Follow-up of the Neutron Star Bearing Gravitational-wave Candidate Events S190425z and S190426c with MMT and SOAR

, , , , , , , , , , , , , , , , , , , , , , , , , and

Published 2019 July 18 © 2019. The American Astronomical Society. All rights reserved.
, , Citation G. Hosseinzadeh et al 2019 ApJL 880 L4 DOI 10.3847/2041-8213/ab271c

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

2041-8205/880/1/L4

Abstract

On 2019 April 25.346 and 26.640 UT the Laser Interferometer Gravitational-Wave Observatory (LIGO) and Virgo gravitational-wave (GW) observatory announced the detection of the first candidate events in Observing Run 3 that contained at least one neutron star (NS). S190425z is a likely binary neutron star (BNS) merger at dL = 156 ± 41 Mpc, while S190426c is possibly the first NS–black hole (BH) merger ever detected, at dL = 377 ± 100 Mpc, although with marginal statistical significance. Here we report our optical follow-up observations for both events using the MMT 6.5 m telescope, as well as our spectroscopic follow-up of candidate counterparts (which turned out to be unrelated) with the 4.1 m SOAR telescope. We compare to publicly reported searches, explore the overall areal coverage and depth, and evaluate those in relation to the optical/near-infrared (NIR) kilonova emission from the BNS merger GW170817, to theoretical kilonova models, and to short gamma-ray burst (SGRB) afterglows. We find that for a GW170817-like kilonova, the partial volume covered spans up to about 40% for S190425z and 60% for S190426c. For an on-axis jet typical of SGRBs, the search effective volume is larger, but such a configuration is expected in at most a few percent of mergers. We further find that wide-field γ-ray and X-ray limits rule out luminous on-axis SGRBs, for a large fraction of the localization regions, although these searches are not sufficiently deep in the context of the γ-ray emission from GW170817 or off-axis SGRB afterglows. The results indicate that some optical follow-up searches are sufficiently deep for counterpart identification to about 300 Mpc, but that localizations better than 1000 deg2 are likely essential.

Export citation and abstract BibTeX RIS

1. Introduction

The joint detection of gravitational waves (GWs) and electromagnetic (EM) radiation from the binary neutron star (BNS) merger GW170817 was a watershed event. The merger was accompanied by a weak short gamma-ray burst (SGRB), by ultraviolet (UV)/optical/near-infrared (NIR) emission due to a kilonova (during the first month), and by radio, X-ray, and long-term optical emission due to an off-axis jet (LIGO Scientific Collaboration & Virgo Collaboration 2017, 2017a; LIGO Scientific Collaboration & Virgo Collaboration 2017b). GW170817 was localized to a region of about 30 deg2 and to a distance of 40 ± 8 kpc, which enabled both galaxy-targeted and wide-field searches to rapidly identify the EM counterpart, within about 11 hr of merger (Arcavi et al. 2017; Coulter et al. 2017; Lipunov et al. 2017; Soares-Santos et al. 2017; Tanvir et al. 2017; Valenti et al. 2017).

Observing run 3 (O3) of the Advanced Laser Interferometer Gravitational-Wave Observatory (LIGO) and Advanced Virgo (ALV) commenced on 2019 April 1, with a 50% increase in sensitivity compared to Observing Runs 1 and 2. The resulting BNS merger detection distances in O3 are on average about 140 Mpc for LIGO Livingston, 110 Mpc for LIGO Hanford, and 50 Mpc for Virgo. Given the volumetric merger rate inferred from GW170817, 110–3840 Gpc−3 yr−1 (LIGO Scientific Collaboration & Virgo Collaboration 2018), the expected number of BNS merger detections in the year-long O3 is ∼1–20 (assuming 70% duty cycle for LIGO). For neutron star (NS)–black hole (BH) mergers the upper bound on the rate based on non-detections in O1 and O2 is ≲600 Gpc−3 yr−1 (LIGO Scientific Collaboration & Virgo Collaboration 2018); however, given their larger detection volume relative to BNS mergers the observed NS–BH merger rate in O3 may exceed that of BNS mergers.

On 2019 April 25 at 08:18:05.017 UTC ALV detected a GW candidate event, designated S190425z, with a false alarm rate (FAR) of 1 in 7 × 104 yr, a probability of being a BNS merger of >99%, and a luminosity distance of 155 ± 45 Mpc (LIGO Scientific Collaboration & Virgo Collaboration 2019a). Because the event was detected only by LIGO Livingston and marginally by Virgo (LIGO Hanford was offline at the time of detection), the localization region had an initial area of about 104 deg2 (90% confidence; LIGO Scientific Collaboration & Virgo Collaboration 2019a), which was refined to 7460 deg2 about 31 hr post merger (LIGO Scientific Collaboration & Virgo Collaboration 2019b). The most up-to-date sky localization region and distance estimate is shown in Figure 1.

Figure 1.

Figure 1. GW localization regions of S190425z (left) and S190426c (right) overlaid with the locations of follow-up observations from our search with MMT, and all publicly reported searches that provided telescope pointing information. We note that the searches include both galaxy-targeted and wide-field imaging; we do not plot the fields of view of the individual telescopes.

Standard image High-resolution image

Soon after, on 2019 April 26 at 15:21:55.337 UTC, ALV detected another GW candidate event, designated S190426c, with an FAR of 1 in 1.7 yr, a probability of containing a NS of >99%, a luminosity distance of 375 ± 108 Mpc, and an initial localization region of about 1260 deg2 (90% confidence; LIGO Scientific Collaboration & Virgo Collaboration 2019c), which was refined to 1130 deg2 about 20 hr post merger (LIGO Scientific Collaboration & Virgo Collaboration 2019d). Given the relatively high FAR, we cannot be certain that the event was astrophysical. However, under the assumption that it was, the latest parameter estimation gives a 60% probability that the more massive binary component was >5 M, a 25% probability that it was 3–5 M, and a 15% probability that it was <3 M, suggesting that the NS–BH classification is most likely (LIGO Scientific Collaboration & Virgo Collaboration 2019e). The most up-to-date sky localization region and distance estimate is shown in Figure 1.

Here we report our optical follow-up of both events, using the MMT 6.5 m telescope to target galaxies within their localization volumes. In Section 2 we present our MMT observations. In Section 3 we collate searches reported publicly via the GRB Coordinates Network (GCN) circulars to explore a few aspects of the follow-up and announced candidates, as well as our spectroscopic follow-up of two candidates. In Sections 4 and 5 we compare the results to the kilonova emission of GW170817, to theoretical kilonova models, and to on-axis and slightly off-axis SGRB afterglow models. In Section 6 we collate γ-ray and X-ray searches and compare to the same SGRB models. We summarize and draw some initial conclusions in Section 7.

2. Galaxy-targeted Follow-up with MMT

Since the beginning of O3 we have been using the MMT 6.5 m telescope at Fred L. Whipple Observatory in Arizona to carry out follow-up observations of galaxies within the localization volumes of GW alerts. Upon receipt of an alert, our automated software generates a list of galaxies in the Galaxy List for the Advanced Detector Era (GLADE) catalog (Dálya et al. 2018) that are located within the 90% confidence volume, ranked by probability within the volume. The software also downloads reference images and catalogs from the Pan-STARRS1 (PS1) 3π database (Chambers et al. 2016) and collates the locations of all previously reported transients and moving objects from the Zwicky Transient Facility Public Survey (Bellm et al. 2019; Graham et al. 2019; Masci et al. 2019, via the MARS broker19 ), all other public time-domain surveys (via the Transient Name Server20 ), and the Minor Planet Center (via the SkyBoT service; Berthier et al. 2006). A custom data reduction pipeline processes each image as it is read out and performs image subtraction (using PyZOGY; Zackay et al. 2016; Guevel & Hosseinzadeh 2017). The reference, science, and subtracted images are then inspected for new transients using a custom web interface based on Flask and JS9 (Mandel & Vikhlinin 2018). An example from our search in the localization region of S190425z is shown in Figure 2.

Figure 2.

Figure 2. The first galaxy imaged with our MMT program in the field of S190425z (left panel), along with the corresponding reference image from PS1 3π (middle panel; Chambers et al. 2016), and the resulting subtraction (right panel). The difference image exhibits only astrometric noise and cosmic-ray artifacts; no possible counterpart is identified in this image to a limit of g = 22.3 mag.

Standard image High-resolution image

For S190425z, we commenced observations using the MMTCam imager on 2019 April 25 at 11:39:23 UT, 3.4 hr post merger, and continued until morning twilight, with our last exposure ending at 12:06:46 UT (Hosseinzadeh et al. 2019b). We obtained 30-s g-band exposures of 17 galaxies.21 On the following night, from 08:30:29 to 10:55:00 UT (24.2–26.6 hr post merger), we imaged 50 additional galaxies in i-band, to minimize moonlight contamination (Hosseinzadeh et al. 2019a). No transient sources were uncovered in these observations to median 3σ limiting magnitudes of g = 22.0 and i = 22.5. We provide the information for all of the individual galaxies in Table 1.

Table 1.  Log of MMT Follow-up Observations

Name R.A. Decl. Date Time Filter Limiting Mag.a
S190425z
2MASX J16545364–1657072 16h54m53fs65 −16d57m07fs3 2019 Apr 25 11:39:23 g 22.3
2MASX J16530485–1617273 16h53m04fs86 −16d17m27fs4 2019 Apr 25 11:40:50 g 22.4
2MASX J16571426–0613510 16h57m14fs26 −06d13m51fs0 2019 Apr 25 11:42:30 g 22.4
2MASX J16520774–1703135 16h52m07fs75 −17d03m13fs5 2019 Apr 25 11:44:22 g 22.5
PGC 58929 16h45m54fs64 −23d27m06fs0 2019 Apr 25 11:47:33 g 22.0
NGC 6234 16h51m57fs34 +04d23m00fs8 2019 Apr 25 11:49:28 g 22.4
PGC 59064 16h49m33fs16 +06d00m58fs5 2019 Apr 25 11:51:21 g 22.2
PGC 59201 16h53m24fs09 +04d14m10fs5 2019 Apr 25 11:52:39 g 22.3
2MASX J16564688–0142052 16h56m46fs89 −01d42m05fs3 2019 Apr 25 11:54:12 g 22.0
2MASX J16504669+0436170 16h50m46fs70 +04d36m17fs0 2019 Apr 25 11:55:46 g 22.2
UGC 10426 16h30m50fs16 +16d15m02fs5 2019 Apr 25 11:57:27 g 21.9
PGC 90265 16h57m26fs82 −10d11m27fs9 2019 Apr 25 11:59:14 g 21.7
NGC 6225 16h48m21fs57 +06d13m22fs0 2019 Apr 25 12:00:51 g 21.7
2MASX J16462248+0902154 16h46m22fs49 +09d02m15fs5 2019 Apr 25 12:02:05 g 21.6
PGC 58705 16h39m26fs39 +11d12m37fs7 2019 Apr 25 12:03:15 g 21.8
2MASX J16552449–0715255 16h55m24fs50 −07d15m25fs5 2019 Apr 25 12:04:57 g 21.7
2MASX J16580128–0149216 16h58m01fs29 −01d49m21fs7 2019 Apr 25 12:06:16 g 21.4
2MASX J16590728–0544311 16h59m07fs28 −05d44m31fs2 2019 Apr 26 08:30:29 i 22.2
PGC 58987 16h47m24fs48 −20d08m30fs3 2019 Apr 26 08:32:14 i 22.2
NGC 6224 16h48m18fs55 +06d18m43fs9 2019 Apr 26 08:34:11 i 22.5
NGC 6051 16h04m56fs70 +23d55m58fs3 2019 Apr 26 08:36:37 i 21.8
IC 4572 15h41m54fs20 +28d08m02fs7 2019 Apr 26 08:42:22 i 22.0
UGC 10320 16h18m07fs32 +21d03m59fs0 2019 Apr 26 08:44:30 i 22.5
IC 4569 15h40m48fs35 +28d17m31fs4 2019 Apr 26 08:47:02 i 22.1
PGC 57607 16h14m57fs84 +21d56m17fs9 2019 Apr 26 08:50:11 i 22.6
PGC 57472 16h12m20fs34 +23d00m07fs0 2019 Apr 26 08:51:41 i 22.6
IC 1219 16h24m27fs44 +19d28m57fs3 2019 Apr 26 08:53:28 i 22.6
UGC 10412 16h29m36fs14 +15d39m30fs4 2019 Apr 26 08:55:24 i 22.6
NGC 6001 15h47m45fs96 +28d38m30fs7 2019 Apr 26 08:57:41 i 22.1
PGC 55883 15h43m46fs04 +28d24m54fs6 2019 Apr 26 08:59:39 i 22.3
PGC 57645 16h15m42fs15 +19d38m15fs0 2019 Apr 26 09:02:13 i 22.6
PGC 59121 16h51m21fs66 +07d51m44fs3 2019 Apr 26 09:04:08 i 22.1
PGC 56949 16h04m35fs57 +25d11m23fs4 2019 Apr 26 09:06:30 i 22.2
PGC 58860 16h44m09fs28 +07d26m43fs0 2019 Apr 26 09:08:30 i 22.0
PGC 57542 16h13m46fs01 +22d55m08fs0 2019 Apr 26 09:10:44 i 22.4
UGC 10224 16h08m50fs24 +22d02m33fs5 2019 Apr 26 09:13:36 i 22.6
PGC 58097 16h25m38fs08 +16d27m18fs0 2019 Apr 26 09:16:57 i 22.6
PGC 1717114 16h04m16fs28 +24d48m44fs4 2019 Apr 26 09:18:57 i 22.6
PGC 59239 16h54m24fs03 −09d53m21fs3 2019 Apr 26 09:23:22 i 22.5
UGC 10260 16h11m57fs88 +20d55m24fs5 2019 Apr 26 09:25:28 i 22.7
IC 4570 15h41m22fs56 +28d13m47fs3 2019 Apr 26 09:38:18 i 22.7
UGC 10035 15h47m36fs35 +26d03m49fs2 2019 Apr 26 09:44:32 i 22.3
PGC 57692 16h16m45fs71 +19d31m16fs9 2019 Apr 26 09:46:51 i 22.4
2MASX J16505342–1500143 16h50m53fs42 −15d00m14fs3 2019 Apr 26 09:49:19 i 21.9
NGC 6240 16h52m58fs86 +02d24m03fs5 2019 Apr 26 09:51:27 i 22.3
UGC 10360 16h23m11fs34 +16d55m57fs4 2019 Apr 26 09:53:38 i 22.6
PGC 55774 15h40m36fs64 +28d30m44fs8 2019 Apr 26 09:56:12 i 22.6
PGC 58735 16h40m40fs22 +14d21m05fs3 2019 Apr 26 09:58:50 i 22.5
IC 4621 16h50m51fs19 +08d47m01fs9 2019 Apr 26 10:00:50 i 22.5
NGC 6075 16h11m22fs57 +23d57m54fs5 2019 Apr 26 10:07:05 i 22.5
2MASX J16582619–0319463 16h58m26fs20 −03d19m46fs4 2019 Apr 26 10:11:42 i 22.5
2MASX J16073961+2220315 16h07m39fs62 +22d20m31fs5 2019 Apr 26 10:15:40 i 22.5
PGC 58028 16h24m15fs15 +20d11m01fs0 2019 Apr 26 10:17:30 i 22.6
PGC 54895 15h22m44fs91 +29d46m11fs0 2019 Apr 26 10:19:52 i 22.3
IC 4505 14h46m33fs38 +33d24m31fs2 2019 Apr 26 10:21:52 i 22.7
PGC 58768 16h41m20fs90 +08d54m32fs6 2019 Apr 26 10:24:48 i 22.5
PGC 55373 15h32m46fs54 +28d22m01fs5 2019 Apr 26 10:27:25 i 22.7
IC 4587 15h59m51fs61 +25d56m26fs4 2019 Apr 26 10:29:22 i 22.7
UGC 08145 13h02m18fs28 +32d53m26fs8 2019 Apr 26 10:32:45 i 22.4
PGC 57293 16h09m06fs46 +24d52m13fs1 2019 Apr 26 10:36:12 i 22.5
2MASX J16540875–0738073 16h54m08fs76 −07d38m07fs3 2019 Apr 26 10:38:59 i 22.3
PGC 52138 14h35m18fs42 +35d07m07fs7 2019 Apr 26 10:41:50 i 22.2
2MASX J16153554+1927123 16h15m35fs54 +19d27m12fs4 2019 Apr 26 10:44:25 i 23.0
PGC 59338 16h58m05fs76 −21d16m26fs8 2019 Apr 26 10:46:47 i 22.2
UGC 09233 14h24m35fs03 +35d16m47fs4 2019 Apr 26 10:50:02 i 23.4
PGC 56421 15h56m03fs87 +24d26m52fs7 2019 Apr 26 10:52:12 i 22.6
PGC 1484188 16h28m52fs42 +15d25m14fs8 2019 Apr 26 10:54:31 i 22.5
S190426c
2MASX J18191810+8807285 18h19m18fs11 +88d07m28fs6 2019 Apr 27 08:38:51 i 22.0
2MASX J18215068+8642223 18h21m50fs68 +86d42m22fs4 2019 Apr 27 08:40:44 i 22.0
2MASX J18242867+8642139 18h24m28fs67 +86d42m14fs0 2019 Apr 27 08:42:02 i 22.1
2MASX J19301513+8540516 19h30m15fs13 +85d40m51fs6 2019 Apr 27 08:44:01 i 22.2
2MASX J20412914+8626330 20h41m29fs14 +86d26m33fs1 2019 Apr 27 08:47:16 i 22.3
2MASX J20441724+8654219 20h44m17fs24 +86d54m22fs0 2019 Apr 27 08:48:45 i 21.4
PGC 3085923 20h52m29fs72 +86d11m11fs9 2019 Apr 27 08:50:07 i 22.4
2MASX J20452666+8620428 20h45m26fs66 +86d20m42fs9 2019 Apr 27 08:53:22 i 22.2
2MASX J20592695+8454369 20h59m26fs95 +84d54m37fs0 2019 Apr 27 08:55:00 i 22.3
2MASX J20110295+4637149 20h11m02fs96 +46d37m14fs9 2019 Apr 27 09:00:56 i 22.3
2MASX J20113931+4550035 20h11m39fs32 +45d50m03fs6 2019 Apr 27 09:03:03 i 22.3
2MASX J20114858+4657335 20h11m48fs58 +46d57m33fs6 2019 Apr 27 09:04:39 i 22.2
2MASX J20132761+4630313 20h13m27fs62 +46d30m31fs3 2019 Apr 27 09:06:17 i 22.3
2MASX J20134502+4726333 20h13m45fs03 +47d26m33fs4 2019 Apr 27 09:07:51 i 22.3
2MASX J20152058+4555282 20h15m20fs59 +45d55m28fs3 2019 Apr 27 09:10:17 i 22.3
2MASX J20201548+4720364 20h20m15fs48 +47d20m36fs4 2019 Apr 27 09:12:40 i 22.2
2MASX J20242781+4900526 20h24m27fs81 +49d00m52fs7 2019 Apr 27 09:14:19 i 22.1
2MASX J20354336+4953165 20h35m43fs37 +49d53m16fs6 2019 Apr 27 09:15:47 i 22.4
2MASX J20224302+5636145 20h22m43fs03 +56d36m14fs6 2019 Apr 27 09:17:53 i 22.4
2MASX J20244336+5245430 20h24m43fs37 +52d45m43fs0 2019 Apr 27 09:19:32 i 22.3
2MASX J20245359+5610264 20h24m53fs60 +56d10m26fs4 2019 Apr 27 09:20:53 i 22.2
2MASX J20260256+5552523 20h26m02fs56 +55d52m52fs4 2019 Apr 27 09:22:19 i 22.3
2MASX J20273404+5015483 20h27m34fs04 +50d15m48fs3 2019 Apr 27 09:23:52 i 22.3
2MASX J20273859+5353393 20h27m38fs59 +53d53m39fs3 2019 Apr 27 09:25:24 i 22.4
2MASX J20281516+5641284 20h28m15fs17 +56d41m28fs4 2019 Apr 27 09:26:51 i 22.3
2MASX J20290191+5817016 20h29m01fs92 +58d17m01fs6 2019 Apr 27 09:28:14 i 22.3
2MASX J20291160+5219510 20h29m11fs60 +52d19m51fs0 2019 Apr 27 09:29:56 i 22.3
2MASX J20300804+5415120 20h30m08fs04 +54d15m12fs1 2019 Apr 27 09:31:48 i 22.2
2MASX J20304675+6259395 20h30m46fs76 +62d59m39fs5 2019 Apr 27 09:37:44 i 22.4
2MASX J20322187+5812031 20h32m21fs88 +58d12m03fs2 2019 Apr 27 09:39:26 i 22.3
2MASX J20334424+5403120 20h33m44fs25 +54d03m12fs0 2019 Apr 27 09:41:06 i 22.5
2MASX J20334533+6254178 20h33m45fs34 +62d54m17fs9 2019 Apr 27 09:42:45 i 22.4
2MASX J20352447+5759548 20h35m24fs47 +57d59m54fs9 2019 Apr 27 09:44:34 i 22.4
2MASX J20354995+6208172 20h35m49fs96 +62d08m17fs2 2019 Apr 27 09:47:17 i 22.4
2MASX J20355212+5549587 20h35m52fs13 +55d49m58fs8 2019 Apr 27 09:49:53 i 22.3
2MASX J20370886+5756538 20h37m08fs87 +57d56m53fs8 2019 Apr 27 09:51:24 i 22.5
2MASX J20373532+5628217 20h37m35fs32 +56d28m21fs7 2019 Apr 27 09:52:51 i 22.3
2MASX J20384775+6128473 20h38m47fs75 +61d28m47fs4 2019 Apr 27 09:54:43 i 22.3
2MASX J20385946+5351220 20h38m59fs47 +53d51m22fs0 2019 Apr 27 09:56:30 i 22.3
2MASX J20391360+6454369 20h39m13fs60 +64d54m37fs0 2019 Apr 27 09:58:12 i 22.5
2MASX J20404068+6437003 20h40m40fs69 +64d37m00fs3 2019 Apr 27 09:59:56 i 22.2
2MASX J20431546+5424171 20h43m15fs47 +54d24m17fs1 2019 Apr 27 10:02:36 i 22.3
2MASX J20450027+6441410 20h45m00fs28 +64d41m41fs0 2019 Apr 27 10:04:16 i 22.5
2MASX J20452857+6332249 20h45m28fs58 +63d32m25fs0 2019 Apr 27 10:05:41 i 22.3
2MASX J20453196+6157519 20h45m31fs96 +61d57m52fs0 2019 Apr 27 10:07:12 i 22.1
2MASX J20482612+6414178 20h48m26fs13 +64d14m17fs8 2019 Apr 27 10:08:31 i 22.4
2MASX J20492307+6412331 20h49m23fs08 +64d12m33fs2 2019 Apr 27 10:10:38 i 22.4
2MASX J20495907+6207478 20h49m59fs08 +62d07m47fs9 2019 Apr 27 10:12:07 i 22.3
2MASX J20515127+6309235 20h51m51fs28 +63d09m23fs5 2019 Apr 27 10:13:43 i 22.4
2MASX J20581565+6217178 20h58m15fs66 +62d17m17fs8 2019 Apr 27 10:15:28 i 22.3

Note.

aThese limiting magnitudes correspond to 3 times the sky noise within an aperture of 2.5 times the FWHM, where the sky noise is estimated to be 1.48 times the median absolute deviation of the difference image.

Download table as:  ASCIITypeset images: 1 2

For S190426c, we imaged 50 galaxies with 30-s i-band exposures on 2019 April 27 at 08:38:51 to 10:15:57 UT (17.3–18.9 hr post merger; Hosseinzadeh et al. 2019c). No transient sources were uncovered in these observations to a median 3σ limiting magnitude of i = 22.3 (see Table 1).

3. Summary of Community Follow-up and Our Spectroscopic Follow-up

Multiple teams reported UV, optical, and NIR follow-up imaging of the sky regions of S190425z and S190426c. In Table 2 (S190425z) and Table 3 (S190426c) we collate the available information, and summarize the timing of the observations relative to the merger time, the filter(s) used and limiting magnitudes, the sky area covered or number of galaxies targeted, and the relevant GCN circular references. In Figure 1 we map the searches that reported their telescope pointing coordinates relative to the GW localization regions.

Table 2.  Summary of Community Follow-up Observations of S190425z

GCN Num. Galaxies Area (deg2) Time (UT) Phase (days) Limiting Mag. Filter Instrument/Group
GCN24167 (Lipunov et al. 2019b) 2019 Apr 25 09:14:36 0.036 17.7 C MASTERa
GCN24172 (Lundquist et al. 2019) 60 2019 Apr 25 09:01:00 0.030 21 G SAGUAROb
GCN24175 (Rosell et al. 2019a) 5 22 B HETc
GCN24179 (Zheng et al. 2019a) 101 2019 Apr 25 12:43:52 0.034 19 clear KAITd
GCN24182 (Hosseinzadeh et al. 2019b) 17 2019 Apr 25 12:06:16 0.140 21 g MMTCam
GCN24183 (Im et al. 2019) 30 2019 Apr 25 09:38:57 0.056 20 i SQUEANe
GCN24187 (De et al. 2019b) 2401 2019 Apr 25 09:12:09 0.038 16.75 J Gattini-IR
GCN24188 (Paek et al. 2019a) 13 2019 Apr 25 10:17:06 0.083 19 R LOAOf
GCN24190 (Xu et al. 2019) 79 2019 Apr 25 12:40:09 0.182 18 Xinglong-Schmidt
GCN24191 (Kasliwal et al. 2019) 4327 2019 Apr 25 09:19:07 0.042 20.4 g, r ZTFg
GCN24192 (Sasada et al. 2019) 154 2019 Apr 25 11:46:00 0.144 23.5 r FOCASh
GCN24193 (Tan et al. 2019a) 27 2019 Apr 25 12:27:23 0.173 20 R LOTi
GCN24197 (McBrien et al. 2019) 2652 2019 Apr 25 09:18:02 0.042 19.5 o ATLASj
GCN24198 (Ahumada et al. 2019b) 10 2019 Apr 25 10:12:00 0.079 20.8 r KPEDk
GCN24207 (Hiramatsu et al. 2019a) 21 2019 Apr 25 13:42:17 0.191 21.7 g, i, r LasCumbres-SSOl
GCN24210 (Smith et al. 2019) 1258 2019 Apr 25 09:39:48 0.057 21.7 i Pan-STARRSm
GCN24216 (Kim et al. 2019) 120 2019 Apr 25 12:28:00 0.174 R KMTNetn
GCN24224 (Steeghs et al. 2019a) 2134 2019 Apr 25 20:38:00 0.514 20.1 L GOTOo
GCN24225 (Hiramatsu et al. 2019b) 19 2019 Apr 25 23:15:41 0.471 21.4 g, i, r LasCumbres-SAAOp
GCN24227 (Blazek et al. 2019a) 123 2019 Apr 25 19:43:51 0.277 17 TAROT-GRANDMAq
GCN24238 (Butler et al. 2019) 23 2019 Apr 26 11:40:00 0.808 RATIRr
GCN24239 (Watson et al. 2019a) 128 w COATLIs
GCN24244 (Hosseinzadeh et al. 2019a) 50 2019 Apr 26 10:54:31 1.009 21 i MMTCam
GCN24256 (Howell et al. 2019) 119 2019 Apr 26 01:03:45 0.292 19.2 GRANDMAt
GCN24270 (Hu et al. 2019b) 63 2019 Apr 25 09:57:38 0.069 20.5 clear BOOTES-5/JGTu
GCN24274 (Tan et al. 2019b) 58 20 R LOTi
GCN24285 (Li et al. 2019a) 675 20.983 V, R CNEOSTv
GCN24309 (Shappee et al. 2019a) 5000 2019 Apr 25 09:18:02 0.042 18.25 g ASAS-SNw
GCN24311 (Anand et al. 2019) 4950 21 g, r ZTFg
GCN24315 (Xin et al. 2019) 80 2019 Apr 26 13:58:39 0.164 16.86 R GWAC-F60Ax
GCN24353 (Tohuvavohu et al. 2019b) 408 2019 Apr 25 23:45:27 0.644 21.1 u UVOTy
GCN24367 (Vinko et al. 2019) 5 2019 Apr 25 21:36:00 0.554 21.5 r HETc

Notes. Compilation of all public follow-up searches reported for S190425z. The list includes both galaxy-targeted and wide-field searches, with their respective number of galaxies observed or area covered in square degrees. The start time of each observation, the approximate limiting magnitude of each instrument, and the filters used by each survey are also shown. The entries are sorted by GCN Circular number.

aMobile Astronomical System of Telescope Robots. bSearches After Gravitational Waves Using Arizona's Observatories. cHobby–Eberly Telescope. dKatzmann Automatic Imaging Telescope. eSpectral Energy Distribution (SED) Camera for Quasars in Early Universe. fLemonsan Optical Astronomical Observatory. gZwicky Transient Facility. hFaint Object Camera And Spectrograph. iLulin One-meter Telescope. jAsteroid Terrestrial-impact Last Alert System. kKitt Peak Electron-Multiplying Charge-Coupled Device (EMCCD) Demonstrator. lLas Cumbres Observatory node at Siding Spring Observatory. mPanoramic Survey Telescope and Rapid Response System. nKorean Microlensing Telescope Network. oGravitational-wave Optical Transient Observer. pLas Cumbres Observatory node at the South African Astronomical Observatory. qTélescopes à Action Rapide pour les Objets Transitoires, part of GRANDMAt. rReionization and Transients IR Camera. sCorrector de Óptica Áctiva y de Tilts al Límite de Difracción. tGlobal Rapid Advanced Network Devoted to the Multi-messenger Addicts. uBurst Observer and Optical Transient Exploring System 5, also known as the Javier Gorosabel Telescope. vChina Near Earth Object Survey Telescope. wAll-Sky Automated Survey for Supernovae. x60 cm telescope at Xinglong Observatory. yUV/Optical Telescope.

Download table as:  ASCIITypeset image

Table 3.  Summary of Community Follow-up Observations of S190426c

GCN Num. Galaxies Area (deg2) Time (UT) Phase (days) Limiting Mag. Filter Instrument/Group
GCN24236 (Lipunov et al. 2019c) 2019 Apr 26 16:25:18 0.037 18.3 Clear MASTERa
GCN24247 (Lim et al. 2019) 19 2019 Apr 26 16:27:00 0.045 SNUb
GCN24257 (Goldstein et al. 2019b) 830 2019 Apr 26 23:42:21 0.318 22.9 r, z DECamc
GCN24258 (Bhalerao et al. 2019) 7.5 20.5 r GITd
GCN24278 (Rosell et al. 2019b) 5 22 B HETe
GCN24281 (Zhu et al. 2019) 10 2019 Apr 26 17:19:27 0.082 18.5 Clear NEXT-0.6mf
GCN24283 (Coughlin et al. 2019) 4340 2019 Apr 27 05:45:00 0.599 22 g, r ZTFg
GCN24284 (Hankins et al. 2019a) 2200 2019 Apr 27 03:31:00 0.506 15 J Gattini-IR
GCN24286 (Li et al. 2019b) 774 2019 Apr 26 16:38:56 0.053 20.667 CNEOSTh
GCN24289 (Zheng et al. 2019b) 247 2019 Apr 27 06:26:42 0.525 19 KAITi
GCN24291 (Steeghs et al. 2019b) 755 2019 Apr 26 20:38:00 0.220 19.9 L GOTOj
GCN24292 (Hosseinzadeh et al. 2019c) 50 2019 Apr 27 10:15:28 0.720 21 i MMTCam
GCN24298 (Becerra et al. 2019) 98 2019 Apr 27 11:46:00 0.511 19.5 w COATLIk
GCN24299 (Niino et al. 2019) 39 15.4,17.6,—,17.7,20.0 H,J,K,R,clear J-GEMl
GCN24300 (Troja et al. 2019) 22 2019 Apr 27 11:41:00 0.506 21.2 i, g, Y, H RATIRm
GCN24310 (Watson et al. 2019b) 384 2019 Apr 27 11:01:00 0.612 w DDOTI/OANn
GCN24316 (Waratkar et al. 2019) 12.3 20.6 GITd
GCN24322 (Paek et al. 2019b) 23 2019 Apr 27 09:14:26 0.745 20.8 R LOAOo
GCN24323 (Shappee et al. 2019b) 973 2019 Apr 26 16:22:17 0.042 18.25 g ASAS-SNp
GCN24327 (Blazek et al. 2019b) 43.1 2019 Apr 27 02:32:20 0.263 19.6 r GRANDMAq
GCN24329 (Hankins et al. 2019b) 1900 2019 Apr 28 03:28:00 1.504 15 J Gattini-IR
GCN24331 (Perley et al. 2019b) 4420 22 g, r ZTFg
GCN24336 (Paek et al. 2019c) 17 2019 Apr 28 9:59:34 1.776 19.6 R LOAOo
GCN24340 (Izzo et al. 2019b) 5.0 2019 Apr 26 22:44:49 0.295 19.4 r Asiago-Schmidt
GCN24346 (Sun et al. 2019) 48 2019 Apr 27 14:11:17 0.949 18.40 Rc Yaoan
GCN24353 (Tohuvavohu et al. 2019b) 959 2019 Apr 26 17:44:46 0.099 21.1 u UVOTr

Notes. See notes in Table 2.

aMobile Astronomical System of Telescope Robots. bSeoul National University telescope. cDark Energy Camera. dGlobal Relay of Observatories Watching Transients Happen (GROWTH) India Telescope. eHobby–Eberly Telescope. fNingbo Bureau of Education and Xinjiang Observatory 0.6 m telescope. gZwicky Transient Facility. hChina Near Earth Object Survey Telescope. iKatzmann Automatic Imaging Telescope. jGravitational-wave Optical Transient Observer. kCorrector de Óptica Áctiva y de Tilts al Límite de Difracción. lJapanese Collaboration for Gravitational-wave EM Follow-up. mReionization and Transients IR Camera. nDeca–Degree Optical Transient Imager at the Observatorio Astronómico Nacional (México). oLemonsan Optical Astronomical Observatory. ppAll-Sky Automated Survey for Supernovae. qGlobal Rapid Advanced Network Devoted to the Multi-messenger Addicts. rUV/Optical Telescope.

Download table as:  ASCIITypeset image

3.1. S190425z

In all, 24 telescopes were reported to follow up S190425z, whose 90% confidence localization volume is about 8 × 106 Mpc3. The galaxy-targeted searches observed a combined total of 418 galaxies in this volume, corresponding to about 1% of the total number of galaxies in the GLADE catalog within the volume. Most searches observed galaxies more luminous than MB ≈ −19 mag (373 galaxies). Integrating the galaxy luminosity function down to this limit indicates 2.5 × 104 galaxies within the localization volume, confirming that the GLADE catalog is effectively complete at this luminosity. The fraction of observed bright galaxies was thus about 1.5%. To further quantify the effective coverage of the galaxy-targeted searches, we consider only the galaxies that were imaged to a sufficient depth to detect a GW170817-like kilonova (M ≈ −16 mag; Villar et al. 2017) at the distance of each galaxy. We find that 304 out of the 373 galaxies satisfy this criterion, leading to an effective coverage of about 1.2%. In the UV, Swift/Ultra-Violet/Optical Telescope (UVOT) observed 389 galaxies that are both in the GLADE catalog and have a redshift that would make it possible to detect a GW170817-like kilonova at the depth of the observations; this corresponds to an effective coverage of about 1.5%.

Similarly, for the wide-field searches we determine the effective fractional volume coverage using the distance to which each search would have detected a GW170817-like kilonova at the reported limiting magnitude, and combine this effective distance with the reported areal coverage. We find that most of the wide-field searches had an effective fractional volume coverage of about 0%–8%, while ZTF had a fractional volume coverage of about 40%; the values of zero correspond to searches that reported limiting magnitudes too shallow to have detected a GW170817-like kilonova at the lower distance limit of ≈115 Mpc.

Naturally, the fractional coverage of the galaxy-targeted and wide-field searches would be smaller (larger) for a dimmer (brighter) counterpart than in GW170817. We also note that our calculation is simplified, and likely errs on the side of being too optimistic. For example, we are not taking into account variations in Galactic extinction, moon illumination, and other differential observational effects that would generally serve to reduce the efficiency of the searches. On the other hand, other groups may have conducted follow-up campaigns that have not (yet) been publicly reported, which may increase the overall efficiency of the community effort. Lastly, the numbers in Tables 2 and 3 are uncertain due to possible human errors in real-time GCN composition (as ours had), but we assume these uncertainties are small compared to the uncertainty in the kilonova models.

The various searches returned 69 candidate optical counterparts, reported by ZTF, ATLAS, Pan-STARRS, Swift/UVOT, and Gaia, with candidates ranging in brightness from about 14 to 21.5 mag (Anand et al. 2019; Breeveld et al. 2019; Kasliwal et al. 2019; Kostrzewa-Rutkowska et al. 2019a, 2019b; McBrien et al. 2019; Smith et al. 2019); see Figure 3 for the brightness distribution. Of these, 18 candidates were followed up with at least one targeted observation, including seven events that were spectroscopically classified.

Figure 3.

Figure 3. Top panel: distribution of magnitudes (in u, g, r, i, o bands) for all reported candidates in the S190425z localization region (blue), those with any optical/NIR imaging follow-up (orange), and those with spectroscopic follow-up (red); the latter were all found to be normal supernovae. The target classified through our SOAR program are marked with arrows (Nicholl et al. 2019a). The vertical bar indicates the optical peak brightness of GW170817 for the 90% distance range of S190425z. Bottom panel: the same plot but for S190426c.

Standard image High-resolution image

Two of the classifications were obtained through our spectroscopic follow-up program using the 4.1 m SOAR telescope (Nicholl et al. 2019a). ZTF19aasckkq was selected based on a known spectroscopic redshift of z = 0.0528 for its host galaxy (Anand et al. 2019), which is within the 2σ contour of the the GW localization distance. The absolute magnitude of the source at this distance was −16.3 mag, comparable to GW170817 at peak. We obtained a 1900 s exposure beginning at 2019 April 28 05:13:50 UT using the Goodman High Throughput Spectrograph (Clemens et al. 2004) with the 400 lines/mm grating and a central wavelength of 5750 Å. The data were processed and the one-dimensional spectrum extracted using a custom Iraf pipeline (for details, see Margutti et al. 2019). The spectrum shows broad H i lines at the redshift of the host galaxy, as well as a strong absorption consistent with He i λ5876. Classification using the Supernova Identification code (SNID; Blondin & Tonry 2007) indicates that this transient is likely a young Type IIb supernova (SN IIb; Figure 4). Our second SOAR target was ZTf19aasckwd. This event did not have a reported redshift, but the apparent magnitude of 20.15 was a good match to the brightness of GW170817 at ∼155 Mpc. Moreover, a search of the coordinates in PS1 3π data (Flewelling et al. 2016) showed a likely host galaxy. We obtained a 1500 s exposure in the same setting as above, beginning at 2019 April 28 04:41:03 UT. The spectrum revealed a SN Ia at z = 0.145 (Figure 4). Both spectra are publicly available via the Transient Name Server (ZTF19aasckkq = SN2019eff, ZTf19aasckwd =SN2019eib).

Figure 4.

Figure 4. Spectroscopic classification using SOAR of optical transients in the S190425z localization region (Nicholl et al. 2019a). Top panel: ZTF19aasckkq was selected based on a host spectroscopic redshift of z = 0.0528 and transient absolute magnitude of ≈ −16.3. We classify this event as a SN IIb, as shown by comparison to SN 1993J. Bottom panel: ZTf19aasckwd was selected based on an apparent magnitude consistent with GW170817 at 155 Mpc. We classify this event as a SN Ia at z = 0.145, as shown by comparison to SN 2011fe.

Standard image High-resolution image

The other five candidates classified by the community also turned out to be normal SNe: one SN Ia, three SNe II, and one SN Ib/c (Anand et al. 2019; Buckley et al. 2019; Carini et al. 2019; Castro-Tirado et al. 2019; Dichiara et al. 2019; Dimitriadis et al. 2019; Izzo et al. 2019a; Jencson et al. 2019; Jonker et al. 2019; McCully et al. 2019; Morokuma et al. 2019; Nicholl et al. 2019b, 2019c; Pavana et al. 2019; Perley et al. 2019a; Short et al. 2019a, 2019b; Wiersema et al. 2019). Additionally, a UV candidate uncovered by Swift/UVOT (Breeveld et al. 2019) was shown to be an M dwarf flare (Bloom et al. 2019; Lipunov et al. 2019a). No NIR candidates were announced.

As shown in Figure 3, the bulk of reported candidates overlapped the optical brightness of a GW170817-like kilonova in the 90% confidence distance range (≈19–21 mag). We further find that the subset of events followed up photometrically and/or spectroscopically similarly span the same magnitude range. In terms of the choice of targets for spectroscopic follow-up, four of the seven classified transients (and four of the 11 transients with only photometric follow-up) were selected based on probable associations with galaxies that have secure distance measurements compatible with the GW distance (Anand et al. 2019; Kasliwal et al. 2019; Smith et al. 2019).

Conversely, six of of the 11 sources with only photometric follow-up were announced as apparently host-less (or "orphan") transients (Anand et al. 2019; McBrien et al. 2019). None of these sources were recovered in follow-up imaging (Ahumada et al. 2019a; Nicholl et al. 2019b; Perley & Copperwheat 2019), suggesting that they were potential image artifacts or due to stellar variability. Follow-up of host-less transients therefore appears to be a somewhat risky strategy, although we note that some mergers may occur at large offsets from their hosts: based on the distribution of SGRB offsets (Fong & Berger 2013), about 10% of BNS and/or NS–BH mergers may have offsets of tens of arcseconds from their hosts at the ALV detection distances.

3.2. S190426c

In all, 21 telescopes were reported to follow up S190426c, whose 90% confidence localization volume is about 2 × 107 Mpc3. The searches that targeted individual galaxies observed a combined total of 378 galaxies in this volume, corresponding to about 3.5% of the total number of galaxies in the GLADE catalog within the volume; however, the GLADE catalog is highly incomplete at the distance of S190426c. Instead, integrating the galaxy luminosity function at MB ≲ −19 mag, we find about 3.1 × 104 galaxies within the volume. Thus, the galaxy-targeted searches covered about 1.2% of the galaxies. Comparing the depth of the searches to the expected brightness of a GW170817-like kilonova at the distance of S190426c (21–23 mag), we find an effective fractional coverage of about 0.1%.

For the wide-field searches we determine the effective fractional volume coverage in the same manner as for S190425z. We find that most of the wide-field searches had an effective fractional volume of ≈0% because they did not reach sufficient depth to detect a GW170817-like kilonova even at the lower bound of the distance range. However, ZTF and DECam had effective volume coverages of about 55% and 8% (Goldstein et al. 2019a), respectively.22

In total, 30 candidate optical counterparts were reported by Las Cumbres Observatory, DECam, ZTF, GRAWITA, and Gaia, ranging from about 15–21.5 mag (Andreoni et al. 2019; Arcavi et al. 2019; Coughlin et al. 2019; Izzo et al. 2019b; Kostrzewa-Rutkowska et al. 2019c; Perley et al. 2019b); see Figure 3 for the brightness distribution. Of these, eight were observed spectroscopically, leading to classifications of three SNe Ia, one SN II, one broad-lined SN Ic, and one Galactic cataclysmic variable (Hu et al. 2019a; Sanchez-Ramirez et al. 2019; Valeev et al. 2019). The other two transients were not detected in their follow-up spectra (Cenko et al. 2019; De et al. 2019a). Only one additional candidate was followed up with imaging rather than spectroscopy. About half of the candidate counterparts and the subset classified spectroscopically are brighter than a GW170817-like kilonova in the 90% distance range (≈21–23 mag; Figure 3).

4. Comparison to GW170817 and Kilonova Models

In Figure 5 (left panel) we compare the limiting magnitudes of the various searches (galaxy-targeted and wide-field) to the model optical/NIR light curves of GW170817 (from Villar et al. 2017 23 ) shifted to the 90% distance ranges of S190425z and S190426c. We also show the shock cooling model of Piro & Kollmeier (2018), as it predicts potentially brighter emission in the first few hours post merger (which were missed in the case of GW170817). For both events some of the searches reached sufficient depth to detect a GW170817-like kilonova, although this was more challenging for S190426c.

Figure 5.

Figure 5. Top row: limiting magnitudes as a function of time post merger for UV/optical/NIR searches in the localization region of S190425z (diamonds represent wide-field searches; triangles represent galaxy-targeted searches; colors correspond to different filters). Left panels: a comparison to the kilonova in GW170817 scaled to the 90% confidence distance range of S190425z in r-band (red) and K-band (orange) based on the model fits from Villar et al. (2017). Additionally shown is a model of shock cooling emission for the early emission from GW170817 (blue; Piro & Kollmeier 2018). The vertical line marks the time when the optical counterpart of GW170817 was first detected. Right panels: a comparison to on-axis (blue) and slightly off-axis (purple) afterglow models (based on SGRBs; Fong et al. 2015), to a lanthanide-rich kilonova with an ejecta mass of 0.01 M (red), and to a neutron precursor (green; Metzger et al. 2015). Bottom row: same as top row, but for observations of S190426c, with the models scaled to its 90% confidence distance range.

Standard image High-resolution image

In the right panel of Figure 5 we compare the searches to several other models of early optical/NIR emission. We consider a kilonova that lacks the blue (lanthanide-poor) component and has an ejecta mass of 0.01 M for the red (lanthanide-rich) component (i.e., about 4 times lower than in GW170817); this model represents a more pessimistic possibility, but one that is supported by binary merger simulations (e.g., Hotokezaka et al. 2011). Both models were generated with MOSFiT (Guillochon et al. 2018). We also show a model for a blue precursor powered by the decay of free neutrons from the shock-heated interface between the NSs (Metzger et al. 2015). In the context of these models, which peak at ≈22 mag for S190425z and ≈24 mag for S190426c, we find that few (if any) observations reached sufficient depth to place meaningful constraints.

5. Comparison to On-axis and Off-axis SGRB Afterglows

Another source of early UV/optical/NIR emission is an on-axis or slightly off-axis relativistic jet, as observed in SGRBs (Berger 2014). In the absence of information about the binary inclination from GW data we cannot directly assess the viewing angle of a potential jet, but we note that based on jet opening angle measurements in SGRBs (Fong et al. 2015) we expect at most a few percent of GW mergers to exhibit on-axis jets (Metzger & Berger 2012). Conversely, for substantial off-axis angles (as was the case for GW170817 with θobs ≈ 30°; Alexander et al. 2017, 2018; Margutti et al. 2017, 2018) the optical emission is significantly delayed and exceedingly dim, making this scenario irrelevant for the rapid searches considered here.

We therefore consider two afterglow models: on-axis (θobs = 0°) and slightly off-axis (θobs = 15°). We generate light curves using the BOXFIT code (v2; van Eerten & MacFadyen 2011) for "top hat" jets24 using median values for cosmological SGRBs (Fong et al. 2015): jet opening angle of θj = 10°, isotropic kinetic energy of EK,iso = 2 × 1051 erg, circumburst density of n = 4 × 10−3 cm−3, electron energy power-law index p = 2.4, and fractional post-shock energies in the relativistic electrons and magnetic fields of epsilonE = 0.1 and epsilonB = 0.01, respectively. We note that this model is comparable to the inferred properties of the relativistic jet in GW170817.

The resulting model light curves, scaled to the distances of S190425z and S190426c, are shown in Figure 5 (right panels). We find that for S190425z, the on-axis afterglow model remains brighter than about 20 mag for the first day, exceeding the expected brightness of a GW170817-like kilonova. A substantial fraction of the searches reached sufficient depth to detect such an on-axis jet. In the case of S190426c, however, such an on-axis afterglow would have declined below 20 mag within about 0.3 days, although it would have still been detectable by at least some of the searches in the first day. We stress again that the probability of an on-axis merger is at most a few percent, so even in the case when the observations are sufficiently deep and cover the entire GW localization region, such a detection is unlikely. For the slightly off-axis afterglow model, the peak brightness is about 23 mag for S190425z and about 25 mag for S190426c. These brightness levels are well below the limiting magnitudes of the majority of follow-up observations, indicating the challenge of detecting off-axis afterglows at these distances.

6. Gamma-Ray and X-Ray Follow-up

Multiple γ-ray and X-ray space missions reacted to the GW alerts and/or were observing parts of the relevant sky area at the time of merger, including the Monitor of All-sky X-ray Image (MAXI), the Neil Gehrels Swift Observatory, the INTErnational Gamma-Ray Astrophysics Laboratory (INTEGRAL), the High-Altitude Water Cherenkov Observatory (HAWC), the Fermi Gamma-ray Space Telescope, the AstroRivelatore Gamma a Immagini Leggero (AGILE), the Hard X-ray Modulation Telescope (Insight-HXMT) and the CALorimetric Electron Telescope (CALET). No high-energy counterpart was identified with high statistical significance for either S190425z or S190426c (Axelsson et al. 2019a, 2019b; Barthelmy et al. 2019; Casentini et al. 2019; Chelovekov et al. 2019; Evans et al. 2019; Fletcher 2019a, 2019b; Guan et al. 2019; HAWC Collaboration 2019a, 2019b; Minaev et al. 2019; Piano et al. 2019a, 2019b; Sakamoto et al. 2019; Shimizu et al. 2019; Sugita et al. 2019; Sugizaki et al. 2019; Tamura et al. 2019; Xiao et al. 2019; Yi et al. 2019). The fractional localization coverage at the time of the GW detection varies widely, from a few percent to nearly 100%.

For S190425z there are four measurements of interest: (i) Fletcher (2019a) reported a Fermi-GBM 3σ flux limit in the range ${F}_{10-1000\mathrm{keV}}\lt (0.1-3)\times {10}^{-6}\,\mathrm{erg}\,{{\rm{s}}}^{-1}\,{\mathrm{cm}}^{-2}$ (depending on the assumed spectral model) for observations obtained at ±30 s relative to the merger time using a 1 s integration time. This corresponds to a luminosity limit of ${L}_{1-{10}^{4}\mathrm{keV}}\lt (0.03-6)\,\times {10}^{49}\,\mathrm{erg}\,{{\rm{s}}}^{-1}$ at 155 Mpc. These observations covered 51% of the initial probability map. (ii) Konus–Wind was observing the entire sky at the time of the GW trigger. Svinkin et al. (2019a) report a flux limit of $2.7\times {10}^{-7}\,\mathrm{erg}\,{{\rm{s}}}^{-1}\,{\mathrm{cm}}^{-2}$ (20–1500 keV) for a SGRB-like spectrum. (iii) Martin-Carillo et al. (2019) and Savchenko et al. (2019) reported the presence of a possible excess of γ-ray emission with limited significance in INTEGRAL data acquired ∼6 s after the GW detection. However, this excess is most probably due to background fluctuations. Under this preferred assumption, V. Savchenko et al. (2019, in preparation) estimate a typical 3σ upper limit on the 75–2000 keV fluence within 50% GW probability containment region of $(2-6)\times {10}^{-7}\,\mathrm{erg}\,{\mathrm{cm}}^{-2}$ depending on the sky location and for a burst lasting less than 1 s with a typical SGRB spectrum. (iv) In the time interval of 1054–5520 s post merger MAXI observed an area of the sky corresponding to 81% of the probability map, with a 1σ flux limit of ${F}_{4-10\mathrm{keV}}\lt 2\times {10}^{-10}\,\mathrm{erg}\,{\mathrm{cm}}^{-2}\,{{\rm{s}}}^{-1}$ corresponding to ${L}_{4-10\mathrm{keV}}\lt 6\times {10}^{44}\,\mathrm{erg}\,{{\rm{s}}}^{-1}$ (Sugizaki et al. 2019).

Similarly, for S190426c there are four measurements of interest: (i) Fletcher (2019b) reported a Fermi-GBM 3σ flux limit in the range ${F}_{10-1000\mathrm{keV}}\lt (1-9)\times {10}^{-7}\,\mathrm{erg}\,{{\rm{s}}}^{-1}\,{\mathrm{cm}}^{-2}$ at ±30 s relative to merger for a 1 s integration time, corresponding to ${L}_{1-{10}^{4}\mathrm{keV}}\lt (0.1-10)\times {10}^{49}\,\mathrm{erg}\,{{\rm{s}}}^{-1}$ at 377 Mpc. These observations covered about 100% of the initial probability region. (ii) From Konus–Wind observations covering the entire sky, Svinkin et al. (2019b) reported a flux limit of $7.3\,\times {10}^{-7}\,\mathrm{erg}\,{{\rm{s}}}^{-1}\,{\mathrm{cm}}^{-2}$ (20–1500 keV) for an SGRB-like spectrum. (iii) Swift/BAT observations covering 95% of the initial probability region obtained at ±100 s relative to merger indicate ${F}_{15-350\mathrm{keV}}\,\lt {10}^{-6}\,\mathrm{erg}\,{\mathrm{cm}}^{-2}\,{{\rm{s}}}^{-1}$ for a 1 s integration time, corresponding to ${L}_{15-350\mathrm{keV}}\lt 2\times {10}^{49}\,\mathrm{erg}\,{{\rm{s}}}^{-1}$ (Barthelmy et al. 2019). (iv) MAXI covered 76% of the probability region at 750–4488 s post merger to a 1σ limit of ${F}_{4-10\mathrm{keV}}\,\lt 2\times {10}^{-10}\,\mathrm{erg}\,{\mathrm{cm}}^{-2}\,{{\rm{s}}}^{-1}$, corresponding to ${L}_{4-10\mathrm{keV}}\lt 3\,\times {10}^{45}\,\mathrm{erg}\,{{\rm{s}}}^{-1}$ at d = 377 Mpc (Shimizu et al. 2019).

As shown in Figure 6, the limits on the prompt γ-ray emission for both events can rule out an on-axis SGRB comparable to the bulk of the energetic cosmological population, which have Eγ,iso ≈ 1051–1052 erg. Song et al. (2019) and Saleem et al. (2019) reached a similar conclusion. As indicated in Section 5, however, an on-axis orientation is expected in at most a few percent of the cases.

Figure 6.

Figure 6. Limits on the isotropic prompt γ-ray energy release from S190425z (dark blue) and S190426c (light blue) as constrained by Fermi-GBM, INTEGRAL, and Konus–Wind observations (Fletcher 2019a, 2019b; Svinkin et al. 2019a, 2019b; V. Savchenko et al. 2019, in preparation). The range of luminosity limits reflects the assumed spectral model used for the flux calibration. For S190425z (S190426c) Fermi-GBM covered about 50% (100%) of the initial GW probability map, while Konus–Wind covered the entire sky. These observations can rule out the most energetic on-axis cosmological SGRBs (red circles; LIGO Scientific Collaboration & Virgo Collaboration 2017a). Also shown for completeness is GRB170817 associated with GW170817 (LIGO Scientific Collaboration & Virgo Collaboration 2017a).

Standard image High-resolution image

In Figure 7 we compare the limits from MAXI to the observed X-ray afterglows of cosmological SGRBs, and find that they similarly rule out about half of the observed population, for the fractional areal coverage of each MAXI search. The same caveat about the rarity of on-axis events applies to these limits as well. On the other hand, if we compare the MAXI limits to the same off-axis model described in Section 5, we find that such a model cannot be constrained. We similarly find that Swift/XRT observations carried out for both events (Evans et al. 2019; Tohuvavohu et al. 2019a, 2019b) cannot constrain off-axis jets (Figure 7).

Figure 7.

Figure 7. MAXI X-ray observations (red) of S190425z (left panel, initial probability map coverage of 81%; Sugita et al. 2019), and S190426c (right panel, initial probability map coverage of 76%; Sugizaki et al. 2019). The limits rule out luminous cosmological SGRB X-ray afterglows (gray points). An off-axis jet as described in Section 5 (orange line) cannot be ruled out. Also shown are the luminosity limits for Swift/XRT observations with exposures of about 80 s (dotted line; e.g., Evans et al. 2019; Tohuvavohu et al. 2019a, 2019b), and the X-ray light curve of GW170817 from Margutti et al. (2018), Alexander et al. (2018), and Hajela et al. (2019).

Standard image High-resolution image

7. Conclusions

We presented MMT follow-up observations of the first two ALV candidate events in O3 that appear to contain NSs, and are therefore capable of generating EM emission. Our MMT search targeted 67 and 50 galaxies for S190425z and S190426c, respectively, and did not yield potential counterparts to a limiting magnitude of i ≈ 22.5. We further presented our spectroscopic follow-up with SOAR of two candidate optical counterparts from other searches, which revealed unrelated SNe Ia and IIb. For comparison we further collated information available from the GCN circulars about other galaxy-targeted and wide-field searches. Due to the large localization areas and volumes of both events all searches were far from complete. Still, a combined total of nearly 100 optical candidates were announced for the two events, and 14 were followed up spectroscopically, revealing normal SNe.

Parameterizing the efficacy of the searches relative to the brightness of the kilonova associated with GW170817, we find maximal volume coverage of about about 40% for S190425z (ZTF) and about 60% for S190426c (ZTF plus DECam). Relative to a dimmer kilonova model (0.01 M of lanthanide-rich ejecta), a neutron precursor, or a slightly off-axis SGRB, we find that only a few searches (including our MMT observations) reached sufficient depth. On the other hand, comparing to an on-axis SGRB we find that most searches would have been able to detect such emission, but this is expected in at most a few percent of mergers.

We end with a few general thoughts. First, the open rapid alerts implemented by the LIGO/Virgo Collaboration in O3 work remarkably well in providing rapid access to sky maps, distance estimates, and rudimentary information about the detections (e.g., FAR). Second, the events considered here indicate that due to duty cycle limitations and the larger detection distances, the localization regions (and volumes) for most events will be much larger than for GW170817, and this is likely to reduce the efficiency of counterpart identification. Third, despite the larger distances of the GW events at least some searches are capable of reaching the depth necessary to detect GW170817-like kilonovae (if those are common). Finally, we note that the robustness of the GW detections, as well as the actual properties of the binaries, are difficult to assess with the partial information provided by the LIGO/Virgo Collaboration at the present. In particular, we advocate breaking down the FAR by detector and by search pipeline, which would provide more guidance about the significance of a given event. Furthermore, the chirp mass and the individual component masses potentially provide critical insight about the expected EM signatures (Margalit & Metzger 2019). Early release of this additional information is particularly important as the number of detections increases in order to prioritize follow-up and tailor it to the properties of the transient.

We thank Dallan Porter and Sean Moran for software development that facilitated our rapid follow-up program at MMT, as well as Ben Kunk for operating the MMT during our observations. We also thank Joey Chatelain for suggesting the use of SkyBoT in our analysis software, and Cristiano Guidorzi for help with the BOXFIT simulations. The Berger Time-Domain Group is supported in part by NSF grant AST-1714498 and NASA grant NNX15AE50G. G.H. thanks the LSSTC Data Science Fellowship Program, which is funded by LSSTC, NSF Cybertraining Grant #1829740, the Brinson Foundation, and the Moore Foundation; his participation in the program has benefited this work. P.S.C. is grateful for support provided by NASA through the NASA Hubble Fellowship grant #HST-HF2-51404.001-A awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. M.N. is supported by a Royal Astronomical Society Research Fellowship. W.F. and K.P. acknowledge support by the National Science Foundation under award No. AST-1814782. I.P. acknowledges funding by the Deutsche Forschungsgemeinschaft under grant GE2506/12-1. R.M. acknowledges support for this work provided by the National Aeronautics and Space Administration through Chandra award No. DD8-19101A and DDT-18096A issued by the Chandra X-ray Center, which is operated by the Smithsonian Astrophysical Observatory for and on behalf of the National Aeronautics Space Administration under contract NAS8-03060. K.D.A. is grateful for support provided by NASA through the NASA Hubble Fellowship grant #HST-HF2-51403.001-A awarded by the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., for NASA, under contract NAS 5-26555. Development of the Boxfit code was supported in part by NASA through grant NNX10AF62G issued through the Astrophysics Theory Program and by the NSF through grant AST-1009863. This research was supported in part through the computational resources and staff contributions provided for the Quest high-performance computing facility at Northwestern University, which is jointly supported by the Office of the Provost, the Office for Research, and Northwestern University Information Technology. Observations reported here were obtained at the MMT Observatory, a joint facility of the Smithsonian Institution and the University of Arizona, and the Southern Astrophysical Research (SOAR) telescope, which is a joint project of the Ministério da Ciência, Tecnologia, Inovações e Comunicações (MCTIC) do Brasil, the U.S. National Optical Astronomy Observatory (NOAO), the University of North Carolina at Chapel Hill (UNC), and Michigan State University (MSU).

Facilities: ADS - , MMT (MMTCam) - , SOAR (Goodman). -

Software: Astropy (Astropy Collaboration 2018), BOXFIT (van Eerten & MacFadyen 2011), Flask, healpy (Zonca et al. 2019), JS9 (Mandel & Vikhlinin 2018), ligo.skymap (Singer 2019b), Matplotlib (Hunter 2007), MOSFiT (Guillochon et al. 2018), NumPy (van der Walt et al. 2011), Photutils (Bradley et al. 2019), PyGCN (Singer 2019a), PyZOGY (Guevel & Hosseinzadeh 2017), SciPy (Oliphant 2007), SEP (Bertin & Arnouts 1996; Barbary 2016), SNID (Blondin & Tonry 2007), SQLAlchemy.

Footnotes

  • 19 
  • 20 
  • 21 

    Our first GCN circular accidentally omitted four observed galaxies and included one galaxy that was not observed until the next night.

  • 22 

    The DECam observations covered a southern probability region that was mostly eliminated in the revised localization map released after the DECam observations occurred.

  • 23 

    These models are based on data obtained by Andreoni et al. (2017), Arcavi et al. (2017), Coulter et al. (2017), Cowperthwaite et al. (2017), Díaz et al. (2017), Drout et al. (2017), Evans et al. (2017), Hu et al. (2017), Kasliwal et al. (2017), Lipunov et al. (2017), Pian et al. (2017), Pozanenko et al. (2018), Shappee et al. (2017), Smartt et al. (2017), Troja et al. (2017), Utsumi et al. (2017), and Valenti et al. (2017).

  • 24 

    For the small viewing angles considered here, a more complex jet structure (as was inferred for GW170817; Alexander et al. 2018; Margutti et al. 2018; Wu & MacFadyen 2018) will make little difference.

Please wait… references are loading.
10.3847/2041-8213/ab271c