1: %Version of 8th November 2007 - Douglas had it last
2: %\documentclass[12pt,preprint]{aastex}
3: \documentclass[preprint2]{aastex}
4:
5:
6: %\usepackage{graphics,multicol}
7: %\usepackage{amsmath}
8: %\usepackage{multirow}
9: %\usepackage{subfigure}
10: %\usepackage{color}
11:
12:
13: \newcommand{\myemail}{patanchon@apc.univ-paris7.fr}
14: \newcommand{\ivcg}{{G86~}}
15: \newcommand{\casa}{{Cas A~}}
16:
17: \slugcomment{To appear in the Astrophysical Journal}
18:
19: \shorttitle{Map-making method for timestream data from large arrays}
20: \shortauthors{Patanchon et al.}
21:
22:
23: \begin{document}
24:
25:
26: \title{SANEPIC: A Map-Making Method for Timestream Data From Large Arrays}
27:
28:
29: %\input{author-list.tex}
30: \author{G.~Patanchon,\altaffilmark{1,2,\dag}
31: P.~A.~R.~Ade,\altaffilmark{3}
32: J.~J.~Bock,\altaffilmark{4,5}
33: E.~L.~Chapin,\altaffilmark{1}
34: M.~J.~Devlin,\altaffilmark{6}
35: S.~Dicker,\altaffilmark{6}
36: M.~Griffin,\altaffilmark{3}
37: J.~O.~Gundersen,\altaffilmark{7}
38: M.~Halpern,\altaffilmark{1}
39: P.~C.~Hargrave,\altaffilmark{3}
40: D.~H.~Hughes,\altaffilmark{8}
41: J.~Klein,\altaffilmark{6}
42: G.~Marsden,\altaffilmark{1}
43: P.~G.~Martin,\altaffilmark{9,10}
44: P.~Mauskopf,\altaffilmark{3}
45: C.~B.~Netterfield,\altaffilmark{10,11}
46: L.~Olmi,\altaffilmark{12,13}
47: E.~Pascale,\altaffilmark{11}
48: M.~Rex,\altaffilmark{6}
49: D.~Scott,\altaffilmark{1}
50: C.~Semisch,\altaffilmark{6}
51: M.~D.~P.~Truch,\altaffilmark{14}
52: C.~Tucker,\altaffilmark{3}
53: G.~S.~Tucker,\altaffilmark{14}
54: M.~P.~Viero,\altaffilmark{10}
55: D.~V.~Wiebe\altaffilmark{11}}
56:
57: \altaffiltext{1}{Department of Physics \& Astronomy, University of
58: British Columbia, 6224 Agricultural Road, Vancouver, BC V6T~1Z1,
59: Canada}
60:
61: \altaffiltext{2}{Laboratoire APC, 10, rue Alice Domon et L{\'e}onie Duquet 75205, Paris, France}
62:
63: \altaffiltext{3}{Department of Physics \& Astronomy, Cardiff University, 5 The Parade, Cardiff, CF24~3AA, UK}
64:
65: \altaffiltext{4}{Jet Propulsion Laboratory, Pasadena, CA 91109-8099}
66:
67: \altaffiltext{5}{Observational Cosmology, MS 59-33, California Institute of Technology, Pasadena, CA 91125}
68:
69: \altaffiltext{6}{Department of Physics and Astronomy, University of Pennsylvania, 209 South 33rd Street, Philadelphia, PA 19104}
70:
71: \altaffiltext{7}{Department of Physics, University of Miami, 1320 Campo Sano Drive, Carol Gables, FL 33146}
72:
73: \altaffiltext{8}{Instituto Nacional de Astrof\'isica \'Optica y Electr\'onica (INAOE), Aptdo. Postal 51 y 72000 Puebla, Mexico}
74:
75: \altaffiltext{9}{Canadian Institute for Theoretical Astrophysics, University of Toronto, 60 St. George Street, Toronto, ON M5S~3H8, Canada}
76:
77: \altaffiltext{10}{Department of Astronomy \& Astrophysics, University of Toronto, 50 St. George Street, Toronto, ON M5S~3H4, Canada}
78:
79: \altaffiltext{11}{Department of Physics, University of Toronto, 60 St. George Street, Toronto, ON M5S~1A7, Canada}
80:
81: \altaffiltext{12}{Istituto di Radioastronomia, Largo E. Fermi 5, I-50125, Firenze, Italy}
82:
83: \altaffiltext{13}{University of Puerto Rico, Rio Piedras Campus, Physics Dept., Box 23343, UPR station, San Juan, Puerto Rico}
84:
85: \altaffiltext{14}{Department of Physics, Brown University, 182 Hope Street, Providence, RI 02912}
86:
87: \altaffiltext{\dag}{\url{patanchon@apc.univ-paris7.fr}}
88:
89: \begin{abstract}
90: We describe a map-making method which we have developed for the
91: Balloon-borne Large Aperture Submillimeter Telescope (BLAST)
92: experiment, but which should have general application to data from
93: other submillimeter arrays. Our method uses a Maximum Likelihood
94: based approach, with several approximations, which allows images to
95: be constructed using large amounts of data with fairly modest
96: computer memory and processing requirements. This new approach,
97: Signal And Noise Estimation Procedure Including Correlations
98: (SANEPIC), builds upon several previous methods, but focuses
99: specifically on the regime where there is a large number of detectors
100: sampling the same map of the sky, and explicitly allowing for the
101: the possibility of strong correlations between the detector timestreams.
102: We provide real and simulated examples of how well this method performs
103: compared with more simplistic map-makers based on filtering. We discuss
104: two separate implementations of SANEPIC: a brute-force approach, in which
105: the inverse pixel-pixel covariance matrix is computed;
106: and an iterative approach, which is much more efficient for large maps.
107: SANEPIC has been successfully used to produce maps using data from the
108: 2005 BLAST flight.
109: \end{abstract}
110:
111:
112: \keywords{methods: data analysis --- techniques: image processing ---
113: submillimeter --- balloons}
114:
115: \hyphenation{Sub-milli-meter}
116: \hyphenation{sub-milli-meter}
117:
118: \section{Introduction}
119:
120: The problem of optimal ``map-making'' or ``image reconstruction'' is
121: complex and multi-faceted, with the basic procedures and even the
122: terminology differing dramatically between different sub-fields of
123: astronomy. The method adopted depends on the form in which the data
124: are gathered, and on the dominant source of systematic effects. From
125: the optical to the near-IR one talks about combining ``frames'', along
126: with measurements of ``darks'' and ``flat-fields''. For the reduction
127: of Cosmic Microwave Background (CMB) data the now conventional method
128: is to start from the principle that there is a linear algebra approach
129: to solving the Maximum Likelihood problem. However, this has only
130: been feasible up until now because of the limited number of detectors in the
131: typical CMB experiment, and the fact that correlated signals among the
132: detectors can be effectively ignored. Because of the rapid development of
133: large bolometer arrays, the question which arises is:
134: how does one adapt the CMB approach to dealing with substantial numbers of
135: detectors, and where there are significant cross-correlations of
136: noise between the detector timestreams.
137:
138: Data from the Balloon-borne Large Aperture Submillimeter Telescope
139: (BLAST, Devlin et al.~2004) represent a new challenge for bolometric
140: timestream-to-map algorithms. Recent CMB experiments which use
141: detectors similar to those used on BLAST, such as BOOMERANG
142: \citep{crill} and Archeops \citep{benoit}, only use a handful of
143: separate bolometers. Furthermore, these experiments' off-axis designs
144: lead to small correlations between detectors. Consequently, the
145: correlations could be ignored at the map-making stage, and each
146: detector timestream could be treated as an independent sub-set of the
147: data. This has changed for BLAST, which has up to 139 detectors per
148: band, with significant correlations induced by the on-axis design, as
149: well as the higher frequencies of the observations. Just by itself,
150: the large number of channels increases the impact of even small time
151: stream correlations, the contribution from which does not integrate
152: down with increasing number of detectors, unlike the uncorrelated
153: noise. The high level of correlation (largely induced by temperature
154: drifts in the obscuring secondary supports in BLAST) make it important
155: that the correlations be handled carefully in the map-making process.
156:
157: This paper describes a Signal And Noise Estimation Procedure Including
158: Correlations (SANEPIC) which has been developed for the analysis of
159: BLAST. This algorithm will also have application to many next
160: generation experiments which will involve both noise correlations
161: between channels (including correlations from the atmosphere) and very
162: large numbers of detectors. This includes the next generation of
163: larger-format arrays for use in ground- and balloon-based instruments
164: at microwave and millimeter wavelengths, which will have typically
165: thousands of detectors.
166:
167: This paper is arranged as follows. We next describe the pertinent
168: aspects of BLAST. In Section~3, the longest section, we set out our
169: basic map-making method. Simulations we use for testing the method
170: are described in Section~4, with results presented in Section~5,
171: demonstrating the benefit of accounting for the correlated noise in
172: BLAST-like observations. Finally some of the maps obtained from the
173: June 2005 BLAST flight are presented in Section~6.
174:
175:
176: \section{BLAST observations of the submillimeter sky}
177:
178: The map-making procedure presented in this paper has been used to
179: analyze the data from the BLAST 2005 flight. We will use BLAST as a
180: specific example for the application of SANEPIC throughout the paper.
181: We describe BLAST in Section~\ref{sub:BLAST}, we then summarize the
182: pre-processing of the data prior to map-making in Section
183: \ref{sub:process}. In Section~\ref{sub:model}, we derive a model of
184: the data that we will use for the map-making.
185:
186: \subsection{BLAST instrument and observations}
187: \label{sub:BLAST}
188:
189: The Balloon-borne Large-Aperture Submillimeter Telescope incorporates
190: a 2-m primary mirror, and large format bolometer
191: arrays operating at 250, 350, and 500$\,\mu$m, designed to have 144, 96
192: and 48 bolometers, respectively (of which 139, 88 and 43,
193: respectively, were used). The instrument is described in detail in
194: \cite{pascale07}. The low atmospheric opacity at the operating
195: altitude of ${\sim}\,38\,$km allows BLAST to map the sky very quickly
196: compared to ground-based experiments and to conduct large area shallow
197: surveys as well as very deep surveys of the sky \citep{devlin}. The
198: BLAST wavelengths are near the peak of the spectral energy
199: distribution of cold galactic dust, which gives BLAST the ability to
200: conduct unique extragalactic and Galactic submillimeter surveys with
201: high spatial resolution and sensitivity. BLAST thus enables studies
202: of the distribution of very high-redshift galaxies and of star forming
203: regions in our Galaxy.
204:
205: The typical observing strategy consists of scanning the telescope back
206: and forth in azimuth, covering the entire field by slowly varying the
207: elevation. Cross-linking of the data is assured by scanning the same
208: field at another time of the day. Typical scanning strategies are
209: given in \cite{pascale07}.
210:
211: The first scientific flight of BLAST took place in June 2005 from the
212: Esrange Arctic base in Sweden to the Canadian Arctic. A total of
213: $\sim$ 100 hours of data were taken in a variety of Galactic
214: fields. They include a star forming region (Vulpecula) over $4\,{\rm deg}^2$,
215: described in \cite{chapin}, three other fields of similar size
216: in the Galactic Plane (which will be the focus of future papers), an
217: integration towards the ELAIS-N1 field (see Oliver et al.~2000),
218: the \casa supernovae remnant over about $0.5\,{\rm deg}^2$
219: (Hargrave et al.~in preparation),
220: and several compact Galactic and extra-galactic sources \citep{truch07}.
221: Hereafter we refer to these as the BLAST05 data, to distinguish them from
222: the data taken during the December 2006 Antarctic flight.
223:
224: \subsection{Time-ordered data pre-processing}
225: \label{sub:process}
226:
227: The processing of BLAST data from detector timestreams to the final
228: map product involves several steps prior to map-making. Each of these
229: steps is designed to remove a particular (or several) artifact(s) from
230: the data, and sometimes requires iterating, since some effects need to
231: be removed simultaneously. In the following, we summarize the main
232: processing stages leading to the time-ordered segments which are used
233: as inputs for the map-making process.
234:
235: We start by identifying events in the data which are sharply localized
236: in time, such as spikes from cosmic rays hits and other spurious
237: sources. We use a method which allows us to discriminate between the
238: different events depending on their signature in the data. Spikes
239: which involve only a single sample are flagged and the corrupted
240: samples are replaced by the average value of the samples in the
241: vicinity. The data are deconvolved from the low-pass filter applied by
242: the readout electronics. This filter has a frequency cut off of
243: approximately $35\,$Hz and is designed to avoid high frequency noise
244: aliasing. The deconvolution is performed in Fourier space. In
245: addition, we have applied a low-pass cosine filter which limits the
246: noise power from increasing at very high frequency (above $38\,$Hz) due to
247: deconvolution. We have checked that the noise power spectrum is
248: relatively flat after these deconvolution operations.
249: Finally, cosmic ray hits
250: and other localized artifacts in the data timestreams are detected and cut
251: out. In order to avoid biasing our data products by having
252: systematically more false event detections located where the sky is
253: bright (e.g.,~when scanning a point source), we iterate this process
254: -- we make maps starting from data which has been cleaned using the
255: process described above, subtract the maps from each original data-set,
256: and reprocess the data. The maps calculated at this intermediate
257: stage are obtained by simply rebinning the data into pixels after
258: strongly high-pass filtering. The filter applied is a Butterworth
259: filter with a frequency cut off of $0.5\,$Hz, which is of the order of
260: the knee frequency of the noise. Even if this operation suppresses
261: most of the intermediate to large scales of the sky signal in the
262: maps, it does not very much alter the signal from localized sources, at
263: least to the level of accuracy needed at this stage, and it has the
264: advantage of removing most of the stripes in the maps due to $1/f$
265: noise. We have verified that the resulting bias for bright calibrators
266: is less than 1\%.
267:
268: About 2\% of the data from the BLAST05 flight was removed due to cosmic
269: ray events. Most of the events affected a single detector timestream,
270: although some events affected a whole array at the same time. Detected
271: spikes (from cosmic rays but also other spurious effects) are flagged
272: over small time intervals of typically 1~second in the data. One second is
273: too large an interval to simply ignore, so the corrupted data need to be
274: replaced with random noise generated in a way that as much as possible
275: preserves the statistical properties of the data. In the later
276: map-making stage, which is partly performed in Fourier space, we
277: assume continuity of the data. We could perform this gap-filling with
278: a constrained noise realization (see e.g., Stompor et al. 2002). However,
279: since the gap intervals are significantly smaller than the inverse of the knee
280: frequency of the noise power spectrum ($0.3\,$Hz), the noise can be well
281: approximated by the sum of a white component plus a straight line of
282: some slope across the gap. Specifically, we generate white noise in each gap
283: with a standard deviation measured from the data in the vicinity of
284: the gap, and add a baseline with the parameters fit using 20
285: samples on each side. The white noise generation is for restoring as
286: best as possible the stationarity of the data (generated samples are
287: not reprojected to the map at the end).
288:
289: After having filled the gaps in the timestreams, we filter out very low
290: frequency drifts which are poorly accounted for in the map-making procedure. A
291: fifth-order polynomial is fit to the data and removed from each data
292: segment in order to reduce fluctuations on timescales larger than the
293: length of the considered segment which, depending on the specific
294: case, varies from 30 minutes to a few hours. These fluctuations are
295: poorly described because of the limited number of Fourier modes, and
296: would cause leakage at all timescales (for instance a gradient in the
297: timestream is described by a wide range of Fourier modes), degrading
298: the efficiency of map-making. Note that we experimented with various
299: polynomials and other effective high-pass filters; we found that the
300: results were not very sensitive to precisely how this is done, but
301: some such filtering is certainly required. The degree of the polynomial
302: was chosen empirically as a compromise between suppressing the artifacts
303: and keeping the large scale signals in the final maps.
304: Using simulations,
305: we have checked that the effect on the transfer function of the signal
306: in the final map is weak. The resulting data segments are then corrected
307: for the time-varying calibration (see Truch et al.~2007) using
308: measurements of an internal calibration lamp \citep{pascale07}.
309: Finally, the data segments are apodized at the edges over ${\sim}\,2,000$
310: samples and are high-pass filtered at $5\times 10^{-3}\,$Hz with a
311: Butterworth filter. This filtering has very little effect on the final
312: maps, since modes in the data at lower frequencies that are cut in
313: this way are not expected to contribute significantly to the signals. We
314: discuss the choices of filters further in Section~\ref{sub:implem}.
315:
316: Accurate pointing reconstruction is a complicated procedure for
317: balloon-borne telescopes, and this affects the map-making task through the
318: pointing matrix (defined in Section~\ref{sub:model}).
319: The pointing reconstruction procedure is
320: described in detail in \cite{pascale07}. The next important step in
321: reducing the BLAST data, which is calibration of the detectors, is detailed in
322: \cite{truch07}.
323:
324: \subsection{Model of the data}
325: \label{sub:model}
326:
327: Having performed the cleaning procedure described in the previous
328: sub-section, the resulting data timestreams can be modeled very
329: accurately as the sum of pure signal and pure noise contributions.
330: The data for detector $i$ observing at a given wavelength and at time
331: sample $t$ can be written as
332: \begin{equation}
333: d_{it} = [A_i]_{tp} ~s_{p} + n_{it},
334: \label{eq:linmodel}
335: \end{equation}
336: where $p$ labels the pixels in the final map, $A_i$ is the pointing
337: matrix for bolometer $i$ (whose elements, indexed with time $t$ and
338: pixel $p$, give the weight of the contribution of pixel $p$ to the
339: sample at time $t$ for bolometer $i$), $s_{p}$ is the signal amplitude
340: at pixel $p$, and the noise amplitude at time $t$ for bolometer $i$ is
341: $n_{it}$. Summation over repeated indices is assumed here.
342:
343: Ideally, the element $[A_i]_{tp}$ of the pointing matrix is equal to
344: the value of the beam response $b( R (\vec{r}-\vec{r_0}))$, where
345: $\vec{r}$ points to the pixel $p$ location, $\vec{r_0}$ is the
346: location of the beam center at time $t$, and $R$ is a rotation matrix
347: which depends on the rotation angle at time $t$ between the telescope
348: and sky coordinate systems. In principle one could then recover a map
349: of the sky {\it deconvolved\/} with the instrumental Point Spread
350: Function (PSF). However, in practice this would be unacceptably
351: noisy, as well as computationally intractable, because of the
352: prohibitive volume of data. Even although $A_i$ might have mostly
353: zero elements, it is nevertheless a huge matrix. It {\it may\/} be
354: feasible to deconvolve a non-trivial beam response (e.g.~like the
355: BLAST05 beams, as shown in Truch et al.~2007) at the same time
356: performing the map-making step, through an approximate treatment of
357: the sparse pointing matrix. But we did not pursue that approach here.
358:
359: In the simple case where the beam is
360: symmetric, the map-making problem becomes tractable, provided one restricts
361: oneself to reconstructing a map of the instrument-convolved sky.
362: We can then consider $s_p$ in
363: Equation~\ref{eq:linmodel} as the map of the sky convolved with the
364: beam, and consequently $A_i$ indicates simply where the detector points in
365: the sky at a given time. In this case, $A_i$ is an ultimately sparse
366: matrix with, in the BLAST case, simply a 1 in a single entry of each
367: row. This approach, which has been conventional for CMB map-making (although
368: with some adaptation for chopped data), is what we use in the following
369: analysis. It gives no loss of information provided that the map pixels are
370: sufficiently small, and one can simply
371: assign all the flux from a bolometer's to the map pixel to
372: which it points at each time interval. The
373: requirement for accuracy is that the pixel size is
374: smaller by a factor three or more than the Full Width at Half Maximum
375: (FWHM) of the instrumental PSF; in that case the additional convolution with
376: the pixel shape gives negligible loss of angular resolution.
377:
378: The noise term $n_{it}$ in the model represents the sum of all
379: contributions to the timestreams which do not reproject on the sky.
380: This will in general include
381: instrumental noise, fluctuations in atmospheric emission
382: and other loading and unrecognized cosmic ray hits. In general, some
383: of those noise contributions will induce strong correlations
384: between detector timestreams. In this paper, we adopt a very general
385: model of the noise where the noise covariance matrix:
386: \begin{equation}
387: N_{ii'tt'} = \left\langle n_{it}\cdot n_{i't'}^t\right\rangle,
388: \label{eq:modfullN}
389: \end{equation}
390: (for bolometer indices $i$, $i'$ and time indices $t$, $t'$) has possibly
391: non-zero elements even for $i \neq i'$. A key assumption, as we will
392: see later, is that the noise is Gaussian (so that $N_{ii'tt'}$ is
393: sufficient to describe all the statistical properties of the noise)
394: and stationary (constraining $N_{ii'tt'}$, see Sections~\ref{sub:NN}
395: and \ref{sub:detdetcorr}).
396:
397: In the specific case of BLAST05 observations, a very significant
398: correlation of the noise is found in the timestreams, and we have
399: shown that a more constraining model provides a very good description
400: of the data. An independent component analysis \citep{delabrouille03}
401: of the data enabled us to find that the noise and its correlations can
402: be described to a high degree of accuracy by a noise component which
403: is {\it not\/} correlated between detectors, together with a single
404: {\it common-mode\/} component seen by all the detectors at a given
405: wavelength (some correlations is also seen between detectors from
406: different wavelength but we have chosen to treat each wavelength
407: independently). The common part of the noise is instantaneous, meaning
408: that the same common-mode noise is seen at the same time by all the
409: detectors. In our model, the noise term in Equation~\ref{eq:linmodel}
410: is then decomposed as:
411: \begin{equation}
412: n_{it} = \tilde{n}_{it} + \alpha_i c_{t},
413: \label{eq:lincmod}
414: \end{equation}
415: where the first term is the noise which is uncorrelated between
416: detectors and the second term represents the common-mode component of
417: the noise, rescaled by an amplitude parameter $\alpha_i$ which depends
418: on the detector but not on time. This model can be generalized easily
419: to deal with multiple noise components in timestreams.
420:
421: In the following section, we present a method to reconstruct $s_p$
422: given the data, in the framework of the linear model
423: (Equation~\ref{eq:linmodel})
424: and in the presence of correlated noise between detector timestreams.
425:
426:
427: \section{Map-making method}
428:
429: \subsection{Maximum Likelihood map-making}
430:
431: The use of Maximum Likelihood map-making techniques has been developed
432: by many authors for application to Cosmic Microwave Background data-sets
433: \citep{wright,tegmark,borrill,prunet00,tegmark00,ferreira,dore,natoli,prunet01,dupac02,stompor,hinshaw,yvon}. Some
434: other approach are more specific to destriping for {\sl Planck}-like
435: scanning strategies
436: \citep{delabrouille98,maino02,keihanen,degasperis,macias,poutanen,ashdown07}.
437: Since there is already a large number of publications on the topic, here we
438: present only a very brief overview of the approach of Maximum
439: Likelihood map-making techniques.
440:
441: Assuming the simple linear model given by
442: Equation~\ref{eq:linmodel}, the log-likelihood of the data can be
443: calculated under the assumption that the noise is Gaussian and
444: stationary. The solution is
445: \begin{equation}
446: \log L(d|s) = - \frac12 (d - As)^t N^{-1} (d - As),
447: \end{equation}
448: where $N \equiv \left\langle n.n^t\right\rangle$ is the noise
449: covariance matrix in the time domain, and $.^t$ denotes transpose.
450: Maximizing the above equation with respect to the map parameters $s$
451: (suppressing the pixel indices here for convenience) leads to the
452: following well known estimator:
453: \begin{equation}
454: \hat{s} = (A^tN^{-1}A)^{-1} A^tN^{-1} d.
455: \label{eq:solmap}
456: \end{equation}
457: The inverse pixel-pixel covariance matrix of the noise in the map is the
458: term in brackets in this equation, i.e.
459: \begin{equation}
460: N_{pp'}^{-1} = A^tN^{-1}A.
461: \label{eq:InvNpix}
462: \end{equation}
463: Computation of the solution to Equation~\ref{eq:solmap} is far from
464: trivial for most astronomical applications, due to the large amount of
465: data, and hence this poses a difficult numerical challenge. The noise
466: covariance matrix $N$ is a very large matrix of size the number of
467: samples squared, which could easily be millions, while $N_{pp'}$ may be
468: more reasonable in size but has no obvious symmetries, and so is still
469: difficult to invert. Nevertheless, we have implemented a method aimed
470: at finding the Maximum Likelihood solution given by
471: Equation~\ref{eq:solmap} when there are a large number of detectors
472: and in the presence of possible correlations in the noise between
473: different detector timestreams.
474:
475: \subsection{Implementation}
476: \label{sub:implem}
477:
478: In the simple case of dealing only with independent noise between
479: detectors, our matrix-inversion method is very similar to the MADCAP
480: method, described in \cite{stompor}. However, we have developed our new
481: approach to deal efficiently with multi-detector data in the presence
482: of correlated noise between detectors (described in detail in
483: Section~\ref{sub:detdetcorr}). In this section, we summarize the
484: basic ideas for the simpler 1-detector 1-scan case.
485:
486: In order to find the Maximum Likelihood solution of the map
487: (Equation~\ref{eq:solmap}), we have developed two different
488: algorithms. They both allow us to solve the linear system
489: $N_{pp'} {\hat s} = x$, with $x \equiv A^tN^{-1} d$. The first approach
490: explicitly computes the inverse pixel-pixel covariance matrix $N_{pp'}^{-1}$,
491: and we refer to this as the `brute-force algorithm'. The second approach
492: uses iterations which converge to the Maximum Likelihood map without
493: the need for computing $N_{pp'}^{-1}$, and we refer to this as the `iterative
494: algorithm'. Both approaches require as a first step
495: the computation of the inverse of the time-time noise covariance matrix $N$.
496:
497: \subsubsection{Inverse noise covariance matrix $N_{tt'}^{-1}$}
498: \label{sub:NN}
499:
500: In practice, even when we have knowledge of the statistical properties
501: of the data as described by the power spectrum $P(\omega)$, the
502: brute-force inversion of $N$ is not tractable because of its enormous
503: size -- for a single BLAST detector observing for 10 hours at a
504: data-rate of $100\,$Hz, the matrix has approximately $10^{13}$
505: elements. However, if we make the approximation that each data
506: segment is ``circulant'', meaning that the beginning and the end of a
507: segment are connected without discontinuity and that there are no gaps
508: in the data, then the matrix $N$ is also circulant (see
509: Section~\ref{sub:gaps} for a description of how we treat gaps in the
510: data). Circulant matrices are much easier to store and to invert.
511: With this approximation the matrix can be written
512: \begin{equation}
513: N_{tt'} = C(|t-t'|),
514: \end{equation}
515: where the correlation function $C(|t-t'|)$ between samples $t$ and
516: $t'$ depends only on the separation between the two samples. A
517: circulant matrix has the property of having a diagonal matrix
518: counterpart in Fourier space.
519:
520: Let $F$ be the discrete Fourier operator, we have
521: \begin{equation}
522: N = F^\dagger \Lambda F,
523: \end{equation}
524: where $.^\dagger$ denotes transpose conjugation, and $\Lambda$ is a
525: diagonal matrix whose diagonal is described by the power spectrum of
526: the data segment,
527: \begin{equation}
528: \Lambda_{\omega\omega} = P(\omega).
529: \end{equation}
530: The inverse of the noise covariance matrix is
531: \begin{equation}
532: N^{-1} = F^\dagger \Lambda^{-1} F,
533: \end{equation}
534: and because $\Lambda^{-1}$ is a diagonal matrix, $N^{-1}$ is also
535: circulant, so that knowledge of only one row is enough to describe the
536: entire matrix. Then the inverse covariance matrix can be written
537: \begin{equation}
538: [N^{-1}]_{tt'} = \bar{C}(|t-t'|),
539: \end{equation}
540: with
541: \begin{equation}
542: \bar{C}(\Delta t) = {\cal{F}}^{-1}\left\{{1 \over P(\omega)}\right\}(\Delta t),
543: \label{eq:PktobC}
544: \end{equation}
545: where ${\cal{F}}^{-1}$ represents the inverse Fourier transform. The
546: inverse of the covariance matrix can then be computed directly using
547: the power spectrum of the data. This is a very fast operation
548: ($O(n_{\rm s}\log n_{\rm s})$, with $n_{\rm s}$ the number of samples),
549: and does not require large memory since only one row of the matrix
550: is computed.
551:
552: The approximation that the data segments are ring-shaped or circulant
553: might seem unreasonable, but in the end this only has an effect on a
554: small fraction of the matrix. For data with large-scale correlations
555: (data described by a $1/f$ power spectrum, for instance), the
556: approximation implies an assumption that the two edges of the segment are
557: very correlated. In reality, there is obviously little correlation at
558: long timescales compared to short timescales, so extra striping
559: could be introduced in the maps if one steps across the two edges of the data
560: segment. We have addressed this problem in two ways in order to
561: avoid introducing artifacts in the final map. In the case where we explicitly
562: compute the inverse pixel-pixel covariance matrix $N_{pp'}^{-1}$ (as in
563: the brute-force inversion algorithm), for the estimation of $N$ entering
564: in the computation of $N^{-1}_{pp'}$, we constrain $\bar{C}(\Delta t) =
565: 0$ for $\Delta t > n_{\rm s}/2$, with $n_{\rm s}$ being the number of
566: samples in the segment; this is sometimes known as ``the MADCAP
567: approximation'' in the literature. It is important to note that
568: this approximation cannot be used for
569: the computation of $A^tN^{-1}d$, because it is performed partly in
570: Fourier space (see Section~~\ref{subsub:ANd}). Instead we have apodized
571: the data $d$ at the edges, and we have removed a low-order polynomial
572: (5th order in practice) to reduce fluctuations having typical
573: timescales of order (or larger than) the data segment (see Section
574: \ref{sub:process}). This is reasonable, since those scales are not
575: well described by a limited number of Fourier modes, and hence will
576: always be hard to reconstruct.
577:
578: \subsubsection{Inverse pixel-pixel covariance matrix $N_{pp'}^{-1}$}
579: \label{subsub:InvNpp}
580:
581: The computation of the inverse pixel-pixel covariance matrix, which is
582: described in this section, is required only in the brute-force inversion
583: algorithm, or for an accurate error estimation in the pixel domain.
584:
585: Since the pointing matrix has only one non-zero element per row in our
586: simple model (one data sample is associated with a single map pixel),
587: the matrix multiplication $A^tN^{-1}A$ requires a single loop going
588: across all the non-zero elements of $N^{-1}$. For most cases,
589: a dominant faction of the map-making computing time will be devoted to this
590: operation. If the data are only correlated within a typical length
591: $\lambda_{\rm c}$, we have the property: $\bar{C}(\Delta t)\,{\simeq}\,0$
592: for $\Delta t\,{>}\,\lambda_{\rm c}$, and $N^{-1}$ is a band-diagonal
593: matrix (elements separated from the diagonal by more than
594: $\lambda_{\rm c}$ are negligible). The number of elements to go
595: through in the loop is of the order of, but smaller than $2n_{\rm
596: s}\lambda_{\rm c}$, which is hopefully much smaller than the size of
597: the matrix itself.
598:
599: Unfortunately, if the noise is described by a power spectrum
600: of the form $(1/f)^\beta$, the correlation length of the noise is
601: basically of the order of the length of the whole data-set. However,
602: the amplitude of the correlation is decreasing for very long
603: timescales and becomes negligible beyond a certain scale. The
604: function $\bar{C}(\Delta t)$ can then be artificially set to zero
605: for $\Delta t > \lambda'_{\rm c}$, with $\lambda'_{\rm c}$ chosen
606: empirically so that the correlation of the noise is low enough for
607: scales longer than $\lambda'_{\rm c}$, and also that there is very
608: little constraint on the signal at those scales. For the specific
609: case of BLAST observations, we find that $\lambda'_{\rm c}\simeq
610: 200\,$s is a good compromise, as illustrated in Figure~\ref{fig:invc}.
611: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
612: %\clearpage
613: \begin{figure*}[t!]
614: \begin{center}
615: \includegraphics[width=\columnwidth]{f1a.eps}
616: \includegraphics[width=\columnwidth]{f1b.eps}
617: \end{center}
618: \caption{Left panel: Absolute value of the first row of the inverse
619: covariance matrix given by $\bar{C}(\Delta t)$ (see
620: Section~\ref{sub:NN}) for a typical BLAST observation. The
621: vertical line indicates the value of $\lambda'_{\rm c}$, such that
622: for $\Delta t > \lambda'_{\rm c}$, $\bar{C}(\Delta t)$ is set to
623: zero for the computation of $N^{-1}$. Right panel: Power spectrum
624: of the noise in a typical BLAST observation (dotted curve), and
625: after thresholding at frequencies smaller than $5\times
626: 10^{-3}\,$Hz, i.e.,~1/$\lambda'_{\rm c}$ (solid curve). The
627: dot-dashed curve is obtained by inverting Equation~\ref{eq:PktobC}
628: after forcing the values of $\bar{C}(\Delta t)$ to zero for
629: $\Delta t > \lambda'_{\rm c}$, with $\bar{C}$ obtained from the
630: initial power spectrum of the noise (dotted curve). To find the
631: power spectrum which corresponds to the dashed curve the same
632: procedure is applied, but starting from the thresholded power
633: spectrum (solid curve). All the curves are very similar for
634: frequencies larger than 1/$\lambda'_{\rm c}$, but the dot-dashed
635: curve begins to diverge at smaller frequencies, while the dashed
636: and solid curves lie very close to each other for all frequencies.
637: This shows the tight relation between the power spectrum and the
638: inverse covariance matrix. The noticeable peak in the power
639: spectrum is located at the scanning frequency of about $4\times
640: 10^{-2}\,$Hz.}
641: \label{fig:invc}
642: \end{figure*}
643: %\clearpage
644: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
645:
646: The noise strongly dominates the signal in BLAST observations for
647: scales longer than $\lambda'_{\rm c}$ (corresponding to frequencies
648: smaller than $5\times 10^{-3}\,$Hz), since: (1) this frequency
649: is well below the knee frequency of the noise
650: power spectrum; and (2) there is very little signal at frequencies
651: smaller than $5\times 10^{-3}\,$Hz, which is more than ten times
652: smaller than the scanning frequency. Therefore we have used
653: $\lambda'_{\rm c} = 200\,$s for the computation of the noise
654: covariance matrix. The impact of fixing $\bar{C}(\Delta t) = 0$ for
655: $\Delta t > \lambda'_{\rm c}$ in the initial power spectrum is shown
656: in Figure~\ref{fig:invc}. We can see a tight relation between the
657: power spectrum and the inverse covariance matrix, since getting
658: $\bar{C}(\Delta t) = 0$ for $\Delta t > \lambda'_{\rm c}$ can be
659: obtained only by modifying the power spectrum at frequencies smaller that
660: 1/$\lambda'_{\rm c}$.
661:
662: \subsubsection{Computation of $A^t N^{-1}d$}
663: \label{subsub:ANd}
664:
665: In the computation of $x = A^t N^{-1}d$, which is necessary for both of our
666: algorithms, the multiplication $N^{-1}d$ is performed in Fourier space
667: where the noise covariance matrix is diagonal. We obtain this vector
668: by dividing the Fourier transform of the data by the power spectrum of
669: the noise. Another way to represent this is by considering that since
670: $N^{-1}$ is circulant, $N^{-1}d$ is a convolution operation. Assuming
671: that this model holds, the resulting data vector $\tilde{d}
672: = N^{-1}d$ contains whitened noise. The remaining operation $A^t \tilde{d}$
673: just performs the addition of the filtered data sample onto the pixels of
674: $s$ (i.e.~the map), and hence is very fast.
675:
676: In the case of BLAST observations, since we are not attempting to recover
677: useful information from timescales larger than $200\,$s, we
678: perform a high-pass pre-filtering of the input data $d$ at
679: $5\times 10^{-3}\,$Hz.
680:
681:
682: \subsubsection{Matrix inversion algorithm}
683:
684: In the matrix inversion algorithm, $N_{pp'}^{-1}$ is directly computed,
685: as described in Section~\ref{subsub:InvNpp}. The next step is to solve
686: the linear system ${\hat s} = [N_{pp'}^{-1}]^{-1}x$.
687: For small maps, in which $N_{pp'}^{-1}$
688: can be stored in memory, we perform a Cholesky decomposition. For
689: larger maps, we write the matrix to disk and perform an iterative
690: inversion of the system using a conjugate gradient method with
691: pre-conditioner.
692:
693: This algorithm allows one to easily perform multiple Monte-Carlo simulations,
694: since the pixel-pixel covariance matrix, being independent on the data
695: realization, can be computed once (this is assuming that the noise
696: power spectrum is known and does not have to be estimated from the data
697: themselves),
698: and used for all the realizations of simulated data. Another advantage
699: of this approach is that it allows for the exact computation of the
700: errors in the map, which are given by the covariance matrix.
701:
702: However, the matrix inversion algorithm is generally slower than the
703: iterative algorithm which we will discuss next,
704: and can be used only for relatively small maps; we found that we were
705: limited to around 200{,}000 pixels if the matrix
706: is written to disk, or less than 20{,}000 pixels if the matrix is
707: stored in memory.
708:
709:
710: \subsubsection{Iterative algorithm}
711: \label{subsub:IterAlgo}
712:
713: We now present an iterative algorithm based on conjugate gradient with
714: pre-conditioner to obtain the Maximum Likelihood solution for the map
715: (a similar algorithm has been used in \citet{ashdown}). Let us
716: rewrite Equation~\ref{eq:solmap}, which relates the best estimate of
717: the map with the data, after multiplying both sides by the pixel-pixel
718: covariance matrix:
719: \begin{equation}
720: A^tN^{-1}A~\hat{s} = A^tN^{-1} d.
721: \end{equation}
722: If we define $\hat{s}_k$ as an estimate of the map at iteration $k$,
723: the conjugate gradient method allows to solve the linear system by
724: minimizing iteratively the following criterion:
725: \begin{equation}
726: \Psi = r^t N_{pp'}^{-1}r,
727: \end{equation}
728: where
729: \begin{equation}
730: r \equiv (A^tN^{-1}A~\hat{s}_k - A^tN^{-1} d).
731: \end{equation}
732: This criterion is indeed minimum and equal to zero if $\hat{s}_k$ is
733: the Maximum Likelihood solution.
734:
735: One can interpret $\Psi$ as the weighted variance of the difference
736: between two map vectors. The first of these vectors, $A^tN^{-1}A~
737: \hat{s}_k$, is the inverse pixel-pixel covariance matrix times the
738: current estimate of the map, while the second vector, $A^tN^{-1} d$,
739: is a map constructed by simply co-adding pre-whitened data. We decide
740: that convergence is reached when the quantity $r^tr$ (which is much
741: easier to compute than $\Psi$, and also converges to zero) gets
742: smaller than a predefined value. In practice the number of iterations
743: required for convergence is of the order of 100.
744:
745: The conjugate gradient method is not described here, since it
746: is a fairly standard numerical tool, and the interested reader can find
747: many descriptions in the literature.\footnote{\scriptsize{e.g.~the 1994 article by
748: J.R. Shewchuk: {\tt http://www.cs.cmu.edu/ \~{}quake-papers/painless-conjugate-gradient.pdf}}}
749: Instead of describing the details, we focus here on aspects which are
750: specific to our map-making process. In particular,
751: let us describe the computation of $A^tN^{-1}A~\hat{s}_k$, which is the
752: time-consuming part of the optimization and has to be performed at
753: each iteration (the computation of $A^tN^{-1} d$, also time consuming,
754: but needs to be done only once, since none of the parameters are changing
755: through the iterations); the other operations for updating the map at each
756: iteration are significantly faster.
757:
758: One advantage of this iterative algorithm is that the computation of the full
759: pixel-pixel covariance matrix is not required, and the operation
760: $A^tN^{-1}A~\hat{s}_k$ can be done step by step. Indeed, we start by
761: computing $\hat{d} = A \hat{s}_k$, which is an estimate of a ``signal''
762: timestream. This operation is equivalent to scanning over the current
763: estimate of the map using the pointing solution. The subsequent
764: operation $A^tN^{-1} \hat{d}$ (which should now be familiar), is carried out
765: in Fourier space,
766: as described in Section~\ref{subsub:ANd} (without applying any extra filtering),
767: and in Section~\ref{sub:detdetcorr} for the case of correlated noise between
768: detectors.
769:
770: This iterative approach is in general much faster than the brute-force
771: inversion approach, because the most time-consuming operations are performed in
772: Fourier space. It also requires less memory, since $N_{pp'}^{-1}$ is not
773: explicitly computed. Of course, if there are found to be (or known to be)
774: non-trivial correlations in $N_{pp'}^{-1}$, then it may have to be
775: calculated explicitly, hence requiring the brute-force approach. However,
776: provided the pixel-pixel correlations only involve relatively few pixels,
777: it should be possible to calculate a restricted part of (or perhaps
778: an approximation for) $N_{pp'}^{-1}$ in a modified iterative approach.
779: A related concept is discussed in Section~\ref{sub:errorestim}.
780:
781:
782: \subsection{Multi-scan, multi-detector case}
783:
784: In the previous sub-section, we presented the general method for the
785: simple case where only a single continuous observation is considered.
786: We now describe how we combine observations from different detectors
787: at the same wavelength, as well as different data segments obtained
788: over different ``visits'' during the flight, where by ``visit'' we mean a
789: period in the data which starts after a sufficiently long gap, or
790: after the observation of a different region of the sky.
791:
792: For convenience, the data vector $d$ in our model
793: (Equation~\ref{eq:linmodel}) now contains all the individual data
794: segments from different detectors, and also within a single channel,
795: concatenated end to end. The noise vector $n$ in
796: Equation~\ref{eq:linmodel} is defined in a similar manner. The matrix
797: $A$ in Equation~\ref{eq:linmodel} is then the result of stacking
798: individual pointing matrices. The Maximum Likelihood solution is also
799: written as in Equation~\ref{eq:solmap}, with $N$ becoming the full
800: covariance matrix of the noise, including all the channels and data
801: chunks. To start with, let us assume that there is no correlation of
802: the noise between data segments. This is a very good assumption if we
803: consider data segments obtained over different visits, but is
804: certainly not a good approximation for segments obtained
805: simultaneously with different channels, since we found a very strong
806: common-mode noise between detectors. We will consider the simple
807: no-correlation case first, and then in the next sub-section
808: (Section~\ref{sub:detdetcorr}) we will generalize the map-making
809: method to account for noise correlations from different detectors.
810:
811: In the absence of correlations between data segments, the time-time noise
812: covariance matrix $N$ is block-diagonal and each block can be inverted
813: separately. Defining $N_\ell$ as the sub-covariance matrix for the data in
814: segment number $\ell$, and $A_\ell$ as the sub-pointing matrix going from the
815: map to the data segment $\ell$, the inverse pixel-pixel covariance matrix
816: can be written as
817: \begin{equation}
818: N_{pp'}^{-1} = \sum_\ell A_\ell^t N_\ell^{-1} A_\ell.
819: \label{eq:combscan}
820: \end{equation}
821: The computation time for obtaining this matrix is proportional to the
822: number of data segments. In this simple case the computation of $x =
823: A^tN^{-1}d$ can be written
824: \begin{equation}
825: x = \sum_\ell A_\ell^t N_\ell^{-1} d_\ell,
826: \end{equation}
827: where the computation of each term is fast and can be performed partly
828: in Fourier space.
829:
830: \subsection{Detector-detector correlated noise}
831: \label{sub:detdetcorr}
832:
833: We now allow for the presence of correlations in the noise between
834: different detectors. In the case of multiple visits, the noise
835: covariance matrix, $N$, still has null cross-terms for samples from
836: two different data visits. Therefore, if the data vector is sorted by
837: visit, then $N$ is block diagonal and each block contains the
838: correlation coefficients between all the detectors for the samples
839: within the time interval defined as a single visit. Each visit can be
840: treated independently, since the sub-matrices can be inverted
841: separately, and Equation~\ref{eq:combscan} is still valid, but in this
842: case $\ell$ is the label for the blocks in $N$. In the following, we
843: therefore focus on a single visit, and consider observations by all
844: the detectors; the generalization to multiple visits should be clear.
845:
846: To simplify the notation, let $N$ denote the noise covariance matrix
847: for the visit being considered, with $d$ and $n$ being the data and
848: noise vectors, respectively, containing the timestream segments for
849: all the detectors put end to end, and $N_{ij}$ being a block of $N$ of
850: size $n_{\rm s} \times n_{\rm s}$, corresponding to the noise
851: correlations between detectors $i$ and $j$. Let us define ${\bar{F}}$
852: the multi-channel Fourier transform operator such that
853: \begin{equation}
854: {\tilde{n}} = {\bar{F}}n,
855: \end{equation}
856: with ${\tilde{n}}$ containing end-to-end Fourier transforms of
857: each data segment. ${\bar{F}}$ is a block-diagonal matrix, and each
858: block is the Fourier transform operator $F$ for one data segment.
859:
860: In Fourier space, the noise covariance matrix $R$ can be written
861: \begin{equation}
862: R = {\bar{F}} N {\bar{F}}^\dagger.
863: \label{eq:FourCov}
864: \end{equation}
865: If we consider a single block of the noise covariance matrix for
866: detectors $i$ and $j$, we obtain:
867: \begin{equation}
868: R_{ij} = F N_{ij} F^\dagger.
869: \end{equation}
870: Under the assumption that the data are stationary and continuous at
871: the edges (see Section~\ref{sub:implem} for a discussion), $N_{ij}$ is
872: a circulant matrix, since each element $[N_{ij}]_{tt'}$ depends only
873: on the time interval $|t-t'|$. $R_{ij}$ is then a diagonal matrix
874: with the diagonal given by the cross-power spectrum of the noise
875: between detectors $i$ and $j$:
876: \begin{eqnarray}
877: \nonumber
878: [R_{ij}]_{\omega\omega'} & = & P_{ij}(\omega) ~~~~{\rm if}~(\omega = \omega')\\
879: & = & 0 ~~~~~~~~~~~{\rm {otherwise}.}
880: \label{eq:CovRP}
881: \end{eqnarray}
882: Here $P(\omega)$ is the noise covariance matrix of size $n_{\rm d}
883: \times n_{\rm d}$ for a given mode $\omega$, where $n_{\rm d}$ is the
884: total number of detectors. The computation of the inverse of $R$ is
885: straightforward, since each Fourier mode can be treated independently.
886: If $P^{-1}(\omega)$ is the inverse noise covariance matrix for mode
887: $\omega$, the same relation as in Equation~\ref{eq:CovRP} applies
888: between $R^{-1}$ and all $P^{-1}$.
889:
890: From Equation~\ref{eq:FourCov}, we can calculate the inverse
891: covariance matrix of the noise in real space:
892: \begin{equation}
893: N^{-1} = {\bar{F}}^\dagger R^{-1} {\bar{F}}.
894: \end{equation}
895: Then, a block of $N^{-1}$ between detectors $i$ and $j$ can be written
896: \begin{equation}
897: [N^{-1}]_{ij} = F^\dagger [R^{-1}]_{ij} F.
898: \end{equation}
899: Because $[R^{-1}]_{ij}$ is a diagonal matrix (as discussed previously)
900: $[N^{-1}]_{ij}$ is circulant, and is related to the inverse of the
901: matrix containing the cross- and auto-power spectra of the noise:
902: \begin{equation}
903: [N^{-1}]_{ijtt'} = {\cal F}^{-1}\left\{ [P^{-1}]_{ij}\right\}(t' - t).
904: \label{eq:InvNCorr}
905: \end{equation}
906: From this relation, we can see that in practice $N^{-1}$ is relatively
907: easy to construct, since each of its blocks (referring to each pair of
908: detectors) is a circulant matrix, so only a row of each sub-matrix
909: needs to be calculated using the Fast Fourier Transform. Finally, in the
910: case the noise covariance matrix is used for multiplication in real space
911: (as in the brute-force algorithm), the same approximation
912: described in Section~\ref{sub:implem} is performed on each block of
913: $N^{-1}$, i.e., $[N^{-1}]_{ijtt'}=0$ for $|t - t'|>\lambda'_{\rm c}$
914: (or for $|t - t'|>n_{\rm s}/2$ if $\lambda'_{\rm c}>n_{\rm s}/2$). The
915: global structure of the final inverse noise covariance matrix is
916: illustrated in Figure~\ref{fig:mapInvN}.
917: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
918: %\clearpage
919: \begin{figure}[t!]
920: \begin{center}
921: \includegraphics[width=\columnwidth]{f2.eps}
922: \end{center}
923: \caption{Inverse noise covariance matrix ($N^{-1}$ in the text) for
924: three detectors and for only 10 minutes of data (corresponding to
925: 60{,}000 samples per detector). The matrix is computed following
926: Equation~\ref{eq:InvNCorr}, and the approximations described in
927: Section~\ref{sub:detdetcorr} are applied. The cross- and
928: auto-power spectra of the noise, used for the calculation of
929: $N^{-1}$, are computed from the data themselves (one of the
930: auto-power spectra is shown in Figure~\ref{fig:noiseSP}, the
931: cross-power spectra are an unbiased measure of the common-mode
932: signal shown in the same figure). Each sub-matrix is circulant
933: and corresponds to a particular pair of detectors. One can see
934: that the off-diagonal sub-matrices have amplitudes of the same
935: order as that of the diagonal sub-matrices. This is due to the
936: very high level of noise correlation between detectors.}
937: \label{fig:mapInvN}
938: \end{figure}
939: %\clearpage
940: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
941:
942: In our model of the data (Equation~\ref{eq:lincmod}) in which all the
943: correlations between detectors are described by a single common-mode,
944: $P^{-1}(\omega)$ can be related to the power spectra of the
945: common-mode and of the uncorrelated part of the noise between
946: detectors:
947: \begin{equation}
948: P^{-1}(\omega) = \left(\alpha \left\langle c(\omega)^\dagger
949: c(\omega)\right\rangle \alpha^t
950: + \left\langle{\tilde{n}}(\omega)^\dagger{\tilde{n}}(\omega)\right\rangle
951: \right)^{-1}.
952: \label{eq:modSpCorr}
953: \end{equation}
954: This relation can be generalized if multiple common-mode
955: components are present: $\alpha$ would then be a mixing matrix and
956: $\left\langle c(\omega)^\dagger c(\omega)\right\rangle$ becomes the covariance
957: matrix of these noise components.
958:
959: Having computed the inverse noise covariance matrix of the
960: timestreams, we can express (using Equation~\ref{eq:InvNpix})
961: the noise covariance matrix in the pixel domain:
962: \begin{equation}
963: N_{pp'}^{-1} = \sum_{ij} A_i^t~[N^{-1}]_{ij}~A_j,
964: \end{equation}
965: where, as before, $i$ and $j$ label the detectors. The computation
966: time for $N_{pp'}^{-1}$ is now proportional to the number of detectors
967: squared. The calculation of $x=A^tN^{-1}d$ is straightforward:
968: \begin{equation}
969: x = \sum_{ij} A_i^t~[N^{-1}]_{ij}~d_j,
970: \end{equation}
971: and this is fast, since (as already shown) the operation
972: $[N^{-1}]_{ij} d_j$ is a convolution, which can be performed in Fourier
973: space. One can see from Equation~\ref{eq:InvNCorr} that $x$ can be expressed
974: directly with respect to the cross- and auto-power spectra of the
975: noise:
976: \begin{equation}
977: x = \sum_{ij} A_i^t~{\cal F}^{-1}\left\{ [P^{-1}]_{ij}(\omega) .
978: \bar{d}_j(w)\right\}.
979: \end{equation}
980:
981: The formalism presented above can also be generalized easily to deal
982: with detectors operating at different wavelenths. The map vector $s$
983: could be merging different maps at different wavelengths and the noise
984: matrix $N$ would account for all the correlations of the noise between
985: detectors. The joint multi-band map-making would be suitable in
986: practice when some contaminations from thermal fluctuations in the
987: instrument or atmospheric emission are present, because they correlate
988: the noise at all wavelengths.
989:
990: \subsection{Noise power spectrum}
991: \label{sub:NSP}
992:
993: The Maximum Likelihood solution for the final map depends on the
994: noise power spectra for each data-set (through $N$ in
995: Equation~\ref{eq:solmap}), which are assumed to be perfectly known.
996: However, in practice the noise power spectra have to be inferred from
997: the data themselves and some uncertainties are associated with this
998: iterative process.
999:
1000: In practice a first (approximate) estimate of the noise power spectrum
1001: can be obtained by rebinning the power spectra of each data segment,
1002: neglecting the contribution of the astrophysical signal in the
1003: timelines. Indeed, for most of the fields observed with BLAST, and in
1004: particular for blank extragalactic fields (like the ELAIS-N1 field in
1005: BLAST's flight from Sweden), the noise is highly dominant over the sky
1006: signal at all frequencies. However, this is not true for measurements
1007: of bright regions in the Galactic Plane. Therefore, for the first
1008: iteration's noise estimate we focus on the data taken for about 6 hours
1009: while scanning the deepest extragalactic field (ELAIS-N1) and use this to
1010: estimate the noise power spectra, which then become the noise input
1011: for making the first set of maps of {\it all\/} of our fields. This
1012: approach is not entirely satisfactory, since the noise is not
1013: stationary during the flight -- the noise power spectra are seen to
1014: vary over long timescales, although they are quite constant within a
1015: single visit of each field. This non-stationarity appears most
1016: obvious when there are variations of the scanning strategy between
1017: different visits (a change of the scanning frequency induces a
1018: variation of the location of peaks in the power spectrum), but can
1019: also occur due to variations of detector loading or the detector bias
1020: being changed during the flight.
1021:
1022: Because of the observed non-stationarity of the noise, we would like
1023: an independent estimate of the noise power spectra for each visit for
1024: each of the observed fields. After starting from a first estimate of
1025: the noise power spectra based on ELAIS-N1 data as described above, we adopt
1026: an iterative approach between maps and noise power spectra. At each
1027: iteration, the estimated maps are subtracted from the data, prior to
1028: noise power spectrum estimation. We can summarize our procedure in the
1029: following steps:
1030: \begin{itemize}
1031: \item Estimate $P_0(\omega)$ from $d$ using a field known to have
1032: little signal;
1033: \item Compute ${\hat s}$ from Equation~\ref{eq:solmap} (and also
1034: Equation~\ref{eq:InvNCorr} or \ref{eq:PktobC}, depending on the noise
1035: correlation being considered) using $P_0(\omega)$ as input;
1036: \item Estimate $P(\omega)$ from $d - A{\hat s}$;
1037: \item Re-estimate ${\hat s}$ from Equation~\ref{eq:solmap} using
1038: $P(\omega)$, and iterate on these last two steps until convergence
1039: is achieved.
1040: \end{itemize}
1041: We stop iterating when the noise power spectra do not vary by more
1042: than 1 per thousand from iteration to iteration (and find that in
1043: practice only 3 to 6 iterations are necessary to reach convergence).
1044:
1045: We now focus on how we estimate the noise power spectra $P(\omega)$ in
1046: steps 1 and 3. For the simplest case, where no correlations are
1047: assumed between detectors, we simply compute a bin-averaged power
1048: spectrum for each data segment:
1049: \begin{equation}
1050: P_\ell(q) = {1 \over n_q}\sum_{\omega_{{\rm min}(q)}}^{\omega_{{\rm max}(q)}}
1051: {\tilde d}_{\ell\omega}^* {\tilde d}_{\ell\omega},
1052: \label{eq:binavP}
1053: \end{equation}
1054: where, for bin number $q$, $~n_q=\omega_{\rm max}(q) - \omega_{\rm
1055: min}(q) +1$, $\ell$ labels the data segment, and ${\tilde d}$ is the
1056: data vector from which an estimate of the map has already been subtracted.
1057: We have
1058: chosen logarithmic spacing between bins and an estimate of $P(\omega)$
1059: for each $\omega$ mode is obtained by logarithmic interpolation, which
1060: leads to a smooth power spectrum estimate.
1061:
1062: In the more complicated case where correlations between detectors are
1063: assumed to be important and therefore are not neglected, we must
1064: estimate, for every iteration, each cross- and auto-power spectrum of
1065: the data between detectors, $P_{ij}(\omega)$, which enter into the
1066: computation of the inverse noise covariance matrix
1067: (Equation~\ref{eq:InvNCorr}). Each cross-power spectrum could be
1068: directly estimated as in Equation~\ref{eq:binavP} (using ${\tilde
1069: d}_{i\omega}$ and ${\tilde d}_{j\omega}$ in the formulae for
1070: detectors $i$ and $j$), but instead we choose to reduce the number of
1071: parameters to estimate at each step, by assuming that the data are
1072: described by a common mode between detectors plus {\it independent\/} noise
1073: (Equation~\ref{eq:lincmod}). In the framework of this model, the
1074: expected cross- and auto-power spectra depend directly on the
1075: following parameters: $\alpha$, the amplitude of the common mode in
1076: each channel; $\left\langle c(\omega)^*\cdot c(\omega)\right\rangle$,
1077: the power spectrum of the common-mode part; and
1078: $\left\langle{\tilde{n}}(\omega)^*\cdot {\tilde{n}}(\omega)\right\rangle$,
1079: the power spectrum of the noise component, which is independent
1080: between detectors. The relation between the model of $P(\omega)$ and
1081: the parameters has been shown in Equation~\ref{eq:modSpCorr}. These
1082: parameters are typically not known {\it a priori}, and must be
1083: measured using the data themselves. We use a blind `component separation'
1084: method developed
1085: for an entirely different problem in \cite{delabrouille03}. This allows us
1086: to obtain a single estimate of all the parameters described
1087: previously, by simultaneously using all the observed timestreams of a
1088: given field for all the detectors in a specified channel (i.e.~at a
1089: single frequency). The method is known to be the Maximum Likelihood
1090: solution for a Gaussian and stationary model of both the noise and the
1091: common-mode. The cross- and auto-power spectra $P(\omega)$ are then
1092: computed following Equation~\ref{eq:modSpCorr}, using these same
1093: estimated parameters. Figure~\ref{fig:noiseSP} shows the estimated
1094: noise power spectra in a sample of BLAST data for one representative
1095: detector using three hours of timestreams during scans of the ELAIS-N1
1096: field (which is known to be essentially devoid of signal).
1097: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1098: %\clearpage
1099: \begin{figure}[t!]
1100: \begin{center}
1101: \includegraphics[width=\columnwidth]{f3.eps}
1102: \end{center}
1103: \caption{Total noise power spectrum (solid) for one of
1104: the $250\,\mu$m detectors, using about three hours of data,
1105: corresponding to scans of the ELAIS-N1 field. The dotted curve
1106: correspond to the estimated power spectrum of the common-mode
1107: between detectors at $250\,\mu$m, rescaled by an amplitude factor
1108: for the specific detector being considered ($\alpha_i$ in the
1109: text, where $i$ is the detector index), also estimated using the
1110: data themselves. The dashed curve represents the estimated power
1111: spectrum of the uncorrelated part of the noise for this detector.
1112: The common-mode is very strong in these data and dominates over the
1113: uncorrelated noise at frequencies lower than about $0.1\,$Hz.
1114: Most of the low frequency noise excess comes from this
1115: common-mode. The uncorrelated part of the noise shows a knee
1116: frequency of less than $0.02\,$Hz, which is 5--10 times smaller
1117: than for the total noise power spectrum. The power of the
1118: uncorrelated noise is thus reduced by a factor of about 100 at low
1119: frequencies. The excess signal at the scanning frequency (peak
1120: around $0.04\,$Hz in the power spectrum) is completely common
1121: between detectors. }
1122: \label{fig:noiseSP}
1123: \end{figure}
1124: %\clearpage
1125: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1126: The auto-power spectrum is shown, as well as its decomposition in terms
1127: of the common-mode power spectrum and the uncorrelated noise power
1128: spectrum.
1129:
1130: Because our estimate of the converged map of the sky ${\hat s}$ is not
1131: perfect and contains contributions from residual noise, then in
1132: subtracting a simulated signal timeline from the data to estimate the
1133: noise power spectrum we reintroduce some noise to the data which could
1134: potentially bias our estimate of the noise power spectrum
1135: \citep{ferreira,hamilton}. However this effect is greatly reduced by
1136: the large redundancy in each pixel of the final maps, as a result of
1137: the many repeated scans and the large number of detectors at each
1138: wavelength. The bias can be neglected to first order for BLAST, since
1139: the noise level per pixel in the final map is much smaller than the
1140: noise in individual detector timestreams.
1141:
1142: \subsection{Dealing with gaps in the data}
1143: \label{sub:gaps}
1144:
1145: In order to derive the formalism presented so far, we have assumed
1146: that each data segment is stationary and hence consists of a
1147: continuous series of
1148: data points. However, we have seen in Section~\ref{sub:process} that
1149: the BLAST05 data contain multiple gaps of typical size less than one
1150: second. The amount of data in these gaps is a few percent of the
1151: total. In order to reasonably restore the continuity of the data
1152: we have filled the gaps with random noise, as described in
1153: Section~\ref{sub:process}.
1154: The data samples generated in the gaps are
1155: not reprojected into the final map but are directed to ``dummy'' pixels.
1156: In principal, the optimal approach would be to create one dummy pixel
1157: per flagged data sample, avoiding the possibility of several simulated
1158: samples falling on the same pixel (through the rebinning of flagged
1159: pixels we do not want to introduce any spurious constraints for the
1160: map-making process which could arise from adding crossings over
1161: different time intervals). However, the approach of using one dummy
1162: pixel per gap sample is impractical, because the total number of dummy
1163: pixels would be excessive for typical BLAST observations, and the
1164: pixel-pixel covariance matrix (which has size the total number of
1165: pixels squared) would be prohibitive to store and compute.
1166:
1167: We have adopted a simpler approach, which does not lead to the
1168: mathematically exact solution, but comes very close (as has been shown
1169: in simulations). This consists of rejecting (for the computation of
1170: $N_{pp'}^{-1}$) all the elements of $N_{tt'}^{-1}$ associated with flagged
1171: samples. This is equivalent to removing from $N_{pp'}^{-1}$ the rows
1172: and columns corresponding to the dummy pixels before the inversion of
1173: the matrix, as opposed to after inversion, which would be the correct
1174: treatment discussed above. This approach is also equivalent to
1175: assuming null off-diagonal terms for those rows and columns. However,
1176: such dummy pixels are obviously correlated to some degree with the real
1177: pixels in the map, and hence this cannot be entirely correct.
1178:
1179: Nevertheless, we have verified that this approximation has a small
1180: impact (a few percent only) on the final map for most of the angular
1181: scales which are sampled, although we found some differences at very
1182: large angular scales, sizes of the order of the map size.
1183: For now, we have not put much
1184: effort into recovering those very large scales because they are
1185: subject to other effects, as discussed in
1186: Section~\ref{subsub:sigonlyIVCG}. The minimal impact on the final map has
1187: been verified using pure signal simulations and by comparing the
1188: results obtained between our simple approach and the correct
1189: map-making solution. This approach works to a high degree of accuracy
1190: on most scales because the gaps are small and do not introduce
1191: important discontinuities in the timestreams.
1192:
1193: We have used this simple approach because it gives sufficiently
1194: accurate results over the relevant angular scales, while being simple and
1195: fast to implement.
1196: However, another iterative procedure could be adopted,
1197: which would lead to the exact solution.
1198: In this approach,
1199: we define two maps. The first map ``A''is made from only the uncorrupted
1200: (i.e.~real sky) samples, while the second map ``B'' is obtained from
1201: projecting the simulated (i.e.~for gap-filling) samples. The difficulty
1202: arises in deciding what to do when simulated data
1203: from different scans or detectors fall in the same pixel -- one might want
1204: the ``generated signal'' (and not necessarily the noise) to be identical in
1205: both measurements, in order to satisfy the map-making hypothesis.
1206: If this condition is not satisfied, then some artifacts may be introduced
1207: into both maps. Here is a solution to this problem:
1208:
1209: \begin{enumerate}
1210: \item generate a first set of maps ``A'' and ``B'' after filling the gaps
1211: in the timestreams with white noise + a linear baseline.
1212: \item fill the gaps in the data with the best estimate of the signal in
1213: map ``A'' together with white noise + a baseline which is fitted in the
1214: gap vicinity of the data ``minus'' signal timestream.
1215: \item recompute the maps ``A'' and ``B''. Step 2 ensures that the signal is
1216: the same for each generated sample falling in the same pixel of map ``B''.
1217: \end{enumerate}
1218:
1219: This approach can also be coupled with the procedure for estimating the
1220: noise power spectra described in Section~\ref{sub:NSP}. Preliminary
1221: results indicate that this approach works in practice. Detailed
1222: studies will be presented in a future publication.
1223:
1224:
1225: \subsection{Pixel constraints}
1226: \label{sub:pixconstr}
1227:
1228: For some specific observed fields we may have strong priors about the
1229: sky emission at a given location. For instance, we know that over
1230: some regions the astronomical signal should vary very smoothly or should be
1231: very weak with respect to the noise, at least outside some localized
1232: region. This is the case in particular when we map bright
1233: extragalactic sources in order to calibrate the detectors and estimate
1234: the beams; in these cases, regions beyond some predefined distance
1235: from the beam center can be assumed to have null flux (or a constant
1236: relative flux in the map, since we do not have access to the DC level
1237: in maps). If we really have strong prior knowledge that we are
1238: dealing with a bright localized region, then we can take a further
1239: drastic step -- we can constrain the map to have the same value in some
1240: domains of the sky by defining a single pixel containing all the data
1241: samples falling in that region.
1242:
1243: In practice, we define a small box centered at the source location and
1244: constrain the part of the map outside this box to have a constant
1245: value. This is a very efficient way to remove stripes from the map,
1246: since the extremities of all the paths across the map are re-adjusted.
1247: We have used this technique to make maps of the isolated calibrators
1248: observed by BLAST (Truch et al.~2007).
1249:
1250:
1251: \subsection{Error estimation}
1252: \label{sub:errorestim}
1253:
1254: The variance of the noise in each pixel of the final map and its
1255: correlations are directly given by the pixel-pixel covariance matrix
1256: $N_{pp'} = (A^t N^{-1} A)^{-1}$. This is true given the following assumptions:
1257: that our model of the data holds, in particular that the noise is a purely
1258: Gaussian random process, which may not be the case in practice at low
1259: frequencies; and that our estimate of the sample-sample
1260: noise covariance matrix $N_{tt'}$ is accurate enough that the errors do not
1261: propagate significantly into the final map. As already mentioned, we
1262: never explicitly compute the covariance matrix, but rather its
1263: inverse. The direct inversion would take a prohibitive
1264: computation time for most applications. However, to first order, we can
1265: obtain an
1266: estimate of the errors by inverting the diagonal part of $N_{pp'}^{-1}$
1267: only, neglecting the off-diagonal terms. This is equivalent to assuming
1268: that the noise in the final map is white. We have checked with the
1269: help of simulations that this very simple approximation is accurate to
1270: better than 10 percent for all our BLAST05 fields, even for those with very
1271: poor cross-linking.
1272:
1273: Provided that the size of the map is reasonably small, so that we are
1274: able to explicitly calculate $N_{pp'}$, we can obtain an accurate estimate
1275: of the errors for a limited (and small) number of pixels in the map. The
1276: variance for pixel $p$, and the covariance with respect to the other
1277: pixels of the map, can be computed by solving the linear system:
1278: \begin{equation}
1279: \left\langle n_p^\dagger \cdot n_{p'}\right\rangle = N_{p'p} u_p,
1280: \label{eq:var}
1281: \end{equation}
1282: where $u_p$ is a unitary vector with a single 1 for pixel $p$. If the
1283: Cholesky decomposition of the matrix has already been performed for
1284: the map-making procedure, the computation of
1285: Equation~\ref{eq:var} is relatively fast and hence can be carried out for a
1286: grid of non-adjacent pixels, for example. This can be used to check
1287: the validity of the error prediction approximation described in the
1288: previous paragraph.
1289:
1290: \subsection{Computational requirements}
1291:
1292: For the brute-force inversion algorithm (in which the full inverse
1293: pixel-pixel covariance matrix $N_{pp'}^{-1}$ is computed), five
1294: minutes of computation with a single $3\,$GHz processor are needed to
1295: process 2 hours of data from a single detector at a rate of $100\,$Hz.
1296: The computational time is proportional to the number of samples if
1297: this is longer than the assumed correlation length of the noise in the
1298: data (which has been evaluated to be $\lambda_{\rm c} = 200\,$s in
1299: BLAST05 timestreams). If noise correlations between detectors are
1300: also to be accounted for, the computational time is proportional to
1301: the square of the number of detectors.
1302:
1303: Most of the computing time is spent on calculating $N_{pp'}^{-1}$.
1304: Inversion of the linear system to estimate the map $s$ is relatively
1305: fast (a few minutes to a few hours for maps of several square degrees
1306: in size).
1307:
1308: For the iterative algorithm, about two minutes of computational time
1309: is required for two hours of data (under the same conditions described
1310: above). This assumes that 100 iterations are necessary to reach
1311: convergence (which is a realistic number for most applications), and
1312: the algorithm scales with $n_{\rm s}\log n_{\rm s}$. The situation is
1313: much better than for the brut-force inversion algorithm if
1314: correlations between detectors are included. In that case, the
1315: algorithm scales with the square of the number of detectors if this
1316: exceeds about 40. If there are fewer than about 40 detectors then the
1317: algorithm scales linearly with the number of detectors. As an
1318: example, if there are 100 detectors, then including noise correlations
1319: between the detectors increases the computational time by a factor of
1320: four with respect to the ``no-correlation'' case. The full processing
1321: of 10 hours of BLAST05 data at all wavelengths, including
1322: detector-detector correlations, can be done with a single processor in
1323: a few days.
1324:
1325: In terms of memory, the brute-force inversion algorithm requires
1326: storage of the full $N_{pp'}^{-1}$ matrix. However, for the iterative
1327: algorithm, only vectors of the size of the maps need to be kept in
1328: memory, which is much less demanding.
1329:
1330: \section{Simulations for testing SANEPIC}
1331: \label{sec:appli}
1332:
1333: We now focus on the application of SANEPIC to data. Our aim is to
1334: develop tests to validate our method using simulated BLAST
1335: observations. We derive conclusions about how well low frequency noise
1336: in the maps can be reduced, depending on observational parameters such as
1337: scanning strategy, and we compare the results obtained with those from
1338: simpler methods based on filtering the data, e.g., common-mode
1339: subtraction.
1340: In this section, we describe the simulations performed to test the
1341: SANEPIC method.
1342:
1343: We have generated several different sets of simulations of BLAST timestreams.
1344: Each set of simulated timestreams,
1345: representing one particular observed field, is generated for all the
1346: BLAST detectors used for the analysis of real data (132 at
1347: 250$\,\mu$m, 78 at 350$\,\mu$m and 39 at 500$\,\mu$m) and is the sum
1348: of simulated astrophysical signal, independent noise, and common-mode
1349: noise between all detectors. The noise is generated randomly with
1350: Gaussian statistics, given fixed power spectra derived from real BLAST05
1351: data. Figure~\ref{fig:noiseSP} shows an example of the power spectrum
1352: of the noise in the data used as input to the simulations for one
1353: of the fields. The part of the noise which is independent between
1354: detectors is generated for every detector timestream and has a power
1355: spectrum well described on average by a relatively flat plateau for
1356: frequencies larger than about $0.05\,$Hz, and by a part proportional
1357: to $(1/f)^{2.5}$ for smaller frequencies (these characteristics vary
1358: slightly from detector to detector). Our knowledge of the real
1359: bolometer noise power spectrum at low frequency is limited by the very
1360: dominant common mode. In simulations, the common-mode noise is
1361: generated once for all detectors and has a power spectrum very well
1362: fit by a power law with an index equal to approximately 2.5, together
1363: with some broad peaks, the largest being at the scanning frequency
1364: (the amplitude of the peak depends on the scanning strategy and the
1365: observed field). The common-mode power spectrum has an amplitude such
1366: that it reaches the level of the independent noise at about $0.3\,$Hz
1367: (see Figure~\ref{fig:noiseSP}). The generated common-mode timestream
1368: is multiplied by an amplitude factor which varies from detector to
1369: detector by $\sim 10\%$ and is added to the simulated detector
1370: timestreams. The amplitude factors used for the simulations have been
1371: estimated from the data themselves.
1372:
1373: In order to represent the astrophysical signal, we have simulated
1374: simple maps of diffuse emission with a power spectrum proportional to
1375: $k^{-3}$, as for typical Galactic cirrus emission (e.g.~Miville-Desch{\^e}nes
1376: et al.~2007). Maps are generated
1377: following Gaussian statistics with a resolution of 1$\arcsec$, much
1378: higher than the typical pixel sizes in the final maps, in order to
1379: reduce artifacts due to re-pixelization. The amplitude of the
1380: fluctuations of the simulated map is chosen to match the expected
1381: level of signal in each observed field. The simulated maps are scanned
1382: using BLAST05 pointing, and pure signal timestreams are generated for
1383: each detector. Signal and noise timestreams are added at the end of
1384: the procedure (but see Section~\ref{sub:results} for an explanation of
1385: why this operation is not always carried out).
1386:
1387: We have generated two sets of simulations which correspond to two
1388: different fields observed by BLAST. We selected two fields that were
1389: observed with very different scanning strategies, since the
1390: performance of the map-making procedure is very dependent on scanning
1391: strategy; this allows us to test SANEPIC in two very different
1392: configurations. In the first case the scanning was performed mainly in
1393: a single direction over a short time interval, while in the second case
1394: the field was observed several times during the flight at different
1395: scanning angles, to achieve significant cross-linking in the map.
1396:
1397: The first data-set uses observations of the \casa supernova remnant
1398: emission which comprises about 20 minutes of data. BLAST observations
1399: of this field and derived conclusions will be described in detail in
1400: Hargrave et al.~(in preparation).
1401: The rectangular region mapped has a size of the order of $0.5\,{\rm deg}^2$
1402: and was scanned two times back and forth over a
1403: short time interval. We have generated simulations corresponding to
1404: all the detectors at 250$\,\mu$m (we used a total of 132 detectors).
1405:
1406: The second data-set reproduces the observations of the intermediate
1407: velocity cloud IVC G86.5+59.6 (hereafter `G86'). Simulations include
1408: four different visits of the field performed during the flight at very
1409: different time intervals (ranging from a few hours to more than a
1410: day). Each continuous observation segment has a size which varies from
1411: one to two hours. Two scanning directions are dominant, which form an
1412: angle close to 50$^\circ$. The region covered has a size of about
1413: $2\,{\rm deg}^2$ on the sky. Simulations for this field are performed
1414: specifically for all the 500$\,\mu$m detectors (41 detectors used).
1415: Similar Monte Carlo simulations at 250$\,\mu$m would have taken a factor
1416: of 10 longer, while we believe that the conclusion would remain unchanged.
1417:
1418: A total of 20 sets of simulations of the observations for each field
1419: have been performed. For each set we vary the realization of the noise
1420: and of the signal input map. About four hours of computing time are needed
1421: to create one realization of a full set of simulations of \ivcg with a
1422: single processor, compared to a few minutes for the \casa simulations.
1423: This is using the pre-computed full pixel-pixel covariance matrix,
1424: which was also used to analyze the real data.
1425:
1426:
1427: \section{Results from simulations}
1428: \label{sub:results}
1429:
1430: We now present the results obtained with SANEPIC applied to the two
1431: sets of simulated data. In each case, we compare the final map with
1432: other maps obtained using simpler map-making procedures. For these tests
1433: we have assumed that the noise power spectra are perfectly known, rather
1434: than estimated separately from each data-set; in practice we fix the
1435: noise power spectra to be the ones from the simulations. We have
1436: verified that relaxing this constraint has almost negligible change
1437: on the final maps.
1438:
1439: For these simulated data-sets, we have applied some pre-processing of the
1440: timestreams before applying SANEPIC, just as we do for the real data.
1441: We have systematically
1442: removed a 5th order polynomial from each timestream segment and
1443: weakly high-pass filtered the data, as described in Section~\ref{sub:process}.
1444: Finally, for the gap-filling we flag the simulated data at the
1445: same locations as in the real data in order to check the influence of
1446: the flagging procedure in the final maps.
1447:
1448: In the following we have applied SANEPIC independently to pure noise
1449: timestreams (containing independent noise and common-mode noise, but
1450: without simulated astrophysical signal) and to pure signal
1451: timestreams. This procedure allows us to easily derive conclusions
1452: about the noise properties in the final maps, as well as about the signal,
1453: without biasing the results, because SANEPIC is a linear method (as shown by
1454: Equation~\ref{eq:solmap}). This is only strictly true if the noise power
1455: spectrum is fixed as done here, and not estimated simultaneously along
1456: with the maps. Then, applying SANEPIC on pure noise timestreams and
1457: pure signal timestreams independently and adding the two final maps
1458: is rigorously equivalent to applying SANEPIC on signal plus noise
1459: timestreams. This has been checked numerically, and we find that the
1460: difference is consistent with double floating precision error. An
1461: important consequence of this is that the properties of the noise in
1462: the final map are independent of the signal-to-noise ratio.
1463:
1464: \subsection{Case without cross-linking}
1465: \label{sub:CasACase}
1466:
1467: \subsubsection{Noise-only timestreams}
1468:
1469: We first study the maps resulting from the noise-only timestreams in
1470: the configuration of \casa observations. The chosen pixel size of the
1471: map is 25$\arcsec$ and matches the pixel size of the maps discussed in
1472: Hargrave et al.~(in preparation).
1473: We compare the noise maps obtained from three different procedures:
1474: \begin{itemize}
1475: \item Case 1: use SANEPIC with the correct treatment of the
1476: correlated noise.
1477: \item Case 2: use SANEPIC fixing the correlation of noise between
1478: detectors to zero and fixing the noise power spectrum for each
1479: detector to the power spectrum of the sum of uncorrelated noise and
1480: common mode. This procedure is very similar to more standard
1481: map-makers in the literature (e.g., Stompor et al.~2002).
1482: \item Case 3: make a simple re-projection of the data onto a pixelized
1483: map by simply averaging the data falling in each pixel, after having
1484: filtered the timestream data with the same very weak low-pass
1485: filter used for SANEPIC. This procedure is sometimes called ``co-addition''.
1486: \end{itemize}
1487:
1488: Figure~\ref{fig:mapnCasA_corr} shows computed noise maps for one of the
1489: realizations of the noise in each of the three cases.
1490: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1491: %\clearpage
1492: \begin{figure*}[t!]
1493: \begin{center}\resizebox{16.5cm}{!}{
1494: \includegraphics[scale=0.35]{f4a.eps}
1495: \includegraphics[scale=0.35]{f4b.eps}
1496: \includegraphics[scale=0.35]{f4c.eps}
1497: }
1498: \end{center}
1499: \caption{Final maps computed from simulated pure noise timestreams
1500: in the configuration of the BLAST05 \casa observations, which
1501: have a dominant scan direction. From left to right:
1502: maps obtained with SANEPIC including noise correlations; SANEPIC
1503: with no noise correlations included in the model; and simple pixel
1504: binning (see text for more details). Note the extended dynamic
1505: range of the simple co-added map (right panel). The maps have a
1506: size of about 40$\arcmin$ in the cross-scan direction and about
1507: $1^\circ$ along the scan. The pixel size is 25$\arcsec$.}
1508: \label{fig:mapnCasA_corr}
1509: \end{figure*}
1510: %\clearpage
1511: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1512: As expected, the map obtained with the simple pixel binning approach
1513: contains a very large amount of low frequency noise, with strong striping
1514: visible along the scan direction. Residual low frequency
1515: noise can also be seen in the map obtained using SANEPIC {\it without\/}
1516: accounting for the noise correlations between detectors. We do not expect this
1517: method to be very efficient, since it is very non-optimal in cases
1518: (such as this example)
1519: where a very large fraction of the noise is correlated
1520: between detectors. In contrast, the noise map obtained with SANEPIC is
1521: quite satisfactory, showing reduced power at low frequency as
1522: compared to the previous case. Nevertheless, some very weak excess
1523: power is seen in the cross-scan direction. This is expected, since the
1524: map is not cross-linked, and very poor constraints can be put on the
1525: cross-scan directions at low spatial frequencies (two positions in the
1526: map separated by more than the size of the array in the cross-scan
1527: direction are observed far apart in time).
1528:
1529: In order to quantify the level of low frequency noise in the maps, we
1530: compute the 1-D power spectra of the maps, averaged over the 20
1531: realizations of the simulated data. For the computation of power
1532: spectra, we take into account only the central part of each map, where
1533: the level of redundancy in the observations is high (we use only the
1534: highest signal-to-noise region in the maps). To do so, we apply an
1535: apodized mask to the maps going smoothly from 0 at the edges
1536: to 1. Figure~\ref{fig:noise1dSpCasA} shows the noise power spectra in
1537: the three cases.
1538: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1539: %\clearpage
1540: \begin{figure}[h!]
1541: \begin{center}
1542: \includegraphics[width=\columnwidth]{f5.eps}
1543: \end{center}
1544: \caption{One-dimensional power spectra of the noise (rebinned in
1545: frequency) in the final noise maps after map-making in the BLAST05
1546: \casa configuration. Power spectra are averaged over the 20
1547: realizations of the simulated data. The dashed curve is for the
1548: simple re-projection map, the dot-dashed curve for SANEPIC with
1549: {\it no\/} noise correlation between detectors and the
1550: triple-dot-dashed curve for SANEPIC {\it including\/} a treatment
1551: of the correlations. The straight line indicates the level of
1552: white noise in the map predicted by the map-making procedure (see
1553: Section~\ref{sub:errorestim}). Error bars are computed from the
1554: dispersion of measurements among the realizations. For
1555: comparison, the upper dotted curve (decreasing almost like a power
1556: law at all scales) represents the power spectrum of the pure
1557: simulated signal in the final map. The solid curve
1558: represents the power spectrum of the final map obtained with real
1559: data using SANEPIC, with correlations included. This shows the
1560: benefit of taking into account correlations of the noise between
1561: detectors in the map-making procedure, reducing the noise
1562: structure far below that of the signal in the map. The real data
1563: power spectrum shows that the signal dominates at all angular
1564: scales larger than about 3$\arcmin$ and at smaller scales we can
1565: see that white noise at the expected level dominates in the
1566: map. The drop of power at around a 3$\arcmin$ scale is due to the
1567: BLAST05 beam.}
1568: \label{fig:noise1dSpCasA}
1569: \end{figure}
1570: %\clearpage
1571: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1572: The noise level in the simple re-projection map is obviously very poor
1573: at all scales. Both of the other map-makers reach the white noise
1574: level for scales smaller than 3$\arcmin$ and have excess power at
1575: larger angular scales. Nevertheless, the gain between full SANEPIC and
1576: SANEPIC without correlations is very important at all scales larger than
1577: about 2$\arcmin$ and reaches a maximum value of about 10 at around
1578: 20$\arcmin$ angular scales. An interesting fact is that the knee
1579: frequency of the noise power spectrum in the optimal case here corresponds
1580: to the inverse of the physical scale of the detector array in the
1581: cross-scan direction (which is of the order of 6$\arcmin$). Indeed,
1582: there are no observational redundancies on scales larger than the
1583: array in the cross-scan direction in the absence of cross-linking in
1584: the map. Thus the very long timescale $1/f$ noise present in the
1585: timestreams is not efficiently removed and re-projects in the final
1586: map at large angular scales. This effect is also present along the
1587: scan direction, but with a lower amplitude as the map is scanned back and
1588: forth. The trend of the large angular scale power spectrum of the
1589: noise in the map just follows the trend of the low frequency noise
1590: power spectrum in the timestreams. We will see in Section
1591: \ref{subsub:crosslink} that this effect is reduced when there are
1592: multiple scanning directions in the map.
1593:
1594: In order to determine the direction in which the noise power is
1595: strongest in the map, we have also computed the 2-dimensional noise power
1596: spectrum. The map of the 2-D power spectrum of the noise obtained with
1597: SANEPIC (noise correlations included) is shown in
1598: Figure~\ref{fig:Sp2dnoiseCasA_corr}.
1599: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1600: %\clearpage
1601: \begin{figure}[t!]
1602: \begin{center}
1603: \includegraphics[width=\columnwidth]{f6.eps}
1604: \end{center}
1605: \caption{Two-dimensional power spectrum of the noise maps in the
1606: BLAST05 \casa observational configuration obtained with SANEPIC (noise
1607: correlations included) plotted on a logarithmic contrast scale.}
1608: \label{fig:Sp2dnoiseCasA_corr}
1609: \end{figure}
1610: %\clearpage
1611: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1612: The large bright spot around the center corresponds to a relatively
1613: isotropic component of correlated noise (at least at large angular
1614: scales). It contains a large fraction of the noise power at
1615: large angular scales (seen in the 1-D power spectrum in
1616: Figure~\ref{fig:noise1dSpCasA}). A smaller, but significant fraction of
1617: the correlated noise is concentrated in directions perpendicular to
1618: the scan direction, as can be seen in the figure. As already
1619: discussed, the reason for this excess power is that the noise in the
1620: cross-scan direction is poorly constrained. This cross-scan component
1621: of the noise is significant all the way up to the pixel scale.
1622:
1623: \subsubsection{Signal-only timestreams}
1624: \label{sec:CasAsignal}
1625:
1626: We now focus on the signal-only timestream simulations. In order to
1627: demonstrate the superior performance of SANEPIC relative to simpler
1628: methods based on data filtering, we compare with a map-making method which
1629: consists of the following: we first remove the whole array average
1630: from each detector timestream and then make maps
1631: using SANEPIC, assuming no correlations between
1632: detectors. Removing the array average reduces the signal to almost
1633: zero for scales larger than the array and so we expect no large scale
1634: structures to survive in the map. This SANEPIC ``common-mode
1635: subtraction'' method is still a better procedure than just
1636: reprojecting the data (after common-mode subtraction) with a well
1637: chosen filtering to suppress noise drifts (at $f_{\rm cut} =
1638: 0.02\,$Hz, for instance, since that corresponds to the knee frequency
1639: of the independent part of the noise). This latter method is commonly
1640: used in the submillimeter community and is referred to as ``sky
1641: removal'' in reduction of SCUBA data \citep{jenness}.
1642:
1643: Figure~\ref{fig:mapsCasA} shows the input map for one of the
1644: signal realizations (left panel), as well as the maps obtained with
1645: SANEPIC (correlations included, central panel) and with the common-mode
1646: subtraction method (right panel). Results are
1647: expected to be worse in the second case, because of the extra
1648: filtering and also because SANEPIC gives less weighting to modes at
1649: lower frequency which are more contaminated by independent noise.
1650: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1651: %\clearpage
1652: \begin{figure*}[t!]
1653: \begin{center}\resizebox{16.5cm}{!}{
1654: \includegraphics[scale=0.35]{f7a.eps}
1655: \includegraphics[scale=0.35]{f7b.eps}
1656: \includegraphics[scale=0.35]{f7c.eps}
1657: }
1658: \end{center}
1659: \caption{From left to right: map of simulated signal used as input
1660: for one of the realizations of the simulations in the
1661: configuration of the BLAST05 \casa field; reconstructed map using SANEPIC
1662: accounting for noise correlations; reconstructed map using the simple
1663: common-mode subtraction method (intensity units here are arbitrary).}
1664: \label{fig:mapsCasA}
1665: \end{figure*}
1666: %\clearpage
1667: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1668: We can see from Figure~\ref{fig:mapsCasA}
1669: that part of the very large scale fluctuations with sizes
1670: of the order of the map are removed using SANEPIC, but apart from
1671: those very large scales, the input map and the SANEPIC map look very
1672: similar. More differences can be seen in the map obtained with the
1673: common-mode subtraction method. This is quantified in Figure
1674: \ref{fig:1DpowerCasA}, which compares the 1-D power spectra of the two
1675: output maps, averaged over 20 simulations and multiplied by $k^3$.
1676: Recall that the input spectrum varies as $k^{-3}$ and so
1677: deviations from a flat line are the result of the map-making
1678: reconstruction. Note that the vertical scale is linear in
1679: Figure~\ref{fig:1DpowerCasA}.
1680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1681: %\clearpage
1682: \begin{figure}[h!]
1683: \begin{center}
1684: \includegraphics[width=\columnwidth]{f8.eps}
1685: \end{center}
1686: \caption{1-D power spectra of signal-only maps for the BLAST05 \casa field
1687: reconstructed using SANEPIC (solid curve) and using the common-mode
1688: subtraction method (dot-dashed curve). Power spectra are
1689: multiplied by $k^3$, displayed on a linear scale and averaged
1690: over 20 realizations of the simulations. The large angular scale
1691: behavior shows the effectively filtering of each map-making procedure,
1692: while the drop off at small scales is caused by the PSF.}
1693: \label{fig:1DpowerCasA}
1694: \end{figure}
1695: %\clearpage
1696: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1697:
1698: With SANEPIC, the power of the reconstructed map decreases for scales
1699: larger than about 30$\arcmin$. There are three reasons for this:
1700: the power spectrum is
1701: computed over only a small fraction of the sky (and for an apodized map),
1702: so that structures of the order of the map size are never fully recovered;
1703: there is weak
1704: filtering of the timestreams at $f_{\rm cut}=5\times 10^{-3}\,$Hz and
1705: through the 5th order polynomial subtraction; modes in the maps that
1706: are very weakly constrained in the map-making procedure tend not to be
1707: reconstructed through the matrix inversion procedure, since the matrix is very
1708: ill-conditioned and numerical problems occur. The last two effects are
1709: the dominant ones. As a result, modes which are preferentially
1710: filtered are those which lie perpendicular to the scan direction.
1711:
1712: At angular scales smaller than about 2$\arcmin$, the power slightly
1713: decreases due to the smoothing effect of the pixelization. The
1714: transfer function at those scales is well described by a sinc
1715: function.
1716:
1717: Turning now to the common-mode subtraction method, the power in the map is
1718: significantly reduced for scales larger than about 10$\arcmin$, and
1719: drops rapidly to zero. This is because the common-mode subtraction
1720: removes power on all scales larger than the array. On smaller
1721: scales, the filtering effect is relatively weak, and is reduced when
1722: the number of detectors increases.
1723:
1724: For these particular simulations the common-mode subtraction method (using
1725: SANEPIC, but with no correlations) does not in fact perform {\it very\/}
1726: poorly compared to the SANEPIC optimal approach.
1727: This is because the observed field is
1728: small, with a size just a few times bigger than the array, and
1729: structure at scales smaller than the array size are not strongly
1730: affected. This particular map is also not cross-linked. However, the
1731: situation is different for large cross-linked maps like the Vulpecula
1732: field, as discussed in Section~\ref{sub:Vulpecula}.
1733:
1734:
1735: \subsection{Case with cross-linking}
1736: \label{subsub:crosslink}
1737:
1738: \subsubsection{Noise-only timestreams}
1739:
1740: We now focus on the set of simulations of the \ivcg field
1741: at 500$\,\mu$m. As in the previous example, we first examine the
1742: maps resulting from noise-only simulated timestreams using three
1743: methods: optimal SANEPIC (with noise correlations taken into account); SANEPIC
1744: without considering noise correlations between detectors; and the simple
1745: co-add method. The chosen pixel size for the reconstructed maps is
1746: 1$\arcmin$, which allows for inversion of the covariance matrix with a
1747: single processor and hence rapid Monte Carlo simulations. The
1748: conclusions drawn would remain unchanged if the pixel size was
1749: reduced.
1750:
1751: Figure~\ref{fig:mapnIVCG86} shows the final noise maps in the three
1752: cases for one realization of the simulations.
1753: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1754: %\clearpage
1755: \begin{figure*}[t!]
1756: \begin{center}\resizebox{16.5cm}{!}{
1757: \includegraphics[scale=0.35]{f9a.eps}
1758: \includegraphics[scale=0.35]{f9b.eps}
1759: \includegraphics[scale=0.35]{f9c.eps}
1760: }
1761: \end{center}
1762: \caption{Noise-only simulations (like Figure~\ref{fig:mapnCasA_corr})
1763: for the BLAST05 \ivcg scanning
1764: configuration. The maps have a size of about 100$\arcmin$ across
1765: the short axis and 2.5$^\circ$ across the long axis. The pixel
1766: size here is 1$\arcmin$. The three panels show the map obtained
1767: with SANEPIC including noise correlations (left), SANEPIC with no
1768: noise correlations in the model (middle), and the simple co-added
1769: map (right). Note the different brightness scale chosen for the last map
1770: due to its larger dynamic range.}
1771: \label{fig:mapnIVCG86}
1772: \end{figure*}
1773: %\clearpage
1774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1775: The simple re-projection map is obviously very stripy and would be of little
1776: use as an estimate of the signal;
1777: nevertheless it helps to visualize the directions of scanning in
1778: the map. We can see two main directions covering the central
1779: region of the map, oriented at about 50$^\circ$ to each other. A third
1780: scanning direction is also visible but has a much smaller weight. The
1781: central part of the map is the cross-linked region, where we expect the
1782: more optimal map-making procedures to excel.
1783:
1784: In the map obtained using SANEPIC {\it without\/} considering noise
1785: correlations between detectors (middle panel of Figure~\ref{fig:mapnIVCG86}),
1786: some large scale noise is still visible in the
1787: map, even if almost no residual striping is apparent in the cross-linked
1788: region. Indeed, too much weight is given to the large timescales in
1789: the timestreams (which are basically common between all the
1790: detectors) as compared to the smaller timescales for which there are
1791: more independent measurements, because the degree of correlation of
1792: the noise is weaker. The residual noise at large angular scales is
1793: much weaker in the SANEPIC map in which we account for the proper
1794: correlations of the noise between detectors (left panel of
1795: Figure~\ref{fig:mapnIVCG86}). The noise in this map
1796: looks particularly white.
1797:
1798: The noise level in each simulated map is quantified
1799: in Figure~\ref{fig:noise1dSpIVCG86}, showing the 1-D power spectrum of
1800: residual noise averaged over 20 simulations, and computed over the
1801: cross-linked region (which has a diameter of about 100$\arcmin$).
1802: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1803: %\clearpage
1804: \begin{figure}[t!]
1805: \begin{center}
1806: \includegraphics[width=\columnwidth]{f10.eps}
1807: \end{center}
1808: \caption{Output power spectrum comparison of noise-only simulations
1809: after map-making (like Figure~\ref{fig:noise1dSpCasA}), for the
1810: BLAST05 \ivcg scanning configuration. These power spectra are
1811: computed only in the cross-linked region of the maps, which
1812: forms a large disk of about 100$\arcmin$ diameter and can be easily
1813: identified in
1814: Figure~\ref{fig:mapnIVCG86}. The dashed curve is for the simple
1815: re-projection map (right panel of Figure~\ref{fig:mapnIVCG86}),
1816: the dot-dashed curve is for SANEPIC {\it without\/} consideration of
1817: noise correlations between detectors
1818: (middle panel of Figure~\ref{fig:mapnIVCG86}),
1819: and the triple-dot-dashed curve is for SANEPIC {\it including\/} the correlation
1820: treatment (left panel of Figure~\ref{fig:mapnIVCG86}).
1821: The horizontal
1822: line indicates the level of white noise in the map predicted by
1823: the map-making procedure. Error bars are estimated from the
1824: dispersion among measurements for all the realizations. Residual low
1825: frequency noise in the optimal SANEPIC map is very low, thanks to
1826: the multiple scanning directions of this field. The situation is
1827: much better than for the \casa observational configuration, which
1828: had a single scan direction (Figure~\ref{fig:noise1dSpCasA}).}
1829: \label{fig:noise1dSpIVCG86}
1830: \end{figure}
1831: %\clearpage
1832: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1833: While several orders of magnitude are gained in the noise power at all
1834: scales using the SANEPIC ``no correlation'' method as compared to the simple
1835: re-projection method, accounting for the correlations with SANEPIC
1836: allows us to further reduce the noise power on scales ranging from 20$\arcmin$
1837: to the size of the map by an additional factor of ${\sim}\,5$. Toward smaller
1838: scales, the gain between the SANEPIC correlation versus no correlation test
1839: cases is still very significant down to about 10$\arcmin$, where the both
1840: methods start to approach the white noise level. In the optimal map
1841: obtained with SANEPIC the ratio between the noise power spectrum at
1842: large scales and the white noise level is around 20, which is
1843: relatively small. Figure~\ref{fig:2dspectrumIVCG86} shows the 2-D
1844: power spectra of the noise in the SANEPIC map averaged over 20
1845: realizations. As expected, the large scale noise is more important in
1846: directions perpendicular to the scans.
1847: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1848: %\clearpage
1849: \begin{figure}[h!]
1850: \begin{center}
1851: \includegraphics[width=\columnwidth]{f11.eps}
1852: \end{center}
1853: \caption{2-D power spectrum of a noise-only simulation reduced using
1854: SANEPIC (like Figure~\ref{fig:Sp2dnoiseCasA_corr}) for the BLAST05
1855: \ivcg scanning configuration.}
1856: \label{fig:2dspectrumIVCG86}
1857: \end{figure}
1858: %\clearpage
1859: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1860:
1861: \subsubsection{Signal-only timestreams}
1862: \label{subsub:sigonlyIVCG}
1863:
1864: We now focus on signal-only simulations for G86. As in the case of the
1865: \casa configuration, we compare the performance of SANEPIC with
1866: respect to the simpler common-mode subtraction method.
1867: Figure~\ref{fig:mapsIVCG86} shows the pure signal input map for one
1868: realization of the simulations (left panel) compared with two recovered maps.
1869: The first output map is obtained with SANEPIC
1870: including correlations between detectors (middle panel), while the second is
1871: obtained by subtracting the common mode between all detectors, followed by
1872: applying SANEPIC, but neglecting noise correlations between detectors (right
1873: panel). We see that the very
1874: largest scales are not recovered by SANEPIC. This is because of the
1875: weak filtering applied to the timestreams and, to a greater extent,
1876: because of inversion problems with SANEPIC on scales of the order of
1877: the size of the map (these scales are very poorly constrained by the
1878: map-making procedure). In the common-mode subtraction method, only the
1879: very largest scales are suppressed by the map-making procedure itself,
1880: but the filtering effect is more dramatic and extends to much smaller
1881: scales.
1882: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1883: %\clearpage
1884: \begin{figure*}[t!]
1885: \begin{center}\resizebox{16.5cm}{!}{
1886: \includegraphics[scale=0.35]{f12a.eps}
1887: \includegraphics[scale=0.35]{f12b.eps}
1888: \includegraphics[scale=0.35]{f12c.eps}
1889: }
1890: \end{center}
1891: \caption{Signal-only simulations (like Figure~\ref{fig:mapsCasA})
1892: for the BLAST \ivcg configuration.
1893: The three panels show the input signal map (left), full
1894: SANEPIC (middle) and simple common-mode subtraction followed by
1895: application of SANEPIC without correlations (right).}
1896: \label{fig:mapsIVCG86}
1897: \end{figure*}
1898: %\clearpage
1899: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1900:
1901: The effective filtering is quantified in
1902: Figure~\ref{fig:1DpowerIVCG86}, which shows the power spectra of output
1903: maps averaged over 20 simulations of pure signal timestreams. Again the
1904: power spectra are multiplied by $k^3$ for comparison purposes.
1905: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1906: %\clearpage
1907: \begin{figure}[h!]
1908: \begin{center}
1909: \includegraphics[width=\columnwidth]{f13.eps}
1910: \end{center}
1911: \caption{Power spectrum comparison of the SANEPIC map (solid) versus the
1912: simple common-mode removal map (dot-dashed) for signal-only simulated
1913: data, as in Figure~\ref{fig:1DpowerCasA}, but for the BLAST05 \ivcg
1914: scanning configuration. Power spectra are multiplied by $k^3$, so that
1915: a flat line would indicate no filtering. The drop off at small angular
1916: scales is due to the BLAST05 PSF.}
1917: \label{fig:1DpowerIVCG86}
1918: \end{figure}
1919: %\clearpage
1920: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1921: In this case SANEPIC works well on scales up to about half a degree, above
1922: which it fails to recover structure in the map;
1923: this limit corresponds to scales of about a
1924: quarter of the map and larger. The filtering effect is much more
1925: pronounced in the map produced with the common-mode subtraction method,
1926: being strong for all scales above around
1927: 14$\arcmin$. This is of course expected, since subtracting the average of the
1928: array strongly reduces the signal on scales larger than the array
1929: size. Therefore, in order to recover large and intermediate scale
1930: structure in the maps, it is beneficial to use SANEPIC instead of
1931: other methods that are based on simply filtering the data.
1932:
1933: \subsection{Advantages of cross-linking}
1934:
1935: The relative level of residual noise at low spatial frequency in the
1936: maps is significantly reduced in the \ivcg observational configuration
1937: as compared to the \casa configuration. The fundamental difference is
1938: that the \ivcg observations contain multiple (essentially two for most of the
1939: data) scanning directions, while the \casa observations are realized
1940: with only two passes across the field in the same direction. Multiple
1941: scanning directions give a huge number of additional constraints for
1942: the map-making procedure. In particular, large scale structures in the
1943: map are much better recovered in directions parallel to scans, because
1944: the noise there is smaller. Thus having multiple scanning angles allows for
1945: recovery of the sky fluctuations for all directions, and ends up giving
1946: almost no weight to the individual loosely constrained cross-scan k-modes.
1947:
1948: Differences in the results for maps from these two example scanning strategies
1949: can be quantified in two ways. First of all, for the \ivcg scanning
1950: strategy, the transition between white noise and ``excess'' large
1951: scale noise in the map occurs at a scale around 10$\arcmin$, while the
1952: same transition occurs at a scale of around 3$\arcmin$ for the \casa
1953: scanning strategy. Secondly, the ratio between large scale noise power
1954: and white noise power is larger by more than two orders of magnitude
1955: for \casa than for G86. On the other hand, some caution should be
1956: taken to not over-interpret this comparison, because the pixel size we used
1957: is larger for the \ivcg map (1$\arcmin$) as compared to the \casa map
1958: (25$\arcsec$), and therefore the number of crossings per pixel is
1959: greater for the \ivcg map. Nevertheless, this simulation exercise has
1960: demonstrated that cross-linking in the map is extremely beneficial,
1961: particularly for recovering the large scale structures in the map.
1962:
1963: \subsection{Map-making transfer function}
1964:
1965: When carrying out a complex data processing procedure, it is important
1966: to check whether the results are biased in any way. We have found
1967: that the transfer function of the map-making procedure, defined as the
1968: ratio between the amplitude of fluctuations in the output pure signal
1969: map relative to the input map, is not always exactly unity, even for
1970: intermediate and small angular scales in the map. For example, in the \casa
1971: configuration at 250$\,\mu$m the fluctuation amplitudes in the final
1972: map are reduced by 3\% on average as compared to the input map, almost
1973: uniformly across spatial frequencies and directions. This global
1974: discrepancy reaches the level of 9\% at 500$\,\mu$m. Moreover, it is also
1975: present for the \ivcg configuration. We believe that this reduction
1976: is due to the fact
1977: that the pixel-pixel covariance is ill-conditioned and numerical
1978: imprecision occurs in the matrix inversion. We find that the bias
1979: tends to be smaller when the number of detectors is larger and also
1980: when the number of constraints increases, like when we have multiple
1981: scanning directions, or when we map isolated bright sources (presented
1982: in Truch et al.~2007) and constrain the data outside a defined region
1983: to have a constant flux (see Section~\ref{sub:pixconstr} for details
1984: of this procedure). Since this bias can be estimated using
1985: simulations, it is straightforward to correct for. We have found that
1986: it is not always important, e.g.~for the large Galactic map in the Vulpecula
1987: region analyzed in Chapin et al.~(2007) the bias is negligible.
1988:
1989:
1990: \section{Application to real BLAST05 data}
1991: \label{sec:AppliData}
1992:
1993: Now that we have looked at signal-only and noise-only simulations, we
1994: now turn to real BLAST data.
1995: In this section, we show maps of two example fields from the BLAST05
1996: data which have been obtained using SANEPIC.
1997:
1998: \subsection{Cassiopeia A}
1999: \label{sub:resCasA}
2000:
2001: Figure~\ref{fig:mapCasA250} shows the map obtained from the
2002: observations of the \casa field at 250$\,\mu$m using SANEPIC including full
2003: consideration of the noise correlated between detectors. Detailed analysis
2004: of the maps at the three wavelengths is described in Hargrave
2005: et al.~(in preparation).
2006: The properties of the noise and the transfer function of the signal in the
2007: map have been studied in detail from Monte Carlo simulations in Section
2008: \ref{sub:CasACase}. Results from such simulations are used to characterize
2009: the map. The power spectrum of this map has been compared to results
2010: from simulations in Figure~\ref{fig:noise1dSpCasA}.
2011: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2012: %\clearpage
2013: \begin{figure}[t!]
2014: \begin{center}
2015: \includegraphics[width=\columnwidth]{f14.eps}
2016: \end{center}
2017: \caption{Map of the \casa supernova remnant at 250$\,\mu$m made from BLAST05
2018: data using SANEPIC including noise correlations between
2019: detectors. The map is represented in Galactic coordinates with
2020: 25$\arcsec$ pixels and has a size of about $0.5\,{\rm deg}^2$.}
2021: \label{fig:mapCasA250}
2022: \end{figure}
2023: %\clearpage
2024: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2025: The map structures are relatively smooth, due to the BLAST05 point spread
2026: function, which has a width of the order of 3$\arcmin$, causing the
2027: drop in the 1-D power spectrum below those spatial scales. The signal
2028: clearly dominates over noise on angular scales larger than about
2029: 3$\arcmin$, and the diffuse structure should be reliable up to a large
2030: fraction of the overall map size.
2031:
2032: \subsection{Vulpecula region}
2033: \label{sub:Vulpecula}
2034:
2035: Another field observed during the BLAST05 campaign is centered in
2036: the Galactic Plane close to the open cluster NGC 6823 in the
2037: constellation of Vulpecula. The region mapped has a size of about
2038: $4\,{\rm deg}^2$ and was chosen for its high-mass star formation
2039: activity. Complete analysis of this observed field is presented in
2040: \cite{chapin}.
2041:
2042: A few hours of these data were taken at different time intervals
2043: during the flight. By design, this field has been observed with very
2044: different scanning directions, and is therefore it should be possible to
2045: recover diffuse large scale structures.
2046:
2047: The map of the observed region at 250$\,\mu$m obtained with SANEPIC is
2048: shown in Figure~\ref{fig:mapL5B250} (left panel). For comparison
2049: (right panel), we have computed another map using a much simpler method
2050: which consists of removing the array average signal at each timestep
2051: for all of the timestreams and reprojecting the data onto the map
2052: after filtering. This is like the ``sky removal'' procedure often
2053: carried out for ground-based submillimeter data (see also
2054: Section~\ref{sec:CasAsignal}).
2055: One can see that it
2056: suppresses almost all the diffuse structure in the map.
2057: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2058: %\clearpage
2059: \begin{figure*}[t!]
2060: \begin{center}\resizebox{16.5cm}{!}{
2061: \includegraphics[scale=0.35]{f15a.eps}
2062: \includegraphics[scale=0.35]{f15b.eps}
2063: }
2064: \end{center}
2065: \caption{Maps of a Galactic Plane region near NGC 6823 in the
2066: Vulpecula constellation derived from BLAST05 observations at
2067: 250$\,\mu$m. The two maps are obtained using two different
2068: methods: SANEPIC (left panel); and simple reprojection after
2069: removing the array average signal at each timestep from all the
2070: timestreams, together with filtering (right panel). Maps are presented in
2071: equatorial coordinates with 15$\arcsec$ pixels. The region mapped
2072: has a size of about $4\,{\rm deg}^2$. Numerous point sources have been
2073: identified in the field (see Chapin et al.~2007), their tell-tale
2074: shape in the map resulting from the PSF of the
2075: BLAST05 optics (see Truch et al.~2007).}
2076: \label{fig:mapL5B250}
2077: \end{figure*}
2078: %\clearpage
2079: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2080:
2081: No residual striping is visible in the map obtained with SANEPIC (left
2082: panel of Figure~\ref{fig:mapL5B250}),
2083: mainly due to the presence of multiple scanning directions for this
2084: field. Large scale structures in the map are successfully recovered
2085: with SANEPIC, as can easily be seen by comparing with the right panel
2086: of Figure~\ref{fig:mapL5B250}. This recovery applies to scales which are
2087: significantly larger than the array size.
2088: However, the resulting effective filtering after applying the
2089: ``array average subtraction'' method induces negative signals near
2090: bright sources in the map, while no such filtering effect is seen in the
2091: SANEPIC map (except perhaps near the edges of the map). This shows
2092: that optimal map-making methods (in the sense of least squares) like
2093: SANEPIC are better suited to recover point sources in the maps as well as
2094: diffuse structures.
2095:
2096:
2097: \section{Conclusions}
2098:
2099: Large format detector arrays operating at far-IR and submillimeter
2100: wavelengths are becoming the norm, rather than the exception.
2101: Ground-based instruments are plagued with common-mode emission arising
2102: from the Earth's atmosphere. And as we have found with BLAST, the
2103: same applies to high frequency balloon-borne instruments, where we see
2104: correlated noise from thermal as well as atmospheric effects. There
2105: is an expectation that even upcoming satellites might be faced with
2106: similar issues, because of thermal variations in the spacecraft, for
2107: example. In general, we expect that correlated noise between
2108: detectors will be a major issue which all such experiments have to
2109: deal with, and we expect that the SANEPIC approach, which we have
2110: described here, will be widely applicable. Indeed, there is evidence
2111: from existing arrays (e.g.,~SHARC-II and AzTEC) that once there are
2112: many detectors, there are {\it multiple\/} correlations between sub-sets of
2113: the detectors, as well as an overall common-mode term. Consequently, one sees
2114: correlations between contiguous blocks of detectors on the array, or
2115: sets of detectors which share amplifiers or are otherwise coupled
2116: through the electronics. Provided that these correlations can be
2117: investigated and their behavior modelled, it is straightforward to
2118: extend the SANEPIC approach to deal with several distinct sources of
2119: correlated noise. Hence we expect the SANEPIC approach to be applicable to
2120: future instruments such as SCUBA-2, SPIRE, ACT, {\sl Planck\/} HFI and
2121: others. There are also many experiments being planned which use large
2122: detector arrays to perform sensitive polarization measurements, and we
2123: see no reason why the SANEPIC approach could not also be
2124: extended to polarimetry.
2125:
2126: \acknowledgments
2127:
2128: The BLAST collaboration acknowledges the support of NASA through grant
2129: numbers NAG5-12785, NAG5-13301 and NNGO-6GI11G, the Canadian Space
2130: Agency (CSA), Canada's Natural Sciences and Engineering Research
2131: Council (NSERC), and the UK Particle Physics \& Astronomy Research Council
2132: (PPARC). We would also like to thank the Columbia Scientific Balloon
2133: Facility (CSBF) staff for their outstanding work.
2134: We are grateful to Matthew
2135: Hasselfield and Duncan Hanson for valuable discussions which contributed
2136: to the development of SANEPIC. LO acknowledges
2137: partial support by the Puerto Rico Space Grant Consortium and by the
2138: Fondo Istitucional para la Investigacion of the University of Puerto
2139: Rico. CBN acknowledges support from the Canadian Institute for
2140: Advanced Research. This research made use of Westgrid computing
2141: resources, which are funded in part by the Canada Foundation for Innovation,
2142: Alberta Innovation and Science, BC Advanced Education, and the participating
2143: research institutions.
2144:
2145: \bibliographystyle{apj}
2146:
2147:
2148: \begin{thebibliography}{}
2149:
2150: \bibitem[{{Ashdown} {et~al.}(2007)}]{ashdown}
2151: Ashdown, M., et al. 2007a, \aap, 467, 761
2152:
2153: \bibitem[{{Ashdown} {et~al.}(2007)}]{ashdown07}
2154: Ashdown, M., et al. 2007b, \aap, 471, 361
2155:
2156: \bibitem[{{Benoit} {et al.}(2002)}]{benoit}
2157: Benoit A., et al, 2002, Astropart. Phys., 17, 101
2158:
2159: \bibitem[{{Borrill}(1999)}]{borrill}
2160: Borrill, J.~D. 1999, preprint, astro-ph/9911389
2161:
2162: \bibitem[{{Chapin} {et~al.}(2007)}]{chapin}
2163: {Chapin}, E.~L.~C. {et~al.} 2007, ApJ, submitted
2164:
2165: \bibitem[{{Crill} {et al.}(2003)}]{crill}
2166: Crill B.P., et al., 2003, ApJS, 148, 527
2167:
2168: \bibitem[{{Delabrouille}(1998)}]{delabrouille98}
2169: Delabrouille, J. 1998, A\&A Supl. 127, 555
2170:
2171: \bibitem[{{Delabrouille}, {Cardoso}, \& {Patanchon}(2003)}]{delabrouille03}
2172: Delabrouille, J., Cardoso, J.~F, Patanchon, G. 2003, \mnras, 346, 1089
2173:
2174: \bibitem[{{Devlin} {et al.}(2004)}]{devlin}
2175: Devlin, M., et al. 2004, Proc. SPIE, 5498, 42
2176:
2177: \bibitem[{{Dor{\'e}} {et~al.}(2001)}]{dore}
2178: Dor{\'e}, O., et al. 2001, \aap, 374, 358
2179:
2180: \bibitem[{{Dupac} \& {Giard}(2002)}]{dupac02}
2181: Dupac, X., Giard, M., 2002, \mnras, 330, 497
2182:
2183: \bibitem[{{Ferreira} \& {Jaffe}(2000)}]{ferreira}
2184: Ferreira, P.~G., Jaffe, A.~H. 2000, \mnras, 312, 89
2185:
2186: \bibitem[{{de Gasperis} {et~al.}(2005)}]{degasperis}
2187: de Gasperis, G., et al. 2005, \aap, 436, 1159
2188:
2189: \bibitem[{{Hamilton}(2003)}]{hamilton}
2190: Hamilton, J.-Ch. 2003, preprint, astro-ph/0310787
2191:
2192: %\bibitem[{{Hargrave} {et~al.}(2007)}]{hargrave}
2193: %{Hargrave}, P. {et~al.} 2007, ApJ, submitted
2194:
2195: \bibitem[{{Hinshaw} {et~al.}(2003)}]{hinshaw}
2196: Hinshaw, G., et al. 2003, \apjs, 148, 63
2197:
2198: \bibitem[{{Janssen} {et~al.}(1996)}]{Janssen}
2199: Janssen, M.~A., et al. 1996, preprint, astro-ph/9602009
2200:
2201: \bibitem[{{Jenness} {et~al.}(1998)}]{jenness}
2202: {Jenness}, T., {Lightfoot}, J.~F., {Holland}, W.~S. 1998, Proc. SPIE, 3357, 548
2203:
2204: \bibitem[{{Keih{\"a}nen} {et~al.}(2005)}]{keihanen}
2205: Keih{\"a}nen, E., Kurki-Suonio, H., Poutanen, T. 2005, \mnras, 360, 390
2206:
2207: \bibitem[{{Mac{\'i}as-P{\'e}rez} {et~al.}(2007)}]{macias}
2208: Mac{\'i}as-P{\'e}rez, J., et al., \aap, 467, 1313
2209:
2210: \bibitem[{{Maino} {et~al.}(2002)}]{maino02}
2211: Maino, D., et al. 2002, \aap, 387, 356
2212:
2213: \bibitem[{{Miville-Desch{\^e}nes} {et~al.}(2007)}]{mamd07}
2214: Miville-Desch{\^e}nes, M.-A., Lagache, G., Boulanger, F., Puget, J.-L.,
2215: 2007, \aap, 469, 595
2216:
2217: \bibitem[{{Natoli} {et~al.}(2001)}]{natoli}
2218: Natoli, P., et al. 2001, \aap, 372, 346
2219:
2220: \bibitem[{{Oliver} {et~al.}(2000)}]{oliver00}
2221: Oliver, S., et al. 2000, \mnras, 316, 749
2222:
2223: \bibitem[{{Pascale} {et~al.}(2007)}]{pascale07}
2224: {Pascale}, E. {et~al.} 2007, ApJ, submitted
2225:
2226: \bibitem[{{Poutanen} {et~al.}(2006)}]{poutanen}
2227: Poutanen, T., et al. 2006, \mnras, 449, 1311
2228:
2229: \bibitem[{{Prunet} {et~al.}(2000)}]{prunet00}
2230: Prunet, S., Netterfield, C.~B., Hivon, E., Crill., B.~P. 2000, in ``Proceedings
2231: of the XXXV Rencontres de Moriond (http://moriond.in2p3.fr/J00/ProcMJ2000/),
2232: preprint, astro-ph/0006052
2233:
2234: \bibitem[{{Prunet} {et~al.}(2001)}]{prunet01}
2235: Prunet, S., et al. 2001, in ``Mining the sky'', ed. A.J. Banday, S. Zaroubi,
2236: M. Bartelmann, ESO Astrophysics Symposia, Springer-Verlag, p$.\,421$
2237:
2238: \bibitem[{{Stompor} {et~al.}(2002)}]{stompor}
2239: Stompor, R., et al. 2002, \prd, 022003
2240:
2241: \bibitem[{{Tegmark}(1997)}]{tegmark}
2242: Tegmark, M. 1997, \apjl, 480, 87
2243:
2244: \bibitem[{{Tegmark} {et~al.}(2000)}]{tegmark00}
2245: Tegmark, M., et al., 2000, \apj, 541, 535
2246:
2247: \bibitem[{{Truch} {et~al.}(2007)}]{truch07}
2248: {Truch}, M. {et~al.} 2007, ApJ, submitted
2249:
2250: \bibitem[{{Wright} {et~al.}(1996)}]{wright}
2251: Wright, E.~L., Hinshaw, G., Bennett, C.~L. 1996, \apjl, 458, 53
2252:
2253: \bibitem[{{Yvon} \& {Mayet}(2005)}]{yvon}
2254: Yvon, D., Mayet, F. 2005, \aap, 436, 729
2255:
2256: \end{thebibliography}
2257:
2258:
2259: \end{document}
2260: