-
Notifications
You must be signed in to change notification settings - Fork 0
/
ms.tex
2436 lines (2151 loc) · 114 KB
/
ms.tex
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
769
770
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
908
909
910
911
912
913
914
915
916
917
918
919
920
921
922
923
924
925
926
927
928
929
930
931
932
933
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
% Prerequisites: Ubuntu 12.04, 12.10 or 13.04, texlive-binaries 2012, texlive-latex-base 2012, texlive-publishers,
% texlive-latex-extra, texlive-fonts-recommended, TeXstudio 2.5.2 (SVN 3733M) or newer, biber, biblatex
% USAGE INSTRUCTIONS:
% Get more information about latex package by using Terminal command texdoc <packagename>
% for example: texdoc rotating or texdoc aastex or texdoc titling
% or visit this website: http://www.texdoc.net/ and search for the packagename
% TeXstudio 2.5.2 (SVN 3733M) binary works a lot better and is more stable than texmaker in Ubuntu 13.04
% Build procedure in TeXstudio 2.5.2 (SVN 3733M) (DO NOT USE texmaker):
% step 1) OPTIONAL: press Tools:Clean within texstudio to delete old output files
% step 2) REQUIRED: press F1 to build and view output PDF
% step 3) REQUIRED: press F12 to generate index file and create index chapter
% step 4) REQUIRED: press F11 to generate bibliography using biber (not bibtex)
% step 5) REQUIRED: press F1 to build and view output PDF
% step 6) REQUIRED: press F1 once more to view more details about bibliography
% and to correctly compile this .tex file into a valid .PDF output file.
% always save graphics files in f<sequence-number>.PDF format, not in .JPG or .PNG format
% use convert command to convert picture to .pdf format:
% example: convert hubble-telescope-photos-mohsin-nazir-mughal-freelance-designer.jpg hubble.pdf
% The tocbibind package can automatically add bibliography, index, TOC, LOF, and LOT to the table
% of contents. It's even capable of using numbered headings instead of the default
% unnumbered ones.
% Use \cite[23]{kennedy1998cationic} for example when citing page 23 from bibliography reference contained in the .bib
% bibtex file
% Use Figure \ref{fig:Array} on page \pageref{fig:Array} for example when referring to a figure at the end of the
% document
% Other handy URL's:
% http://detexify.kirelabs.org/classify.html
% http://www.ottobib.com/
% http://scholar.google.be
% Figure list
% 1) fig:Coverage
% 2) fig:Array
% 3) fig:PointingCalibrators
% 4) fig:PointingModel
% 5) fig:SCUBAPointingComparison
% 6) fig:IterativeMapping
% 7) fig:PCA_Graphical
% 8) fig:Glitches
% 9) fig:Flagger
% 10) fig:Deconvolution
% 11) fig:NoiseVsLongitude
% 12) fig:CalibrationCurves
% 13) fig:Flatfield
% 14) fig:PSF
% 15) fig:PCA_Filter
% 16) fig:Bandpass
% 17) fig:FluxComparison
% 18) fig:SampleImages
% 19) fig:BGPSMontage
% 20) fig:Multiwavelength
% 21) fig:HII
% 22) fig:IRDC
% 23) fig:FillingFactor
% Major Sections
% sec:Introduction
% sec:Observations
% sec:FluxCalibration
% sec:Astrometry
% sec:Mapping
% sec:Noise
% sec:Photometry
% sec:FinalMaps
% sec:Discussion
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
% %
% Bolocam Galactic Plane Survey Description %
% %
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
\RequirePackage[l2tabu, orthodox]{nag}
\documentclass[12pt,preprint]{aastex}
%\headheight 0.5in
\usepackage{rotating}
\usepackage{subfigure}
\usepackage{lscape}
\usepackage{tocbibind}
\usepackage{amsmath}
\usepackage[a4paper]{geometry}
\usepackage{graphicx}
\usepackage{microtype}
%\usepackage{siunitx}
\usepackage{booktabs}
%\usepackage[colorlinks=false, pdfborder={0 0 0}]{hyperref}
\usepackage{cleveref}
\citestyle{aa}
\newcommand{\vect}[1]{\mathbf{#1}}
\newcommand{\xad}{\vect{x}}
\newcommand{\ammonia}{\mbox{{\rm NH}$_3$}}
\newcommand{\form}{H$_2$CO}
\newcommand{\td}{\mbox{$T_D$}}
\newcommand{\jj}[2]{\mbox{$J = #1\rightarrow#2$}}
\newcommand{\vdag}{(v)^\dagger}
\newcommand{\myemail}{{\tt [email protected]}}
\newcommand{\lsim}{{_{<}\atop^{\sim}}}
\newcommand{\gsim}{{_{>}\atop^{\sim}}}
\newcommand{\etal}{{et al.\/}}
\newcommand{\ie}{{i.e.\/}}
\newcommand{\eg}{{e.g.\/}}
\newcommand{\cmq}{cm{$^{-3}$}}
\newcommand{\per}{$^{\rm{-1}}$}
\newcommand{\tc}{{$\theta^1$~Orionis~C}}
%\newcommand{\msol}{M{$_{\odot}$}}
\def\msol{\ifmmode {\>M_\odot}\else {$M_\odot$}\fi}
\newcommand{\lsol}{L{$_{\odot}$}}
\newcommand{\kms}{km~s{$^{-1}$}}
\newcommand{\hii}{H{\sc ii}}
\newcommand{\Hii}{H~{\sc ii}}
\newcommand{\Ha}{\mbox{H$\alpha$}}
\newcommand{\sii}{S~{\sc ii}}
\newcommand{\Feii}{Fe~{\sc ii}}
\newcommand{\oi}{O~{\sc i}}
\newcommand{\nii}{N~{\sc ii}}
\newcommand{\oiii}{O~{\sc iii}}
\newcommand{\mgii}{Mg~{\sc ii}}
\newcommand{\tco}{{$^{13}$CO}}
\newcommand{\CO}{{$^{12}$CO}}
\newcommand{\Tco}{{$^{12}$CO}}
\newcommand{\co}{C{$^{18}$O}}
\newcommand{\Lsol}{L$_{\odot}$}
\newcommand{\Msol}{M$_{\odot}$}
%\newcommand{\C18o}{C$^{18}$O($1\rightarrow 0$)}
\def\Check#1{{\tt #1}}
\newcommand{\mum}{\ensuremath{\mu \mathrm{m}}}
\newcommand{\flux}{flux density}
\newcommand{\solar}{\ensuremath{\odot}}
\newcommand{\lon}{\ensuremath{l}}
\newcommand\jyb{Jy~beam$^{-1}$}
\newcommand\mjysr{\ensuremath{{\rm MJy~sr}^{-1}}}
\newcommand\Ros{R10}
\newcommand{\epsi}{\varepsilon}
% -----------------------------------------------------------------------------
% BGPS specific definitions
\def\bgps{BGPS}
\newcommand{\bgpsarea}{170}
\newcommand{\bcamfwhm}{31\arcsec}
\newcommand{\bcamfwhmeff}{33\arcsec}
\newcommand{\ncores}{8454}
\newcommand{\ncoresapprox}{8400}
\newcommand{\bgpsdepthlow}{11}
\newcommand{\bgpsdepthhigh}{53}
% -----------------------------------------------------------------------------
% Missions
\def\spitzer{{\em Spitzer}}
% -----------------------------------------------------------------------------
\newcommand{\TBD}{{\bf TBD}}
\def\Figure#1#2#3#4{
\begin{figure}[htb]
\epsscale{#4}
\plotone{#1}
\caption{#2}
\label{#3}
\end{figure}
}
\def\FigureTwo#1#2#3#4#5{
\begin{figure}[htb]
\epsscale{#5}
\plottwo{#1}{#2}
\caption{#3}
\label{#4}
\end{figure}
}
\def\Table#1#2#3#4#5{
\begin{deluxetable}{#1}
\tablewidth{0pt}
\tablecaption{#2}
\tablehead{#3}
\startdata
\label{#4}
#5
\enddata
\end{deluxetable}
}
\def\TableNote#1#2#3#4#5#6{
\begin{deluxetable}{#1}
\tablewidth{0pt}
\tablecaption{#2}
\tablehead{#3}
\startdata
\label{#4}
#5
\enddata
#6
\end{deluxetable}
}
\newcommand{\penn}{1}
\newcommand{\casa}{2}
\newcommand{\utexas}{3}
\newcommand{\virginia}{4}
\newcommand{\ucf}{5}
\newcommand{\wisc}{6}
\newcommand{\jpl}{7}
\newcommand{\ubc}{8}
\newcommand{\hawaiihilo}{9}
\newcommand{\hawaiimain}{10}
\newcommand{\arizona}{11}
% End definitions
\slugcomment{DRAFT: \today}
\shorttitle{BGPS}
\shortauthors{Aguirre et al.}
\begin{document}
\newcounter{subfig}
%\newcounter{irritating}
\setcounter{subfig}{1}
\title{The Bolocam Galactic Plane Survey: Survey Description and Data
Reduction}
\author{James E.~Aguirre\altaffilmark{\penn},
Adam G. Ginsburg\altaffilmark{\casa},
Miranda K.~Dunham\altaffilmark{\utexas},
Meredith M. Drosback\altaffilmark{\virginia},
John Bally\altaffilmark{\casa},
Cara Battersby\altaffilmark{\casa},
Eric Todd Bradley\altaffilmark{\ucf},
Claudia Cyganowski\altaffilmark{\wisc},
Darren Dowell\altaffilmark{\jpl},
Neal J. Evans II\altaffilmark{\utexas},
Jason Glenn\altaffilmark{\casa},
Paul Harvey\altaffilmark{\utexas,\casa},
Erik Rosolowsky\altaffilmark{\ubc},
Guy S. Stringfellow\altaffilmark{\casa},
Josh Walawender\altaffilmark{\hawaiihilo}, and
Jonathan P.~Williams\altaffilmark{\hawaiimain}
}
\affil{{$^\penn$}{\it{Department of Physics and Astronomy, University of
Pennsylvania, Philadelphia, PA }}}
\email{\myemail}
\affil{{$^\casa$}{\it{CASA, University of Colorado, 389-UCB, Boulder, CO 80309}}}
\affil{{$^\utexas$}{\it{Department of Astronomy, University of Texas,
1 University Station C1400, Austin, TX 78712}}}
\affil{{$^\virginia$}{\it{Department of Astronomy, University of
Virginia, P.O. Box 400325, Charlottesville, VA 22904}}}
\affil{{$^\ucf$}{\it{Department of Physics, University of Central
Florida}}}
\affil{{$^\wisc$}{\it{Department of Astronomy, University of Wisconsin,
Madison, WI 53706}}}
\affil{{$^\jpl$}{\it{Jet Propulsion Laboratory, California Institute
of Technology, 4800 Oak Grove Dr., Pasadena, CA 91104}}}
\affil{{$^{\ubc}$}{\it{Department of Physics and Astronomy, University
of British Columbia, Okanagan }}}
\affil{{$^{\hawaiihilo}$}{\it{Institute for Astronomy, University of
Hawaii,640 N. Aohoku Pl., Hilo, HI 96720}}}
\affil{{$^{\hawaiimain}$}{\it{Institute for Astronomy, University of
Hawaii, 2680 Woodlawn Drive, Honolulu, HI 96822}}
\newpage}
\begin{abstract}
We present the Bolocam Galactic Plane Survey (BGPS), a 1.1 mm
continuum survey at 33\arcsec\ effective resolution of \bgpsarea\
square degrees of the Galactic Plane visible from the northern
hemisphere. The BGPS is one of the first large area, systematic
surveys of the Galactic Plane in the millimeter continuum without
pre-selected targets. The survey is contiguous over the range $-10.5
\le \lon \le 90.5$, $|b| \le 0.5$. Towards the Cygnus X spiral arm,
the coverage was flared to $|b| \le 1.5$ for $75.5 \le \lon \le 87.5$.
In addition, cross-cuts to $|b| \le 1.5$ were made at $\lon =$~3, 15,
30 and 31. The total area of this section is 133 square degrees.
With the exception of the increase in latitude, no pre-selection
criteria were applied to the coverage in this region. In addition to
the contiguous region, four targeted regions in the outer Galaxy were
observed: IC1396 (9 square degrees, $97.5 \le \lon \le 100.5$, $2.25
\le b \le 5.25$), a region towards the Perseus Arm (4 square degrees
centered on $\lon = 111$, $b=0$ near NGC7538), W3/4/5 (18 square
degrees, $132.5 \le l \le 138.5$) and Gem OB1 (6 square degrees,
$187.5 \le l \le 193.5$).
The survey has detected approximately \ncoresapprox\ clumps over the
entire area to a limiting non-uniform 1-$\sigma$ noise level in the
range \bgpsdepthlow\ to \bgpsdepthhigh\ m\jyb\ in the inner Galaxy.
The BGPS source catalog is presented in a companion paper
\citep[page 23]{kennedy1998cationic}. This paper details the survey observations and
data reduction methods for the images. We discuss in detail the
determination of astrometric and flux density calibration
uncertainties and compare our results to the literature. Data
processing algorithms that separate astronomical signals from
time-variable atmospheric fluctuations in the data time-stream are
presented. These algorithms reproduce the structure of the
astronomical sky over a limited range of angular scales and produce
artifacts in the vicinity of bright sources. Based on simulations, we
find that extended emission on scales larger than about 5\farcm9 is
nearly completely attenuated ($>90\%$) and the linear scale at which
the attenuation reaches 50\% is 3\farcm8. Comparison with other
millimeter-wave data sets implies a possible systematic offset in flux
calibration, for which no cause has been discovered.
This presentation serves as a companion and guide to the public data
release\footnote{\tt
http://irsa.ipac.caltech.edu/Missions/bolocam.html} through NASA's
Infrared Processing and Analysis Center (IPAC) Infrared Science
Archive (IRSA). New data releases will be provided through IPAC-IRSA
with any future improvements in the reduction. The BGPS provides a
complementary long-wavelength spectral band for the ongoing ATLASGAL
and {\it Herschel}-SPIRE surveys, and an important database and
context for imminent observations with SCUBA-2 and ALMA.
\end{abstract}
\keywords{
methods: data analysis --
surveys --
ISM: clouds --
submillimeter: ISM --
stars: formation --
stars: massive
\newpage
}
\tableofcontents
\newpage
\listoffigures
\newpage
\listoftables
\newpage
\section{Introduction}
\label{sec:Introduction}
Millimeter-wavelength continuum surveys of the Galactic plane provide
the most efficient way to find molecular clumps that are the likely
formation sites of massive stars and star clusters. The development of
detector arrays has made blind surveys of large areas possible. Such
surveys bypass the need for selection based on the presence of
embedded stars or star clusters, infrared sources, masers, or radio
continuum emission. In particular, they can locate molecular clumps
before stars form, providing vital information on the initial
conditions of star formation. These surveys can also provide valuable
constraints on the physical properties of the clumps, especially
masses and mean densities, when combined with distance information.
Galaxy-wide surveys are essential for measuring the impacts of the
environment on clump properties and star formation activity. Do clump
properties vary with Galactocentric distance, or with location with
respect to spiral arms? Do they depend on the level of nearby star
formation activity? Answering these questions in our Galaxy will
provide the essential ``ground truth'' required for the analysis of
distant galaxies where individual clouds and clumps are not resolved
and only galaxy-wide average quantities can be measured.
Surveys of the galaxy in low-lying rotational transitions of the CO
molecule \citep{dame01,jackson06} have identified the locations of
large molecular clouds and traced the transition from atomic gas in
the outer galaxy to molecular gas in the inner galaxy. Surveys in CO
have detected a strong preference for GMCs to form along spiral arms
\citep{stark06}. Recent improved measurements of kinematic distances
\citep[e.g.][]{RotationCurve} and accurate distances from VLBI
\citep{reid09} have refined our view of the spiral structure of the
Milky Way. This new increase in data now allows a more complete
investigation of the relationship between spiral density waves and
star formation.
Locating the early stages of star formation is quite difficult, and
often relies on the serendipitous location of a cold, dark cloud which
appears as an IRDC. While dark clouds were known from optically
obscured regions, the ``discovery'' of IRDCs from the MSX (Mid-course
Space Experiment) satellite was presented by \citet{egan98} and
\citet{carey98}. IRDCs have now been extensively cataloged with MSX
\citep{simon06a} and GLIMPSE 8 \mum\ data \citep{peretto09}. In the
past decade, compelling evidence has emerged which suggests that IRDCs
are the pre-cursors to massive stars, and therefore, stellar clusters
(see, e.g. \citet{rathborne06,rathborne08}). IRDCs, however, require the
favorable viewing condition of being in front of a bright, mid-IR
background for detection. Millimeter wave surveys, however, provide
an efficient means of identifying cold, dense dust throughout the
Galaxy independent of the Galactic background, and thus may be seen at
large distances. Moreover, the millimeter-wave emission is optically
thin, allowing the properties of the cloud to be determined.
Millimeter wave surveys such as BGPS are essential for understanding
massive star and cluster formation on a Galactic scale, from the inner
to the outer Galaxy.
Studies of nearby clouds have demonstrated that star formation is far
from uniform over molecular clouds, but is concentrated in unusually
dense regions called clumps or cores, which fill a small fraction of
the area of most clouds \citep{lada91,enoch07,evans09}. Following
\citet{williams00} and \citet{mckee07}, we use the term `core' to
refer to a very dense region destined to form an individual star or
small multiple star system and the term `clump' to refer to a region
likely to form a group or cluster of stars. While clumps forming
massive clusters may be even denser than cores forming low mass stars
\citep{mueller02}, in general, clumps are larger and have a lower mean
density \citep{mckee07}.
To delineate clumps from their parent clouds as traced in CO, line
emission from molecules which are high-density tracers, such as
\ammonia, \form, CS, HCN, etc. are extremely useful. Spectral lines
also provide excellent diagnostics of line-of-sight motions,
temperatures, and densities. However, the interpretation of these gas
tracers is complicated by variations in tracer abundances caused by
freeze-out onto grains, sublimation caused by star formation, and
subsequent complex chemical processing
\citep{vandishoeck98}. Interpretation is further complicated by
uncertainties in optical depths and excitation conditions and the
impacts of radiation fields and shocks. These complications make the
derivation from such molecular line data of column densities, masses,
and other physical properties of star-forming clumps very difficult.
A further limitation of spectral line surveys is that the tracers of
denser gas tend to be weak, and multi-element receiver arrays have only
a modest number of elements, making blind surveys very time-consuming.
Instruments are beginning to be developed to address this limitation,
such as NRAO's K-band Focal Plane Array\footnote{\tt
https://safe.nrao.edu/wiki/bin/view/Kbandfpa/WebHome} and JCMT's
HARP-B \citep{buckle09}.
In contrast, focal-plane arrays containing hundreds of individual
bolometers sensitive to millimeter and sub-millimeter (sub-mm)
radiation are now available. This advance in technology enables blind
surveys of the Galactic plane, providing a uniform inventory of
massive star-forming and starless clumps. Catalogs resulting from
these surveys will provide the data base for subsequent observations
in tracers such as hard-to-excite molecular lines and higher
resolution studies with instruments like CARMA and ALMA. A
significant advantage of continuum observations of dust at long
wavelengths is the low optical depth of the dust \citep{johnstone06},
permitting a straight-forward estimate of the mass of the emitting
region.
Submillimeter observations of Galactic sources have been conducted by
a variety of different groups focusing on small regions. The
Submillimeter Common-User Bolometer Array (SCUBA; \citet{holland99})
on the 15-m James Clerk Maxwell Telescope mapped many significant
regions in the Galactic plane over its 8 year lifetime (see
\citet{difrancesco08} for a summary), but never completed a contiguous
survey of the Galactic plane. The Bolocam instrument has been used on
the 10.4-m Caltech Submillimeter Observatory (CSO\footnote{The Caltech
Submillimeter Observatory is supported by the NSF.}) to map nearby
large molecular clouds at 1.1 mm and identify hundreds of cores
\citep{enoch07}. MAMBO on the 30-m IRAM telescope has been used at
1.2 mm to map both small, nearby cores \citep{kauffmann08} and more
distant regions of more massive star formation
\citep{rathborne06,motte07}. None of these surveys covered a
significant fraction of the Galactic plane.
This situation has begun to change dramatically. Recent wide-area
surveys have been carried out in the submillimeter bands 250 \mum\ -
500 \mum\ by the Balloon-borne Large Aperture Submillimeter Telescope
(BLAST) \citep{olmi09,roy10,netterfield09,chapin08}. This work
provides an unprecedented view of a wide range of star forming
environments and demonstrates the richness of submillimeter datasets,
creating great anticipation for the imminent results from instruments
aboard {\em Herschel}, particularly the SPIRE camera. From the
ground, the ATLASGAL survey \citep{schuller09}, using the LABOCA
instrument at 870 \micron\ with 19\farcs2 resolution on the APEX
telescope in Chile, has covered 95 square degrees in the Galactic
Plane, $-30 \le l \le 11.5$ and $15 \le l \le 21$ with $|b| \le 1$.
The survey will eventually encompass $-60 \le l \le 60$ with $|b| \le
1.5$.
The BGPS is the first millimeter survey of a substantial fraction of
the Galactic Plane in the northern hemisphere. The BGPS maps in the
first quadrant overlap with the VLA Galactic Plane Survey (VGPS) in
H{\sc i} and radio continuum \citep{stil06}, the {\it Spitzer}-GLIMPSE
and GLIMPSE-II fields \citep{benjamin03}, MIPSGAL \citep{carey09}, the
BU-FCRAO $^{13}$CO $J=1 \to 0$ Galactic Ring Survey \citep{jackson06},
and ATLASGAL \citep{schuller09}, among others. The BGPS coverage will
also overlap with the upcoming {\it Herschel} HiGAL \citep{molinari10}
and SCUBA-2/JCMT Plane Surveys \citep{difrancesco08b}. The \bgps\
provides a long-wavelength dust continuum data point, complementing
the spectral energy distribution provided by other surveys. By
covering the both the Galactic center and anti-center regions with a
single instrument, a consistent comparison over the widest possible
range of Galactocentric radius is possible. The BGPS gives a view of
massive star and cluster formation throughout the Galaxy, and provides
an important finder chart for future high-resolution observations with
facilities such as CARMA and ALMA.
This paper presents the imaging data from the Bolocam Galactic Plane
Survey (BGPS). The source catalog is described in a companion paper
(\citet{rosolowsky10}; hereafter \Ros). The outline of this paper is
as follows. Section \ref{sec:Observations} describes the instrument
and the observations. The data analysis is described in Sections
\ref{sec:Astrometry} (astrometry), \ref{sec:Mapping} (the map-making
algorithm), and \ref{sec:Calibration} (flux calibration). Key
discussions are in Section \ref{sec:Photometry}, which explains the
effect of the data processing on the accurate recovery of extended
structure, and in Section \ref{sec:FluxComparison}, which compares the
BGPS flux with other surveys, including a discussion of a possible
systematic offset. The final image data products and the public
release are described in Section \ref{sec:FinalMaps}. We conclude in
Section \ref{sec:Discussion} with a brief discussion of the broad
features of the BGPS in comparison with other surveys.
\section{Observations}
\label{sec:Observations}
We used Bolocam\footnote{{\tt http://www.cso.caltech.edu/bolocam}} to
survey more than \bgpsarea\ square degrees of the northern Galactic
Plane. Bolocam is the facility 144-element bolometer array camera
mounted at the Cassegrain focus of the 10.4-m mirror of the CSO on the
summit of Mauna Kea. We used the filter configuration with a band
center of 271.1~GHz (hereafter 1.1~mm) and fractional bandwidth
$\Delta \nu/\nu = 0.17$ (46 GHz). The detectors are silicon nitride
micromesh absorbers with NTD germanium thermistors, operated at a
temperature of $\sim$250 mK. The Bolocam array field-of-view is
7\farcm5, with individual detectors having nearly Gaussian beams of
\bcamfwhm\ FWHM. The spacing between the individual detector beams on
the sky is 38\arcsec, so the focal plane is not instantaneously fully
sampled. The Bolocam instrument is described in greater detail in
\citet{glenn03}.
The maps presented here were acquired during six separate observing
sessions at the CSO over the course of two years between June 2005 and
September 2007. The observing epochs are given in Table
\ref{tab:Observing}. Bolocam observations were typically scheduled
when $\tau_{225} > 0.06$. Between observing epochs, Bolocam was
removed from its mount at the re-imaged Cassegrain focus and stored
warm. Thus the flux calibration and pointing model were re-computed
for each epoch to allow for variations in the instrument and optics.
The final data products were aligned to the well-constrained pointing
model created from the Epoch V and VI data (Section
\ref{sec:Astrometry}). We found that the flux calibration did not in
fact differ significantly between epochs; see Section
\ref{sec:FluxCalibration} for more details.
Figure \ref{fig:Coverage} on page \pageref{fig:Coverage} shows the coverage of the \bgps. The survey
is contiguous over the range $-10.5 \le \lon \le 90.5, |b| \le 0.5$.
Towards the Cygnus X spiral arm, the coverage was flared to $|b| \le
1.5$ for $75.5 \le \lon \le 87.5$. In addition, cross-cuts to $|b|
\le 1.5$ were made at $\lon = 3, 15, 30$ and 31. The total area of
this section is 133 square degrees. With the exception of the
increase in latitude, no pre-selection criteria were applied to the
coverage in this region. In addition to the contiguous region, four
targeted regions in the outer Galaxy were observed: IC1396 (9 square
degrees, $97.5 \le \lon \le 100.5$, $2.25 \le b \le 5.25$), a region
towards the Perseus Arm (4 square degrees centered on $\lon = 111$,
$b=0$ near NGC7538), W3/4/5 (18 square degrees, $132.5 \le l \le
138.5$) and Gem OB1 (6 square degrees, $187.5 \le l \le 193.5$). The
total area of good coverage for the BGPS is \bgpsarea\ square degrees.
Our basic observing strategy was to raster scan a field by moving the
primary mirror of the CSO. This was done to modulate the
astrophysical signal faster than fluctuations in atmospheric opacity.
Each field was scanned in a pattern with alternating rasters along
lines of constant $l$, followed by a series of rasters along lines of
constant $b$. Fields were observed several times with such raster
scans to improve signal-to-noise (S/N). Thus each point in the final
map of a field was observed multiple times with the array moving in
different directions with respect to the astrophysical emission,
helping to separate it from atmospheric emission (see Section
\ref{sec:Mapping} for further details on this process).
In Epochs I and II, the fundamental observing block was a region
$1\arcdeg \times 1\arcdeg$, covered by 23 rasters along each of \lon\
or $b$. Starting with Epoch III, the fundamental block was changed to
$3\arcdeg \times 1\arcdeg$ for increased mapping efficiency (i.e.,
less time spent in turnarounds at the end of a raster). Each
$3\arcdeg \times 1\arcdeg$ block was covered using either 23 rasters
along lines of constant $b$, or 67 rasters along lines of constant
$l$. In both cases, the spacing between adjacent rasters was
162\arcsec. The total time for observing the entire block was 39 (48)
minutes for rasters along constant $b$ ($l$). The data were
electronically sampled at 10 Hz along the scan direction, slightly
higher than the Nyquist rate for the scan speed of 120\arcsec\
s$^{-1}$.
These fundamental observing blocks were stitched together to make 29
large, contiguous maps. These images define the ``fields'' used for
comparison. Each ``field'' includes all observations covering a
region between 1 and 6 square degrees.
Because the Bolocam detector beams do not fully overlap on the sky, if
the array is scanned along detector rows, gaps will be left in the
resulting map. To ameliorate this, we began using in Epoch II a field
rotator to adjust the rotation angle of the array with respect to the
scan direction so that as the scan proceeded the full extent of the
FOV in the direction orthogonal to the scan was sampled by at least
one detector. This is shown in Figure \ref{fig:Array} on page \pageref{fig:Array}.
Achieving a uniform sampling along the scan-orthogonal direction is further
complicated by a number of missing bolometers. A simulation was
performed to determine the optimal angle to rotate the array with
respect to the scan direction to account for both the effects of the
beam spacing and missing bolometers. During the turnaround following
a raster, the field rotator is adjusted to this optimum angle for the
subsequent raster. Without the field rotator, the coverage shows
variations of 100\% from pixel to pixel in a single raster of a field.
With the rotator, this is reduced to $\sim40\%$.
\section{Astrometry}
\label{sec:Astrometry}
\subsection{Absolute Reference Sources and Pointing Model}
\label{sec:PointingModel}
We constructed an absolute reference system for the BGPS by observing
bright quasars and blazars near the Galactic Plane with small raster
maps. The sources were chosen from the SMA Submillimeter Calibrator
List\footnote{{\tt http://sma1.sma.hawaii.edu/callist/callist.html}},
since they have positions reliably determined from interferometric
measurements and are point sources at the scale of the Bolocam beam.
The distribution of absolute pointing sources over the sky is shown in
Figure \ref{fig:PointingCalibrators}.
Pointing observations were performed approximately once every two
hours over each night. The elevation offsets showed a deviation from
zero which was empirically well-modeled by a quadratic function of
altitude. No systematic deviation was observed for the azimuth
offsets. The derived model is shown in Figure
\ref{fig:PointingModel}. An RMS scatter of $\sim 6\arcsec$ for the
model over the entire observed range was achieved, with somewhat worse
scatter in altitude than in azimuth.
\subsection{Pixel Positions in the FOV}
In addition to the model for the pointing center, it is necessary to
empirically determine the actual projected pattern of the array on the
sky, and measure the rotation angle between the focal plane and the sky
coordinate system. This is done by making observations which track
all detectors across a bright source (e.g. a planet) and making maps
from each bolometer individually.
In our first data release, the projected positions of the detectors on
the sky are made using the nominal positions without correction for
optical distortion. This results in a slightly reduced effective
resolution (\bcamfwhmeff\ instead of \bcamfwhm). The effective beam
size was measured from fully-sampled beam-mapping observations of
planets, and is a good description of the blurring of the beam for the
portions of a field observed by all the detectors. Within about one
FOV of the field edges, however, only a subset of the array observes
the field, and so the effective beam size varies slightly at these
positions, and may be asymmetric. Because this distortion is
asymmetric, it can also result in a small pointing offset. However,
this effect is strongly mitigated by the low number of hits at field
edges: this data is generally flagged out. The residual offset is
found to be within the overall pointing model error, but caution
should be used when measuring source locations at the field edges in
the first data release. Future releases will have the distortion
correction applied, removing this effect and improving the effective
resolution.
\subsection{Relative Alignment and Mosaicking}
A master pointing field was constructed from observations taken in the
epoch with the best-constrained pointing model for that field (either
Epoch V or VI) for every field observed by the \bgps. Relative
alignment between the master and other observations of the same field
was performed by finding the peak of the cross-correlation between
images. The cross-correlation peak was fit with a Gaussian profile and
the difference between the Gaussian peak and the image center was used
as the pointing offset. The offset may be determined accurately to
within the error in finding the peak of the Gaussian, typically
$<1\arcsec$. To create the final image, all observations of a field
were merged into a single timestream with these pointing offsets
applied.
This method of alignment makes use of all the information available in
the maps, and avoids the ambiguities inherent in using extracted
sources to align fields, as the BGPS sources are rarely point-like.
It further avoids the slightly larger effective beam and loss of peak
flux density which would result if the maps of individual observations
were combined using the pointing model alone, where each individual
observation would be co-added with the 6\arcsec\ RMS uncertainty of
the model. In some fields with few sources, particularly those in the
range \lon=65 to 75, there was not enough signal to acquire a pointing
offset using cross-correlation. In these fields, the pointing model
was used directly, and so the above mentioned effects affect these
fields.
\subsection{Comparison to the SCUBA Legacy Catalog}
\label{sec:SCUBAPointingComparison}
All SCUBA 850 and 450 \mum\ data has been re-processed in a uniform
manner and made publicly available as the SCUBA Legacy Catalog
\citep[][hereafter SLC]{difrancesco08}. These maps allow a
cross-check of the accuracy of the BGPS pointing model. We have
applied the same cross-correlation procedure used to obtain the
relative alignment of BGPS observations to compare the BGPS to the SLC
850 \mum\ images. \citet{difrancesco08} claim a typical pointing
accuracy for SCUBA of $\sim3\arcsec$, with offsets as large as
$\sim6\arcsec$ occurring occasionally. No further adjustment to the
nominal JCMT pointing model was performed in producing the SLC maps,
and thus we assume these numbers are typical.
There are a number of differences in the Bolocam and SLC maps which
might lead to difficulties in comparing their astrometry. In
particular, different scan strategies were used (SCUBA observations
were typically taken in jiggle-map mode, which tends to remove
extended structure), and of course there is the difference in
wavelength and beam size between the instruments. Nevertheless, the
by-eye morphological comparison between SLC 850 \mum\ and BGPS sources
is generally excellent, as exemplified in Figure
\ref{fig:SCUBAPointingComparison-a}. A cross-correlation technique
was used in order to minimize these limitations, since it does not
require the establishment of a single position for an object.
The SLC has the advantage of having observations over a large portion
of the Plane, allowing the pointing to be checked (in spots) for
deviations as a function of position. To test this, 21 of the 29 BGPS
fields were used which overlapped with the SLC. For these fields, we
find $\Delta l = 1\farcs8 \pm 1\farcs2$, $\Delta b = -0\farcs4 \pm
0\farcs8$, nearly consistent with zero, and an RMS dispersion about
the mean BGPS position of 6\farcs2, consistent with the errors derived
for the Bolocam pointing model. The offset for each field is shown in
Figure \ref{fig:SCUBAPointingComparison-b}. We have {\it not}
corrected the Bolocam pointing based on the SLC or any external data
set, but have produced our final maps with the internally derived
model described in Section \ref{sec:PointingModel}.
\section{Mapping Algorithm}
\label{sec:Mapping}
At wavelengths near 1 mm, the emission from the atmosphere dominates
any astrophysical emission. For an atmospheric zenith optical depth
of $\tau$ = 0.1, the sky brightness temperature is around 30 to 40
Kelvin at typical zenith angles. This background loading determines
the instrument calibration (see Section \ref{sec:FluxCalibration}),
but the primary time-variable signal is due to few percent
fluctuations in atmospheric opacity. The conversion $J$ from \jyb\
to Kelvin for Bolocam is
\begin{equation}
\label{eq:JyPerK}
J = 1\times10^{26} \frac{2k}{A_{\rm eff}} =
58 \, {\rm \frac{{\rm Jy\,beam}^{-1}}{{\rm K}}}
\end{equation}
where $k$ is Boltzmann's constant and $A_{\rm eff}$ is the effective
collecting area of the telescope. Thus atmosphere fluctuations are
10's of \jyb, as compared to the typical brightness of Galactic
sources, which lie in the range of 0.1 - 1 \jyb. The essential
signal-processing problem to be solved by the mapping algorithm, then,
is the estimation of the astrophysical emission in the presence of the
fluctuating atmosphere signal, without {\it a priori} knowledge of
either.
For this work, a custom pipeline was developed to address issues
specific to the BGPS. It is based on, but significantly improves, the
method of \citet{enoch06} for generating maps and characterizing their
properties, and is similar in many respects to the methods used by
other bolometer cameras, {e.g.}, \citet{cotton09} (MUSTANG/GBT) or
\citet{kovacs08} (SHARC-2/CSO). It also incorporates some of the
technical developments of \citet{sayers09} developed for Bolocam data
at 2.1 mm. A key element of the reduction is the iterative estimation
of the atmospheric fluctuations and the astrophysical signal. The
atmospheric model is developed from a set of principal components of
the bolometer signal time series under the assumption that the bulk of
the correlated signal is atmospheric. After subtraction of the
atmosphere model, the astrophysical emission is estimated, then
subtracted, and the process is repeated.
\subsection{Algorithm}
We assume the raw timestream data $d$ for each bolometer (indexed by
$i$) at discrete time points (indexed by $t$) can be written as
\begin{equation}
\label{eq:TimeDomainModel}
d_{it} = s_{it}+a_{it}+c_{it}+e_{it}+p_{it}+\epsi_{it}
\end{equation}
where $s$ is the astrophysical signal, $a$ is the median atmospheric
fluctuation seen by all detectors, $c$ are terms correlated between
bolometers that are distinct from $a$, $e$ are non-random signals due
to the instrument itself, $p$ are long timescale fluctuations not
modeled by $a$ or $c$, and $\epsi$ is irreducible Gaussian noise due
to photon fluctuations and detector noise. (Each of these terms is
discussed in greater detail in the following sections.)
In the presence of purely Gaussian noise, the maximum likelihood map
$m$ can be obtained by minimizing the goodness-of-fit statistic
\begin{equation}
\label{eq:ChiSquared}
\chi^2 = (d - Am)^T W (d - Am)
\end{equation}
Here $d$ is the concatenated data from all bolometers (of dimension
$N_{tod}$=number time samples of all detectors). $A$ provides the
mapping between a given sample in the time-ordered data and the pixel
in the sky map to which it corresponds. $A$ is a matrix of dimension
$N_{tod} \times N_{pix} =$number of map pixels. Each element of $A$
is either 1 or 0. Note that we are considering the data $d$ and the
map $m$ as vectors (the mapping to a two-dimensional map being
implicit in the matrix $A$). $W$ is an $N_{tod} \times N_{tod}$
matrix which is the inverse of the covariance matrix $N$ of the time
domain noise $\epsi$,
\begin{equation}
\label{eq:NoiseCovariance}
N_{tt'} = \langle \epsi_t \epsi_{t'} \rangle,
\end{equation}
where $\langle \rangle$ denotes an ensemble average over many
realizations of the random time\-stream noise process $\epsi_t$.
Further detail about the way in which $W$ was estimated is given in
Section \ref{sec:Noise}.
The result of the $\chi^2$ minimization is
\begin{equation}
\label{eq:LSMap}
m = (A^T W A) A^T W d
\end{equation}
Thus, the mapping from data to map is a linear operator, $M =(A^T W A)
A^T W$. For compactness, Equation \ref{eq:LSMap} will be written as
$M[d] = m$. Note that the mapping operator is not invertible.
However, given a map $m$ and an observing matrix $A$, there exists a
linear operation which makes predictions about the observed
timestreams, namely $d = A m$; this will be denoted $T[m] = d$.
Because the model $Am$ in Equation \ref{eq:ChiSquared} only includes
the celestial contribution to the observed data $d$, direct use of
Equation \ref{eq:LSMap} will produce a map containing artifacts due to
the unmodeled components $a$, $c$, $e$, and $p$. One approach would
be to include templates for these terms as additional rows in $A$ and
proceed with a simultaneous fit. In general, however, the correct
forms to use are not known {\it a priori}. The goal, then, is to
produce a time series for each bolometer which as closely as possible
approximates $s + \epsi$ so that we may produce the best estimate of
the astrophysical signal $m = M[s+\epsi] = S + N$. We proceed
iteratively, estimating each term in Equation \ref{eq:TimeDomainModel}
in the order of its relative strength. In the following, we denote
the best estimate of a time series $x$ in the $n^{th}$ iteration as
$\tilde{x}^{(n)}$.
In the first iteration, we estimate the largest signal, $a$, according
to the method in Section \ref{sec:AtmosphereModel}, and assume
$\tilde{s}^{(0)}_{it}=\tilde{c}^{(0)}_{it}=\tilde{p}^{(0)}_{it}=0$.
We then form
\begin{equation}
\label{eq:SignalEstimate}
\xi^{(n)}_{it} = d_{it} - \tilde{a}^{(n)}_{it} - \tilde{c}^{(n)}_{it}
- \tilde{e}^{(n)}_{it} - \tilde{p}^{(n)}_{it} \approx s_{it} +
\epsi_{it}
\end{equation}
This is the best estimate of signal plus irreducible noise at
iteration step $n$, and is made into a map
\begin{equation}
M[\xi^{(n)}_{it}] = \tilde{m}^{(n)}
\end{equation}
The current best map $\tilde{m}^{(n)}$ is then deconvolved to provide
a relatively low-noise, smooth map from which to generate a timestream
\begin{equation}
\label{eq:MapToTime}
T[{\cal D}[\tilde{m}^{(n)}]] = \tilde{s}^{(n)}_{it}
\end{equation}
where ${\cal D}$ represents the deconvolution operation. At this
stage, we can subtract both $\tilde{a}^{(n)}_{it}$ and
$\tilde{s}^{(n)}_{it}$ from the original data, and estimate
$\tilde{c}^{(n)}_{it}$, $\tilde{e}^{(n)}_{it}$, and
$\tilde{p}^{(n)}_{it}$. The iterative process begins again with
Equation \ref{eq:SignalEstimate}. An example time series showing the
successive removal of these models is shown in Figure
\ref{fig:IterativeMapping}. We discuss the estimation of each of the
various terms $a$, $c$, $e$, and $p$ in the following sections.
It is useful at each step to produce a ``residual map''
\begin{equation}
\label{eq:ResidualMap}
E = M[d_{it} - \tilde{s}^{(n)}_{it} - \tilde{a}^{(n)}_{it}
- \tilde{c}^{(n)}_{it}
- \tilde{e}^{(n)}_{it}
- \tilde{p}^{(n)}_{it} ]
\end{equation}
This residual map serves as a visualization of the progress of the
iterative method, since in a perfect process it would be map of
$\epsi$.
\subsection{Atmosphere Fluctuation Model}
\label{sec:AtmosphereModel}
Since atmosphere fluctuations are the largest signal in the raw data,
they are modeled first. By construction, the Bolocam beams as they
pass through the dominant layer of atmospheric water vapor a few km
above the telescope are still highly overlapped, and therefore sample
a nearly identical region of atmosphere. The simplest model for the
atmosphere fluctuations makes use of this fact by assuming that the
largest correlated signals between detectors are due to to the
atmosphere fluctuations; this is the term $a$ in Equation
\ref{eq:TimeDomainModel}. We construct a model of the atmosphere
fluctuations which are seen in common by all detectors, which we
denote the ``median atmosphere model''. In each iteration $n$, we
construct the non-atmosphere-fluctuation part of the time series as
\begin{equation}
\tilde{a}^{(n)}_{it} =
d_{it} - \tilde{s}^{(n)}_{it} -
\tilde{c}^{(n)}_{it} - \tilde{e}^{(n)}_{it} - \tilde{p}^{(n)}_{it}
\end{equation}
The median atmosphere model is then simply
\begin{equation}
\label{eq:MeanAtmosphereModel}
\tilde{a}^{(n)}_t =
{\rm median}_i[{\tilde{a}^{(n)}_{it}}]
\end{equation}
Each bolometer's timestream is then fit to the median atmosphere model,
\ie, we find $r_i^{(n)}$ such that $\tilde{a}^{(n)}_{it}$ most closely
approximates $r_i^{(n)} \tilde{a}^{(n)}_t$. This defines the
relative gains of the detectors as
\begin{equation}
\label{eq:RelativeGains}
r_i^{(n)} = \frac{\sum_{t'} \tilde{a}^{(n)}_{it'} \tilde{a}^{(n)}_{t'}}
{\sum_{t'} (\tilde{a}^{(n)}_{t'})^2}
\end{equation}
The $r_i$ converge fairly rapidly. (See the discussion in Section
\ref{sec:FluxCalibration}.)
The simple common mode model of Equation \ref{eq:MeanAtmosphereModel}
does not remove all of the signal correlated at zero time lag between
the bolometers. Some of this remaining correlated signal may also be
atmosphere. \citet{sayers10} found for the 2.1 mm Bolocam data that
the subtraction of a second order spatial polynomial across the array
accounted for nearly all of their residual noise; however, the 1.1 mm
data seem to have additional correlated components that are not
completely described by such a model. Therefore, to further remove
the atmosphere (and any other correlated instrument noise) without a
detailed physical model, we use a principal component analysis (PCA)
as given in \citet{laurent05}. While this method was developed for
point source detection in deep, high-redshift surveys, the technique
works equally well for this survey, with the caveat that it removes
large-scale structure on the size of the array FOV. This effect is
mitigated by the iterative process (the \citet{laurent05} analysis did
not iterate). Figure \ref{fig:PCA_Graphical} shows an example of the
modes removed by the PCA model. These are a graphical visualization
of the terms $c$ in Equation \ref{eq:TimeDomainModel}.
\subsection{Instrument Error Signals}
\label{sec:InstrumentErrors}
Most of the instrument error signals have characteristic features
which aid in their removal. In particular, unlike the atmosphere
emission and correlated signals described above, these signals are
{\it not} correlated between detector channels. The error signals
include the following:
\begin{enumerate}
\item Pickup from the 60 Hz AC power. This appears as narrow lines in
the power spectral density (PSD) of the data. The second harmonic of
60 Hz is aliased via beating against the 130 Hz bolometer bias
frequency into 10 Hz with sidebands split at $\sim1$ Hz. It is
removed by replacing components at these frequencies in the Fourier
transform with Gaussian noise matched to the local mean amplitude.
\item Spikes in voltage due to cosmic ray strikes on the bolometers
(``glitches''). An example of a glitch in the time series is shown in
Figure \ref{fig:Glitches}, along with the extracted distribution.
\item Microphonic pickup due to vibrations of the receiver. The
microphonic effect is due to a change in the capacitive coupling of
the readout wires to the circuit ground, which is converted by the
high impedance of the detectors into a measurable signal, much like a
condenser microphone. The most noticeable microphonics occur at the
end of each scan when the telescope is turning around and the field
rotator is adjusting. This leads to broad spikes in the time series,
whose long decay must be removed from the data, particularly during
the beginnings of scans.
\end{enumerate}
Fortunately, most of the AC powerline pick-up occurs at frequencies
where there is no astrophysical signal, given the beam size and scan
speed (c.f. Figure 1 of \citet{sayers09}). Thus this error is dealt
with by first notch filtering at the line frequencies and then
low-pass filtering the data. Because of the low correlation with
astrophysical signal, this step is performed only once and is not
iterated.
Both glitches and microphonic pickup from the scan turnarounds are
degenerate with astrophysical signal and must be estimated as part of
the iterative process. Glitches are identified as large excursions
from the RMS level after subtraction of the best atmosphere and bright
source model, and the data there is excluded in
subsequent maps. The turnaround microphonics are explicitly modeled
as decaying exponentials at the beginnings and ends of scans.
Residuals from this model may also be removed by the PCA cleaning.