1:
2: %% --------------------------------------------------------------------
3: %% Fri May 15 10:08:42 2009
4: %% This file was generated automagically from the files
5: %% Bullet_main.bbl and Bullet_main.tex using
6: %% nat2jour.pl
7: %% All citations have been inlined and dependencies on the natbib
8: %% package have been removed so that this file (together with
9: %% Bullet_main-aas.bbl) should be suitable for submission to journals with
10: %% the citation styles of ApJ or MNRAS.
11: %% --------------------------------------------------------------------
12:
13: \documentclass[preprint2]{aastex}
14: %\documentclass{aastex}
15: \usepackage{graphicx}
16: \usepackage{graphics}
17: \usepackage{color}
18: %\usepackage{cite} % bibliographic sitations
19: %\usepackage{natbib}
20: %\citestyle{aa}
21: %\usepackage{floatflt}
22: %\graphicspath{{figures/}}
23: \DeclareGraphicsExtensions{.pdf}
24:
25: \slugcomment{Accepted for publication in the Astrophysical Journal}
26:
27: \shortauthors{Halverson et al.}
28: \shorttitle{Bullet Cluster SZE Observations with APEX-SZ}
29:
30: \begin{document}
31:
32: %%%%%%%%%%%%%%%%%%%%%%%
33: % Define new commands %
34: %%%%%%%%%%%%%%%%%%%%%%%
35: \newcommand{\oscdemod}{oscillator/demodulator}
36: \newcommand{\fMUX}{fMUX}
37: \newcommand{\squid}{SQUID}
38: \newcommand{\rtHz}{$\sqrt{\mbox{Hz}}$}
39: \newcommand{\phinot}{\mbox{$\Phi_0$}}
40: \newcommand{\degree}{\mbox{$^{\circ}$}}
41: \newcommand{\fortran}{{\tt Fortran~77}}
42: \newcommand{\CXX}{C++}
43: \newcommand{\order}{\mbox{${\cal O}$}}
44: \newcommand{\const}{\mbox{\sc\small Const}}
45: %\newcommand{\mycomment}[1]{\marginpar{\it\tiny #1}} % used for margin comments
46: %\newcommand{\mycomment}[1]{\typeout{#1}}
47: \newcommand{\mycomment}[1]{{\bf\it\color{red} #1}} % Use for bold in-tex comments.
48:
49: %--------------------------------------------------------------
50:
51: \title{Sunyaev-Zel'dovich Effect Observations of the Bullet Cluster (1E 0657--56) with APEX-SZ}
52:
53: % Note: to compare present version with first submitted version:
54: %>cvs diff -r1.190 Bullet_main.tex
55:
56: \author{N.~W.~Halverson,\altaffilmark{1,2} T.~Lanting,\altaffilmark{3}
57: P.~A.~R.~Ade,\altaffilmark{4} K.~Basu,\altaffilmark{5}
58: A.~N.~Bender,\altaffilmark{1}
59: B.~A.~Benson,\altaffilmark{6}
60: F.~Bertoldi,\altaffilmark{5}
61: H.-M.~Cho,\altaffilmark{7} G.~Chon,\altaffilmark{8}
62: J.~Clarke,\altaffilmark{6,9} M.~Dobbs,\altaffilmark{3}
63: D.~Ferrusca,\altaffilmark{6} R.~G\"usten,\altaffilmark{8}
64: W.~L.~Holzapfel,\altaffilmark{6} A.
65: Kov\'acs,\altaffilmark{8} J.~Kennedy,\altaffilmark{3}
66: Z.~Kermish,\altaffilmark{6} R.~Kneissl,\altaffilmark{8}
67: A.~T.~Lee,\altaffilmark{6,9} M.~Lueker,\altaffilmark{6}
68: J.~Mehl,\altaffilmark{6} K.~M.~Menten,\altaffilmark{8}
69: D.~Muders,\altaffilmark{8} M.~Nord,\altaffilmark{5,8}
70: F.~Pacaud,\altaffilmark{5} T.~Plagge,\altaffilmark{6}
71: C.~Reichardt,\altaffilmark{6} P.~L.~Richards,\altaffilmark{6}
72: R.~Schaaf,\altaffilmark{5} P.~Schilke,\altaffilmark{8}
73: F.~Schuller,\altaffilmark{8} D.~Schwan,\altaffilmark{6}
74: H.~Spieler,\altaffilmark{9} C.~Tucker,\altaffilmark{4}
75: A.~Weiss,\altaffilmark{8} O.~Zahn\altaffilmark{6}}
76:
77: \altaffiltext{1} {Center for Astrophysics and Space Astronomy, Department of Astrophysical and Planetary Sciences, University of Colorado, Boulder, CO, 80309}
78: \altaffiltext{2} {Department of Physics, University of Colorado, Boulder, CO, 80309}
79: \altaffiltext{3}{Department of Physics, McGill University, Montr\'{e}al, Canada, H3A 2T8}
80: \altaffiltext{4}{School of Physics and Astronomy, Cardiff University, CF24 3YB Wales, UK}
81: \altaffiltext{5}{Argelander Institute for Astronomy, Bonn University, Bonn, Germany}
82: \altaffiltext{6}{Department of Physics, University of California, Berkeley, CA, 94720}
83: \altaffiltext{7}{National Institute of Standards and Technology, Boulder, CO, 80305}
84: \altaffiltext{8}{Max Planck Institute for Radioastronomy, 53121 Bonn, Germany}
85: \altaffiltext{9}{Lawrence Berkeley National Laboratory, Berkeley, CA, 94720}
86:
87: \begin{abstract}
88: We present observations of the Sunyaev-Zel'dovich effect (SZE) in
89: the Bullet cluster (1E 0657--56) using the APEX-SZ instrument at 150$\,$GHz
90: with a resolution of $1\arcmin$. The main results are maps of the
91: SZE in this massive, merging galaxy cluster. The cluster is
92: detected with $23\,\sigma$ significance
93: within the central 1\arcmin\ radius of the source position.
94: The SZE map has a broadly similar morphology to that in existing X-ray
95: maps of this system, and
96: we find no evidence for significant contamination of the SZE emission by
97: radio or IR sources.
98: In order to make simple quantitative comparisons with cluster
99: gas models derived from X-ray observations,
100: we fit our data to an isothermal elliptical $\beta$ model, despite the
101: inadequacy of such a model for this complex merging system.
102: With an X-ray derived prior on the power-law index, $\beta = 1.04^{+0.16}_{-0.10}$,
103: we find a core radius $r_c =142 \pm 18 \arcsec$, an axial
104: ratio of $0.889 \pm 0.072$, and a central temperature decrement
105: of $-771 \pm 71\,\mu{\rm K_{CMB}}$, including a $\pm 5.5\%$ flux calibration uncertainty.
106: Combining the APEX-SZ map with a map of projected electron surface density from Chandra X-ray
107: observations, we determine the mass-weighted temperature of the cluster gas to be
108: $T_{mg}=10.8 \pm 0.9\,$keV, significantly lower than some previously reported X-ray spectroscopic
109: temperatures.
110: Under the assumption of an isothermal cluster gas distribution in hydrostatic equilibrium,
111: we compute the gas mass fraction for prolate and oblate spheroidal geometries and find it to
112: be consistent with previous results from X-ray and weak lensing observations.
113: This work is the first result from the APEX-SZ experiment, and
114: represents the first reported scientific result from
115: observations with a large array of multiplexed superconducting
116: transition-edge sensor bolometers.
117: \end{abstract}
118:
119: \keywords{cosmic microwave background --- cosmology:observations ---
120: galaxies: clusters: individual (1E 0657--56)}
121:
122: %\maketitle
123:
124: %--------------------------------------------------------------
125:
126: \section{Introduction}
127: \label{SEC:introduction}
128:
129: Clusters of galaxies are a unique probe of the growth and dynamics of
130: structure in the Universe.
131: In particular, active mergers of sub-clusters provide a window to the
132: processes by which massive clusters are assembled.
133: In these systems, the galaxies and associated dark
134: matter are essentially collisionless. In contrast, the ionized
135: intracluster gas, typically at temperatures of $T \sim 10^8$~K,
136: is strongly interacting and experiences complex dynamics.
137: In extreme cases, the normally associated dark matter and
138: intracluster gas can be significantly separated.
139:
140: The Bullet cluster (1E 0657--56) at $z=0.296$, is a massive cluster
141: consisting of two
142: sub-clusters in the process of merging. The smaller sub-cluster or
143: ``bullet" has passed through the larger main cluster.
144: X-ray observations infer a bow shock velocity of $\sim$4700 km/s~\markcite{markevitch2006}({Markevitch} 2006),
145: while simulations of the collision yield a substantially lower speed for the
146: sub-cluster \markcite{springel2007}({Springel} \& {Farrar} 2007).
147: This collision is perpendicular to the line of sight,
148: providing an ideal system for studying interacting sub-clusters~\markcite{clowe2006}({Clowe} {et~al.} 2006).
149:
150: The mass surface density of the Bullet cluster has been measured using
151: weak and strong gravitational lensing of light from background
152: galaxies. There are significant angular offsets between the peaks of
153: the X-ray surface brightness, which trace the baryonic gas through
154: thermal bremsstrahlung emission,
155: and the peaks of the lensing surface
156: density, which are associated with the majority of the mass. The
157: combined weak and strong lensing analyses of \markcite{bradac2006}{Brada{\v c}} {et~al.} (2006) show
158: that the main cluster and sub-cluster are separated from their
159: associated X-ray peaks at $10\,\sigma$ and $6\,\sigma$ significance
160: respectively. This result has been recognized as providing direct
161: evidence for the presence of collisionless dark matter in this
162: system~\markcite{clowe2006}({Clowe} {et~al.} 2006).
163:
164: The Sunyaev-Zel'dovich effect (SZE) provides an independent probe of the intracluster gas.
165: In the SZE, a small fraction
166: ($\sim1\%$) of cosmic microwave background (CMB) photons undergo
167: inverse Compton scattering from intracluster
168: electrons~\markcite{sunyaev1970,birkinshaw1999}({Sunyaev} \& {Zel'dovich} 1970; {Birkinshaw} 1999). This process distorts
169: the Planck blackbody spectrum of the CMB and produces a signal
170: proportional to the gas pressure integrated along the line of sight.
171: At $150\,$GHz, the SZE produces a temperature decrement with respect to the
172: unperturbed CMB intensity.
173: Early detections of the SZE in the Bullet cluster include
174: work by \markcite{andreani1999}{Andreani} {et~al.} (1999) and \markcite{gomez2003}{Gomez} {et~al.} (2004).
175:
176: Unlike the X-ray surface brightness, the peak SZE surface brightness for a
177: given cluster is independent of redshift.
178: Therefore, the SZE has the potential to be an effective probe of intracluster gas
179: out to the redshifts at which clusters are assembled.
180: SZE measurements of galaxy clusters provide
181: complementary constraints on cluster properties typically derived
182: from X-ray measurements such as central electron density, core
183: radius of the intracluster gas, cluster gas mass, and fraction of the total cluster mass
184: in gas. Since the SZE and X-ray signals are proportional to the line-of-sight integral of
185: the electron density and electron density squared respectively, SZE results will be less
186: sensitive to clumping of the intracluster gas.
187: For all comparisons between SZ and X-ray data, we assume
188: a $\Lambda$CDM cosmology, with $h = 0.7$, $\Omega_{\rm{m}} = 0.27$, and $\Omega_{\Lambda} = 0.73$.
189:
190: In this paper, we present a 1\arcmin\ resolution SZE image of the
191: Bullet cluster at $150\,$GHz made with the APEX-SZ instrument. It is the
192: first reported scientific result from observations with a large array of
193: multiplexed superconducting transition-edge sensor bolometers. In
194: \S\,\ref{SEC:observations}, we discuss the instrument and
195: observations. Calibration is discussed in \S\,\ref{SEC:calibration}.
196: In \S\,\ref{SEC:datareduction}, we describe the data
197: reduction procedure, and in \S\,\ref{SEC:results}, we present the
198: results of fits to the SZE surface brightness with cluster models,
199: including mass-weighted electron temperature and gas mass
200: fraction calculations.
201: We summarize the conclusions and discuss
202: future work in \S\,\ref{SEC:conclusions}.
203:
204:
205:
206: %--------------------------------------------------------------
207: \section{Observations}
208: \label{SEC:observations}
209:
210: APEX-SZ is a receiver designed specifically for SZE galaxy cluster
211: surveys~\markcite{schwan2003,dobbs2006,schwan2008}({Schwan} {et~al.} 2003; {Dobbs} {et~al.} 2006; Schwan {et~al.} 2009, in preparation). It
212: is mounted on the 12-meter diameter APEX telescope, located on the
213: Atacama plateau in northern Chile~\markcite{guesten2006}({G{\"u}sten} {et~al.} 2006). The observing
214: site was chosen for its extremely dry and stable atmospheric
215: conditions. The mean atmospheric transmittance is frequently better
216: than 95$\%$ in the APEX-SZ frequency band at
217: 150~GHz~\markcite{peterson2003,chamberlain1995}({Peterson} {et~al.} 2003; {Chamberlain} \& {Bally} 1995). The telescope is capable
218: of round-the-clock observations.
219:
220: Three reimaging mirrors in the Cassegrain cabin couple the APEX
221: telescope to the focal plane of APEX-SZ. We achieve the diffraction
222: limited performance of the telescope across the entire 0.4\degree\
223: field of view with a mean measured beam full width half maximum (FWHM)
224: of 58\arcsec, and a measured beam solid angle of 1.5~arcmin$^2$,
225: including measured sidelobes at the $-14$~dB level.
226:
227: The APEX-SZ receiver houses a cryogenic focal plane, operating at 0.3~K.
228: The focal plane contains 330 horn-fed absorber-coupled superconducting
229: transition-edge sensor
230: bolometers~\markcite{ref:richards-bolometer-review,lee1996}(Richards 1994; {Lee} {et~al.} 1996),
231: with 55 detectors on each of six sub-array wafers. Of the 330
232: detectors, 280 are read out with the current frequency-domain multiplexed
233: readout hardware. We measure
234: the median individual pixel Noise Equivalent Power (NEP) to be $10^{-16}
235: $~W/$\sqrt{{\rm Hz}}$ and the median Noise Equivalent Temperature (NET)
236: to be $860~\mu$K$_{\rm{CMB}}\sqrt{{\rm s}}$.
237: The measured optical bandwidth of the receiver is 40\% narrower
238: than the design goal of 38~GHz, resulting in lower sensitivity than
239: anticipated.
240:
241: The large field of view of the APEX-SZ instrument is designed for
242: surveying large areas of sky. In order to efficiently observe a single
243: target, we use the circular scan pattern illustrated in
244: Figure~\ref{FIG:scanpattern}. The circle center is fixed in AZ/EL
245: coordinates for twenty circular sub-scans, with a total duration of 100
246: seconds. This choice has a number of important advantages.
247: The sky signal is modulated so that it appears in the timestream at frequencies
248: higher than atmospheric drifts and readout $1/f$ noise.
249: In addition, the circle scan has a moderate continuous acceleration; the lack
250: of high acceleration turn-arounds makes it possible to achieve a high observing
251: efficiency.
252: Approximately $20\%$ of the total observing time is spent moving the telescope
253: to a new center position before the start of the next scan.
254: Every bolometer maps a $12\arcmin \times 25\arcmin$ sub-field,
255: with a combined map field of $36\arcmin \times 48\arcmin$ every 100
256: seconds.
257:
258: \begin{figure}[h]\centering
259: \includegraphics[width=0.45\textwidth]{f1.pdf}
260: \caption[]{ The 100-second circular drift scan pattern. The solid
261: line shows the track of the center of the array. One circle has a
262: period of five seconds. The dashed line shows the instantaneous
263: field of view of the bolometer array. The $+$ marker indicates the source
264: position with respect to the scan pattern. The center of the
265: circles is constant in azimuth and elevation as the source drifts
266: across the field. The small disk in the lower left indicates the
267: 58\arcsec\ mean FWHM beam for a single bolometer.}
268: \label{FIG:scanpattern}
269: \end{figure}
270:
271: Observations of the Bullet cluster were conducted over a period of
272: seven days in August 2007, when the cluster was visible between the
273: hours of 03:00 and 15:00 local time. The weather over this period was
274: typical for the site, with precipitable water vapor varying between
275: 0.25 and 1.5 mm, and a median atmospheric transmittance of $97\%$. For
276: the analysis in this paper, 235 scans are used, each scan consisting
277: of twenty 5-s circular sub-scans, for a total of 6.4 hours of on-source
278: data.
279:
280:
281: %--------------------------------------------------------------
282: \section{Calibration}
283: \label{SEC:calibration}
284:
285: The response of the receiver to astronomical sources is measured with
286: daily raster scans of Mars over every bolometer in the
287: array. For each bolometer, the observations provide a primary flux
288: calibration and a high signal-to-noise beam profile from which we
289: determine beam parameters such as size, ellipticity, and position with
290: respect to the array-center pointing.
291: Additional observations of RCW57
292: and RCW38 are used to monitor gain stability, and frequent
293: observations of bright quasars near the cluster source are used to
294: monitor pointing stability.
295:
296: The WMAP satellite has been used to calibrate the brightness
297: temperature of Mars at $93\,$GHz in five measurement periods spanning
298: several years \markcite{hill08}({Hill} {et~al.} 2009). The WMAP Mars temperatures are tied to
299: the CMB dipole moment and are accurate to better than $1.0\%$. The
300: brightness temperature of Mars changes significantly ($\sim 15\%$) as
301: a function of its orbit and orientation. We use a version of the Rudy
302: Model \markcite{rudy87,muhleman91}({Rudy} {et~al.} 1987; {Muhleman} \& {Berge} 1991), that has been updated and maintained
303: by Bryan
304: Butler,\footnote{http://www.aoc.nrao.edu/\~{}bbutler/work/mars/model/}
305: to transfer the WMAP Mars temperature results to the APEX-SZ frequency
306: band and specific times of our Mars observations.
307:
308: After applying a constant scaling factor, we find the Rudy model predictions
309: for the Mars brightness temperature to be in excellent agreement with
310: the WMAP measurements.
311: We find that the Rudy model
312: brightness temperatures at $93\,$GHz are systematically a factor of
313: $1.052 \pm 0.010$ higher than those measured by WMAP in the five
314: published observation periods.
315: In contrast, repeating the same exercise with the thermal model
316: developed by \markcite{wright76,wright07}{Wright} (1976, 2007), as implemented in the
317: online JCMT-FLUXES program,\footnote{http://www.jach.hawaii.edu/jac-bin/planetflux.pl}
318: results in a scaling factor of $1.085\pm 0.043$.
319: This is consistent with the 10\% rescaling of this model
320: called for in \markcite{hill08}{Hill} {et~al.} (2009), but the scaling factor exhibits significantly
321: larger rms scatter than that of the Rudy Model.
322:
323: We therefore use the WMAP $93\,$GHz calibrated Rudy Model to compute
324: the Mars brightness temperatures at $150\,$GHz for the specific times
325: of our Mars observations by reducing the Rudy model $150\,$GHz
326: temperatures by a factor of 1.052. The Rudy model $93\,$GHz to $150\,$GHz frequency scaling factor
327: is $1.016 \pm 0.009$ at the times of our Mars observations, and we
328: adopt the rms scatter in this frequency scaling factor as an estimate of its
329: uncertainty. Combining the uncertainties in the WMAP Mars calibration,
330: the WMAP to Rudy model scaling factor at $93\,$GHz, and the Rudy model frequency
331: scaling factor, we estimate the uncertainty in Mars temperature to be $\pm 1.7\%$.
332:
333: The measured signals from the calibrators are corrected for atmospheric
334: opacity, which is measured with a sky dip observation at the beginning and end of each day's
335: observations. Measured zenith transmittance over the observing period ranged
336: between 0.92 and 0.98, with a median of 0.97. Based on the
337: observed temporal variability of
338: the opacity, drifts in atmospheric opacity between the sky dip and
339: observation contribute $<0.4\%$ to the overall calibration uncertainty.
340: After correcting for the atmospheric opacity, we find that the
341: Mars temperature measured by APEX-SZ varies from the model prediction
342: by up to $\sim 3\%$ over the course of the observation period.
343: This gain variation is included as a source of error
344: in the final calibration uncertainty.
345: The APEX-SZ observing band center is measured with a Fourier
346: transform spectrometer to be $152\pm2$ GHz.
347: The uncertainty in the band center results in a $\pm 1.4\%$ uncertainty
348: in extrapolation of the Mars based calibration to CMB temperature.
349:
350: The beam shape, including near sidelobes, is characterized by creating
351: a beam map from Mars observations, combining the same bolometer
352: channels that are used to make the science maps. We adjust the
353: calibration and measured beam size for the small ($\sim 1\%$)
354: correction due to the $8^{\prime \prime}$ angular size of Mars. We estimate
355: a fractional uncertainty in the beam solid angle of $\pm 4\%$.
356:
357: The APEX-SZ detectors operate in a state of strong negative
358: electrothermal feedback which results in a linear response to changes
359: in the input optical power.
360: We have measured the response of the detectors during sky dips
361: between $90^\circ$ and $30^\circ$ elevation (antenna temperature
362: difference $\sim13\,$K),
363: and find no significant deviation from the expected linear response to loading.
364: We therefore conclude that detector
365: non-linearity makes a negligible contribution to the calibration
366: uncertainty.
367:
368: Slowly changing errors in telescope pointing result in both a
369: pointing uncertainty and a flux calibration
370: uncertainty due a broadening of the effective beam pattern.
371: To measure pointing errors during our
372: observations, we observe a bright quasar within a few degrees of the
373: Bullet cluster every 1--2 hours, and apply a pointing correction
374: as needed.
375: The typical RMS pointing variations of the APEX telescope between
376: quasar observations is $\sim 4 \arcsec$.
377: This pointing uncertainty results in a slightly larger effective
378: beam for the coadded maps than is measured with the individual calibrator
379: maps.
380: The correction to the flux calibration of the coadded maps due to
381: pointing uncertainty is negligible,
382: particularly for the observation of extended objects such as the
383: Bullet cluster. We estimate the pointing uncertainty
384: in the coadded maps to be $\pm 4\arcsec$ in both
385: the RA and DEC directions.
386:
387: The uncertainty in the CMB temperature calibration of the APEX-SZ
388: maps is summarized in Table~\ref{TBL:cal}.
389: The combination of all contributions to the calibration uncertainty
390: described above results in an overall point source flux uncertainty of
391: $\pm 5.5\%$.
392:
393: \begin{deluxetable}{ll}
394: %\tabletypesize{\small}
395: \tablecaption{\label{TBL:cal} APEX-SZ Flux Calibration Uncertainty}
396: \tablewidth{0pt}
397: \tablehead{
398: \colhead{Source} & \colhead{Uncertainty}}
399: \startdata
400: WMAP Mars temperature at $93\,$GHz & $\pm1.0\%$ \\
401: Rudy model to WMAP scaling factor at $93\,$GHz & $\pm1.0\%$ \\
402: $93\,$GHz to $150\,$GHz frequency scaling factor & $\pm0.9\%$ \\
403: Frequency band center & $\pm1.4\%$ \\
404: Beam solid angle & $\pm4.0\%$ \\
405: Atmospheric attenuation & $\pm0.4\%$ \\
406: Gain variation & $\pm3.0\%$ \\\hline
407: Total & $\pm5.5\%$ \\
408:
409: \enddata
410: \end{deluxetable}
411:
412:
413: \section{Data Reduction}
414: \label{SEC:datareduction}
415:
416: The data consist of 280 bolometer timestreams sampled at 100
417: Hz, telescope pointing data interpolated to the same rate,
418: housekeeping thermometry data, bolometer bias and readout
419: configuration data, and other miscellaneous monitoring data. The
420: fundamental observation unit is a scan comprising twenty 5-s
421: circular sub-scans in AZ/EL coordinates, allowing the source to drift
422: through the field of view (FOV), as described in \S\,\ref{SEC:observations} above.
423:
424: Data reduction consists of cuts to remove
425: poor-quality data, filtering of $1/f$ and correlated noise due to
426: atmospheric fluctuations, and binning the bolometer data into maps.
427: These steps are described in more detail below.
428:
429:
430: \subsection{Timestream Data Cuts}
431:
432: Timestream data are first parsed into individual circular sub-scans.
433: We reject $\sim 7$\% of the data at the beginning and end of the scan
434: where the telescope deviates from the constant angular velocity
435: circular pattern.
436: We reject bolometer channels that are optically or electronically unresponsive, or lack high-quality flux calibration data;
437: typically, 160--200 of the 280 bolometer channels remain after these
438: preliminary cuts. The large number of rejected channels is due primarily
439: to low fabrication yield for two of the six bolometer sub-array wafers.
440:
441: We reject spikes and step-like glitches caused by cosmic rays or
442: electrical interference. These are infrequent and occur on time scales
443: faster than the detector optical time constant. We use a simple
444: signal-to-noise cut on the data to reject these, since the
445: timestream is noise dominated even for the $\sim$1~mK
446: Bullet cluster signal. Step-like glitches are often correlated
447: across many channels in the array, so we reject data from all channels whenever a
448: spike or glitch in $\geq 2\%$ of the channels is detected. These cuts result
449: in a loss of 8\% of the remaining data.
450:
451: For each circular sub-scan, we also reject channels that have excessive
452: noise in signal band, resulting in a loss of 19\% of the remaining data.
453:
454: \subsection{Atmospheric Fluctuation Removal}
455: \label{SEC:atmremoval}
456:
457: After the timestream data cuts, fluctuations in atmospheric emission
458: produce the dominant signal in the raw bolometer timestreams. The
459: atmospheric signal is highly correlated across the array, which can be
460: exploited to remove the signal.
461: Principal component analysis (PCA) has been used by some groups
462: to reduce the atmospheric signal \markcite{scott2008,laurent2005}(see, e.g., {Scott} {et~al.} 2008; {Laurent} {et~al.} 2005).
463: However, the effect of PCA filtering on the source is difficult to predict and
464: is a function of the atmospheric conditions.
465: We have developed an analysis strategy that reduces the atmospheric signal
466: through the application of spatial filters that have a constant and well
467: understood effect on the signals we are attempting to measure.
468:
469: Atmospheric fluctuation power is expected to follow a Kolmogorov
470: spatial power spectrum, with most power present on scales larger
471: than the separation between beams as they pass through the atmosphere,
472: resulting in an atmospheric signal that is highly correlated across the array.
473: To reduce these fluctuations, we first remove a polynomial and an elevation dependent
474: airmass opacity model from each channel's timestream, then remove
475: a first-order two-dimensional spatial polynomial across the
476: array for each time-step. This algorithm is described in detail
477: in the two following subsections. This atmospheric fluctuation removal
478: strategy requires that both the spatial extent of the scan
479: pattern and the instantaneous array FOV are larger than the
480: source. The $6\arcmin$ radius circular scan and the 23$\arcmin$ array
481: FOV allow us to recover most of the Bullet cluster's flux,
482: but some extended emission is lost as is described in \S\,\ref{SEC:results}.
483:
484: \subsubsection{Timestream Atmosphere Removal}\label{subsec:circle}
485:
486:
487: We observe scan-synchronous signals in the bolometer timestreams due
488: to elevation-dependent atmospheric emission. The optical path length
489: $L$ through the atmosphere is proportional to the cosecant of the
490: elevation angle $\epsilon$, $L \propto \csc(\epsilon)$. The change in
491: optical path-length is nearly a linear function of elevation angle
492: over the $6\arcmin$ circular scan radius. In the circular drift scans,
493: this modulation of the elevation-dependent opacity produces an
494: approximately sinusoidal modulation in the bolometer timestream. For
495: each channel in each scan, we simultaneously fit and remove an
496: atmospheric model consisting of this cosecant function plus an order
497: 20 polynomial (one degree of freedom per circular sub-scan) to remove
498: slow drifts in the atmospheric opacity and readout $1/f$ noise.
499: This scheme effectively
500: removes the common scan-induced atmospheric signal as well as most of
501: the atmospheric fluctuation power below the frequency of the circular
502: sub-scan (0.20 Hz), while only modestly affecting the central Bullet
503: cluster signal.
504:
505: \subsubsection{Spatially Correlated Atmosphere Removal}
506:
507: Removing the cosecant-plus-polynomial model
508: from the timestream data reduces low-frequency
509: atmospheric fluctuation power, but not higher frequency power
510: corresponding to smaller spatial scales near those where the cluster
511: signal occurs. To reduce these fluctuations, the atmosphere can be modeled
512: as a spatially correlated signal across the array pixel positions on
513: the sky with a low-order two-dimensional polynomial function. At each
514: time step, we fit and remove a low-order
515: two-dimensional polynomial
516: function across the array, similar to the procedure described in
517: \markcite{sayers07}Sayers (2007).
518: The relative gain coefficients for each bolometer channel are
519: calculated by taking the ratio of each channel's timestream, which is
520: dominated by correlated atmospheric noise, to a median timestream
521: signal generated from all channels. With the favorable atmospheric
522: conditions of these observations, we find that a first order spatial
523: polynomial (offset and tilt) is adequate to remove most of the atmospheric signal while
524: preserving the cluster signal.
525:
526: Bolometer channels with excess uncorrelated noise are
527: more easily identified after removing the correlated atmospheric
528: noise component; we reject these noisy channels, then perform the
529: spatially correlated signal removal a second time.
530:
531: \subsection{Map-Making}
532: \label{SEC:mapmaking}
533:
534: The atmospheric removal algorithms described above act as a high-pass
535: filter. They suppress signals on scales comparable to the scan length
536: or the focal plane FOV. The cluster emission can be quite
537: extended, and therefore the data reduction filtering process attenuates
538: diffuse flux in the cluster signal and produces small
539: positive sidelobes around the cluster decrement. The data reduction
540: pipeline filters can be tailored, within limits, to meet various
541: scientific objectives. Thus, our primary data products consist of two
542: different high signal-to-noise maps of the cluster.
543:
544: For one map, we mask a circular region centered on the cluster source
545: prior to fitting the timestream and spatial filters described in
546: \S\,\ref{SEC:atmremoval}, then apply the resulting filter functions to
547: the entire data set, including the source region. The source-mask
548: procedure prevents the cluster signal within the masked region from
549: influencing the baseline fits, and thus reduces attenuation of the
550: source central decrement and extended emission at the expense of
551: increasing the contribution of low-frequency noise in the map
552: center. We choose a source-mask radius of $4.75\arcmin$ as a
553: compromise between attenuation of diffuse emission and increased map
554: noise. We use the source-masked map to visually interpret the
555: morphology and extended emission in the cluster. These results are discussed in \S\,\ref{subsec:tempmap}.
556:
557: We also produce a map in which we do not mask the source when applying
558: filters. The non-source-masked
559: map is used for model parameter estimation because it has higher signal to noise
560: in the central region of the map. In addition, it is easier to take
561: into account the effects of the data reduction filters, or transfer
562: function, on the underlying sky intensity distribution, which is
563: necessary for comparing the data to the model for parameter
564: estimation. The fitting procedure and results are described in more detail in
565: \S\,\ref{subsec:fitting}.
566:
567: For each of the two maps, the post-cut, filtered timestream data are
568: binned in angular sky coordinates to create maps. For a given scan, a
569: map is created from each bolometer channel, applying the channel's
570: pointing offset and flux calibration.
571: A coadded scan map is created by combining individual channel maps
572: with minimum variance weighting in each pixel, using the sample variance
573: of the conditioned timestream data in the scan.
574: The final coadded map
575: is created by combining all scan maps, again with minimum variance
576: weighting in each pixel. We bin maps at a resolution of 10\arcsec\ to
577: oversample the beam. The source-masked map that we present
578: in \S\,\ref{SEC:results} is
579: convolved with a 1\arcmin\ FWHM Gaussian to smooth noise fluctuations
580: to the angular size of the beam. However, the radial profiles
581: presented below and the non-source-masked map used for model
582: fitting do not include this additional smoothing.
583:
584:
585: %--------------------------------------------------------------
586: \section{Results}
587: \label{SEC:results}
588:
589: \begin{figure*}[p!]\centering
590: \includegraphics[width=0.7\textwidth]{f2a.pdf}
591: \includegraphics[width=0.7\textwidth]{f2b.pdf}
592: \caption[]{ (Top) Temperature map of the Bullet cluster system from
593: the source-masked data reduction, with scale in CMB temperature
594: units. The circle in the lower left corner represents the 85\arcsec\
595: FWHM map resolution which is the result of the instrument beam and
596: data reduction filter convolved with the 1\arcmin\ FWHM Gaussian
597: smoothing applied to the map. (Bottom) Difference map made by
598: multiplying alternate scan maps by $+1$ and $-1$,
599: respectively, then coadding all scan maps with minimum variance
600: weighting, in the same manner as was used to produce the temperature
601: map shown in the top panel. The contour interval is 100~$\mu \rm{K}_{\rm{CMB}}$
602: in both maps.}
603: \label{FIG:bulletcoadd}
604: \end{figure*}
605:
606: \subsection{Temperature Map}
607: \label{subsec:tempmap}
608:
609: Figures \ref{FIG:bulletcoadd}~and~\ref{FIG:bulletcoaddzoom} show the
610: source-masked temperature map from our observations of the Bullet
611: cluster. The map has a resolution of 85\arcsec\ FWHM which results from the
612: combination of the 58\arcsec\ instrumental beam, the data reduction filters, and
613: a final 1\arcmin\ FWHM Gaussian smoothing of the map. The
614: source-masked map is shown in order to provide a more accurate representation of the
615: extended emission and cluster morphology.
616: The noise in the central region of the source-masked map
617: is 55~$\mu \rm{K}_{\rm{rms}}$ per 85\arcsec\ FWHM resolution
618: element.
619: Near the cluster center, the emission hints at elongation
620: in the East-West direction, which is along the axis between the main
621: and sub-cluster gas detected in the X-ray, see \S\,\ref{sec:xray}.
622: The more extended emission appears to be elongated
623: in the Northwest-Southeast direction, which is the major axis of the best-fit
624: elliptical $\beta$ model discussed in \S\,\ref{subsec:fitting}.
625: Figure~\ref{FIG:bulletcoaddzoom} shows the centroid position of the
626: best-fit elliptical $\beta$ model, and the position of the dust
627: obscured, lensed galaxy detected at 270~GHz by \markcite{wilson2008}{Wilson} {et~al.} (2008).
628: As discussed in \S\,\ref{SEC:source_contributions},
629: we see no evidence for emission from this source in our 150~GHz map.
630:
631: Radial profiles for the unsmoothed source-masked and non-source-masked
632: maps are shown in Figure~\ref{FIG:radial_plots}. The source-masked map
633: has a signal-to-noise of 10 within the central 1\arcmin\ radius,
634: compared to 23 for the non-source-masked map, due to the fact that
635: source-masking allows more large-scale atmospheric fluctuation noise
636: to remain in the map. However, the source-masked map preserves signal on
637: larger spatial scales than the non-source-masked map. In both the
638: source-masked and non-source-masked maps, the sky intensity
639: distribution is filtered by the instrument beam and data reduction
640: pipeline described in \S\,\ref{SEC:datareduction}. We do not
641: renormalize the map amplitudes, since the source is extended and an
642: assumption would need to be made about the shape of the sky-brightness
643: distribution to do so. However, in order to accurately estimate
644: cluster parameters such as the central temperature decrement, a model
645: for cluster emission must be adopted, and the instrument beam and data
646: reduction filtering must be taken into account.
647:
648: \begin{figure}[h!]\centering
649: \includegraphics[width=0.45\textwidth]{f3.pdf}
650: \caption[]{ Temperature map detail from
651: Figure~\ref{FIG:bulletcoadd}, with color scale adjusted to the
652: limits of the detail region, and a contour interval of 100~$\mu \rm{K}_{\rm{CMB}}$.
653: The $+$ marker indicates the centroid
654: position of the best-fit elliptical $\beta$ model, see
655: \S\,\ref{subsec:fitting}. The $*$ marker indicates the position of
656: the bright, dust obscured, lensed galaxy detected at 270~GHz by
657: \markcite{wilson2008}{Wilson} {et~al.} (2008), see \S\,\ref{SEC:source_contributions}.}
658: \label{FIG:bulletcoaddzoom}
659: \end{figure}
660:
661: \subsection{Fit to Elliptical $\beta$ Model}\label{subsec:fitting}
662:
663: We fit an elliptical $\beta$ model to the non-source-masked temperature map
664: to allow a straightforward comparison of cluster gas properties
665: derived from our measurements to those derived from X-ray
666: observations. In all analyses here we assume the cluster gas is
667: well-described by an isothermal $\beta$ model, and is in hydrostatic
668: equilibrium. These assumptions are unphysical in the case of the
669: Bullet cluster, which is a dynamically complex merging system where
670: the gas is separated from the rest of the mass~\markcite{clowe2006}({Clowe} {et~al.} 2006).
671: However, we find that with the sensitivity and spatial
672: resolution of the observations, these assumptions yield
673: an adequate description of the observed emission.
674:
675: We model the three-dimensional radial profile of the
676: electron density with an isothermal $\beta$ model
677: \markcite{cavaliere1978}({Cavaliere} \& {Fusco-Femiano} 1978):
678:
679: \begin{equation}
680: \label{EQN:betamodeldensity}
681: n_e(r) = n_{e0}\left( 1 + \frac{r^2}{r_c^2}\right)^{-3\beta/2}.
682: \end{equation}
683: Here, $n_{e0}$ is the central electron number density, $r_c$ is the
684: core radius of the gas distribution, and $\beta$ describes the
685: power-law index at large radii.
686:
687: The radial surface temperature
688: profile of the SZE takes a simple analytic form:
689: \begin{equation}
690: \label{EQN:betamodeltemperature}
691: \Delta T_{\rm{SZ}} = \Delta
692: T_{0}\left( 1 + \frac{\theta^2}{\theta_c^2} \right)^{(1-3\beta)/2},
693: \end{equation}
694: where $\Delta T_{0}$ is the central temperature decrement, and
695: $\theta_{c} = r_c/D_A$ is the core radius divided by the angular-diameter distance. A similar form exists for the
696: X-ray surface brightness.
697:
698:
699: Because of the significant ellipticity in the measured SZE intensity
700: profile, we generalize the cluster gas model to be a spheroidal rather
701: than spherical function of the spatial coordinates:
702: %
703: \begin{equation}
704: \label{EQN:ellipticalbeta}
705: \Delta T_{\rm{SZ}} = \Delta T_{0}\left(1 + A + B \right)^{(1-3\beta)/2},
706: \end{equation}
707: %
708: with
709: $$
710: A = \frac{(\cos(\Phi)(X - X_{\rm{0}}) + \sin(\Phi)(Y - Y_{\rm{0}}))^2}{\theta_{\rm{c}}^2},
711: $$
712: $$
713: B = \frac{(-\sin(\Phi)(X - X_{\rm{0}}) + \cos(\Phi)(Y - Y_{\rm{0}}))^2}{(\eta \theta_{\rm{c}})^2}.
714: $$
715: Here $(X - X_{\rm{0}})$ and $(Y - Y_{\rm{0}})$ are angular offsets on the sky in the RA and
716: DEC directions, with respect to center positions $X_{\rm{0}}$ and
717: $Y_{\rm{0}}$. The axial ratio, $\eta$, is the ratio between the minor and major axis
718: core radii, $\Phi$ is the angle
719: between the major axis and the RA ($X$) direction.
720: $\Delta T_{0}$ is given by the gas pressure integrated along the
721: central line of sight through the cluster:
722: %
723: \begin{equation}
724: \label{EQN:integratedpressure} \frac{\Delta T_{0}}{T_{\rm{CMB}}} =
725: \frac{k_{\rm{B}} \sigma_T}{m_e c^2} \int f(x,T_e) n_e(l) T_e(l) dl
726: \end{equation}
727: %
728: where $x = h \nu / k T$, $f(x,T_e)$ describes the frequency dependence
729: of the SZE, $\sigma_T$ is the Thomson scattering
730: cross-section, and $T_{\rm{CMB}} = 2.728$ K.
731: For all results in this paper, we use the relativistic SZE spectrum $f(x,T_e)$
732: provided by ~\markcite{nozawa2000}{Nozawa} {et~al.} (2000), and neglect the kinematic effect.
733: At 150~GHz and $T_e = 13.9$~keV (see \S\,\ref{sec:xray}), this is a 9\% correction to
734: the non-relativistic value.
735:
736: To accurately estimate $\beta$ model parameters for the cluster, the
737: instrument beam and data reduction filters, or transfer function, must
738: be applied to the model before comparing it with the data \markcite{benson2003,reese00}(see,
739: e.g., {Benson} {et~al.} 2003; {Reese} {et~al.} 2000). We characterize this transfer function
740: by creating a map from a simulated point source, convolved with the
741: instrument beam, and inserted into a noiseless timestream, similar to
742: the method described in \markcite{scott2008}{Scott} {et~al.} (2008). The point source transfer
743: function map $\mathcal{K}$ is then convolved with a simulated $\beta$
744: model cluster map $\mathcal{B}$ to generate a filtered model map
745: $\mathcal{B^\prime}$, which is a noiseless simulated APEX-SZ
746: observation of a $\beta$ model cluster. The filtered model map,
747: $\mathcal{B^\prime}$, is then differenced with the data map,
748: $\mathcal{M}$, and model parameters are estimated by minimizing a
749: $\chi^2$ statistic.
750:
751: Simulating maps of many different cluster models is required for model
752: parameter fitting. Convolving the cluster model with the point source
753: transfer function map is much faster than processing each model
754: through the reduction pipeline. We find that the resulting simulated
755: maps from both methods agree sufficiently well to have negligible
756: effect on the parameter estimation results.
757:
758: We use the unsmoothed non-source-masked map with $10\arcsec$
759: pixelization described in \S\,\ref{SEC:mapmaking} for all parameter
760: estimation described below, since this map has lower noise and a more
761: easily characterized transfer function than the source-masked map
762: shown in Figure~\ref{FIG:bulletcoadd}. Diffuse cluster emission is
763: more attenuated in the non-source-masked map, but this
764: is taken into account using the point source transfer function.
765:
766: Map noise properties are assessed in the spatial frequency domain
767: using jackknife noise maps \markcite{sayers07,sayers08}(see Sayers 2007; {Sayers} {et~al.} 2009).
768: To estimate the noise covariance $C_n$, we assume that the noise is
769: stationary in the map basis. With this assumption, the Fourier
770: transform of the noise covariance matrix, $\widetilde{C}_n$, is
771: diagonal, and the diagonal elements are equal to the noise map power
772: spectral density (PSD). For each of 500 jackknife noise map
773: realizations, we find the two-dimensional Fourier transform, then
774: average over all realizations.
775: This averaged map PSD is the experimental estimate of the diagonal elements of
776: $\widetilde{C}_n$.
777: However, these jackknife maps do not include fluctuations due to the
778: primary CMB anisotropies. We estimate the CMB signal covariance from
779: the WMAP5 best-fit power spectrum \markcite{nolta09}({Nolta} {et~al.} 2009) convolved with the point source
780: transfer function described earlier and add it to the jackknife noise
781: PSD to determine the total covariance matrix.
782:
783:
784: We construct a $\chi^2$ statistic for the
785: model fit using the transform of the filtered $\beta$ model,
786: $\widetilde{\mathcal{B^\prime}}$, and the transform of the
787: central $14\arcmin \times 14\arcmin$ portion of the data map
788: $\widetilde{\mathcal{M}}$ as:
789: \begin{equation}
790: \chi^2 = (\widetilde{\mathcal{M}}-\widetilde{\mathcal{B^\prime}})^T \widetilde{C}_n^{-1} (\widetilde{\mathcal{M}}-\widetilde{\mathcal{B^\prime}}).
791: \end{equation}
792: Using Markov Chain Monte Carlo (MCMC) methods, the likelihood,
793: $\mathcal{L} = e^{-\frac{1}{2} \chi^2}$, is sampled in the 7-dimensional
794: model parameter space and integrated to find
795: the marginal likelihood distributions of the $\beta$ model parameters.
796: The model parameter estimates and uncertainties that we report are the
797: maximum likelihood values and constant-likelihood 68\% confidence
798: intervals, respectively, of the marginal likelihood distributions.
799:
800: The above approach to noise covariance estimation is chosen for its
801: simplicity and because we do not have enough linear combinations of
802: individual scan maps to fully sample the noise covariance matrix using
803: jackknife noise maps. But, the method relies on several simplifying
804: assumptions, including that the bolometer noise is stationary for each
805: 100-s scan, the timestream noise is uncorrelated from scan to scan,
806: and the map coverage is uniform. Our map coverage is not actually
807: uniform, but we find through simulations of non-uniform Gaussian
808: noise maps that the $\chi^2$ statistic is not significantly affected.
809: In addition, the validity of the approach is tested by inserting
810: simulated clusters into the real timestream data; the simulated
811: cluster parameters are accurately recovered within the estimated
812: uncertainties.
813:
814: Results of the $\beta$ model parameter estimation are given in
815: Table~\ref{TBL:betamodelfits}. Due to the degeneracy between the core
816: radius $\theta_{\rm{c}}$ and $\beta$ parameters, we assume a prior probability
817: density on $\beta$ of $1.04^{+0.16}_{-0.10}$, which is found from fits to ROSAT X-ray data
818: by \markcite{ota2004}{Ota} \& {Mitsuda} (2004). \markcite{hallman2007}{Hallman} {et~al.} (2007) find that in hydro/N-body simulations,
819: $\beta$ derived from fits to SZE profiles is higher than that from
820: X-ray, with $\beta_{\mathrm{SZ}}/\beta_{\mathrm{X}\mbox{-}\mathrm{ray}} = 1.21 \pm 0.13$ for fits
821: within $r_{500}$. We do not account for that factor here due to the
822: significant uncertainty in the X-ray derived $\beta$ value, but we
823: note that our SZE data prefer a higher value for $\beta$ than
824: the peak value of the prior. We further discuss this choice of prior in
825: \S\,\ref{subsec:mwte}.
826:
827: The best-fit $\beta$ model fits the data well, with a reduced $\chi^2$
828: value of 1.008, and with 7219 degrees of freedom (DOF) has a probability
829: to exceed (PTE) of 31.5\%. The difference map between the data map,
830: $\mathcal{M}$, and the best-fit filtered $\beta$ model,
831: $\mathcal{B^\prime}$, shows no evidence of residual cluster structure
832: or point sources.
833:
834: \begin{deluxetable}{llll}
835: %\tabletypesize{\small}
836: \tablecaption{\label{TBL:betamodelfits} $\beta$ Model Fit Results}
837: \tablewidth{0pt}
838: \tablehead{
839: \colhead{Parameter} & \colhead{Description} & \colhead{Value} & \colhead{Uncertainty$^a$}
840: }
841: \startdata
842: $X_{\rm{0}}$ & RA centroid position &$06^{\rm h}58^{\rm m}30.86^{\rm s}$ (J2000) & $\pm 7.4\arcsec$ \\
843: $Y_{\rm{0}}$ & DEC centroid position & $-55\degree56\arcmin46.2\arcsec$ (J2000) & $\pm 7.3\arcsec$ \\
844: $\Delta T_{\rm{0}}$ & Central temperature decrement & $-771\ \mu$K$_{\rm{CMB}}$ & $\pm 71\ \mu$K$_{\rm{CMB}}$ \\
845: $y_0$ & Central Comptonization$^b$($T_e = 13.9$~keV) & $3.31 \times 10^{-4}$ & $\pm 0.30 \times 10^{-4}$ \\
846: $y_0$ & Central Comptonization$^b$($T_e = 10.6$~keV) & $3.24 \times 10^{-4}$ & $\pm 0.30 \times 10^{-4}$ \\
847: $\theta_{\rm{c}} $ & Core radius & $142 \arcsec$ & $\pm 18 \arcsec$ \\
848: $\eta$ & Ellipse minor/major core radius ratio & 0.889 & $\pm 0.072$ \\
849: $\Phi $ & Ellipse orientation angle & $-52\degree$ & $\pm 20\degree$ \\
850: $\beta$ & Power-law index & 1.15 & $\pm 0.13$ \\
851:
852: \enddata
853: \tablenotetext{a}{Quoted uncertainties are 68\% confidence intervals
854: in the marginal likelihood distribution for each parameter. The
855: uncertainty in $\Delta T_{\rm{0}}$ includes a statistical uncertainty
856: from the fit of $\pm 57\,\mu$K, and a $\pm 5.5\%$ flux calibration
857: uncertainty. The uncertainties in the centroid parameters $X_{\rm{0}}$
858: and $Y_{\rm{0}}$ include a $\pm 4\arcsec$ pointing uncertainty
859: and are given in units of arcseconds on the sky.}
860: \tablenotetext{b}{Central Comptonization, $y_0$, is a derived parameter,
861: assuming an electron temperature of 13.9 keV from \markcite{govoni2004}{Govoni} {et~al.} (2004)
862: and 10.6~keV from \markcite{zhang2006}{Zhang} {et~al.} (2006), an SZE observation frequency of 152~GHz, and
863: $T_{\rm CMB} = 2.728$~K. It is provided to facilitate comparisons with data at other
864: wavelengths.}
865:
866:
867:
868: %%%%%%%%%%%
869: \end{deluxetable}
870:
871:
872: Radial profile plots of the best-fit $\beta$ model, $\mathcal{B}$, and
873: the filtered $\beta$ model map, $\mathcal{B^\prime}$, are shown in
874: Figure~\ref{FIG:radial_plots}. Also plotted for comparison are the
875: radially binned data from the unsmoothed non-source-masked map
876: $\mathcal{M}$, used for model fitting, and the unsmoothed
877: source-masked map, used to visualize extended emission (without the
878: 1\arcmin\ Gaussian smoothing used in
879: Figures~\ref{FIG:bulletcoadd}~\&~\ref{FIG:bulletcoaddzoom}).
880: Uncertainties in both sets of radially binned data are highly
881: correlated due to large-spatial-scale correlated noise in the
882: maps. The coincidence of the non-source-masked data ($\mathcal{M}$,
883: filled circles) and the filtered best-fit $\beta$ model
884: ($\mathcal{B^\prime}$, red solid line) show that the data and best-fit
885: $\beta$ model are in good agreement. The source-masked map preserves
886: signal on larger spatial scales than the non-source-masked map, and is
887: useful for visualizing extended emission on larger spatial
888: scales. But, as expected, even the source-masked map attenuates signal
889: on scales exceeding the $4.75^\prime$ radius of the source masking, and thus has
890: a lower signal amplitude when compared with the unfiltered $\beta$ model
891: ($\mathcal{B}$, blue dashed line).
892:
893: \begin{figure}[h!]\centering
894: \includegraphics[width=0.45\textwidth]{f4.pdf}
895: \caption[]{Radial profile of Sunyaev-Zel'dovich effect in the Bullet
896: cluster compared to the best-fit $\beta$ model. Points with error
897: bars are the SZE data binned in 1\arcmin\ radial bins, from the
898: non-source-masked map, ($\mathcal{M}$, filled circles) and the
899: source-masked map (open circles). The lines show the radial profile
900: of the best-fit $\beta$ model, unfiltered ($\mathcal{B}$, blue
901: dashed line) and after convolving with the instrument beam and
902: non-source-masked data reduction filters ($\mathcal{B^\prime}$, red
903: solid line). The non-source-masked map radial profile is reasonably
904: well fit by the filtered $\beta$ model. The source-masked map
905: preserves signal on larger spatial scales than the non-source-masked
906: map, but still attenuates signal on scales exceeding the
907: $4.75^\prime$ radius of the source masking. The source-masked data
908: thus have a lower signal amplitude when compared with the unfiltered
909: $\beta$ model, as expected. See text for details.}
910: \label{FIG:radial_plots}
911: \end{figure}
912:
913: \subsection{Radio and IR source contributions}
914: \label{SEC:source_contributions}
915:
916: Radio sources associated with a galaxy cluster and background IR
917: galaxy sources can have a significant impact on the measurement of the
918: SZE emission at 150~GHz. We interpret the published results of observations
919: of the Bullet cluster at other frequencies and conclude that the measured
920: SZE decrement is not significantly contaminated.
921:
922: The most important source of potential confusion is a bright, dust obscured,
923: lensed galaxy in the direction of the Bullet cluster recently reported
924: by \markcite{wilson2008}{Wilson} {et~al.} (2008). This source has a flux density of $13.5 \pm
925: 0.5$~mJy at an observing frequency of 270~GHz, and is centered at
926: RA $06^{\rm h}58^{\rm m}37.31^{\rm s}$, DEC $-55\degree57\arcmin1.5\arcsec$ (J2000),
927: $\approx 56^{\prime \prime}$ to
928: the east of the measured SZE centroid position, see Figure~\ref{FIG:bulletcoaddzoom}.
929: Assuming a spectral index
930: $\alpha=3$, where $S \propto \nu^\alpha$, we expect a flux density of
931: $1.94\,$mJy at $150\,$GHz corresponding to a temperature increment of
932: $\Delta T = 38\ \mu{\rm K_{CMB}}$ in the 1.5~arcmin$^2$ APEX-SZ beam solid angle.
933: This lensed
934: source is expected to be the dominant contribution to positive flux in
935: the direction of the cluster and we have repeated the $\beta$ model
936: fit taking it into account. We first add the source at its
937: measured position with the predicted 150~GHz flux to the SZE $\beta$
938: model and repeat the model fit. As expected, including the point
939: source results in a slightly ($\sim\sigma/3$) deeper decrement;
940: however, the $\chi^2$ of the model fit slightly increases.
941: We next allow the flux of the point source to vary along
942: with the other model parameters and find that values of positive flux
943: are a poorer fit than no source at all.
944: Therefore, we have no evidence for significant emission from this source
945: at 150~GHz. For the results in this paper, we use
946: cluster model parameters derived from fits that do not include this IR
947: source.
948:
949: The Bullet cluster is also associated with a number of relatively
950: compact radio sources and one of the brightest cluster radio halos yet
951: discovered. However, these sources are predicted to produce negligible
952: temperature increments in the APEX-SZ beam when extrapolated to
953: 150~GHz.
954:
955: \markcite{liang2000}{Liang} {et~al.} (2000) report the detection of eight radio point sources
956: all of which have steeply falling spectra. Only two of these sources
957: were detected with ACTA at 8.8~GHz, and they were found to have flux
958: densities of $3.2 \pm 0.5$~mJy and $3.3 \pm 0.5$~mJy. The spectra of
959: these sources, measured between $4.9$ and $8.8\,$GHz are falling with
960: $\alpha = -0.93$ and $\alpha=-1.33$, respectively. Extrapolating to
961: $150\,$GHz, the flux of these sources are expected to be $0.24$ and
962: $0.07\,$mJy, corresponding to CMB temperature increments of $\Delta T
963: = 4.7$ and $1.4\,\mu{\rm K_{CMB}}$ in the 1.5~arcmin$^2$ APEX-SZ
964: beam solid angle.
965:
966: The radio halo in the Bullet cluster is very luminous, but has a
967: characteristically steeply falling spectrum. \markcite{liang2000}{Liang} {et~al.} (2000)
968: measure the flux and spectra for the two main spatial components of
969: the halo. At $8.8\,$GHz, they find the two components to have fluxes
970: of $3.5$ and $0.55\,$mJy, with spectral indices of $\alpha=-1.3$ and
971: $\alpha=-1.4$, respectively. Extrapolating to $150\,$GHz, the combined flux from the
972: radio halo is expected to be $\sim 0.1\,$mJy. This emission is spread
973: over an area comparable to the size of the cluster and therefore
974: corresponds to a temperature increment $< 1\,\mu{\rm K_{CMB}}$ in the
975: APEX-SZ beam.
976:
977: \subsection{Comparison with X-ray Data}\label{sec:xray}
978:
979: X-ray emission in the ionized intracluster gas is dominated by thermal
980: bremsstrahlung. The X-ray surface brightness can be written
981: \begin{equation}
982: \label{xraysurfacebrightness}
983: S_{X} = \frac{1}{4 \pi (1+z)^4} \int
984: n_{e} n_{i} \Lambda_{ei} dl,
985: \end{equation}
986: where $n_{e,i}$ are the electron and ion densities in this gas,
987: $\Lambda_{ei}$ is the X-ray cooling function, and the
988: integral is taken along the line of sight. The X-ray flux is
989: proportional to the line-of-sight integral of the square of the
990: electron density, resulting in emission that is more sensitive
991: to local density concentrations than the SZE emission.
992:
993: The Bullet sub-cluster and the bow shock are apparent in the X-ray
994: surface brightness map shown in Figure~\ref{FIG:bullet-overlay}.
995: The SZE contour map of the Bullet cluster in
996: Figures~\ref{FIG:bulletcoadd}~\&~\ref{FIG:bulletcoaddzoom} is
997: overlaid on an X-ray map and weak lensing
998: surface mass density reconstruction from
999: \markcite{clowe2006}{Clowe} {et~al.} (2006).\footnote{Data are publicly available at
1000: http://flamingos.astro.ufl.edu/1e0657/public.html.} The X-ray map is
1001: made from XMM data (observation Id: 0112980201) extracted in the
1002: [0.5-2]~keV band, corresponding to Bullet rest frame energies
1003: where the X-ray cooling function for hot gas is relatively
1004: insensitive to temperature. The map is smoothed with a 12\arcsec\
1005: Gaussian kernel.
1006:
1007: The SZE contours do not resolve the sub-cluster. However, an
1008: elongation of the inner contours to the West suggests that a
1009: contribution from it may be detected. The observed SZE map is
1010: consistent with expectations, given the 85\arcsec\ resolution of
1011: the SZE map, the different dependence of the X-ray and
1012: SZE signals on gas density, and the mass and temperature difference between the two
1013: merging components which predict a factor of $\sim10$ lower
1014: integrated pressure from the sub-cluster. There is no evidence in the
1015: SZE contours of a contribution from the lensed sub-mm bright galaxy
1016: discussed in \S\,\ref{SEC:source_contributions}.
1017:
1018: \begin{figure}[h]\centering
1019: \includegraphics[width=0.45\textwidth]{f5.pdf}
1020: \caption[]{The SZE map of the Bullet system from this work, in white
1021: contours, overlaid on an X-ray map from XMM observations. The green
1022: contours show the weak lensing surface mass density reconstruction
1023: from \markcite{clowe2006}{Clowe} {et~al.} (2006). The SZE contour interval is 100~$\mu \rm{K}_{\rm{CMB}}$.}
1024: \label{FIG:bullet-overlay}
1025: \end{figure}
1026:
1027: \subsection{Mass-Weighted Temperature}\label{subsec:mwte}
1028:
1029: The combination of cluster SZE and X-ray measurements can be used to place
1030: constraints on the thermal structure of the intracluster gas.
1031: The SZE intensity is proportional to the product of the electron density and the
1032: electron temperature along the line of sight, see equation~(\ref{EQN:integratedpressure}). Therefore, if
1033: the electron density is known from another measurement, the SZE
1034: can be used to measure a mass-weighted temperature.
1035: For simplicity, we assume here that the intracluster gas is isothermal, but a more detailed
1036: comparison of the SZE surface brightness and projected density could be used to
1037: constrain the thermal structure in the cluster.
1038:
1039: We perform this calculation with two different descriptions of the intracluster gas density.
1040: First, we model the spatial distribution of the intracluster gas as a
1041: spherical $\beta$ model following equation~(\ref{EQN:betamodeldensity}).
1042: We use $\beta$ model parameters from \markcite{ota2004}{Ota} \& {Mitsuda} (2004), derived from ROSAT HRI
1043: ($\sim 2\arcsec$ resolution) measurements of the inner $6\arcmin$ radius of the
1044: Bullet cluster: $\beta = 1.04^{+0.16}_{-0.10}$, $\theta_c =
1045: 112.5^{+15.6}_{-10.4} \arcsec$, and $n_{e0} = 7.2^{+0.3}_{-0.3}
1046: \times 10^{-3}$ cm$^{-3}$.
1047:
1048: We construct an X-ray derived SZE surface brightness model from
1049: the $\beta$ model electron surface density profile using
1050: equation~(\ref{EQN:integratedpressure}). To account for $\beta$ model
1051: uncertainties, we incorporate the values and
1052: uncertainties for $\beta$, $\theta_c$, and $n_{e0}$ as independent Gaussian
1053: priors. We then use the analysis method described in
1054: \S\,\ref{subsec:fitting} to minimize $\chi^2$ on the difference
1055: between the X-ray derived SZE model, convolved with the point source
1056: transfer function, and the APEX-SZ non-source-masked data. The free
1057: parameters in the fit are the three $\beta$ model parameters, the mass-weighted electron
1058: temperature $T_{mg}$, and the relative map alignment in RA and DEC.
1059:
1060: We find $T_{mg}=11.4 \pm 1.4\,$keV after marginalizing over the other
1061: parameters in the fit and including the SZE
1062: flux calibration uncertainty and the
1063: effect of the APEX-SZ band center frequency uncertainty on the
1064: relativistic SZE spectrum $f(x,T_{mg})$. The reduced $\chi^2$ of the
1065: best fit model is $1.008$ with an associated PTE of $31.3\%$,
1066: indicating that the spherical $\beta$ density model and the assumption
1067: of isothermality produce an acceptable fit to the data.
1068:
1069: Given the complex morphology of this merging system, the validity of
1070: the spherical $\beta$ model is questionable. We therefore repeat the
1071: determination of the mass-weighted temperature by directly comparing
1072: X-ray measurements of the projected intracluster gas density with the
1073: measured SZE signal in order to produce a less model-dependent
1074: measurement of the mass-weighted temperature.
1075: We make use of the publicly
1076: available\footnote{http://flamingos.astro.ufl.edu/1e0657/public.html}
1077: electron surface density map derived from Chandra X-ray satellite data presented
1078: in \markcite{clowe2006}{Clowe} {et~al.} (2006).
1079: Using the same analysis as above, and marginalizing over the relative
1080: map alignment parameters, we find mass-weighted electron temperature
1081: $T_{mg} = 10.8 \pm 0.9\,$keV. The fit to the data is again good, with a
1082: reduced $\chi^2= 1.037$ and a PTE of 20.6\%. This is in excellent
1083: agreement with the value $T_{mg}=11.4 \pm 1.4 \,$keV found from the above
1084: $\beta$ model analysis.
1085:
1086: Given the complex dynamics in the Bullet cluster, there have been several
1087: studies of the temperature structure
1088: \markcite{finoguenov2005, markevitch2006,andersson2007}(e.g., {Finoguenov}, {B{\"o}hringer}, \& {Zhang} 2005; {Markevitch} 2006; {Andersson}, {Peterson}, \& {Madejski} 2007).
1089: There have also been several published results for the spectroscopic
1090: temperature within annuli about the cluster center.
1091: Chandra data was used by \markcite{govoni2004}{Govoni} {et~al.} (2004) to determine a spectroscopic
1092: X-ray temperature of $T_{spec}=13.9 \pm 0.7\,$keV within $0.75\,$Mpc of the
1093: cluster center.
1094: From the analysis of XMM data within an annulus of $0.14-0.7\,$Mpc
1095: radius, \markcite{zhang2006}{Zhang} {et~al.} (2006) find a temperature of $T_{spec}=10.6\pm 0.2\,$keV.
1096: Analyzing the combination of XMM and RXTE data, \markcite{petrosian06}{Petrosian}, {Madejski}, \& {Luli} (2006), find
1097: $T_{spec}=12.1\pm0.2\,$keV within a radius of $0.95\,$Mpc.
1098: The published X-ray spectroscopic temperatures span a range much larger
1099: than the stated uncertainties in the measurements.
1100: Given the complex thermal structure for the cluster, and the presence of
1101: gas at temperatures corresponding to energies at or above the upper limits
1102: of the Chandra and XMM energy response, the variation in the measured
1103: X-ray spectroscopic temperature is not surprising.
1104: The mass-weighted temperature found with APEX-SZ falls near the lowest of the
1105: reported X-ray spectroscopic temperatures.
1106: However, we do not expect exact agreement between the mass-weighted and
1107: spectroscopic temperatures.
1108: Using Chandra data for a sample of 13 relaxed clusters, \markcite{vikhlinin2006}{Vikhlinin} {et~al.} (2006)
1109: find that, due to the presence of thermal structure in the intracluster gas, the
1110: X-ray spectroscopic temperature is typically a factor of
1111: $T_{spec}/T_{mg}=1.11 \pm 0.06$
1112: larger than the X-ray derived mass-weighted electron temperature.
1113: This is consistent with the simulation results of \markcite{nagai07}{Nagai}, {Vikhlinin}, \& {Kravtsov} (2007) who find
1114: $T_{spec}/T_{mg} \approx 1.14$ for relaxed clusters and
1115: $T_{spec}/T_{mg} \approx 1.12$, with a somewhat larger scatter,
1116: for unrelaxed systems.
1117: Naively applying this correction to the published X-ray spectroscopic
1118: temperatures, we infer results for mass-weighted temperatures that bracket the
1119: APEX-SZ result.
1120:
1121:
1122: \subsection{Gas Mass Fraction}\label{sec:gmass}
1123:
1124: Using the SZE measurements, we construct a model for the
1125: intracluster gas distribution which, when combined with X-ray
1126: measurements, can be used to determine the gas mass, total mass, and
1127: therefore gas mass fraction of the cluster. The gas mass is estimated by
1128: integrating a spheroidal model for the cluster gas, following
1129: \markcite{laroque2006}{LaRoque} {et~al.} (2006).
1130:
1131: Several assumptions about the model must be made to estimate the gas
1132: mass. We assume that the cluster gas is isothermal in order to convert
1133: pressure to density. We also assume spheroidal symmetry for the gas
1134: distribution in order to convert the two-dimensional SZE integrated
1135: pressure measurement to a three-dimensional gas distribution.
1136:
1137: We consider two simple cases, an oblate spheroid generated by rotation
1138: about the minor axis and a prolate spheroid generated by rotation
1139: about the major axis, where the symmetry axis is in the sky plane.
1140: The gas mass, under these assumptions, becomes:
1141: %
1142: \begin{eqnarray}
1143: \label{EQN:gasmass}
1144: \lefteqn{M_{\rm{gas}}(r) =
1145: 8 \mu_e n_{e0} m_p {D_{A}}^3 \int_0^{r/D_{A}} dX \, dY \, dZ } \nonumber \\
1146: & & \left(1+\left(\frac{X}{\theta_{\rm{c}}}\right)^2+ \right.
1147: \left.\left(\frac{Y}{\eta\theta_{\rm{c}}}\right)^2 +
1148: \left(\frac{Z}{\zeta\theta_{\rm{c}}}\right)^2 \right)^{-3\beta/2}
1149: \end{eqnarray}
1150: %
1151: where $\mu_e$ is the nucleon/electron ratio, taken to be
1152: 1.16~\markcite{grego2001}({Grego} {et~al.} 2001). The factor of eight is due to integrating over
1153: only one octant of the spheroid. The factor $\zeta$ is set to unity in
1154: the case of oblate spheroidal symmetry, while in the case of prolate spheroidal
1155: symmetry, $\zeta$ is set to $\eta$.
1156:
1157: The total cluster mass is estimated by assuming hydrostatic
1158: equilibrium and integrating the inferred gas distribution
1159: \markcite{grego2000}({Grego} {et~al.} 2000) to find:
1160: %
1161: \begin{equation}
1162: \label{EQN:hydroeq}
1163: \rho_{\rm{total}} = -\frac{kT_{\rm{e}}}{4\pi G \mu m_{\rm{p}}} \nabla^2 \ln \rho_{\rm{gas}}.
1164: \end{equation}
1165: %
1166: Here $\mu$ is the mean molecular weight of the intracluster gas,
1167: which is assumed to be 0.62 \markcite{zhang2006}({Zhang} {et~al.} 2006). Using
1168: equations~(\ref{EQN:gasmass}) \&~(\ref{EQN:hydroeq}), and our model parameters in
1169: Table~\ref{TBL:betamodelfits}, we calculate the gas mass, total
1170: mass, and gas mass fraction for the cluster.
1171: In Table~\ref{TBL:massestimates}, we give these results.
1172: The gas mass fraction results for a prolate gas
1173: distribution model are $\sim3\%$ larger than those for
1174: an oblate model, while the total mass and gas mass are $\lesssim18\%$ larger.
1175: We quote only the oblate spheroidal results.
1176:
1177: We calculate our results within two different radii. The first is
1178: the radius of the cluster at which its mean density is equal to 2500
1179: times the critical density at the redshift of the cluster,
1180: $r_{2500}$. The second radius is 1.42~Mpc, which is the same radius
1181: used by \markcite{zhang2006}{Zhang} {et~al.} (2006) for their gas mass fraction calculation.
1182: This will allow for a more direct comparison to their result, and is
1183: also near where our measured SZE radial profile has unity signal to
1184: noise.
1185: For all results, we assume a $\Lambda$CDM cosmology,
1186: with $h = 0.7$, $\Omega_{\rm{m}} = 0.27$, and $\Omega_{\Lambda} =
1187: 0.73$.
1188: The results of this analysis are summarized in Table~\ref{TBL:massestimates}.
1189: Under the assumption of a $10.6\,$keV mass-weighted temperature (the lowest of the published
1190: X-ray spectroscopic temperatures and near our mass-weighted temperature results
1191: in \S\,\ref{subsec:mwte}), we find gas mass fractions
1192: $f_g=0.216 \pm 0.031$ and $0.179 \pm 0.036$ within $r_{2500}$ and 1.42~Mpc, respectively.
1193: The fact that the computed gas fraction in the central region significantly
1194: exceeds the cosmic average determined by WMAP5 \markcite{dunkley09}($f_g=0.165\pm 0.009$, {Dunkley} {et~al.} 2009),
1195: and a lower value observed in relaxed clusters
1196: \markcite{vikhlinin09}($f_g \simeq 0.12$, see, e.g., {Vikhlinin} {et~al.} 2009), is likely due to deviations of the
1197: intracluster gas from isothermal hydrostatic equilibrium.
1198: On larger scales, baryon fractions produced for the range of reported
1199: X-ray temperatures
1200: bracket the published results using X-ray and weak lensing data.
1201: \markcite{bradac2006}{Brada{\v c}} {et~al.} (2006) measure a gas
1202: mass fraction $f_g=0.14 \pm 0.03$ by comparing the gas mass
1203: calculated from Chandra X-ray measurements to weak lensing total
1204: mass measurements in a $4.9\arcmin \times 3.2\arcmin$ box roughly
1205: centered around the cluster. \markcite{zhang2006}{Zhang} {et~al.} (2006), measured a gas mass
1206: fraction $f_g=0.161 \pm 0.018$ within a radius of 1.42~Mpc.
1207: Despite the limitations of applying a hydrostatic equilibrium model
1208: to this merging cluster, the APEX-SZ results for the
1209: gas mass fraction are in good agreement with previous work.
1210:
1211: \begin{deluxetable}{ccccccc}
1212: \tabletypesize{\tiny}
1213: \tablecaption{\label{TBL:massestimates} Mass Estimates for the Bullet Cluster System}
1214: \tablewidth{0pt}
1215: \tablehead{
1216: \colhead{$T_e (keV)^a$} & \colhead{Mean Overdensity} & \colhead{$r_{\rm{int}}$ (\arcmin)$^b$} & \colhead{$r_{\rm{int}}$ (Mpc)$^c$} & \colhead{Gas Mass Fraction} & \colhead{Gas Mass ($10^{14} M_{\sun}$)} & \colhead{Total Mass ($10^{14} M_{\sun}$)}
1217: }
1218: \startdata
1219: $13.9 \pm 0.7$ & $2506 \pm 233$ & 2.77 & 0.739 & $0.124 \pm 0.022$ & $0.944 \pm 0.105$ & $7.56 \pm 0.70$ \\
1220: $13.9 \pm 0.7$ & $961 \pm 98$ & 5.32 & 1.42 & $0.106 \pm 0.024$ & $2.20 \pm 0.33$ & $20.6 \pm 2.1 $ \\
1221: $10.6 \pm 0.2$ & $2521 \pm 230$ & 2.15 & 0.572 & $0.216 \pm 0.031$ & $0.765 \pm 0.072$ & $3.54 \pm 0.32$ \\
1222: $10.6 \pm 0.2$ & $734 \pm 66$ & 5.32 & 1.42 & $0.179 \pm 0.036$ & $2.83 \pm 0.40$ & $15.7 \pm 1.4 $ \\
1223: \enddata
1224: \tablecomments{Two different isothermal electron temperatures are assumed, in order to bracket the range of X-ray spectroscopic temperatures
1225: reported in the literature. The top two rows assume an isothermal electron temperature of $13.9 \pm 0.7$ keV. The bottom two rows assume an isothermal electron temperature of $10.6 \pm 0.2$ keV. For each electron temperature, we integrate to $r_{2500}$, the radius within which the mean cluster density is 2500
1226: times greater than the critical density at the redshift of the cluster.
1227: For each electron temperature, we also integrate to a fixed radius of 1.42~Mpc, allowing a direct comparison
1228: to results in \markcite{zhang2006}{Zhang} {et~al.} (2006). This radius is also near where our
1229: measured SZE radial profile has unity signal to
1230: noise. For all results, we assume a $\Lambda$CDM cosmology,
1231: with $h = 0.7$, $\Omega_{\rm{m}} = 0.27$, and $\Omega_{\Lambda} =
1232: 0.73$. Uncertainties in the gas mass and gas mass fraction include
1233: a $\pm 5.5\%$ SZE flux calibration uncertainty.}
1234: \tablenotetext{a}{Isothermal electron temperature}
1235: \tablenotetext{b}{Angular integration radius.}
1236: \tablenotetext{c}{Physical integration radius.}
1237: \end{deluxetable}
1238:
1239: %--------------------------------------------------------------
1240:
1241: \section{Conclusions}
1242: \label{SEC:conclusions} Measurements of the SZE provide a robust and
1243: independent probe of the intracluster gas properties in galaxy
1244: clusters. The APEX-SZ $150\,$GHz observations detect the Bullet system
1245: with $23\,\sigma$ significance within the central 1\arcmin\ radius of the
1246: SZE centroid position.
1247: We do not expect to see a resolved signal from the Bullet
1248: sub-cluster in the $150\,$GHz 85\arcsec\ FWHM resolution SZE maps,
1249: and no obvious feature, such as a secondary peak, is present.
1250: We expect no significant contamination of the observed SZE decrement
1251: due to radio sources, and there is no evidence for
1252: significant contamination by a known bright lensed dusty galaxy.
1253:
1254: We process an elliptical $\beta$ model through the observation
1255: transfer function and fit it to the measured temperature decrement map.
1256: We also measure the cluster mass-weighted
1257: electron temperature and gas mass fraction with the SZE data.
1258: Combining the APEX-SZ map with a map of projected electron surface density from
1259: Chandra X-ray observations, we determine the mass-weighted temperature of the
1260: cluster gas to be $T_{mg}=10.8 \pm 0.9\,$keV.
1261: This value is consistent with the lowest X-ray spectroscopic temperatures reported
1262: for this cluster and should be less sensitive to the details of the cluster
1263: thermal structure.
1264: The derived baryon fraction is also found to be in reasonable agreement with
1265: previous X-ray and weak lensing determinations.
1266:
1267: Throughout this work, we make an assumption of isothermal cluster gas.
1268: Clearly, incorporating thermal structure, measured by X-ray
1269: observations, in the analysis of the SZE data would improve the
1270: determination of the gas distribution and gas mass fraction.
1271: Ultimately, a more sophisticated analysis could be implemented that
1272: combines X-ray, SZE, and weak lensing data and relaxes assumptions of
1273: hydrostatic equilibrium between the gas and dark matter components of
1274: the cluster. This is particularly important for a detailed
1275: understanding of actively merging systems such as the Bullet cluster.
1276:
1277: \acknowledgments
1278:
1279: We thank the staff at the APEX telescope site, led by David
1280: Rabanus and previously by Lars-\AA ke Nyman, for their dedicated and exceptional support. We also thank
1281: LBNL engineers John Joseph and Chinh Vu for their work on the
1282: readout electronics. APEX-SZ is funded by the National Science
1283: Foundation under Grant Nos.\ AST-0138348 \& AST-0709497.
1284: Work at LBNL is supported
1285: by the Director, Office of Science, Office of High Energy and
1286: Nuclear Physics, of the U.S. Department of Energy under Contract No.
1287: DE-AC02-05CH11231. Work at McGill is supported by the Natural Sciences
1288: and Engineering Research Council of Canada and the Canadian Institute
1289: for Advanced Research. RK acknowledges partial financial support from
1290: MPG Berkeley-Munich fund. NWH acknowledges support from an Alfred P. Sloan
1291: Research Fellowship.
1292:
1293: %%\bibliographystyle{abbrvnat}
1294: %\bibliographystyle{apj}
1295: \bibliography{}
1296:
1297: \end{document}
1298:
1299: