshift. This determination depends on detailed knowledge of the
206: scaling between mass and light in clusters. It is critical to understand how cluster
207: observables correlate with cluster total mass in order to use clusters
208: of galaxies as precision cosmological tools.
209: 
210: \subsection{The Role of Simulations in Understanding Cluster Surveys}
211: Recent results indicate that high resolution N-body simulations
212: \citep{warren,heitman,reed} generate mass functions
213: which differ significantly from the Press-Schechter result and also from
214: the subsequent modifications of \citet{sheth} and
215: \citet{jenkins}. Since there is strong evidence that purely analytic
216: methods are insufficient, ``precision cosmology'' requires the use of
217: numerical simulations. In other words, in order to make predictions
218: which match the observed cluster population to percent-level
219: precision, analytic methods are inadequate. 
220: 
221: The output of numerical simulations of clusters can be compared to the
222: observed cluster mass function.  This comparison is non-trivial, however, due to
223: the uncertain nature of the conversion between observable quantities
224: (e.g., X-ray luminosity, SZE Compton y parameter, lensing shear) and cluster total
225: mass. Observations of the cluster gas typically depend on the detailed properties of
226: the hot baryons in clusters. It has been shown by our group and
227: others that cluster observables have a strong dependence on the baryonic physics in the ICM
228: \citep{hall06,nagai}, and are subject to an array of
229: uncertainties. However, the SZE signal integrated over the projected
230: cluster area (as defined in
231: Equation 2) inside $r_{500}$ is unique in that
232: the general scaling with mass is relatively independent of the assumed 
233: gas physics \citep{motl05,nagai}. While the normalization of the Y-M relation has some
234: dependence on ICM physics, the slope and tightness of the correlation
235: are unaffected \citep{nagai}. Most recent simulated light cone
236: calculations use the dark matter mass function generated by large N-body simulations like
237: the Hubble Volume simulation or the Millennium run, with
238: ``painted on'' baryons to generate mock surveys \citep[e.g.,][]{evrard02,paint2}. These studies
239: typically assume the gas is isothermal and in hydrostatic equilibrium
240: with the dark matter potential. 
241: 
242: Both simulations including relevant physics \citep[e.g.,][]{white03, rasia06, hall06} and high resolution X-ray
243: observations of galaxy clusters \citep[e.g.,][]{utp_obs, 1E06, a168}
244: suggest that many clusters depart strongly
245: from both equilibrium and isothermality. These deviations
246: can have a strong impact on both the observable and derived properties
247: of clusters. Thus, in order to properly simulate sky surveys, it is
248: critically important to self-consistently include baryons in numerical
249: simulations. While some work has been done in this area
250: \citep[e.g.,][]{springel, white02, roncar, ronc07}, the largest volumes simulated were
251: small ($\sim$100-200 h$^{-1}$Mpc), and only sufficient to
252: generate synthetic light cones of roughly 1-4 deg$^2$. The simulation performed for this study
253: models a significantly larger physical volume than previous efforts,
254: has a higher peak physical resolution, and fully incorporates baryons
255: self-consistently. This allows us to perform much larger synthetic
256: surveys ($\sim$100 deg$^{2}$) than could be done with earlier
257: N-body/hydro simulations. 
258: 
259: Cosmological N-body/hydro simulations have advanced significantly in the last decade, such that the
260: simulation output now compares quite well to observations of galaxy
261: clusters \citep{sfw}. In particular, our group and others have shown that there is good agreement in simulated and
262: observed scaling relations between bulk
263: cluster ICM properties (e.g., cluster mass, X-ray luminosity, X-ray spectral
264: temperature) \citep{motlchar}. There remain important differences, particularly
265: for lower mass clusters, which indicates the need for a better
266: understanding of the details of baryonic cluster physics. The advance
267: of realism in simulations is a result of diligent, iterative efforts by various
268: groups of investigators to directly compare simulations to
269: observations. This study uses a large volume high resolution adiabatic
270: simulation, and serves as a template for more complex runs involving
271: additional non-gravitational physical processes which are likely important to
272: accurately modeling cluster surveys. These results should be
273: relatively robust in any case, as it has been shown that SZE survey
274: yields are relatively independent of cluster physics details \citep{white02}.
275: 
276: \subsection{Modeling SZE Surveys}
277: A variety of approaches have been taken to model SZE surveys. Most
278: involve either semi-analytic prescriptions or N-body
279: simulations where the gas is added in a post processing step
280: \citep[e.g.,][]{schulz}. As discussed in the previous section, there are limitations to
281: these methods, particularly in the assumptions of hydrostatic
282: equilibrium and isothermality. Some studies \citep[e.g.,][]{white02,holder}
283: have discussed the contribution of unresolved clusters/groups to the
284: signature of detected clusters. The presence of gas outside the cluster virial radii in low density
285: unbound structures such as filaments is potentially also quite important. This gas is completely absent in
286: non-hydrodynamic treatments of the simulation volume, appears
287: naturally in our calculation, and is expected to contain 40-50\% of
288: the baryons in the universe \citep{whim}. Since the SZE does not diminish with
289: distance, and results from a line of sight integral of the gas
290: pressure, the sum of all the flux from unbound gas could be a
291: significant contributor to the total flux in any cluster detection.
292: 
293: In this paper, we have examined the contribution of line-of-sight
294: baryonic gas to the expected SZE signal of simulated clusters. We
295: have stacked a (512 h$^{-1}$Mpc)$^3$ volume (comoving) adaptive mesh refinement (AMR)
296: N-body/hydro simulation to generate a survey of a light cone
297: subtending 100 square degrees on the sky. Our simulated
298: survey covers a larger sky area, and at higher angular resolution,
299: than any previous N-body + hydro simulated survey.  
300: 
301: There is extensive work in the literature on cluster detection
302: algorithms for SZE surveys,
303: \citep[e.g.,][]{diego,herranz,hobson,schafer,melin}. These methods
304: involve various techniques designed to spatially filter out the
305: primary CMB anisotropies (so called matched filtering), wavelet techniques, and application of public
306: tools such as SExtractor\footnote{http://terapix.iap.fr/rubrique.php?id\_rubrique=91/} \citep{sex}. We find the existing work to be quite
307: detailed, and do not introduce new algorithms of this type
308: here. Indeed, it is important to step back from the analyses which
309: have attempted to include all the relevant contaminants and
310: instrumental effects and explore the intrinsic difficulties resulting
311: from the cluster population as projected on the sky. There is a
312: limiting precision one can expect from cluster surveys irrespective of
313: the ability to remove instrumental effects, point source
314: confusion and sky backgrounds. This limit results from the possibly irreducible
315: confusion due to clusters, groups, lower mass halos and unbound gas, all
316: of which contribute to the SZE signal with a nearly identical spectral
317: signature. This study examines these effects with a more realistic
318: cosmological calculation than has typically been done, including the
319: full complement of baryons expected in the real universe. 
320: 
321: We explore the intrinsic limitations of SZE surveys by comparing and
322: characterizing the contribution of unresolved halos and unbound
323: filamentary gas to the cluster signal in samples that might result
324: from upcoming surveys
325: using a full hydro/N-body simulation of the cluster sky. In future
326: work, we will model backgrounds and instrumental characteristics as
327: has been done in the literature recently \citep[e.g.,][]{sehgal,
328:   melin, schafer} with a focus on techniques for accurately extracting
329: cluster properties and abundance. 
330: 
331: We discuss our methodology of simulating these large surveys in
332: Section 2, analytic predictions of SZE observables in Section 3,
333: present results in Section 4, discussion in Section 5, and summarize
334: our work in Section 6. 
335: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
336: \section{Methodology}\label{sec:Methodology}
337: 
338: %---------------------------------------------------------------------------
339: \subsection{The Enzo Code}\label{sec:enzocode}
340: 
341: `Enzo'\footnote{http://lca.ucsd.edu/portal/software/enzo/} is a publicly available, extensively tested 
342: adaptive mesh refinement
343: cosmology code developed by Greg Bryan and colleagues \citep{bryan97,bryan99,norman99,oshea04,
344: 2005ApJS..160....1O}.
345: The specifics of the Enzo code are described in detail in these papers (and references therein),
346: but we present a brief description here for clarity.
347: 
348: The Enzo code couples an N-body particle-mesh (PM) solver \citep{Efstathiou85, Hockney88} 
349: used to follow the evolution of a collisionless dark
350: matter component with an Eulerian AMR method for ideal gas dynamics by \citet{Berger89}, 
351: which allows high dynamic range in gravitational physics and hydrodynamics in an 
352: expanding universe.  This AMR method (referred to as \textit{structured} AMR) utilizes
353: an adaptive hierarchy of grid patches at varying levels of resolution.  Each
354: rectangular grid patch (referred to as a ``grid'') covers some region of space in its
355: \textit{parent grid} which requires higher resolution, and can itself become the 
356: parent grid to an even more highly resolved \textit{child grid}.  Enzo's implementation
357: of structured AMR places no fundamental restrictions on the number of grids at a 
358: given level of refinement, or on the number of levels of refinement.  However, owing 
359: to limited computational resources it is practical to institute a maximum level of 
360: refinement, $\ell_{max}$.  Additionally, the Enzo AMR implementation allows arbitrary 
361: integer ratios of parent
362: and child grid resolution, though in general for cosmological simulations (including the 
363: work described in this paper) a refinement ratio of 2 is used.
364: 
365: Since the addition of more highly refined grids is adaptive, the conditions for refinement 
366: must be specified.  In Enzo, the criteria for refinement can be set by the user to be
367: a combination of any or all of the following:  baryon or dark matter overdensity
368: threshold, minimum resolution of the local Jeans length, local density gradients,
369: local pressure gradients, local energy gradients, shocks, and cooling time.
370: A cell reaching
371: any or all of the user-specified criteria will then be flagged for refinement.  Once all 
372: cells of a given level have been flagged, rectangular solid boundaries are determined which 
373: minimally 
374: encompass them.  A refined grid patch is then introduced within each such bounding 
375: volume, and the results are interpolated to a higher level of resolution.
376: 
377: In Enzo, resolution of the equations being solved is adaptive in time as well as in
378: space.  The timestep in Enzo is satisfied on a level-by-level basis by finding the
379: largest timestep such that the Courant condition (and an analogous condition for 
380: the dark matter particles) is satisfied by every cell on that level.  All cells
381: on a given level are advanced using the same timestep.  Once a level $L$ has been
382: advanced in time $\Delta t_L$, all grids at level $L+1$ are 
383: advanced, using the same criteria for timestep calculations described above, until they
384: reach the same physical time as the grids at level $L$.  At this point grids at level
385: $L+1$ exchange baryon flux information with their parent grids, providing a more 
386: accurate solution on level $L$.  Cells at level $L+1$ are then examined to see 
387: if they should be refined or de-refined, and the entire grid hierarchy is rebuilt 
388: at that level (including all more highly refined levels).  The timestepping and 
389: hierarchy rebuilding process is repeated recursively on every level to the 
390: maximum existing grid level in the simulation.
391: 
392: Two different hydrodynamic methods are implemented in Enzo: the piecewise parabolic
393: method (PPM) \citep{Woodward84}, which was extended to cosmology by 
394: \citet{Bryan95}, and the hydrodynamic method used in the ZEUS magnetohydrodynamics code
395: \citep{stone92a,stone92b}.  We direct the interested reader to the papers describing 
396: both of these methods for more information, and note that PPM is the preferred choice
397: of hydro method since it is higher-order-accurate and is based on a technique that 
398: does not require artificial viscosity, which smoothes shocks and can smear out 
399: features in the hydrodynamic flow.
400: 
401: 
402: %---------------------------------------------------------------------------
403: \subsection{Simulation Setup and Analysis}\label{sec:simsetup}
404: 
405: The simulation discussed in this paper is set up as follows.  We initialize
406: our calculation at $z=99$ assuming a cosmological model with  $\Omega_m = 0.3$, 
407: $\Omega_b = 0.04$, $\Omega_{CDM} = 0.26$, $\Omega_\Lambda = 0.7$, $h=0.7$ (in units of 100 km/s/Mpc), 
408: $\sigma_8 = 0.9$, and using an \citet{eishu99} power spectrum
409: with a spectral index of $n = 1$.  The simulation is of a volume of the 
410: universe 512~h$^{-1}$~Mpc (comoving) on a side with a $512^3$ root grid.  The dark matter
411: particle mass is $7.228 \times 10^{10}$~h$^{-1}$~M$_\odot$ and the mean baryon mass resolution
412: is $1.112 \times 10^{10}$~h$^{-1}$~M$_\odot$.  The simulation was then evolved to $z=0$ with
413: a maximum of $7$ levels of adaptive mesh refinement (a maximum spatial resolution of 
414: $7.8$~h$^{-1}$ comoving kpc), refining on dark matter 
415: and baryon overdensities of $8.0$ (to ensure an approximately Lagrangian mass
416: resolution in baryonic structures).  The equations of hydrodynamics were
417: solved with the Piecewise Parabolic Method (PPM) using the dual energy formalism.
418: The entire grid hierarchy (including both particle and baryon information) was 
419: written out at regular intervals, and in particular, data was output at intervals
420: of $\Delta z = 0.25$ between $z=3$ and $z=2.5$ (inclusive), and $\Delta z = 0.1$ between $z=2.5$ 
421: and $z=0.1$ (inclusive).
422: 
423: Analysis was performed on every data output between $z=3$ and $z=0.1$ in an
424: identical way.  The HOP halo-finding algorithm \citep{eishut98} was applied to the 
425: dark matter particle distribution to produce a dark matter halo catalog.
426: Spherically-averaged, mass-weighted radial profiles of various baryonic and dark matter
427: quantities including density, temperature, and pressure were then generated 
428: for every halo in the catalog with an
429: estimated halo mass greater than $4 \times 10^{13}$~M$_\odot$.  These radial 
430: profiles were used to calculate more accurate virial masses and radii as well 
431: as an estimate for the Compton $y$ parameter as a function of impact parameter on
432: the halo.  Projections of the integrated Compton $y$ value along the line of sight were created
433: for each of the three axes along the simulation volume, with two projections per axis --
434: one of the front half of the simulation volume, and one along the back half.  Each
435: projection has an approximate depth of $\Delta z = 0.1$.  These projections have a 
436: resolution of $2048$ pixels on a side.
437: 
438: %---------------------------------------------------------------------------
439: \subsection{Generation of the ``Santa Fe'' Light Cone}\label{sec:lightcone}
440: 
441: Mock SZE observations of a $10^{\circ} \times 10^{\circ}$ patch of sky are 
442: generated by stacking projections from the simulation discussed in 
443: Section~\ref{sec:simsetup}.  These ``light cones'' are created by stacking 
444: projections of the Compton $y$ parameter at each redshift output.  At each redshift, the
445:  projection is chosen to be along a random axis, and has been randomly shifted in space
446: such that the positions of large scale structure is uncorrelated.  Additionally, 
447: the projections have been rescaled to the resolution of the light cone, which is 2048 pixels
448: per side, or a resolution of $17.58$ arc seconds per pixel.  This scaling may involve 
449: tiling (for redshifts
450: where 512 h$^{-1}$ Mpc comoving corresponds to less than $10^{\circ}$ on the sky) or
451: interpolating (for redshifts where  512 h$^{-1}$ Mpc comoving corresponds to more
452:  than $10^{\circ}$ on the sky). Secondary maps are created which
453:  include only the
454: Compton $y$ parameter contributed by gas within the virial radius of halos with
455: masses above $5 \times 10^{13}$~M$_\odot$, and only gas outside of the virial radius
456: of these objects.  200 of these mock ``light cones'' at this size and
457: resolution are created using different random 
458: seeds.  These light cones (named ``Santa Fe'' light cones due to the location
459: where the project was conceived) have angular resolution which
460: is significantly higher than any current or proposed SZE observational
461: campaign. \textit{The goal of this
462: analysis is not to determine an optimal method for source finding, but
463: to determine the contamination from unresolved halos and unbound gas for a simple method.} 
464: %---------------------------------------------------------------------------
465: 
466: 
467: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
468: \section{Analytic Predictions for SZE Observables}\label{sec:theory}
469: Here we describe some of the theory behind the use of SZE cluster
470: observations in constraining cosmology. Though we are aware that these
471: types of analytic calculations have been performed previously, we show
472: them here to motivate not just the current analysis, but that which
473: will be performed for subsequent papers in this series. 
474:  
475: One of the most useful methods for retrieving cosmological information from
476: SZE observations of galaxy clusters is by the calculation of galaxy cluster
477: counts as a function of redshift.  The number of galaxy clusters above some
478: given minimum mass $M_{min}(z)$ in a redshift bin of width $dz$ and solid angle 
479: $d\Omega$ can be defined using the Press-Schechter formalism 
480: \citep{ps} as
481: \begin{equation}
482: \frac{dN}{dz d\Omega}(z) = \frac{dV}{dz d\Omega}(z) \int_{M_{min}(z)}^{\infty} dM \frac{dn}{dM}(M,z)
483: \label{eqn-Nofz}
484: \end{equation}
485: where $dV/dzd\Omega$ is the cosmological comoving volume element at a given redshift, 
486: and $\frac{dn}{dM} dM$ is the comoving halo number density as 
487: a function of mass and redshift.  The latter is expressed as by \citet{jenkins} as
488: \begin{eqnarray}
489: \frac{dn}{dm} (M,z) & = & -0.315 \frac{\rho_0}{M} \frac{1}{\sigma_M} \frac{d\sigma_M}{dm} \nonumber\\
490: & exp & \left[ -|0.61-log(D(z)\sigma_M)|^{3.8} \right]
491: \label{eqn-dndm}
492: \end{eqnarray}
493: where $\sigma_M$ is the RMS density fluctuation, computed on mass scale M from the $z=0$ linear power spectrum \citep{eishu99}, 
494: $\rho_0$ is the mean matter density of the universe, defined as $\rho_0 \equiv \Omega_m \rho_c$ (with $\rho_c$ being 
495: the cosmological critical density, defined as $\rho_c \equiv 3 H_0^2 / 8 \pi G$), and $D(z)$ is the linear growth function, given by this fitting function:
496: \begin{eqnarray}
497: & &D(z) = \frac{1}{1+z} \frac{5 \Omega_m(z)}{2} \nonumber\\
498: & &\left\{ \Omega_m(z)^{4/7} - \Omega_\Lambda(z) + [1+\frac{\Omega_m(z)}{2}][1+\frac{\Omega_\Lambda(z)}{70}]\right\}^{-1}
499: \label{eqn-DofZ}
500: \end{eqnarray}
501: \citep{1992ARA&A..30..499C}, with $\Omega_m(z)$ and $\Omega_\Lambda(z)$ defined as:
502: \begin{equation}
503: \Omega_m(z) = \Omega_{m,0} (1+z)^3 E^{-2}(z)
504: \label{eqn-omofz}
505: \end{equation}
506: and
507: \begin{equation}
508: \Omega_\Lambda(z) = \Omega_{\Lambda,0} E^{-2}(z)
509: \label{eqn-olofz}
510: \end{equation}
511: where $\Omega_{m,0}$ and $\Omega_{\Lambda,0}$ are the density of matter and cosmological constant at the
512: present day, expressed in units of the critical density.  The cosmological volume element is given by:
513: \begin{equation}
514: \frac{dV}{dz d\Omega}(z) = \frac{c}{H_0} \frac{ (1+z)^2 D_A^2}{E(z)}
515: \label{eqn-dV}
516: \end{equation}
517: where $D_A(z)$ is the angular diameter distance as a function of redshift, c is the speed of light, 
518: $H_0$ is the Hubble parameter at $z=0$, and E(z) is given by:
519: \begin{equation}
520: E^2(z) = \Omega_{m,0} (1+z)^3 + \Omega_\Lambda
521: \label{eqn-Eofz}
522: \end{equation}
523: in a flat universe with a cosmological constant \citep{peebles-book}.  The RMS amplitude of the 
524: density fluctuations as function of mass
525: at any redshift, as smoothed by a spherically symmetric window function with a characteristic comoving
526: radius R, can be computed from the matter power spectrum using the relation:
527: \begin{equation}
528: \sigma^2(M,z) = \int_0^\infty \frac{dk}{k} \frac{k^3}{2 \pi^2} P(k,z) |\mbox{\~{W}}_R(k)|^2
529: \label{eqn-sigma}
530: \end{equation}
531: where $\mbox{\~{W}}_R(k)$ is the Fourier transform of the real-space top hat smoothing function:
532: \begin{equation}
533: \mbox{\~{W}}_R(k) = \frac{3}{k^3 R^3} \left[ sin(kR)-kRcos(kR)\right]
534: \label{eqn-W}
535: \end{equation}
536: 
537: The radius R is calculated for a given mass by using the relation $M = \frac{4}{3} \pi R^3 \Omega_m \rho_c$,
538: and $\sigma(M,z)$ is normalized to $\sigma_8$, defined as the RMS density fluctuation when smoothed by a sphere with a comoving radius of $8$~h$^{-1}$ Mpc at $z=0$, using observations
539: of large-scale structure or the cosmic microwave background. The matter power spectrum is expressed as:
540: \begin{equation}
541: \frac{k^3}{2 \pi^2} P(k,z) = \left( \frac{c k}{H_0} \right)^{3+n} T^2(k) \frac{D^2(z)}{D^2(0)}
542: \label{eqn-pofk}
543: \end{equation}
544: where $T(k)$ is the matter transfer function which describes the way in which the processing of the 
545: initial spectrum of matter density fluctuations during the radiation-dominated era~\citep{peebles-book} and
546: $D(z)$ is the fitting function for the linear growth function, as given in Equation~\ref{eqn-DofZ}.
547: In the calculations discussed in this paper, we use the transfer function $T(k)$ provided by \citet{eishu99}. 
548: 
549: Figure~\ref{fig.dNdz} shows the number of galaxy clusters per square degree as a function of redshift with
550: $M_{tot} \geq 1 \times 10^{14}$~h$^{-1}$~M$_\odot$ in the WMAP Year III ``most favored''
551: cosmology ($\Omega_m = 0.268$,~$\sigma_8 = 0.776$) and in the cosmology used in the simulation in this paper
552: ($\Omega_m = 0.3$,~$\sigma_8 = 0.9$).  Due to the higher $\Omega_m$ and $\sigma_8$, significantly more 
553: galaxy clusters are expected to be seen in this simulation than one would expect given the WMAP Year III result.
554: 
555: Figure~\ref{fig.dNdz-cosmo} shows the number of galaxy clusters per square degree as a function of redshift
556: with $M_{tot} \geq 1 \times 10^{14}$~h$^{-1}$~M$_\odot$ for a variety of cosmological models.
557: Panel (a) shows a sequence of cosmologies where all  parameters except $\sigma_8$ are held constant, using the same
558: cosmological parameters as in the simulation described in this paper ($\Omega_m = 0.3$, $\Omega_b = 0.04$, 
559: $\Omega_\Lambda = 0.7$, $h=0.7$, $n=1$), but varying $\sigma_8$ from $0.6$ to $1.2$ in steps of $0.1$ (bottom to top lines).
560: Panel (b) shows a sequence of cosmologies where all parameters except $\Omega_m$ and $\Omega_\Lambda$ are held constant 
561: using the same cosmological parameters
562: as in the simulation described in this paper ($\Omega_b = 0.04$, $h=0.7$, $n=1$, $\sigma_8 = 0.9$), but varying $\Omega_m$
563: from 0.2 to 0.4 in steps of 0.05 (bottom to top lines), and keeping $\Omega_\Lambda = 1 - \Omega_m$.  It is interesting
564: to note that varying $\sigma_8$ while holding all other cosmological parmaters results in a change in both the overall
565: number of halos (a
566: factor of more than 3 between WMAP and the cosmology in our
567: simulation) and the redshift at which the distribution peaks .
568: Both effects can be explained by examining the term in the 
569: exponent in equation~\ref{eqn-dndm}.  The comoving number density of halos is maximized when $0.61-log(D(z)\sigma_M) = 0$,
570: or $D(z) \times \sigma_M \simeq 4.07$.  For a given set of cosmological parameters, increasing $\sigma_8$ increases $\sigma_M$ for 
571: a given mass value, and thus maximizes the number density of halos at a given mass at a smaller linear growth factor (or, more
572: intuitively, a higher redshift).  Variation in $\Omega_m$ and $\Omega_\Lambda$ while holding all other parameters constant
573: effectively results in a change in the normalization of the overall halo number density, while keeping the redshift at which
574: this distribution peaks roughly constant.
575: 
576: These results are of great interest for upcoming SZE surveys, which
577: will sample clusters to relatively low mass at high redshifts (z =
578: 0.5-1.0) compared to optical and X-ray surveys. This redshift range is
579: where we should expect the largest difference between a low $\sigma_8$
580: cosmology preferred by WMAP and a higher value typically used in
581: cosmological simulations. The large difference between the abundance
582: of clusters in these different cosmologies should lead to large
583: differences in the number of identified clusters in surveys. This
584: means that very early on in any SZE survey it should become fairly
585: obvious which cosmology is preferred by cluster observations. This
586: will be an even stronger constraint when optical follow up
587: observations are used to
588: determine redshifts, which will break the degeneracy between $\Omega_M$ and
589: $\sigma_8$. 
590: 
591:  
592: \begin{figure}
593: \begin{center}
594: \includegraphics[width=0.45\textwidth]{./f1.eps}
595: \end{center}
596: \caption{Number of galaxy clusters per square degree as a function of redshift with 
597: $M_{tot} \geq 1 \times 10^{14}$~h$^{-1}$~M$_\odot$ in the WMAP Year III ``most favored''
598: cosmology (solid line; $\Omega_m = 0.268$,~$\sigma_8 = 0.776$) and the cosmology used in the simulation in this paper
599: (dashed line; $\Omega_m = 0.3$,~$\sigma_8 = 0.9$).  The distribution peaks at 
600: $z \simeq 0.55$ for the WMAP Year III cosmology and at $z \simeq 0.64$ for the cosmology used in the simulations in this paper.
601: }
602: \label{fig.dNdz}
603: \end{figure}
604: 
605: 
606: \begin{figure}
607: \begin{center}
608: \includegraphics[width=0.45\textwidth]{f2a.eps}
609: \includegraphics[width=0.45\textwidth]{f2b.eps}
610: \end{center}
611: \caption{Number of galaxy clusters per square degree as a function of redshift with 
612: $M_{tot} \geq 1 \times 10^{14}$~h$^{-1}$~M$_\odot$ for a variety of cosmologies.
613: Panel (a):  All  parameters except $\sigma_8$ are held constant, using the same
614: cosmological parameters as in the simulation described in this paper ($\Omega_m = 0.3$, $\Omega_b = 0.04$, 
615: $\Omega_\Lambda = 0.7$, $h=0.7$, $n=1$), but varying $\sigma_8$ from $0.6$ to $1.2$ in steps of $0.1$ (bottom to top lines).
616: Panel (b):  All parameters except $\Omega_m$ and $\Omega_\Lambda$ are held constant using the same cosmological parameters
617: as in the simulation described in this paper ($\Omega_b = 0.04$, $h=0.7$, $n=1$, $\sigma_8 = 0.9$), but varying $\Omega_m$
618: from 0.2 to 0.4 in steps of 0.05 (bottom to top lines) and keeping $\Omega_\Lambda = 1 - \Omega_m$.
619: }
620: \label{fig.dNdz-cosmo}
621: \end{figure}
622:  
623: 
624: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
625: \section{Results}\label{sec:results}
626: 
627: %---------------------------------------------------------------------------
628: \subsection{Simulated Mass Function}\label{sec:massfctn}
629: 
630: One of the most basic tests of the correctness of a cosmological simulation 
631: is whether or not it can match the predicted halo mass function for a given 
632: cosmology.  This is particularly important in the context of creating simulated
633: sky maps for cosmological surveys of any kind, given that the number of 
634: halos and their redshift distribution is the most basic test of cosmological 
635: properties.  Figure~\ref{fig.cumnumdens} shows the cumulative number density
636: of cosmological halos as a function of mass for several redshifts which span the
637: range of interest for the topics discussed in this work ($z = 0.1 -3$).  
638: The halos were found as described in Section~\ref{sec:simsetup}, and 
639: the masses used are the total
640: halo virial mass (including both baryons and dark matter), rather than the mass returned
641: by the Hop halo finding algorithm~\citep{eishut98}.  We also show the fitting function 
642: for cumulative halo number density obtained by \citet{warren} as a reference.  At 
643: low redshifts ($z=0.1-1$) the mass function of halos from the simulation agrees quite well with 
644: the fitting function over the mass range of interest.  This is encouraging, as the bulk of 
645: galaxy clusters in the universe are at
646: these low redshifts (as shown in Figure~\ref{fig.dNdz}).  At higher redshifts ($z=2$), the 
647: fitting function and halo mass function only agree at the highest masses 
648: ($M_{halo} \geq 10^{14}h^{-1}$~M$_\odot$).  This is to be expected: grid-based codes, including adaptive
649: mesh codes, tend to suppress low-mass halo formation, particularly at high redshift, as has
650: been seen in recent code comparisons~\citep{2005ApJS..160....1O,
651:   2005ApJS..160...28H}. This suppression in our study only exists due
652: to the choice of simulation setup. In order to model such a large physical
653: volume with both N-body + hydrodynamics, we must sacrifice mass
654: resolution due to computational effort concerns. Given
655: that the cosmological surveys of interest will only be sensitive to halos in the mass range
656: where this simulation agrees well with the Warren et al. fit, and the relative paucity of these
657: halos at $z \ga 2$, there is little cause for concern.
658: 
659:  
660: \begin{figure}
661: \begin{center}
662: \includegraphics[width=0.45\textwidth]{f3.eps}
663: \end{center}
664: \caption{Cumulative number density at several redshifts.  Solid: $z=0.1$.  Short-dashed: $z=1.0$.  Long-dashed: $z=2$.
665:  Thick lines are simulation results calculated using the halo virial masses (dm+gas) and the thin
666: lines are the Warren fitting function \citep{warren}.
667: }
668: \label{fig.cumnumdens}
669: \end{figure}
670:  
671: 
672: %---------------------------------------------------------------------------
673: \subsection{SZE Angular Power Spectrum}\label{sec:szpspec}
674: We have calculated angular power spectra for the 200 survey images to
675: determine the cosmic variance in this field, and the general form of
676: the power spectrum. The calculation involves determining the
677: power spectral density of the image as a function of image
678: scale. Additionally, we have performed the same calculation on maps
679: smoothed with the beam size at 144 GHz for each of four upcoming SZE
680: survey instruments. We have modified the images to model the limitations
681: of these instruments in a simple way.  First, for each instrument, we
682: have gaussian smoothed the image with the FWHM of the beam at a wavelength
683: (frequency) of 2.1mm (144 GHz) corresponding to the maximum SZE
684: decrement. All the upcoming SZE single-dish surveys will have the
685: capability of operating at this wavelength. We have also rebinned each
686: image such that the beam diameter is represented by at least two pixels in
687: order to be Nyquist sampled. A ``background'' in each case is
688: generated by adding a gaussian distributed variation with a FWHM equal
689: to each survey's stated limiting sensitivity. 
690: 
691: \begin{figure*}
692: \plotone{f4.eps}
693: \caption{Upper Left: Light cone survey image of 100 square degrees
694:   modified to simulate the beam size and limiting sensitivity of the
695:   Planck Surveyor all-sky survey at 144 GHz (7.1$\arcmin$,
696:   6$\mu$K). Upper right: Angular power spectra generated from these images. Lower left: Light cone image modified to
697:   simulate APEX-SZ/SPT survey characteristics at $\sim$144 GHz
698:   (1.0$\arcmin$, 10$\mu$K). Lower right: Light cone image modified to
699:   simulate ACT survey (1.7$\arcmin$, 2$\mu$K). Numbers in parentheses
700:   indicate (beam size, survey sensitivity/beam) at 144 GHz.}
701: \label{images}
702: \end{figure*}
703: The result of our analysis of the power spectrum of the simulated SZE
704: surveys is shown in Figure \ref{images}. The solid lines are the
705: mean values of the power from the 200 stacking realizations of the
706: light cone, and the dotted lines indicate the range in which 90\% of
707: our simulated light cone power spectra fall. The color indicates
708: which survey's characteristics were used to generate the result. 
709: 
710: The location of the peak of each curve is a function of the resolving
711: power of the survey, as is its amplitude, in that the power at the
712: smallest angular scale for each survey is different.  The cyan line
713: (labeled SWH01) is the result obtained by \citet{springel} with a 1 degree
714: angular scale light cone generated from an SPH N-body/hydro
715: simulation run with cosmological parameters matching our simulation. The raw power spectrum is generated with no smoothing from
716: the raw light cone survey images. The SWH01 result is different from
717: ours, possibly a result of lower spatial resolution in the simulation
718: (resulting in a deficit in the small scale angular power) and had a smaller
719: angular field (1 square degree) than ours, leading to a higher $\ell$
720: cutoff ($\ell$=400) than we show ($\ell$=70). As has been shown
721: previously, and specifically in Figure 4 of \citet{springel}, analytic
722: predictions of the SZE power vary widely depending on the calculation
723: method. They neither agree with the simulation results, nor in many
724: cases with each other. It seems that the analytic result has not
725: converged, and so to avoid confusion we do not plot it in this work.  
726: 
727: It is clear that each survey will sample a slightly different range of scales, though
728: they obviously all are able to measure the large angular scale power. Each power
729: spectrum peaks and turns over where the angular scale of the beam
730: begins to limit the measurement at high $\ell$. The power spectrum is not very sensitive to
731: non-gravitational gas physics at low multipole numbers, but primarily
732: is sensitive to cosmology, particularly the value of $\sigma_8$
733: \citep{white02,holder}. The addition of non-gravitational physics does impact
734: the small scale power however, for example \citet{holder} show that
735: preheating results in reduced small scale power in their simulated
736: images. 
737: 
738: What is also of interest in this analysis is the size of the variance shown
739: by the 90\% range error bars. On 100 square degree patches of the sky
740: at $\ell \la 2000$ the deduced power can be different by factors of
741: 5-8. This indicates that the power spectrum can be quite different
742: from one area of the sky to another, and clearly requires greater sky
743: coverage to be well constrained. The cosmic variance range does
744: not become very small until $\ell \ga $ a few thousand. 
745: 
746: %---------------------------------------------------------------------------
747: \subsection{SZE Source Identification}\label{sec:szsourceid}
748: To identify objects in the light cone images, we simply locate the
749: projected clusters from the three-dimensional halo finding in the
750: image plane. Since we design the shifting and stacking strategy, it is
751: trivial to determine the image plane location of each cluster in each
752: redshift slice of the light cone. We also have calculated the
753: spherically averaged radial profiles projected into the image plane of
754: the Compton $y$-parameter. For each cluster with M $\geq 10^{14}
755: M_{\odot}$, we can then calculate the integrated value of Y (when
756: comparisons to analytic results are not performed, we use true masses
757: on the simulation grid, without the $h^{-1}$ modifier). The
758: result is shown in Figure \ref{logNS}. In this case, we have
759: integrated out to the virial radius of each cluster. We show the
760: variance in 200 stacking realizations of the simulated light cone,
761: indicating the increase in variation at high flux, as expected due to
762: rare, very massive objects projected into some fraction of the
763: images. 
764: \begin{figure}
765: \includegraphics[width=0.45\textwidth]{f5.eps}
766: \caption{Plot of number of objects per square degree in the simulated
767:   survey image as a
768:   function of integrated Compton $y$ parameter. Dotted lines are 90\% variance as calculated from
769:   200 stacking realizations of the light cone. }
770: \label{logNS}
771: \end{figure}
772: 
773: We are also interested in the redshift distribution of the objects,
774: which can be compared to analytic estimates. 
775: The result of this analysis is shown
776: in Figure \ref{fzplot}. In each case, we have taken the mean value and
777: plotted it as a solid line, and the 90\% variance in 200 light cone
778: realizations as dotted lines. The black lines indicate all clusters
779: with total masses above $10^{14} M_{\odot}$ in the simulation which are in the projected
780: field of the survey. Blue and red lines are for clusters above higher
781: mass cutoffs, blue for M$\geq 3.0 \times 10^{14} M_{\odot}$ and red
782: for M$\geq 5.0\times 10^{14} M_{\odot}$. These give a rough indication of
783: the expected redshift distribution of identified clusters in upcoming
784: surveys.  
785: \begin{figure}
786: \includegraphics[width=0.5\textwidth]{f6.eps}
787: \caption{Angular density of clusters in the light cone images as a
788:   function of redshift. Black solid line is the redshift distribution
789:   of clusters with M$\geq 10^{14} M_{\odot}$. Blue solid line is for
790:   M$\geq 3.0 \times 10^{14} M_{\odot}$, red is for M$\geq 5.0\times 10^{14} M_{\odot}$. Dotted lines are
791:   90\% variance of 200 independent stacking realizations of the light cone.}
792: \label{fzplot}
793: \end{figure}
794: 
795: For one projected light cone image, we show the
796: Y vs M relationship in Figure \ref{yvm}. In this case, $Y$ is
797: calculated by integrating the value of Compton $y$ in the image out to
798: each cluster's \textit{projected} virial radius. The value of $Y$ is corrected for redshift since it
799: depends on E(z)$^{-2/3}$ \citep[see][]{nagai} due to the cosmological
800: dependence of the cluster M-T relation, and the angular scale is converted to Mpc
801: through use of the value of $D_A$ for each cluster. Contrast this plot with Figure \ref{yvmtrue}, for a single
802: realization of the light cone image, where the true projected Y is plotted against
803: cluster true mass from the simulation. The true value for Y is
804: calculated from the projected spherically averaged radial profiles of
805: each simulated cluster, but includes the gas out only to the
806: virial radius \textit{in three dimensions}. This true relation has
807: extraordinarily tight scatter, as has been shown previously
808: \citep{dasilva, motl05, nagai}. The difference in the two plots is
809: effectively the difference between the cluster's true integrated SZE
810: and the SZE integrated in a cylinder with radius equal to the cluster's
811: virial radius. For a narrow mass bin around 3.0$\times$10$^{14}
812: M_{\odot}$, the median bias in Y due to projection is 79\%, and the
813: scatter is +32\%/-16\% about that median, a significant increase over
814: the scatter in the true Y-M relationship. Note that some clusters in Figure \ref{yvm} appear to
815: have lower value of Y in projection than the true value for that
816: cluster. These are the clusters which lie near the edge of the
817: simulated survey image, and the cluster extends beyond the image
818: edge. As identified by other studies,
819: errors in extracting the correct value of Y should dominate the error
820: budget for this relation \citep[e.g.,][]{white02,melin}. These figures
821: illustrate the difficulty in accurately estimating Y from observations. A similar result is noted by
822: \citet{white02} using data from a smoothed particle hydrodynamics (SPH)
823: simulation. We have corrected the primary object in the
824: source region for cosmological evolution in the M-T relation, so the
825: remaining scatter results from line of sight projection effects. This
826: means that there are secondary bound objects (and unbound gas) which are projected
827: into the cluster's source region as will be described in the next
828: section. 
829: \begin{figure}
830: \includegraphics[width=0.5\textwidth]{f7.eps}
831: \caption{Projected integrated Y value plotted against cluster mass. Y
832:   is integrated from the simulated light cone survey images from the
833:   center of each cluster out to the projected virial radius. 
834:   Y is converted from angular units using the angular diameter
835:   distance appropriate to the redshift of the matched simulated
836:   cluster. Y is also scaled with E(z)$^{-2/3}$ to account for the
837:   cosmological variation of the mass scaling relation.}
838: \label{yvm}
839: \end{figure}
840: \begin{figure}
841: \includegraphics[width=0.5\textwidth]{f8.eps}
842: \caption{Integrated Compton $y$ vs Total mass relationship extracted
843:   directly from the simulation data. Y is the projected integrated SZE
844:   $y$ parameter from each cluster out the the virial radius, total mass is
845:   from the simulation grid for each object. Y is also scaled with E(z)$^{-2/3}$ to account for the
846:   cosmological variation of the mass scaling relation.}
847: \label{yvmtrue}
848: \end{figure}
849:  
850: \subsection{Confusion Problems}
851: In this section we address the confusion resulting from clusters,
852: groups, and smaller mass halos, as opposed to confusion resulting from
853: radio and millimeter wave point sources in the cluster fields. This
854: work also includes the additional flux contributed by unbound gas, which contributes at some level in
855: the real universe, but has been ignored in most simulations of SZE
856: surveys. While there are several definitions of
857: confusion in the literature, even for SZE surveys, we have chosen to
858: define cluster confusion as the number of true cluster- or group-mass objects
859: in the simulation whose centers are within the source
860: region in projection. Since upcoming SZE surveys are unlikely to
861: detect cluster gas out to the virial radius, we have chosen a smaller
862: radius ($r_{500}$) as the source region. For each of the upcoming
863: surveys, we have defined the source region as a radius of a full beam
864: diameter, presuming that if two objects were imaged by two separate
865: non-overlapping beams, that there is a possibility they would not be
866: confused. This also presumes that the secondary object is bright
867: enough to be detected on its own. One can also define confusion as an error in recovered
868: flux \citep{holder} from sources found by progressive matched filtering at
869: different angular scales corresponding to variation of the angular size
870: of clusters as a function of redshift \citep[e.g.,][]{melin}. 
871: 
872: Whether this type of confusion can be mitigated
873: depends on a variety of factors, including the mass of the additional
874: secondary objects in the source region, redshift distribution and
875: angular scale of the objects and the observing beam and multiple
876: wavelength identfication of the objects (e.g., optical, X-ray,
877: lensing). We endeavor here then simply to characterize the amount of
878: said confusion, leaving the mitigation of this problem to future
879: work. Figure \ref{all_source} shows for each of the surveys considered
880: a histogram of the number of objects above 10$^{14}
881: M_{\odot}$ in the source regions (as defined above) of all clusters
882: with M $\geq 3.0\times10^{14} M_{\odot}$. These mass limits are chosen
883: somewhat arbitrarily, though for the ground-based surveys,
884: $3.0\times10^{14} M_{\odot}$ is close
885: to the expected mass limit. $10^{14} M_{\odot}$ is chosen for the
886: contaminating objects on the expectation that one or more object above
887: that mass in the source region should lead to a significant bias in
888: the SZE flux from that expected from a single cluster in the
889: detectable mass range. As in Section 4.3, we have identified each
890: cluster by its projected image plane position associated with its true
891: three-dimensional location in the simulation after stacking.  It is clear that for Planck, there are a
892: high number of objects per source region above this relatively high
893: mass. In fact, roughly 40\% of source regions have only one object above
894: this mass cut, while nearly 60\% have more than one. Contrast this
895: result with that of APEX/SPT, where 95\% of source regions have only one
896: massive cluster projected into them. This result is
897: clearly correlated directly with beam size, but also depends weakly on
898: sensitivity, since a deeper survey will lead to larger source regions
899: (when limited by signal-to-noise) and will increase the chance of
900: objects overlapping into the source, and can also increase
901: ``bridging'' between sources. Also, it is likely that SZE experiments
902: will sometimes detect a source where there is no cluster in the mass
903: range at that location (spurious detections). Therefore our estimates
904: shown in Figure \ref{all_source} may be optimistic. 
905: \begin{figure}
906: \includegraphics[width=0.5\textwidth]{f9.eps}
907: \caption{Histogram of number of simulated clusters with M $>10^{14}
908:   M_{\odot}$ per source region for each of survey modified light
909:   cones. Source region is defined as inside $r_{500}$ projected for the solid
910:   line, and as within 1 beam diameter (at 144 GHz) distance for each of the survey
911:   instruments listed. Cluster identifications for all 200 light cones are tallied
912:   and used in the fraction.}
913: \label{all_source}
914: \end{figure}
915:  
916: %---------------------------------------------------------------------------
917: \subsection{Contribution of Unresolved Halos and Unbound Gas}\label{sec:flux_break}
918: 
919: One possibly significant difference between this and other similar
920: simulation-based studies is the inclusion of adiabatic hydrodynamics
921: in addition to N-body dynamics in the calculation. This results in
922: several advantages over N-body calculations that include the effects
923: of baryons in the post-processing phase. The first is that our clusters
924: need not be in hydrostatic equilibrium (HSE) which is a standard
925: assumption in dark matter-only simulations which have been post-processed. Both simulations and
926: observations indicate that hydrostatic equilibrium is not a safe
927: assumption for many clusters \citep[see e.g.,][]{rasia,1E06}. A significant amount of scatter in cluster
928: observables results from the disequilibrium caused by mergers
929: \citep{roettmass, ricker, randall}. This
930: scatter is absent in N-body + HSE type studies, and is naturally
931: included in our work. Secondly, our simulations include baryons
932: which are outside virialized objects, including gas in filaments and voids. The
933: SZE in particular is sensitive to this additional gas, since the
934: effect is only linear in the gas density and is relatively redshift
935: independent. Thus any gas along the line of sight contributes to the
936: SZE integral, and is not diminished by distance. 
937: 
938: \begin{figure*}[t]
939: \begin{center}
940: \plottwo{f10a.eps}{f10b.eps}
941: \end{center}
942: \caption{Left panel: 100 deg$^2$ projected light cone image of the
943:   Compton y-parameter from a $512^3$ \textit{Enzo} simulation of a
944:   $(512h^{-1}$ Mpc$)^3$ volume with 7 dynamic levels of refinement. Light
945:   cone includes tiles at 27 discrete redshift intervals between z=3 and z=0.1. Right
946: panel: Same image as left panel, but with clusters with M $\geq 5 \times
947:   10^{13} M_{\odot}$ cut from the data. Roughly one third of the total flux in
948:   the image comes from the objects that remain after the removal of
949:   the massive clusters, including poor groups and filaments. We
950:   predict that such observations could provide the first detection of
951:   the WHIM gas over large sky areas.}
952: \label{sz_lc}
953: \end{figure*} 
954: While several authors have noted that the angular power in the SZE
955: from the cluster subtracted field is small \citep[see,
956: e.g.,][]{holder}, it is not necessarily true that the total SZE flux
957: (or decrement) from unresolved halos and unbound gas is negligible.  Figure \ref{sz_lc} shows an image of the full field
958: of light cone in projected Compton $y$ parameter next to an
959: image of the field where clusters with M $> 5.0\times 10^{13}M_{\odot}$ have
960: been removed from the field. We
961: show in Figure \ref{power_noc} the angular power from the SZE in the
962: cluster subtracted images compared to the power due to the full SZE
963: image. At small scales ($\ell >$ 5000), the power in the cluster
964: subtracted image is more than an order of magnitude lower than that of
965: the full image. At larger scales, however, the difference narrows, and
966: in fact the 90\% variance overlaps for the two in some regions. This
967: is likely due partly to incomplete cluster subtraction in the image.
968: \begin{figure}
969: \begin{center}
970: \includegraphics[width=0.5\textwidth]{f11.eps}
971: \end{center}
972: \caption{Angular power from 200 full light cone images (solid line) compared to
973:   angular power from images where M $\geq 5 \times
974:   10^{13}M_{\odot}$ halos are subtracted (dashed line).  Dotted lines
975:   indicate 90\% variance range for the 200 independent stacking
976:   realizations of the light cone images.}
977: \label{power_noc}
978: \end{figure}
979:  
980: A significant result from this analysis is that
981: roughly one third of the SZE flux in the image comes from objects with M
982:   $< 5.0\times 10^{13}M_{\odot}$ and filamentary structures made up of
983:   gas in the Warm-Hot Intergalactic Medium (WHIM) phase. Figure \ref{szy_histo} shows a histogram for 200 independent light cone
984: realizations from our simulations of the ratio of the total SZE flux
985: (or integrated Compton $y$ parameter) in the 100 deg$^2$ field from
986: only gas within the virial radius of those clusters to that of all other gas
987: in the field. Thus, we predict that upcoming
988: SZE instruments are the only near-term telescopes that will possibly be capable
989: of detecting WHIM over large sky areas.  This result is consistent
990: with \citet{verde}, who have performed a similar study with a fixed
991: grid N-body/hydro simulation with considerably lower peak resolution
992: (195 h$^{-1}$ kpc) than our work.
993: 
994: It is important to note that in an adiabatic simulation, gas fractions
995: are relatively constant with cluster/group mass. In the real universe,
996: as well as in more realistic simulations such as those we performed
997: for \citet{hall06}, the ICM gas radiatively cools and forms stars,
998: lowering the gas fraction of the cluster, and effectively attenuating
999: its SZE signal. This effect is also cluster-mass dependent, lower mass
1000: objects have comparatively lower gas fractions. Within a
1001: simulation with radiative cooling and star formation prescriptions, we find gas
1002: fractions 30-50\% lower in clusters with M $\approx 10^{14} M_{\odot}$
1003: than the average value for clusters with M $> 3\times 10^{14}
1004: M_{\odot}$. Additionally, recent observational studies
1005: \citep[e.g.,][]{mcc07} point out that
1006: gas fractions deduced from X-ray data decrease with cluster temperature (therefore with mass),
1007: by of order 50\% for 1-2 keV clusters from a flat value above $\sim$4
1008: keV.  Therefore we expect our result here to be an upper limit
1009: on the amount of flux from low mass objects and WHIM gas, possibly
1010: above the true value by as much as a factor of two. Our future light
1011: cone simulations will be run with non-gravitational physics, and we
1012: will explore the effect on the SZE background. 
1013: \begin{figure}
1014: \begin{center}
1015: \includegraphics[width=0.45\textwidth]{f12.eps}
1016: \end{center}
1017: \caption{Histogram of the ratio of total flux in the SZE y
1018:     parameter images from clusters with M $\geq 5 \times
1019:   10^{13}M_{\odot}$ and from images with clusters subtracted. The histogram is
1020:     generated from 200 independent realizations of the
1021:     light cone using the same simulation. Roughly two thirds of the flux in the
1022:     image comes from clusters with M$\geq 5 \times
1023:   10^{13}M_{\odot}$, and the other third comes from the WHIM and poor groups.}
1024: \label{szy_histo}
1025: \end{figure}
1026:  
1027: \subsection{Contribution per Source of Cluster Subtracted Images}
1028: Since unresolved halos (M $< 5 \times
1029:   10^{13} M_{\odot}$) and unbound gas in this simulation clearly contribute flux to
1030:   the image, it is important to ask how much additional
1031:   flux per source is added.  This extra flux is a bias, in that
1032:   it is always additive. Therefore it should boost the emission of all
1033:   clusters in the field by some amount which may vary from cluster to
1034:   cluster, adding both bias and scatter to the cluster SZE
1035:   observable. This effect is critical to understand, since photometric
1036:   accuracy of the SZE in clusters is key to calibrating a Y-M
1037:   relationship which will be useful in determining cosmological paramteters. The precision of the calibration
1038:   of the Y-M relationship depends strongly on its scatter
1039:   \citep{melin}. 
1040: 
1041:   Figure \ref{whim_ratio} shows the ratio of the integrated Compton $y$
1042:   parameter inside a cylinder of radius $r_{500}$ around each cluster
1043:   above 10$^{14} M_{\odot}$ from the
1044:   cluster subtracted image to the value for the full image. The plot
1045:   shows values for clusters in a single light cone image, but there is
1046:   very little variation in the mean, median and scatter across all the
1047:   realizations. The large scatter is partly a result of
1048:   incomplete subtraction of the cluster SZE signature from the image,
1049:   but also results from variations in large scale structure in the
1050:   various source regions. What is clear from the plots is that there
1051:   is a systematic bias in integrated Compton $y$ resulting from low
1052:   mass objects and unbound gas. The mean value of this ratio is $16.3^{+7.0}_{-6.4}$\%. The
1053:   1$\sigma$ scatter is $\pm$30-40\%. It is unclear
1054:   that this bias is reducible, since identifying objects of very low
1055:   mass in source fields, particularly at high redshift, seems
1056:   prohibitively difficult. Additionally, making direct observation of
1057:   filamentary gas, particularly to locate its position on the sky, has
1058:   been nearly impossible. 
1059: \begin{figure}
1060: \begin{center}
1061: \epsscale{.9}
1062: \includegraphics[width=0.5\textwidth]{f13.eps}
1063: \end{center}
1064: \caption{Ratio of integrated Compton $y$ inside a cylinder projected to $r_{500}$ from the
1065:   cluster-subtracted image to integrated Compton $y$ from the
1066: same projected cylinder in the full light cone image. Includes all
1067:   clusters projected into the survey image which have M $\geq 10^{14}
1068:   M_{\odot}$. }
1069: \label{whim_ratio}
1070: \end{figure}
1071:  
1072: There will be some variation from survey to survey in this
1073: additional flux. Depending on how much of the cluster's radial extent
1074: is sampled, the mean value for all clusters will change. The scatter though,
1075: is quite large, which means that accounting for this flux is not as
1076: simple as removing a uniform background from each cluster's SZE
1077: signal. In our study we also performed the identical analysis for the
1078: clusters, assuming detection out to $r_{2500}$, and found no change in
1079: the scatter, though the mean value of the additional flux dropped to
1080: roughly 8-10\%. As discussed in previous sections, a more realistic
1081: modeling of the heating and cooling in the ICM should result in a
1082: reduction in this additional flux (or decrement) by as much as
1083: 30-50\%. Even with a reduction of this size, we still expect it to be
1084: a few percent to 10\% effect with of order $\pm$30-40\% scatter, creating
1085: challenges for a percent-level calibration of the Y-M relationship. 
1086: 
1087: 
1088: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1089: \section{Discussion and Summary}\label{sec:discuss}
1090: In this study, we have taken an important step missing from previous work in the
1091: literature on characterizing selection functions of SZE surveys. In
1092: earlier work, investigators have attempted to evaluate methods of
1093: removing contamination of the galaxy cluster SZE signal due to the CMB
1094: and other point sources and backgrounds. However, they have included
1095: only the baryons present in clusters and groups, and artificially
1096: inserted the gas in hydrostatic equilibrium with the dark matter
1097: distribution from an N-body only simulation. Here, we have explored
1098: the often neglected contribution of gas in low mass halos and unbound
1099: filamentary gas in aggregate to determine the effect on the cluster
1100: signal at a single frequency. 
1101: 
1102: The presence of gas outside the cluster virial radii in low density
1103: structures such as filaments is potentially an important
1104: contributor to the cluster SZE signal. This gas is completely absent in
1105: non-hydrodynamic treatments, but appears
1106: naturally in our calculation. Since the SZE does not diminish with
1107: distance and results from a line of sight integral of the gas
1108: pressure, the sum of all the flux from unbound gas could be a
1109: significant contributor to the total flux in any cluster
1110: detection. Additionally, our simulation includes
1111: the cluster gas in the full array of dynamically active states. Though
1112: the integrated SZE is less sensitive to cluster-cluster merging than an X-ray
1113: observation would be, merging contributes non-zero scatter to the Y-M
1114: relationship. Thus, for all these reasons it is critically important to self-consistently include
1115: baryons in numerical simulations in order to properly simulate sky surveys.  
1116: 
1117: We have shown that on 100 square degree patches of the sky
1118: at $l < 2000$ or so the deduced power can be different by factors of
1119: 5-8. This indicates that the power spectrum can be quite different
1120: from one area of the sky to another, and clearly requires much larger
1121: angular areas to be well constrained. The effect of cosmic variance does
1122: not become very small until $\ell > $ a few thousand. 
1123: 
1124: We have shown that projection effects can create a large bias and
1125: additional scatter in the value of Y measured for clusters of
1126: galaxies. For clusters of M $\approx$3.0$\times$10$^{14}
1127: M_{\odot}$, the median bias in Y due to projection is 79\%, and the
1128: scatter is +32\%/-16\% about that median, a significant increase over
1129: the scatter in the true Y-M relationship. Additionally, we have shown
1130: that the contribution of low-mass unresolved
1131: halos and unbound gas to the flux (or decrement) of identified sources can be
1132: significant in some cases, and certainly varies widely from source to
1133: source. We find that there
1134: is a contribution from gas outside clusters of $16.3^{+7.0}_{-6.4}$\% per object
1135: on average for upcoming surveys. This indicates both a bias and an additional source of scatter
1136: in the determination of the true SZE signal from any given cluster. As identified by other studies,
1137: errors in extracting the correct value of Y should dominate the error
1138: budget for the Y-M relation. This effect is critical to understand,
1139: since photometric accuracy of the SZE in clusters is key to calibrating a Y-M
1140: relationship which can be useful for the precision determination of
1141: cosmological parameters. While the intrinsic Y-M relation
1142: has very small scatter, what matters in practice is the ability to
1143: determine the value of Y accurately. The precision of the calibration
1144: of the Y-M relationship depends strongly on its scatter. 
1145: 
1146: We also show results from an analysis of the source confusion for each
1147: instrument based on how many massive (M $>10^{14}$ M$_{\odot}$)
1148: clusters lie within an identified source region (within a radius of
1149: $r_{500}$ or a beam diameter at 144 GHz). It
1150: appears that pure cluster/group confusion in these surveys will be a
1151: significant problem, particularly for Planck Surveyor, but also to a
1152: lesser extent for the other single-dish surveys. Smaller
1153: beam surveys ($\sim$1$\arcmin$) have more than one massive
1154: cluster within a beam diameter 5-10\% of the time, and a larger beam
1155: experiment like Planck has multiple clusters per beam 60\% of the
1156: time. We may have slightly
1157: overestimated the problem, since the use of higher resolution (shorter
1158: wavelength) bands on some of the survey instruments will help to
1159: alleviate this issue. On the other hand, we have not accounted for
1160: spurious detections which may result in a field with real backgrounds
1161: and instrument noise. Whether this type of confusion can be mitigated
1162: depends on a variety of factors, including the mass of the additional
1163: secondary objects in the source region, the redshift distribution and
1164: angular scale of the objects and the observing beam, and multiple
1165: wavelength identfication of the objects (e.g., using optical, X-ray,
1166: and lensing measurements).
1167: 
1168: This study uses a large volume, high resolution adiabatic
1169: simulation, which serves as a template for more complex runs involving
1170: additional non-gravitational physics that is likely important to
1171: accurately modeling cluster surveys. There remain important
1172: differences between simulation outputs and cluster observations, particularly
1173: for lower mass clusters, which indicates the need for a better
1174: understanding of the details of baryonic cluster physics. In addition, deviations from isothermality and
1175: hydrostatic equilibrium in the cluster gas can have a strong impact on
1176: both the observable and derived properties of clusters. 
1177: It is also important to note that there is some dependence of the SZE signal
1178: on the details of the ICM physics (heating, cooling, conduction, etc.)
1179: which is not modeled in this work. In future work, we will explore the
1180: impact of this additional physics, modeled self-consistently within
1181: the hydrodynamic framework of the simulation code, on a selection of SZE
1182: clusters from surveys. 
1183: 
1184: This work will be expanded in future papers by a detailed treatment of
1185: the point source confusion and instrumental and observing
1186: limitations. This will include adding to our synthetic surveys the
1187: contribution of the CMB, dusty galaxies, AGN, and atmospheric
1188: foreground. We are now working on modeling these, in particular for
1189: APEX-SZ and SPT, and will then experiment with matched filtering and
1190: the use of multiwavelength coverage provided by SPT for example to
1191: mitigate confusion and remove the atmospheric and CMB signals. Matched
1192: filtering is a process by which the images are filtered with a kernel
1193: matched to the presumed size and shape of the expected sources in the
1194: field. In the case of the SZE, this procedure filters out information
1195: on larger scales where the primary CMB anisotropies will be a source
1196: of confusion. It also maximizes the contrast of the image for the
1197: objects at that scale. Since the angular scale subtended by massive
1198: clusters is a function of redshift and a weak function of cluster
1199: mass, spatial filtering will need to be done at a variety of angular
1200: scales to get a complete cluster sample as in \citet{melin}. A similar type
1201: of analysis has been performed by \citet{sehgal} for ACT's survey. 
1202: 
1203: We also are currently performing additional synthetic light cone sky
1204: surveys at X-ray wavelengths, to take a first look at the limitations
1205: of current (e.g. XMM-LSS) and upcoming (e.g. eRosita) X-ray surveys in
1206: extracting cosmological parameters. 
1207: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1208: \acknowledgments
1209: BWO and MLN have been supported in part by NASA
1210: grant NAG5-12140 and NSF grant AST-0307690. 
1211: BWO has been funded in part
1212: under the auspices of the U.S.\ Dept.\ of Energy, and supported by its
1213: contract W-7405-ENG-36 to Los Alamos National Laboratory.  The
1214: simulations were by performed at SDSC and NCSA with computing time provided by 
1215: NRAC allocation MCA98N020. EJH and JOB have been supported in part by a grant 
1216: from the U.S. National Science Foundation (AST-0407368). EJH also
1217: acknowledges support from NSF AAPF AST-0702923. We thank Yoel Rephaeli for useful discussions on the
1218: SZE angular power spectrum. We thank John Carlstrom, Nils Halverson, Jeremiah Ostriker
1219: ,Douglas Rudd and an anonymous referee for useful comments and critiques. We thank the Aspen Center
1220: for Physics for hosting three of the authors during the writing of this manuscript,
1221: and also thank Maria's New Mexican Kitchen in Santa Fe for providing an atmosphere conducive
1222: to the conception of this project.
1223: 
1224: 
1225: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1226: \bibliographystyle{apj}
1227: %\bibliography{ms}  % looks in ms.bib for bibliography info
1228: \input{ms.bbl}
1229: \end{document}  
1230: