1: \documentclass[useAMS,usenatbib]{mn2e}
2: \usepackage{graphics}
3: \usepackage{times}
4: \usepackage{amssymb}
5:
6: \def\clover{{\sevensize C}$_\ell${\sevensize OVER}}
7: \def\wmap{{\sevensize WMAP}}
8: \def\sdss{{\sevensize SDSS}}
9: \def\quad{{\sevensize QUaD}}
10: \def\bicep{{\sevensize BICEP}}
11: \def\boomerang{{\sevensize BOOMERANG}}
12: \def\capmap{{\sevensize CAPMAP}}
13: \def\spud{{\sevensize SPUD}}
14: \def\quiet{{\sevensize QUIET}}
15: \def\piper{{\sevensize PIPER}}
16: \def\polarbear{{\sevensize POLARBEAR}}
17: \def\ebex{{\sevensize EBEX}}
18: \def\spider{{\sevensize SPIDER}}
19: \def\cbi{{\sevensize CBI}}
20: \def\dasi{{\sevensize DASI}}
21: \def\brain{{\sevensize BRAIN}}
22: \def\planck{{\sevensize PLANCK}}
23: \def\maxipol{{\sevensize MAXIPOL}}
24: \def\cmbpolbpol{{\sevensize CMBPOL/BPOL}}
25:
26: \def\healpix{{\sevensize HEALPIX}}
27: \def\camb{{\sevensize CAMB}}
28: \def\fftw{{\sevensize FFTW}}
29: \def\camgrid{{\sevensize CAMGRID}}
30:
31: \def\be{\begin{equation}}
32: \def\ee{\end{equation}}
33: \def\ba{\begin{eqnarray}}
34: \def\ea{\end{eqnarray}}
35: \def\nn{\nonumber}
36: \def\lgl{\langle}
37: \def\rgl{\rangle}
38: \def\vnhat{\hat{\bmath{n}}}
39: \def\vx{\bmath{x}}
40: \def\vl{\bmath{l}}
41: \def\vL{\bmath{L}}
42: \def\valpha{\bmath{\alpha}}
43: \def\vbeta{\bmath{\beta}}
44: \def\vgrad{\bmath{\nabla}}
45: \def\vgamma{\bmath{\gamma}}
46: \def\mC{\mathbfss{C}}
47: \def\lsim{\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}}
48: \raise1pt\hbox{$<$}}} % less than or approx. symbol
49: \def\gsim{\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$}}
50: \raise1pt\hbox{$>$}}} % greater than or approx. symbol
51: \pagerange{\pageref{firstpage}--\pageref{lastpage}}
52: \pubyear{2009}
53:
54: \setlength{\textheight}{23.5cm}
55: \setlength{\topmargin}{-1.65cm}
56:
57: \begin{document}
58:
59: \label{firstpage}
60:
61: \title[Impact of modulation on CMB $B$-mode experiments]{Impact of
62: modulation on CMB $\bmath B$-mode polarization experiments}
63: \author[M.L. Brown et al.]{Michael L. Brown$^{1}$\thanks{E-mail: mbrown@mrao.cam.ac.uk},
64: Anthony Challinor$^{2,3}$, Chris
65: E. North$^{4}$\thanks{Current address: School of Physics and
66: Astronomy, Cardiff University, Queen's
67: Buildings, The Parade, Cardiff CF24 3A}, Bradley
68: R. Johnson$^{4}$\thanks{Current address: Department of Physics, University of
69: California, Berkeley, CA, 94720, USA},
70: \newauthor Daniel O'Dea$^{1}$ and David Sutton$^{4}$\\
71: $^{1}$ Astrophysics Group, Cavendish Laboratory, University of
72: Cambridge, Cambridge CB3 OHE \\
73: $^{2}$ Institute of Astronomy, University of Cambridge, Madingley
74: Road, Cambridge CB3 OHA \\
75: $^{3}$ DAMTP, Centre for Mathematical Sciences, University of
76: Cambridge, Wilberforce Road, Cambridge CB3 OWA \\
77: $^{4}$ Oxford Astrophysics, University of Oxford, Denys Wilkinson
78: Building, 1 Keble Road, OX1 3RH}
79: \date{\today}
80:
81: \maketitle
82:
83: \begin{abstract}
84: We investigate the impact of both slow and fast polarization
85: modulation strategies on the science return of upcoming ground-based
86: experiments aimed at measuring the $B$-mode polarization of the
87: cosmic microwave background. Using detailed simulations of
88: the \clover\, experiment, we compare the
89: ability of modulated and un-modulated observations to recover the
90: signature of gravitational waves in the polarized CMB sky in the
91: presence of a number of anticipated systematic effects. The general
92: expectations that fast modulation is helpful in
93: mitigating low-frequency detector noise, and that the additional
94: redundancy in the projection of the instrument's polarization
95: sensitivity directions onto the sky when modulating reduces the
96: impact of instrumental polarization, particularly for fast
97: modulation, are borne out by our simulations. Neither low-frequency
98: polarized atmospheric fluctuations nor systematic errors in the
99: polarization sensitivity directions are mitigated by
100: modulation. Additionally, we find no significant reduction in the
101: effect of pointing errors by modulation. For a \clover-like
102: experiment, pointing jitter should be negligible but any systematic
103: mis-calibration of the polarization coordinate reference system
104: results in significant $E$-$B$ mixing on all angular scales and will
105: require careful control. We also stress the importance of combining
106: data from multiple detectors in order to remove the effects of
107: common-mode systematics (such as un-polarized $1/f$ atmospheric
108: noise) on the measured polarization signal. Finally we compare the
109: performance of our simulated experiment with the predicted
110: performance from a Fisher analysis. We find good agreement between
111: the (optimal) Fisher predictions and the simulated experiment except
112: for the very largest scales where the power spectrum estimator we
113: have used introduces additional variance to the $B$-mode signal
114: recovered from our simulations. In terms of detecting the total
115: $B$-mode signal, including lensing, the Fisher analysis and the
116: simulations are in excellent agreement. For a detection of the
117: primordial $B$-mode signal only, using an input tensor-to-scalar
118: ratio of $r=0.026$, the Fisher analysis predictions are $\sim
119: 20$~per cent.
120: better than the simulated performance.
121: \end{abstract}
122:
123: \begin{keywords}
124: methods: statistical - methods: analytical -
125: cosmology: theory - cosmic microwave background
126: \end{keywords}
127:
128: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
129: % INTRO %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
130: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
131:
132: \section{Introduction}
133: \label{sec:intro}
134: There is currently a great deal of interest in the rapidly evolving
135: field of observations of the polarization of the cosmic microwave
136: background (CMB). This interest stems from the fact that such
137: observations have the potential to discriminate between inflation and
138: other early-universe models through their ability to constrain an
139: odd-parity $B$-mode polarization component induced by a stochastic
140: background of gravitational waves at the time of last
141: scattering~\citep{kamionkowsky97,seljak97}.
142:
143: From an observational point of view, we are still a long way off from a
144: detection of this $B$-mode signature of inflation. However, much
145: progress has been made recently with the detection of the much
146: stronger $E$-mode polarization signal on large scales by the \wmap\,
147: experiment \citep{page07,nolta08}.
148: On smaller scales, a growing number of
149: balloon-borne and ground-based experiments have also measured $E$-mode
150: polarization including \dasi\, \citep{leitch05}, \cbi\,
151: \citep{sievers07}, \boomerang\, \citep{montroy06}, \maxipol\,
152: \citep{wu07}, \capmap\, \citep{bischoff08} and \quad\,
153: \citep{pryke09}. Most recently, the high precision measurement of
154: small-scale polarization by the \quad\, experiment has, for the first
155: time, revealed a characteristic series of acoustic peaks in the
156: $E$-mode spectrum and put the strongest upper limits to date on the
157: small-scale $B$-mode polarization signal expected from gravitational
158: lensing by large-scale structure.
159:
160: Building on the experience gained from these pioneering experiments, a
161: new generation of experiments is now under
162: construction with the ambitious goal of observing the primordial
163: $B$-mode signal. Observing this signal is one of the most challenging goals
164: of modern observational cosmology. There are a number of reasons why
165: these types of observations are so difficult. First and foremost, the
166: sought-after signal is expected to be extremely small -- in terms of the
167: tensor-to-scalar ratio\footnote{%
168: Our normalisation conventions follow those adopted in the
169: \camb\, code~\citep{lewis00}, so that $r$ is the ratio of primordial
170: power spectra for gravitational waves and curvature perturbations.
171: Explicitly, for slow-roll inflation in a potential $V(\phi)$,
172: $2 r\approx M_{\rm Pl}^2 [V'(\phi)/V(\phi)]^2$ where
173: $M_{\rm Pl}= 2.436 \times 10^{18}\, \mathrm{GeV}/c^2$ is the reduced
174: Planck mass.}, the RMS polarization signal from primordial $B$-modes is
175: $0.4 \sqrt{r}\, \mu \mathrm{K}$, and the current 95 per cent limit $r< 0.22$ from
176: \wmap\, temperature and $E$-mode polarization plus distance
177: indicators~\citep{komatsu09} implies an RMS $<180\,\mathrm{nK}$. Secondly,
178: polarized emission from
179: our own galaxy and from extra-Galactic objects act as a foreground
180: contaminant in observing the CMB polarized sky. Although our
181: knowledge of such polarized foregrounds is currently limited,
182: particularly at the higher frequencies $\gsim 100\,\mathrm{GHz}$
183: of relevance to bolometer experiments, models suggest that such contamination
184: could be an order of magnitude larger than the sought-after signal on
185: the largest scales (e.g. \citealt{amblard07}).
186: Thirdly, gravitational lensing by large-scale structure
187: converts $E$-modes into $B$-modes on small to medium scales
188: (see~\citealt{lewis06} for a review) and acts
189: as a source of confusion in attempts to measure the primordial
190: $B$-mode signal \citep{knox02, kesden02}. Note however that the
191: lensing $B$-mode signal is a valuable source of cosmological
192: information in its own right and can be used to put unique constraints
193: on dark energy and massive neutrinos
194: (e.g. \citealt{kaplinghat03,smithchallinor06}).
195: Unfortunately these latter two effects (foreground contamination and
196: weak gravitational lensing) contrive in such a way as to render $B$-mode
197: polarization observations subject to contamination on all angular
198: scales (the primordial $B$-mode signal is dominated by foregrounds on
199: large scales whilst on smaller scales it is swamped by the lensing
200: signal). Last, but not least, exquisite control of systematic and
201: instrumental effects will be required, to much better than $100\, \mathrm{nK}$,
202: before any detection of $B$-modes can be claimed.
203: The sought-after signal is so small that
204: systematic and instrumental effects considered negligible for an
205: exquisitely precise measurement of $E$-modes say, could potentially
206: ruin a detection of $B$-modes, if left uncorrected. One possible
207: approach to mitigating some of these systematics in hardware is to
208: modulate the incoming polarization signal such that it is shifted to
209: higher frequency and thus away from low-frequency systematics which
210: would otherwise contaminate it. There are a number of
211: techniques for achieving this including the use of a rotating
212: half-wave plate (HWP; e.g.~\citealt{johnson07}),
213: phase-switching, or Faraday rotation modulators~\citep{keating03}.
214:
215: In this paper, we investigate the ability of
216: modulation techniques that are either slow or fast
217: with respect to the temporal variation of the
218: signal
219: to mitigate a range of possible systematic
220: effects. Our analysis is based on simulations, and the
221: subsequent analysis, of data from a ground-based CMB $B$-mode
222: polarization experiment. Previous investigations of the impact of
223: systematic effects on $B$-mode observations include the analytic works
224: of \citet{hu03}, \citet{odea07} and~\citet{shimon08},
225: as well as the simulation-based
226: analysis of \cite{mactavish08} who based their study on signal-only
227: simulations of the \spider\, experiment. The simulation work presented
228: here is complementary to these previous analyses but we also take our
229: analysis further by including realistic noise in our simulations ---
230: we are thus able to quantify not only any bias found, but also any
231: degradation of performance due to the presence of systematic
232: effects. Our work makes use of a detailed simulation pipeline which we
233: have created in the context of the \clover\, experiment. Although the
234: precise details of our simulations are specific to \clover\,, our
235: general conclusions regarding the impact of modulation on a variety of
236: systematic effects are relevant to all upcoming ground-based $B$-mode
237: experiments and many of them are also relevant for both balloon-borne
238: and space-based missions.
239:
240: The paper is organised as follows. In Section \ref{sec:cmb_expts}, we
241: review the relevant upcoming $B$-mode polarization
242: experiments. Section \ref{sec:sims} describes our simulation technique
243: and the systematic effects we have considered. In Section
244: \ref{sec:analysis}, we describe the map-making and power spectrum
245: estimation techniques that we use to analyse the simulated
246: data. Section \ref{sec:results} presents the results from our main
247: analysis of systematic effects. We discuss our results in Section
248: \ref{sec:discussion} where, for clarity, we group the possible
249: systematic effects considered into those that are, and those that are
250: not, mitigated by a modulation scheme. In this section, we also
251: demonstrate the importance of combining information from multiple
252: detectors during analysis and compare the simulated performance of
253: \clover\, to the predicted performance from a Fisher
254: analysis. Our conclusions are summarised in Section
255: \ref{sec:conclusions}. Finally, Appendix~\ref{app:pointing} develops
256: a simple model of the spurious $B$-mode power produced by
257: pointing jitter for experiments with a highly redundant scan strategy.
258:
259: \section{CMB polarization experiments}
260: \label{sec:cmb_expts}
261: A host of experiments are currently under construction (one is, in
262: fact, already observing) with their primary goal being to constrain the
263: tensor-to-scalar ratio, and thus the energy scale of inflation,
264: through observations of the $B$-mode component of the CMB. Here, we
265: give a short summary of the planned experiments.
266: \vspace{-3mm}
267: \subsubsection*{(i) Ground-based experiments:}
268: \begin{itemize}
269: \item{{\sevensize BICEP/BICEP-2/KECK} array:} The \bicep\,
270: experiment \citep{yoon06} has recently completed its third and final season of
271: operation from the South Pole. The experiment consisted of a total of 98 polarization-sensitive
272: bolometers (PSBs) at 100 and 150~GHz. The optical design is
273: very clean but with the downside of poor resolution -- 45 arcmin
274: full-width at half-maximum (FWHM) at 150~GHz -- which limits the
275: target multipole range to $\ell <300$. The stated target sensitivity
276: is $r = 0.1$. In the first observing season, three 100~GHz pixels and
277: three 150~GHz pixels were equipped with Faraday rotation modulators
278: and two pixels operated at 220~GHz. {\sevensize
279: BICEP-2} will consist of an upgrade to the \bicep\, telescope with a
280: 150~GHz 512-element array of antenna-coupled detectors. It will be
281: deployed to the South Pole in November, 2009. The {\sevensize KECK}
282: array will consist of three {\sevensize BICEP-2}-like telescopes
283: (at 100, 150 and 220~GHz). It is hoped to be installed on the
284: \dasi\, mount (previously occupied by \quad) in November 2010. The
285: nominal goal of this array is $r=0.01$.
286:
287: \vspace{3mm}\item{{\clover}:} For an up-to-date overview
288: see~\citet{north08}. \clover\, is a
289: three-frequency (97, 150 and 225~GHz) instrument to be sited at Pampa La
290: Bola in the Atacama desert in Chile. It will have 576 single-polarization
291: transition-edge sensors (TES), split equally among the three frequencies.
292: The beam size of $\sim$ 5.5 arcmin FWHM at 150~GHz will sample the
293: multipole range $25 < \ell < 2000$. The target sensitivity is
294: $r \sim 0.03$ and the polarization signal will be modulated with a
295: HWP. The 97~GHz instrument is expected to be deployed to Chile in
296: late 2009 with the combined 150/225~GHz instrument to follow soon
297: after in 2010.
298:
299: \vspace{3mm}\item{{\quiet}:} See~\citet{samtleben08} for a recent overview.
300: \quiet\, is unique among
301: planned $B$-mode experiments in that it uses pseudo-correlation
302: HEMT-based receivers rather than bolometers. It will observe from Chile
303: using the CBI mount -- a planned second phase will involve upgrading
304: to $\sim 1000$ element arrays and relocation of the 7-m Crawford Hill
305: antenna from New Jersey to Chile. \quiet\, will observe
306: at 40 and 90~GHz. The target sensitivity for the second phase
307: is $r \sim 0.01$.
308:
309: \vspace{3mm}\item{{\polarbear}:} A three-frequency (90, 150 and
310: 220~GHz) single-dish instrument to be sited in the Inyo Mountains, CA
311: for its first year of operation, after which it will be relocated to
312: the Atacama desert in Chile. It will use 1280 TES bolometers at each
313: frequency with polarization modulation from a HWP. The planned beam
314: size is 4 arcmin (FWHM) at 150~GHz. The target sensitivity is
315: $r=0.015$ for the full instrument.
316:
317: \vspace{3mm}\item{{\brain}:} See~\cite{charlassier08} for a recent
318: review. \brain\, is a unique
319: bolometric interferometer project (c.f. \dasi, \cbi) to be sited on
320: the Dome-C site in Antarctica. The final instrument will have $\sim 1000$
321: bolometers observing at 90, 150 and 220~GHz. \brain\, will be primarily
322: sensitive to multipoles $50 < \ell < 200$. The full experiment is planned
323: to be operational in 2011 and the stated target sensitivity is $r = 0.01$.
324: \end{itemize}
325: \vspace{-5mm}
326: \subsubsection*{(ii) Balloon-borne experiments}
327: \begin{itemize}
328: \item{{\ebex}:} See \cite{oxley04} for a summary. \ebex\, will
329: observe at 150, 250 and 410~GHz and will fly a total of
330: $1406$ TES with HWP modulation.
331: The angular resolution is 8 arcmin and
332: the target multipole range is $20 < \ell < 2000$. The stated target
333: sensitivity is $r = 0.02$. A test
334: flight is planned for 2009 and a long-duration balloon (LDB)
335: flight is expected soon after.
336:
337: \vspace{3mm}\item{{\spider}:} See~\cite{crill08} for a recent description.
338: \spider\, will deploy $\sim 3000$ antenna-coupled TES observing
339: at 96, 145, 225 and 275~GHz, with a beam size of $\sim 40$ arcmin at 145~GHz.
340: The target multipole range is $10 < \ell < 300$. A 2-6 day first
341: flight is planned for 2010. The target sensitivity is
342: $r=0.01$. Signal modulation will be provided by a (slow) stepped HWP
343: and fast gondola rotation.
344:
345: \vspace{3mm}\item{{\piper}:} This balloon experiment will deploy a
346: focal plane of 5120 TES bolometers in a backshort-under-grid (BUG)
347: configuration. Each flight of \piper\, will observe at a different
348: frequency, covering 200, 270, 350 and 600~GHz after the four planned
349: flights. The beam size is $\sim15$ arcmin, corresponding to a target
350: multipole range $\ell < 800$. The first element of the optical system is
351: a variable polarization modulator (VPM). The entire optical chain,
352: including the modulators, are cooled to 1.5 K so that \piper\, observes
353: at the background limit for balloon altitudes. Including removal of
354: foregrounds, the experiment has the sensitivity to make a $2\sigma$
355: detection of $r = 0.007$. The first flight is scheduled for 2013.
356:
357: \end{itemize}
358: \vspace{-5mm}
359: \subsubsection*{(iii) Space missions}
360: \begin{itemize}
361: \item{{\planck}:} See the publication of the~\citet{planck06} for a detailed
362: review of the science programme. \planck\, will measure the temperature
363: in nine frequency bands, seven of which will have (some) polarization
364: sensitivity. The polarized channels (100, 143, 217 and 353~GHz)
365: of the high-frequency instrument
366: (HFI) use similar PSBs to those deployed on
367: \boomerang\, and have beam sizes 5--9.5 arcmin. For low $r$, sensitivity to
368: primordial gravitational waves will
369: mostly come from the large-angle reionisation
370: signature~\citep{zaldarriaga97b} and
371: $r=0.05$ may be possible if foregrounds allow. \planck\, will
372: be sensitive to the multipole range $2 < \ell <3000$ and is
373: scheduled to launch in 2009. The HFI has no active or fast signal modulation
374: (i.e.\ other than scanning).
375:
376: \vspace{3mm}\item{{\cmbpolbpol}:}
377: Design studies have been conducted
378: for satellite mission(s) dedicated to measuring primordial
379: $B$-modes comprising $\sim 2000$ detectors with the ability to measure
380: $0.001 < r < 0.01$ if foregrounds allow it~\citep{bock08,deBernardis08}.
381: The timescale for launch of any selected mission is likely beyond 2020.
382: \end{itemize}
383:
384: \section{Simulations}
385: \label{sec:sims}
386:
387: The \clover\, experiment will consist of two telescopes -- a low frequency
388: instrument with a focal plane consisting of 192
389: single-polarization 97~GHz TES detectors, and a high frequency instrument with
390: a combined focal plane of 150 and 225~GHz detectors (192 of
391: each). Note that we have not included foreground contamination in our
392: simulations, so for this analysis, we consider only the 150~GHz
393: detector complement -- the corresponding reduction in sensitivity will
394: approximate the effect of using the multi-frequency observations to
395: remove foregrounds. Figure~\ref{fig:hf_focal_plane} shows the
396: arrangement of the 150 and 225~GHz detectors on the high frequency
397: focal plane. The detectors are arranged in detector blocks consisting
398: of eight pixels each. Each pixel consists of two TES detectors which are
399: sensitive to orthogonal linear polarizations. The polarization
400: sensitivity of the eight detector pairs within a block are along, and at
401: right angles to, the major axis of their parent block. The detector
402: complement, both at 150 and 225~GHz, therefore consists of three
403: `flavours' of pixels with different polarization sensitivity
404: directions. The 97~GHz focal plane (not shown) has a similar mix of
405: detector orientations.
406: \begin{figure}
407: \centering
408: \resizebox{0.48\textwidth}{!}{
409: \rotatebox{-90}{\includegraphics{fig1.ps}}}
410: \caption{Layout of detectors on the \clover\, high frequency focal
411: plane. Each point indicates a pixel comprising two TES detectors
412: sensitive to orthogonal linear polarizations. The polarization
413: sensitivity directions of detectors within each block are along, and
414: at right angles to, the major axis of the block. The outlined
415: 150~GHz detector blocks at the centre of the array are used in the
416: simulations described here. The field of view of the entire array is
417: $\sim 5 \, {\rm deg}^2$.}
418: \label{fig:hf_focal_plane}
419: \end{figure}
420:
421: \subsection{Simulation parameters}
422: \label{sec:sims_params}
423: Because simulating the full \clover\, experiment is computationally
424: demanding (a single simulation of a two-year campaign
425: at the full \clover\, data rate would require $\sim 10^4$ CPU-hours),
426: we have scaled some of the simulation parameters in order to make our
427: analysis feasible.
428: \begin{enumerate}
429: \item We simulate only half the 150~GHz detector complement and have
430: scaled the noise accordingly. We have verified with a restricted
431: number of simulations using all the 150~GHz detectors that the
432: marginally more even coverage obtained across our field using all detectors
433: has little or no impact on our results or
434: conclusions. The 150~GHz detector blocks used in our simulations are
435: indicated in Fig.~\ref{fig:hf_focal_plane} and include all three
436: possible orientations of pixels on the focal plane.
437: \item The \clover\, detectors will have response times of $\sim 200\,
438: \mu$s and so the data will be sampled at $\sim 1$~kHz in order to
439: sample the detector response function adequately. Simulating at this
440: rate is prohibitive so we simulate at a reduced data rate of
441: $100$~Hz. For the \clover\, beam size (FWHM = 5.5 arcmin at 150~GHz) and
442: our chosen scan speed ($0.25^{\circ}$/s), this data rate is still fast
443: enough to sample the sky signal adequately.
444: \item \clover\, will observe four widely separated fields on the sky,
445: each covering an area of $\sim 300 \, {\rm deg}^2$, over the course
446: of two years. Two of the fields are in the southern sky and two
447: lie along the equator. For our analysis, we observe each of the four fields
448: for a single night only and, again, we have scaled the noise levels to those
449: appropriate for the full two-year observing campaign.
450: \end{enumerate}
451:
452: \subsection{Observing strategy}
453: \label{sec:obs_strategy}
454: Although optimisation of the observing strategy is not the focus of
455: this work, a number of possible strategies have been investigated by
456: the \clover\, team. For our analysis, we use the most favoured scan
457: strategy at the time of writing. To minimise rapid variations in
458: atmospheric noise, the two telescopes will scan back and forth in
459: azimuth at constant elevation, allowing the field to rise through the
460: chosen observing elevation. Every few hours, the elevation angle of
461: the telescopes will be re-pointed to allow for
462: field-tracking. Although the precise details of the scan are likely to
463: change, the general characteristics of the scan and resulting field
464: coverage properties will remain approximately the same due to the
465: limitations imposed by constant-elevation scanning and observing from
466: Atacama. The \clover\, telescopes are designed with the capability of
467: scanning at up to $10^{\circ}$/s. However, for our analysis where we
468: have considered \clover\, operating with a rotating HWP (Section
469: \ref{sec:modulation}), we have chosen a relatively slow scan speed of
470: $0.25^{\circ}$/s in light of the HWP rotation frequency which we have
471: employed ($f_\lambda = 3$~Hz). Although the mode of operation of a HWP
472: on \clover\, is still under development, a continuously rotating HWP
473: is likely to be restricted to rotation frequencies of $f_\lambda <
474: 5$~Hz due to mechanical constraints (with current cryogenic rotation
475: technologies, fast rotation, $\gsim 5$~Hz, could possibly result in
476: excessive heat generation). Figure~\ref{fig:hitmaps} shows the coverage maps for a
477: single day's observing on one of the southern fields and on one of the
478: equatorial fields. The corresponding maps for the other two fields are
479: broadly similar. Note that, in the real experiment, we expect that
480: somewhat more uniform field coverage than that shown in
481: Fig.~\ref{fig:hitmaps} will be achievable by employing slightly
482: different scan patterns on different days.
483: \begin{figure*}
484: \centering
485: \resizebox{0.80\textwidth}{!}{
486: \rotatebox{-90}{\includegraphics{fig2.ps}}}
487: \caption{Hit-maps for one of the southern fields (left; RA
488: 09:30 hrs , Dec -40.00$^{\circ}$) and one of the equatorial fields
489: (right; RA 04:00 hrs, Dec 0.00$^{\circ}$) for a single day's
490: observation with half the 150~GHz detector complement. The central
491: part of the fields (shown in yellow and red) are roughly $20^{\circ}$ in
492: diameter. These maps have been constructed using a \healpix\, resolution of
493: $N_{\rm side} = 1024$ corresponding to a pixel size $\sim$3.4 arcmin.}
494: \label{fig:hitmaps}
495: \end{figure*}
496:
497: \subsection{Signal simulations}
498: \label{sec:signal_sims}
499: We generate model $TT$, $EE$, $TE$ and $BB$ CMB power spectra using
500: \camb\, \citep{lewis00}. The input cosmology used consists
501: of the best-fit standard $\Lambda$CDM model to the 5-year \wmap\, data
502: set \citep{hinshaw09}, but with a tensor-to-scalar ratio of $r=0.026$,
503: chosen to match the \clover\, `target' value. Realisations of CMB
504: skies from these power spectra are then created using a modified
505: version of the \healpix\footnote{%
506: See http://healpix.jpl.nasa.gov}
507: software \citep{gorski05}. Our simulations include weak gravitational lensing
508: but ignore its non-Gaussian aspects. Using only Gaussian simulations
509: means that we slightly mis-estimate the covariance matrices of our
510: power spectrum estimates, particularly the $B$-mode
511: covariances~\citep{smith04,smithchallinor06}. For the \clover\, noise
512: levels, this is expected to have a negligible impact on the
513: significance level of the total $B$-mode signal.
514:
515: As part of the simulation process, the input CMB signal is convolved with a
516: perfect Gaussian beam with FWHM $=5.5$ arcmin. Note that an important
517: class of systematic effects which we do not consider in this paper are
518: those caused by imperfect optics; see the discussion in
519: Section~\ref{sec:systematics}.
520:
521: For our analysis of the simulated data sets (Section
522: \ref{sec:analysis}) we have chosen to reconstruct maps of the Stokes
523: parameters with a map resolution of $3.4$ arcmin (\healpix\, $N_{\rm
524: side} = 1024$). Note that this pixel size does not fully sample the
525: beam and it is likely that we will adopt $N_{\rm side} = 2048$ for the
526: analysis of real data. In order to isolate the effects of the various
527: systematics we have considered, our simulated CMB skies have therefore
528: been created at this same resolution --- we can then be sure that any
529: bias found in the recovered CMB signals is due to the systematic
530: effect under consideration rather than due to a poor choice of map
531: resolution\footnote{%
532: We have verified that spurious $B$-modes
533: generated through the pixelisation are negligible for the \clover\,
534: noise levels.}.
535: Note that our adopted procedure of simulating and map-making at
536: identical resolutions, although useful for the specific aims of this
537: paper, is not a true representation of a CMB observation. For real
538: observations, pixelisation of the CMB maps will introduce a bias to
539: the measured signal on scales comparable to the pixel size adopted.
540:
541: Using the pointing registers as provided by the scan strategy and
542: after applying the appropriate focal plane offsets for each detector,
543: we create simulated time-streams according to
544: \be
545: d_i = \left[ T(\theta) + Q(\theta) \cos(2\phi_i) + U(\theta) \sin(2\phi_i) \right]/2,
546: \label{eqn:signal_timestream}
547: \ee
548: where $\theta$ denotes the pointing and $T, Q$ and $U$ are the sky
549: signals as interpolated from the
550: input CMB sky map. The polarization angle, $\phi_i$ is, in general, a
551: combination of the polarization sensitivity direction of each detector,
552: any rotation
553: of the telescope around its boresight, the direction of travel of the
554: telescope in RA--Dec space and the orientation of the half-wave plate,
555: if present.
556:
557: \subsection{Noise simulations}
558: \label{sec:noise_sims}
559: The \clover\, data will be subject to several different noise
560: sources. Firstly, photon loading from the telescope, the atmosphere
561: and the CMB itself will subject the data to uncorrelated random
562: Gaussian noise. Secondly, the TES detectors used in \clover\, are
563: subject to their own sources of noise which will possibly
564: include low-frequency $1/f$ behaviour and correlations between
565: detectors. Thirdly, the atmosphere also has a very strong $1/f$
566: component which will be heavily correlated across the detector
567: array. Fortunately, the $1/f$ component of the atmosphere is known to
568: be almost completely un-polarized and so can be removed from the
569: polarization analysis by combining data from multiple detectors (see
570: Section \ref{sec:differencing}).
571:
572: The white-noise levels due to loading from the instrument, atmosphere
573: and CMB have been carefully modelled for the case of \clover\,
574: observations from Atacama. We will not present the details here, but
575: for realistic observing conditions and scanning elevations, we have
576: calculated the expected noise-equivalent temperature (NET) due to
577: photon noise alone to be $\approx 146 \, \mu {\rm K} \sqrt{\rm s}$. We
578: add this white noise component to our simulated signal time-streams
579: for each detector as
580: \be
581: d_i \rightarrow d_i + \frac{\rm NET}{2} \sqrt{f_{\rm samp}} g_i,
582: \ee
583: where $f_{\rm samp}$ is the sampling frequency and $g_i$ is a Gaussian
584: random number with $\mu = 0$ and $\sigma = 1$. Note that the white-noise
585: level in the detector time streams is ${\rm NET} / 2$ since the
586: \clover\, detectors are half-power detectors (equation
587: \ref{eqn:signal_timestream}).
588:
589: Using instrument parameters appropriate for the \clover\, detectors,
590: we use the small-signal TES model of \cite{irwin05} to create a
591: model noise power spectrum for the detector noise. This model includes
592: both a contribution from the super-conducting SQUIDs which will be used to
593: record the detector signals (e.g. \citealt{reintsema03}) and a
594: contribution from aliasing in the Multi Channel Electronics (MCE;
595: \citealt{battistelli08}) which will be used to read out the signals.
596: For the instrument parameters we have chosen, the effective
597: NET of the detector noise in our simulations is approximately equal
598: to the total combined photon noise contribution from the atmosphere, the
599: instrument and the CMB. Note however that for the final instrument, it
600: is hoped that the detector NET can be reduced to half that of the
601: total photon noise, thus making \clover\, limited by irreducible
602: photon loading.
603: The \cite{irwin05} small-signal TES model does not
604: include a $1/f$ component to the detector noise so in order
605: to investigate the impact of modulation on possible low-frequency
606: detector noise, we add a heuristically chosen $1/f$ component to the
607: detector noise model with knee frequencies in the range, $0.01 <
608: f_{\rm knee} < 0.1$~Hz.
609: The MCE system which will be used to read out the \clover\, data
610: should have low cross-talk between different channels. However,
611: correlations will be present at some level and so we include
612: 10 per cent correlations between all of our simulated detector noise
613: time-streams. Generally, to simulate stationary noise that is correlated in time and
614: across the $N_{\rm det}$ detectors we proceed as follows.
615: Let the noise cross-power spectrum
616: between detector $d$ and $d'$ be $P_{d,d'}(f)$.
617: Taking the Cholesky decomposition of this matrix at each frequency,
618: $L_{d,d'}(f)$, defined by
619: \be
620: P_{d,d'}(f) = \sum_{d''} L_{d,d''}(f) L_{d',d''}(f),
621: \ee
622: we apply $L$ to $N_{\rm det}$ independent, white-noise time streams
623: $g_{d}(f)$ in Fourier space,
624: \be
625: g_d(f) \rightarrow \sum_{d'} L_{d,d'}(f) g_{d'}(f) .
626: \ee
627: The resulting time-streams, transformed to
628: real space then possess the desired correlations between detectors. Here,
629: we assume that the correlations are independent of frequency so that
630: the noise cross-power spectrum takes the form
631: \be
632: P_{d,d'}(f) = C_{d,d'} P(f) ,
633: \ee
634: where the correlation matrix $C_{d,d'} = 1$ for $d = d'$ and $C_{d,d'} = 0.1$
635: otherwise. In practice, we use discrete Fourier transforms to synthesise
636: noise with periodic boundary conditions (and hence circulant time-time
637: correlations).
638:
639: We use the same technique to simulate the correlated $1/f$ component
640: of the atmosphere. We have measured the noise properties of the atmosphere
641: from data from the \quad\, experiment \citep{hinderks09}. The 150~GHz
642: frequency channel of \quad\, is obviously well matched to the
643: \clover\, 150~GHz channel although \quad\, observed the CMB from the
644: South Pole rather than from Atacama. Although there are significant
645: differences between the properties of the atmosphere at the South Pole
646: and at Atacama (e.g. \citealt{bussmann05}), the \quad\, observations
647: still represent the best estimate of the $1/f$ noise properties of
648: the atmosphere available at present. A rough fit of the \quad\, data to the model,
649: \be
650: P(f) = {\rm NET^2} \left[1 + \left( \frac{f_{\rm knee}}{f} \right)^\alpha \right],
651: \label{eqn:pk_atms}
652: \ee
653: yields a knee frequency, $f_{\rm knee} = 0.45$~Hz and spectral index,
654: $\alpha = 2.5$. Using this model power spectrum we simulated $1/f$
655: atmospheric noise correlated across the array in exactly the same way
656: as was used for the detector noise. Fortunately, the $1/f$ component
657: in the atmosphere is almost completely un-polarized.
658: If there were no instrumental polarization, detectors
659: within the same pixel (which always look in
660: the same direction) would therefore be completely correlated
661: with one another and detectors from different pixels would also be
662: heavily correlated. For the correlated atmosphere, we therefore use a
663: correlation matrix given by $C_{d,d'} = 1.0$ for $|d-d'| \le 1$
664: (i.e. for detector pairs) and $C_{d,d'} = 0.5$ otherwise. In the
665: following sections, as one of the systematics we have investigated, we
666: relax the assumption that the atmosphere is un-polarized.
667:
668: Figure~\ref{fig:noise_compare} compares the photon, atmospheric $1/f$
669: and detector noise contributions to our simulated data in frequency
670: space. At low frequencies, the noise is completely dominated by the
671: atmospheric $1/f$ while the white-noise contributions from photon
672: loading (including the uncorrelated component of the atmosphere) and
673: detector noise are approximately equal. Note that for observations
674: without active modulation, and for the scan speed and observing
675: elevations which we have adopted, the ``science band'' for the
676: multipole range $20 < \ell < 2000$ corresponds roughly to $0.01 < f <
677: 1$~Hz in time-stream frequency. In contrast, for our simulations which
678: include a continuously rotating HWP, the temperature signal remains within
679: the $0.01 < f < 1$~Hz frequency range but the polarized sky signal is
680: moved to a narrow band centred on $\sim$12~Hz, well away from both the
681: detector and atmospheric $1/f$ noise components (see
682: Section~\ref{sec:modulation} and Fig.~\ref{fig:mod_frequency}). Note also that although the
683: atmospheric $1/f$ dominates the detector $1/f$ at low frequency, the
684: atmosphere is heavily correlated across detectors and can therefore be
685: removed by combining detectors (e.g. differencing detectors within a
686: pixel) but this is not true for the detector noise which is only
687: weakly correlated between detectors. We demonstrate this in
688: Fig.~\ref{fig:tod_noise} where we plot a five-minute sample of simulated
689: atmospheric and detector noise for the two constituent detectors
690: within a pixel. Including the atmospheric $1/f$ component, the
691: effective total NET per detector measured from our simulated
692: time-streams is $293 \, \mu {\rm K} \sqrt{\rm sec}$ whilst excluding
693: atmospheric $1/f$, we measure $210 \, \mu {\rm K} \sqrt{\rm s}$.
694:
695: \begin{figure}
696: \centering
697: \resizebox{0.47\textwidth}{!}{
698: \rotatebox{-90}{\includegraphics{fig3.ps}}}
699: \caption{Frequency space comparison between the different noise
700: sources in the simulations. The grey line shows the detector noise
701: power spectrum (here with a $1/f$ component with knee frequency,
702: $f_{\rm knee} = 0.1$~Hz). The correlated component to the
703: atmosphere is shown as the dashed line and the total photon noise
704: (including atmospheric loading) is shown as the dotted line. For
705: the simulation parameters we have adopted, the temperature sky-signal from
706: multipoles $20 < \ell < 2000$ appears in the time-stream in the
707: frequency range $0.01 < f < 1$~Hz. In the absence of fast modulation,
708: the polarized sky signal also appears in this frequency range
709: whereas in our simulations including fast modulation, the polarized
710: sky signal appears in a narrow band centred on 12~Hz.}
711: \label{fig:noise_compare}
712: \end{figure}
713: \begin{figure}
714: \centering
715: \resizebox{0.48\textwidth}{!}{
716: \rotatebox{-90}{\includegraphics{fig4.ps}}}
717: \caption{A five-minute sample of simulated noise time-stream for the two
718: detectors within a pixel (denoted ``A'' and ``B'') for both the
719: atmospheric noise simulations (\emph{top}) and for the detector
720: noise simulations (\emph{bottom}). The $1/f$ component in the
721: atmospheric noise time-streams is 100 per cent correlated between the
722: A and B detectors and can be removed entirely by
723: differencing. In contrast the $1/f$ component in the detector noise
724: (which is much weaker and not noticeable on this plot)
725: is only weakly correlated between A and B and is not removed by
726: differencing. For comparison, the signal-only time-streams are
727: also plotted in the lower panels as the red curves.}
728: \label{fig:tod_noise}
729: \end{figure}
730:
731: \subsection{Detector response}
732: \label{sec:detector_response}
733: The \cite{irwin05} TES small-signal model mentioned above also
734: provides us with an estimate of the detector response (the conversion
735: from incident power to resultant current in the detectors). In this
736: model, the power-to-current responsivity, $s_I(\omega)$, is given by
737: \be
738: s_I(\omega) \propto \frac{1 - \tau_+/\tau_I}{1 + i\omega\tau_+}
739: \frac{1 - \tau_-/\tau_I}{1 + i\omega\tau_-},
740: \label{eqn:response}
741: \ee
742: where $\omega = 2 \pi f$ is the angular frequency and $\tau_+$ and
743: $\tau_-$ are the ``rise time'' and ``fall time'' (relaxation to steady
744: state) after a delta-function temperature impulse. Here, $\tau_I$ is
745: the current-biased thermal time constant. The impulse-response in
746: the time domain is
747: \be
748: s_I(t) \propto \frac{e^{-t/\tau_+} - e^{-t/\tau_-}}{\tau_+ - \tau_-}
749: \Theta(t) ,
750: \ee
751: where $\Theta(t)$ is the Heaviside step function.
752: Note that the constant of
753: proportionality in equation~(\ref{eqn:response}) (which is also
754: predicted by the model) is essentially the calibration (gain) of the
755: detectors. The simulated time-streams of
756: equation~(\ref{eqn:signal_timestream}) are converted to detector
757: time-streams through convolution with this response
758: function. Note that the photon noise and correlated atmospheric noise
759: are added to the time-stream before convolution (and so are also
760: convolved with the response function) while the detector noise is
761: added directly to the convolved time-streams. The \clover\, detectors
762: are designed to be extremely fast with time-constants $\tau_\pm < 1 \,
763: {\rm ms}$. For our simulations, we have used time-constants
764: predicted by the small-signal TES model for \clover\, instrument
765: parameters of $\tau_+ = 300 \, \mu{\rm s}$ and $\tau_- = 322 \, \mu
766: {\rm s}$.
767:
768: Note that for our chosen scan speed of $0.25^{\circ}$/s, the effect of
769: the response function on the signal component in the simulations is
770: small in the absence of fast modulation -- the \clover\,
771: detectors are so fast that signal attenuation
772: and phase differences introduced by convolution with the detector
773: response only become important at high frequencies, beyond the
774: frequency range of the sky signals. For our simulations without
775: fast modulation, sky signals from multipoles $\ell \sim 2000$ will
776: appear in the time-stream at $\sim$ 1~Hz where the amplitude of the
777: normalised response function is effectively unity and the associated
778: phase change is $-0.2^{\circ}$. For our simulations
779: including fast modulation, it becomes more important to correct the
780: time-stream data for the detector response. In this case, the polarized sky signals (from all
781: multipoles) appear within a narrow band centred on $12$~Hz in the
782: time-stream. Here, the amplitude of the response function is still very
783: close to unity but the phase change has grown to $-2.7^{\circ}$. We
784: include a deconvolution step in the analysis of all of our simulated
785: data to correct for this effect. Finally, we note that for the $\sim 10$
786: per cent errors in the time-constants which we have considered (see
787: Section~\ref{sec:systematics}), the resulting mis-estimation of the
788: polarization signal will again be small, even for the case of
789: fast modulation.
790:
791: In principle we should also include the effect of sample integration: each
792: discrete observation is an integral of a continuous signal over the sample period.
793: Of course, in the case where down-sampled data is simulated
794: the sample integration should include the effect of the sample
795: averaging. For integration over a down-sampled period $\Delta$, there is an additional
796: phase-preserving filtering by $\mathrm{sinc}(\omega \Delta/2)$ where
797: $\omega \equiv 2\pi f$. For our scan
798: parameters, the filter is negligible (i.e.\ unity) for unmodulated data
799: while for modulated data the filter can be approximated at the frequency
800: $4 \omega_\lambda$, where $\omega_\lambda$ is the angular rotation frequency
801: of the HWP. This acts like a small decrease in the polarization efficiency,
802: with the discretised signal at the detector being
803: %
804: \begin{eqnarray}
805: d_i &\approx & \frac{1}{2}\big\{T(\theta) +
806: \mathrm{sinc}(2\omega_\lambda \Delta) \nonumber \\
807: &&\mbox{} \times \left[Q(\theta)\cos(2\phi_i)
808: + U(\theta) \sin(2\phi_i)\right]\big\} .
809: \end{eqnarray}
810: %
811: For $\omega_\lambda = 2\pi \times 3\,\mathrm{Hz}$ and $\Delta =
812: 10\,\mathrm{ms}$, the effective polarization efficiency is $0.98$ and
813: this has the effect of raising the noise level in the polarization maps by
814: 2 per cent. We do not include this small effect in our simulations, but could
815: easily do so.
816:
817: \subsection{Systematic effects}
818: \label{sec:systematics}
819: In the analysis that follows, we will investigate the impact of
820: several systematic effects on the ability of a \clover-like
821: experiment to recover an input $B$-mode signal. For a reference, we
822: use a suite of simulations which contain no systematics. This ideal simulation
823: contains the input signal, photon noise, $1/f$ atmospheric noise
824: correlated across the array (but un-polarized) and TES
825: detector noise with no additional correlated $1/f$
826: component. Additionally, for our reference simulation all pointing
827: registers and detector polarization sensitivity angles use the nominal
828: values and the signal is convolved with the detector response function
829: using the nominal time constants.
830:
831: We then perform additional sets of simulations with the following
832: systematic effects included in isolation:
833: \begin{itemize}
834: \item $1/f$ detector noise. We have considered an additional
835: correlated $1/f$ component to the detector noise with $1/f$ knee
836: frequencies of $0.1$, $0.05$ and $0.01$~Hz.
837: \item Polarized atmosphere. In addition to the un-polarized atmosphere
838: present in the reference simulation, we consider a \emph{polarized}
839: $1/f$ component in the atmosphere. To simulate polarized atmosphere,
840: we proceed as described in Section~\ref{sec:noise_sims} but now
841: we add correlated $1/f$ atmospheric noise to the $Q$ and $U$ sky signal
842: time-streams such that equation~(\ref{eqn:signal_timestream})
843: becomes
844: \begin{eqnarray}
845: d_i &=& \frac{1}{2} \left[T + \left(Q + Q^{\rm atms}_i\right) \cos(2 \phi_i)
846: \nonumber \right. \\
847: && \mbox{} \phantom{xxxx} \left.
848: + \left(U + U^{\rm atms}_i\right) \sin(2 \phi_i)\right].
849: \end{eqnarray}
850: We take the $Q^{\rm atms}$ and $U^{\rm atms}$ atmospheric signals to
851: have the same power spectrum as the common-mode atmosphere
852: (equation~\ref{eqn:pk_atms}) but a factor ten smaller in magnitude.
853: \item Detector gain errors. We consider three types of gain errors:
854: (i) random errors in the gain that are constant in time, uncorrelated
855: between detectors and have a 1 per cent RMS; (ii) gain drifts
856: in each detector corresponding to a 1 per cent drift over the course of a
857: two-hour observation -- the start and end
858: gains for each detector are randomly distributed about the nominal
859: gain value with an RMS of 1 per cent; (iii) systematic A/B gain
860: mis-matches (1 per cent mis-match) between the two detectors within each
861: pixel. For this latter systematic, we have applied a constant 1 per
862: cent A/B mis-match to all pixels on the focal plane but the direction
863: of the mis-match (that is, whether the gain of A is greater or smaller
864: than B) is chosen randomly.
865: \item Mis-estimated polarization sensitivity directions. Random
866: errors uncorrelated between detectors (including those with the
867: same feedhorn) with RMS 0.5$^{\circ}$ and which are constant in time,
868: and a systematic mis-estimation of the
869: instrument polarization coordinate reference system by 0.5$^{\circ}$ are
870: considered.
871: \item Mis-estimated half-wave plate angles. For the case where we
872: consider an experiment which includes polarization modulation with a
873: half-wave plate (see Section \ref{sec:modulation}), we also
874: consider random errors (with RMS 0.5$^{\circ}$) in the recorded HWP angle
875: which we apply to each 100~Hz time-sample. In addition, we consider
876: a 0.5$^{\circ}$ systematic offset in the half-wave plate angle measurements.
877: \item Mis-estimated time-constants. The analysis that follows
878: includes a deconvolution step to undo the response function of the
879: detectors and return the deconvolved sky signal. In all cases, we
880: use the nominal time-constant values of $\tau_+ = 300 \, \mu{\rm s}$
881: and $\tau_- = 322 \, \mu {\rm s}$ to perform the deconvolution. To
882: simulate the effect of mis-estimated time-constants, we introduce
883: both a random scatter (with RMS = 10 per cent across detectors) and a
884: systematic offset ($\tau_\pm$ identically offset by 10 per cent for all
885: detectors) in the time-constants when creating the simulated data.
886: \item Pointing errors. We simulate the effects of both a random jitter and a
887: slow wander in the overall pointing of the telescope by introducing
888: a random scatter uncorrelated between time samples (with RMS 30 arcsec)
889: and an overall drift in the
890: pointing (1 arcmin drift from true pointing over the course of a
891: two-hour observation) when creating the simulated time-stream. Once
892: again, the simulated data is subsequently analysed assuming perfect
893: pointing registers.
894: \item Differential transmittance in the HWP. As a simple example of a
895: HWP-induced systematic, we have considered a differential
896: transmittance of the two linear polarizations by the
897: HWP. Preliminary measurements of the \clover\, HWPs suggest the level
898: of differential transmittance will be in the region of
899: $1\!\!\!\relbar\!\!\!2$ per cent. For
900: this work, we consider a 2 per cent differential transmittance in the HWP.
901: \end{itemize}
902:
903: The range of systematic effects we include is not exhaustive. In
904: particular, we ignore effects in the HWP, when present, except for a
905: mis-estimation of the rotation angle and a differential transmittance
906: of the two linear polarizations. In addition, we ignore all optical
907: effects. In practice, there are many possible HWP-related systematic
908: effects which we have not yet considered. In general, a
909: thorough analysis of HWP-related systematics require detailed physical
910: optics modelling which is beyond the scope of our current analysis. We
911: therefore leave a detailed investigation of HWP-related systematics to future
912: work and simply urge the reader to bear in mind that where our
913: analysis has included a HWP, we have, in most cases, assumed a perfect one. For other
914: optical effects, we note that \citet{odea07} have already investigated
915: some relevant effects using analytic and numerical techniques and we
916: are currently adapting their flat-sky numerical analysis to work with
917: the full-sky simulations described here. Our conclusions on the
918: ability of modulation to mitigate systematic effects associated with
919: imperfect optics, which will be based on detailed physical optics
920: simulations of the \clover\, beams (Johnson et al., in prep), will be
921: presented in a future paper. Note that there are important
922: instrument-specific issues to consider in such a study to do with
923: where the modulation is performed in the instrument. In \clover, the
924: modulation will be provided by a HWP between the horns and mirrors and
925: this may lead to a difference in the relative rotation of the field
926: directions on the sky and the beam shapes as the HWP rotates compared
927: to a set-up, as in \spider~\citep{crill08}, where the HWP is after
928: (thinking in emission) the beam-defining elements.
929:
930: For a \clover-like receiver, consisting of a HWP followed by a polarization
931: analyser (e.g.\ an orthomode transducer) and detectors, we include
932: essentially all
933: relevant systematic effects introduced \emph{by the receiver}.
934: To see this, note that
935: the most general Jones matrix describing propagation of the two linear
936: (i.e.\ $x$ and $y$) polarization states through the polarization analyser
937: is~\citep{odea07}
938: %
939: \be
940: \mathbfss{J} = \left(
941: \begin{array}{cc}
942: 1+g_1 & \epsilon_1 \\
943: -\epsilon_2 & (1+g_2)e^{i\alpha}
944: \end{array}
945: \right) ,
946: \ee
947: %
948: where $g_1$, $g_2$ and $\alpha$ are small real parameters and
949: $\epsilon_1$ and $\epsilon_2$ are small and complex-valued. The detector
950: outputs are proportional to the power in the
951: $x$ and $y$-components of the transmitted field (after convolution
952: with the detector response function).
953: To first-order in small
954: parameters, only $g_1$, $g_2$ and the real parts of $\epsilon_1$ and
955: $\epsilon_2$ enter the detected power. In this limit, the perturbed Jones
956: matrix is therefore equivalent in terms of the detected power to
957: %
958: \be
959: \mathbfss{J} \sim \left(
960: \begin{array}{cc}
961: 1+g_1 & 0 \\
962: 0 & 1+g_2
963: \end{array}
964: \right)
965: \left(
966: \begin{array}{cc}
967: \cos \alpha_1 & \sin \alpha_1 \\
968: -\sin\alpha_2 & \cos\alpha_2
969: \end{array}
970: \right)
971: \ee
972: %
973: where the small angles $\alpha_1 \approx \Re\epsilon_1$ and $\alpha_2 \approx
974: -\Re\epsilon_2$ denote the perturbations in the polarization-sensitivity
975: directions introduced above, and $g_1$ and $g_2$ are the gain errors.
976: Note that instrument polarization (i.e.\ leakage from $T$ to detected
977: $Q$ or $U$) is only generated in the receiver through mismatches in the gain at first
978: order, but also through $|\epsilon_1|^2$ and
979: $|\epsilon_2|^2$ in an exact calculation. The latter effect is not
980: present in the simplified description in terms of offsets in the
981: polarization-sensitivity directions. Note also that if we
982: difference the outputs of
983: the two detectors in the same pixel, in the presence of perturbations
984: $\alpha_1$ and $\alpha_2$ to the polarization sensitivity directions we
985: find
986: \begin{eqnarray}
987: d_1 - d_2 &=& \cos(\alpha_1-\alpha_2)\left(
988: Q\cos[2(\phi-\bar{\alpha})] \nonumber \right. \\
989: &&\mbox{} \phantom{xxxxxxxxxxxx} \left. + U \sin[ 2(\phi-\bar{\alpha})] \right) ,
990: \end{eqnarray}
991: where $\bar{\alpha} = (\alpha_1 + \alpha_2)/2$. This is equivalent
992: to a common rotation of the pair by $\bar{\alpha}$ and a decrease in the
993: polarization efficiency to $\cos(\alpha_1-\alpha_2)$.
994:
995: \subsection{Polarization modulation}
996: \label{sec:modulation}
997:
998: For our reference simulation, and for each of the systematic effects
999: listed above, we simulate the experiment using three different
1000: strategies for modulating the polarization signal. Firstly, we
1001: consider the case where no explicit modulation of the polarization
1002: signal is performed -- in this case, the only modulation achieved is
1003: via telescope scanning and the relatively small amount of sky rotation
1004: provided by the current \clover\, observing strategies. In addition,
1005: we also consider the addition of a half-wave plate, either continuously
1006: rotating or ``stepped'', placed in front of the focal plane. A
1007: half-wave plate modulates the polarization signal such that the
1008: output of a single detector (in the detector's local polarization
1009: coordinate frame) is
1010: \be d_i = \frac{1}{2}\left[ T + Q \cos ( 4\phi_i )
1011: + U \sin ( 4\phi_i ) \right],
1012: \label{eqn:pol_mod}
1013: \ee
1014: where, here, $\phi_i$ is the angle between the detector's local
1015: polarization frame and the principal axes of the wave plate.
1016:
1017: For a continuously rotating HWP, the polarized-sky signal is thus
1018: modulated at $4 f_\lambda$ where $f_\lambda$ is the rotation frequency
1019: of the HWP. As well as allowing all three Stokes parameters to be
1020: measured from a single detector, modulation with a continuously
1021: rotating HWP (which we term ``fast'' modulation in this paper) moves
1022: the polarization sky-signal to higher frequency and thus away from any
1023: low-frequency $1/f$ detector noise that may be
1024: present; see Fig.~\ref{fig:mod_frequency}.
1025: (Note that the temperature signal is not
1026: modulated and one needs to rely on telescope scanning and analysis
1027: techniques to mitigate $1/f$ noise in $T$.) This ability to mitigate
1028: the effect of $1/f$ noise on the polarization signal is the prime
1029: motivation for including a continuously rotating HWP in a CMB
1030: polarization experiment\footnote{%
1031: The ability to measure all three Stokes
1032: parameters from a single detector has also been suggested as
1033: motivation for including a modulation scheme in CMB polarization
1034: experiments. However, we will argue later in
1035: Section~\ref{sec:differencing} that, at least for ground-based
1036: experiments, an analysis based on extracting all three Stokes
1037: parameters from individual detectors in isolation using a real-space
1038: demodulation technique may be a poor choice.}
1039: Systematic effects that generate an apparent polarization signal that is
1040: not modulated at $4f_\lambda$ can also be mitigated almost
1041: completely with fast modulation.
1042: Most notably, instrument polarization generated in the receiver will
1043: not produce a spurious polarization signal in the recovered maps
1044: unless the gain and
1045: time-constant mismatches vary sufficiently rapidly ($\sim 4 f_\lambda$)
1046: to move the \emph{scan}-modulated temperature leakage up into the
1047: polarization signal band.
1048:
1049: \begin{figure}
1050: \centering
1051: \resizebox{0.47\textwidth}{!}{
1052: \rotatebox{-90}{\includegraphics{fig5.ps}}}
1053: \caption{Frequency-space representation of polarization modulation
1054: with a continuously rotating HWP. The plotted power spectrum is that
1055: for a single azimuth scan from one of our signal-only simulations
1056: with the HWP continuously rotating at $f_\lambda = 3$~Hz. The power
1057: in the frequency range $0.01 < k < 1$~Hz is the unmodulated
1058: temperature signal from sky multipoles in the range $20 < \ell <
1059: 2000$. The power spike at $4 f_\lambda = 12$~Hz is the modulated
1060: polarized signal. The dashed (dotted) line shows the expected
1061: temperature (polarization) signal band (with arbitrary
1062: normalisation) appropriate for the scan speed, modulation frequency
1063: and beam size we have used. The residual power between $1$ and
1064: $6$~Hz and on either side of the polarization power spike is due to
1065: pixelisation effects.}
1066: \label{fig:mod_frequency}
1067: \end{figure}
1068:
1069: As mentioned in the previous section, as an example of a HWP-induced
1070: systematic, we have considered the case of a 2 per cent differential
1071: transmittance by the HWP of the two incoming linear polarizations. We
1072: model this effect using a non-ideal Jones matrix for the HWP of the
1073: form,
1074: %
1075: \be
1076: \mathbfss{J} = \left(
1077: \begin{array}{cc}
1078: 1 & 0 \\
1079: 0 & -(1+\delta)
1080: \end{array}
1081: \right) ,
1082: \label{eqn:diff_trans_jones}
1083: \ee
1084: %
1085: where $\delta$ describes the level of differential
1086: transmittance. Propagating through to detected power, for the
1087: difference in output of the two detectors within a pixel, we find
1088: %
1089: \ba
1090: d_1 - d_2 \, &=& \left[1 + \delta + \frac{\delta^2}{4} \right]
1091: (Q\cos(4\phi) + U\sin(4\phi) ) \nonumber \\
1092: &-& \left[\delta + \frac{\delta^2}{2}\right] I\cos(2\phi)
1093: + \frac{\delta^2}{4} Q.
1094: \label{eqn:diff_trans_power}
1095: \ea
1096: Note that, in this expression, both the HWP angle, $\phi$, and the
1097: Stokes parameters are defined in the pixel basis. The first term in
1098: equation (\ref{eqn:diff_trans_power}) is the ideal
1099: detector-differenced signal but mis-calibrated by a factor, $\delta +
1100: \delta^2/4$. For reasonable values of $\delta$, this mis-calibration
1101: is small ($\lsim 2$ per cent) and, in any case, is easily dealt with during
1102: a likelihood analysis of the power spectra. The
1103: potential problem term is the middle term which contains the
1104: total intensity signal modulated at $2 f_\lambda$. Note that there
1105: will be contributions to this term from the CMB monopole, dipole and
1106: the atmosphere, which, for our simulations, we have taken to be 5 per cent
1107: emissive. Even for small values of $\delta$ therefore, these
1108: HWP-synchronous signals will completely dominate the raw detector data
1109: and need to be removed from the data prior to the map-making step.
1110:
1111: With a HWP operating in ``stepped'' mode where the angle of the HWP is
1112: changed at regular time intervals (e.g. at the end of each scan), the
1113: gains are less clear. The polarization sky signal is not shifted to
1114: higher frequencies so $1/f$ detector noise can only be dealt with by
1115: fast scanning. What stepping the waveplate can potentially do is to increase
1116: the range of polarization sensitivity directions with which a given pair of detectors
1117: samples any pixel on the sky. This has two important effects: (i)
1118: it reduces the correlations between the errors in the reconstructed $Q$ and $U$
1119: Stokes parameters in each sky pixel; and (ii) it can mitigate somewhat those
1120: systematic effects that do not transform like a true polarization under
1121: rotation of the waveplate. Of course, one of the strongest motivations
1122: for stepped, slow modulation is the avoidance of systematic effects
1123: associated with the continuous rotation of the HWP. If these effects are
1124: sufficiently well understood, then the resulting spurious signals can
1125: be rejected during analysis. However, if they are not well understood,
1126: a stepped HWP, while not as effective in mitigating systematics, may
1127: well be the preferred option.
1128:
1129: Note that for a perfect optical system, rotation of the waveplate is
1130: equivalent to rotation (by twice the angle) of the instrument.
1131: However, this need not hold with imperfect optics. For example,
1132: suppose the beam patterns for the two polarizations of a given
1133: feedhorn are purely co-polar (i.e.\ the polarization sensitivity
1134: directions are ``constant'' across the beams and orthogonal), but the
1135: beam shapes are orthogonal ellipses. This set-up generates instrument
1136: polarization with the result that a temperature distribution that is
1137: locally quadrupolar on the sky will generate spurious polarization
1138: that transforms like true sky polarization under rotation of the
1139: instrument~\citep{hu03,odea07}. However, for an optical arrangement
1140: like that in \spider, where the HWP is after (in emission) the
1141: beam-defining optics, as the HWP rotates the polarization directions
1142: rotate on the sky but the beam shapes remain fixed. The spurious
1143: polarization from the mis-match of beam shapes is then constant as the
1144: HWP rotates for any temperature distribution on the sky, and so the
1145: quadrupolar temperature leakage can be reduced.
1146:
1147: In our analysis, in addition to simulations without explicit
1148: modulation we have also simulated an experiment with a HWP continuously
1149: rotating at $3$~Hz (thus modulating the polarization signal at
1150: $12$~Hz) and an experiment where a HWP is stepped (by 20$^{\circ}$) at the end
1151: of each azimuth scan (for the scan strategy and scan speed we are
1152: using, this corresponds to stepping the HWP roughly every $\sim$ 90
1153: s).
1154:
1155: We end this section with a comment on the ability of a continuously
1156: rotating HWP to mitigate un-polarized $1/f$ atmospheric noise. The polarization
1157: signal band is still, of course, moved to higher frequency and thus
1158: away from the $1/f$ noise but the atmospheric $1/f$ noise is so strong
1159: that extremely rapid HWP rotation would be required to move the
1160: polarization band far enough into the tail of the $1/f$ spectrum.
1161: Such rapid rotation is not an option in practice as it would introduce its
1162: own systematic effects (e.g. excessive heat generation). This is the basis of our argument
1163: mentioned above that extracting all three Stokes parameters from a
1164: single detector may be a poor choice of analysis technique. However,
1165: since the $1/f$ atmospheric noise is un-polarized, it can be removed
1166: \emph{completely} by combining data from multiple detectors. We
1167: revisit this issue again with simulations in
1168: Section~\ref{sec:differencing}. Finally, we note that if the atmosphere
1169: does contain a polarized $1/f$ component, then we expect that this
1170: will not be mitigated by modulation -- the polarized atmosphere would
1171: be modulated in the same way as the sky signal and would
1172: shift up in frequency accordingly.
1173:
1174: \section{Analysis of simulated data}
1175: \label{sec:analysis}
1176: For our reference simulation, and for each of the systematic effects
1177: and modulation strategies described in the previous section, we have
1178: created a suite of 50 signal-only, noise-only and signal-plus-noise
1179: simulated datasets. Our analysis of the signal-only data will be used
1180: to investigate potential biases caused by the systematics while our
1181: signal-plus-noise realisations are used together with the noise-only
1182: simulations to investigate any degradation of the sensitivity of the
1183: experiment due to the presence of the systematic effects.
1184:
1185: Our analysis of each dataset consists of processing the data through
1186: the stages of deconvolution for the bolometer response function,
1187: polarization demodulation and map-making, and finally estimation of the
1188: $E$- and $B$-mode power spectra. For any given single realisation these
1189: processes are performed separately for each of the four observed
1190: \clover\, fields -- that is, we make maps and measure power spectra for
1191: each field separately. Since our fields are widely separated on the
1192: sky, we can treat them as independent and combine the power spectra
1193: measured from each using a simple weighted average to produce a single
1194: set of $E$- and $B$-mode power spectra for each realisation of the
1195: experiment. Note that even if our fields were not widely separated,
1196: our procedure would still be unbiased (but sub-optimal) and
1197: correlations between the fields would be automatically taken
1198: into account in our error analysis since, for any given realisation,
1199: the input signal for all four fields is taken from the same simulated
1200: CMB sky.
1201:
1202: \subsection{Time-stream processing and map-making}
1203: \label{sec:map-making}
1204: We first deconvolve the time-stream data for the detector response in
1205: frequency space using the response function of
1206: equation~(\ref{eqn:response}) and using the nominal time-constants in
1207: all cases. Once this is done, the data from detectors within each
1208: pixel are differenced in order to remove both the CMB temperature signal and
1209: the correlated $1/f$ component of the atmospheric noise. For the case
1210: where the $1/f$ atmosphere is completely correlated between the two
1211: detectors and in the absence of instrumental polarization and/or
1212: calibration systematics, this process will remove the CMB $T$ signal and the
1213: $1/f$ atmosphere completely.
1214:
1215: For the case where we have simulated the effect of a differential
1216: transmittance in the HWP, a further time-stream processing step is
1217: required at this point to fit for and remove the HWP-synchronous
1218: signals from the time-stream. To do this, we have implemented a simple
1219: iterative least-squares estimator to fit, in turn, for the amplitudes
1220: of both a cosine and sine term at the second harmonic of the
1221: HWP-rotation frequency, $f_\lambda$. For our simulations containing
1222: both signal and noise, the accuracy with which we are able to remove
1223: the HWP-synchronous signals is determined by the noise level in the
1224: data. A demonstration of the performance of this procedure is given
1225: in Fig.~\ref{fig:hwp_sys_removal} where we plot the power spectra of
1226: six minutes of simulated time-stream data (containing both signal
1227: and noise) before and after the removal of the HWP-synchronous signal.
1228:
1229: \begin{figure}
1230: \centering
1231: \resizebox{0.455\textwidth}{!}{
1232: \rotatebox{-90}{\includegraphics{fig6.ps}}}
1233: \caption{Power spectra of a six minute segment of time-stream
1234: data before (upper panel) and after (lower panel) fitting for and
1235: removing the HWP-synchronous signal. A 2 per cent differential
1236: transmittance in the HWP was used to create the simulated data. The
1237: resulting spurious signal appears at $2 f_\lambda = 6$~Hz in the upper
1238: panel and has a peak power of $\sim 10^{10}$ in the units
1239: plotted. No obvious residuals are apparent in the lower panel after
1240: applying our procedure for removing the contamination.}
1241: \label{fig:hwp_sys_removal}
1242: \end{figure}
1243:
1244: After detector differencing, and the removal of the HWP-synchronous
1245: signals if present, the resulting differenced time stream is then a
1246: pure polarization signal:
1247: \be d_i = Q \cos(2\phi_i) + U \sin(2\phi_i),
1248: \label{eqn:diff_timestream}
1249: \ee
1250: where again, the angle $\phi_i$ is, in the most general case, a
1251: combination of detector orientation, sky crossing angle, telescope
1252: boresight rotation and the orientation of the HWP if present. In
1253: order to construct maps of the $Q$ and $U$ Stokes parameters, these
1254: quantities need to be decorrelated from the differenced time-stream
1255: using multiple observations of the same region of sky taken with
1256: different values of $\phi_i$. For an experiment which does not
1257: continuously modulate the polarization signal, the $Q$ and $U$ signals
1258: have to be demodulated as part of the map-making step. Note however
1259: that for an experiment where the polarization signal is continuously
1260: modulated, there are a number of alternative techniques to demodulate
1261: $Q$ and $U$ at the time-stream level. In separate work, one of us has
1262: compared the performance of a number of such demodulation schemes and
1263: our results will be presented in a forthcoming paper (Brown, in prep.).
1264: For the purposes of our current analysis
1265: however, we have applied the same map-based demodulation scheme to all
1266: three experiments which we have simulated. In this scheme, once the
1267: time-streams from each detector pair have been differenced, $Q$ and
1268: $U$ maps are constructed as
1269: \ba
1270: \left( \begin{array}{c} Q \\ U \end{array} \right) &=& \left(
1271: \begin{array}{cc}
1272: \lgl\cos^2(2\phi_i)\rgl & \lgl\cos(2\phi_i)\sin(2\phi_i)\rgl \\
1273: \lgl\cos(2\phi_i)\sin(2\phi_i)\rgl & \lgl\sin^2(2\phi_i)\rgl \\
1274: \end{array} \right)^{-1} \nn \\
1275: &&\mbox{} \times \left( \begin{array}{c}
1276: \lgl\cos(2\phi_i)d_i\rgl \\
1277: \lgl\sin(2\phi_i)d_i\rgl \\ \end{array} \right),
1278: \label{eqn:qu_mapmaking}
1279: \ea
1280: where the angled brackets denote an average over all data falling within
1281: each map pixel. For the work presented here, this averaging is
1282: performed using the data from all detector pairs in one operation.
1283: One could alternatively make maps per detector pair which could then
1284: be co-added later. For the case where the noise properties of each
1285: detector pair are similar, the two approaches should be
1286: equivalent. Note that the map-maker we use for all of our analyses is
1287: a na\"{\i}ve one -- that is, we use simple binning to implement
1288: equation~(\ref{eqn:qu_mapmaking}). There are, of course, more optimal
1289: techniques available (e.g. \citealt{sutton09} and references therein)
1290: which would out-perform a
1291: na\"{\i}ve map-maker in the presence of non-white noise. However, for our
1292: purposes, where we wish to investigate the impact of modulation in
1293: isolation, it is more appropriate to apply the same na\"{\i}ve map-making
1294: technique to all of our simulated data. We can then be sure that any
1295: improvement we see in results from our simulations including
1296: modulation are solely due to the modulation scheme employed.
1297:
1298: \subsection{Power spectrum estimation}
1299: \label{subsec:power}
1300: We measure $E$- and $B$-mode power spectra from each of our
1301: reconstructed maps using the ``pure'' pseudo-$C_\ell$ method of
1302: \cite{smith06}. We will not describe the method in detail here and refer the
1303: interested reader to \cite{smith06} and \cite{smith07} for
1304: further details. Here, we simply note that the ``pure''
1305: pseudo-$C_\ell$ framework satisfies most of the requirements of a power
1306: spectrum estimator for a mega-pixel CMB polarization experiment with
1307: complicated noise properties targeted at constraining $B$-modes: it
1308: is, just like normal pseudo-$C_\ell$, a fast estimator scaling as
1309: $N_{\rm pix}^{3/2}$ where $N_{\rm pix}$ is the number of map pixels
1310: (as opposed to a maximum likelihood estimator which scales as $N_{\rm
1311: pix}^3$); it is a Monte-Carlo based estimator relying on simulations
1312: of the noise properties of the experiment to remove the noise bias and
1313: estimate band-power errors and covariances -- it is thus naturally
1314: suited to experiments with complicated noise properties for which
1315: approximations to the noise cannot be made; and it is near-optimal in the
1316: sense that it eliminates excess sample variance from $E \rightarrow B$ mixing
1317: due to ambiguous modes which result from incomplete sky observations
1318: \citep{lewis02, bunn02}, and which renders simple pseudo-$C_\ell$ techniques
1319: unsuitable for small survey areas~\citep{challinor05}.
1320:
1321: \subsubsection{Power spectrum weight functions}
1322:
1323: With normal pseudo-$C_\ell$ estimators, one multiplies the data with a function
1324: $W(\vnhat)$ that is chosen heuristically and apodizes the edge of the survey
1325: (to reduce mode-coupling effects). For example, if one
1326: is signal dominated, uniform weighting (plus apodization) is a reasonable
1327: choice, whereas an inverse-variance weight is a good choice in the
1328: noise-dominated regime. Similar reasoning applies for the
1329: pure pseudo-$C_\ell$ technique, but here one weights the spherical
1330: harmonic functions rather than the data themselves.
1331: To see this, compare the
1332: definition of the ordinary and pure pseudo harmonic $B$-modes:
1333: \ba
1334: \widetilde a_{\ell m}^B = &-&\frac{i}{2}
1335: \sqrt{\frac{(l-2)!}{(l+2)!}} \int d^2\vnhat \bigg[ \Pi_+(\vnhat)
1336: W(\vnhat) \bar{\eth}\bar{\eth} Y_{\ell m}^*(\vnhat) \nonumber \\
1337: &&\mbox{} - \Pi_-(\vnhat) W(\vnhat) \eth\eth Y_{\ell m}^*(\vnhat) \bigg]
1338: \label{eq:almbdef} \\
1339: \widetilde a_{\ell m}^{B\,{\rm pure}} = &-&\frac{i}{2}
1340: \sqrt{\frac{(l-2)!}{(l+2)!}} \int d^2\vnhat \bigg[ \Pi_+(\vnhat)
1341: \bar{\eth}\bar{\eth} \big( W(\vnhat) Y_{\ell m}^*(\vnhat) \big)
1342: \nonumber \\
1343: &&\mbox{} - \Pi_-(\vnhat) \eth\eth \big(
1344: W(\vnhat) Y_{\ell m}^*(\vnhat) \big) \bigg].
1345: \label{eq:almbpuredef}
1346: \ea
1347: Here, $\Pi_\pm(\hat{n}) = (Q\pm
1348: iU)(\vnhat)$ is the complex polarization and $\eth, \bar{\eth}$ are
1349: the spin raising and lowering operators defined
1350: in~\citet{zaldarriaga97}. If $W(\vnhat)$ is chosen so that it vanishes
1351: along with its first derivative on the survey boundary, then the
1352: $\widetilde a_{\ell m}^{B\,{\rm pure}}$ couple only to $B$-modes and
1353: the excess sample variance due to $E$-$B$ mixing is eliminated.
1354: The action of $\eth, \bar{\eth}$ on the spin spherical
1355: harmonics is simply to convert between different spin-harmonics but
1356: their action on a general weight function is non-trivial for
1357: $W(\vnhat)$ defined on an irregular pixelisation such as \healpix\footnote{%
1358: One possibility that we have yet to explore is performing the derivatives
1359: directly in spherical-harmonic space. Since $W(\vnhat)$ is typically
1360: smooth, its spherical transform will be band-limited and straightforward
1361: to handle.}. To
1362: get around this problem, we choose to calculate the derivatives of
1363: $W(\vnhat)$ in the flat-sky approximation where the differential
1364: operators reduce to
1365: \ba
1366: \eth \approx -(\partial_x + i \partial_y), \\
1367: \bar{\eth} \approx -(\partial_x - i \partial_y).
1368: \ea
1369: The derivatives are then trivially calculated on a regular Cartesian
1370: grid using finite differencing~\citep{smith07}.
1371:
1372: The most optimal weighting scheme for a pseudo-$C_\ell$ analysis
1373: involves different weight functions for each $C_\ell$ band-power
1374: according to the signal-to-noise level expected in that band.
1375: However, this is a costly solution (requiring $3 N_{\rm band}$
1376: spherical harmonic transforms) and the indications are, from some
1377: restricted tests that we have carried out, that the improvement in
1378: resulting error-bars is small, at least for the specific noise
1379: properties of our simulations. For the analysis presented here, we
1380: have therefore chosen a simpler scheme whereby we have used a uniform
1381: weight, appropriately apodized at the boundaries for the entire $\ell$
1382: range for $E$-modes and for $\ell \le 200$ for $B$-modes. For $\ell >
1383: 200$ our simulated experiment is completely noise dominated for a
1384: measurement of $B$-modes and so here we use an inverse-variance
1385: weighting, again, appropriately apodized at the boundaries. For
1386: simplicity, we have approximated the boundary of the map as a circle
1387: of radius $11^{\circ}$. Note that restricting our power spectrum analysis
1388: to this central region of our maps means we are effectively using only
1389: $\sim 70$ per cent of the available data. To calculate the derivatives of the
1390: weight functions, we project our weight maps (defined in \healpix)
1391: onto a Cartesian grid using a gnomonic projection. Once the
1392: derivatives of the weight maps have been constructed on the grid using
1393: finite differencing, they are transformed back to the original
1394: \healpix\, grid. An example of the inverse variance weight maps we
1395: have used and the resulting spin-1 and spin-2 weight functions, $\eth
1396: W$ and $\eth\eth W$, for one of our fields are shown in
1397: Fig.~\ref{fig:ppcl_weights}.
1398: \begin{figure*}
1399: \centering
1400: \resizebox{0.55\textwidth}{!}{
1401: \rotatebox{0}{\includegraphics{fig7.ps}}}
1402: \caption{Inverse-variance weight functions used for power spectrum
1403: estimation for one of the southern fields. For the noise properties
1404: of our simulated data, the hit-map shown in the top left panel
1405: closely approximates the inverse-variance map. This map is heavily
1406: smoothed and apodized at the boundary of the map to produce the
1407: spin-0 weight function shown in the top right panel.
1408: The spin-1 and spin-2 weight functions, $\eth W$ and $\eth\eth W$,
1409: are shown (as vector fields) in the bottom left and right-hand
1410: panels respectively.}
1411: \label{fig:ppcl_weights}
1412: \end{figure*}
1413:
1414: \section{Results from simulations}
1415: \label{sec:results}
1416: The map-making and power spectrum estimation procedures described
1417: above have been applied to each of our simulated datasets treating
1418: each of our four observing fields independently. For any given
1419: simulation set, we have 50 Monte-Carlo simulations so we can estimate the
1420: uncertainties on the band-powers measured from each field. For each
1421: realisation, we can then combine our measurements from the four fields
1422: using inverse-variance weights to produce a final single estimate of
1423: the power spectrum for each realisation. When presenting our results
1424: below, in all cases, we plot the mean of these final estimates. For
1425: our reference simulation, and for the $1/f$ noise systematics, the
1426: error-bars plotted are calculated from the scatter among the
1427: realisations and are those appropriate for a single realisation. When
1428: investigating the $B$-mode bias from systematics, we plot the results
1429: from signal-only simulations and the error-bars plotted are the
1430: standard error on the mean. For some of the noise-related systematics,
1431: we will also examine the impact of modulation in the map domain where
1432: the effects are already clear.
1433: \subsection{Reference simulation}
1434: \begin{figure*}
1435: \centering
1436: \resizebox{0.55\textwidth}{!}{
1437: \includegraphics{fig8.ps}}
1438: \caption{Sample maps constructed from simulated time-stream
1439: containing noise only (left panels) and both signal and noise
1440: (right panels). Temperature maps are shown in the top panel and
1441: $U$-polarization maps are shown in the bottom panels. These maps
1442: are for one of our reference simulations with no explicit
1443: modulation scheme and no systematics included. Note the striping
1444: in the noise-only $T$ map which is completely absent from the $U$
1445: maps due to differencing of detector pairs before map-making.}
1446: \label{fig:maps_reference}
1447: \end{figure*}
1448: \begin{figure*}
1449: \centering
1450: \resizebox{0.70\textwidth}{!}{
1451: \rotatebox{-90}{\includegraphics{fig9.ps}}}
1452: \caption{Mean recovered $E$-mode (top) and $B$-mode (bottom) power
1453: spectra for the reference simulations without explicit
1454: modulation. The errors plotted are those appropriate for a single
1455: realisation. The input CMB power spectra used to create the signal
1456: component of the simulations are shown as the red curves. In the
1457: bottom panel the total $B$-mode input signal (including lensing) for
1458: a tensor-to-scalar ratio of $r=0.026$ is shown as the red curve and
1459: the $B$-mode signal due to lensing alone is shown as the dashed
1460: curve. The $\ell < 200$ multipole range is shown in detail in the
1461: inset plots.}
1462: \label{fig:cls_reference}
1463: \end{figure*}
1464: To provide a reference for the results which follow, in
1465: Figs.~\ref{fig:maps_reference} and \ref{fig:cls_reference} we show the results from our
1466: suite of simulations with no explicit modulation and with no
1467: systematics included. In Fig.~\ref{fig:maps_reference} we plot examples of the
1468: reconstructed noise-only and signal-plus-noise $T$ and $U$ Stokes
1469: parameter maps (the reconstructed $Q$ maps -- not shown -- are
1470: qualitatively similar to $U$). The raw collecting power of an
1471: experiment like \clover\, is apparent from the top two panels in this
1472: figure. Although the noise $T$ map shown in the top-left panel is
1473: clearly dominated by striping due to the correlated noise from the
1474: atmosphere, only signal is apparent in the signal-plus-noise map shown in
1475: the top-right panel. (In fact, the noise contribution to the $T$
1476: signal-plus-noise map is significant, particularly on large scales, and
1477: so would need to be accounted for when measuring the temperature power
1478: spectrum.) Conversely, the $U$ noise map is dominated by
1479: white noise; the correlated component of the atmosphere has been
1480: removed completely from the polarization time-streams (as has the $T$
1481: sky signal) by differencing detector pairs before map-making.
1482:
1483: Figure~\ref{fig:cls_reference} shows the mean recovered $E$- and $B$-mode power
1484: spectra from our reference simulations for the case of no explicit
1485: modulation. Here, we see that our analysis is unbiased and recovers the input
1486: polarization power spectra correctly. For an input tensor-to-scalar
1487: ratio of $r=0.026$, we recover a detection of $B$-modes \emph{in
1488: excess} of the lensing signal of $1.54\sigma$.
1489: (We argue in Section~\ref{sec:fisher} that this is an under-estimate of
1490: the detection significance by around 10 per cent due to our ignoring small
1491: anti-correlations between the errors in adjacent band-powers.)
1492:
1493: The corresponding plots for the stepped and continuously rotating HWP
1494: are very similar apart from the reconstructed polarization maps at the
1495: very edges of the fields where a modulation scheme increases the
1496: ability to decompose into the $Q$ and $U$ Stokes parameters. Since the
1497: edges of the field are excluded in our power spectrum analysis in any
1498: case (see Fig.~\ref{fig:ppcl_weights}), we find that the performance
1499: (in terms of $C_\ell$ errors) of all three types of experiments which
1500: we have considered is qualitatively the same in the absence of
1501: systematic effects. Note that for all the systematics we have
1502: considered, the effects on the recovery of the $E$-mode spectrum is
1503: negligible and so, in the following sections, we plot only the
1504: recovered $B$-mode power spectra which are the main focus of this
1505: paper.
1506:
1507: \subsection{$1/f$ detector noise}
1508:
1509: Figure~\ref{fig:maps_det_noise} shows the recovered noise-only maps from
1510: a simulation containing a correlated $1/f$ detector noise
1511: component with a knee frequency of $f_{\rm knee} = 0.1$~Hz.
1512: In this figure, we have plotted the noise-only maps from
1513: the three types of experiment we have considered: no modulation; a
1514: HWP which is stepped by 20$^{\circ}$ at the end of each azimuth scan; and
1515: a HWP continuously rotating at $3$~Hz. The impact of modulation on $1/f$
1516: detector noise is clear from this plot -- as described in
1517: Section~\ref{sec:modulation}, the continuously rotating HWP shifts
1518: the polarization band in frequency away from the $1/f$ detector noise
1519: leaving only white noise in the resulting map. A stepped HWP, on the
1520: other hand, does not mitigate $1/f$ detector noise in this way
1521: and so noise striping is apparent in the middle panel of
1522: Fig.~\ref{fig:maps_det_noise}.
1523: \begin{figure*}
1524: \centering
1525: \resizebox{0.9\textwidth}{!}{
1526: \rotatebox{-90}{\includegraphics{fig10.ps}}}
1527: \caption{Sample noise-only $U$-maps from simulations containing a
1528: $1/f$ component correlated to the detector noise}. For display
1529: purposes only, the maps have been smoothed with a Gaussian with a FWHM
1530: of 7 arcmin. In the case where explicit modulation is either absent (left
1531: panel) or slow (stepped HWP; middle panel), the $1/f$ noise
1532: results in faint residual stripes in the polarization maps. In the case of
1533: a continuously rotating HWP, the polarization signal is modulated
1534: away from the low frequency $1/f$ resulting in white-noise behaviour
1535: in the polarization map (right panel).
1536: \label{fig:maps_det_noise}
1537: \end{figure*}
1538:
1539: The $B$-mode power spectra measured from our signal-plus-noise simulations
1540: including $1/f$ detector noise are shown in
1541: Fig.~\ref{fig:cls_det_noise} (again for $f_{\rm knee} = 0.1$~Hz) where
1542: we show the results from all three types of experiment.
1543: \begin{figure*}
1544: \centering
1545: \resizebox{0.9\textwidth}{!}{
1546: \rotatebox{-90}{\includegraphics{fig11.ps}}}
1547: \caption{Recovered $B$-mode power spectra for the simulations
1548: including a correlated $1/f$ component to the detector noise with
1549: $f_{\rm knee} = 0.1$~Hz for no modulation (top), a stepped HWP
1550: (middle) and a HWP continuously rotating at 3~Hz (bottom). See Table
1551: \ref{tab:simsummary} for the significances with which each
1552: experiment detects the input signals.}
1553: \label{fig:cls_det_noise}
1554: \end{figure*}
1555: Examination of the figure suggests that the presence of a $1/f$
1556: component in the detector noise leads to a significant degradation in
1557: the ability of the unmodulated and stepped-HWP experiments to recover
1558: the input $B$-mode signal. This degradation happens at all multipoles
1559: but is particularly acute on the largest scales ($\ell < 200$) where
1560: the primordial $B$-mode signal resides. The marginal detection of
1561: primordial $B$-modes (for $r = 0.026$) which we saw in our reference
1562: simulation (Fig.~\ref{fig:cls_reference}) is now completely destroyed
1563: by the presence of the $1/f$ correlated detector noise. Furthermore,
1564: note that our analysis of the simulated data sets is optimistic in the
1565: sense that we have assumed that any correlated noise can be modelled
1566: accurately -- that is, our noise-only simulations, which we use to
1567: measure the noise bias, are generated from the same model noise power
1568: spectrum used to generate the noise component in our signal-plus-noise
1569: simulations. (This is the reason that our recovered spectra are
1570: unbiased.) However, for the analysis of a real experiment, the noise
1571: properties need to be measured from the real data and there are
1572: uncertainties and approximations inherent in this process. Any
1573: correlated detector noise encountered in a real experiment is unlikely
1574: to be understood to the level which we have assumed in our analysis
1575: and so will likely result not only in the increased uncertainties we
1576: have demonstrated here but also in a biased result at low
1577: multipoles. Estimating cross-spectra between maps made from subsets of
1578: detectors for which the $1/f$ detector noise is measured to be
1579: uncorrelated is a simple way to avoid this noise bias issue, at the
1580: expense of a small increase in the error-bars~\citep{hinshaw03}.
1581:
1582: The results from the simulations where we have continuously modulated
1583: the polarization signal recover the input $B$-mode signal to the same
1584: precision that we saw with our reference simulation -- our marginal
1585: detection of $r=0.026$ is retained even in the presence of the
1586: correlated detector noise. In Section \ref{sec:significances} and
1587: Table \ref{tab:simsummary} we show quantitatively that, for a detector
1588: knee frequency of $0.1$~Hz, the significance with which the
1589: continuously modulated experiment detects the primordial $B$-mode
1590: signal is roughly twice that found for the un-modulated and
1591: stepped-HWP experiments. Also detailed in Table \ref{tab:simsummary}
1592: are the results from our $1/f$ noise simulations with knee frequencies
1593: of $0.05$ and $0.01$~Hz. We see, as expected, that the impact of fast
1594: modulation is less for a lower knee frequency --- for $f_{\rm knee} =
1595: 0.05$, rapid modulation still significantly out-performs the
1596: un-modulated and stepped-HWP experiments while for $f_{\rm knee} =
1597: 0.01$~Hz, there is essentially no difference between the performance
1598: of the three types of experiment.
1599:
1600: Note that, in the case of rapid modulation, because the polarization
1601: signal is moved completely away from the $1/f$ frequency regime, the
1602: recovered spectra should be immune to the issues of mis-estimation or
1603: poor knowledge of the noise power spectrum at low frequencies
1604: mentioned above. Although detector $1/f$ noise can be mitigated by
1605: other methods (e.g. using a more sophisticated map-maker;
1606: \citealt{sutton09}), these usually require accurate knowledge of the
1607: low-frequency noise spectrum unlike the hardware approach of fast
1608: modulation.
1609:
1610: \subsection{Polarized atmospheric $1/f$}
1611: In contrast to the addition of $1/f$ detector noise, which can be
1612: successfully dealt with by rapid modulation, all three types of
1613: experiment are degraded similarly by polarized low-frequency noise in
1614: the atmosphere. In particular, the errors at low multipoles are
1615: inflated by a large factor since the large amount of polarized
1616: atmosphere which we have input to the simulations swamps the input
1617: $B$-mode signal for $r=0.026$. We stress that the levels of polarized
1618: atmosphere we have used in these simulations were deliberately chosen to
1619: demonstrate the point that modulation does not help and the levels are
1620: certainly pessimistic.
1621:
1622: \subsection{Calibration errors}
1623: The power spectra recovered from signal-only simulations where we
1624: introduced random gain errors (constant in time) across the focal plane,
1625: or 1 per cent systematic A/B gain mis-matches between the two detectors
1626: within each pixel are shown in Fig.~\ref{fig:cls_gain_errors}. In both
1627: cases, we see a clear bias in the recovered $B$-mode signal in the
1628: absence of fast modulation, but the bias is mitigated entirely by the
1629: presence of a HWP rotating at 3~Hz. The bias is also mitigated to some
1630: degree by the stepped HWP but not completely. In our simulations, the bias is
1631: generally larger for the case of random gain errors since the
1632: variance (across the focal plane) of the gain mismatches is twice as large
1633: in the former case. For our simulations where we allowed detector gains to drift
1634: over the course of a two-hour observation, we found a similarly
1635: behaved $B$-mode bias to those shown in Fig.~\ref{fig:cls_gain_errors}
1636: but with a smaller magnitude (due to the two-hour drifts averaging down
1637: over the eight-hour observation).
1638:
1639: \begin{figure*}
1640: \centering
1641: \resizebox{0.90\textwidth}{!}{
1642: \rotatebox{-90}{\includegraphics{fig12.ps}}}
1643: \caption{Mean recovered $B$-mode power spectra for the simulations
1644: including random gain errors across the focal plane (black points) or
1645: systematic 1 per cent A/B gain mis-matches between detector pairs (blue points)
1646: for no modulation (top), a stepped
1647: HWP (middle) and a continuously rotating HWP (bottom). These spectra
1648: are measured from our suite of signal-only simulations. Our
1649: simulations containing both signal and noise exhibit the same biased
1650: recovery for the no-modulation and stepped-HWP cases. The presence
1651: of a fast modulation scheme (bottom panel) mitigates entirely the
1652: bias caused by these gain mis-matches. The standard errors in these
1653: mean recovered spectra are smaller than the plotted symbols.}
1654: \label{fig:cls_gain_errors}
1655: \end{figure*}
1656:
1657: A mis-match between the gains of the two detectors within a pixel
1658: corresponds to a $T \rightarrow Q$ leakage in the detector
1659: basis. The projection of this instrumental polarization onto the sky
1660: will therefore be suppressed if a wide range of sensitivity directions
1661: $\phi_i$ contribute to each sky pixel, as is the case for fast
1662: modulation. Note that in the case of a stepped HWP, one should be
1663: careful to design the stepping strategy in such a way that it does not
1664: undo some of the effect of sky rotation. During our analysis, we have
1665: found that the performance of a stepped HWP in mitigating systematic
1666: effects can depend critically on the direction, magnitude and
1667: frequency of the HWP step applied. In fact, for some set-ups we have
1668: investigated, a stepped HWP actually worsened the performance in
1669: comparison to the no modulation case due to interactions between the
1670: scan strategy and HWP stepping strategy. However, the results plotted in
1671: Fig.~\ref{fig:cls_gain_errors} for the stepped HWP case are for a
1672: HWP step of $20^{\circ}$ between each azimuth scan which is large and
1673: frequent enough to ensure that such interactions between the stepping strategy
1674: and the scan strategy are sub-dominant.
1675:
1676: \subsection{Mis-estimated polarization angles}
1677: The next set of systematics we have considered concern a
1678: mis-estimation of both the detector orientation angles (i.e.\ the
1679: direction of linear polarization to which each detector is sensitive
1680: to) and, for the case where a HWP is employed, a mis-estimation of the
1681: HWP orientation. We have performed simulations including both a random
1682: scatter (with an RMS of $0.5^{\circ}$) and a systematic offset of
1683: $0.5^{\circ}$ in the simulated detector and HWP angles. Note that for
1684: the systematic offset in the detector angles, the same offset is
1685: applied to all detectors. For both the detector angles and the HWP,
1686: the offset introduced corresponds to a systematic error in the
1687: estimation of the global polarization coordinate frame of the
1688: experiment and the effects are therefore degenerate.
1689:
1690: For the simulations which included a random scatter in the angles
1691: (both detector angles and HWP orientation), we found neither a bias in
1692: the recovered $B$-mode power spectra, nor a degradation in the
1693: error-bars from the simulations containing both signal and
1694: noise. Following the discussion in Section~\ref{sec:systematics},
1695: common errors in the detector angles for the pair of detectors in a single
1696: focal-plane pixel give rise to a rotation of the polarization sensitivity
1697: direction of the pixel, while differential errors reduce the polarization
1698: efficiency.
1699: For a typical differential scatter of $\sqrt{2}\times 0.5^{\circ}$,
1700: the reduction in the polarization efficiency ($\sim 10^{-4}$) is
1701: negligible. For a given pixel on the sky, the impact of the polarization
1702: rotation is suppressed by $\sqrt{N_\mathrm{sample}}$, where
1703: $N_\mathrm{sample}$ is the total number of samples contributing to that
1704: pixel with independent errors in the angles. The combination of a
1705: large number of detectors and, in the case of random HWP angle errors,
1706: their assumed short correlation time in our simulations renders the
1707: effect of small and random scatter in the angles negligible.
1708:
1709: The results from simulations which included a systematic error in the
1710: angles are shown in Fig.~\ref{fig:cls_pol_angles}, where we plot the recovered
1711: $B$-mode power spectrum from our signal-only simulations. In contrast
1712: to the simulations with random scatter, there is a clear mixing between $E$
1713: and $B$ due to the systematic mis-calibration of the polarization
1714: coordinate reference system of the instrument. A global mis-estimation
1715: of the polarization direction by an angle $\psi$ in the reconstructed maps
1716: leads to spurious $B$-modes with
1717: \be
1718: C_\ell^B = \sin^2(2\psi) C_\ell^E \approx 4 \psi^2 C_\ell^E .
1719: \ee
1720: Note that in
1721: Fig.~\ref{fig:cls_pol_angles}, we show the results from the detector angle
1722: systematic only for the case of the non-modulated experiment but the
1723: plot is identical for both the stepped and continuously rotating HWP --
1724: polarization modulation cannot mitigate a mis-calibration of detector
1725: angles. The fact that the mixing apparent in Fig.~\ref{fig:cls_pol_angles} is
1726: greater for the HWP mis-calibration is simply because rotating the
1727: waveplate by $\psi$ rotates the polarization direction by $2\psi$.
1728: Although the spurious $B$-mode power is most
1729: noticeable at high multipoles, where $\ell(\ell+1)C_\ell^E$ is largest,
1730: it is also present on large scales and,
1731: as is clear from the plot, would bias a measurement of the $B$-mode
1732: spectrum at all multipoles.
1733: \begin{figure}
1734: \centering
1735: \resizebox{0.48\textwidth}{!}{
1736: \rotatebox{-90}{\includegraphics{fig13.ps}}}
1737: \caption{Mean recovered $B$-mode power spectra for the signal-only simulations
1738: including mis-estimated detector polarization sensitivity angles (black
1739: points) and mis-estimated HWP angles (blue points) where the angles have been
1740: systematically offset by 0.5$^{\circ}$ in both cases. The standard errors in these
1741: mean recovered spectra are smaller than the plotted symbols.}
1742: \label{fig:cls_pol_angles}
1743: \end{figure}
1744:
1745: \subsection{Mis-estimated time-constants}
1746:
1747: The power spectra measured from simulations which included random and
1748: systematic errors in the detector time-constants displayed neither a
1749: bias, nor a degradation in error-bars. This was to be expected for the
1750: slow scan speed and extremely fast time-constants we have considered
1751: in this analysis -- the response function of the \clover\, detectors
1752: is effectively phase-preserving with zero attenuation in the frequency
1753: band which contains the sky-signal in our simulations. Note that this
1754: would not necessarily have been the case had we considered a much
1755: faster scan speed or more rapid polarization modulation.
1756:
1757: \subsection{Pointing errors}
1758:
1759: Our analysis of simulations where we have introduced a jitter in the
1760: pointing and/or an overall wander in the pointing suggest that these
1761: systematics have only a very small effect on the recovered $B$-mode
1762: power spectra, at least for the levels which we have considered
1763: (i.e. a $30$ arcsec random jitter in the pointing and/or an overall
1764: wander of the pointing by $1$ arcmin over the course of a two-hour
1765: observation). The only observed effect was a slight suppression of the
1766: recovered $B$-mode signal at high multipoles consistent with a
1767: slight smearing of the effective beam. We note however that the effect
1768: we observed was extremely small and was only noticeable in our
1769: signal-only simulations. For our simulations containing noise, the
1770: effect was completely swamped by the errors due to random noise.
1771:
1772: In principle, pointing errors can also lead to leakage
1773: from $E$ to $B$~\citep{hu03,odea07}. In Appendix~\ref{app:pointing}
1774: we develop a toy-model for
1775: the leakage expected from random pointing jitter in the case of
1776: a scan/modulation strategy that produces a uniform spread of polarization
1777: sensitivity directions in each sky pixel (such as by fast modulating
1778: with a HWP). The result is a white-noise spectrum of $B$-modes but,
1779: for the simulation parameters adopted here, the effect is very small --
1780: less than $1$ per cent of the $B$-mode power induced by weak gravitational lensing
1781: on large scales.
1782:
1783: \subsection{Differential transmittance in the HWP}
1784: The power spectra reconstructed from our simulations which
1785: included a 2 per cent differential transmittance in the HWP exhibited
1786: no degradation in the accuracy of the recovered $B$-mode
1787: signal --- even the relatively simple recipe which we have used to
1788: remove the HWP-synchronous signals from the time-stream
1789: (see Section \ref{sec:map-making}) appears sufficient to recover the
1790: $B$-mode signal to the same accuracy as was seen in our reference
1791: simulations. (We quantify this statement in the next section where we
1792: estimate the detection significances with which the different simulations
1793: detect the $E$ and $B$-mode signals). As mentioned in
1794: Section~\ref{sec:modulation}, the recovered polarization signal is
1795: mis-calibrated by $\sim 2$ per cent in amplitude ($4$ per cent in
1796: power). Compared to the random noise however, this mis-calibration is
1797: a small effect and is easily dealt with during, e.g. a cosmological
1798: parameter analysis by marginalising over it.
1799:
1800: Note that no prior information on the level of differential
1801: transmittance was used during our analysis of the data. Our technique
1802: for removing the HWP-synchronous signals is a blind one in this
1803: sense and should work equally well for other HWP-systematic effects
1804: that result in spurious signals at harmonics of the HWP rotation
1805: frequency, $f_\lambda$.
1806:
1807: \section{Discussion}
1808: \label{sec:discussion}
1809:
1810: \subsection{Controlling systematics with polarization modulation}
1811: \label{sec:significances}
1812: The main goal of the analysis presented in this paper is to
1813: demonstrate and quantify with simulations the impact of two types of
1814: polarization modulation (slow modulation using a stepped HWP and rapid
1815: modulation with a continuously rotating HWP) on the science return of
1816: upcoming CMB $B$-mode experiments in the presence of various
1817: systematic effects. Although our list of included systematics is not
1818: an exhaustive one (in particular, we are still investigating the case
1819: of imperfect optics), we are nevertheless in a position to draw some
1820: rather general conclusions regarding the usefulness of modulation in
1821: mitigating systematics. It is, of course, important to bear in mind
1822: that we have only considered two examples of a HWP-related systematic
1823: effect (imperfect HWP angles and differential transmittance in the
1824: plate). The are many more possible effects which will need to be
1825: well understood and strictly controlled if fast polarization
1826: modulation with HWPs is to realise its potential.
1827:
1828: \subsubsection*{(i) Systematics mitigated by modulation}
1829: \begin{itemize}
1830: \item{\bf Correlated $1/f$ detector noise:} As expected by the general
1831: reasoning of Section~\ref{sec:modulation}, and further borne out
1832: by our results from simulations, rapid polarization modulation is
1833: extremely powerful at mitigating a correlated $1/f$ component in the
1834: detector noise. Any such $1/f$ component is not mitigated by
1835: a HWP operating in stepped mode.
1836: \item{\bf Calibration errors:} Our results demonstrate that fast
1837: modulation is also useful for mitigating against possible
1838: calibration errors since it greatly increases the range of
1839: directions over which sky polarization is measured in a given pixel.
1840: For example, the clear bias introduced in our simulations by random
1841: gain drifts or systematic mis-calibrations between detectors was
1842: mitigated entirely by the HWP continuously rotating at $3$~Hz. This
1843: bias was also partly (but not completely) mitigated by stepping the
1844: HWP by 20$^{\circ}$ between each of our azimuth scans. For some stepping
1845: strategies we have investigated, the bias actually increased --- a
1846: poor choice of stepping strategy can actually be worse than having
1847: no modulation because of interactions between the sky rotation and
1848: the HWP orientations.
1849: \end{itemize}
1850:
1851: \begin{table*}
1852: \caption{Detection significances (in units of $\sigma$) for our
1853: reference simulations, for our simulations with $1/f$ noise
1854: systematics and for our simulations with a 2 per cent differential
1855: transmittance in the HWP. Also included for comparison are the predicted
1856: detection significances from a Fisher matrix analysis of the power
1857: spectrum errors (see Section \ref{sec:fisher}) and from the
1858: simulations containing isotropic and uniform Gaussian noise (see
1859: text). The rightmost column displays the significance of the
1860: detection of the $B$-mode signal in excess of the lensing signal
1861: which corresponds directly to the significance with which each
1862: simulation detects the input tensor-to-scalar ratio of $r=0.026$.}
1863: \begin{center}
1864: \begin{tabular}{c|c|c|c|c}
1865: Simulation & Modulation & $E$-mode & $B$-mode & $r = 0.026$ \\
1866: \hline
1867: Fisher predictions & --- & $128.6$ & $10.24$ & $1.90$ \\
1868: \hline
1869: Uniform noise & --- & $126.4$ & $10.30$ & $1.45$ \\
1870: \hline
1871: Reference simulation & None & $127.8$ & $10.41$ & $1.54$ \\
1872: & Stepped HWP & $130.1$ & $10.11$ & $1.41$ \\
1873: & Rotating HWP & $127.7$ & $10.72$ & $1.45$ \\
1874: \hline
1875: $1/f$ detector noise & None & $124.2$ & $7.29$ & $0.83$ \\
1876: ($f_{\rm knee} = 0.1$~Hz) & Stepped HWP & $124.6$ & $7.09$ & $0.79$ \\
1877: & Rotating HWP & $126.8$ & $9.96$ & $1.47$ \\
1878: \hline
1879: $1/f$ detector noise & None & $125.1$ & $8.61$ & $0.95$ \\
1880: ($f_{\rm knee} = 0.05$~Hz) & Stepped HWP & $127.5$ & $8.59$ & $1.14$ \\
1881: & Rotating HWP & $125.6$ & $10.31$ & $1.45$ \\
1882: \hline
1883: $1/f$ detector noise & None & $126.8$ & $9.78$ & $1.30$ \\
1884: ($f_{\rm knee} = 0.01$~Hz) & Stepped HWP & $128.0$ & $9.99$ & $1.44$ \\
1885: & Rotating HWP & $127.1$ & $10.28$ & $1.40$ \\
1886: \hline
1887: Polarized atmosphere & None & $122.9$ & $6.86$ & $0.12$ \\
1888: & Stepped HWP & $124.9$ & $6.99$ & $0.10$ \\
1889: & Rotating HWP & $126.0$ & $8.24$ & $0.21$ \\
1890: \hline
1891: Differential transmittance & Rotating HWP & $127.6$ & $10.92$ & $1.59$ \\
1892: \hline
1893: \end{tabular}
1894: \end{center}
1895: \label{tab:simsummary}
1896: \end{table*}
1897:
1898: \vspace{-6mm}
1899: \subsubsection*{(ii) Systematics not mitigated by modulation}
1900: \begin{itemize}
1901: \item{\bf Polarized $1/f$ atmosphere:} No amount of modulation (rapid or
1902: slow) will mitigate a polarized $1/f$ component in the
1903: atmosphere. The results from our simulations containing polarized
1904: atmosphere are summarised in Table \ref{tab:simsummary}.
1905: \item{\bf Pointing errors:} For our simulations which included pointing
1906: errors, the effect on the recovered $B$-mode power spectra was
1907: extremely small and was equivalent to a slight smoothing of the
1908: effective beam. Although the same amount of smoothing was observed
1909: in all the simulations (and so the effect is not mitigated by
1910: polarization modulation), the effect is negligible for the
1911: sensitivities and beam sizes considered here. The further leakage of
1912: $E$-mode power into $B$ modes due to pointing errors was, as expected
1913: (see Appendix~\ref{app:pointing}) unobservably small in our simulations.
1914: \item{\bf Mis-calibration of polarization angles: } A polarization
1915: modulation scheme does not mitigate a systematic error in the
1916: calibration of the polarization sensitivity directions. Experiments
1917: using a HWP will require precise and accurate measurements of the
1918: HWP angle at any given time to avoid the $E \rightarrow B$ mixing
1919: apparent in Fig.~\ref{fig:cls_pol_angles}.
1920: \end{itemize}
1921:
1922: Our results are in broad agreement with those of a similar study by
1923: \cite{mactavish08} who based their analysis on signal-only simulations
1924: of the \spider\, experiment. Both \cite{mactavish08} and this study find
1925: that polarization modulation with a continuously rotating HWP is
1926: extremely effective in mitigating the effects of $1/f$ detector noise
1927: but that, in the presence of significant $1/f$ noise, the
1928: analysis of an experiment where modulation is either absent or slow
1929: will require near-optimal map-making techniques. In addition, both studies find that the effect
1930: of small and random pointing errors on the science return of upcoming
1931: $B$-mode experiments is negligible given the experimental
1932: sensitivities. The two analyses also find that the effect of random errors
1933: (with $\sim 0.5^{\circ}$ RMS) in the detector polarization sensitivity
1934: angles is negligible but that the global polarization coordinate frame
1935: of the experiment needs to be measured carefully --- \cite{mactavish08}
1936: quote a required accuracy of $< 0.25^{\circ}$ for \spider\, which is
1937: consistent with the requirement for an unbiased measurement of
1938: the $B$-mode signal (for $r = 0.026$) at $\ell < 300$ with \clover. Finally, both
1939: studies suggest that, in the absence of fast modulation, relative gain errors
1940: will also need to be controlled to the $< 1$ per cent level (although, in
1941: this paper, we have demonstrated that such gain errors are almost
1942: entirely mitigated using a fast modulation scheme; see
1943: Fig.~\ref{fig:cls_gain_errors}).
1944:
1945: In Table~\ref{tab:simsummary}, we quantify the impact of modulation on
1946: the $1/f$ noise systematics we have considered in this work by
1947: considering the significances with which we detect the $E$-mode and
1948: $B$-mode signals. For comparison, the detection significances in the
1949: presence of a 2 per cent differential transmittance in the HWP are also
1950: presented. To calculate the total significance of the detection
1951: we compute the Fisher error on the amplitude of a fiducial
1952: spectrum,
1953: \be
1954: \frac{S}{N} = \left( \sum_{bb'} P_b^{\rm fid} {\rm cov}^{-1}_{bb'}
1955: P_{b'}^{\rm fid} \right)^{1/2},
1956: \ee
1957: where ${\rm cov}^{-1}_{bb'}$ is the inverse of the band-power covariance
1958: matrix for the given spectrum.
1959: For the total significance of a detection of $E$ or $B$-modes, the
1960: fiducial band-powers are simply the binned input power spectra. In
1961: order to estimate the significance of a detection of primordial
1962: $B$-modes, we subtract the lensing contribution from the input
1963: $B$-mode power spectra. Because the primordial $B$-mode power spectrum
1964: is directly proportional to the tensor-to-scalar ratio, $r$, the
1965: significance with which we detect the $B$-mode signal in excess of the
1966: lensing signal translates directly to a significance for the detection
1967: of our input tensor-to-scalar ratio of $r=0.026$.
1968: When analysing the results of the
1969: simulations, we approximate ${\rm cov}^{-1}_{bb'} \approx \delta_{bb'}/
1970: \sigma_b^2$ since we are unable to estimate the off-diagonal elements
1971: from our small number of realisations (50) in each simulation set.
1972: We know from a Fisher-based analysis (see Section~\ref{sec:fisher}),
1973: the results of which are also reported in Table~\ref{tab:simsummary},
1974: that neighbouring band-powers on the largest scales are, in fact,
1975: $\sim 10$ per cent
1976: anti-correlated, and the diagonal approximation therefore
1977: \emph{underestimates} the detection significance by $\sim 10$ per cent.
1978: For consistency, the numbers quoted for the Fisher analysis ignore the
1979: off-diagonal elements of the covariance matrix. Including the correlations
1980: increases the $E$-mode significance to $144.6$ (from $128.6$),
1981: the total $B$-mode significance to $11.57$ (from $10.24$) and
1982: the primordial $B$-mode significance to $2.04$ (from $1.90$).
1983:
1984: In comparing the entries in Table~\ref{tab:simsummary} one should
1985: keep in mind that the significances reported for the simulations are
1986: subject to a Monte-Carlo error due to the finite number ($N_{\mathrm{sim}}=50$)
1987: of simulations used to estimate the band-power errors. Approximating the
1988: band-powers as uncorrelated and Gaussian distributed, the
1989: sampling error in our estimates of the $S/N$ is
1990: %
1991: \be
1992: \Delta (S/N) \approx \frac{1}{S/N} \frac{1}{\sqrt{2N_{\mathrm{sim}}}}
1993: \left[\sum_b \left(\frac{P_b^{\mathrm{fid}}}{\sigma_b}\right)^4\right]^{1/2} .
1994: \ee
1995: %
1996: For the reference simulation, this gives an error of $0.15$ in the
1997: significance of a detection of $r$ and $0.22$ in the significance of the
1998: total $B$-mode spectrum. The size of these errors likely explain the
1999: apparent anomalies that rotating the HWP degrades the detection of $r$
2000: in the reference simulation, and that adding $1/f$ detector noise
2001: improves the detection of $r$ over the reference simulation for the case
2002: of a rotating HWP.
2003:
2004: Also included in Table~\ref{tab:simsummary} is the performance of our
2005: experiment as estimated from a set
2006: of simple map-based simulations where we have injected uniform and
2007: isotropic white noise into signal-only $T$, $Q$ and $U$ maps
2008: directly. In these simulations, and also for the Fisher analysis,
2009: the white-noise levels were chosen to
2010: match the noise levels in our main analysis and so they have identical
2011: raw sensitivity to the time-stream simulations but with perfectly
2012: behaved noise properties. The broad agreement between our full
2013: time-stream simulations and these simple map-based simulations
2014: suggests that the anisotropic noise distribution introduced by the
2015: \clover\, scan strategy does not have a large impact on the
2016: performance of the experiment. This agreement also suggests that the
2017: slightly poorer performance of the simulations in recovering the
2018: $r=0.026$ primordial $B$-mode signal as compared to the Fisher
2019: predictions is due to the sub-optimal performance on large scales of the
2020: (pure) pseudo-$C_\ell$ estimator we have used.
2021:
2022: \subsection{Importance of combining data from multiple detectors}
2023: \label{sec:differencing}
2024: For all of our analyses up to this point, in order to remove the
2025: correlated $1/f$ atmospheric noise from the polarization
2026: analysis, we have differenced detector pairs before
2027: map-making. However, as mentioned in Section~\ref{sec:modulation}, for
2028: the case of a continuously modulated experiment, it is possible to
2029: measure all three Stokes parameters from a single detector in
2030: isolation. Here, we argue that this may be a poor choice of analysis
2031: technique in the presence of a highly correlated and common-mode
2032: systematic such as atmospheric $1/f$, at least when one employs
2033: real-space demodulation techniques such as those that we have used in
2034: this analysis. The key point to appreciate here is that, even with a
2035: rapid modulation scheme, and for an ideal experiment, it is impossible
2036: to separate completely the temperature and polarization signals in
2037: real space using only a single detector.\footnote{We note that this is
2038: not necessarily true for the case of genuinely band-limited
2039: temperature and polarization signals when a classical lock-in
2040: technique (such as that used in the analysis of the \maxipol\, data;
2041: \citealt{johnson07}) is used to perform the demodulation. We are
2042: currently working to integrate such a technique into our analysis.}
2043: In contrast, the technique of detector differencing achieves this
2044: complete separation of the temperature and polarization signals (again
2045: for the case of an ideal experiment). Note that this is true even in
2046: the signal-only case. Consider again the modulated signal, in the
2047: absence of noise, from a single detector sensitive to a single
2048: polarization: \be d_i = \left[ T(\theta) + Q(\theta) \cos(2\phi_i) +
2049: U(\theta)\sin(2\phi_i) \right]/2. \ee If for each observed data
2050: point, $d_i$, the true sky signals, $T$, $Q$ and $U$ are different (as
2051: is the case for a scanning experiment), there is clearly no way to
2052: recover the true values of $T$, $Q$ and $U$ at each point in time.
2053: The approximation that one must make in order to demodulate the data
2054: in real space goes to the very heart of map-making -- that the true
2055: continuously varying sky signal can be approximated as a pixelised
2056: distribution where $T$, $Q$ and $U$ are taken to be constant within
2057: each map-pixel. Armed with this assumption, all three Stokes
2058: parameters can be reconstructed from a single detector time-stream
2059: using a generalisation of equation~(\ref{eqn:qu_mapmaking}):
2060: \be
2061: \left( \begin{array}{c} T \\ Q \\ U \end{array} \right) = 2 \, {\mathsf M}^{-1} \cdot
2062: \left( \begin{array}{c}
2063: \lgl d_i \rgl \\
2064: \lgl\cos(2\phi_i)d_i\rgl \\
2065: \lgl\sin(2\phi_i)d_i\rgl \\ \end{array} \right).
2066: \label{eqn:tqu_mapmaking}
2067: \ee
2068: where the decorrelation matrix, ${\mathsf M}$ is now given by
2069: \be
2070: {\mathsf M} = \left( \begin{array}{ccc}
2071: \!\!\! 1 & \!\!\! \lgl\cos(2\phi_i)\rgl & \!\!\!\!\! \lgl\sin(2\phi_i)\rgl \\
2072: \!\!\! \lgl\cos(2\phi_i)\rgl & \!\!\! \lgl\cos^2(2\phi_i)\rgl &
2073: \!\!\!\!\! \lgl\cos(2\phi_i)\sin(2\phi_i)\rgl \\
2074: \!\!\! \lgl\sin(2\phi_i)\rgl & \!\!\! \lgl\cos(2\phi_i)\sin(2\phi_i)\rgl &
2075: \!\!\!\!\! \lgl\sin^2(2\phi_i)\rgl \\ \end{array} \!\!\!\! \right).
2076: \ee
2077: \begin{figure}
2078: \centering
2079: \resizebox{0.48\textwidth}{!}{
2080: \rotatebox{-90}{\includegraphics{fig14.ps}}}
2081: \caption{Recovered noise only $U$-polarization maps from one of our
2082: reference simulations with continuous modulation. The map on the
2083: left is reconstructed from demodulated single detector ``pure'' $U$
2084: time streams and has not used information from multiple detectors to
2085: separate the $T$ and polarization signals. The map on the right is
2086: made from demodulated detector-pair ``pure'' $U$ time-streams and
2087: explicitly combines information from the two detectors within each
2088: pixel to separate $T$ from $Q/U$. Although striping is absent from
2089: both maps, the white-noise level in the detector-differenced map is
2090: reduced compared to that made using the non-differencing
2091: analysis.}
2092: \label{fig:maps_diff_nodiff}
2093: \end{figure}
2094: If, on the other hand, the data from different detectors are combined
2095: (e.g.\ when detector differencing is used), the situation is different
2096: -- because the two detectors within a pixel observe exactly the same
2097: un-polarized component of the sky signal at exactly the same time,
2098: differencing the detectors removes
2099: the $T$ signal completely without any assumptions regarding the scale
2100: over which the true sky signal is constant. In this case, the
2101: decorrelation of $Q$ and $U$ using equation~(\ref{eqn:qu_mapmaking})
2102: still requires an assumption regarding the constancy of the $Q$ and
2103: $U$ signals over the scale of a map-pixel but now the much larger
2104: temperature signal has been removed from the polarization analysis
2105: completely.
2106:
2107: Now, if in addition to the sky signal, we have a common-mode
2108: time-varying systematic such as an un-polarized $1/f$ component in the
2109: atmosphere, this contaminant will again be removed entirely with the
2110: detector differencing technique (as long as it is completely
2111: correlated between the two detectors) whilst it will introduce a
2112: further approximation into any attempt to decorrelate all three Stokes
2113: parameters from a single detector time-stream using
2114: equation~(\ref{eqn:tqu_mapmaking}).
2115:
2116: To illustrate this point, and to stress the importance of combining
2117: data from multiple detectors, we have re-analysed the simulated data
2118: from our set of continuously modulated reference simulations but now
2119: we perform the demodulation at the time-stream level for either single
2120: detectors or single detector-pairs in isolation. Firstly, we
2121: demodulate each detector time-stream individually using
2122: equation~(\ref{eqn:tqu_mapmaking}) but now the averaging is performed
2123: over short segments in time rather than over all data falling within a
2124: map-pixel. This procedure results in ``pure'' $T$, $Q$ and $U$
2125: time-streams for each detector but at a reduced data rate determined
2126: by the number of data samples over which the averaging of
2127: equation~(\ref{eqn:tqu_mapmaking}) is performed. For our second
2128: analysis, we first difference each detector pair and then demodulate
2129: the differenced time-streams using equation~(\ref{eqn:qu_mapmaking}),
2130: again applied over short segments of time, resulting in ``pure'' $Q$
2131: and $U$ time-streams for each detector pair, once again at a reduced
2132: data rate. Maps of the $Q$ and $U$ Stokes parameters are then
2133: constructed by simple binning of the demodulated $Q$ and $U$
2134: time-streams from all detectors or detector pairs. In the case where
2135: detector pairs are differenced, we are explicitly combining
2136: information from multiple detectors to separate the temperature and
2137: polarization signals whilst when we do not difference, we are
2138: attempting to separate the $T$ and $Q/U$ signals present in detector
2139: time-stream in isolation.
2140:
2141: The results of these tests are shown in Figs.~\ref{fig:maps_diff_nodiff} and
2142: \ref{fig:cls_diff_nodiff}. Figure~\ref{fig:maps_diff_nodiff} shows the noise-only
2143: $U$-polarization maps recovered using the two different analysis
2144: techniques. Although striping from the atmospheric fluctuations is not
2145: present in either of the maps, the extra uncertainty introduced when
2146: one attempts to separate the $T$ and $Q/U$ signals from individual
2147: detectors in isolation clearly results in an increased white noise
2148: level in the polarization maps. This results in a degradation factor
2149: of $\sim 2$ in the resulting measurements of the $B$-mode power
2150: spectrum on all angular scales (Fig.~\ref{fig:cls_diff_nodiff}) with
2151: a corresponding degradation in the detection significances for both
2152: a measurement of the total $B$-mode signal and for a detection of $r=0.026$
2153: (Table~\ref{tab:diff_nodiff}). These results are in excellent
2154: agreement with those of \cite{sutton09} who found it necessary to
2155: apply optimal mapping techniques to rapidly modulated single-detector
2156: time-streams in the presence of $1/f$ atmospheric noise.
2157:
2158: We emphasise that the results presented in
2159: Fig.~\ref{fig:cls_diff_nodiff} and in Table~\ref{tab:diff_nodiff} for
2160: the case where we have analysed single detectors in isolation would
2161: likely be improved if the Fourier domain filtering used in
2162: \cite{johnson07} was implemented. We are currently working to
2163: integrate this step into our algorithm, and we expect to report the
2164: subsequent improvement in future publications.
2165:
2166: \begin{figure}
2167: \centering
2168: \resizebox{0.48\textwidth}{!}{
2169: \rotatebox{-90}{\includegraphics{fig15.ps}}}
2170: \caption{Comparison between the $B$-mode power spectra recovered
2171: using an analysis based on differencing detector pairs (black
2172: points) and one based on demodulating each detector individually
2173: (blue points). In the presence of a time-varying common-mode
2174: systematic, such as the $1/f$ atmospheric noise we have considered
2175: here, the analysis based on detector differencing is far superior to
2176: the analysis based on demodulating each detector individually.}
2177: \label{fig:cls_diff_nodiff}
2178: \end{figure}
2179:
2180: \begin{table}
2181: \caption{Detection significances (in units of $\sigma$) from the
2182: analysis of identical simulated data with a HWP rotating
2183: continuously at $3$~Hz. The first analysis is based on detector
2184: differencing, the second based on demodulation of individual
2185: detectors in isolation.}
2186: \begin{center}
2187: \begin{tabular}{c|c|c|c}
2188: Analysis & $E$-mode & $B$-mode & $r = 0.026$ \\
2189: \hline
2190: Detector differencing & $132.9$ & $10.14$ & $1.43$ \\
2191: Demodulation & $123.4$ & $5.11$ & $0.68$ \\
2192: \hline
2193: \end{tabular}
2194: \end{center}
2195: \label{tab:diff_nodiff}
2196: \end{table}
2197:
2198: \subsection{Comparison of simulated and predicted \clover\, performance}
2199: \label{sec:fisher}
2200: It is common practice to make predictions of the performance of
2201: upcoming experiments using a Fisher-matrix analysis which
2202: attempts to predict the achievable errors on, for example, power
2203: spectra or cosmological parameters under some simplifying
2204: assumptions. Generally these assumptions will include uniform coverage
2205: of the observing fields and uncorrelated Gaussian noise resulting in
2206: an isotropic and uniform white-noise distribution across the observing
2207: field. In contrast, the work described in this paper has made use of a
2208: detailed simulation pipeline which we have created for the \clover\,
2209: experiment. Our simulation pipeline includes the \clover\,
2210: focal-plane designs as well as a realistic scan strategy appropriate
2211: for observing the four chosen \clover\, fields from the telescope site
2212: in Chile. In addition we have employed a detailed model of the TES
2213: detector noise properties and responsivity, and $1/f$ atmospheric
2214: noise correlated across the focal-plane array. Moreover, our errors
2215: are calculated using a Monte-Carlo analysis and so should
2216: automatically include any effects due to correlations between map
2217: pixels etc. An interesting exercise therefore is to compare the
2218: expected errors from a Fisher-matrix analysis to those obtained
2219: from our simulation analysis.
2220:
2221: The polarization band-power Fisher matrix is (e.g.~\citealt{tegmark01})
2222: %
2223: \be
2224: \mathrm{cov}^{-1}_{(bP)(bP)'} = \frac{1}{2}\mathrm{tr}\left(
2225: \mC^{-1} \frac{\partial \mC}{\partial \mathcal{C}_b^P}
2226: \mC^{-1} \frac{\partial \mC}{\partial \mathcal{C}_{b'}^{P'}}
2227: \right)
2228: \ee
2229: %
2230: where $P$ and $P'$ are $E$ or $B$, and $b$ labels the bandpower. Here,
2231: $\mC$ is the covariance matrix of the noisy Stokes maps and
2232: $\mathcal{C}_b^P$ are bandpowers of $\ell (\ell +1)C_\ell^P/(2\pi)$.
2233: We analyse a single circular field with area equal to that retained
2234: in the pseudo-$C_\ell$ analysis described in Section~\ref{subsec:power},
2235: and multiply the Fisher matrix by four to account for the
2236: number of fields observed (which are thus assumed to be fully independent).
2237: We ignore inhomogeneity of the noise in the maps so that our problem
2238: has azimuthal symmetry about the field centre. This allows us to
2239: work in a basis where the data is Fourier transformed in azimuth and
2240: the covariance matrix becomes block diagonal, thus speeding up the
2241: computation of the Fisher matrix considerably. The Fisher matrix
2242: takes full account of band-power correlations (both between $b$ and
2243: polarization type) and the effect of ambiguities in isolating
2244: $E$ and $B$-modes given the survey geometry. We deal with power
2245: on scales larger than the survey by including a junk band-power for
2246: each of $E$ and $B$ whose contribution to $\mathrm{cov}_{(bP)(bP)'}$
2247: we remove before computing detection significances.
2248:
2249: The comparison between the predicted and simulated performance of
2250: \clover\, is shown in Fig.~\ref{fig:clover_compare}. In this plot, we
2251: also include the predicted $B$-mode errors from a na\"{\i}ve mode-counting
2252: argument based on the fraction of sky observed, $f_{\rm sky}$. For
2253: these estimates, we assume independent measurements of the power
2254: spectrum in bands of width $\Delta \ell$ given by
2255: \be
2256: (\Delta C_\ell)^2 = \frac{2}{(2\ell + 1) f_{\rm sky} \Delta \ell}
2257: (C_{\ell} + N_{\ell})^2,
2258: \ee
2259: where $C_\ell$ is the band-averaged input signal and $N_\ell$ is the
2260: band-averaged noise. For uncorrelated and isotropic Gaussian random
2261: noise, the latter is given by $N_\ell = w^{-1} B_\ell^{-2}$
2262: where $B_\ell = \exp( - \ell (\ell + 1) \sigma_B^2 / 2)$ is the
2263: transform of the beam with $\sigma_B = \theta_B / \sqrt{8 \ln 2}$ for
2264: a beam with FWHM of $\theta_B$. The weight $w^{-1} =
2265: \Omega_{\rm pix} \sigma^2_{\rm pix}$, where the pixel noise in the $Q$
2266: and $U$ maps is
2267: \begin{equation}
2268: \sigma^2_{\rm pix} = \frac{ ({\rm NET}/\sqrt{2})^2 \Theta^2}{t_{\rm obs}
2269: (N_{\rm det}/4) \Omega_{\rm pix}}.
2270: \label{eqn:pixel_noise}
2271: \end{equation}
2272: Here $\Theta^2$ is the total observed area, $t_{\rm obs}$ is
2273: the total observation time and $\Omega_{\rm pix}$ is the pixel
2274: size. In equation~(\ref{eqn:pixel_noise}), we have used ${\rm
2275: NET}/\sqrt{2}$ to account for the fact that a single measurement of
2276: $Q$ or $U$ requires a measurement from two detectors (or,
2277: alternatively, two measurements from a single detector) and we use
2278: $N_{\rm det}/4$ as the effective number of $Q$ and $U$ detectors.
2279:
2280: Over most of the $\ell$ range, the agreement between the Fisher matrix
2281: predictions and the simulated performance is rather good -- the only
2282: significant discrepancy is for the lowest band-power where the
2283: simulations fail to match the predicted Fisher error. This is almost
2284: certainly due to the relatively poor performance of our power spectrum
2285: estimator on the very largest scales where pseudo-$C_\ell$ techniques
2286: are known to be sub-optimal (compared to, for example, a maximum
2287: likelihood analysis). In terms of a detection of the total $B$-mode
2288: signal, the Fisher analysis predicts a detection for \clover\, of
2289: $\sim 12.0\sigma$. For comparison, the na\"{\i}ve $f_{\rm sky}$
2290: analysis predicts a $12.4\sigma$ detection. For our assumed
2291: tensor-to-scalar ratio, the Fisher matrix analysis predicts $r=0.026
2292: \pm 0.013$ (a $2.04\sigma$ detection) and the na\"{\i}ve analysis
2293: yields $r=0.026 \pm 0.011$ (a $2.39\sigma$ detection). Comparing to
2294: the detection significances quoted in Table~\ref{tab:simsummary}, we
2295: see that the detections recovered from the simulations fail to match
2296: these numbers. For the case of the total $B$-mode amplitude, this
2297: discrepancy is entirely due to the fact that we are unable to measure
2298: and include in our analysis (anti-)correlations between the
2299: band-powers measured from our small number (50) of simulations --
2300: when we neglect the correlations in the Fisher matrix analysis, the
2301: Fisher prediction drops to a $10.24\sigma$ detection, in excellent
2302: agreement with our measured value from simulations. For the primordial
2303: $B$-mode signal only, the discrepancy found is also partly due to the
2304: same effect (ignoring correlations in the Fisher analysis reduces the
2305: Fisher prediction for primordial $B$-modes to $1.9\sigma$). As
2306: mentioned above, we suspect that the additional decrease in sensitivity to
2307: primordial $B$-modes seen in the simulations is due to the slightly
2308: sub-optimal performance of our implementation of the pure
2309: pseudo-$C_\ell$ estimator on the largest scales.
2310:
2311: We should point out that in this work, we have made no attempt to optimise the
2312: survey strategy in light of recent instrument developments. In
2313: particular, the survey size we have adopted for these simulations was
2314: optimised for a measurement of $r=0.01$ with \clover\, when the
2315: experiment was expected to have twice the number of detectors now
2316: planned. For the instrument parameters we have adopted in this analysis
2317: (which are a fair representation of the currently envisaged
2318: experiment), the optimal survey area for a measurement of $r=0.026$
2319: would be significantly smaller than the $\sim 1500$ deg$^2$ we have
2320: used here due to the increased noise levels from the reduced number of
2321: detectors. Alternatively, if we had assumed a larger input value of
2322: $r$, the optimal survey size would increase. Optimisation of both the
2323: survey area and the scan strategy in light of these changes in the
2324: instrument design is the subject of on-going work.
2325: \begin{figure}
2326: \centering
2327: \resizebox{0.48\textwidth}{!}{
2328: \rotatebox{-90}{\includegraphics{fig16.ps}}}
2329: \caption{Comparison between the predicted performance of \clover\,
2330: as calculated using a Fisher-matrix analysis and the simulated
2331: performance from our Monte-Carlo pipeline (for our reference
2332: simulation). Also shown for comparison are the errors predicted from
2333: a na\"{\i}ve $f_{\rm sky}$ analysis.}
2334: \label{fig:clover_compare}
2335: \end{figure}
2336: There are, of course, many other sources of uncertainty which we have
2337: not yet accounted for in our simulation pipeline and so both the
2338: predicted and simulated performance numbers should be taken only as
2339: guidelines at this time. However, it is encouraging that the extra
2340: sources of uncertainty which are included in our simulation pipeline
2341: (realistic instrument parameters, a realistic scan strategy,
2342: correlated noise), in addition to any uncertainties introduced as part
2343: of our subsequent analysis of the simulated data, do not degrade the
2344: expected performance of \clover\, by a large amount.
2345:
2346: \section{Conclusions}
2347: \label{sec:conclusions}
2348: We have performed a detailed investigation of the ability of both slow
2349: and fast polarization modulation schemes to mitigate possible
2350: systematic effects in upcoming CMB polarization experiments, targeted
2351: at measuring the $B$-mode signature of gravitational waves in the
2352: early universe. To do this we have used a simulation pipeline
2353: developed in the context of the \clover\, experiment, which includes
2354: realistic instrument and observation parameters as well as $1/f$
2355: detector noise and $1/f$ atmospheric noise correlated across the
2356: \clover\, focal-plane array. Using this simulation tool, we have
2357: performed simulations of \clover\, operating with no explicit
2358: modulation, with a stepped HWP and with a HWP rotating continuously at
2359: $3$~Hz. We have analysed the resulting time-stream simulations using
2360: the technique of detector differencing coupled with a na\"{\i}ve map-making
2361: scheme, and finally have reconstructed the $E$ and $B$-mode power
2362: spectra using an implementation of the near-optimal ``pure''
2363: pseudo-C$_\ell$ power spectrum estimator.
2364:
2365: As expected, we find that fast modulation via a continuously rotating
2366: HWP is extremely powerful in mitigating a correlated $1/f$ component
2367: in the detector noise but that a stepped HWP is not. In addition, we
2368: have demonstrated that a polarized $1/f$ component in the atmosphere
2369: is not mitigated by any amount of modulation and if present, would
2370: need to be mitigated in the analysis using a sophisticated map-making
2371: technique. We have further verified with simulations that fast
2372: modulation is very effective in mitigating instrumental polarization
2373: that is fixed relative to the instrument basis, for example the $T
2374: \rightarrow Q$ leakage caused by systematic gain errors and
2375: mis-matches between detectors, in agreement with the conclusions
2376: of~\citet{odea07}. We have also demonstrated that modulation does not
2377: mitigate a systematic mis-calibration of polarization angles and that
2378: these angles will need to be measured accurately in order to avoid a
2379: systematic leakage between $E$ and $B$-modes. The other systematics
2380: which we have investigated (pointing errors, mis-estimated
2381: time-constants) have a negligible impact on the recovered power
2382: spectra for the parameters adopted in our simulations.
2383:
2384: In addition to our investigation of systematic effects, we have
2385: stressed the importance of combining data from multiple detectors and
2386: have demonstrated the superior performance of a differencing technique
2387: as opposed to one based on measuring all three Stokes parameters from
2388: single detectors in isolation. We suggest that the latter technique,
2389: although possible in the presence of rapid modulation, is likely a
2390: poor choice of analysis technique, at least in the presence of a
2391: common-mode systematic effect such as atmospheric $1/f$ noise.
2392:
2393: Finally, we have compared the simulated performance of the \clover\,
2394: experiment with the expected performance from a simplified
2395: Fisher-matrix analysis. For all but the very lowest multipoles, where
2396: the simulations fail to match the Fisher predictions, we find
2397: excellent agreement between the predicted and simulated
2398: performance. In particular, despite the highly anisotropic noise
2399: distribution present in our simulated maps, our measurement of the
2400: total $B$-mode signal matches closely with the Fisher matrix
2401: prediction (the latter assuming isotropic noise). On the other hand,
2402: the measurement of the large scale $B$-mode signal (and thus of the
2403: tensor-to-scalar ratio, $r$) from the simulations is around 20 per
2404: cent worse
2405: than the Fisher prediction. This is almost certainly due to the
2406: sub-optimality of our power spectrum analysis on large scales. It is
2407: possible that the Fisher matrix predictions could be recovered from
2408: the simulations by using a more optimal weighting scheme in the pure
2409: pseudo-$C_\ell$ analysis, or, more likely, by using a
2410: maximum-likelihood $C_\ell$ estimator for the low multipoles.
2411:
2412: One important class of systematic effects which we have not considered
2413: in this paper are those associated with imperfect optics. Additionally,
2414: we have considered only two effects associated with an imperfect HWP.
2415: The efficacy of fast modulation to mitigate systematic
2416: effects from imperfect optics, for example instrumental polarization
2417: due to beam mis-match, is expected to depend critically on the optical
2418: design (such as the HWP location). We are currently working to
2419: include such optical systematics in our simulation pipeline, along
2420: with a more detailed physical model of the atmosphere and models of
2421: the expected polarized foreground emission. In future work, in
2422: addition to investigating further systematic effects, we will extend
2423: our simulations to multi-frequency observations and will use these to
2424: test alternative foreground removal techniques. We will also apply the
2425: ``destriping'' map-making technique of \cite{sutton09} to our
2426: simulations to assess the relative merits of destriping in analysis as
2427: opposed to a hardware based approach for mitigating $1/f$ noise.
2428:
2429: \section*{Acknowledgments}
2430: We are grateful to the \clover\, collaboration for useful
2431: discussions. We thank Kendrick Smith for making his original pure
2432: pseudo-$C_\ell$ code available which we adapted to carry out some of
2433: the analysis in this paper. We thank John Kovac and Jamie Hinderks for
2434: the up-to-date descriptions of the {\sevensize BICEP-2/KECK} and \piper\,
2435: experiments respectively. The simulation work described in this paper was carried
2436: out on the University of Cambridge's distributed computing facility,
2437: \camgrid. We acknowledge the use of the \fftw\, \citep{frigo05},
2438: \camb\, \citep{lewis00} and \healpix\, \citep{gorski05} packages.
2439:
2440: \setlength{\bibhang}{2.0em}
2441: \vspace{-3mm}
2442: \begin{thebibliography}{999}
2443:
2444: \bibitem[\protect\citeauthoryear{Amblard, Cooray \& Kaplinghat}{2007}]{amblard07}
2445: Amblard A., Corray A., Kaplinghat M., 2007, PRD, 75, 083508
2446:
2447: \bibitem[\protect\citeauthoryear{Battistelli et al.}{2008}]{battistelli08}
2448: Battistelli E. et al., 2008, ``Proc. of the 12th International
2449: Workshop on Low Temperature Detectors'', Journal of Low Temperature Physics,
2450: 151, 908
2451:
2452: \bibitem[\protect\citeauthoryear{de Bernardis et al.}{2008}]{deBernardis08}
2453: de Bernardis P. et al., 2008, Experimental Astronomy, 23, 5
2454:
2455: \bibitem[\protect\citeauthoryear{Bischoff et al.}{2008}]{bischoff08}
2456: Bischoff C. et al., 2008, ApJ, 684, 771
2457:
2458: \bibitem[\protect\citeauthoryear{Bock et al.}{2008}]{bock08}
2459: Bock J. et al., 2008, Arxiv astrophysics e-prints, 0805.4207
2460:
2461: \bibitem[\protect\citeauthoryear{Bunn}{2002}]{bunn02}
2462: Bunn E.F., 2002, PRD, 65, 043003
2463:
2464: \bibitem[\protect\citeauthoryear{Bussmann, Holzapfel \& Kuo}{2005}]{bussmann05}
2465: Bussmann R.S., Holzapfel W.L., Kuo C.L., 2005, ApJ, 622, 1343
2466:
2467: \bibitem[\protect\citeauthoryear{Challinor \& Chon}{2005}]{challinor05}
2468: Challinor A., Chon G., 2005, MNRAS, 360, 509
2469:
2470: \bibitem[\protect\citeauthoryear{Charlassier}{2008}]{charlassier08}
2471: Charlassier R., 2008, Arxiv astrophysics e-prints, 0805.4527
2472:
2473: \bibitem[\protect\citeauthoryear{Crill et al.}{2008}]{crill08}
2474: Crill B.P, et al., 2008, Proc.\ SPIE, 7010, 70102P
2475:
2476: \bibitem[\protect\citeauthoryear{Frigo \& Johnson}{2005}]{frigo05}
2477: Frigo M., Johnson S.G., 2005, Proceedings of the IEEE, 93, 216
2478:
2479: \bibitem[\protect\citeauthoryear{G\'orski et al.}{2005}]{gorski05}
2480: G\'orski, K.M., Hivon E., Banday A.J., Wandelt B.D., Hansen F.K.,
2481: Reinecke M., Bartelmann M., 2005, ApJ, 622, 759
2482:
2483: \bibitem[\protect\citeauthoryear{Hinderks et al.}{2009}]{hinderks09}
2484: Hinderks J. et al., 2009, ApJ, 692, 1221
2485:
2486: \bibitem[\protect\citeauthoryear{Hinshaw et al.}{2003}]{hinshaw03}
2487: Hinshaw G. et al., 2003, ApJS, 148, 135
2488:
2489: \bibitem[\protect\citeauthoryear{Hinshaw et al.}{2009}]{hinshaw09}
2490: Hinshaw G. et al., 2009, ApJS, 180, 225
2491:
2492: \bibitem[\protect\citeauthoryear{Hu, Hedman \& Zaldariagga}{2003}]{hu03}
2493: Hu W., Hedman M.M., Zaldariagga M., 2003, PRD, 67, 043004
2494:
2495: \bibitem[\protect\citeauthoryear{Irwin \& Hilton}{2005}]{irwin05}
2496: Irwin K.D., Hilton G.C., 2005, ``Transition Edge Sensors'', in
2497: ``Cryogenic Particle Detectors'', Topics Appl. Phys., 99, 63
2498:
2499: \bibitem[\protect\citeauthoryear{Johnson et al.}{2007}]{johnson07}
2500: Johnson B.R. et al., 2007, ApJ, 665, 42
2501:
2502: \bibitem[\protect\citeauthoryear{Kamionkowsky, Kosowksy \& Stebbins}{1997}]{kamionkowsky97}
2503: Kamionkowski M., Kosowsky A., Stebbins A., 1997, PRL, 78, 2058
2504:
2505: \bibitem[\protect\citeauthoryear{Kaplinghat, Knox \& Song}{2003}]{kaplinghat03}
2506: Kaplinghat M., Knox L., Song Y-S., 2003, PRL, 91, 241301
2507:
2508: \bibitem[\protect\citeauthoryear{Keating et al.}{2003}]{keating03}
2509: Keating B.G., Ade P.A.R., Bock J.J., Hivon E., Holzapfel W.L., Lange A.E.,
2510: Nguyen H., Yoon K.W., 2003, Proc.\ SPIE, 4843, 314
2511:
2512: \bibitem[\protect\citeauthoryear{Kesden, Cooray \& Kamionkowski}{2002}]{kesden02}
2513: Kesden M., Cooray A., Kamionkowski M., 2002, PRL, 89, 011304
2514:
2515: \bibitem[\protect\citeauthoryear{Knox \& Song}{2002}]{knox02}
2516: Knox L., Song Y., 2002, PRL, 89, 011303
2517:
2518: \bibitem[\protect\citeauthoryear{Komatsu et al.}{2009}]{komatsu09}
2519: Komatsu E. et al., 2009, ApJS, 180, 330
2520:
2521: \bibitem[\protect\citeauthoryear{Leitch et al.}{2005}]{leitch05}
2522: Leitch E.M., Kovac J.M., Halverson N.W., Carlstrom J.E., Pryke C.,
2523: Smith M.W.E., ApJ, 624, 10
2524:
2525: \bibitem[\protect\citeauthoryear{Lewis \& Challinor}{2006}]{lewis06}
2526: Lewis A., Challinor A., 2006, Physics Reports, 429, 1
2527:
2528: \bibitem[\protect\citeauthoryear{Lewis, Challinor \& Lasenby}{2000}]{lewis00}
2529: Lewis A., Challinor A., Lasenby A., 2000, ApJ, 538, 473
2530:
2531: \bibitem[\protect\citeauthoryear{Lewis, Challinor \& Turok}{2002}]{lewis02}
2532: Lewis A., Challinor A., Turok N., 2002, PRD, 65, 023505
2533:
2534: \bibitem[\protect\citeauthoryear{MacTavish et al.}{2008}]{mactavish08}
2535: MacTavish C.J. et al., 2008, ApJ, 689, 655
2536:
2537: \bibitem[\protect\citeauthoryear{Montroy et al.}{2006}]{montroy06}
2538: Montroy T.E. et al., 2006, ApJ, 647, 813
2539:
2540: \bibitem[\protect\citeauthoryear{Nolta et al.}{2008}]{nolta08}
2541: Nolta M.R. et al., 2009, ApJS, 180, 296
2542:
2543: \bibitem[\protect\citeauthoryear{North et al.}{2008}]{north08}
2544: North C.E., et al., 2008, Arxiv astrophysics e-prints, 0805.3690
2545:
2546: \bibitem[\protect\citeauthoryear{O'Dea, Challinor \& Johnson}{2007}]{odea07}
2547: O'Dea D., Challinor A., Johnson B.R., 2007, MNRAS, 376, 1767
2548:
2549: \bibitem[\protect\citeauthoryear{Oxley et al.}{2004}]{oxley04}
2550: Oxley P., et al., 2004, Proc.\ SPIE, 5543, 320
2551:
2552: \bibitem[\protect\citeauthoryear{Page et al.}{2007}]{page07}
2553: Page L. et al., 2007, ApJS, 170, 335
2554:
2555: \bibitem[\protect\citeauthoryear{Planck Collaboration}{2006}]{planck06}
2556: Planck Collaboration, 2006, Arxiv astrophysics e-prints, 0604069
2557:
2558: \bibitem[\protect\citeauthoryear{Pryke et al.}{2009}]{pryke09}
2559: Pryke C. et al., 2009, ApJ, 692, 1247
2560:
2561: \bibitem[\protect\citeauthoryear{Reintsema et al.}{2003}]{reintsema03}
2562: Reintsema C.D. et al., 2003, Review of Scientific Instruments, 74, 4500
2563:
2564: \bibitem[\protect\citeauthoryear{Samtleben}{2008}]{samtleben08}
2565: Samtleben D., 2008, Arxiv astrophysics e-prints, 0806.4334
2566:
2567: \bibitem[\protect\citeauthoryear{Seljak \& Zaldarriaga}{1997}]{seljak97}
2568: Seljak U., Zaldarriaga M., 1997, PRL, 78, 2054
2569:
2570: \bibitem[\protect\citeauthoryear{Shimon et al.}{2008}]{shimon08}
2571: Shimon M., Keating B., Ponthieu N., Hivon E., 2008, PRD, 77, 083003
2572:
2573: \bibitem[\protect\citeauthoryear{Sievers et al.}{2007}]{sievers07}
2574: Sievers J.L. et al., 2007, ApJ, 660, 976
2575:
2576: \bibitem[\protect\citeauthoryear{Smith, Hu \& Kaplinghat}{2004}]{smith04}
2577: Smith K.M., Hu W., Kaplinghat M., 2004, PRD, 70, 043002
2578:
2579: \bibitem[\protect\citeauthoryear{Smith}{2006}]{smith06}
2580: Smith K.M., 2006, PRD, 74, 083002
2581:
2582: \bibitem[\protect\citeauthoryear{Smith \& Zaldarriaga}{2007}]{smith07}
2583: Smith K.M., Zaldarriaga M., 2007, PRD, 76, 043001
2584:
2585: \bibitem[\protect\citeauthoryear{Smith, Challinor \& Rocha}{2006}]{smithchallinor06}
2586: Smith S., Challinor A., Rocha G., 2006, PRD, 73, 023517
2587:
2588: \bibitem[\protect\citeauthoryear{Sutton et al.}{2009}]{sutton09}
2589: Sutton D., Johnson B.R., Brown M.L., Cabella P., Ferreira P.G.,
2590: Smith K.M., 2009, MNRAS, 393, 894
2591:
2592: \bibitem[\protect\citeauthoryear{Tegmark \& de Oliveira-Costa}{2001}]{tegmark01}
2593: Tegmark M., de Oliveira-Costa A., 2001, PRD, 64, 063001
2594:
2595: \bibitem[\protect\citeauthoryear{Wu et al.}{2007}]{wu07}
2596: Wu J.H.P. et al., 2007, ApJ, 665, 55
2597:
2598: \bibitem[\protect\citeauthoryear{Yoon et al.}{2006}]{yoon06}
2599: Yoon K.Y., et al., 2006, Proc.\ SPIE, 6275, 62751K
2600:
2601: \bibitem[\protect\citeauthoryear{Zaldarriaga \& Seljak}{1997}]{zaldarriaga97}
2602: Zaldarriaga M., Seljak U., 1997, PRD, 55, 1830
2603:
2604: \bibitem[\protect\citeauthoryear{Zaldarriaga}{1997}]{zaldarriaga97b}
2605: Zaldarriaga M., 1997, PRD, 55, 4, 1822
2606:
2607: \end{thebibliography}
2608:
2609: \appendix
2610: \section{$E$-$B$ leakage from pointing jitter}
2611: \label{app:pointing}
2612:
2613: For the case of random pointing jitter and a scan/modulation
2614: strategy that ensures every sky pixel is visited with a wide range
2615: of polarization sensitivity directions, it is straightforward to estimate
2616: the spurious $B$-mode power that is induced from $E$-modes for a single
2617: focal-plane pixel. We work in the flat-sky approximation where
2618: directions on the sky are denoted by positions $\vx$ in the tangent plane.
2619: A given sky pixel $p$ receives $N_p$ hits with polarization sensitivity
2620: directions $\{ \phi_i \}$, where $i=1,\ldots,N_p$, and assume these
2621: angles are uniformly distributed.
2622: If $d_i$ is the $i$th differential measurement from the detector pair
2623: contributing to pixel $p$, then we can approximate the recovered
2624: Stokes parameters in that pixel (equation~\ref{eqn:qu_mapmaking})
2625: as
2626: %
2627: \begin{eqnarray}
2628: \hat{Q}(\vx_p) &\approx& \frac{2}{N_p} \sum_i d_i \cos (2\phi_i) \nonumber \\
2629: \hat{U}(\vx_p) &\approx& \frac{2}{N_p} \sum_i d_i \sin (2\phi_i) .
2630: \end{eqnarray}
2631: %
2632: If the $i$th observation has a pointing error $\valpha_i$, then
2633: the signal contribution to $\hat{Q}(\vx_p)$, for example, is
2634: %
2635: \begin{eqnarray}
2636: \hat{Q}(\vx_p) &\approx& \frac{2}{N_p} \sum_i \left[Q(\vx_p+\valpha_i)
2637: \cos^2(2\phi_i) \right. \nonumber \\
2638: &&\mbox{} \left. \phantom{xxxxxxx}
2639: + U(\vx_p+\valpha_i)\sin(2\phi_i) \cos(2\phi_i)\right] .
2640: \end{eqnarray}
2641: %
2642: Note that here, $Q$ and $U$ are the beam-smoothed fields.
2643: Approximating the displaced Stokes parameters by a gradient approximation,
2644: we find
2645: %
2646: \be
2647: \hat{Q}(\vx_p) \approx Q(\vx_p) + \valpha_{\mathrm{eff}} \cdot \vgrad Q
2648: |_{\vx_p}
2649: + \vbeta_{\mathrm{eff}} \cdot \vgrad U |_{\vx_p},
2650: \label{eq:gradientQ}
2651: \ee
2652: %
2653: where the effective angles
2654: %
2655: \begin{eqnarray}
2656: \valpha_{\mathrm{eff}} &\equiv& \frac{2}{N_p} \sum_i \cos^2(2\phi_i)
2657: \valpha_i \nonumber \\
2658: \vbeta_{\mathrm{eff}} &\equiv& \frac{2}{N_p} \sum_i \cos(2\phi_i)\sin(2\phi_i)
2659: \valpha_i .
2660: \end{eqnarray}
2661: %
2662: Repeating for the recovered $U$ Stokes parameter, we find
2663: %
2664: %
2665: \be
2666: \hat{U}(\vx_p) \approx U(\vx_p) + \vgamma_{\mathrm{eff}} \cdot \vgrad U
2667: |_{\vx_p} + \vbeta_{\mathrm{eff}} \cdot \vgrad Q |_{\vx_p}
2668: \label{eq:gradientU}
2669: \ee
2670: %
2671: where
2672: %
2673: \be
2674: \vgamma_{\mathrm{eff}} \equiv \frac{2}{N_p} \sum_i \sin^2(2\phi_i)
2675: \valpha_i .
2676: \ee
2677: %
2678: Note that for pointing errors that are not constant in time,
2679: $Q$ and $U$ are generally displaced differently and there is
2680: non-local $Q$-$U$ coupling through $\vbeta_{\mathrm{eff}}$.
2681: We recover the simple map-based model of pointing errors used
2682: in~\citet{hu03} and~\citet{odea07},
2683: %
2684: \be
2685: (\hat{Q}\pm i \hat{U})(\vx_p) \approx (Q\pm i U)(\vx_p) +
2686: \valpha \cdot \vgrad (Q\pm i U)|_{\vx_p} ,
2687: \ee
2688: %
2689: for the case of constant pointing errors, i.e.\ $\valpha_i = \valpha$.
2690:
2691: In the limit that the pointing jitter is independent between time samples,
2692: the pointing errors in the map, $\valpha_{\mathrm{eff}}$,
2693: $\vbeta_{\mathrm{eff}}$ and $\vgamma_{\mathrm{eff}}$ are independent
2694: between pixels. (Note that this will not hold if maps are made
2695: from more than one focal-plane pixel.) These pointing errors are
2696: mean zero, and have independent $x$- and $y$-components. The
2697: non-vanishing correlators are
2698: %
2699: \begin{eqnarray}
2700: \langle \valpha_{\mathrm{eff}} \cdot \valpha_{\mathrm{eff}}
2701: \rangle &\approx& 3 \alpha_{\mathrm{tod}}^2 / (2N_p) \\
2702: \langle \vbeta_{\mathrm{eff}} \cdot \vbeta_{\mathrm{eff}}
2703: \rangle &\approx& \alpha_{\mathrm{tod}}^2 / (2N_p) \\
2704: \langle \vgamma_{\mathrm{eff}} \cdot \vgamma_{\mathrm{eff}}
2705: \rangle &\approx& 3 \alpha_{\mathrm{tod}}^2 / (2N_p) \\
2706: \langle \valpha_{\mathrm{eff}} \cdot \vgamma_{\mathrm{eff}}
2707: \rangle &\approx& \alpha_{\mathrm{tod}}^2 / (2N_p) .
2708: \end{eqnarray}
2709: %
2710: Here, $\alpha_{\mathrm{tod}}^2 = \langle \valpha_i \cdot \valpha_i \rangle$
2711: is the variance of the pointing jitter, and we have approximated
2712: discrete sums over the angles $\phi_i$ by their integrals.
2713: We shall further assume that the
2714: number of hits per pixel is uniform over the map in which case
2715: the map-based pointing errors have homogeneous statistics through the map.
2716: The quantity $\alpha_{\mathrm{tod}}^2 \Omega_{\mathrm{pix}} / N_p$
2717: will arise in our final result for the spurious $B$-mode power,
2718: where $\Omega_{\mathrm{pix}}$ is the pixel area. This product is independent
2719: of pixel area since $N_p \propto \Omega_{\mathrm{pix}}$.
2720: For 30-arcsec jitter that is random between 100~Hz samples,
2721: $\alpha_{\mathrm{tod}} \sqrt{\Omega_{\mathrm{pix}}/N_p}
2722: \approx 3\times 10^{-8}\,\mathrm{rad}^2$ for an
2723: eight-hour observation of a 380-$\mathrm{deg}^2$ field.
2724:
2725: The calculation of the $B$-mode power due to leakage of
2726: $E$-modes follows from equations~(\ref{eq:gradientQ}) and
2727: (\ref{eq:gradientU}) with the standard techniques used, for example,
2728: in the calculation of gravitational lensing on the CMB
2729: (see, e.g.~\citealt{lewis06}). We shall not give the details here
2730: but simply note that the final result is a white-noise contribution
2731: with
2732: %
2733: \be
2734: C_\ell^B \approx \frac{\alpha_{\mathrm{tod}}^2\Omega_{\mathrm{pix}}}{2N_p}
2735: \int \frac{\ell^3 d\ell}{2\pi} C_\ell^E
2736: e^{-\ell^2 \sigma_B^2},
2737: \ee
2738: %
2739: where $\sigma_B$ is the beam size.
2740: The integral here approximates the mean-squared gradient of the beam-smoothed
2741: polarization which, for a $5.5$-arcmin beam
2742: $\approx 1.8\times 10^7\,\mu\mathrm{K}^2$.
2743: For our simulation parameters we therefore expect
2744: $C_\ell^B < 10^{-8}\, \mu\mathrm{K}^2$
2745: from pointing jitter which is $< 1$ per cent of
2746: the large-scale power induced by gravitational lensing.
2747:
2748: \label{lastpage}
2749:
2750: \end{document}
2751:
2752: