Draft version October 21, 2020 Typeset using LATEX RNAAS style in AASTeX62

GW190521: orbital eccentricity and signatures of dynamical formation in a binary merger signal Isobel Romero-Shaw,1, 2 Paul D. Lasky,1, 2 Eric Thrane,1, 2 and Juan Calderón Bustillo1, 2, 3

1School of Physics and Astronomy, Monash University, Clayton VIC 3800, Australia 2OzGrav: The ARC Centre of Excellence for Discovery, Clayton VIC 3800, Australia 3Department of Physics, The Chinese University of Hong Kong, Shatin, N.T., Hong Kong

ABSTRACT

Pair instability supernovae are thought to restrict the formation of black holes in the mass range ∼ 50−135 M . However, black holes with masses within this “high mass gap” are expected to form as the remnants of binary black hole mergers. These remnants can merge again dynamically in densely populated environments such as globular clusters. The hypothesis that the binary black hole merger GW190521 formed dynamically is supported by its high mass. Orbital eccentricity can also be a signature of dynamical formation, since a binary that merges quickly after becoming bound may not circularize before merger. In this work, we measure the orbital eccentricity of GW190521. We find that the data prefer a signal with eccentricity 푒 ≥ 0.1 at 10 Hz to a non-precessing, quasi-circular signal, with a log Bayes factor ln B = 5.0. When compared to precessing, quasi-circular analyses, the data prefer a non-precessing, 푒 ≥ 0.1 signal, with log Bayes factors ln B ≈ 2. Using injection studies, we find that a non-spinning, moderately eccentric (푒 = 0.13) GW190521-like binary can be mistaken for a quasi-circular, precessing binary. Conversely, a quasi-circular binary with spin-induced precession may be mistaken for an eccentric binary. We therefore cannot confidently determine whether GW190521 was precessing or eccentric. Nevertheless, since both of these properties support the dynamical formation hypothesis, our findings support the hypothesis that GW190521 formed dynamically.

INTRODUCTION Bromberg 2019; Bouffanais et al. 2019) and/or active galactic The first and second observing runs of the Advanced nuclei disks (Yang et al. 2019; McKernan et al. 2020; Gröb- LIGO (Abbott et al. 2018) and Virgo (Acernese et al. ner et al. 2020). Young star clusters may create something of 2015) gravitational-wave observatories yielded ten obser- a hybrid channel, with dynamical interactions perturbing the vations of stellar-mass black-hole binaries (Abbott et al. evolution of primordial stellar binaries, which evolve to make 2016a,b), reported in their first gravitational-wave transient merging double compact objects (Ziosi et al. 2014; Di Carlo catalogue (GWTC-1; Abbott et al. 2019a). The question et al. 2019a; Rastello et al. 2020). of how these binaries came to merge within the age of the The component masses and spins of a black-hole binary can Universe remains unanswered. Proposed formation channels illuminate its formation history, as can its orbital eccentricity typically fall into two categories: isolated, in which two stars (e.g., (Vitale et al. 2017; Stevenson et al. 2015; Rodriguez evolve side-by-side until they form black holes and coalesce et al. 2016; Farr et al. 2017; Fishbach & Holz 2017; Talbot (see, e.g., Livio & Soker 1988; Bethe & Brown 1998; de Mink & Thrane 2017)). Information about these parameters can be et al. 2010; Ivanova et al. 2013; Kruckow et al. 2016; de Mink extracted from the gravitational-wave signal. Both isolated & Mandel 2016), and dynamical, in which two black holes evolution and dynamical formation can produce black-hole bi- arXiv:2009.04771v3 [astro-ph.HE] 20 Oct 2020 become bound due to gravitationally-driven interactions in- naries with properties like those presented in GWTC-1, with side dense star clusters (e.g., Sigurdsson & Hernquist 1993; component masses 푚1, 푚2 . 50M , dimensionless compo- Portegies Zwart & McMillan 2000; O’Leary et al. 2006; Sam- nent spins 푎1, 푎2 consistent with 0, and eccentricities 푒 con- sing et al. 2014; Morscher et al. 2015; Gondán et al. 2018; sistent with 0 at 10 Hz (Abbott et al. 2019a; Romero-Shaw Samsing 2018; Rodriguez et al. 2018a; Randall & Xianyu et al. 2019). Dynamical formation is the preferred pathway 2018a,b; Samsing & D’Orazio 2018; Samsing et al. 2018; for binaries with more extreme masses (Gerosa & Berti 2017; Rodriguez et al. 2018b; Fragione & Kocsis 2018; Fragione & Rodriguez et al. 2019; Bouffanais et al. 2019; Fragione et al. 2020; Fragione et al. 2020), isotropically distributed spin tilt angles (Rodriguez et al. 2016; Talbot & Thrane 2017), and [email protected] non-zero orbital eccentricities (Zevin et al. 2017; Rodriguez 2 et al. 2018a; Gondán & Kocsis 2019; Samsing et al. 2018; gap.1 The data exhibit a modest preference (log Bayes factor Zevin et al. 2019a). ln B ≈ 2.4) for spin-induced precession of the orbital plane, The mass distribution of black holes that form as the rem- suggesting that the black-hole spin vectors may be signifi- nants of massive stars is thought to deplete between ∼ 50 and cantly misaligned from the orbital angular momentum axis. ∼ 135 M due to pair-instability supernovae (PISN; Heger If confirmed, the signature of precession would lend support & Woosley 2002; Özel et al. 2010; Belczynski et al. 2016; for the dynamical hypothesis. Marchant et al. 2016; Fishbach & Holz 2017; Woosley 2017) In this work, we show that GW190521 is consistent with unless exotic physics is invoked (Sakstein et al. 2020). The an eccentric merger. For brevity, we hereafter refer to the precise lower limit of the PISN mass gap is an area of active eccentricity measured at a gravitational-wave frequency of reseach; see Belczynski(2020), and references therein. The 10 Hz as 푒10. Our method allows us to study eccentricities remnants of binary black hole merger events can have masses up to 푒10 = 0.2, beyond which the waveform is not avail- within the PISN gap; see, e.g., (Abbott et al. 2016a; Fishbach able. Our analysis reveals overwhelming support for a spin- et al. 2017; Chatziioannou et al. 2019; Kimball et al. 2019, aligned eccentric signal with 푒10 ≥ 0.1 over a spin-aligned 2020). Second-generation mergers—where at least one of the quasi-circular signal. We use simulated events to demonstrate binary components is a remnant of a previous merger, poten- that precession and eccentricity cannot be distinguished for a tially within the mass gap—can occur in the high-density en- GW190521-like signal. We end with a discussion of the im- vironments conducive to dynamical mergers (Gerosa & Berti plications of our results on the potential formation mechanism 2017; Rodriguez et al. 2019; Bouffanais et al. 2019; Fragione of GW190521. et al. 2020). Prior to the detection of GW190521, no convinc- ing evidence has emerged for hierarchical mergers (Fishbach METHOD et al. 2017; Chatziioannou et al. 2019; Abbott et al. 2019b; We construct eccentric posterior probability density distri- Kimball et al. 2019, 2020). butions using the method developed in Romero-Shaw et al. Isolated binaries are thought to circularize efficiently, lead- (2019), which is built on those introduced by Payne et al. ing to negligible eccentricity close to merger (Peters 1964; (2019) and Lower et al.(2018). We use the Bayesian inference Hinder et al. 2008). While it is possible that the late-inspiral library Bilby (Ashton et al. 2019; Romero-Shaw et al. 2020) eccentricity of field mergers can be increased by Kozai-Lidov to perform an analysis using our “proposal” model: the spin- resonance (Kozai 1962; Lidov 1962) during three-body (Sils- aligned quasi-circular waveform model IMRPhenomD (Khan bee & Tremaine 2017; Antonini et al. 2017; Fishbach et al. et al. 2016). We reweight our IMRPhenomD posteriors to our 2017; Rodriguez & Antonini 2018; Fragione & Kocsis 2019a; “target” model: the spin-aligned eccentric waveform SEOB- Liu et al. 2019) and four-body (Liu & Lai 2019; Fragione & NRE (Cao & Han 2017; Liu et al. 2019). Our prior on eccen- Kocsis 2019b) interactions in the field, the relative rate of such tricity is log-uniform in the range −6 ≤ log10 (푒10) ≤ −0.7. events is expected to be small, assuming moderate progenitor The upper limit arises from waveform limitations, although metallicities and black-hole natal kicks (Silsbee & Tremaine even a model allowing higher eccentricities would be re- 2017; Antonini et al. 2017; Rodriguez & Antonini 2018; Fra- stricted by the reweighting method. In order to reweight gione & Kocsis 2019a; Liu et al. 2019). In contrast, some posterior samples efficiently, the samples obtained using the dynamically-formed binaries merge so rapidly after becom- proposal model must cover the same region of the multidi- ing bound that they retain non-negligible eccentricity in the mensional parameter space as would be obtained by direct LIGO–Virgo band (Zevin et al. 2017; Rodriguez et al. 2018b; sampling with our target model. The overlap between eccen- Samsing et al. 2018; Gondán & Kocsis 2019; Zevin et al. tric and quasi-circular waveforms with otherwise-identical 2019b). Multiple authors (e.g., Samsing 2018; Samsing & parameters falls drastically for higher eccentricities, so their Ramirez-Ruiz 2017; Samsing & D’Orazio 2018; Rodriguez posterior samples would not reside in the same region of the et al. 2018a,b) show that we can expect O(5%) of all dynami- parameter space. We marginalise over the time and phase of cal mergers in globular clusters to have eccentricities 푒 > 0.1 coalescence as in Payne et al.(2019) to account for differ- at a gravitational-wave frequency of 10 Hz. ing definitions of these parameters between our proposal and The LIGO-Virgo Collaboration recently announced the target models. detection of GW190521, a gravitational-wave signal from the merger of a black hole binary with component masses ANALYSIS OF GW190521 +21 +17 푚1 = 85−14 M , 푚2 = 66−18 M (Abbott et al. 2020b,c). We analyze publicly-available data and noise power spectral The median and 90% confidence intervals quoted for these densities from Abbott et al.(2020b,a). We reproduce the masses place at least one component within the PISN mass

1 Fishbach & Holz(2020) find that 푚1 is above the mass gap if 푚2 < 48 M , below the mass gap. 3

GW190521.1 ) d | ) 0 ) 1 d e | ( p 0 1 ( g p o l ( p

6 5 4 3 2 1 0.2 0.4 0.6 0.8 log10(e10) p

Figure 1. Results of analysis of GW190521 using SEOBNRE and IMRPhenomPv2. Left: posterior probability density distribution for eccentricity at 10 Hz for GW190521, recovered using SEOBNRE. At 90% confidence, 푒10 ≥ 0.11. The posterior rails against the upper limit of the prior, 푒10 = 0.2, suggesting that the true value lies beyond this waveform-enforced constraint. Right: posterior probability density distribution for the precession parameter 휒푝 for GW190521, recovered using IMRPhenomPv2. The prior probability for each parameter is shown in gray.

Table 1. Recovered GW190521 parameter values obtained using eccentric waveform model SEOBNRE, precessing waveform models IMRPhenomPv2 and NRSur7dq4, and NRSur7dq4 constrained to have aligned spins. For the SEOBNRE analysis, we give the 90% confidence lower limit on eccentricity at 10 Hz. For other parameters, the median of the posterior is given along with the 90% credible interval. In the final column, we state the values inferred from the LIGO–Virgo analysis, read from the public posterior samples obtained using NRSur7dq4 (Abbott et al. 2020a). In the final row, we provide the log Bayes factor of each analysis against the signal-to-noise log Bayes factor obtained for 푒10 ≥ 0.1 using SEOBNRE (ln B푆/푁 = 85.7). Parameter (source frame) SEOBNRE IMRPhenomPv2 NRSur7dq4 NRSur7dq4 aligned NRSur7dq4 LVC +26 +61 +18 +22 +21 Primary mass, 푚1 [M ] 92−16 126−41 86−13 85−14 85−14 +18 +32 +18 +15 +17 Secondary mass, 푚2 [M ] 69−19 59−24 69−17 61−17 66−18 +1.8 +2.3 +2.2 +1.6 +2.4 Luminosity distance, 푑L [Gpc] 4.1−1.8 2.4−1.0 4.7−2.2 4.7−1.5 5.3−2.6 +2.7 +1.9 +2.9 +2.6 +2.8 Right ascension, 훼 [rad] 3.6−3.5 4.3−4.3 3.4−3.4 3.7−3.7 3.5−3.4 − +1.4 − +1.5 − +1.5 − +1.6 − +1.5 Declination, 훿 [rad] 0.7−0.5 0.7−0.4 0.8−0.4 0.9−0.3 0.8−0.4 +2.9 +3.0 +2.6 +2.9 +2.6 Reference phase, 휙 [rad] 3.1−2.7 3.0−2.7 3.2−2.6 3.1−2.8 3.4−3.2 +1.5 +1.3 +1.2 +1.4 +1.2 Polarisation, 휓 [rad] 1.5−1.4 1.6−1.5 1.8−1.5 1.6−1.4 1.8−1.6 +1.6 +1.0 +2.0 +2.2 +2.1 Inclination, 휃JN [rad] 1.3−1.0 1.4−0.7 0.8−0.6 0.7−0.5 0.8−0.6 min Eccentricity lower limit at 10 Hz, 푒10 0.11 N/A N/A N/A N/A +0.2 +0.4 +0.3 +0.2 +0.3 Effective spin, 휒eff 0.0−0.2 0.1−0.4 0.0−0.3 0.0−0.3 0.1−0.4 +0.2 +0.2 +0.3 Effective precession, 휒푝 N/A 0.7−0.3 0.6−0.3 N/A 0.7−0.4 Log Bayes factor against SEOBNRE, ln B푋/퐸 0.0 −2.0 −1.8 −5.0 −1.2

max Table 2. The 90% credible upper limit on eccentricity at 10 Hz, 푒10 , and recovered precession parameter 휒푝 for different injections with varying waveform model, 푒10 and 휒푝 settings. For the recovered 휒푝 we quote the posterior median and 90% credible interval. max Injected waveform model Injected 푒10 Injected 휒푝 Recovered 푒10 with SEOBNRE Recovered 휒푝 with IMRPhenomPv2 +0.37 IMRPhenomD 0 0 0.025 0.39−0.29 +0.40 NRSur7dq4 0 0 0.032 0.33−0.25 +0.36 SEOBNRE 0 0 0.055 0.42−0.30 +0.35 IMRPhenomPv2 0 0.63 0.077 0.43−0.32 +0.29 NRSur7dq4 0 0.63 0.118 0.53−0.37 +0.26 SEOBNRE 0.13 0 0.136 0.57−0.39 4

Table 3. Parameters shared by all injected waveforms. implement an astrophysically-motivated prior on eccentric- Parameter (source frame) Value ity. Theoretical studies robustly predict that ∼ 5% of bi- naries that form dynamically in globular clusters will have Primary mass, 푚 [M ] 84 1 푒 ≥ 0.1 (e.g., Samsing 2018; Samsing et al. 2018; Ro- Secondary mass, 푚 [M ] 62 10 2 driguez et al. 2018b; Kremer et al. 2020). To investigate Luminosity distance, 푑 [Gpc] 5.0 L this hypothesis, we assume a log-uniform distribution for Right ascension, 훼 [rad] 3.3 log10 푒10 ∈ (−1, −0.7). Using this astrophysically-motivated Declination, 훿 [rad] 0.5 prior, the eccentric model is mildly preferred to the pre- Reference phase, 휙 [rad] 6.2 cessing model by a factor of ln B퐸/푃 = 2.0. If we re- Polarisation, 휓 [rad] 1.6 peat the same calculation using the (less well-motivated) Inclination, 휃JN [rad] 0.3 prior range log10 푒10 ∈ (−6, −0.7) as in Fig.1, the ec- Geocent time, 푡0 [s] 1242442967.46 centric (E) and precessing (P) waveform models are almost equally well-supported by the data, with a log Bayes factor settings of the LVC analysis for our parameter estimation, ln B퐸/푃 = −0.35. with a data segment of 8 s, a frequency band 11–512 Hz, and Finally, we perform computationally-intensive analyses sampling frequency 1024 Hz. In order to assess the role of using the precessing, higher-order-model waveform NR- waveform systematics, we perform four analyses using three Sur7dq4, using parallel Bilby (Smith et al. 2019) to manage different waveform models (one waveform is used twice with computational costs. We run two versions of the NRSur7dq4 two different spin priors). The results of these analyses are analysis: one assuming aligned black-hole spins (no preces- summarized in Table1. sion) and one allowing arbitrary spin orientations (allowing First, we analyze the data using the aligned-spin eccen- precession). Otherwise, the assumptions are identical to the tric waveform model SEOBNRE. We present the posterior IMRPhenomPv2 analysis above. While the two NRSur7dq4 analyses obtain near-identical results, the analysis that in- distribution on the 푒10 of GW190521 in the left-hand panel of Fig.1. The posterior drastically deviates from the log- cludes precession (P) is preferred over the no-precession hy- pothesis with a moderate ln B = 3.2. The eccentric uniform prior, strongly favouring eccentricities 푒10 ≥ 0.1. 푃/푁 푃 SEOBNRE hypothesis (with 푒 > 0.1) is preferred to the There is little support for 푒10 < 0.1, with 90% of the poste- 10 precessing and non-precessing NRSur7dq4 analyses by log rior at 푒10 ≥ 0.11. For other parameters, we obtain median posterior values similar to those given in Table 1 of Ab- Bayes factors of ln B퐸/푃 = 1.8 and ln B퐸/푁 푃 = 5.0, respec- bott et al.(2020b), with a median source-frame total mass tively. +28 +0.2 +0.2 We perform two additional analyses, identical in almost 푀 = 161−20 M , mass ratio 푞 = 0.7−0.3, and 휒eff = 0.0−0.2. +1.9 all aspects to the NRSur7dq4 studies described above, but We obtain a luminosity distance, 푑L = 4.0−1.7 Gpc, which is slightly lower than (but consistent with) the value of without including higher-order modes. If we assume aligned +2.4 spin, we obtain results similar to the SEOBNRE analysis. 5.3−2.6 Gpc from the LIGO–Virgo analysis. Eccentricity causes a faster merger, reducing the signal power. Thus, If we allow for precession, we obtain results similar to the +2.2 in order to match the observed signal-to-noise ratio with an IMRPhenomPv2 results with luminosity distance 2.8−1.5 Gpc eccentric template, we may require a closer source. Addition- (90% credibility) and 푞 ≈ 0.5. ally, models like SEOBNRE, IMRPhenomD and IMRPhe- nomPv2, which do not contain higher-order modes, cannot INJECTION STUDIES rule out edge-on binaries, which reduces the median distance Ideally, one would analyze gravitational-wave signals us- estimate (Abbott et al. 2020c). Posterior distribution plots for ing models that include both precession and eccentricity. This all other parameters are available online.2 would allow simultaneous measurements of 휒푝 and 푒10, as Next, we perform an analysis using the precessing wave- well as illuminating the full extent of the degeneracy be- form IMRPhenomPv2 (Schmidt et al. 2012) with otherwise- tween the two parameters and how that degeneracy changes identical settings. In Fig.1, we show the posterior distri- with mass. Unfortunately, such models do not yet exist; bution for 휒푝 of GW190521 obtained with IMRPhenomPv2. see Healy et al.(2009); Levin et al.(2011) for a theoreti- This analysis recovers a smaller median 푑L than the SEOB- cal background of eccentric and precessing binary dynamics ≈ NRE analysis, with a more extreme mass ratio, 푞 0.5. and waveforms. Thus, we use numerical tests to explore how In order to carry out model selection comparing the IMR- our limited waveform models affect what we infer about ec- PhenomPv2 results to those obtained with SEOBNRE, we centricity and precession. We generate six GW190521-like waveform templates using different waveform models, each 2 git..org/isobel.romero-shaw/gw190521.1 with different values of 푒10 and 휒푝; see Table2. Other pa- rameters are identical to those in Table3. Using Bilby, we 5

e10 = 0, p = 0, SEOBNRE )

d , NRSur7dq2 | e10 = 0, p = 0 ) 0 )

1 , IMRPhenomD e10 = 0, p = 0 d e | ( p 0 1 ( g p o l ( p

e10 = 0, p = 0.63, NRSur7dq2 )

d , IMRPhenomPv2 | e10 = 0, p = 0.63 ) 0 ) 1 d e | ( p 0 1 ( g p o l ( p

e10 = 0.13, p = 0, SEOBNRE ) d | ) 0 ) 1 d e | ( p 0 1 ( g p o l ( p

6 5 4 3 2 1 0.2 0.4 0.6 0.8 log10(e10) p

Figure 2. Results of SEOBNRE and IMRPhenomPv2 analysis of simulated data using GW190521-like injections. Left: Posterior distributions for eccentricity at 10 Hz for GW190521-like injection studies with varying 푒10 and 휒푝, obtained using SEOBNRE. Right: Posterior distribution for precession parameter 휒푝 for GW190521-like injection studies with varying 푒10 and 휒푝, recovered using IMRPhenomPv2. The prior distributions are shown in gray. inject these waveforms into simulated detector networks con- ries, precession may be mistaken for eccentricity, and that the sisting of LIGO Hanford, LIGO Livingston, and Virgo, with imprint of eccentricity may be mistaken for that of precession. noise power spectral densities matching those used for anal- We provide the full posterior distributions for all parameters ysis of GW190521 (Abbott et al. 2020b). For each injection, in these injection studies online.3 we recover the signal using both the aligned-spin eccentric model SEOBNRE and the quasi-circular precessing model IMRPhenomPv2. DISCUSSION In Fig.2, we compare the posterior distributions for 푒10 (ob- Assuming the aligned-spin SEOBNRE waveform model, tained using SEOBNRE) and 휒푝 (obtained using IMRPhe- we infer an eccentricity 푒10 & 0.1 for GW190521. We nomPv2) for all injections. When circular, non-precessing find that the SEOBNRE waveform is slightly preferred over waveforms are injected, the SEOBNRE analysis recovers pos- the circular-waveform models NRSur7dq4 and IMRPhe- terior distributions for 푒10 consistent with the prior below nomPv2, both of which allow for precession. While we lack max ≤ the 90% credible upper limit, 푒10 0.025 (0.032, 0.055) a waveform model that can simultaneously account for pre- for injected IMRPhenomD (NRSur7dq4, SEOBNRE) wave- cession and eccentricity, GW190521 could be later verified forms. For these same waveforms, IMRPhenomPv2 analysis as the first detection of a binary black hole with 푒10 ≥ 0.1. recovers posteriors consistent with the prior on 휒푝. When we The presence of either precession or eccentricity adds weight increase only 휒푝, the posteriors on both 휒푝 and 푒10 skew away to the hypothesis that the progenitor of GW190521 formed from their priors. This is most notable for the NRSur7dq4 dynamically. injections, suggesting that higher-order modes (included in Samsing(2018) predicts there are ∼ 19 dynamical mergers NRSur7dq4, but not in IMRPhenomPv2) may be important with 푒10 < 0.1 for every merger with 푒10 ≥ 0.1—a predic- for distinguishing precession and eccentricity. When we in- tion thought to be robust to details about the globular cluster crease 푒10, both posteriors deviate from their priors, more model; see also Rodriguez et al.(2018b) and Martinez et al. significantly than for the increased-precession case. These injection studies demonstrate that, for GW190521-like bina- 3 git.ligo.org/isobel.romero-shaw/gw190521.1/injection_studies 6

(2020). From the public alerts listed on GraceDB4, there are For GW190521-like signals, we highlight the degeneracy O(30) binary black hole mergers from the first half of LIGO– between eccentricity and precession5. This complements Virgo’s third observing run (O3a). Combining these with the results of Bustillo et al.(2020), who found that for the the results of Abbott et al.(2019a) and Venumadhav et al. gravitational-wave signal of a head-on black-hole collision (2019); Zackay et al.(2019); Venumadhav et al.(2020), the (푒10 = 1) with total mass in the range 푀 ∈ (130, 300) M total number of binary black holes observed in gravitational can be indistinguishable from the signal of a much more dis- waves is O(50). If globular cluster mergers dominate LIGO tant quasi-circular precessing binary. Recently, a candidate +2.0 and Virgo’s observed black hole mergers, we expect 2.5−2.5 electromagnetic counterpart for GW190521 was observed at mergers with 푒10 ≥ 0.1 from the first 50 binary black hole ≈ 2.8 Gpc and reported by Graham et al.(2020), who pro- observations. Thus, it would not be surprising if GW190521 pose that a binary black hole merger in an AGN disk might is determined to be highly eccentric. Moreover, if GW190521 have such a counterpart. Extrapolating between the 푒10 = 1 +2.0 is eccentric, then O3a may provide us with another 1.5−1.5 results from Bustillo et al.(2020) and the results shown here, events with 푒10 ≥ 0.1, assuming that O3a searches did not the detected distance of GW190521 in gravitational waves is miss them; the signals of highly eccentric binaries may be consistent with the electromagnetic counterpart if GW190521 missed by CBC and burst searches (East et al. 2013). had an eccentricity in the range 0.2 < 푒10 < 1.0, a region of We note that while GW190521 may have formed within a parameter space that cannot be fully explored with existing globular cluster, this is not its only viable formation pathway. gravitational waveform models. However, new developments Dynamical formation may also occur in active galactic nu- in eccentric waveforms (see, e.g., Chiaramello & Nagar 2020) clei (e.g., Kocsis & Levin 2012; Yang et al. 2019; Gröbner may allow us to start probing previously unexplored param- et al. 2020), nuclear star clusters (e.g., Mapelli et al. 2020), eter space in the near future. If the transient reported by young open clusters (e.g., Di Carlo et al. 2019b) and young Graham et al.(2020) is truly an electromagnetic counterpart massive clusters (Banerjee 2017, 2018a,b, 2020; Kremer et al. emanating from an AGN disk merger, it would be consistent 2020). Mergers in young star clusters are likely to take place with the hypothesis that GW190521 was an eccentric binary, after ejection, giving the binary ample time to circularise and since orbital eccentricity vastly increases the merger rate from making young star clusters a less promising explanation for such environments; see Gröbner et al.(2020). eccentric binaries. Both active galactic nuclei and globular Note added.—During the final stages of preparation of this clusters may produce binary black holes with misaligned spin manuscript, we became aware of the work of Gayathri et al. and/or eccentricity, and so it is not clear which dynamical (2020), who compare numerical-relativity waveform simula- formation pathway is favoured for GW190521. Regardless, tions to GW190521. waveforms are too both precession and eccentricity are signatures of dynamical computationally expensive to be used for Bayesian parameter formation; therefore, GW190521 is likely to have formed in estimation. However, the fact that Gayathri et al.(2020) find a dense stellar environment conducive to dynamical interac- that eccentric numerical-relativity simulations are consistent tions. with GW190521 supports the conclusions drawn in our work. In dense environments like those mentioned above, binary black hole merger remnants may have masses within the mass ACKNOWLEDGEMENTS gap. If these mergers are retained within the cluster, then they We thank the anonymous referee for their thoughtful may merge again, producing intermediate-mass black holes. suggestions, which improved the manuscript. We thank As an alternative to hierarchical black hole mergers, Roupas Colm Talbot, Max Isi, Alan Weinstein, Tito Dal Canton, & Kazanas(2019) argue that black holes may accrete enough Christopher Berry and Chase Kimball for fruitful sugges- gas in proto-clusters to enter into the mass gap. Another op- tions and illuminating discussions. We thank Rory Smith tion is the direct collapse of stellar merger remnants to mass- for assistance with parallel Bilby. This work is supported gap black holes (e.g., Spera et al. 2019; Kremer et al. 2020). through Australian Research Council (ARC) Future Fellow- These black holes may undergo subsequent dynamical merg- ships FT150100281, FT160100112, Centre of Excellence ers if their environments are sufficiently densely populated. CE170100004, and Discovery Project DP180103155. JCB Although the high masses of GW190521 render it incompat- acknowledges support from the Direct Grant of the CUHK ible with current models of isolated binary evolution, these Research Committee, Project ID: 4053406 Computing was masses can be produced in models where various model as- performed on the OzSTAR Australian national facility at sumptions are substantially relaxed (see, e.g., Stevenson et al. Swinburne University of Technology, which receives funding 2019; Farmer et al. 2019; Marchant & Moriya 2020). in part from the Astronomy National Collaborative Research

4 gracedb.ligo.org/superevents/public/O3/ 5 The degeneracy between eccentricity and precession is less pronounced for less massive systems, which have longer signals in-band. 7

Infrastructure Strategy (NCRIS) allocation provided by the ogy with funding from the National Science Foundation and Australian Government, and the LIGO Laboratory comput- operates under cooperative agreement PHY-1764464. Virgo ing cluster at California Institute of Technology, supported is funded by the French Centre National de Recherche Sci- by National Science Foundation Grants PHY-0757058 and entifique (CNRS), the Italian Istituto Nazionale della Fisica PHY-0823459. LIGO was constructed by the California In- Nucleare (INFN) and the Dutch Nikhef, with contributions stitute of Technology and Massachusetts Institute of Technol- by Polish and Hungarian institutes.

REFERENCES

Abbott, et al. 2020a, Gravitational Wave Open Science Center Chiaramello, D., & Nagar, A. 2020, Phys. Rev. D, 101, 101501, Strain Data Release for GW190521, LIGO Open Science Center. doi: 10.1103/PhysRevD.101.101501 www.gw-openscience.org/eventapi/html/O3_Discovery_Papers/ de Mink, S. E., & Mandel, I. 2016, Mon. Not. Roy. Astron. Soc., GW190521/v2/ 460, 3545, doi: 10.1093/mnras/stw1219 Abbott, B. P., et al. 2016a, Phys. Rev. Lett., 116, 061102, de Mink, S. E., et al. 2010, in American Institute of Physics doi: 10.1103/PhysRevLett.116.061102 Conference Series, Vol. 1314, American Institute of Physics —. 2016b, Phys. Rev. X, 6, 041015, Conference Series, ed. V. Kalogera & M. van der Sluys, 291–296 doi: 10.1103/PhysRevX.6.041015 Di Carlo, U. N., Giacobbo, N., Mapelli, M., et al. 2019a, MNRAS, Abbott, B. P., et al. 2018, Living Rev. Rel., 21, 3, 487, 2947, doi: 10.1093/mnras/stz1453 doi: 10.1007/s41114-018-0012-9,10.1007/lrr-2016-1 —. 2019b, MNRAS, 487, 2947, doi: 10.1093/mnras/stz1453 —. 2019a, Phys. Rev. X, 9, 031040, East, W. E., McWilliams, S. T., Levin, J., & Pretorius, F. 2013, doi: 10.1103/PhysRevX.9.031040 Phys. Rev. D, 87, 043004, doi: 10.1103/PhysRevD.87.043004 —. 2019b, arXiv e-prints, arXiv:1907.09384. Farmer, R., Renzo, M., de Mink, S. E., Marchant, P., & Justham, S. https://arxiv.org/abs/1907.09384 2019, ApJ, 887, 53, doi: 10.3847/1538-4357/ab518b Abbott, R., Abbott, T. D., Abraham, S., et al. 2020b, Phys. Rev. Farr, W. M., et al. 2017, Nature, 548, 426, Lett., 125, 101102, doi: 10.1103/PhysRevLett.125.101102 doi: 10.1038/nature23453 —. 2020c, doi: 10.3847/2041-8213/aba493 Fishbach, M., & Holz, D. E. 2017, Astrophys. J., 851, L25, Acernese, F., et al. 2015, Class. Quant. Grav., 32, 024001, doi: 10.3847/2041-8213/aa9bf6 doi: 10.1088/0264-9381/32/2/024001 Fishbach, M., & Holz, D. E. 2020, arXiv e-prints, Antonini, F., et al. 2017, Astrophys. J., 841, 77, arXiv:2009.05472. https://arxiv.org/abs/2009.05472 doi: 10.3847/1538-4357/aa6f5e Fishbach, M., et al. 2017, Astrophys. J., 840, L24, Ashton, G., et al. 2019, Astrophys. J. Suppl., 241, 27, doi: 10.3847/2041-8213/aa7045 doi: 10.3847/1538-4365/ab06fc Fragione, G., & Bromberg, O. 2019, arXiv e-prints, Banerjee, S. 2017, MNRAS, 467, 524, doi: 10.1093/mnras/stw3392 arXiv:1903.09659. https://arxiv.org/abs/1903.09659 —. 2018a, MNRAS, 473, 909, doi: 10.1093/mnras/stx2347 Fragione, G., & Kocsis, B. 2018, Phys. Rev. Lett., 121, —. 2018b, MNRAS, 481, 5123, doi: 10.1093/mnras/sty2608 doi: 10.1103/PhysRevLett.121.161103 —. 2020, MNRAS, doi: 10.1093/mnras/staa2392 Fragione, G., & Kocsis, B. 2019a, arXiv e-prints, Belczynski, K. 2020, arXiv e-prints, arXiv:2009.13526. arXiv:1910.00407. https://arxiv.org/abs/1910.00407 https://arxiv.org/abs/2009.13526 —. 2019b, MNRAS, 486, 4781, doi: 10.1093/mnras/stz1175 Belczynski, K., Heger, A., Gladysz, W., et al. 2016, A&A, 594, Fragione, G., Loeb, A., & Rasio, F. A. 2020, arXiv e-prints, A97, doi: 10.1051/0004-6361/201628980 arXiv:2009.05065. https://arxiv.org/abs/2009.05065 Bethe, H. A., & Brown, G. E. 1998, Astrophys. J., 506, 780, Fragione, G., et al. 2020. https://arxiv.org/abs/2002.11278 doi: 10.1086/306265 Gayathri, V., Healy, J., Lange, J., et al. 2020, arXiv e-prints, Bouffanais, Y., et al. 2019. https://arxiv.org/abs/1905.11054 arXiv:2009.05461. https://arxiv.org/abs/2009.05461 Bustillo, J. C., Sanchis-Gual, N., Torres-Forné, A., & Font, J. A. Gerosa, D., & Berti, E. 2017, Phys. Rev., D95, 124046, 2020, Confusing head-on and precessing intermediate-mass doi: 10.1103/PhysRevD.95.124046 binary black hole mergers Gondán, L., & Kocsis, B. 2019, Astrophys. J., 871, 178, Cao, Z., & Han, W.-B. 2017, Phys. Rev., D96, 044028, doi: 10.3847/1538-4357/aaf893 doi: 10.1103/PhysRevD.96.044028 Gondán, L., et al. 2018, Astrophys. J., 860, 5, Chatziioannou, K., et al. 2019, Phys. Rev. D, 100, 104015 doi: 10.3847/1538-4357/aabfee 8

Graham, M. J., Ford, K. E. S., McKernan, B., et al. 2020, Morscher, M., et al. 2015, ApJ, 800, 9, Phys. Rev. Lett., 124, 251102, doi: 10.1088/0004-637X/800/1/9 doi: 10.1103/PhysRevLett.124.251102 O’Leary, R. M., et al. 2006, Astrophys. J., 637, 937, Gröbner, M., Ishibashi, W., Tiwari, S., Haney, M., & Jetzer, P. doi: 10.1086/498446 2020, A&A, 638, A119, doi: 10.1051/0004-6361/202037681 Özel, F., Psaltis, D., Narayan, R., & McClintock, J. E. 2010, ApJ, Healy, J., Levin, J., & Shoemaker, D. 2009, Phys. Rev. Lett., 103, 725, 1918, doi: 10.1088/0004-637X/725/2/1918 131101, doi: 10.1103/PhysRevLett.103.131101 Payne, E., Talbot, C., & Thrane, E. 2019, Phys. Rev. D, 100, Heger, A., & Woosley, S. E. 2002, Astrophy. J., 567, 532, 123017, doi: 10.1103/PhysRevD.100.123017 doi: 10.1086/338487 Peters, P. C. 1964, Phys. Rev., 136, B1224, Hinder, I., et al. 2008, Phys. Rev. D, D77, 081502, doi: 10.1103/PhysRev.136.B1224 doi: 10.1103/PhysRevD.77.081502 Portegies Zwart, S. F., & McMillan, S. 2000, Astrophys. J., 528, Ivanova, N., et al. 2013, A&A Rev., 21, 59, L17, doi: 10.1086/312422 doi: 10.1007/s00159-013-0059-2 Randall, L., & Xianyu, Z.-Z. 2018a, Astrophys. J., 853, Khan, S., et al. 2016, Phys. Rev., D93, 044007, doi: 10.3847/1538-4357/aaa1a2 doi: 10.1103/PhysRevD.93.044007 —. 2018b, Astrophys. J., 864, 134, doi: 10.3847/1538-4357/aad7fe Kimball, C., Berry, C., & Kalogera, V. 2019, RNAAS, 4, 2 Rastello, S., Mapelli, M., Di Carlo, U. N., et al. 2020, MNRAS, Kimball, C., et al. 2020 doi: 10.1093/mnras/staa2018 Kocsis, B., & Levin, J. 2012, Phys. Rev. D, 85, 123005, Rodriguez, C. L., & Antonini, F. 2018, Astrophys. J., 863, 7, doi: 10.3847/1538-4357/aacea4 doi: 10.1103/PhysRevD.85.123005 Rodriguez, C. L., et al. 2016, Astrophys. J., 832, L2, Kozai, Y. 1962, Astrophys. J., 67, 591, doi: 10.1086/108790 doi: 10.3847/2041-8205/832/1/L2 Kremer, K., Spera, M., Becker, D., et al. 2020, arXiv e-prints, —. 2018a, Phys. Rev., D98, 123005, arXiv:2006.10771. https://arxiv.org/abs/2006.10771 doi: 10.1103/PhysRevD.98.123005 Kruckow, M. U., et al. 2016, Astron. Astrophys., 596, A58, —. 2018b, Phys. Rev. Lett., 120, 151101, doi: 10.1051/0004-6361/201629420 doi: 10.1103/PhysRevLett.120.151101 Levin, J., McWilliams, S. T., & Contreras, H. 2011, Classical and —. 2019. https://arxiv.org/abs/1906.10260 Quantum Gravity, 28, 175001, Romero-Shaw, I. M., Lasky, P. D., & Thrane, E. 2019, MNRAS, doi: 10.1088/0264-9381/28/17/175001 490, 5210, doi: 10.1093/mnras/stz2996 Lidov, M. L. 1962, Planetary and Space Science, 9, 719, Romero-Shaw, I. M., Talbot, C., Biscoveanu, S., et al. 2020, arXiv doi: 10.1016/0032-0633(62)90129-0 e-prints, arXiv:2006.00714. https://arxiv.org/abs/2006.00714 Liu, B., & Lai, D. 2019, MNRAS, 483, 4060, Roupas, Z., & Kazanas, D. 2019, A&A, 632, L8, doi: 10.1093/mnras/sty3432 doi: 10.1051/0004-6361/201937002 Liu, B., et al. 2019. https://arxiv.org/abs/1905.00427 Sakstein, J., Croon, D., McDermott, S. D., Straight, M. C., & Liu, X., et al. 2019, arXiv e-prints, arXiv:1910.00784. Baxter, E. J. 2020, arXiv e-prints, arXiv:2009.01213. https://arxiv.org/abs/1910.00784 https://arxiv.org/abs/2009.01213 Livio, M., & Soker, N. 1988, Astrophys. J., 329, 764, Samsing, J. 2018, Phys. Rev. D, D97, 103014, doi: 10.1086/166419 doi: 10.1103/PhysRevD.97.103014 Lower, M., et al. 2018, Phys. Rev. D, 98, Samsing, J., & D’Orazio, D. J. 2018, Mon. Not. Roy. Astron. Soc., doi: 10.1103/PhysRevD.98.083028 481, doi: 10.1093/mnras/sty2334 Mapelli, M., Santoliquido, F., Bouffanais, Y., et al. 2020, arXiv Samsing, J., & Ramirez-Ruiz, E. 2017, Astrophys. J., 840, L14, e-prints, arXiv:2007.15022. https://arxiv.org/abs/2007.15022 doi: 10.3847/2041-8213/aa6f0b Marchant, P., Langer, N., Podsiadlowski, P., Tauris, T. M., & Samsing, J., et al. 2014, Astrophys. J., 784, 71, Moriya, T. J. 2016, A&A, 588, A50, doi: 10.1088/0004-637X/784/1/71 doi: 10.1051/0004-6361/201628133 —. 2018, arXiv e-prints, arXiv:1802.08654. Marchant, P., & Moriya, T. 2020, arXiv e-prints, https://arxiv.org/abs/1802.08654 arXiv:2007.06220. https://arxiv.org/abs/2007.06220 Schmidt, P., et al. 2012, Phys. Rev. D, 86, 104063 Martinez, M. A. S., Fragione, G., Kremer, K., et al. 2020, arXiv Sigurdsson, S., & Hernquist, L. 1993, Nature, 364, 423, e-prints, arXiv:2009.08468. https://arxiv.org/abs/2009.08468 doi: 10.1038/364423a0 McKernan, B., Ford, K. E. S., & O’Shaughnessy, R. 2020, arXiv Silsbee, K., & Tremaine, S. 2017, Astrophys. J., 836, 39, e-prints, arXiv:2002.00046. https://arxiv.org/abs/2002.00046 doi: 10.3847/1538-4357/aa5729 9

Smith, R., Ashton, G., Vajpeyi, A., & Talbot, C. 2019, arXiv Woosley, S. E. 2017, ApJ, 836, 244, e-prints, arXiv:1909.11873. https://arxiv.org/abs/1909.11873 doi: 10.3847/1538-4357/836/2/244 Spera, M., Mapelli, M., Giacobbo, N., et al. 2019, MNRAS, 485, Yang, Y., Bartos, I., Haiman, Z., et al. 2019, ApJ, 876, 122, 889, doi: 10.1093/mnras/stz359 doi: 10.3847/1538-4357/ab16e3 Stevenson, S., Sampson, M., Powell, J., et al. 2019, ApJ, 882, 121, Yang, Y., Bartos, I., Gayathri, V., et al. 2019, Phys. Rev. Lett., 123, doi: 10.3847/1538-4357/ab3981 181101, doi: 10.1103/PhysRevLett.123.181101 Stevenson, S., et al. 2015, Astrophys. J., 810, 58, Zackay, B., Dai, L., Venumadhav, T., Roulet, J., & Zaldarriaga, M. doi: 10.1088/0004-637X/810/1/58 2019, arXiv e-prints, arXiv:1910.09528. Talbot, C., & Thrane, E. 2017, Phys. Rev. D, 96, 023012 https://arxiv.org/abs/1910.09528 Venumadhav, T., Zackay, B., Roulet, J., Dai, L., & Zaldarriaga, M. Zevin, M., et al. 2017, Astrophys. J., 846, 82, 2020, Phys. Rev. D, 101, 083030, doi: 10.3847/1538-4357/aa8408 doi: 10.1103/PhysRevD.101.083030 —. 2019a, Astrophys. J., 871, 91, doi: 10.3847/1538-4357/aaf6ec Venumadhav, T., et al. 2019, Phys. Rev. D, 100, 023011, —. 2019b. https://arxiv.org/abs/1906.11299 doi: 10.1103/PhysRevD.100.023011 Ziosi, B. M., Mapelli, M., Branchesi, M., & Tormen, G. 2014, Vitale, S., et al. 2017, Class. Quant. Grav., 34, 03LT01, Monthly Notices of the Royal Astronomical Society, 441, 3703 doi: 10.1088/1361-6382/aa552e