Issue |
A&A
Volume 659, March 2022
|
|
---|---|---|
Article Number | A84 | |
Number of page(s) | 25 | |
Section | Stellar structure and evolution | |
DOI | https://doi.org/10.1051/0004-6361/202141452 | |
Published online | 16 March 2022 |
Search for intermediate-mass black hole binaries in the third observing run of Advanced LIGO and Advanced Virgo
1
LIGO Laboratory, California Institute of Technology, Pasadena, CA 91125, USA
2
Louisiana State University, Baton Rouge, LA 70803, USA
3
Dipartimento di Farmacia, Università di Salerno, 84084 Fisciano, Salerno, Italy
4
INFN, Sezione di Napoli, Complesso Universitario di Monte S. Angelo, 80126 Napoli, Italy
5
OzGrav, School of Physics and Astronomy, Monash University, Clayton, 3800 Victoria, Australia
6
LIGO Livingston Observatory, Livingston, LA 70754, USA
7
University of Wisconsin-Milwaukee, Milwaukee, WI 53201, USA
8
OzGrav, Australian National University, Canberra, ACT 0200, Australia
9
Max Planck Institute for Gravitational Physics (Albert Einstein Institute), 30167 Hannover, Germany
10
Leibniz Universität Hannover, 30167 Hannover, Germany
11
Inter-University Centre for Astronomy and Astrophysics, Pune, 411007, India
12
University of Cambridge, Cambridge, CB2 1TN, UK
13
Theoretisch-Physikalisches Institut, Friedrich-Schiller-Universität Jena, 07743 Jena, Germany
14
University of Birmingham, Birmingham, B15 2TT, UK
15
Center for Interdisciplinary Exploration and Research in Astrophysics (CIERA), Northwestern University, Evanston, IL 60208, USA
16
Instituto Nacional de Pesquisas Espaciais, 12227-010 São José dos Campos, São Paulo, Brazil
17
Gravity Exploration Institute, Cardiff University, Cardiff, CF24 3AA, UK
18
INFN, Sezione di Pisa, 56127 Pisa, Italy
19
International Centre for Theoretical Sciences, Tata Institute of Fundamental Research, Bengaluru, 560089, India
20
Gravitational Wave Science Project, National Astronomical Observatory of Japan (NAOJ), Mitaka, Tokyo 181-8588, Japan
21
Advanced Technology Center, National Astronomical Observatory of Japan (NAOJ), Mitaka, Tokyo 181-8588, Japan
22
INFN Sezione di Torino, 10125 Torino, Italy
23
Università di Napoli “Federico II”, Complesso Universitario di Monte S. Angelo, 80126 Napoli, Italy
24
Université de Lyon, Université Claude Bernard Lyon 1, CNRS, Institut Lumière Matière, 69622 Villeurbanne, France
25
Department of Physics, The University of Tokyo, Bunkyo-ku, Tokyo 113-0033, Japan
26
Research Center for the Early Universe (RESCEU), The University of Tokyo, Bunkyo-ku, Tokyo 113-0033, Japan
27
Institut de Ciències del Cosmos (ICCUB), Universitat de Barcelona, C/ Martí i Franquès 1, Barcelona, 08028, Spain
28
Laboratoire d’Annecy de Physique des Particules (LAPP), Univ. Grenoble Alpes, Université Savoie Mont Blanc, CNRS/IN2P3, 74941 Annecy, France
29
Gran Sasso Science Institute (GSSI), 67100, L’Aquila, Italy, SUPA, University of Strathclyde, Glasgow, G1 1XQ, UK
30
Dipartimento di Scienze Matematiche, Informatiche e Fisiche, Università di Udine, 33100 Udine, Italy
31
INFN, Sezione di Trieste, 34127 Trieste, Italy
32
Embry-Riddle Aeronautical University, Prescott, AZ 86301, USA
33
Université de Paris, CNRS, Astroparticule et Cosmologie, 75006 Paris, France
34
Institute for Cosmic Ray Research (ICRR), KAGRA Observatory, The University of Tokyo, Kashiwa, Chiba 277-8582, Japan
35
Accelerator Laboratory, High Energy Accelerator Research Organization (KEK), Tsukuba, Ibaraki 305-0801, Japan
36
Earthquake Research Institute, The University of Tokyo, Bunkyo-ku, Tokyo 113-0032, Japan
37
California State University Fullerton, Fullerton, CA 92831, USA
38
Université Paris-Saclay, CNRS/IN2P3, IJCLab, 91405 Orsay, France
39
European Gravitational Observatory (EGO), 56021 Cascina, Pisa, Italy
40
Chennai Mathematical Institute, Chennai, 603103, India
41
Department of Mathematics and Physics, Gravitational Wave Science Project, Hirosaki University, Hirosaki, Aomori 036-8561, Japan
42
Columbia University, New York, NY 10027, USA
43
Kamioka Branch, National Astronomical Observatory of Japan (NAOJ), Kamioka-cho, Hida, Gifu 506-1205, Japan
44
The Graduate University for Advanced Studies (SOKENDAI), Mitaka, Tokyo 181-8588, Japan
45
Università degli Studi di Urbino “Carlo Bo”, 61029 Urbino, Italy
46
INFN, Sezione di Firenze, 50019 Sesto Fiorentino, Firenze, Italy
47
INFN, Sezione di Roma, 00185 Roma, Italy
48
Université catholique de Louvain, 1348 Louvain-la-Neuve, Belgium
49
Nikhef, Science Park 105, 1098 XG Amsterdam, The Netherlands
50
King’s College London, University of London, London, WC2R 2LS, UK
51
Korea Institute of Science and Technology Information (KISTI), Yuseong-gu, Daejeon 34141, Korea
52
National Institute for Mathematical Sciences, Yuseong-gu, Daejeon 34047, Korea
53
Christopher Newport University, Newport News, VA 23606, USA
54
International College, Osaka University, Toyonaka, Osaka 560-0043, Japan
55
School of High Energy Accelerator Science, The Graduate University for Advanced Studies (SOKENDAI), Tsukuba, Ibaraki 305-0801, Japan
56
University of Oregon, Eugene, OR 97403, USA
57
Syracuse University, Syracuse, NY 13244, USA
58
Université de Liège, 4000 Liège, Belgium
59
University of Minnesota, Minneapolis, MN 55455, USA
60
Università degli Studi di Milano-Bicocca, 20126 Milano, Italy
61
INFN, Sezione di Milano-Bicocca, 20126 Milano, Italy
62
INAF, Osservatorio Astronomico di Brera sede di Merate, 23807 Merate, Lecco, Italy
63
LIGO Hanford Observatory, Richland, WA 99352, USA
64
Dipartimento di Medicina, Chirurgia e Odontoiatria “Scuola Medica Salernitana”, Università di Salerno, 84081 Baronissi, Salerno, Italy
65
SUPA, University of Glasgow, Glasgow, G12 8QQ, UK
66
LIGO Laboratory, Massachusetts Institute of Technology, Cambridge, MA 02139, USA
67
Wigner RCP, RMKI, Konkoly Thege Miklós út 29-33, 1121 Budapest, Hungary
68
University of Florida, Gainesville, FL 32611, USA
69
Stanford University, Stanford, CA 94305, USA
70
Università di Pisa, 56127 Pisa, Italy
71
INFN, Sezione di Perugia, 06123 Perugia, Italy
72
Università di Perugia, 06123 Perugia, Italy
73
Università di Padova, Dipartimento di Fisica e Astronomia, 35131 Padova, Italy
74
INFN, Sezione di Padova, 35131 Padova, Italy
75
Montana State University, Bozeman, MT 59717, USA
76
Institute for Plasma Research, Bhat, Gandhinagar 382428, India
77
Nicolaus Copernicus Astronomical Center, Polish Academy of Sciences, 00-716 Warsaw, Poland
78
Dipartimento di Ingegneria, Università del Sannio, 82100 Benevento, Italy
79
OzGrav, University of Adelaide, Adelaide, South Australia, 5005, Australia
80
California State University, Los Angeles, 5151 State University Dr, Los Angeles, CA 90032, USA
81
INFN, Sezione di Genova, 16146 Genova, Italy
82
OzGrav, University of Western Australia, Crawley, Western Australia, 6009, Australia
83
RRCAT, Indore, Madhya Pradesh, 452013, India
84
GRAPPA, Anton Pannekoek Institute for Astronomy and Institute for High-Energy Physics, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands
85
Missouri University of Science and Technology, Rolla, MO 65409, USA
86
Faculty of Physics, Lomonosov Moscow State University, Moscow, 119991, Russia
87
Università di Trento, Dipartimento di Fisica, 38123 Povo, Trento, Italy
88
INFN, Trento Institute for Fundamental Physics and Applications, 38123 Povo, Trento, Italy
89
SUPA, University of the West of Scotland, Paisley, PA1 2BE, UK
90
Bar-Ilan University, Ramat Gan, 5290002, Israel
91
Artemis, Université Côte d’Azur, Observatoire de la Côte d’Azur, CNRS, 06304 Nice, France
92
Dipartimento di Fisica “E.R. Caianiello”, Università di Salerno, 84084 Fisciano, Salerno, Italy
93
INFN, Sezione di Napoli, Gruppo Collegato di Salerno, Complesso Universitario di Monte S. Angelo, 80126 Napoli, Italy
94
Università di Roma “La Sapienza”, 00185 Roma, Italy
95
Univ. Rennes, CNRS, Institut FOTON – UMR6082, 3500 Rennes, France
96
Indian Institute of Technology Bombay, Powai, Mumbai 400 076, India
97
INFN, Laboratori Nazionali del Gran Sasso, 67100 Assergi, Italy
98
Laboratoire Kastler Brossel, Sorbonne Université, CNRS, ENS-Université PSL, Collège de France, 75005 Paris, France
99
Astronomical Observatory Warsaw University, 00-478 Warsaw, Poland
100
University of Maryland, College Park, MD 20742, USA
101
Max Planck Institute for Gravitational Physics (Albert Einstein Institute), 14476 Potsdam, Germany
102
L2IT, Laboratoire des 2 Infinis – Toulouse, Université de Toulouse, CNRS/IN2P3, UPS, 31062 Toulouse Cedex 9, France
103
School of Physics, Georgia Institute of Technology, Atlanta, GA 30332, USA
104
IGFAE, Campus Sur, Universidade de Santiago de Compostela, 15782 Santiago de Compostela, Spain
105
The Chinese University of Hong Kong, Shatin, NT, Hong Kong
106
Stony Brook University, Stony Brook, NY 11794, USA
107
Center for Computational Astrophysics, Flatiron Institute, New York, NY 10010, USA
108
NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA
109
Dipartimento di Fisica, Università degli Studi di Genova, 16146 Genova, Italy
110
Institute for Gravitational and Subatomic Physics (GRASP), Utrecht University, Princetonplein 1, 3584 CC Utrecht, The Netherlands
111
RESCEU, University of Tokyo, Tokyo, 113-0033, Japan
112
Department of Astronomy, Beijing Normal University, Beijing, 100875, PR China
113
OzGrav, University of Melbourne, Parkville, Victoria 3010, Australia
114
Università degli Studi di Sassari, 07100 Sassari, Italy
115
INFN, Laboratori Nazionali del Sud, 95125 Catania, Italy
116
Università di Roma Tor Vergata, 00133 Roma, Italy
117
INFN, Sezione di Roma Tor Vergata, 00133 Roma, Italy
118
University of Sannio at Benevento, 82100, Benevento, Italy and INFN, Sezione di Napoli, 80100 Napoli, Italy
119
Villanova University, 800 Lancaster Ave, Villanova, PA 19085, USA
120
Departamento de Astronomía y Astrofísica, Universitat de València, 46100 Burjassot, València, Spain
121
Universität Hamburg, 22761 Hamburg, Germany
122
Rochester Institute of Technology, Rochester, NY 14623, USA
123
National Tsing Hua University, Hsinchu, 30013, Taiwan
124
Department of Applied Physics, Fukuoka University, Jonan, Fukuoka, Fukuoka 814-0180, Japan
125
OzGrav, Charles Sturt University, Wagga Wagga, New South Wales, 2678, Australia
126
Department of Physics, Tamkang University, Danshui Dist., New Taipei, 25137, Taiwan
127
Department of Physics and Institute of Astronomy, National Tsing Hua University, Hsinchu, 30013, Taiwan
128
Department of Physics, Center for High Energy and High Field Physics, National Central University, Zhongli, Taoyuan, 32001, Taiwan
129
CaRT, California Institute of Technology, Pasadena, CA 91125, USA
130
Department of Physics, National Tsing Hua University, Hsinchu, 30013, Taiwan
131
Dipartimento di Ingegneria Industriale (DIIN), Università di Salerno, 84084 Fisciano, Salerno, Italy
132
Institute of Physics, Academia Sinica, Nankang, Taipei 11529, Taiwan
133
Université Lyon, Université Claude Bernard Lyon 1, CNRS, IP2I Lyon/IN2P3, UMR 5822, 69622 Villeurbanne, France
134
Seoul National University, Seoul, 08826, South Korea
135
Pusan National University, Busan, 46241, South Korea
136
INAF, Osservatorio Astronomico di Padova, 35122 Padova, Italy
137
University of Arizona, Tucson, AZ 85721, USA
138
Rutherford Appleton Laboratory, Didcot, OX11 0DE, UK
139
OzGrav, Swinburne University of Technology, Hawthorn, VIC 3122, Australia
140
Université Libre de Bruxelles, Avenue Franklin Roosevelt 50, 1050 Bruxelles, Belgium
141
Universitat de les Illes Balears, IAC3–IEEC, 07122 Palma de Mallorca, Spain
142
Université Libre de Bruxelles, Brussels, 1050, Belgium
143
Departamento de Matemáticas, Universitat de València, 46100 Burjassot, València, Spain
144
Texas Tech University, Lubbock, TX 79409, USA
145
The Pennsylvania State University, University Park, PA 16802, USA
146
University of Rhode Island, Kingston, RI 02881, USA
147
The University of Texas Rio Grande Valley, Brownsville, TX 78520, USA
148
Bellevue College, Bellevue, WA 98007, USA
149
Scuola Normale Superiore, Piazza dei Cavalieri, 7, 56126 Pisa, Italy
150
MTA-ELTE Astrophysics Research Group, Institute of Physics, Eötvös University, Budapest, 1117, Hungary
151
Maastricht University, PO Box 616, 6200 MD Maastricht, The Netherlands
152
University of Portsmouth, Portsmouth, PO1 3FX, UK
153
The University of Sheffield, Sheffield, S10 2TN, UK
154
Université Lyon, Université Claude Bernard Lyon 1, CNRS, Laboratoire des Matériaux Avancés (LMA), IP2I Lyon/IN2P3, UMR 5822, 69622 Villeurbanne, France
155
Dipartimento di Scienze Matematiche, Fisiche e Informatiche, Università di Parma, 43124 Parma, Italy
156
INFN, Sezione di Milano Bicocca, Gruppo Collegato di Parma, 43124 Parma, Italy
157
Physik-Institut, University of Zurich, Winterthurerstrasse 190, 8057 Zurich, Switzerland
158
University of Chicago, Chicago, IL 60637, USA
159
Université de Strasbourg, CNRS, IPHC UMR 7178, 67000 Strasbourg, France
160
West Virginia University, Morgantown, WV 26506, USA
161
Montclair State University, Montclair, NJ 07043, USA
162
Colorado State University, Fort Collins, CO 80523, USA
163
Institute for Nuclear Research, Hungarian Academy of Sciences, Bem t’er 18/c, 4026 Debrecen, Hungary
164
Department of Physics, University of Texas, Austin, TX 78712, USA
165
CNR-SPIN, c/o Università di Salerno, 84084 Fisciano, Salerno, Italy
166
Scuola di Ingegneria, Università della Basilicata, 85100 Potenza, Italy
167
Gravitational Wave Science Project, National Astronomical Observatory of Japan (NAOJ), Mitaka, Tokyo 181-8588, Japan
168
Observatori Astronòmic, Universitat de València, 46980 Paterna, València, Spain
169
The University of Utah, Salt Lake, UT 84112, USA
170
Kenyon College, Gambier, OH 43022, USA
171
Vrije Universiteit Amsterdam, 1081 HV Amsterdam, The Netherlands
172
Department of Astronomy, The University of Tokyo, Mitaka, Tokyo 181-8588, Japan
173
Faculty of Engineering, Niigata University, Nishi-ku, Niigata, Niigata 950-2181, Japan
174
State Key Laboratory of Magnetic Resonance and Atomic and Molecular Physics, Innovation Academy for Precision Measurement Science and Technology (APM), Chinese Academy of Sciences, Xiao Hong Shan, Wuhan 430071, PR China
175
University of Szeged, Dóm tér 9, Szeged, 6720, Hungary
176
Universiteit Gent, 9000 Gent, Belgium
177
Cornell University, Ithaca, NY 14850, USA
178
University of British Columbia, Vancouver, BC V6T 1Z4, Canada
179
Tata Institute of Fundamental Research, Mumbai, 400005, India
180
INAF, Osservatorio Astronomico di Capodimonte, 80131 Napoli, Italy
181
The University of Mississippi, University, Mississippi, MS 38677, USA
182
University of Michigan, Ann Arbor, MI 48109, USA
183
Texas A&M University, College Station, TX 77843, USA
184
Department of Physics, Ulsan National Institute of Science and Technology (UNIST), Ulju-gun, Ulsan 44919, Korea
185
Applied Research Laboratory, High Energy Accelerator Research Organization (KEK), Tsukuba, Ibaraki 305-0801, Japan
186
Dipartimento di Fisica, Università di Trieste, 34127 Trieste, Italy
187
Shanghai Astronomical Observatory, Chinese Academy of Sciences, Shanghai, 200030, PR China
188
American University, Washington, DC 20016, USA
189
Faculty of Science, University of Toyama, Toyama, Toyama 930-8555, Japan
190
Institute for Cosmic Ray Research (ICRR), KAGRA Observatory, The University of Tokyo, Kamioka-cho, Hida, Gifu 506-1205, Japan
191
Carleton College, Northfield, MN 55057, USA
192
University of California, Berkeley, CA 94720, USA
193
Maastricht University, 6200 MD Maastricht, The Netherlands
194
College of Industrial Technology, Nihon University, Narashino, Chiba 275-8575, Japan
195
Graduate School of Science and Technology, Niigata University, Nishi-ku, Niigata, Niigata 950-2181, Japan
196
Department of Physics, National Taiwan Normal University, Sec. 4, Taipei, 116, Taiwan
197
Astronomy & Space Science, Chungnam National University, Yuseong-gu, Daejeon 34134, Korea
198
Department of Physics and Mathematics, Aoyama Gakuin University, Sagamihara, Kanagawa 252-5258, Japan
199
Kavli Institute for Astronomy and Astrophysics, Peking University, Haidian, Beijing, 100871, PR China
200
Yukawa Institute for Theoretical Physics (YITP), Kyoto University, Sakyou-ku, Kyoto, Kyoto 606-8502, Japan
201
Graduate School of Science and Engineering, University of Toyama, Toyama, Toyama 930-8555, Japan
202
Department of Physics, Graduate School of Science, Osaka City University, Sumiyoshi-ku, Osaka, Osaka 558-8585, Japan
203
Nambu Yoichiro Institute of Theoretical and Experimental Physics (NITEP), Osaka City University, Sumiyoshi-ku, Osaka, Osaka 558-8585, Japan
204
Institute of Space and Astronautical Science (JAXA), Chuo-ku, Sagamihara, Kanagawa 252-0222, Japan
205
Directorate of Construction, Services & Estate Management, Mumbai, 400094, India
206
Vanderbilt University, Nashville, TN 37235, USA
207
Universiteit Antwerpen, Prinsstraat 13, 2000 Antwerpen, Belgium
208
University of Białystok, 15-424 Białystok, Poland
209
Department of Physics, Ewha Womans University, Seodaemun-gu, Seoul 03760, Korea
210
National Astronomical Observatories, Chinese Academic of Sciences, Chaoyang, Beijing, PR China
211
School of Astronomy and Space Science, University of Chinese Academy of Sciences, Chaoyang, Beijing, PR China
212
University of Southampton, Southampton, SO17 1BJ, UK
213
Institute for Cosmic Ray Research (ICRR), The University of Tokyo, Kashiwa, Chiba 277-8582, Japan
214
Faculty of Science, University of Toyama, Toyama, Toyama 930-8555, Japan
215
Chung-Ang University, Seoul, 06974, South Korea
216
Institut de Física d’Altes Energies (IFAE), Barcelona Institute of Science and Technology, and ICREA, 08193 Barcelona, Spain
217
Graduate School of Science, Tokyo Institute of Technology, Meguro-ku, Tokyo 152-8551, Japan
218
University of Washington Bothell, Bothell, WA 98011, USA
219
Institute of Applied Physics, Nizhny Novgorod, 603950, Russia
220
Ewha Womans University, Seoul, 03760, South Korea
221
Inje University Gimhae, South Gyeongsang, 50834, South Korea
222
Department of Physics, Myongji University, Yongin, 17058, Korea
223
Korea Astronomy and Space Science Institute, Daejeon, 34055, South Korea
224
National Institute for Mathematical Sciences, Daejeon, 34047, South Korea
225
Ulsan National Institute of Science and Technology, Ulsan, 44919, South Korea
226
Department of Physical Science, Hiroshima University, Higashihiroshima, Hiroshima 903-0213, Japan
227
School of Physics and Astronomy, Cardiff University, Cardiff, CF24 3AA, UK
228
Institute of Astronomy, National Tsing Hua University, Hsinchu, 30013, Taiwan
229
Bard College, 30 Campus Rd, Annandale-On-Hudson, NY 12504, USA
230
Institute of Mathematics, Polish Academy of Sciences, 00656 Warsaw, Poland
231
National Center for Nuclear Research, 05-400 Świerk-Otwock, Poland
232
Instituto de Fisica Teorica, 28049 Madrid, Spain
233
Department of Physics, Nagoya University, Chikusa-ku, Nagoya, Aichi 464-8602, Japan
234
Université de Montréal/Polytechnique, Montreal, Quebec H3T 1J4, Canada
235
Laboratoire Lagrange, Université Côte d’Azur, Observatoire Côte d’Azur, CNRS, 06304 Nice, France
236
Department of Physics, Hanyang University, Seoul, 04763, Korea
237
Sungkyunkwan University, Seoul, 03063, South Korea
238
NAVIER, École des Ponts, Univ. Gustave Eiffel, CNRS, Marne-la-Vallée, France
239
Department of Physics, National Cheng Kung University, Tainan, 701, Taiwan
240
National Center for High-performance computing, National Applied Research Laboratories, Hsinchu Science Park, Hsinchu, 30076, Taiwan
241
Institute for High-Energy Physics, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands
242
NASA Marshall Space Flight Center, Huntsville, AL 35811, USA
243
University of Washington, Seattle, WA 98195, USA
244
Dipartimento di Matematica e Fisica, Università degli Studi Roma Tre, 00146 Roma, Italy
245
INFN, Sezione di Roma Tre, 00146 Roma, Italy
246
ESPCI, CNRS, 75005 Paris, France
247
Concordia University Wisconsin, Mequon, WI 53097, USA
248
Università di Camerino, Dipartimento di Fisica, 62032 Camerino, Italy
249
School of Physics Science and Engineering, Tongji University, Shanghai, 200092, PR China
250
Southern University and A&M College, Baton Rouge, LA 70813, USA
251
Centre Scientifique de Monaco, 8 Quai Antoine Ier, 98000, Monaco
252
Institute for Photon Science and Technology, The University of Tokyo, Bunkyo-ku, Tokyo 113-8656, Japan
253
Indian Institute of Technology Madras, Chennai, 600036, India
254
Saha Institute of Nuclear Physics, Bidhannagar, West Bengal, 700064, India
255
The Applied Electromagnetic Research Institute, National Institute of Information and Communications Technology (NICT), Koganei, Tokyo 184-8795, Japan
256
Institut des Hautes Etudes Scientifiques, 91440 Bures-sur-Yvette, France
257
Faculty of Law, Ryukoku University, Fushimi-ku, Kyoto, Kyoto 612-8577, Japan
258
Indian Institute of Science Education and Research, Kolkata, Mohanpur, West Bengal, 741252, India
259
Department of Astrophysics/IMAPP, Radboud University Nijmegen, PO Box 9010, 6500 GL Nijmegen, The Netherlands
260
Department of Physics, University of Notre Dame, Notre Dame, IN 46556, USA
261
Consiglio Nazionale delle Ricerche – Istituto dei Sistemi Complessi, Piazzale Aldo Moro 5, 00185 Roma, Italy
262
Korea Astronomy and Space Science Institute (KASI), Yuseong-gu, Daejeon 34055, Korea
263
Hobart and William Smith Colleges, Geneva, NY 14456, USA
264
International Institute of Physics, Universidade Federal do Rio Grande do Norte, Natal, RN 59078-970, Brazil
265
Museo Storico della Fisica e Centro Studi e Ricerche “Enrico Fermi”, 00184 Roma, Italy
266
Lancaster University, Lancaster, LA1 4YW, UK
267
Università di Trento, Dipartimento di Matematica, 38123 Povo, Trento, Italy
268
Indian Institute of Science Education and Research, Pune, Maharashtra 411008, India
269
Dipartimento di Fisica, Università degli Studi di Torino, 10125 Torino, Italy
270
Indian Institute of Technology, Palaj, Gandhinagar, Gujarat 382355, India
271
Department of Physics, Kyoto University, Sakyou-ku, Kyoto, Kyoto 606-8502, Japan
272
Department of Electronic Control Engineering, National Institute of Technology, Nagaoka College, Nagaoka, Niigata 940-8532, Japan
273
Departamento de Matemática da Universidade de Aveiro and Centre for Research and Development in Mathematics and Applications, Campus de Santiago, 3810-183 Aveiro, Portugal
274
Marquette University, 11420 W. Clybourn St., Milwaukee, WI 53233, USA
275
Graduate School of Science and Engineering, Hosei University, Koganei, Tokyo 184-8584, Japan
276
Faculty of Science, Toho University, Funabashi, Chiba 274-8510, Japan
277
Faculty of Information Science and Technology, Osaka Institute of Technology, Hirakata, Osaka 573-0196, Japan
278
Università di Firenze, Sesto Fiorentino, 50019, Italy
279
INAF, Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, 50125 Firenze, Italy
280
Indian Institute of Technology Hyderabad, Sangareddy, Khandi, Telangana 502285, India
281
iTHEMS (Interdisciplinary Theoretical and Mathematical Sciences Program), The Institute of Physical and Chemical Research (RIKEN), Wako, Saitama 351-0198, Japan
282
INAF, Osservatorio di Astrofisica e Scienza dello Spazio, 40129 Bologna, Italy
283
Department of Space and Astronautical Science, The Graduate University for Advanced Studies (SOKENDAI), Sagamihara, Kanagawa 252-5210, Japan
284
Andrews University, Berrien Springs, MI 49104, USA
285
Research Center for Space Science, Advanced Research Laboratories, Tokyo City University, Setagaya, Tokyo 158-0082, Japan
286
Institute for Cosmic Ray Research (ICRR), Research Center for Cosmic Neutrinos (RCCN), The University of Tokyo, Kashiwa, Chiba 277-8582, Japan
287
National Metrology Institute of Japan, National Institute of Advanced Industrial Science and Technology, Tsukuba, Ibaraki 305-8568, Japan
288
Dipartimento di Scienze Aziendali – Management and Innovation Systems (DISA-MIS), Università di Salerno, 84084 Fisciano, Salerno, Italy
289
Van Swinderen Institute for Particle Physics and Gravity, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands
290
Faculty of Science, Department of Physics, The Chinese University of Hong Kong, Shatin, NT, Hong Kong
291
Vrije Universiteit Brussel, Boulevard de la Plaine 2, 1050 Ixelles, Belgium
292
Department of Communications Engineering, National Defense Academy of Japan, Yokosuka, Kanagawa 239-8686, Japan
293
Department of Physics, University of Florida, Gainesville, FL 32611, USA
294
Department of Information and Management Systems Engineering, Nagaoka University of Technology, Nagaoka, Niigata 940-2188, Japan
295
Vrije Universiteit Amsterdam, 1081 HV Amsterdam, The Netherlands
296
Department of Physics and Astronomy, Sejong University, Gwangjin-gu, Seoul 143-747, Korea
297
Department of Electrophysics, National Chiao Tung University, Hsinchu, Taiwan
298
Department of Physics, Rikkyo University, Toshima-ku, Tokyo 171-8501, Japan
Received:
2
June
2021
Accepted:
9
September
2021
Intermediate-mass black holes (IMBHs) span the approximate mass range 100−105 M⊙, between black holes (BHs) that formed by stellar collapse and the supermassive BHs at the centers of galaxies. Mergers of IMBH binaries are the most energetic gravitational-wave sources accessible by the terrestrial detector network. Searches of the first two observing runs of Advanced LIGO and Advanced Virgo did not yield any significant IMBH binary signals. In the third observing run (O3), the increased network sensitivity enabled the detection of GW190521, a signal consistent with a binary merger of mass ∼150 M⊙ providing direct evidence of IMBH formation. Here, we report on a dedicated search of O3 data for further IMBH binary mergers, combining both modeled (matched filter) and model-independent search methods. We find some marginal candidates, but none are sufficiently significant to indicate detection of further IMBH mergers. We quantify the sensitivity of the individual search methods and of the combined search using a suite of IMBH binary signals obtained via numerical relativity, including the effects of spins misaligned with the binary orbital axis, and present the resulting upper limits on astrophysical merger rates. Our most stringent limit is for equal mass and aligned spin BH binary of total mass 200 M⊙ and effective aligned spin 0.8 at 0.056 Gpc−3 yr−1 (90% confidence), a factor of 3.5 more constraining than previous LIGO-Virgo limits. We also update the estimated rate of mergers similar to GW190521 to 0.08 Gpc−3 yr−1.
Key words: gravitational waves / stars: black holes / black hole physics
© ESO 2022
1. Introduction
Black holes are classified according to their masses: stellar-mass black holes (BHs) are those with mass below ∼100 M⊙, formed by stellar collapse, while supermassive BHs (Ferrarese & Ford 2005) at the centers of galaxies have masses above 105 M⊙. Between stellar-mass and supermassive BHs is the realm of IMBHs and BHs with masses in the 100−105 M⊙ range (van der Marel 2004; Miller & Colbert 2004; Ebisuzaki et al. 2001; Koliopanos 2017; Inayoshi et al. 2020).
Stellar evolution models suggest that BHs with masses up to ∼65 M⊙ are the result of core-collapse in massive stars (Woosley 2017, 2019; Giacobbo et al. 2018; Farmer et al. 2019, 2020; Mapelli et al. 2020). The final fate of the star is determined by the mass of the helium core alone. Stars with helium core masses in the ∼32−64 M⊙ range undergo pulsational pair instability, leaving behind remnant BHs of masses below ∼65 M⊙ (Fowler & Hoyle 1964; Barkat et al. 1967). When the helium core mass is in the ∼64−135 M⊙ range pair-instability drives the supernova explosion and leaves no remnant; while stars with a helium core mass greater than ∼135 M⊙ are expected to directly collapse to intermediate-mass BHs. Thus, pair-instability (PI) prevents the formation of heavier BHs from core-collapse, and suggests a mass gap between ∼65−120 M⊙ in the BH population known as PI supernova (PISN) mass gap (Bond et al. 1984; Woosley et al. 2007; Woosley & Heger 2021). Possible IMBH formation channels also include the direct collapse of massive first-generation, low-metallicity Population III stars (Fryer et al. 2001; Heger et al. 2003; Spera & Mapelli 2017; Madau & Rees 2001; Heger & Woosley 2002), and multiple, hierarchical collisions of stars in dense young star clusters (Miller & Hamilton 2002; O’Leary et al. 2006; Giersz et al. 2015; Mapelli 2016), among others. It is not currently known how supermassive black holes form. The hierarchical merger of IMBH systems in a dense environment is among the putative formation channels for supermassive BHs (King & Dehnen 2005; Volonteri 2010; Mezcua 2017; Koliopanos 2017).
Several IMBH candidates are suggested by electromagnetic observations, but these lack conclusive confirmation (Greene et al. 2020). Observations include direct kinematical measurement of the mass of the central BH in massive star clusters and galaxies (Mezcua 2017; Miller & Hamilton 2002; Atakan Gurkan et al. 2004; Anderson & van der Marel 2010; Baumgardt et al. 2003; Pasham et al. 2015; Vitral & Mamon 2021). Other possible evidence of IMBHs includes extrapolation of scaling relations between the masses of host galaxies and their central supermassive BH to the mass range of globular clusters (Graham 2012; Graham & Scott 2013; Kormendy & Ho 2013). In addition, observations of characteristic imprints on the surface brightness, mass-to-light ratio and/or line-of-sight velocities also suggest that dense globular clusters harbour IMBHs (van den Bosch et al. 2006; Gebhardt et al. 2005; Noyola et al. 2008; Lützgendorf et al. 2011; Kızıltan et al. 2017). Controversy exists regarding the interpretation of these observations, as some of them can also be explained by a high concentration of stellar-mass BHs or the presence of binaries (Baumgardt et al. 2003; Anderson & van der Marel 2010; Lanzoni et al. 2013). Empirical mass scaling relations of quasi-periodic oscillations in luminous X-ray sources have also provided evidence of IMBHs (Remillard & McClintock 2006). Ultraluminous X-ray sources exceed the Eddington luminosity of an accreting stellar-mass BH (Kaaret et al. 2017; Farrell et al. 2009). An accreting IMBH is a favored explanation in several cases (Kaaret et al. 2001; Miller & Colbert 2004). However, neutron stars or stellar-mass BH emitting above their Eddington luminosity could also account for such observations (Bachetti et al. 2014; Israel et al. 2017). The strongest IMBH candidate among them is HLX-1, an hyper-luminous X-ray source indicating an IMBH mass of ∼0.3−30 × 104 M⊙ (Farrell et al. 2009; Godet et al. 2009; Servillat et al. 2011; Webb et al. 2012; Cseh et al. 2015; Soria et al. 2012). In Lin et al. (2018), an intermediate-mass black hole candidate was found in a tidal disruption event in a massive star cluster. More recently Paynter et al. (2021) claimed an IMBH detection through a gravitationally lensed gamma-ray burst.
The Advanced LIGO (Aasi et al. 2015) and Advanced Virgo (Acernese et al. 2015) interferometric gravitational wave (GW) detectors completed three observing runs between September 2015 and March 2020. The third observing run of Advanced LIGO and Advanced Virgo, O3, extended from April 1, 2019, 15:00 UTC to March 27, 2020 17:00 UTC. The recently released second gravitational-wave transient catalog provided a comprehensive summary of significant compact binary coalescence events observed up to October 1st, 2019 (Abbott et al. 2021a), reporting a total of 50 events. The corresponding binary black hole (BBH) population analysis of Abbott et al. (2021b) indicates that 99% of primary BH masses lie below m99% ∼ 60 M⊙; thus, the large majority of merging BHs have masses below a limit of ∼65 M⊙, consistently with expectations from PI.
Near the beginning of O3, an unusually high mass black hole coalescence, GW190521 (Abbott et al. 2020a), was detected. This GW signal was consistent with a coalescence of black holes of and which resulted in a remnant black hole of falling in the mass range of intermediate-mass black holes1. GW190521 provided the first conclusive evidence for the formation of an IMBH below 103 M⊙. It is a massive binary black hole system with an IMBH remnant and a primary BH in the PISN mass gap with high confidence (Abbott et al. 2020b; although, see Fishbach & Holz 2020; Nitz & Capano 2021 for an alternative interpretation). The discovery triggered a variety of investigations regarding the evolution models and the subsequent mass gap in the BH population. It also suggested the possibility of the formation of massive BHs (> 100 M⊙) via a hierarchical merger scenario in a dense environment (Abbott et al. 2020b; Kimball et al. 2021).
The Advanced LIGO and Advanced Virgo detectors are sensitive to the lower end of the IMBH binary mass range, potentially making IMBHs detectable out to cosmological distances, as is evident from GW190521. Observation of IMBH binary systems are not only interesting for massive BH formation channels, but they act as a perfect laboratory to test general relativity (Abbott et al. 2016a; Yunes et al. 2016; Yunes & Siemens 2013; Gair et al. 2013). Massive BH coalescences produce louder mergers and ringdown signals in the sensitive band of the advanced GW detectors. Furthermore, these can display prominent higher order modes that confer GWs a more complex morphology that can significantly deviate from a canonical chirp (Calderon Bustillo et al. 2020). Observations of higher order modes help to test general relativity and fundamental properties of BHs such as the no-hair theorem (Kamaretsos et al. 2012; Meidam et al. 2014; Thrane et al. 2017; Carullo et al. 2018) and BH kick measurements (Gonzalez et al. 2007; Campanelli et al. 2007; Calderón Bustillo et al. 2018). These IMBHs might be multiband events observable by both LIGO and Virgo and LISA (Amaro-Seoane et al. 2017), and could provide novel probes of cosmology and contribute to the stochastic background (Fregeau et al. 2006; Miller 2009; Jani et al. 2020; Ezquiaga & Holz 2021).
The GW signal from a massive BBH coalescence is evident as a short-duration waveform with little inspiral and mostly merger-ringdown signal, falling in the low-frequency region of the advanced detectors. With initial GW detectors (Abadie et al. 2012; Aasi et al. 2014), the IMBH binary searches were restricted in order to probe the merger-ringdown phase of the coalescing BBH system using the model waveform-independent coherent WaveBurst (cWB) (Klimenko & Mitselmakher 2004; Klimenko et al. 2005, 2006) and a ringdown templated search (Aasi et al. 2014). Improvement in the detector sensitivity at low frequencies in the advanced era made IMBH binaries a target for a matched filtering search that would probe the short inspiral phase. In Abbott et al. (2017a), we used a combined search with the matched filtering GstLAL (Messick et al. 2017; Hanna et al. 2020; Sachdev et al. 2019) search and model-independent cWB (Klimenko et al. 2011, 2016). This combined search was further extended with an additional matched filtering PyCBC search (Usman et al. 2016; Allen 2005; Dal Canton et al. 2014; Nitz et al. 2017) in Abbott et al. (2019a) using the data from the first two observing runs. No significant IMBH binary event was found in these searches.
While all the previous matched filtering searches were generic BBH searches, the improvements in the detector sensitivity at low frequencies and the IMBH merger signals’ short duration nature motivated us to use matched filter searches targeted to the IMBH mass-spin parameter space. We carried out an IMBH binary search using the entire year-long third observing run, O3, of the Advanced LIGO and Advanced Virgo detector networks with a combined search using three search algorithms: two matched-filtering-based IMBH binary searches using the PyCBC and GstLAL libraries, and the minimally modeled time-frequency-based cWB search. We searched for massive binary systems with at least one component above the expected PISN mass gap limit of 65 M⊙, and with an IMBH remnant. GW190521 remains the most significant candidate in the combined search; no other event is comparably significant. We provide the results from the combined search with the next most significant events and follow-up investigations to assess their origins.
The increased sensitivity of the O3 run allows us to set more stringent bounds on the binary merger rate density. The lack of a confirmed IMBH population as well as possible formation channels of IMBH distinct from those of stellar-mass BHs preclude us from using an overall mass model for the IMBH population. Thus, we confined all the upper limit studies to a suite of discrete points in the IMBH parameter space. We incorporated more detailed physics in selecting the suite of IMBH binary waveforms as compared to earlier upper limit studies. In Abbott et al. (2017a), we simulated a limited set of discrete mass and aligned-spin binary waveforms in the first advanced detector observation data to obtain upper limits on the merger rate. The study with the first two observation runs used the most realistic numerical relativity (NR) simulation set with aligned spins for the upper limit study (Abbott et al. 2019a). The most recent stringent merger rate upper limit is 0.2 Gpc−3 yr−1; this is for the equal mass binary system with a component mass of 100 M⊙ and component spins of dimensionless magnitude 0.8 aligned with the binary orbital angular momentum. Recently, Chandra et al. (2020) used IMBH binary systems with generically spinning BHs with total masses from 210−500 M⊙ and obtained a most stringent upper limit of 0.28 Gpc−3 yr−1 for equal-mass binaries with total mass of 210 M⊙.
We used a suite of NR simulations of GW emission from an IMBH binary system with generically spinning BHs in order to estimate our search sensitivity over the O3 data. We placed the most stringent 90% merger rate upper limit on equal mass and aligned spin BH binary of total mass 200 M⊙ and with individual BH spins of 0.8 as 0.056 Gpc−3 yr−1. The revised limit is a factor of ∼3.5 more stringent than that obtained with the first two observing runs. We also updated the merger rate for systems compatible with the source parameters of GW190521 (first estimated in Abbott et al. 2020b) at Gpc−3 yr−1 using the combined search method applied to simulated signals injected over the entire O3 data.
The paper is organized as follows: Sect. 2 summarizes the data being used for the search. Section 3 summarizes the combined search approach from the results from three distinct IMBH binary search algorithms. Section 4 discusses the search results and follow-up of the most significant candidate events. Section 5 provides a detailed discussion about the NR GW injection set used and the rate upper limits studied including the updated rate on the most significant GW190521-like systems.
2. Data summary
We carried out the analysis using O3 data from both LIGO detectors (LHO-LIGO Hanford Observatory and LLO-LIGO Livingston Observatory) and the Virgo detector. We conditioned the data in multiple steps before performing our search (Abbott et al. 2020c). The strain data, recorded from each detector, were calibrated in near-real time to produce an online data set (Viets et al. 2018; Acernese et al. 2018). A higher latency offline calibration stage identifies systematic errors and calibration configuration changes (Sun et al. 2020; Estevez et al. 2021). The analyses presented here use the offline recalibrated data from the LIGO detectors, and the Virgo detector’s online data. For this search, we considered 246.2 days, 254.1 days, and 250.8 days of observing-mode data from the Hanford, Livingston, and Virgo detectors, respectively. The joint observation time for the full network of three detectors is 156.4 days.
We then linearly subtracted spectral features of known instrumental origin using auxiliary witness sensors; that is, sensors that indicate the presence of noise causing these features. The subtraction removes calibration lines in all detectors, as well as 60 Hz harmonics produced by power mains coupling in the LIGO detectors (Driggers et al. 2019; Davis et al. 2019). Low-frequency modulation of the power mains coupling also results in side-bands around the 60 Hz line; we applied an additional non-linear noise subtraction to remove these side bands (Vajente et al. 2020).
Periods of poor data quality are marked using data quality flags separated into three categories (Abbott et al. 2020c; Fisher et al. 2020; Davis et al. 2021), which we used to exclude time segments from different searches, as described below. Category 1 flags indicate times when a detector is not operating or recording data in its nominal state; these periods are not analyzed by any search. Category 2 flags indicate periods of excess noise that are highly likely to be caused by known instrumental effects. The cWB and PyCBC searches use different sets of category 2 flags. The GstLAL search does not use category 2 flags, as discussed in Sect. 3. Category 3 flags are based on statistical correlations with auxiliary sensors. Of the analyses presented here, only the cWB search uses category 3 flags.
The candidate events in this paper are vetted in the same way as past GW events (Abbott et al. 2016b, 2021a). This validation procedure identifies data quality issues such as nonstationary noise or glitches of instrumental origin appearing in the strain data. Auxiliary sensors that monitor the detectors and environmental noise are used to check for artifacts that may either have accounted for, or contaminated, the candidate signal (Nguyen et al. 2021). For candidate events that coincide with glitches, subtraction of the glitches from the strain data was performed if possible (Cornish & Littenberg 2015; Littenberg et al. 2016; Pankow et al. 2018); otherwise recommendations are made to exclude the relevant time or frequency ranges from parameter estimation analyses. Validation assessments for individual candidate events are provided in Sect. 4 and Appendix A.
3. Search methods
In this section, we describe the analysis methods’ algorithms (pipelines) used to search the LIGO-Virgo data from O3 for IMBH binary merger signals. Such signals have short durations in the detectors’ sensitive frequency band of typically less than 1 s. Thus, methods for the detection of generic short transient GW events (bursts) may be competitive compared to search methods that use parameterized models of the expected signals (templates) from binary coalescences (e.g. Chandra et al. 2020). As in the IMBH binary search of O1 and O2 (Abbott et al. 2017a, 2019a), we employed both generic transient search methods and modeled template searches. We first describe the generic transient pipeline, cWB, in the configuration used here, and then the two template based pipelines, GstLAL and PyCBC, which have been adapted to maximize sensitivity to IMBH binary mergers. We then summarize the method used to combine the search outputs into a single candidate list and finally discuss selection criteria to distinguish IMBH binary candidates from the known heavy stellar-mass BBH population (Abbott et al. 2021a,b).
The output of a transient search algorithm or pipeline is a set of candidate events, each with an estimated time of peak strain at the participating detector(s)2. Each event is also assigned a ranking statistic value, and its significance is quantified by estimating the corresponding false alarm rate (FAR), which is the expected number of events per time that have an equal or higher ranking statistic value caused by detector noise.
The sensitivity of a search to a population of IMBH mergers can be evaluated by adding simulated signals (injections) to real GW detector strain data and analyzing the resulting data streams to output the ranking statistic and estimated FAR that each simulated signal would be assigned if present in an actual search. Specific simulation campaigns will be described in detail in Sect. 5 and sensitivity estimates from individual search pipelines are included in a public data release.
3.1. cWB model waveform-independent search for IMBH binaries
cWB (Klimenko & Mitselmakher 2004; Klimenko et al. 2005, 2006, 2011; Klimenko et al. 2016; Drago et al. 2020) is a GW search that uses minimal assumptions on signal morphology to detect and reconstruct GW transients. The search identifies coincident energy across the network of detectors to classify GW signals. The cWB search has been participating in the search for IMBH signals since Initial LIGO’s fifth science run (Abadie et al. 2012). The algorithm uses a multi-resolution wavelet transform, known as the Wilson Daubechies Meyer wavelet transform (Necula et al. 2012), to map the multi-detector data into the time-frequency domain, as blocks of a fixed time-frequency area known as pixels. The algorithm selects pixels with excess energy above the expected noise fluctuation and groups them into clusters, referred to as candidate events. The collection and clustering of pixels differ based on the target source (Klimenko et al. 2016). Each candidate event is ranked according to its coherent ignalto- noise ratio (S/N) statistic (Klimenko et al. 2016), which incorporates the estimated coherent energy and residual noise energy. An additional threshold is applied to the network correlation which provides the measure on the event correlation across multiple detectors in the network. The cWB algorithm reconstructs the source sky location and whitened signal waveforms using the constrained maximum likelihood method (Klimenko et al. 2016).
We estimated the FAR of a search event with time lag analysis: data from one or more detectors were time-shifted by more than 1 s with respect to other detectors in the network; then, cWB identifies events in this time-shifted data. Since the time shift is greater than the GW time of flight between detectors, this analysis estimates the rate and distribution of false alarms. The analysis is repeated many times with different time-shifts, yielding a total analyzed background time Tbkg. For a given search event, the FAR value is estimated as the number of background events with coherent S/N greater than the value assigned to the event, divided by Tbkg.
The model-independent nature of cWB search makes it susceptible to the incorrect classification of noise artifacts. We applied a series of signal-dependent vetoes based on the time-frequency morphology and energy distribution properties to remove spurious noisy transients. We tuned the veto values based on the extensive simulation of IMBH binary signals (see Appendix A of Gayathri et al. 2019). We divided the cWB search for quasi-circular BBH signals into two separate configurations: a high-mass search and a low-mass search, depending on the central frequency fc of the GW signal. For a compact binary merger signal, fc is inversely proportional to the redshifted total mass Mz = (1 + z)M, where M is the source frame total mass and z is the source redshift. We then optimized the low-mass search sensitivity for signals with fc > 80 Hz (the BBH regime), and the high-mass search sensitivity for signals with fc < 80 Hz (the IMBH regime). In practice, a cut fc > 60 Hz was imposed in the low-mass search and fc < 100 Hz in the high-mass search, resulting in an overlap region covering 60−100 Hz. In the O3 search, we combined the two searches by applying a trials factor of 2 to the estimated FAR for events in the overlap region. This improved the overall search sensitivity to borderline IMBH events (Szczepańczyk et al. 2021).
The cWB search analyzes data from all three detectors in low latency. However, the follow-up offline cWB analysis does not improve detection efficiency with the inclusion of Virgo. This is primarily due to the additional noise in the Virgo detector. Thus, at a given time, the cWB search uses the best available (most sensitive) two-detector network configuration. This ensures that the cWB search does not analyze the same data with multiple detector configurations. In the case where any event showed high significance in the low latency cWB analysis with the three-detector network and low significance in the offline cWB analysis with the best two-detector configuration, we reanalyzed that observing time with both the LLO-LHO-Virgo and LLO-LHO networks and applied a trial factor of 2 to the minimum FAR over the two networks for the final significance.
3.2. Template-based searches for IMBH mergers
For GW signals whose forms are known or can be theoretically predicted, search sensitivity is optimized by the use of matched filter templates that suppress noise realizations inconsistent with the predicted signals. Since the binary parameters are a priori unknown, a discrete set (bank) of templates is used in order to cover signal parameter values within a predetermined range with a specified minimum waveform accuracy (Sathyaprakash & Dhurandhar 1991; Owen 1996). General binary black hole coalescence signals bear the imprint of component spins misaligned with the orbital axis, causing orbital precession, and potentially also of orbital eccentricity. It is a so far unsolved problem to implement an optimal search over such a complex space of signals.
Instead, the searches presented here restrict the signal model to the dominant mode of GW emission from quasi-circular, non-precessing binaries (Ajith et al. 2011), that is, with component spins perpendicular to the orbital plane. Both the GstLAL and PyCBC searches use the SEOBNRv4 waveform approximate (Bohé et al. 2017) as template waveforms, implemented as a reduced-order model (Pürrer 2016) for computational speed. These templates may still have high matches to signals from precessing or eccentric binaries; however, in general, sensitivity to such signals will be reduced due to lower matches with template waveforms.
Each detector’s strain time series is then correlated with each template to produce a matched filter time series. Single-detector candidates were generated by identifying maxima of the matched filter S/N above a predetermined threshold value. However, during times of known disturbances in detector operation, or during very high amplitude non-Gaussian excursions in the strain data, candidates are either not produced or are discarded, since such high-S/N maxima are very likely to be artifacts. Signal consistency checks such as chi-squared (Allen 2005) verifications are also calculated, and single-detector candidates may also be discarded for excessive deviation from the expected range of values.
If two or more detectors are operating, their single-detector candidates are compared in order to identify multi-detector candidate events that are consistent in the template parameters, time of arrival, amplitude, and waveform phase over the detector network. The resulting multi-detector events are then ranked via a statistic that depends on the properties of single-detector candidates and their consistency over the network. Finally, the statistical significance of each multi-detector event is obtained by comparing its statistic value to the distribution expected for noise events, resulting in an estimate of its FAR. In the following subsection, we briefly summarize the methods specific to each of the matched filter pipelines.
3.2.1. GstLAL search
The search for IMBH mergers executed by the matched-filter-based GstLAL pipeline (Messick et al. 2017; Hanna et al. 2020; Sachdev et al. 2019; Cannon et al. 2021) uses a template bank covering a parameter space of binaries with (redshifted) total masses in the [50, 600] M⊙ range. The mass ratios (q = m2/m1) of the binary systems covered lie between 1 and 1/10, while their spins are either aligned or anti-aligned with the total angular momentum of the system, with the dimensionless spin magnitude being less than 0.98. The analysis starts at a frequency of 10 Hz.
The S/N threshold applied for single-detector triggers is 4 for the Hanford and Livingston detectors and 3.5 for the Virgo detector. The GstLAL search pipeline applies a signal-consistency test based on the template’s auto-correlation over time. The search also uses a signal model to describe the prior probability of a binary from a given source population being detected by each template: the signal model used for this search is uniform in the log of the reduced mass of the binary.
The ranking statistic applied to candidate events is an estimate of the relative probability of the event’s parameters being caused by a GW signal as compared to noise, that is, the likelihood ratio. In addition to events formed from triggers from multiple detectors, triggers found in a single detector are also included in the search, albeit with a penalty applied to their ranking to account for the higher probability of noise origin.
The GstLAL search does not use data-quality-based vetoes of category 2 and above. Instead, the search uses data quality information known as iDQ (Essick et al. 2020; Godwin et al. 2020), from auxiliary channels monitoring the detector, to compute a penalty term in the denominator of the ranking statistic. This has been computed for both single-detector and multi-detector triggers found by the search. The non-coincident and noise-like triggers are then used to estimate the background noise probability density, which is sampled to find the estimated FAR, corresponding to the likelihood ratio for a candidate (Messick et al. 2017; Sachdev et al. 2019).
3.2.2. PyCBC search
The PyCBC-IMBH search used here (Chandra et al. 2021) covers a target space of redshifted total masses between 100 and 600 M⊙, with component masses greater than 40 M⊙ and mass ratio between 1/1 and 1/10. The components have dimensionless spins projected onto the orbital axis between −0.998 and 0.998. To reduce false alarms arising from short-duration noise transients (Cabero et al. 2019), we discarded any templates with a duration of less than 0.07 s, measured from the fixed starting frequency of 15 Hz.
The analysis preprocesses the data from each detector by windowing out very high amplitude excursions (> 50σ deviation from Gaussian noise) in the whitened strain time series (Usman et al. 2016). This gating step significantly suppresses the noise background. The S/N threshold for trigger generation was chosen as 4; any triggers in time marked by a category 2 data quality veto were discarded. We also removed LIGO triggers within ( − 1, +2.5) s of the center of a gating window, since empirically such times contain many lower-amplitude noise transients correlated with the central high amplitude glitch (Chandra et al. 2021).
Signal-consistency and sine-Gaussian discriminant tests were applied to the remaining triggers (Allen 2005; Nitz 2018). The single-detector S/N were corrected for short-term variation in the detector power spectral density (PSD) (Nitz et al. 2019; Mozzon et al. 2020), and a penalty was applied to triggers with a short-term PSD measure over 10 times the expectation from stationary noise (Chandra et al. 2021). The analysis also penalizes triggers with values above 10, where the expectation for a well-matched signal is unity. These vetoes significantly reduce the background.
The search identifies candidates by checking the consistency between triggers in 2 or 3 detectors. The resulting candidates are ranked using the expected distribution of astrophysical signal S/N, phases and times over multiple detectors, as well as models of the non-Gaussian noise distribution in each template and detector (Nitz et al. 2017; Davies et al. 2020). A FAR is assigned to each candidate event by simulating the background noise distribution using time-shifted analyses (Usman et al. 2016), similarly to cWB. The FARs for events involving different detector combinations are finally combined, as in Davies et al. (2020).
Other PyCBC-based searches overlapping the IMBH parameter region were recently presented in Nitz et al. (2019) and Abbott et al. (2021a) using two strategies: a broad parameter space search covering compact binaries with components as neutron stars up to IMBH, and a “focused” search for BBH covering a restricted range of masses and with strict cuts to suppress noise artifacts. The sensitivity of the present IMBH search (at a FAR threshold of 0.01 yr−1) to a set of simulated, generically spinning binary merger signals is increased relatively to the broad (BBH) PyCBC searches of Abbott et al. (2021a) by a factor of ∼1.5 (∼1.1) in volume time (VT) for a redshifted total mass of Mz ∈ [100−200] M⊙, and up to a factor ∼2.8 (∼12.6) for Mz ∈ [450, 600] M⊙ (Chandra et al. 2021).
3.3. Combined search
Each of our three targeted searches produces a list of candidates characterized by GPS times and FAR values. The p value for each candidate in a given search, defined as the probability of observing one or more events from the noise alone with a detection statistic as high as that of the candidate, is then
where T is the total duration of data analyzed by the search. We combined these lists to form a single list of candidates by first checking whether any events from different searches fall within a 0.1 s time window of each other; if so, we only selected the event with the lowest p value pmin. The resulting clustered events are ranked by a combined p value:
where m denotes the trial (look-elsewhere) factor (Abbott et al. 2017a, 2019a). We took m = 3 under the assumption that the noise backgrounds of our searches are independent of one another. If there is any correlation between these backgrounds, the effective trial factor will be lower, which makes m = 3 a conservative choice.
3.4. Selection of intermediate-mass black hole binaries
As noted in Abbott et al. (2021a,b), LIGO-Virgo observations include a population of black hole binaries with component masses extending up to 60 M⊙ or above and remnant masses extending up to ∼100 M⊙; thus, there is a priori no clear separation between such heavy BBH systems and the lightest IMBH binaries. We also expect search pipelines tuned for sensitivity to IMBH mergers to be capable of detecting such heavy stellar-mass BBH, since the overlap of their GW signals with those of IMBH binaries may be large. We indeed find that many such BBH systems occurring within O3a are recovered with high significance by our search.
Here, we only selected those events for which, under the assumption that the signals were produced by a quasi-circular binary black hole merger, we have clear evidence that the remnant is an IMBH of mass above 100 M⊙, and at least the primary black hole has a mass greater than the lower bound of the pair-instability mass gap. For this, which we made a conservative estimate under the standard assumptions of the 12C(α, γ)16O reaction rate, etc. to be (see e.g., Woosley 2017; Stevenson et al. 2019; Belczynski et al. 2020; Costa et al. 2021; Tanikawa et al. 2021; Vink et al. 2021). Strong evidence that this is the case for the primary component of GW190521 was presented in Abbott et al. (2020a,b). Furthermore, we can say with confidence that the maximum component mass of the black hole binary population seen up to O3a is above ∼50 M⊙, even if GW190521 is excluded from the analysis (see Fig. 2 of Abbott et al. 2021b); thus, we placed a more stringent condition on the primary mass to identify potential outliers of this population.
The selection criteria are evaluated as follows. We begin by defining the hypothesis H according to which the detector output time series d(t) is given by
where n(t) is the noise time series, taken as a realization of a zero-mean wide-sense stationary stochastic process, and h(t; θ) is the gravitational wave-signal model dependent on a set of parameters θ. We estimated the parameters θ by computing their posterior probability distribution p(θ|d, H) using Bayes’ theorem:
where p(θ|H) is the prior probability distribution, p(d|θ, H) is the likelihood function – taken as the normal distribution in the frequency domain, with the variance given by the power spectral density of the data d(t) thanks to the wide-sense stationarity assumption – and
is the evidence for the hypothesis H.
Each potential candidate is followed up with a coherent Bayesian parameter estimation analysis (Veitch et al. 2015; Lange et al. 2017; Wysocki et al. 2019). In these analyses, we modeled the GW signal as represented by precessing quasi-circular waveforms from three different families: NRSur7dq4 (Varma et al. 2019), SEOBNRv4PHM (Ossokine et al. 2020), and IMRPhenomXPHM (Pratten et al. 2021). All considered models include the effects of higher order multipole moments as well as orbital precession due to misaligned BH spins. Details of the analysis configuration follow previously published ones (Abbott et al. 2021a) and are documented in a separate paper (Abbott et al., in prep.). In particular, we considered uniform priors on the redshifted component masses, the individual spin magnitudes, and the luminosity distance proportional to its square modulus. For the source orientation and spin vectors, we employed isotropic priors.
As a quantitative criterion to select a GW event as an IMBH binary, we considered the support of the joint posterior distributions for the primary mass m1 and of the remnant mass Mf. For reference values M* = 100 M⊙ and , we labeled a candidate as an intermediate-mass black hole binary if
where p* is a reference probability threshold, chosen to be p* = 0.9. To perform the integral in Eq. (6), we constructed a Gaussian kernel density estimate to interpolate the posterior p(Mf, m1|d, H) that we used to perform the integral on a grid. Thus, the main list of candidates presented in the following section does not correspond to the complete set of events recovered by the searches, but only to those relevant to a potential astrophysical IMBH population. However, for comparison with earlier results (Abbott et al. 2021a), we also report a full list of events detected by the combined IMBH search in O3a data, including BBH events that do not fall into the IMBH region (see Sect. 4.2). A catalog of confidently detected heavy BBH systems over the complete O3 run will be provided in a subsequent publication as an update to GWTC-2.
The data for some events may not be consistent with the quasi-circular BBH signal plus Gaussian noise model, either because they contain a signal that deviates significantly from this standard BBH model, or they are affected by detector noise artifacts that cannot be removed or mitigated. In such cases, the values of p(Mf|d, H) and p(m1|d, H) extracted from the Bayesian analysis may either be inaccurate or meaningless for events arising from instrumental noise or even from a putative astrophysical source that is not a compact binary merger. Such events will not be excluded from results presented here: they are discussed individually in the following sections.
4. Search results
4.1. Candidate IMBH events
The individual searches are applied on the full O3 data with the analysis time of 0.734 yr, 0.747 yr, and 0.874 yr for the cWB, PyCBC-IMBH, and GstLAL-IMBH searches, respectively. In Table 1, we report events satisfying the criteria for potential IMBH binaries from Sect. 3.This is done in increasing order of combined p value (see Eq. (3)) up to and including the most significant event with unambiguous evidence of instrumental noise origin, 200214_224526. For completeness, we also list marginal triggers found by our combined search in Appendix A.
Events from the combined search for intermediate-mass black hole binary mergers in O3 data, sorted by their combined p-value p̄.
The top-ranked event is GW190521, and it has a highly significant combined p-value of 4.5 × 10−4. If this signal is from a quasi-circular merger, then the signal is found to be consistent with the merger of two black holes in a mildly precessing orbit with component masses of and and a remnant black hole of . This falls in the mass range of intermediate-mass black holes. A full description of GW190521 and its implications can be found in Abbott et al. (2020a, b).
The second-ranked candidate, 200114_020818, was observed on 14 January 2020 at 02:08:18 UTC and identified by the low-latency cWB search in the LHO-LLO-Virgo detector network configuration, with a FAR of < 0.04 yr−1. The event was publicly reported via GCN minutes after the event was observed (Abbott et al. 2019b). Given the significance of the low-latency alert with the three-detector configuration, we employed both LHO-LLO and LHO-LLO-Virgo networks in cWB to estimate the significance for this event: we find FARs of 15.87 yr−1 and 0.029 yr−1 for these configurations, respectively. The S/N reconstructed by cWB for each network configuration is 12.3 and 14.5, respectively. As mentioned above in Sect. 3.1, we applied a trial factor of 2 to the most significant result, obtaining a FAR of 0.058 yr−1 for the cWB search. The combined p value of this event, 0.12, is marginally significant.
We then examined possible environmental or instrumental causes for the candidate signal. Excess vibration noise could have contributed to the signal in the LIGO Hanford detector, as discussed in Appendix B.1. Furthermore, the morphology of 200114_020818 is consistent with a well-studied class of glitches known as Tomtes (Buikema et al. 2020; Davis et al. 2021). These occur multiple times per hour in LIGO Livingston. However, we are not currently able to exclude a putative, morphologically similar astrophysical signal, as there are no known instrumental auxiliary channels that couple to this glitch type. We undertook detailed model-independent event reconstruction and parameter estimation (PE) studies, which are summarized in Appendix B. Although model-independent methods and algorithms produce mutually consistent reconstructions of the event, our analysis using the available quasi-circular BBH merger waveforms does not support a consistent interpretation of the event as a binary merger signal present across the detector network. We cannot conclusively rule out an astrophysical origin for the event; however, it also appears consistent with an instrumental artifact in LLO, coincidentally with noise fluctuations in LHO and Virgo.
The third-ranked event was observed by the cWB pipeline on 14 February, 2020 at 22:45:26 UTC with a combined S/N of 13.1 in the two Advanced LIGO detectors. The event has a pcWB = 0.092 and thus a p̄ = 0.251. In addition to its marginal significance, the event has characteristics consistent with an instrumental noise transient. Excess noise due to fast scattered light (Soni et al. 2021) is present in both LLO and LHO data. At Livingston, the excess noise extends up to 70 Hz and lasts many seconds before and after the event. The Hanford scattering noise is weaker in amplitude but still overlaps completely with the duration of the event. Since it is the most significant noise event obtained in the combined search with the cWB pipeline in its production configuration considering only two-detector events, we used 200214_224526 to establish a threshold of significance for inference of IMBH merger rates (for which see Sect. 5). As 200214_224526 is likely caused by detector noise, any events with lower significance may be assumed to have a high probability of noise origin. For completeness, we discuss some marginal events from the combined search in Appendix A.
4.2. Complete O3a search results including BBH
As noted earlier in Sect. 3, the template-based searches have high sensitivity to the known population of heavy stellar-mass BBH mergers, which may be compared to searches deployed in GWTC-2 (Abbott et al. 2021a). Here, we record the complete list of significant events recovered by the combined IMBH search from O3a data in Table 2, and we supply corresponding search results from GWTC-2 for comparison. Specifically, we show outputs from the PyCBC broad parameter space and focused BBH searches (Nitz et al. 2019) and the GstLAL broad parameter space search (Sachdev et al. 2019).
Candidate events from this search for IMBH mergers in O3a data, including binary black hole mergers outside the IMBH parameter space, and a comparison with previously obtained GWTC-2 results from the template based search algorithms (Abbott et al. 2021a).
For GW190521, the PyCBC IMBH search yields a FAR of 1.4 × 10−3 yr−1, as compared to 1.1 yr−1 for the broad parameter space analysis of Abbott et al. (2020a, 2021a). This significant change was made in part because the PyCBC IMBH search is optimized for shorter duration signals, and it does not consider potential signals of total mass significantly below 100 M⊙; the mass and spin values of GW190521 are also likely not covered by the templates used in earlier PyCBC searches, which imposed a minimum duration of 0.15 s. A similar change in statistical significance is also observed for GW190602_175927 for the same reasons. However, the IMBH search results assign lower significance to GW190519_153544 and GW190706_222641 as compared to the GWTC-2 results.
The GstLAL pipeline recovers the GW190521 event at a FAR of 1.9 × 10−3 yr−1 across the O3 data. This was reported earlier (Abbott et al. 2020a, 2021a) at a FAR of 1.2 × 10−3 yr−1 over O3a. As described in Sect. 3.2.1, the GstLAL pipeline employed a dedicated search for IMBH binaries with better coverage for the heavier mass binaries than the catalog search. Also, the iDQ-based data quality information used to inform the calculation of the ranking statistics now incorporates multi-detector triggers, going against the only single-detector triggers that were used before. Differences in the significance of the events found by the IMBH specific GstLAL search presented here, compared to what was reported for O3a in Abbott et al. (2021a), can be attributed to the differences in the search settings and the data spanning over all of O3.
5. Astrophysical rates of IMBH binary coalescence
Improved detector sensitivity, updated search methods, and the detection of GW190521 allow us to obtain revised bounds on the merger rate (strictly rate density) of IMBH binaries. Due to a lack of knowledge of specific formation channels for IMBH binaries, even more so than for stellar-mass BH binaries, and the sparse observational evidence of any IMBH population, we did not consider any overall mass model for such a population. Instead, we simulated a suite of IMBH binary waveforms for discrete points in parameter space, including generically spinning component BHs derived from NR simulations. A similar campaign was carried out in Abbott et al. (2019a) using NR waveforms for IMBH binaries with component BH spins aligned with the binary orbital axis and injected into the O1 and O2 data.
5.1. Injection set
We now report on the merger rate of IMBH binary sources based on NR simulations computed by the SXS (Mroué et al. 2013), RIT (Healy et al. 2017), and GeorgiaTech (Jani et al. 2016) codes. These simulations include higher order multipoles, which may make important contributions to the detection of high-mass- and low-mass-ratio (q ≤ 1/4) binaries (Calderón Bustillo et al. 2016). Based on previous studies that measured the agreement between different NR codes (Abbott et al. 2019a), we include the following harmonic modes in our analysis: (ℓ,m)={(2, ±1),(2, ±2),(3, ±2),(3, ±3),(4, ±2),(4, ±3),(4, ±4)}.
We considered 43 IMBH binary sources with fixed source frame masses and spins, shown in Table 3. These 43 sources include a subset of 16 sources investigated in the O1-O2 IMBH binary search. This updated search includes sources with a total mass of up to 800 M⊙, and it expands the range of targeted mass ratio q to between 1/1−1/10. We also further explored the effects of the component spins on detection efficiency. Of the 43 targeted IMBH sources, four have spins aligned with the orbital axis, with effective total spin (Ajith et al. 2011) χeff ≡ (χ1, ∥ + qχ2, ∥)/(1 + q)=0.8, where χ∥ denotes the BH spin resolved along the orbital axis, and four have anti-aligned spins with χeff = −0.8. A further 11 have precessing spins: χp ≠ 0, where χp is the effective spin-precession parameter of Hannam et al. (2014), Schmidt et al. (2015).
Summary of the source frame parameters, sensitive volume time, and merger rate density upper limit at 90% confidence.
The simulated signals for each targeted source point are uniformly distributed in sky location (θ, ϕ) and inclination angle cos(ι). The source redshift z is uniformly distributed in comoving volume, according to the TT+lowP+lensing+ext cosmological parameters given in Table 4 of Planck Collaboration XIII (2016), up to a maximum redshift zmax. The signals are added to the O3 strain data, that is, injected with a uniform spacing in time approximately every 100 s over the full observing time, T0 = 363.38 days.
To avoid generating injections that are well outside any possible detection range, zmax is independently calculated for each IMBH source point. We consider values of redshift z in increments of 0.05 and calculate a conservative upper bound on the optimal three-detector network S/N, S/Nnet, for each z. To bound the optimal S/N in a single detector, we assume the source is face-on cos(ι)=1, located directly overhead the detector, and we estimate the detector’s PSD using ∼8 h of typical O3 data. For precessing waveforms, the ι is set at 10 Hz. We determined the maximum redshift by requiring the following: S/Nnet(zmax)∼5. This results in a range of zmax across all targeted sources from 0.05 for the (400 + 400) M⊙ anti-aligned spin source to 2.75 for the (100 + 100) M⊙ aligned-spin source, as in Table 3.
When generating the injection parameters, we imposed an additional threshold S/Nnet > 5 to limit the number of simulations injected into detection pipelines that have a negligibly small probability of detection. For this purpose, the S/Nnet is reestimated, taking into account the randomly selected source position and orientation. We thus assumed simulated events with S/Nnet < 5 are missed by the search pipelines; these events are, however, accounted for in the calculation of sensitive volume and merger rates.
As stated in Sect. 3, the searches process the remaining injections with the same configuration as used for results from O3 data. This is necessary to obtain unbiased rate estimates. In the case of cWB, injections were processed with the most sensitive two-detector configuration; thus, for consistency, we only considered events recovered in the corresponding offline two-detector search results.
5.2. Sensitive volume time and merger Rate
Here, we calculate limits on the merger rate for points in the binary component mass and spin parameter space described in Table 3 using the loudest event method (Biswas et al. 2009; Abbott et al. 2016c).
To derive the upper limit on merger rate for a given point in source parameter space, we consider the sensitive volume-time, ⟨VT⟩sen, of our combined search to such sources at a p-value threshold of 0.251, which is determined by 200214_224526, the most significant event due to noise in the combined search results. For mergers with given intrinsic parameters, the expected number of detected signals N is related to the merger rate R and to the sensitive volume time as ⟨N⟩=R⟨VT⟩sen. For each source point, we estimate ⟨VT⟩sen as a fraction of the total volume time out to its maximum injection redshift zmax by counting injected signals that are detected with a combined p value below the threshold and dividing by the total number of injections generated.
Then, taking a uniform prior on R and using the Poisson probability of zero detected signals as a likelihood, we obtain the 90% credible upper limit R90% = 2.3/⟨VT⟩sen. The only significant IMBH binary signal in the combined search results is GW190521. However, there is only one mass-spin (marked with † in Table 3) point that is consistent with both its component mass and spin χeff − χp90% credible regions. Therefore, for that source point, we conservatively used the Poisson probability of having one IMBH binary detection and thus took R90% = 3.9/⟨VT⟩sen.
Injections with component masses (60 + 60) M⊙ were performed. However, since this parameter point is within the stellar-mass BBH distribution characterized in Abbott et al. (2021b), to which several heavy BBH systems detected in O3a may contribute, we do not quote an upper rate limit. We do, however, state the search sensitivity for such systems in our data release products.
Table 3 summarizes the sensitive volume-time and upper limit on the merger rate for our chosen set of injections. For simulated non-spinning sources, the sensitive volume-time decreases with an increase in total mass but increases with increasing mass ratio q. There are multiple reasons for these trends. First, for a fixed mass ratio, the duration of a signal within the detector bandwidth decreases with increased total mass, even though its overall intrinsic luminosity increases. This is evident if one compares the sensitive volume time obtained for (80 + 40) M⊙, (100 + 50) M⊙, and (133 + 67) M⊙ systems. Second, the amplitude of a source goes down with a decrease in the mass ratio for a fixed total mass. Hence, the sensitivity drops with a decrease in mass ratio. Last, a decrement in mass ratio also increases the contribution coming from subdominant emission multipoles. This significantly affects the GstLAL and PyCBC searches that filter using dominant multipole templates only.
Concerning the dependence on spins, for more positive (negative) values of the effective inspiral spin of a system, when keeping the source frame component masses fixed, the duration of the merger signal within the detector bandwidth increases (decreases) compared to a non-spinning counterpart. Also, for a binary with fixed mass and fixed non-zero effective precession spin, the total emitted GW energy is higher for aligned spins (χeff > 0) compared to anti-aligned spins (χeff < 0). Hence, the sensitivity improves (degrades) for systems with positive (negative) effective inspiral spin (Abbott et al. 2019a; Tiwari et al. 2018). All precessing systems used in this analysis have positive χeff; hence, the combined search can observe them at a greater distance compared to their non-spinning counterpart. Furthermore, current searches are expected to have a higher sensitive volume for systems with aligned spins, and consequently the merger rate limits for such systems would be lower than in the case of anti-aligned spins.
Figure 1 shows this trend visually. The panels show the sensitive volume time for non-precessing and precessing simulated sources, respectively. Each circle corresponds to one class of IMBH binaries in the source frame. The IMBH binaries with aligned and anti-aligned BH spins, χ1, 2, are labeled and shown as displaced circles. In general, we find an increase in the sensitive volume time of the combined search compared to results in Abbott et al. (2019a). This increase is due to an overall increase in the analysis time, detector sensitivity, and the contributing searches’ sensitivity.
Fig. 1. Averaged sensitive time volume ⟨VT⟩sen in Gpc−3 yr−1 for the targeted IMBH binary sources in the m1 − m2 plane. The values are rounded where necessary for visualization. Left panel: is for χp = 0 and right panel: is for χp = 0.42. Each circle corresponds to one class of IMBH binaries in the source frame. The χeff values of injection sets are labeled and shown as displaced circles. |
Figure 2 shows the 90% upper limit on merger rate, R90%, in Gpc−3 yr−1 for the targeted 43 IMBH binary sources in the m1 − m2 plane. As before, the left panel shows the result for non-precessing simulated sources, whereas the right panel shows the same for precessing simulated sources. We set our most stringent upper limit 0.06 Gpc−3 yr−1 for equal-mass IMBH binaries with total mass 200 M⊙ and spin χ1, 2 = 0.8, which is ∼3.5 times more stringent than in the previous study (Abbott et al. 2019a).
Fig. 2. 90% upper limit on merger rate density R90% in Gpc−3 yr−1 for the targeted IMBH binary sources in the m1 − m2 plane. The values are rounded where necessary for display. Left panel: χp = 0 and right panel: for χp = 0.42. Each circle corresponds to one class of IMBH binaries in the source frame. The χeff values of injection sets are labeled and shown as displaced circles. |
5.3. Updated GW190521 merger rate estimate
We reestimated the merger rate of a GW190521-like population. As in Abbott et al. (2016c, 2017b, 2020a,b), we considered a simulated signal to be detected if it is recovered with a FAR of less than 100 yr−1. This corresponds to a combined p-value threshold of 0.009. We considered the maximum observed time (Ta = 0.874 yr) across the three pipelines as the analysis time of the combined search. The population is generated by drawing the intrinsic parameters from the posterior distribution inferred using the NRSur7dq4 waveform model (Abbott et al. 2020b) and then distributing them isotropically over binary orientation parameters and the sky position, and uniformly over comoving volume time up to a max redshift z = 1.5.
The sensitive volume time of the combined search to GW190521-like mergers over the entire O3 data is 14.35 Gpc−3 yr−1. As in Abbott et al. (2020b), we took a Jeffreys prior proportional to R−1/2, where R is the astrophysical merger rate. Given the count of one detection above the threshold, we obtain an estimate of Gpc−3 yr−1, which is more constraining than the estimate given in Abbott et al. (2020b) and consistent with the higher observing time and increased sensitivity of the searches.
6. Discussion and conclusions
The Advanced LIGO and Advanced Virgo detector network concluded their year-long third observing run in March 2020. The first part of the run witnessed a massive black hole merger event (GW190521; Abbott et al. 2020a) with a remnant firmly falling in the IMBH parameter space; the mass was and the components were and , thus a primary BH was inferred to be in the PISN mass gap. In this work, we present the IMBH binary search carried out on the entire range of data of the third observing run. We used three GW search algorithms for this search; the two template-based searches GstLAL and PyCBC, and one model waveform-independent cWB search. The template-based matched filter searches use the dedicated template bank designed for the massive black hole binary coalescences, while the model waveform independent cWB uses a single detection approach for the entire range of BBHs masses including IMBH binaries. We ranked the events based on the combined significance computed combining the three searches. The search shows that GW190521 is still the most significant IMBH binary event. Besides that, we do not find any other significant event in the combined search. We update the significance of already published O3a events with this combined search. We report the discussion on the candidate events, including the marginal events.
Amongst the remaining events, 200114_020818 shows marginal significance in the offline cWB search. As the detector characterization study does not conclusively demonstrate the origin of this event to be terrestrial, we carried out a follow-up investigation on this event in Appendix B. This includes event reconstruction and residual analysis. While the premerger dynamics of 200114_020818 are barely accessible, we analysed 200114_020818 under the quasi-circular BBH hypothesis. Unlike the case of GW190521, here we find strongly inconsistent results across different waveform approximants. However, the model-independent event reconstructions are consistent with each other. Hence, either the event is not consistent with the available quasi-circular binary black hole waveforms or its origin is non-astrophysical in nature.
We update the merger rate density on a suite of simulated signals for IMBH binary systems with generic BH spins using the numerical relativity waveforms provided by SXS, RIT, and GeorgiaTech catalogs. With the year-long O3 data and improved sensitivity of the combined search, we computed merger rate limits using the method of the loudest confirmed noise trigger. The most stringent merger rate, a 90% upper limit, is placed on equal-mass IMBH binaries with a total mass of 200 M⊙ and spin χ1, 2 = 0.8, as 0.056 Gpc−3 yr−1. This is a factor ∼3.5 below the previous limit using O1-O2 data. We further revise the astrophysical merger rate estimate for binary systems comparable to GW190521 to Gpc−3 yr−1, a factor 1.6 more stringent than the earlier result in Abbott et al. (2020b). A semi-analytical estimate of IMBH merger rate density in globular cluster environments under certain assumptions was performed in Rasskazov et al. (2020) and Fragione et al. (2018). In this study, the calculation of the expected S/N considered quadrupolar GW emission at Newtonian order from quasi-circular, non-spinning binaries; however, the inclusion of the merger and ringdown phases of the signal is expected to have a significant impact on the predicted detection rate. The observational rate upper limits obtained here can provide input into such semi-analytical approaches.
We emphasize that the IMBH binaries pose an extreme challenge to interpret. All current GW observations have so far been interpreted within the canonical scenario of an inspiraling quasi-circular BBH. While this is a safe assumption when the (pre-merger) inspiral process is clearly visible in the band, the low frequency of IMBH binary signals makes such a putative inspiral barely visible. This leaves the pre-merger dynamics and even the very nature of the colliding objects open to further interpretation, in turn making some conclusions obtained through canonical analyses less robust (Bustillo et al. 2021a). For instance, alternative scenarios have been proposed that GW190521 is consistent with an eccentric binary merger (Romero-Shaw et al. 2020; Gayathri et al. 2022). Moreover, Bustillo et al. (2021b) showed that GW190521 is consistent with the merger of exotic objects that have dramatically different astrophysical conclusions. The primordial BH scenario is explored in De Luca et al. (2021) and Clesse & Garcia-Bellido (2020). An alternative scenario for forming BHs in the mass gap includes gas accretion onto stellar mass BHs in dense molecular clouds or in primordial dense clusters (Roupas & Kazanas 2019; Safarzadeh & Haiman 2020; Rice & Zhang 2021).
With continuous improvement in the sensitivity of advanced detectors, especially in the low-frequency region, we expect to observe higher numbers of inspiral cycles from massive BBH mergers, and hence better probe inspiral dynamics, including the effects of orbital eccentricity and precession on the signal. In addition, systems with asymmetric component masses may be observed, enabling us to probe higher order multipole emissions. Such gravitational-wave signals including higher order multipoles and precession have a complex morphology: future developments in template-based searches to incorporate such complex dynamics (Harry et al. 2018) can be expected to improve the detectability of IMBH binaries with matched-filter based methods. Similarly to the current interferometer network, we expect future detector noise to include a variety of artifacts, including transients (glitches) at a low frequency that may mimic some aspects of IMBH binary signals. Improved veto methods to distinguish between short-duration noise transients and IMBH signals will thus represent a valuable step toward increasing the detection rate, and they may benefit from novel machine-learning-based classification approaches (Jadhav et al. 2021).
The detection of massive BHs in the GW window has provided crucial observational input for the stellar evolutionary models. While we used the conservative limit of the lower edge of the mass gap in the BH population, it is highly uncertain; it might be as low as ∼40 M⊙ or above ∼70 M⊙, depending on uncertainties surrounding the nuclear reaction rates (e.g., Farmer et al. 2019, 2020; Costa et al. 2021), the collapse of the residual stellar envelope (e.g., Mapelli et al. 2020; Costa et al. 2021), the impact of stellar rotation (e.g., Mapelli et al. 2020; Marchant & Moriya 2020; Woosley & Heger 2021), the result of stellar mergers (e.g., Di Carlo et al. 2019, 2020a; Renzo et al. 2020a), the efficiency of accretion from companion stars (e.g., van Son et al. 2020), the model of convection (e.g., Renzo et al. 2020b; Farrell et al. 2021; Tanikawa et al. 2021), and the onset of dredge-up episodes (e.g., Costa et al. 2021; Tanikawa et al. 2021; Umeda et al. 2020). Recently, Costa et al. (2021) proposed that the mass gap might even disappear if a low rate for the 12C(α, γ)16O reaction is assumed and if a mild envelope under-shooting is included in stellar evolution calculations. Theoretical and numerical models show that black holes with mass in the pair instability gap could be the result of hierarchical mergers of smaller black holes (Miller & Hamilton 2002; Antonini & Rasio 2016; Gerosa & Berti 2017; Fishbach et al. 2017; Rodriguez et al. 2019; Kimball et al. 2021; Doctor et al. 2020) or the outcome of stellar collisions in dense star clusters (Di Carlo et al. 2019, 2020b; Kremer et al. 2020; Renzo et al. 2020a). Hierarchical mergers appear to be particularly efficient in nuclear star clusters (Antonini et al. 2019; Fragione et al. 2020; Baibhav et al. 2020; Mapelli et al. 2021) and in the dense gaseous disks of active galactic nuclei (McKernan et al. 2012, 2018; Bartos et al. 2017; Yang et al. 2019; Tagawa et al. 2020, 2021). The detection of more massive BH binaries in the advanced detector era will provide constraints on all the formation channels. In addition, future observations of IMBH binaries across the GW spectrum (Sathyaprakash & Schutz 2009) could strengthen the possible evolutionary link between stellar-mass BHs and supermassive BHs in the galactic centers in coming decades (Mezcua 2017; Koliopanos 2017; King & Dehnen 2005).
The parameter estimated values for GW190521 reported in GWTC-2 are slightly different from that of the detection paper. The estimation in the detection paper is based on the NRSur7dq4 waveform model, and GWTC-2 values are obtained from the estimates averaged over three waveforms; SEOBNRv4PHM, NRSur7dq4, and IMRPhenomPv3HM, respectively.
Acknowledgments
This material is based upon work supported by NSF’s LIGO Laboratory which is a major facility fully funded by the National Science Foundation. The authors also gratefully acknowledge the support of the Science and Technology Facilities Council (STFC) of the United Kingdom, the Max-Planck-Society (MPS), and the State of Niedersachsen/Germany for support of the construction of Advanced LIGO and construction and operation of the GEO600 detector. Additional support for Advanced LIGO was provided by the Australian Research Council. The authors gratefully acknowledge the Italian Istituto Nazionale di Fisica Nucleare (INFN), the French Centre National de la Recherche Scientifique (CNRS) and the Netherlands Organization for Scientific Research, for the construction and operation of the Virgo detector and the creation and support of the EGO consortium. The authors also gratefully acknowledge research support from these agencies as well as by the Council of Scientific and Industrial Research of India, the Department of Science and Technology, India, the Science & Engineering Research Board (SERB), India, the Ministry of Human Resource Development, India, the Spanish Agencia Estatal de Investigación, the Vicepresidència i Conselleria d’Innovació, Recerca i Turisme and the Conselleria d’Educació i Universitat del Govern de les Illes Balears, the Conselleria d’Innovació, Universitats, Ciència i Societat Digital de la Generalitat Valenciana and the CERCA Programme Generalitat de Catalunya, Spain, the National Science Centre of Poland and the Foundation for Polish Science (FNP), the Swiss National Science Foundation (SNSF), the Russian Foundation for Basic Research, the Russian Science Foundation, the European Commission, the European Regional Development Funds (ERDF), the Royal Society, the Scottish Funding Council, the Scottish Universities Physics Alliance, the Hungarian Scientific Research Fund (OTKA), the French Lyon Institute of Origins (LIO), the Belgian Fonds de la Recherche Scientifique (FRS-FNRS), Actions de Recherche Concertées (ARC) and Fonds Wetenschappelijk Onderzoek – Vlaanderen (FWO), Belgium, the Paris Île-de-France Region, the National Research, Development and Innovation Office Hungary (NKFIH), the National Research Foundation of Korea, the Natural Science and Engineering Research Council Canada, Canadian Foundation for Innovation (CFI), the Brazilian Ministry of Science, Technology, and Innovations, the International Center for Theoretical Physics South American Institute for Fundamental Research (ICTP-SAIFR), the Research Grants Council of Hong Kong, the National Natural Science Foundation of China (NSFC), the Leverhulme Trust, the Research Corporation, the Ministry of Science and Technology (MOST), Taiwan, the United States Department of Energy, and the Kavli Foundation. The authors gratefully acknowledge the support of the NSF, STFC, INFN and CNRS for provision of computational resources. This work was supported by MEXT, JSPS Leading-edge Research Infrastructure Program, JSPS Grant-in-Aid for Specially Promoted Research 26000005, JSPS Grant-in-Aid for Scientific Research on Innovative Areas 2905: JP17H06358, JP17H06361 and JP17H06364, JSPS Core-to-Core Program A. Advanced Research Networks, JSPS Grant-in-Aid for Scientific Research (S) 17H06133, the joint research program of the Institute for Cosmic Ray Research, University of Tokyo, National Research Foundation (NRF) and Computing Infrastructure Project of KISTI-GSDC in Korea, Academia Sinica (AS), AS Grid Center (ASGC) and the Ministry of Science and Technology (MoST) in Taiwan under grants including AS-CDA-105-M06, Advanced Technology Center (ATC) of NAOJ, and Mechanical Engineering Center of KEK. We would like to thank all of the essential workers who put their health at risk during the COVID-19 pandemic, without whom we would not have been able to complete this work.
References
- Aasi, J., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2014, Phys. Rev. D, 89, 102006 [NASA ADS] [CrossRef] [Google Scholar]
- Aasi, J., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2015, Class. Quant. Grav., 32, 074001 [Google Scholar]
- Abadie, J., Abbott, B. P., Abbott, R., et al. 2012, Phys. Rev. D, 85, 102004 [NASA ADS] [CrossRef] [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2016a, Phys. Rev. Lett., 116, 221101 [Erratum: Phys. Rev. Lett. 121, 129902 (2018)] [NASA ADS] [CrossRef] [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2016b, Class. Quant. Grav., 33, 134001 [NASA ADS] [CrossRef] [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2016c, ApJ, 832, L21 [CrossRef] [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2017a, Phys. Rev. D, 96, 022001 [NASA ADS] [CrossRef] [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2017b, Phys. Rev. Lett., 119, 161101 [CrossRef] [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2019a, Phys. Rev. D, 100, 064064 [NASA ADS] [CrossRef] [Google Scholar]
- Abbott, R., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2019b, GCN, 25330, 1 [NASA ADS] [Google Scholar]
- Abbott, R., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2020a, Phys. Rev. Lett., 125, 101102 [Google Scholar]
- Abbott, R., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2020b, ApJ, 900, L13 [NASA ADS] [CrossRef] [Google Scholar]
- Abbott, B. P., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2020c, Class. Quant. Grav., 37, 055002 [NASA ADS] [CrossRef] [Google Scholar]
- Abbott, R., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2021a, Phys. Rev. X, 11, 021053 [Google Scholar]
- Abbott, R., et al. (LIGO Scientific Collaboration, Virgo Collaboration) 2021b, ApJ, 913, L7 [NASA ADS] [CrossRef] [Google Scholar]
- Acernese, F., Agathos, M., Agatsuma, K., et al. 2015, Class. Quant. Grav., 32, 024001 [Google Scholar]
- Acernese, F., Adams, T., Agatsuma, K., et al. 2018, Class. Quant. Grav., 35, 205004 [NASA ADS] [CrossRef] [Google Scholar]
- Ajith, P., Hannam, M., Husa, S., et al. 2011, Phys. Rev. Lett., 106, 241101 [NASA ADS] [CrossRef] [Google Scholar]
- Allen, B. 2005, Phys. Rev. D, 71, 062001 [NASA ADS] [CrossRef] [Google Scholar]
- Amaro-Seoane, P., Audley, H., Babak, S., et al. 2017, ArXiv e-prints [arXiv:1702.00786] [Google Scholar]
- Anderson, J., & van der Marel, R. P. 2010, ApJ, 710, 1032 [NASA ADS] [CrossRef] [Google Scholar]
- Antonini, F., & Rasio, F. A. 2016, ApJ, 831, 187 [Google Scholar]
- Antonini, F., Gieles, M., & Gualandris, A. 2019, MNRAS, 486, 5008 [NASA ADS] [CrossRef] [Google Scholar]
- Atakan Gurkan, M., Freitag, M., & Rasio, F. A. 2004, ApJ, 604, 632 [NASA ADS] [CrossRef] [Google Scholar]
- Babak, S., Taracchini, A., & Buonanno, A. 2017, Phys. Rev. D, 95, 024010 [NASA ADS] [CrossRef] [Google Scholar]
- Bachetti, M., Harrison, F. A., Walton, D. J., et al. 2014, Nature, 514, 202 [NASA ADS] [CrossRef] [Google Scholar]
- Baibhav, V., Gerosa, D., Berti, E., et al. 2020, Phys. Rev. D, 102, 043002 [NASA ADS] [CrossRef] [Google Scholar]
- Barkat, Z., Rakavy, G., & Sack, N. 1967, Phys. Rev. Lett., 18, 379 [Google Scholar]
- Bartos, I., Kocsis, B., Haiman, Z., & M’arka, S. 2017, ApJ, 835, 165 [CrossRef] [Google Scholar]
- Baumgardt, H., Hut, P., Makino, J., McMillan, S., & Portegies Zwart, S. F. 2003, ApJ, 582, L21 [NASA ADS] [CrossRef] [Google Scholar]
- Belczynski, K., Hirschi, R., Kaiser, E. A., et al. 2020, ApJ, 890, 113 [NASA ADS] [CrossRef] [Google Scholar]
- Biswas, R., Brady, P. R., Creighton, J. D. E., & Fairhurst, S. 2009, Class. Quant. Grav., 26, 175009 [Erratum: Class. Quant. Grav. 30, 079502 (2013)] [NASA ADS] [CrossRef] [Google Scholar]
- Bohé, A., Shao, L., Taracchini, A., et al. 2017, Phys. Rev. D, 95, 044028 [CrossRef] [Google Scholar]
- Bond, J. R., Arnett, W. D., & Carr, B. J. 1984, ApJ, 280, 825 [NASA ADS] [CrossRef] [Google Scholar]
- Buikema, A., Cahillane, C., Mansell, G. L., et al. 2020, Phys. Rev. D, 102, 062003 [NASA ADS] [CrossRef] [Google Scholar]
- Bustillo, J. C., Sanchis-Gual, N., Torres-Forné, A., & Font, J. A. 2021a, Phys. Rev. Lett., 126, 201101 [CrossRef] [Google Scholar]
- Bustillo, J. C., Sanchis-Gual, N., Torres-Forné, A., et al. 2021b, Phys. Rev. Lett., 126, 081101 [CrossRef] [Google Scholar]
- Cabero, M., Lundgren, A., Nitz, A. H., et al. 2019, Class. Quant. Grav., 36, 155010 [NASA ADS] [CrossRef] [Google Scholar]
- Calderón Bustillo, J., Husa, S., Sintes, A. M., & Pürrer, M. 2016, Phys. Rev. D, 93, 084019 [CrossRef] [Google Scholar]
- Calderón Bustillo, J., Clark, J. A., Laguna, P., & Shoemaker, D. 2018, Phys. Rev. Lett., 121, 191102 [CrossRef] [Google Scholar]
- Calderon Bustillo, J., Evans, C., Clark, J. A., et al. 2020, Commun. Phys., 3, 176 [NASA ADS] [CrossRef] [Google Scholar]
- Campanelli, M., Lousto, C. O., Zlochower, Y., & Merritt, D. 2007, Phys. Rev. Lett., 98, 231102 [Google Scholar]
- Cannon, K., Caudill, S., Chan, C., et al. 2021, SoftwareX, 14, 100680 [NASA ADS] [CrossRef] [Google Scholar]
- Carullo, G., van der Schaaf, L., London, L., et al. 2018, Phys. Rev. D, 98, 104020 [NASA ADS] [CrossRef] [Google Scholar]
- Chandra, K., Gayathri, V., Bustillo, J. C., & Pai, A. 2020, Phys. Rev. D, 102, 044035 [CrossRef] [Google Scholar]
- Chandra, K., Villa-Ortega, V., Dent, T., et al. 2021, Phys. Rev. D, 104, 042004 [NASA ADS] [CrossRef] [Google Scholar]
- Clesse, S., & Garcia-Bellido, J. 2020, ArXiv e-prints [arXiv:2007.06481] [Google Scholar]
- Cornish, N., & Littenberg, T. 2015, Class. Quant. Grav., 32, 135012 [NASA ADS] [CrossRef] [Google Scholar]
- Cornish, N. J., Littenberg, T. B., Bécsy, B., et al. 2021, Phys. Rev. D, 103, 044006 [CrossRef] [Google Scholar]
- Costa, G., Bressan, A., Mapelli, M., et al. 2021, MNRAS, 501, 4514 [NASA ADS] [CrossRef] [Google Scholar]
- Cseh, D., Webb, N. A., Godet, O., et al. 2015, MNRAS, 446, 3268 [NASA ADS] [CrossRef] [Google Scholar]
- Dal Canton, T., Nitz, A. H., Lundgren, A. P., et al. 2014, Phys. Rev. D, 90, 082004 [NASA ADS] [CrossRef] [Google Scholar]
- Dal Canton, T., Nitz, A. H., Gadre, B., et al. 2021, ApJ, 923, 254 [CrossRef] [Google Scholar]
- Davies, G. S., Dent, T., Tápai, M., et al. 2020, Phys. Rev. D, 102, 022004 [CrossRef] [Google Scholar]
- Davis, D., Massinger, T. J., Lundgren, A. P., et al. 2019, Class. Quant. Grav., 36, 055011 [NASA ADS] [CrossRef] [Google Scholar]
- Davis, D., Areeda, J. S., Berger, B. K., et al. 2021, Class. Quant. Grav., 38, 135014 [NASA ADS] [CrossRef] [Google Scholar]
- De Luca, V., Desjacques, V., Franciolini, G., Pani, P., & Riotto, A. 2021, Phys. Rev. Lett., 126, 051101 [CrossRef] [Google Scholar]
- Di Carlo, U. N., Giacobbo, N., Mapelli, M., et al. 2019, MNRAS, 487, 2947 [NASA ADS] [CrossRef] [Google Scholar]
- Di Carlo, U. N., Mapelli, M., Giacobbo, N., et al. 2020a, MNRAS, 498, 495 [NASA ADS] [CrossRef] [Google Scholar]
- Di Carlo, U. N., Mapelli, M., Bouffanais, Y., et al. 2020b, MNRAS, 497, 1043 [NASA ADS] [CrossRef] [Google Scholar]
- Doctor, Z., Wysocki, D., O’Shaughnessy, R., Holz, D. E., & Farr, B. 2020, ApJ, 893, 35 [NASA ADS] [CrossRef] [Google Scholar]
- Drago, M., Gayathri, V., Klimenko, S., et al. 2020, ArXiv e-prints [arXiv:2006.12604] [Google Scholar]
- Driggers, J. C., Vitale, S., Lundgren, A. P., et al. 2019, Phys. Rev. D, 99, 042001 [NASA ADS] [CrossRef] [Google Scholar]
- Ebisuzaki, T., Makino, J., Tsuru, T. G., et al. 2001, ApJ, 562, L19 [NASA ADS] [CrossRef] [Google Scholar]
- Essick, R., Godwin, P., Hanna, C., Blackburn, L., & Katsavounidis, E. 2020, ArXiv e-prints [arXiv:2005.12761] [Google Scholar]
- Estevez, D., Lagabbe, P., Masserot, A., et al. 2021, Class. Quantum Grav., 38, 075007 [NASA ADS] [CrossRef] [Google Scholar]
- Ezquiaga, J. M., & Holz, D. E. 2021, ApJ, 909, L23 [CrossRef] [Google Scholar]
- Farmer, R., Renzo, M., de Mink, S. E., Marchant, P., & Justham, S. 2019, ApJ, 887, 53 [Google Scholar]
- Farmer, R., Renzo, M., de Mink, S. E., Fishbach, M., & Justham, S. 2020, ApJ, 902, L36 [CrossRef] [Google Scholar]
- Farrell, S., Webb, N., Barret, D., Godet, O., & Rodrigues, J. 2009, Nature, 460, 73 [CrossRef] [Google Scholar]
- Farrell, E., Groh, J. H., Hirschi, R., et al. 2021, MNRAS, 502, L40 [NASA ADS] [CrossRef] [Google Scholar]
- Ferrarese, L., & Ford, H. 2005, Space Sci. Rev., 116, 523 [Google Scholar]
- Fishbach, M., & Holz, D. E. 2020, ApJ, 904, L26 [Google Scholar]
- Fishbach, M., Holz, D. E., & Farr, B. 2017, ApJ, 840, L24 [NASA ADS] [CrossRef] [Google Scholar]
- Fisher, R. P., Hemming, G., Bizouard, M. A., et al. 2020, ArXiv e-prints [arXiv:2008.11316] [Google Scholar]
- Fowler, W. A., & Hoyle, F. 1964, ApJS, 9, 201 [Google Scholar]
- Fragione, G., Ginsburg, I., & Kocsis, B. 2018, ApJ, 856, 92 [NASA ADS] [CrossRef] [Google Scholar]
- Fragione, G., Loeb, A., & Rasio, F. A. 2020, ApJ, 902, L26 [NASA ADS] [CrossRef] [Google Scholar]
- Fregeau, J. M., Larson, S. L., Miller, M. C., O’Shaughnessy, R., & Rasio, F. A. 2006, ApJ, 646, L135 [NASA ADS] [CrossRef] [Google Scholar]
- Fryer, C. L., Woosley, S. E., & Heger, A. 2001, ApJ, 550, 372 [NASA ADS] [CrossRef] [Google Scholar]
- Gair, J. R., Vallisneri, M., Larson, S. L., & Baker, J. G. 2013, Liv. Rev. Relativ., 16, 7 [CrossRef] [Google Scholar]
- Gayathri, V., Bacon, P., Pai, A., et al. 2019, Phys. Rev. D, 100, 124022 [NASA ADS] [CrossRef] [Google Scholar]
- Gayathri, V., Healy, J., Lange, J., et al. 2022, Nat. Astron., in press [arXiv:2009.05461] [Google Scholar]
- Gebhardt, K., Rich, R. M., & Ho, L. C. 2005, ApJ, 634, 1093 [NASA ADS] [CrossRef] [Google Scholar]
- Gerosa, D., & Berti, E. 2017, Phys. Rev. D, 95, 124046 [NASA ADS] [CrossRef] [Google Scholar]
- Ghonge, S., Chatziioannou, K., Clark, J. A., et al. 2020, Phys. Rev. D, 102, 064056 [NASA ADS] [CrossRef] [Google Scholar]
- Giacobbo, N., Mapelli, M., & Spera, M. 2018, MNRAS, 474, 2959 [Google Scholar]
- Giersz, M., Leigh, N., Hypki, A., Lützgendorf, N., & Askar, A. 2015, MNRAS, 454, 3150 [Google Scholar]
- Godet, O., Barret, D., Webb, N. A., Farrell, S. A., & Gehrels, N. 2009, ApJ, 705, L109 [Google Scholar]
- Godwin, P., Essick, R., Hanna, C., et al. 2020, ArXiv e-prints [arXiv:2010.15282] [Google Scholar]
- Gonzalez, J. A., Sperhake, U., Bruegmann, B., Hannam, M., & Husa, S. 2007, Phys. Rev. Lett., 98, 091101 [NASA ADS] [CrossRef] [Google Scholar]
- Graham, A. W. 2012, ApJ, 746, 113 [NASA ADS] [CrossRef] [Google Scholar]
- Graham, A. W., & Scott, N. 2013, ApJ, 764, 151 [Google Scholar]
- Greene, J. E., Strader, J., & Ho, L. C. 2020, ARA&A, 58, 257 [Google Scholar]
- Hanna, C., Caudill, S., Messick, C., et al. 2020, Phys. Rev. D, 101, 022003 [NASA ADS] [CrossRef] [Google Scholar]
- Hannam, M., Schmidt, P., Bohé, A., et al. 2014, Phys. Rev. Lett., 113, 151101 [Google Scholar]
- Harry, I., Bustillo, J. C., & Nitz, A. 2018, Phys. Rev. D, 97, 023004 [NASA ADS] [CrossRef] [Google Scholar]
- Healy, J., Lousto, C. O., Zlochower, Y., & Campanelli, M. 2017, Class. Quant. Grav., 34, 224001 [NASA ADS] [CrossRef] [Google Scholar]
- Heger, A., & Woosley, S. E. 2002, ApJ, 567, 532 [Google Scholar]
- Heger, A., Fryer, C. L., Woosley, S. E., Langer, N., & Hartmann, D. H. 2003, ApJ, 591, 288 [CrossRef] [Google Scholar]
- Inayoshi, K., Visbal, E., & Haiman, Z. 2020, ARA&A, 58, 27 [NASA ADS] [CrossRef] [Google Scholar]
- Israel, G. L., Belfiore, A., Stella, L., et al. 2017, Science, 355, 817 [NASA ADS] [CrossRef] [Google Scholar]
- Jadhav, S., Mukund, N., Gadre, B., Mitra, S., & Abraham, S. 2021, Phys. Rev. D, 104, 064051 [CrossRef] [Google Scholar]
- Jani, K., Healy, J., Clark, J. A., et al. 2016, Class. Quant. Grav., 33, 204001 [NASA ADS] [CrossRef] [Google Scholar]
- Jani, K., Shoemaker, D., & Cutler, C. 2020, Nat. Astron., 4, 260 [NASA ADS] [CrossRef] [Google Scholar]
- Kaaret, P., Prestwich, A. H., Zezas, A., et al. 2001, MNRAS, 321, L29 [Google Scholar]
- Kaaret, P., Feng, H., & Roberts, T. P. 2017, ARA&A, 55, 303 [Google Scholar]
- Kamaretsos, I., Hannam, M., Husa, S., & Sathyaprakash, B. S. 2012, Phys. Rev. D, 85, 024018 [NASA ADS] [CrossRef] [Google Scholar]
- Kimball, C., Talbot, C., Berry, C. P. L., et al. 2021, ApJ, 915, L35 [CrossRef] [Google Scholar]
- King, A. R., & Dehnen, W. 2005, MNRAS, 357, 275 [Google Scholar]
- Kızıltan, B., Baumgardt, H., & Loeb, A. 2017, Nature, 542, 203 [Google Scholar]
- Klimenko, S., & Mitselmakher, G. 2004, Class. Quant. Grav., 21, S1819 [NASA ADS] [CrossRef] [Google Scholar]
- Klimenko, S., Mohanty, S., Rakhmanov, M., & Mitselmakher, G. 2005, Phys. Rev. D, 72, 122002 [NASA ADS] [CrossRef] [Google Scholar]
- Klimenko, S., Mohanty, S., Rakhmanov, M., & Mitselmakher, G. 2006, J. Phys. Conf. Ser., 32, 12 [CrossRef] [Google Scholar]
- Klimenko, S., Vedovato, G., Drago, M., et al. 2011, Phys. Rev. D, 83, 102001 [NASA ADS] [CrossRef] [Google Scholar]
- Klimenko, S., Vedovato, G., Drago, M., et al. 2016, Phys. Rev. D, 93, 042004 [NASA ADS] [CrossRef] [Google Scholar]
- Koliopanos, F. 2017, Proceedings of the XII Multifrequency Behaviour of High Energy Cosmic Sources Workshop, 12–17 June, 51 [Google Scholar]
- Kormendy, J., & Ho, L. C. 2013, ARA&A, 51, 511 [Google Scholar]
- Kremer, K., Spera, M., Becker, D., et al. 2020, ApJ, 903, 45 [NASA ADS] [CrossRef] [Google Scholar]
- Lange, J., O’Shaughnessy, R., Boyle, M., et al. 2017, Phys. Rev. D, 96, 104041 [NASA ADS] [CrossRef] [Google Scholar]
- Lanzoni, B., Mucciarelli, A., Origlia, L., et al. 2013, ApJ, 769, 107 [NASA ADS] [CrossRef] [Google Scholar]
- Lin, D., Strader, J., Carrasco, E. R., et al. 2018, Nat. Astron., 2, 656 [Google Scholar]
- Littenberg, T. B., Kanner, J. B., Cornish, N. J., & Millhouse, M. 2016, Phys. Rev. D, 94, 044050 [NASA ADS] [CrossRef] [Google Scholar]
- Lützgendorf, N., Kissler-Patig, M., Noyola, E., et al. 2011, A&A, 533, A36 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Madau, P., & Rees, M. J. 2001, ApJ, 551, L27 [Google Scholar]
- Mapelli, M. 2016, MNRAS, 459, 3432 [NASA ADS] [CrossRef] [Google Scholar]
- Mapelli, M., Spera, M., Montanari, E., et al. 2020, ApJ, 888, 76 [NASA ADS] [CrossRef] [Google Scholar]
- Mapelli, M., Dall’Amico, M., Bouffanais, Y., et al. 2021, MNRAS, 505, 339 [CrossRef] [Google Scholar]
- Marchant, P., & Moriya, T. J. 2020, A&A, 640, L18 [EDP Sciences] [Google Scholar]
- McKernan, B., Ford, K. E. S., Lyra, W., & Perets, H. B. 2012, MNRAS, 425, 460 [NASA ADS] [CrossRef] [Google Scholar]
- McKernan, B., Ford, K. E. S., Bellovary, J., et al. 2018, ApJ, 866, 66 [Google Scholar]
- Meidam, J., Agathos, M., Van Den Broeck, C., Veitch, J., & Sathyaprakash, B. S. 2014, Phys. Rev. D, 90, 064009 [NASA ADS] [CrossRef] [Google Scholar]
- Messick, C., Blackburn, K., Brady, P., et al. 2017, Phys. Rev. D, 95, 042001 [NASA ADS] [CrossRef] [Google Scholar]
- Mezcua, M. 2017, Int. J. Mod. Phys. D, 26, 1730021 [Google Scholar]
- Miller, M. C. 2009, Class. Quant. Grav., 26, 094031 [NASA ADS] [CrossRef] [Google Scholar]
- Miller, M. C., & Colbert, E. J. M. 2004, Int. J. Mod. Phys. D, 13, 1 [NASA ADS] [CrossRef] [Google Scholar]
- Miller, M. C., & Hamilton, D. P. 2002, MNRAS, 330, 232 [Google Scholar]
- Mozzon, S., Nuttall, L. K., Lundgren, A., et al. 2020, Class. Quant. Grav., 37, 215014 [NASA ADS] [CrossRef] [Google Scholar]
- Mroué, A. H., Scheel, M. A., Szilágyi, B., et al. 2013, Phys. Rev. Lett., 111, 241104 [CrossRef] [Google Scholar]
- Necula, V., Klimenko, S., & Mitselmakher, G. 2012, J. Phys. Conf. Ser., 363, 012032 [CrossRef] [Google Scholar]
- Nguyen, P., Schofield, R. M. S., Effler, A., et al. 2021, Class. Quant. Grav., 38, 145001 [NASA ADS] [CrossRef] [Google Scholar]
- Nitz, A. H. 2018, Class. Quant. Grav., 35, 035016 [NASA ADS] [CrossRef] [Google Scholar]
- Nitz, A. H., & Capano, C. D. 2021, ApJ, 907, L9 [NASA ADS] [CrossRef] [Google Scholar]
- Nitz, A. H., Dent, T., Dal Canton, T., Fairhurst, S., & Brown, D. A. 2017, ApJ, 849, 118 [CrossRef] [Google Scholar]
- Nitz, A. H., Dent, T., Davies, G. S., et al. 2019, ApJ, 891, 123 [Google Scholar]
- Noyola, E., Gebhardt, K., & Bergmann, M. 2008, ApJ, 676, 1008 [NASA ADS] [CrossRef] [Google Scholar]
- O’Leary, R. M., Rasio, F. A., Fregeau, J. M., Ivanova, N., & O’Shaughnessy, R. 2006, ApJ, 637, 937 [CrossRef] [Google Scholar]
- Ossokine, S., Buonanno, A., Marsat, S., et al. 2020, Phys. Rev. D, 102, 044055 [NASA ADS] [CrossRef] [Google Scholar]
- Owen, B. J. 1996, Phys. Rev. D, 53, 6749 [NASA ADS] [CrossRef] [Google Scholar]
- Pankow, C., Chatziioannou, K., Chase, E. A., et al. 2018, Phys. Rev. D, 98, 084016 [NASA ADS] [CrossRef] [Google Scholar]
- Pasham, D. R., Strohmayer, T. E., & Mushotzky, R. F. 2015, Nature, 513, 74 [Google Scholar]
- Paynter, J., Webster, R., & Thrane, E. 2021, Nat. Astron., 5, 560 [NASA ADS] [CrossRef] [Google Scholar]
- Planck Collaboration XIII. 2016, A&A, 594, A13 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Pratten, G., García-Quirós, C., Colleoni, M., et al. 2021, Phys. Rev. D, 103, 104056 [NASA ADS] [CrossRef] [Google Scholar]
- Pürrer, M. 2016, Phys. Rev. D, 93, 064041 [CrossRef] [Google Scholar]
- Rasskazov, A., Fragione, G., & Kocsis, B. 2020, ApJ, 899, 149 [NASA ADS] [CrossRef] [Google Scholar]
- Remillard, R. A., & McClintock, J. E. 2006, ARA&A, 44, 49 [Google Scholar]
- Renzo, M., Cantiello, M., Metzger, B. D., & Jiang, Y. F. 2020a, ApJ, 904, L13 [NASA ADS] [CrossRef] [Google Scholar]
- Renzo, M., Farmer, R. J., Justham, S., et al. 2020b, MNRAS, 493, 4333 [NASA ADS] [CrossRef] [Google Scholar]
- Rice, J. R., & Zhang, B. 2021, ApJ, 908, 59 [NASA ADS] [CrossRef] [Google Scholar]
- Rodriguez, C. L., Zevin, M., Amaro-Seoane, P., et al. 2019, Phys. Rev. D, 100, 043027 [Google Scholar]
- Romero-Shaw, I., Lasky, P. D., Thrane, E., & Calder’on Bustillo, J. 2020, ApJ, 903, L5 [NASA ADS] [CrossRef] [Google Scholar]
- Roupas, Z., & Kazanas, D. 2019, A&A, 632, L8 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Sachdev, S., Caudill, S., Fong, H., et al. 2019, Phys. Rev. D, submitted [arXiv:1901.08580] [Google Scholar]
- Safarzadeh, M., & Haiman, Z. 2020, ApJ, 903, L21 [NASA ADS] [CrossRef] [Google Scholar]
- Salemi, F., Milotti, E., Prodi, G. A., et al. 2019, Phys. Rev. D, 100, 042003 [NASA ADS] [CrossRef] [Google Scholar]
- Sathyaprakash, B. S., & Dhurandhar, S. V. 1991, Phys. Rev. D, 44, 3819 [CrossRef] [Google Scholar]
- Sathyaprakash, B. S., & Schutz, B. F. 2009, Liv. Rev. Relativ., 12, 2 [CrossRef] [Google Scholar]
- Schmidt, P., Ohme, F., & Hannam, M. 2015, Phys. Rev. D, 91, 024043 [NASA ADS] [CrossRef] [Google Scholar]
- Servillat, M., Farrell, S. A., Lin, D., et al. 2011, ApJ, 743, 6 [Google Scholar]
- Soni, S., Berry, C. P. L., Coughlin, S. B., et al. 2021, Class. Quant. Grav., 38, 19 [Google Scholar]
- Soria, R., Hakala, P. J., Hau, G. K. T., Gladstone, J. C., & Kong, A. K. H. 2012, MNRAS, 420, 3599 [NASA ADS] [CrossRef] [Google Scholar]
- Spera, M., & Mapelli, M. 2017, MNRAS, 470, 4739 [Google Scholar]
- Stevenson, S., Sampson, M., Powell, J., et al. 2019, ApJ, 882, 121 [Google Scholar]
- Sun, L., Goetz, E., Kissel, J. S., et al. 2020, Class. Quant. Grav., 37, 225008 [NASA ADS] [CrossRef] [Google Scholar]
- Szczepańczyk, M., Klimenko, S., O’Brien, B., et al. 2021, Phys. Rev. D, 103, 082002 [CrossRef] [Google Scholar]
- Tagawa, H., Haiman, Z., Bartos, I., & Kocsis, B. 2020, ApJ, 899, 26 [Google Scholar]
- Tagawa, H., Kocsis, B., Haiman, Z., et al. 2021, ApJ, 908, 194 [NASA ADS] [CrossRef] [Google Scholar]
- Tanikawa, A., Kinugawa, T., Yoshida, T., Hijikawa, K., & Umeda, H. 2021, MNRAS, 505, 2170 [CrossRef] [Google Scholar]
- Thrane, E., Lasky, P. D., & Levin, Y. 2017, Phys. Rev. D, 96, 102004 [NASA ADS] [CrossRef] [Google Scholar]
- Tiwari, V., Fairhurst, S., & Hannam, M. 2018, ApJ, 868, 140 [NASA ADS] [CrossRef] [Google Scholar]
- Umeda, H., Yoshida, T., Nagele, C., & Takahashi, K. 2020, ApJ, 905, L21 [NASA ADS] [CrossRef] [Google Scholar]
- Usman, S. A., Nitz, A. H., Harry, I. W., et al. 2016, Class. Quant. Grav., 33, 215004 [NASA ADS] [CrossRef] [Google Scholar]
- Vajente, G., Huang, Y., Isi, M., et al. 2020, Phys. Rev. D, 101, 042003 [NASA ADS] [CrossRef] [Google Scholar]
- van den Bosch, R., de Zeeuw, T., Gebhardt, K., Noyola, E., & van de Ven, G. 2006, ApJ, 641, 852 [Google Scholar]
- van der Marel, R. P. 2004, Coevolution of Black Holes and Galaxies (Cambridge: Cambridge University Press), 37 [Google Scholar]
- van Son, L. A. C., De Mink, S. E., Broekgaarden, F. S., et al. 2020, ApJ, 897, 100 [Google Scholar]
- Varma, V., Field, S. E., Scheel, M. A., et al. 2019, Phys. Rev. Res., 1, 033015 [NASA ADS] [CrossRef] [Google Scholar]
- Veitch, J., & Vecchio, A. 2008, Phys. Rev. D, 78, 022001 [NASA ADS] [CrossRef] [Google Scholar]
- Veitch, J., Raymond, V., Farr, B., et al. 2015, Phys. Rev. D, 91, 042003 [NASA ADS] [CrossRef] [Google Scholar]
- Viets, A., Wade, M., Urban, A. L., et al. 2018, Class. Quant. Grav., 35, 095015 [NASA ADS] [CrossRef] [Google Scholar]
- Vink, J. S., Higgins, E. R., Sander, A. A. C., & Sabhahit, G. N. 2021, MNRAS, 504, 146 [NASA ADS] [CrossRef] [Google Scholar]
- Vitral, E., & Mamon, G. A. 2021, A&A, 646, A63 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
- Volonteri, M. 2010, A&ARv, 18, 279 [Google Scholar]
- Webb, N., Cseh, D., Lenc, E., et al. 2012, Science, 337, 554 [NASA ADS] [CrossRef] [Google Scholar]
- Woosley, S. E. 2017, ApJ, 836, 244 [Google Scholar]
- Woosley, S. E. 2019, ApJ, 878, 49 [Google Scholar]
- Woosley, S. E., & Heger, A. 2021, ApJ, 912, L31 [NASA ADS] [CrossRef] [Google Scholar]
- Woosley, S. E., Blinnikov, S., & Heger, A. 2007, Nature, 450, 390 [NASA ADS] [CrossRef] [PubMed] [Google Scholar]
- Wysocki, D., O’Shaughnessy, R., Lange, J., & Fang, Y.-L. L. 2019, Phys. Rev. D, 99, 084026 [NASA ADS] [CrossRef] [Google Scholar]
- Yang, Y., Bartos, I., Gayathri, V., et al. 2019, Phys. Rev. Lett., 123, 181101 [CrossRef] [Google Scholar]
- Yunes, N., & Siemens, X. 2013, Liv. Rev. Relativ., 16, 9 [CrossRef] [Google Scholar]
- Yunes, N., Yagi, K., & Pretorius, F. 2016, Phys. Rev. D, 94, 084002 [NASA ADS] [CrossRef] [Google Scholar]
Appendix A: Other marginal candidate events
Table A.1 summarizes the other marginal candidates identified by our combined IMBH search in O3 data that are not reported elsewhere in a catalog of compact binary coalescence events and that satisfy the criteria described in Sect. 3.4. These triggers were reported by at least one of the contributing searches with a FAR below 0.5 yr−1 and have a p̄ ≤ 0.7.
Other marginal candidate event list. We find three candidate events that passed a FAR threshold of 0.5 yr−1 in at least one of the three dedicated searches and additionally have a combined p-value less than 0.7.
The event 191225_215715 was first reported by PyCBC Live (Dal Canton et al. 2021), a low-latency matched-filter search with a FAR of 0.4 yr−1. When the contributing searches conducted a dedicated offline analysis, the PyCBC-IMBH search identified it with a FAR of 0.47 yr−1. However, the cWB and GstLAL-IMBH searches did not identify the event. The transient did not pass cWB veto threshold as discussed in Sect. 3.1. The event is, however, identified by the most general cWB search for GW bursts of short duration with a FAR of ∼2 yr−1, which is still consistent with the noise origin. The model-agnostic BayesWave (BW) analysis (Cornish & Littenberg 2015; Cornish et al. 2021) also identified the event with a FAR of ∼1 yr−1. The event morphologically resembles the Tomte class of glitches (Davis et al. 2021), which are common but of unknown origin. Follow-up of LIGO Livingston data also showed the existence of multiple comparable glitches within 100 s of the event time.
The remaining two events, 190924_232654 and 191223_014159, were likely caused by instrumental noise. In the case of 190924_232654, fast light scattering noise extending up to about 60 Hz is present in the LIGO Livingston data around the time of the event, and there is a high-S/N glitch in the Virgo data. The time-frequency morphology of the 191223_014159 signal in the LIGO Livingston data matches an instrumental glitch (Davis et al. 2021).
Appendix B: Follow-up studies of 200114_020818
The cWB offline search detects 200114_020818 with a combined FAR of 0.058 yr−1. The network S/N with an LHO-LLO network is 12.3, and the three-detector-network S/N is 14.5. Although we cannot exclude the terrestrial origin of 200114_020818, we did perform several follow-up studies on this candidate, which we summarize here. The studies include event reconstruction by BayesWave (BW) and cWB, parameter estimation (PE) with models of black hole binary merger including effects of orbital precession and higher order multipole emission, reconstruction of PE sample waveforms and a comparison to the event reconstruction with cWB, and residual analysis with BW.
B.1. Investigation of instrumental noise
As mentioned in Sect. 4, an instrumental noise transient at the Hanford observatory coincides with 200114_020818. The noise originates from a fan on a laser controller located on top of a squeezed light optics enclosure. At the time of the event, an accelerometer detected a second-long frequency dip in the 76 Hz fan motion. Such vibration transients can weakly couple to the strain data through the squeezing system. We performed a follow-up investigation using the methods described in Nguyen et al. (2021) to acquire accurate estimates of the vibration coupling between the table accelerometer and the strain channel. We estimate the expected noise in the strain channel at the fundamental frequency to be over an order of magnitude below background levels, so the fan is highly unlikely to account for the event candidate; however, the estimated noise at the first harmonic (152 Hz) is about a factor of two below the background and could potentially impact parameter estimation.
B.2. Event reconstruction by model-independent analysis
We reconstructed the signal using two model-independent analyses, namely cWB (used in the searches) and BW. The BW algorithm constructs the signal as a linear combination of sine-Gaussian wavelets and does not use any astrophysical model. The cWB reconstructs the multi-detector maximum likelihood signal by using the inverse wavelet transformation with selected pixels. In Fig. B.1, the red colored solid and dotted blue curves correspond to the whitened reconstructed signal from cWB and BW, respectively. The cWB event reconstruction is within the 90% credible region of the event reconstruction by BW (blue shaded region) for all three detectors. The BW S/N is 4, 14, and 5 in LHO, LLO, and Virgo, respectively, while cWB S/N is 5, 12, and 6, obtained from the reconstructed event.
B.3. PE analysis
Here, we investigate the possibility that 200114_020818 may be described by the merger of a quasi-circular BBH system. We thus carried out a parameter estimation with up-to-date waveform models including effects of precession and higher order multipole moments. Specifically, we used three quasi-circular BBH waveform models h(t; θ): i) the numerical relativity surrogate model NRSur7dq4 (Varma et al. 2019); ii) the effective one-body model SEOBNRv4PHM (Ossokine et al. 2020; Babak et al. 2017); and iii) the phenomenological model IMRPhenomXPHM (Pratten et al. 2021). We performed the analysis on 8 s of data centered around 200114_020818. All analyses were performed on C01 60 Hz subtracted data with a lower cutoff frequency of 10 Hz and reference frequency of 11 Hz. For the IMRPhenomXPHM analysis, we used the nested sampling algorithm as implemented in LALInference (Veitch et al. 2015), while for SEOBNRv4PHM and NRSur7dq4 analyses we use the RIFT (Lange et al. 2017) analysis tool instead. Both algorithms are designed to compute the joint 15-dimensional posterior distribution p(θ|D, H). The evidence computed via Eq. 5 may be employed for model selection; the evidence for the signal hypothesis H, described by a specific model with given priors, is compared to the evidence for the null hypothesis N, according to which no signal is present, by computing the (log10) Bayes factor as given by
Fig. B.1. Consistency of the waveform reconstruction by cWB with three waveform models in the time domain: IMRPhenomXPHM (upper panel), SEOBNRv4PHM (middle panel), and NRSur7dq4 (lower panel). The colored solid red and dashed blue curves correspond to the whitened reconstructed event by cWB and Bayeswave, respectively. The blue shaded region corresponds to the 90% credible region from the event construction by BW. The gray shaded belts are reconstructed waveforms by cWB for the 90% credible interval corresponding to the PE runs. |
The log10BSN3 is tabulated in Table B.1 for all the three runs. The values of log10BSN indicate a preference for the hypothesis H that a signal is present over the alternative of only Gaussian noise.
Summary of median and 90% credible intervals of 200114_020818 for different waveform models. The columns show the waveform model used for parameter estimation, the source frame component masses mi, effective spin parameters χeff and χp, luminosity distance DL, the angle between the total angular momentum and the direction of propagation of the gravitational wave signal θJN, and the log10Bayes Factor between the signal and Gaussian noise given the model.
These results do not, however, address the possibility that excess power in one or more detectors may be due to an instrumental artifact (glitch). As a diagnostic, we therefore performed a coherence test (Veitch & Vecchio 2008) using the IMRPhenomXPHM waveform model. The coherence test computes the Bayes factor for the coherent signal hypothesis against the hypothesis of an incoherent signal in the network of detectors. It can be thus interpreted loosely as an indicator of the presence of accidentally coincident noise artefacts that could mimic an astrophysical signal. The resulting log10 Bayes factor for coherent versus incoherent signal 0.2, providing little to no evidence in support of the coherent signal hypothesis. Such insignificant evidence is easily understood by looking at the log10 Bayes factors computed from analyses of each individual detector’s data: both Hanford-only and Virgo-only analyses recover a log10 Bayes factor for the signal versus Gaussian noise hypothesis of 0.2. As a consequence, the posterior distributions from the Hanford- and Virgo-only analyses are largely uninformative. On the other hand, a Livingston-only analysis finds a log10 Bayes factor of 25. Hence, from the parameter estimation point of view, 200114_020818 is essentially a single detector event. Returning to the results under the hypothesis H of a quasi-circular merger signal plus Gaussian noise, we summarize the resulting median and symmetric 90% credible regions for a few astrophysically relevant parameters from each of the models in Table B.1.
Figure B.2 shows the joint posterior distribution for the component masses m1 and m2 of the source according to each waveform model. The three models infer BH masses that are largely inconsistent. In particular, the inferred values – median and 90% credible intervals – show little overlap (see Table B.1). Moreover, the result from SEOBNRv4PHM shows a hint of bimodality in the mass posterior distributions.
The posterior distributions for the spin parameters (Fig. B.2) tell a similar story. If we compare the joint posterior distributions for the effective spin parameter χeff along the direction of the orbital angular momentum and the in-plane effective spin parameter χp, we find that the IMRPhenomXPHM and SEOBNRv4PHM probability distributions are disjointed at the 90% credible level. NRSur7dq4 instead recovers a posterior distribution that is much broader and encompasses both the posterior from IMRPhenomXPHM and SEOBNRv4PHM. With reference to Table B.1, all the three results indicate a preference toward the system being pr