-
Notifications
You must be signed in to change notification settings - Fork 0
/
Copy paththesis.tex
1941 lines (1551 loc) · 221 KB
/
thesis.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
\documentclass[12pt]{article}
\usepackage{amsmath}
\usepackage[top=1in, bottom=1in, left=1.25in, right=1.25in]{geometry}
% \geometry{ papersize={8.5in,11in} }
\usepackage{setspace}
\usepackage{pdfpages}
\setcounter{secnumdepth}{1}
% \usepackage{tocstyle}
% \usetocstyle{standard}
\renewcommand{\contentsname}{\centerline {\normalsize Table of Contents}}
\usepackage{tocloft}
\renewcommand{\cftsecfont}{\normalsize}
\renewcommand{\cftsecleader}{\cftdotfill{\cftsecdotsep}}
\renewcommand\cftloftitlefont{\normalsize\bfseries}
\renewcommand\cftlottitlefont{\normalsize\bfseries}
% \addtocontents{toc}{\vspace{\normalbaselineskip}}
\usepackage{sectsty}
\sectionfont{\normalsize}
\usepackage{enumitem}
\usepackage{natbib}
\usepackage{mathtools}
\usepackage{dsfont}
\usepackage{bm}
%\usepackage{cite}
\usepackage{float}
\usepackage{placeins}
\usepackage{adjustbox}
\usepackage{lscape}
\usepackage{caption}
%\usepackage{graphicx} % Including figure files
\usepackage{amsmath} % Advanced maths commands
\usepackage{amssymb} % Extra maths symbols
\def\dd#1#2{\frac{d #1}{d #2}}
\def\2dd#1#2{\frac{d^2 #1}{d #2^2}}
\def\pd#1#2{\frac{\partial #1}{\partial #2}}
\def\2pd#1#2{\frac{\partial^2 #1}{\partial #2^2}}
\newcommand\sq{\mbox{\rlap{$\sqcap$}$\sqcup$}}%
\newcommand\arcdeg{\mbox{$^\circ$}}%
\newcommand\arcmin{\mbox{$^\prime$}}%
\newcommand\arcsec{\mbox{$^{\prime\prime}$}}%
\newcommand\fd{\mbox{$.\!\!^{\mathrm d}$}}%
\newcommand\fh{\mbox{$.\!\!^{\mathrm h}$}}%
\newcommand\fm{\mbox{$.\!\!^{\mathrm m}$}}%
\newcommand\fs{\mbox{$.\!\!^{\mathrm s}$}}%
\newcommand\diameter{\ooalign{\hfil/\hfil\crcr\mathhexbox20D}}%
\newcommand\degr{\arcdeg}%
\newcommand\Sun{\sun}%
\newcommand\Sol{\sun}%
\newcommand\sun{\odot}%
\newcommand\Mercury{\astro{\char1}}% Mercury symbol, "1"
\newcommand\Venus{\astro{\char2}}% Venus symbol, "2"
\newcommand\Earth{\earth}%
\newcommand\Terra{\earth}%
\newcommand\earth{\oplus}%
\newcommand\Mars{\astro{\char4}}% Mars symbol, "4"
\newcommand\Jupiter{\astro{\char5}}% Jupiter symbol, "5"
\newcommand\Saturn{\astro{\char6}}% Saturn symbol, "6"
\newcommand\Uranus{\astro{\char7}}% Uranus symbol, "7"
\newcommand\Neptune{\astro{\char8}}% Neptune symbol, "8"
\newcommand\Pluto{\astro{\char9}}% Pluo symbol, "9"
\newcommand\Moon{\astro{\char10}}% Moon symbol, "M"
\newcommand\Luna{\Moon}%
\newcommand\Aries{\astro{\char11}}%
\newcommand\VEq{\Aries}% vernal equinox (Aries)
\newcommand\Taurus{\astro{\char12}}%
\newcommand\Gemini{\astro{\char13}}%
\newcommand\Cancer{\astro{\char14}}%
\newcommand\Leo{\astro{\char15}}%
\newcommand\Virgo{\astro{\char16}}%
\newcommand\Libra{\astro{\char17}}%
\newcommand\AEq{\Libra}% autumnal equinox (Libra)
\newcommand\Scorpius{\astro{\char18}}%
\newcommand\Sagittarius{\astro{\char19}}%
\newcommand\Capricornus{\astro{\char20}}%
\newcommand\Aquarius{\astro{\char21}}%
\newcommand\Pisces{\astro{\char22}}%
\newcommand\ion[2]{#1$\;${%
\ifx\@currsize\normalsize\small \else
\ifx\@currsize\small\footnotesize \else
\ifx\@currsize\footnotesize\scriptsize \else
\ifx\@currsize\scriptsize\tiny \else
\ifx\@currsize\large\normalsize \else
\ifx\@currsize\Large\large
\fi\fi\fi\fi\fi\fi
\rmfamily\@Roman{#2}}\relax}%
\newcommand\sbond{\chem@bnd{\@sbnd}}%
\newcommand\dbond{\chem@bnd{\@dbnd}}%
\newcommand\tbond{\chem@bnd{\@tbnd}}%
\graphicspath{{./}{Figures/}}
% Standard journal abbreviations
% Mostly as used by ADS, with a few additions for journals where MNRAS does not
% follow normal IAU style.
\newcommand\aap{A\&A} % Astronomy and Astrophysics
\let\astap=\aap % alternative shortcut
\newcommand\aapr{A\&ARv} % Astronomy and Astrophysics Review (the)
\newcommand\aaps{A\&AS} % Astronomy and Astrophysics Supplement Series
\newcommand\actaa{Acta Astron.} % Acta Astronomica
\newcommand\afz{Afz} % Astrofizika
\newcommand\aj{AJ} % Astronomical Journal (the)
\newcommand\ao{Appl. Opt.} % Applied Optics
\let\applopt=\ao % alternative shortcut
\newcommand\aplett{Astrophys.~Lett.} % Astrophysics Letters
\newcommand\apj{ApJ} % Astrophysical Journal
\newcommand\apjl{ApJ} % Astrophysical Journal, Letters
\let\apjlett=\apjl % alternative shortcut
\newcommand\apjs{ApJS} % Astrophysical Journal, Supplement
\let\apjsupp=\apjs % alternative shortcut
% The following journal does not appear to exist! Disabled.
%\newcommand\apspr{Astrophys.~Space~Phys.~Res.} % Astrophysics Space Physics Research
\newcommand\apss{Ap\&SS} % Astrophysics and Space Science
\newcommand\araa{ARA\&A} % Annual Review of Astronomy and Astrophysics
\newcommand\arep{Astron. Rep.} % Astronomy Reports
\newcommand\aspc{ASP Conf. Ser.} % ASP Conference Series
\newcommand\azh{Azh} % Astronomicheskii Zhurnal
\newcommand\baas{BAAS} % Bulletin of the American Astronomical Society
\newcommand\bac{Bull. Astron. Inst. Czechoslovakia} % Bulletin of the Astronomical Institutes of Czechoslovakia
\newcommand\bain{Bull. Astron. Inst. Netherlands} % Bulletin Astronomical Institute of the Netherlands
\newcommand\caa{Chinese Astron. Astrophys.} % Chinese Astronomy and Astrophysics
\newcommand\cjaa{Chinese J.~Astron. Astrophys.} % Chinese Journal of Astronomy and Astrophysics
\newcommand\fcp{Fundamentals Cosmic Phys.} % Fundamentals of Cosmic Physics
\newcommand\gca{Geochimica Cosmochimica Acta} % Geochimica Cosmochimica Acta
\newcommand\grl{Geophys. Res. Lett.} % Geophysics Research Letters
\newcommand\iaucirc{IAU~Circ.} % IAU Cirulars
\newcommand\icarus{Icarus} % Icarus
\newcommand\japa{J.~Astrophys. Astron.} % Journal of Astrophysics and Astronomy
\newcommand\jcap{J.~Cosmology Astropart. Phys.} % Journal of Cosmology and Astroparticle Physics
\newcommand\jcp{J.~Chem.~Phys.} % Journal of Chemical Physics
\newcommand\jgr{J.~Geophys.~Res.} % Journal of Geophysics Research
\newcommand\jqsrt{J.~Quant. Spectrosc. Radiative Transfer} % Journal of Quantitiative Spectroscopy and Radiative Transfer
\newcommand\jrasc{J.~R.~Astron. Soc. Canada} % Journal of the RAS of Canada
\newcommand\memras{Mem.~RAS} % Memoirs of the RAS
\newcommand\memsai{Mem. Soc. Astron. Italiana} % Memoire della Societa Astronomica Italiana
\newcommand\mnassa{MNASSA} % Monthly Notes of the Astronomical Society of Southern Africa
\newcommand\mnras{MNRAS} % Monthly Notices of the Royal Astronomical Society
\newcommand\na{New~Astron.} % New Astronomy
\newcommand\nar{New~Astron.~Rev.} % New Astronomy Review
\newcommand\nat{Nature} % Nature
\newcommand\nphysa{Nuclear Phys.~A} % Nuclear Physics A
\newcommand\pra{Phys. Rev.~A} % Physical Review A: General Physics
\newcommand\prb{Phys. Rev.~B} % Physical Review B: Solid State
\newcommand\prc{Phys. Rev.~C} % Physical Review C
\newcommand\prd{Phys. Rev.~D} % Physical Review D
\newcommand\pre{Phys. Rev.~E} % Physical Review E
\newcommand\prl{Phys. Rev.~Lett.} % Physical Review Letters
\newcommand\pasa{Publ. Astron. Soc. Australia} % Publications of the Astronomical Society of Australia
\newcommand\pasp{PASP} % Publications of the Astronomical Society of the Pacific
\newcommand\pasj{PASJ} % Publications of the Astronomical Society of Japan
\newcommand\physrep{Phys.~Rep.} % Physics Reports
\newcommand\physscr{Phys.~Scr.} % Physica Scripta
\newcommand\planss{Planet. Space~Sci.} % Planetary Space Science
\newcommand\procspie{Proc.~SPIE} % Proceedings of the Society of Photo-Optical Instrumentation Engineers
\newcommand\rmxaa{Rev. Mex. Astron. Astrofis.} % Revista Mexicana de Astronomia y Astrofisica
\newcommand\qjras{QJRAS} % Quarterly Journal of the RAS
\newcommand\sci{Science} % Science
\newcommand\skytel{Sky \& Telesc.} % Sky and Telescope
\newcommand\solphys{Sol.~Phys.} % Solar Physics
\newcommand\sovast{Soviet~Ast.} % Soviet Astronomy (aka Astronomy Reports)
\newcommand\ssr{Space Sci. Rev.} % Space Science Reviews
\newcommand\zap{Z.~Astrophys.} % Zeitschrift fuer Astrophysik
\def\av#1{\left\langle #1 \right\rangle}
\def\braket#1#2{\left\langle#1|#1\right\rangle}
\def\opbraket#1#2#3{\left\langle #1\left|#2\right|#3\right\rangle}
\def\ket#1{\left|#1\right\rangle}
\def\cev#1{\reflectbox{\ensuremath{\vec{\reflectbox{\ensuremath{#1}}}}}}
\def\beq{\begin{equation}}
\def\eeq{\end{equation}}
\def\beqs#1\eeqs{\beq\begin{split} #1 \end{split}\eeq}
\usepackage{graphicx}
\graphicspath{{./}{Figures/}}
\usepackage{deluxetable}
%% Units
\def\fm {\,{\tt fm}}
\def\MeV {\,{\tt MeV}}
\def\GeV {\,{\tt GeV}}
\def\degC{\,{^\circ{\tt C}}}
\def\degK{\,{\tt K}}
\usepackage{microtype}
\usepackage[colorlinks=true,backref=false, linktocpage=true,
citecolor=blue,urlcolor=blue,linkcolor=blue,pdfpagemode=UseOutlines]{hyperref}
\hypersetup{%
bookmarksnumbered=true,
pdftitle = {Characterizing the Radio Transient Sky},
pdfsubject = {Physics},
pdfauthor = {Sarah I Chastain},
pdfkeywords = {radio astronomy, transients, Monte-Carlo simulations}
}
\begin{document}
\thispagestyle{empty}
\vspace*{45pt}
\begin{center}
\center{Characterizing the Radio Transient Sky\vspace*{36pt}}
\center{Sarah I. Chastain\vspace*{48pt}}
\centerline{M.S. in Physics, May 2021, George Washington University}
\centerline{B.S. in Physics, May 2017, University of Memphis}\vspace*{24pt}
\center{A Dissertation submitted to\vspace*{36pt}}
\begin{center}
The Faculty of\\The Columbian College of Arts and Sciences \\ of The George
Washington University\\ in partial fulfillment of the requirements\\ for the degree
of Doctor of Philosophy\\[36pt]\vspace*{12pt}
May 21, 2023\\[36pt] %date of degree conferral
Dissertation directed by\\[\baselineskip]
Alexander J. van der Horst\\
Associate Professor of Physics\\[\baselineskip]
%Name of Co-director\\
%Position
\end{center}
\end{center}
% display page numbers in the footer and centered. Start with roman numerals %
\pagestyle{plain}
\setcounter{page}{1}
\pagenumbering{roman}
\newpage
\begin{doublespace}
\noindent
The Columbian College of Arts and Sciences of The George Washington University certifies that Sarah I. Chastain has passed the Final Examination for the degree of Doctor of Philosophy as of \texttt{April 6, 2023}. This is the final and approved form of the dissertation.
\end{doublespace}
\vspace*{24pt}
\begin{center}
Characterizing the Radio Transient Sky
\vspace*{36pt}
Sarah I. Chastain
\vspace*{36pt}
\end{center}
Dissertation Research Committee:
\vspace{12pt}
\indent Alexander J. van der Horst, Associate Professor of Physics, Dissertation Director
\vspace{12pt}
%\indent Name of Co-director, Position, Dissertation Co-Director
%\vspace{12pt}
\indent Oleg Kargaltsev, Associate Professor of Physics, Committee Member
\vspace{12pt}
\indent Bethany Cobb Kung, Associate Professor of Honors and Physics, Committee
\indent Member
\vspace{24pt}
% \noindent Examiners:
% \vspace{12pt}
% \indent Michael D\"oring, Associate Professor of Physics
% \vspace{12pt}
% \indent Sarah Burke-Spolaor, Assistant Professor of Physics and Astronomy, West Virginia University
%only readers and directors are listed per CCAS. No examiners or chairs.
% % \newpage
% % \phantomsection \label{dedication}
% % \begin{center}
% % \section*{\textbf{Dedication}}
% % \end{center}
% % \addcontentsline{toc}{section}{\textbf{Dedication}}
% % \begin{center}
% % \vspace*{6pt}
% % %\vspace*{-24pt}
% % %Dedication text goes here.
% \end{center}
\doublespacing
\newpage
\phantomsection \label{acknowledgements}
\begin{center}
\section*{Acknowledgments}
\end{center}
\vspace*{6pt}
\addcontentsline{toc}{section}{\textbf{Acknowledgements}}
\doublespacing
This dissertation is based in part (Chapter 2) on the previously published article, \citet{2022A&C....4000629C}. I have permission from my publishers and co-authors to use the work listed below in my dissertation.
This work was made possible through a collaboration with other researchers, many of whom are a part of the ThunderKAT collaboration. The author would like to acknowledge their assistance in making the project possible and refining the content of the thesis itself.
Additionally, the author would like to thank the individuals who made it possible for her to make it past the first few years of graduate school and to the thesis defense. In particular, Dr. Evie Downie for her advice that helped the author make it through some particularly difficult years.
The author would like to thank her thesis advisor Dr. Alexander J. van der Horst whose support, guidance, and care made this project possible, all while making it a thoroughly enjoyable project!
\newpage
\begin{center}
\section*{Abstract of Dissertation}
\end{center}
\phantomsection \label{abstract}
\addcontentsline{toc}{section}{\textbf{Abstract of Dissertation}}
\vspace*{-40pt}
\vspace{24pt}
\begin{singlespace}
\center{Characterizing the Radio Transient Sky}
\end{singlespace}
\vspace{-12pt}
\vspace*{24pt}
Since the beginning of history, humans have looked up to the sky for possible changes and their significance. Transient phenomena, that is, phenomena that appear for a time and then disappear, have always been a particular curiosity. We now know a bit more about these phenomena, but they nonetheless continue to be a source of amazing extremes. We know some of these transients to be merging neutron stars, collapsing massive stars, neutron stars with extremely high magnetic field strength, matter accreting onto black holes, flares from giant planets and stars, and much more. The study of these phenomena has been undergoing a shift recently: advances in radio interferometers such as the MeerKAT observatory in South Africa have allowed for the radio regime to be studied in unprecedented detail. MeerKAT has a field of view of almost two square degrees at an observing frequency of 1.3~GHz. This capability, together with its great sensitivity, opens the possibility for finding transients and variables in image searches in the radio regime, a task previously only possible in the optical, x-rays and gamma-rays. With these searches, parameters such as the transient surface density and the transient rate can be found for the radio sky, which is useful for population studies and finding counterparts to rare transients. Furthermore, this opens the possibility for finding new phenomena and understanding existing phenomena at a deeper level. In addition, radio continues to be essential in the follow-up observations of transient phenomena such as gravitational wave sources and short gamma-ray bursts, due to the ability for radio observations to uncover the physics of the outflows and environments of these extreme events. This thesis presents follow-up searches of short gamma-ray bursts, commensal transient searches in observations at various timescales, and detailed simulations to accurately determine transient rates from such commensal searches.
\doublespacing
\newpage
\tableofcontents
\newpage
\cleardoublepage
%\fi
%\iffalse
\phantomsection \label{listoffig}
\begin{center}
% \addcontentsline{lof}{section}{\vspace*{24pt}}
% \cftaddtitleline{lof}{section}{\vspace*{12pt}}{}
\addcontentsline{toc}{section}{\textbf{List of Figures}}
\listoffigures
\end{center}
\newpage
%\fi
%\iffalse
\cleardoublepage
\phantomsection \label{listoftab}
\begin{center}
% \addcontentsline{toc}{section}{\vskip24pt}
% \cftaddtitleline{lot}{section}{\vspace*{12pt}}{}
\addcontentsline{toc}{section}{\textbf{List of Tables}}
% \vskip12pt
\listoftables
\end{center}
\newpage
%\fi
%
%\begin{center}
%\section*{\textbf{List of Symbols}}
%\end{center}
%\addcontentsline{toc}{section}{\textbf{List of Symbols}}
%
%\subsection*{General Units}
%\begin{itemize}
% \item[]pc Parsec ($3.26$ly)
% \item[]ly Lightyear ($9.461\times 10^{17}$cm)
% \item[]$c$ Speed of Light
% \item[]$M_{\odot}$ Solar Mass ($1.99\times10^{33}$g)
% \item[]$R_{\odot}$ Solar Radius ($6.96\times10^{10}$cm)
% \item[]$L_{\odot}$ Solar Luminosity ($3.826\times10^{33}$erg/s)
% \item[]foe $10^{51}$erg
%\end{itemize}
%
%\newpage
%
%\iffalse
%\section {\protect \centering Glossary of Terms(optional)}
%\doublespacing
%\begin{itemize}
%\addtolength{\itemindent}{0.25in}
%\item[Term 1]: Start Definiton here
%\item[Term 2]: Start Definiton here
%\end{itemize}
%\newpage
%\pagenumbering{arabic}
%\fi
%----------------------------------------------------------------------------------------
% ARTICLE CONTENTS
%----------------------------------------------------------------------------------------
\doublespacing
\setcounter{page}{1}
%\vspace*{-2cm}
\pagenumbering{arabic}
\section{Introduction}
\label{introlabel}
A burst of gamma-rays from merging neutron stars, a star exploding at the end of its lifetime, a bright flare caused by a star's extremely strong magnetic field: these are all examples of astrophysical transients. Transient phenomena like these examples can be seen in all parts of the electromagnetic spectrum, from low frequency radio using arrays of dipole antennae to the gamma-ray detectors in space at GeV energies and on the ground at TeV energies. Just as early scientists like Tycho Brahe and Johannes Kepler observed what appeared to be a ``new star'' in the sky~\citep{1969dnen.book.....B,1606dsnip.book.....K}, we are filled with wonder and scientific inspiration by the perplexing, changing sky above us. With the history of the study of astrophysical transients being almost as old as human written history itself, new advances require new instruments and techniques to propel the field towards new discoveries.
These new instruments need to have a large view of the sky, also known as the field of view, and good sensitivity to detect even the faintest sources of light. These requirements made it for a long time only possible to find many of these phenomena in certain parts of the electromagnetic spectrum, such as in gamma-rays, X-rays, and optical wave bands. Now, with upgraded radio facilities, searching for transients in the radio regime has become easier and more promising than ever before~\citep{2011BASI...39..315F}.
Observing transient phenomena at radio wavelengths allows for probing some of the most extreme environments in the universe. Radio is an ideal regime for studying the synchrotron emission from high-energy phenomena, probing the fastest ejecta and with less of an absorption effect due to interstellar dust in the Milky Way galaxy when compared to the optical and X-ray regimes. In addition, for many objects the shape of the synchrotron spectrum can be well defined and tracked over time in the radio~\citep[e.g.,][]{2014PASA...31....8G}. In other cases, low-frequency radio observations may probe coherent emission from transients \citep[e.g.,][]{2015MNRAS.446.3687P}. It is possible that there may be many more kinds of transient objects that are not found yet. With the coming of new large radio interferometers, not only will the objects we already know about be studied in more detail, but we will also be able to study objects that may not have previously been characterized. These discoveries and observations will expand our understanding of transient phenomena and their environments.
\subsection{Searching for Transients}
Due to recently built and upgraded facilities in the radio regime, there is a tremendous opportunity for finding and studying transient phenomena. Figure~\ref{murphy2017} \citep{2017MNRAS.466.1944M} shows how complicated the observational parameter space is when searching for transients: one has to consider the sensitivity of the observations, the timescale or approximate duration of the transients, the observing frequency within the radio band, and the shape of the rise and fall of the emission, otherwise known as the light curve. In this figure, we can see how various transient searches in radio have been able to place constraints on the possible number of times a transient will be seen in a given part of the sky. This number is called the transient surface density and is shown on the vertical axis. From this plot we also see how different surveys place different limits on this quantity depending on the assumed timescale or duration of the transient (left-hand side of Figure~\ref{murphy2017}) and the sensitivity of the observations that were taken (right-hand side of the figure).
\begin{figure}
\includegraphics[width=\textwidth]{rates.png}
\caption{These plots show the limits on the transient rate observed in radio surveys. The plot on the left shows the transient rate surface density for the timescale probed by the survey; the plot on the right shows the transient surface density for the sensitivity of the survey \citep{2017MNRAS.466.1944M}. Various transient searches around 1 GHz have probed a portion of parameter space, but there is much more that can be done. The lines for the various kinds of objects are from \citet{2015ApJ...806..224M}. These objects are magnetars, tidal disruption events (TDE), long gamma-ray burst events (LGRB), off-axis short gamma-ray burst events (SGRB), and neutron star mergers resulting in a black hole (NSM).}
\label{murphy2017}
\end{figure}
Figure~\ref{dario_var_lum} \citep{2015MNRAS.446.3687P} shows the kinds of objects that may be variable or transient in the radio regime. The horizontal axis shows the observing frequency multiplied by the pulse width, the latter being a proxy for the transient duration, and the vertical axis shows the spectral luminosity. Overplotted are dotted lines for different brightness temperatures: the brightness temperature is the temperature of a black body with its peak at the observing frequency\footnote{In the Rayleigh-Jeans limit, this can be represented as $I_{\nu}=\frac{2\nu^2}{c^2}kT_b$, where $T_b$ is the brightness temperature, $c$ is the speed of light, $\nu$ is the observing frequency, and $I_{\nu}$ is the intensity at a given observing frequency. See e.g. \citet{1986rpa..book.....R} for more information.}. As can be seen, different types of transients can be grouped together on this plot. On the shortest timescales are astrophysical phenomena such as fast radio bursts (FRBs), extremely short (about ten milliseconds) and bright bursts of radio emission of which the origins are still mysterious and being investigated~\citep{2021ApJS..257...59C}; and pulsars, rapidly rotating neutron stars which emit bright pulses in radio wavelengths. On longer timescales, we see phenomena like gamma-ray bursts (GRBs), mass-accreting supermassive black holes known as active galactic nuclei (AGN), supernovae from exploding stars, and X-ray binaries from stars that are in orbit with a compact object like a black hole or neutron star.
\begin{figure}
\begin{center}
\includegraphics[width=0.8\textwidth]{dario_var_lum.png}
\caption{From \cite{2015MNRAS.446.3687P}, spectral luminosity plotted against the observing frequency multiplied by pulse width. The diagonal lines represent different brightness temperatures.}
\label{dario_var_lum}
\end{center}
\end{figure}
Figure~\ref{dario_var_lum} \citep{2015MNRAS.446.3687P} shows that by observing at various frequencies and time scales, many different kinds of objects can be probed and perhaps some of the gaps can be filled. Furthermore, moving to lower frequencies and/or shorter duration allows for observing a broader diversity of objects, including observing coherent emission processes. This coherent emission is emission in which the properties of the electromagnetic waves are uniform, usually generated in very small and highly structured environments. At very short timescales, i.e., shorter than about 1 second, it is difficult to make a radio image, and as a result many coherent emission transients are observed using time series analysis in a high time resolution mode.
On longer timescales it is possible to make radio images that are sensitive to the faintest of emission across multiple frequencies. This kind of detail allows for observing and characterizing synchrotron emission. Synchrotron emission is created when electrons are accelerated to relativistic speeds, producing a predictable pattern of emission over time and across the radio frequencies. This kind of emission is possible when powerful outflows and jets interact with their environments, creating shocks. By studying the exact way in which the emission behaves as a function of time and frequency, we can learn many different properties of the exact microphysics within these shocks, the global energetics, and the density and structure of the surrounding material in the environment (or medium).
In this thesis, we focus on studying transients in radio images, which means that most of the transients we would usually expect to observe would have some sort of synchrotron emission. However, with the powerful new telescopes at our disposal we are able to make images at shorter and shorter timescales, down to even seconds. As can be seen in Figure~\ref{dario_var_lum}, this part of parameter space has very few types of known transients, and the transients that are on these short timescales include the very bright FRBs. At these short timescales, if the instantaneous sensitivity and image fidelity of the telescope is sufficient, it may be possible to detect a very bright event such as an FRB even if its duration is actually much shorter than the integration time, or the time that the radio telescope spends looking in one place. Such a detection would be possible if the transient is so bright that even after averaging out the emission over the integration time, it can still be detected. Exciting developments in radio observatories and time-domain astronomy make this an ideal moment in time to study radio transients and search for possible new types of transients.
\subsubsection{New Radio Telescopes}
Up until fairly recently, transient searches were mostly practical in the optical, X-ray or gamma-ray regime. The ability to do transient searches in radio has come with upgrades to existing facilities, such as the improved sensitivity and frequency coverage of the Karl G. Jansky Very Large Array \citep[VLA or JVLA;][]{2011ApJ...739L...1P}; and the rise of new radio facilities, such as the MeerKAT Karoo Area Telescope \citep[MeerKAT;][]{2016mks..confE...1J}, with its wide field of view and unparalleled sensitivity in its observing bands; the Low Frequency Array \citep[LOFAR;][]{2013A&A...556A...2V}, with excellent sensitivity in low radio frequencies and capabilities for resolving very small objects by using the antennae spanning the European continent; the Murchison Widefield Array \citep[MWA][]{2013PASA...30....7T}, with its excellent sensitivity at low frequencies and wide field of view; and the Australian Square Kilometer Array Precursor \citep[ASKAP;][]{2014PASA...31...41H}, with its wide field of view and excellent survey speed, just to name a few. Another crucial component to this new era of radio astronomy is the fact that computers have become advanced enough to process the immense amount of data produced by these advanced telescopes. For example, the Square Kilometer Array (SKA) \citep{2009IEEEP..97.1482D}, for which many of the aforementioned telescopes are pathfinders, is expected to need exascale-level computing power and efficiency to operate effectively, that is computing power capable of performing $10^{18}$ double precision floating point operations per second \citep{doi:10.1177/1094342014549059}.
The upgraded VLA \citep{2011ApJ...739L...1P} began observing in 2012 and provided a vast improvement over the previous generation. It allowed for increased sensitivity, broader observing frequency bands, and vastly improved correlators and other hardware and software, allowing for better flagging and finer frequency resolution. Sensitivity increases help a great deal for transient follow-up observations. For instance, for gamma-ray bursts (GRBs), the VLA is used to follow up on detections made in other bands. Observing in the radio can provide a much clearer picture of the physics behind these events (see also the next section), and the upgrade makes that possible for a much larger sample than before (which was sensitivity limited \cite{Chandra_2012}). The VLA was a valuable resource in following up on the first electromagnetic counterpart to a gravitational wave event, GW170817, an event that was quite remarkable in the amount of new knowledge gained in several areas of physics. The peak flux of this event was at the border of what the old VLA could have observed. After this detection with the VLA, I participated in a collaboration using the VLA to follow up on gravitational wave events. This resulted in an observation for which we obtained upper limits~\citep{2019GCN.26527....1C}. In addition to following up on gravitational waves, using the VLA~\citep{2021Natur.589..207R}, I also assisted with the search for a radio counterpart to an extragalactic magnetar, a neutron star with an extremely strong magnetic field that was detected in the gamma-rays. Both of these projects are not part of the thesis presented here, since it focuses on advances made with the new MeerKAT observatory.
MeerKAT \citep{2016mks..confE...1J} is a radio interferometer in South Africa that has recently come online (summer of 2018), will be part of SKA Phase 1, and observes at mainly in the L~band (856-1711 MHz), but has recently also started observing in the UHF (544-1087 MHz) and S~bands (1750-3499 MHz). With a field of view of almost two square degrees at 1.3~GHz, and a very good image fidelity at short timescales, MeerKAT provides a unique opportunity to find transients in this observing band. It is also well suited for transient follow up since it has better sensitivity and image quality than the VLA between 1 and 2 GHz.
In light of the excellent capabilities of MeerKAT, ThunderKAT \citep{2016mks..confE..13W} was proposed as a Large Survey Project on MeerKAT, with the goal of studying radio transients. Many important results have come from observations by the ThunderKAT collaboration, such as follow-up observations of several X-ray binaries showing jets \citep[e.g.,][]{2020NatAs...4..697B}; new transient discoveries in commensal searches, such as in~\citet{2020MNRAS.491..560D} and~\citet{2022MNRAS.513.3482A}; discoveries of variable sources and limits on transient rates such as in~\citet{2022MNRAS.517.2894R} and ~\citet{commensal1}, and many more. The wealth of deep observations taken by MeerKAT is ideal for studying and searching for radio transients like never before, even engaging citizen scientists in new ways as well~\footnote{https://www.zooniverse.org/projects/alex-andersson/bursts-from-space-meerkat/}.
\subsection{Following Up on Transients}
In addition to searching for transients, the radio regime is used to follow up on transient phenomena that have been discovered in other parts of the electromagnetic spectrum. GRBs are prime examples of transients followed up in the radio and their value to our understanding of high-energy astrophysical phenomena.
\subsubsection{Gamma-Ray Burst Afterglows}
GRB afterglows were discovered to follow the prompt gamma-ray emission of GRBs in 1997, and the first radio afterglow emission was found for GRB~970508 \citep{1997Natur.389..261F}. GRB afterglows have a distinct synchrotron spectrum with multiple segments and characteristic frequencies \citep{1998ApJ...497L..17S,1999ApJ...523..177W}. These characteristic frequencies correspond to the minimum energy for the electron energy distribution, the synchrotron self-absorption frequency, and the cooling frequency of the most energetic electrons. The characteristic frequencies and the peak flux evolve with a certain time dependence and typically follow the Blandford-McKee solution for a relativistic blast wave at early times \citep{1976PhFl...19.1130B}. The behavior of the spectra can reveal a number of physical properties of the collimated outflow, or jet, produced in the GRB explosion, the emission processes at play, and the environment. Radio follow up is essential for getting a clear picture of the long term behavior, since the afterglow is visible in radio long after the optical and X-ray emission have disappeared; and in some cases, even until the non-relativistic phase \citep{1959sadm.book.....S, doi:10.1098/rspa.1950.0050}, such as for GRB~030329 \citep{2008A&A...480...35V}. In particular in the radio, the breaks in the synchrotron spectrum related to the minimum energy of the electron energy distribution and the self-absorption are more clearly detectable than in other bands \citep{2014PASA...31....8G}. In addition, the effects of both the forward and reverse shock are visible for a much longer time. The ability to observe the reverse shock can prove valuable for learning more about the properties of the jet itself, since the reverse shock is propagating through it \citep{2014MNRAS.444.3151V, 2013ApJ...776..119L, 2014ApJ...781...37P}.
From multi-wavelength observations, in which the radio band plays a crucial role, we can determine physical parameters of the explosion (e.g., the energy), the environment (e.g., the ambient medium density), and the electrons and magnetic fields necessary to produce the synchrotron emission \citep{2014PASA...31....8G}. These parameters can be constrained for both long and short GRBs. Long GRBs are those which have $T_{90}>2$ seconds, where $T_{90}$ is the time during which 90\% of the gamma-ray counts are detected above background \citep{1993ApJ...413L.101K}, and short GRBs have $T_{90}<2$ seconds. Long GRBs are associated with the collapse of a massive star \citep{1998Natur.395..670G, 2003Natur.423..847H, 1993ApJ...405..273W} and short GRBs with the merger of two compact objects \citep{1989Natur.340..126E, 1992ApJ...395L..83N,2017PhRvL.119p1101A}. While radio detections of afterglows from long GRBs have been occurring with regularity for some time now, radio detections of afterglows from short GRBs are still quite rare~\citep{2015ApJ...815..102F}. It is still not completely clear why it is so challenging to detect these afterglows in radio, i.e., which physical parameters drive the radio dimness and overall dimness of short GRBs. More radio detections will be important for pinning down the causes for this.
\subsection{Structure of the Thesis}
In this thesis I will present follow-up searches of short GRBs, commensal transient searches in observations at various timescales, and detailed simulations to accurately determine transient rates from such commensal searches.
In Chapter 2 we will present accurate transient rate calculations by using Monte-Carlo simulations. In particular, we show how starting with either a real or simulated observational setup, the simulations code calculates a transient rate as a function of transient duration and peak flux. These simulations allow for replicating a wide variety of realistic scenarios including observations with varying sensitivities and durations, multiple overlapping telescope pointings, and a wide variety of light curve shapes; and this simulations toolkit is easily adaptable for a variety of different science cases.
In Chapter 3 we report on a commensal search in deep observations of short gamma-ray burst fields carried out with the MeerKAT radio telescope. These four-hour observations of eight different fields span survey lengths of weeks to months. We also carry out transient searches in time slices of the full observations, at timescales of 15 minutes and 8 seconds. We present a large number of variable sources and discuss their nature, both intrinsic to the sources as well as external, such as interstellar scintillation effects. We also place constraints on transient rates based on the transient simulations code presented in Chapter~2.
In Chapter 4 we report on another commensal transient search, using methodology established in Chapter~3, on long observations of supernova and short GRB fields. We search for transients in images with 30 minute integration times, finding several variable sources. We again explore the variability and its nature, both intrinsic and external. Also in this case, we will establish accurate upper limits on the transient rate using transient simulations.
In Chapter 5, we use deep MeerKAT observations of seven short GRB fields to search for radio afterglow emission. We use these observations to place constraints on astrophysical parameters, in particular the efficiency of particle acceleration and emission processes, and the density of the circumburst medium. We also discuss the impact that future observatories such as the SKA will have on determining these physical parameters. Furthermore, we report the detection of possible host galaxies associated with some of these short GRBs and estimate the star formation rate, assuming the observed radio emission is indeed from star formation.
Finally, in Chapter 6, we summarize our overall results from this thesis work and provide some thoughts on future work that can be done in this area of research.
\newpage
\section{Simulating Transients for Radio Surveys}
\label{sec:transientsims}
\label{sec:sample1}
We are entering an exciting era in time-domain astronomy. New and upgraded facilities such as the Vera C. Rubin Observatory~\citep{2019ApJ...873..111I} and Zwicky Transient Facility~\citep{2019PASP..131a8002B} in the optical, and the MeerKAT~\citep{2016mks..confE...1J} and Australian Square Kilometer Array Pathfinder (ASKAP)~\citep{2021PASA...38...54M} radio telescopes, have been finding, or are expected to find, transients and variables in images at rates that are orders of magnitude higher than ever before. This is in addition to exciting new transients found in time series data, such as the wealth of Fast Radio Bursts (FRBs) found using the Canadian Hydrogen Intensity Mapping Experiment (CHIME)~\citep{2019Natur.566..235C}.
Many transients are discovered in blind searches, found by examining large portions of the sky for new sources or known sources that display significant flux changes. There are also transients found in a targeted way, such as those associated with gravitational wave events~\citep{2017PhRvL.119p1101A,2017Natur.551...71T,2018ApJ...868L..11M}, gamma ray bursts~\citep{1997Natur.389..261F,2018MNRAS.473.1512A}, tidal disruption events~\citep{2011Sci...333..199L,2016Sci...351...62V}, and outbursts from X-ray binaries~\citep{2004MNRAS.355.1105F,2017MNRAS.469.3141T}. Considering that transients can be found in both blind and targeted searches brings up important questions: if we are doing a targeted transient search, what is the chance that a detection may be a different transient source that happens to be in the same area of the sky, even within the same uncertainty region of the transient of interest? How many transients of a certain type or with a specific light curve shape would we expect to find in a given survey? Finding the answers to these questions is important for a variety of applications in time-domain astronomy and requires calculating transient rates with high accuracy.
The most straightforward approach to calculating a transient rate is to use the Poisson distribution to find a rate given the number of detections in a survey, but there are shortcomings in this simplified approach~\citep{2016MNRAS.459.3161C}. This transient rate does not account for a number of important factors such as the relative timescales of the transients and the observations, and some of the confounding observational effects such as gaps within an observation or a survey. In addition, it does not account for the distribution of sensitivities present in the observations of a real survey. These effects can be partly mitigated in an analytical approach~\citep{2016MNRAS.459.3161C}, but Monte-Carlo simulations provide a way to more easily account for the issues presented by real observations and surveys in transient rate calculations~\citep{2017MNRAS.465.4106C}.
\citet{2017MNRAS.465.4106C} examined two light curve shapes: the tophat light curve, a light curve that instantaneously rises to its peak flux and at some point in time instantaneously decays; and the fast rise exponential decay (FRED) light curve, a light curve that instantaneously rises to its peak flux and exponentially decays thereafter. The differences between the resulting transient rates from these two light curve shapes indicate how the wide variety of real light curves can affect transient rate calculations. This is in addition to the previously mentioned observational effects that should be accounted for.
Observational radio surveys present a number of challenges for computing transient rates. Although radio observations can be calibrated using a sky model, many radio observations require the use of calibrator fields that need to be observed at certain time intervals before, after, and during an observation of a science target field. This means that the telescope does not continuously point at a target for an entire observation. Often when using a calibrator source, a radio observation would be broken down into observing a very bright, well-known source to calibrate the bandpass, followed by alternately observing a bright source close to the target for gain calibration and the science target. This means that the time on target is less than the total observing time, and that there are gaps in the target observations. This also means that there is the possibility of searching calibrator fields for transients \citep{2011ApJ...728L..14B}. Furthermore, typical radio observations can be broken down into shorter timescales for imaging. In addition, in order to explore a wider field of view, a survey may consist of multiple adjacent pointings on the sky with some degree of overlap between pointings. These pointings may have different limits on transient rates due to differing observing cadences, and the overlap regions will provide different transient rates as well.
The goal of this work is to calculate transient rates while accounting for the aforementioned features and complexities of radio surveys. Corrections for observational effects such as gaps in observations, systematic errors in flux measurements, different kinds of transient light curves, multiple overlapping pointings, and a distribution of observational sensitivities are accounted for. Mitigating all the aforementioned effects makes the transient rate calculations more accurate. In addition, the publicly available simulations code is relatively simple in its use, has Python 3 support, and is designed for modularity so that the user can easily add new items such as other light curve shapes than those already provided.
In the Design section, we will go into detail on how the code is written and its features implemented. We will in the Results and Discussion section present and discuss results from several example radio surveys illustrating the various features. In the Performance section, we will look at the computational performance of the simulations code. In the Future Applications section we will discuss the ways in which this code can be expanded in the future, and we draw conclusions in the final section.
\subsection{Design}
\label{design}
\subsubsection{Language and Libraries}
The code was written in Python\footnote{http://www.python.org} and designed for its most up-to-date versions ($>3.6$). It uses several libraries: Astropy~\citep{2013A&A...558A..33A} to provide accurate angular source separation calculations and any necessary coordinate system changes; Scipy~\citep{2020NatMe..17..261V} for a few special mathematical functions; Bitarray\footnote{https://github.com/ilanschnell/bitarray} for storing large amounts of information efficiently; tqdm\footnote{https://tqdm.github.io} for easy-to-use progress bars; and Numpy\footnote{https://numpy.org/} for the vast majority of the numerical computations. In addition, the script to assist with creating input files uses Common Astronomy Software Applications (CASA)~\citep{2007ASPC..376..127M} to read and extract metadata from radio measurement sets.
% Is it possible that I'm listing to many libraries/details here?
By using an interpreted language that allows for the use of classes, the code is easy to modify or extend for different use cases, or to increase accuracy. Adding new light curves can be done by creating a Python file with the name of the light curve and a class with the essential information. Using Numpy partially makes up for Python's lack of speed compared to a compiled language such as C. The information on whether a simulated source is detected is stored and written using bit arrays, in order to reduce memory usage so that these computations can be performed on a regular desktop or laptop.
\subsubsection{Input}
In order to accurately simulate transient rates, it is necessary to provide detailed information on the survey that will be simulated. This information includes observation times, pointings, field of view, sensitivity, and any gaps in the observations. This information is either supplied by the user as a comma-separated values (CSV) file or it can be generated using a separate script that extracts information from the metadata in the measurement sets of the survey observations.
In addition, the simulations code base contains a configuration file with settings that can be adjusted depending on the use case. These settings include items such as number of transients to simulate, number of transient detections in the survey, flux and duration ranges to simulate, detection threshold, light curve type, confidence level, output filename, and options for simulating a survey such as number of observations, sensitivity, mean and standard deviation of simulated normally-distributed error in sensitivity, interval between observations, and the duration of the observations. The light curve type can be any of the included ones or a new light curve created by the user.
%%%%%%% NEW FIGURE
\begin{figure}
\includegraphics[width=0.8\columnwidth]{multilcexample.png}
\caption{Examples of all light curve shapes included in the simulations in this paper, plotted using arbitrary units of time and flux.}
\label{multilc}
\end{figure}
%%%%%%%%
\subsubsection{Light Curves}
In Figure~\ref{multilc} we show the light curves included in the simulations, which are the tophat, fast rise exponential decay (FRED), exponential rise fast decay (Wilma), exponential rise exponential decay (ERED), Gaussian, and parabola. The tophat light curve is defined to have an instantaneous rise to the peak flux, followed at some point in time by an instantaneous decay. This light curve represents the classic case of a transient that turns on and off, and the simplest form of transient light curves. The FRED light curve instantaneously rises to the peak flux and exponentially decays. This light curve is commonly observed in a variety of X-ray and gamma-ray transients. The Wilma is simply the time reversed FRED: it exponentially rises to the peak flux and then instantaneously decays. Including the ERED light curve is a convenient way to introduce the simplest form of light curve with no definite start or end. The ERED is such a light curve that is formed by putting the previous two together: it exponentially rises to a peak flux and then exponentially decays. The Gaussian is a light curve that has the shape of a Gaussian function with the mean being located at the peak flux and the duration given by the standard deviation. Similar light curves can be seen arising from, for example, binary systems and magnetar bursts. The parabolic light curve is a concave down parabola that reaches the peak flux at the vertex and the duration being the range of time in which the flux is positive; its inclusion provides an example of a light curve with a definite duration but with a profile that rises and falls below the peak flux in a symmetric way (i.e., one step more complex than the tophat). More details about these light curves, including their mathematical definitions, can be found in the appendix to this chapter.
For a given radio survey, the simulated light curve has implications on the part of parameter space that the survey probes. One of the clearest ways to examine these implications is by looking at probability contour plots. Figures~\ref{tophat}-\ref{fred} show the probability contours for example light curves included in the simulations code. The horizontal axis shows the characteristic duration of the transient, which is defined slightly differently for each light curve shape: the tophat and parabolic light curves' characteristic duration is the duration that the transient's flux is non-zero; the characteristic duration for the Gaussian is the standard deviation; and the duration of the FRED, Wilma, and ERED light curves is the e-folding time. The vertical axis shows the characteristic flux, which is the peak flux for all light curves that are currently implemented (but could vary for more exotic light curve shapes). The color legend shows the probability of detecting a source as a transient at a given duration and flux. Note that a source that is detected in every observation would not be a transient. A probability of 1 means that the survey detects every transient source at the particular flux and duration. Note how the region where the transient is always detected changes for the different light curve shapes. The reason for some of these differences is discussed in detail in the Light Curves section.
\begin{figure}
\includegraphics[width=\columnwidth]{tophat.png}
\caption{Probability contours for the tophat light curve. The leftmost vertical line marks the shortest observation in the survey, the middle line corresponds to length of the longest gap between observations, and the rightmost line corresponds to approximately the length of the survey itself. The horizontal line is a line marking the flux value that is greater than 99\% of the flux values of the false transient sources.}
\label{tophat}
\end{figure} \begin{figure}
\includegraphics[width=\columnwidth]{parabolic.png}
\caption{Probability contours for the parabolic light curve. The meaning of the horizontal and vertical lines is the same as in Figure~\ref{tophat}.}
\label{parabolic}
\end{figure} \begin{figure}
\includegraphics[width=\columnwidth]{gaussian.png}
\caption{Probability contours for the Gaussian light curve. The leftmost vertical line marks the shortest observation in the survey, the curve in the middle marks a boundary in the duration of sources to the left of which these sources can fall in the longest gap between observations and go undetected in any observations, and the rightmost curve is a boundary to the right of which sources can be detected as a constant source by being detected in every observation. The meaning of the horizontal line is the same as in Figure~\ref{tophat}.}
\label{gaussian}
\end{figure}\begin{figure}
\includegraphics[width=\columnwidth]{fred.png}
\caption{Probability contours for the FRED light curve. The meaning of the horizontal and vertical lines is the same as in Figure~\ref{gaussian}.}
\label{fred}
\end{figure}
\subsubsection{Main Detection Algorithm}
In the transient simulations, a large number of sources need to be generated based on the user's settings. Parameters such as the source flux, duration, and the start, end or critical time (depending on the light curve type) are generated in a uniformly random fashion in log10 space via the random number generator in Numpy.
The main detection algorithm tests whether or not the simulated sources will be detected in the observations. For this step, the code iterates over each observation, calculating the integrated flux for all the simulated sources, and testing if these integrated fluxes are greater than the sensitivity of the observation multiplied by a user-specified detection threshold. After this detection step, the sources that are detected in every observation are removed from the detection list, since they are constant sources and not transients.
The number of detected transient sources together with the number of simulated sources are used to generate probabilities of detection for each flux and duration bin. Assuming that transients are distributed as a Poisson distribution, the probabilities are used to calculate limits on transient surface densities and rates. In the case of no transient detections in a survey, the Poisson probability mass function can be inverted to give an upper limit. In case of transient detections in the survey, the code uses the $\chi^2$ distribution \citep[for a review see][]{12005udd3.inbook.....JKK} to calculate the upper and lower limits on the transients rates, by inputting the user-provided confidence level and the number of transient detections in the survey.
%% Poisson explainer HERE
% I considered doing a section on ``incorporating detections'' here, but I decided not to because it would be only one sentence: ``Finding limits on transient rates in the case of purely non-detections is straightforward: one can merely invert the Poisson Distribution (if one assumes transients to be distributed thusly); however, finding upper and lower limits is slightly more complicated: one has to use the inverse incomplete gamma function as shown in~/ref{whatever textbook}.''
\subsubsection{Gaps}
An important ingredient in calculating transient rates accurately is taking into account gaps of varying sizes during observations and surveys. These gaps may exist for a variety of reasons. In the case of radio observations, a long observation on a particular source has to be broken up into scans that are briefly interrupted by observations of a calibrator source. For measuring the flux of a particular source of interest, these gaps are usually unimportant, but for the purposes of calculating transient rates, especially transient rates in a regime where the transients may be shorter than the size of the gaps, it is important to account for these gaps.
Gaps are accounted for in the simulations code base by specifying a gaps file. This file contains all the sub-observations, also known as scans, that make up the full length observation. By running the simulations over the scans, and averaging together the measured flux in each scan, we are able to account for realistic gaps in observations. By correcting for these gaps, we are able to account for multiple different timescales and different sensitivities present in the same survey in an accurate way that would not be possible, or at least very challenging, to do in an analytic fashion \citep{2016MNRAS.459.3161C}.
\subsubsection{False Detections}
False transient detections is an issue that affects real transient searches and should therefore be included in transient simulations. When an astronomical source is close to the detection threshold, any small amount of measurement error, either statistical or systematic, can change it from a detection to a non-detection or vise-versa. Since this can be true for every observation in the survey, there can exist a fairly wide distribution of false transient detections, governed by the sensitivities of the observations. These sources will be flagged as transients, which is an issue because they are not real transients but merely faint sources of constant flux. Therefore, it is necessary to find a way to eliminate these sources from consideration as transients.
In order to solve this problem, a second run through the detection algorithm is performed using sources with tophat light curves along with duration and start times that ensure that they ought to be constant sources. After the false detections of these constant sources are calculated, the number of sources detected are counted from the minimum flux simulated until 99\% of the falsely detected sources are accounted for. At the flux level where 99\% is reached, we define this to be the false detection limit. This is shown in all of the probability contour plots, such as in figures~\ref{tophat}-\ref{fred}, and transient rate plots as a horizontal line.
\subsubsection{Multiple Pointings}
Real surveys can involve multiple pointings that overlap, resulting in an uneven probing of the sky. This creates opportunities and challenges for determining transient rates in these regions of the sky, due to the differences in timescales and observed area.
In order to account for this, the simulations accurately calculate the area of each region on the sky, and then determine the transient rates for each region on the sky. This is currently implemented for a maximum of three overlapping pointings with possibly varying observing timescales and cadences, but an expansion of this is easily doable.
% The journal wants the figures separately
\begin{figure}
\includegraphics[width=\columnwidth]{Figure_1_blank.png}
\caption{An example of three overlapping pointings with red circles representing three different telescope pointings that overlap}
\label{threepointings}
\end{figure}
Figure \ref{threepointings} shows a simulated example of three overlapping pointings. Each red circle represents a pointing of the telescope. It can be seen that there are also three double overlap regions and one triple overlap region. For each of these regions, the transient rates will be different due to the differences in observing cadence and time.
\subsection{Results and Discussion}\label{sec:results1}
\subsubsection{A Realistic Survey}
\begin{figure}
\includegraphics[width=\columnwidth]{figure2.png}
\caption{Lower limits (left) and upper limits (middle) on transient rate for a realistic survey setup. The red line indicates the false detection limit. The plot on the right shows the limits on the transient rate at 0.424 Jy, which is marked on the two plots to the left with a dashed line at the very top of the plots.}
\label{realscen}
\end{figure}
For the purpose of demonstrating the capabilities of the transient simulations code, a survey setup similar to that in \citet{10.1093/mnras/stz3027} is used: 46 weekly observations of 13 minutes in duration, with the rms noise of the observations being varied as a Gaussian with mean 35 $\mu$Jy and standard deviation 5 $\mu$Jy and a detection threshold of $5\sigma$. Given that transients were detected in \citet{10.1093/mnras/stz3027} , we also demonstrate the ability of the simulations code to calculate transient rates based on detections. Assuming Poisson statistics, one can calculate the upper and lower limits on the transient rate, as explained in the previous section. In the configuration for this simulations run, two transient detections are used as input to calculate the rates along with a 95\% confidence interval. The light curve type used for this example is the Gaussian.
Figure \ref{realscen} shows the results of these simulations. The left-side plot shows the lower limits on the transient rate, the middle plot shows the upper limits on the transient rate, and the plot on the right shows the example of transient rate limits at 0.424 Jy as a function of transient duration. The horizontal red lines indicate the 99\% false detection rate. These plots show the transient rate limits: we can see that at 0.424 Jy and around a transient duration of 10 days, the transient rate is between $7\times10^{-4}$ and $5\times10^{-3}$ transients per day per square degree.
\subsubsection{Light Curves}\label{lcdiscussion}
Figures~\ref{tophat}-\ref{fred} show the probability contours for the tophat, parabolic, Gaussian, and FRED light curves. These probabilities are the ratios of the detections to total simulated sources for each duration and flux bin. Each plot has a region of parameter space where all of the transients are detected. As shown by \citet{2017MNRAS.465.4106C}, in the tophat case this is bounded on the left by the duration of the longest gap between consecutive observations. The boundary on the right corresponds to the longest possible duration transient that will still be considered a transient and not a constant source. In other words, this duration is slightly less than the length of the entire survey, since a transient of this length would be detected in every observation except for one. For the FRED, Gaussian, Wilma, and ERED light curves, we observe that the boundaries around this same region are curves. In the appendix, we go into detail on finding the equations for these curves.
Examining the probability contours for the parabolic light curve in Figure~\ref{parabolic} shows a plot that looks closer to the tophat than the other light curves due to the vertical boundaries on the region where the probability is equal to 1. While this may seem counter-intuitive, a similarity between the parabolic and tophat light curve is that they both have a fixed start and stop time at which the flux drops to zero. All the other example light curves approach but never reach zero. For this reason, we use a value to characterize the duration such as the e-folding time for the FRED, or the standard deviation in the case of the Gaussian light curve. Using these values to characterize the duration is what causes the difference in these probability contours. As an example, for the FRED light curve, if the transient has a low flux compared to the sensitivity of the observations, then the duration of the transient that would be detected might be something closer to its e-folding time. In contrast, a very bright transient would be detected well past its e-folding time. This is the reason why these light curves seem to curve away to the left as flux increases in these probability contour plots: the actual duration that is detected in the survey becomes longer. If we were able to define the duration of the transient by the duration that is actually detectable in the survey, then we could make all of the probability contour plots have the kinds of vertical boundaries that we see in the tophat and parabolic light curves. However, defining the durations this way, would make the simulations much more computationally and mathematically complex to the extent that it makes this prohibitive.
\subsubsection{False Detections} \label{fddiscussion}
Figure~\ref{fig9} shows an example of a survey with a large number of false detections of transients. This can happen when including images that are grouped around very different sensitivity scales. In the example shown here, the survey included observations on three very different time scales and sensitivities: 4 hour observations with an rms noise of around $9~\mu$Jy; 15 minute observations with an rms noise of around $30~\mu$Jy; and 8 second observations with a noise around $350~\mu$Jy. Including all of these images in one run of the simulations creates many false detections. In this example, it is better to run simulations of these three different time and sensitivity scales separately. In figures~\ref{sample4hr}-\ref{sampleint}, the probability contours for the three different timescales are shown separately. From these plots, we can clearly see that the false detection limit is much lower on two of the three timescales and a little higher on the shortest timescale.
\subsubsection{Gaps} \label{Gaps}
In order to demonstrate the capability of including observations with gaps, an observation file was created with weekly 4 hour observations (instead of 13 minutes), containing gaps within the weekly observation, and a total survey duration of 46 weeks (as in the previous example). The gaps were typical for a target-gain calibration loop in radio observations: 5 minutes on a calibrator field followed by 15 minutes on a science target field. The noise in the 15-minute scans was simulated like before, with a mean of 35 $\mu$Jy and a standard deviation of 5 $\mu$Jy in the target observations. For the full four hour observations, the noise was scaled as $1/\sqrt{time}$ and simulated as a Gaussian with a mean of 8 $\mu$Jy and a standard deviation of 1 $\mu$Jy. Due to the nature of having a bright calibrator source in a field, the noise for the calibrator observations was higher than would be suggested by scaling by $1/\sqrt{time}$. The noise of the 5 minute scans of the calibrator observation was simulated to be 100 $\mu$Jy with a standard deviation of 15 $\mu$Jy. The noise of the combined image of the calibrator scans was simulated to be 25 $\mu$Jy with a standard deviation of 4 $\mu$Jy. For this example we assume that there are no detected transients in this simulated survey.
\begin{figure}
\includegraphics[width=\columnwidth]{figure9.png}
\caption{Probability contours for a tophat light curve in a survey with very different observation sensitivities (see main text for details).}
\label{fig9}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{sample4hr.png}
\caption{Probability contours for a tophat light curve in a survey with only 4 hour observations at an rms noise of around $9~\mu$Jy.}
\label{sample4hr}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{samplescan.png}
\caption{Probability contours for a tophat light curve in a survey with only 15 minute observations at an rms noise of around $30~\mu$Jy.}
\label{samplescan}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{sampleint.png}
\caption{Probability contours for a tophat light curve in a survey with only 8 second observations at an rms noise of around $350~\mu$Jy.}
\label{sampleint}
\end{figure}
Using these simulations, we can show how accounting for gaps results in more accurate transient rate calculations. This is particularly important on timescales close to the length of the gap itself. In order to test the gaps algorithm, three observational scenarios were used: the calibrator field with 15 minute gaps, the target field with 5 minute gaps, and a full four hour observation with no gaps. These scenarios provide a comparison between different extremes of gaps in observations. These simulations were done with both tophat and FRED light curves. Figures \ref{fig3}-\ref{fig8} show the results of these scenarios. Figures~\ref{fig3} and~\ref{fig4} show upper limits on transient rates in the color legend, with transient duration on the horizontal axis and characteristic flux on the vertical axis. Figure~\ref{fig3} is for a tophat light curve and Figure~\ref{fig4} is for a FRED light curve. The three plots in each figure show the difference in transient rates when not accounting for gaps in a weekly survey with 4 hour observations (top), when accounting for 5 minute gaps in a 4 hour science target observation (middle), and when accounting for 15 minute gaps in calibrator observations (bottom).
In figures~\ref{fig3} and~\ref{fig4} we can see a diagonal trend at the shortest durations below which there are no colored contours. This boundary marks the transients that are shortest in duration and lowest in flux to possibly be detected. It is a diagonal because it is the fluence that determines if a transient is detected \citep{2017MNRAS.465.4106C}; and in the FRED case, for short durations the integrated flux becomes identical to the tophat case. The blank space in the bottom left of the plots represents the region of transient parameter space that cannot be probed by the simulated survey. The red vertical lines on these plots mark 5 minutes, the length of the gaps in the target observations.
Differences in the transient rate are small and difficult to distinguish between the target gap and no gap plots in figures~\ref{fig3} and~\ref{fig4}. The calibrator gap shows a bit of a departure from the others: examining closely reveals a slightly different trend to the left of the red line for both light curves. This departure from the case of having no gaps or the case of a smaller gap only shows in the part of parameter space that has the smallest duration transients. When transients are longer in duration, they are not likely to fall in the gaps and are more likely to be detected in an observation.
Figures~\ref{fig5} through~\ref{fig8} show the differences between the different gaps in a different way. Figures~\ref{fig5} and~\ref{fig6} show the transient rates from Figures~\ref{fig3} and~\ref{fig4} on the vertical axis at a constant flux of 0.464 Jy. Figures~\ref{fig7} and~\ref{fig8} show the percent difference in transient rate between the gaps and no gaps cases. The top panel of Figures~\ref{fig7} and~\ref{fig8} shows the difference between the target gap and no gap cases, and the bottom plot shows the difference between the calibrator gap and no gap cases. As we can see there is an appreciable difference when accounting for 5 minute gaps in a target observation, and a significant difference of nearly 300\% when accounting for 15 minute gaps in the calibrator observation.
\begin{figure}
\includegraphics[width=0.75\columnwidth]{figure3.png}
\caption{Upper limits on transient rate for transients with a tophat light curve in a survey with 4 hour observations with no gaps (top), 5-minute gaps in between 15-minute observations (middle), and 15-minute gaps in between 5-minute observations (bottom); see main text for sensitivities of observations. The dashed black line marks where two different simulations were combined into a single plot. The vertical red line indicates 5 minutes, and the horizontal red line indicates the false detection limit.}
\label{fig3}
\end{figure}
\begin{figure}
\includegraphics[width=0.75\columnwidth]{figure4.png}
\caption{Upper limits on transient rate for transients with a FRED light curve in a survey with 4 hour observations with no gaps (top), 5-minute gaps in between 15-minute observations (middle), and 15-minute gaps in between 5-minute observations (bottom); see main text for sensitivities of observations. The dashed black line marks where two different simulations were combined into a single plot. The vertical red line indicates 5 minutes, and the horizontal red line indicates the false detection limit.}
\label{fig4}
\end{figure}
\begin{figure}
\includegraphics[width=0.75\columnwidth]{figure5.png}
\caption{Upper limits on transient rates for transients with a tophat light curve in a survey with 4 hour observations with no gaps (top), 5-minute gaps in between 15-minute observations (middle), and 15-minute gaps in between 5-minute observations (bottom), for transients with a peak flux of 0.464 Jy.}
\label{fig5}
\end{figure}
\begin{figure}
\includegraphics[width=0.75\columnwidth]{figure6.png}
\caption{Upper limits on transient rates for transients with a FRED light curve in a survey with 4 hour observations with no gaps (top), 5-minute gaps in between 15-minute observations (middle), and 15-minute gaps in between 5-minute observations (bottom), for transients with a peak flux of 0.464 Jy.}
\label{fig6}
\end{figure}
\begin{figure}
\includegraphics[width=0.75\columnwidth]{figure7.png}
\caption{Percent difference in upper limits on transient rates for transients with a tophat light curve for transients with a peak flux of 0.464 Jy, from no gaps in a survey with 4 hour observations to 5-minute gaps (top) and 15-minute gaps (bottom).}
\label{fig7}
\end{figure}
\begin{figure}
\includegraphics[width=0.75\columnwidth]{figure8.png}
\caption{Percent difference in upper limits on transient rates for transients with a FRED light curve for transients with a peak flux of 0.464 Jy, from no gaps in a survey with 4 hour observations to 5-minute gaps (top) and 15-minute gaps (bottom).}
\label{fig8}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{threeregion_comboplot.png}
\caption{A probability contour plot for a tophat light curve in a region with no overlap (left), double overlap (center), and triple overlap (right), as seen in Figure~\ref{threepointings}. }
\label{threeregion}
\end{figure}
\subsubsection{Multiple Pointings}
Calculating transient rates for multiple overlapping pointings gives a more complete picture of how a survey can probe transient parameter space. An example of such a survey is used here and illustrated in Figure~\ref{threepointings}: three circular fields of view each with a radius of 1.4 degrees. The details are summarized in Tables~\ref{simsurvalt} and~\ref{simsurvseq} below. This setup has seven different regions: three that are probed only by one of the pointings, three that are probed by two pointings, and one that is probed by all three pointings. The three different fields may be observed at various cadences that affect the transient rates in the different areas. If one calculates the probability contours for a tophat transient, as is shown in Figure~\ref{threeregion}, one can compare the single pointing (left) with a double overlapping pointing (middle) and a triple overlapping pointing (right). Note how regions with more overlap have a larger region in the transient duration space where the probability of detecting the transient is equal to 1.
For a comparison of different survey cadences, one example survey shown in Table~\ref{simsurvalt} alternates between each of the three pointings each week, and another one shown in Table~\ref{simsurvseq} is set up to observe each pointing exclusively before moving to the next one. Figure~\ref{multirgnprob} shows that the probability contours for the triple overlapping region, labelled 0\&1\&2, are the same for both scenarios, as expected. We do, however, see slight differences in the regions with no overlapping pointings in the part of parameter space where transients are best detected. This difference is due to the variations in the maximum gap and survey length in the two survey setups. One particular region with two overlapping pointings, labelled 0\&1, shows the most striking differences between the survey setups. Survey setup 2 produces two of the double overlap regions with good limits on transient rate and one double overlap region with poor limits on the transient rate. In this case, the region that is observed in both the first observed field and the last observed field will have an extremely large gap. Survey setup 1 produces much more consistent detection regions which may suggest that it is the better choice if more uniform transient rate limits are the goal.
% Relabel survey numbers to make survey one the one on top of the figures
\begin{figure}
\includegraphics[width=\columnwidth]{multirgnprob.png}
\caption{Probability contours for all of the regions in survey setup 1 (top) and survey setup 2 (bottom). Each pointing is labeled 0, 1 or 2 and the overlap between two pointings are indicated with an `\&', for example 0\&1.}
\label{multirgnprob}
\end{figure}
\clearpage
\begin{landscape}
\begin{deluxetable}{|c|c|c|c|c|c|c|}
\tablecolumns{7}
\tablewidth{0pc}
\tablecaption{Simulated survey moving between pointings weekly \label{simsurvalt}}
\tablehead{\colhead{RA (J2000)} & \colhead{DEC }& \colhead{ID }&\colhead{ Area (deg$^2$)} & \colhead{Duration (days)} & \colhead{Start (MJD)} &\colhead{ End}}
\startdata
274.7345 & 7.7974 & 0 & 6.1572 & 294.01 & 58997.54 & 58703.53 \\
275.0913 & 7.1851 & 1 & 6.1572 & 294.01 & 59011.54 & 58717.53 \\
275.4482& 7.7974 & 2 & 6.1572 & 294.01 & 59004.54 & 58710.53 \\
274.9130 & 7.4913 & 0\&1 & 4.1987 & 308.01 & 59011.54 & 58703.53 \\
275.0913 & 7.7975 & 0\&2 & 4.1987 & 301.01 & 59004.54 & 58703.53 \\
275.2696 & 7.4913 & 1\&2 & 4.1987 & 301.01 & 59011.54 & 58710.53 \\
275.0913 & 7.5926 & 0\&1\&2 & 3.4360 & 308.01 & 59011.54 & 58703.53 \\
\enddata
\end{deluxetable}
\begin{deluxetable}{|c|c|c|c|c|c|c|}
\tablecolumns{7}
\tablewidth{0pc}
\tablecaption{Simulated survey moving between pointings sequentially \label{simsurvseq}}
\tablehead{\colhead{RA (J2000)} & \colhead{DEC }& \colhead{ID }&\colhead{ Area (deg$^2$)} & \colhead{Duration (days)} & \colhead{Start (MJD)} &\colhead{ End}}
\startdata
274.7345 & 7.7974 & 0 & 6.1572 & 98.16 & 58801.70 & 58703.53 \\
275.0913 & 7.1851 & 1 & 6.1572 & 98.16 & 59011.70 & 58913.53 \\
275.4482 & 7.7974 & 2 & 6.1572 & 98.16 & 58906.70 & 58808.53 \\
274.9130 & 7.4913 & 0\&1 & 4.1987 & 308.16 & 59011.70 & 58703.53 \\
275.0913 & 7.7975 & 0\&2 & 4.1987 & 203.16 & 58906.70 & 58703.53 \\
275.2696 & 7.4913 & 1\&2 & 4.1987 & 203.16 & 59011.70 & 58808.53 \\
275.0913 & 7.5926 & 0\&1\&2 & 3.4360 & 308.16 & 59011.70 & 58703.53 \\
\enddata
\end{deluxetable}
\end{landscape}
% \geometry{ papersize={8.5in,11in} }
\clearpage
\doublespacing
\subsection{Performance}\label{performance}
Figure~\ref{fig10} shows how the simulations scale in execution time as a function of the number of sources simulated, for the example of 46 observations of 13 minutes with all observations have an identical field-of-view. Figure~\ref{fig11} shows scaling in execution time as a function of the number of observations when the number of sources is held constant at $4.3\times10^{5}$. All of the simulations for this example were performed on a 2020 Apple Macbook Pro 13 inch model with the M1 chip. In addition to showing the total execution time, key portions of the code are shown as well. The conditionals and flux filtering steps take place when the algorithm is determining if the sources are detected in observations. These steps are so-named because they filter sources based on the calculated integrated flux, and do a large number of boolean and bit operations to calculate and store each transient's state as either detected or not-detected. The stats step is the step that aggregates the detections into probabilities. Finally, the plotting step is where all of the detection statistics, observation information, and false detection information is plotted. Since the data is broken down into a grid in order to plot, the plotting step has no scaling with the number of sources, so it takes a constant amount of time.
Not shown here are the impacts of a few features and algorithms. The false detection algorithm re-runs the conditionals, flux filtering, and stats steps. In the case of a tophat light curve, this means the false detection algorithm would slightly less than double the amount of time (the plotting step is not doubled). In the case of other light curves, it could have a different impact, usually lower, since the false detection algorithm always uses tophat light curves to simulate constant sources. Another factor not shown here is the impact of having multiple pointings. For example, in a survey setup with two overlapping pointings there will be three regions. Therefore, the run time will be about three times longer than for a single region, assuming equal numbers of observations in each region.
\begin{figure}
\includegraphics[width=\columnwidth]{figure10.png}
\caption{Execution time as a function of the number of simulated sources.}
\label{fig10}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth]{figure11.png}
\caption{Execution time as a function of the number of observations.}
\label{fig11}
\end{figure}
% Make sources vs execution time
% Address impacts of algorithms,
% Each time per region * regions
% False Det doubles
%
\subsection{Future Applications} \label{futureapplications}
\citet{2017MNRAS.465.4106C} shows the existence of a region of 100\% detection probability for the FRED and tophat light curves. The work presented here shows that this region of 100\% detection can be found in a wide variety of light curves. Optimizing the survey parameters to maximize this region of 100\% detection has many potential applications for future surveys. In addition, transient searches can use the false detection calculations to decide the best manner to plan a transient search in a given survey. The wide variety of data outputs can assist with a number of scientific goals that one might have for a survey, and the flexibility of the code can be easily adapted to those purposes.
One application of the code is using the transient simulations to optimize resource allocation. Such optimizations can be done both in terms of survey design and also the data reduction after the observation itself. Simulating a variety of survey setups, such as in section~\ref{Gaps}, is one way to ensure that the survey will accomplish its goals most effectively. In addition, simulations can also be done for varying aspects related to the data reduction, such as the timescale that images are made on. By simulating a combining or splitting up of the observations in a survey into multiple different timescales, one can find an optimal way to search for transients on these multiple time scales with a minimum of re-imaging. These optimizations are becoming increasingly more important with radio facilities such as MeerKAT~\citep{2016mks..confE...1J} and the LOw-Frequency Array (LOFAR)~\citep{2013A&A...556A...2V} which can easily use terabytes of disk space and considerable other computer resources as well. The Square Kilometer Array~\citep{2009IEEEP..97.1482D} and next-generation Very Large Array~\citep{2018ASPC..517....3M}, and other upcoming facilities, will surely require even more resources.
The previously mentioned tools for planning surveys have potential to be expanded to be even more helpful in the future. A potential future update to these simulations could include a tool to help calculate optimal pointings for uniformly probing a large area of the sky in both space and time.
The simulations code presented in this paper accounts for a large number of realistic effects that complicate transient searches and calculate transient rates from surveys. For research into particular kinds of sources, future upgrades can be made for particular light curves and population numbers that reflect certain sources of interest.
Finally, even though this simulations code has been designed for surveys in the radio regime, and the examples in this paper are based on this particular use case, it can easily be adapted and applied to other spectral regimes.
\subsection{Conclusions} \label{conclusions}
Simulating transients, following the methodology and code case presented here, allows for calculating transient rates that are highly accurate due to the implementation of a variety of observational effects. The simulations presented here account for a variety of observing sensitivities, pointings, survey cadences, and gaps within observations and surveys. Furthermore, it has been made easy to obtain, since it will be freely available for download through Github, and easy to use through the use of a modular design, the inclusion of scripts to extract metadata from observations, and updates for modern versions of Python.
\subsection{Acknowledgements}
The authors would like to thank the referee for their constructive comments that helped improve the paper.
The authors would like to acknowledge the ThunderKAT collaboration for the valuable sharing of knowledge and resources, and Michael Moss for his helpful comments and feedback on this manuscript.
This work was completed in part with resources provided by the High Performance Computing Cluster
at The George Washington University, Information Technology, Research Technology Services.
\subsection{Appendix: Included Light Curves}
%\subsection{Included Light Curves}
\label{sec:lc:appendix}
Here we present and briefly discuss the light curve shapes that are currently included in the simulations code base.
\subsubsection{Tophat}
The tophat is the simplest transient light curve in concept:
\begin{equation}
F=F_{pk}~\text{for}~t_{start} \le t \le t_{end}
\end{equation}
It is simply at the peak flux for the entire duration of the transient. The probability contour plot shown in Figure~\ref{tophat} has a region in parameter space in which all transients are always detected, which can be referred to as a region of guaranteed detection. This region has vertical boundaries that can be found to have a quite straightforward interpretation \citep{2017MNRAS.465.4106C}. The left-most boundary is the longest gap in the observations or, in other words, the longest duration of a tophat transient that could go undetected. The right-most vertical bounding line corresponds to the longest time scale that a transient could have while not being detected as a constant source. This quantity is the duration of the entire survey minus either the first or last observation.
\subsubsection{Fast Rise Exponential Decay}
The fast rise exponential decay (FRED) light curve is defined as instantaneously rising to the peak flux and exponentially decaying with an characteristic duration $\tau$, defined as its e-folding time:
\begin{equation}
F=F_{pk}\,\exp\left[\frac{-(t-t_{start})}{\tau}\right]~\text{for}~t\ge t_{start}
\end{equation}
This light curve produces a slightly different probability contour, seen in Figure~\ref{fred}, in which the bounding lines for the region of guaranteed detection can be interpreted as follows. The left boundary corresponds to the boundary due to the longest gap, like the tophat. However, unlike the tophat, the flux of the FRED light curve approaches but never actually reaches zero as time progresses. Therefore, brighter transients can be detected for longer than the characteristic duration of the transient, making this boundary a curve instead of a vertical line. The boundary condition can be expressed as $F_{int}=S_{gap}$, i.e. the integrated flux of the transient needs to be equal to the sensitivity of the observation the transient would be detected in, which would be the observation after the gap. We can find the integrated flux:
\begin{equation}
F_{int} = F_{pk}\,\tau\,\frac{\exp\left[-\frac{\max(T_{start},t_{start})}{\tau}\right] - \exp\left[\frac{T_{end}}{\tau}\right]}{T_{end}-T_{start}}
\end{equation}
Since we consider the case where the transient starts in the gap, the start of the observation that detects the transient is equal to the length of time from the start of the transient until the end of the gap, which we label $T_{gap}$: $T_{start}=T_{gap}$.
Therefore, $T_{end}=T_{start}+\Delta T_{gap}$, where $\Delta T_{gap}$ is the duration of the observation. Inserting the integrated flux into the previous equation and solving for $F_{pk}$ yields:
\begin{equation}
F_{pk}(\tau) = \frac{S_{gap}\,\Delta T_{gap}}{\tau\,\left(\exp\left[-\frac{T_{gap}}{\tau}\right] - \exp\left[\frac{(T_{gap} + \Delta T_{gap})}{\tau}\right]\right)}
\end{equation}
The right boundary is the boundary for the longest timescale. We can follow the same procedure as the left boundary, finding that $S_{obs} = S_{last}$, the sensitivity of the last observation in the survey. We also find the following modifications:
\begin{equation}
T_{start} = \tau_{survey} - \Delta T_{last}
\end{equation}
\begin{equation}
T_{end} = \tau_{survey}
\end{equation}
\begin{equation}
F_{pk}(\tau) = \frac{S_{last}\,\Delta T_{last}}{\tau\,\left(\exp\left[-\frac{(\tau_{survey} - \Delta T_{last})}{\tau}\right] - \exp\left[\frac{\tau_{survey}}{\tau}\right]\right)}
\end{equation}
\begin{figure}
\includegraphics[width=\columnwidth]{wilma.png}
\caption{Probability contours for the Wilma light curve}
\label{wilma}
\end{figure}
\subsubsection{Exponential Rise Fast Decay (Wilma)}
In light of the FRED light curve, a natural extension would be to examine the reverse FRED light curve. The light curve ends at the peak flux and has no definite start:
\begin{equation}F=F_{pk}\,\exp\left[{\frac{(t-t_{end})}{\tau}}\right]\text{ for }t\le t_{end}\end{equation}
The probability contour plot for this light curve is shown in Figure~\ref{wilma}. As one can see, it is identical to the FRED light curve in Figure~\ref{fred}. This makes sense when one realizes that if the entire survey were time-reversed, the light curve would be a FRED. For this reason, the lines bounding the region of guaranteed detection are the same as for the FRED light curve.
\begin{figure}
\includegraphics[width=\columnwidth]{ered.png}
\caption{Probability contours for the ERED light curve}
\label{ered}
\end{figure}
\subsubsection{Exponential Rise Exponential Decay}
The Exponential Rise Exponential Decay (ERED) light curve (Figure~\ref{ered}) does not have a definite beginning nor end, only a characteristic time at which the flux is the peak flux and $\tau$, which is its e-folding time:
\begin{equation}F=F_{pk}\,\exp\left[{\frac{(t-t_{char})}{\tau}}\right]\text{ for }t< t_{char}\end{equation}
\begin{equation}F=F_{pk}\text{ for }t=t_{char}\end{equation}
\begin{equation}F=F_{pk}\,\exp\left[{\frac{-(t-t_{char})}{\tau}}\right]\text{ for }t> t_{char}\end{equation}
The transients from this light curve also behave similarly to the previous two cases, since this light curve is a Wilma light curve immediately followed by a FRED light curve. Therefore, if we use the integrated flux to find the curve marking the boundary corresponding to the shortest duration transient that will always be detected, we find:
\begin{equation}F_{pk}(\tau) = \frac{ S _{gap}\,\Delta T_{gap}}{\frac{\tau}{2}\,\left(\exp\left[-\frac{T_{gap}}{\tau}\right] - \exp\left[\frac{2\Delta T_{gap} + T_{gap}}{\tau}\right]\right)}\end{equation}
Similarly, for the flux limit on the longest timescale, we have:
\begin{equation}F_{pk}(\tau) = \frac{S_{first/last}\,\Delta T_{first/last}}{\frac{\tau}{2}\,\left(\exp\left[-\frac{\tau_{survey}}{\tau}\right] - \exp\left[-\frac{2\Delta T_{first/last} + \tau_{survey}}{\tau}\right]\right)}\end{equation}
In this equation, $S_{first/last}$ and $\Delta T_{first/last}$ would correspond to either the first or last observation in the survey depending on which is more sensitive.
\subsubsection{Parabola}
Also included is a parabolic light curve defined as follows:
\begin{equation}F = F_{pk}\,\left(1 - \frac{4}{\tau^2}\left(t - \frac{\tau}{2} - t_{crit}\right)^2\right)\end{equation}
$t_{crit}$ is the peak of the light curve, which occurs at half of the duration of the light curve. Since the parabolic light curve starts and ends at zero flux, rather than approaching zero like the exponential or Gaussian light curves, this light curve has a definite duration like the tophat. The light curves with a definite duration all have boundaries in the probability contour plots that are derived in the same way as those for the tophat light curves. The boundaries in the case of the the exponential or Gaussian light curves come about from there being a difference between the characteristic duration and the duration that the transient is actually detected in the observations.
\subsubsection{Gaussian}
The final light curve included is a Gaussian-shaped one:
\begin{equation}F=F_{pk}\,\exp\left[\frac{-(t-t_{crit})^2}{2\left(\frac{\tau}{2}\right)^2}\right]\end{equation}
In order to find the boundaries for the region of parameter space where the transients are always detected, we follow the same process as we did for the FRED light curve, and find an equation for the integrated flux (with erf being the Gauss error function):
\begin{equation}
\begin{split}
F_{int} = F_{pk}\,\tau\,\sqrt{\frac{\pi}{8}}\,(\text{erf}\left[\frac{\sqrt{2}(T_{end}-t_{crit})}{\tau}\right]\\
-\text{erf}\left[\frac{\sqrt{2}(T_{start}-t_{crit})}{\tau}\right])/(T_{end}-T_{start})
\end{split}
\end{equation}
We can define the boundary on the transient flux needed to be detected as $F_{int}=S_{gap}$.
Using the equation for integrated flux, we find the boundary for the shortest possible duration transient that will always be detected:
\begin{equation}
\begin{split}
F_{pk} = \frac{S_{obs}\,\Delta T_{obs}}{\tau\, \sqrt{\frac{\pi}{8}}}\frac{1}{\text{erf}\left[\frac{-\Delta T_{gap}}{\sqrt{2}\,\tau}\right] - \text{erf}\left[\frac{(-2\Delta T_{obs} - \Delta T_{gap})}{\sqrt{2}\,\tau}\right]}
\end{split}
\end{equation}
We also find the boundary on the right, for the longest possible duration transient before it is considered a constant source:
\begin{equation}
\begin{split}
F_{pk} = \frac{S_{obs}\,\Delta T_{obs}}{\tau\, \sqrt{\frac{\pi}{8}}}\frac{1}{\text{erf}\left[\frac{-(\Delta T_{survey} + 2\Delta T_{obs})}{\sqrt{2}\,\tau}\right] - \text{erf}\left[\frac{-\sqrt{2}\,\Delta T_{survey}}{\tau}\right]}
\end{split}
\end{equation}
\newpage
\section{Commensal Transient Searches in 8 Short Gamma-Ray Burst Fields}
Transient searches at radio wavelengths are now yielding an unprecedented number of transients of all kinds. For some time now, transient searches in other wavebands such as the optical and X-rays have yielded a large number of results, and now this is starting to be true for the radio regime as well. Some searches in time series analysis have found transients like fast radio bursts (FRBs) with timescales on the order of milliseconds~\citep[e.g.,][]{2007Sci...318..777L,2021ApJS..257...59C}. Other searches have been performed in radio images, with the number of transients and variables found this way increasing and yielding interesting results. For example, a transient was found in the LOFAR Multi-frequency Snapshot Survey on a timescale of around 10 minutes at 60 MHz~\citep{2016MNRAS.456.2321S}. The CHILES Variable and Explosive Radio Dynamic Evolution Survey~\citep{2021ApJ...923...31S} spent hundreds of hours observing the COSMOS field at 1.4 GHz and found a number of variable sources at timescales from days to years. There have also been transients found as part of a commensal search, that is, a search of data taken as part of a different scientific objective. In commensal transient searches with MeerKAT at 1.3 GHz,~\citet{2020MNRAS.491..560D} find a transient with a timescale of weeks and a variable pulsar on sub-week timescales; and in this same field,~\citet{2022MNRAS.512.5037D} find variable sources on timescales of weeks to months. Similarly,~\citet{2022MNRAS.517.2894R} found four variable sources with timescales spanning from seconds up to over a year. \citet{2022MNRAS.513.3482A} also found a radio transient source in a commensal search of MeerKAT data. As part of the Deeper, Wider, Faster,~\citet{2023MNRAS.519.4684D} have found multiple transients and variables with ASKAP. Additionally, the Variables and Slow Transients survey (VAST) using ASKAP~\citep{2021PASA...38...54M} has found multiple radio transients~\citep{2021ApJ...920...45W,2022MNRAS.516.5972W} and the Very Large Array Sky Survey (VLASS) using the VLA at frequencies around 3 GHz~\citep{2020PASP..132c5001L} promises to find a large number of transients and variables due to their large sky coverage and multi-epoch observing strategies.
Commensal searches for transients and variables is proving to be a valuable way of probing the radio sky, in particular with facilities that have a large field of view. Not only are commensal searches an efficient use of pre-existing scientific data, they also have the potential to find new and interesting sources as well as increasing our knowledge of the populations of sources on the radio sky by constraining transient rates~\citep[e.g.,][]{2011ApJ...728L..14B,2016MNRAS.459.3161C}. The number of detections along with the survey properties, if used in conjunction with accurate transient rate calculations, can uncover more information about sources with unknown associations. In addition, with calculations that allow for calculating different transient rates for different parts of the sky, such as~\citep{2022ascl.soft04007C}, it is possible to reveal differences in transients and their behavior in different parts of the sky, such as galactic versus extragalactic sources.
Enabling all of these aforementioned new transient discoveries, with their excellent sensitivity and large field of view, are new facilities such as MeerKAT, ASKAP, and LOFAR~\citep{2009IEEEP..97.1522J,2008ExA....22..151J,2013A&A...556A...2V}. Due to the excellent instantaneous {\it uv} coverage of these instruments, these searches are also able to probe increasingly shorter timescales, with the capability to image on timescales down to seconds, or to create deep images that combine many hours' worth of data. All of these improvements are creating a wealth of new opportunities for commensal transient searches in radio images.
ThunderKAT~\citep{2016mks..confE..13F} is a large survey project for image plane radio transients with MeerKAT. Taking advantage of the new opportunities provided by MeerKAT is a key part of its mission, as it includes conducting commensal transient searches in MeerKAT imaging data (besides performing follow-up observations of specific transients found in other wavebands). Part of the challenge of these searches are that it requires analyzing a large amount of data, on the order of hundreds of gigabytes to terabytes. In order to search through these images, we use the LOFAR transients pipeline \citep[{\sc TraP};][]{2015A&C....11...25S}, which creates a catalog of sources and their light curves, and tracks the variability of all the sources in the images. Using TraP we conduct a commensal transient search on multiple timescales of short gamma-ray burst observations taken as part of the ThunderKAT project. We establish methodologies and techniques to find new variable sources among the large quantity of sources in this dataset. We also look into whether the variability of these sources is intrinsic or extrinsic (e.g., interstellar scintillation), and draw conclusions to guide future similar studies.
We will describe the observations and overall data set in the Observations section, and the methodology for the transient search in the Methods section. The results are presented in the Results section and discussed in the Discussion section, with a summary and concluding remarks in the Conclusion.
\subsection{Observations}
We performed a commensal transient search in observations of eight short gamma-ray burst fields originally taken as part of the ThunderKAT project for the purpose of searching for afterglow emission from short gamma-ray bursts. Each observation was about 4 hours in duration including overhead, such as calibrator observations, with the number of observations per field varying. Each 4-hour observation consisted of 15-minute scans with calibrator measurements of a few minutes interspersed. Table~\ref{tab:allobs} lists all the observations that are a part of our transient survey. The observations were calibrated using version 1.1 of the ProcessMeerKAT pipeline~\citep[{\sc ProcessMeerKAT};][]{pminprep}. Observations were taken at L-band, centerted at 1.28 GHz and spanning frequency range from 856 MHz to 1.7 GHz. As part of this calibration process, parts of the spectrum with a large amount of known radio frequency interference (RFI) were flagged including 933 to 960 MHz, 1163 to 1299 MHz, 1524 to 1630 MHz, and 1680 to 1711 MHz. Calibration was performed in parallel by separating the measurement set into 11 spectral windows. The bandpass and complex gain calibration was performed using Common Astronomy Software Applications \citep[{\sc CASA;}][]{2022arXiv221002276T} tools and the calibrators listed in Table~\ref{tab:allobs}. Automated RFI flagging was performed using tfcrop and rflag. After two rounds of calibration and flagging, the spectral windows were recombined into a single measurement set for imaging.
All images were made using tclean. The 4-hour images were made by producing a shallow image with the cleaning process stopping based on a threshold of 1 mJy; and then self-calibration and flagging for RFI was performed before making the final, deep image stopping at a threshold of around 80 $\mu$Jy. The 15-minute images were made using the self-calibrated measurement set. The imaging parameters used include the multi-term multi-scale imaging algorithm with the w-project gridder. For the 4-hour images, 128 w-planes were used; and for the 15-minute images, 64 w-planes were used. The latter resulted in increased correlated noise in the 15-minute images.
The shortest imaging timescale of the data is determined by the integration time of the observations, which is 8 seconds for every observation in this survey. On this timescale the imaging parameters were slightly different. The quality of images made using w-projection and those not using w-projection were seen to be quite similar, apart from a slight offset in the spatial coordinates between the two. Therefore, in an effort to save processing time and computational resources, the images were made using the standard gridder without w-projection, using the multi-term multi-scale imaging algorithm that is a part of tclean.
\subsubsection{Image Quality}
The typical noise values roughly follow the expected scaling for noise as a function of observation time $t$, that is $~1/\sqrt(t)$, and are summarized in Table~\ref{tab:noisesummary} below. As the timescales go shorter, the trend is for the variance in the noise values to go larger, with the images on the 8-second timescale showing a large range of values. 31 out of the 47,964 images at the 8-second timescale had a noise that was many orders of magnitude higher than the typical noise distribution, skewing the statistics in a way that is possibly misleading, and therefore we exclude these highest 31 values for noise on this timescale. In the actual analysis we did not perform any additional quality control steps in the version of TraP we used.
\newgeometry{margin=1.25in} % modify this if you need even more space
\begin{landscape}
\begin{deluxetable}{lllll} \tablecolumns{5}
\tablewidth{0pc}
\tablecaption{All observations used in this study, indicating the observations' start and end times, phase center position, and time spent on the target.
\label{tab:allobs}}
\tablehead{\colhead{Name} & \colhead{Observation Time }& \colhead{RA } & \colhead{DEC} & \colhead{Time On Target (hrs) } }
\startdata
GRB 200219A & 2020-02-21T12:28:44 to 16:28:44 & 342.6385 & -59.1196 & 3.2031 \\
GRB 200219A & 2020-02-23T12:07:21 to 16:07:21 & 342.6385 & -59.1196 & 3.1986 \\
GRB 200219A & 2020-02-27T13:52:43 to 17:51:43 & 342.6385 & -59.1196 & 3.1986 \\
GRB 200411A & 2020-04-12T07:07:12 to 11:09:12 & 47.6641 & -52.3176 & 3.2053 \\
GRB 200411A & 2020-04-14T11:31:27 to 15:32:27 & 47.6641 & -52.3176 & 3.2031 \\
Sculptor & 2020-04-16T05:15:42 to 09:17:42 & 11.8875 & -25.2886 & 3.2031 \\
GRB 200411A & 2020-04-18T07:27:35 to 11:29:35 & 47.6641 & -52.3176 & 3.2009 \\
GRB 200522A & 2020-05-23T06:56:37 to 11:10:37 & 5.6820 & -0.2832 & 3.4496 \\
GRB 200522A & 2020-05-24T06:01:09 to 10:15:09 & 5.6820 & -0.2832 & 3.4496 \\
GRB 200522A & 2020-05-29T02:11:13 to 06:24:13 & 5.6820 & -0.2832 & 3.4452 \\
GRB 200522A & 2020-06-06T02:01:14 to 06:15:44 & 5.6820 & -0.2832 & 3.4474 \\
GRB 200907B & 2020-09-08T01:03:47 to 05:23:17 & 89.0290 & 6.9062 & 3.4430 \\
GRB 200907B & 2020-09-10T01:47:12 to 06:05:42 & 89.0290 & 6.9062 & 3.4541 \\
GRB 200907B & 2020-09-14T01:35:52 to 05:54:16 & 89.0290 & 6.9062 & 3.4585 \\
GRB 200907B & 2020-09-25T02:17:12 to 06:34:54 & 89.0290 & 6.9062 & 3.4563 \\
GRB 210323A & 2021-03-25T06:17:56 to 10:37:26 & 317.9461 & 25.3699 & 3.4519 \\
GRB 210323A & 2021-03-27T06:03:55 to 10:23:16 & 317.9461 & 25.3699 & 3.4519 \\
GRB 210323A & 2021-04-01T05:37:48 to 09:57:17 & 317.9461 & 25.3699 & 3.4563 \\
GRB 210726A & 2021-07-28T14:24:49 to 17:50:28 & 193.2909 & 19.1875 & 2.7122 \\
GRB 210726A & 2021-08-01T12:28:16 to 16:47:14 & 193.2909 & 19.1875 & 3.4519 \\
GRB 210726A & 2021-08-07T12:07:14 to 16:26:20 & 193.2909 & 19.1875 & 3.4519 \\
GRB 210726A & 2021-08-19T12:18:07 to 16:36:33 & 193.2909 & 19.1875 & 3.4519 \\
GRB 210726A & 2021-09-06T11:38:11 to 15:56:29 & 193.2909 & 19.1875 & 3.4496 \\
GRB 210919A & 2021-09-20T01:22:10 to 05:40:20 & 80.2545 & 1.3115 & 3.4519 \\
GRB 210919A & 2021-09-24T02:49:58 to 07:08:40 & 80.2545 & 1.3115 & 3.4541 \\
GRB 210726A & 2021-09-26T09:35:20 to 13:54:18 & 193.2909 & 19.1875 & 3.4563 \\
GRB 210919A & 2021-09-27T01:23:09 to 05:41:27 & 80.2545 & 1.3115 & 3.4541 \\
GRB 210323A & 2021-09-30T17:35:55 to 21:56:06 & 317.9461 & 25.3699 & 3.4541 \\
GRB 210726A & 2021-12-27T02:48:11 to 07:07:17 & 193.2909 & 19.1875 & 3.4430 \\
GRB 210726A & 2021-09-26T09:35:20 to 13:54:18 & 193.2909 & 19.1875 & 3.4563 \\
GRB 210919A & 2021-09-27T01:23:09 to 05:41:27 & 80.2545 & 1.3115 & 3.4541\\
GRB 210323A & 2021-09-30T17:35:55 to 21:56:06 & 317.9461 & 25.3699 & 3.4541 \\
GRB 210726A & 2021-12-27T02:48:11 to 07:07:17 & 193.2909 & 19.1875 & 3.4430 \\
\enddata
\end{deluxetable}
\end{landscape}
\begin{deluxetable}{lll} \tablecolumns{3}
\tablewidth{0pc}
\tablecaption{Calibrators used for each field in this study.
\label{tab:allobscals}}
\tablehead{\colhead{Name} & \colhead{Bandpass Calibrator} & \colhead{Gain Calibrator}}
\startdata
GRB 200219A & J0408-6545 & J2329-4730 \\
GRB 200411A & J0408-6545 & J0210-5101 \\
Sculptor & J1939-6342 & J0025-2602 \\
GRB 200411A & J0408-6545 & J0210-5101 \\
GRB 200522A & J1939-6342 & J0022+0014 \\
GRB 200907B & J0408-6545 & J0521+1638 \\
GRB 210323A & J1939-6342 & J2236+2828 \\
GRB 210726A & J1331+3030 & J1330+2509 \\
GRB 210919A & J0408-6545 & J0503+0203 \\
\enddata
\end{deluxetable}
\clearpage
\restoregeometry
\doublespacing
\begin{deluxetable}{llll}
\tablecolumns{4}
\tablewidth{0pc}
\tablecaption{Summary of the mean, median and range of the image noise distributions at each timescale in our study.
Note that the 8-second timescale statistics are computed with the highest 31 noise values excluded.
\label{tab:noisesummary}}
\tablehead{\colhead{Timescale} & \colhead{Range ($\mu$Jy)} & \colhead{Median ($\mu$Jy)} & \colhead{Mean ($\mu$Jy)}}
\startdata
4 hours & 6 to 32 & 10 & 13 \\
15 minutes & 19 to 184 & 30 & 43 \\
8 seconds & 106 to 17709 & 176 & 205 \\
\enddata
\end{deluxetable}
\begin{deluxetable}{l|c|c|c}
\tablecolumns{4}
\tablewidth{0pc}
\tablecaption{Summary of number of images of each field at each timescale. \label{tab:obstimescales}}
\tablehead{\colhead{Target} & \colhead{4 hour images} & \colhead{15 minute images} & \colhead{8 second images}}
\startdata
GRB200219A & 4& 51& 5552\\
GRB200411A & 3& 39& 4365\\
Sculptor &1 &13& 1455\\
GRB200522A &4& 54& 6265\\
GRB200907B& 4& 56& 6274\\
GRB210323A& 4& 56& 6275\\
GRB210726A& 7& 95& 9072\\
GRB210919A& 3& 42& 4706\\
\enddata
\end{deluxetable}
\subsection{Methods}
\subsubsection{Transient Searches with the LOFAR Transients Pipeline}
After calibrating the data and producing images, the latter were run through the TraP~\citep{2015A&C....11...25S} version 4. While originally designed for LOFAR, the TraP is telescope-agnostic and well suited for any kind of image based radio transient search. When running the images through the pipeline, a detection threshold of 5$\sigma$ was used, which is the threshold for blind detection of a source, along with an analysis threshold of $3\sigma$, which is the threshold used for analyzing information about the source (such as position and flux, and uncertainties in those quantities). The detection threshold was set to 5$\sigma$ instead of the default 8$\sigma$ so that more sources would be detected and analyzed by TraP. We later increase this threshold and reduce the number of candidate transients and variables through additional analysis, as described in section ``Determining Candidate Variables and Transients.'' This process of starting with a lower threshold and increasing it later was beneficial for capturing longer portions of variable light curves since, when a variable source reaches the detection threshold in TraP, the TraP does not go back to previous images to measure the flux of the source before detection. The TraP calculates variability statistics $V$ and $\eta$, which the user can use to classify a source as constant or varying. These statistics are as defined in \citet{2015A&C....11...25S}:
\begin{equation}\label{Veqn}
V_{\nu} = \frac{1}{\bar{I}_{\nu}}\sqrt{\frac{N}{N-1} (\bar{I_{\nu}^2} - \bar{I_{nu}}^2)}
\end{equation}
\begin{equation}\label{etaeqn}