1:
2: % WMAP Pass-3 Overview
3:
4: %\documentclass[10pt,preprint]{aastex}
5: \documentclass[12pt,preprint]{aastex}
6: \usepackage{natbib}
7: %\input MAP_cosmo
8: \singlespace
9:
10: \shorttitle{\WMAP\ 5-year Overview}
11: \shortauthors{Hinshaw et al.}
12:
13: % definitions:
14: \newcommand{\map} {{\sl WMAP}}
15: \newcommand{\cobe} {{\sl COBE}}
16: \newcommand{\n} {{\bf{n}}}
17: \newcommand{\EV}[1] {\langle#1\rangle}
18: \newcommand{\lmax} {l_{\rm max}}
19: \newcommand{\mmax} {m_{\rm max}}
20: \newcommand{\uKsq} {\mbox{$\mu{\rm K}^2$}}
21: \newcommand{\dg} {\mbox{$^{\circ}$}}
22: \newcommand{\lsim} {\mbox{$_<\atop^{\sim}$}}
23: \newcommand{\gsim} {\mbox{$_>\atop^{\sim}$}}
24: \newcommand{\lt} {\mbox{$<$}}
25: \newcommand{\gt} {\mbox{$>$}}
26: \newcommand{\order} {{\cal O}}
27: \newcommand{\JJ} {{\cal J}}
28: \newcommand{\<} {\langle}
29: \renewcommand{\>} {\rangle}
30: \newcommand{\amin} {\mbox{$^\prime\ $}}
31: \newcommand{\asec} {\mbox{$^{\prime\prime}\ $}}
32: \newcommand{\ddeg} {\mbox{${\rlap.}^\circ$}}
33: \newcommand{\beq} {\begin{equation}}
34: \newcommand{\eeq} {\end{equation}}
35: \newcommand{\beqa} {\begin{eqnarray}}
36: \newcommand{\eeqa} {\end{eqnarray}}
37: \newcommand{\ylm}[2] {{Y_{#1#2}}}
38: \newcommand{\ylmcc}[2]{{Y^{*}_{#1#2}}}
39: \newcommand{\fnlKS} {f_{NL}^{\rm local}}
40: \newcommand{\fnleq} {f_{NL}^{\rm equil}}
41:
42: %% references
43: \newcommand{\refeqnp}[1]{(eq.~[\ref{#1}])}
44: \newcommand{\refeqnt}[1]{{equation~(\ref{#1})}}
45: \newcommand{\refeqnT}[1]{{Equation~(\ref{#1})}}
46: \newcommand{\reffigp}[1]{(Fig.~\ref{#1})}
47: \newcommand{\reffigt}[1]{{Figure~\ref{#1}}}
48:
49: \slugcomment{Astrophysical Journal Supplement Series, in press}
50:
51: \begin{document}
52:
53: \title{Five-Year Wilkinson Microwave Anisotropy Probe
54: (WMAP\altaffilmark{1}) Observations:\\
55: Data Processing, Sky Maps, \& Basic Results}
56:
57: \author{{G. Hinshaw} \altaffilmark{2},
58: {J. L. Weiland} \altaffilmark{3},
59: {R. S. Hill} \altaffilmark{3},
60: {N. Odegard} \altaffilmark{3},
61: {D. Larson} \altaffilmark{4},
62: {C. L. Bennett} \altaffilmark{4},
63: {J. Dunkley} \altaffilmark{5,6,7},
64: {B. Gold} \altaffilmark{4},
65: {M. R. Greason} \altaffilmark{3},
66: {N. Jarosik} \altaffilmark{5},
67: {E. Komatsu} \altaffilmark{8},
68: {M. R. Nolta} \altaffilmark{9},
69: {L. Page} \altaffilmark{5},
70: {D. N. Spergel} \altaffilmark{6,10},
71: {E. Wollack} \altaffilmark{2},
72: {M. Halpern} \altaffilmark{11},
73: {A. Kogut} \altaffilmark{2},
74: {M. Limon} \altaffilmark{12},
75: {S. S. Meyer} \altaffilmark{13},
76: {G. S. Tucker} \altaffilmark{14},
77: {E. L. Wright} \altaffilmark{15}}
78:
79: \altaffiltext{1}{\map\ is the result of a partnership between Princeton
80: University and NASA's Goddard Space Flight Center. Scientific
81: guidance is provided by the \map\ Science Team.}
82: \altaffiltext{2}{{Code 665, NASA/Goddard Space Flight Center, Greenbelt, MD 20771}}
83: \altaffiltext{3}{{Adnet Systems, Inc., 7515 Mission Dr., Suite A100, Lanham, Maryland 20706}}
84: \altaffiltext{4}{{Dept. of Physics \& Astronomy, The Johns Hopkins University, 3400 N. Charles St., Baltimore, MD 21218-2686}}
85: \altaffiltext{5}{{Dept. of Physics, Jadwin Hall, Princeton University, Princeton, NJ 08544-0708}}
86: \altaffiltext{6}{{Dept. of Astrophysical Sciences, Peyton Hall, Princeton University, Princeton, NJ 08544-1001}}
87: \altaffiltext{7}{{Astrophysics, University of Oxford, Keble Road, Oxford, OX1 3RH, UK}}
88: \altaffiltext{8}{{Univ. of Texas, Austin, Dept. of Astronomy, 2511 Speedway, RLM 15.306, Austin, TX 78712}}
89: \altaffiltext{9}{{Canadian Institute for Theoretical Astrophysics, 60 St. George St, University of Toronto, Toronto, ON Canada M5S 3H8}}
90: \altaffiltext{10}{{Princeton Center for Theoretical Physics, Princeton University, Princeton, NJ 08544}}
91: \altaffiltext{11}{{Dept. of Physics and Astronomy, University of British Columbia, Vancouver, BC Canada V6T 1Z1}}
92: \altaffiltext{12}{{Columbia Astrophysics Laboratory, 550 W. 120th St., Mail Code 5247, New York, NY 10027-6902}}
93: \altaffiltext{13}{{Depts. of Astrophysics and Physics, KICP and EFI, University of Chicago, Chicago, IL 60637}}
94: \altaffiltext{14}{{Dept. of Physics, Brown University, 182 Hope St., Providence, RI 02912-1843}}
95: \altaffiltext{15}{{UCLA Physics \& Astronomy, PO Box 951547, Los Angeles, CA 90095-1547}}
96:
97: \email{Gary.F.Hinshaw@nasa.gov}
98:
99: \begin{abstract}
100:
101: We present new full-sky temperature and polarization maps in five frequency
102: bands from 23 to 94 GHz, based on data from the first five years of the \map\
103: sky survey. The new maps are consistent with previous maps and are more
104: sensitive. The five-year maps incorporate several improvements in data
105: processing made possible by the additional years of data and by a more complete
106: analysis of the instrument calibration and in-flight beam response. We present
107: several new tests for systematic errors in the polarization data and conclude
108: that W band polarization data is not yet suitable for cosmological studies, but
109: we suggest directions for further study. We {\em do} find that Ka band data is
110: suitable for use; in conjunction with the additional years of data, the addition
111: of Ka band to the previously used Q and V band channels significantly reduces
112: the uncertainty in the optical depth parameter, $\tau$. Further scientific
113: results from the five year data analysis are presented in six companion papers
114: and are summarized in \S\ref{sec:science} of this paper.
115:
116: With the 5 year \map\ data, we detect no convincing deviations from the minimal
117: 6-parameter $\Lambda$CDM model: a flat universe dominated by a cosmological
118: constant, with adiabatic and nearly scale-invariant Gaussian fluctuations.
119: Using \map\ data combined with measurements of Type Ia supernovae (SN) and
120: Baryon Acoustic Oscillations (BAO) in the galaxy distribution, we find (68\% CL
121: uncertainties):
122: \ensuremath{\Omega_bh^2 = 0.02267^{+ 0.00058}_{- 0.00059}},
123: \ensuremath{\Omega_ch^2 = 0.1131\pm 0.0034},
124: \ensuremath{\Omega_\Lambda = 0.726\pm 0.015},
125: \ensuremath{n_s = 0.960\pm 0.013},
126: \ensuremath{\tau = 0.084\pm 0.016}, and
127: \ensuremath{\Delta_{\cal R}^2 = (2.445\pm 0.096)\times 10^{-9}} at $k=0.002~{\rm Mpc^{-1}}$.
128: From these we derive:
129: \ensuremath{\sigma_8 = 0.812\pm 0.026},
130: \ensuremath{H_0 = 70.5\pm 1.3}~${\rm
131: km~s^{-1}~Mpc^{-1}}$,
132: \ensuremath{\Omega_b = 0.0456\pm 0.0015},
133: \ensuremath{\Omega_c = 0.228\pm 0.013},
134: \ensuremath{\Omega_mh^2 = 0.1358^{+ 0.0037}_{- 0.0036}},
135: \ensuremath{z_{\rm reion} = 10.9\pm 1.4},
136: and
137: \ensuremath{t_0 = 13.72\pm 0.12\ \mbox{Gyr}}.
138: The new limit on the tensor-to-scalar ratio is
139: \ensuremath{r < 0.22\ \mbox{(95\% CL)}},
140: while the evidence for a running spectral index is insignificant,
141: \ensuremath{dn_s/d\ln{k} = -0.028\pm 0.020} (68\% CL).
142: We obtain tight, simultaneous limits on the (constant)
143: dark energy equation of state and the spatial curvature of the
144: universe:
145: \ensuremath{-0.14<1+w<0.12\ \mbox{(95\% CL)}}
146: and
147: \ensuremath{-0.0179<\Omega_k<0.0081\ \mbox{(95\% CL)}}.
148: The number of relativistic degrees of freedom, expressed in units of the
149: effective number of neutrino species, is found to be
150: \ensuremath{N_{\rm eff} = 4.4\pm 1.5} (68\% CL),
151: consistent with the standard value of 3.04. Models with $N_{\rm eff} = 0$ are
152: disfavored at $\gt$99.5\% confidence. Finally, new limits on physically
153: motivated primordial non-Gaussianity parameters are $-9 < \fnlKS <111$ (95\% CL)
154: and $-151 < \fnleq < 253$ (95\% CL) for the local and equilateral models,
155: respectively.
156:
157: \end{abstract}
158:
159: \keywords{cosmic microwave background, cosmology: observations, early universe,
160: dark matter, space vehicles, space vehicles: instruments,
161: instrumentation: detectors, telescopes}
162:
163:
164: \section{INTRODUCTION}\label{sec:intro}
165:
166: The Wilkinson Microwave Anisotropy Probe (\map) is a Medium-Class Explorer
167: (MIDEX) satellite aimed at elucidating cosmology through full-sky observations
168: of the cosmic microwave background (CMB). The \map\ full-sky maps of the
169: temperature and polarization anisotropy in five frequency bands provide our most
170: accurate view to date of conditions in the early universe. The multi-frequency
171: data facilitate the separation of the CMB signal from foreground emission
172: arising both from our Galaxy and from extragalactic sources. The CMB angular
173: power spectrum derived from these maps exhibits a highly coherent acoustic peak
174: structure which makes it possible to extract a wealth of information about the
175: composition and history of the universe, as well as the processes that seeded
176: the fluctuations.
177:
178: \map\ data \citep{bennett/etal:2003, spergel/etal:2003, hinshaw/etal:2007,
179: spergel/etal:2007}, along with a host of pioneering CMB experiments
180: \citep{miller/etal:1999, lee/etal:2001, netterfield/etal:2002,
181: halverson/etal:2002, pearson/etal:2003, scott/etal:2003, benoit/etal:2003b}, and
182: other cosmological measurements \citep{percival/etal:2001, tegmark/etal:2004a,
183: cole/etal:2005, tegmark/etal:2006, eisenstein/etal:2005, percival/etal:2007c,
184: astier/etal:2006, riess/etal:2007, wood-vasey/etal:2007} have established
185: $\Lambda$CDM as the standard model of cosmology: a flat universe dominated by
186: dark energy, supplemented by dark matter and atoms with density fluctuations
187: seeded by a Gaussian, adiabatic, nearly scale invariant process. The basic
188: properties of this universe are determined by five numbers: the density of
189: matter, the density of atoms, the age of the universe (or equivalently, the
190: Hubble constant today), the amplitude of the initial fluctuations, and their
191: scale dependence.
192:
193: By accurately measuring the first few peaks in the angular power spectrum and
194: the large-scale polarization anisotropy, \map\ data have enabled the following
195: inferences:
196: \begin{itemize}
197:
198: \item A precise (3\%) determination of the density of atoms in the universe. The
199: agreement between the atomic density derived from \map\ and the density inferred
200: from the deuterium abundance is an important test of the standard big bang
201: model.
202:
203: \item A precise (3\%) determination of the dark matter density. (With five
204: years of data and a better determination of our beam response, this measurement
205: has improved significantly.) Previous CMB measurements have shown that the dark
206: matter must be non-baryonic and interact only weakly with atoms and radiation.
207: The \map\ measurement of the density puts important constraints on
208: supersymmetric dark matter models and on the properties of other dark matter
209: candidates.
210:
211: \item A definitive determination of the acoustic scale at redshift $z=1090$.
212: Similarly, the recent measurement of baryon acoustic oscillations (BAO) in the
213: galaxy power spectrum \citep{eisenstein/etal:2005} has determined the acoustic
214: scale at redshift $z \sim 0.35$. When combined, these standard rulers
215: accurately measure the geometry of the universe and the properties of the dark
216: energy. These data require a nearly flat universe dominated by dark energy
217: consistent with a cosmological constant.
218:
219: \item A precise determination of the Hubble Constant, in conjunction with BAO
220: observations. Even when allowing curvature ($\Omega_0 \ne 1$) and a free dark
221: energy equation of state ($w \ne -1$), the acoustic data determine the Hubble
222: constant to within 3\%. The measured value is in excellent agreement with
223: independent results from the Hubble Key Project \citep{freedman/etal:2001},
224: providing yet another important consistency test for the standard model.
225:
226: \item Significant constraint of the basic properties of the primordial
227: fluctuations. The anti-correlation seen in the temperature/polarization (TE)
228: correlation spectrum on 4$^{\circ}$ scales implies that the fluctuations are
229: primarily adiabatic and rule out defect models and isocurvature models as the
230: primary source of fluctuations \citep{peiris/etal:2003}.
231: \end{itemize}
232:
233: Further, the \map\ measurement of the primordial power spectrum of matter
234: fluctuations constrains the physics of inflation, our best model for the origin
235: of these fluctuations. Specifically, the 5 year data provide the best
236: measurement to date of the scalar spectrum's amplitude and slope, and place the
237: most stringent limits to date on the amplitude of tensor fluctuations. However,
238: it should be noted that these constraints assume a smooth function of scale,
239: $k$. Certain models with localized structure in $P(k)$, and hence additional
240: parameters, are not ruled out, but neither are they required by the data; see
241: e.g. \citet{shafieloo/souradeep:2007, hunt/sarkar:2007}.
242:
243: The statistical properties of the CMB fluctuations measured by \map\ are close
244: to Gaussian; however, there are several hints of possible deviations from
245: Gaussianity, e.g. \citet{eriksen/etal:2007c, copi/etal:2007,
246: land/magueijo:2007, yadav/wandelt:2008}. Significant deviations would be a
247: very important signature of new physics in the early universe.
248:
249: Large-angular-scale polarization measurements currently provide our best window
250: into the universe at $z \sim 10$. The \map\ data imply that the universe was
251: reionized long before the epoch of the oldest known quasars. By accurately
252: constraining the optical depth of the universe, \map\ not only constrains the
253: age of the first stars but also determines the amplitude of primordial
254: fluctuations to better than 3\%. This result is important for
255: constraining the growth rate of structure.
256:
257: This paper summarizes results compiled from 5 years of \map\ data that are fully
258: presented in a suite of 7 papers (including this one). The new results improve
259: upon previous results in many ways: additional data reduces the random noise,
260: which is especially important for studying the temperature signal on small
261: angular scales and the polarization signal on large angular scales; five
262: independent years of data enable comparisons and null tests that were not
263: previously possible; the instrument calibration and beam response have been much
264: better characterized, due in part to improved analyses and to additional years
265: of data; and, other cosmological data have become available.
266:
267: In addition to summarizing the other papers, this paper reports on changes in
268: the \map\ data processing pipeline, presents the 5 year temperature and
269: polarization maps, and gives new results on instrument calibration and on
270: potential systematic errors in the polarization data. \citet{hill/etal:prep}
271: discuss the program to derive an improved physical optics model of the \map\
272: telescope, and use the results to better determine the \map\ beam response.
273: \citet{gold/etal:prep} present a new analysis of diffuse foreground emission in
274: the \map\ data and update previous analyses using 5 year data.
275: \citet{wright/etal:prep} analyze extragalactic point sources and provide an
276: updated source catalog, with new results on source variability.
277: \citet{nolta/etal:prep} derive the angular power spectra from the maps,
278: including the TT, TE, TB, EE, EB, and BB spectra. \citet{dunkley/etal:prep}
279: produce an updated likelihood function and present cosmological parameter
280: results based on 5 year \map\ data. They also develop an independent analysis
281: of polarized foregrounds and use those results to test the reliability of the
282: optical depth inference to foreground removal errors. \citet{komatsu/etal:prep}
283: infer cosmological parameters by combining 5 year \map\ data with a host of
284: other cosmological data and discuss the implications of the results. Concurrent
285: with the submission of these papers, all 5 year \map\ data are made available to
286: the research community via NASA's Legacy Archive for Microwave Background Data
287: Analysis (LAMBDA). The data products are described in detail in the \map\
288: Explanatory Supplement \citep{limon/etal:prep}, also available on LAMBDA.
289:
290: The \map\ instrument is composed of 10 differencing assemblies (DAs) spanning 5
291: frequencies from 23 to 94 GHz \citep{bennett/etal:2003}: 1 DA each at 23 GHz
292: (K1) and 33 GHz (Ka1), 2 each at 41 GHz (Q1,Q2) and 61 GHz (V1,V2), and 4 at 94
293: GHz (W1-W4). Each DA is formed from two differential radiometers which are
294: sensitive to orthogonal linear polarization modes; the radiometers are
295: designated 1 or 2 (e.g., V11 or W12) depending on polarization mode.
296:
297: In this paper we follow the notation convention that flux density is $S\sim
298: \nu^\alpha$ and antenna temperature is $T\sim \nu^\beta$, where the spectral
299: indices are related by $\beta=\alpha-2$. In general, the CMB is expressed in
300: terms of thermodynamic temperature, while Galactic and extragalactic foregrounds
301: are expressed in antenna temperature. Thermodynamic temperature differences
302: are given by $\Delta T = \Delta T_A [(e^x-1)^2/x^2e^x]$, where $x=h\nu/kT_0$,
303: $h$ is the Planck constant, $\nu$ is the frequency, $k$ is the Boltzmann
304: constant, and $T_0=2.725$ K is the CMB temperature \citep{mather/etal:1999}.
305: A \map\ band-by-band tabulation of the conversion factors between thermodynamic
306: and antenna temperature is given in Table~\ref{tab:radiometers}.
307:
308:
309: \section{CHANGES IN THE 5 YEAR DATA ANALYSIS}
310: \label{sec:change}
311:
312: The 1 year and 3 year data analyses were described in detail in previous
313: papers. In large part, the 5 year analysis employs the same methods, so we do
314: not repeat a detailed processing description here. However, we have made
315: several improvements that are summarized here and described in more detail later
316: in this paper and in a series of companion papers, as noted. We list the
317: changes in the order they appear in the processing pipeline:
318:
319: \begin{itemize}
320:
321: \item There is a $\sim 1'$ temperature-dependent pointing offset between the
322: star tracker coordinate system (which defines spacecraft coordinates) and the
323: instrument boresights. In the 3 year analysis we introduced a correction to
324: account for the elevation change of the instrument boresights in spacecraft
325: coordinates. With additional years of data, we have been able to refine our
326: thermal model of the pointing offset, so we now include a small ($\lt 1'$)
327: correction to account for the azimuth change of the instrument boresights.
328: Details of the new correction are given in the 5 year Explanatory Supplement
329: \citep{limon/etal:prep}.
330:
331: \item We have critically re-examined the relative and absolute intensity
332: calibration procedures, paying special attention to the absolute gain recovery
333: obtainable from the modulation of the CMB dipole due to \map's motion. We
334: describe the revised procedure in \S\ref{sec:cal_improve} and note that the sky
335: map calibration uncertainty has decreased from 0.5\% to 0.2\%.
336:
337: \item The \map\ beam response has now been measured in 10 independent
338: ``seasons'' of Jupiter observations. In the highest resolution W band channels,
339: these measurements now probe the beam response $\sim$44 dB down from the beam
340: peak. However, there is still non-negligible beam solid angle below this level
341: ($\sim$0.5\%) that needs to be measured to enable accurate cosmological
342: inference. In the 3 year analysis we produced a physical optics model of the
343: A-side beam response starting with a pre-flight model and fitting in-flight
344: mirror distortions to the flight Jupiter data. In the 5 year analysis we have
345: extended the model to the B-side optics and, for both sides, we have extended
346: the fit to include distortion modes a factor of 2 smaller in linear scale (4
347: times as many modes). The model is used to augment the flight beam maps below a
348: given threshold. The details of this work are given in \citet{hill/etal:prep}.
349:
350: \item The far sidelobe response of the beam was determined from a combination of
351: ground measurements and in-flight lunar data taken early in the mission
352: \citep{barnes/etal:2003}. For the current analysis, we have replaced a small
353: fraction of the far sidelobe data with the physical optics model described
354: above. We have also made the following changes in our handling of the far
355: sidelobe pickup \citep{hill/etal:prep}: 1) We have enlarged the ``transition
356: radius'' that defines the boundary between the main beam and the far sidelobe
357: response. This places a larger fraction of the total beam solid angle in the
358: main beam where uncertainties are easier to quantify and propagate into the
359: angular power spectra. 2) We have moved the far sidelobe deconvolution into the
360: combined calibration and sky map solver (\S\ref{sec:cal_improve}). This
361: produces a self-consistent estimate of the intensity calibration and the
362: deconvolved sky map. The calibrated time-ordered data archive has had an
363: estimate of the far sidelobe response subtracted from each datum (as it had in
364: the 3 year processing).
365:
366: \item We have updated the optimal filters used in the final step of map-making.
367: The functional form of the filter is unchanged \citep{jarosik/etal:2007}, but
368: the fits have been updated to cover years 4 and 5 of the flight data.
369:
370: \item Each \map\ differencing assembly consists of two radiometers that are
371: sensitive to orthogonal linear polarization states. The sum and difference of
372: the two radiometer channels split the signal into intensity and polarization
373: components, respectively. However, the noise levels in the two radiometers are
374: not equal, in general, so more optimal sky map estimation is possible in theory,
375: at the cost of mixing intensity and polarization components in the process. For
376: the current analysis, we investigated one such weighted algorithm and found that
377: the polarization maps were subject to unacceptable contamination by the
378: intensity signal in cases where the beam response was non-circular and the
379: gradient of the intensity signal was large, e.g., in K band. As a result, we
380: reverted to the unweighted (and unbiased) estimator used in previous work.
381:
382: \item We have improved the sky masks used to reject foreground contamination.
383: In previous work, we defined masks based on contours of the K band data. In the
384: 5 year analysis we produce masks based jointly on K band and Q band contours.
385: For a given sky cut fraction, the new masks exclude flat spectrum (e.g.
386: free-free) emission more effectively. The new masks are described in detail in
387: \citet{gold/etal:prep} and are provided with the 5 year data release. In
388: addition, we have modified the ``processing'' mask used to exclude very bright
389: sources during sky map estimation. The new mask is defined in terms of
390: low-resolution (r4) HEALPix sky pixels \citep{gorski/etal:2005} to facilitate a
391: cleaner definition of the pixel-pixel inverse covariance matrices, $N^{-1}$.
392: One side effect of this change is to introduce a few r4-sized holes around the
393: brightest radio sources in the analysis mask, which incorporates the processing
394: mask as a subset.
395:
396: \item We have amended our foreground analysis in the following ways: 1)
397: \citet{gold/etal:prep} perform a pixel-by-pixel analysis of the joint
398: temperature and polarization data to study the breakdown of the Galactic
399: emission into physical components. 2) We have updated some aspects of the
400: Maximum Entropy (MEM) based analysis, as described in \citet{gold/etal:prep}.
401: 3) \citet{dunkley/etal:prep} develop a new analysis of polarized foreground
402: emission using a Gibbs sampling approach that yields a cleaned CMB polarization
403: map and an associated covariance matrix. 4) \citet{wright/etal:prep} update the
404: \map\ point source catalog and present some results on variable sources in the 5
405: year data. However, the basic cosmological results are still based on maps that
406: were cleaned with the same template-based procedure that was used in the 3 year
407: analysis.
408:
409: \item We have improved the final temperature power spectrum, $C_l^{TT}$, by
410: using a Gibbs-based maximum likelihood estimate for $l \le 32$
411: \citep{dunkley/etal:prep} and a pseudo-$C_l$ estimate for higher $l$
412: \citep{nolta/etal:prep}. As with the 3 year analysis, the pseudo-$C_l$ estimate
413: uses only V- and W-band data. With 5 individual years of data and six V- and
414: W-band differencing assemblies, we can now form individual cross-power spectra
415: from 15 DA pairs within each of 5 years and from 36 DA pairs across 10 year
416: pairs, for a total of 435 independent cross-power spectra.
417:
418: \item In the 3 year analysis we developed a pseudo-$C_l$ method for evaluating
419: polarization power spectra in the presence of correlated noise. In the present
420: analysis we additionally estimate the TE, TB, EE, EB, \& BB spectra and their
421: errors using an extension of the maximum likelihood method in
422: \citet{page/etal:2007}. However, as in the 3 year analysis, the likelihood of a
423: given model is still evaluated directly from the polarization maps using a
424: pixel-based likelihood.
425:
426: \item We have improved the form of the likelihood function used to infer
427: cosmological parameters from the Monte Carlo Markov Chains
428: \citep{dunkley/etal:prep}. We use an exact maximum likelihood form for the $l
429: \le 32$ TT data \citep{eriksen/etal:2007}. We have investigated theoretically
430: optimal methods for incorporating window function uncertainties into the
431: likelihood, but in tests with simulated data we have found them to be biased.
432: In the end, we adopt the form used in the 3 year analysis
433: \citep{hinshaw/etal:2007}, but we incorporate the smaller 5 year window function
434: uncertainties \citep{hill/etal:prep} as inputs. We now routinely account for
435: gravitational lensing when assessing parameters, and we have added an option to
436: use low-$l$ TB and EB data for testing non-standard cosmological models.
437:
438: \item For testing nongaussianity, we employ an improved estimator for $f_{NL}$
439: \citep{creminelli/etal:2006, yadav/etal:prep}. The results of this analysis are
440: described in \citet{komatsu/etal:prep}.
441:
442: \end{itemize}
443:
444:
445: \section{OBSERVATIONS AND MAPS}
446: \label{sec:maps}
447:
448: The 5 year \map\ data encompass the period from 00:00:00 UT, 10 August 2001 (day
449: number 222) to 00:00:00 UT, 9 August 2006 (day number 222). The observing
450: efficiency during this time is roughly 99\%; Table~\ref{tab:baddata} lists the
451: fraction of data that was lost or rejected as unusable. The Table also gives
452: the fraction of data that is flagged due to potential contamination by thermal
453: emission from Mars, Jupiter, Saturn, Uranus, and Neptune. These data are not
454: used in map-making, but are useful for in-flight beam mapping
455: \citep{hill/etal:prep, limon/etal:prep}.
456:
457: After performing an end-to-end analysis of the instrument calibration,
458: single-year sky maps are created from the time-ordered data using the procedure
459: described by \citet{jarosik/etal:2007}. Figure~\ref{fig:i_maps} shows the 5
460: year temperature maps at each of the five \map\ observing frequencies: 23, 33,
461: 41, 61, and 94 GHz. The number of independent observations per pixel, $N_{\rm
462: obs}$, is qualitatively the same as Figure~2 of \citet{hinshaw/etal:2007} and is
463: not reproduced here. The noise per pixel, $p$, is given by $\sigma(p) =
464: \sigma_0 N_{\rm obs}^{-1/2}(p)$, where $\sigma_0$ is the noise per observation,
465: given in Table~\ref{tab:radiometers}. To a very good approximation, the noise
466: per pixel in the 5 year maps is a factor of $\sqrt{5}$ times lower than in the
467: single-year maps. Figures~\ref{fig:q_maps} and \ref{fig:u_maps} show the 5 year
468: polarization maps in the form of the Stokes parameters Q and U, respectively.
469: Maps of the relative polarization sensitivity, the Q and U analogs of $N_{\rm
470: obs}$, are shown in Figure~13 of \citet{jarosik/etal:2007} and are not updated
471: here. A description of the low-resolution pixel-pixel inverse covariance
472: matrices used in the polarization analysis is also given in
473: \citet{jarosik/etal:2007}, and is not repeated here. The polarization maps are
474: dominated by foreground emission, primarily synchrotron emission from the Milky
475: Way. Figure~\ref{fig:p_maps} shows the polarization maps in a form in which the
476: color scale represents polarized intensity, $P = \sqrt{Q^2+U^2}$, and the line
477: segments indicate polarization direction for pixels with a signal-to-noise ratio
478: greater than 1. As with the temperature maps, the noise per pixel in the 5 year
479: polarization maps is $\sqrt{5}$ times lower than in the single-year maps.
480:
481: Figure~\ref{fig:diff_maps_i_recal} shows the difference between the 5 year
482: temperature maps and the corresponding 3 year maps. All maps have been smoothed
483: to 2$^{\circ}$ resolution to minimize the noise difference between them (due to
484: the additional years of data). The left column shows the difference without any
485: further processing, save for the subtraction of a relative offset between the
486: maps. Table~\ref{tab:i_diff} gives the value of the relative offset in each
487: band. Recall that \map\ is insensitive to absolute temperature, so we adopt a
488: convention that sets the zero level in each map based on a model of the
489: foreground emission at the galactic poles. While we have not changed
490: conventions, our 3 year estimate was erroneous due to the use of a preliminary
491: CMB signal map at the time the estimate was made. This error did not affect any
492: cosmological results, but it probably explains the offset differences noted by
493: \citet{eriksen/etal:prep} in their recent analysis of the 3 year data.
494:
495: The dominant structure in the left column of Figure~\ref{fig:diff_maps_i_recal}
496: consists of a residual dipole and galactic plane emission. This reflects the
497: updated 5 year calibration which has produced changes in the gain of order 0.3\%
498: compared to the 3 year gain estimate (see \S\ref{sec:cal_improve} for a more
499: detailed discussion of the calibration). Table~\ref{tab:i_diff} gives the
500: dipole amplitude difference in each band, along with the much smaller quadrupole
501: and octupole power difference. (For comparison, we estimate the CMB power at
502: $l=2,3$ to be $l(l+1)C_l / 2\pi = 211, 1041$ $\mu$K$^2$, respectively.) The
503: right column of Figure~\ref{fig:diff_maps_i_recal} shows the corresponding sky
504: map differences after the 3 year map has been rescaled by a single factor (in
505: each band) to account for the mean gain change between the 3 and 5 year
506: calibration determinations. The residual galactic plane structure in these maps
507: is less than 0.2\% of the nominal signal in Q band, and less than 0.1\% in all
508: the other bands. The large scale structure in the band-averaged temperature
509: maps is quite robust.
510:
511: \subsection{CMB Dipole}
512: \label{sec:dipole}
513:
514: The dipole anisotropy stands apart from the rest of the CMB signal due to its
515: large amplitude and to the understanding that it arises from our peculiar motion
516: with respect to the CMB rest frame. In this section we present CMB dipole
517: results based on a new analysis of the 5 year sky maps. Aside from an absolute
518: calibration uncertainty of 0.2\% (see \S\ref{sec:cal_improve}), the dominint
519: source of uncertainty in the dipole estimate arises from uncertainties in
520: Galactic foreground subtraction. Here we present results for two different
521: removal methods: template-based cleaning and an internal linear combination
522: (ILC) of the \map\ multifrequency data \citep{gold/etal:prep}. Our final
523: results are based on a combination of these methods with uncertainties that
524: encompass both approaches.
525:
526: With template-based foreground removal, we can form cleaned maps for each of the
527: 8 high frequency DA's, Q1-W4, while the ILC method produces one cleaned map from
528: a linear combination of all the \map\ frequency bands. We analyze the residual
529: dipole moment in each of these maps (a nominal dipole based on the 3 year data
530: is subtracted from the time-ordered data prior to map-making) using a Gibbs
531: sampling technique which generates an ensemble of full-sky CMB realizations that
532: are consistent with the data, as detailed below. We evaluate the dipole moment
533: of each full-sky realization and compute uncertainties from the scatter of the
534: realizations.
535:
536: We prepared the data for the Gibbs analysis as follows. The $N_{\rm side}=512$,
537: template-cleaned maps were zeroed within the KQ85 mask, smoothed with a
538: $10^{\circ}$ FWHM Gaussian kernel, and degraded to $N_{\rm side}=16$. Zeroing
539: the masked region prior to smoothing prevents residual cleaning errors within
540: the mask from contaminating the unmasked data. We add random white noise (12
541: $\mu$K rms per pixel) to each map to regularize the pixel-pixel covariance
542: matrix. The $N_{\rm side}=512$ ILC map was also smoothed with a $10^{\circ}$
543: FWHM Gaussian kernel and degraded to $N_{\rm side}=16$, but the data within the
544: sky mask were not zeroed prior to smoothing. We add white noise of 6 $\mu$K per
545: pixel to the smoothed ILC map to regularize its covariance matrix. Note that
546: smoothing the data with a $10^{\circ}$ kernel reduces the residual dipole in the
547: maps by $\sim$0.5\%. We ignore this effect since the residual dipole is only
548: $\sim$0.3\% of the full dipole amplitude to start with.
549:
550: The Gibbs sampler was run for 10,000 steps for each of the 8 template-cleaned
551: maps (Q1-W4) and for each of 6 independent noise realizations added to the ILC
552: map. In both cases we applied the KQ85 mask to the analysis and truncated the
553: CMB power at $l_{\rm max}=32$. The resulting ensembles of 80,000 and 60,000
554: dipole samples were analyzed independently and jointly. The results of this
555: analysis are given in Table~\ref{tab:dipole}. The first row combines the
556: results from the template-cleaned DA maps; the scatter among the 8 DA's was well
557: within the noise scatter for each DA, so the Gibbs samples for all 8 DA's were
558: combined for this analysis. The results for the ILC map are shown in the second
559: row. The two methods give reasonably consistent results, however, the Galactic
560: longitude of the two dipole axis estimates differ from each other by about
561: 2$\sigma$. Since we cannot reliably identify one cleaning method to be superior
562: to the other, we have merged the Gibbs samples from both methods to produce the
563: conservative estimate shown in the bottom row. This approach, which enlarges
564: the uncertainty to emcompass both estimates, gives
565: \beq
566: (d,l,b) = (3.355 \pm 0.008 \; {\rm mK}, 263.99^{\circ} \pm 0.14^{\circ}, 48.26^{\circ} \pm 0.03^{\circ}),
567: \eeq
568: where the amplitude estimate includes the 0.2\% absolute calibration
569: uncertainty. Given the CMB monopole temperature of 2.725 K
570: \citep{mather/etal:1999}, this amplitude implies a Solar System peculiar
571: velocity of $369.0 \pm 0.9$ km s$^{-1}$ with respect to the CMB rest frame.
572:
573:
574: \section{CALIBRATION IMPROVEMENTS}
575: \label{sec:cal_improve}
576:
577: With the 5 year processing we have refined our procedure for evaluating the
578: instrument calibration, and have improved our estimates for the calibration
579: uncertainty. The fundamental calibration source is still the dipole anisotropy
580: induced by \map's motion with respect to the CMB rest frame
581: \citep{hinshaw/etal:2003b, jarosik/etal:2007}, but several details of the
582: calibration fitting have been modified. The new calibration solution is
583: consistent with previous results in the overlapping time range. We estimate the
584: uncertainty in the absolute calibration is now 0.2\% per differencing assembly.
585:
586: The basic calibration procedure posits that a single channel of time-ordered
587: data, $d_i$, may be modeled as
588: \beq
589: d_i = g_i \left[\Delta T_{vi} + \Delta T_{ai} \right] + b_i,
590: \eeq
591: where $i$ is a time index, $g_i$ and $b_i$ are the instrument gain and baseline,
592: at time step $i$, $\Delta T_{vi}$ is the differential dipole anisotropy induced
593: by \map's motion, and $\Delta T_{ai}$ is the differential sky anisotropy. We
594: assume that $\Delta T_{vi}$ is known exactly and has the form
595: \beq
596: \Delta T_{vi} = \frac{T_0}{c} {\bf v}_i \cdot [(1+x_{\rm im}){\bf n}_{A,i} - (1-x_{\rm im}){\bf n}_{B,i}],
597: \label{eq:Tv}
598: \eeq
599: where $T_0 = 2.725$ K is the CMB temperature \citep{mather/etal:1999}, $c$ is
600: the speed of light, ${\bf v}_i$ is \map's velocity with respect to the CMB rest
601: frame at time step $i$, $x_{\rm im}$ is the loss imbalance parameter
602: \citep{jarosik/etal:2007}, and ${\bf n}_{A,i}$, and ${\bf n}_{B,i}$ are the unit
603: vectors of the A- and B-side lines of sight at time step $i$ (in the same frame
604: as the velocity vector). The velocity may be decomposed as
605: \beq
606: {\bf v}_i = {\bf v}_{\rm WMAP-SSB,i} + {\bf v}_{\rm SSB-CMB},
607: \eeq
608: where the first term is \map's velocity with respect to the solar system
609: barycenter, and the second is the barycenter velocity with respect to the CMB.
610: The former is well determined from ephemeris data, while the latter has been
611: measured by COBE-DMR with an uncertainty of 0.7\% \citep{Kogut/etal:1996d}.
612: Since the latter velocity is constant over \map's life span, any error in our
613: assumed value of ${\bf v}_{\rm SSB-CMB}$ will, in theory, be absorbed into a
614: dipole contribution to the anisotropy map, $T_a$. We test this hypothesis
615: below. The differential sky signal has the form
616: \beq
617: \Delta T_{ai} = (1+x_{\rm im})[I_a(p_{A,i})+P_a(p_{A,i},\gamma_{A,i})]
618: - (1-x_{\rm im})[I_a(p_{B,i})+P_a(p_{B,i},\gamma_{B,i})],
619: \eeq
620: where $p_{A,i}$ is the pixel observed by the A-side at time step $i$ (and
621: similarly for B), $I_a(p)$ is the temperature anisotropy in pixel $p$ (the
622: intensity Stokes parameter, $I$), and $P_a(p,\gamma)$ is the polarization
623: anisotropy in pixel $p$ at polarization angle $\gamma$
624: \citep{hinshaw/etal:2003b} which is related to the linear Stokes parameters $Q$
625: and $U$ by
626: \beq
627: P_a(p,\gamma) = Q(p) \cos 2\gamma + U(p) \sin 2\gamma.
628: \eeq
629: We further note that, in general, $I_a$ and $P_a$ depend on frequency owing to
630: Galactic emission.
631:
632: A main goal of the data processing is to simultaneously fit for the calibration
633: and sky signal. Unfortunately, since the data model is nonlinear and the number
634: of parameters is large, the general problem is intractable. In practice, we
635: proceed iteratively as follows. Initially we assume the gain and baseline are
636: constant for a given time interval, typically between 1 and 24 hours, \beqa g_i
637: = G_k & & \tau_k < t_i < \tau_{k+1} \\ b_i = B_k & & \tau_k < t_i <
638: \tau_{k+1}, \eeqa where $t_i$ is the time of the $i$th individual observation,
639: and $\tau_k$ is the start time of the $k$th calibration interval. Throughout
640: the fit we fix the velocity-induced signal, equation~(\ref{eq:Tv}), using ${\bf
641: v}_{\rm SSB-CMB} = [-26.29, -244.96, +275.93]$ km s$^{-1}$ (in Galactic
642: coordinates), and, for the first iteration, we assume no anisotropy signal,
643: $\Delta T_a = 0$. Then, for each calibration interval $k$ we perform a linear
644: fit for $G_k$ and $B_k$ with fixed $\Delta T_v + \Delta T_a$. As we proceed
645: through the intervals, we apply this calibration to the raw data and accumulate
646: a new estimate of the anisotropy map as per equation~19 of
647: \citet{hinshaw/etal:2003b}. The procedure is repeated with each updated
648: estimate of $\Delta T_a$. Once the calibration solution has converged, we fit
649: the gain data, $G_k$, to a model that is parameterized by the instrument
650: detector voltage and the temperatures of the receiver's warm and cold stages,
651: equation~2 of \citet{jarosik/etal:2007}. This parametrization still provides a
652: good fit to the $G_k$ data, so we have not updated its form for the 5 year
653: analysis. The updated best-fit parameters are given in the 5 year Explanatory
654: Supplement \citep{limon/etal:prep}. Note that for each radiometer, the relative
655: gain vs. time over 5 years is determined by just two parameters.
656:
657: For the 5 year processing we have focused on the veracity of the ``raw''
658: calibration, $G_k$ and $B_k$. Specifically, we have improved and/or critically
659: reexamined several aspects of the iterative fitting procedure:
660: \begin{itemize}
661: \item We have incorporated the effect of far sidelobe pickup directly into the
662: iterative calibration procedure, rather than as a fixed correction
663: \citep{jarosik/etal:2007}. We do this by segregating the differential signal
664: into a main beam contribution and a sidelobe contribution,
665: \beq
666: \Delta T_i = \Delta T_{\rm main,i} + \Delta T_{\rm side,i}.
667: \eeq
668: (\citealt{hill/etal:prep} discuss how this segregation is defined in the 5 year
669: processing.) After each iteration of the calibration and sky map estimation, we
670: (re)compute a database of $\Delta T_{\rm side}$ on a grid of pointings using the
671: new estimate of $I_a$. We then interpolate the database to estimate $\Delta
672: T_{\rm side,i}$ for each time step $i$. Note that $\Delta T_{\rm side}$
673: includes contributions from both the velocity-induced signal and the intrinsic
674: anisotropy. Ignoring sidelobe pickup can induce gain errors of up to 1.5\% in K
675: band, 0.4\% in Ka band, and $\sim$0.25\% in Q-W bands.
676:
677: \item In general, the different channels within a DA have different center
678: frequencies \citep{jarosik/etal:2003}; hence the different channels measure a
679: slightly different anisotropy signal due to differences in the Galactic signal.
680: We assess the importance of accounting for this in the calibration procedure.
681:
682: \item A single DA channel is only sensitive to a single linear polarization
683: state. (\map~measures polarization by differencing orthogonal polarization
684: channels.) Thus we cannot reliably solve for both $P_a$ and for $I_a$ at each
685: channel's center frequency. We assess the relative importance of accounting for
686: one or the other on both the gain and baseline solutions.
687:
688: \item We examine the sensitivity of the calibration solution to the choice of
689: ${\bf v}_{\rm SSB-CMB}$ and to assumptions of time-dependence in the gain.
690:
691: \end{itemize}
692:
693:
694: \subsection{Calibration Tests}
695: \label{sec:cal_test}
696:
697: We use a variety of end-to-end simulations to assess and control the systematic
698: effects noted above. We summarize a number of the key tests in the remainder of
699: this section.
700:
701: The first case we consider is a noiseless simulation in which we generate
702: time-ordered data from an input anisotropy map which includes CMB and Galactic
703: foreground signal (one map per channel, evaluated at the center frequency of
704: each channel) and a known dipole amplitude. The input gain for each channel is
705: fixed to be constant in time. We run the iterative calibration and sky map
706: solver allowing for an independent sky map solution at each channel (but no
707: polarization signal). When fitting for the calibration, we assume that ${\bf
708: v}_{\rm SSB-CMB}$ differs from the input value by 1\% to see if the known,
709: modulated velocity term, ${\bf v}_{\rm WMAP-SSB}$, properly ``anchors'' the
710: absolute gain solution. The results are shown in the top panel of
711: Figure~\ref{fig:gain_converge} where it is shown that the absolute gain recovery
712: is robust to errors in ${\bf v}_{\rm SSB-CMB}$. We recover the input gain to
713: better than 0.1\% in this instance.
714:
715: The second case we consider is again a noiseless simulation that now includes
716: {\it only} dipole signal (with Earth-velocity modulation), but here we vary the
717: input gain using the flight-derived gain model \citep{jarosik/etal:2007}. The
718: iterative solver was run on the K band data for 1400 iterations, again starting
719: with an initial guess that was in error by 1\%. The results are shown in the
720: bottom panel of Figure~\ref{fig:gain_converge}, which indicate systematic
721: convergence errors of $\gt$0.3\% in the fitted amplitude of the recovered gain
722: model. Since the input sky signal in this case does not have any Galactic
723: foreground or polarization components, we cannot ascribe the recovery errors to
724: the improper handling of those effects in the iterative solver. We have also
725: run numerous other simulations that included various combinations of instrument
726: noise, CMB anisotropy, Galactic foreground signal (with or without individual
727: center frequencies per channel), polarization signal, and input gain
728: variations. The combination of runs are too numerous to report on in detail,
729: and the results are not especially enlightening. The most pertinent trend we
730: can identify is that when the input value of ${\bf v}_{\rm SSB-CMB}$ is assumed
731: in the iterative solver, the recovered gain is in good agreement with the input,
732: but when the initial guess is in error by 1\%, the recovered gain will have
733: comparable errors. We believe the lack of convergence is due to a weak
734: degeneracy between gain variations and the sky map solution. Such a degeneracy
735: is difficult to diagnose in the context of this iterative solver, especially
736: given the computational demands of the system, so we are assessing the
737: system more directly with a low-resolution parameterization of the gain and sky
738: signal, as outlined in Appendix~\ref{app:cal_fisher}.
739:
740: Since the latter effort is still underway, we have adopted a more pragmatic
741: approach to evaluating the absolute gain and its uncertainty for the 5 year data
742: release. We proceed as follows: after 50 iterations of the calibration and sky
743: map solver, the dominant errors in the gain and sky map solution are 1) a dipole
744: in the sky map, and 2) a characteristic wave form that reflects a relative error
745: between ${\bf v}_{\rm SSB-CMB}$ and ${\bf v}_{\rm WMAP-SSB}$. At this point we
746: can calibrate the amplitude of the gain error wave form to the magnitude of the
747: velocity error in ${\bf v}_{\rm SSB-CMB}$. We can then fit the gain solution to
748: a linear combination of the gain model of \citet{jarosik/etal:2007} and the
749: velocity error wave form. See Appendix~\ref{app:cal_model_fit} for details on
750: this fitting procedure. In practice this fit is performed simultaneously on
751: both channels of a radiometer since those channels share one gain model
752: parameter. We have tested this procedure on a complete flight-like simulations
753: that includes every important effect known, including input gain variations.
754: The results of the gain recovery are shown in Figure~\ref{fig:gain_recover}, and
755: based on this we conservatively assign an absolute calibration uncertainty of
756: 0.2\% per channel for the 5 year \map\ archive.
757:
758:
759: \subsection{Summary}
760: \label{sec:cal_summ}
761:
762: The series of steps taken to arrive at the final 5 year calibration are as
763: follows:
764: \begin{itemize}
765:
766: \item Run the iterative calibration and sky map solver over the full 5 year data
767: set for 50 iterations, using 24 hour calibration intervals. This run starts
768: with $I_a = P_a = 0$ and updates $I_a$ for each individual channel of data.
769: $P_a$ is assumed to be 0 throughout this run. We keep the gain solution, $G_k$,
770: from this run and discard the baseline solution.
771:
772: \item Run the iterative calibration and sky map solver over the full 5 year data
773: set for 50 iterations, using 1 hour calibration intervals. This run starts with
774: $I_a = P_a = 0$ and updates both using the intensity and polarization data in
775: the two radiometers per DA, as per Appendix~D of \citet{hinshaw/etal:2003b}. We
776: keep the baseline solution, $B_k$, from this run and discard the gain solution.
777: Both of these runs incorporate the sidelobe correction as noted above.
778:
779: \item Fit the gain solution, $G_k$ simultaneously for the gain model and for an
780: error in the velocity, ${\bf \Delta v}_{\rm SSB-CMB}$, as described in
781: Appendix~\ref{app:cal_model_fit}. This fit is performed on two channels per
782: radiometer with the gain model parameter $T_0$ common to both channels.
783:
784: \item We average the best-fit velocity error over all channels within a
785: frequency band under the assumption that the dipole is the same in each of these
786: channels. We then fix the velocity error to a single value per frequency band
787: and re-fit the gain model parameters for each pair of radiometer channels.
788:
789: \end{itemize}
790:
791: Based on end-to-end simulations with flight-like noise, we estimate the absolute
792: gain error per radiometer to be 0.2\%. We believe the limiting factor in this
793: estimate is a weak degeneracy between thermal variations in the instrument gain,
794: which are annually modulated, and annual variations induced by errors in ${\bf
795: v}_{\rm SSB-CMB}$. Since there is a small monotonic increase in the spacecraft
796: temperature, additional years of data should allow improvements in our ability
797: to separate these effects.
798:
799: Once we have finalized the gain model, we form a calibrated time-ordered data
800: archive using the gain model and the 1 hour baseline estimates to calibrate the
801: data. This archive also has a final estimate of the far sidelobe pickup
802: subtracted from each time-ordered data point. However, we opt not to subtract a
803: dipole estimate from the archive at this stage in the processing.
804:
805:
806: \section{BEAM IMPROVEMENTS}
807: \label{sec:beams}
808:
809: In addition to reassessing the calibration, the other major effort undertaken to
810: improve the 5 year data processing was to extend the physical optics model of
811: the \map\ telescope based on flight measurements of Jupiter. This work is
812: described in detail in \citet{hill/etal:prep} so we only summarize the key
813: results with an emphasis on their scientific implications. The basic aim of the
814: work is to use the flight beam maps from all 10 DA's to determine the in-flight
815: distortion of the mirrors. This program was begun for the A-side mirror during
816: the 3 year analysis; for the 5 year analysis we have quadrupled the number of
817: distortion modes we fit (probing distortion scales that are half the previous
818: size), and we have developed a completely new and independent model of the
819: B-side distortions, rather than assuming that they mirror the A-side
820: distortions. We have also placed limits on smaller scale distortions by
821: comparing the predicted beam response at large angles to sidelobe data collected
822: during \map's early observations of the Moon.
823:
824: Given the best-fit mirror model, we compute the model beam response for each DA
825: and use it in conjunction with the flight data to constrain the faint tails of
826: the beams, beyond $\sim 1^{\circ}$ from the beam peak. These tails are
827: difficult to constrain with flight data alone because the Jupiter signal to
828: noise ratio is low, but, due to their large areal extent they contain a
829: non-negligible fraction (up to 1\%) of the total beam solid angle. An accurate
830: determination of the beam tail is required to properly measure the ratio of
831: sub-degree-scale power to larger-scale power in the diffuse CMB emission (and to
832: accurately assign point source flux).
833:
834: Figure~14 in \citet{hill/etal:prep} compares the beam radial profiles used in
835: the 3 year and 5 year analyses, while Figure~13 compares the $l$-space transfer
836: functions derived from the Legendre transform of the radial profile. The
837: important changes to note are the following.
838:
839: \begin{enumerate}
840:
841: \item In both analyses we split the beam response into main beam and far
842: sidelobe contributions. In the 5 year analysis we have enlarged the radius at
843: which this transition is made \citep{hill/etal:prep}. In both cases, we correct
844: the time-ordered data for far sidelobe pickup prior to making sky maps, while
845: the main beam contribution is only accounted for in the analysis of sky maps,
846: e.g., in power spectrum deconvolution. As a result, the sky maps have a
847: slightly different effective resolution which is most apparent in K band, as in
848: Figure~\ref{fig:diff_maps_i_recal}. However, in each analysis, the derived
849: transfer functions are appropriate for the corresponding sky maps.
850:
851: \item In the 3 year analysis, the main beam profile was described by a Hermite
852: polynomial expansion fit to the observations of Jupiter in the time-ordered
853: data. This approach was numerically problematic in the 5 year analysis due to
854: the larger transition radius; as a result, we now simply co-add the time-ordered
855: data into radial bins to obtain the profiles. In both cases, the underlying
856: time-ordered data is a hybrid archive consisting of flight data for points where
857: the beam model predicts a value above a given contour, and model values for
858: points below the contour \citep{hill/etal:prep}. With the improved beam models
859: and a new error analysis, we have adjusted these hybrid contours down slightly,
860: with the result that we use proportionately more flight data (per year) in the
861: new analysis. The radius at which the 5 year profile becomes model dominated
862: ($\gt$50\% of the points in a bin) is indicated by dotted lines in Figure~14 of
863: \citet{hill/etal:prep}.
864:
865: \item The right column of Figure~14 in \citet{hill/etal:prep} shows the
866: fractional change in solid angle due to the updated profiles. The main point to
867: note is the $\sim$1\% increase in the V2 and W band channels, primarily arising
868: in the bin from 1 to 2 degrees off the beam peak. As can be seen in Figure~3 of
869: \citet{hill/etal:prep}, this is the angular range in which the new beam models
870: produced the most change, owing to the incorporation of smaller distortion modes
871: in the mirror model. The 3 year analysis made use of the model in this angular
872: range which, in hindsight, was suppressing up to $\sim$1\% of the solid angle in
873: the V and W band beams. (The longer wavelength channels are less sensitive to
874: distortions in this range, so the change in solid angle is smaller for K-Q
875: bands.) In the 5 year analysis, we use relatively more flight data in this
876: regime, so we are less sensitive to any remaining model uncertainties.
877: \citet{hill/etal:prep} place limits on residual model errors and propagate those
878: errors into the overall beam uncertainty.
879:
880: \item Figure~13 in \citet{hill/etal:prep} compares the beam transfer functions,
881: $b_l$, derived by transforming the 3 year and 5 year radial profiles. (To
882: factor out the effect of changing the transition radius, the 3 year profiles
883: were extended to the 5 year radius using the far sidelobe data, for this
884: comparison.) Since the transfer functions are normalized to 1 at $l=1$, the
885: change is restricted to high $l$. In V and W bands, $b_l$ has decreased by
886: $\sim$0.5 - 1\% due largely to the additional solid angle picked up in the 1-2
887: degree range. This amounts to a $\sim$1 $\sigma$ change in the functions, as
888: indicated by the red curves in the Figure.
889:
890: \end{enumerate}
891:
892: The calibrated angular power spectrum is proportional to $1/g^2 b_l^2$, where
893: $g$ is the mean gain and $b_l$ is the beam transfer function, thus the net
894: effect of the change in gain and beam determinations is to increase the power
895: spectrum by $\sim$0.5\% at $l \la 100$, and by $\sim$2.5\% at high $l$.
896: \citet{nolta/etal:prep} give a detailed evaluation of the power spectrum while
897: \citet{dunkley/etal:prep} and \citet{komatsu/etal:prep} discuss the implications
898: for cosmology.
899:
900:
901: \section{LOW-$l$ POLARIZATION TESTS}
902: \label{sec:pol_test}
903:
904: The 3 year data release included the first measurement of microwave polarization
905: over the full sky, in the form of Stokes Q and U maps in each of 5 bands. The
906: analysis of \map\ polarization data is complicated by the fact that the
907: instrument was not designed to be a true polarimeter, thus a number of
908: systematic effects had to be understood prior to assigning reliable error
909: estimates to the data. \citet{page/etal:2007} presented the 3 year polarization
910: data in great detail. In this section we extend that analysis by considering
911: some additional tests that were not covered in the 3 year analysis. We note that
912: all of the tests described in this section have been performed on the
913: template-cleaned reduced-foreground maps except for the final test of the Ka
914: band data, described at the end of the section, which tests an alternative
915: cleaning method.
916:
917: \subsection{Year-to-Year Consistency Tests}
918: \label{sec:pol_year}
919:
920: With 5 years of data it is now possible to subject the data to more stringent
921: consistency tests than was previously possible. In general, the number of
922: independent cross-power spectra we can form within a band with $N_d$
923: differencing assemblies is $N_d(N_d-1)/2 \times N_y + N_d^2 \times
924: N_y(N_y-1)/2$. With 5 years of data, this gives 10 independent estimates each
925: in K and Ka band, 45 each in Q and V band, and 190 in W band. For cross power
926: spectra of distinct band pairs, with $N_{d1}$ and $N_{d2}$ DA's in each band,
927: the number is $N_{d1}N_{d2} \times N_y^2$. This gives 50 each in KaQ and KaV,
928: 100 each in KaW and QV, and 200 each in QW and VW. (For comparison, the
929: corresponding numbers are 3, 15, \& 66, and 18, 36, \& 72 with 3 years of
930: data.)
931:
932: We have evaluated these individual spectra from the 5 year data and have
933: assigned noise uncertainties to each estimate using the Fisher formalism
934: described in \citet{page/etal:2007}. We subject the ensemble to an internal
935: consistency test by computing the reduced $\chi^2$ of the data at each multipole
936: $l$ within each band or band pair, under the hypothesis that the data at each
937: multipole and band measures the same number from DA to DA and year to year. The
938: results of this test are given in Table~\ref{tab:chi2_pol_l} for the
939: foreground-cleaned EE, EB, and BB spectra from $l=2-10$ for all band pairs from
940: KaKa to WW. There are several points to note in these results.
941:
942: \begin{enumerate}
943:
944: \item For $l \ge 6$, the most significant deviation from 1 in reduced $\chi^2$,
945: in any spectrum or band, is 1.594 in the $l=7$ BB spectrum for KaQ. With 50
946: degrees of freedom, this is a 3 $\sigma$ deviation, but given that we have 150
947: $l \ge 6$ samples in the table, we expect of order 1 such value. Thus we
948: conclude that the Fisher-based errors provide a good description of the DA-to-DA
949: and year-to-year scatter in the $l \ge 6$ polarization data. If anything, there
950: is a slight tendency to overestimate the uncertainties at higher $l$.
951:
952: \item For $l \le 5$, we find 37 out of 120 points where the reduced $\chi^2$
953: deviates from 1 at more than 4 $\sigma$ significance, indicating excessive
954: internal scatter in the data relative to the Fisher errors. However, all but 5
955: of these occur in cross-power spectra in which one or both of the bands contain
956: W band data. If we exclude combinations with W band, the remaining 72 points
957: have a mode in the reduced $\chi^2$ distribution of 1 with a slight positive
958: skewness due to the 5 points noted above, which all contain Q band data. This
959: may be a sign of slight foreground residuals contributing additional noise to
960: the Q band data, though we do not see similar evidence in the Ka band spectra
961: which would be more foreground contaminated prior to cleaning. For Ka-V bands,
962: we believe that the Fisher errors provide an adequate description of the scatter
963: in this $l \le 5$ polarization data, but we subject polarization sensitive
964: cosmological parameter estimates, e.g., the optical depth, to additional
965: scrutiny in \S\ref{sec:pol_Ka}.
966:
967: \item Of special note is $l=3$ BB which, as noted in \citet{page/etal:2007}, is
968: the power spectrum mode that is least modulated in the \map\ time-ordered data.
969: This mode is therefore quite sensitive to how the instrument baseline is
970: estimated and removed and, in turn, to how the 1/f noise is modeled. In the
971: accounting above, the $l=3$ BB data have the highest internal scatter of any
972: low-$l$ polarization mode. In particular, {\em every} combination that includes
973: W band data is significantly discrepant; and the two most discrepant non-W band
974: points are also estimates of $l=3$ BB. We comment on the W band data further
975: below, but note here that the final co-added BB spectrum (based on Ka, Q, and V
976: band data) does {\em not} lead to a significant detection of tensor modes.
977: However, we caution that any surprising scientific conclusions which rely
978: heavily on the \map\ $l=3$ BB data should be treated with caution.
979:
980: \end{enumerate}
981:
982: Based on the analysis presented above, we find the W band polarization data is
983: still too unstable at low-$l$ to be reliably used for cosmological studies. We
984: cite more specific phenomenology and consider some possible explanations in the
985: remainder of this section.
986:
987: The 5 year co-added W band EE spectrum is shown in Figures~\ref{fig:ww_ee}, in
988: the form of likelihood profiles from $l=2-7$. At each multipole we show two
989: curves: an estimate based on evaluating the likelihood multipole by multipole,
990: and an estimate based on the pseudo-$C_l$ method \citep{page/etal:2007}. The
991: best-fit model EE spectrum, based on the combined Ka, Q, and V band data is
992: indicated by the dashed lines in each panel. Both spectrum estimates show
993: excess power relative to the model spectrum, with the most puzzling multipole
994: being $l=7$ which, as shown in Table~\ref{tab:chi2_pol_l}, has an internal
995: reduced $\chi^2$ of 1.015, for 190 degrees of freedom. This data has the
996: hallmark of a sky signal, but that hypothesis is implausible for a variety of
997: reasons \citep{page/etal:2007}. It is more likely due to a systematic effect
998: that is common to a majority of the W band channels over a majority of the 5
999: years of data. We explore and rule out one previously neglected effect in
1000: \S\ref{sec:pol_emiss}. It is worth recalling that $l=7$ EE, like $l=3$ BB, is a
1001: mode that is relatively poorly measured by \map, as discussed in
1002: \citet{page/etal:2007}; see especially Figure~16 and its related discussion.
1003:
1004: The W band BB data also exhibit unusual behavior at $l=2,3$. In this case,
1005: these two multipoles have internal reduced $\chi^2$ greater than 6, and the
1006: co-added $l=2$ point is nearly 10 $\sigma$ from zero. However, with 190 points
1007: in each 5 year co-added estimate it is now possible to look for trends within
1008: the data that were relatively obscure with only 3 years of data. In particular,
1009: we note that in the $l=2$ estimate, there are 28 points that are individually
1010: more than 5 $\sigma$ from zero and that {\em all} of them contain W1 data in one
1011: or both of the DA pairs in the cross power spectrum. Similarly for $l=3$, there
1012: are 14 points greater than 5 $\sigma$ and {\em all} of those points contain W4
1013: data in one or both of the DA pairs. We have yet to pinpoint the significance
1014: of this result, but we plan to study the noise properties of these DA's beyond
1015: what has been reported to date, and to sharpen the phenomenology with additional
1016: years of data.
1017:
1018:
1019: \subsection{Emissivity Tests}
1020: \label{sec:pol_emiss}
1021:
1022: In this section we consider time dependent emission from the \map\ optics as a
1023: candidate for explaining the excess W band ``signal'' seen in the EE spectrum,
1024: mostly at $l=7$. In the end, the effect proved not to be significant, but it
1025: provides a useful illustration of a common-mode effect that we believe is still
1026: present in the W band polarization data.
1027:
1028: From a number of lines of reasoning, we know that the microwave emissivity of
1029: the mirrors is a few percent in W band, and that it scales with frequency
1030: roughly like $\nu^{1.5}$ across the \map\ frequency range, as expected for a
1031: classical metal \citep{born/wolf:POO:6e}. Hence this mechanism has the
1032: potential to explain a common-mode effect that is primarily seen in W band.
1033: Further, Figure~1 in \citet{jarosik/etal:2007} shows that the physical
1034: temperature of the primary mirrors are modulated at the spin period by $\sim$200
1035: $\mu$K, with a dependence on solar azimuth angle that is highly repeatable from
1036: year to year. We believe this modulation is driven by solar radiation
1037: diffracting around the \map\ sun shield reaching the tops of the primary
1038: mirrors, which are only a few degrees within the geometric shadow of the sun
1039: shield. In contrast, the secondary mirrors and feed horns are in deep shadow
1040: and show no measurable variation at the spin period, so that any emission they
1041: produce only contributes to an overall radiometer offset, and will not be
1042: further considered here.
1043:
1044: As a rough estimate, the spin modulated emission from the primary mirrors could
1045: produce as much as $\sim 0.02 \times 200 = 4$ $\mu$K of radiometric response in
1046: W band, but the actual signal depends on the relative phase of the A and B-side
1047: mirror variations and the polarization state of the emission. In more detail,
1048: the differential signal, $d(t)$, measured by a radiometer with lossy elements is
1049: \beq
1050: d(t) = (1 - \epsilon_A)\,T_A(t) - (1 - \epsilon_B)\,T_B(t)
1051: + \epsilon^{\rm p}_A \, T^{\rm p}_A(t) - \epsilon^{\rm p}_B \, T^{\rm p}_B(t)
1052: \label{eq:loss}
1053: \eeq
1054: where $\epsilon_A\ = \epsilon^{\rm p}_A + \epsilon^{\rm s}_A + \epsilon^{\rm
1055: f}_A$ is the combined loss in the A-side optics: (p)rimary plus
1056: (s)econdary mirrors, plus the (f)eed horn, and likewise for the B-side.
1057: $T_{A,B}$ is the sky temperature in the direction of the A or B-side
1058: line-of-sight; and $T^{\rm p}_{A,B}$ is the physical temperature of the A or
1059: B-side primary mirror.
1060:
1061: The first two terms are the sky signal attenuated by the overall loss in the A
1062: and B side optics, respectively. The effects of loss imbalance, which arise when
1063: $\epsilon_A \ne \epsilon_B$, have been studied extensively
1064: \citep{jarosik/etal:2003, jarosik/etal:2007}. We account for loss imbalance in
1065: the data processing and we marginalize over residual uncertainties in the
1066: imbalance coefficients when we form the pixel-pixel inverse covariance matrices
1067: \citep{jarosik/etal:2007}. Updated estimates of the loss imbalance coefficients
1068: based on fits to the 5 year data are reported in Table~\ref{tab:x_im}.
1069:
1070: In the remainder of this section we focus on the last two emissive terms in
1071: Equation~\ref{eq:loss}. Recall that a \map\ differencing assembly consists of
1072: two radiometers, 1 and 2, that are sensitive to orthogonal linear polarization
1073: modes. The temperature and polarization signals are extracted by forming the
1074: sum and difference of the two radiometer outputs; thus, the emission terms we
1075: need to evaluate are
1076: \beq
1077: d^{\rm p}_1(t) \pm d^{\rm p}_2(t)
1078: = \left(\frac{\epsilon^{\rm p}_{A1} \pm \epsilon^{\rm p}_{A2}}{1 - \epsilon}\right) \, T^{\rm p}_A
1079: - \left(\frac{\epsilon^{\rm p}_{B1} \pm \epsilon^{\rm p}_{B2}}{1 - \epsilon}\right) \, T^{\rm p}_B
1080: \eeq
1081: where $\epsilon^{\rm p}_{A1}$ is the A-side primary mirror emissivity measured
1082: by radiometer 1, and so forth. The factor of $1-\epsilon$ in the denominator
1083: applies a small correction for the mean loss, $\epsilon \equiv (\epsilon_A +
1084: \epsilon_B)/2$, and arises from the process of calibrating the data to a known
1085: sky brightness temperature (\S\ref{sec:cal_improve}). Note that we only pick up
1086: a polarized response if $\epsilon_1 \ne \epsilon_2$.
1087:
1088: We have simulated this signal in the time-ordered data using the measured
1089: primary mirror temperatures as template inputs. The emissivity coefficients
1090: were initially chosen to be consistent with the loss imbalance constraints.
1091: However, in order to produce a measurable polarization signal, we had to boost
1092: the emissivity differences to the point where they became unphysical, that is $|
1093: \epsilon_1 - \epsilon_2 | > | \epsilon_1 + \epsilon_2 |$. Nonetheless, it was
1094: instructive to analyze this simulation by binning the resulting data (which also
1095: includes sky signal and noise) as a function of solar azimuth. The results are
1096: shown in the top panel of Figure~\ref{fig:spin_bin} which shows 3 years of
1097: co-added W band polarization data, the $d_1-d_2$ channel; the input emissive
1098: signal is shown in red for comparison. We are clearly able to detect such a
1099: signal with this manner of binning. We also computed the low-$l$ polarization
1100: spectra and found that, despite the large spin modulated input signal, the
1101: signal induced in the power spectrum was less than 2 $\mu$K$^2$ in
1102: $l(l+1)C^{EE}_l/2\pi$, which is insufficient to explain the $l=7$ feature in the
1103: W band EE spectrum.
1104:
1105: In parallel with the simulation analysis, we have binned the flight radiometer
1106: data by solar azimuth angle to search for spin modulated features in the
1107: polarization data. The results for W band are shown in the bottom panel of
1108: Figure~\ref{fig:spin_bin} for the 5 year data. While the $\chi^2$ per degree of
1109: freedom relative to zero is slightly high, there is no compelling evidence for a
1110: coherent spin modulated signal at the $\sim$2 $\mu$K level. In contrast, the
1111: simulation yielded spin modulated signals of 5-10 $\mu$K and still failed to
1112: produce a significant effect in the EE spectrum. Hence we conclude that thermal
1113: emission from the \map\ optics cannot explain the excess W band EE signal. In
1114: any event, we continue to monitor the spin modulated data for the emergence of a
1115: coherent signal.
1116:
1117: \subsection{Ka Band Tests}
1118: \label{sec:pol_Ka}
1119:
1120: The analysis presented in \S\ref{sec:pol_year} shows that the Ka band
1121: polarization data is comparable to the Q and V band data in its internal
1122: consistency. That analysis was performed on data that had been foreground
1123: cleaned using the template method discussed in \citet{page/etal:2007} and
1124: updated in \citet{gold/etal:prep}. In order to assess whether or not this
1125: cleaned Ka band data is suitable for use in cosmological parameter estimation we
1126: subject it to two further tests: 1) a null test in which Ka band data is
1127: compared to the combined Q and V band data, and 2) a parameter estimation based
1128: solely on Ka band data.
1129:
1130: For the null test, we form polarization maps by taking differences,
1131: $\frac{1}{2}S_{\rm Ka} - \frac{1}{2}S_{\rm QV}$, where S = Q,U are the
1132: polarization Stokes parameters, $S_{\rm Ka}$ are the maps formed from the Ka
1133: band data, and $S_{\rm QV}$ are the maps formed from the optimal combination of
1134: the Q and V band data. We evaluate the EE power spectrum from these null maps
1135: by evaluating the likelihood mode by mode while holding the other multipoles
1136: fixed at zero. The results are shown in Figure~\ref{fig:null_EE}, along with
1137: the best-fit model spectrum based on the final 5 year $\Lambda$CDM analysis.
1138: The spectrum is clearly consistent with zero, but to get a better sense of the
1139: power of this test, we have also used these null maps to estimate the optical
1140: depth parameter, $\tau$. The result of that analysis is shown as the dashed
1141: curve in Figure~\ref{fig:tau_test}, where we find that the null likelihood peaks
1142: at $\tau=0$ and excludes the most-likely cosmological value with $\sim$95\%
1143: confidence.
1144:
1145: As a separate test, we evaluate the $\tau$ likelihood using only the
1146: template-cleaned Ka band signal maps. The result of that test is shown as the
1147: blue curve in Figure~\ref{fig:tau_test}. While the uncertainty in the Ka band
1148: estimate is considerably larger than the combined QV estimate (shown in red),
1149: the estimates are highly consistent. The result of combining Ka, Q, and V band
1150: data is shown in the black curve.
1151:
1152: \citet{dunkley/etal:prep} present a complementary method of foreground cleaning
1153: that makes use of Ka band data, in conjunction with K, Q, and V band data.
1154: Using a full 6 parameter likelihood evalutaion, they compare the optical depth
1155: inferred from the two cleaning methods while using the full combined data sets
1156: in both cases: see Figure~9 of \citet{dunkley/etal:prep} for details. Based
1157: on these tests, we conclude that the Ka band data is sufficiently free of
1158: systematic errors and residual foreground signals that it is suitable for
1159: cosmological studies. The use of this band significantly enhances the overall
1160: polarization sensitivity of \map.
1161:
1162:
1163: \section{SUMMARY OF 5-YEAR SCIENCE RESULTS}
1164: \label{sec:science}
1165:
1166: Detailed presentations of the scientific results from the 5 year data are given
1167: by \citet{gold/etal:prep}, \citet{wright/etal:prep}, \citet{nolta/etal:prep},
1168: \citet{dunkley/etal:prep}, and \citet{komatsu/etal:prep}. Starting with the 5
1169: year temperature and polarization maps, with their improved calibration,
1170: \citet{gold/etal:prep} give a new Markov Chain Monte Carlo-based analysis of
1171: foreground emission in the data. Their results are broadly consistent with
1172: previous analyses by the \map\ team and others \citep{eriksen/etal:2007}, while
1173: providing some new results on the microwave spectra of bright sources in the
1174: Galactic plane that aren't well fit by simple power-law foreground models.
1175: Figure~\ref{fig:ilc_map} shows the 5 year CMB map based on the internal linear
1176: combination (ILC) method of foreground removal.
1177:
1178: \citet{wright/etal:prep} give a comprehensive analysis of the extragalactic
1179: sources in the 5 year data, including a new analysis of variability made
1180: possible by the multi-year coverage. The 5 year \map\ source catalog now
1181: contains 390 objects and is reasonably complete to a flux of 1 Jy away from the
1182: Galactic plane. The new analysis of the \map\ beam response
1183: \citep{hill/etal:prep} has led to more precise estimates of the point source
1184: flux scale for all 5 \map\ frequency bands. This information is incorporated in
1185: the new source catalog \citep{wright/etal:prep}, and is also used to provide new
1186: brightness estimates of Mars, Jupiter, and Saturn \citep{hill/etal:prep}. We
1187: find significant (and expected) variability in Mars and Saturn over the course
1188: of 5 years and use that information to provide a preliminary recalibration of a
1189: Mars brightness model \citep{wright:2007}, and to fit a simple model of Saturn's
1190: brightness as a function of ring inclination.
1191:
1192: The temperature and polarization power spectra are presented in
1193: \citet{nolta/etal:prep}. The spectra are all consistent with the 3 year results
1194: with improvements in sensitivity commensurate with the additional integration
1195: time. Further improvements in our understanding of the absolute calibration and
1196: beam response have allowed us to place tighter uncertainties on the power
1197: spectra, over and above the reductions from additional data. These changes are
1198: all reflected in the new version of the \map\ likelihood code. The most notable
1199: improvements arise in the third acoustic peak of the TT spectrum, and in all of
1200: the polarization spectra; for example, we now see unambiguous evidence for a 2nd
1201: dip in the high-$l$ TE spectrum, which further constrains deviations from the
1202: standard $\Lambda$CDM model. The 5 year TT and TE spectra are shown in
1203: Figure~\ref{fig:tt_te}. We have also generated new maximum likelihood estimates
1204: of the low-$l$ polarization spectra: TE, TB, EE, EB, and BB to complement our
1205: earlier estimates based on pseudo-$C_l$ methods \citep{nolta/etal:prep}. The
1206: TB, EB, and BB spectra remain consistent with zero.
1207:
1208: The cosmological implications of the 5 year \map\ data are discussed in detail
1209: in \citet{dunkley/etal:prep} and \citet{komatsu/etal:prep}. The now-standard
1210: cosmological model: a flat universe dominated by vacuum energy and dark matter,
1211: seeded by nearly scale-invariant, adiabatic, Gaussian random-phase fluctuations,
1212: continues to fit the 5 year data. \map\ has now determined the key parameters
1213: of this model to high precision; a summary of the 5 year parameter results is
1214: given in Table~\ref{tab:best_param}. The most notable improvements are the
1215: measurements of the dark matter density, $\Omega_c h^2$, and the amplitude of
1216: matter fluctuations today, $\sigma_8$. The former is determined with 6\%
1217: uncertainty using \map\ data only \citep{dunkley/etal:prep}, and with 3\%
1218: uncertainty when \map\ data is combined with BAO and SNe constraints
1219: \citep{komatsu/etal:prep}. The latter is measured to 5\% with \map\ data, and
1220: to 3\% when combined with other data. The redshift of reionization is
1221: \ensuremath{z_{\rm reion}} = \ensuremath{11.0\pm 1.4}, if
1222: the universe were reionized instantaneously. The 2 $\sigma$ lower limit is
1223: $z_{\rm reion} \gt 8.2$, and instantaneous reionization at $z_{\rm reion} = 6$
1224: is rejected at 3.5 $\sigma$. The \map\ data continues to favor models with a
1225: tilted primordial spectrum,
1226: \ensuremath{n_s} = \ensuremath{0.963^{+ 0.014}_{- 0.015}}.
1227: \citet{dunkley/etal:prep} discuss how the $\Lambda$CDM model continues to fit a
1228: host of other astronomical data as well.
1229:
1230: Moving beyond the standard $\Lambda$CDM model, when \map\ data is combined with
1231: BAO and SNe observations \citep{komatsu/etal:prep}, we find no evidence for
1232: running in the spectral index of scalar fluctuations,
1233: \ensuremath{dn_s/d\ln{k} = -0.028\pm 0.020} (68\% CL).
1234: The new limit on the tensor-to-scalar ratio is
1235: \ensuremath{r < 0.22\ \mbox{(95\% CL)}},
1236: and we obtain tight, simultaneous limits on the (constant) dark energy equation
1237: of state and the spatial curvature of the universe:
1238: \ensuremath{-0.14<1+w<0.12\ \mbox{(95\% CL)}}
1239: and
1240: \ensuremath{-0.0179<\Omega_k<0.0081\ \mbox{(95\% CL)}}.
1241: The angular power spectrum now exhibits the signature of the cosmic neutrino
1242: background: the number of relativistic degrees of freedom, expressed in units of
1243: the effective number of neutrino species, is found to be
1244: \ensuremath{N_{\rm eff} = 4.4\pm 1.5} (68\% CL),
1245: consistent with the standard value of 3.04. Models with $N_{\rm eff} = 0$ are
1246: disfavored at $\gt$99.5\% confidence. A summary of the key cosmological
1247: parameter values is given in Table~\ref{tab:best_param}, where we provide
1248: estimates using \map\ data alone and \map\ data combined with BAO and SNe
1249: observations. A complete tabulation of all parameter values for each model and
1250: dataset combination we studied is available on LAMBDA.
1251:
1252: The new data also place more stringent limits on deviations from Gaussianity,
1253: parity violations, and the amplitude of isocurvature fluctuations
1254: \citep{komatsu/etal:prep}. For example, new limits on physically motivated
1255: primordial non-Gaussianity parameters are $-9 < \fnlKS <111$ (95\% CL) and $-151
1256: < \fnleq < 253$ (95\% CL) for the local and equilateral models, respectively.
1257:
1258:
1259: \section{CONCLUSIONS}
1260: \label{sec:conclusions}
1261:
1262: We have presented an overview of the 5 year \map\ data and have highlighted the
1263: improvements we have made to the data processing and analysis since the 3 year
1264: results were presented. The most substantive improvements to the processing
1265: include a new method for establishing the absolute gain calibration (with
1266: reduced uncertainty), and a more complete analysis of the \map\ beam response
1267: made possible by additional data and a higher fidelity physical optics model.
1268: Numerous other processing changes are outlined in \S\ref{sec:change}.
1269:
1270: The 5 year sky maps are consistent with the 3 year maps and have noise levels
1271: that are $\sqrt{5}$ times less than the single year maps. The new maps are
1272: compared to the 3 year maps in \S\ref{sec:maps}. The main changes to the
1273: angular power spectrum are as follows: at low multipoles ($l \la 100$) the
1274: spectrum is $\sim$0.5\% higher than the 3 year spectrum (in power units) due to
1275: the new absolute gain determination. At higher multipoles it is increased by
1276: $\sim$2.5\%, due to the new beam response profiles, as explained in
1277: \S\ref{sec:beams} and in \citet{hill/etal:prep}. These changes are consistent
1278: with the 3 year uncertainties when one accounts for both the 0.5\% gain
1279: uncertainty (in temperature units) and the 3 year beam uncertainties, which were
1280: incorporated into the likelihood code.
1281:
1282: We have applied a number of new tests to the polarization data to check internal
1283: consistency and to look for new systematic effects in the W band data
1284: (\S\ref{sec:pol_test}). As a result of these tests, and of new analyses of
1285: polarized foreground emission \citep{dunkley/etal:prep}, we have concluded that
1286: Ka band data can be used along with Q and V band data for cosmological
1287: analyses. However, we still find a number of features in the W band
1288: polarization data that preclude its use, except in the Galactic plane where the
1289: signal to noise is relatively high. We continue to investigate the causes of
1290: this and have identified new clues to follow up on in future studies
1291: (\S\ref{sec:pol_year}).
1292:
1293: Scientific results gathered from the suite of 5 year papers are summarized in
1294: \S\ref{sec:science}. The highlights include smaller uncertainties in the
1295: optical depth, $\tau$, due to a combination of additional years of data and to
1296: the inclusion of Ka band polarization data: instantaneous reionization at
1297: $z_{\rm reion} = 6$ is now rejected at 3.5 $\sigma$. New evidence favoring a
1298: non-zero neutrino abundance at the epoch of last scattering, made possible by
1299: improved measurements of the third acoustic peak; and new limits on the
1300: nongaussian parameter $f_{NL}$, based on additional data and the application of
1301: a new, more optimal bispectrum estimator. The 5 year data continue to favor a
1302: tilted primordial fluctuation spectrum, in the range $n_s \sim 0.96$, but a
1303: purely scale invariant spectrum cannot be ruled out at $\gt$3 $\sigma$
1304: confidence.
1305:
1306: The \map\ observatory continues to operate at L2 as designed, and the addition
1307: of two years of flight data has allowed us to make significant advances in
1308: characterizing the instrument. Additional data beyond 5 years will give us a
1309: better understanding of the instrument, especially with regards to the W band
1310: polarization data since the number of jackknife combinations scales like the
1311: square of the number of years of operation. If W band data can be incorporated
1312: into the EE power spectrum estimate, it would become possible to constrain a
1313: second reionization parameter and thereby further probe this important epoch in
1314: cosmology. The \map\ data continues to uphold the standard $\Lambda$CDM model
1315: but more data may reveal new surprises.
1316:
1317:
1318: \section{DATA PRODUCTS}
1319: \label{sec:data}
1320:
1321: All of the \map\ data is released to the research community for further analysis
1322: through the Legacy Archive for Microwave Background Data Analysis (LAMBDA) at
1323: http://lambda.gsfc.nasa.gov. The products include the complete 5 year
1324: time-ordered data archive (both raw and calibrated); the calibrated sky maps in
1325: a variety of processing stages (single year by DA, multi-year by band, high
1326: resolution and low resolution, smoothed, foreground-subtracted, and so forth);
1327: the angular power spectra and cosmological model likelihood code; a full table
1328: of model parameter values for a variety of model and data sets (including the
1329: best-fit model spectra and Markov chains); and a host of ancillary data to
1330: support further analysis. The \map\ Explanatory Supplement provides detailed
1331: information about the \map\ in-flight operations and data products
1332: \citep{limon/etal:prep}.
1333:
1334:
1335: \acknowledgements
1336:
1337: The \map\ mission is made possible by the support of the Science Mission
1338: Directorate Office at NASA Headquarters. This research was additionally
1339: supported by NASA grants NNG05GE76G, NNX07AL75G S01, LTSA03-000-0090,
1340: ATPNNG04GK55G, and ADP03-0000-092. EK acknowledges support from an Alfred P.
1341: Sloan Research Fellowship. This research has made use of NASA's Astrophysics
1342: Data System Bibliographic Services. We acknowledge use of the HEALPix, CAMB,
1343: CMBFAST, and CosmoMC packages.
1344:
1345:
1346: %\bibliographystyle{apj}
1347: %\input ms.bbl
1348: \bibliographystyle{wmap}
1349: \bibliography{wmap}
1350:
1351: \appendix
1352:
1353: \section{FISHER MATRIX ANALYSIS OF CALIBRATION AND SKY MAP FITS}
1354: \label{app:cal_fisher}
1355:
1356: \subsection{Least Squares Calibration and Sky Model Fitting}
1357:
1358: Let $i$ be a time index in the time ordered data. Let $g^j$ be
1359: parameters for the gain, $a_{l m}$ be parameters for the
1360: temperature anisotropy and $b^k$ be parameters for the baseline
1361: offset.
1362:
1363: The model of the time-ordered data (TOD) is
1364: \beq
1365: m_i = g_i \left[\Delta T_{vi} + \Delta T_{ai} \right] + b_i,
1366: \eeq
1367: where $i$ is a time index, $\Delta T_{vi}$ is the differential dipole signal at
1368: time step $i$, including the CMB dipole, and $\Delta T_{ai}$ is the differential
1369: anisotropy signal at time step $i$. The parameters of the model are the hourly
1370: gain and baseline values, and the sky map pixel temperatures (which goes into
1371: forming $\Delta T_a$. We fit for them by minimizing
1372: \beq
1373: \chi^2 = \sum_i \frac{(c_i - m_i)^2}{\sigma_i^2},
1374: \eeq
1375: where $c_i$ is the raw data, in counts, and $\sigma_i$ is the {\em rms} of the
1376: $i$th observation, in counts. The Fisher matrix requires taking the second
1377: derivative of $\chi^2$ with respect to all parameters being fit. In order to
1378: reduce the dimensionality of the problem to something manageable, we expand the
1379: calibration and sky signal in terms of a small number of parameters. We can
1380: write
1381: \beqa
1382: g_i & = & \sum_j g^j G_{ji}, \\
1383: b_i & = & \sum_k b^k B_{ki}, \\
1384: \Delta T_{ai} & = & \sum_{lm} a_{lm} \left[Y_{lm}(\hat{n}_{Ai})
1385: - Y_{lm}(\hat{n}_{Bi})\right],
1386: \eeqa
1387: where $G$ and $B$ are function of time (defined below), $a_{lm}$ are the
1388: harmonic coefficients of the map, and $\hat{n}_{Ai}$ is the unit vector of the
1389: $A$-side feed at time step $i$, and likewise for $B$.
1390:
1391: A reasonable set of basis functions for the gain and baseline allow for an
1392: annual modulation and a small number of higher harmonics. Note that this does
1393: not include power at the spin or precession period, which might be an important
1394: extension to consider. For now we consider the trial set
1395: \beq
1396: G_{ji} = \left\{
1397: \begin{array}{ll}
1398: 1 & j=0 \\
1399: \cos j\theta_i & j=1,\ldots,j_{\rm max} \\
1400: \sin (j-j_{\rm max})\theta_i & j=j_{\rm max}+1,\ldots,2j_{\rm max}
1401: \end{array}
1402: \right. ,
1403: \eeq
1404: and
1405: \beq
1406: B_{ki} = \left\{
1407: \begin{array}{ll}
1408: 1 & k=0 \\
1409: \cos k\theta_i & k=1,\ldots,k_{\rm max} \\
1410: \sin (k-k_{\rm max})\theta_i & k=k_{\rm max}+1,\ldots,2k_{\rm max}
1411: \end{array}
1412: \right. ,
1413: \eeq
1414: where $\theta = \tan^{-1}(\hat{n}_y/\hat{n}_x)$. Here $\hat{n}$ is the unit
1415: vector from \map\ to the Sun, and the components are evaluated in ecliptic
1416: coordinates.
1417:
1418: \subsection{Evaluation of the Fisher Matrix}
1419:
1420: We wish to evaluate the 2nd derivative
1421: \beq
1422: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial p_i \partial p_j}
1423: \eeq
1424: where $p_i$ and $p_j$ are the parameters we are trying to fit. The needed first
1425: derivatives are
1426: \beq
1427: \frac{1}{2}\,\frac{\partial \chi^2}{\partial g^{j'}}
1428: = -\sum_i \frac{(c_i - m_i) G_{j'i} \left[\Delta T_{vi} + \Delta T_{ai}\right]}
1429: {\sigma_i^2},
1430: \eeq
1431: \beq
1432: \frac{1}{2}\,\frac{\partial \chi^2}{\partial b^{k'}}
1433: = -\sum_i \frac{(c_i - m_i) B_{k'i}}
1434: {\sigma_i^2},
1435: \eeq
1436: \beq
1437: \frac{1}{2}\,\frac{\partial \chi^2}{\partial a_{l'm'}}
1438: = -\sum_i \frac{(c_i - m_i) g_i
1439: \left[Y_{l'm'}(\hat{n}_{Ai}) - Y_{l'm'}(\hat{n}_{Bi})\right]}
1440: {\sigma_i^2}.
1441: \eeq
1442: Then
1443: \beq
1444: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial g^{j'} \partial g^{j''}}
1445: = \sum_i \frac{G_{j'i} \left[\Delta T_{vi} + \Delta T_{ai}\right]
1446: \,G_{j''i} \left[\Delta T_{vi} + \Delta T_{ai}\right]}
1447: {\sigma_i^2}
1448: \eeq
1449: \beq
1450: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial g^{j'} \partial a_{l'm'}}
1451: = \sum_i \frac{g_i \left[Y_{l'm'}(\hat{n}_{Ai}) - Y_{l'm'}(\hat{n}_{Bi})\right]
1452: G_{j'i} \left[\Delta T_{v i} + \Delta T_{ai}\right]}
1453: {\sigma_i^2} + {\cal O} \sum_i (c_i - m_i)
1454: \eeq
1455: \beq
1456: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial g^{j'} \partial b^{k'}}
1457: = \sum_i \frac{B_{k'i} G_{j'i} \left[\Delta T_{vi} + \Delta T_{ai}\right]}
1458: {\sigma_i^2}
1459: \eeq
1460: \beq
1461: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial a_{l'm'} \partial a_{l''m''}}
1462: = \sum_i \frac{g_i \left[Y_{l'm'}(\hat{n}_{Ai}) - Y_{l'm'}(\hat{n}_{Bi})\right]
1463: \, g_i \left[Y_{l''m''}(\hat{n}_{Ai}) - Y_{l''m''}(\hat{n}_{Bi})\right]}
1464: {\sigma_i^2}
1465: \eeq
1466: \beq
1467: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial a_{l'm'} \partial b^{k'}}
1468: = \sum_i \frac{g_i B_{k'i} \left[Y_{l'm'}(\hat{n}_{Ai}) - Y_{l'm'}(\hat{n}_{Bi})\right]}
1469: {\sigma_i^2}
1470: \eeq
1471: \beq
1472: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial b^{k'} \partial b^{k''}}
1473: = \sum_i \frac{B_{k''i} B_{k'i}}{\sigma_i^2}
1474: \eeq
1475:
1476: From this we can form the inverse covariance matrix
1477: \beq
1478: C^{-1} = \left(
1479: \begin{array}{ccc}
1480: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial g^{j'} \partial g^{j''}} &
1481: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial g^{j'} \partial a_{l''m''}} &
1482: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial g^{j'} \partial b^{k''}}
1483: \vspace{3mm} \\
1484: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial a_{l'm'} \partial g^{j''}} &
1485: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial a_{l'm'} \partial a_{l''m''}} &
1486: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial a_{l'm'} \partial b^{k''}}
1487: \vspace{3mm} \\
1488: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial b^{k'} \partial g^{j''}} &
1489: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial b^{k'} \partial a_{l''m''}} &
1490: \frac{1}{2}\frac{\partial^2 \chi^2}{\partial b^{k'} \partial b^{k''}}
1491: \end{array}
1492: \right),
1493: \eeq
1494: where the gain and baseline blocks are $(2j_{\rm max}+1) \times (2j_{\rm
1495: max}+1)$, and the sky map block is $(l_{\rm max}+1)^2 \times (l_{\rm max}+1)^2$.
1496:
1497: If we decompose $C^{-1}$ using SVD the parameter covariance matrix can be
1498: inverted to have the form
1499: \beq
1500: C = \sum_i \frac{1}{w_i} V_{(i)} \otimes V_{(i)}
1501: \eeq
1502: where the $w_i$ are the singular values, and the $V_{(i)}$ are the columns of
1503: the orthogonal matrix $V$. In this form, the uncertainty in the linear
1504: combination of parameters defined by $V_{(i)}$ is $1/w_i$.
1505:
1506:
1507: \section{CALIBRATION MODEL FITTING WITH GAIN ERROR TEMPLATES}
1508: \label{app:cal_model_fit}
1509:
1510: \subsection{Gain Error From Calibration Dipole Error}
1511:
1512: Consider a simple model where the input sky consists of only a pure fixed (CMB)
1513: dipole, described by the vector ${\bf d}_c$, and a dipole modulated by the
1514: motion of \map\ with respect to the Sun, described by the time-dependent vector
1515: ${\bf d}_v(t)$. The raw data produced by an experiment observing this signal is
1516: \beq
1517: c(t_i) = g(t_i)[\Delta t_c(t_i) + \Delta t_v(t_i)]
1518: \eeq
1519: where $c(t_i)$ is the TOD signal in counts, $g(t_i)$ is the true gain of the
1520: instrument and $\Delta t_m(t_i)$ is the differential signal produced by each
1521: dipole component ($m = c,v$) at time $t_i$ given the instrument pointing at that
1522: time. Note that we have suppressed the explicit baseline and noise terms here
1523: for simplicity.
1524:
1525: Now suppose we calibrate the instrument using an erroneous CMB dipole, ${\bf
1526: d}'_c = r{\bf d}_c = (1+\Delta r){\bf d}_c$, where $r$ is a number of order one
1527: (and $\Delta r \ll 1$ so we can ignore terms of order $\Delta r^2$). The fit
1528: gain, $g_f(t)$, will then roughly have the form
1529: \beq
1530: g_f(t) = \frac{c(t)}{|{\bf d}'_c + {\bf d}_v(t)|}
1531: = g(t)\frac{|{\bf d}_c + {\bf d}_v(t)|}{|r{\bf d}_c + {\bf d}_v(t)|},
1532: \eeq
1533: where the vertical bars indicate vector magnitude. Now define ${\bf d} \equiv
1534: {\bf d}_c + {\bf d}_v$ and expand to 1st order in $\Delta r$ to get
1535: \beq
1536: g_f(t) = g(t)\left[1 - \Delta r \frac{{\bf d}(t)\cdot{\bf d}_c}{{\bf d}(t)\cdot{\bf d}(t)}\right].
1537: \eeq
1538: Note that the term $({\bf d}\cdot{\bf d}_c)/({\bf d}\cdot{\bf d})$ is dominated
1539: by a constant component of order $d_c^2/(d_c^2+d_v^2) \sim 0.99$, followed by an
1540: annually modulated term that is suppressed by a factor of order $d_v/d_c$. Thus
1541: an erroneous calibration dipole induces a specific error in the fit gain that
1542: can be identified and corrected for, assuming the time dependence of the true
1543: gain is orthogonal to this form.
1544:
1545: \subsection{Gain Model Fitting}
1546:
1547: In theory, the way to do this is as follows. We have a set of data in the form
1548: of the fit gains, $g_{f,i}$ for each calibration sequence $i$, and we have a
1549: gain model, $G(t;p_n)$, which is a function of time and a set of model
1550: parameters $p_n$. Ideally we would like to fit the model to the true gain,
1551: $g(t)$, but since we don't know the true gain, the next best thing is to modify
1552: the gain model to have the same modulation form as the dipole gains have and to
1553: fit for this modulation simultaneously with the other gain model parameters.
1554: Thus $\chi^2$ takes the form
1555: \beq
1556: \chi^2 = \sum_i \frac{\left[g_i - G_i(p_n)\right]^2}{\sigma_i^2}
1557: = \sum_i \frac{\left[g_{f,i} - G_i(p_n)(1 - \Delta r f_{d,i})\right]^2}{\sigma_i^2},
1558: \eeq
1559: where $f_{d,i} \equiv ({\bf d}\cdot{\bf d}_c)/({\bf d}\cdot{\bf d})$ evaluated
1560: at time $t_i$, or is a function generated from simulations.
1561:
1562: Since the system is nonlinear, it must be minimized using a suitable nonlinear
1563: least squares routine. However, we can analyze the parameter covariance matrix
1564: directly by explicitly evaluating the 2nd derivative of $\chi^2$ with respect to
1565: the model parameters
1566: \beq
1567: C^{-1} = \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial p_j \partial p_k}.
1568: \eeq
1569: First compile the necessary 1st derivatives
1570: \beq
1571: \frac{1}{2}\,\frac{\partial \chi^2}{\partial \Delta r}
1572: = \sum_i \frac{\left[g_{f,i} - G_i(p_n)(1 - \Delta r f_{d,i})\right]\,
1573: (G_i \, f_{d,i})}
1574: {\sigma_i^2}
1575: \eeq
1576: \beq
1577: \frac{1}{2}\,\frac{\partial \chi^2}{\partial p_m}
1578: = \sum_i \frac{\left[g_{f,i} - G_i(p_n)(1 - \Delta r f_{d,i})\right]\,
1579: (-\partial G_i / \partial p_m)(1 - \Delta r f_{d,i})}
1580: {\sigma_i^2}
1581: \eeq
1582: (We evaluate the individual $\partial G / \partial p_m$ terms below.) Next the
1583: various 2nd derivatives are
1584: \beq
1585: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial \Delta r \partial \Delta r}
1586: = \sum_i \frac{(G_i \, f_{d,i})(G_i \, f_{d,i})}
1587: {\sigma_i^2},
1588: \eeq
1589: \beq
1590: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial \Delta r \partial p_m}
1591: = \sum_i \frac{(G_i \, f_{d,i})(-\partial G_i / \partial p_m)(1 - \Delta r f_{d,i})}
1592: {\sigma_i^2}
1593: + {\cal O} \sum_i (g_i - G_i),
1594: \eeq
1595: \beq
1596: \frac{1}{2}\,\frac{\partial^2 \chi^2}{\partial p_m \partial p_n}
1597: = \sum_i \frac{(\partial G_i / \partial p_m)(1 - \Delta r f_{d,i})
1598: (\partial G_i / \partial p_n)(1 - \Delta r f_{d,i})}
1599: {\sigma_i^2}
1600: + {\cal O} \sum_i (g_i - G_i).
1601: \eeq
1602: In the last two expressions, we neglect the term proportional to $\partial^2G /
1603: \partial p_m \partial p_n$ because the prefactor of $(g_i-G_i)$ is statistically
1604: zero for the least squares solution.
1605:
1606: Finally, we evaluate the $\partial G / \partial p_m$ terms. The gain model has
1607: the form \citep{jarosik/etal:2007}
1608: \beq
1609: G_i = \alpha\frac{\bar{V}(t_i)-V_0-\beta(T_{\rm RXB}(t_i)-T^0_{\rm RXB})}{T_{\rm FPA}(t_i)-T^0_{\rm FPA}},
1610: \eeq
1611: where $T^0_{\rm RXB} \equiv 290$ K, and $\alpha$, $V_0$, and $T^0_{\rm FPA}$ are
1612: parameters to be fit. The necessary 1st derivatives are
1613: \beq
1614: \partial G_i / \partial \alpha
1615: = \frac{\bar{V}(t_i)-V_0-\beta(T_{\rm RXB}(t_i)-T^0_{\rm RXB})}{T_{\rm FPA}(t_i)-T^0_{\rm FPA}},
1616: \eeq
1617: \beq
1618: \partial G_i / \partial V_0
1619: = \frac{-\alpha}{T_{\rm FPA}(t_i)-T^0_{\rm FPA}},
1620: \eeq
1621: \beq
1622: \partial G_i / \partial \beta
1623: = \frac{-\alpha(T_{\rm RXB}(t_i)-T^0_{\rm RXB})}{T_{\rm FPA}(t_i)-T^0_{\rm FPA}},
1624: \eeq
1625: \beq
1626: \partial G_i / \partial T^0_{\rm FPA}
1627: = \alpha\frac{\bar{V}(t_i)-V_0-\beta(T_{\rm RXB}(t_i)-T^0_{\rm RXB})}
1628: {(T_{\rm FPA}(t_i)-T^0_{\rm FPA})^2}.
1629: \eeq
1630:
1631:
1632: %----- Figures -----
1633:
1634: \clearpage
1635: \begin{figure}
1636: \epsscale{1.0}
1637: \plotone{f01.eps}
1638: \caption{Five-year temperature sky maps in Galactic coordinates smoothed with a
1639: $0.2\dg$ Gaussian beam, shown in Mollweide projection. {\it top}: K band (23
1640: GHz), {\it middle-left}: Ka band (33 GHz), {\it bottom-left}: Q band (41 GHz),
1641: {\it middle-right}: V band (61 GHz), {\it bottom-right}: W band (94 GHz).}
1642: \label{fig:i_maps}
1643: \end{figure}
1644:
1645: \clearpage
1646: \begin{figure}
1647: \epsscale{1.0}
1648: \plotone{f02.eps}
1649: \caption{Five-year Stokes Q polarization sky maps in Galactic coordinates
1650: smoothed to an effective Gaussian beam of $2.0\dg$, shown in Mollweide
1651: projection. {\it top}: K band (23 GHz), {\it middle-left}: Ka band (33 GHz),
1652: {\it bottom-left}: Q band (41 GHz), {\it middle-right}: V band (61 GHz), {\it
1653: bottom-right}: W band (94 GHz).}
1654: \label{fig:q_maps}
1655: \end{figure}
1656:
1657: \clearpage
1658: \begin{figure}
1659: \epsscale{1.0}
1660: \plotone{f03.eps}
1661: \caption{Five-year Stokes U polarization sky maps in Galactic coordinates
1662: smoothed to an effective Gaussian beam of $2.0\dg$, shown in Mollweide
1663: projection. {\it top}: K band (23 GHz), {\it middle-left}: Ka band (33 GHz),
1664: {\it bottom-left}: Q band (41 GHz), {\it middle-right}: V band (61 GHz), {\it
1665: bottom-right}: W band (94 GHz).}
1666: \label{fig:u_maps}
1667: \end{figure}
1668:
1669: \clearpage
1670: \begin{figure}
1671: \epsscale{1.0}
1672: \plotone{f04.eps}
1673: \caption{Five-year polarization sky maps in Galactic coordinates smoothed to an
1674: effective Gaussian beam of $2.0\dg$, shown in Mollweide projection. The color
1675: scale indicates polarized intensity, $P = \sqrt{Q^2+U^2}$, and the line segments
1676: indicate polarization direction in pixels whose signal-to-noise exceeds 1. {\it
1677: top}: K band (23 GHz), {\it middle-left}: Ka band (33 GHz), {\it bottom-left}: Q
1678: band (41 GHz), {\it middle-right}: V band (61 GHz), {\it bottom-right}: W band
1679: (94 GHz).}
1680: \label{fig:p_maps}
1681: \end{figure}
1682:
1683: \clearpage
1684: \begin{figure}
1685: \epsscale{0.9}
1686: \plotone{f05.eps}
1687: \caption{Difference between the 5 year and 3 year temperature maps. {\it left
1688: column}: the difference in the maps, as delivered, save for the subtraction of a
1689: relative offset (Table~\ref{tab:i_diff}), {\it right column}: the difference
1690: after correcting the 3 year maps by a scale factor that accounts for the mean
1691: gain change, $\sim 0.3$\%, between the 3 year and 5 year estimates. {\it top to
1692: bottom}: K, Ka, Q, V, W band. The differences before recalibration are
1693: dominated by galactic plane emission and a dipole residual: see
1694: Table~\ref{tab:i_diff}, which also gives the changes for $l=2,3$.}
1695: \label{fig:diff_maps_i_recal}
1696: \end{figure}
1697:
1698: \clearpage
1699: \begin{figure}
1700: \epsscale{0.75}
1701: \plotone{f06.eps}
1702: \caption{Gain convergence tests using the iterative sky map \& calibration
1703: solver run on a pair of simulations with known, but different, inputs. Both
1704: panels show the recovered gain as a function of iteration number for a 4-channel
1705: K band simulation. The initial calibration guess was chosen to be in error by
1706: 1\% to test convergence; the output solutions, extrapolated with an exponential
1707: fit, are printed in each panel. {\it top}: Results for a noiseless simulation
1708: that includes a dipole signal (with Earth-velocity modulation) plus CMB and
1709: foreground anisotropy (the former is evaluated at the center frequency of each
1710: channel). The input gain was set to be constant in time. The extrapolated
1711: solutions agree with the input values to much better than 0.1\%. {\it bottom}:
1712: Results for a noiseless simulation that includes only dipole signal (with
1713: Earth-velocity modulation) but no CMB or foreground signal. In this case the
1714: input gain was set up to have flight-like thermal variations. The extrapolated
1715: absolute gain recovery was in error by $\gt$0.3\%, indicating a small residual
1716: degeneracy between the sky model and the time-dependent calibration.}
1717: \label{fig:gain_converge}
1718: \end{figure}
1719:
1720: \clearpage
1721: \begin{figure}
1722: \epsscale{1.0}
1723: \plotone{f07.eps}
1724: \caption{Gain error recovery test from a flight-like simulation that includes
1725: every effect known to be important. Using the daily dipole gains recovered from
1726: the iterative sky map \& calibration solver as input, the gain convergence
1727: error, shown here, is fit simultaneously with the gain model parameters, not
1728: shown, following the procedure outlined in Appendix~\ref{app:cal_model_fit}.
1729: The red trace indicates the true gain error for each \map\ channel, based on the
1730: known input gain and the gain solution achieved by the iterative solver on its
1731: final iteration. The black trace shows the gain error recovered by the fit,
1732: averaged by frequency band. The channel-to-channel scatter within a band is
1733: $\lt$0.1\%, though the mean of Ka band error is of order 0.1\%.}
1734: \label{fig:gain_recover}
1735: \end{figure}
1736:
1737: \clearpage
1738: \begin{figure}
1739: \epsscale{1.0}
1740: \plotone{f08.eps}
1741: \caption{W band EE power spectrum likelihood from $l=2-7$ using two separate
1742: estimation methods: {\it black}: maximum likelihood and {\it red}:
1743: pseudo-$C_l$. The vertical dashed lines indicate the best-fit model power
1744: spectrum based on fitting the combined Ka, Q, and V band data. The two spectrum
1745: estimates are consistent with each other, except at $l=3$. The maximum
1746: likelihood estimates are wider because they include cosmic variance whereas the
1747: pseudo-$C_l$ estimates account for noise only. Both estimates show excess power
1748: in the W band data relative to the best-fit model, and to the combined KaQV band
1749: spectrum, shown in Figure~6 of \citet{nolta/etal:prep}. The extreme excess in
1750: the $l=7$ pseudo-$C_l$ estimate is not so severe in the maximum likelihood, but
1751: both methods are still inconsistent with the best-fit model.}
1752: \label{fig:ww_ee}
1753: \end{figure}
1754:
1755: \clearpage
1756: \begin{figure}
1757: \epsscale{1.0}
1758: \plotone{f09.eps}
1759: \caption{{\it top}: Simulated W band data with a large polarized thermal
1760: emission signal injected, binned by solar azimuth angle. The red trace shows
1761: the input waveform based on the flight mirror temperature profile and a model of
1762: the polarized emissivity. The black profile is the binned co-added data which
1763: follows the input signal very well. The thickness of the points represents the 1
1764: $\sigma$ uncertainty due to white noise. {\it bottom}: Same as the top panel
1765: but for the 5 year flight data. The reduced $\chi^2$ of the binned data with
1766: respect to zero is 2.1 for 36 degrees of freedom, but this does not account for
1767: 1/f noise, so the significance of this result requires further investigation.
1768: However, the much larger signal in the simulation did not produce an EE spectrum
1769: with features present in the flight W band EE spectrum, so the feature in the
1770: binned flight data cannot account for the excess $l=7$ emission.}
1771: \label{fig:spin_bin}
1772: \end{figure}
1773:
1774: \clearpage
1775: \begin{figure}
1776: \epsscale{1.0}
1777: \plotone{f10.eps}
1778: \caption{The EE power spectrum computed from the null sky maps,
1779: $\frac{1}{2}S_{\rm Ka} - \frac{1}{2}S_{\rm QV}$, where S = Q,U are the
1780: polarization Stokes parameters, and $S_{\rm QV}$ is the optimal combination of
1781: the Q and V band data. The pink curve is the best-fit theoretical spectrum
1782: from \citet{dunkley/etal:prep}. The spectrum derived from the null maps is
1783: consistent with zero.}
1784: \label{fig:null_EE}
1785: \end{figure}
1786:
1787: \clearpage
1788: \begin{figure}
1789: \epsscale{1.0}
1790: \plotone{f11.eps}
1791: \caption{Estimates of the optical depth from a variety of data combinations.
1792: The dashed curve labeled Null uses the same null sky maps used in
1793: Figure~\ref{fig:null_EE}. The optical depth obtained from Ka band data alone
1794: (blue) is consistent with independent estimates from the combined Q and V band
1795: data (red). The final 5 year analysis uses Ka, Q, and V band data combined
1796: (black). These estimates all use a 1-parameter likelihood estimation, holding
1797: other parameters fixed except for the fluctuation amplitude, which is adjusted
1798: to fit the first acoustic peak in the TT spectrum \citep{page/etal:2007}. The
1799: degeneracy between $\tau$ and other $\Lambda$CDM parameters is small: see
1800: Figure~7 of \citet{dunkley/etal:prep}.}
1801: \label{fig:tau_test}
1802: \end{figure}
1803:
1804: \clearpage
1805: \begin{figure}
1806: \epsscale{0.80}
1807: \plotone{f12.eps}
1808: \caption{The foreground-reduced Internal Linear Combination (ILC) map based on
1809: the 5 year \map\ data.}
1810: \label{fig:ilc_map}
1811: \end{figure}
1812:
1813: \clearpage
1814: \begin{figure}
1815: \epsscale{1.0}
1816: \plotone{f13.eps}
1817: \caption{The temperature (TT) and temperature-polarization correlation (TE)
1818: power spectra based on the 5 year \map\ data. The addition of 2 years of data
1819: provide more sensitive measurements of the third peak in TT and the high-$l$ TE
1820: spectrum, especially the second trough.}
1821: \label{fig:tt_te}
1822: \end{figure}
1823:
1824:
1825:
1826: %----- Tables -----
1827:
1828: \begin{deluxetable}{cccccccccc}
1829: \tablecaption{Differencing Assembly (DA) Properties \label{tab:radiometers}}
1830: \tablewidth{0pt}
1831: \tablehead{
1832: \colhead{DA} &
1833: \colhead{$\lambda$\tablenotemark{a}} &
1834: \colhead{$\nu$\tablenotemark{a}} &
1835: \colhead{$g(\nu)$\tablenotemark{b}} &
1836: \colhead{$\theta_{\rm FWHM}$\tablenotemark{c}} &
1837: \colhead{$\sigma_0$(I)\tablenotemark{d}} &
1838: \colhead{$\sigma_0$(Q,U)\tablenotemark{d}} &
1839: \colhead{$\nu_{\rm s}$\tablenotemark{e}} &
1840: \colhead{$\nu_{\rm ff}$\tablenotemark{e}} &
1841: \colhead{$\nu_{\rm d}$\tablenotemark{e}} \\
1842: \colhead{} &
1843: \colhead{(mm)} &
1844: \colhead{(GHz)} &
1845: \colhead{} &
1846: \colhead{($^\circ$)} &
1847: \colhead{(mK)} &
1848: \colhead{(mK)} &
1849: \colhead{(GHz)} &
1850: \colhead{(GHz)} &
1851: \colhead{(GHz)}}
1852: \startdata
1853: K1 & 13.17 & 22.77 & 1.0135 & 0.807 & 1.436 & 1.453 & 22.47 & 22.52 & 22.78 \\
1854: Ka1 & 9.079 & 33.02 & 1.0285 & 0.624 & 1.470 & 1.488 & 32.71 & 32.76 & 33.02 \\
1855: Q1 & 7.342 & 40.83 & 1.0440 & 0.480 & 2.254 & 2.278 & 40.47 & 40.53 & 40.85 \\
1856: Q2 & 7.382 & 40.61 & 1.0435 & 0.475 & 2.141 & 2.163 & 40.27 & 40.32 & 40.62 \\
1857: V1 & 4.974 & 60.27 & 1.0980 & 0.324 & 3.314 & 3.341 & 59.65 & 59.74 & 60.29 \\
1858: V2 & 4.895 & 61.24 & 1.1010 & 0.328 & 2.953 & 2.975 & 60.60 & 60.70 & 61.27 \\
1859: W1 & 3.207 & 93.49 & 1.2480 & 0.213 & 5.899 & 5.929 & 92.68 & 92.82 & 93.59 \\
1860: W2 & 3.191 & 93.96 & 1.2505 & 0.196 & 6.565 & 6.602 & 93.34 & 93.44 & 94.03 \\
1861: W3 & 3.226 & 92.92 & 1.2445 & 0.196 & 6.926 & 6.964 & 92.34 & 92.44 & 92.98 \\
1862: W4 & 3.197 & 93.76 & 1.2495 & 0.210 & 6.761 & 6.800 & 93.04 & 93.17 & 93.84 \\
1863: \enddata
1864: \tablenotetext{a}{Effective wavelength and frequency for a thermodynamic
1865: spectrum.}
1866: \tablenotetext{b}{Conversion from antenna temperature to thermodynamic
1867: temperature, $\Delta T = g(\nu) \Delta T_A$.}
1868: \tablenotetext{c}{Full-width-at-half-maximum from radial profile of A- and
1869: B-side average beams. Note: beams are not Gaussian.}
1870: \tablenotetext{d}{Noise per observation for resolution 9 and 10 $I$, $Q$, \& $U$
1871: maps, to $\sim$0.1\% uncertainty. $\sigma(p)=\sigma_0 N_{\rm obs}^{-1/2}(p)$.}
1872: \tablenotetext{e}{Effective frequency for synchrotron (s), free-free (ff), and
1873: dust (d) emission, assuming spectral indices of $\beta = -2.9, -2.1, +2.0$,
1874: respectively, in antenna temperature units.}
1875: \end{deluxetable}
1876:
1877:
1878: \begin{deluxetable}{lccccc}
1879: \tablecaption{Lost and Rejected Data\label{tab:baddata}}
1880: \tablewidth{0pt}
1881: \tablehead{
1882: \colhead{Category} & \colhead{K-band} & \colhead{Ka-band} & \colhead{Q-band} &
1883: \colhead{V-band} & \colhead{W-band}}
1884: \startdata
1885: Lost or incomplete telemetry(\%) & 0.12 & 0.12 & 0.12 & 0.12 & 0.12 \\
1886: Spacecraft anomalies(\%) & 0.44 & 0.46 & 0.52 & 0.44 & 0.48 \\
1887: Planned stationkeeping maneuvers(\%) & 0.39 & 0.39 & 0.39 & 0.39 & 0.39 \\
1888: Planet in beam (\%) & 0.11 & 0.11 & 0.11 & 0.11 & 0.11 \\
1889: &------&------&------&------&------\\
1890: Total lost or rejected (\%) & 1.06 & 1.08 & 1.14 & 1.06 & 1.10 \\
1891: \enddata
1892: \end{deluxetable}
1893:
1894:
1895: \begin{deluxetable}{lrrrr}
1896: \tablecaption{Change in low-$l$ Power from 3 year Data \label{tab:i_diff}}
1897: \tablewidth{0pt}
1898: \tablehead{
1899: \colhead{Band} &
1900: \colhead{$l=0$\tablenotemark{a}} &
1901: \colhead{$l=1$\tablenotemark{a}} &
1902: \colhead{$l=2$\tablenotemark{b}} &
1903: \colhead{$l=3$\tablenotemark{b}} \\
1904: \colhead{} &
1905: \colhead{($\mu$K)} &
1906: \colhead{($\mu$K)} &
1907: \colhead{($\mu$K$^2$)} &
1908: \colhead{($\mu$K$^2$)} }
1909: \startdata
1910: K & 9.3 & 5.1 & 4.1 & 0.7 \\
1911: Ka & 18.9 & 2.1 & 2.8 & 0.2 \\
1912: Q & 18.3 & 0.4 & 2.5 & 0.5 \\
1913: V & 14.4 & 7.3 & 1.2 & 0.0 \\
1914: W & 16.4 & 3.5 & 1.0 & 0.0 \\
1915: \enddata
1916: \tablenotetext{a}{$l=0,1$ - Amplitude in the difference map, outside the
1917: processing cut, in $\mu$K.}
1918: \tablenotetext{b}{$l=2,3$ - Power in the difference map, outside the processing
1919: cut, $l(l+1)\,C_l/2\pi$, in $\mu$K$^2$.}
1920: \end{deluxetable}
1921:
1922:
1923: \begin{deluxetable}{lrrrrrr}
1924: \tablecaption{\map\ 5 year CMB Dipole Anisotropy\tablenotemark{a}\label{tab:dipole}}
1925: \tablewidth{0pt}
1926: \tablehead{
1927: \colhead{Cleaning} &
1928: \colhead{$d_x$\tablenotemark{b}} &
1929: \colhead{$d_y$} &
1930: \colhead{$d_z$} &
1931: \colhead{$d$\tablenotemark{c}} &
1932: \colhead{$l$} &
1933: \colhead{$b$} \\
1934: \colhead{method} &
1935: \colhead{(mK)} &
1936: \colhead{(mK)} &
1937: \colhead{(mK)} &
1938: \colhead{(mK)} &
1939: \colhead{($^{\circ}$)} &
1940: \colhead{($^{\circ}$)} }
1941: \startdata
1942: Templates & $-0.229 \pm 0.003$ & $-2.225 \pm 0.003$ & $2.506 \pm 0.003$ & $3.359 \pm 0.008$ & $264.11 \pm 0.08$ & $48.25 \pm 0.03$ \\
1943: ILC & $-0.238 \pm 0.003$ & $-2.218 \pm 0.002$ & $2.501 \pm 0.001$ & $3.352 \pm 0.007$ & $263.87 \pm 0.07$ & $48.26 \pm 0.02$ \\
1944: Combined & $-0.233 \pm 0.005$ & $-2.222 \pm 0.004$ & $2.504 \pm 0.003$ & $3.355 \pm 0.008$ & $263.99 \pm 0.14$ & $48.26 \pm 0.03$ \\
1945: \enddata
1946: \tablenotetext{a}{The CMB dipole components for two different galactic cleaning
1947: methods are given in the first two rows. The Gibbs samples from each set are
1948: combined in the last row to produce an estimate with conservative uncertainties
1949: that encompasses both cases.}
1950: \tablenotetext{b}{The cartesian dipole components are given in Galactic
1951: coordinates. The quoted uncertainties reflect the effects of noise and sky cut,
1952: for illustration. An absolute calibration uncertainty of 0.2\% should be added
1953: in quadrature.}
1954: \tablenotetext{c}{The spherical components of the dipole are given in Galactic
1955: coordinates. In this case the quoted uncertainty in the magnitude, $d$,
1956: includes the absolute calibration uncertainty.}
1957: \end{deluxetable}
1958:
1959:
1960: \begin{deluxetable}{lcccccccccc}
1961: \tablecaption{Polarization $\chi^2$ Consistency Tests\tablenotemark{a}\label{tab:chi2_pol_l}}
1962: \tablewidth{0pt}
1963: \tablehead{\colhead{Multipole} &
1964: \colhead{KaKa} & \colhead{KaQ} & \colhead{KaV} & \colhead{KaW} & \colhead{QQ} &
1965: \colhead{QV} & \colhead{QW} & \colhead{VV} & \colhead{VW} & \colhead{WW} \\
1966: \colhead{} &
1967: \colhead{(10)\tablenotemark{b}} & \colhead{(50)} & \colhead{(50)} & \colhead{(100)} & \colhead{(45)} &
1968: \colhead{(100)} & \colhead{(200)} & \colhead{(45)} & \colhead{(200)} & \colhead{(190)}}
1969: \startdata
1970: \multicolumn{11}{l}{EE} \\
1971: 2 & 0.727 & 1.059 & 1.019 & 1.301 & 1.586 & 0.690 & 1.179 & 0.894 & 1.078 & 1.152 \\
1972: 3 & 1.373 & 0.994 & 1.683 & 1.355 & 1.092 & 1.614 & 1.325 & 1.005 & 1.386 & 1.519 \\
1973: 4 & 1.561 & 1.816 & 1.341 & 2.033 & 0.993 & 1.126 & 1.581 & 1.195 & 1.596 & 1.724 \\
1974: 5 & 0.914 & 1.313 & 1.062 & 1.275 & 1.631 & 1.052 & 1.155 & 0.589 & 0.881 & 1.252 \\
1975: 6 & 1.003 & 0.847 & 0.688 & 1.124 & 0.740 & 0.856 & 1.049 & 1.384 & 1.168 & 1.142 \\
1976: 7 & 0.600 & 0.671 & 0.689 & 0.936 & 0.936 & 0.780 & 0.864 & 0.900 & 1.064 & 1.015 \\
1977: 8 & 1.578 & 1.262 & 1.337 & 1.212 & 1.080 & 0.763 & 0.608 & 1.025 & 0.871 & 0.749 \\
1978: 9 & 0.760 & 0.710 & 0.891 & 0.820 & 0.582 & 0.726 & 0.651 & 0.791 & 0.821 & 0.795 \\
1979: 10 & 0.494 & 0.821 & 0.996 & 0.914 & 0.656 & 0.763 & 0.806 & 0.676 & 0.891 & 0.943 \\[2.0mm]
1980: \multicolumn{11}{l}{EB} \\
1981: 2 & 0.900 & 1.297 & 1.179 & 2.074 & 1.006 & 0.915 & 2.126 & 1.242 & 2.085 & 2.309 \\
1982: 3 & 0.719 & 1.599 & 0.651 & 2.182 & 1.295 & 0.986 & 2.739 & 1.095 & 3.276 & 3.157 \\
1983: 4 & 0.746 & 1.702 & 1.378 & 1.777 & 1.926 & 1.110 & 1.435 & 1.028 & 1.279 & 1.861 \\
1984: 5 & 1.161 & 0.948 & 0.945 & 1.003 & 1.149 & 1.232 & 1.468 & 0.699 & 1.122 & 1.516 \\
1985: 6 & 0.475 & 1.183 & 0.651 & 0.687 & 0.829 & 1.023 & 0.814 & 1.201 & 1.136 & 0.960 \\
1986: 7 & 1.014 & 1.007 & 0.829 & 0.700 & 0.817 & 0.759 & 1.112 & 0.616 & 0.802 & 1.233 \\
1987: 8 & 0.849 & 0.897 & 1.279 & 0.861 & 0.681 & 0.689 & 0.955 & 1.021 & 0.954 & 0.996 \\
1988: 9 & 0.743 & 0.734 & 1.007 & 1.112 & 0.820 & 0.798 & 0.686 & 0.882 & 0.808 & 0.824 \\
1989: 10 & 0.413 & 1.003 & 1.316 & 0.859 & 0.722 & 0.900 & 0.693 & 1.124 & 0.836 & 0.852 \\[2.0mm]
1990: \multicolumn{11}{l}{BB} \\
1991: 2 & 2.038 & 1.570 & 1.244 & 2.497 & 1.340 & 1.219 & 2.529 & 0.694 & 1.631 & 9.195 \\
1992: 3 & 0.756 & 0.868 & 0.808 & 1.817 & 3.027 & 1.717 & 3.496 & 0.601 & 2.545 & 5.997 \\
1993: 4 & 1.058 & 1.455 & 1.522 & 2.144 & 1.007 & 0.905 & 1.786 & 0.752 & 1.403 & 1.984 \\
1994: 5 & 1.221 & 1.659 & 1.742 & 2.036 & 0.889 & 1.057 & 1.271 & 1.078 & 1.660 & 1.255 \\
1995: 6 & 0.379 & 0.805 & 0.483 & 0.812 & 1.009 & 0.861 & 1.238 & 0.800 & 0.767 & 0.955 \\
1996: 7 & 1.925 & 1.594 & 0.967 & 1.332 & 1.074 & 0.817 & 0.928 & 0.772 & 0.994 & 1.024 \\
1997: 8 & 0.804 & 1.005 & 0.999 & 0.912 & 1.069 & 0.782 & 0.831 & 0.997 & 0.879 & 0.943 \\
1998: 9 & 0.320 & 0.489 & 0.502 & 0.450 & 0.884 & 0.491 & 0.729 & 0.748 & 0.664 & 0.959 \\
1999: 10 & 1.181 & 1.162 & 1.028 & 0.980 & 1.218 & 1.165 & 0.951 & 1.079 & 0.621 & 0.791 \\
2000: \enddata
2001: \tablenotetext{a}{Table gives $\chi^2$ per degree of freedom of the independent
2002: spectrum estimates per multipole per band or band-pair, estimated from the
2003: template-cleaned maps. See text for details.}
2004: \tablenotetext{b}{Second header row indicates the number of degrees of freedom
2005: in the reduced $\chi^2$ for that spectrum. See text for details.}
2006: \end{deluxetable}
2007:
2008:
2009: \begin{deluxetable}{lrr}
2010: \tablecaption{Loss Imbalance Coefficients\tablenotemark{a}}
2011: \tablewidth{0pt}
2012: \tablehead{
2013: \colhead{DA} & \colhead{$x_{\rm im,1}$} & \colhead{$x_{\rm im,2}$} \\
2014: \colhead{} & \colhead{(\%)} & \colhead{(\%)} }
2015: \startdata
2016: K1 & 0.012 & 0.589 \\
2017: Ka1 & 0.359 & 0.148 \\
2018: Q1 & -0.031 & 0.412 \\
2019: Q2 & 0.691 & 1.048 \\
2020: V1 & 0.041 & 0.226 \\
2021: V2 & 0.404 & 0.409 \\
2022: W1 & 0.939 & 0.128 \\
2023: W2 & 0.601 & 1.140 \\
2024: W3 & -0.009 & 0.497 \\
2025: W4 & 2.615 & 1.946 \\
2026: \enddata
2027: \tablenotetext{a}{{Loss} imbalance is defined as $x_{\rm im} = (\epsilon_A -
2028: \epsilon_B)/(\epsilon_A + \epsilon_B)$. See \S\ref{sec:pol_emiss} and
2029: \citet{jarosik/etal:2007} for details.}
2030: \label{tab:x_im}
2031: \end{deluxetable}
2032:
2033:
2034: \begin{deluxetable}{lccc}
2035: \tabletypesize{\footnotesize}
2036: \tablecaption{Cosmological Parameter Summary \label{tab:best_param}}
2037: \tablewidth{0pt}
2038: \tablehead{
2039: \colhead{Description} & \colhead{Symbol} & \colhead{\map-only} & \colhead{\map+BAO+SN}}
2040: \startdata
2041: \multicolumn{4}{c}{Parameters for Standard $\Lambda$CDM Model \tablenotemark{a}}
2042: \\[3.0mm]
2043: Age of universe
2044: &\ensuremath{t_0}
2045: &\ensuremath{13.69\pm 0.13\ \mbox{Gyr}}
2046: &\ensuremath{13.72\pm 0.12\ \mbox{Gyr}}
2047: \\[1.5mm]
2048: Hubble constant
2049: &\ensuremath{H_0}
2050: &\ensuremath{71.9^{+ 2.6}_{- 2.7}\ \mbox{km/s/Mpc}}
2051: &\ensuremath{70.5\pm 1.3\ \mbox{km/s/Mpc}}
2052: \\[1.5mm]
2053: Baryon density
2054: &\ensuremath{\Omega_b}
2055: &\ensuremath{0.0441\pm 0.0030}
2056: &\ensuremath{0.0456\pm 0.0015}
2057: \\[1.5mm]
2058: Physical baryon density
2059: &\ensuremath{\Omega_bh^2}
2060: &\ensuremath{0.02273\pm 0.00062}
2061: &\ensuremath{0.02267^{+ 0.00058}_{- 0.00059}}
2062: \\[1.5mm]
2063: Dark matter density
2064: &\ensuremath{\Omega_c}
2065: &\ensuremath{0.214\pm 0.027}
2066: &\ensuremath{0.228\pm 0.013}
2067: \\[1.5mm]
2068: Physical dark matter density
2069: &\ensuremath{\Omega_ch^2}
2070: &\ensuremath{0.1099\pm 0.0062}
2071: &\ensuremath{0.1131\pm 0.0034}
2072: \\[1.5mm]
2073: Dark energy density
2074: &\ensuremath{\Omega_\Lambda}
2075: &\ensuremath{0.742\pm 0.030}
2076: &\ensuremath{0.726\pm 0.015}
2077: \\[1.5mm]
2078: Curvature fluctuation amplitude, $k_0=0.002$ Mpc$^{-1}$ \tablenotemark{b}
2079: &\ensuremath{\Delta_{\cal R}^2}
2080: &\ensuremath{(2.41\pm 0.11)\times 10^{-9}}
2081: &\ensuremath{(2.445\pm 0.096)\times 10^{-9}}
2082: \\[1.5mm]
2083: Fluctuation amplitude at $8h^{-1}$ Mpc
2084: &\ensuremath{\sigma_8}
2085: &\ensuremath{0.796\pm 0.036}
2086: &\ensuremath{0.812\pm 0.026}
2087: \\[1.5mm]
2088: $l(l+1)C^{TT}_{220}/2\pi$
2089: &\ensuremath{C_{220}}
2090: &\ensuremath{5756\pm 42} $\mu$K$^2$
2091: &\ensuremath{5751^{+ 42}_{- 43}} $\mu$K$^2$
2092: \\[1.5mm]
2093: Scalar spectral index
2094: &\ensuremath{n_s}
2095: &\ensuremath{0.963^{+ 0.014}_{- 0.015}}
2096: &\ensuremath{0.960\pm 0.013}
2097: \\[1.5mm]
2098: Redshift of matter-radiation equality
2099: &\ensuremath{z_{\rm eq}}
2100: &\ensuremath{3176^{+ 151}_{- 150}}
2101: &\ensuremath{3253^{+ 89}_{- 87}}
2102: \\[1.5mm]
2103: Angular diameter distance to matter-radiation eq.\tablenotemark{c}
2104: &\ensuremath{d_A(z_{\rm eq})}
2105: &\ensuremath{14279^{+ 186}_{- 189}\ \mbox{Mpc}}
2106: &\ensuremath{14200^{+ 137}_{- 140}\ \mbox{Mpc}}
2107: \\[1.5mm]
2108: Redshift of decoupling
2109: &\ensuremath{z_{*}}
2110: &\ensuremath{1090.51\pm 0.95}
2111: &\ensuremath{1090.88\pm 0.72}
2112: \\[1.5mm]
2113: %Thickness of decoupling (FWHM)
2114: % &
2115: % &
2116: % &
2117: % \\[1.5mm]
2118: Age at decoupling
2119: &\ensuremath{t_{*}}
2120: &\ensuremath{380081^{+ 5843}_{- 5841}\ \mbox{yr}}
2121: &\ensuremath{376971^{+ 3162}_{- 3167}\ \mbox{yr}}
2122: \\[1.5mm]
2123: %Duration of decoupling
2124: % &
2125: % &
2126: % &
2127: % \\[1.5mm]
2128: Angular diameter distance to decoupling \tablenotemark{c,d}
2129: &\ensuremath{d_A(z_{*})}
2130: &\ensuremath{14115^{+ 188}_{- 191}\ \mbox{Mpc}}
2131: &\ensuremath{14034^{+ 138}_{- 142}\ \mbox{Mpc}}
2132: \\[1.5mm]
2133: Sound horizon at decoupling \tablenotemark{d}
2134: &\ensuremath{r_s(z_*)}
2135: &\ensuremath{146.8\pm 1.8\ \mbox{Mpc}}
2136: &\ensuremath{145.9^{+ 1.1}_{- 1.2}\ \mbox{Mpc}}
2137: \\[1.5mm]
2138: Acoustic scale at decoupling \tablenotemark{d}
2139: &$l_A(z_*)$
2140: &\ensuremath{302.08^{+ 0.83}_{- 0.84}}
2141: &\ensuremath{302.13\pm 0.84}
2142: \\[1.5mm]
2143: Reionization optical depth
2144: &\ensuremath{\tau}
2145: &\ensuremath{0.087\pm 0.017}
2146: &\ensuremath{0.084\pm 0.016}
2147: \\[1.5mm]
2148: Redshift of reionization
2149: &\ensuremath{z_{\rm reion}}
2150: &\ensuremath{11.0\pm 1.4}
2151: &\ensuremath{10.9\pm 1.4}
2152: \\[1.5mm]
2153: Age at reionization
2154: &\ensuremath{t_{\rm reion}}
2155: &$427^{+88}_{-65}$ Myr
2156: &$432^{+90}_{-67}$ Myr
2157: \\[1.5mm]
2158: %Physical baryon density (cm$^{-3}$)
2159: %Baryon to photon ratio
2160: %Baryon to matter ratio
2161: \multicolumn{4}{c}{Parameters for Extended Models \tablenotemark{e}}
2162: \\[3.0mm]
2163: Total density \tablenotemark{f}
2164: &\ensuremath{\Omega_{\rm tot}}
2165: &\ensuremath{1.099^{+ 0.100}_{- 0.085}}
2166: &\ensuremath{1.0050^{+ 0.0060}_{- 0.0061}}
2167: \\[1.5mm]
2168: Equation of state \tablenotemark{g}
2169: &\ensuremath{w}
2170: &\ensuremath{-1.06^{+ 0.41}_{- 0.42}}
2171: &\ensuremath{-0.992^{+ 0.061}_{- 0.062}}
2172: \\[1.5mm]
2173: Tensor to scalar ratio, $k_0=0.002$ Mpc$^{-1}$ \tablenotemark{b,h}
2174: &\ensuremath{r}
2175: &\ensuremath{< 0.43\ \mbox{(95\% CL)}}
2176: &\ensuremath{< 0.22\ \mbox{(95\% CL)}}
2177: \\[1.5mm]
2178: Running of spectral index, $k_0=0.002$ Mpc$^{-1}$ \tablenotemark{b,i}
2179: &\ensuremath{dn_s/d\ln{k}}
2180: &\ensuremath{-0.037\pm 0.028}
2181: &\ensuremath{-0.028\pm 0.020}
2182: \\[1.5mm]
2183: Neutrino density \tablenotemark{j}
2184: &\ensuremath{\Omega_\nu h^2}
2185: &\ensuremath{< 0.014\ \mbox{(95\% CL)}}
2186: &\ensuremath{< 0.0071\ \mbox{(95\% CL)}}
2187: \\[1.5mm]
2188: Neutrino mass \tablenotemark{j}
2189: &\ensuremath{\sum m_\nu}
2190: &\ensuremath{< 1.3\ \mbox{eV}\ \mbox{(95\% CL)}}
2191: &\ensuremath{< 0.67\ \mbox{eV}\ \mbox{(95\% CL)}}
2192: \\[1.5mm]
2193: Number of light neutrino families \tablenotemark{k}
2194: &\ensuremath{N_{\rm eff}}
2195: &\ensuremath{> 2.3\ \mbox{(95\% CL)}}
2196: &\ensuremath{4.4\pm 1.5}
2197: \\[1.5mm]
2198: \enddata
2199: \tablenotetext{a}{The parameters reported in the first section assume the 6
2200: parameter $\Lambda$CDM model, first using \map\ data only
2201: \citep{dunkley/etal:prep}, then using \map+BAO+SN data
2202: \citep{komatsu/etal:prep}.}
2203: \tablenotetext{b}{$k=0.002$ Mpc$^{-1}$ $\longleftrightarrow$ $l_{\rm eff} \approx 30$.}
2204: \tablenotetext{c}{Comoving angular diameter distance.}
2205: \tablenotetext{d}{$l_A(z_*) \equiv \pi \, \ensuremath{d_A(z_{*})} \, \ensuremath{r_s(z_*)}^{-1}$.}
2206: \tablenotetext{e}{The parameters reported in the second section place limits on
2207: deviations from the $\Lambda$CDM model, first using \map\ data only
2208: \citep{dunkley/etal:prep}, then using \map+BAO+SN data
2209: \citep{komatsu/etal:prep}. A complete listing of all parameter values and
2210: uncertainties for each of the extended models studied is available on LAMBDA.}
2211: \tablenotetext{f}{Allows non-zero curvature, $\Omega_k \ne 0$.}
2212: \tablenotetext{g}{Allows $w \ne -1$, but assumes $w$ is constant.}
2213: \tablenotetext{h}{Allows tensors modes but no running in scalar spectral index.}
2214: \tablenotetext{i}{Allows running in scalar spectral index but no tensor modes.}
2215: \tablenotetext{j}{Allows a massive neutrino component, $\Omega_{\nu} \ne 0$.}
2216: \tablenotetext{k}{Allows $N_{\rm eff}$ number of relativistic species. The last
2217: column adds the HST prior to the other data sets.}
2218: %\tablenotetext{z}{Assumes $T_{\rm CMB} = 2.725 \pm 0.001$ K and $n_{\gamma} =
2219: %410.4 \pm 0.9$ cm$^{-3}$ as inferred from \cobe\ \citep{mather/etal:1999}.}
2220: \end{deluxetable}
2221:
2222: \end{document}
2223: