<<

Observation of a multimode quasi-normal spectrum from a perturbed Collin D. Capano*1,2, Miriam Cabero3, Julian Westerweck1,2, Jahed Abedi1,2, Shilpa Kastha1,2, Alexander H. Nitz1,2, Alex B. Nielsen4, and Badri Krishnan1,2,5

1Albert-Einstein-Institut, Max-Planck-Institut fur¨ Gravitationsphysik, Callinstraße 38, 30167 Hannover, Germany 2Leibniz Universitat¨ Hannover, 30167 Hannover, Germany 3Department of Physics and Astronomy, The University of British Columbia, Vancouver, BC V6T 1Z4, Canada 4Department of Mathematics and Physics, University of Stavanger, NO-4036 Stavanger, Norway 5Institute for Mathematics, Astrophysics and Particle Physics, Radboud University, Heyendaalseweg 135, 6525 AJ Nijmegen, The Netherlands

ABSTRACT

When two black holes merge, the late stage of emission is a superposition of exponentially damped sinusoids. According to the black hole no-hair theorem, this ringdown spectrum depends only on the mass and angular momentum of the final black hole. An observation of more than one ringdown mode can test this fundamental prediction of . Here we provide strong observational evidence for a multimode black hole ringdown spectrum using the gravitational wave event GW190521, with a Bayes factor of ∼ 40 preferring two fundamental modes over one. The dominant mode is the ` = m = 2 harmonic, and the sub-dominant mode corresponds to the ` = m = 3 harmonic. We estimate the redshifted mass and dimensionless spin of +30 +0.05 the final black hole as 330−40 M and 0.87−0.10, respectively. The detection of the two modes disfavors a binary progenitor with +0.2 equal masses; the mass ratio is constrained to 0.4−0.3. We find that the final black hole is consistent with the no hair theorem and +0.07 constrain the fractional deviation from general relativity of the sub-dominant mode’s frequency to be −0.01−0.11.

1 Introduction holes which merge at a low frequency relative to the detector sensitivity band. As such, it has a barely observable inspiral A perturbed black hole approaches equilibrium by emitting a and the signal is dominated by the merger and ringdown phase. 1–3 spectrum of damped sinusoidal gravitational-wave signals. GW190521 was initially reported as the merger of two Unlike other astrophysical objects, the ringdown spectrum of comparable mass black holes,9, 10 in which case one would a black hole is remarkably simple. General relativity predicts not expect to detect sub-dominant ringdown modes. Subse- that the frequencies and damping times of the entire spectrum quent re-analysis of the data suggested the possibility that of damped sinusoids, or “quasi-normal modes”, are fully de- the progenitor could be an intermediate mass-ratio binary,13 termined by just two numbers: the black hole mass M and 4 suggesting the possibility of detectable sub-dominant modes. angular momentum J, as described by the Kerr solution. This Here we find strong evidence for multimode damped sinu- prediction, a consequence of the black hole “no-hair theorem”, 5 soids in the ringdown phase of the gravitational wave event does not hold in many alternate theories. If astrophysical GW190521. black holes are observed to violate this property, it indicates new physics beyond standard general relativity. In order to observationally test this prediction using binary 2 Multimode agnostic search black hole mergers, an important observational challenge must arXiv:2105.05238v2 [gr-qc] 26 May 2021 A quasi-normal mode description of the gravitational wave 6 be met: at least two ringdown modes must be observed. The from a is not expected to be valid until higher the binary mass ratio asymmetry, the more likely it after the binary has merged to form a perturbed black hole. is that sub-dominant ringdown modes are observable. How- On the flip side, the damping time of an O(100M ) black ever, more asymmetric binary systems are less likely to be hole is O(10ms), leaving a window of only a few tens of formed, and also lead to weaker signals. Population studies milliseconds after merger in which the ringdown is detectable suggested that such multimode ringdown modes were unlikely above noise. Accurate identification of the merger time is to be observed until the next generation of gravitational-wave therefore crucial to extract quasi-normal modes from the data. 7,8 observatories, since black-hole population models did not To account for uncertainty in the merger time of GW190521 anticipate observations of massive, asymmetric binaries. due to modelling systematics, we perform a series of analyses Here we confound this expectation with the gravitational- in short time increments starting at a geocentric GPS reference wave event GW190521, detected by the two LIGO detectors time tref = 1242442967.445. This time is taken from the and Virgo at 03:02:29 UTC on May 21st, 2019.9, 10 This is the maximum likelihood merger time obtained via the analysis in heaviest black-hole merger event observed to date.11, 12 The Nitz & Capano13. We also fix the sky location to the maximum signal is consistent with the merger of two high mass black likelihood values from the same analysis.

1 Figure 1. Marginal posterior probability distributions on frequency and damping time from an agnostic quasi-normal mode analysis of GW190521 at 6ms after tref. A single mode is searched for in each of the shown frequency ranges, range A 50 to 80Hz and range B 80 to 256Hz. Top panels show the marginal posterior on the mode frequencies, with priors indicated by dotted lines; the gray region in the top left panel shows the highest 90th percentile density interval of the dominant mode. White dotted (dashed) contours in the bottom panels show the 50th (90th) credible regions. Assuming the dominant mode in range A corresponds to the (220) mode of a Kerr black hole, we estimate what the frequency and damping times would be of the (330), (440), and (550) modes (blue, green, and red regions, respectively). The mode in range B is clearly consistent with the expected frequency and damping time of the (330) mode. Here we do not see the (440) and (550) modes, indicating they are weaker than the (330) mode. This is consistent with an asymmetric binary black hole merger.

The ringdown spectrum of a Kerr black hole consists of an relation is assumed between the modes. infinite set of frequencies f`mn and damping times τ`mn labeled We repeat this analysis at time steps of tref + 0, 6, 12, 18, by three integers (`,m,n). Here ` and m are the usual angular and 24ms. As expected from the visual inspection of the data, harmonic numbers. The third index n ≥ 0 denotes overtones, we find a significant mode in range A at all grid points, which with n = 0 being the fundamental mode. The most agnostic decreases in amplitude at later times. A clear second mode way to search for quasi-normal modes from a perturbed black is found in range B. This mode is most visible at tref + 6ms, hole is to search for them individually, without assuming any the result of which is shown in Fig.1 (results at other times relation between them. Such a search is complicated by the na- are shown in Supplemental Fig. S.2). The frequency of the ture of quasi-normal modes: they are not orthogonal, meaning secondary mode at this time is ∼ 98Hz with a damping time that modes that overlap in time must be sufficiently separated of ∼ 30ms, while the primary mode has frequency of ∼ 63Hz in frequency or damping time in order to be distinguishable. and damping time ∼ 26ms. The signal-to-noise ratio of the Simulations of binary black hole mergers have shown that the primary and secondary modes of the maximum likelihood fundamental ` = m = 2 mode is typically significantly louder waveform is 12.2 and 4.1, respectively. Results from range C than other modes. In order to extract sub-dominant modes (not shown) are consistent with noise. from noisy data in an agnostic search it is useful to separate the dominant mode in frequency from the others. The dominant mode found at ∼ 63Hz is expected to be the quadrupolar ` = m = 2, n = 0 fundamental mode. Mea- A visual inspection of the time- and frequency-domain data surement of f220 and τ220 provides an estimate of the mass taken at the reference time revealed significant power in the and angular momentum of the remnant black hole.14 This in two LIGO detectors between 60-70 Hz (see Supplemental turn predicts the entire ringdown spectrum of subdominant Fig. S.1). In order to isolate this and search for sub-dominant modes. Figure1 shows that the subdominant mode at ∼ 98Hz modes we constructed three frequency ranges: “range A” be- is consistent with the ` = m = 3, n = 0 mode. This is also in tween 50−80Hz, “range B” between 80−256Hz, and “range agreement with expectations from numerical simulations of C” between 15−50Hz. We search for one quasi-normal mode binary black hole mergers.15, 16 in each range using Bayesian inference. We use uniform pri- ors on the relative amplitudes of the modes in range B and C between 0 and 0.9 times the mode in range A. No other

2/14 Figure 2. Bayes factor of models with the indicated modes compared to the stronger of the (220) or the (220) + (221) Figure 3. Posterior distribution of final redshifted mass modes model. The Bayes factor for the (220) + (221) model (1 + z)M and dimensionless spin χ measured at 7ms after is calculated against the (220)-only model. The hexagon f f tref assuming the identified modes are the (220) and (330) marks where no-hair tests are performed in section3. modes of a Kerr black hole. Dashed lines indicate the 90% credible interval. For the Kerr with δ(330) results, we use fitting formulae14 to convert the frequency f (1 + δ f ) 3 Consistency with the Kerr solution 330 330 and damping time τ330(1 + δτ330) into mass and spin. The search for damped sinusoids in section2 assumed no par- ticular relation between different modes, with a corresponding large prior parameter volume. In this section, we assume that with only the (220) mode. For models that include the (330) the frequency and damping times of the damped sinusoids (or other sub-dominant modes), the Bayes factor is evaluated are related as in the ringdown of a Kerr black hole. This re- against the stronger of the (220) or (220)+(221) models. duces the prior parameter volume and focuses in on particular The Bayes factors for the various multimode Kerr mod- modes. The amplitudes and phases of the modes are left as els are shown in Fig.2. Consistent with the agnostic re- free parameters, since they depend on the specific initial state sults, we find strong evidence for the presence of the (330) of the remnant black hole immediately after the merger. mode around 6ms, with the Bayes factor for the (220)+(330) +6 For this analysis, we model the ringdown signal based on model peaking at 44−5 one millisecond later, at tref + 7ms. the final Kerr black hole mass, Mf , and dimensionless spin, The maximum likelihood ringdown waveforms at this time 2 are shown in Supplemental Fig. S.3. Only moderate evidence χ f = Jf /Mf . We expect only a subset of the entire spectrum of quasi-normal modes to be visible above noise. Including is found for other fundamental modes, although at 19ms after all possible modes in our signal model can lead to overfitting tref the most favoured model contains the (220), (330), (440) the data. For this reason we perform several analyses which and (210) modes. Extending the analysis to earlier times, we include different combinations of the (330), (440), (210), and find increasing support for the presence of the (221) overtone (550) modes, in addition to the dominant (220) mode. Nu- (see Supplemental Fig. S.4), with the Bayes factor reaching merical simulations of binary black hole mergers have gen- a maximum at tref − 5ms. This would be consistent with the erally shown these modes to be the strongest.16 Giesler et merger happening prior to tref, in agreement with a recent 24 al.25 found that overtones of the dominant harmonic are sig- reanalysis of GW190521. nificant close to the merger time, and that including them Figure3 shows the redshifted mass and the dimensionless allows a quasi-normal mode description of the signal to be spin of the final black hole, measured with the (220) + (330) used at earlier times. We therefore also perform analyses in Kerr model at 7ms after tref. We find that the remnant black +30 which we include the first overtone of the dominant harmonic hole has a redshifted mass (1 + z)Mf = 330−40 M and di- +0.05 (`mn) = (221). mensionless spin χ f = 0.87−0.10. We repeat these analyses in 1ms intervals between tref + If a quasi-normal model without overtones is used too close [−9ms, 24ms]. We use Bayes factors to determine which to merger, the resulting final mass estimate can be biased model is most favored at each time step. For the model that toward larger values.3, 25 Supplemental Fig. S.4 shows the includes the fundamental dominant harmonic (220) and its stability of the final mass estimate between 6 and 12ms using overtone (221), the Bayes factor is evaluated against the model the (220) + (330) model, and its agreement with the mass

3/14 estimate at earlier times using the (221) overtone. This indi- cates that by this time the black hole has reached a regime of constant ringdown frequency, a requirement for the validity of linear-regime, quasi-normal modes. Given the strong evidence for the presence of the (330) mode at tref + 7ms, we can perform the classic no-hair the- 6 orem test. We keep the dependence of f220 and τ220 on (Mf , χ f ) as in the Kerr solution but introduce fractional devia- tions δ f330 and δτ330 of f330 and τ330, respectively. Figure3 shows the Kerr black hole mass Mf and dimensionless spin χ f associated to the (330) mode frequency f330(1 + δ f330) and damping time τ330(1 + δτ330) measured at 7ms after tref. Posterior distributions on the fractional deviations are shown in Supplemental Fig. S.5. We constrain the fractional devia- +0.07 tion from Kerr to δ f330 = −0.01−0.11. The damping time is +2.0 only weakly constrained, with δτ330 = 0.6−1.1.

4 Discussion and conclusions Figure 4. Posterior distribution for (top) the amplitude ratio The detection of a (330) mode indicates that the progeni- of the (330) mode, A /A , and (bottom) the mass ratio of tor black holes in GW190521 had asymmetric masses, since 330 220 the binary, m /m < 1, obtained using numerical fits between equal-mass binaries are not expected to excite the (330) mode. 2 1 the (330) amplitude and mass ratio.16 Vertical dashed lines Numerical fits16 provide the relation between the amplitude indicate the 90% credible interval. In the bottom panel we of the (330) mode and the ratio of the initial black hole masses assume a prior uniform in mass ratio between 1/25 – 1. For m and m for quasi-circular, aligned spin binaries. We find 1 2 comparison, we also show the mass ratio obtained from the m /m = 0.4+0.2 from the (330) amplitude measured by the 2 1 −0.3 full signal using NR Surrogate13 and PhenomXPHM (220) + (330) Kerr model at 7ms after t . Figure4 shows ref models12 assuming the same prior. the posterior distribution on the amplitude of the (330) mode and on the corresponding mass ratio. The redshifted final mass of GW190521 measured by Black hole spectroscopy tests showed consistency with the the LVC11 using a (220) ringdown fit is (1 + z)M = f Kerr hypothesis for these events.22, 23 However, the resulting 282.2+50.0M , or 259.2+36.6M when analyzing the full sig- −61.9 −29.0 constraints were weaker than what we find with the (330) nal. The low-mass-ratio part of the posterior of Nitz and Ca- mode here. Furthermore, while there is strong numerical 13 ( + z)M ∼ 17, 18 pano found 1 f 260M using the full signal . evidence for the presence of ringdown overtones close to the These results are somewhat in tension with the final mass and merger,25 a number of theoretical questions remain as to the spin inferred from the ringdown modes found here. validity of a quasi-normal description of the black hole close However, the complete waveform models used in the above to merger.26–30 analyses may not include all relevant physical effects. This, Given the evidence we find for the (221) model at t − coupled with the fact that GW190521 has a very short inspiral ref 5ms, we also perform a no-hair theorem test on the (221) signal, can lead to systematic errors for parameter estimation. overtone at this time. The results are shown in Supplemental For example, the waveform models used in the LVC analysis Fig. S.6. We find poor constraints, with δ f = −0.11+0.33 and Nitz & Capano assume quasi-circular orbits, but several 221 −0.04 +0.31 studies have indicated that the binary may have been eccentric and δτ221 = 0.45−0.66. This highlights the benefits of black at merger19–21. These studies have also found slightly larger hole spectroscopy using the fundamental modes. estimates for the binary’s total mass, making them more con- The true nature of the gravitational wave event GW190521 sistent with our estimate for the final mass. The ringdown has been the subject of much speculation.31–33 The interpreta- waveforms used in this paper are simpler and more robust tion of GW190521 as a head-on collision of two highly spin- than full inspiral-merger-ringdown models for signals like ning Proca stars33 predicts the presence of a (200) mode.34 GW190521, provided they are applied sufficiently late in the We do not find evidence for such a mode. Additionally, the post-merger regime. This argument would tend to favor the high-mass, multiple-mode ringdown signal observed here estimates derived in this paper for the total mass. Neverthe- does not agree with the scenario of a very massive star collaps- less, a full resolution of this tension is beyond the scope of ing to a black hole of mass ∼ 50M and an unstable massive this work. disk.35 Evidence for overtones of the (220) mode very close to Expectations based on population models were that black merger were previously found for the events GW15091422 and hole ringdown signals with multiple modes were unlikely to GW190521 07435923 (not to be confused with GW190521). be observed with the Advanced LIGO and Virgo detectors7,8

4/14 (although those population models did not include massive 21. J. Calderon-Bustillo,´ N. Sanchis-Gual, A. Torres-Forne´ and J. A. Font, binaries). Our results here show that, remarkably, GW190521 “Confusing Head-On Collisions with Precessing Intermediate-Mass displays a distinct subdominant mode and that this mode is Binary Black Hole Mergers,” Phys. Rev. Lett. 126 (2021), 201101 consistent with the ringdown of a Kerr black hole. 22. M. Isi, M. Giesler, W. M. Farr, M. A. Scheel and S. A. Teukolsky, “Testing the no-hair theorem with GW150914,” Phys. Rev. Lett. 123 (2019) no.11, 111102 References 23. R. Abbott et al., “Tests of General Relativity with Binary Black Holes from the second LIGO-Virgo Gravitational-Wave Transient Catalog,” 1. C. V. Vishveshwara, ”Scattering of Gravitational Radiation by a [arXiv:2010.14529 [gr-qc]]. Schwarzschild Black-hole”, Nature 227, 936–938 (1970) 24. H. Estelles,´ S. Husa, M. Colleoni, M. Mateu-Lucena, M. d. Planas, 2. S. Chandrasekhar and S.L. Detweiler, ”The quasi-normal modes of C. Garc´ıa-Quiros,´ D. Keitel, A. Ramos-Buades, A. K. Mehta and the Schwarzschild black hole”, Proc. Roy. Soc. Lond. A 344 (1975), A. Buonanno, et al. [arXiv:2105.06360 [gr-qc]]. 441–452” 25. M. Giesler, M. Isi, M. A. Scheel and S. Teukolsky, “Black Hole Ring- 3. B. P. Abbott et al., “Tests of general relativity with GW150914,” Phys. down: The Importance of Overtones,” Phys. Rev. X 9, no.4, 041060 Rev. Lett. 116, 221101 (2016) [erratum: Phys. Rev. Lett. 121, no.12, (2019) 129902 (2018)] 26. H.P. Nollert, ”About the significance of quasinormal modes of black 4. R. P. Kerr, “Gravitational field of a spinning mass as an example of holes”, Phys. Rev. D, 53 (1996), 4397–4402”. algebraically special metrics,” Phys. Rev. Lett. 11 (1963), 237-238 27. H.P. Nollert and R. Price, ”Quantifying excitations of quasinormal mode 5. V. Cardoso and P. Pani, “Testing the nature of dark compact objects: a systems”, J. Math. Phys. 40 (1999), 980–1010. status report,” Living Rev. Rel. 22 (2019) no.1, 4 28. J.L. Jaramillo, P. Macedo and L. Al Sheikh, ”Pseudospectrum and 6. O. Dreyer et al., “Black hole spectroscopy: Testing general relativity black hole quasi-normal mode (in)stability”, To appear in Phys. Rev. X through gravitational wave observations,” Class. Quant. Grav. 21 (2004), [arXiv:2004.06434 [gr-qc]]. 787-804 29. R.G. Daghigh, M.D. Green and J.C. Morey, ”Significance of Black 7. E. Berti, A. Sesana, E. Barausse, V. Cardoso and K. Belczynski, “Spec- Hole Quasinormal Modes: A Closer Look”, Phys. Rev. D 101 (2020), troscopy of Kerr black holes with Earth- and space-based interferome- 104009 ters,” Phys. Rev. Lett. 117 (2016) no.10, 101102 30. W. Qian, K. Lin, C. Shao, B. Wang and R. Yue, ”Asymptotical quasi- 8. M. Cabero et al., “Black hole spectroscopy in the next decade,” Phys. normal mode spectrum for piecewise approximate effective potential”, Rev. D 101 (2020) no.6, 064044 Phys. Rev. D 103 (2021), 024019 9. R. Abbott et al., “GW190521: A Binary Black Hole Merger with a 31. J. Sakstein, D. Croon, S. D. McDermott, M. C. Straight and E. J. Baxter, Total Mass of 150M ,” Phys. Rev. Lett. 125 (2020) no.10, 101102 “Beyond the Standard Model Explanations of GW190521,” Phys. Rev. Lett. 125 (2020) no.26, 261105 10. R. Abbott et al., “Properties and Astrophysical Implications of the 150 M Binary Black Hole Merger GW190521,” Astrophys. J. Lett. 900 32. M. Safarzadeh and Z. Haiman, “Formation of GW190521 via gas accre- (2020) no.1, L13 tion onto Population III remnants born in high- minihalos,” Astrophys. J. Lett. 903 (2020) no.1, L21 11. R. Abbott et al., “GWTC-2: Compact Binary Coalescences Observed by LIGO and Virgo During the First Half of the Third Observing Run,” 33. J. C. Bustillo et al., “GW190521 as a Merger of Proca Stars: A Potential [arXiv:2010.14527 [gr-qc]]. New Vector Boson of 8.7×10−13 eV,” Phys. Rev. Lett. 126 (2021) no.8, 081101 12. A. H. Nitz et al, “3-OGC: Catalog of gravitational waves from compact- binary mergers,” [arXiv:2105.09151 [astro-ph.HE]]. 34. C. Palenzuela, I. Olabarrieta, L. Lehner and S. L. Liebling, “Head-on collisions of boson stars,” Phys. Rev. D 75 (2007), 064005 13. A. H. Nitz and C. D. Capano, “GW190521 may be an intermediate mass ratio inspiral,” Astrophys. J. Lett. 907 (2021) no.1, L9 35. M. Shibata et al., “Alternative possibility of GW190521: Gravitational waves from high-mass black hole-disk systems,” Phys. Rev. D 103 14. E. Berti, V. Cardoso and C. M. Will, “On gravitational-wave spec- (2021) no.6, 063037 troscopy of massive black holes with the space interferometer LISA,” Phys. Rev. D, 73 (2006), 064030 15. I. Kamaretsos, M. Hannam, S. Husa and B. S. Sathyaprakash, “Black- Acknowledgments hole hair loss: learning about binary progenitors from ringdown signals,” Phys. Rev. D 85 (2012), 024018 The authors thank Ofek Birnholtz, Jose Luis Jaramillo, Rein- hard Prix, Bruce Allen, Evan Goetz, Saul Teukolsky, Max- 16. S. Borhanian, K. G. Arun, H. P. Pfeiffer and B. S. Sathyaprakash, “Com- parison of post-Newtonian mode amplitudes with imiliano Isi, Juan Calderon-Bustillo,´ Abhay Ashtekar and simulations of binary black holes,” Class. Quant. Grav. 37 (2020) no.6, Bangalore Sathyaprakash for useful discussions. We thank 065006 also the Atlas Computational Cluster team at the Albert Ein- 17. V. Varma et al., “Surrogate model of hybridized numerical relativity stein Institute in Hanover for assistance. MC acknowledges binary black hole waveforms,” Phys. Rev. D 99 (2019) no.6, 064045 funding from the Natural Sciences and Engineering Research 18. G. Pratten et al., “Let’s twist again: computationally efficient models for Council of Canada (NSERC). This research has made use of the dominant and sub-dominant harmonic modes of precessing binary data obtained from the Gravitational Wave Open Science Cen- black holes,” [arXiv:2004.06503 [gr-qc]]. ter (https://www.gw-openscience.org/ ), a service of LIGO 19. I. M. Romero-Shaw, P. D. Lasky, E. Thrane and J. C. Bustillo, Laboratory, the LIGO Scientific Collaboration and the Virgo “GW190521: orbital eccentricity and signatures of dynamical formation in a binary black hole merger signal,” Astrophys. J. Lett. 903 (2020) Collaboration. LIGO Laboratory and Advanced LIGO are no.1, L5 funded by the United States National Science Foundation 20. V.Gayathri et al. “GW190521 as a Highly Eccentric Black Hole Merger,” (NSF) who also gratefully acknowledge the Science and Tech- [arXiv:2009.05461 [astro-ph.HE]]. nology Facilities Council (STFC) of the United Kingdom,

5/14 the Max-Planck-Society (MPS), and the State of Niedersach- For a given ` and n, the +m and −m modes are related to 14 sen/Germany for support of the construction of Advanced each other by ω`−mn = −ω`mn and τ`−mn = τ`mn . Further- LIGO and construction and operation of the GEO600 detector. more, if the ringdown waveform is circularly polarized, the Additional support for Advanced LIGO was provided by the amplitude and phase of the ±m modes are related to each iφ ` −iφ Australian Research Council. Virgo is funded, through the other by A`−mne `−mn = −(1) A`mne `mn . By “circularly po- European Gravitational Observatory (EGO), by the French larized”, we mean that an observer at an inclination angle of 0 Centre National de Recherche Scientifique (CNRS), the Italian or π (above or below the black hole, as determined by the di- Istituto Nazionale di Fisica Nucleare (INFN) and the Dutch rection of its spin) would detect gravitational waves in which Nikhef, with contributions by institutions from Belgium, Ger- the plus and cross polarizations have equal amplitudes; the many, Greece, Hungary, Ireland, Japan, Monaco, Poland, polarization for observers at other positions are determined Portugal, Spain. by the spheroidal harmonics. Numerical simulations of black hole binaries have shown Author contributions: Conceptualization, CDC, MC, JA, that it is difficult to get non-circular polarization in the ring- ABN, BK; Data release, CDC, MC, JW; Formal analysis, down14. To simultaneously sum over the ±m modes for a CDC, MC, JA, JW; Funding acquisition: AHN, BK; Method- given `n, we parameterize the waveform as ology: CDC, MC, SK, JW, AHN, ABN, BK; Project admin- istration: CDC, ABN, BK; Resources: AHN, BK; Software: 0 −t/τ`|m|n h`|m|n = A`|m|ne × CDC, MC, SK, JW, AHN; Supervision: CDC, MC, AHN, (+) i(ω`|m|nt+φ`|m|n) ABN, BK; Validation: CDC, MC, JA, SK, JW, AHN, ABN, [−2S`mn(ι,ϕ)A`|m|ne BK; Visualization: CDC, MC, JA, SK, JW; Writing–original (−) + S (ι,ϕ)A e−i(ω`|m|nt−φ`|m|n)], draft: CDC, ABN, BK; Writing–review and editing: CDC, −2 `−mn `|m|n MC, JA, SK, JW, AHN, ABN, BK. 0 where A`|m|n is the intrinsic amplitude of the (`mn) mode and Competing interests: The authors declare no competing in- √ terests. (+) Al|m|n = 2cos(π/4 + ∆βl|m|n) (2) Correspondence: C.D. Capano, [email protected]. √ (−) i(lπ+∆φl|m|n) Al|m|n = 2sin(π/4 + ∆βl|m|n)e . (3) 5 Methods If the parameters ∆β`mn and ∆φ`mn are both zero, the wave- 5.1 The ringdown signal model form reduces to circular polarization. In the quasi-normal mode (QNM) spectrum of a perturbed When using the Kerr model, we do two analyses at several Kerr black hole, the allowed frequencies f`mn = ω`mn/(2π) times, a circularly polarized one in which ∆β`mn and ∆φ`mn are and damping times τ`mn are labeled by three integers ` = both set to zero, and one in which a common ∆β and ∆φ for 2,3,..., −` ≤ m ≤ `, and n = 0,1,2,.... These can be com- all modes are allowed to vary uniformly between [−π/4,π/4) bined together in a complex frequency Ω`mn = ω`mn + i/τ`mn and [−π,π), respectively. In all cases we find that the circu- such that the ringdown signal of a perturbed Kerr black hole larly polarized analysis is favored over the arbitrary polarized can be expressed as a sum of damped sinusoids: analysis; we therefore only report results from the former here. In all analyses we fix the azimuthal angle ϕ = 0, as it is de- Mf h + ih = S (ι,ϕ)A ei(Ω`mnt+φ`mn) , (1) + × D ∑ −2 `mn `mn generate with modes’ initial phases. To obtain the frequency L `mn and damping times for a given mass and spin we use tabulated 14 where h+ and h× are the plus and cross polarizations of the values from Berti et al., which we interpolate using a cubic gravitational wave, Mf is the mass of the black hole in the spline. For the spheroidal harmonics we use tabulated values 14 detector frame and DL is the luminosity distance to the source. of the angular separation constants (also from Berti et al. ) and solve the recursion formula given in Leaver.37 Our code The functions −2 S`mn(ι,ϕ) are the spin-weighted spheroidal harmonics of spin weight −2, which depend on the inclination for doing this is publicly available on GitHub.38 angle ι between the black hole spin and the line-of-sight from For the agnostic analysis, we do not assume any mode the observer to the source, and the azimuth angle ϕ between corresponds to any particular `mn. We therefore use arbitrary iψ the black hole and the observer. complex numbers X`±mn = e `±mn in place of the −2S`±mn. The complex QNM frequencies Ω`mn can be determined Here, the ψ`±mn are allowed to vary uniformly in [0,2π). We from the Teukolsky equation36, 37. According to the no-hair vary a common ∆β parameter, but fix the ∆φ parameter to theorem, the frequencies and damping times are determined by zero, as it is degenerate with the X`mn. the mass Mf and spin χ f of the black hole, with χ f ∈ (−1,1). Positive (negative) spin means the perturbation is co(counter)- 5.2 Computational analysis methods rotating with respect to the black hole. The amplitudes A`mn Standard parameter estimation with gravitational waves be- and phases φ`mn depend on the initial perturbation and take gins with Bayes’ theorem. Given some data s and a signal different values for different `mn modes. model h that depends on some set of parameters λ, we wish

6/14 to know the posterior probability density function p(λ|s,h). hood function for the noise,41 Applying Bayes’ theorem we have " # 1 K 1 p(s |n) ∝ exp − hn ,n i . (6) p(λ|s,h) = p(s|λ,h)p(λ|h), net 2 ∑ d d Z d=1 where p(s|λ,h) is the likelihood function, p(λ|h) is the prior, Here, the inner product h·,·i is defined as and Z is a normalization constant known as the evidence. Es- timates on a single parameter are obtained by marginalizing ( N/2−1 ∗ ) 1 u˜d[p]v˜d[p] the posterior over all other parameters; marginalizing over all hud,vdi ≡ 4ℜ , (7) T ∑ (d) parameters yields the evidence. Taking the ratio of evidences p=1 Sn [p] Z /Z for two different signal models yields the “Bayes fac- A B whereu ˜ is the discrete Fourier transform of the time series u. tor”. If our prior belief for the validity of the two models is the The signal hypothesis is that the data consists of the signal same, the Bayes factor gives the odds that model A is favored plus the noise. The likelihood function p(s| ,h) is there- over model B. Using a scale by Kass and Raftery,39 a Bayes λ fore Eq. (6) with the n replaced by the residuals s − h . factor greater than 3.2 is considered “substantial” evidence in d d d However, this assumes that h is an accurate model of the favor of model A; greater than 10 is “strong” evidence; greater signal across the entire observation time T. As stated above, than 100 is “decisive”. quasi-normal modes only model the gravitational wave from Evaluating the posterior requires a likelihood function a binary black hole after the merger, when the two compo- p(s|λ,h). Consider a gravitational-wave detector, which we nent black holes have formed a single, perturbed black hole. sample every ∆t seconds over a time T to obtain N = bT/∆tc Performing Bayesian inference using quasi-normal modes as time-ordered samples s = {s ,...,s }. A network of K de- 0 N the signal model therefore requires excising times from the tectors sampled in this way will produce a set of samples data when the ringdown prescription is not valid. In other s = {s ,...,s }. To obtain a likelihood function we first net 0 K words, instead of considering the full set of time samples consider the hypothesis that the set of samples only contain s = {s ,...,s }, we wish to only evaluate the truncated set noise p(s |n) = p(n ). 0 N net net s = {s ,...,s ,s ,...,s }, with M > 1. The data between In gravitational-wave astronomy it is common to assume tr 0 a a+M N the time steps [a,a + M) is said to be ”gated”. that, in the absence of a signal, the detectors output stochastic Gaussian noise that has zero mean and is independent across The probability density function of the truncated noise is detectors. Under this assumption the probability density func- still a multivariate normal distribution (excising dimensions tion describing the network of time-ordered noise samples from a multivariate normal is equivalent to marginalizing over those dimensions), and so Eq. (4) still applies. The nnet is a product of KN−dimensional multivariate normal distributions, challenge is that the covariance matrix of the truncated noise Ctr is no longer Toeplitz. Its eigenvectors can no longer  1 K > −1  exp − 2 ∑d=1 nd Cd nd be approximated by that of a circulant matrix, and so the p(nnet) = q . (4) expression for the likelihood Eq. (6) is no longer valid. The NK K (2π) ∏d=1 detCd inverse of the covariance matrix needs to be found by other means. One possibility is to numerically invert the covariance Here, C is the covariance matrix of the noise in detector d. d matrix. However, this is numerically unstable due to the large In order to evaluate this function it is necessary to know what dynamic range of the matrix’s elements, and computationally the inverse of the covariance matrix is. impractical for observation times of more than a ∼second. If we assume that a detector’s noise is wide-sense stationary Instead, we use “gating and in-painting” to find the likelihood and ergodic, then its covariance C is a symmetric Toeplitz of the truncated time series. This was applied to the problem matrix with elements given by the autocorrelation function of matched filtering in Zackay et al.42 Here we apply it to of the data. If the autocorrelation function goes to zero in parameter estimation. some finite amount of time that is less than T/2 (for the LIGO Define n0 = n + x, where n is the noise with the gated and Virgo detectors, this typically happens within a few sec- g g times t ∈ [a,a+M)∆t zeroed out, and x is a vector that is zero onds), then the eigenvectors of the covariance matrix can be everywhere except in the gated times. If the non-zero elements well-approximated by that of a circulant matrix. All circulant √ of x are such that (C−1n0)[k] = 0 for all k ∈ [a,a + M), then matrices have the same eigenvectors, e−2πikp/N/ N.40 Solv- n0TC−1n0 will be the same as the truncated version nTC−1n . ing for the eigenvalues yields an analytic expression for C−1: tr tr tr Our aim is to solve the equation C−1(n + x) = 0 in the gated the j,k-th element is the discrete inverse Fourier transform of g region. Since x is zero outside of the gated region, C−1x only 1/S evaluated at the k − j time step, n involves the [a,a+M) rows and columns of C−1, which form −1 2 −1 −1 an M × M Toeplitz matrix (cf. Eq. (5)). We therefore solve C [ j,k] ≈ 2∆t F (Sn )[k − j], (5) for x such that where Sn is the power spectral density of the detector’s noise. −1 −1 Substituting this back into Eq. (4), yields a canonical likeli- C x = −C ng, (8)

7/14 where the overbar indicates the [a,a+M) rows (and columns) pykerr, available at https://github.com/cdcapano/ of the given vector (matrix). This can be solved numerically pykerr. using a Toeplitz solver.47, 48 Adding x to the gated noise (“in painting”) will then yield the same result as if we had References for Methods truncated the noise and the covariance matrix. 36. S. A. Teukolsky, “Rotating black holes - separable wave equations for Note that if the gate spans the entire beginning of the data gravitational and electromagnetic perturbations,” Phys. Rev. Lett. 29 segment, the truncated covariance matrix Ctr is Toeplitz, and (1972), 1114-1118 so could be inverted numerically using a Toeplitz solver. This 37. E. W. Leaver, “An Analytic representation for the quasi normal modes is the method used by Isi et al.22 The advantage of using in- of Kerr black holes,” Proc. Roy. Soc. Lond. A 402 (1985), 285-298 painting is that it involves solving for an M × M matrix rather 38. C. D. Capano, 2021, pykerr, than an (N − M) × (N − M) matrix. Gating and in-painting https://github.com/cdcapano/pykerr, GitHub also have other applications beyond what we use it for here, 39. R. Kass and A. E. Raftery, “Bayes Factors,” J. Am. Statist. Assoc. 90 such as excising glitches from data when doing parameter (1995) no. 430, 773-795 estimation. 40. R. M. Gray, “Toeplitz and Circulant Matrices: A Review,” Foundations and Trends in Communications and Information Theory 2 (2006) n0. 3, To evaluate the likelihood for a signal, we use ng = sg − 155-239 hg (i.e., the residual with the gated region zeroed out) in Eq. (8) and solve for x. We can then use x + s − h in the 41. L. S. Finn, “Detection, measurement and gravitational radiation,” Phys. g g Rev. D 46, 5236-5249 (1992) standard likelihood, Eq. (6). We do not attempt to normalize 42. B. Zackay, T. Venumadhav, J. Roulet, L. Dai and M. Zaldarriaga, the likelihood, which would involve finding the determinant of “Detecting Gravitational Waves in Data with Non-Gaussian Noise,” the truncated covariance matrix. For this reason, we calculate [arXiv:1908.05644 [astro-ph.IM]]. and report Bayes factors by comparing models that start at the 43. A. H. Nitz et al., 2021, PyCBC Software, same time offset from tref, for which the determinant cancels. https://github.com/gwastro/pycbc, GitHub We use the open source PyCBC Inference library for per- 44. C. M. Biwer et al., “PyCBC Inference: A Python-based parameter forming Bayesian inference,43, 44 to which we have added the estimation toolkit for compact binary coalescence signals,” Publ. Astron. gated likelihood described above. For all analyses we use Soc. Pac. 131 (2019) no.996, 024503 a gate of two seconds, ending at the start time of the ring- 45. J. S. Speagle, “dynesty: a dynamic nested sampling package for estimat- down. For sampling the parameter space we use the dynesty ing Bayesian posteriors and evidences,” Mon. Not. Roy. Astron. Soc. 493 (2020) no. 3, 3132-3158 nested sampler.45 We use data for the event GW190521 made 46. R. Abbott et al., “Open data from the first and second observing runs of publicly available by the Gravitational Wave Open Science Advanced LIGO and Advanced Virgo,” [arXiv:1912.11716 [gr-qc]]. Center.46 We fix the sky location to the values given by the 13 47. P. Virtanen et al., “SciPy 1.0–Fundamental Algorithms for Scientific maximum likelihood result of Nitz and Capano, although Computing in Python,” Nature Meth. 17, 261 (2020) we have obtained similar results using the LVC’s maximum 48. C. R. Harris et al., “Array programming with NumPy,” Nature 585, 10 likelihood sky location. We use a geocentric GPS reference no.7825, 357-362 (2020) 13 time of tref = 1242442967.445. With the sky location used in our analyses, this corresponds to the detector GPS refer- 6 Supplemental ence times 1242442967+0.4259 at LIGO Hanford, +0.4243 at LIGO Livingston and +0.4361 at Virgo. Credible intervals in the text are quoted to 90%.

5.3 Data availability Posterior data samples and data necessary to reproduce the fig- ures are available at https://github.com/gwastro/ BH-Spectroscopy-GW190521. The gravitational-wave data used in this work were obtained from the Gravitational Wave Open Science Center (GWOSC) at https://www. gw-openscience.org.

5.4 Code availability All software used in this analysis is open source. Bayesian inference was performed with the PyCBC li- brary, available at https://github.com/gwastro/ . Configuration files used to perform all analy- ses can be found at https://github.com/gwastro/ BH-Spectroscopy-GW190521. Spheroidal harmonics, Kerr frequencies, and Kerr damping times generated using

8/14 100 H1 C A B H1 tref H1 t + 1s 0 ref 100 L1 tref L1 tref + 1s 100

100 L1

0

100

100 V1 10 1 whitened gated strain

0 whitened frequency-domain strain whitened frequency-domain

100 0.10 0.05 0.00 0.05 0.10 0.14 101 102 103 time - detector tref (s) frequency (Hz)

Figure S.1. Left: Whitened data in each detector. Dark traces show the data gated at the reference time tref in each detector, which is indicated by the gray vertical lines. Right: Frequency domain representation of the Hanford and Livingston data shown in the left panel. Also shown is the frequency domain representation at an off-source time, one second later. The signal is clearly visible in the LIGO time-domain data, and is seen as a spike in the frequency domain data between ∼ 60 − 70Hz. The primary frequency bin (“A”) boundaries were set to isolate this spike. Frequencies below (region “C”) and above (region “B”) were searched for additional QNMs in the agnostic analysis.

9/14 Figure S.2. Spectra plots at 0, 6, 12, and 18ms showing marginal posterior distributions from frequency range A and B in the agnostic analysis. Also shown are the expected regions for the (330) (blue), (440) (green), and (550) (red) modes, assuming the peak in region A is the (220) mode.

10/14

10/14 100

0

H1 100

220 220+330 100

0

100 L1

100 whitened gated strain

0

V1 100 0.10 0.05 0.00 0.05 0.14 time - detector tref (s)

Figure S.3. Whitened data in each detector, with a gate applied at 7ms (gray lines) after the detector reference time tref (gray dotted lines). Semi-transparent traces show the whitened data without the gate. Plotted are the maximum likelihood waveforms using just the (220) mode (black) and the (220) plus (330) mode (orange).

11/14 Figure S.4. Bayes factors as shown in Fig.2 for an extended range of times. ( Top) Bayes factor of models with the indicated modes compared to the stronger of the (220) or the (220) + (221) modes model. The Bayes factor for the (220) + (221) model is calculated against the (220)-only model. (Center) Median values for the frequency of the (220) mode for the model with the indicated modes. The shading shows the 90% credible interval. (Bottom) Median values and 90% credible intervals for the redshifted final mass. Hexagons mark where no-hair tests are performed in section3 for the (220) + (330) model and in section4 for the (220) + (221) model. All values are shown for start times of the analysis relative to the reference time tref.

12/14 Figure S.5. Posterior on the deviation from Kerr of the (330) frequency δ f330 and damping time δτ330, as well as the resulting modified (330) frequency, using a model in which we include the (220)+(330) modes at tref + 7ms. Quoted values are the median and 90% credible interval, which is indicated by the dashed vertical lines. The fractional deviations are expected to be zero for a Kerr black hole (indicated by the green lines). We use a prior (black dotted lines) that is uniform over δ f330 ∈ [−0.3,0.3), with the constraint that f330(1 + δ f330) > 75 Hz. This constraint is necessary to avoid label switching with the (220) mode; even with the constraint we clearly measure a lower bound on δ f330. For the damping time we use a prior that is uniform over δτ330 ∈ [−0.9,3), and find that the damping time is only weakly constrained.

13/14 Figure S.6. Posterior on the deviation from Kerr of the (221) frequency δ f221 and damping time δτ221, as well as the resulting modified (221) frequency, using a model in which we include the (220)+(221) modes at tref − 5ms. Quoted values are the median and 90% credible interval, which is indicated by the dashed vertical lines. The fractional deviations are expected to be zero for a Kerr black hole (indicated by the green lines). We use a prior (black dotted lines) that is uniform over δ f221 ∈ [−0.16,0.3), with the constraint that f221(1 + δ f221) > 55 Hz. This constraint is used to try to exclude additional noise that is present in the Hanford and Virgo detectors at ∼ 50Hz. For the damping time we use a prior that is uniform over δτ221 ∈ [−0.8,0.8). Despite the tighter prior constraints than that used for the (330) mode, we obtain a larger 90% credible interval on both δ f221 and δτ221. The posterior also peaks toward the prior boundaries. This may be due to the noise at low frequency, or may indicate that the signal is not fully captured by a sum of quasi-normal modes at this time.

14/14