You are on page 1of 12

Hints of spin-orbit resonances in the binary black hole population

Vijay Varma,1, 2, ∗ Sylvia Biscoveanu,3, 4 Maximiliano Isi,3, 4, † Will M. Farr,5, 6 and Salvatore Vitale3, 4
1
Department of Physics, Cornell University, Ithaca, New York 14853, USA
2
Cornell Center for Astrophysics and Planetary Science, Cornell University, Ithaca, New York 14853, USA
3
LIGO Laboratory, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA
4
Department of Physics and Kavli Institute for Astrophysics and Space Research,
Massachusetts Institute of Technology, 77 Massachusetts Ave, Cambridge, MA 02139, USA
5
Department of Physics and Astronomy, Stony Brook University, Stony Brook NY 11794, USA
6
Center for Computational Astrophysics, Flatiron Institute, New York NY 10010, USA
(Dated: July 22, 2021)
Binary black hole spin measurements from gravitational wave observations can reveal the binary’s
evolutionary history. In particular, the spin orientations of the component BHs within the orbital
arXiv:2107.09693v1 [astro-ph.HE] 20 Jul 2021

plane, φ1 and φ2 , can be used to identify binaries caught in the so-called spin-orbit resonances. In a
companion paper, we demonstrate that φ1 and φ2 are best measured near the merger of the two
black holes. In this work, we use these spin measurements to constrain the distribution of φ1 and
∆φ = φ1 − φ2 over the population of merging binary black holes. We find that there is a preference
for ∆φ ∼ ±π in the population, which can be a signature of spin-orbit resonances. We also find a
preference for φ1 ∼ −π/4 with respect to the line of separation near merger, which has not been
predicted for any astrophysical formation channel. However, we are unable to constrain the widths
of the φ1 and ∆φ distributions, and more observations may be necessary to confirm these trends.
Finally, we derive constraints on the distribution of recoil kicks in the population, and use this to
estimate the fraction of merger remnants retained by globular and nuclear star clusters.

Introduction.— Binaries of spinning black holes (BHs) z ̂ = L̂


serve as a unique astrophysical laboratory for a range of χ1
relativistic phenomena. For example, if the BH spins χ1
and χ2 are aligned with the orbital angular momentum χ2 θ1
L, the orientations of the orbital plane and the spins
remain fixed during the inspiral (cf. Fig. 1 for definitions
θ2
of the binary BH spin parameters). However, if the spins
are tilted with respect to L, relativistic spin-orbit and
spin-spin coupling cause the orbital plane and the spins
to precess [1, 2].
̂
While the tilt angles θ1 and θ2 control precession, the x̂ = l ϕ ϕ1 ŷ
orbital-plane spin angles φ1 and φ2 play a central role 2 Δϕ
in binaries undergoing spin-orbit resonances [3–6]. For
these binaries, the χ1 , χ2 and L vectors become locked
into a common resonant-plane such that ∆φ = φ1 − φ2 is Figure 1. The BH spins are represented by 3-vectors χ1 and
fixed at 0 or ±π as the binary precesses. While evidence χ2 , with index 1 (2) denoting the heavier (lighter) BH. We
parameterize the spins by their dimensionless magnitudes
for precession has been found in the astrophysical binary
χ1 , χ2 ≤ 1, tilts θ1 , θ2 w.r.t the orbital angular momentum
BH population [7], spin-orbit resonances have not yet L [10], and orbital-plane spin angles φ1 , φ2 w.r.t the line
been observed even though they are expected in some of separation l from the lighter to the heavier BH. Finally,
astrophysical scenarios. For example, stellar binaries can ∆φ = φ1 − φ2 .
cluster near these resonances if supernova natal kicks and
stellar tides are significant [8, 9].
Another important relativistic effect that gets amplified
to constrain the formation of heavy BHs via successive
for spinning binaries is the gravitational recoil. Gravi-
mergers [20]. However, the kick depends very sensitively
tational waves (GWs) can carry away linear momentum
on the orbital-plane spin angles [21].
from the binary, imparting a recoil or kick velocity to the
merger remnant [11–14]. These velocities can reach values GW observations by LIGO [22] and Virgo [23] have en-
up to ∼ 5000 km/s for precessing binaries [15–17], large abled increasingly precise constraints on the astrophysical
enough to be ejected from any host galaxy [18]. Kick distributions of BH spin magnitudes and tilts [7, 24], but
measurements from GW observations [19] can be used the distributions of the orbital-plane spin angles remain
unconstrained. Constraining these distributions will allow
us to understand the prevalence of spin-orbit resonances
and merger kicks in nature. The biggest obstacle for this,
∗ vvarma@cornell.edu; Klarman fellow however, is the difficulty in measuring φ1 , φ2 and ∆φ from
† NHFP Einstein fellow individual GW events with current detectors [25–30].
2

30
Jeffreys-σφ φ1 ∆φ

0.
0.75 Jeffreys-σφ
Flat-σφ

15
0.
0.50

p(φ)
µ∆φ [rad]

2
π

0.
2

0.25
0

1
0.
π
2

φ [rad]
σφ1 [rad]

Flat-σφ

4
0.
0.20

2
0.
π

p(φ)
0.15
σ∆φ [rad]

4
0.
2
0.10

0.
π

−π
0


−π/2 0 π/2 π
2

2

µφ1 [rad] µ∆φ [rad] σφ1 [rad] σ∆φ [rad] φ [rad]

Figure 2. Constraints on the φ1 and ∆φ populations at tref /M = −100. Left: Posteriors for the mean and width parameters.
The shaded regions show 90% credible bounds on joint 2D posteriors. The diagonal plots show 1D marginalized posteriors, with
the priors shown as dashed black lines. We consider two prior choices (Jeffreys-σφ and Flat-σφ ) for the width parameters.
Right: Constraints on the posterior population distributions p(φ1 ) and p(∆φ) for the two prior choices. Shaded regions show
90% credible bounds, while the solid lines show the mean. The dashed grey lines show the 90% prior bounds.

However, in a companion paper, Varma et al. [31], we of the full spin degrees of freedom in the binary BH
show that this can be greatly improved by measuring the population. Next, given the spin population, we derive
spins near the merger, in particular, at a fixed dimension- constraints on the kick population. Finally, we use the kick
less reference time tref /M = −100 before the peak of the constraints to estimate the fraction of merger remnants
GW amplitude, rather than the traditional choice of a retained by globular and nuclear star clusters.
fixed GW frequency of fref = 20 Hz. Here M = m1 + m2 is
the total (redshifted) mass of the binary with component Spin population.— To estimate the distribution of BH
masses m1 ≥ m2 , and we set G = c = 1. Ref. [31] shows spins in the astrophysical population, we assume that the
that this improvement can be attributed to the waveform true value of the spins for each event is drawn from a
being more sensitive to variations in the orbital-plane common underlying distribution. Then, using hierarchical
spin angles near the merger. In particular, measuring Bayesian inference, we estimate the properties of this
the spins near the merger leads to improved constraints distribution by collectively analyzing all 31 events. The
for φ1 and φ2 for several events in the latest GWTC-2 full details of our hierarchical analysis are provided in the
catalog [32–36] released by the LIGO-Virgo Collabora- Supplement [38], which we summarize below.
tion. While the ∆φ measurements are not significantly We first reweight the mass and spin posterior sam-
impacted, Ref. [31] shows that this parameter will also ples for each event to account for known astrophysi-
be better constrained with the louder signals we expect cal constraints on the primary mass and mass ratio
in the future. (q = m2 /m1 ) populations [7]. Using the reweighted
These spin measurements are obtained using the posteriors, we constrain the hyperparameters govern-
numerical relativity (NR) surrogate waveform model ing the distribution of the full spin degrees of freedom,
NRSur7dq4 [37]. This model accurately reproduces pre- S = {χ1 , χ2 , θ1 , θ2 , φ1 , ∆φ}. The spin magnitudes and
cessing NR simulations and is necessary to reliably mea- tilts are modeled following the “Default spin” model of
sure the orbital-plane spin angles [31]. GWTC-2 in- Ref. [7]. The orbital-plane spin angles φ1 and ∆φ are
cludes a total of 46 binary BH events. However, be- modeled as being drawn from independent von Mises dis-
cause NRSur7dq4 only encompasses ∼20 orbits before tributions [39]. The von Mises distribution is an approxi-
merger, it can only be applied to the shorter signals with mation of a Gaussian distribution with periodic boundary
M & 60 M [37]. This reduces our set of events to 31; conditions and is parameterized by a mean and a standard
these events are listed in Tab. I of Ref. [31]. deviation (or simply, width). The remaining angle, φ2 , is
In this Letter, we use the spin measurements at tref /M = redundant given φ1 and ∆φ; we choose to work with ∆φ
−100 from those 31 signals to constrain the distribution as it is relevant for spin-orbit resonances.
3

The left panel of Fig. 2 shows our constraints on the


mean (µφ1 and µ∆φ ) and width (σφ1 and σ∆φ ) for the φ1
and ∆φ distributions. Our prior for the mean is always
uniform in (−π, π). However, we consider two different Jeffreys-σφ
3
prior choices for the width: (i) A Jeffreys prior [40] that is Flat-σφ
log-uniform in σφ1 and σ∆φ between (0.3, 4π), henceforth Infinite-σφ
referred to as the Jeffreys-σφ prior. (ii) A prior that

p(vf ) [10−3 ]
Globular clusters
is uniform in σφ1 and σ∆φ between (0.3, 4π), henceforth 2
Nuclear star clusters
referred to as the Flat-σφ prior. The Jeffreys prior is a
uninformative prior choice often used for scale parame- Elliptical galaxies
ters [40]. The flat prior may be considered a control case Milky Way
to understand the impact of the prior. 1
For both prior choices, we find in Fig. 2 that the 1D
marginalized posteriors for the width parameters σφ1 and
σ∆φ are dominated by the prior itself. However, the 1D
0 500 1000 1500 2000 2500 3000
posteriors for the mean parameters show a preference
for µφ1 ∼ −π/4 and µ∆φ ∼ ±π. This is reflected in the vf [km/s]
corresponding constraints on the posterior population
distributions, p(φ1 ) and p(∆φ), shown in the right panel Figure 3. Constraints on the kick magnitude population for
of Fig. 2. These represent our constraints on the astro- different prior choices for σφ1 and σ∆φ . Shaded regions show
physical distributions for φ1 and ∆φ; they are generated the central 90% credible bounds for the posterior, while the
by evaluating the von Mises model using draws from the solid lines show the mean. The dashed lines show the 90%
joint posterior of the mean and width parameters. bounds for the prior. For comparison, we show known ranges
We interpret the population constraint in Fig. 2 as for the escape velocities for various types of host environments.
follows. For the Flat-σφ prior, examining the 2D posterior
for µ∆φ − σ∆φ , we note that when σ∆φ → 0, only the
region around µ∆φ ∼ ±π is allowed in the 90% credible leaving one event out from the dataset and repeating our
region. This means that, if there is a sharp peak in the analysis, we check that our results are not driven by any
∆φ population, it is only allowed near ∼ ±π. Similarly, single event.
examining the 2D posterior for µφ1 − σφ1 , we find that One limitation of this work is the restriction to the 31
when σφ1 → 0, there is a preference for µφ1 ∼ −π/4. signals with M & 60M so that we can use the NRSur7dq4
These features get amplified for the Jeffreys-σφ prior, model. We also repeat our analysis for all 46 binary BH
as this prior already prefers small widths. In short, the events from GWTC-2, using the phenomenological wave-
data disfavor peaks at regions other than φ1 ∼ −π/4 form model IMRPhenomTPHM [41] for the remaining 15
and ∆φ ∼ ±π, and this leads to p(φ1 ) and p(∆φ) peaks events. Interestingly, we find that there is some infor-
in these regions. However, the data are not informative mation gain in the width parameters in this case, with
enough to constrain the widths of these peaks. We further a preference for small widths. However, as noted in
note that both populations are still consistent with a Ref. [31], IMRPhenomTPHM can have biases in recovering
uniform distribution at the 90% credible level. the orbital-plane spin angles. Therefore, while we include
these results in the Supplement [38] for completeness, we
It is important to recognize that the location of the φ1
treat Fig. 2 as our main result.
peak in Fig. 2 depends strongly on our choice of reference
point. This is because φ1 changes on the orbital timescale Kick population.— Having constrained the full spin de-
as it is defined with respect to the line-of-separation grees of freedom for the binary BH population, we can
(cf. Fig. 1). On the other hand, ∆φ only changes on now derive constraints on the kick population. We begin
the longer precession time scale, and we find that repeat- by generating one realization of the q, χ1 and χ2 popula-
ing our analysis using spins measured at 20Hz leads to tions (for q, we use the same model that was used in the
consistent results for the ∆φ population [38]. However, initial posterior reweighting). We simply draw one hyper-
the biggest gain in measuring the spins at tref /M = −100 parameter sample from the posterior of the hierarchical
is in the φ1 population constraint, as φ1 is significantly analysis and evaluate the q, χ1 and χ2 population models
better measured there [31]. Constraining both φ1 and ∆φ at that point. Next, we draw a large number samples
is necessary to constrain the kick population below. for q, χ1 and χ2 from this population realization and
For completeness, we include our constraints on the compute the corresponding kick magnitudes using the
spin magnitude and tilt populations, along with full model NRSur7dq4Remnant model [37, 42]. Repeating these steps
hyperparameter posteriors in the Supplement [38]. To over many draws of q, χ1 and χ2 populations, we generate
gain further confidence in our results, we also conduct an ensemble of kick population distributions p(vf ). For
some mock population studies [38], which suggest that at comparison, we also evaluate the prior p(vf ) by repeating
least some φ1 and ∆φ populations can be reliably recov- this procedure using prior hyperparameter samples.
ered at current detector sensitivity. Finally, by iteratively Figure 3 shows the 90% credible constraints on p(vf )
4

for the Jeffreys-σφ and Flat-σφ prior choices. In ad- will be retained for the Jeffreys-σφ (Flat-σφ ) prior. For
max
dition, we consider a prior choice where the φ1 and ∆φ nuclear star clusters, assuming vesc = 600 km/s [45], the
populations are restricted to be uniformly distributed. retention fraction is constrained to 54+19 +16
−17 (57−15 ) % for
We refer to this prior choice as Infinite-σφ , as the other the Jeffreys-σφ (Flat-σφ ) prior. Averaging over the two
priors reduce to this when σφ1 = σ∆φ = ∞. This restricted prior choices, we estimate the retention fraction to be
prior was also used in Ref. [43] to constrain the kick popu- ∼ 3 − 10 % for globular clusters, and ∼ 39 − 73 % for
lation. Note that the mass ratio, spin magnitude and tilt nuclear star clusters. All constraints are quoted at 90%
population models are the same for all three choices. We credibility. Our constraints on the retention fraction are
compare our kick population constraints against fiducial consistent with those of Refs. [43, 47].
escape velocities for globular clusters [44, 45], nuclear star
clusters [45], elliptical galaxies [18] and Milky Way-like Conclusion.— We constrain the distribution of the
galaxies [46]. orbital-plane spin orientations φ1 and ∆φ in the binary
The three prior choices lead to consistent kick popu- black hole population. We find that there is a preference
lations in Fig. 3, with the Infinite-σφ prior leading to for ∆φ ∼ ±π, which can be a signature of spin-orbit reso-
the tightest constraint. This is expected as the Infinite- nances. In addition, we find a preference for φ1 ∼ −π/4
σφ prior is a special case of the other two. This is also in the population, which has not been predicted for any
reflected in the more restrictive p(vf ) prior in Fig. 3 for astrophysical formation channel. Finally, we constrain the
Infinite-σφ . It is somewhat surprising that the kick distribution of recoil kicks in the population, and estimate
population is not hugely influenced by the prior choices the fraction of merger remnants retained by globular and
on φ1 and ∆φ population, even though the kick is known nuclear star clusters.
to be very sensitive to these parameters [21]. This is
Observational evidence for spin-orbit resonances has far
explained by the fact that the φ1 and ∆φ distributions
reaching implications for black hole astrophysics. While
in Fig. 2 are still consistent with a uniform distribution
our population constraints suggest the influence of spin-
at 90% credibility. It will be interesting to see how this
orbit resonances, we are unable to constrain the widths of
changes with future observations.
the φ1 and ∆φ distributions with the current dataset of
Astrophysical implications.— While the location of events. Therefore, further observations are necessary to
the φ1 peak in Fig. 2 is not particularly important, the confirm these trends. With LIGO and Virgo approaching
fact that there is a peak at all is indeed interesting. This their design sensitivities [48], our constraints are certain
feature has not been predicted for any formation channel. to improve in the near future.
Therefore, our naive expectation is that this peak will get
smoothed over as more data are added. However, it will Acknowledgments.— We thank Davide Gerosa and Ka-
be interesting to see if there are alternative explanations. terina Chatziioannou for useful discussions. V.V. is sup-
On the other hand, the preference for ∆φ ∼ ±π in ported by a Klarman Fellowship at Cornell. S.B., M.I.
Fig. 2 is expected in some formation channels where spin- and S.V. acknowledge support of the National Science
orbit resonances [3] are important. In particular, stellar Foundation and the LIGO Laboratory. S.B. is also sup-
binaries with significant supernova kicks and efficient ported by the NSF Graduate Research Fellowship under
stellar tides can be driven towards these resonances [8, 9]. Grant No. DGE-1122374. M.I. is supported by NASA
In the standard scenario where the heavier star becomes through the NASA Hubble Fellowship grant No. HST-
the heavier BH, the ∆φ = ±π resonant mode is expected HF2-51410.001-A awarded by the Space Telescope Science
to be dominant. However, if mass transfer between the Institute, which is operated by the Association of Univer-
two components is significant, a mass-ratio reversal occurs sities for Research in Astronomy, Inc., for NASA, under
and the ∆φ = 0 mode becomes dominant [8]. contract NAS5-26555. This research made use of data,
While our ∆φ population constraint can be interpreted software and/or web tools obtained from the Gravitational
as coming from spin-orbit resonances, this does not yet Wave Open Science Center [49], a service of the LIGO
constitute conclusive evidence for them—especially not Laboratory, the LIGO Scientific Collaboration and the
without a measurement of the width of the distribution Virgo Collaboration. LIGO was constructed by the Cali-
to confirm this feature. Furthermore, as pointed out by fornia Institute of Technology and Massachusetts Institute
Ref. [27], some nonprecessing binaries can be confused of Technology with funding from the National Science
for precessing binaries with ∆φ = ±π. A more detailed Foundation and operates under Cooperative Agreement
analysis may be necessary to break such degeneracies. No. PHY-1764464. Computations were performed on the
Finally, as an application of our p(vf ) constraints in Alice cluster at ICTS; the Nemo cluster at University of
Fig. 3, we estimate the fraction of merger remnants that Wisconsin-Milwaukee, which is supported by NSF Grant
would be retained by various host environments. As- PHY-1626190; the Wheeler cluster at Caltech, which is
max
suming a maximum escape velocity vesc = 100 km/s for supported by the Sherman Fairchild Foundation and by
+5 +4
globular clusters [44, 45], 6−3 (6−3 ) % of the remnants Caltech; and the High Performance Cluster at Caltech.
5

[1] Theocharis A. Apostolatos, Curt Cutler, Gerald J. Suss- 1049–1062 (1983), http://oup.prod.sis.lan/mnras/article-
man, and Kip S. Thorne, “Spin-induced orbital precession pdf/203/4/1049/18223796/mnras203-1049.pdf.
and its modulation of the gravitational waveforms from [15] Manuela Campanelli, Carlos O. Lousto, Yosef Zlochower,
merging binaries,” Phys. Rev. D 49, 6274–6297 (1994). and David Merritt, “Maximum gravitational recoil,” Phys.
[2] Lawrence E. Kidder, “Coalescing binary systems of com- Rev. Lett. 98, 231102 (2007), arXiv:gr-qc/0702133 [GR-
pact objects to postNewtonian 5/2 order. 5. Spin effects,” QC].
Phys. Rev. D 52, 821–847 (1995), arXiv:gr-qc/9506022. [16] J. A. Gonzalez, M. D. Hannam, U. Sperhake, Bernd
[3] Jeremy D. Schnittman, “Spin-orbit resonance and the Bruegmann, and S. Husa, “Supermassive recoil veloci-
evolution of compact binary systems,” Phys. Rev. D 70, ties for binary black-hole mergers with antialigned spins,”
124020 (2004), arXiv:astro-ph/0409174. Phys. Rev. Lett. 98, 231101 (2007), arXiv:gr-qc/0702052
[4] Michael Kesden, Davide Gerosa, Richard O’Shaughnessy, [GR-QC].
Emanuele Berti, and Ulrich Sperhake, “Effective poten- [17] Carlos O. Lousto and Yosef Zlochower, “Hangup Kicks:
tials and morphological transitions for binary black-hole Still Larger Recoils by Partial Spin/Orbit Alignment of
spin precession,” Phys. Rev. Lett. 114, 081103 (2015), Black-Hole Binaries,” Phys. Rev. Lett. 107, 231102 (2011),
arXiv:1411.0674 [gr-qc]. arXiv:1108.2009 [gr-qc].
[5] Davide Gerosa, Michael Kesden, Ulrich Sperhake, [18] David Merritt, Milos Milosavljevic, Marc Favata, Scott A.
Emanuele Berti, and Richard O’Shaughnessy, “Multi- Hughes, and Daniel E. Holz, “Consequences of gravita-
timescale analysis of phase transitions in precessing tional radiation recoil,” Astrophys. J. 607, L9–L12 (2004),
black-hole binaries,” Phys. Rev. D 92, 064016 (2015), arXiv:astro-ph/0402057 [astro-ph].
arXiv:1506.03492 [gr-qc]. [19] Vijay Varma, Maximiliano Isi, and Sylvia Biscoveanu,
[6] Khun Sang Phukon, Anuradha Gupta, Sukanta Bose, “Extracting the Gravitational Recoil from Black Hole
and Pankaj Jain, “Effect of orbital eccentricity on the Merger Signals,” Phys. Rev. Lett. 124, 101104 (2020),
dynamics of precessing compact binaries,” Phys. Rev. D arXiv:2002.00296 [gr-qc].
100, 124008 (2019), arXiv:1904.03985 [gr-qc]. [20] Davide Gerosa and Maya Fishbach, “Hierarchical mergers
[7] R. Abbott et al. (LIGO Scientific, Virgo), “Population of stellar-mass black holes and their gravitational-wave
Properties of Compact Objects from the Second LIGO- signatures,” (2021), arXiv:2105.03439 [astro-ph.HE].
Virgo Gravitational-Wave Transient Catalog,” Astrophys. [21] Bernd Bruegmann, Jose A. Gonzalez, Mark Hannam,
J. Lett. 913, L7 (2021), arXiv:2010.14533 [astro-ph.HE]. Sascha Husa, and Ulrich Sperhake, “Exploring black
[8] Davide Gerosa, Michael Kesden, Emanuele Berti, Richard hole superkicks,” Phys. Rev. D77, 124047 (2008),
O’Shaughnessy, and Ulrich Sperhake, “Resonant-plane arXiv:0707.0135 [gr-qc].
locking and spin alignment in stellar-mass black-hole bi- [22] J. Aasi et al. (LIGO Scientific), “Advanced LIGO,” Class.
naries: a diagnostic of compact-binary formation,” Phys. Quant. Grav. 32, 074001 (2015), arXiv:1411.4547 [gr-qc].
Rev. D87, 104028 (2013), arXiv:1302.4442 [gr-qc]. [23] F. Acernese et al. (Virgo), “Advanced Virgo: a second-
[9] Davide Gerosa, Emanuele Berti, Richard O’Shaughnessy, generation interferometric gravitational wave detector,”
Krzysztof Belczynski, Michael Kesden, Daniel Wysocki, Class. Quant. Grav. 32, 024001 (2015), arXiv:1408.3978
and Wojciech Gladysz, “Spin orientations of merging [gr-qc].
black holes formed from the evolution of stellar bina- [24] B. P. Abbott et al. (LIGO Scientific, Virgo), “Binary Black
ries,” Phys. Rev. D98, 084036 (2018), arXiv:1808.02491 Hole Population Properties Inferred from the First and
[astro-ph.HE]. Second Observing Runs of Advanced LIGO and Advanced
[10] More precisely, we use the coorbital frame defined in Virgo,” (2018), arXiv:1811.12940 [astro-ph.HE].
Ref. [37]. In this frame, the z-axis is along the direction [25] Salvatore Vitale, Ryan Lynch, John Veitch, Vivien Ray-
that maximises the power in the (2,2) mode, which is taken mond, and Riccardo Sturani, “Measuring the spin of
to be the direction of the orbital angular momentum [50]. black holes in binary systems using gravitational waves,”
The x-axis is along the line of separation from the lighter Phys. Rev. Lett. 112, 251101 (2014), arXiv:1403.0129
to the heavier BH, and the y-axis completes the right- [gr-qc].
handed triad. Note that this frame is defined using the [26] P. Schmidt, F. Ohme, and M. Hannam, “Towards mod-
gauge-invariant waveform at future null infinity, rather els of gravitational waveforms from generic binaries II:
than the gauge-dependent BH trajectories. (). Modelling precession effects with a single effective pre-
[11] W. B. Bonnor, M. A. Rotenberg, and Rosenhead Louis, cession parameter,” Phys. Rev. D 91, 024043 (2015),
“Transport of momentum by gravitational waves: the arXiv:1408.1810 [gr-qc].
linear approximation,” Proceedings of the Royal Society [27] Sylvia Biscoveanu, Maximiliano Isi, Vijay Varma, and
of London Series A. 265 (1961), 10.1098/rspa.1961.0226. Salvatore Vitale, “Measuring the spins of heavy binary
[12] Asher Peres, “Classical radiation recoil,” Phys. Rev. 128, black holes,” (2021), arXiv:2106.06492 [gr-qc].
2471–2475 (1962). [28] Davide Gerosa, Richard O’Shaughnessy, Michael Kesden,
[13] Jacob D. Bekenstein, “Gravitational-Radiation Recoil and Emanuele Berti, and Ulrich Sperhake, “Distinguishing
Runaway Black Holes,” The Astrophysical Journal 183, black-hole spin-orbit resonances by their gravitational-
657–664 (1973). wave signatures,” Phys. Rev. D 89, 124025 (2014),
[14] M. J. Fitchett, “The influence of gravitational arXiv:1403.7147 [gr-qc].
wave momentum losses on the centre of mass [29] Daniele Trifirò, Richard O’Shaughnessy, Davide Gerosa,
motion of a Newtonian binary system,” Monthly Emanuele Berti, Michael Kesden, Tyson Littenberg, and
Notices of the Royal Astronomical Society 203, Ulrich Sperhake, “Distinguishing black-hole spin-orbit
6

resonances by their gravitational wave signatures. II: Full hole binaries in galactic nuclei: implications for advanced-
parameter estimation,” Phys. Rev. D 93, 044071 (2016), LIGO detections,” Astrophys. J. 831, 187 (2016),
arXiv:1507.05587 [gr-qc]. arXiv:1606.04889 [astro-ph.HE].
[30] Chaitanya Afle et al., “Detection and characterization [46] G. Monari, B. Famaey, I. Carrillo, T. Piffl, M. Steinmetz,
of spin-orbit resonances in the advanced gravitational R. F. G. Wyse, F. Anders, C. Chiappini, and K. Janßen,
wave detectors era,” Phys. Rev. D 98, 083014 (2018), “The escape speed curve of the Galaxy obtained from Gaia
arXiv:1803.07695 [gr-qc]. DR2 implies a heavy Milky Way,” A&A 616, L9 (2018),
[31] Vijay Varma, Maximiliano Isi, Sylvia Biscoveanu, Will arXiv:1807.04565 [astro-ph.GA].
Farr, and Salvatore Vitale, “Measuring binary black hole [47] Parthapratim Mahapatra, Anuradha Gupta, Marc Fa-
orbital-plane spin orientations,” (2021), in preparation. vata, K. G. Arun, and B. S. Sathyaprakash, “Remnant
[32] R. Abbott et al. (LIGO Scientific, Virgo), “GWTC-2: black hole kicks and implications for hierarchical mergers,”
Compact Binary Coalescences Observed by LIGO and (2021), arXiv:2106.07179 [astro-ph.HE].
Virgo During the First Half of the Third Observing Run,” [48] B. P. Abbott et al. (KAGRA, LIGO Scientific, VIRGO),
Phys. Rev. X 11, 021053 (2021), arXiv:2010.14527 [gr-qc]. “Prospects for Observing and Localizing Gravitational-
[33] B. P. Abbott et al. (LIGO Scientific, Virgo), “GWTC- Wave Transients with Advanced LIGO, Advanced
1: A Gravitational-Wave Transient Catalog of Compact Virgo and KAGRA,” Living Rev. Rel. 21, 3 (2018),
Binary Mergers Observed by LIGO and Virgo during arXiv:1304.0670 [gr-qc].
the First and Second Observing Runs,” Phys. Rev. X9, [49] LIGO Scientific Collaboration and Virgo Collaboration,
031040 (2019), arXiv:1811.12907 [astro-ph.HE]. “Gravitational Wave Open Science Center,” https://www.
[34] Rich Abbott et al. (LIGO Scientific, Virgo), “Open data gw-openscience.org.
from the first and second observing runs of Advanced [50] Michael Boyle, Robert Owen, and Harald P. Pfeiffer, “A
LIGO and Advanced Virgo,” SoftwareX 13, 100658 (2021), geometric approach to the precession of compact binaries,”
arXiv:1912.11716 [gr-qc]. Phys. Rev. D84, 124011 (2011), arXiv:1110.2965 [gr-qc].
[35] LIGO Scientific Collaboration and Virgo Collaboration, [51] Eric Thrane and Colm Talbot, “An introduction to
“GWTC-1,” https://doi.org/10.7935/82H3-HH23 (2018). Bayesian inference in gravitational-wave astronomy: Pa-
[36] LIGO Scientific Collaboration and Virgo Collaboration, rameter estimation, model selection, and hierarchical mod-
“GWTC-2,” https://doi.org/10.7935/99gf-ax93 (2020). els,” Publications of the Astronomical Society of Australia
[37] Vijay Varma, Scott E. Field, Mark A. Scheel, Jonathan 36, e010 (2019), arXiv:1809.02293 [astro-ph.IM].
Blackman, Davide Gerosa, Leo C. Stein, Lawrence E. Kid- [52] Shanika Galaudage, Colm Talbot, and Eric Thrane,
der, and Harald P. Pfeiffer, “Surrogate models for precess- “Gravitational-wave inference in the catalog era: evolving
ing binary black hole simulations with unequal masses,” priors and marginal events,” Phys. Rev. D 102, 083026
Phys. Rev. Research. 1, 033015 (2019), arXiv:1905.09300 (2020), arXiv:1912.09708 [astro-ph.HE].
[gr-qc]. [53] I. M. Romero-Shaw et al., “Bayesian inference for compact
[38] See Supplemental Material here, for details about our hi- binary coalescences with bilby: validation and applica-
erarchical analysis, posterior for the full spin population, tion to the first LIGO–Virgo gravitational-wave transient
population constraints at 20 Hz, a mock population study, catalogue,” Mon. Not. Roy. Astron. Soc. 499, 3295–3319
and results for all GWTC-2 events, using a phenomeno- (2020), arXiv:2006.00714 [astro-ph.IM].
logical model when the surrogate is not applicable. This [54] Gregory Ashton et al., “BILBY: A user-friendly Bayesian
further includes Refs. [51–59] (). inference library for gravitational-wave astronomy,” Astro-
[39] K. V. Mardia and P. E. Jupp, Directional Statistics (John phys. J. Suppl. 241, 27 (2019), arXiv:1811.02042 [astro-
Wiley and Sons, Ltd., West Sussex, 1999). ph.IM].
[40] Harold Jeffreys, “An Invariant Form for the Prior Proba- [55] Joshua S. Speagle, “DYNESTY: a dynamic nested sam-
bility in Estimation Problems,” Proceedings of the Royal pling package for estimating Bayesian posteriors and ev-
Society of London Series A 186, 453–461 (1946). idences,” Monthly Notices of the Royal Astronomical
[41] Héctor Estellés, Marta Colleoni, Cecilio García-Quirós, Society 493, 3132–3158 (2020), arXiv:1904.02180 [astro-
Sascha Husa, David Keitel, Maite Mateu-Lucena, Maria ph.IM].
de Lluc Planas, and Antoni Ramos-Buades, “New twists [56] Daniel Wysocki, Jacob Lange, and Richard
in compact binary waveform modelling: a fast time do- O’Shaughnessy, “Reconstructing phenomenological
main model for precession,” (2021), arXiv:2105.05872 distributions of compact binaries via gravitational
[gr-qc]. wave observations,” Phys. Rev. D 100, 043012 (2019),
[42] Vijay Varma, Davide Gerosa, Leo C. Stein, François arXiv:1805.06442 [gr-qc].
Hébert, and Hao Zhang, “High-accuracy mass, spin, and [57] Colm Talbot and Eric Thrane, “Determining the popu-
recoil predictions of generic black-hole merger remnants,” lation properties of spinning black holes,” Phys. Rev. D
Phys. Rev. Lett. 122, 011101 (2019), arXiv:1809.09125 96, 023012 (2017), arXiv:1704.08370 [astro-ph.HE].
[gr-qc]. [58] R. Abbott et al. (LIGO Scientific, Virgo), “GW190521: A
[43] Zoheyr Doctor, Ben Farr, and Daniel E. Holz, “Black Hole Binary Black Hole Merger with a Total Mass of 150 M ,”
Leftovers: The Remnant Population from Binary Black Phys. Rev. Lett. 125, 101102 (2020), arXiv:2009.01075
Hole Mergers,” (2021), arXiv:2103.04001 [astro-ph.HE]. [gr-qc].
[44] Oleg Y. Gnedin, HongSheng Zhao, J. E. Pringle, [59] J. Veitch et al., “Robust parameter estimation for compact
S. Michael Fall, Mario Livio, and Georges Meylan, “The binaries with ground-based gravitational-wave observa-
unique history of the globular cluster omega centauri,” tions using the LALInference software library,” Phys. Rev.
Astrophys. J. Lett. 568, L23–L26 (2002), arXiv:astro- D 91, 042003 (2015), arXiv:1409.7215 [gr-qc].
ph/0202045.
[45] Fabio Antonini and Frederic A. Rasio, “Merging black
7

SUPPLEMENTAL MATERIALS described in Ref. [52]. This ensures that the event j is
not double-counted in the analysis.
I. HIERARCHICAL ANALYSIS We perform an additional reweighting to switch to a
more astrophysically motivated redshift prior. Note that,
A. Parameter estimation for individual events when obtaining the posterior samples in Eq. (S1), Ref. [31]
used a prior that is uniform in comoving volume [32, 53]:
Binary BH parameters can be estimated from GW sig- dVc
nals following Bayes’ theorem (see Ref. [51] for a review): π(z) ∝ , (S3)
dz
p(Θ|d) ∝ L(d|Θ) π(Θ), (S1) where dVc /dz is the differential comoving volume. We
now apply the weights
where p(Θ|d) is the posterior probability distribution
of the binary parameters Θ given the observed data d, w(zj ) = (1 + zj )−1 . (S4)
L(d|Θ) is the likelihood of the data given Θ, and π(Θ) is
the prior probability distribution for Θ. The full set of to these posteriors, effectively switching to a prior that
binary parameters Θ is 15 dimensional [32], and includes that is uniform in comoving volume and source frame
the masses and spins of the component BHs as well as time [32, 53]:
extrinsic properties such as the distance and sky location.
In this paper, we use the parameter estimates for dVc
GWTC-2 events from Ref. [31]. Ref. [31] analyzed π(z) ∝ (1 + z)−1 . (S5)
dz
31 events with total mass M & 60M (listed in
Tab.I of Ref. [31]) using the NRSur7dq4 [37] model at The additional (1 + z)−1 factor accounts for cosmological
both tref /M = −100 and fref = 20 Hz. In addition, time dilation. This matches the redshift prior assumed
IMRPhenomTPHM [41] was used to analyze all 46 GWTC-2 for the “Power Law + Peak” model in Ref. [7].
binary BH events at fref = 20 Hz. Our main results use We perform both reweighting steps simultaneously, by
the NRSur7dq4 posteriors at tref /M = −100. However, the applying the weights
spin measurements at fref = 20 Hz are used in Sec. II B
and Sec. II D for comparison. w(Θj ) = w(Γj ) w(zj ), (S6)

to the posterior samples Θj for each event j in our dataset.


B. Reweighting to an astrophysical mass and Post reweighting, Eq. (S1) can be rewritten as
redshift prior
p(Θ|d, R) ∝ L(d|Θ) π(Θ|R), (S7)
We are interested in constraining the population for the
full spin degrees of freedom, S = {χ1 , χ2 , θ1 , θ2 , φ1 , ∆φ} where R indicates that these are the mass, mass-ratio and
(cf. Fig. 1). Recently, Ref. [7] analyzed the GWTC-2 cat- redshift reweighted posteriors.
alog to place constraints on the populations of component
BH masses, spin magnitudes and tilts (but not φ1 and
∆φ). For simplicity, we only model the spin degrees of C. Hierarchical parameter estimation
freedom in this work, but incorporate the astrophysical
mass and mass-ratio population constraints from Ref. [7]. In order to constrain the population of S =
We denote Γ = {msrc src
1 , q}, where m1 = m1 /(1 + z) is {χ1 , χ2 , θ1 , θ2 , φ1 , ∆φ}, which is a subset of Θ, we make
the mass of the heavier BH in the source frame, and z the assumption that the true value of S for each event
is the source redshift. To account for the astrophysical is drawn from a common underlying distribution π(S|Λ),
constraints on Γ, we apply the following weights to the which is conditional on a set of hyperparameters Λ. We use
posterior samples for each individual event the notation π(S|Λ) as this embodies our prior assumption
for the underlying distribution.
p(Γj |{d}i6=j ) Once again, we employ Bayes’ theorem to infer the
w(Γj ) = , (S2)
π(Γj ) hyperparameters Λ from the data (see Ref. [51] for a
review):
where j indicates the particular GW event, π() is the
same prior as in Eq. (S1), and p(Γj |{d}i6=j ) denotes the p(Λ|{di }) ∝ L({di }|Λ) π(Λ), (S8)
posterior population distribution (cf. Eq. (46) of Ref. [52])
for the Γ population obtained using the data from all where p(Λ|{di }) is the hyper-posterior distribution for Λ
events other than j. We use the public data release given a set of observations {di }, L({di }|Λ) is the hyper-
for the “Power Law + Peak” model from Ref. [7] for likelihood of this dataset given Λ, and π(Λ) is the hyper-
the Γ population constraints. p(Γj |{d}i6=j ) is obtained prior distribution for Λ. In our case, {di } with i = 1 . . . N
from these results using the “leave-one-out” computation represents the observed data for our set of N = 31 GW
8

events. The hyper-likelihood is obtained by coherently Parameter Prior


combining the data from from all events [51]: µχ U(0, 1)
σχ2 U(0, 0.25)
N Z
Y π(Si | Λ) ξθ U(0,1)
L({di }|Λ) ∝ dSi p(Si |di , R) . (S9)
i
π(Si |R) σθ U(0.01,4)
µφ 1 U(−π, π)
We use the Bilby [54] package with the dynesty [55] µ∆φ U(−π, π)
sampler to draw posterior samples for the hyperparame- σφ1 See Tab. S2.
ters Λ from p(Λ|{di }). Finally, the posterior distribution σ∆φ See Tab. S2.
for the S population, also referred to as the posterior
population distribution, is obtained by averaging over Λ: Table S1. Priors on hyperparameters for our spin population
Z model. In addition, following Ref. [7], we exclude µχ , σχ2 values
p(S) = dΛ π(S|Λ) p(Λ|{di }). (S10) where the Beta distribution becomes singular. Here, U(a, b)
indicates a uniform random distribution on the interval (a, b).

In practice, this is done by drawing samples from the


hyper-posterior p(Λ|{di }) and evaluating π(S|Λ) on an Name Prior on σφ1 and σ∆φ
array of S values for each Λ sample. This gives us an Jeffreys-σφ LU(0.3, 4π)
ensemble of probability distributions on S, which we use Flat-σφ U(0.3, 4π)
to compute the mean and 90% credible widths. Infinite-σφ δ(∞)

Table S2. The prior choices we consider for σφ1 and σ∆φ .
D. Population model LU(a, b) indicates a log-uniform random distribution on the
interval (a, b), while δ(a) indicates a Dirac delta distribution
where the parameter is fixed at a. Note that the Infinite-
We use the following joint distribution for the underly-
σφ prior restricts the φ1 and ∆φ populations to be uniform
ing spin distribution: random.
π(S|Λ) = p(χ1,2 | µχ , σχ2 ) p(θ1,2 | ξθ , σθ )
p(φ1 | µφ1 , σφ1 ) p(∆φ | µ∆φ , σ∆φ ). (S11) E. Correlations and selection effects

Here, p(χ1,2 | µχ , σχ2 ) is a Beta distribution in the spin


magnitudes, parameterized by its mean µχ and variance Ref. [7] models the mass degrees of freedom simulta-
σχ2 [56], and p(θ1,2 | ξθ , σθ ) is an isotropic tilt distribution neously with the spin degrees of freedom. Using the
with a Gaussian peak component, parameterized by the mass-population-reweighted posteriors (cf. Sec. I B) is
standard deviation σθ of the Gaussian and the mixing equivalent to Ref. [7], except that this does not account
fraction ξθ coming from the Gaussian component [57]. for any possible correlations between the mass and spin
Note that we assume the distributions for the two com- hyperparameters. However, Ref. [7] found that these
ponent BHs are the same for the spin magnitude and correlations are not significant.
tilt. This model for the spin magnitudes and tilts is the Ref. [7] also distinguishes between between the astro-
same as “Default spin” model described in App.D.1 of physical distribution of a parameter— the distribution
Ref. [7]. On top of this model, we include p(φ1 | µφ1 , σφ1 ) as it is in nature—and the observed distribution of a pa-
and p(∆φ | µ∆φ , σ∆φ ) as independent von Mises distri- rameter—the distribution as it appears among detected
butions parameterized by their corresponding mean and events due to selection effects, because of which binaries
standard deviations. Our choices for the hyper-prior π(Λ) with certain parameters may be easier to detect than
imposed on the hyperparameters are described in Tab. S1. others. These effects can be accounted for by modifying
The von Mises distribution [39] is defined as the hyper-likelihood in Eq. (S9) as done in Eq. (1) of
Ref. [7]. While Ref. [7] includes selection effects for their
exp (κ cos (φ − µ)) mass population models, they are ignored for the “Default
p(φ|µ, κ) = , (S12) spin” model as they are not expected to be significant at
2π I0 (κ)
current detector sensitivity.
where µ is the mean, κ is a shape parameter, and I0 is In our case, the mass-reweighted posteriors (cf. Sec. I B)
the modified Bessel function of order 0. The von Mises already account for selection effects for the mass popu-
distribution is a close approximation of a Gaussian with lation. As we simply extend the spin population model
periodic boundary conditions at φ = ±π, making it an of Ref. [7] with the φ1 and ∆φ models, we also ignore
appropriate choice for phase parameters like φ1 and ∆φ. selection effects for our spin population model. As noted
The variance of the von Mises distribution can be approx- in Ref. [7], it will be important to include spin selection
imated as 1/κ;
√ therefore we define the standard deviation effects as detectors sensitivity improves. However, at cur-
to be σ ≡ 1/ κ. rent sensitivity, assuming spin selection effects are not
9

p(µχ )

4
3

2
2

p(χ)
p(σχ2 )
1
2

16
0.
σχ2

0
1

8
0.0 0.2 0.4 0.6 0.8 1.0
0.

p(ξθ ) χ
1.00
8
0.

0
1.
ξθ

0.75

p(cos θ)
4
0.

5
0.
p(σθ ) 0.50

0.25

50
0

0.
3.
σθ

0.00
−1.0 −0.5 0.0 0.5 1.0

25
5
1.

0.
cos θ
p(µφ1 )

30
5
2.
µφ1 [rad]

0.
0

15
0.

0.
p(µ∆φ )
5
2.

5
µ∆φ [rad]
2.

2
0.
0
0.

1
0.
p(σφ1 )
5
2.
σφ1 [rad] −

4
10

0.
2
5

0.
p(σ∆φ )
σ∆φ [rad]
10

4
0.
5

2
0.
4

5
5

10

10
0.

0.

0.

0.

0.

0.

1.

3.

2.

0.

2.
2.

0.

2.

µχ σχ2 ξθ σθ σφ1 [rad] σ∆φ [rad]


µφ1 [rad] µ∆φ [rad]

Figure S1. Posterior distribution for hyperparameters for our spin model with a Jeffreys prior for σφ1 and σ∆φ . The shaded
regions in the lower-triangle subplots represent 50% and 90% credible bounds on joint 2d posteriors. The diagonal subplots show
the 1d marginalized posteriors and priors (black dashed lines). In the top-right, we show constraints on the spin magnitude and
tilt populations. Shaded regions show the central 90% credible bounds, while the solid lines show the mean. The dashed grey
lines show the 90% prior bounds. The corresponding population constraints on φ1 and ∆φ are shown in the top half of the
right-panel of Fig. 2.

significant (as done in Ref. [7]), our results can be treated II. ADDITIONAL INVESTIGATIONS
as constraints on the astrophysical distribution rather
than the observed distribution. A. Full spin population

In Fig. 2, we only show the population constraints on


φ1 and ∆φ. For completeness, we now show the hyper-
parameter posteriors and population constraints on the
spin magnitudes and tilts in Fig. S1 (for the Jeffreys-σφ
prior) and Fig. S2 (for the Flat-σφ prior). Our popu-
10

p(µχ )

4
3

2
2

p(χ)
p(σχ2 )
1

16
2
0.
σχ2

8
1

0.0 0.2 0.4 0.6 0.8 1.0


0.

p(ξθ ) χ
1.00
8

2
0.

1.
ξθ

0.75

p(cos θ)
4

6
0.

0.
p(σθ ) 0.50

0.25

50
0
3.

0.
σθ

0.00
−1.0 −0.5 0.0 0.5 1.0

25
5
1.

0.
cos θ
p(µφ1 )
5

16
2.
µφ1 [rad]

0.
0
0.

08
0.
p(µ∆φ )
5
2.

5
µ∆φ [rad]
2.

16
0.
0
0.

08
0.
p(σφ1 )
5
2.
σφ1 [rad] −

08
10

0.
04
5

0.
p(σ∆φ )
σ∆φ [rad]

08
10

0.
04
5

0.
4

5
5

10

10
0.

0.

0.

0.

0.

0.

1.

3.

2.

0.

2.
2.

0.

2.

µχ σχ2 ξθ σθ σφ1 [rad] σ∆φ [rad]


µφ1 [rad] µ∆φ [rad]

Figure S2. Same as Fig. S1, but using a Flat prior for σφ1 and σ∆φ . The corresponding φ1 and ∆φ population constraints are
shown in the bottom half of the right-panel of Fig. 2.

lation constraints on the spin magnitudes and tilts are improvement in φ1 measurements for individual events at
consistent with Ref. [7], but somewhat broader as we only tref /M = −100 (cf. Fig.3 of Ref. [31]), the φ1 population
use 31 of the available 46 binary BH events in GWTC-2. is much better constrained in Fig. 2 compared to Fig. S3.
It is important to note that this does not imply that the
astrophysical φ1 distribution is flatter at fref = 20 Hz com-
B. Population constraints at fref = 20 Hz pared to tref /M = −100. Instead, this is because the φ1
measurements at fref = 20 Hz are very poor. In fact, even
Figure S3 shows constraints on the φ1 and ∆φ popu- the mild peak near φ1 ∼ 0 for the Jeffreys-σφ prior in
lations when measured at fref = 20 Hz. Because the ∆φ Fig. S3 is driven entirely by GW190521 [58]. As discussed
measurements for individual events do not change signifi- in Ref. [31], GW190521 is the only event with a good
cantly between fref = 20 Hz and tref /M = −100 (cf. Fig.4 of measurement of φ1 at fref = 20 Hz as this binary happens
Ref. [31]), the ∆φ populations are also consistent between to merge near 20 Hz.
Fig. S3 and Fig. 2. By contrast, as there is significant
11

ity; however, we rescale the injected distance such that the


φ1 ∆φ
0.75 Jeffreys-σφ SNR matches that of the observed event. Therefore, our
mock populations approximately mimic the parameters
and detector sensitivity for these events.
0.50
p(φ)

Figure S4 shows the results from our hierarchical analy-


sis on these mock populations. In both cases, the φ1 and
0.25 ∆φ populations show a clear preference for the region near
the injected values. Interestingly, the ∆φ = 0 population
is better recovered than the ∆φ = π population. This
Flat-σφ φ [rad] suggests that it is easier to constrain ∆φ for binaries with
0.20 ∆φ = 0, in agreement with Refs [28, 29].
D. Results using all 46 GWTC-2 events
p(φ)

0.15
All results shown so far were restricted to the 31 sig-
nals with M & 60M so that we can use the NRSur7dq4
0.10 model. We now repeat our analysis for all 46 binary BH
events from GWTC-2, using the IMRPhenomTPHM poste-
−π −π/2 0 π/2 π riors from Ref. [31] for the remaining 15 events (listed
φ [rad] in Tab. II of Ref. [31]). We use spins at fref = 20 Hz for
all events, as spins at tref /M = −100 are not available for
IMRPhenomTPHM. Figure S5 shows constraints on the ∆φ
Figure S3. Same as Fig. 2, but for spins measured at fref = population using using all 46 events. While we simulta-
20 Hz. While the ∆φ constraints are consistent between fref = neously model all spin degrees of freedom, we only show
20 Hz and tref /M = −100, φ1 is much better constrained at the ∆φ population for simplicity. Similar to Sec. II B, the
tref /M = −100.
φ1 population is not well constrained when the spins are
measured at fref = 20 Hz.
C. Mock population study The left-panel of Fig. S5 shows the posteriors for the
µ∆φ and σ∆φ parameters. Compared to Fig. 2, we now see
To test the fidelity of the von Mises model in recovering that the 1d σ∆φ posterior is distinguishable from the prior
φ1 and ∆φ populations, we conduct a mock population for both prior choices. In particular, there is a preference
study. For each of our 31 events, we pick the maximum- for small widths, while the µ∆φ distribution still peaks
likelihood posterior sample, but we rotate the orbital- at ∼ ±π. As shown in the right-panel of Fig. S5, this
plane spin angles and set them to constant values at leads to a stronger peak near ∼ ±π in the ∆φ population,
tref /M = −100 for all events. We construct two such compared to Fig. 2.
populations, one with φ1 = −π/2 and ∆φ = ±π and While this reinforces our results using only NRSur7dq4,
another with φ1 = π/2 and ∆φ = 0. We inject the it is important to consider that IMRPhenomTPHM can have
corresponding signals in simulated detector noise and biases in recovering the orbital-plane spin angles [31]. In
recover the spin population using our hierarchical analysis. addition, Ref. [31] found significant differences between
We use the NRSur7dq4 waveform model for the injec- the ∆φ posteriors for NRSur7dq4 and IMRPhenomTPHM for
tions as well as the parameter inference (with the LALIn- GW190521 [58]. This suggests that a detailed study of
ference package [59]). The signals are injected into noise the impact of waveform systematics on the ∆φ population
from a simulated LIGO-Virgo network at design sensitiv- is necessary. We leave this exploration to future work.
12

Fixed φ1 = −π/2, ∆φ = ±π Fixed φ1 = π/2, ∆φ = 0

1.5 Jeffreys-σφ φ1 ∆φ 1.5 Jeffreys-σφ φ1 ∆φ

1.0 1.0
p(φ)

p(φ)
0.5 0.5

Flat-σφ φ [rad] Flat-σφ φ [rad]


1.0 1.0
p(φ)

p(φ)
0.5 0.5

−π −π/2 0 π/2 π −π −π/2 0 π/2 π


φ [rad] φ [rad]

Figure S4. Same as Fig. 2, but for the two mock populations described in Sec. II C. The left panels correspond to injections
with fixed φ1 = −π/2 and ∆φ = ±π at tref /M = −100, while the right panels have fixed φ1 = π/2 and ∆φ = 0. The injected
values are indicated by star markers. For both mock populations, the true values are reasonably well recovered, but the ∆φ = 0
population is better constrained than the ∆φ = ±π one, especially for the Flat-σφ prior.

Jeffreys-σφ ∆φ
1.00
30

0.75
0.

p(φ)

Jeffreys-σφ
Flat-σφ 0.50
15
0.

0.25

φ [rad]
Flat-σφ
0.4
8
0.

σ∆φ [rad]

0.3
p(φ)

0.2
4
0.
π

0.1
π

−π −π/2 0 π/2 π
0


2

2

µ∆φ [rad] σ∆φ [rad] φ [rad]

Figure S5. Constraints on the ∆φ population at fref = 20 Hz using all 46 binary BH events from GWTC-2. We use the NRSur7dq4
model for the 31 events with M & 60M , and the IMRPhenomTPHM model for the remaining 15 events. Left: Posteriors for the
mean and width parameters for ∆φ. The shaded regions show 50% (dark shade) and 90% (light shade) credible bounds on the
joint 2d posterior. The top and right subplots show 1d marginalized posteriors as solid lines. The prior on µ∆φ is shown as
a dashed black line, while the two prior choices for σ∆φ are shown as colored dashed lines. Unlike Fig. 2, we now see some
information gain in the 1d posteriors for the width parameters, with a preference towards small widths. Right: Corresponding
constraints on the p(∆φ) population distribution. The peak at ∆φ ∼ ±π is amplified compared to Fig. 2.

You might also like