1: %\documentclass[apj]{emulateapj}
2: \documentclass[12pt,preprint]{aastex}
3:
4: \usepackage{amsmath,epsf,multicol,natbib,graphicx}
5: \usepackage[alpha,z,expert]{tmsfnts}
6: \usepackage{ulem}
7: \usepackage{color}
8: \usepackage{pdfcolmk}
9: \bibliographystyle{apj}
10:
11: \definecolor{DarkGreen}{rgb}{0.0, 0.5, 0.0}
12: \definecolor{purple}{rgb}{0.5, 0.0, 0.5}
13: \definecolor{red}{rgb}{1, 0.0, 0.0}
14: \definecolor{green}{rgb}{0, 1.0, 0.0}
15:
16: %% Definitions of useful commands
17: \newcommand{\sigT}{\mbox{$\sigma_{\mbox{\tiny T}}$}}
18: \newcommand{\Tcmb}{\mbox{$T_{\mbox{\tiny CMB}}$}}
19: \newcommand{\kB}{\mbox{$k_{\mbox{\tiny B}}$}}
20: \newcommand{\nH}{\mbox{$n_{\mbox{\tiny H}}$}}
21: \newcommand{\NH}{\mbox{$N_{\mbox{\tiny H}}$}}
22: \newcommand{\LameH}{\mbox{$\Lambda_{e \mbox{\tiny H}}$}}
23: \newcommand{\Lamee}{\mbox{$\Lambda_{ee}$}}
24: \newcommand{\rhogas}{\mbox{$\rho_{\mbox{\scriptsize gas}}$}}
25: \newcommand{\Mgas}{\mbox{$M_{\mbox{\scriptsize gas}}$}}
26: \newcommand{\Mtot}{\mbox{$M_{\mbox{\scriptsize tot}}$}}
27: \newcommand{\Yint}{\mbox{$Y_{\mbox{\scriptsize int}}$}}
28: \newcommand{\Ycyl}{\mbox{$Y_{\mbox{\scriptsize cyl}}$}}
29: \newcommand{\Ysph}{\mbox{$Y_{\mbox{\scriptsize sph}}$}}
30: \newcommand{\fgas}{\mbox{$f_{\mbox{\scriptsize gas}}$}}
31: \newcommand{\LCDM}{\mbox{$\Lambda$CDM}}
32:
33: \begin{document}
34: \normalem
35:
36: \title{\bf Application of a Self-Similar Pressure Profile to
37: Sunyaev-Zel'dovich Effect Data from Galaxy Clusters}
38:
39: \author{
40: Tony~Mroczkowski,\altaffilmark{1,2,3}
41: Max~Bonamente,\altaffilmark{4,5}
42: John~E.~Carlstrom,\altaffilmark{6,7,8,9}
43: Thomas~L.~Culverhouse,\altaffilmark{6,7}
44: Christopher~Greer,\altaffilmark{6,7}
45: David~Hawkins,\altaffilmark{10}
46: Ryan~Hennessy,\altaffilmark{6,7}
47: Marshall~Joy,\altaffilmark{5}
48: James~W.~Lamb,\altaffilmark{10}
49: Erik~M.~Leitch,\altaffilmark{6,7}
50: Michael~Loh,\altaffilmark{6,9}
51: Ben~Maughan,\altaffilmark{11,12}
52: Daniel~P.~Marrone,\altaffilmark{6,8,13}
53: Amber~Miller,\altaffilmark{1,14,15}
54: Stephen~Muchovej,\altaffilmark{2}
55: Daisuke~Nagai,\altaffilmark{16,17}
56: Clem~Pryke,\altaffilmark{6,7,8}
57: Matthew~Sharp,\altaffilmark{6,9}
58: and David~Woody\altaffilmark{10}}
59:
60: \altaffiltext{1}{Columbia Astrophysics Laboratory, Columbia University, New York, NY 10027}
61: \altaffiltext{2}{Department of Astronomy, Columbia University, New York, NY 10027}
62: \altaffiltext{3}{Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104}
63: \altaffiltext{4}{Department of Physics, University of Alabama, Huntsville, AL 35899}
64: \altaffiltext{5}{Department of Space Science, VP62, NASA Marshall Space Flight Center, Huntsville, AL 35812}
65: \altaffiltext{6}{Kavli Institute for Cosmological Physics, University of Chicago, Chicago, IL 60637}
66: \altaffiltext{7}{Department of Astronomy and Astrophysics, University of Chicago, Chicago, IL 60637}
67: \altaffiltext{8}{Enrico Fermi Institute, University of Chicago, Chicago, IL 60637}
68: \altaffiltext{9}{Department of Physics, University of Chicago, Chicago, IL 60637}
69: \altaffiltext{10}{Owens Valley Radio Observatory, California Institute of Technology, Big Pine, CA 93513}
70: \altaffiltext{11}{Department of Physics, University of Bristol, Tyndall Ave, Bristol BS8 1TL, UK.}
71: \altaffiltext{12}{Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138}
72: \altaffiltext{13}{Jansky Postdoctoral Fellow, National Radio Astronomy Observatory}
73: \altaffiltext{14}{Department of Physics, Columbia University, New York, NY 10027}
74: \altaffiltext{15}{Alfred P. Sloan Fellow}
75: \altaffiltext{16}{Department of Physics, Yale University, New Haven, CT 06520}
76: \altaffiltext{17}{Yale Center for Astronomy \& Astrophysics, Yale University, New Haven, CT 06520}
77:
78:
79: \begin{abstract}
80:
81: We investigate the utility of a new, self-similar pressure profile for
82: fitting Sunyaev-Zel'dovich (SZ) effect observations of galaxy
83: clusters. Current SZ imaging instruments---such as the Sunyaev-Zel'dovich Array
84: (SZA)---are capable of probing clusters over a large
85: range in physical scale. A model is therefore required that can accurately describe a
86: cluster's pressure profile over a broad range of radii, from
87: the core of the cluster out to a significant fraction of the virial radius.
88: In the analysis presented here, we fit a
89: radial pressure profile derived from simulations and
90: detailed X-ray analysis of relaxed clusters to SZA observations of
91: three clusters with exceptionally high quality X-ray data: A1835, A1914, and CL~J1226.9+3332. From the
92: joint analysis of the SZ and X-ray data, we derive physical
93: properties such as gas mass, total mass, gas fraction and the intrinsic, integrated
94: Compton $y$-parameter.
95: We find that parameters derived from the joint fit to the SZ and X-ray data
96: agree well with a detailed, independent X-ray-only analysis of the
97: same clusters.
98: In particular, we find that, when combined with X-ray imaging data,
99: this new pressure profile yields an independent electron radial temperature
100: profile that is in good agreement with spectroscopic X-ray
101: measurements.
102:
103: \end{abstract}
104:
105: \keywords{cosmology: observations --- clusters: individual (Abell 1835, Abell 1914, CL~J1226.9+3332)
106: --- Sunyaev-Zel'dovich Effect}
107:
108: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
109:
110: \section{Introduction}
111:
112: The expansion history of the universe and the growth of large-scale
113: structure are two of the most compelling topics in cosmology. As
114: galaxy clusters are the largest collapsed objects in the universe,
115: taking a Hubble time to form, their abundance and evolution are
116: critically sensitive to the details of that expansion history.
117: Cluster surveys can therefore provide fundamental clues to the nature
118: and abundance of dark matter and dark energy \citep[see,
119: e.g.,][]{white1993b,frenk1999,haiman2001,weller2002}.
120:
121: While clusters have been extensively studied using X-ray observations
122: of the hot gas that comprises the intracluster medium (ICM), radio
123: measurements of the same gas via the Sunyaev-Zel'dovich (SZ) effect \citep{sunyaev1972}
124: provide an independent and complementary probe of the ICM \cite[e.g.,][]{carlstrom2002}. The SZ effect
125: arises from inverse Compton scattering of cosmic microwave background
126: (CMB) photons by the electrons in the ICM, imparting a detectable
127: spectral signature to the CMB that is independent of the redshift of
128: the cluster. As the SZ effect is a measure of the integrated line-of-sight
129: electron density, weighted by temperature, i.e., the integrated
130: pressure (see \S~\ref{xray_sze}), it probes different properties than
131: the X-ray emission, which is proportional to the square of the
132: electron density. Cluster surveys exploiting the redshift independence
133: of the SZ effect are now being conducted by a variety of instruments,
134: including the SZA \citep{muchovej2007}, the South Pole Telescope
135: \citep{ruhl2004}, the Atacama Cosmology Telescope \citep{kosowsky2003}
136: and the Atacama Pathfinder Experiment (APEX) SZ instrument \citep{dobbs2006}.
137:
138: To maximize the utility of clusters as cosmological probes we must
139: understand how to accurately relate their observable properties to
140: their total masses. The integrated SZ signal from a cluster is
141: proportional to the total thermal energy of the cluster, and is
142: therefore a measure of the underlying gravitational potential, and
143: ultimately the dark matter content, within a cluster.
144: The SZ flux of a cluster thereby should provide a robust, low scatter
145: proxy for the total cluster mass, \Mtot\
146: \citep[see, for example,][]{dasilva2004,motl2005,nagai2006,reid2006}.
147:
148: To date, cosmological studies combining SZA and X-ray data have relied
149: almost exclusively on the isothermal $\beta$-model
150: \citep[first used in][]{cavaliere1976,cavaliere1978} to fit the SZ signal in the
151: region interior to $r_{2500}$
152: \citep[see][for applications of the isothermal $\beta$-model to SZ+X-ray data]{grego2000,
153: reese2002,laroque2006,bonamente2006,bonamente2008}.
154: Here $r_{2500}$ is the radius within which the mean cluster density is a
155: factor of 2500 over the critical density of the universe at the
156: cluster redshift. While the isothermal $\beta$-model recovers global
157: properties of clusters quite accurately in this regime
158: \citep{laroque2006}, deep X-ray observations of nearby
159: clusters show that isothermality is a poor description of the cluster
160: outskirts ($r\sim r_{500}$) \citep[see e.g.][and references therein]{piffaretti2005,vikhlinin2005a,pratt2007}. An improved model
161: for the cluster gas, accurate to large radii, is therefore critical for
162: the analysis and cosmological interpretation of SZ data obtained with
163: new instruments that are capable of probing the outer regions ($r \sim r_{500}$) of clusters.
164: Such a model must be simple enough that it can be constrained by SZ
165: data with limited angular resolution and sensitivity typical of data sets acquired by SZ survey
166: instruments optimized for detection, rather than imaging. While the
167: $\beta$-model has the virtue of simplicity, previous attempts to relax the
168: assumption of isothermality typically required high-significance,
169: spatially-resolved X-ray spectroscopy; such data are seldom obtained in
170: short X-ray exposures of high-redshift clusters. This is particularly
171: true for the cluster outskirts \citep[see, e.g.,][]{laroque2006}.
172: Attempts to move beyond the $\beta$-model have typically improved the
173: modeling of the gas density only within the core ($r\lesssim0.15\,r_{500}$).
174:
175: In this work, we investigate a new model for the cluster gas pressure
176: by using it to fit SZ data from the SZA and X-ray data from {\em Chandra}
177: to three well-studied clusters: Abell 1835, Abell 1914, and CL~J1226.9+3332. The model was
178: derived by \citet{nagai2007b} from simulations and from detailed
179: analysis of deep {\it Chandra} measurements of nearby relaxed
180: clusters. The simplicity of this model---and the fact that SZ data
181: are inherently sensitive to the integrated electron pressure---allow
182: it to be used either in conjunction with X-ray imaging data, or fit to
183: SZ data alone. The outline of the paper is as follows: in
184: \S~\ref{modeling}, we present the details of the model, and couple it
185: with an ICM density model that allows the inclusion of X-ray imaging
186: data. In \S~\ref{realobs}, we apply this method to three clusters,
187: combining new data from the SZA with {\em Chandra} X-ray imaging
188: data. We demonstrate the utility of this model by applying it to
189: SZ+X-ray data without relying on X-ray spectroscopic information.
190: Results from the joint SZ+X-ray analysis are then compared to results
191: from an X-ray-only analysis, including spectroscopic data, in \S~\ref{results}.
192: We offer our conclusions in \S~\ref{conclusions}.
193:
194:
195: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
196:
197: \section{Cluster Gas Models}\label{modeling}
198:
199: \subsection{Sunyaev Zel'dovich Effect and X-ray Emission}\label{xray_sze}
200:
201: The thermal SZ effect is a small ($< 10^{-3}$) distortion
202: in CMB intensity caused by inverse Compton scattering of CMB
203: photons by energetic electrons in the hot intracluster gas
204: \citep{sunyaev1970,sunyaev1972}. This spectral distortion can be
205: expressed, for dimensionless frequency $x \equiv h\nu/\kB \Tcmb$, where
206: $h$ is Planck's constant, $\nu$ is frequency, $k_B$ is Boltzmann's constant, and
207: \Tcmb\ is the primary CMB temperature, as
208: the change $\Delta I_{SZ}$ relative to the primary CMB intensity normalization $I_0$,
209: \begin{eqnarray}
210: \label{eq:thermal_sz}
211: \frac{\Delta I_{SZ}}{I_0} &=& \frac{k_B \, \sigma_T}{m_e c^2} \int \!\! g(x,T_e) \, n_e T_e \,d\ell \\
212: \label{eq:thermal_sz2}
213: &=& \frac{\sigma_T}{m_e c^2} \int \!\! g(x,T_e) \, P_e \,d\ell.
214: \end{eqnarray}
215: Here \sigT\ is the Thomson scattering cross-section of the electron,
216: $\ell$ is the line of sight, and $m_e c^2$ is an electron's rest energy.
217: The factor $g(x,T_e)$ encapsulates the frequency dependence of the SZ effect intensity.
218: For non-relativistic electrons,
219: \begin{equation}
220: \label{eq:g_x}
221: g(x) = \frac{x^4 e^x}{(e^x-1)^2} \left(x \frac{e^x + 1}{e^x - 1} - 4\right).
222: \end{equation}
223: We use the \citet{itoh1998} relativistic corrections to Eq.~\ref{eq:g_x}, which are appropriate
224: for fitting thermal SZ observations, and discuss their impact on the fit ICM profiles in \S~\ref{profiles}.
225: Note that we have used the ideal gas law ($P_e = n_e k_B T_e$)
226: to obtain Eq.~\ref{eq:thermal_sz2} from Eq.~\ref{eq:thermal_sz}; we use
227: this to relate SZ intensity directly to the ICM electron pressure.
228:
229: The X-ray emission from massive clusters arises
230: predominantly as thermal bremsstrahlung from electrons. The
231: X-ray surface brightness produced by a cluster at redshift $z$ is given by \citep[e.g.,][]{sarazin1988}
232: \begin{equation}
233: \label{eq:xray_sb}
234: S_X = \frac{1}{4\pi (1+z)^4} \! \int \!\! n_e^2 \Lamee(T_e,Z) \,d\ell,
235: \end{equation}
236: where the integral is evaluated along the line of sight and the X-ray
237: cooling function $\Lamee$ is proportional to $T_e^{1/2}$. The X-ray
238: emission is therefore proportional to the square of the gas density,
239: with a relatively weak dependence on $T_e$. Separate spectroscopic
240: observations of X-ray line emission can be used to measure the gas
241: temperature. Throughout this work, we use the Raymond-Smith plasma
242: emissivity code \citep{raymond1977} with constant metallicity $Z=0.3 Z_{\odot}$,
243: yielding $\mu_e=1.17$ and $\mu = 0.61$ for the mean molecular weight
244: of the electrons and gas, respectively, using the elemental abundances provided by \citet{anders1989}.
245: The SZ and X-ray observables $\Delta I_{SZ}$ and $S_X$ are computed by evaluating the integrals
246: in Eqs.~\ref{eq:thermal_sz2} \& \ref{eq:xray_sb} numerically.
247:
248:
249: \subsection{Three-dimensional Models of the ICM Profiles}
250: \label{models}
251:
252: %% Pressure profile
253: In this work, we adopt an analytic parameterization of the cluster radial
254: pressure profile proposed by \citet{nagai2007b}~(hereafter N07),
255: %
256: \begin{equation}
257: P_e(r) = \frac{P_{e,i}}{(r/r_p)^c
258: \left[1+(r/r_p)^a\right]^{(b-c)/a}},
259: \label{eq:press}
260: \end{equation}
261: %
262: where $P_{e,i}$ is a scalar normalization of the pressure profile,
263: $r_p$ is a scale radius (typically $r_p \approx r_{500}/1.3$), and the
264: parameters $(a,b,c)$ respectively describe the slopes at intermediate
265: ($r\approx r_p$), outer ($r>r_p$), and inner ($r \ll r_p$) radii.
266: Note that Eq.~\ref{eq:press} is a generalization of the analytic
267: fitting formula obtained in numerical simulations as a parameterization of
268: the distribution of mass in a dark matter halo \citep[][NFW profile]{navarro1997}.
269: This choice is reasonable because
270: the gas pressure distribution is primarily determined by the dark
271: matter potential. The use of a parameterized
272: pressure profile is further motivated by the fact that self-similarity
273: is best preserved for pressure, as demonstrated by the low
274: cluster-to-cluster scatter seen when using these parameters to fit
275: numerical simulations. The NFW profile -- in its pure form -- has been applied by
276: \citet{atriobarandela2008} to fit the electron pressure profiles of SZ observations of clusters,
277: who demonstrated an improvement over the application of the isothermal $\beta$-model to SZ cluster studies.
278: In this work, we adopt the fixed slopes
279: $(a,b,c)=(0.9,5.0,0.4)$, which N07 found to closely match the observed profiles of
280: the \emph{Chandra} X-ray clusters and the results of numerical simulations
281: in the outskirts of a broad range of relaxed clusters.\footnote{The original values
282: published in N07 were $(a,b,c)=(1.3,4.3,0.7)$. These have recently been updated, however,
283: and will be published in an erratum to N07.}
284:
285: %% Gas density profile
286: The density model used to fit the X-ray image data is a
287: simplified version of the model employed by \cite{vikhlinin2006}~(hereafter V06) is
288: %
289: \begin{equation}
290: n_e^2(r) = n_{e0}^2\;\frac{(r/r_c)^{-\alpha}}{\left[1+(r/r_c)^2\right]^{3\beta-\alpha/2}}\\
291: \frac{1}{\left[1+(r/r_s)^\gamma\right]^{\varepsilon/\gamma}}\\
292: + \frac{n_{e02}^2}{\left[1+(r/r_{c2})^2\right]^{3\beta_2}}.
293: \label{eq:Vikhlinin}
294: \end{equation}
295: %
296: We refer to Eq.~\ref{eq:Vikhlinin}, used in the independent X-ray analysis
297: to which we compare our SZ+X-ray results, as the ``V06 density model.'' Since the cluster core
298: contributes negligibly to the SZ signal observed by the SZA, we
299: exclude the inner 100~kpc of the cluster from the X-ray images used in
300: the joint SZ+X-ray analysis. Recognizing $\alpha$ as the component introduced
301: by \cite{pratt2002} to fit the inner slope of a cuspy cluster density
302: profile, and that the second $\beta$-model component ($\beta_2$) is present
303: explicitly to fit the cluster core, we simplify the V06 density model
304: to
305: %
306: \begin{equation}
307: n_e(r) = n_{e0}
308: \left[1+(r/r_c)^2\right]^{-3\beta/2}
309: \left[1+(r/r_s)^{\gamma}\right]^{-0.5\varepsilon/\gamma},
310: \label{eq:svm}
311: \end{equation}
312: %
313: where $r_c$ is the core radius, and $r_s$ is the radius at which the
314: density profile steepens with respect to the traditional
315: $\beta$-model. Following V06, we fix $\gamma=3$, as this provides an
316: adequate fit to all clusters in the V06 sample. We refer to this
317: model as the ``Simplified Vikhlinin Model'' (SVM). Note that in the
318: limit $\varepsilon \rightarrow 0$, Eq.~\ref{eq:svm} reduces to
319: the standard $\beta$-model. The SVM is thus a
320: modification to the $\beta$-model that has the additional freedom to
321: extend from the core ($r \gtrsim r_c$) to the outer regions of cluster gas,
322: spanning intermediate ($r \gtrsim r_s$) to large radii ($r \sim r_{500}$).
323:
324: %% Gas temperature
325: With the electron pressure and density in hand, we may also derive the
326: electron temperature of the ICM using the ideal gas law,
327: $T_e(r)=P_e(r)/k_B n_e(r)$, where $P_e$ and $n_e$ are given by
328: Eqs.~\ref{eq:press} and \ref{eq:svm}, respectively. Note that $T_e(r)$
329: derived in this way is used in the analysis of X-ray surface brightness
330: (Eq.~\ref{eq:xray_sb}).
331: Hereafter, we refer to this
332: jointly-fit cluster gas model as the N07+SVM profile.
333:
334: For comparison with previous work \citep[e.g.][]{bonamente2008}, we
335: also employ the isothermal $\beta$-model for joint analysis of the SZ
336: and X-ray data. In this model the density is given by Eq.~\ref{eq:svm}
337: with $\varepsilon=0$ and $T_e(r)$ is a constant equal to the
338: spectroscopically-measured temperature, $T_X$. The shape parameters of
339: the isothermal $\beta$-model, $r_c$ and $\beta$, are jointly fit to
340: the SZ and X-ray data, while the X-ray surface brightness
341: (Eq.~\ref{eq:xray_sb}) and SZ intensity profile
342: (Eq.~\ref{eq:thermal_sz2}) normalizations are independently determined
343: from the X-ray and SZ data, respectively.
344:
345:
346: \subsection{Parameter Estimation Using the Markov Chain Monte Carlo Method \label{mcmc}}
347: Our models have five free parameters to describe the radial distribution of the gas density (see
348: Eq.~\ref{eq:svm}) and two parameters for the electron
349: pressure (see Eq.~\ref{eq:press}). Additional parameters such as the
350: cluster centroid, X-ray background level, as well as the positions, fluxes
351: and spectral indices of compact radio sources are also included
352: where necessary. The Markov chain Monte Carlo (MCMC) method is used to
353: extract the model parameters from the SZ and X-ray data, as described by \citet{bonamente2004}.
354: In this section we provide a brief overview of
355: this method, focusing on the changes to accommodate the N07 pressure model.
356:
357: The first step in fitting the SZ data is to compute the model image over
358: a regular grid, sampled at less than half the
359: smallest scale the SZA can probe. This image is multiplied by the
360: primary beam of the SZA, transformed via FFT to Fourier space (where the data
361: are naturally sampled by an interferometer; see \S~\ref{sze_visfits}), and interpolated to the Fourier-space
362: coordinates of the SZ data. The likelihood function for the SZ data is
363: then computed directly in the Fourier plane, where the noise properties of
364: the interferometric data are well-characterized.
365:
366: The first step of the MCMC method is the calculation of the joint
367: likelihood $\mathcal{L}$ of the X-ray and SZ data with the model.
368: The SZ likelihood is given by
369: \begin{equation}
370: \ln(\mathcal{L}_{SZ})=\displaystyle \sum_i \left[-\frac{1}{2} \left(\Delta R_i^2+\Delta I_i^2\right)\right] W_i,
371: \end{equation}
372: where $\Delta R_i$ and $\Delta I_i$ are the differences between model
373: and data for the real and imaginary components at each point $i$ in
374: the Fourier plane, and $W_i$ is a measure of the Gaussian noise
375: ($1/\sigma^2$).
376:
377: Since the X-ray counts, treated in image space, are distributed
378: according to Poisson statistics, the likelihood of the model fit is given by
379: \begin{equation}
380: \ln(\mathcal{L}_{image})=\displaystyle \sum_i \left[ D_i \ln(M_i) - M_i -\ln(D_i!) \right],
381: \label{eq:L_image}
382: \end{equation}
383: where $M_i$ is the model prediction (including cluster and background
384: components), and $D_i$ is the number of counts detected in pixel
385: $i$. The inner 100~kpc of the X-ray images---as well as any detected X-ray point sources---are
386: excluded from the fits by excluding these regions from the calculation
387: of $\ln(\mathcal{L}_{image})$ in Eq.~\ref{eq:L_image}.
388:
389: The joint likelihood of the spatial and spectral models is given by
390: $\mathcal{L}=\mathcal{L}_{SZ}\mathcal{L}_{Xray}$. For the N07+SVM
391: fits, $\mathcal{L}_{Xray}$ is simply $\mathcal{L}_{image}$. Following
392: \citet{bonamente2004,bonamente2006}, the X-ray likelihood for the
393: $\beta$-model fits is $\mathcal{L}_{image}\mathcal{L}_{Xspec}$, as
394: these must incorporate the likelihood $\mathcal{L}_{Xspec}$ of the
395: spectroscopic determination of $T_X$. The likelihood is used to
396: generate the Markov parameter chains, and convergence of the chain to
397: a stationary distribution is established using the Raftery-Lewis and Geweke tests
398: \citep{raftery1992,gilks1996}.
399:
400: \subsection{Calculation of \Mgas, \Mtot, and \Yint}\label{derived}
401: % mention get prob dist from mcmc, using each iteration.
402: With knowledge of the three-dimensional gas profiles, we compute
403: global properties of galaxy clusters as follows. The gas mass
404: $\Mgas(r)$ enclosed within radius $r$ is obtained by integrating the
405: gas density $\rhogas \equiv \mu_e m_p n_e(r)$ over a spherical
406: volume:
407: %
408: \begin{equation}
409: \Mgas(r) = 4 \pi \! \int_{0}^{r} \! \rhogas (r') r'^2 dr'.
410: \label{eq:gasmass}
411: \end{equation}
412: %
413: The total mass, \Mtot, can be obtained by solving the hydrostatic
414: equilibrium equation as:
415: \begin{equation}
416: \Mtot(r) = -\frac{r^2}{G \rhogas(r)} \frac{dP(r)}{dr},
417: \label{eq:hse}
418: \end{equation}
419: where $P=(\mu_e/\mu)P_e$ is the total gas pressure. We then compute
420: the gas mass fractions as $\fgas =\Mgas/\Mtot$.
421:
422: The line of sight Compton $y$-parameter, which characterizes the strength of
423: Compton scattering by electrons, is defined
424: \begin{equation}
425: y \equiv \frac{k_B \sigma_T}{m_e c^2} \int \!\! n_e T_e \,d\ell.
426: \end{equation}
427: We compute the volume-integrated Compton $y$-parameter, $Y$, from the
428: pressure profile fit to the SZ observations for both
429: cylindrical and spherical volumes of integration. The
430: cylindrically-integrated quantity, \Ycyl, is calculated within an
431: angle $\theta$ on the sky, corresponding to physical radius $R=\theta d_A$ at the
432: redshift of the cluster,
433: \begin{equation}
434: \Ycyl\left(R\right) \equiv 2\pi\,{d_A^2} \int_0^\theta \!\! y\left(\theta\right) \, \theta' d\theta' =
435: 2\pi \int_0^R \!\! y\left(r'\right) \, r' dr' = \frac{2\pi\,\sigma_T}{m_e \, c^2}
436: \int_{-\infty}^{\infty} \!\! d\ell \int_0^R \! P_e\left(r'\right) \, r' dr'.
437: \label{eq:Ycyl}
438: \end{equation}
439: The last form makes explicit the infinite limits of integration in the
440: line of sight direction, originating with the definition of $y$
441: (Eqs.~\ref{eq:thermal_sz} and \ref{eq:thermal_sz2}). The
442: spherically-integrated quantity, \Ysph, is obtained by integrating the
443: pressure profile within a radius $r$ from the cluster center,
444: \begin{equation}
445: \Ysph(r) = \frac{4 \pi \,\sigma_T}{m_e \, c^2} \int_{0}^{r} \!\! P_e (r') \, r'^2 dr.
446: \label{eq:Ysph}
447: \end{equation}
448: The parameter $\Ysph(r)$ is thus proportional to the thermal energy content of the ICM.
449:
450: To compute the global cluster properties described above, one needs to
451: define a radius out to which all quantities will be calculated.
452: Following \citet{laroque2006} and \citet{bonamente2006}, we compute global
453: properties of clusters enclosed within the overdensity radius $r_{\Delta}$,
454: within which the average density $\langle \rho \rangle$ of the cluster is a specified
455: fraction $\Delta$ of the critical density, via
456: %
457: \begin{equation}
458: \frac{4}{3} \pi \, \rho_c(z) \, \Delta \, r_{\Delta}^3 = \Mtot(r_{\Delta}),
459: \label{rvir}
460: \end{equation}
461: where $\rho_c(z)$ is the critical density at cluster redshift $z$,
462: and $\Delta \equiv \langle \rho \rangle /\rho_c(z)$.
463: In this work, we evaluate cluster properties at density
464: contrasts of $\Delta=2500$ and $\Delta=500$.
465: The overdensity radius $r_{2500}$ has been used in previous OVRO and BIMA
466: interferometric SZ studies \citep[e.g.][]{laroque2006,bonamente2008} as well as in many
467: X-ray cluster studies \citep[e.g.][]{vikhlinin2006, allen2007}, while $r_{500}$ is a radius
468: reachable with SZA and deep \emph{Chandra} X-ray data.
469:
470: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
471:
472: \section{Data\label{realobs}}
473:
474: \subsection{Cluster Sample \label{clusterselection}}
475:
476: % Samples (why we chose these three clusters)
477: For this work, we selected three massive clusters that are well studied at
478: X-ray wavelengths, and span a range of redshifts
479: ($z$~=~0.17--0.89) and morphologies, on which to test the
480: joint analysis of the {\em Chandra} X-ray and SZA data.
481: We assume a \LCDM\ cosmology with $\Omega_M=0.3$, $\Omega_\Lambda=0.7$, and $h=0.7$
482: throughout our analysis.
483:
484: Located at $z=0.25$, Abell 1835 (A1835) is an intermediate-redshift, relaxed
485: cluster, as evidenced by its circular morphology in the X-ray images
486: and its cooling core \citep{peterson2001}. To demonstrate the
487: applicability of our technique for high redshift clusters, we analyzed
488: CL~J1226.9+3332 (CL1226), an apparently relaxed cluster at $z=0.89$
489: \citep{maughan2004b,maughan2007b}. To assess how this method performs
490: on somewhat disturbed clusters, we also analyzed Abell 1914 (A1914), an
491: intermediate redshift ($z=0.17$) cluster with a hot subclump near the core.
492: When the subclump is not excluded from the X-ray analysis,
493: \citet[][hereafter M08]{maughan2008} find a large X-ray centroid shift
494: in the density profile, which they use as an indicator of an unrelaxed
495: dynamical state.
496:
497: In the following sections, we discuss the instruments, data
498: reduction, and analysis of the SZ and X-ray data. Details of these observations,
499: including the X-ray fitting regions, the unflagged, on-source
500: integration times, and the pointing centers used for the SZ observations, are
501: presented in Tables \ref{obsTable} and \ref{xrayTable}.
502: We also review an independent, detailed X-ray-only analysis, with which
503: we compare the results of our joint SZ+X-ray analyses.
504:
505: \subsection{Sunyaev-Zel'dovich Array Observations\label{sza}}
506:
507: The Sunyaev-Zel'dovich Array is an interferometric array comprising eight 3.5-meter telescopes,
508: and is located at the Owens Valley Radio Observatory.
509: For the observations presented here, the instrument was configured to operate in an 8-GHz-wide band
510: covering 27--35~GHz using the 26--36~GHz receivers
511: (hereafter referred to as the ``30-GHz'' band) or covering 90--98~GHz using the 80--115~GHz receivers
512: (the ``90-GHz'' band).
513: See \citet{muchovej2007} and Marrone et al. (in preparation) respectively
514: for more details about commissioning observations performed with the
515: 30-GHz and the 90-GHz SZA instruments. Details of the observations presented here, including the sensitivity and
516: effective resolution (the {\it synthesized beam}) of the long and
517: short baselines, are given in Table \ref{obsTable}.
518:
519: An interferometer directly measures the amplitude and phase of Fourier
520: modes of the sky intensity, with sensitivity to a range of angular
521: scales on the sky given by $\sim\lambda/B$, where $B$ is the projected separation
522: of any pair of telescopes, i.e., a {\it baseline}. The
523: field of view of the SZA is given by the primary beam of a
524: single telescope, approximately 10.7\arcmin\ at the
525: center of the 30-GHz band. At 30~GHz, optimal detection of
526: the arcminute-scale bulk SZ signal from clusters requires the short
527: baselines of a close-packed array configuration; six of the SZA
528: telescopes were arranged in this configuration for the observations
529: presented here, yielding 15 baselines
530: with sensitivity to $\sim$ 1--5\arcmin\ scales. The two outer antennas,
531: identical to the inner six, provide an additional 13 long baselines in this configuration,
532: with sensitivity to small-scale structure, allowing simultaneous
533: measurement of compact radio sources unresolved by the long
534: baselines (angular size $\lesssim 20\arcsec$) which could otherwise contaminate the SZ
535: signal. Observations at 90~GHz with the SZA probe scales at three times the resolution of the
536: 30-GHz observations for the same array configuration. The short baselines of the
537: 90-GHz observations thereby bridge the gap between long and short baseline coverage at
538: 30-GHz.
539:
540: % SZA data reduction
541: SZA data are processed in a complete pipeline for the reduction and
542: calibration of interferometric data, developed within the SZA
543: collaboration. Absolute flux calibrations are derived from
544: observations of Mars, scaled to the predictions
545: of \citet{rudy1987}. Data from each observation are bandpass-calibrated
546: using a bright, unresolved, flat-spectrum radio source, observed at
547: the start or end of an observation. The data are regularly
548: phase-calibrated using radio sources near the targeted field; these
549: calibrators are also used to track small variations in the antenna
550: gains. Data are flagged for corruption due to bad weather, sources of
551: radio interference and other instrumental effects that could impact
552: data quality. A more detailed account of the SZA data reduction
553: pipeline is presented in \citet{muchovej2007}.
554:
555: % Point sources
556: In the SZA cluster observations presented here, the A1835 field
557: contains three detectable compact sources at 31~GHz: a $2.8\pm0.3$~mJy
558: central source, a $1.1\pm0.4$~mJy source about one arcminute from
559: the cluster center, and a $0.8\pm0.4$~mJy source 5.5 arcminutes from center.
560: The positions of these sources are
561: in good agreement with those from the NVSS (which only contains the
562: central source) and the FIRST surveys.
563: The SZA observation of CL1226 contains one detectable compact source,
564: identified in both FIRST and NVSS, with flux at 31~GHz of $3.9\pm0.2$~mJy.
565: This source is 4.3 arcminutes from the cluster center \citep[see also][]{muchovej2007}
566: and therefore lies outside the field of view of the SZA 90~GHz observations.
567: Two compact sources, with positions constrained by NVSS and FIRST, were detected
568: at 31~GHz in the A1914 field. The fluxes of these sources are $1.8\pm0.4$~mJy and
569: $1.2\pm0.3$~mJy. The stronger was detected in both the NVSS and the FIRST surveys,
570: while the weaker was only detected in the FIRST survey.
571: Table \ref{ptsrcTable} summarizes the properties of the compact radio
572: sources detected in the SZA observations.
573:
574: When fitting compact sources detected in the SZ observations, we calculated
575: the spectral indices from the measured flux at 31~GHz and 1.4~GHz,
576: where the latter was constrained by either the NVSS \citep{condon1998} or
577: FIRST \citep{white1997} survey, respectively.
578: The source fluxes and
579: approximate coordinates are first identified using the interferometric
580: imaging package Difmap \citep{shepherd1997}. We first refine the source
581: positions by fitting a model in the MCMC routine, and then fix these
582: positions to their best-fit values when fitting the cluster SZ model.
583: We leave the source flux a free parameter, so that the cluster SZ
584: flux and any compact sources are fit simultaneously.
585:
586: \subsection{X-ray Observations \label{xrayobs_details}}
587: \label{sec:xray}
588:
589: All X-ray imaging data used in this analysis were obtained with the {\em
590: Chandra} ACIS-I detector, which provides spatially resolved X-ray
591: spectroscopy and imaging with an angular resolution of
592: $\sim0.5\arcsec$ and energy resolution of $\sim$ 100--200~eV.
593: Table~\ref{xrayTable} summarizes the X-ray observations of individual clusters.
594:
595: For the X-ray data used in the joint SZ+X-ray analysis, images
596: were limited to 0.7--7~keV in order to exclude the
597: data most strongly affected by
598: background and by calibration uncertainties. The X-ray images---which
599: primarily constrain the ICM density profiles---were
600: binned in 1.97$\arcsec$ pixels; this binning sets the limiting
601: angular resolution of our processed X-ray data, as the {\em Chandra} point
602: response function in the center of the X-ray image is smaller than our
603: adopted pixel size. The X-ray background was measured for each
604: cluster exposure, using source-free, peripheral regions of the adjacent detector
605: ACIS-I chips. Additional details of the {\em
606: Chandra} X-ray data analysis are presented in \citet{bonamente2004,bonamente2006}.
607:
608: In \S~\ref{profiles} \& \ref{global} we compare the results of our joint SZ+X-ray
609: analysis to the results of independent X-ray analyses. For A1914 and CL1226
610: we use the data and analysis described in detail in M08 and
611: \cite{maughan2007b} (hereafter M07), respectively. For A1835, the
612: ACIS-I observations used became public after the M08 sample was published, and we
613: therefore present its analysis here for the first time.
614: The observation of A1835 was calibrated and
615: analyzed using the same methods and routines described in M08, which we now
616: briefly review.
617:
618: In the X-ray-only analyses, blank-sky fields are used to
619: estimate the background for both the imaging and spectral analysis. The
620: imaging analysis (primarily used to obtain the gas emissivity profile)
621: is performed in the 0.7--2~keV energy band to maximize signal to noise.
622: Similar to the joint SZ+X-ray analyses, these images were also binned into 1.97$\arcsec$ pixels.
623:
624: For the spectral analysis, spectra extracted from a region of interest were
625: fit in the 0.6--9~keV band with an absorbed, redshifted
626: Astrophysical Plasma Emission Code (APEC) \citep{smith2001}
627: model. This absorption was fixed at the Galactic value.
628: This spectral analysis was used to derive both the global temperature
629: $T_X$, determined within the annulus $r \in [0.15,1.0] \, r_{500}$, and the radial temperature
630: profile fits of the V06 temperature profile, given by
631: \begin{equation}
632: T_{\mathrm{3D}}(r) = T_0
633: \left[\frac{(r/r_{\text{cool}})^{a_{\text{cool}}}+T_{\text{min}}/T_0}
634: {(r/r_{\text{cool}})^{a_{\text{cool}}}+1}\right]
635: \left[\frac{(r/r_t)^{-a}}{(1+(r/r_t)^b)^{c/b}}\right].
636: \label{eq:V06_tprof}
637: \end{equation}
638: We refer the reader to V06 for details, but note that this temperature profile
639: is the combination of a cool core component (first set of square brackets, where
640: the core temperature declines to $T_{\text{min}}$) and a decline at large radii
641: (second set of square brackets, where temperature falls at $r \gtrsim r_t$).
642:
643: An important consideration when using a blank-sky background method is
644: that the count rate at soft energies can be significantly different in
645: the blank-sky fields than in the target field, due to differences between
646: the level of the soft Galactic foreground emission in the target field
647: and that in the blank-sky field.
648: This was accounted for in the imaging analysis by normalizing the
649: background image to the count rate in the target image in regions
650: far from the cluster center. In the spectral analysis, this was modeled by an
651: additional thermal component that was fit
652: to a soft residual spectrum (the difference between spectra extracted in
653: source free regions of the target and background datasets; see \citet{vikhlinin2005a}).
654: The exception to this was the \emph{XMM-Newton} data
655: used in addition to the {\em Chandra} data for CL1226. As discussed
656: in M07, a local background was found to be more reliable
657: for the spectral analysis in this case, thus requiring no correction for
658: the soft Galactic foreground.
659:
660: The M07/M08 X-ray analysis methods exploit the full V06 density and temperature models
661: (Eqs.~\ref{eq:Vikhlinin} \& \ref{eq:V06_tprof}, respectively) to fit the emissivity and temperature profiles derived for each cluster,
662: and the results of these fits are used to derive the total hydrostatic mass profiles of each system.
663: Uncertainties for the independent, X-ray-only analysis method are derived using a
664: Monte Carlo randomization process. These fits involved typically
665: $\sim$1000 realizations of the temperature and surface brightness profiles, fit
666: to data randomized according to the measured noise.
667: We refer the reader to M07 and M08, where this fitting procedure is described in detail.
668:
669: \section{Results \label{results}}
670:
671: \subsection{SZ Cluster Visibility Fits}\label{sze_visfits}
672:
673: Interferometric SZ data are in the form of visibilities $V_\nu(u,v)$ \citep[see, for example,][]{thompson2001},
674: which for single, targeted cluster observations with the SZA can be expressed in the small angle approximation as
675: \begin{equation}
676: V_\nu(u,v) = \int \! \int A_\nu(x,y) \, I_\nu(x,y) \, {e^{- i 2 \pi (ux + vy)}} \, dx \, dy.
677: \label{eq:visibility}
678: \end{equation}
679: Here $u$ and $v$ (in number of wavelengths) are the Fourier conjugates of the direction cosines
680: $x$ and $y$ (relative to the observing direction), $A_\nu(x,y)$ is the angular power sensitivity pattern of each antenna at frequency $\nu$,
681: and $I_\nu(x,y)$ is the intensity pattern of the sky (also at $\nu$).
682: Eq.~\ref{eq:visibility} is recognizable as a 2-D Fourier transform, so the visibilities
683: give the flux for the Fourier mode for the corresponding $u,v$-coordinate.
684:
685: By combining Eqs.~\ref{eq:thermal_sz} and \ref{eq:visibility}, we can remove the frequency
686: dependence from the measured cluster visibilities, just as we have related SZ intensity to the
687: frequency-independent Compton $y$-parameter.
688: We define the frequency-independent cluster visibilities $Y(u,v)$ as
689: \begin{equation}
690: \label{eq:Yuv}
691: V_\nu(u,v) \equiv g(x) \, I_0 \, Y(u,v),
692: \end{equation}
693: where $I_0$ (in units of flux per solid angle) is
694: \begin{equation}
695: \label{eq:I0_cmb}
696: I_0 = \frac{2 (k_B \Tcmb)^3}{(h c)^2}.
697: \end{equation}
698: Additionally, we rescale $Y(u,v)$ by the square of the angular diameter distance, $d_A^2$, in order
699: to remove the redshift dependence from the cluster SZ signal.
700: Note that, while we use the non-relativistic $g(x)$ (Eq.~\ref{eq:g_x}) to compute $Y(u,v)$
701: (Eq.~\ref{eq:Yuv}) , we only use this for display purposes. The effects of assuming the classical
702: SZ frequency dependence are discussed in \S~\ref{profiles}.
703:
704: Figure~\ref{fig:uvplots} shows the maximum-likelihood fits of the
705: N07 profile and the isothermal $\beta$-model to each cluster's visibility data,
706: from which we have subtracted the detected radio sources (Table~\ref{ptsrcTable}).
707: We also removed the frequency dependence of the SZ effect by rescaling the
708: cluster visibilities to $Y(u,v)$ (Eq.~\ref{eq:Yuv}).
709: For the purposes of plotting, this rescaling is useful when binning the SZ signal
710: across 8~GHz of bandwidth as well as when plotting the 30-GHz and 90-GHz SZA data
711: taken on CL1226.
712:
713: As indicated in Fig.~\ref{fig:uvplots}, both the isothermal $\beta$-model and N07 model (which was fit jointly
714: with the SVM) fit the available data equally well. However, as $\sqrt{u^2 + v^2} \rightarrow 0$ (large
715: scales on the sky), the isothermal $\beta$-model extrapolates to a much larger value of $Y(u,v)$.
716: This corresponds to the much larger values of \Ycyl\ that are computed at large radii
717: using fit $\beta$-models (see \S \ref{global}).
718:
719: The middle panel of Fig.~\ref{fig:uvplots} shows the combined 30+90~GHz observations of CL1226,
720: which has a smaller angular extent than A1835 or A1914 due to its distance (compare, for example, the values
721: of $r_{500}$ for each cluster, listed in Table~\ref{table:derivedQuants_r500}).
722: The dynamic range and \emph{u,v}-space coverage provided by the 30-GHz SZA observations (black points) on CL1226
723: were insufficient for constraining the radial pressure profile of the cluster.
724: The short baselines of the 90-GHz SZA observations (middle three points) help to provide
725: more complete \emph{u,v}-space coverage, as discussed in \S \ref{sza}.
726:
727:
728: \subsection{X-ray Surface Brightness Fits}\label{xray_surffits}
729:
730: The X-ray surface brightness (Eq.~\ref{eq:xray_sb}), ignoring the data within a 100 kpc radius,
731: was modeled separately with both the isothermal
732: $\beta$-model, using the spectroscopically-determined, global $T_X$ (measured within
733: $r \in [0.15,1.0] \, r_{500}$), and the SVM, using the temperature derived from the N07 pressure
734: profile fit to the SZ data.
735: Figure~\ref{fig:Sx_profiles} shows the maximum-likelihood fits to the surface
736: brightness of each cluster for both the SVM and isothermal $\beta$-model. For plotting purposes,
737: the X-ray data are radially-averaged around the cluster centroid, which is determined in the
738: joint SZ+X-ray analysis by fitting the two-dimensional X-ray imaging data with the
739: spherically-symmetric SVM and isothermal $\beta$-model profiles.
740:
741: \subsection{ICM Profiles \label{profiles}}
742:
743: Figure~\ref{fig:profiles} shows the three-dimensional ICM radial
744: profiles derived from the joint analysis of SZA + \emph{Chandra} X-ray
745: surface brightness data for A1835, CL1226 and A1914 (from left to
746: right). From top to bottom, we show the electron pressure, the gas
747: density and the derived electron temperature profiles, each as a function of cluster
748: radius. In all panels, we compare the ICM profiles derived from the
749: N07+SVM model to the results of a traditional isothermal $\beta$-model
750: analysis, indicated by solid and dot-dashed lines, respectively. The
751: hatched regions indicate the 68\% confidence interval for each derived
752: parameter.
753:
754: As shown in the top panels of Fig.~\ref{fig:profiles}, the pressure
755: profiles derived from the N07 model and the isothermal $\beta$-model
756: show good agreement within $r_{2500}$, but deviate by
757: $\sim$3--5-$\sigma$ in the cluster outskirts. This is a consequence
758: of the fact that clusters exhibit a significant decline in
759: temperature beyond $r_{2500}$, as determined from spectroscopic X-ray
760: observations (see the bottom panel of Fig.~\ref{fig:profiles}). The
761: pressure profile derived from the isothermal $\beta$-model analysis
762: is therefore biased systematically high beyond $r_{2500}$. In
763: contrast, the N07 model, which fits the pressure directly, is free to
764: capture the true shape of the pressure profile well beyond the radius
765: at which the assumption of isothermality becomes invalid.
766:
767: For all three clusters there is little evidence for the second
768: component of the electron density allowed by the SVM; the density is
769: fit equally well by either the SVM or a single-component
770: $\beta$-model, as illustrated by the center row of panels in
771: Fig.~\ref{fig:profiles} (see also Table~\ref{xrayTable}). For all
772: three clusters, the fits of the SVM agree to within 1--2\% of the full V06
773: density profile fits (not shown) outside the core; discrepancies at
774: this level are easily attributed to differences in the fitting of the X-ray background, and
775: to the differences between the APEC and Raymond-Smith emissivity
776: models used respectively in the M07/M08 X-ray analysis and the joint SZ+X-ray
777: analyses.
778:
779: In the bottom panels, the electron temperature profiles inferred from
780: the N07+SVM profiles are compared to temperature profiles derived from
781: deep spectroscopic \emph{Chandra} X-ray observations (and
782: \emph{XMM-Newton} in the case of CL1226; see M07). The comparison
783: shows that the radial electron temperature profiles derived from the N07+SVM
784: profiles are in good agreement with independent X-ray measurements for
785: the relaxed clusters A1835 and CL1226, which exhibit radially
786: decreasing temperature profiles in the cluster outskirts \citep[see
787: also][]{markevitch1998,vikhlinin2005a}. The disturbed cluster A1914,
788: however, shows less overall agreement between the derived N07+SVM radial
789: temperature profile and the M08 fit of the V06 temperature profile.
790: Since the N07 pressure profile fit to the SZA observation of A1914 agrees
791: with that derived from M08 within their respective 68\% confidence intervals,
792: the temperature discrepancy is likely due to deviations from the spherical symmetry
793: implicitly assumed in this analysis.
794: Additionally, scales greater than about six arcminutes are beyond the radial
795: extents probed by the SZA; it is unsurprising the agreement becomes poorer
796: at radii larger than this.
797:
798: We note that we fit the SZ data using relativistic corrections to the SZ frequency dependence
799: provided by \citet{itoh1998}. These corrections are appropriate for the thermal SZ
800: effect at the SZA observing frequencies of 30 and 90~GHz.
801: Compared to fits assuming the classical SZ frequency dependence, a pressure profile fit using the
802: relativistic corrections has both a higher normalization and larger upper error bars.
803: This is noticeable when including higher frequency data, where the relativistic correction is larger
804: ($\sim 5\%$ at 90~GHz versus $\sim 3\%$ at 30~GHz for cluster temperatures $\sim$~8~keV).
805: This increase in the pressure fit is due to the diminished magnitude of the SZ effect
806: when using these corrections (for frequencies below the null in the SZ spectrum,
807: $\lesssim 218$~GHz; see e.g. \citet{itoh1998}).
808: The pressure profile therefore must adjust to fit the observed SZ flux.
809:
810: The larger upper error bars on the fit pressure profile arise from a more subtle effect. Since the
811: temperature is derived from the simultaneously fit pressure and density profiles, and the
812: SZ effect diminishes as electrons become more relativistic (i.e. hotter), the upper error
813: bar of the pressure fit must increase to fit the same noise in the observation (compared
814: to the non-relativistic case).
815: The lower error bar is less affected, as lower electron temperatures require smaller
816: relativistic corrections. The resulting asymmetric error bars can be seen in the derived
817: temperature profile of CL1226, which relied on 90~GHz data, in Figure~\ref{fig:profiles}.
818:
819: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
820:
821: \subsection{Measurements of $Y$, \Mgas, \Mtot, and \fgas \label{global}}
822:
823: In Tables \ref{table:derivedQuants_r2500} and
824: \ref{table:derivedQuants_r500}, we report global properties of
825: individual clusters derived from the N07+SVM model fits to the
826: SZ+X-ray data. We calculate all quantities at overdensity radii
827: $r_{2500}$ and $r_{500}$, and compare them to results from both the
828: isothermal $\beta$-model analysis of the same data, as well as to the
829: X-ray-only analysis.
830:
831: The N07 pressure profile has just two free parameters---$P_{e,i}$ and
832: $r_p$---which exhibit a degeneracy. Figure~\ref{fig:a1835_yint_joint_SZ_degen} shows this
833: degeneracy in fits of the N07 profile to the SZA observations of A1835.
834: Similar to the $r_c-\beta$ degeneracy of the $\beta$-model
835: \citep[see][for example]{grego2001}, these two quantities are not
836: individually constrained by our SZ observations, but they are tightly
837: correlated and the preferred region in the $P_{e,i}-r_p$ plane encloses
838: approximately constant \Ycyl. As a result, the 68\% confidence region for \Ycyl\
839: is more tightly constrained than the large variation in $P_{e,i}$ or $r_p$ might
840: individually indicate.
841: Figure~\ref{fig:a1835_yint_joint_SZ_degen} also shows that the inclusion of
842: X-ray data has only a marginal effect on the value of \Ycyl\ derived from the SZ fit. This is
843: as expected, due to
844: the weak dependence of the X-ray surface brightness on temperature (see \S~\ref{xray_sze})
845: and the fact that the N07 profile is not linked to the SVM density profile's shape.
846: This indicates that X-ray data are not necessary to constrain \Ycyl, although they do limit
847: the range of accepted radial pressure profiles.
848:
849: At both $r_{2500}$ and $r_{500}$, the measurements of \Ycyl\ derived from the joint
850: N07+SVM and the X-ray-only analysis are consistent at the 1-$\sigma$
851: level, for all three clusters. The isothermal $\beta$-model analysis,
852: however, overestimates \Ycyl\ by $\sim$~20\%--40\% at $r_{2500}$, and
853: by $\sim$~30\%--115\% at $r_{500}$. This is due to the large
854: contribution to \Ycyl\ from the cluster outskirts, at every projected radius,
855: where the $\beta$-model significantly overestimates the
856: pressure.
857:
858: In contrast to the systematic variations in \Ycyl, the determinations of $\Ysph(r_{2500})$ are
859: consistent among the three analyses. At $r_{500}$, however, the
860: median \Ysph\ values from the isothermal $\beta$-model can be as much as
861: $\sim$ 60\% higher than either the N07+SVM or M08 results, due to
862: the fact that isothermality is a poor description of the cluster
863: outskirts.
864:
865: The gas mass estimates computed using either the jointly-fit SVM or
866: the isothermal $\beta$-model agree with the gas masses derived from
867: the Maughan X-ray fits (Tables \ref{table:derivedQuants_r2500} and
868: \ref{table:derivedQuants_r500}). This agreement is not surprising,
869: given that the gas mass is determined in all cases from density fits
870: to the X-ray data. It demonstrates, however, that the 100~kpc core
871: makes a negligible contribution to the total gas mass even at
872: $r_{2500}$, and that excluding the core from the joint analysis does
873: not therefore introduce any significant bias in our estimate of
874: \Mgas. Incidentally, it also shows that the additional component
875: allowed by the SVM and the full V06 density models is not indicated in
876: these clusters.
877:
878: In Tables \ref{table:derivedQuants_r2500} and
879: \ref{table:derivedQuants_r500}, we also present estimates of the total
880: masses, computed using each model's estimate of the overdensity radius
881: ($r_\Delta$) for each Monte Carlo realization of the fit parameters.
882: For two of the clusters, we find that the error bars are significantly
883: larger for \Mtot\ determined from the N07+SVM fits than for the
884: isothermal $\beta$-model or M08 fits. This is a consequence of the fact
885: that the $\beta$-model analysis, with fewer free parameters and the
886: assumption of isothermality, effectively places stronger but poorly-justified
887: priors on the total mass.
888: We find that the N07+SVM and M08 total mass estimates agree
889: at both $r_{500}$ and $r_{2500}$, leading to good overall agreement
890: between gas fractions computed using the N07+SVM profiles and those
891: from the Maughan X-ray fits.
892:
893: Since the HSE estimate for \Mtot\ is sensitive to the change in slope in the pressure
894: profile ($dP/dr$), \Mtot\ is not as well constrained by the N07+SVM profiles as is \Ycyl,
895: which scales directly with integrated SZ flux, a parameter that is more directly linked
896: to what the SZA observes (see \S~\ref{sze_visfits}).
897: Figure~\ref{fig:1dhists} shows a comparison of the N07+SVM estimates for \Mtot\ and \Ycyl,
898: revealing that \Ycyl\ has a more tightly constrained and centrally peaked distribution than
899: \Mtot\ does.
900:
901: The isothermal $\beta$-model, on the other hand, is over-constrained such that
902: it cannot agree with the non-isothermal estimates of \Mtot\ at both $r_{2500}$ and $r_{500}$,
903: Its estimates of \Mtot\ are moreover sensitive to the annulus within
904: which $T_X$ is determined (see \S~\ref{sec:xray}). This trend can be
905: seen in Fig.~\ref{fig:massprofiles}, which shows \Mtot$(r)$ for each
906: cluster.
907:
908: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
909:
910: \section{Conclusions}\label{conclusions}
911: In this work, we demonstrated the application of a new pressure
912: profile, proposed by \citet{nagai2007b}, in fitting SZ data taken by the Sunyaev-Zel'dovich Array.
913: By combining the pressure profile constraints from the SZ data with constraints on density from \emph{Chandra} X-ray
914: imaging, using a simplified form of the density model proposed by \citet{vikhlinin2006},
915: we were able to determine the properties of three clusters (A1835, A1914, and CL~J1226.9+3332)
916: spanning a wide range in redshift and dynamical states.
917:
918: This technique was compared with two others:
919: a joint analysis of the same SZ + X-ray data relying on the commonly used isothermal $\beta$-model,
920: which assumes the pressure and density profiles have the same form,
921: and an X-ray only analysis that independently models temperature and density
922: using both X-ray imaging and spectroscopic measurements (but excluding SZ data).
923: We find that the global cluster properties at both $r_{2500}$
924: and $r_{500}$ determined from ICM profiles fit in the joint pressure analysis are in excellent agreement
925: with properties determined in the X-ray analysis.
926: By contrast, the isothermal $\beta$-model tends to
927: overestimate with respect to the other models the gas pressure at $r > r_{2500}$, where
928: isothermality is an increasingly poor assumption. The $\beta$-model
929: thus leads to an overestimate of the cylindrically-integrated Compton $y$-parameter
930: at $r_{500}$.
931: Since the isothermal $\beta$-model does not provide a good description of
932: the ICM profile in the cluster outskirts, we caution
933: against its use in deriving global properties of clusters even at $r \sim r_{500}$,
934: which is a large fraction of the virial radius.
935:
936: We tested the ability to recover the ICM electron pressure profiles
937: from SZ data by analyzing the SZ data together with the X-ray imaging data alone,
938: ignoring the X-ray spectroscopic information.
939: Assuming the ideal gas law, we derive electron temperature profiles by
940: coupling the pressure fits to the SZ data with the density fits to the X-ray imaging data.
941: We find that these derived temperature profiles show broad agreement with those
942: determined spectroscopically from deep X-ray observations, even for
943: the highest redshift cluster in our sample, at $z=0.89$. This method
944: therefore provides an independent technique for determining the radial
945: electron temperature distribution in high-redshift clusters, for which
946: deep spectroscopic X-ray data may be unavailable and are impractical to
947: obtain.
948:
949:
950: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
951:
952: \acknowledgments
953:
954: We thank John Cartwright, Ben Reddall and Marcus Runyan for their significant contributions
955: to the construction and commissioning of the \facility{SZA} instrument. We thank Esra Bulbul and
956: Nicole Hasler for their insights, comments, and help with \emph{Chandra} X-ray Observatory
957: (\facility{CXO}) data reduction and figure production.
958: We thank the staff of the Owens Valley Radio Observatory and CARMA for their outstanding support.
959: We thank Samuel LaRoque for his help with the modeling code.
960: We gratefully acknowledge the James S.\ McDonnell Foundation, the National Science Foundation
961: and the University of Chicago for funding to construct the SZA.
962: The operation of the SZA is supported by NSF Division of Astronomical Sciences through
963: grant AST-0604982. Partial support is provided by NSF Physics Frontier Center grant PHY-0114422
964: to the Kavli Institute of Cosmological Physics at the University of Chicago, and by NSF grants
965: AST-0507545 and AST-05-07161 to Columbia University. AM acknowledges support from a Sloan Fellowship,
966: DM from an NRAO Jansky Fellowship, BM from a Chandra Fellowship, and CG, SM, and MS from NSF Graduate Research Fellowships.
967:
968: \bibliography{genNFW}
969: %%%%%%%%%%%%%%%%%%%%%%
970: % Tables
971: \input{obs_table.tex}
972: \input{xray_table.tex}
973: \input{ptsrc_table.tex}
974: \input{derived_table_r2500.tex}
975: \input{derived_table_r500.tex}
976:
977: \pagebreak
978: %%%%%%%%%%%%%%%%%%%%%%
979: % Figures
980: \begin{figure}
981: \centerline{
982: \includegraphics[height=1.8in]{a1835}
983: \includegraphics[height=1.8in]{cl1226}
984: \includegraphics[height=1.8in]{a1914}
985: }
986: \caption{SZ profiles for A1835 (left), CL1226 (middle), and A1914 (right),
987: plotted as a function of \emph{u,v}-radius ($\sqrt{u^2 + v^2}$).
988: The panels show the real component of the visibilities over the full range of the fit models,
989: using a broken axis to capture the model predictions as the \emph{u,v}-radius
990: approaches zero k$\lambda$.
991: Using Eq.~\ref{eq:Yuv}, the visibilities were rescaled to $Y(u,v) \, d_A^2$, which removes
992: both the frequency and redshift dependence of the cluster visibilities.
993: The black points with error bars (1-$\sigma$) are the binned 30-GHz SZA visibility data.
994: Models for the compact sources have been subtracted from the visibility
995: data before rescaling.
996: The blue solid line is a high likelihood N07 model fit, while the red dashed line is a similarly-chosen
997: fit of the isothermal $\beta$-model.
998: For the available data points, both cluster models fit equally well. However, note that as the
999: \emph{u,v}-radius approaches zero k$\lambda$ (corresponding to large spatial scales on the sky)---where
1000: there are no data to constrain the models---the $\beta$-model predicts much more flux than the N07 model.
1001: The middle three (green) data points for CL1226 are taken from 90-GHz
1002: SZA observations.
1003: }\label{fig:uvplots}
1004: \end{figure}
1005: %%%%%%%%%%%%%%%%%%%%%%
1006: \begin{figure}
1007: \begin{center}
1008: \includegraphics[width=1.8in,angle=270]{sbprofile_log_A1835}
1009: \includegraphics[width=1.8in,angle=270]{sbprofile_log_CL1226}
1010: \includegraphics[width=1.8in,angle=270]{sbprofile_log_A1914}
1011: \end{center}
1012: \caption{X-ray surface brightness profiles for A1835 (left), CL1226 (middle), and A1914 (right).
1013: The vertical dashed line denotes the 100~kpc core cut.
1014: In each panel, the blue, solid line is the surface brightness computed using a high-likelihood fit of the
1015: N07+SVM profiles (analogous to the SZ models plotted in Fig.~\ref{fig:uvplots}), while the red,
1016: dot-dashed line is that from an isothermal $\beta$-model fit.
1017: Both fit lines include the X-ray background that was simultaneously fit with the cluster emission model.
1018: The black squares are the annularly-binned X-ray data, where the widths of the bins are denoted
1019: by horizonal error bars. The vertical error bars are the 1-$\sigma$ errors on the measurements.
1020: Arrows indicate $r_{2500}$ and $r_{500}$ derived from the N07+SVM profiles (see Tables
1021: \ref{table:derivedQuants_r2500} \& \ref{table:derivedQuants_r500}).}\label{fig:Sx_profiles}
1022: \end{figure}
1023: %%%%%%%%%%%%%%%%%%%%%%
1024: \begin{figure}
1025: \includegraphics[width=6in]{clusterplots}
1026: \caption{Three-dimensional ICM radial profiles derived from the joint
1027: analysis of SZA+{\em Chandra} X-ray surface brightness data for
1028: A1835, CL1226, and A1914 (from left to right). From top to
1029: bottom: the electron pressure $P_e(r)$, the electron density $n_e(r)$,
1030: the derived electron temperature $T_e(r)$, profiles as a function of the
1031: cluster-centric radius. The lines show the median deprojected
1032: quantity derived from data using the N07+SVM (solid lines, vertically-hatched regions) and the
1033: isothermal $\beta$-model (dot-dashed lines, horizontally-hatched regions). The derived electron
1034: temperature profiles are compared to the spectroscopically-determined
1035: radial temperature profiles (black dashed lines, slanted hatching) obtained
1036: according to methods presented in M07/M08.
1037: The hatched region indicate the 68\% confidence on parameters derived from each
1038: model. The arrows denote the median values of $r_{2500}$ and $r_{500}$ from the
1039: N07+SVM fits.}\label{fig:profiles}
1040: \end{figure}
1041: %%%%%%%%%%%%%%%%%%%%%%
1042: \begin{figure}
1043: \centerline{\includegraphics[width=4in]{yint_joint_SZ_a1835}}
1044: \caption{\Ycyl\ computed within 6\arcmin\ ($\sim 1.4~\rm Mpc$, which is $\approx r_{500}$
1045: for this cluster) for fits to SZA observation of A1835. The bold, black contours contain 68\% and
1046: 95\% of the accepted iterations on the jointly-fit {\em Chandra} + SZA data, while the
1047: thinner, blue contours are those for fits to SZA data alone.}\label{fig:a1835_yint_joint_SZ_degen}
1048: \end{figure}
1049: %%%%%%%%%%%%%%%%%%%%%%
1050: \begin{figure}
1051: \centerline{\includegraphics{1Dhist_joint}}
1052: \caption{1-D histograms of \Mtot\ (cyan/light gray region
1053: with dashed outline) and \Ycyl\ (vertically-hatched region with solid outline), normalized by their
1054: respective median values, derived from the joint fits of the N07+SVM profiles to A1835.
1055: Both \Mtot\ and \Ycyl\ are computed within a fixed radius of $\theta=6\arcmin$.
1056: The derived \Ycyl, which scales directly with integrated SZ flux, has a more tightly constrained
1057: and centrally peaked distribution than that of \Mtot, as \Mtot\ is sensitive to the change
1058: in slope in the pressure profile ($dP/dr$).}\label{fig:1dhists}
1059: \end{figure}
1060: %%%%%%%%%%%%%%%%%%%%%%
1061: \begin{figure}
1062: \includegraphics[width=6in]{clustermasses}
1063: \caption{Total mass profiles derived from the joint analysis of
1064: SZA+{\em Chandra} X-ray surface brightness data for A1835,
1065: CL1226, and A1914 (from left to right). The line types are the same
1066: as those in Fig.~\ref{fig:profiles}.}\label{fig:massprofiles}
1067: \end{figure}
1068: %%%%%%%%%%%%%%%%%%%%%%
1069:
1070: \end{document}
1071: