Testing General Relativity Using Higher

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Testing general relativity using higher-order modes of gravitational waves from binary

black holes

Anna Puecher1,2 , Chinmay Kalaghatgi1,2,3 , Soumen Roy1,2 , Yoshinta


Setyawati1,2 , Ish Gupta4 , B.S. Sathyaprakash4,5,6 , and Chris Van Den Broeck1,2
1
Nikhef – National Institute for Subatomic Physics,
Science Park 105, 1098 XG Amsterdam, The Netherlands
2
Institute for Gravitational and Subatomic Physics (GRASP),
Utrecht University, Princetonplein 1, 3584 CC Utrecht, The Netherlands
3
Institute for High-Energy Physics, University of Amsterdam,
Science Park 904, 1098 XH Amsterdam, The Netherlands
4
Institute for Gravitation and the Cosmos, Department of Physics,
Penn State University, University Park, PA 16802, USA
5
Department of Astronomy and Astrophysics, Penn State University, University Park, PA 16802, USA and
6
School of Physics and Astronomy, Cardiff University, Cardiff, CF24 3AA, United Kingdom
(Dated: May 19, 2022)
Recently, strong evidence was found for the presence of higher-order modes in the gravitational
wave signals GW190412 and GW190814, which originated from compact binary coalescences with
significantly asymmetric component masses. This has opened up the possibility of new tests of gen-
arXiv:2205.09062v1 [gr-qc] 18 May 2022

eral relativity by looking at the way in which the higher-order modes are related to the basic signal.
Here we further develop a test which assesses whether the amplitudes of sub-dominant harmonics
are consistent with what is predicted by general relativity. To this end we incorporate a state-of-
the-art waveform model with higher-order modes and precessing spins into a Bayesian parameter
estimation and model selection framework. The analysis methodology is tested extensively through
simulations. We investigate to what extent deviations in the relative amplitudes of the harmonics
will be measurable depending on the properties of the source, and we map out correlations between
our testing parameters and the inclination of the source with respect to the observer. Finally, we
apply the test to GW190412 and GW190814, finding no evidence for violations of general relativity.

I. INTRODUCTION categories. In the first case, one tests the phase evolu-
tion, e.g. by testing for deviations in the way parameters
The Advanced LIGO [1] and Advanced Virgo [2] grav- like the chirp mass and symmetric mass ratio enter into
itational wave (GW) observatories have by now detected the expressions for the different harmonics [33]. This
90 candidate signals from coalescing binary black holes kind of test has already been applied to GW190412 and
[3–8], binary neutron stars [9, 10], and neutron star-black GW190814 in Ref. [37]. A second test looks for anoma-
hole systems [11]. A battery of tests of general relativ- lies in the amplitudes of the sub-dominant modes [34];
ity (GR) were performed [12–17], including tests of the the latter test is the focus of this paper.
spacetime dynamics as inferred from the binary coales- Specifically, defining h(t) ≡ h+ (t)−ih× (t) with h+ , h×
cence process [18–24]. the two polarizations, the GW signal from a coalescing
Recently, strong evidence was obtained for the pres- binary can be written as
ence of higher-order modes [25] in the gravitational wave
signals GW190412 and GW190814 [26–28], which were
emitted by coalescing binary compact objects with sig- ℓ
∞ X
nificantly different component masses. Measuring these h(t; ι, φ0 , ~λ) =
X
ℓm
Y−2 (ι, φ0 ) hℓm (t; ~λ), (1)
sub-dominant harmonics of the basic signal enables more
ℓ=2 m=−ℓ
precise measurements of the source parameters, and can
allow for stronger constraints on certain deviations from
GR [29, 30]. Several tests of GR that directly probe the ℓm
harmonic structure for binary black hole (BBH) coales- where the Y−2 are spin-weighted spherical harmonics of
cences1 were proposed in [33–36]. These fall into two weight −2, (ι, φ0 ) indicate the direction of the radiation
in the source frame, and ~λ collects all other parameters in
the problem. The latter are the total mass M ≡ m1 +m2
(with m1 , m2 the component masses), the mass ratio
1 Given the low mass of the lighter component of GW190814 (≃ q ≡ m1 /m2 (where we assume m1 ≥ m2 ), the dimen-
2.6 M⊙ ), there is a possibility that it was a signal from a neutron
sionless spin vectors S1 and S2 at some reference time
star-black hole rather than a binary black hole coalescence [27],
but studies based on the known properties of neutron stars make tref , a reference phase ϕref , and the luminosity distance
a BBH origin much more likely [31, 32]. For the purposes of DL . Taking the contribution with ℓ = 2, m = ±2 to
this paper we will assume that GW190814 came from a BBH constitute the fundamental mode in the signal, the test
coalescence. of GR considered here follows Ref. [34] to allow for devi-
2

ations in the amplitudes of the higher-order modes: • At zeroth post-Newtonian order (0PN) in ampli-
tude there is the harmonic with ℓ = |m| = 2, which
h(t; ι, φ0 , ~λ) is the most dominant of all multipole modes.
(ι, φ0 ) h2m (t; ~λ)
X
2m
= Y−2
• At 0.5PN order in amplitude, harmonics with
m=±2
(ℓ, |m|) = (2, 1), (3, 3), (3, 1) appear. In this pa-

per we will be the most interested in the (2, 1) and
(ι, φ0 ) hℓm (t; ~λ),
X X
ℓm
+ (1 + cℓm ) Y−2 (2) (3, 3) harmonics, since the (3, 1) harmonic is sup-
HOM m=−ℓ
pressed with respect to the others due to its small
where HOM stands for the ℓ labels of the higher-order overall numerical prefactor. For purposes of testing
modes. The cℓm are free parameters, to be measured GR we will also not consider harmonics that only
together with all other parameters in the problem; the appear at higher PN order.
case where GR is valid corresponds to cℓm = 0. Since
hℓ −m = (−1)ℓ h∗ℓm , we set cℓ −m = cℓm . Here we will per- • The (2, 1) and (3, 3) modes are proportional to the
form parameterized tests where the cℓ|m| are allowed to relative mass difference ∆ = (m1 − m2 )/M , so that
vary one by one, as in the phase-based tests performed they are more prominent for systems with a higher
in [12–16], and we will focus on modes that will usually value of q = m1 /m2 .
be the strongest, namely the ones with (ℓ, |m|) = (3, 3) • The fact that the harmonics enter the polariza-
and (ℓ, |m|) = (2, 1). We will not only perform parame- tions through the spin-weighted spherical harmon-
ter estimation, as was done in Ref. [34], but also model ℓm
ics Y−2 (ι, φ0 ) causes their prominence to depend
selection; as we shall see, the latter will be of particular sensitively on the inclination angle ι, as illustrated
importance here. in Fig. 1. For systems that are “face-on” (ι = 0) or
The observed strengths of the higher harmonics are set “face-off” (ι = 180◦ ), only the dominant harmonic
by the total mass M , the inclination angle ι, and the rel- is visible. The sub-dominant harmonics on which
ative mass difference ∆ ≡ (m1 − m2 )/M [25, 38]. One we will focus on in this work are strongest around
aim of this paper is to investigate to what extent devia- ι ≃ 50◦ and ι ≃ 130◦ . In the figure we also indicate
tions in amplitudes of the harmonics can be determined the peak likelihood values of ι for GW190412 and
depending on the values of these parameters, in terms GW190814.
of both parameter estimation and model selection. Sec-
ondly, when performing tests that allow for non-zero cℓm , • Finally, the observed power in the sub-dominant
there will be correlations between these and the angular modes relative to that in the (2, 2) mode increases
parameters, notably ι, which will affect both the measur- with the total mass, due to a combination of
ability of the deviations from GR and the shapes of the beyond-leading-order contributions to their ampli-
posterior distributions. We will map out this interplay, tudes, how much of the signal is in the detectors’
which is necessary to interpret the results of our tests. Fi- sensitive band, and the shape of the noise power
nally, for the first time we apply this test to GW190412 spectral density Sn (f ).
and GW190814. To make the latter point more concrete, let us define
The rest of this paper is structured as follows. In Sec. II the quantities
we recall the basic properties of higher harmonics, to- 2 2
gether with the waveform model we will use. In Sec. III Z fhigh h̃ℓm (f ; ~λ) . Z fhigh h̃22 (f ; ~λ)
we set up the Bayesian analysis framework used in this αℓm ≡ df df,
study, and explain our choices for simulated signals (or flow Sn (f ) flow Sn (f )
injections), which will be used to understand the behav- (3)
ior of our analysis depending on the properties of the where h̃ℓm (f ; ~λ) is the Fourier transform of the real part
GW source. Sec. IV shows the results of our simulations of hℓm (t; ~λ), and Sn (f ) denotes the one-sided detector
and of measurements on GW190412 and GW190814. A noise power spectral density, which we take to be the one
summary and conclusions are provided in Sec. V. for Advanced LIGO at design sensitivity [1]. The inte-
grals are evaluated from a lower cut-off frequency flow =
20 Hz to an upper cut-off frequency fhigh = 2048 Hz,
II. PROPERTIES OF HIGHER HARMONICS which amply suffices for the kinds of signals considered in
AND WAVEFORM MODEL this paper. The waveform model is taken to be the most
up-to-date phenomenological inspiral-merger-ringdown
Let us start by recalling some properties of the har- model IMRPhenomXPHM [39, 40], which incorporates
monics hℓm in Eq. (1), which we will need to interpret harmonics with (ℓ, |m|) = (2, 2), (2, 1), (3, 3), (3, 2), (4, 4)
the results in subsequent sections. In doing so we limit modes, as well as effects of spin-induced precession.
ourselves to qualitative statements, mostly referring to Fig. 2 shows the dependence of the αℓm on total mass
the inspiral regime; for explicit dependences on the pa- M and mass ratio q, for (ℓ, m) = (2, 1), (3, 3), where for
rameters in the problem we refer to Refs. [25, 38]. The simplicity we have focused on binaries composed of non-
salient features relevant to us here are: spinning black holes.
3

tail the choices made for simulations that were performed


to understand the response of the analysis framework to
possible violations of GR in the amplitudes of different
harmonics.

A. Analysis framework

Consider detector data d and a hypothesis H, where,


for practical purposes, the latter corresponds to a wave-
~ in our case this could be the GR
form model h̃(f ; θ);
model for binary black hole coalescence, or one that al-
lows for deviations from GR in the amplitudes of one
of the harmonics. Then, in a Bayesian setting, measur-
ing the parameters θ~ of the source amounts to obtaining
~ H). From Bayes’
the posterior probability density p(θ|d,
theorem,
~ H) p(θ|H)
p(d|θ, ~
~ H) =
p(θ|d, , (4)
FIG. 1. The absolute values of spin-weighted spherical har- p(d|H)
monics of weight −2 as function of the inclination angle ι.
The vertical lines indicate the peak likelihood values of ι for where the evidence p(d|H) for the hypothesis H is given
GW190412 (black dashed) and GW190814 (red dashed), lo- by
cated at ≃ 47◦ and ≃ 49◦ , respectively [26, 27]. Z
p(d|H) = dθ~ p(d|θ,~ H) p(θ|H).
~ (5)

~
In the above, p(θ|H) is the prior probability density, and
~ H) takes the form [41]
the likelihood p(d|θ,
 
~ H) ∝ exp − 1 hd − h(θ)|d
p(d|θ, ~ − h(θ)i~ , (6)
2
where the noise-weighted inner product h · | · i is given by
Z fhigh ∗
ã (f ) b̃(f )
ha|bi ≡ 4 Re df. (7)
flow Sn (f )

Eq. (4) together with Eqs. (5)-(7) allow us to calculate


~ H) from the data.
the posterior probability density p(θ|d,
The posterior probability density p(θk |d, H) for a partic-
ular parameter in θ~ is obtained by integrating out all the
other parameters ξ~ in θ~ = (θk , ξ):
~
Z
p(θk |d, H) = dξ~ p(θk , ξ|d,
~ H). (8)
FIG. 2. The relative signal power in the real part of hℓm for
some of the higher-order modes with respect to the dominant Additionally, we will want to rank hypotheses: the GR
(2, 2) mode, as a function of the total mass M of the binary, hypothesis HGR versus hypotheses HNonGR , which allow
for three different values of the mass ratio q and assuming one of the cℓm to be non-zero. To this end we calculate
Advanced LIGO at design sensitivity. Bayes factors, or ratios of evidences,

NonGR p(d|HNonGR )
BGR ≡ , (9)
III. ANALYSIS FRAMEWORK AND SETUP OF
p(d|HGR )
SIMULATIONS where p(d|HNonGR ) and p(d|HGR ) are obtained using
Eq. (5), taking H to be HNonGR or HGR , respectively.
We now explain our data analysis methodology for In practice it is usually convenient to focus on the loga-
NonGR
measuring source parameters and to rank hypotheses rithm of the Bayes factor, ln BGR , as will also be done
based on the available detector data. Next we will de- here.
4

It will also be important to consider the loudness of a IV. RESULTS OF SIMULATIONS, AND
signal as it appears in a detector. The optimal signal- ANALYSES OF GW190412 AND GW190814
to-noise ratio (SNR) is defined as ρ ≡ hh(θ)|h(~ ~ 1/2 .
θ)i
For a network of detectors, the combined optimal SNR We now describe the results for our simulations, as
is obtained by summing in quadrature the SNRs in the well as for the real events GW190412 and GW190814, in
individual detectors. terms of parameter estimation and hypothesis ranking.
Finally, for estimating the evidence integrals as in In doing so, it will be useful to make a distinction between
Eq. (5), and obtaining samples for posterior density dis- the more massive BBHs (M = 65, 120 M⊙ ), the injections
~ H), we used nested sampling [41, 42] as
tributions p(θ|d, with parameters similar to those of the real events, and
implemented in the LALInference package [43] of the of course the real events themselves.
LIGO Algorithms Library (LAL) software suite [44].

A. More massive binary black holes


B. Setup of the simulations

Let us first look at results for injections with M =


To understand the response of our analysis pipeline to
65 M⊙ and M = 120 M⊙ . To have an easier overview it
GR violations in mode amplitudes with various strengths,
is convenient to first look at the behavior of log Bayes
we add simulated signals, or injections, to synthetic sta- NonGR
factors, ln BGR , which we do in Fig. 3. The trends
tionary, Gaussian noise for a network of Advanced LIGO
are as follows:
and Virgo detectors following the predicted noise spectral
densities at design sensitivity [1, 2]. Since higher-order
1. As expected, for a larger injected cℓm , the log Bayes
modes are more prominent for larger total masses, we
factor is larger. The cases c33 = 0.5 and c21 = 1
will start by considering heavier BBH systems. Later in NonGR
lead to ln BGR that tend to be consistent with
the paper we will analyze the real GW events GW190412
zero, meaning that the data are not sufficiently in-
and GW190814 to look for GR violations. To this end,
formative to clearly distinguish between hypothe-
for lower-mass systems we will perform injections whose
ses. However, starting from c33 = 1.5 or c21 = 3,
GR parameter values and SNRs are set to the maximum- NonGR
the ln BGR are significantly away from zero, and
likelihood values obtained from analyses on these events
as will be seen in terms of parameter estimation be-
that assumed GR to be correct. Specificallly:
low, here the GR deviations tend to be detectable.
• We will inject signals with M = 65 M⊙ and M =
NonGR
120 M⊙ , for mass ratios q = 3, 6, 9. Here the incli- 2. Higher values of M lead to higher ln BGR , con-
nation angle is fixed to be ι = 45◦ , and the network sistent with there being more power in the higher-
SNR to 25. order modes relative to the (2, 2) mode; see Fig. 2.

• For GW190412-like injections, M = 46.6 M⊙ , q = 3. Again as expected, on the whole a larger mass ratio
4.2, ι = 47◦ , and the network SNR is 19.8. q tends to lead to a higher ln BGRNonGR
, consistent
with there being more power in the higher-order
• For GW190814-like injections, M = 27.6 M⊙ , q = modes. We do see that the ln BGRNonGR
tend to differ
9.3, ι = 49◦ , and the network SNR is 25. less between q = 6 and q = 9 than between q = 3
and q = 6; in fact, for M = 65 M⊙ and c33 , the
We also need to choose values for the deviation pa-
log Bayes factors for the higher two values of q are
rameters c33 and c21 in the injections. Since the (3, 3)
nearly equal. Again pointing to Fig. 2, we note
mode will tend to be the strongest (see Fig. 2), we can
that the cases q = 6 and q = 9 are closer to each
expect smaller values of c33 to lead to detectable GR vi-
other than to q = 3 in terms of the power present
olations than for c21 , where “detectable” can be taken
in higher-order modes.
to mean that the 90% credible region of the posterior
density function has support that excludes zero. We
Fig. 4 shows posterior probability densities for the cor-
found that, at least for the higher masses listed above,
responding injections. The trends show broad consis-
the following choices constitute examples ranging from
tency with what we saw for the log Bayes factors. In
non-detectability to easy detectability of the GR viola-
particular, for the injected values c33 = 0.5 and c21 = 1,
tions:
posterior densities either include the GR value of zero,
• c33 = 0.5, 1.5, 3. or extend to quite close to it, while for higher injected
values, the GR value tends to be outside the support of
• c21 = 1, 3, 6. the distribution. Also, the 90% confidence intervals tend
to be tighter for higher total mass and for higher mass
NonGR
Hence these are the values for which we will show results ratio, again consistent with the behavior of the ln BGR
in the next section. in Fig. 3, and indeed with Fig. 2.
5

NonGR
FIG. 3. ln BGR for M = 65 M⊙ (top row) and M = 120 M⊙ (bottom row), for different mass ratios q indicated by the
differently shaped markers. The horizontal axes show the injected values of c33 (left column) and c21 (right column). In each
case, the non-GR hypothesis has the corresponding cℓm as free parameter.

B. Injections with parameters similar to those of Event GW190412 GW190814


GW190412 and GW190814 c33 -1.25 -3.96
c21 -2.48 -1.77

NonGR
TABLE I. Values of ln BGR for analyses of the real events
GW190412 and GW190814.
Next we turn to injections with GR parameters close
to those of the real events GW190412 and GW190814.
NonGR
Fig. 5 shows results for ln BGR . Here too the trends C. Results for GW190412 and GW190814
are as expected: the log Bayes factor increases with in-
creasing injected values for c33 and c21 . Note that al- Finally we turn to the real events GW190412 and
though GW190412 had a higher mass than GW190814 GW190814. Table I shows the results for ln BGR NonGR
when
(M = 46.6 M⊙ versus M = 27.6 M⊙ ), the mass ra- comparing the hypothesis of a non-zero c33 or c22 with
tio of GW190412 was considerably smaller than that of the GR hypothesis. All the log Bayes factors are nega-
GW190814 (q = 4.2 versus q = 9.3). The log Bayes fac- tive, so that we have no reason to suspect a violation of
tors are higher for the latter event, consistent with Fig. 2. GR in the amplitudes of sub-dominant modes.
NonGR
We see that for GW190412 one has ln BGR < 0 for More interesting are the posterior distributions for c21
c33 = 0.5, and the same is true for both injections in the and especially c33 , which are shown in Figs. 7 and 8.
cases c21 = 1 and c21 = 3, presumably due to the lower For both events, the posterior for c21 is unimodal, and
total masses. consistent with the GR value of zero. However, just like
in the simulations of the previous section, the posterior
for c33 is bimodal, also for both events.
Fig. 6 shows posterior probability distributions for the As it turns out, this bimodality results from a degen-
same injections. In all cases, the injected value for c33 eracy between c33 and the inclination angle ι. The lower
and c21 lies within the support of the posterior. For c21 panels of Figs. 7 and 8 show mismatches between (a)
the results look like what one might expect, but for c33 a reference waveform h̃ref (f ), which is a GR waveform
the posteriors are bimodal, with the true value not al- with maximum-likelihood parameters for the respective
ways lying in the strongest mode. As will be clarified signals, and (b) a waveform h̃(cℓm , ι; f ) in which cℓm and
in the next section, this behavior results from a partial ι can take on arbitrary values, but all other parameters
degeneracy between c33 and the inclination angle ι. are the maximum-likelihood ones from the GR analysis.
6

FIG. 4. Violin plots for the posterior density distributions of c33 (top two rows) and c21 (bottom two rows), for M = 65, 120 M⊙ ,
and q = 3 (left column), q = 6 (middle column), and q = 9 (right column). In each case the black horizontal bars indicate 90%
confidence intervals, and the red horizontal bar the injected value; the black vertical line shows the support of the posterior.
7

NonGR
FIG. 5. ln BGR for injections with GR parameters similar to those of GW190412 and GW190814.

FIG. 6. Violin plots for the posterior density distributions of c33 (top row) and c21 (bottom row), for injections similar to
GW190412 (left column) and GW190814 (right column). In each case the black horizontal bars indicate 90% confidence
intervals, and the red horizontal bar the injected value; the black vertical line shows the support of the posterior.
8

Specifically, we compute associated with significantly unequal component masses,


and in which strong evidence for sub-dominant mode con-
hhref |h(cℓm , ι)i tent had been found [26–28]. Log Bayes factors indicated
MM = 1 − max p p ,
tref ,ϕrefhhref |href i hh(cℓm , ι)|h(cℓm , ι)i no evidence for a GR violation in either the (2, 1) or
(10) (3, 3) mode. In the case where the (3, 3) mode was be-
where the maximization is over the above mentioned ref- ing investigated, the posterior density function for c33 ,
erence time and phase of the waveform, and we focus on while being consistent with the GR value c33 = 0, did
(ℓ, m) = (2, 1) and (ℓ, m) = (3, 3). exhibit bimodality, but this was shown to result from cor-
In the bottom panels of Figs. 7, 8, these mismatches relations between c33 and the inclination angle ι. Since
are indicated with color coding, with dark colors signify- the bimodality was also present in c33 posterior densi-
ing small mismatch. Overlaid are dashed lines indicating ties for injections with parameters similar to the ones of
the peak-likelihood values in the (bimodal) posterior dis- GW190412 and GW190814 and c33 6= 0, some caution
tribution for ι obtained when analyzing the events with is called for in interpreting such posteriors, at least for
either c21 or c33 as additional free parameters. Focus- BBHs with total mass M . 50 M⊙ . However, our results
NonGR
ing first on the case of c33 and GW190412 in Fig. 7, we show that log Bayes factors ln BGR , which were not
see that there are two regions in the (c33 , ι) plane where considered in previous work in this context [34], are ro-
mismatches are low: one region that contains the GR bust indicators for or against the presence of a violation
value c33 = 0 and is consistent with the lower value of ι, of GR.
and another region consistent with the higher ι value and Even in systems with significantly asymmetric masses
c33 6= 0. In either region, waveforms h(c33 , ι) are consis- and high total mass, with second-generation detectors,
tent with the reference waveform href , which explains the GR violations have to be sizeable (c33 & 1.5 and c21 & 3)
bimodality in the posterior for c33 . By contrast, based in order to be confidently detected. It will be of interest
on the analogous plot for (c21 , ι), no such bimodality is to see how the sensitivity of our method will improve go-
to be expected, and indeed, the posterior for c21 is uni- ing towards Einstein Telescope [45–48], Cosmic Explorer
modal. The corresponding Fig. 8 for GW190814 leads to [48–51], and the space-based LISA [52], but this is left
similar conclusions. for future work.

V. SUMMARY AND CONCLUSIONS ACKNOWLEDGMENTS

We have set up a Bayesian analysis framework to test A.P., C.K., S.R., Y.S., and C.V.D.B. are supported
GR by looking at the amplitudes of sub-dominant modes by the research programme of the Netherlands Organisa-
in GW signals from BBH coalescences, using a state-of- tion for Scientific Research (NWO). I.G. and B.S.S. are
the-art waveform model. Specifically, we allow for modi- supported by NSF grant numbers PHY-2012083 and
fications in the amplitudes of the (3, 3) and (2, 1) modes, AST-2006384. The authors are grateful for computa-
which tend to be the strongest among the sub-dominant tional resources provided by the LIGO Laboratory and
modes. Apart from performing parameter estimation on supported by the National Science Foundation Grants
the associated testing parameters c33 and c21 , this allows No. PHY-0757058 and No. PHY-0823459. This research
for hypothesis ranking between the presence and absence has made use of data, software and/or web tools ob-
of such anomalies in the modes. tained from the Gravitational Wave Open Science Cen-
Results from simulations involving injected waveforms ter (https://www.gw-openscience.org), a service of LIGO
in stationary, Gaussian noise largely follow the trends one Laboratory, the LIGO Scientific Collaboration and the
would expect based on the dependence of mode ampli- Virgo Collaboration. LIGO is funded by the U.S. Na-
tudes on total mass and mass ratio: for similar SNRs, tional Science Foundation. Virgo is funded by the French
heavier and more asymmetric systems make it easier to Centre National de Recherche Scientifique (CNRS), the
find violations of GR of the type studied here. Italian Istituto Nazionale della Fisica Nucleare (INFN)
We then performed the first analysis of this kind on and the Dutch Nikhef, with contributions by Polish and
the real events GW190412 and GW190814, which were Hungarian institutes.

[1] J. Aasi et al. (LIGO Scientific), Class. Quant. Grav. 32, Lett. 116, 241103 (2016), arXiv:1606.04855 [gr-qc].
074001 (2015), arXiv:1411.4547 [gr-qc]. [5] B. P. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev.
[2] F. Acernese et al. (VIRGO), Class. Quant. Grav. 32, X 6, 041015 (2016), [Erratum: Phys.Rev.X 8, 039903
024001 (2015), arXiv:1408.3978 [gr-qc]. (2018)], arXiv:1606.04856 [gr-qc].
[3] B. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev. [6] B. P. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev.
Lett. 116, 061102 (2016), arXiv:1602.03837 [gr-qc]. X 9, 031040 (2019), arXiv:1811.12907 [astro-ph.HE].
[4] B. P. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev. [7] R. Abbott et al. (LIGO Scientific, Virgo), (2020),
9

FIG. 7. Top panels: Posterior density functions for c21 (left) and c33 (right) for GW190412. Bottom panels: Contours of
constant mismatch between the maximum-likelihood GR waveform, and a waveform in which ι and c21 (left) or c33 (right) are
varied while keeping all other parameters the same. The dashed vertical lines indicate the GR values c21 = 0 and c33 = 0,
respectively, and the dashed horizontal lines indicate the peak-likelihood values for ι obtained from the analyses of GW190412
with respectively c21 and c33 as free parameters.

FIG. 8. The same as in Fig. 7 but for GW190814.


10

arXiv:2010.14527 [gr-qc]. [28] S. Roy, A. S. Sengupta, and K. G. Arun, Phys. Rev. D


[8] R. Abbott et al. (LIGO Scientific, VIRGO, KAGRA), 103, 064012 (2021), arXiv:1910.04565 [gr-qc].
(2021), arXiv:2111.03606 [gr-qc]. [29] C. Van Den Broeck and A. S. Sengupta, Class. Quant.
[9] B. P. Abbott et al. (Virgo, LIGO Scientific), Phys. Rev. Grav. 24, 155 (2007), arXiv:gr-qc/0607092.
Lett. 119, 161101 (2017), arXiv:1710.05832 [gr-qc]. [30] C. Van Den Broeck and A. S. Sengupta, Class. Quant.
[10] B. P. Abbott et al. (LIGO Scientific, Virgo), (2020), Grav. 24, 1089 (2007), arXiv:gr-qc/0610126.
arXiv:2001.01761 [astro-ph.HE]. [31] R. Essick and P. Landry, Astrophys. J. 904, 80 (2020),
[11] R. Abbott et al. (LIGO Scientific, KAGRA, VIRGO), As- arXiv:2007.01372 [astro-ph.HE].
trophys. J. Lett. 915, L5 (2021), arXiv:2106.15163 [astro- [32] I. Tews, P. T. H. Pang, T. Dietrich, M. W. Coughlin,
ph.HE]. S. Antier, M. Bulla, J. Heinzel, and L. Issa, Astrophys.
[12] B. P. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev. J. Lett. 908, L1 (2021), arXiv:2007.06057 [astro-ph.HE].
Lett. 116, 221101 (2016), [Erratum: Phys.Rev.Lett. 121, [33] S. Dhanpal, A. Ghosh, A. K. Mehta, P. Ajith, and
129902 (2018)], arXiv:1602.03841 [gr-qc]. B. S. Sathyaprakash, Phys. Rev. D 99, 104056 (2019),
[13] B. P. Abbott et al. (LIGO Scientific, Virgo), Astrophys. arXiv:1804.03297 [gr-qc].
J. Lett. 882, L24 (2019), arXiv:1811.12940 [astro-ph.HE]. [34] T. Islam, A. K. Mehta, A. Ghosh, V. Varma, P. Ajith,
[14] B. P. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev. and B. S. Sathyaprakash, Phys. Rev. D 101, 024032
Lett. 123, 011102 (2019), arXiv:1811.00364 [gr-qc]. (2020), arXiv:1910.14259 [gr-qc].
[15] B. P. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev. [35] S. Kastha, A. Gupta, K. G. Arun, B. S. Sathyaprakash,
D 100, 104036 (2019), arXiv:1903.04467 [gr-qc]. and C. Van Den Broeck, Phys. Rev. D 98, 124033 (2018),
[16] R. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev. D arXiv:1809.10465 [gr-qc].
103, 122002 (2021), arXiv:2010.14529 [gr-qc]. [36] S. Kastha, A. Gupta, K. G. Arun, B. S. Sathyaprakash,
[17] R. Abbott et al. (LIGO Scientific, VIRGO, KAGRA), and C. Van Den Broeck, Phys. Rev. D 100, 044007
(2021), arXiv:2112.06861 [gr-qc]. (2019), arXiv:1905.07277 [gr-qc].
[18] K. G. Arun, B. R. Iyer, M. S. S. Qusailah, and B. S. [37] C. D. Capano and A. H. Nitz, Phys. Rev. D 102, 124070
Sathyaprakash, Class. Quant. Grav. 23, L37 (2006), (2020), arXiv:2008.02248 [gr-qc].
arXiv:gr-qc/0604018. [38] L. Blanchet, Living Rev. Rel. 17, 2 (2014),
[19] K. G. Arun, B. R. Iyer, M. S. S. Qusailah, and arXiv:1310.1528 [gr-qc].
B. S. Sathyaprakash, Phys. Rev. D 74, 024006 (2006), [39] G. Pratten et al., Phys. Rev. D 103, 104056 (2021),
arXiv:gr-qc/0604067. arXiv:2004.06503 [gr-qc].
[20] N. Yunes and F. Pretorius, Phys. Rev. D 80, 122003 [40] A. Ramos-Buades, P. Schmidt, G. Pratten, and S. Husa,
(2009), arXiv:0909.3328 [gr-qc]. Phys. Rev. D 101, 103014 (2020), arXiv:2001.10936 [gr-
[21] T. G. F. Li, W. Del Pozzo, S. Vitale, C. Van Den Broeck, qc].
M. Agathos, J. Veitch, K. Grover, T. Sidery, R. Stu- [41] J. Veitch and A. Vecchio, Phys. Rev. D 81, 062003
rani, and A. Vecchio, Phys. Rev. D 85, 082003 (2012), (2010), arXiv:0911.3820 [astro-ph.CO].
arXiv:1110.0530 [gr-qc]. [42] J. Skilling, Bayesian Analysis 1, 833 (2006).
[22] T. G. F. Li, W. Del Pozzo, S. Vitale, C. Van Den Broeck, [43] J. Veitch et al., Phys. Rev. D 91, 042003 (2015),
M. Agathos, J. Veitch, K. Grover, T. Sidery, R. Sturani, arXiv:1409.7215 [gr-qc].
and A. Vecchio, J. Phys. Conf. Ser. 363, 012028 (2012), [44] LIGO Scientific Collaboration, “LIGO Algorithm Li-
arXiv:1111.5274 [gr-qc]. brary - LALSuite,” free software (GPL) (2018).
[23] M. Agathos, W. Del Pozzo, T. G. F. Li, C. Van [45] M. Punturo et al., Class. Quant. Grav. 27, 084007 (2010).
Den Broeck, J. Veitch, and S. Vitale, Phys. Rev. D 89, [46] S. Hild et al., Class. Quant. Grav. 28, 094013 (2011),
082001 (2014), arXiv:1311.0420 [gr-qc]. arXiv:1012.0908 [gr-qc].
[24] J. Meidam et al., Phys. Rev. D 97, 044033 (2018), [47] M. Maggiore et al., JCAP 03, 050 (2020),
arXiv:1712.08772 [gr-qc]. arXiv:1912.02622 [astro-ph.CO].
[25] L. Blanchet, G. Faye, B. R. Iyer, and S. Sinha, [48] V. Kalogera et al., (2021), arXiv:2111.06990 [gr-qc].
Class. Quant. Grav. 25, 165003 (2008), [Erratum: [49] B. P. Abbott et al. (LIGO Scientific), Class. Quant. Grav.
Class.Quant.Grav. 29, 239501 (2012)], arXiv:0802.1249 34, 044001 (2017), arXiv:1607.08697 [astro-ph.IM].
[gr-qc]. [50] D. Reitze et al., Bull. Am. Astron. Soc. 51, 035 (2019),
[26] R. Abbott et al. (LIGO Scientific, Virgo), Phys. Rev. D arXiv:1907.04833 [astro-ph.IM].
102, 043015 (2020), arXiv:2004.08342 [astro-ph.HE]. [51] M. Evans et al., (2021), arXiv:2109.09882 [astro-ph.IM].
[27] R. Abbott et al. (LIGO Scientific, Virgo), Astrophys. J. [52] S. Babak, A. Petiteau, and M. Hewitson, (2021),
Lett. 896, L44 (2020), arXiv:2006.12611 [astro-ph.HE]. arXiv:2108.01167 [astro-ph.IM].

You might also like