1: \documentclass[12pt,preprint]{aastex}
2:
3: %\usepackage{natbib}
4:
5: \shorttitle{Acrminute CMB Anisotropies with Bolocam}
6: \shortauthors{Sayers et al.}
7:
8: \begin{document}
9:
10: \title{A Search for Cosmic Microwave Background Anisotropies
11: on Arcminute Scales with Bolocam}
12:
13: \author{J.~Sayers\altaffilmark{1,6}, S.~R.~Golwala\altaffilmark{1},
14: P.~Rossinot\altaffilmark{1}, P.~A.~R.~Ade\altaffilmark{2},
15: J.~E.~Aguirre\altaffilmark{3,4}, J.~J.~Bock\altaffilmark{5},
16: S.~F.~Edgington\altaffilmark{1}, J.~Glenn\altaffilmark{4},
17: A.~Goldin\altaffilmark{5}, D.~Haig\altaffilmark{2},
18: A.~E.~Lange\altaffilmark{1}, G.~T.~Laurent\altaffilmark{4},
19: P.~D.~Mauskopf\altaffilmark{2}, and~H.~T.~Nguyen\altaffilmark{5}}
20: \altaffiltext{1}
21: {Division of Physics, Mathematics, \& Astronomy,
22: California Institute of Technology,
23: Mail Code 59-33, Pasadena, CA 91125}
24: \altaffiltext{2}
25: {Physics and Astronomy, Cardiff University, 5 The Parade,
26: P. O. Box 913, Cardiff CF24 3YB, Wales, UK}
27: \altaffiltext{3}
28: {Jansky Fellow, National Radio Astronomy Observatory}
29: \altaffiltext{4}
30: {Center for Astrophysics and Space Astronomy \& Department of
31: Astrophysical and Planetary Sciences,
32: University of Colorado, 389 UCB, Boulder, CO 80309}
33: \altaffiltext{5}
34: {Jet Propulsion Laboratory, California Institute of Technology,
35: 4800 Oak Grove Drive, Pasadena, CA 91109}
36: \altaffiltext{6}
37: {jack@caltech.edu}
38:
39: \begin{abstract}
40: We have surveyed two science fields totaling one square degree with
41: Bolocam at 2.1 mm to search for secondary CMB anisotropies caused by
42: the Sunyaev-Zel'dovich effect (SZE). The fields are in the Lynx and
43: Subaru/XMM SDS1 fields. Our survey is sensitive to angular scales
44: with an effective angular multipole of $\ell_{eff} = 5700$ with
45: FWHM$_{\ell} = 2800$ and has an angular resolution of 60 arcseconds
46: FWHM. Our data provide no evidence for anisotropy. We are able to
47: constrain the level of total astronomical anisotropy, modeled as a
48: flat band power in $\mathcal{C}_\ell$, with frequentist 68\%, 90\%, and
49: 95\% CL upper limits of 590, 760, and 830 $\mu K_{CMB}^2$. We
50: statistically subtract the known contribution from primary CMB
51: anisotropy, including cosmic variance, to obtain constraints on the
52: SZE anisotropy contribution. Now including flux calibration
53: uncertainty, our frequentist 68\%, 90\% and 95\% CL upper limits on a
54: flat band power in $\mathcal{C}_\ell$ are 690, 960, and 1000 $\mu
55: K_{CMB}^2$. When we instead employ the analytic spectrum suggested by
56: \citet{komatsu02}, and account for the non-Gaussianity of the
57: SZE anisotropy signal, we obtain upper limits on the average amplitude of
58: their spectrum weighted by our transfer function of 790, 1060, and 1080
59: $\mu K_{CMB}^2$. We obtain a 90\% CL upper limit on $\sigma_8$, which
60: normalizes the power spectrum of density fluctuations, of 1.57. These
61: are the first constraints on anisotropy and $\sigma_8$ from survey
62: data at these angular scales at frequencies near 150~GHz.
63:
64: % The observations were flux-calibrated and pointing-corrected using
65: % beam maps of Uranus and Neptune and other secondary calibrators taken
66: % frequently during observing. Internal uncertainty on the pointing and
67: % flux calibration contributes negligible uncertainty to the final
68: % result; calibration uncertainty in the final result is dominated by
69: % uncertainty in models for the absolute brightness temperatures of
70: % Mars, Uranus, and Neptune. We developed several algorithms to
71: % subtract atmospheric noise from our data and a fast pseudo
72: % least-squares map-maker. We used simulations to calibrate the
73: % transfer function of our data-taking and analysis pipeline and
74: % map-maker, and we determined the expected noise properties of our
75: % final maps using jackknife realizations of the data.
76: \end{abstract}
77: \keywords{cosmology: observation --- cosmic microwave background ---
78: methods: data analysis --- large-scale structure of the universe ---
79: cosmological parameters}
80:
81: \section{Introduction}
82:
83: \subsection{Background}
84:
85: % One observational approach to learn more
86: % about the structure of the universe involves the SZE.\footnote{
87: % Throughout this paper SZE refers to the thermal SZE.}
88: % SZE observations have been used to determine the
89: % value of the Hubble parameter without relying on
90: % the standard distance ladder
91: % approach \citep{bonamente06,udomprasert04,reese02},
92: % and can be used to constrain the values of
93: % $\sigma_8$, $\Omega_m$, $\Omega_{\Lambda}$,
94: % and $w$~\citep{holder00, haiman00, komatsu02}.
95: % Additionally, the SZE is a powerful tool for
96: % understanding the largest bound objects in the
97: % universe, clusters of galaxies, at any redshift.
98:
99: The SZE\footnote{Throughout this paper SZE refers
100: to the thermal SZE.}
101: is the inverse Compton scattering of CMB photons
102: with a distribution of hot electrons, causing a net increase in
103: the energy of the photons \citep{sunyaev72}.
104: Since the background CMB is redshifted along with the
105: SZE-induced distortion, the relative amplitude of the
106: distortion, $\Delta T_{CMB} / T_{CMB}$, is
107: independent of redshift.
108: The distortion caused by the SZE is proportional
109: to the Comptonization parameter $y$, which is
110: a measure of the integral of the electron thermal energy density
111: along the line of sight and is given by
112: \begin{displaymath}
113: y =
114: \frac{\sigma_T}{m_e c^2} \int dl \mbox{ } n_e k_B T_e,
115: % \label{eqn:SZ_y}
116: \end{displaymath}
117: where $\sigma_T$ is the Thomson cross section,
118: $m_e$ is the electron mass, $c$ is the speed of light,
119: $k_B$ is Boltzmann's constant, $T_e$ is the temperature of
120: the electrons, and
121: $n_e$ is the number density of electrons.
122: Since the scattering process conserves photon number, the
123: thermal spectrum of the CMB is distorted by the SZE;
124: there is a negative temperature shift at low
125: frequency and a positive temperature shift at high
126: frequency.
127: The cross-over point where there is no distortion
128: of the CMB occurs at approximately 218~GHz.
129: The temperature shift caused by the SZE, $\Delta T_{CMB}$, is
130: \begin{displaymath}
131: \frac{\Delta T_{CMB}}{T_{CMB}} = f(x) y,
132: % \label{eqn:SZ_temp}
133: \end{displaymath}
134: where
135: \begin{displaymath}
136: f(x) = x \frac{e^x+1}{e^x-1} - 4,
137: % \label{eqn:SZ_spec}
138: \end{displaymath}
139: and $x = h \nu / k_B T_{CMB}$, $h$ is Planck's constant,
140: $\nu$ is the frequency, and $T_{CMB} = 2.73$~K is the
141: temperature of the CMB.
142: For reference, excellent reviews of the SZE and its relevance to
143: cosmology are given by \citet{birkinshaw99} and
144: \citet{carlstrom02}.
145:
146: \subsection{Untargeted SZE Surveys}
147:
148: % Since the SZE signal is roughly proportional to the
149: % total mass of a cluster, and the surface brightness
150: % of the SZE signal is independent of the cluster redshift,
151: % the SZE offers an ideal tool to conduct a mass-limited
152: % survey to high redshift\footnote{
153: % In practice, the SZE surface brightness depends
154: % strongly on the cluster core radius and density,
155: % and is not a good characterization of the
156: % total cluster mass.
157: % The total integrated flux of the cluster does provide
158: % a good measurement of the mass of the cluster,
159: % but it is not independent of redshift due to
160: % the factor of $1/D_A^2$.
161: % However, this angular diameter distance factor
162: % is largely canceled out by the evolution of the
163: % cluster virial temperature with redshift,
164: % since clusters that form earlier in the universe
165: % will be denser and hotter.
166: % The net result is a mass selection function that
167: % varies by less than a factor of two for redshifts
168: % between $z \simeq 0.1$ and $z \simeq 3.0$ \citep{holder00}.}.
169: % Such a survey is an excellent way to
170: % constrain $\Omega_m$
171: % because the formation history of large scale structure is
172: % sensitive to
173: % the density of matter in the universe.
174: % SZE number counts can also be used to
175: % constrain the values of $\Omega_{\Lambda}$, $w$,
176: % and $\sigma_8$ \citep{holder00, haiman00}.
177:
178: To date, there have been no detections of previously unknown
179: clusters using the SZE.
180: However,
181: unresolved objects in SZE surveys will produce anisotropies
182: in the CMB that are expected to dominate the CMB power
183: spectrum at small angular scales corresponding
184: to angular multipoles above $\ell \simeq 2500$.
185: The overall normalization of these SZE-induced CMB anisotropies
186: is extremely sensitive to $\sigma_8$ and can
187: be used to constrain the value of this cosmological
188: parameter \citep{komatsu02}.
189: Several experiments have conducted SZE surveys
190: that have produced tentative detections of the SZE-induced
191: anisotropies in the CMB.
192: At 30~GHz CBI has measured an excess CMB power between
193: $\ell = 2000$ and $\ell = 3500$ at a significance of
194: 3.1$\sigma$ \citep{mason03}.
195: Also at 30~GHz, BIMA/OVRO has measured
196: a CMB anisotropy of $220^{+140}_{-120}$~$\mu$K$^2_{CMB}$
197: at an angular multipole of $\ell = 5237$ \citep{dawson06}.
198: ACBAR, at 150~GHz and $2000 < \ell < 3000$, has measured an
199: excess power of $34 \pm 20$~$\mu$K$_{CMB}^2$ \citep{reichardt08}.
200: A joint analysis of the CBI and ACBAR excesses
201: shows that they are six times more likely to be caused
202: by the SZE than primordial fluctuations \citep{reichardt08}.
203:
204:
205: Additionally, these tentative anisotropy detections have
206: been used to constrain cosmological parameters;
207: the CBI data are consistent with
208: $\sigma_8 \simeq 1$ \citep{bond05},
209: and the BIMA/OVRO data measure
210: $\sigma_8 = 1.03^{+0.20}_{-0.29}$ \citep{dawson06}.
211: \citet{reichardt08} combine various data sets to place
212: constraints on $\sigma_8$ via an excess contribution
213: to anisotropy at high $\ell$, $\ell > 1950$.
214: % ACBAR plus WMAP3 data only, with the amplitude
215: % of the SZE contribution at high $\ell$ not slaved
216: % to its contribution at low $\ell$,
217: % indicate $\sigma_8^{SZ} = 0.96^{+0.08}_{-0.12}$.
218: % However, they caution that the constraint is highly
219: % sensitive to the prior distribution used.
220: % When
221: % ACBAR and WMAP3,
222: % plus CBI and BIMA/OVRO data at high $\ell$ and lower
223: % frequency are included, the unslaved $\sigma_8^{SZ}$
224: % constraint is narrowed to $0.95^{+0.03}_{-0.04}$.
225: ACBAR and WMAP3,
226: plus CBI and BIMA/OVRO data at high $\ell$ and lower
227: frequency, combine to indicate
228: $\sigma_8^{SZ} = 0.95^{+0.03}_{-0.04}$
229: when the amplitude of the SZE contribution is not
230: slaved to its contribution at low $\ell$.
231: Since this result is inconsistent with other
232: constraints on $\sigma_8$, including those from the
233: lower $\ell$ portions of the CMB spectrum,
234: \citet{reichardt08} also consider a case in which
235: the high-$\ell$ SZE contribution is slaved to to
236: the contribution at lower $\ell$ via
237: $\sigma_8^{SZ} = \sigma_8$.
238: In this case the excess power
239: is produced by point sources.
240: Fitted to the CMBall data set, which excludes
241: CBI high-$\ell$ and BIMA/OVRO data,
242: this model results in $\sigma_8$ values consistent
243: with other measurements,
244: $\simeq 0.80-0.81 \pm 0.03-0.04$ depending on the
245: assumptions and data sets included.
246: % In this case the excess power is dominated by
247: % point sources with a contribution of
248: % $35-37^{+13-12}_{-35-37}$, again depending on
249: % assumptions.
250: No attempt is made to explain the CBI and BIMA/OVRO
251: excesses.
252: Overall, the current results suggest two possibilities:
253: there are point source contributions to all the
254: high-$\ell$ data (ACBAR, CBI, BIMA/OVRO) that have
255: not been properly included;
256: or the SZ contribution calculated from theory is
257: underestimated.
258:
259: The survey presented here is the first such survey
260: at 150~GHz and at $\ell \simeq 6000$.
261: As we shall explain, contributions from primary
262: CMB anisotropies, SZE, radio, and submillimeter point
263: sources are all expected to be comparable,
264: each at a level of $\mathcal{C}_{\ell} \simeq
265: 50$~$\mu$K$^2_{CMB}$.
266:
267: % The situation will likely be resolved more satisfactorily
268: % by the upcoming APEX-SZ, ACT, and SPT experiments,
269: % which
270: % should detect hundreds or thousands of previously
271: % unknown clusters and
272: % measure the SZE-induced CMB anisotropies to high
273: % precision \citep{dobbs06, kosowsky03, ruhl04}.
274: % The multiple spectral bands available to ACT and SPT
275: % will significantly help to resolve the amplitude
276: % and identity of the high-$\ell$ excess;
277: % contributions from SZE, as well as submillimeter and
278: % radio point sources, may be present.
279:
280: \section{Observations}
281:
282: \subsection{Instrument Description}
283:
284: Bolocam is a large format, 144 detector, millimeter-wave camera designed
285: to be operated at the Caltech Submillimeter Observatory (CSO).
286: For these observations the array was comprised of 115 optical
287: and 6 dark detectors.
288: Each detector is housed within its own integrating cavity,
289: formed by a frontshort plate and a backshort plate \citep{glenn02}.
290: Smooth-walled conical feedhorns separated by 0.7 (f/\#)$\lambda$,
291: a cold (4~K)
292: high-density polyethylene (HDPE) lens,
293: and a room-temperature ellipsoidal mirror are used to
294: couple the detectors to the CSO optics.
295: Each feedhorn terminates into a cylindrical waveguide,
296: which defines the low-frequency cutoff of the system;
297: the final filter in
298: a series of six cold metal-mesh filters determines the
299: high-frequency cutoff.
300: The resulting passband is centered at 143~GHz, and has
301: an effective width of 21~GHz.
302: A cold (4~K) Lyot stop is used to define
303: the illumination of the 10.4~m primary mirror,
304: and the resulting far-field beams
305: have FWHMs of 60~arcseconds.
306: Bolocam can also observe at 270~GHz, and has been
307: used in this mode for several types of observations,
308: including surveys for submillimeter
309: galaxies and protostellar cores \citep{laurent05, enoch06, young06}.
310:
311: The detector array has a hexagonal geometry, and utilizes
312: silicon nitride micromesh (spider-web) bolometers
313: \citep{mauskopf97} which
314: are cooled to 260~mK using a three-stage $^4$He/$^3$He/$^3$He
315: sorption refrigerator \citep{bhatia00, bhatia02}.
316: JFETs located near the array and operated at 140~K are used
317: to buffer the high-impedance bolometer signals from sources
318: of current noise.
319: In order to avoid the $1/f$ noise from the JFETs, the bolometers
320: are biased at 130~Hz and read out using room-temperature
321: lockin amplifiers.
322: More details of the Bolocam instrument can be found
323: in \citet{golwala08}, \citet{glenn98}, \citet{glenn03},
324: and \citet{haig04}.
325:
326: \subsection{Observing Strategy}
327:
328: The data described in this paper was collected during a forty
329: night observing run in late 2003.
330: % \footnote{
331: % A similar size data set was also collected in 2004,
332: % although a problem with the electronics made this data
333: % extremely noisy and therefore useless.
334: % More details are given in \citet{sayers07}.}.
335: During the first half of each night we observed a
336: 0.5~deg$^2$ region centered at 02h18m00s, \mbox{-5d00m00s} (J2000),
337: which coincides with the Subaru/XMM Deep Survey
338: (SXDS or SDS1); and
339: during the second half of each night we observed
340: a 0.5~deg$^2$ region centered on the Lynx field
341: at 08h49m12s, +44d50m24s (J2000).
342: These fields were selected because they have extremely
343: low dust emission and a large amount of optical/X-ray
344: data that could be used to follow up any
345: SZE cluster candidates found in the maps.
346:
347: Two nights at the start of the run were used to analyze
348: different scan strategies for mapping the science fields.
349: % which could be implemented to map the
350: % science fields.
351: The maps were made by repeatedly raster scanning across the field,
352: stepping perpendicular to the scan, then
353: rastering across the field in the opposite direction
354: until the entire field has been covered.
355: Our studies showed that the time-stream noise
356: is independent of the angle of the raster scan and the
357: turnaround time between scans, so we chose to scan
358: parallel to RA or dec and turnaround as quickly
359: as the telescope would allow ($\simeq 10$~seconds).
360: Additionally, we found that our sensitivity to
361: astronomical signals is maximized when we raster scan
362: at a speed of 240~arcseconds/second.\footnote{
363: Faster speeds were not attempted due to fears
364: that the CSO would not function properly and/or
365: would be damaged.}
366: %JS 2008/07/31 addressing referee comment 2
367: At this speed it takes approximately 12.5 seconds to
368: complete one scan across the field, which
369: means we were on-source approximately 56\% of the time
370: during an observation.
371: Although scanning at this relatively quick speed
372: reduces our time on-source
373: because a larger fraction of time is spent on turnarounds
374: between scans, it also puts a larger amount of
375: our signal band above the $1/f$ atmospheric noise.
376: Given the scan speed and turnaround time mentioned above,
377: along with our step size of 162~arcseconds
378: ($\simeq 1/3$ of the field of view)
379: % based on the array geometry,
380: a complete map of the field was made in
381: approximately eight minutes.
382: % Note that the array angle relative to the horizon was fixed
383: % at 80 degrees.
384: % It was not possible to rotate between observations, and this
385: % angle was chosen
386: % because it provides the most uniform coverage for alt/az scans.
387:
388: \section{Data Reduction}
389:
390: \subsection{Initial Processing}
391: \label{sec:merge}
392:
393: % Several data time-streams are recorded by the Bolocam
394: % data acquisition system (DAS)
395: % at 50~Hz, including the bolometer signals and
396: % a logic signal that transitions at the start and
397: % end of each scan.
398: % Additionally, several data streams are recorded by the
399: % telescope computer at 100~Hz, including pointing information
400: % and the same logic signal that transitions at the start
401: % and end of each scan.
402: % After the telescope data are down-sampled to 50~Hz, we use
403: % the logic signal recorded by both computers to align the
404: % data streams and merge them
405: % into a single netCDF file\footnote{
406: % Network Common Data Form (netCDF) is a set of machine-independent
407: % software libraries that can be used to store and access
408: % array-oriented scientific data.
409: % See the netCDF website at
410: % http://www.unidata.ucar.edu/software/netcdf/.}.
411: After merging the bolometer time-streams recorded by
412: the data acquisition system with the pointing information
413: recorded by the telescope,
414: we parse the data into files that
415: contain a single observation.
416: Each single observation contains a set of scans that
417: completely map the astronomical field or object, and they
418: are typically around ten minutes in length.
419: Parsing the data by observation is useful because individual
420: observations are statistically independent, have a small
421: enough number of data samples to be easily manageable
422: from an analysis standpoint, and provide a convenient
423: division of the data for the sake of bookkeeping.
424:
425: % \subsection{Filtering and Down-Sampling}
426:
427: Once the initial merging and parsing of the data is complete,
428: we begin the process of refining the data.
429: The first step in this process is to remove the effects
430: of the lockin amplifier electronics filters and to
431: down-sample the data from 50~Hz to 10~Hz.
432: We down-sample the data because essentially no
433: astronomical signal is lost, while a large amount
434: of 60-Hz pickup noise is removed.
435: See Figure~\ref{fig:pickup_60hz} for an illustration
436: of the noise spectrum and the shape of the
437: beam in frequency space.
438:
439: % To perform the above steps, we first
440: % Fourier transform the time-stream from
441: % the entire observation for each bolometer.
442: % Prior to Fourier transforming, we remove some
443: % data samples before the first scan and/or
444: % after the final scan to speed up the transform
445: % by making the number of time-stream
446: % samples divisible by $2^{N>8}$.
447: % In Fourier space, we remove the effects of the filtering applied by
448: % the lockin electronics by dividing the transformed data
449: % by the effective lockin filter.
450: % Next, we multiply the transformed bolometer time-stream
451: % by an anti-aliasing filter.
452: % Finally, we transform the data back to time space
453: % and down-sample by a factor of five.
454: % The anti-aliasing filter is given by
455: % \begin{equation}
456: % \mathcal{F} = \frac{2}{1 + 10^{(f/f_N)^3}},
457: % \label{eqn:aa_filt}
458: % \end{equation}
459: % where $f$ is the frequency in Hz and $f_N$ is the Nyquist
460: % frequency of the data after down-sampling, equal to 5~Hz.
461: % An ideal anti-aliasing filter would attenuate all of the signal
462: % above $f_N$, but still be slowly varying so that it
463: % does not produce ringing in the time domain.
464: % In practice, there is always a trade off between high-frequency
465: % attenuation and ringing, and the filter in
466: % Equation~\ref{eqn:aa_filt} proved to be the best
467: % combination of good attenuation and minimal ringing
468: % for our bolometer time-streams.
469: % After the data are down-sampled, they are Fourier transformed
470: % again and divided by the filter in Equation~\ref{eqn:aa_filt}
471: % to remove any artifacts caused by the filter.
472: % Finally, the data are transformed back to time space,
473: % and we are left with bolometer time-streams that are sampled
474: % at 10~Hz and are free from any filtering effects.
475:
476: \subsection{Noise Removal}
477:
478: % The next step in the data processing involves removing several
479: % types of correlated noise from the data.
480: There are several forms of correlated noise present in the
481: raw bolometer data which contaminate the astronomical signal
482: and therefore must be modeled and removed.
483: First, the emission from the atmosphere changes as a
484: function of telescope elevation angle due to the
485: changing path length through the atmosphere.
486: The path length through the atmosphere relative to the
487: zenith path length is called the airmass, $A$, and is described by
488: \begin{displaymath}
489: A = 1 / \sin(\epsilon),
490: \end{displaymath}
491: where $\epsilon$ is the elevation angle.
492: For a typical observation the range of elevation angles is
493: approximately one degree, which corresponds to a
494: change in airmass between 0.005 and 0.060 for elevation
495: angles between 75 and 30 degrees.
496: For reference, a change of 0.060 in airmass corresponds to
497: a change of approximately 0.5~K of optical loading
498: from the atmosphere, or a change in surface brightness
499: of a little less than 1~K$_{CMB}$.
500: To remove this elevation-dependent signal, we calculate
501: a linear fit of bolometer signal versus airmass.
502: We build up a
503: fit using each 12.5-second-long scan within the observation,
504: after removing the mean signal level and airmass for the scan.
505: This process yields one set of linear fit coefficients for
506: each bolometer for the entire observation,
507: which is used to create a template that is removed from
508: the bolometer time-streams.
509:
510: Next, we create a template from the
511: bias voltage monitors to account for the small amount of noise from the
512: bias electronics.
513: Note that the bias applied to the bolometers is monitored through amplifier
514: electronics identical in design to those used to monitor the
515: bolometer signals.
516: This template is then correlated and removed
517: from each of the the bolometer time-streams.
518: A template is also created from the dark bolometer signals
519: and removed from the bolometer time-streams.
520: Note that both the bias template and dark bolometer template
521: have an RMS of $\lesssim 1$~mK$_{CMB}$.
522: % It is not clear why the regular bolometer time-streams are correlated
523: % with the dark bolometer time-streams, but thermal fluctuations
524: % in the array substrate temperature or refrigerator
525: % temperature may be the cause because there is no
526: % temperature regulation of the focal plane.
527:
528: Finally, and most importantly, we remove a template describing
529: the fluctuations in emission from the atmosphere
530: (\emph{i.e.}, the atmospheric noise).
531: Since atmospheric noise is the dominant signal in our data,
532: and the beams from the individual detectors overlap to
533: a high degree while passing through the atmosphere,
534: a template for the atmospheric signal is created
535: by averaging the signals from all the bolometers.\footnote{
536: To quantify the degree of overlap,
537: note that the far-field distance for Bolocam at the CSO is
538: approximately 30~km; at a height of 20~km,
539: well above most of the water vapor in the atmosphere,
540: the ray bundles from nearby detectors have only
541: separated by one half-width.}
542: Three different algorithms are used to construct this
543: template, one for which the atmospheric signal is assumed to
544: be constant over the array, one for which the atmospheric
545: signal is allowed to vary linearly with bolometer
546: location on the array, and one for which the atmospheric
547: signal is allowed to vary quadratically with
548: bolometer location on the array.
549:
550: For the most basic case of an average template, the
551: algorithm proceeds as follows.
552: Initially, a template is constructed according to
553: \begin{equation}
554: T_n = \frac{\sum_{i=1}^{i=N_{b}}
555: c_i d_{in}}{\sum_{i=1}^{i=N_{b}} c_i}
556: \label{eqn:avg_template}
557: \end{equation}
558: where $n$ is the sample number, $N_{b}$ is the number
559: of bolometers, $c_i$ is the relative responsivity
560: of bolometer $i$, $d_{in}$ is the signal recorded by
561: bolometer $i$ at time sample number $n$, and $T_n$ is the
562: template.
563: The template generally has
564: an RMS between 10 and 100~mK$_{CMB}$, depending on
565: the observing conditions.
566: A separate template is computed for each 12.5-second-long scan.
567: After the template is computed, it is correlated with
568: the signal from each bolometer to determine the
569: correlation coefficient, with
570: \begin{equation}
571: \tilde{c_i} = \frac{\sum_{j=1}^{j=N_{s}}
572: T_n d_{in}}{\sum_{j=1}^{j=N_{s}} T_n^2}.
573: \label{eqn:skysub_corr}
574: \end{equation}
575: $\tilde{c_i}$ is the correlation coefficient
576: of bolometer $i$ and $N_s$ is the number of samples
577: in the 12.5-second-long scan.\footnote{
578: The best fit correlation coefficients change
579: from one scan to the next, typically by a
580: couple percent.}
581: % It is not clear what causes these fluctuations
582: % in the correlation coefficients, but there is
583: % a noticeable improvement in the amount of noise
584: % removed from the data when the coefficients
585: % are floated for each scan, rather
586: % than fixing them for all twenty scans
587: % in the observation.}
588: Next, the $c_i$ in Equation~\ref{eqn:avg_template} are
589: set equal to the values of $\tilde{c_i}$ found from
590: Equation~\ref{eqn:skysub_corr}, and a new template is
591: computed.
592: The process is repeated until the values of $c_i$
593: stabilize.
594: We generally iterate until the average fractional change
595: in the $c_i$s is less than $1 \times 10^{-8}$, which
596: usually takes five to ten iterations.
597: If the $c_i$s fail to converge after 100 iterations,
598: then the scan is discarded from the data.
599: For the more advanced planar and quadratic algorithms,
600: the process proceeds in the same way except linear
601: and quadratic variations with bolometer position are
602: allowed when the template is constructed.
603: These algorithms, along with adaptive PCA
604: and time-lagged average template subtraction,
605: are described and compared in more
606: detail in \citet{sayers08}.
607:
608: Each of the three different atmospheric noise
609: removal algorithms,
610: average, planar, and quadratic template
611: removal, was applied to each observation.
612: Therefore, three different atmospheric-noise-cleaned
613: time-streams are generated for
614: each observation.
615: A figure of merit is calculated for each of the three
616: files for each observation, based on the noise level of the
617: data and the expected astronomical signal shape.
618: Details of the calculation of this figure of merit
619: are given in Section~\ref{sec:opt_skysub}.
620: For each observation, the file with the best figure of
621: merit value will be the one used to create the final map
622: of the data.
623: Weather is the main criteria that determines which algorithm will
624: be selected as optimal for a given observation;
625: more aggressive algorithms (planar or quadratic) are selected
626: in poor weather conditions and more benign algorithms (average or planar)
627: are selected in good weather conditions.
628: However, there is some dependence on the profile of the source,
629: and observations of compact objects tend to be
630: optimally processed using more aggressive algorithms
631: than observations of extended objects.
632:
633: \section{Calibration}
634:
635: \subsection{Pointing reconstruction}
636:
637: Pointing reconstruction consists of determining the
638: location of each detector's beam on the sky at
639: each instant in time.
640: We compute this location in two steps:
641: 1) we calculate the location of each bolometer relative
642: to the center of the array and
643: 2) we then determine the absolute coordinates of the
644: center of the array.
645:
646: To determine the relative locations of the bolometers, we
647: observed Uranus or Neptune for approximately fifteen minutes
648: every other night.
649: These planets are bright enough to appear at high
650: signal-to-noise in a map made from a single bolometer,
651: so they can be used to determine the position of each detector
652: relative to the array center.
653: Since Bolocam was held at a fixed angle in the alt/az coordinate system
654: for the entire observing run, each bolometer
655: views the optics in the same way for the entire run
656: and the coordinates on the sky in alt/az units remain fixed.
657: Therefore, we combined the data from all the planet observations
658: to determine
659: the average position of each beam on the sky.
660: See Figure~\ref{fig:beam_locations}.
661: The uncertainties on these average positions were
662: $\sim 1$~arcsecond,
663: which is negligible when compared to
664: the 60~arcsecond FWHM
665: of a Bolocam beam.
666: We found no evidence for a systematic difference in the
667: beam positions derived from any single observation to
668: the average beam position found from all the observations.
669: This indicates that the optical system was very stable
670: over the entire observing run, including a wide range of
671: telescope elevation angles.
672:
673: To determine the absolute location of the center of the
674: array, we observed a bright
675: quasar with a known position
676: near the science field for approximately ten minutes once every
677: two hours.
678: Three different quasars were used for the SDS1 field
679: (0106+013, 0113-118, and 0336-019), and two different
680: quasars were used for the Lynx field
681: (0804+499 and 0923+392).
682: Each source was observed for five minutes while scanning parallel to
683: RA, then for five minutes scanning parallel to dec (analogous to how the
684: science fields were observed).
685: We found no systematic offset based on scan direction;
686: the maps made while scanning parallel to RA produce the same
687: source location as the maps made while scanning parallel to dec.
688: The difference in the centroid location for these consecutive
689: observations was then used to determine the
690: measurement uncertainty for the centroided location of each source.
691: As expected, the uncertainty in the centroided location
692: of the five sources is inversely
693: proportional to the flux of the source.
694: Additionally, we found no evidence that the measurement
695: uncertainty degrades or improves as a function of time
696: during the night for our typical observing times between
697: 20:00 and 07:00 local time.
698:
699: The pointing data were broken up into three distinct subsets
700: corresponding to the azimuthal position of the telescope:
701: SDS1 was observed between azimuth angles of 90 and 270 (in the south),
702: while Lynx was observed between azimuth angles
703: of -90 and 90 (in the north approaching from the east), and
704: also between azimuth angles of 270 and 360
705: (in the north approaching from the west).
706: Most of the Lynx data were taken between an azimuth angle
707: of -90 to 90,
708: so the third subset of data is considerably smaller than the first
709: two (about 1/5 the size).
710: Note that the slewing limits of the telescope are roughly
711: equal to azimuth angles of -90 to 360.
712: There is a correlation between the elevation angle of the
713: telescope and the pointing offset for each of these subsets.
714: We attempted to model
715: this correlation with several low-order polynomials,
716: % but a quadratic fit of elevation versus
717: % pointing offset is used because
718: % the residual scatter of the data do not improve significantly
719: % when a higher-order fit is used.
720: but we found that a quadratic fit of pointing offset
721: versus elevation was sufficient since higher-order fits
722: did not significantly reduce the scatter of the data.
723: We found no correlation between the telescope azimuth angle
724: and the residual offset, other than the slight difference
725: between the pointing models determined for the three subsets.
726: Therefore, a simple quadratic fit of
727: pointing offsets versus telescope elevation angle
728: served as our only pointing model.
729: Plots of this final model can be found in
730: Figure~\ref{fig:raw_pointing}.
731:
732: For each of the three subsets, we calculated the uncertainty
733: in the pointing model by analyzing the residual offset
734: of each centroid location from the model.
735: Some residual scatter is expected due to the measurement
736: uncertainty of each centroid,
737: however the scatter we find is slightly larger.
738: The difference between the actual scatter and the
739: predicted scatter is consistent for all three subsets,
740: and translates to an uncertainty in the pointing
741: model of 4.9~arcseconds.
742: This uncertainty is small compared to our beam size
743: and thus made a negligible difference in the beam shape
744: used in the final science analysis.
745:
746: \subsection{Flux Calibration}
747: \label{sec:flux_cal}
748:
749: Our flux calibration technique,
750: summarized below,
751: has been used previously with Bolocam
752: to calibrate 1.1~mm data \citep{laurent05}.
753: Since the amount of astronomical signal attenuation by the
754: atmosphere is a function of opacity and airmass,
755: the standard flux calibration technique for millimeter-wave
756: instruments requires frequent observations of calibration
757: sources that are close to the science field.
758: However, we were able to use a more advanced technique
759: with Bolocam because we continuously monitor the
760: operating resistance of the bolometers using the
761: carrier amplitude measured by the bolometer voltage
762: at the bias frequency.
763: When the atmospheric transmission decreases, the optical
764: loading from the atmosphere increases, which
765: lowers the bolometer resistance.
766: Additionally, the bolometer responsivity is a monotonically
767: decreasing function of the bolometer resistance.
768: Therefore, by fitting the flux calibration as a function
769: of the bolometer operating resistance, we can simultaneously account
770: for changes in the atmospheric transmission and
771: bolometer responsivity.
772:
773: Six different flux calibrations were needed for our data set.
774: The base temperature
775: of the sub-Kelvin refrigerator
776: was changed on November 4, 2003, and the
777: bias voltage applied
778: to the bolometers was changed on November 5, 8 (twice),
779: and 10, 2003.
780: % These bias voltage
781: % changes were not intentional; they were caused by an instability
782: % in the electronics used to generate our bias signal.
783: Each of the bias changes caused a change in
784: the responsivity of the bolometers, so a different flux calibration
785: is needed after each change.
786: Since the first five data sets are relatively short in duration,
787: the observing conditions were relatively constant
788: within each set.
789: Therefore, a constant flux calibration, rather than a flux
790: calibration that varies as a function of bolometer operating resistance,
791: was adequate to describe the data for these five sets.
792: However, a fit of the flux calibration as a function
793: of bolometer operating resistance was required for the final data set.
794:
795: The relative calibration of the detectors was determined
796: from the science field observations.
797: Since these observations covered regions of the sky with negligible
798: amounts of astronomical flux,
799: the fluctuations in thermal emission from the atmosphere are the
800: dominant source of the signal recorded by each bolometer.
801: Additionally, this signal should be the only one that is correlated
802: among all the bolometers
803: since the beams from
804: all bolometers overlap to a high degree when passing
805: through the atmosphere.
806: Therefore, this signal should be the same in each
807: bolometer, weighted by the responsivity of that bolometer.
808: So, by determining how correlated the data from each bolometer
809: is with this common signal, it is possible to determine the relative
810: calibration of each bolometer.
811: The uncertainties in the relative calibrations determined
812: using this method are less than 1\%.
813:
814: The absolute flux calibration was determined from
815: observations of Uranus, Neptune,
816: 0923+392, and NGC2071IR.
817: Since we did not have enough observations of Uranus and Neptune
818: to adequately determine the shape of the calibration versus
819: bolometer operating resistance, we used 0923+392 and NGC2071IR as
820: secondary calibrators.
821: These two sources are known to have minimal
822: variations in emitted flux as a function of time, so they are well
823: suited to be used for determining the functional form
824: of the flux calibration versus bolometer operating
825: resistance relationship \citep{peng00,sandell94}.
826: Note that we did not use any of the published fluxes
827: for 0923+392 or NGC2071IR, rather the fluxes were left as
828: free parameters and they were used to determine
829: the shape of the calibration curve versus
830: bolometer operating resistance.
831: We used the peak signal and median bolometer operating resistance
832: from each observation to determine the fit parameters in the
833: function
834: \begin{displaymath}
835: V_{j}(R_{bolo}) = F_{j} ( \alpha_1 + \alpha_2 R_{bolo} ),
836: \end{displaymath}
837: where $V_{j}$ is the peak bolometer signal (in nV) recorded
838: for the $j^{th}$ source, $R_{bolo}$ is the bolometer operating resistance,
839: $F_j$ is
840: equal to the flux of the $j^{th}$ source (known for Uranus and
841: Neptune, left as a free parameter for NGC2071IR and 0923+392),
842: and $\alpha_1$ and $\alpha_2$ free parameters.
843: The planet fluxes were determined from the
844: temperature spectra
845: given in \citet{griffin93} or \citet{orton86},
846: along with the planet solid angles
847: calculated from the planet flux calculator at the
848: James Clerk Maxwell Telescope website.\footnote{
849: http://www.jach.hawaii.edu/jacbin/planetflux.pl.}
850: For reference,
851: the absolute calibration ranges from approximately 180~nV/Jy up to
852: 280~nV/Jy over the range of
853: bolometer operating resistances recorded during
854: our observing run.
855: See Figure~\ref{fig:flux_slope}.
856:
857: Our flux calibration uncertainty was determined as follows.
858: First, the temperature profiles of Uranus and Neptune were derived
859: by Griffin and Orton using Mars as an absolute
860: calibrator \citep{griffin93}.
861: To determine the surface brightness of Mars at millimeter
862: wavelengths, Griffin and Orton used the model
863: developed by Wright based on observations
864: made at far-infrared wavelengths \citep{wright76},
865: along with the logarithmic interpolation to
866: longer wavelengths described by \citet{griffin86}.
867: The estimated uncertainty on this interpolated model is
868: approximately 5\% \citep{wright76}.\footnote{
869: There is also a brightness model based on a
870: physical model of the dielectric properties
871: of the Martian surface that was developed
872: by Rudy \citep{rudy87, rudy87_2}.
873: This model was constrained by measurements at
874: centimeter wavelengths, and also needs to be
875: extrapolated to millimeter wavelengths.
876: Griffin and Orton, along with Goldin, et al.,
877: compared the results of these
878: two models at millimeter wavelengths, and
879: found that they agree within their estimated
880: uncertainties \citep{griffin93, goldin97}.
881: Based on the comparison of these two models,
882: Griffin and Orton conclude the the
883: uncertainty in the Martian brightness based
884: on the Wright model is 5\%.}
885: Second, the uncertainties on the temperature profiles of Uranus and
886: Neptune are estimated to be less than
887: 1.5\% relative to Mars \citep{griffin93}.\footnote{
888: Griffin and Orton find that the uncertainty is 1.7~K
889: for both their Uranus and Neptune models.
890: Since the temperature of these planets in our band
891: is approximately 115~K, this translates to an
892: uncertainty of $\simeq 1.5$\%.}
893: Additionally, the observations of
894: Uranus and Neptune were taken with a precipitable
895: water vapor of $1.5 \pm 0.5$~mm, which results in a calibration
896: uncertainty of $\sim 1.4$\%.
897: Finally, the error inferred by the scatter of our measurements
898: results in calibration uncertainties between 0.6\% and 3.0\% for
899: each of the data sets.
900: The end result is an overall flux calibration uncertainty
901: of approximately 5.5\%, limited by the uncertainty
902: in the temperature of Mars.
903:
904: \subsection{Beam Calibration}
905:
906: Since the astronomical signals in our maps are inherently
907: smoothed based on the profile of the Bolocam
908: beams, it is important to understand their shapes.
909: Additionally,
910: our flux calibration is based on observations of point sources,
911: so our maps have units of flux density.
912: However, since the CMB or SZE signal we are looking for is a surface
913: brightness or temperature, we need to know the area of our beam
914: in solid angle to convert our maps to
915: surface brightness units.
916: Therefore, any error in our determination of the beam area will
917: show up as a surface brightness or temperature calibration
918: error.
919: To determine the profile of our beam, we used the
920: observations of Uranus and Neptune.
921: % These observations work well
922: % because both Uranus and Neptune have semi-diameters of
923: % $\simeq 1$~arcsecond, so they are nearly ideal point sources
924: % for our 60~arcsecond FWHM beams.
925: % This means that Uranus and Neptune will appear in our maps
926: % with a shape given by our beam profile.
927: These planets are well suited for measuring our
928: 60~arcsecond FWHM beams;
929: they have semi-diameters of $\simeq 1$~arcsecond,
930: which means they are essentially point-like and thus
931: will appear in our maps with shapes given
932: by our beam profile.
933:
934: % Since there was no a priori reason to assume that the beam
935: % profile of each individual bolometer is the same,
936: Based on simulations, we expected all of the beams
937: to have a similar profile.
938: However,
939: we first calculated the beam for each bolometer
940: separately to validate this expectation.
941: There was not enough data from a single planet observation
942: to make a high signal-to-noise measurement of the beam
943: for an individual bolometer, so we averaged the data
944: for groups of four bolometers that are close to each
945: other on the focal plane.
946: Nearby bolometers have beams with similar paths through the
947: optics, so they should also have similar profiles.
948: Each bolometer was grouped into four distinct sets, each
949: of which contained four nearby bolometers, and the
950: average profile from these four sets was determined.
951: The measurement uncertainty on these profiles can be
952: quantified by the standard deviation
953: of the peak-normalized
954: areas of the beam profiles, which was approximately
955: 3.1\%.
956: See Figure~\ref{fig:beam_area_variation}.
957: Within our measurement uncertainty, all of the individual
958: bolometer beam profiles were consistent, so a single
959: beam profile can be used to describe every bolometer.
960: To measure this single beam profile, we averaged
961: the data from all of the planet observations
962: for all of the bolometers.
963: The peak-normalized area of this profile is 3970~arcseconds$^2$,
964: which is the area of a Gaussian beam with a FWHM
965: of 59.2~arcseconds.
966: However, the beam profile is not exactly Gaussian, and
967: the measured profile was used for all of our analysis.
968: Since we cannot rule out systematic variations in the beam area
969: from one bolometer to the next at the level
970: of our single bolometer measurement uncertainty,
971: we have conservatively
972: estimated the uncertainty in this beam area measurement to be 3.1\%.
973:
974: \section{Map Making}
975:
976: \subsection{Least Squares Map Making Theory}
977:
978: The astronomical signals we seek can be
979: thought of as two-dimensional objects, which
980: can be represented by a map with finite pixelization.
981: For simplicity, this two-dimensional map can be thought
982: of as a vector, $\vec{m}$.
983: This map is stored in the bolometer time-streams,
984: $\vec{d}$, according to
985: \begin{equation}
986: \vec{d} = \mathbf{p} \vec{m} + \vec{n},
987: \label{eqn:TOD_map}
988: \end{equation}
989: where $\mathbf{p}$ is a matrix containing the pointing
990: information and $\vec{n}$ is noise.
991: Note that we represent matrices with a bold symbol, and
992: vectors with an arrow.
993: Since $\vec{m}$ is what we are fundamentally interested
994: in obtaining, we need to find a solution to
995: Equation~\ref{eqn:TOD_map} that yields the optimum unbiased
996: estimate of $\vec{m}$ given $\vec{d}$.
997: There are several methods that can be used to estimate
998: $\vec{m}$, including the commonly used least
999: squares method described below \citep{tegmark97, wright96}.
1000:
1001: Solving the least squares problem for Equation~\ref{eqn:TOD_map}
1002: requires minimizing
1003: \begin{equation}
1004: \chi^2 = (\vec{d} - \mathbf{p}\vec{m})^T
1005: \mathbf{w} (\vec{d} - \mathbf{p}\vec{m}),
1006: \label{eqn:chi_map}
1007: \end{equation}
1008: where $\mathbf{w}$ is the inverse of the time-stream noise
1009: covariance matrix, $\left< \vec{n}\vec{n}^T \right>^{-1}$.
1010: The estimator for $\vec{m}$ derived from
1011: Equation~\ref{eqn:chi_map} is
1012: \begin{equation}
1013: \vec{m}' = \mathbf{c}
1014: \mathbf{p}^T \mathbf{w} \vec{d},
1015: \label{eqn:map_est}
1016: \end{equation}
1017: where $\mathbf{c} = (\mathbf{p}^T \mathbf{w} \mathbf{p})^{-1}$
1018: is the map-space noise covariance matrix.
1019: If the time-stream noise, $\vec{n}$, has a white spectrum, then
1020: the various terms in Equation~\ref{eqn:map_est} are easy to
1021: understand because $\mathbf{w}$ and $\mathbf{c}$ are both
1022: diagonal.
1023: $\mathbf{w}$ is the inverse of the time-stream noise
1024: variance, and applies the appropriate weight to each sample
1025: in the time-stream.
1026: $\mathbf{p}^T$ then bins the data time-stream into a map,
1027: and $\mathbf{c}$ corrects for the fact that $\mathbf{p}^T$ sums
1028: all of the data in a single map bin instead of averaging it.
1029: The general idea is the same for non-white time-stream noise,
1030: but $\mathbf{w}$
1031: will mix time samples and $\mathbf{c}$ will mix map pixels.
1032:
1033: If the time-stream noise is stationary, then
1034: the time-stream noise covariance matrix can be diagonalized
1035: by applying the Fourier transform operator,
1036: $\mathbf{F}$.
1037: % \citep{natoli01}.
1038: In this case, any element
1039: of the inverse time-stream noise covariance matrix can be described by
1040: \begin{displaymath}
1041: \mathbf{w}(t_1,t_2) =
1042: \left< \vec{n(t_1)} \vec{n(t_2)}^T \right>^{-1} =
1043: \mathbf{w}(\Delta t),
1044: \end{displaymath}
1045: where $t_1$ and $t_2$ are any two time samples separated
1046: by $\Delta t$.
1047: The corresponding elements of the Fourier transform
1048: of the inverse covariance matrix,
1049: $\mathbf{W} = \mathbf{F} \mathbf{w} \mathbf{F}^{-1}$,
1050: can be written as
1051: \begin{displaymath}
1052: \mathbf{W}(f_1, f_2) =
1053: \mathbf{W}(f_1) \delta_{f_1,f_2},
1054: \end{displaymath}
1055: where $\delta_{f_1,f_2}$ represents a Kronecker delta and
1056: $f$ is frequency in Hz.\footnote{
1057: Note that physical space values are denoted with a lower
1058: case letter, and the corresponding frequency space
1059: values are denoted with an upper case letter.}
1060: The diagonal elements of $\mathbf{W}$ are equal to
1061: 1/(PSD*$\Delta f$), where PSD is the noise power spectral density
1062: and $\Delta f$ is the frequency resolution of the time-stream.
1063: The Kronecker delta ensures that all of the off-diagonal
1064: elements are equal to zero.
1065: Returning to Equation~\ref{eqn:map_est},
1066: the estimate for $\vec{m}$ can be rewritten as
1067: \begin{equation}
1068: \vec{m}' = ((\mathbf{p}^T \mathbf{F}^{-1})
1069: (\mathbf{F} \mathbf{w} \mathbf{F}^{-1})
1070: (\mathbf{F} \mathbf{p}))^{-1}
1071: (\mathbf{p}^T \mathbf{F}^{-1})
1072: (\mathbf{F} \mathbf{w} \mathbf{F}^{-1})
1073: (\mathbf{F} \vec{d}),
1074: \label{eqn:map_est_fourier1}
1075: \end{equation}
1076: using the fact that $\mathbf{F}^{-1} \mathbf{F} = 1$.
1077: Finally, taking the Fourier transform of the various terms in
1078: Equation~\ref{eqn:map_est_fourier1} yields
1079: \begin{displaymath}
1080: \vec{m}' = (\mathbf{P}^T \mathbf{W} \mathbf{P})^{-1}
1081: \mathbf{P}^T \mathbf{W} \vec{D}
1082: % \label{eqn:map_est_fourier}
1083: \end{displaymath}
1084: as an alternate expression to estimate the value of $\vec{m}$,
1085: where $\mathbf{P} = \mathbf{F}\mathbf{p}$,
1086: ${\vec{D}} = \mathbf{F}{\vec{d}}$,
1087: and $\mathbf{W} = \mathbf{F}\mathbf{w}\mathbf{F}^{-1}$.
1088: Note that
1089: $\mathbf{c} = (\mathbf{p}^T \mathbf{w} \mathbf{p})^{-1} =
1090: (\mathbf{P}^T \mathbf{W} \mathbf{P})^{-1}$ does not
1091: in general simplify as a result of Fourier transforming.
1092:
1093: \subsection{The Bolocam Algorithm: Theory}
1094: \label{sec:bolo_map_theory}
1095:
1096: The science field maps produced by Bolocam each contain
1097: $n_p \simeq 20000$ pixels, and an extremely large matrix
1098: must be inverted to calculate $\mathbf{c}$ since
1099: $\mathbf{P}^T\mathbf{WP} = \mathbf{p}^T\mathbf{wp}$
1100: has dimensions of $n_p \times n_p$.
1101: Direct inversion of such a matrix is possible, but is not practical
1102: on a typical high-end desktop computer.
1103: The map could be determined on a desktop computer via
1104: a conjugate gradient solver, but determining the covariance
1105: matrix would require a significant amount of simulation
1106: power.
1107: Therefore, we developed an algorithm to approximate
1108: $\vec{m}'$ by exploiting the simplicity of our
1109: scan pattern, which involved raster scanning
1110: parallel to either the RA or dec axis.
1111: This approximation allows us to make maps in a relatively
1112: short amount of time using a standard desktop computer,
1113: which is extremely convenient.
1114:
1115: To illustrate this simplification, consider the map
1116: made from a single bolometer for a single scan
1117: within an observation.
1118: This scan will produce a one-dimensional map
1119: at a single dec value (for an RA scan) or
1120: a single RA value (for a dec scan).
1121: Each data point in the time-stream is separated
1122: by 24~arcseconds in map-space
1123: since our data are sampled at 10~Hz and
1124: the telescope scans at 240~arcseconds/sec.
1125: Therefore, our data is approximately Nyquist sampled
1126: for Bolocam's $\simeq 60$~arcsecond FWHM beams.
1127: The maps are binned with 20~arcsecond pixels
1128: (1/3 of the beam FWHM, and slightly finer than
1129: Nyquist sampled),
1130: so $\mathbf{p}^T$ will map either one or zero time-stream
1131: samples to each map pixel.
1132: Note that $n_s$, the number of time-stream samples,
1133: will be slightly less than $n_p$, the number of map-space pixels.
1134: Since $\mathbf{p}^T$ has dimensions of $n_s \times n_p$,
1135: the sum of each row in $\mathbf{p}^T$ is either one or zero and
1136: the sum of each column is one.
1137: Consequently, we will make the approximation that
1138: $\mathbf{p}^T = 1$.
1139: From Equation~\ref{eqn:map_est},
1140: this means that
1141: \begin{equation}
1142: \mathbf{c} = \mathbf{w}^{-1},
1143: \label{eqn:map_corr_single_scan}
1144: \end{equation}
1145: and therefore $\vec{m} = \vec{d}$ for a single scan of
1146: time-stream data.
1147: If we Fourier transform Equation~\ref{eqn:map_corr_single_scan},
1148: then we find that
1149: \begin{displaymath}
1150: \mathbf{C} = \mathbf{W}^{-1}.
1151: % \label{eqn:map_corr_single2}
1152: \end{displaymath}
1153: Since $\mathbf{W}$ is diagonal, the inversion is trivial,
1154: and the result is that
1155: the Fourier transform of the map-space noise covariance
1156: matrix is diagonal with elements equal to the time-stream
1157: PSD*$\Delta f$.
1158:
1159: The next step is to consider a map made from a
1160: single bolometer for a full
1161: observation, which contains twenty scans.
1162: We move the telescope in the orthogonal direction to
1163: the scan between scans by more than the size of
1164: a single map pixel,
1165: so we can still approximate $\mathbf{p}^T \approx 1$.
1166: There are almost no correlations between scans
1167: because the atmospheric-noise subtraction coefficients
1168: are calculated scan-by-scan along with subtraction
1169: of the mean signal level.
1170: The covariance of maps made for a single
1171: observation from alternate scans is negligible, supporting
1172: this assumption that individual scans are uncorrelated.
1173: % We have confirmed that the correlations between scans
1174: % are negligible by making two maps for each observation:
1175: % one from the right-going (even) scans and one from
1176: % the left-going (odd) scans.
1177: % The correlations between these two maps are consistent
1178: % with noise, indicating that the data from
1179: % different scans are independent.}.
1180: % There will be almost no correlations between the data
1181: % in one scan and the data in all the other scans because the
1182: % time-stream data are effectively high-pass filtered on the
1183: % time scale of a scan by the atmospheric noise
1184: % subtraction algorithms.
1185: Therefore, the time-stream data and map-space data
1186: for different scans are essentially independent.\footnote{
1187: To verify that the data from different scans are independent,
1188: we created maps for each observation from all the
1189: odd-numbered (right-going) scans and from all the
1190: even-numbered (left-going) scans.
1191: The cross PSDs of the right-going maps with the left-going
1192: maps were consistent with noise, indicating that the
1193: data from separate scans are independent.}
1194: Consequently, the noise in map-space will be
1195: stationary, which means that the noise covariance matrix
1196: can be diagonalized by Fourier transforming it.
1197: The Fourier transform of the full-map noise covariance
1198: matrix, $\mathbf{C}$, can be visualized by noting that each
1199: diagonal element corresponds to a single Fourier-space map
1200: pixel (or equivalently, a single Fourier-space time-stream sample).
1201: So, this visualization of $\mathbf{C}$ will be
1202: equal to the single scan time-stream PSD*$\Delta f$ for rows of
1203: map-space pixels that are parallel to the scan
1204: direction, and will have a white spectrum
1205: for columns of map-space pixels that are perpendicular
1206: to the scan direction.
1207: Alternatively, since there is a one-to-one correspondence
1208: between time-stream samples and map-space pixels,
1209: this visualization of the diagonal elements of $\mathbf{C}$ is
1210: equal to the full map-space PSD*$\Delta f_{\Omega}$,
1211: where $\Delta f_{\Omega}$ is the angular frequency
1212: resolution of the map.
1213:
1214: At this point, we need to add together all of the
1215: individual observations to make a single map.
1216: Since we have shown that the map-space data are
1217: equivalent to the time-stream data for a single
1218: observation, the easiest way to co-add data from
1219: separate observations is to use the single
1220: observation maps.
1221: Since the noise in separate observations is uncorrelated,
1222: the maps can be co-added according to
1223: \begin{equation}
1224: \vec{m} = \left( \sum_i \mathbf{c}^{-1}_i \right)^{-1}
1225: \sum_j \mathbf{c}^{-1}_j \vec{m}_j,
1226: \label{eqn:map_obs_coadd}
1227: \end{equation}
1228: where the subscripts $i$ and $j$ refer to observation number.
1229: The easiest way to evaluate Equation~\ref{eqn:map_obs_coadd}
1230: is to Fourier transform it so that the noise
1231: covariance matrices are all diagonal.
1232: The result is
1233: \begin{equation}
1234: \vec{M} = \left( \sum_i \mathbf{C}^{-1}_i \right)^{-1}
1235: \sum_j \mathbf{C}^{-1}_j \vec{M}_j,
1236: \label{eqn:map_obs_coadd2}
1237: \end{equation}
1238: where $\vec{M}$ is the Fourier transform of the map
1239: and $\mathbf{C}$ is the Fourier transform of the noise covariance
1240: matrix, with diagonal elements equal to the PSD*$\Delta f_{\Omega}$
1241: of the map.
1242: Since all of the $\mathbf{C}$s are diagonal, we can simplify
1243: Equation~\ref{eqn:map_obs_coadd2} to
1244: \begin{equation}
1245: M = \left( \sum_i \frac{1}{\mathcal{P}_i} \right)^{-1}
1246: \sum_j \frac{M_j}{\mathcal{P}_j},
1247: \label{eqn:map_obs_coadd3}
1248: \end{equation}
1249: where $M$ is the two-dimensional Fourier transform of the
1250: map and $\mathcal{P}$ is the two-dimensional PSD
1251: of the noise in the map.
1252: At this point we have dropped the vector and matrix notation
1253: since $M_i$ and $P_i$ have the same dimensions.
1254: Note that $\Delta f_{\Omega}$ is the same for every map, so the constant
1255: factor of $\Delta f_{\Omega}$ from the first sum in
1256: Equation~\ref{eqn:map_obs_coadd3} cancels the factor
1257: of 1/$\Delta f_{\Omega}$ from the second sum in
1258: Equation~\ref{eqn:map_obs_coadd3}.
1259:
1260: Finally, to make a map using all of our data,
1261: we need to consider every bolometer, not just a single
1262: detector.
1263: To properly weight the data from each bolometer
1264: prior to co-adding, we
1265: calculate the expected variance, $(\sigma_{pf})^2_i$,
1266: in measuring
1267: the peak flux of a point-like source from a single
1268: scan through the center of the source for
1269: bolometer $i$.
1270: This variance is calculated using the scan-averaged
1271: time-stream PSD for each bolometer,
1272: $PSD_i(f)$, and the Fourier transform
1273: of the expected signal shape of a point-like
1274: astronomical signal, $S(f)$,
1275: according to
1276: \begin{equation}
1277: (\sigma_{pf})^2_i = \left( \int df \frac{S(f)^2}{PSD_i(f)}
1278: \right)^{-1}
1279: \label{eqn:relsens}
1280: \end{equation}
1281: where $f$ is temporal frequency.
1282: Note that $S(f)$ is the beam profile, not a delta function.
1283: Then, the data from each bolometer is weighted
1284: by a factor proportional to $1/(\sigma^2_{pf})$
1285: prior to co-adding it with data from other bolometers.
1286: This is the optimal way to co-add the
1287: data for point-like signals; it is nearly
1288: optimal for signals of any shape if the
1289: PSDs have similar profiles for every bolometer,
1290: which is largely true for our bolometer signals since they
1291: are dominated by atmospheric noise.
1292:
1293: However, due to atmospheric noise, along with our noise
1294: removal algorithms, there are correlations between
1295: the bolometers.
1296: But, most of these correlations are instantaneous in time
1297: and constant over the observation.\footnote{
1298: We have been able to find a small amount of correlated
1299: atmospheric signal that is not time-instantaneous.
1300: However, the time lag of these correlations is generally
1301: much less than one time sample, which means they
1302: will also be less than one map pixel.}
1303: Additionally, the relative positions of the bolometers
1304: do not change during the observation, so the
1305: map-space separation of the correlations does not change.
1306: % Therefore, the Fourier transform
1307: % of the time-stream noise covariance matrix will still
1308: % be diagonal.
1309: % The reason that this $\mathbf{W}^{-1}$ is diagonal
1310: % is because
1311: Therefore, the correlations are stationary in time
1312: with separations that are fixed in map-space,
1313: so the correlations are an additional time-independent
1314: covariance between map pixels that are sampled at the
1315: same time by different bolometers.
1316: This additional covariance is approximately stationary
1317: over the entire map, except where it breaks down near
1318: the edges because part of the focal plane is outside
1319: the map region.
1320: Since this additional covariance between map pixels is
1321: approximately stationary in space, its contribution to
1322: $\mathbf{W}^{-1}$ will be diagonal.
1323: % the correlations do not depend on
1324: % the time-stream sample number of either bolometer,
1325: % nor on the relative difference in time-stream samples
1326: % for each bolometer.
1327: Since $\mathbf{W}^{-1}$ is still diagonal, co-addition of the maps of
1328: individual observations can proceed according to
1329: Equation~\ref{eqn:map_obs_coadd3}.
1330: Therefore, Equation~\ref{eqn:map_obs_coadd3} can
1331: be used as the algorithm to produce our final
1332: science field maps.
1333: % , with
1334: % \begin{displaymath}
1335: % M = \left( \sum_i \frac{1}{\mathcal{P}_i} \right)^{-1}
1336: % \sum_j \frac{M_j}{\mathcal{P}_j},
1337: % \end{displaymath}
1338: % where $M$ is the Fourier transform of the full-data map,
1339: % $\mathcal{P}_i$
1340: % is the two-dimensional PSD (in $\mu$K$^2_{CMB}$ ster)
1341: % of the noise in the map made from observation $i$,
1342: % and $M_i$ is the Fourier transform of the map made
1343: % from observation $i$.
1344:
1345: Note that we were forced to make several simplifying
1346: assumptions in order to develop Equation~\ref{eqn:map_obs_coadd3}.
1347: We have assumed that the pointing matrix, $\mathbf{p}^T$,
1348: is equal to one.
1349: We have also assumed that the noise in our time-streams
1350: is stationary for each eight-minute-long observation.
1351: Additionally, we have assumed that the PSD
1352: of the correlations between bolometers is white, and
1353: that all of the correlations are time-instantaneous.
1354: Finally, we have assumed that the map coverage
1355: (\emph{i.e.}, the number of time-stream samples that are
1356: binned in each map-space pixel) is uniform,
1357: so that the Fourier transform of the map is
1358: a valid description of the time-stream data.
1359: Deviations from these assumptions will alter the
1360: map estimate we compute from the optimal least squares
1361: map estimate.
1362: But, these deviations only affect how each time-stream
1363: sample is weighted before it is mapped.
1364: This means our final map will have more noise than
1365: an optimal map, but it will not be biased in any way.
1366: In other words, since the map-making operation is linear,
1367: the resulting map will be unbiased no matter what
1368: weightings are used to co-add the data, as
1369: long as the weights are properly normalized.
1370: We have confirmed this lack of bias via simulation,
1371: as we discuss below.
1372:
1373: \subsection{The Bolocam Algorithm: Implementation}
1374: \label{sec:mapmaking}
1375:
1376: To start, we must first produce a map from the time-stream
1377: data for each eight-minute-long observation.
1378: As mentioned in Section~\ref{sec:bolo_map_theory},
1379: this is done by calculating the variance in measuring
1380: the peak flux of a point-like source under the assumption that
1381: the profile of the time-stream PSD
1382: is similar for every bolometer.
1383: To determine this variance, we calculate the
1384: PSD for each bolometer for each scan.
1385: These spectra are then averaged over all twenty scans
1386: for each bolometer, thereby making the assumption that the
1387: noise properties do not change over the course
1388: of the observation.
1389: Then, we determine the expected shape of a point-like
1390: source using our measured beam profile and scan speed.
1391: Finally, Equation~\ref{eqn:relsens} is used to determine
1392: the variance in measuring the peak flux
1393: of a point-like source for each bolometer,
1394: which is inversely proportional to the weighting factor
1395: applied to the time-stream data
1396: for that bolometer.
1397:
1398: At this point, we have individual observation maps
1399: for every observation,
1400: and we can make a map from all of the data using
1401: Equation~\ref{eqn:map_obs_coadd3}.
1402: But, one of the main assumptions made in developing
1403: Equation~\ref{eqn:map_obs_coadd3} was that the
1404: map coverage is uniform for each observation.
1405: If this assumption fails, then the Fourier transform of the
1406: map is not a good description of the time-stream data.
1407: Our scan strategy produced highly uniform
1408: coverage in the central region of the map, and this
1409: coverage falls rapidly to zero at the edges of the map.
1410: See Figure~\ref{fig:mapcov_single_obs}.
1411: % If we restrict our map to a square region in the center, with
1412: To obtain sufficiently uniform coverage, we restrict our map
1413: to have
1414: sides of 42~arcminutes; the fractional RMS variations in
1415: coverage within this region for a single eight-minute-long observation
1416: are only about 8 - 9\%.
1417: Since the coverage variations are minimal, we will assume
1418: that this square central region has uniform coverage,
1419: and therefore uniform noise properties.
1420: This assumption of uniform coverage allows us to directly
1421: compute the Fourier transform and noise properties
1422: of the map.
1423: We emphasize that, even if the assumption of uniform coverage fails
1424: and our algorithm is non-optimal, it is never biased
1425: because Equation~\ref{eqn:map_obs_coadd3} is linear
1426: in the map-space maps.
1427:
1428: We now have a uniform coverage map for each observation,
1429: which can easily be Fourier transformed to produce
1430: the $M_i$s needed in Equation~\ref{eqn:map_obs_coadd3}.
1431: But, we still need to determine the two-dimensional
1432: PSD of each single observation map.
1433: Due to residual correlations between bolometers,
1434: we do not understand the noise properties of our
1435: data well enough to determine the map PSD from simulation,
1436: so we instead estimate the PSD by generating a large
1437: number of jackknifed maps from our real data.
1438: In each jackknifed map, a different subset of
1439: the time-streams from half
1440: of the scans within each observation was multiplied by $-1$.
1441: Note that the data from all of the bolometers within a single
1442: scan are multiplied by $-1$, so the residual atmospheric noise
1443: that is correlated between bolometers is preserved.
1444: This multiplication leaves the noise properties of the
1445: map unchanged,\footnote{
1446: Each time-stream sample (and therefore each map-space pixel)
1447: can be expressed as the sum of two signals:
1448: 1) an astronomical signal and 2) a random noise signal that
1449: is drawn from the underlying distribution of the noise in the
1450: Bolocam system. The astronomical signal
1451: corresponding to a particular map-space pixel will be the
1452: same for any scan, and will disappear in the jackknife
1453: realizations when
1454: time-stream data from half of the scans is multiplied
1455: by $-1$. But, if the underlying distribution of the noise
1456: is Gaussian, then the distribution of signals it
1457: will produce is symmetric about 0.
1458: Therefore, the statistical properties of the noise
1459: will be unchanged when half of the data are multiplied
1460: by -1.}
1461: while allowing us to produce a large
1462: number of noise realizations for each map.
1463: Note that the residual atmospheric noise correlations are
1464: time-instantaneous, so they remain in the
1465: jackknifed realizations.
1466: We then generate 100 realizations for each observation,
1467: and we set the true PSD for each observation
1468: equal to the average of the map-space PSD computed for
1469: each realization.
1470: See Equation~\ref{eqn:map_psd_est}.
1471: Examples of the PSDs we calculated are given in
1472: Figure~\ref{fig:map_psd_single_obs}.
1473: This method of determining the map-space PSDs assumes that
1474: the time-stream data for each scan is uncorrelated with the
1475: data from all other scans, which we argued in
1476: Section~\ref{sec:bolo_map_theory}.
1477: % We make the same assumption in developing our map-making
1478: % algorithm, and it is reasonable since the time between scans
1479: % is larger than the cutoff of the high-pass filter
1480: % we apply to the time-stream data.
1481:
1482: To determine the validity of the map-space PSDs we estimated
1483: from the jackknifed map realizations,
1484: we examined the distribution of PSD values for each
1485: realization.
1486: If the noise properties of the data are Gaussian, as we have assumed,
1487: then the PSD measured at any given Fourier map-space pixel
1488: will be drawn from
1489: \begin{equation}
1490: f(X_{i,\vec{\nu}}) = (1/\mathcal{P}_{\vec{\nu}})
1491: e^{(-X_{i,\vec{\nu}}/\mathcal{P}_{\vec{\nu}})},
1492: \label{eqn:PDF_map_PSD}
1493: \end{equation}
1494: where $X_{i,\vec{\nu}}$ is the measured PSD for
1495: realization $i$ at pixel $\vec{\nu}$,
1496: $\mathcal{P}_{\vec{\nu}}$ is the true PSD for
1497: pixel $\vec{\nu}$, and $f(X_{i,\vec{\nu}})$ is the probability
1498: density function of $X_{i,\vec{\nu}}$.
1499: Note that $\vec{\nu}$ has units of spatial frequency
1500: (i.e., radians$^{-1}$), and describes
1501: a pixel in the spatial Fourier transform of the map.
1502: See Appendix~\ref{sec:map_var} for a derivation of $f(X_{i,\vec{\nu}})$.
1503: The true PSD is estimated from
1504: \begin{equation}
1505: \widehat{\mathcal{P}_{\vec{\nu}}} =
1506: \frac{1}{N_r} \sum_{i=1}^{i=N_r} X_{i,\vec{\nu}},
1507: \label{eqn:map_psd_est}
1508: \end{equation}
1509: where $N_r = 100$ is the number of realizations.
1510: To compare our measured PSDs to the probability density
1511: function (PDF) given in Equation~\ref{eqn:PDF_map_PSD},
1512: we created the dimensionless value
1513: \begin{equation}
1514: Y_{i,\vec{\nu}} =
1515: \frac{X_{i,\vec{\nu}}}{\widehat{\mathcal{P}_{\vec{\nu}}}},
1516: \label{eqn:dimless_PSD}
1517: \end{equation}
1518: with associated PDF
1519: \begin{equation}
1520: f'(Y_{i,\vec{\nu}}) = e^{-Y_{i,\vec{\nu}}}.
1521: \label{eqn:PDF_map_PSD2}
1522: \end{equation}
1523: Then, we compared our measured values of $Y_{i,\vec{\nu}}$ to
1524: the PDF in Equation~\ref{eqn:PDF_map_PSD2}.
1525: In general, we found that
1526: our measured $Y_{i,\vec{\nu}}$ follow a
1527: distribution extremely close to $f'(Y_{i,\vec{\nu}})$,
1528: except that the number of $Y_{i,\vec{\nu}}$
1529: with values near zero is slightly less than expected.
1530: Therefore, the map-space PSDs estimated from the jackknife
1531: realizations should be a good estimate of the true map-space
1532: PSDs.
1533:
1534: %start Sunil's addition for referee
1535: Let us consider the possible effects of imperfect signal removal in
1536: the single-observation jackknife maps.
1537: Because we only use the single-observation noise estimates as weights
1538: for coadding,
1539: the result of
1540: residual signal in the jackknife maps will be
1541: coaddition weights that are non-optimal.
1542: This non-optimality may degrade the noise of the final maps,
1543: but will not cause them to be biased.
1544: We may estimate the size of the signal leakage
1545: to determine how large the deviation
1546: from optimality could be.
1547: Our final flat band power anisotropy upper limit is
1548: approximately 1000~$\mu$K$_{CMB}^2$.
1549: With our effective $\Delta \ln(\ell)$ of 0.63
1550: (derived in Section~\ref{sec:xfer}),
1551: this upper limit corresponds to an excess
1552: variance in our final maps of
1553: $\simeq 500$~$\mu$K$_{CMB}^2$.
1554: These final maps have a variance of
1555: 10000~$\mu$K$_{CMB}^2$ (see Figure~\ref{fig:final_maps}).
1556: Given that $\simeq 500$ observations contribute to each map,
1557: the single-observation map variance is
1558: $\simeq 5 \times 10^6$~$\mu$K$_{CMB}^2$,
1559: or approximately 10000 times larger than our upper limit
1560: on the astronomical signal contribution.
1561: Even if we did not remove the astronomical signal using
1562: the jackknifing procedure,
1563: it would affect the single-observation PSDs,
1564: and therefore the weights,
1565: at only the 0.01\% level.
1566: Using jackknife-generated PSDs reduces the
1567: effect of signal contamination further.
1568: Therefore, the effect of signal leakage
1569: into the single-observation jackknife maps is negligible.
1570:
1571: % First, note that we only
1572: % use the single-observation noise estimate as a weight for coadding.
1573: % Therefore,
1574: % residual signal in the jackknife maps
1575: % will produce non-optimal coadding weights, but
1576: % such weights will cause no bias in the final map.
1577: % However, they will cause the final map to have slightly more
1578: % noise compared to an optimal map.
1579: % We emphasize that we do not estimate
1580: % the expected final map noise using these
1581: % single-observation noise PSDs.
1582: % Second, we have calculated the size of the effect for the sake of
1583: % completeness.
1584: % Our final flat band power anisotropy upper limit is
1585: % approximately 1000~$\mu$K$_{CMB}^2$.
1586: % With our effective $\Delta \ln(\ell)$ of 0.63
1587: % (derived in Section~\ref{sec:xfer}),
1588: % this upper limit corresponds to an excess
1589: % variance in our final maps of
1590: % $\simeq 500$~$\mu$K$_{CMB}^2$. Our final
1591: % maps have a variance of approximately
1592: % 10000~$\mu$K$_{CMB}^2$ (see Figure~\ref{fig:final_maps}),
1593: % so a single-observation map has a variance of
1594: % $\simeq 5 \times 10^6$~$\mu$K$_{CMB}^2$
1595: % since there are $\sim 500$ observations in each map.
1596: % Thus, our upper limit on the signal is
1597: % a factor of approximately $1 \times 10^4$
1598: % smaller in $\mu$K$_{CMB}^2$ than the
1599: % single-observation
1600: % noise variance; it would affect the single-observation
1601: % noise variance estimate at the 0.01\% level.
1602: % Therefore, the effect of imperfect signal removal on the noise
1603: % estimates we obtain from our jackknife
1604: % maps is negligible.
1605: %end Sunil's addition
1606:
1607: \section{Transfer Functions}
1608: \label{sec:xfer}
1609:
1610: The transfer function describes the fraction
1611: of the astronomical signal that
1612: remains after processing as a function of
1613: map-space Fourier mode.
1614: In order to determine the transfer function of our
1615: data processing algorithms, we first generate
1616: a simulated map of the expected astronomical signal.
1617: This map is then reverse-mapped into
1618: a time-stream using the pointing
1619: information in a real observation.
1620: Next, this simulated time-stream is added to the
1621: real bolometer time-streams from the observation,
1622: and then processed and mapped in the standard way.
1623: A map made from data that did not have a simulated
1624: signal added to it is then subtracted from this map,
1625: producing a map with the simulated signal after
1626: processing.
1627: Finally, the PSD of this map is divided by the PSD
1628: of the original simulated signal map to determine
1629: how much of the signal remains.
1630: Note that we are computing the transfer function
1631: for a PSD because we are interested in measuring an
1632: excess noise and not a specific signal shape,
1633: which means we do not need the phase of the transfer function.
1634:
1635: This transfer function was computed for twenty randomly selected
1636: observations, ten taken while scanning parallel to
1637: RA and ten taken while scanning parallel to dec.
1638: Realizations of the expected flat-band power anisotropy signal were
1639: used as the simulated signal.
1640: These realizations were generated in Fourier map-space
1641: assuming Gaussian fluctuations and a flat
1642: band power in $\mathcal{C}_{\ell} = C_{\ell} \ell (\ell+1) / 2\pi$
1643: of 50~$\mu$K$_{CMB}^2$.
1644: Note that although a flat band power of 50~$\mu$K$_{CMB}^2$ was used
1645: for the simulated signal maps, we found that the
1646: transfer function is independent of
1647: the amplitude of the flat-band power anisotropy signal.
1648: For each observation, we averaged the transfer
1649: function obtained from 100 different signal realizations
1650: to determine the average transfer function.
1651: We then compared the average transfer function for each of the ten
1652: observations taken with a similar scan pattern.
1653: The result is that
1654: the transfer functions were the same within our measurement
1655: uncertainty for all of the observations.
1656: Therefore, we averaged the transfer function from all
1657: ten observations to produce a high signal-to-noise
1658: measurement for each atmospheric-noise removal method:
1659: average, planar, and quadratic.
1660: See Figure~\ref{fig:RA_xfer}.
1661:
1662: Since all of the data processing is performed on the time-streams,
1663: the attenuation caused by the processing has a preferred
1664: orientation based on the scan strategy.
1665: The result is a transfer function that is not azimuthally
1666: symmetric because of the large amount of attenuation
1667: at low frequencies parallel to the scan direction
1668: due to atmospheric noise removal.
1669: Additionally, there is massive attenuation on scales
1670: larger than the Bolocam focal plane ($\simeq 500$~radians$^{-1}$)
1671: because of the atmospheric noise removal algorithms.
1672: This occurs because these algorithms are designed to remove
1673: all time-instantaneous signals at each data sample,
1674: which is equivalent to subtracting any signals
1675: that vary slowly compared to the size of the focal plane.
1676:
1677: In addition to the signal attenuation caused by the
1678: data processing, the Bolocam system also attenuates
1679: some of the astronomical signal.
1680: By scanning across the sky, we are effectively convolving
1681: any signal with the profile of a Bolocam beam;
1682: since the beams have a non-zero width, this convolution
1683: will act like a low-pass filter on all of the
1684: astronomical signals.
1685: This filter will be approximately symmetric because
1686: the Bolocam beam profiles have a high degree
1687: of rotational symmetry.
1688: Additionally, since the beams are nearly Gaussian,
1689: the filter will be approximately Gaussian
1690: with a HWHM in variance of about 1000~radians$^{-1}$
1691: (which is equivalent to a HWHM$_{\ell} \simeq 6000$
1692: in angular multipole space).
1693: See Figure~\ref{fig:RA_xfer}.
1694:
1695: In order to quantify the amount of signal
1696: attenuation by each atmospheric noise removal algorithm,
1697: it is useful to determine the effective bandwidth
1698: of the transfer function.
1699: The effective bandwidth describes the range of angular
1700: multipoles to which we are sensitive,
1701: as quantified by the transfer function,
1702: and can be used to convert an angular
1703: power, $C_{\ell}$, to a map-space variance in $\mu$K$^2_{CMB}$.
1704: In general, the effective bandwidth is calculated by
1705: integrating the transfer function over all
1706: angular multipoles.
1707: However, since the expected SZE power spectrum is
1708: approximately flat in $\mathcal{C}_{\ell}$,
1709: which results in a spectrum in $C_{\ell}$
1710: that falls like $1/\ell(\ell+1)$,
1711: it is more useful to weight the transfer function
1712: by a factor of $1/\ell(\ell+1)$.
1713: This weighting will produce an effective logarithmic,
1714: rather than linear, bandwidth, and can
1715: be used to convert an angular power
1716: in $\mathcal{C}_{\ell}$ to a map-space variance.
1717: This effective logarithmic bandwidth,
1718: ${\rm BW}_{eff}$, is defined as
1719: \begin{displaymath}
1720: {\rm BW}_{eff} = \int_{\vec{\nu}} d\vec{\nu} S_{\vec{\nu}}^2
1721: T_{\vec{\nu}} B_{\vec{\nu}}^2,
1722: \end{displaymath}
1723: where $\vec{\nu}$ is the two-dimensional spatial frequency,
1724: $S_{\vec{\nu}}$ is the expected signal spectrum,
1725: $T_{\vec{\nu}}$ is the transfer function of the
1726: data processing in squared units, and $B_{\vec{\nu}}$ is the
1727: profile of the Bolocam beam.
1728: Since the expected anisotropy signal has a flat band power
1729: in $\mathcal{C}_{\ell}$,
1730: \begin{displaymath}
1731: S_{\vec{\nu}}^2 \propto \frac{1}{\ell(\ell + 1)}
1732: \end{displaymath}
1733: for $\ell = 2 \pi |\vec{\nu}|$.\footnote{
1734: We have used the
1735: small-scale flat sky approximation, $\ell = 2\pi |\vec{\nu}|$.}
1736: Assuming this spectrum for $S_{\vec{\nu}}^2$,
1737: a top-hat window between $\ell = \ell_{min}$ and
1738: $\ell = \ell_{max}$ will produce a bandwidth approximately
1739: equal to
1740: \begin{displaymath}
1741: {\rm BW}_{eff} \propto \ln(\ell_{max}) - \ln(\ell_{min}) =
1742: \Delta \ln(\ell).
1743: \end{displaymath}
1744: Although the Bolocam transfer functions are
1745: not azimuthally symmetric, it is still useful to determine
1746: the effective $\Delta \ln(\ell)$ for each
1747: of the atmospheric noise removal algorithms,
1748: with $\Delta \ln(\ell) = 0.98$, 0.58, and 0.37
1749: for average, planar, and quadratic subtraction.
1750: Note that our final map, which consists of observations
1751: processed with different atmospheric noise removal
1752: algorithms as described in Section~\ref{sec:opt_skysub},
1753: has a bandwidth of $\Delta \ln(\ell) \simeq 0.63$.
1754:
1755: \section{Optimal Atmospheric Noise Subtraction}
1756: \label{sec:opt_skysub}
1757:
1758: Each of the science field observations were processed
1759: with average, planar, and quadratic sky subtraction,
1760: creating three separate files for each observation.
1761: Quadratic subtraction removes the most atmospheric noise,
1762: while average subtraction retains the most astronomical
1763: signal, so there is an optimal sky subtraction
1764: algorithm for each observation
1765: based on the type of astronomical signal we are
1766: looking for.
1767: To determine which algorithm is optimal, we computed
1768: a figure of merit, FOM, for each subtraction method.
1769: Since the anisotropy signal appears as a variance in the map,
1770: the variance on the amplitude of the
1771: anisotropy signal will be proportional
1772: to the square of the map PSD divided by the
1773: transfer function of the experiment.
1774: This can be seen
1775: in Equations~\ref{eqn:est_a}~and~\ref{eqn:est_var_a}.
1776: Therefore, the FOM is defined as the inverse
1777: of this variance on the anisotropy signal summed over
1778: all angular scales according to
1779: \begin{equation}
1780: {\rm FOM} = \sum_{\vec{\nu}}
1781: \frac{(S_{\vec{\nu}}^2)^2 T_{\vec{\nu}}^2 (B_{\vec{\nu}}^2)^2}
1782: {\mathcal{P}_{\vec{\nu}}^2},
1783: \label{eqn:fom}
1784: \end{equation}
1785: where $\vec{\nu}$ is a two-dimensional spatial frequency with
1786: units of radians$^{-1}$, $S_{\vec{\nu}}^2$ is the expected
1787: anisotropy power spectrum, $T_{\vec{\nu}}$ is the transfer
1788: function of the data processing in squared units, $B_{\vec{\nu}}$
1789: is the profile of the Bolocam beam, and
1790: $\mathcal{P}_{\vec{\nu}}$ is the PSD of the noise
1791: in the map in squared units.
1792: Note that we have included the $\simeq 5$~arcsecond
1793: uncertainty in our pointing model in $B_{\vec{\nu}}$,
1794: and this pointing uncertainty effectively broadens
1795: the beam.
1796: To be precise,
1797: \begin{displaymath}
1798: B_{\vec{\nu}} = \mathsf{B}_{\vec{\nu}}
1799: e^{-|\vec{\nu}|^2 / 2 \sigma_{\nu}^2},
1800: \end{displaymath}
1801: where $\mathsf{B}_{\vec{\nu}}$ is the measured beam profile,
1802: and $\sigma_{\nu} = 1/ 2 \pi \sigma_p$ for
1803: a pointing uncertainty of $\sigma_p$.
1804: For the anisotropy spectrum, we assumed a flat band power in
1805: $\mathcal{C}_{\ell}$, so
1806: \begin{displaymath}
1807: S_{\vec{\nu}}^2 = \frac{1}{\ell(\ell+1)}
1808: \end{displaymath}
1809: for $\ell = 2 \pi |\vec{\nu}|$.
1810: The figure of merit is inversely proportional to the
1811: variance on an estimate of the anisotropy amplitude
1812: (in $\mu$K$_{CMB}^2$),
1813: so it characterizes the signal-to-noise ratio of the map.
1814:
1815: In the end, average subtraction was the optimal method for just over
1816: 50\% of the observations, planar subtraction was the
1817: optimal method for just over 40\% of the observations,
1818: and quadratic subtraction was the optimal method for
1819: just under 10\% of the observations.
1820: We can calculate how much the observations optimally cleaned
1821: by each method contribute to our final S/N from
1822: \begin{displaymath}
1823: \textrm{S/N} = \sqrt{\frac{\sum_{i \, \epsilon \, T} FOM_i^{-2}}
1824: {\sum_{i} FOM_i^{-2}}},
1825: \end{displaymath}
1826: where $T$ denotes the set of observations optimally cleaned
1827: by a given method
1828: and $FOM_i$ is the figure of merit from Equation~\ref{eqn:fom}.
1829: The S/N contributed by the average/planar/quadratic
1830: observations is 70/29/1\%.
1831: These ratios are different from the number of observations
1832: optimally cleaned by each method
1833: because the amount of atmospheric noise in the data
1834: generally determines which subtraction algorithm is optimal,
1835: and the observations optimally cleaned with average
1836: subtraction were made in the best observing conditions.
1837: Note that quadratic subtraction is the optimal method
1838: only when the weather conditions are extremely poor.
1839: This is because the anisotropy power
1840: spectrum falls quickly at high frequency,
1841: and the quadratic subtraction algorithm attenuates a large
1842: amount of signal at low frequency.
1843: For point-like sources, whose spectra are flatter, quadratic
1844: subtraction is the optimal processing method slightly more often.
1845:
1846: \section{Final Map Properties}
1847:
1848: Once the FOM is determined for each subtraction method
1849: for each observation, we can then
1850: produce a map of all of the data using the optimally
1851: processed map for each observation.
1852: To produce this final map, we need to make a slight
1853: modification to Equation~\ref{eqn:map_obs_coadd3} to
1854: account for the transfer function of the data processing
1855: and the Bolocam beam.
1856: We need to account for these effects because the transfer
1857: function depends on the scan direction and optimal
1858: sky subtraction algorithm for each observation.
1859: Therefore, the amount of astronomical signal in the map
1860: is in general different for each observation.
1861: To account for the amount of signal attenuation in each
1862: observation, the map PSD
1863: needs to be divided by the transfer function and
1864: the Fourier transform of the map needs to be divided
1865: by the square root of the transfer function.
1866: After making these modifications to Equation~\ref{eqn:map_obs_coadd3},
1867: we have
1868: \begin{equation}
1869: \mathcal{M} = \frac{\sum_i \left( \frac{M_i}{\sqrt{T_i B_i^2}} \right)
1870: \left( \frac{T_i B_i^2}{\mathcal{P}_i} \right) }
1871: {\sum_j \left( \frac{T_j B_j^2}{\mathcal{P}_j} \right) }
1872: \label{eqn:opt_sig_map}
1873: \end{equation}
1874: as the Fourier transform of the optimal map estimate, $\mathcal{M}$.
1875: $T_i$ is the transfer function of the data processing for
1876: observation $i$ in squared units, $B_i$ is the Bolocam beam profile
1877: for observation $i$,
1878: $M_i$ is the Fourier transform of the map from observation $i$,
1879: and $\mathcal{P}_i$ is the noise PSD for observation $i$
1880: in squared units.
1881: Note that the astronomical signal in $\mathcal{M}$ will be
1882: equal to the true astronomical signal, because we have
1883: divided the Fourier transform of each single observation
1884: map, $M_i$, by the appropriate attenuation factor,
1885: $\sqrt{T_i B_i^2}$.\footnote{
1886: We have not included any phase information in the factor
1887: $\sqrt{T_i B_i^2}$ because both the signal and the noise
1888: PSD contain only noise; the phase is irrelevant.}
1889: However, for some pixels in Fourier space,
1890: $T_i$ and/or $B_i$ take on extremely small values,
1891: which means that some pixels in both the
1892: numerator and denominator of $\mathcal{M}$
1893: have extremely small values.
1894: Therefore, before taking the ratio of the numerator
1895: and denominator in Equation~\ref{eqn:opt_sig_map}
1896: we apply a regularizing factor, so that
1897: \begin{equation}
1898: M' = \sqrt{\mathcal{R}} \mathcal{M} =
1899: \frac{\frac{1}{\sqrt{\mathcal{R}}}
1900: \sum_i \left( \frac{M_i}{\sqrt{T_i B_i^2}} \right)
1901: \left( \frac{T_i B_i^2}{\mathcal{P}_i} \right) }
1902: {\frac{1}{\mathcal{R}}
1903: \sum_j \left( \frac{T_j B_j^2}{\mathcal{P}_j} \right) },
1904: \label{eqn:regularizing}
1905: \end{equation}
1906: for
1907: \begin{equation}
1908: \sqrt{\mathcal{R}} = \frac{\sum_i \left( \sqrt{T_i B_i^2} \right)
1909: \left( \frac{T_i B_i^2}{\mathcal{P}_i} \right)}
1910: {\sum_j
1911: \left( \frac{T_j B_j^2}{\mathcal{P}_j} \right)}.
1912: \label{eqn:regularizing2}
1913: \end{equation}
1914: Although $M'$ will be biased (\emph{i.e.}, it is not the
1915: Fourier transform of the true map of the sky),
1916: this bias is accounted for by the final transfer
1917: function we calculate in Section~\ref{sec:sig_att}.\footnote{
1918: In Equations~\ref{eqn:regularizing} and
1919: \ref{eqn:regularizing2}, $T_iB_i^2/\mathcal{P}_i$
1920: acts as a weighting factor for each observation.
1921: Therefore, $\mathcal{M}$ represents the
1922: weighted mean of the Fourier transform of
1923: each single observation map divided by the square
1924: root of the transfer function for that map,
1925: $\overline{(M/\sqrt{TB^2})}$.
1926: Similarly, $\sqrt{R}$ represents the weighted
1927: mean of the square root of the transfer function
1928: for each observation, $\overline{\sqrt{TB^2}}$.
1929: So, $M' =
1930: (\overline{\sqrt{TB^2}}) (\overline{(M/\sqrt{TB^2})})$,
1931: which reduces to the weighted mean
1932: of all the single observation map Fourier
1933: transforms, $M' \simeq \overline{M}$, in the limit
1934: that all of the single observation transfer
1935: functions, $T_iB_i^2$, are the same.}
1936: Note that $M'$ can be Fourier transformed back to
1937: map-space to produce a map $m'$,
1938: although $m'$ will be biased.
1939: The maps, $m'$, for each science field are given in
1940: Figure~\ref{fig:final_maps}.
1941:
1942: \subsection{Noise PSDs}
1943:
1944: Analogous to the case of a single observation, we used
1945: jackknifed realizations of our data to estimate the
1946: noise PSD of $m'$.
1947: In this case, each realization is
1948: generated by multiplying a randomly
1949: selected set of half the observations in $m'$ by
1950: $-1$.
1951: The map-space PSD from
1952: 1000 realizations were averaged to
1953: determine the best estimate of the
1954: noise PSD for
1955: each science field,
1956: with the results shown in Figure~\ref{fig:final_psd}.
1957: In this section we establish that the noise PSD
1958: estimated in this way is statistically well-behaved
1959: (Gaussian) and unbiased.
1960: These characteristics are critical to the remainder of our analysis.
1961:
1962: We analyzed the distribution of individual realization
1963: PSDs to determine if the underlying probability distribution
1964: describing the noise is Gaussian.
1965: As in the single observation case, we computed a dimensionless
1966: PSD value according to Equation~\ref{eqn:dimless_PSD},
1967: and compared the distribution of these values to the PDF
1968: given in Equation~\ref{eqn:PDF_map_PSD}.
1969: In general, the agreement is good, indicating
1970: the underlying noise distribution is well approximated
1971: by a Gaussian.
1972: % Specifically, the probability of getting a worse $\chi^2$
1973: % is 88\% and 25\% for the Lynx field and
1974: % SDS1 field, respectively.
1975: See Figure~\ref{fig:final_PSD_PDF}.
1976: The Gaussianity of the noise PSDs of the jackknife maps is
1977: important because it justifies the form of the likelihood
1978: function we use, Equation~\ref{eqn:log_l}
1979: (presented in Section~\ref{sec:overview_analysis}).
1980: That form assumes the Fourier coefficients of the final map
1981: are Gaussian-distributed random variables with variance given
1982: by the noise PSD estimated from the jackknife maps.
1983:
1984: %begin Sunils response to referee
1985:
1986: Next, we show that the noise PSD estimated from the jackknifes
1987: is unbiased under two assumptions:
1988: 1) the covariance of any pair of distinct observations
1989: vanishes on average; and
1990: 2) negligible signal leaks into the jackknife maps.
1991: % We will justify the first assumption below.
1992: The first assumption is equivalent to the statement that there is
1993: no scan-synchronous or fixed-pattern noise in the maps.
1994: We have checked this assumption empirically and find that the
1995: average fractional covariance of distinct observations is
1996: $\sim 2 \times 10^{-5}$, which is consistent
1997: with noise.
1998: We will discuss below how our non-detection of signal in the final
1999: map further justifies this assumption. With these assumptions, we can
2000: prove lack of bias of the noise estimate in a straightforward fashion
2001: by a simulation that obeys the assumptions.
2002: We generate
2003: % $N_{real}$ ensembles of simulated
2004: a set of $N_{obs}=515$ single-observation maps
2005: using the single-observation noise PSDs; these obviously have no
2006: signal and are uncorrelated with one another.
2007: % There will be
2008: % a total of $N_{real} \times N_{obs}$ single-observation maps.
2009: % We may coadd these maps to obtain $N_{real}$ simulated final maps,
2010: % which obviously have no signal and thus can be used to estimate the
2011: % simulated coadd map noise PSD.
2012: Then,
2013: % for any one of the realizations,
2014: we construct $N_{jack}=1000$
2015: simulated jackknife final maps and calculate
2016: the noise PSD of these maps, which also obviously
2017: have no signal.
2018: % We conducted this simulation for $N_{real}=1000$ and $N_{jack}=1000$,
2019: % identical to the number of jackknives used to estimate the
2020: % noise in our actual data.
2021: % We find that
2022: Next, we average these noise PSD estimates
2023: over all of the jackknife final
2024: maps, and divide by the input noise spectrum to determine how
2025: accurately we have recovered that input spectrum.
2026: The average (over all Fourier space pixels) of this normalized PSD
2027: is $0.9997 \pm 0.0006$, showing that we indeed recover the
2028: input noise spectrum within the measurement uncertainty of
2029: our simulation.
2030: % The fractional statistical precision of the
2031: % noise PSD estimates are 0.055 and 0.032, respectively, for
2032: % our simulation with $N_{jack}=1000$ and $N_{real}=1000$.
2033: % $\sqrt{2/N_{real}}$ and $\sqrt{2/N_{jack}}$,
2034: % respectively, which are X\% and Y\% for our simulation.
2035: % We find that the two noise PSDs estimated in this way are identical
2036: % to within this statistical precision.
2037: This exercise thus shows that the simulated final map noise PSD
2038: estimated by jackknife maps is an unbiased estimate of the
2039: simulated final map noise PSD.
2040: Note that we do not claim that the
2041: noise PSD generated in this fashion is the noise PSD of our true
2042: final map; the simulation is
2043: intended only to show that the jackknife noise
2044: PSD estimate method is unbiased.
2045:
2046: % One can show that the noise estimate obtained from the jackknife
2047: % maps is unbiased with only one assumption: independent observations
2048: % (not jackknifes) are uncorrelated (no coherently adding or canceling
2049: % scan-synchronous or fixed-pattern noise).
2050: % Consider the following toy model, where an ensemble of
2051: % independent noise maps are generated using spatially white
2052: % noise for simplicity.
2053: % Each noise map can be thought of as a single observation map.
2054: % Jackknifed coadds of these noise maps can then be made
2055: % in the same way that the jackknifed coadds of our real
2056: % data are produced.
2057: % The resulting distribution of jackknife map variances has
2058: % a mean equal to the expected value based on the noise maps,
2059: % indicating that the noise estimate from the jackknife maps
2060: % is unbiased.
2061: % Additionally, the standard deviation
2062: % of the distribution of variances is equal to
2063: % the variance times $\sqrt{2/N_{pix}}$,
2064: % where $N_{pix}$ is the number of pixels in the map
2065: % (\emph{i.e.}, the number of independent spatial modes).
2066: % This is the same width one would expect from a set of
2067: % independent maps, and indicates that the jackknife map
2068: % variances are distributed according the underlying PDF
2069: % given by the noise.
2070: % For reference, we also note that
2071: % one can show that the correlation coefficient of a pair of jackknife
2072: % maps vanishes on average in spite of the fact that the same data are
2073: % used to make the two jackknife maps.
2074:
2075: % One might worry that, because the jackknife maps make use of the
2076: % same data, the width of the distribution of A_sim will be wider than
2077: % if completely independent simulated maps were used. One can see this
2078: % worry is unfounded using a toy simulation. Our measurement is
2079: % conceptually equivalent to testing a map for an excess variance. If
2080: % one generates an ensemble of independent noise maps to be coadded
2081: % (using spatially white noise for simplicity) and then generates from
2082: % that set of noise maps an ensemble of jackknifes in the same way we
2083: % do and measures the variances of the jackknife maps, one finds that
2084: % the distribution of jackknife map variances has mean equal to the
2085: % expected variance of the coadded map (which is the single-observation
2086: % variance divided by the number of observations) and standard
2087: % deviation equal to var sqrt(2/N_pix) where N_pix is the number of
2088: % pixels in the map (the number of independent spatial modes). The
2089: % mean and
2090:
2091: % Additionally,
2092: % let us consider the issue of signal leakage into the jackknife
2093: % final maps used to estimate the expected noise in our final map.
2094: Let us now justify the assumption that negligible signal leaks
2095: into the jackknife final maps.
2096: % Note that
2097: In generating jackknife final maps, negative signs were applied to
2098: exactly one half of the observations.
2099: There are no fluctuations allowed in the number of negative signs,
2100: only in which observations have them applied.
2101: Therefore,
2102: residual signal can arise in the jackknife final maps in only two ways:
2103: 1) if the relative
2104: calibrations of the different observations are imperfectly known or
2105: 2) if the weights of the
2106: different observations are unequal.
2107:
2108: In the first case, consider a single-observation fractional relative
2109: calibration error of $\psi$, but assume all the component observations
2110: would otherwise be weighted equally
2111: (\emph{i.e.}, no variation in noise between
2112: observations).
2113: This fluctuating relative calibration error does not
2114: cause a bias;
2115: in an ensemble of experiments, the final map has an
2116: expected signal
2117: value equal to the
2118: signal value of the true final map and the jackknife final
2119: map has an expected signal value of exactly zero.
2120: But, the calibration fluctuations
2121: will cause an imperfect coaddition or cancellation
2122: of the signal
2123: in any given final map or jackknife final map realization,
2124: which will produce a fractional spread in the signal level
2125: of $\psi/\sqrt{N_{obs}}$ relative to the true signal.
2126: Given that $\psi \lesssim 3$\%\footnote{
2127: In Section~\ref{sec:flux_cal} we calculated our flux
2128: calibration uncertainty to be approximately 5.5\%.
2129: However, most of this uncertainty is due to systematics
2130: that will not change from one observation to the next.
2131: The uncertainty caused by fluctuations in the atmospheric
2132: opacity and the fit of our model are $\simeq 2-3$\%.}
2133: and $N_{obs} \simeq 500$ observations per map, this error
2134: in both the final map and jackknife final maps is very small compared
2135: to the signal. Since we are not attempting a high signal-to-noise
2136: measurement, the vanishingly small size of the error relative to
2137: the signal is thus not a concern in the final map. The error
2138: affects the jackknifes in a more subtle way because it effectively
2139: adds noise to the jackknife final maps,
2140: which means our noise estimate is slightly higher than the
2141: true noise level of our final map.
2142: However, this bias is negligible: the magnitude of the error
2143: is of order the signal times $\psi/\sqrt{N_{obs}}$.
2144: We know $\psi \lesssim 3$\%, $N_{obs} \simeq 500$,
2145: and the signal is less than 2\% of the noise in the final maps
2146: (in RMS units, c.f. Section~\ref{sec:mapmaking}),
2147: so this bias is $< 0.003$\%
2148: of the final map noise level in RMS units, or
2149: $< 0.006$\% in variance units. The small size of the effect is
2150: not surprising: it is proportional to the signal size, and we have
2151: no detection of signal. Moreover, even if the effect were not
2152: negligible, it would result in an overestimate of the noise PSD and
2153: thus would result in an overly conservative upper limit.
2154: This kind of effect would only be problematic if signal were
2155: visible at high significance.
2156: % This is not a problem because: 1)
2157: % it makes our noise estimate more conservative, essentially making it
2158: % more difficult to claim detection of a signal; and 2) the magnitude
2159: % of the error is of order the signal times $\psi/\sqrt{N_{obs}}$,
2160: % which, since
2161: % $\psi \lesssim 3$\% and
2162: % $N_{obs} \simeq 500$, is very small compared to any purported
2163: % signal. The conservativeness of the noise estimate is thus very
2164: % small compared to the signal and hence is not a cause for concern
2165: % given that no signal is seen. It would become important if we were
2166: % attempting to measure the amplitude of a signal with high fractional
2167: % precision.
2168:
2169: In the second case, the argument is very similar, but now what
2170: matters is the fractional variation in observation weight.
2171: This fractional variation is large, $\lesssim 2$,
2172: due to the significant
2173: differences in atmospheric noise between observations.
2174: Here, the large number of
2175: observations and the fact that the spread is proportional to the
2176: signal size render the effect negligible.
2177: The RMS spread of the
2178: residual signal in the jackknifes will be
2179: $\lesssim 2/\sqrt{N_{obs}} = 0.09$ times the
2180: signal size.
2181: Again, because of the small size of the error relative
2182: to the signal and the lack of detected signal,
2183: the error due to this
2184: effect is insignificant.
2185:
2186: If the noise estimation approach has underestimated the noise
2187: % by
2188: (for example, by failing to account for non-stationarity
2189: or correlations, or in any other manner) then there
2190: will be more noise in the final map than expected from the
2191: jackknifes.
2192: % This added noise would be the signature of anisotropy that we look
2193: % for, and thus we would be forced to claim a detection.
2194: However, because we find in Section~\ref{sec:CMB_results} that
2195: our 90\% CL interval on the amplitude of astronomical anisotropy,
2196: $\hat{A}$, includes $\hat{A} = 0$, we do not see
2197: any significant excess of noise in the true map above what is
2198: expected from the noise PSD estimate.
2199: % final map is consistent with the jackknife noise
2200: % realizations.
2201: % This explicitly shows that the final map contains no
2202: % such unaccounted-for noise sources at the level of interest for this
2203: % analysis.
2204: This explicitly rules out scan-synchronous or
2205: fixed-pattern
2206: noise that would be averaged away in jackknife maps but would
2207: remain in the coadded map at the level of interest for this
2208: analysis.
2209: % There may be such features at a much lower level.
2210: Had there been an excess above the expectation from the noise
2211: estimate, we would have had to show more explicitly that the
2212: noise estimate was correct in order to claim a detection.
2213:
2214: Alternatively, we consider the effect of overestimating the noise.
2215: The resulting final map PSD would be too low to be consistent
2216: with the noise PSD estimate. In our analysis
2217: (see Section~\ref{sec:sci_anal}),
2218: this would yield a best fit value for the astronomical anisotropy
2219: $\hat{A}$ of zero. We do not find
2220: this to be true: since the best-fit value of $\hat{A}$ must
2221: lie inside the confidence interval of any value,
2222: and our 68\% CL interval on $\hat{A}$ does
2223: not include $\hat{A} = 0$
2224: (see Table~\ref{tab:SZE_result}), the best-fit value of
2225: $\hat{A}$ must therefore differ from zero.
2226:
2227: % Alternatively, we also
2228: % consider what would happen if the noise estimation
2229: % approach somehow overestimates the noise (though it is hard to think
2230: % of how this might occur, it would have to be fixed pattern noise that
2231: % somehow acquires a sign flip between observations).
2232: % In that case, the
2233: % final map would have less noise than it ought to.
2234: % Again, though,
2235: % there is no evidence for such an effect since our final maps
2236: % are consistent with the jackknife noise realizations.
2237:
2238: % It may amuse the reader to note that these kinds of effects are far
2239: % less problematic in the low or vanishing signal-to-noise regime. Only
2240: % if one detects noise above what is expected from the jackknifes is
2241: % one forced to demonstrate that it is not spurious!
2242: %end Sunils response
2243:
2244: \subsection{Astronomical Signal Attenuation}
2245: \label{sec:sig_att}
2246:
2247: Now that the noise properties of the maps are well
2248: described, we need to determine the amount of astronomical
2249: signal attenuation due to data processing, the Bolocam
2250: beam, and the regularizing factor in Equation~\ref{eqn:regularizing}.
2251: The method for calculating the transfer
2252: function of the data processing and regularizing factor
2253: is analogous to the method described in Section~\ref{sec:xfer}
2254: for single observations.
2255: % First, a simulated map of the expected anisotropy signal is generated,
2256: % then reverse-mapped and added into the time-stream for every observation
2257: % that is co-added into $m'$.
2258: % Next, these time-streams are processed and mapped in the same
2259: % way that the original data were processed and mapped,
2260: % including application of the regularizing factor\footnote{
2261: % Since the single observation PSDs, transfer functions, and
2262: % beam profiles are the same for the original files and the
2263: % files with simulated anisotropy signal, the regularizing factor
2264: % will be the same for both files.
2265: % Therefore, the transfer function computed in the end will
2266: % account for the effects of the regularizing factor.}.
2267: % The map made from the unmodified time-streams is then subtracted
2268: % from the map made from the time-streams that include simulated signal,
2269: % and this map is compared to the original
2270: % simulated anisotropy map.
2271: % Since the beam profile is the same for every observation,
2272: % the beam profiles, $B_i$s, will cancel out in
2273: % Equations~\ref{eqn:opt_sig_map},~\ref{eqn:regularizing},~and~\ref{eqn:regularizing2}.
2274: % Therefore, the beam profile in the final maps is the same
2275: % as the beam profile used for single observation maps.
2276: Contour plots of the total astronomical signal attenuation
2277: are given in Figure~\ref{fig:final_xfer}.
2278: Compared to a single observation, the transfer functions
2279: for the final maps are much closer to being
2280: rotationally symmetric.
2281: The difference
2282: in the transfer functions
2283: is at low spatial frequencies parallel to either
2284: RA or dec, and is caused by adding observations made while
2285: scanning in perpendicular directions.
2286: This is because the modes in single observation
2287: maps, where there is
2288: a large amount of astronomical signal
2289: attenuation (\emph{i.e.},
2290: at low frequency parallel to the scan direction),
2291: do not contribute much to the final map.
2292: Therefore, most of the signal at low frequency along the
2293: RA direction is obtained from maps made while scanning parallel
2294: to dec, and vice versa.
2295: This effect can be seen by comparing the plots in
2296: Figure~\ref{fig:RA_xfer} with
2297: the plots in Figure~\ref{fig:final_xfer}.
2298:
2299: \section{Noise from Astronomical Sources}
2300:
2301: \label{sec:astr_noise}
2302:
2303: Since the noise PSD of the final map is estimated from
2304: jackknifed realizations of the data, all of the
2305: astronomical signal will be absent from the noise PSD.
2306: This is fine for the anisotropy signal we are looking for,
2307: because we want to understand the noise of our system
2308: in the absence of our signal of interest.
2309: However, we need to estimate
2310: the amount of noise produced by sources other than the
2311: SZE-induced CMB anisotropies,
2312: including galactic dust emission, radio point-source
2313: emission, emission from dusty submillimeter galaxies,
2314: and primary CMB anisotropies.
2315:
2316: The amount of galactic dust emission can be estimated
2317: from maps of our science fields taken from
2318: the full-sky 100~$\mu$m DIRBE/IRAS dust
2319: map \citep{dirbe_website, schlegel98}.
2320: To extrapolate the 100~$\mu$m data to our band at
2321: 143~GHz~$\simeq$~2.1~mm, we have used the
2322: ``model 8'' extrapolation given in
2323: \citet{finkbeiner99}.
2324: % This extrapolation is known to underestimate the
2325: % dust emission below $\simeq 500$~GHz \citep{finkbeiner99}, but it will
2326: % provide an order of magnitude estimate of the
2327: % signal in our band.
2328: At 100~$\mu$m, the typical surface brightness of the dust
2329: emission in our science fields
2330: is just over 1~MJy/ster, which corresponds to
2331: a surface brightness of around 5 -- 15~nK$_{CMB}$ for Bolocam.
2332: Using the maps that have been converted to a
2333: thermodynamic temperature at
2334: 143~GHz, we determined the map-space PSD of the
2335: dust emission, which corresponds to
2336: a $\mathcal{C}_{\ell}$ less than $10^{-6}$~$\mu$K$_{CMB}^2$
2337: for $\ell \gtrsim 1000$.\footnote{
2338: Note that the resolution of the DIRBE/IRAS dust map
2339: is 6.1~arcminutes, which corresponds to
2340: HWHM in $\ell$-space of $\lesssim 2000$.
2341: Therefore, we have no direct knowledge of the power
2342: spectrum on scales smaller than $\simeq 6$~arcminutes,
2343: which are the angular scales Bolocam is most
2344: sensitive to.
2345: However, the power spectrum of the dust falls rapidly
2346: at small angular scales, so the estimate at $\ell < 2000$
2347: should provide a reasonable upper limit.}
2348: % for the expected power spectrum.}
2349: % at the scales
2350: % Bolocam is most sensitive to.}
2351: Since this is well below the expected SZE-induced CMB anisotropy
2352: we are looking for, it is safe to conclude that the signal
2353: from the dust emission in our maps is negligible.
2354:
2355: Emission from radio point sources will also contribute to
2356: the astronomical signal in our maps.
2357: The power spectrum from these sources can be calculated
2358: from
2359: \begin{equation}
2360: C_{\ell} = \int_0^{S_{cut}} S^2 N(S) dS + w_{\ell} I^2,
2361: \label{eqn:ptsrc_ps}
2362: \end{equation}
2363: where $S$ is the flux of the source, $N(S)$ is the differential
2364: number of sources at a given flux in a given solid angle,
2365: $S_{cut}$ is an estimate of the source-detection
2366: threshold in the map (\emph{i.e.}, the level at which
2367: sources may be detected and removed),
2368: $C_{\ell}$ is the angular power spectrum,
2369: $w_{\ell}$ is the Legendre transform of the two-point correlation
2370: function of the sources, and
2371: \begin{displaymath}
2372: I = \int_0^{S_{cut}} S N(S) dS
2373: \end{displaymath}
2374: is the background contributed by the
2375: sources \citep{white04, scott99}.
2376: We will assume $w_{\ell} = 0$, since there is a large amount of
2377: uncertainty in the clustering of these sources.\footnote{
2378: Note that the total number of sources and total integrated
2379: power in $\ell$-space will not change if $w_{\ell}$ is
2380: non-zero; the clustering modeled by $w_{\ell}$
2381: will only shift power from high-$\ell$ to low-$\ell$.}
2382: Differential number counts have been determined from measurements
2383: at 1.4, 5, and 8.44~GHz \citep{toffolatti98, danese87}, with
2384: \begin{equation}
2385: N(S)_{5 {\rm GHz}} = 150 \textrm{ } S^{-2.5} \textrm{ }
2386: {\rm Jy}^{-1} {\rm ster}^{-1}.
2387: \label{eqn:ns_cm}
2388: \end{equation}
2389: Since the spectrum of the sources is nearly flat
2390: (i.e., $S_{\nu} \propto \nu^{\beta}$ with $\beta = 0$),
2391: this equation is valid over a
2392: wide range of frequencies.
2393: Additionally, the WMAP K, Ka, and Q bands have been used to
2394: determine the differential number counts at 22, 30,
2395: and 40~GHz \citep{bennett03}.
2396: $N(S)$ is similar for all three WMAP bands, and is
2397: $\lesssim 70$\% of the value of the model in Equation~\ref{eqn:ns_cm}.
2398: The differential number counts at 40~GHz are described by
2399: \begin{displaymath}
2400: N(S)_{40{\rm GHz}} = 32 \textrm{ } S^{-2.7} \textrm{ }
2401: {\rm Jy}^{-1} {\rm ster}^{-1}.
2402: \end{displaymath}
2403: To extrapolate this equation to the Bolocam band center at
2404: 143~GHz, we will use the method described in \citet{white04}.
2405: Since there is evidence of the power law for $N(S)$ flattening
2406: out at higher frequencies, they describe the differential
2407: number counts according to\footnote{
2408: There is some uncertainty in the spectrum of $S_{\nu}$
2409: for these radio sources between 40~GHz and 143~GHz.
2410: White and Majumdar quote two spectra, one with
2411: $\beta = 0$, and one with $\beta = -0.3$.
2412: This uncertainty in the spectrum of the radio point sources
2413: results in a finite range for the normalization of the
2414: number counts after extrapolating to 143~GHz.}
2415: \begin{equation}
2416: N(S)_{143 {\rm GHz}} = (20-32) \textrm{ } S^{-2.3} \textrm{ }
2417: {\rm Jy}^{-1} {\rm ster}^{-1}.
2418: \label{eqn:ns_150}
2419: \end{equation}
2420:
2421: We also need to estimate $S_{cut}$ in order to evaluate the
2422: power spectrum in Equation~\ref{eqn:ptsrc_ps}.
2423: This cutoff flux will necessarily be somewhat arbitrary, but, since
2424: $C_{\ell}$ is only weakly dependent on $S_{cut}$, it will not
2425: significantly alter our result.
2426: We have chosen $S_{cut} = 10$~mJy, which is approximately four
2427: times the RMS fluctuations per beam in maps made from
2428: data that have been optimally filtered for point sources.\footnote{
2429: From Equation~\ref{eqn:ns_150}, we only expect $1-2$ sources
2430: brighter than 10~mJy in our entire survey of 1 square
2431: degree, which is why we have not attempted to subtract
2432: out any sources prior to our anisotropy analysis.
2433: Additionally, the largest excursions in our maps are
2434: $\simeq 10$~mJy, further justifying our choice to set
2435: $S_{cut} = 10$~mJy.}
2436: Inserting this value of $S_{cut}$ into Equation~\ref{eqn:ptsrc_ps},
2437: along with Equation~\ref{eqn:ns_150}, yields
2438: $C_{\ell} \simeq 1.1-1.9$~Jy$^2$~ster$^{-1}$, or
2439: $C_{\ell} \simeq 7-12 \times 10^{-6}$~$\mu$K$_{CMB}^2$.
2440: To compare this angular power spectrum to the expected SZE-induced
2441: CMB anisotropies, we determine the amplitude
2442: of a flat band power, $\mathcal{C}^{\rm eff}_{\ell}$,
2443: that is required
2444: to cause the same temperature fluctuation as
2445: $C_{\ell}$ given our transfer function, $T_{\ell} B_{\ell}^2$,
2446: % we compare the temperature fluctuation
2447: % caused by $C_{\ell}$ to the temperature fluctuation
2448: % caused by a constant band power in $\mathcal{C}_{\ell}$ given
2449: % our full transfer function, $W_{\ell} B_{\ell}$.
2450: % For this comparison, we compute the value of a constant
2451: % $\mathcal{C}_{\ell}^{\rm eff}$ that would produce the same temperature
2452: % fluctuation as $C_{\ell}$, given by
2453: according to
2454: \begin{equation}
2455: \mathcal{C}_{\ell}^{\rm eff} = \frac{\sum_{\ell}
2456: C_{\ell}
2457: \frac{2\ell + 1}{4 \pi}
2458: T_{\ell} B_{\ell}^2}{
2459: \sum_{\ell} \frac{2\pi}{\ell(\ell+1)} \frac{2\ell + 1}{4 \pi}
2460: T_{\ell}B_{\ell}^2}.
2461: \label{eqn:c_l_eff}
2462: \end{equation}
2463: For the radio point sources with
2464: $C_{\ell} = 7-12 \times 10^{-6}$~$\mu$K$_{CMB}^2$,
2465: the effective $\mathcal{C}_{\ell}$ given the Bolocam
2466: transfer function is
2467: $\mathcal{C}_{\ell}^{\rm eff} \simeq 35-60$~$\mu$K$_{CMB}^2$,
2468: which is comparable to the expected signal from the
2469: SZE-induced CMB anisotropies.
2470:
2471: Additionally, emission from dusty submillimeter galaxies will
2472: be present in our maps.
2473: The same method used to determine the power spectrum from
2474: radio point sources can also be used to
2475: estimate the power spectrum of these sources.
2476: % To calculate the differential number counts we used
2477: We used the number counts distribution
2478: determined by \citet{aguirre08}, with
2479: \begin{displaymath}
2480: N(S)_{268 {\rm GHz}} = 1619 \textrm{ } S^{-2.26} e^{-303 S}
2481: \textrm{ }
2482: {\rm Jy}^{-1} {\rm ster}^{-1}.
2483: % \label{eqn:ns_268}
2484: \end{displaymath}
2485: The spectrum of these objects can be described by
2486: $S_{\nu} \propto \nu^{\beta}$, where
2487: $2.5 \lesssim \beta \lesssim 3.5$
2488: \citep{borys03},
2489: % To be conservative, we will chose $\beta =3$,
2490: which gives a differential number count at 143~GHz of
2491: \begin{displaymath}
2492: N(S)_{143 {\rm GHz}} = (100-220) \textrm{ } S^{-2.26}
2493: e^{-(2730-1460) S} \textrm{ }
2494: {\rm Jy}^{-1} {\rm ster}^{-1}.
2495: \end{displaymath}
2496: Inserting the above formula into Equation~\ref{eqn:ptsrc_ps}
2497: gives $C_{\ell} = 0.4-1.2$~Jy$^2$~ster$^{-1}$, or
2498: $C_{\ell} = 3 - 9 \times 10^{-6}$~$\mu$K$_{CMB}^2$.
2499: Equation~\ref{eqn:c_l_eff} can again be used to convert
2500: this to an effective constant $\mathcal{C}_{\ell}$ for
2501: our transfer function, giving
2502: $\mathcal{C}_{\ell}^{\rm eff} \simeq 15-45$~$\mu$K$_{CMB}^2$.
2503: Alternatively, we can compute a power spectrum
2504: using the differential number
2505: counts derived from SHADES data at 350~GHz \citep{coppin06},
2506: which is described by
2507: \begin{displaymath}
2508: N(S)_{350 {\rm GHz}} = 2.2 \times 10^4 \textrm{ }
2509: \left[S^2 + (5.9 \times 10^7) S^{5.8} \right]^{-1} \textrm{ }
2510: {\rm Jy}^{-1} {\rm ster}^{-1}.
2511: \end{displaymath}
2512: Converting this $N(S)$ to a differential number count at 143~GHz
2513: using the average spectrum of $\nu^3$
2514: % spectrum used
2515: % to interpolate Equation~\ref{eqn:ns_268}
2516: yields a similar
2517: power spectrum,
2518: with $C_{\ell} = 1.0$~Jy$^2$~ster$^{-1}$, or
2519: $C_{\ell} = 8 \times 10^{-6}$~$\mu$K$_{CMB}^2$,
2520: which is consistent with the result from the number
2521: counts given by \citet{aguirre08}.
2522:
2523: % Since there is a wide range of uncertainty in the power
2524: % spectrum for both the radio and submillimeter point
2525: % sources, we have not attempted to correct for this
2526: % contamination in our anisotropy amplitude estimates.
2527: % This means that the upper limits we find for
2528: % the anisotropy amplitude will be conservative.
2529:
2530: Finally, there will also be a signal in our map due to
2531: the primary CMB anisotropies,
2532: which are distinct from the SZE-induced anisotropies
2533: we are searching for.
2534: % that we are looking for.
2535: The power spectrum of the primary CMB anisotropies
2536: has been measured to high precision by
2537: WMAP at $\ell \lesssim 800$ \citep{nolta08},
2538: % by BOOMERANG at $500 \lesssim \ell \lesssim 1100$ \citep{jones06},
2539: and by ACBAR at $500 \lesssim \ell \lesssim 2500$
2540: \citep{reichardt08}.
2541: This measured power spectrum is well fit by theory, with
2542: only a small number of free parameters.
2543: Therefore, we have generated a template of the primary
2544: CMB power spectrum using the theoretical prediction generated
2545: by CMBFAST \citep{seljak96, zaldarriaga98,
2546: zaldarriaga00},
2547: with the best fit values to the free parameters from
2548: the WMAP 5-year data \citep{dunkley08}.
2549: % , BOOMERANG, ACBAR, and other CMB measurements\footnote{
2550: % The best fit values are:
2551: % primordial helium fraction, $Y_{He}$ = 0.248;
2552: % baryon fraction, $\Omega_{b}$ = 0.0422;
2553: % cold dark matter fraction, $\Omega_{CDM}$ = 0.203;
2554: % dark energy fraction, $\Omega_{\lambda}$ = 0.76;
2555: % Hubble constant, $H_0$ = 73~km~sec$^{-1}$~Mpc$^{-1}$;
2556: % number of effective neutrino species, $N_{\nu}$ = 3.29;
2557: % and an optical depth to the surface of last
2558: % scattering of $\tau$ = 0.09.} \citep{kuo06, spergel07}.
2559: Since the CMBFAST routine only computes the power spectrum
2560: up to $\ell = 3000$, we fit a decaying exponential
2561: to the $\mathcal{C}_{\ell}$ versus $\ell$ to extrapolate
2562: the primary CMB power spectrum to higher $\ell$.
2563: We can again use Equation~\ref{eqn:c_l_eff} to convert
2564: this power spectrum to an effective constant $\mathcal{C}_{\ell}$
2565: given our transfer function,
2566: with $\mathcal{C}_{\ell}^{\rm eff} \simeq 45$~$\mu$K$_{CMB}^2$.
2567: This band power is similar to what is expected
2568: from the SZE-induced CMB anisotropies.
2569: % However, in contrast to the radio and submillimeter point sources,
2570: % the primary CMB power spectrum is precisely known,
2571: % and can be accounted for in our analysis.
2572: A summary of the expected signal from the various
2573: astronomical sources is given in Figure~\ref{fig:astr_noise}.
2574:
2575: \section{Science Analysis}
2576: \label{sec:sci_anal}
2577:
2578: % \subsection{Procedure}
2579: % \label{sec:proceedure}
2580:
2581: \subsection{Overview of Analyses}
2582: \label{sec:overview_analysis}
2583:
2584: In addition to instrumental noise from the
2585: bolometers, electronics, etc.,
2586: our maps will contain an excess noise from astronomical sources,
2587: including
2588: anisotropies due to primary CMB fluctuations, fluctuations due to the
2589: SZE, and fluctuations due to unresolved astronomical point sources.
2590: %will result in our maps having more noise than expected from
2591: %non-astronomical sources;
2592: It is our goal to constrain the level of
2593: these
2594: %possible excess,
2595: astronomically sourced noises, which we will
2596: specify as the amplitude of flat band power anisotropy power spectrum
2597: contributions
2598: in $\mathcal{C}_\ell$. To obtain such a constraint, we must calculate
2599: the difference between the observed and expected power spectra of our
2600: maps and obtain a best estimate of the excess noise, goodness-of-fit
2601: of the data to the model, including any possible excess noise, and
2602: confidence intervals for the amount of excess noise. This section
2603: describes how we obtain the estimate and intervals.
2604:
2605: The first analysis we perform will simply constrain the total
2606: astronomical anisotropy in the maps, without any interpretation of the
2607: source, assuming only that the astronomical noise has a spectral shape
2608: flat in $\mathcal{C}_\ell$.
2609:
2610: The second analysis will statistically subtract the primary CMB
2611: anisotropy power spectrum by using the precise constraints
2612: placed on it by a
2613: variety of measurements \citep{reichardt08, nolta08}.
2614: The result will be a constraint
2615: on the non-primary-CMB contributions to anisotropy, and will be mildly
2616: more sensitive because of the subtraction.
2617: We will do this by
2618: adding the expected ``noise'' from the primary CMB to
2619: our model of the instrumental noise.
2620: %including in our expected PSD both non-astronomical noise as well as
2621: %the ``noise'' due to primary CMB.
2622: This expectation will fully take
2623: into account cosmic variance on the primary CMB anisotropy in a manner
2624: that we will explain below.
2625: %We shall see that this analysis yields an
2626: %upper limit.
2627:
2628: In the end, this analysis will yield an upper limit
2629: on the astronomical noise.
2630: Because it yields an upper limit, it is conservative to
2631: immediately interpret the constraint as a limit on SZE anisotropy: if
2632: there are point source contributions, as we expect there are, then the SZE
2633: contribution
2634: %can then only
2635: will be smaller than the upper limit we
2636: %will
2637: obtain by the assumption that the point source contributions are
2638: negligible. The situation would of course be different, and that
2639: assumption would not be conservative, were we claiming a detection of
2640: excess non-primary-CMB anisotropy.
2641:
2642: One could extend this methodology to statistical subtraction of the
2643: non-negligible submillimeter and radio point source contributions, but the
2644: large uncertainties in those contributions as well as the possibly
2645: unknown systematic uncertainties lead us to conclude that the
2646: improvement in sensitivity will be negligible and somewhat
2647: untrustworthy.
2648: %, so we do not perform such an analysis.
2649:
2650: \subsection{Deficiencies of a Bayesian Analysis}
2651: \label{sec:bayesian_dificiencies}
2652:
2653: We have chosen to model astronomical anisotropies using a flat
2654: band power in $\mathcal{C}_\ell$, which corresponds to $C_{\ell}
2655: = A S_{\vec{\nu}}^2$, for $\ell = 2 \pi |\vec{\nu}|$ and
2656: $S_{\vec{\nu}}^2 = 2 \pi / \ell(\ell+1)$.
2657: With these definitions,
2658: and assuming the noise PSD, $\mathcal{P}_{\vec{\nu}}$, fully
2659: describes the noise properties of the data for the reasons we
2660: have explained in Section~\ref{sec:mapmaking}, the best fit
2661: amplitude for an astronomical anisotropy signal is determined by
2662: maximizing Equation~\ref{eqn:log_l},
2663: \begin{displaymath}
2664: {\rm log}(\mathcal{L}) = \sum_{\vec{\nu} \epsilon V} \left(
2665: - {\rm log}
2666: ({\mathcal{P}_{\vec{\nu}} + A S_{\vec{\nu}}^2
2667: B_{\vec{\nu}}^2 T_{\vec{\nu}}})
2668: - \frac{X_{\vec{\nu}}}{
2669: {\mathcal{P}_{\vec{\nu}} + A S_{\vec{\nu}}^2
2670: B_{\vec{\nu}}^2 T_{\vec{\nu}}}}
2671: \right),
2672: \end{displaymath}
2673: with respect to $A$, where $X_{\vec{\nu}}$ is the measured PSD
2674: of the science field map in squared units,
2675: $T_{\vec{\nu}}$ is the transfer
2676: function of our data processing in squared units,
2677: and $B_{\vec{\nu}}$ is the
2678: profile of our beam.\footnote{
2679: Note that we are calculating the anisotropy amplitude for
2680: a single bin in $\ell$-space.
2681: However, the technique can be applied to multiple bins
2682: in $\ell$-space by windowing the appropriate terms
2683: in Equation~\ref{eqn:log_l}
2684: (\emph{i.e.}, if an $\ell$-space bin is described by the
2685: transfer function $\mathcal{T}_{\vec{\nu}}$, then
2686: $\mathcal{P}_{\vec{\nu}} \rightarrow \mathcal{T}_{\vec{\nu}}
2687: \mathcal{P}_{\vec{\nu}}$,
2688: $S_{\vec{\nu}}^2 \rightarrow \mathcal{T}_{\vec{\nu}}
2689: S^2_{\vec{\nu}}$,
2690: and $X_{\vec{\nu}}
2691: \rightarrow \mathcal{T}_{\vec{\nu}} X_{\vec{\nu}}$.)}
2692: Since our maps are real,
2693: $X_{\vec{\nu}} = X_{-\vec{\nu}}$, $\mathcal{P}_{\vec{\nu}} =
2694: \mathcal{P}_{-\vec{\nu}}$, etc., so the sum only
2695: includes half the
2696: $\vec{\nu}$-space pixels, denoted by the set $V$.
2697: For reference, a detailed
2698: derivation of the above equation is given in
2699: Appendix~\ref{sec:map_var}.
2700: Note that Equation~\ref{eqn:log_l}
2701: allows for $A < 0$. Although such values are not physical,
2702: fluctuations in the noise can cause the most likely value of $A$
2703: to be less than zero when the expected value of $A$ is small
2704: compared to the non-astronomical noise.
2705:
2706: The above expression is incorrect at some level because the
2707: $\vec{\nu}$-space pixels are slightly correlated,
2708: approximately
2709: 1 - 4\% for nearest-neighbor pairs of pixels and less than 1\%
2710: for all other pairs of pixels, while Equation~\ref{eqn:log_l}
2711: treats all Fourier modes as independent.
2712: % The correlations
2713: % between Fourier space modes are small and the correlation
2714: Note that the correlation
2715: function, $c_{\vec{\nu},\vec{\nu}'}$, is largely translation
2716: invariant ($c_{\vec{\nu},\vec{\nu}'} \approx c(\vec{\nu} -
2717: \vec{\nu}')$).
2718: This error due to pixel correlations raises
2719: three questions: 1) Does maximization of the likelihood given
2720: above result in an unbiased estimator of $A$? 2) Is this
2721: an approximately minimum variance estimator? and 3) Can we derive
2722: Bayesian credibility intervals on $A$ from it?
2723: We have demonstrated using simulations that
2724: Equation~\ref{eqn:log_l} remains an unbiased and
2725: approximately minimum variance
2726: estimator for $A$ in spite of these correlations, presumably
2727: because ignoring these fairly uniform correlations does not
2728: shift the peak of $\mathcal{L}$.
2729: See Table~\ref{tab:log_l}.
2730: However, the width of
2731: $\mathcal{L}$ is certainly dependent on these correlations: we
2732: are essentially over-counting the number of independent data
2733: points entering the likelihood and thus assuming more
2734: statistical power than we really have.
2735:
2736:
2737: We can make an approximate, unrigorous correction for the
2738: effective number of independent modes
2739: % We may determine the
2740: % number of independent modes
2741: by calculating
2742: \begin{displaymath}
2743: N_{\rm eff} = N_{\rm true} \left(
2744: \frac{1}{2 N_{\rm true}}
2745: \left( \sum_{\vec{\nu} \epsilon V} \sum_{\vec{\nu}' \epsilon V}
2746: c_{\vec{\nu},\vec{\nu}'} + \sum_{\vec{\nu} \epsilon V}
2747: c_{\vec{\nu},\vec{\nu}} \right)
2748: \right)^{-1},
2749: \end{displaymath}
2750: where $N_{\rm true}$ is the
2751: total number of $\vec{\nu}$-space pixels in $V$, $N_{\rm eff}$ is the
2752: effective number of $\vec{\nu}$-space pixels, and
2753: $c_{\vec{\nu},\vec{\nu}'}$ is the correlation between pixel
2754: $\vec{\nu}$ and pixel $\vec{\nu}'$.
2755: The factor of 1/2 inside the
2756: parentheses arises from the fact that we have
2757: double counted the correlations with our
2758: sums over $\vec{\nu}$ and $\vec{\nu}'$;
2759: the factor of $1/N_{\rm true}$ is a normalization factor.
2760: $c_{\vec{\nu},\vec{\nu}'}$
2761: is calculated from the Fourier transform of the map, $M$,
2762: according to
2763: \begin{displaymath}
2764: c_{\vec{\nu},\vec{\nu}'} = \left|
2765: \frac{\left< M_{\vec{\nu}}^{*} M_{\vec{\nu}'} \right>}{
2766: \left<|M_{\vec{\nu}}|\right> \left<|M_{\vec{\nu}'}|\right>} \right|,
2767: \end{displaymath}
2768: where the averages are taken over jackknife realizations.
2769: Equation~\ref{eqn:log_l} is then multiplied
2770: by $N_{\rm eff}$/$N_{\rm true}$
2771: to account for these correlations when
2772: calculating the Bayesian likelihood, with $N_{\rm eff}/N_{\rm
2773: true} \simeq 0.43$ for our data.
2774: % \footnote{ This factor of 2.3 is
2775: % due entirely to these correlations, and does not include the
2776: % factor of 2 due to the fact that the map is real.}.
2777: When
2778: $\ln(\mathcal{L})$ is exponentiated, this scaling factor will
2779: cause $\mathcal{L}$ to fall off less quickly than it would with
2780: $N_{\rm eff} = N_{\rm true}$, thereby increasing the width of
2781: $\mathcal{L}$.
2782:
2783: The lack of rigor behind the above correction implies
2784: that there will be statistical problems in placing
2785: constraints using the above likelihood function.
2786: Were the above likelihood function correct, we could
2787: use it to set a $\alpha$\% Bayesian credibility interval
2788: on the parameter $A$ by finding an interval
2789: $[A_1,A_2]$, $A_1, A_2 \ge 0$, such that
2790: \begin{displaymath}
2791: \frac{\alpha}{100} =
2792: \frac{\int_{A_1}^{A_2} dA\, \mathcal{L}(X_{\vec{\nu}}|A)}
2793: {\int_0^{\infty} dA\, \mathcal{L}(X_{\vec{\nu}}|A)}
2794: \end{displaymath}
2795: where we have assumed a flat prior $A \ge 0$
2796: % , the latter
2797: because
2798: of the non-physical nature of $A < 0$. If the likelihood function's
2799: width in $A$ is not to be trusted, then such credibility
2800: intervals are not valid. Not even a simulation permits
2801: one to set a credibility interval because $\mathcal{L}$ is
2802: simply not the correct likelihood, even if its distribution can
2803: be characterized by simulation.
2804: % Moreover, even if the
2805: % likelihood function were correct, the use of the prior $A \ge 0$
2806: % has no rigorous justification (in spite of the general use of
2807: % such priors when constraining parameters near such physical
2808: % boundaries).
2809:
2810: Additionally,
2811: determining the goodness-of-fit for the best-fit value of $A$,
2812: $\hat{A}$, will require simulation. That is, if
2813: Equation~\ref{eqn:log_l} was correct, then
2814: we should be able to determine an analytic expression
2815: for the distribution of $\ln(\mathcal{L})$ for $\hat{A}$ that would
2816: allow us to calculate the goodness-of-fit of the
2817: data to the model.
2818: But, since the above likelihood
2819: function is incorrect,
2820: we must simulate an ensemble of measurements,
2821: with appropriate correlations in the Fourier modes, to determine
2822: the distribution of $\ln(\mathcal{L})$ for the value
2823: $\hat{A}$.
2824: Therefore, the Bayesian approach offers no simplifications
2825: or reductions in computing time relative to the
2826: simulation-based frequentist technique we employ below.
2827:
2828: \subsection{Overview of Frequentist, Feldman-Cousins Analysis
2829: Technique}
2830:
2831: It is possible to deal with all of the above problems
2832: approximately correctly with a frequentist technique for
2833: establishing goodness-of-fit confidence levels and frequentist
2834: confidence (as opposed to Bayesian credibility) intervals on $A$
2835: that incorporate the prescriptions of Feldman and Cousins for
2836: dealing with a physical boundary~\citep{feldman98}. The
2837: technique has two main features: \begin{enumerate}
2838:
2839: \item First, we use jackknife maps with signal added based on an
2840: input value $A_{\rm sim}$ in the physically allowed region
2841: $A_{\rm sim} \ge 0$ to determine the distribution of
2842: $\mathcal{L}(X_{\vec{\nu},i}(A_{\rm sim})|A_{\rm sim})$ as
2843: defined in Equation~\ref{eqn:log_l} for an ensemble of
2844: experiments with outcomes $X_{\vec{\nu},i}(A_{\rm sim})$
2845: % (The
2846: % simulate outcomes are of course a function of the simulation
2847: % input parameter value $A_{\rm sim}$).
2848: With this distribution,
2849: we may determine whether $\mathcal{L}(X_{\vec{\nu}}|A_{\rm
2850: sim})$, the value of $\mathcal{L}$ for the true data and the
2851: value $A_{\rm sim}$, is among the $\alpha$\% most likely
2852: outcomes for that input value $A_{\rm sim}$, thereby determining
2853: a goodness-of-fit confidence level. In doing this, we make the
2854: reasonable approximation that, although $\mathcal{L}$ is not a
2855: rigorously correct likelihood, it maps in a one-to-one,
2856: monotonic fashion to the true likelihood function
2857: $\mathcal{L}_{\rm true}$. Specifically, if we consider two
2858: realizations $X_{\vec{\nu},1}$ and $X_{\vec{\nu},2}$, we assume
2859: that the sign of $\ln \mathcal{L}(X_{\vec{\nu},1}|A) - \ln
2860: \mathcal{L}(X_{\vec{\nu},2}|A)$ is the same as that of $\ln
2861: \mathcal{L_{\rm true}}(X_{\vec{\nu},1}|A) - \ln \mathcal{L_{\rm
2862: true}}(X_{\vec{\nu},2}|A)$. This assumption is far looser than
2863: the assumption that rescaling $\ln \mathcal{L}$ by $N_{\rm
2864: eff}/N_{\rm true}$ is correct; we are only assuming that the
2865: ordering of realizations in $\mathcal{L}$ and $\mathcal{L}_{\rm
2866: true}$ are the same, even if the numerical values are not the
2867: same.
2868:
2869: \item Second, we want to define a confidence interval of
2870: confidence level $\alpha$\% on $A$. Since we are taking a
2871: frequentist approach, these confidence intervals are defined to
2872: include the values of $A$ for which, if $A$ is the true value of
2873: the anisotropy amplitude, then the observed outcome
2874: $X_{\vec{\nu}}$ is within the $\alpha$\% most likely
2875: outcomes (as defined below, a definition that is different than
2876: the usual likelihood) for that value of $A$. We use the same
2877: set of simulations with the following procedure based on the
2878: Neyman construction as modified by \citet{feldman98}.
2879: We now calculate for each simulation
2880: realization $i$ for each input parameter value $A_{\rm sim}$ the
2881: ratio
2882: \begin{displaymath}
2883: R_i(A_{\rm sim}) = \frac{\mathcal{L}(X_{\vec{\nu},i}(A_{\rm
2884: sim})|A_{\rm sim})}
2885: {\mathcal{L}(X_{\vec{\nu},i}(A_{\rm sim})|\hat{A}_i)}
2886: \end{displaymath}
2887: where $A_{\rm sim}$ is the simulation input parameter value
2888: ($A_{\rm sim} \ge 0$) and $\hat{A}_i$ is the best-fit value of $A$
2889: {in the physically allowed region $A \ge 0$} for the given
2890: realization $i$. We order the realizations in order of
2891: decreasing $R_i(A_{\rm sim})$ until $\alpha$\% of the realizations
2892: have been included; the value of $R_i(A_{\rm sim})$ defining this
2893: boundary is denoted by $R_\alpha(A_{\rm sim})$. The input parameter
2894: value $A_{\rm sim}$ is then included in the $\alpha$\% confidence
2895: interval if the likelihood ratio for the real data, $R_{\rm
2896: data}(A_{\rm sim})$, is among the $\alpha$\% largest $R_i(A_{\rm sim})$
2897: values, $R_{\rm data}(A_{\rm sim}) \ge R_\alpha(A_{\rm sim})$. The
2898: interpretation is that, for values $A_{\rm sim}$ belonging to the
2899: confidence interval of confidence level $\alpha$\%, the data is
2900: among the $\alpha$\% most likely outcomes, where ``likely'' is
2901: quantified by $R_{\rm data}(A_{\rm sim})$ {instead of}
2902: $\mathcal{L}(X_{\vec{\nu},i}|A_{\rm sim})$.
2903:
2904: The above procedure can be conveniently visualized as follows.
2905: The simulations indicate that there is a smooth relationship
2906: between $R_i(A_{\rm sim})$ and $\hat{A}_i$ at a given value of
2907: $A_{\rm sim}$. This is generically true, not specific to this
2908: analysis. Therefore, each simulation realization may be
2909: labeled by its value of $\hat{A}$ and we may write $R(A_{\rm
2910: sim},\hat{A})$ in place of $R_i(A_{\rm sim})$. We may visualize
2911: $R(A_{\rm sim},\hat{A})$ as a function of $\hat{A}$ for a given
2912: value of $A_{\rm sim}$; the cutoff value $R_\alpha(A_{\rm sim})$
2913: is a horizontal line in this plot, and so points with $R(A_{\rm
2914: sim},\hat{A}) > R_\alpha(A_{\rm sim})$ map to a set of intervals
2915: in $\hat{A}$; in fact, in our case, there is a single interval
2916: for each $A_{\rm sim}$. These intervals, called {confidence
2917: belts}, can be displayed as intervals $[\hat{A}_1, \hat{A}_2]$
2918: at the given value $A_{\rm sim}$ in a plot of $A_{\rm sim}$
2919: vs. $\hat{A}$, as illustrated in
2920: Figure~\ref{fig:full_conf_belts}. Do not confuse
2921: confidence belts, which are intervals along the $\hat{A}$ axis,
2922: with confidence intervals, which are intervals along the $A_{\rm
2923: sim}$ axis as defined below.
2924:
2925: Then, to determine the confidence interval of confidence level
2926: $\alpha$\% on $A$ given a data set $X_{\vec{\nu}}$, one finds
2927: $\hat{A}_{\rm data}$, draws a vertical line on the plot of
2928: confidence belts at the value $\hat{A}_{\rm data}$ on the
2929: horizontal ($\hat{A}$) axis, and includes all values of
2930: $A_{\rm sim}$ for which the vertical line lies inside the confidence
2931: belt at that value of $A_{\rm sim}$. This confidence belt
2932: construction is equivalent to the above description based on
2933: $R(A_{\rm sim},\hat{A})$ because the smooth relationship between
2934: $R(A_{\rm sim},\hat{A})$ and $\hat{A}$ ensures that, for a given
2935: $A_{\rm sim}$, if $\hat{A}_{\rm data}$ is inside the confidence belt
2936: at a given value $A_{\rm sim}$, then $R_{\rm data} \ge
2937: R_{\alpha}(A_{\rm sim})$ for that value $A_{\rm sim}$.
2938:
2939: \end{enumerate}
2940:
2941: We comment on two important aspects of this construction of the
2942: confidence intervals. First is the ordering of the simulation
2943: realizations by $R$, not by $\mathcal{L}(X_{\vec{\nu},i}(A_{\rm
2944: sim})|A_{\rm sim})$. Feldman and Cousins discuss both possible
2945: constructions (the latter originally proposed by~\citet{crow59})
2946: and argue that the latter has a
2947: serious deficiency in that it ties the confidence level of the
2948: confidence interval to the goodness-of-fit confidence level;
2949: essentially, it is possible for the confidence interval to not
2950: provide the advertised frequentist coverage if the
2951: goodness-of-fit is poor. This typically happens when the
2952: experimental outcomes yield best-fit parameter values near or
2953: outside a physical boundary. In our application, this can occur
2954: if the simulation realization has a bit less anisotropy than
2955: expected, which would yield a best-fit $\hat{A}$ that is
2956: negative. In such a case, the (approximate) likelihood of the
2957: data set, $\mathcal{L}(X_{\vec{\nu}} | A_{\rm sim})$, will in
2958: general be small. However, the approximate likelihood of that
2959: data set may not be small compared to the approximate
2960: likelihood, $\mathcal{L}(X_{\vec{\nu}} | 0)$, of the most
2961: probable physically allowed alternative hypothesis of $\hat{A}_i
2962: = 0$. Feldman and Cousins show that, with this ordering
2963: principle, the confidence intervals never contain unphysical
2964: values for the observable. Additionally, there is a smooth
2965: transition from the case of an upper limit to a central
2966: confidence region, eliminating intervals that under-cover due to
2967: choosing between an upper limit and a central region based on
2968: the result. There is not room here to reproduce their arguments
2969: in detail, we refer the reader to \citet{feldman98}.
2970:
2971: The second important aspect is that the construction is done
2972: entirely by simulation so that the only way in which we depend
2973: on $\mathcal{L}$, which we know to be deficient, is in the
2974: ordering it provides. We have assumed above that, in spite of
2975: its inaccuracy, $\mathcal{L}$ provides the same ordering of
2976: points as $\mathcal{L}_{\rm true}$, and hence $\hat{A}$ and
2977: $R(A_{\rm sim},\hat{A})$ for a given (simulated or real) data
2978: realization and $R_\alpha(A_{\rm sim})$ for a given $A_{\rm
2979: sim}$ will be the same regardless of whether we use
2980: $\mathcal{L}$ or $\mathcal{L}_{\rm true}$.
2981:
2982: \subsection{Construction of Simulated Data Sets for
2983: Frequentist Technique}
2984:
2985: To apply this method to our data, we first create a simulated
2986: map of the astronomical anisotropy for a given value of the
2987: astronomical anisotropy amplitude, $A_{\rm sim}$, using our assumed
2988: profile $S_{\vec{\nu}}^2$. This simulation is produced by drawing
2989: a value for each pixel, $\vec{\nu}$, from an underlying Gaussian
2990: distribution,\footnote{
2991: We have also determined confidence intervals using
2992: non-Gaussian distributions for the SZE-induced anisotropy
2993: signal.
2994: The results are described in Section~\ref{sec:SZE_results}.}
2995: then multiplying it by $A_{\rm sim} S_{\vec{\nu}}^2$.
2996: The PSD of this simulated map is multiplied by our full transfer
2997: function and added to the jackknifed realization of our data,
2998: $X_{i,\vec{\nu}}$.\footnote{ The reason we add the simulated
2999: astronomical anisotropy map to the jackknifed realization map
3000: instead of the time-streams is to reduce the amount of
3001: computational time required. Since the transfer functions of
3002: the maps are well measured, there is no reason to
3003: go all the way back to the time-streams to add the simulated
3004: signal.}
3005: % simulated data directly to the time-streams.}
3006: Note that a
3007: different simulated map is created for each jackknifed
3008: realization of the data to allow for cosmic variance. Then,
3009: we use Equation~\ref{eqn:log_l} to determine the most likely
3010: value of the astronomical anisotropy amplitude, $\hat{A}_{i}$,
3011: for realization $i$. By using jackknifes of our actual data, we
3012: are including all of the correlations between pixels, and by
3013: simulating the astronomical anisotropy maps we are accounting
3014: for cosmic variance in the astronomical anisotropy. For a given
3015: value of $A_{\rm sim}$, we repeat this process for each jackknifed
3016: realization of the data.
3017:
3018: The data sets are then ordered based on the ratio of their
3019: likelihood to the likelihood of the most probable physically
3020: allowed outcome, $R_i(A_{\rm sim})$, as defined above. The procedure
3021: outlined in the previous section for defining
3022: $R_\alpha(A_{\rm sim})$, finding confidence belts for each
3023: $A_{\rm sim}$, and then determining a confidence interval in
3024: $A_{\rm sim}$ is then employed as described.
3025:
3026: To determine the goodness-of-fit of our data to the
3027: model given by $S_{\vec{\nu}}^2$, we compare the
3028: likelihood of the actual data at the best fit value
3029: of $\hat{A}_{data}$, $\mathcal{L}(X_{\vec{\nu}}|\hat{A}_{data})$,
3030: to the likelihoods of a set of jackknifed realizations
3031: of our data with simulated spectra added according to
3032: $S_{\vec{\nu}}^2$ with amplitude $\hat{A}_{data}$,
3033: $\mathcal{L}(X_{\vec{\nu},i}(\hat{A}_{data})|
3034: \hat{A}_{data})$.
3035: For the observations of the Lynx field
3036: $\mathcal{L}(X_{\vec{\nu}}|\hat{A}_{data})$ is greater than
3037: $\mathcal{L}(X_{\vec{\nu},i}(\hat{A}_{data})|
3038: \hat{A}_{data})$ for 17\% of the realizations,
3039: and for the observations of the SDS1 field
3040: $\mathcal{L}(X_{\vec{\nu}}|\hat{A}_{data})$ is greater than
3041: $\mathcal{L}(X_{\vec{\nu},i}(\hat{A}_{data})|
3042: \hat{A}_{data})$ for 43\% of the realizations.
3043: Therefore, we can conclude that our model provides
3044: an adequate description of the data.
3045:
3046: \subsection{Total Anisotropy Amplitude Results}
3047:
3048: \label{sec:CMB_results}
3049:
3050: To determine the confidence intervals for the full data set,
3051: we make a joint estimate of $A$ using both the
3052: Lynx and SDS1 data sets.
3053: % The calculation was performed according to the same
3054: % methods described in Section~\ref{sec:proceedure}
3055: % for the single field/season data sets.
3056: A plot of the Bayesian likelihood, along with
3057: confidence belts computed using the Feldman and
3058: Cousins method are given in Figure~\ref{fig:full_conf_belts}.
3059: Uncertainties in our pointing model have already been included
3060: in these calculations by an effective broadening of the
3061: Bolocam beam.
3062: % \footnote{
3063: % Uncertainties in the flux calibration have not been
3064: % included here due to the standard convention.
3065: % However, these uncertainties are included in our
3066: % estimates of the SZE-induced anisotropies described
3067: % in Section~\ref{sec:SZE_results}.}.
3068: Our upper limits on the total anisotropy amplitude are equal to
3069: 590, 760, and 830~$\mu$K$^2_{CMB}$ at confidence levels
3070: of 68\%, 90\%, and 95\%.
3071: Note that the uncertainty on these limits due to the finite
3072: number of simulations we have run is $\simeq 10-15$~$\mu$K$_{CMB}^2$.
3073: % A summary of this result is given in Table~\ref{tab:final_result}.
3074:
3075: To determine the effective angular scale of our
3076: anisotropy amplitude measurements we have computed
3077: our band power window function, $W^B_{\ell}/\ell$,\footnote{
3078: This band power window function is defined such that $
3079: % \begin{equation}
3080: \left< \mathcal{C}_B \right> = \sum_{\ell}
3081: (W^B_{\ell}/\ell) \mathcal{C}_{\ell}
3082: % \end{equation}
3083: $, where $\left< \mathcal{C}_B \right>$ is the
3084: experimental band power measurement for
3085: the power spectrum, $\mathcal{C}_{\ell}$.
3086: Note that the transfer function of our data
3087: processing, $T_{\ell}$,
3088: is not the same as the band power window function,
3089: $W_{\ell}^B/\ell$.}
3090: using
3091: the method given by \citet{knox99}.
3092: A plot of the peak-normalized band power window function
3093: for the full data set is given in Figure~\ref{fig:knox}.
3094: From this band power window function we have calculated
3095: an effective angular multipole for our data
3096: set, $\ell_{eff}$, given by
3097: \begin{displaymath}
3098: \ell_{eff} = \frac{\sum_{\ell} \ell (W^B_{\ell}/\ell)}
3099: {\sum_{\ell} W^B_{\ell}/\ell},
3100: \end{displaymath}
3101: and equal to 5700.
3102: Additionally, the full-width half-maximum of the window
3103: function, FWHM$_{\ell}$, is equal to 2800.
3104: A plot comparing our result to other measurements of the
3105: CMB on similar scales is shown in Figure~\ref{fig:cmb_power}.
3106:
3107: \subsection{SZE-Induced CMB Anisotropy Results and Constraints on
3108: $\sigma_8$}
3109: \label{sec:SZE_results}
3110:
3111: In order to determine the amplitude of the SZE-induced
3112: CMB power spectrum, we follow the same methods described
3113: above
3114: % in Section~\ref{sec:proceedure}
3115: to determine the total
3116: amplitude of the anisotropy power spectrum.
3117: However, we now have to statistically subtract the
3118: signal due to the primary CMB anisotropies by accounting
3119: for both the amplitude and fluctuations of its
3120: expected power spectrum in the likelihood;
3121: these primary CMB anisotropies are effectively
3122: an additional noise in the map.
3123: % Following the notation from Section~\ref{sec:proceedure},
3124: The noise contributed to the map from the Bolocam
3125: system is given by $\mathcal{P}_{\vec{\nu}}$.
3126: Since the spectrum of the primary anisotropies in the CMB
3127: is well understood,
3128: we can calculate the
3129: expected noise from the primary CMB anisotropies.
3130: % $\mathcal{P}_{\vec{\nu}}^{ \left[ CMB \right] }$.
3131: To calculate this noise
3132: % $\mathcal{P}_{\vec{\nu}}^{ \left[ CMB \right] }$
3133: we first create a simulated map of the primary CMB, assuming
3134: that the underlying distribution of
3135: $\vec{\nu}$-space pixel values is Gaussian.
3136: This simulation is produced by drawing a value for each
3137: pixel, $\vec{\nu}$, from an underlying Gaussian distribution,
3138: then multiplying it by the best fit primary CMB
3139: spectrum given in Section~\ref{sec:astr_noise}.
3140: The PSD of this map is then multiplied by $T_{\vec{\nu}} B_{\vec{\nu}}^2$
3141: and added to a jackknifed realization of our data, $X_{i,\vec{\nu}}$,
3142: to give $X_{i,\vec{\nu}}^{[SZE]}$.
3143: A different simulated map is generated for each
3144: jackknifed realization of the data to account for
3145: the cosmic variance in the CMB spectrum.
3146: These modified jackknifed realizations of the data
3147: are then be used to determine the expected PSD,
3148: $\mathcal{P}_{\vec{\nu}}^{ \left[ SZE \right] }$,
3149: for the noise contributed by the Bolocam system
3150: and the primary CMB anisotropies.
3151: We note that such a simulation of the primary CMB
3152: contribution is more correct than simply adding the
3153: primary CMB power spectrum to the non-astronomical noise
3154: power spectrum because it correctly reproduces
3155: pixelization and Fourier-mode correlation effects.
3156:
3157: Next, we select a model spectrum for the SZE anisotropies,
3158: $S_{\vec{\nu}}^{[SZE]}$.
3159: Using these new definitions, the
3160: Bayesian likelihood function in Equation~\ref{eqn:log_l}
3161: can be written as
3162: \begin{displaymath}
3163: {\rm log}(\mathcal{L}) = \sum_{\vec{\nu}} \left(
3164: - {\rm log}
3165: ({\mathcal{P}_{\vec{\nu}}^{[SZE]} +
3166: A^{[SZE]} (S_{\vec{\nu}}^{[SZE]})^2 B_{\vec{\nu}}^2 T_{\vec{\nu}}})
3167: - \frac{X_{\vec{\nu}}^{[SZE]}}{
3168: {\mathcal{P}_{\vec{\nu}}^{[SZE]} +
3169: A^{[SZE]}
3170: (S_{\vec{\nu}}^{[SZE]})^2 B_{\vec{\nu}}^2 T_{\vec{\nu}}}}
3171: \right),
3172: \end{displaymath}
3173: where $A^{[SZE]}$ is the amplitude of the SZE-induced
3174: CMB anisotropies, $B_{\vec{\nu}}^2$ is the profile of our beam,
3175: and $T_{\vec{\nu}}$
3176: is the transfer function of our data processing in squared units.
3177: % At this point we can proceed exactly as in
3178: % Section~\ref{sec:proceedure} to determine confidence intervals
3179: % for $A^{[SZE]}$.
3180: As before, we create simulated SZE maps with an amplitude
3181: $A_{sim}^{[SZE]}$, add these to our jackknifed realizations
3182: after multiplying by $(S_{\vec{\nu}}^{[SZE]})^2
3183: B^2_{\vec{\nu}} T_{\vec{\nu}}$,
3184: then use the ordering method developed by
3185: \citet{feldman98} to determine
3186: the width of the confidence belt at $A_{sim}^{[SZE]}$.
3187: By repeating this procedure for a range of physically
3188: allowed values of $A_{sim}^{[SZE]}$, we can construct
3189: a full confidence belt that can be used
3190: to determine our confidence intervals.
3191: We emphasize that, while we use the Bayesian likelihood to
3192: construct our best estimator for $A$, a procedure that
3193: we have already demonstrated by simulation is unbiased and
3194: approximately
3195: minimum variance, we in no way rely on the Bayesian likelihood
3196: to determine confidence intervals on $A$.
3197: The frequentist Feldman-Cousins method is used for the
3198: latter task.
3199:
3200: Additionally, we need to account for the flux calibration
3201: uncertainty.
3202: % \footnote{
3203: % The flux calibration uncertainty was not included in
3204: % our total anisotropy amplitude estimate because that
3205: % is the standard convention.}.
3206: The uncertainty in the flux calibration model derived from
3207: point sources is 5.5\%, and the uncertainty in the area
3208: of our beam is 3.1\%.
3209: Therefore, the uncertainty in our surface brightness
3210: calibration is 6.3\%.
3211: To determine the effect of this flux calibration error
3212: on our confidence intervals, we multiplied each
3213: simulated primary and SZE-induced CMB map by
3214: $\phi_i = 1 + y$, where $y$ is drawn from
3215: a Gaussian distribution with a standard deviation
3216: equal to our flux uncertainty of 0.063.
3217: A different $\phi_i$ was generated for each simulated
3218: CMB map.
3219: This means that each simulated map has a different
3220: flux calibration, distributed
3221: according to our uncertainty in the calibration.
3222: New confidence belts were then calculated using
3223: the same procedure described above.
3224: % in Section~\ref{sec:proceedure}.
3225: We have also determined the confidence intervals
3226: assuming that there is no uncertainty in the
3227: known flux of Uranus and Neptune
3228: (\emph{i.e.}, the only flux calibration uncertainties are due
3229: to our measurement errors and observational
3230: techniques).
3231: In this case, the flux calibration uncertainty is
3232: 3.5\% instead of 6.3\%.
3233:
3234: These flux calibration uncertainties produce
3235: non-negligible changes to the confidence intervals we determine
3236: for the anisotropy amplitude since it is a variance
3237: (\emph{i.e.}, it depends
3238: quadratically on the flux calibration).
3239: Therefore, for a simulated amplitude $A_{sim}$,
3240: a fractional flux calibration uncertainty of
3241: $\sigma_f$ will increase/decrease the upper/lower
3242: bounds of the 68\% CL confidence belt by an amount
3243: roughly proportional to $A_{sim} (1 + \sigma_f)^2$.
3244: The resulting fractional change to the
3245: confidence interval limits
3246: will in general be non-trivial, but should be
3247: approximately equal to $(1 + \sigma_f)^2$ for
3248: 68\% CL limits.
3249: So, for a 3.5\% flux calibration uncertainty
3250: we expect the 68/90/95\% CL upper limits to increase
3251: by approximately 7/12/14\% compared to the
3252: case of no flux calibration uncertainty.
3253: Similarly, for a 6.3\% flux calibration
3254: uncertainty we expect the 68/90/95\% CL upper
3255: limits to increase by approximately
3256: 13/22/27\% compared to the case of no
3257: flux calibration uncertainty.
3258: After fully simulating the effect of the flux calibration
3259: uncertainty on our upper limits, we find results
3260: that are comparable to the predictions given above.
3261: See Table~\ref{tab:SZE_result}.
3262:
3263: We have computed confidence intervals for two different
3264: SZE spectra: a flat
3265: spectrum, $(S_{\vec{\nu}}^{[SZE]})^2 = 2\pi/\ell(\ell+1)$
3266: for $\ell = 2\pi|\vec{\nu}|$ and
3267: the analytic spectrum calculated by \citet{komatsu02}.
3268: The results for both of these spectra are similar,
3269: which is reasonable since the analytic spectrum
3270: is nearly flat at the scales to which we are most
3271: sensitive ($4000 \lesssim \ell \lesssim 7000$).
3272: See Table~\ref{tab:SZE_result}.
3273: In addition to the analytic spectrum calculated
3274: by Komatsu and Seljak, several SZE power
3275: spectra have been determined via hydrodynamic simulations using
3276: either MMH (moving-mesh
3277: hydrodynamic)
3278: or SPH
3279: (smoothed-particle hydrodynamic) algorithms.
3280: Examples of MMH simulations can be found in
3281: \citet{zhang02}, \citet{seljak01},
3282: \citet{refregier00}, and \citet{refregier00_2}.
3283: Examples of SPH simulations can be found in
3284: \citet{dasilva01} and \citet{springel01}.
3285: Since most of the simulated SZE spectra
3286: are approximately flat at the
3287: angular scales we are most sensitive to, we have
3288: not determined confidence levels using any
3289: of these spectra.
3290: See Figure 1 in \citet{komatsu02}.
3291:
3292: Komatsu and Seljak determined that the amplitude
3293: of the SZE-induced CMB anisotropies scales
3294: according to $\sigma_8^7 (\Omega_b h)^2$
3295: and is relatively insensitive to all other
3296: cosmological parameters \citep{komatsu02}.
3297: Using the results from the WMAP 5-year data,
3298: % and other CMB experiments,
3299: the best fit values for $\sigma_8$,
3300: $\Omega_b$, and $h$ are
3301: 0.796, 0.0440, and 0.719 \citep{dunkley08}.
3302: These values produce a maximum SZE anisotropy
3303: amplitude of less than 10~$\mu$K$_{CMB}^2$
3304: at our band center of 143~GHz for the
3305: analytic Komatsu and Seljak spectrum.
3306: For comparison, the 90\% confidence level
3307: upper limit on the average value of the analytic spectrum
3308: weighted by the Bolocam transfer function
3309: is 950~$\mu$K$_{CMB}^2$,
3310: including our flux calibration error.
3311: See Table~\ref{tab:SZE_result}.
3312: Based on this upper limit,
3313: assuming the scaling relation given
3314: by Komatsu and Seljak and holding
3315: all other parameters fixed, the corresponding
3316: 90\% confidence level upper limit on the
3317: three cosmological parameters is
3318: $\sigma_8^7 (\Omega_b h)^2 < 2.13$.
3319: Individually, the best constraint can be placed
3320: on $\sigma_8$ since the amplitude depends most
3321: strongly on this parameter, with
3322: $\sigma_8 < 1.55$ at a confidence level
3323: of 90\%.
3324:
3325: However, this upper limit has been derived by assuming
3326: the SZE-induced anisotropy signal is Gaussian, which
3327: is a poor assumption.
3328: To account for the non-Gaussianity of the signal, we
3329: have used a method similar to the one described by
3330: \citet{goldstein03} to analyze data collected with
3331: ACBAR.
3332: Based on the results of numerical simulations by
3333: \citet{white02} and \citet{zhang02}, along with
3334: calculations of the trispectrum term from
3335: \citet{cooray01} and \citet{komatsu02},
3336: they determined that the sample variance of the
3337: SZE-induced anisotropy signal should be a
3338: factor of three larger than the Gaussian
3339: equivalent for the $\ell$-range that ACBAR
3340: is most sensitive to.
3341: For our data, at $\ell \simeq 6000$, the sample
3342: variance is approximately four times
3343: larger than the expectation for a Gaussian.
3344: When we account for this increased sample variance
3345: our 68\%, 90\%, and 95\% confidence level upper limits for the average
3346: amplitude of the Komatsu and Seljak spectrum
3347: % weighted by our transfer function
3348: are 790, 1060, and 1080~$\mu$K$_{CMB}^2$, which are
3349: approximately 10\% higher than the upper limits obtained
3350: from assuming the fluctuations in the SZE anisotropy signal
3351: are Gaussian.
3352: The changes in the upper limits we determine are relatively
3353: minor because our uncertainty is dominated by
3354: Gaussian instrument noise rather than sample variance
3355: on the anisotropy signal.
3356: When we convert these upper limits on the anisotropy signal
3357: to an upper limit on $\sigma_8$, we find $\sigma_8 < 1.57$
3358: at a 90\% confidence level.
3359:
3360: \section{Conclusions}
3361:
3362: We have surveyed two science fields totaling one square degree with
3363: Bolocam at 2.1 mm to search for secondary CMB anisotropies caused by
3364: the Sunyaev-Zel'dovich effect. The fields are in the Lynx and
3365: Subaru/XMM SDS1 fields. Our survey is sensitive to angular scales
3366: with an effective angular multipole of $\ell_{eff} = 5700$ with
3367: FWHM$_{\ell} = 2800$ and has an angular resolution of 60 arcseconds
3368: FWHM. Our data provide
3369: no evidence for anisotropy. We are able to constrain the level of
3370: total astronomical anisotropy, modeled as a flat band power in
3371: $\mathcal{C}_\ell$, with frequentist 68\%, 90\%, and 95\% CL upper
3372: limits of 560, 760, and 830 $\mu K_{CMB}^2$. We statistically
3373: subtract the known contribution from primary CMB anisotropy, including
3374: cosmic variance, to obtain constraints on the SZE anisotropy
3375: contribution. Now including flux calibration uncertainty, our
3376: frequentist 68\%, 90\% and 95\% CL upper limits on a flat band power in
3377: $\mathcal{C}_\ell$ are 690, 960, and 1000 $\mu K_{CMB}^2$. When we
3378: instead employ the analytic spectrum suggested by \citet{komatsu02},
3379: and account for the non-Gaussianity of the SZE anisotropy signal,
3380: we obtain upper limits on the average amplitude of their spectrum
3381: weighted by our transfer function of 790, 1060, and 1080 $\mu
3382: K_{CMB}^2$. We obtain a 90\% CL upper limit on $\sigma_8$, which
3383: normalizes the power spectrum of density fluctuations, of 1.57. These
3384: are the first constraints on anisotropy and $\sigma_8$ from survey
3385: data at these angular scales at frequencies near 150~GHz.
3386:
3387: To calibrate the observations, beam maps were obtained using Uranus
3388: and Neptune. Pointing reconstruction was performed using frequent
3389: pointing observations of bright sources near our science fields. The
3390: data were flux-calibrated using techniques similar to those developed
3391: to analyze earlier Bolocam survey data collected at 1.1
3392: mm~\citep{laurent05}, using Uranus and Neptune as absolute calibrators
3393: and a number of other sources as transfer calibrators. Internal
3394: uncertainty on the pointing and flux calibration contributes
3395: negligible uncertainty to the final result; calibration uncertainty in
3396: the final result is dominated by uncertainty in models for the
3397: absolute brightness temperatures of Mars, Uranus, and Neptune.
3398:
3399: Our time-streams are dominated by fluctuations in atmospheric thermal
3400: emission (sky noise) and we developed several algorithms to subtract
3401: this noise from our data. We made use of our simple yet cross-linked
3402: scan strategy to develop a pseudo least-squares map-maker that can be
3403: run in moderate amounts of time on a single desktop computer. We used
3404: simulations to calibrate the transfer function of our data-taking and
3405: analysis pipeline and map-maker. We determined the expected noise
3406: properties of our final maps using jackknife realizations of the data
3407: obtained by randomly signed combinations of the $\sim$500 independent
3408: observations contributing to each science field map. Our final
3409: confidence intervals on anisotropy level are determined using these
3410: jackknife realizations combined with the measured transfer function
3411: for anisotropies.
3412:
3413: \section{Acknowledgements}
3414:
3415: We acknowledge the assistance of: Minhee Yun and Anthony D. Turner of
3416: NASA's Jet Propulsion Laboratory, who fabricated the Bolocam science
3417: array; Toshiro Hatake of the JPL electronic packaging group, who
3418: wirebonded the array; Marty Gould of Zen Machine and Ricardo Paniagua
3419: and the Caltech PMA/GPS Instrument Shop, who fabricated much of the
3420: Bolocam hardware; Carole Tucker of Cardiff University, who tested
3421: metal-mesh reflective filters used in Bolocam; Ben Knowles of
3422: the University of Colorado, who contributed to the software pipeline,
3423: the day crew and Hilo
3424: staff of the Caltech Submillimeter Observatory, who provided
3425: invaluable assistance during commissioning and data-taking for this
3426: survey data set; and Kathy Deniston, who provided effective
3427: administrative support at Caltech. Bolocam was constructed and
3428: commissioned using funds from NSF/AST-9618798, NSF/AST-0098737,
3429: NSF/AST-9980846, NSF/AST-0229008, and NSF/AST-0206158. JS and GL
3430: were partially supported by
3431: NASA Graduate Student Research Fellowships and
3432: SG was partially supported by a R.~A.~Millikan Postdoctoral Fellowship
3433: at Caltech.
3434:
3435: %\acknowledgments
3436:
3437: %% To help institutions obtain information on the effectiveness of their
3438: %% telescopes, the AAS Journals has created a group of keywords for telescope
3439: %% facilities. A common set of keywords will make these types of searches
3440: %% significantly easier and more accurate. In addition, they will also be
3441: %% useful in linking papers together which utilize the same telescopes
3442: %% within the framework of the National Virtual Observatory.
3443: %% See the AASTeX Web site at http://www.journals.uchicago.edu/AAS/AASTeX
3444: %% for information on obtaining the facility keywords.
3445:
3446: %% After the acknowledgments section, use the following syntax and the
3447: %% \facility{} macro to list the keywords of facilities used in the research
3448: %% for the paper. Each keyword will be checked against the master list during
3449: %% copy editing. Individual instruments or configurations can be provided
3450: %% in parentheses, after the keyword, but they will not be verified.
3451:
3452: {\it Facilities:} \facility{CSO}.
3453:
3454: %% Appendix material should be preceded with a single \appendix command.
3455: %% There should be a \section command for each appendix. Mark appendix
3456: %% subsections with the same markup you use in the main body of the paper.
3457:
3458: %% Each Appendix (indicated with \section) will be lettered A, B, C, etc.
3459: %% The equation counter will reset when it encounters the \appendix
3460: %% command and will number appendix equations (A1), (A2), etc.
3461:
3462: \appendix
3463:
3464: \section{Appendix material}
3465: \label{sec:map_var}
3466:
3467: The goal of our analysis is to determine the
3468: amplitude of the power spectrum due to emission from astronomical
3469: sources by measuring
3470: an excess noise
3471: in the maps of the science fields.
3472: This excess noise is the difference between the
3473: actual noise of the map, and the
3474: expected noise of the map based on measurements
3475: of the noise in the Bolocam system and knowledge
3476: of the expected signal spectrum.
3477: Therefore, we need measurements of the following quantities:
3478: \begin{itemize}
3479: \item $X_{\vec{\nu}}$: The measured PSD of the science field map
3480: at pixel $\vec{\nu}$ in units of $\mu$K$_{CMB}^2$.
3481: $\vec{\nu}$ is a two-dimensional value,
3482: $\vec{\nu} = (\nu_{RA}, \nu_{dec})$, describing a location
3483: in the spatial Fourier transform of the map,
3484: and has units of 1/radians.
3485: \item $\mathcal{P}_{\vec{\nu}}$:
3486: The predicted PSD of the science field map
3487: at pixel $\vec{\nu}$ in the absence of
3488: the desired astronomical signal.
3489: $\mathcal{P}_{\vec{\nu}}$ is estimated from jackknife
3490: realizations, along with the PSDs of unwanted astronomical
3491: sources in the map (\emph{i.e.}, primary CMB anisotropies in our case).
3492: \item $\mathcal{S}_{\vec{\nu}}^2$: The spatial power spectral profile
3493: of the expected astronomical signal. For a flat band power
3494: $\mathcal{S}_{\vec{\nu}}^2 = 2 \pi / (\ell(\ell+1))$,
3495: where the angular multipole $\ell$ is
3496: described by $\ell = 2 \pi |\vec{\nu}|$.
3497: \item $B_{\vec{\nu}}^2$: The peak-normalized square of the
3498: $\vec{\nu}$-space
3499: Bolocam beam profile.
3500: Since astronomical signals are attenuated by the beam,
3501: $B_{\vec{\nu}}^2$ acts like a transfer function or filter.
3502: Note that the broadening of the beam in
3503: map-space due to our pointing uncertainty is included
3504: in $B_{\vec{\nu}}^2$.
3505: \item $T_{\vec{\nu}}$: The effective transfer function, or window
3506: function, of the data processing applied to the
3507: time-stream data. Analogous to $B_{\vec{\nu}}^2$,
3508: $T_{\vec{\nu}}$ describes how much astronomical signal
3509: is attenuated.
3510:
3511: \end{itemize}
3512: With this convention, the expected PSD of the map can
3513: be described by
3514: \begin{equation}
3515: \left< X_{\vec{\nu}} \right> =
3516: \mathcal{P}_{\vec{\nu}} + A S^2_{\vec{\nu}} B^2_{\vec{\nu}}
3517: T_{\vec{\nu}},
3518: \label{eqn:exp_map_var}
3519: \end{equation}
3520: where $A$ is the amplitude of the excess anisotropy power,
3521: in $\mu$K$_{CMB}^2$.
3522:
3523: The anisotropy amplitude can be estimated by determining what value of $A$
3524: maximizes the likelihood of the measured map PSD, $X_{\vec{\nu}}$.
3525: Therefore, we need to determine the probability density function (PDF)
3526: describing $X_{\vec{\nu}}$, given $A$.
3527: First, note that
3528: \begin{displaymath}
3529: {X}_{\vec{\nu}} = \left| \alpha + i \beta \right|^2,
3530: \end{displaymath}
3531: where $\alpha$ is the real part of the Fourier transform of the
3532: science field map and $\beta$ is the imaginary part of the
3533: Fourier transform of the science field map.
3534: If we assume that the noise properties of the map are
3535: Gaussian,\footnote{
3536: This is an extremely good assumption.
3537: See Figure~\ref{fig:final_PSD_PDF}.
3538: Although the anisotropy signal may not follow a Gaussian
3539: distribution,
3540: $A S^2_{\vec{\nu}} B^2_{\vec{\nu}} T_{\vec{\nu}} \ll
3541: P_{\vec{\nu}}$ for a single $\nu$-space pixel,
3542: so the underlying distribution function for
3543: $X_{\vec{\nu}}$ will still be well approximated by a Gaussian.}
3544: then the PDFs for $\alpha$ and $\beta$ are the same and are given by
3545: \begin{eqnarray}
3546: f(\alpha) = \frac{1}{\sqrt{2 \pi \sigma^2}} e^{-\alpha^2/2\sigma^2}
3547: & {\rm and} &
3548: f(\beta) = \frac{1}{\sqrt{2 \pi \sigma^2}} e^{-\beta^2/2\sigma^2},
3549: \label{eqn:PDF_real_mapft}
3550: \end{eqnarray}
3551: where $\sigma^2 = \left< X_{\vec{\nu}} \right>/2$.
3552: Next, after a change of variables to
3553: $\alpha = r \cos(\theta)$ and $\beta = r \sin(\theta)$, the
3554: PDF in Equation~\ref{eqn:PDF_real_mapft} becomes
3555: \begin{displaymath}
3556: f(r,\theta) = \frac{1}{2 \pi \sigma^2}
3557: r e^{-r^2/2\sigma^2}.
3558: \end{displaymath}
3559: Since the $\theta$ dependence of $f(r,\theta)$ is trivial,
3560: we can reduce the above PDF to $f(r) = 2 \pi f(r,\theta)$,
3561: with
3562: \begin{displaymath}
3563: f(r) = \frac{r}{\sigma^2}e^{-r^2/2\sigma^2}.
3564: \end{displaymath}
3565: Finally, after one more change of variables
3566: using the relation $X_{\vec{\nu}} = r^2$, we find that the
3567: PDF for $X_{\vec{\nu}}$ is equal to
3568: \begin{equation}
3569: f(X_{\vec{\nu}}) =
3570: % \left(\frac{1}{N_p} \right)
3571: \frac{1}{2 \sigma^2}
3572: e^{-X_{\vec{\nu}} / 2\sigma^2},
3573: \label{eqn:PDF_x}
3574: \end{equation}
3575: where the factor of $r$ has been replaced by $1/2$
3576: due to the change in the differential element.
3577: Equation~\ref{eqn:PDF_x}
3578: can be written in terms of our measured parameters as
3579: \begin{equation}
3580: f(X_{\vec{\nu}}|A) =
3581: % \left( \frac{1}{N_p} \right)
3582: \frac{1}{\mathcal{P}_{\vec{\nu}} +
3583: A S^2_{\vec{\nu}} B^2_{\vec{\nu}} T_{\vec{\nu}}}
3584: {\rm exp} \left( \frac{-X_{\vec{\nu}}}
3585: {\mathcal{P}_{\vec{\nu}} + A S^2_{\vec{\nu}} B^2_{\vec{\nu}}
3586: T_{\vec{\nu}}}
3587: \right)
3588: \label{eqn:PDF_X_a}
3589: \end{equation}
3590: using Equation~\ref{eqn:exp_map_var}.
3591: Note that we have made use of the fact that
3592: $2\sigma^2 = \left< X_{\vec{\nu}} \right> =
3593: \mathcal{P}_{\vec{\nu}} + A S^2_{\vec{\nu}} B^2_{\vec{\nu}} T_{\vec{\nu}}$
3594: to go from Equation~\ref{eqn:PDF_x} to Equation~\ref{eqn:PDF_X_a}.
3595: % The factor of $1/N_p$, where $N_p$ is the number of pixels
3596: % in the $\vec{\nu}$-space map, has
3597: % been added to keep the PDF normalized so that
3598: % \begin{equation}
3599: % \sum_{\vec{\nu}} \int_0^{\infty} d X_{\vec{\nu}} f(X_{\vec{\nu}}|A) = 1.
3600: % \end{equation}
3601:
3602: The next step is to calculate a likelihood function,
3603: $\mathcal{L}$, from
3604: Equation~\ref{eqn:PDF_X_a} by multiplying $f(X_{\vec{\nu}}|A)$
3605: over all
3606: of the $\vec{\nu}$-space pixels.
3607: This product can be turned into a sum by taking the
3608: logarithm of $\mathcal{L}$, with
3609: \begin{equation}
3610: {\rm log}(\mathcal{L}) = \sum_{\vec{\nu}} \left(
3611: - {\rm log}
3612: ({\mathcal{P}_{\vec{\nu}} + A S^2_{\vec{\nu}} B^2_{\vec{\nu}}
3613: T_{\vec{\nu}}})
3614: - \frac{X_{\vec{\nu}}}{
3615: {\mathcal{P}_{\vec{\nu}} + A S^2_{\vec{\nu}} B^2_{\vec{\nu}}
3616: T_{\vec{\nu}}}}
3617: \right).
3618: \label{eqn:log_l}
3619: \end{equation}
3620: Note that half of the $\nu$-space is discarded from the sum
3621: in Equation~\ref{eqn:log_l} since our maps are real
3622: ($X_{\vec{\nu}} = X_{-\vec{\nu}}$).
3623:
3624: Then, the most probable value of the anisotropy amplitude
3625: for our measured map PSD can be determined
3626: by maximizing ${\rm log}(\mathcal{L})$ with respect to $A$.
3627: In practice, we maximize Equation~\ref{eqn:log_l} by evaluating
3628: ${\rm log}(\mathcal{L})$ at a range of values for $A$.
3629: Since the number of $\vec{\nu}$-space pixels is $\lesssim 10000$,
3630: the computational time required to evaluate ${\rm log}(\mathcal{L})$
3631: at each value of $A$ is minimal, which means that we
3632: can determine the best fit value of $A$ to almost
3633: any desired precision using this numerical method.
3634:
3635: However, it is also instructive to analytically approximate the value
3636: of $A$ that maximizes Equation~\ref{eqn:log_l}.
3637: To start, we take the derivative of ${\rm log}(\mathcal{L})$ with
3638: respect to $A$, yielding
3639: \begin{equation}
3640: \left. \frac{\partial {\rm log}(\mathcal{L})}{\partial A}
3641: \right|_{A = \hat{A}} = \left.
3642: \sum_{\vec{\nu}} \frac{\Theta_{\vec{\nu}}}
3643: {(\mathcal{P}_{\vec{\nu}} + A \Theta_{\vec{\nu}})^2}
3644: \left(X_{\vec{\nu}} - \mathcal{P}_{\vec{\nu}} - A \Theta_{\vec{\nu}}
3645: \right) \right|_{A = \hat{A}} = 0,
3646: \label{eqn:partial_l}
3647: \end{equation}
3648: where $\Theta_{\vec{\nu}} = S^2_{\vec{\nu}} B^2_{\vec{\nu}}
3649: T_{\vec{\nu}}$ and
3650: $\hat{A}$ is the best fit value of $A$.
3651: For any given $\vec{\nu}$-space pixel, $\mathcal{P}_{\vec{\nu}} \gg
3652: A \Theta_{\vec{\nu}}$ for any physically reasonable value of $A$.
3653: Therefore, we can simplify
3654: Equation~\ref{eqn:partial_l} to
3655: \begin{displaymath} \left.
3656: \sum_{\vec{\nu}}
3657: \frac{\Theta_{\vec{\nu}}}{\mathcal{P}_{\vec{\nu}}^2}
3658: \left(1 - \frac{2 A \Theta_{\vec{\nu}}}{\mathcal{P}_{\vec{\nu}}}
3659: % + \frac{3 A^2 \Theta_{\vec{\nu}}^2}{\mathcal{P}_{\vec{\nu}}^2}
3660: + \mathcal{O}
3661: \left( \frac{A^2 \Theta_{\vec{\nu}}^2}{\mathcal{P}_{\vec{\nu}}^2}
3662: \right) \right)
3663: \left(X_{\vec{\nu}} - \mathcal{P}_{\vec{\nu}} - A \Theta_{\vec{\nu}}
3664: \right) \right|_{A = \hat{A}} \simeq 0.
3665: \end{displaymath}
3666: If we rearrange some terms, and again keep only the lowest
3667: order terms in $A \Theta_{\vec{\nu}} / \mathcal{P}_{\vec{\nu}}$,
3668: then we find
3669: \begin{displaymath}
3670: \hat{A} \simeq
3671: \frac{\sum_{\vec{\nu}}
3672: \frac{\Theta_{\vec{\nu}}^2}{\mathcal{P}^2}
3673: \left( \frac{X_{\vec{\nu}} - \mathcal{P}_{\vec{\nu}}}{\Theta_{\vec{\nu}}}
3674: \right)}
3675: {\sum_{\vec{\nu}}
3676: \frac{\Theta_{\vec{\nu}}^2}{\mathcal{P}^2}
3677: \left( \frac{ 2 X_{\vec{\nu}} - \mathcal{P}_{\vec{\nu}}}
3678: {\mathcal{P}_{\vec{\nu}}} \right) }.
3679: \end{displaymath}
3680: Finally, because
3681: $A \Theta_{\vec{\nu}} \ll \mathcal{P}_{\vec{\nu}}$,
3682: we can make the approximation that $\left< X_{\vec{\nu}} \right> \simeq
3683: \mathcal{P}_{\vec{\nu}}$,
3684: which means that $\left< 2 X_{\vec{\nu}} - \mathcal{P}_{\vec{\nu}} \right>
3685: \simeq \mathcal{P}_{\vec{\nu}}$.
3686: With this approximation we find
3687: \begin{equation}
3688: \hat{A} \simeq
3689: \frac{\sum_{\vec{\nu}}
3690: \frac{\Theta_{\vec{\nu}}^2}{\mathcal{P}^2}
3691: \left( \frac{X_{\vec{\nu}} - \mathcal{P}_{\vec{\nu}}}{\Theta_{\vec{\nu}}}
3692: \right)}
3693: {\sum_{\vec{\nu}}
3694: \frac{\Theta_{\vec{\nu}}^2}{\mathcal{P}^2}}.
3695: \label{eqn:est_a}
3696: \end{equation}
3697: To understand this result, consider that
3698: for a single $\vec{\nu}$-space pixel the best estimate of $A$
3699: is $(X_{\vec{\nu}} - \mathcal{P}_{\vec{\nu}})/ \Theta_{\vec{\nu}}$.
3700: Therefore, Equation~\ref{eqn:est_a} determines the
3701: weighted mean of $A$ over all pixels, assuming that the
3702: uncertainty on the value of $A$ for
3703: each $\vec{\nu}$-space pixel is proportional to
3704: $\mathcal{P}_{\vec{\nu}}/\Theta_{\vec{\nu}}$,
3705: which is a reasonable assumption.
3706: This means that the variance on $\hat{A}$ implied by
3707: Equation~\ref{eqn:est_a} is proportional to
3708: \begin{equation}
3709: \sigma^2_{\hat{A}} \propto \frac{1}{\sum_{\vec{\nu}}
3710: \frac{\Theta_{\vec{\nu}}^2}{\mathcal{P}^2}}.
3711: \label{eqn:est_var_a}
3712: \end{equation}
3713:
3714: %% The reference list follows the main body and any appendices.
3715: %% Use LaTeX's thebibliography environment to mark up your reference list.
3716: %% Note \begin{thebibliography} is followed by an empty set of
3717: %% curly braces. If you forget this, LaTeX will generate the error
3718: %% "Perhaps a missing \item?".
3719: %%
3720: %% thebibliography produces citations in the text using \bibitem-\cite
3721: %% cross-referencing. Each reference is preceded by a
3722: %% \bibitem command that defines in curly braces the KEY that corresponds
3723: %% to the KEY in the \cite commands (see the first section above).
3724: %% Make sure that you provide a unique KEY for every \bibitem or else the
3725: %% paper will not LaTeX. The square brackets should contain
3726: %% the citation text that LaTeX will insert in
3727: %% place of the \cite commands.
3728:
3729: %% We have used macros to produce journal name abbreviations.
3730: %% AASTeX provides a number of these for the more frequently-cited journals.
3731: %% See the Author Guide for a list of them.
3732:
3733: %% Note that the style of the \bibitem labels (in []) is slightly
3734: %% different from previous examples. The natbib system solves a host
3735: %% of citation expression problems, but it is necessary to clearly
3736: %% delimit the year from the author name used in the citation.
3737: %% See the natbib documentation for more details and options.
3738:
3739: \begin{thebibliography}{}
3740: %\bibitem[Aguirre(2003)]{aguirre03}
3741: % Aguirre, J. E., 2003, PhD Thesis, University of Chicago
3742: \bibitem[Aguirre(2008)]{aguirre08}
3743: Aguirre, J. E., et al., 2008, in preparation
3744: \bibitem[Bennett et al.(2003)]{bennett03}
3745: Bennett, C. L. et al., 2003, \apjs, 148, 97
3746: \bibitem[Bhatia et al.(2000)]{bhatia00}
3747: Bhatia, R. S. et al., 2000, Cryogenics, 40, 685
3748: \bibitem[Bhatia et al.(2002)]{bhatia02}
3749: Bhatia, R. S., Chase, S. T., Jones, W. C.,
3750: Keating, B. G., Lange, A. E., Mason, P. V.,
3751: Philhour, B. J., and Sirbi, G., 2002, Cryogenics, 42, 113
3752: \bibitem[Birkinshaw(1999)]{birkinshaw99}
3753: Birkinshaw, M., 1999, \physrep, 310, 97
3754: %\bibitem[Bonamente et al.(2006)]{bonamente06}
3755: % Bonamente, M., Joy, M. K., LaRoque, S. J.,
3756: % Carlstrom, J. E., Reese, E. D., and Dawson, K. S.,
3757: % 2006, \apj, 647, 25
3758: \bibitem[Bond et al.(2005)]{bond05}
3759: Bond, J. R. et al., 2005, \apj, 626, 12
3760: \bibitem[Borys et al.(2003)]{borys03}
3761: Borys, C., Chapman, S., Halpern, M., and Scott, D.,
3762: 2003, \mnras, 344, 385
3763: \bibitem[Carlstrom et al.(2002)]{carlstrom02}
3764: Carlstrom, J. E., Holder, G. P., and
3765: Reese, E. D., 2001, \araa, 40, 643
3766: \bibitem[Cooray(2001)]{cooray01}
3767: Cooray, A., 2001, \prd, 64, 063514
3768: \bibitem[Coppin et al.(2006)]{coppin06}
3769: Coppin, K. et al., 2006, \mnras, 372, 1621
3770: \bibitem[Crow and Gardner(1959)]{crow59}
3771: Crow, E. L. and Gardner, R. S.,
3772: 1959, Biometrika, 46, 441
3773: \bibitem[da Silva et al.(2001)]{dasilva01}
3774: da Silva, A. C., Kay, S. T., Liddle, A. R.,
3775: Thomas, P. A., Pearce, F. R., and Barbosa, D.,
3776: 2001, \apjl, 561, L15
3777: \bibitem[Danese et al.(1987)]{danese87}
3778: Danese, L., Franceschini, A., Toffolatti, L., and
3779: de Zotti, G., 1987, \apjl, 318, L15
3780: \bibitem[Dawson et al.(2006)]{dawson06}
3781: Dawson, K. S., Holzapfel, W. L., Carlstrom, J. E.,
3782: Joy, M., and LaRoque, S. J., 2006,
3783: \apj, 647, 13
3784: \bibitem[DIRBE website()]{dirbe_website}
3785: DIRBE 100 $\mu$m full sky maps (available at
3786: http://astro.berkeley.edu/marc/dust/data/data.html)
3787: %\bibitem[Dobbs et al.(2006)]{dobbs06}
3788: % Dobbs, M. et al., 2006, New Astr. Rev., 50, 960
3789: \bibitem[Dunkley et al.(2008)]{dunkley08}
3790: Dunkley, J. et al., 2008 preprint (astro-ph/08030586)
3791: \bibitem[Enoch et al.(2006)]{enoch06}
3792: Enoch, M. L. et al., 2006, \apj, 638, 293
3793: \bibitem[Feldman and Cousins(1998)]{feldman98}
3794: Feldman, G. J. and Cousins, R. D.,
3795: 1998, \prd, 57, 3873
3796: \bibitem[Finkbeiner et al.(1999)]{finkbeiner99}
3797: Finkbeiner, D. P., Davis, M., and Schlegel, D. J.,
3798: 1999, \apj, 524, 867
3799: \bibitem[Glenn et al.(1998)]{glenn98}
3800: Glenn, J. et al., 1998, \procspie, 3357, 326
3801: \bibitem[Glenn et al.(2002)]{glenn02}
3802: Glenn, J., Chattopadhyay, G., Edgington, S. F.,
3803: Lange, A. E., Bock, J. J., Mauskopf, P. D.,
3804: and Lee, A. T., 2002, \applopt, 41, 136
3805: \bibitem[Glenn et al.(2003)]{glenn03}
3806: Glenn, J. et al., 2003, \procspie, 4855, 30
3807: \bibitem[Goldin et al. (1997)]{goldin97}
3808: Goldin, A. B. et al., 1997, \apjl, 488, L161
3809: \bibitem[Goldstein et al.(2003)]{goldstein03}
3810: Goldstein, J. H. et al., 2003,
3811: \apj, 599, 773
3812: \bibitem[Golwala et al.(2008)]{golwala08}
3813: Golwala et al., 2008, in preparation
3814: \bibitem[Griffin et al.(1986)]{griffin86}
3815: Griffin, M. J., Ade, P. A., R., Orton, G. S.,
3816: Robson, E. I., Gear, W. K., Nolt, I. G.,
3817: and Radostitz, J. V., 1986,
3818: Icarus, 65, 244
3819: \bibitem[Griffin and Orton(1993)]{griffin93}
3820: Griffin, M. J. and Orton, G. S.,
3821: 1993, Icarus, 105, 537
3822: \bibitem[Haig et al.(2004)]{haig04}
3823: Haig, D. J. et al., 2004, \procspie, 5498, 78
3824: %\bibitem[Haiman et al.(2000)]{haiman00}
3825: % Haiman, Z., Mohr, J. J., and Holder, G. P.,
3826: % 2000, \apj, 553, 545
3827: %\bibitem[Hinshaw et al.(2007)]{hinshaw07}
3828: % Hinshaw, G. et al., 2007, \apjs, 170, 288
3829: %\bibitem[Holder et al.(2000)]{holder00}
3830: % Holder, G. P., Mohr, J. J., Carlstrom, J. E., Evrard, A. E.,
3831: % Leitch, E. M., 2000, \apj, 544, 629
3832: %\bibitem[Jones et al.(2006)]{jones06}
3833: % Jones, W. C. et al., 2006, \apj, 647, 823
3834: \bibitem[Laurent et al.(2005)]{laurent05}
3835: Laurent, G. T. et al., 2005, \apj, 623, 742
3836: \bibitem[Knox(1999)]{knox99}
3837: Knox, L., 1999, \prd, 60, 103516
3838: \bibitem[Komatsu and Seljak(2002)]{komatsu02}
3839: Komatsu, E. and Seljak, U., 2002,
3840: \mnras, 336, 1256
3841: %\bibitem[Kosowsky(2003)]{kosowsky03}
3842: % Kosowsky, A., 2003, New Astr. Rev., 47, 939
3843: %\bibitem[Kuo et al.(2007)]{kuo06}
3844: % Kuo, C. L., et al., 2007, \apj, 664, 687
3845: %\bibitem[Maloney et al.(2005)]{maloney05}
3846: % Maloney, P. R. et al., 2005, \apj, 635, 1044
3847: \bibitem[Mauskopf et al.(1997)]{mauskopf97}
3848: Mauskopf, P. D., Bock, J. J., Del Castillo, H.,
3849: Holzapfel, W. L., and Lange, A. E.,
3850: 1997, \ao, 36, 765
3851: \bibitem[Mason et al.(2003)]{mason03}
3852: Mason, B. S. et al., 2003, \apj, 591, 540
3853: %\bibitem[Natoli et al.(2001)]{natoli01}
3854: % Natoli, P., de Gasperis, G., Gheller, C., and
3855: % Vittorio, N., 2001, \aap, 372, 346
3856: \bibitem[Nolta et al.(2008)]{nolta08}
3857: Nolta, M. R. et al., 2008 preprint (astro-ph/08030593)
3858: \bibitem[Orton et al.(1986)]{orton86}
3859: Orton, G. S., Griffin, M. J., Ade, P. A. R.,
3860: Nolt, I. G., and Radostitz, J. V.,
3861: 1986, Icarus, 67, 289
3862: \bibitem[Peng et al.(2000)]{peng00}
3863: Peng, B., Kraus, A., Krichbaum, T. P., and Witzel, A.,
3864: 2000, \aaps, 145, 1
3865: %\bibitem[Reese et al.(2002)]{reese02}
3866: % Reese, E. D., Carlstrom, J. E., Joy, M.,
3867: % Mohr, J. J., Grego, L., and Holzapfel, W. L.,
3868: % 2002, \apj, 581, 53
3869: \bibitem[Refregier et al.(2000)]{refregier00}
3870: Refregier, A., Komatsu, E., Spergel, D. N., and Pen, U.-L.,
3871: 2000, \prd, 61, 123001
3872: \bibitem[Refregier and Teyssier(2002)]{refregier00_2}
3873: Refregier, A. and Teyssier, R.,
3874: 2002, \prd, 66, 043002
3875: \bibitem[Reichardt et al.(2008)]{reichardt08}
3876: Reichardt, C. R. et al., 2008 preprint (astro-ph/08011491)
3877: \bibitem[Rudy(1987)]{rudy87}
3878: Rudy, D. J., 1987, PhD Thesis, Caltech
3879: \bibitem[Rudy et al.(1987)]{rudy87_2}
3880: Rudy, D. J., Muhleman, D. O., Berge, G. L.,
3881: Jakosky, B. M., and Christensen, P. R.,
3882: 1987, Icarus, 71, 159
3883: %\bibitem[Ruhl et al.(2004)]{ruhl04}
3884: % Ruhl, J. et al., 2004, \procspie, 5498, 11
3885: \bibitem[Sandell(1994)]{sandell94}
3886: Sandell, G., 1994, \mnras, 271, 75
3887: %\bibitem[Sayers(2007)]{sayers07}
3888: % Sayers, J., 2007, PhD Thesis, Caltech
3889: \bibitem[Sayers et al.(2008)]{sayers08}
3890: Sayers, J. et al., 2008, in preparation
3891: \bibitem[Schlegel et al.(1998)]{schlegel98}
3892: Schlegel, D. J., Finkbeiner, D. P., and Davis, M.,
3893: 1998, \apj, 500, 525
3894: \bibitem[Scott and White(1999)]{scott99}
3895: Scott, D. and White, M., 1999,
3896: \aap, 346, 1
3897: \bibitem[Seljak and Zaldarriaga(1996)]{seljak96}
3898: Seljak, U. and Zaldarriaga, M., 1996,
3899: \apj, 469, 437
3900: \bibitem[Seljak et al.(2001)]{seljak01}
3901: Seljak, U., Burwell, J., and Pen, U.-L.,
3902: 2001, \prd, 63, 063001
3903: %\bibitem[Spergel et al.(2007)]{spergel07}
3904: % Spergel, D. N. et al., 2007, \apjs, 170, 377
3905: \bibitem[Springel et al.(2001)]{springel01}
3906: Springel, V., White, M., and Hernquist, L.,
3907: 2001, \apj, 549, 681
3908: \bibitem[Sunyaev and Zel'dovich(1972)]{sunyaev72}
3909: Sunyaev, R. A. and Zel'dovich, Y. B., 1972
3910: Comm. Astr. and Space Phys., 4, 173
3911: \bibitem[Tegmark(1997)]{tegmark97}
3912: Tegmark, M., 1997, \apjl, 480, L87
3913: \bibitem[Toffolatti et al.(1998)]{toffolatti98}
3914: Toffolatti, L., Argueso Gomez, F., de Zotti, G.,
3915: Mazzei, P., Franceschini, A., Danese, L.,
3916: and Burigana, C., 1998, \mnras, 297, 117
3917: %\bibitem[Udomprasert et al.(2004)]{udomprasert04}
3918: % Udomprasert, P. S., Mason, B. S., Readhead, A. C. S.,
3919: % and Pearson, T. J., 2004, \apj, 615, 63
3920: \bibitem[White et al.(2002)]{white02}
3921: White, M. J., Hernquist, L., and Springel, V.,
3922: 2002, \apj, 577, 555
3923: \bibitem[White and Majumdar(2004)]{white04}
3924: White, M. and Majumdar, S., 2004,
3925: \apj, 602, 565
3926: \bibitem[Wright(1976)]{wright76}
3927: Wright, E. L., 1976, \apj, 210, 250
3928: \bibitem[Wright(1996)]{wright96}
3929: Wright, E. L., 1996, paper presented at the IAS CMB
3930: Data Analysis Workshop (astro-ph/9612006)
3931: \bibitem[Young et al.(2006)]{young06}
3932: Young, K. E. et al., 2006, \apj, 644, 326
3933: \bibitem[Zaldarriaga et al.(1998)]{zaldarriaga98}
3934: Zaldarriaga, M., Seljak, U., and Bertschinger, E.,
3935: 1998, \apj, 494, 491
3936: \bibitem[Zaldarriaga and Seljak(2000)]{zaldarriaga00}
3937: Zaldarriaga, M. and Seljak, U., 2000,
3938: \apjs, 129, 431
3939: \bibitem[Zhang et al.(2002)]{zhang02}
3940: Zhang, P., Pen, U.-L., and Wang, B.,
3941: 2002, \apj, 577, 555
3942: \end{thebibliography}
3943:
3944: \clearpage
3945: \begin{deluxetable}{ccccc}
3946: \tablewidth{0pt}
3947: \tablecaption{SZE-induced CMB anisotropy results}
3948: \tablehead{\colhead{spectrum} & \colhead{flux uncertainty} &
3949: \colhead{68\% CL interval} & \colhead{90\% CL interval} &
3950: \colhead{95\% CL interval}}
3951: \startdata
3952: flat-total & 0 & $100 - 590$ $\mu$K$^2_{CMB}$ &
3953: $0 - 760$ $\mu$K$^2_{CMB}$ &
3954: $0 - 830$ $\mu$K$^2_{CMB}$ \\
3955: flat-SZE & 0 & $90 - 580$ $\mu$K$^2_{CMB}$ &
3956: $0 - 750$ $\mu$K$^2_{CMB}$ &
3957: $0 - 830$ $\mu$K$^2_{CMB}$ \\
3958: flat-SZE & 3.5\% (meas) & $90 - 630$
3959: $\mu$K$^2_{CMB}$ & $0 - 790$ $\mu$K$^2_{CMB}$ &
3960: $0 - 880$ $\mu$K$^2_{CMB}$ \\
3961: flat-SZE & 6.3\% (total) & $80 - 690$ $\mu$K$^2_{CMB}$ &
3962: $0 - 960$ $\mu$K$^2_{CMB}$ &
3963: $0 - 1000$ $\mu$K$^2_{CMB}$ \\
3964: KS-SZE & 0 & $80 - 540$ $\mu$K$^2_{CMB}$ &
3965: $0 - 690$ $\mu$K$^2_{CMB}$ &
3966: $0 - 770$ $\mu$K$^2_{CMB}$ \\
3967: KS-SZE & 3.5\% (meas) & $80 - 570$ $\mu$K$^2_{CMB}$ &
3968: $0 - 740$ $\mu$K$^2_{CMB}$ &
3969: $0 - 830$ $\mu$K$^2_{CMB}$ \\
3970: KS-SZE & 6.3\% (total) & $70 - 730$ $\mu$K$^2_{CMB}$ &
3971: $0 - 950$ $\mu$K$^2_{CMB}$ &
3972: $0 - 990$ $\mu$K$^2_{CMB}$ \\
3973: KS-SZE (nG) & 6.3\% (total) & $90 - 790$ $\mu$K$^2_{CMB}$ &
3974: $0 - 1060$ $\mu$K$^2_{CMB}$ &
3975: $0 - 1080$ $\mu$K$^2_{CMB}$ \\
3976: \enddata
3977: \tablecomments{Confidence intervals
3978: for our estimates of the total and
3979: SZE-induced CMB anisotropy
3980: amplitude for both a flat SZE band power in
3981: $\mathcal{C}_{\ell}$ and the SZE spectrum given
3982: by the analytic model of Komatsu and Seljak \citep{komatsu02}.
3983: The limits for the analytic model refer to the
3984: average amplitude of the SZE spectrum weighted by
3985: our transfer function.
3986: % , and therefore
3987: % make the upper limits for the analytic spectrum
3988: % appear artificially high compared to the upper
3989: % limits for the flat spectrum.
3990: The three rows for each SZE spectrum give the upper
3991: limits for no uncertainty in our flux calibration,
3992: the 3.5\% uncertainty in our flux calibration
3993: due to measurement error, and the
3994: 6.3\% uncertainty in our flux calibration
3995: due to the combination of measurement error
3996: and uncertainty in the surface brightness
3997: of Uranus and Neptune.
3998: The final row gives the confidence intervals when
3999: the non-Gaussianity of the SZE anisotropy signal
4000: is accounted for.}
4001: \label{tab:SZE_result}
4002: \end{deluxetable}
4003:
4004: \clearpage
4005: \begin{deluxetable}{ccccc}
4006: \tablewidth{0pt}
4007: \tablecaption{Bias and efficiency of Equation~\ref{eqn:log_l}}
4008: \tablehead{\colhead{input $A_{sim}$} & \colhead{average $\hat{A}$} &
4009: \colhead{$\sigma_{\hat{A}}$} & \colhead{$\sigma_{\hat{A}}/N_{real}$} &
4010: \colhead{min. var. $\sigma_{\hat{A}}$}}
4011: \startdata
4012: 0 $\mu$K$_{CMB}^2$ & -2 $\mu$K$_{CMB}^2$ & 365 $\mu$K$_{CMB}^2$ &
4013: 12 $\mu$K$_{CMB}^2$ & 270 $\mu$K$_{CMB}^2$ \\
4014: 100 $\mu$K$_{CMB}^2$ & 96 $\mu$K$_{CMB}^2$ & 366 $\mu$K$_{CMB}^2$ &
4015: 12 $\mu$K$_{CMB}^2$ & 270 $\mu$K$_{CMB}^2$ \\
4016: 200 $\mu$K$_{CMB}^2$ & 194 $\mu$K$_{CMB}^2$ & 367 $\mu$K$_{CMB}^2$ &
4017: 12 $\mu$K$_{CMB}^2$ & 270 $\mu$K$_{CMB}^2$ \\
4018: 400 $\mu$K$_{CMB}^2$ & 389 $\mu$K$_{CMB}^2$ & 371 $\mu$K$_{CMB}^2$ &
4019: 12 $\mu$K$_{CMB}^2$ & 270 $\mu$K$_{CMB}^2$ \\
4020: 800 $\mu$K$_{CMB}^2$ & 806 $\mu$K$_{CMB}^2$ & 380 $\mu$K$_{CMB}^2$ &
4021: 12 $\mu$K$_{CMB}^2$ & 270 $\mu$K$_{CMB}^2$ \\
4022: \enddata
4023: \tablecomments{A comparison between the amplitude
4024: of a simulated power spectrum added to a
4025: jackknifed realization of the
4026: data, $A_{sim}$, and the most likely
4027: amplitude determined from Equation~\ref{eqn:log_l},
4028: $\hat{A}$.
4029: In each case 1000 jackknifed realizations of the Lynx data were
4030: used,
4031: and the table lists the average value of
4032: $\hat{A}$ for these realizations along with the
4033: standard deviation of the values of $\hat{A}$.
4034: In each case the average value of $\hat{A}$ is
4035: consistent with $A_{sim}$, indicating that
4036: Equation~\ref{eqn:log_l} is an unbiased estimator
4037: of $A$.
4038: Additionally,
4039: to determine whether Equation~\ref{eqn:log_l} is an
4040: efficient (minimum variance) estimator for $A$,
4041: we calculate the standard deviation of the estimates
4042: $\hat{A}$ for each input $A_{sim}$ and compare them to
4043: the standard deviation one would estimate using the
4044: Bayesian likelihood, Equation~\ref{eqn:est_var_a}.
4045: The latter underestimates the minimum possible standard
4046: deviation because the Bayesian likelihood is incorrect
4047: for the reasons presented in Section~\ref{sec:bayesian_dificiencies}.
4048: Thus, the fact that the observed standard deviation is only
4049: 40\% larger than the Equation~\ref{eqn:est_var_a}-based
4050: estimate gives us confidence that our estimator
4051: for $A$ is reasonably efficient.}
4052: % we have estimated the minimum
4053: % variance we can expect for $\sigma_{\hat{A}}$
4054: % using Equation~\ref{eqn:est_var_a}.
4055: % This estimate for the minimum variance is
4056: % based on the Bayesian likelihood and is probably
4057: % optimistic,
4058: % so although the values of $\sigma_{\hat{A}}$ are slightly
4059: % larger than this minimum variance estimate,
4060: % Equation~\ref{eqn:log_l} is still an
4061: % approximately minimum variance estimator of $\hat{A}$.}
4062: \label{tab:log_l}
4063: \end{deluxetable}
4064:
4065: \clearpage
4066: \begin{figure}
4067: \plotone{f1.eps}
4068: % \includegraphics[width=\textwidth]{f1.eps}
4069: \caption{The solid black line represents
4070: a pre-down-sampled time-stream PSD,
4071: which has 60~Hz pickup at
4072: frequencies above $\simeq 10$~Hz.
4073: The dashed red line shows the time-stream PSD
4074: after downsampling and processing, including
4075: removal of an atmospheric noise template.
4076: %JS 2008/07/31 addressing referee comment 4
4077: The sharp increase in this PSD near 5~Hz is
4078: caused by the small amount of noise that is
4079: aliased in from frequencies just above 5~Hz
4080: during the downsampling process.
4081: Since there is approximately no astronomical
4082: signal at the frequencies
4083: where this noise increase occurs, this
4084: noise does not have a noticeable effect on
4085: our sensitivity.
4086: Overlaid as a dot-dashed green line is the beam profile,
4087: showing that very little astronomical
4088: signal will be present above a few Hz.}
4089: \label{fig:pickup_60hz}
4090: \end{figure}
4091:
4092: \clearpage
4093: \begin{figure}
4094: \plotone{f2.eps}
4095: % \includegraphics[width=\textwidth]{f2.eps}
4096: \caption{Location of the beam center of every detector
4097: relative to the center of the array.
4098: The red rings around each beam center represent the
4099: approximate FWHM of the beam.}
4100: \label{fig:beam_locations}
4101: \end{figure}
4102:
4103: \clearpage
4104: \begin{figure}
4105: \plottwo{f3a.eps}{f3b.eps}
4106: % \includegraphics[width=.49\textwidth]{f3a.eps}
4107: % \includegraphics[width=.49\textwidth]{f3b.eps}
4108: \caption{All of the
4109: raw pointing data for Lynx at azimuth angles between
4110: -90~and~90 degrees.
4111: The pointing model (quadratic fit) is overlaid.
4112: Similar models were fit to the SDS1 pointing data and the
4113: Lynx data at azimuth angles between 270~and~360 degrees.}
4114: \label{fig:raw_pointing}
4115: \end{figure}
4116:
4117: \clearpage
4118: \begin{figure}
4119: \plotone{f4.eps}
4120: % \includegraphics[width=\textwidth]{f4.eps}
4121: \caption{Flux calibration for one of the six calibration data
4122: sets, overlaid with
4123: a linear fit of calibration versus
4124: bolometer operating resistance measured by the
4125: bolometer voltage at the bias frequency.
4126: Note that the bolometer voltage at the bias frequency is
4127: a monotonic function of the atmospheric opacity and
4128: the bolometer responsivity.}
4129: \label{fig:flux_slope}
4130: \end{figure}
4131:
4132: \clearpage
4133: \begin{figure}
4134: \plotone{f5.eps}
4135: % \includegraphics[width=\textwidth]{f5.eps}
4136: \caption{A histogram of
4137: the beam area calculated
4138: for each bolometer, with a Gaussian
4139: fit overlaid.}
4140: \label{fig:beam_area_variation}
4141: \end{figure}
4142:
4143: \clearpage
4144: \begin{figure}
4145: \plottwo{f6a.eps}{f6b.eps}
4146: % \centering \includegraphics[width=.6\textwidth]{f6a.eps}
4147: % \centering \includegraphics[width=.6\textwidth]{f6b.eps}
4148: \caption{Map coverage, quantified by the number of time-stream
4149: samples that correspond to a particular map-space pixel
4150: for a single observation of the Lynx field and for
4151: the co-add of all observations of the Lynx field.
4152: % The top plot shows a single eight-minute-long observation
4153: % made while scanning in the RA direction; the bottom plot
4154: % shows the total of all $\simeq 500$ observations made
4155: % of the Lynx field.
4156: The white square has sides of
4157: approximately 42~arcminutes and
4158: contains the region of the map
4159: defined to have uniform coverage.
4160: The RMS deviations in coverage within this region
4161: relative to the
4162: average coverage within the region
4163: are approximately 8 -- 9\% for a single observation
4164: and around 1.5\% for the co-add of all observations.}
4165: \label{fig:mapcov_single_obs}
4166: \end{figure}
4167:
4168: \clearpage
4169: \begin{figure}
4170: \plottwo{f7a.eps}{f7b.eps}
4171: % \centering \includegraphics[width=.7\textwidth]{f7a.eps}
4172: % \centering \includegraphics[width=.7\textwidth]{f7b.eps}
4173: \caption{Map-space PSDs, $\mathcal{P}_{\vec{\nu}}$, for single observations.
4174: % The plot on the left shows the PSD for an observation
4175: % made while scanning in the RA direction in good
4176: % weather conditions.
4177: % The plot on the right shows the PSD for an observation
4178: % made while scanning in the dec direction
4179: % in poor weather conditions.
4180: The plot on the left shows an observation made in relatively good weather
4181: while scanning in the RA direction, and the plot on the right
4182: shows an observation made in relatively poor weather
4183: while scanning in the dec direction.
4184: In each case, note that there is a stripe of increased
4185: noise at low frequency along the scan direction,
4186: due to time-stream noise with a $1/f$ spectrum.}
4187: \label{fig:map_psd_single_obs}
4188: \end{figure}
4189:
4190: \clearpage
4191: \begin{figure}
4192: \plottwo{f8a.eps}{f8b.eps}
4193: % \centering \includegraphics[width=.5\textwidth]{f8a.eps}
4194: % \centering \includegraphics[width=.5\textwidth]{f8b.eps}
4195: \caption{Contour plots of the
4196: transfer function, $T_{\vec{\nu}} B^2_{\vec{\nu}}$,
4197: for observations made while scanning
4198: parallel to RA for average subtraction and
4199: quadratic subtraction.
4200: % with each contour representing 0.1.
4201: % The plot on the left shows average sky subtraction and the
4202: % plot on the right shows quadratic sky subtraction.
4203: % Rotating these transfer functions by 90 degrees
4204: % gives the transfer functions for observations made while scanning
4205: % parallel to dec.
4206: % Each transfer function has been multiplied
4207: % by the effective transfer function of the beam, which
4208: % attenuates the signal at high-$\vec{\nu}$.
4209: Note the large amount of attenuation at low frequencies along
4210: the scan direction and at scales larger than the focal plane
4211: size of approximately 500~radians$^{-1}$.}
4212: \label{fig:RA_xfer}
4213: \end{figure}
4214:
4215: \clearpage
4216: \begin{figure}
4217: \plottwo{f9a.eps}{f9b.eps}
4218: % \centering \includegraphics[width=.7\textwidth]{f9a.eps}
4219: % \centering \includegraphics[width=.7\textwidth]{f9b.eps}
4220: \caption{Maps of the science fields.
4221: Note that the astronomical signal in
4222: each map has been convolved with the
4223: transfer functions of the data processing
4224: and the beam, $\sqrt{TB^2}$, but the noise
4225: has not been filtered in any way.
4226: The RMS of these unfiltered maps
4227: %JS 2008/07/31 JS addressing referee comment 5
4228: is approximately 90~$\mu$K$_{CMB}$ per beam,
4229: and the RMS after optimally filtering
4230: for point sources is $\simeq 70$~$\mu$K$_{CMB}$
4231: per beam.}
4232: \label{fig:final_maps}
4233: \end{figure}
4234:
4235: \clearpage
4236: \begin{figure}
4237: \plottwo{f10a.eps}{f10b.eps}
4238: % \centering \includegraphics[width=.65\textwidth]{f10a.eps}
4239: % \centering \includegraphics[width=.65\textwidth]{f10b.eps}
4240: \caption{The map-space PSDs, $P_{\vec{\nu}}$,
4241: of the maps made from co-adding
4242: all observations for a each science field.
4243: % The plot on the left shows the Lynx field, and
4244: % the plot on the right shows the SDS1 field.
4245: Note that relative to these PSDs,
4246: the power spectra of
4247: any astronomical signals will have been multiplied by
4248: the transfer functions of the data processing and
4249: the beam, $T_{\vec{\nu}} B^2_{\vec{\nu}}$.}
4250: \label{fig:final_psd}
4251: \end{figure}
4252:
4253: \clearpage
4254: \begin{figure}
4255: \plottwo{f11a.eps}{f11b.eps}
4256: % \centering \includegraphics[width=.5\textwidth]{f11a.eps}
4257: % \centering \includegraphics[width=.5\textwidth]{f11b.eps}
4258: \caption{Contour plots showing the
4259: transfer functions, $T_{\vec{\nu}} B^2_{\vec{\nu}}$,
4260: for the maps made from all observations
4261: of each science field.
4262: % The astronomical signal is attenuated both by our processing
4263: % and map-making algorithms and by the Bolocam beam.
4264: % Each contour line represents 0.1.
4265: There is slightly more attenuation along the
4266: $\nu_{RA}$ axis compared to the $\nu_{dec}$ axis in the
4267: maps because more observations were taken while scanning
4268: parallel to RA compared to scanning parallel to dec.}
4269: % This asymmetry between RA and dec scans
4270: % was caused by an observing error near the
4271: % start of the season.}
4272: \label{fig:final_xfer}
4273: \end{figure}
4274:
4275: \clearpage
4276: \begin{figure}
4277: \plotone{f12a.eps}
4278: \plotone{f12b.eps}
4279: % \centering \includegraphics[width=.75\textwidth]{f12a.eps}
4280: % \centering \includegraphics[width=.75\textwidth]{f12b.eps}
4281: \caption{A comparison between the distribution of PSD values
4282: from the jackknifed realizations to
4283: a Gaussian PDF for the data
4284: co-added over all observations for each
4285: Science field.
4286: See Equation~\ref{eqn:PDF_map_PSD}.
4287: The agreement is good, indicating
4288: that the underlying noise distribution is
4289: approximately Gaussian.}
4290: \label{fig:final_PSD_PDF}
4291: \end{figure}
4292:
4293: \clearpage
4294: \begin{figure}
4295: \plotone{f13.eps}
4296: % \includegraphics[width=\textwidth]{f13.eps}
4297: \caption{The power spectra from the primary CMB anisotropies
4298: (short green dashes),
4299: high and low estimates for
4300: radio point sources (red dash-dot),
4301: high and low estimates for submillimeter
4302: point sources (blue dot-dot-dot-dash), and the
4303: analytically predicted
4304: SZE-induced CMB anisotropies from \citet{komatsu02} using the best fit
4305: value of $\sigma_8$ from \citet{dawson06}
4306: (long orange dashes).
4307: Note that the point-source power spectra assume
4308: unclustered distributions.
4309: Also included as a solid black line
4310: is the transfer function of the final
4311: map of the Lynx field
4312: with arbitrary normalization.}
4313: \label{fig:astr_noise}
4314: \end{figure}
4315:
4316: \clearpage
4317: \begin{figure}
4318: \plottwo{f14a.eps}{f14b.eps}
4319: % \centering \includegraphics[width=.65\textwidth]{f14a.eps}
4320: % \centering \includegraphics[width=.65\textwidth]{f14b.eps}
4321: \caption{The Bayesian likelihood given by Equation~\ref{eqn:log_l}
4322: for each science field.
4323: The likelihoods have all been normalized to one at
4324: the peak.
4325: These plots should only be considered as rough estimates
4326: for determining confidence intervals
4327: because the cosmic variance of the CMB
4328: spectra, correlations among map pixels, and the physical
4329: boundary that the anisotropy amplitude must be greater
4330: than or equal to zero have not been fully accounted for
4331: in the likelihood function.}
4332: \label{fig:liklihood}
4333: \end{figure}
4334:
4335: \clearpage
4336: \begin{figure}
4337: \plottwo{f15a.eps}{f15b.eps}
4338:
4339: \plottwo{f15c.eps}{f15d.eps}
4340: % \includegraphics[width=.5\textwidth]{f15a.eps}
4341: % \includegraphics[width=.5\textwidth]{f15b.eps}
4342: % \includegraphics[width=.5\textwidth]{f15c.eps}
4343: % \includegraphics[width=.5\textwidth]{f15d.eps}
4344: \caption{The first plot shows the Bayesian likelihood for
4345: a range of anisotropy amplitudes for the full data
4346: set, which includes all of the observations
4347: of both science fields.
4348: The remaining three plots show the frequentist
4349: confidence belts for the full data set for confidence
4350: levels of 68\%, 90\%, and 95\%.}
4351: \label{fig:full_conf_belts}
4352: \end{figure}
4353:
4354: \clearpage
4355: \begin{figure}
4356: \plotone{f16.eps}
4357: % \includegraphics[width=\textwidth]{f16.eps}
4358: \caption{The band power window function, $W_{\ell}^B/\ell$,
4359: for the full data set.
4360: We have arbitrarily peak normalized the window function.}
4361: \label{fig:knox}
4362: \end{figure}
4363:
4364: \clearpage
4365: \begin{figure}
4366: \plotone{f17.eps}
4367: % \resizebox{\columnwidth}{!}{
4368: % \includegraphics{f17.eps}}
4369: \caption{A plot of all of the current CMB anisotropy measurements
4370: above $\ell = 2000$.
4371: Solid lines represent observations made near 150~GHz,
4372: and dashed lines represent observations made near
4373: 30~GHz.
4374: The primary CMB anisotropies are
4375: represented by a solid black line on
4376: the left side of the plot, and the predicted SZE-induced CMB
4377: anisotropies are shown as solid (150~GHz) and
4378: dashed (30~GHz) black lines.
4379: The primary CMB anisotropies were calculated using
4380: the parameters given in Section~\ref{sec:astr_noise};
4381: the analytic model of \citet{komatsu02},
4382: along with the best estimate of $\sigma_8$ from
4383: \citet{dawson06}, were used to estimate
4384: the SZE-induced CMB anisotropies.
4385: % Note that the best fit value of $\sigma_8$ from
4386: % the WMAP 5-year data at lower $\ell$ produces
4387: % a SZE-induced CMB power spectrum
4388: % with an amplitude of less than 10~$\mu$K$_{CMB}^2$
4389: % at 150~GHz~\citep{dunkley08}.
4390: All of the data are plotted with 1$\sigma$ error bars,
4391: except for the Bolocam upper limit at $\ell=5700$ and
4392: the BIMA upper limit at $\ell=8748$, which are given
4393: as 90\% and 95\% confidence level upper limits, respectively.
4394: The ACBAR data were taken from \citet{reichardt08},
4395: the BIMA data were taken from \citet{dawson06},
4396: and the CBI data were taken from \citet{mason03}.}
4397: % Note that the Bolocam 68\% and 95\% CL upper limits
4398: % do not appear much different from the 90\%
4399: % CL upper limit for the scale of the logarithmic plot above.}
4400: \label{fig:cmb_power}
4401: \end{figure}
4402:
4403: %\clearpage
4404: %\begin{figure}
4405: % \resizebox{\textwidth}{!}{
4406: % \includegraphics*[.3in,.2in][2.3in,1.6in]{komatsu02_modified2.ps}}
4407: % \caption{SZE spectra calculated from various MMH or SPH
4408: % simulations (dashed lines)
4409: % \citep{zhang02, seljak01, refregier00, refregier00_2,
4410: % springel01, dasilva01} and from the
4411: % analytic model given by \citet{komatsu02} (solid black lines).
4412: % Each of the simulations was run with slightly different
4413: % input parameters, and an analytic spectrum was
4414: % calculated for each set of parameters.
4415: % All of the spectra have been scaled
4416: % by $\sigma_8^7 (\Omega_bh)^2$, since this
4417: % combination of parameters approximates
4418: % the amplitude of the spectra to good precision.
4419: % The highlighted region between $\ell = 4000$
4420: % and $\ell = 7000$ indicates the range
4421: % of angular scales Bolocam is most sensitive to,
4422: % and most of the spectra are reasonably
4423: % flat within this region.
4424: % Figure adapted from \citet{komatsu02}.}
4425: % \label{fig:sim_sz_spectra}
4426: %\end{figure}
4427:
4428:
4429:
4430: \end{document}
4431:
4432: