19
Parametrized tests of the strong-field dynamics of general relativity using gravitational wave signals from coalescing binary black holes: Fast likelihood calculations and sensitivity of the method Jeroen Meidam, 1 Ka Wa Tsang, 1 Janna Goldstein, 2 Michalis Agathos, 3 Archisman Ghosh, 1 Carl-Johan Haster, 4 Vivien Raymond, 5 Anuradha Samajdar, 1 Patricia Schmidt, 6 Rory Smith, 7 Kent Blackburn, 8 Walter Del Pozzo, 9 Scott E. Field, 10 Tjonnie Li, 11 Michael Pürrer, 5 Chris Van Den Broeck, 1,12 John Veitch, 13 and Salvatore Vitale 14 1 NikhefNational Institute for Subatomic Physics, 105 Science Park, 1098 XG Amsterdam, The Netherlands 2 School of Physics and Astronomy, University of Birmingham, Birmingham, B15 2TT, United Kingdom 3 DAMTP, Centre for Mathematical Sciences, University of Cambridge, Wilberforce Road, Cambridge CB3 0WA, United Kingdom 4 Canadian Institute for Theoretical Astrophysics, University of Toronto, Toronto, Ontario M5S 3H8, Canada 5 Albert-Einstein-Institut, Max-Planck-Institut für Gravitationsphysik, D-14476 Golm, Germany 6 Department of Astrophysics / IMAPP, Radboud University Nijmegen, P.O. Box 9010, 6500 GL Nijmegen, The Netherlands 7 OzGrav, School of Physics and Astronomy, Monash University, Clayton 3800, Victoria, Australia 8 LIGO, California Institute of Technology, Pasadena, California 91125, USA 9 Dipartimento di Fisica Enrico Fermi, Universit ` a di Pisa, Pisa I-56127 and INFN sezione di Pisa, Italy 10 Mathematics Department, University of Massachusetts Dartmouth, Dartmouth, Massachusetts 02747, USA 11 Department of Physics, The Chinese University of Hong Kong, Shatin, NT, Hong Kong 12 Van Swinderen Institute for Particle Physics and Gravity, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands 13 Institute for Gravitational Research, University of Glasgow, Glasgow G12 8QQ, United Kingdom 14 LIGO, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA (Received 23 December 2017; published 22 February 2018) Thanks to the recent discoveries of gravitational wave signals from binary black hole mergers by Advanced Laser Interferometer Gravitational Wave Observatory and Advanced Virgo, the genuinely strong-field dynamics of spacetime can now be probed, allowing for stringent tests of general relativity (GR). One set of tests consists of allowing for parametrized deformations away from GR in the template waveform models and then constraining the size of the deviations, as was done for the detected signals in previous work. In this paper, we construct reduced-order quadratures so as to speed up likelihood calculations for parameter estimation on future events. Next, we explicitly demonstrate the robustness of the parametrized tests by showing that they will correctly indicate consistency with GR if the theory is valid. We also check to what extent deviations from GR can be constrained as information from an increasing number of detections is combined. Finally, we evaluate the sensitivity of the method to possible violations of GR. DOI: 10.1103/PhysRevD.97.044033 I. INTRODUCTION Since 2015, the twin Advanced Laser Interferometer Gravitational Wave Observatories (Advanced LIGOs) [1] have routinely been detecting gravitational wave signals from coalescing binary black holes [26], recently also in conjunction with Advanced Virgo [7,8]. Later in the decade, the worldwide gravitational wave detector network will be extended with the Japanese KAGRA [9], to be followed by LIGO-India [10]. In the course of the next several years, tens to hundreds more binary black hole detections are expected to be made [4]. Coalescences of stellar mass binary black holes (BBHs) are ideal laboratories for testing the genuinely strong-field dynamics of general relativity (GR) [11,12]: they are likely to be pure spacetime events, involving stronger curvatures and shorter dynamical time scales than in any other experiment or observation, by many orders of magnitude [13]. The process starts with two black holes that are orbiting each other, gradually losing orbital energy and PHYSICAL REVIEW D 97, 044033 (2018) 2470-0010=2018=97(4)=044033(19) 044033-1 © 2018 American Physical Society

PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

Parametrized tests of the strong-field dynamics of general relativity usinggravitational wave signals from coalescing binary black holes:

Fast likelihood calculations and sensitivity of the method

Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan Haster,4

Vivien Raymond,5 Anuradha Samajdar,1 Patricia Schmidt,6 Rory Smith,7 Kent Blackburn,8

Walter Del Pozzo,9 Scott E. Field,10 Tjonnie Li,11 Michael Pürrer,5 Chris Van Den Broeck,1,12

John Veitch,13 and Salvatore Vitale141Nikhef—National Institute for Subatomic Physics, 105 Science Park, 1098 XG Amsterdam,

The Netherlands2School of Physics and Astronomy, University of Birmingham, Birmingham, B15 2TT, United Kingdom

3DAMTP, Centre for Mathematical Sciences, University of Cambridge, Wilberforce Road,Cambridge CB3 0WA, United Kingdom

4Canadian Institute for Theoretical Astrophysics, University of Toronto, Toronto,Ontario M5S 3H8, Canada

5Albert-Einstein-Institut, Max-Planck-Institut für Gravitationsphysik, D-14476 Golm, Germany6Department of Astrophysics / IMAPP, Radboud University Nijmegen,

P.O. Box 9010, 6500 GL Nijmegen, The Netherlands7OzGrav, School of Physics and Astronomy, Monash University, Clayton 3800, Victoria, Australia

8LIGO, California Institute of Technology, Pasadena, California 91125, USA9Dipartimento di Fisica “Enrico Fermi”, Universita di Pisa, Pisa I-56127 and INFN sezione di Pisa, Italy

10Mathematics Department, University of Massachusetts Dartmouth, Dartmouth,Massachusetts 02747, USA

11Department of Physics, The Chinese University of Hong Kong, Shatin, NT, Hong Kong12Van Swinderen Institute for Particle Physics and Gravity, University of Groningen,

Nijenborgh 4, 9747 AG Groningen, The Netherlands13Institute for Gravitational Research, University of Glasgow, Glasgow G12 8QQ, United Kingdom

14LIGO, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA

(Received 23 December 2017; published 22 February 2018)

Thanks to the recent discoveries of gravitational wave signals from binary black hole mergers by AdvancedLaser Interferometer Gravitational Wave Observatory and Advanced Virgo, the genuinely strong-fielddynamics of spacetime can now be probed, allowing for stringent tests of general relativity (GR). One set oftests consists of allowing for parametrized deformations away from GR in the template waveform models andthen constraining the size of the deviations, as was done for the detected signals in previous work. In thispaper, we construct reduced-order quadratures so as to speed up likelihood calculations for parameterestimation on future events. Next, we explicitly demonstrate the robustness of the parametrized tests byshowing that they will correctly indicate consistency with GR if the theory is valid. We also check to whatextent deviations from GR can be constrained as information from an increasing number of detections iscombined. Finally, we evaluate the sensitivity of the method to possible violations of GR.

DOI: 10.1103/PhysRevD.97.044033

I. INTRODUCTION

Since 2015, the twin Advanced Laser InterferometerGravitational Wave Observatories (Advanced LIGOs)[1] have routinely been detecting gravitational wavesignals from coalescing binary black holes [2–6],recently also in conjunction with Advanced Virgo[7,8]. Later in the decade, the worldwide gravitationalwave detector network will be extended with theJapanese KAGRA [9], to be followed by LIGO-India[10]. In the course of the next several years, tens to

hundreds more binary black hole detections are expectedto be made [4].Coalescences of stellar mass binary black holes (BBHs)

are ideal laboratories for testing the genuinely strong-fielddynamics of general relativity (GR) [11,12]: they are likelyto be pure spacetime events, involving stronger curvaturesand shorter dynamical time scales than in any otherexperiment or observation, by many orders of magnitude[13]. The process starts with two black holes that areorbiting each other, gradually losing orbital energy and

PHYSICAL REVIEW D 97, 044033 (2018)

2470-0010=2018=97(4)=044033(19) 044033-1 © 2018 American Physical Society

Page 2: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

orbital angular momentum through the emission of gravi-tational waves (GWs). By the time the GW frequency ishigh enough to be in the sensitive band of Earth-baseddetectors, the binary will likely have shed almost all ofits original eccentricity [14] and will be undergoing quasi-circular inspiral. Eventually, the inspiral becomes non-adiabatic, after which the components of the binaryundergo a plunge followed by merger, leading to theformation of a single, highly excited black hole. The latterundergoes ringdown as it asymptotes to a quiescent, Kerrblack hole. The GW signal that is emitted can, at largedistances, be described as a small metric perturbationpropagating at the speed of light on a Minkowski back-ground; however, the shape of the wave encodes detailedinformation about the strong-field, dynamical inspiral-merger-ringdown (IMR) process it originated from.A number of techniques have been developed to under-

stand inspiral-merger-ringdown in GR, including large-scale numerical relativity (NR) simulations resulting fromdirect integration of the Einstein equations [15–17] andthe construction of (semi)analytic waveform models. Theeffective one-body (EOB) formalism [18–22] has beenextended to combine the post-Newtonian (PN) descriptionof inspiral [23] with NR results for the merger, as well asblack hole perturbation theory for the ringdown [24–26],leading to high-quality IMR waveforms in the time domain[27]. In the frequency domain, phenomenological IMRmodels [28–30] were developed based on a frequencydomain PN expansion together with hybridized EOB/NRwaveforms [31–33].A variety of possible deviations from GR have been

considered in the context of binary coalescence, includingscalar-tensor theories, a varying Newton constant, modifiedGW dispersion relations, e.g., arising from “massivegraviton” models, violations of the no-hair hypothesis,violations of cosmic censorship, and parity violatingtheories (see, e.g., [13] and references therein). Even withinthe GR paradigm, one can think of alternative compactobjects to black holes (e.g. boson stars, dark matter stars, orgravastars), which may exhibit tidal effects during inspiral(see [34,35]) and will also have a different ringdown signalfrom a black hole. For some alternative theories, it has beenworked out how the post-Newtonian inspiral would bemodified to leading order [36], and for certain exoticobjects, the ringdown spectrum has been computed [37].However, what seems to be lacking in all cases are the kindof high-accuracy IMR waveforms that are available forBBH coalescence in GR. Thus, given observational GWdata for a detected compact binary coalescence event, atpresent, it is not possible to compare GR with alternativetheories or BBH coalescences with those of alternativecompact objects, while making use of the full informationin the IMR signal. Moreover, GR might be violated in analtogether different way that is yet to be envisaged.

Given these restrictions, at the present time, it isexpedient to devise tests of the theory of general relativityitself, which to the largest extent possible are generic and asaccurate as we can make them. Following the recent binaryblack hole merger detections, a battery of such tests weredeployed [4,5,12]: looking for coherent excess signalpower in the data after subtraction of the best-fitting GRwaveform [38,39], checking for consistency with GRbetween the pre- and postmerger signals in terms of massesand spins [40,41], evaluating consistency of the postmergersignal with the presence of a least-damped ringdown mode[12], constraining anomalous GW propagation with a viewon bounding the mass of the graviton as well as violationsof local Lorentz invariance [42,43] (the latter also using thebinary neutron star detection [44–46]), looking for evi-dence of nonstandard polarization states [47], and meas-uring a series of judiciously chosen coefficients associatedwith parametrized deformations of IMR waveforms awayfrom GR [4,5,12,36,48–54]. This paper deals primarilywith the latter tests.As mentioned above, a number of IMR waveform

models have been developed. For parametrized tests ofGR, we use the phenomenological models, which have aclosed expression in the frequency domain and hence canbe generated fast on a computer (which is important fordata analysis purposes when exploring high-dimensionalparameter spaces), capture the essential physics of theproblem (including, e.g., spin-induced precession), andallow for some amount of analytic insight into the meaningof the induced deformations. In particular, we use themodel which in the LIGO Algorithm Library is designatedas IMRPhenomPv2 [31–33]. The IMRPhenomPv2 wave-form phase is characterized by a number of parametersfpig: (i) in the adiabatic inspiral regime, PN coefficientsfφ0;…;φ7g and fφ5l;φ6lg; (ii) in the intermediate regimebetween adiabatic inspiral and merger, phenomenologicalcoefficients fβ0;…; β3g; and (iii) in the merger-ringdownregime fα0;…; α5g. In the most relevant of these coef-ficients, parametrized deformations are introduced byallowing for relative deviations: pi → ð1þ δpiÞpi. Theδpi will be referred to as our testing parameters.We then perform a series of tests, in each of which some

testing parameter δpj is allowed to vary freely along withall other parameters entering the phase (component massesand spins, which enter through the GR expressions for thepi themselves), but δpk ¼ 0 for k ≠ j. In principle, multipleδpi could be allowed to vary at the same time, but this willlead to a degradation in the measurement accuracy for all ofthem [12]; statistical errors will be much smaller when theδpi are varied one at a time. Note that, in most alternativetheories of gravity, a violation will likely show up in morethan one coefficient. However, as already demonstrated in[55] in a PN context, looking for a deviation from zero in asingle testing parameter is an efficient way to search forGR violations that occur at multiple PN orders, and one can

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-2

Page 3: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

even find violations at powers of frequency that are distinctfrom the one that the testing parameter is associated with[52,53]. Of course, if a deviation is present, then theindividual measurements of the δpi will not necessarilyreflect the predicted values of the correct alternative theory.Should one want to measure or constrain, e.g., extracharges or coupling constants that may be present in one’sfavorite alternative theory using an IMR signal, then anaccurate IMR waveform model would need to be con-structed for that particular theory. However, this is not theaim of the framework presented here; what we want to do istest Einstein’s theory itself by constraining deviations fromthe theory.Even though the IMRPhenomPv2 waveform model has

an explicit analytic expression, in the case of low-massbinary mergers, which leave a long signal in the detectors’sensitive band, the analyses are computationally costly andcan take more than a month of time to complete, due to thelarge number of likelihood evaluations [Oð108Þ] that mustbe performed. Given the large number of detections that areexpected to be made in the coming LIGO-Virgo observingruns, ways must be found to reduce the computationalburden. One solution is to speed up the likelihood calcu-lation by constructing reduced-order quadratures (ROQs)[56–58], which in turn are based on reduced-order models[59–66]. In the method of reduced-order quadratures, thediscrete overlap calculation involved in the likelihoodevaluation is split up into a data dependent sum, whichonly needs to be evaluated once for each detection, and amuch shorter sum that takes care of the parameter depen-dent part of the overlap calculation that must be performedmany times during the sampling over parameter space. Inline with the method outlined above, a series of ROQs iscreated, in each of which a single testing parameter δpi isallowed for.Results of the parametrized tests for the LIGO-Virgo

detections of binary black hole coalescences have alreadyappeared elsewhere [4,5,12]. The aim of this paper istwofold: (a) to construct ROQs for IMRPhenomPv2 wave-forms with parametrized deformations, for use on futuredetections, and (b) to explicitly demonstrate the robustnessand sensitivity of the method as a whole, which had not yetbeen done in previous publications.The structure of this paper is as follows. In Sec. II, we

briefly recall the waveform model used and exploreanalytically how the phase varies with the chosen defor-mations and as a function of mass. We describe the setupof the parametrized tests and explain how results frommultiple detections can be combined to arrive at strongerbounds on GR violations. Section III describes the con-struction of the reduced-order quadrature for waveformswith testing parameters. Next, in Sec. IV, we presentsome checks of the correctness and robustness of thedata analysis pipeline. In Sec. V, we show how well theparametrized tests can bound GR violations by combining

information from all available sources. Furthermore, weinvestigate how testing parameters in the template wave-forms respond when deviations in one or more parametersare present in the signal. Section VI provides a summaryand conclusions.Throughout this paper, we set c ¼ G ¼ 1 unless speci-

fied otherwise.

II. WAVEFORM MODEL ANDPARAMETRIZED TESTS

A. Waveform model

The starting point for the parametrized tests is thephenomenological frequency domain waveform modelwhich in the LIGO Algorithm Library is designated asIMRPhenomPv2 [33]. This waveform model describes anapproximate signal of a precessing binary by applying arotation transformation [33,67] to an underlying alignedspin waveform model, here taken to be IMRPhenomD[31,32]. The orbital precession dynamics are given in termsof an effective spin parametrization [33,68]. For an in-depthdescription, we refer to these papers; here we only give aquick overview.The phasing of IMRPhenomPv2 consists of three regimes,

whose physical meaning and parametrization are as follows:(1) The inspiral regime is parametrized by post-

Newtonian coefficients fφ0;…;φ7g and fφ5l;φ6lg,as well as phenomenological parameters fσ0;…;σ4g.The latter are contributions at high effective PNorder (up to 5.5 PN) to correct for nonadiabaticityin late inspiral and for unknown high-order PNcoefficients in the adiabatic regime.

(2) The intermediate regime transitions between inspiraland merger-ringdown; it is parametrized by thephenomenological coefficients fβ0;…; β3g.

(3) The merger-ringdown regime is parametrizedby a combination of phenomenological and ana-lytical black hole perturbation theory parametersfα0;…; α5g.

Note that the PN coefficients fφ0;…;φ7g and fφ5l;φ6lghave their usual dependences on the binary’s componentmasses and spins. The other phenomenological parametersare fixed by calibration against numerical relativity wave-forms. For the functions of frequency in which the aboveparameters appear, we refer to [32]; see also Table I in [12].The transition from the inspiral to the intermediate regimehappens at a frequency f ¼ f1 ¼ 0.018/M (whereM is thetotalmass) and from the intermediate to themerger-ringdownregime at f ¼ f2 ¼ 0.5fRD, with fRD a “ringdown fre-quency”, in such a way that the waveform is C1 continuous.Figure 1 shows the modulus of the waveform jhðfÞj,

highlighting the three regimes; also shown is the Fouriertransform to the time domain hðtÞ and the correspondinginstantaneous frequency as a function of time.

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-3

Page 4: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

Not all of the coefficients mentioned above will be usedin the parametrized tests. In the inspiral regime, φ5 iscompletely degenerate with the phase at coalescence φc;similarly, the pair ðσ0; σ1Þ is degenerate with ðφc; tcÞ, withtc as the time at coalescence. The pairs ðβ0; β1Þ and ðα0; α1Þare set by the requirement of C1 continuity between thedifferent regimes. We also omit α5, which occurs in thesame term as α4, meaning that there will be some amountof degeneracy between the two. Finally, we do not usefσ2; σ3; σ4g, whose fractional calibration uncertaintieswere larger [ða fewÞ × 10−1] than those of the otherphenomenological parameters [at most ða fewÞ × 10−2],though all calibration uncertainties were observed to bebelow measurement uncertainties for the binary black holecoalescence detections that were made [69].The way our parametrized tests are implemented is by

allowing for fractional deviations from the GR values for allof the remaining coefficients pi in turn

pGRi ðm1; m2;S1;S2Þ → ð1þ δpiÞpGR

i ðm1; m2;S1;S2Þ;ð1Þ

where m1, m2 are the component masses and S1, S2 are thecomponent spins; one has

fδpigi ¼ fδφ0;…; δφ7; δφ5l; δφ6l; δβ2; δβ3; δα2; δα3; δα4g:ð2Þ

We note that in GR, φ1 ≡ 0, so that as an exception, we letδφ1 be an absolute rather than a relative deviation.Including extrinsic parameters coming from the detector

response, in practice, the full parameter sets of the resultingwaveform models will be

λ ¼ ftc;φc; DL; θ;ϕ;ψ ; m1; m2; χ1; χ2; χp; θJ; α0; δpig:ð3Þ

Here, tc and φc are, respectively, the time and phase atcoalescence; DL is the luminosity distance; ðθ;ϕÞ give thesky position; ψ is a polarization angle; m1 and m2 are thecomponent masses; χ1 and χ2 are spin magnitudes; and χpis an “effective” spin precession parameter given by [68]

χp ¼ maxðA1m21χ1⊥; A2m2

2χ2⊥ÞA21m

21

; ð4Þ

where A1 ¼ 2þ 3m2/2m1, A2 ¼ 2þ 3m1/2m2, and χ1⊥,χ2⊥ are the projections of the spin vectors onto the orbitalplane, i.e., orthogonal to the direction of the orbital angularmomentum L at a specific reference frequency fref. θJ is theangle between the line of sight n and the total angularmomentum J at fref , and α0 indicates the azimuthalorientation of L at fref [33].

B. Effect of testing parameters on the phase

Before going on to evaluate the sensitivity of para-metrized tests given stellar mass BBHs as seen in theadvanced detectors, we first illustrate the effect on thephase of varying the δpi. As it turns out, one of the best-determined PN testing parameters tends to be δφ3; in theintermediate regime this is δβ2 and in the merger-ringdownregime it is δα2; these are the parameters we focus on.Figure 2 shows how the phase as a function of frequency

varies with testing parameters Ψðδpi; fÞ, as well as thedifference with the phase in GR, ΔΨðδpi; fÞ ¼ Ψðδpi; fÞ−ΨGRðfÞ, for tc ¼ φc ¼ 0. Two kinds of sources are con-sidered, with masses and spins chosen to be the means ofthe posterior density functions for the signals that weredesignated GW150914 [2] and GW151226 [3]. The phasesand their differences are plotted from flow ¼ 20 Hz and upto a frequency where the dominant ðl ¼ 2; m ¼ 2Þmode ofthe ringdown signal can be safely assumed to have ended(600 Hz for GW150914 and 800 Hz for GW151226). Thequalitative behavior is as expected given the differencesbetween the two. GW150914, being more massive,had a short inspiral regime and the merger occurred atf ∼ 130 Hz, close to the frequency where the detectors arethe most sensitive. By contrast, GW151226 had a much

FIG. 1. The three regimes of the IMRPhenomPv2 model. (Top)The modulus of the waveform as a function of frequency for asignal similar to GW150914. (Bottom) Fourier transform to thetime domain (top) and instantaneous frequency as a function oftime (bottom).

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-4

Page 5: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

longer inspiral (with ∼55 cycles in band) and its mergeroccurred at f ∼ 450 Hz. For the PN testing parameter δφ3, amuch larger phase difference is accumulated in the case ofGW151226, which will cause this parameter to be muchbetter measured in the latter case despite the overall smallersignal-to-noise ratio. The intermediate regime parameter δβ2exhibits a relatively slowly increasing phase difference andlevels out between 100 and 200 Hz for both events, which iswhere the detectors are the most sensitive; hence, we canexpect it to be roughly equally well measurable for bothevents. Finally, for δα2, in the case of GW150914, the phase

difference reaches a maximum at some point before decreas-ing again, while for GW151226, the difference keepsincreasing up to high frequencies, but not with a largermaximum phase difference; hence, this parameter will bebetter measurable with GW150914, for which the merger-ringdown regime occurs at frequencies closer to the range ofbest detector sensitivity. These expectations are borne out bythe published results for the two events [4,5,12]. Note thatvarying the δpi has an effect at all frequencies; this is aconsequence of the C1 junction conditions between theinspiral, intermediate, and merger-ringdown regimes.

FIG. 2. The effect of varying testing parameters on the phase as a function of frequency for tc ¼ φc ¼ 0. In the top of each panel, weplot the GR phase (black) as well as the way the phase varies with a testing parameter (colors); the bottom shows the difference. In theleft column, we show results for an event with parameters like that of GW150914 and similarly in the right column for GW151226.

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-5

Page 6: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

In Fig. 3, we illustrate the phase differences as chirpmass Mc ¼ Mη3/5 is varied (where M ¼ m1 þm2 andη ¼ m1m2/M2), for particular values of the δpi; thesymmetric mass ratio is fixed at η ¼ 0.2 and againtc ¼ φc ¼ 0; shown are the ΔΨðδpi; fÞ for f ¼ 150 Hz,i.e., close to the frequency of optimal sensitivity for theAdvanced LIGO detectors. Again, the behavior is asexpected. Low Mc corresponds to waveforms with signifi-cant inspiral in band; deviations in the φi then have a large

effect on the observable phase.Deviations in the intermediateregime parameters βi have the largest effect when thisregime occurs at frequencies where the detectors are themost sensitive, which corresponds to Mc ¼ 10–20 M⊙.Finally, shifts in the merger-ringdown parameters αi havetheir largest effect for Mc ≳ 20 M⊙, which brings thisregime in the detectors’ most sensitive band. For complete-ness, we show results for zero spins, as well as aligned spinswith jS1;2j ¼ 0.9; thewell-known effect of “orbital hang-up”[70], which prolongs the duration of waveforms in the timedomain, then causes similar features to occur at higherMc.

C. Parameter estimation

Parameter estimation is done using the LALInferenceframework [71,72], in which the posterior density distri-bution for the parameters λ is obtained as

pðλjHi; d; IÞ ¼pðλjHi; IÞpðdjHi; λ; IÞ

pðdjIÞ : ð5Þ

Here, Hi is the hypothesis corresponding to the waveformmodel in which δpi is an extra free parameter, d are thedata, and I denotes whatever background information wemay have; pðdjHi; λ; IÞ is the likelihood function, which upto an overall prefactor is given by

pðdjHi; λ; IÞ ∝ exp½−hd − hðλÞjd − hðλÞi/2�; ð6Þ

with hðλÞ the signal model described in Sec. II A and h·j·i asthe noise-weighted inner product

hajbi ¼ 4ℜZ

fhigh

flow

a�ðfÞbðfÞSnðfÞ

df; ð7Þ

where ℜ denotes the real part, and SnðfÞ is the one-sidednoise spectral density. For the second-generation detectors,the lower cutoff frequency is taken to be flow ¼ 20 Hz,while fhigh ¼ 2048 Hz suffices as an upper cutoff fre-quency for stellar mass BBH coalescences. The likelihoodfunction is evaluated using the efficient nested samplingalgorithm [71]. pðλjHi; IÞ is the prior probability densityfor the free parameters; for those parameters that alsoappear in the GR waveform, these are chosen in the sameway as in [72], while for δpi, we choose priors uniform inan interval that is wide enough to contain the supports ofthe posterior densities; suitable ranges are given in Sec. IIIB below. Finally, pðdjIÞ is the probability of the data,which can be absorbed into an overall normalization factorfor the posterior density pðλjHi; d; IÞ.To obtain one-dimensional posterior densities for the

parameters δpi, one marginalizes over all other parameters

pðδpijHi; d; IÞ ¼Z

dθpðθ; δpijHi; d; IÞ; ð8Þ

FIG. 3. The differences at f ¼ 150 Hz between the GR phaseand the phase for given values of testing parameters, for inspiral(top), the intermediate regime (middle), and the merger-ringdownregime (bottom). For definiteness, we again set tc ¼ φc ¼ 0. Thesolid lines are for zero spins, and the dashed lines for alignedspins with jS1;2j ¼ 0.9.

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-6

Page 7: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

where the integration is performed over all parameters in(3) except for δpi.Finally, posterior densities from individual events can be

conveniently combined to arrive at stronger bounds on theδpi under the assumption that the fractional deviations arethe same in each event. Assuming independent detectionsd1; d2;…; dN , it is easy to see that

pðδpijHi; d1; d2;…; dN; IÞ

¼ pðδpijIÞ1−NYNn¼1

pðδpijHi; dn; IÞ: ð9Þ

For events with similar signal-to-noise ratios and in theabsence of measurement offsets, one can expect the widthsof these posteriors to decrease roughly with

ffiffiffiffiN

p.

III. REDUCED-ORDER QUADRATURES FORFAST LIKELIHOOD CALCULATIONS

A. Basic method

We now proceed to constructing reduced-order quad-ratures. The technical underpinnings have already beendiscussed in detail elsewhere [57,58]; here we will onlygive an overview.The first step is to approximate the waveform hðλ; fÞ as

hðλ; fÞ ≃ PEn ½hðλ; fÞ�

≡Xni¼1

ðeijhðλÞÞeiðfÞ; ð10Þ

where the vectors in the reduced basis En ¼ feiðfÞgni¼1 areorthonormal with respect to the inner product

ðajbÞ≡Z

fmax

fmin

a�ðfÞbðfÞdf; ð11Þ

and the approximation is good to within a greedy projectionerror ϵ:

jjhðλ; fÞ − PEn ½hðλ; fÞ�jj2 < ϵ; ð12Þ

for λ ∈ T N, where T N is a suitably large training set andjjajj≡ ffiffiffiffiffiffiffiffiffiffiffiðajaÞp

. From this, one constructs an empiricalinterpolant to approximate the waveform

In½h�ðλ; fÞ≡Xni¼1

xiðλÞeiðfÞ; ð13Þ

where the coefficients xi are solutions to

In½h�ðλ;F kÞ ¼ hðλ;F kÞ ð14Þat interpolation points fF kgnk¼1. Defining the matrixAij ¼ ejðF iÞ, one has

In½h�ðλ;F kÞ ¼Xni¼1

Xnk¼1

ðA−1Þikhðλ;F kÞeiðfÞ

¼Xnk¼1

BLk ðfÞhðλ;F kÞ; ð15Þ

where

BLk ðfÞ ¼

Xni¼1

ðA−1ÞikeiðfÞ: ð16Þ

The fF kgnk¼1 are chosen from a set ffigLi¼1, where L isrelated to the duration T of the longest waveform consid-ered through

L ¼ ðfmax − fminÞT þ 1; ð17Þ

and the fi are spaced by Δf ¼ 1/T. The first interpolationpoint F 1 is chosen such that it maximizes the amplitude ofthe first reduced basis vector, i.e., je1ðF 1Þj ≥ je1ðfiÞj forall fi. Next, one builds an interpolant of e2ðfÞ using only e1and F 1, and one finds an F 2 that maximizes the pointwiseinterpolation error, i.e., jI1½e2�ðF 2Þ − e2ðF 2Þj ≥ jI1½e2�ðfiÞ − e2ðfiÞj for all fi. One then continues in this fashionuntil n interpolation points have been obtained.Though the interpolant In½h�ðλ; fÞ can be evaluated at

any parameter values λ, the underlying reduced basis En

satisfies the tolerance criterion (12) only for λ ∈ T N. Nextcomes the validation step, where the accuracy of theinterpolant is evaluated also for values λ that lie outsidethe training set (though inside the same ranges as forthe training set, where the waveform approximant isdeemed valid). Arbitrary values are picked, for which itis checked that

jjhðλ; fÞ − In½h�ðλ; fÞjj2 < β ð18Þfor some choice of maximum interpolation error β. All“bad points” λ for which this is not the case get collectedand added to the training set T N , thus creating a newtraining set on which the algorithm is repeated, leading to anew interpolant In0 ½h�ðλ; fÞ. The validation step is repeateduntil no more bad points are found, leading to the finalinterpolant INL

½h�ðλ; fÞ.Recall that the aim is to speed up the calculation of the

likelihood L ¼ pðdjHi; λ; IÞ, the logarithm of which takesthe form

logL ¼ 1

2½2hdjhðλÞi − hhðλÞjhðλÞi − hdjdi�: ð19Þ

First, consider the term hdjhi. Substituting for hðλ; fÞ theempirical interpolant INL

½h�ðλ; fÞ and discretizing theintegral in the definition of the inner product, one gets

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-7

Page 8: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

hdjhðλÞi ¼ 4ΔfℜXLi¼1

d�ðfiÞhðλ; fiÞSnðfiÞ

≃ 4ΔfℜXLi¼1

XNL

k¼1

BLk ðfiÞhðλ;F kÞ

d�ðfiÞSnðfiÞ

¼ 4ΔfℜXNL

k¼1

�XLi¼1

BLk ðfiÞ

d�ðfiÞSnðfiÞ

�hðλ;F kÞ

¼XNL

k¼1

wkhðλ;F kÞ; ð20Þ

where

wk ≡ 4ΔfℜXLi¼1

BLk ðfiÞ

d�ðfiÞSnðfiÞ

: ð21Þ

An important point is now that, typically, NL ≪ L, and thelikelihood calculations (20), which during the nestedsampling process need to be performed “on the fly” formany different parameter values, now only involve theevaluation of NL expressions hðλ;F kÞ, rather than the Levaluations of hðλ; fÞ that were required originally. While itis true that the calculation of the ROQ weights wk stillinvolves a sum over L terms, they only need to be evaluatedonce for every detection. This means that, with the ROQ,this part of the likelihood calculation will be sped up by afactor L/NL.Finally, in Eq. (19), there is also the term hhðλÞjhðλÞi,

which can be approximated by an expression of the form

hhðλÞjhðλÞi ¼XNQ

j¼1

wQj jhðλ;FQ

j Þj2; ð22Þ

where

wQj ¼ 4Δfℜ

XLi¼1

BQj ðfiÞ

SnðfiÞ; ð23Þ

for some BQj , and typically NQ ≪ L. The BQ

j are obtainedthrough a similar procedure as in the linear case. Here too, theweights wQ

j will only have to be calculated once perdetection. Note that L and Q in the superscripts refer tothe linear and quadratic parts of the likelihood respectively.The above only gives an overview of the rationale behind

ROQs. In practice, one needs to write the waveform hðλ; fÞin terms of the þ and × polarizations as Fþhþðλ; fÞ þF×h×ðλ; fÞ, with Fþ and F× as the beam pattern functions.However, it turns out that a single set of functions fBL

j g is

sufficient to represent hþ and h×, and a single set fBQj g is

sufficient to represent the products jhþj2, jh×j2, andℜh�þh× [57,58].

B. A ROQ for IMRPhenomPv2 with parametrizeddeformations

A ROQ for IMRPhenomPv2 in the GR case was alreadyconstructed in [58]. Here, we want to build a series ofROQs for IMRPhenomPv2, each including a single testingparameter δpi. As a starting point, we use the final reducedbasis for the GR waveform T N (where N can be either theNL or the NQ of the linear and quadratic bases, respec-tively), and for each basis element we introduce Nδpi

¼500 samples placed uniformly in the δpi direction (seeFig. 4). The ranges for the various δpi are chosen such thatthey accommodate the widths of posterior density functionsof the LIGO-Virgo events that were recorded so far (withthe exception of δα2, δα3, δα4, which are essentiallyunmeasurable for low-mass events)

δφ0 ∈ ½−2; 2�; δφ1 ∈ ½−5; 5�; δφ2 ∈ ½−10; 10�;δφ3 ∈ ½−10; 10�; δφ4 ∈ ½−10; 10�; δφ5l ∈ ½−10; 10�;δφ6 ∈ ½−10; 10�; δφ6l ∈ ½−20; 20�; δφ7 ∈ ½−20; 20�;δβ2 ∈ ½−5; 5�; δβ3 ∈ ½−5; 5�;δα2 ∈ ½−5; 5�; δα3 ∈ ½−5; 5�; δα4 ∈ ½−5; 5�: ð24Þ

FIG. 4. Schematic illustration of how the original basis forIMRPhenomPv2 from [58] (triangles) is extended in the addi-tional parameter dimension δpi (in this example δφ3) to form anew training set. The plot only shows a three-dimensional slice ofthe full parameter space. Note that points with −0.02 ≤ δφ3 ≤0.02 are not shown to aid visualization.

TABLE I. The different chirp mass bins for which ROQs werebuilt, with the ranges of waveform durations in the GR case aswell as sampling in frequency.

Bin Mc (M⊙) GR waveform duration (sec) Δf (Hz)

A [12.3, 45] [0.4, 4] 1/4B [7.9, 14.8] [3, 8] 1/8C [5.2, 9.5] [6, 16] 1/16D [3.4, 6.2] [12, 32] 1/32E [2.2, 4.2] [23.8, 64] 1/64

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-8

Page 9: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

The resulting set T N×Nδpithen becomes a training set

for the construction of a new ROQ, as outlined in theprevious subsection. As for the ROQ of the GR waveform,this is done independently for waveforms in differentoverlapping chirp mass bins, so as to obtain better like-lihood calculation speedups than when all chirp masseswould be lumped together. The chirp mass ranges roughlycorresponding to different ranges for the length of the

waveform in the time domain. We note that away from theGR case there is no clear mapping from chirp mass towaveform length, as the latter is also partially determinedby the values of the δpi. Even so, in each bin we a priori setΔf ¼ 1/Tmax, where Tmax is the longest GR waveform inthe bin; though waveforms can become longer whenδpi ≠ 0, in the end what counts is that all interpolationerrors are below the given threshold. To reduce the burden

FIG. 5. Distributions of interpolation errors at different validation steps for some of the testing parameters and chirp mass bins. The leftcolumn is for the linear bases, the right column for the quadratic ones. (First row) δφ3 for chirp mass bin A in Table I; (second row) δβ2for bin C; (third row) δα2 for bin C. In some cases no bad points are found, so that the basis does not need to be enlarged and no furthervalidation steps are needed.

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-9

Page 10: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

on computer memory required, we perform multibandingas explained in [58]: an adaptive frequency resolutionΔfðfÞ is applied such that waveforms are sampled lessdensely at higher frequencies, where there is less power perfrequency bin due to the faster frequency sweep (seeFig. 1). However, once a basis has been obtained, we upsample by direct evaluation of the waveform model.The various bins are shown in Table I. Note that no ROQs

were made for systems with Mc > 45 M⊙, since for suchbinaries the signal will be short enough that parameterestimation is sufficiently fast and not much speedup canbe expected from a ROQ. The bin with the lowest chirpmasses considered here isMc ∈ ½2.2; 5.2�M⊙, correspond-ing to a lowest total mass of M ≃ 5 M⊙ for m1/m2 ¼ 1,which should suffice for the lightest astrophysical binaryblack holes. For the other parameters appearing inIMRPhenomPv2, we use the same ranges as in [58]:1 ≤ m1/m2 ≤ 9; ð−0.9;−0.9;0Þ≤ ðχ1L;χ2L;χpÞ≤ ð0.9;0.9;0.9Þ, where χ1L, χ2L are the spin components along thedirection of angular momentum L; ð0; 0Þ ≤ ðθJ; α0Þ ≤ðπ; 2πÞ; and m1 ≥ m2 ≥ 1 M⊙. In the validation steps,we also impose the bound χ1L ≥ 0.4–7η, as was done in[58]; this is needed to avoid clustering of bad points in aparticular region, indicating a limitation of the originalIMRPhenomPv2 waveform model. For the ROQs with theδpi, it turned out to be necessary to impose an additional

boundffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiχ21L þ χ2p

q≤ 0.98.

Like for the GR version of IMRPhenomPv2, the greedyprojection error is set to ϵ ¼ 10−8 and the maximuminterpolation error is set to β ¼ 10−6. Some representativedistributions of the interpolation error at different validationsteps are shown in Fig. 5. As it turns out, the addition of a

testing parameter δφi typically increases the sizes of thefinal bases in the different chirp mass bins by only a factorof a few, though with some exceptions; the largest increasehappens to be for δφ1 and M ∈ ½3.4; 6.2�M⊙, where thelinear basis size went from 524 to 5264.Table II shows the speedups in likelihood calculations—

defined as ½ðfmax − fminÞT þ 1�/ðNL þ NQÞ—that areachievable with the ROQs. The speedup is greatest forlong signals where analyses are the most involved. Theseare the theoretical speedups; the actual speedups in prac-tical parameter estimation will vary, but tend to be the sameas the theoretical ones within a factor of 2 or less.Finally, the ROQs were interfaced with the aforemen-

tioned LALInference framework. Figure 6 compares someparameter estimation results obtained with and without theROQ on the same simulated signal. We see that the resultsare consistent, with posterior density functions not differingby more than what is expected given uncertainties in thesampling process [72]. The robustness of the infrastructurewill be tested in more detail in the next section.

IV. ROBUSTNESS OF THE TESTS

We now perform some checks of the correctness of thedata analysis pipeline and its robustness against waveformsystematics and instrumental noise. We do this in two ways.One is to construct so-called p-p plots, which quantify thestatistical inconsistencies of the posterior density distribu-tions. Another consists of analyzing a numerical waveforminjected in many different stretches of real detector noise, asa check that the pipeline behaves as it should under thecombined effects of the injected waveform being differentfrom the template waveform model and the presence ofinstrumental glitches in the detector output.

A. Reliable measurement of testing parameters

A requirement for a parameter estimation algorithm isthat it is capable of measuring parameters in a statisticallyreliable way. Detector noise can cause offsets in posteriordensity functions, but given a large number of signals, itshould be the case that the correct parameter value isrecovered with a confidence p in a fraction p of the cases.Specifically, assuming GR is correct, for any of the para-metrized tests, it should be the case that the value δpi ¼ 0lies in a confidence interval of width p for a fraction p ofthe measurements. We check this by adding 100 simulatedGR signals (injections) to synthetic, stationary, Gaussiannoise, with the predicted power spectral density at designsensitivity for the two Advanced LIGO detectors [73].The signals have randomly chosen sky positions andorientations and are placed uniformly in comoving volumewith DL ∈ ½250; 750� Mpc, with component masses m1,m2 ∈ ½6; 40�M⊙, and arbitrarily oriented spins with mag-nitudes jS1j, jS2j ∈ ½0; 0.9�. Injections are analyzed with theROQs whose chirp mass bins they fall into; in reality, onewould look at the chirp mass measured with GR templates.

TABLE II. Theoretical speedups of likelihood calculations dueto the ROQs, for different testing parameters and the chirp massbins of Table I. Note how these are larger for longer signals,where they are the most needed. Speedups in practical parameterestimation will vary, but tend to be the same as the theoreticalones within a factor of 2 or less.

δpi A B C D E

δφ0 4.3 7.6 28.3 38.7 47.1δφ1 3.0 4.6 8.6 11.7 27.1δφ2 4.2 6.8 24.8 42.4 56.5δφ3 4.0 6.1 20.8 36.9 55.8δφ4 3.9 10.1 40.3 76.5 111.2δφ5l 4.1 7.6 29.9 62.9 97.9δφ6 3.7 9.8 39.0 76.0 114.0δφ6l 3.8 10.1 42.1 78.1 117.1δφ7 3.7 9.1 39.5 74.7 112.6δβ2 3.0 8.7 34.9 78.0 117.5

δβ3 3.5 6.8 28.5 69.8 111.2δα2 2.8 9.2 39.4 88.2 124.6δα3 2.9 10.8 44.8 87.5 128.3δα4 2.8 10.4 43.3 88.1 131.6

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-10

Page 11: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

p-p plots for a few of the testing parameters are shown inFig. 7. As an indicator of consistency of the results withabsence of bias in the measurements, one can calculate theKolmogorov-Smirnov (K-S) statistic, which is defined asthe maximum (in absolute value) of the difference betweendistributions; in our case, the latter are simply the p-pdistribution on the one hand and the diagonal on the other.We find K-S values of 0.04, 0.09, and 0.04 for δφ3, δβ2, andδα2, respectively. We conclude that the analyses work asexpected.

B. Numerical relativity injections in real detector noise

Finally, we investigate the response of the parametrizedtests to a numerical relativity waveform injected in detector

noise that contains instrumental glitches. In particular, weuse real data from the S6 data set, but “recolored” to theearly Advanced LIGO noise curve from [74]; this pro-cedure changes the average power spectral density butretains (and in fact enhances) any instrumental nonstatio-narities that were present in the original data. Sinceinstrumental glitches will have a larger effect for short-duration signals, we focus on GW150914. We consider anumerical relativity waveform from the SXS catalog, whosemass ratio and spins are close to the measured means forGR150914; specifically, we pick SXS:BBH:0307 [75]. Theintrinsic parameters were ðm1; m2Þ ¼ ð40.83; 33.26ÞM⊙,and S1¼ð0.092;0.038;0.326Þ, S2¼ð0.215;0.301;−0.558Þat fref¼20Hz. This same waveform is then injected in 21

FIG. 6. A comparison of parameter estimation results on a simulated signal with parameters Mc ¼ 8.9 M⊙, q ¼ 1.99,DL ¼ 200 Mpc, and δpi ¼ 0, in synthetic, stationary, Gaussian noise, analyzed with and without the ROQ. Results are shown forthe cases δφ3 (top row), δβ2 (middle row), and δα2 (bottom row). In each case, we show the posterior density function for the testingparameter itself (left column) and for chirp mass (right column). The values of the parameters in the signal are indicated by the verticaldashed lines. Results with and without ROQ agree to within sampling uncertainties [72].

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-11

Page 12: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

different stretches of noise [76,77] and the parametrized testsare performed. In choosing these stretches, care was taken topick ones that did not exhibit egregiously large glitches(which can be done by visual inspection of time-frequency

spectrograms), since the presence of a sufficiently sizeabledeparture from Gaussianity of the noise may preclude anevent being detected in the first place. The strategy is similarto what was followed in [78] (see their Sec. III E), where theeffect of possible nonstationarities on parameter estimation—in the GR case—was also assessed by injecting a particularnumerical relativity waveform in different stretches of realdetector noise.As a diagnostic, we define the “GR quantile” as the

cumulative probability of a given δpi being nonpositive

Qi ≡Z

0

−∞pðδpijHi; d; IÞdδpi: ð25Þ

If the GR quantile is close to zero, then the posteriorpðδpijHi; d; IÞ exhibits a significant offset towards positiveδpi; if it is close to one, then there is a large offset towardsnegative values. Given many measurements on the samesignal in different noise realizations, we expect theQi to bedistributed uniformly on the interval [0, 1].In Fig. 8, we first of all show the 90% credible intervals for

the PN testing parameters fδφ0;…; δφ7g and fδφ5l; δφ6lg

0.0 0.2 0.4 0.6 0.8 1.0GR confidence level

0.0

0.2

0.4

0.6

0.8

1.0

Fra

ctio

nof

even

tsbel

owle

vel

0.0 0.2 0.4 0.6 0.8 1.0GR confidence level

0.0

0.2

0.4

0.6

0.8

1.0

Fra

ctio

nof

even

tsbel

owle

vel

0.0 0.2 0.4 0.6 0.8 1.0GR confidence level

0.0

0.2

0.4

0.6

0.8

1.0

Fra

ctio

nof

even

tsbel

owle

vel

FIG. 7. Fraction of simulated signals in stationary Gaussian noise for which the value of zero for δpi is within a given confidence level.Shown are p-p plots for δφ3 (left), δβ2 (middle), and δα2 (right). The dark and light gray bands indicate the 1-σ and 2-σ departures fromthe diagonal that can be expected on theoretical grounds. The results are consistent with a general absence of bias in the measurements.

FIG. 8. Ninety percent credible intervals for the PN testing parameters obtained by performing the parametrized tests on a numericalrelativity injection in 21 different stretches of realistic detector data. Note how offsets tend to alternate from one PN testing parameter tothe next; this is due to partial correlation between them and the alternating signs of the PN parameters themselves.

FIG. 9. Histograms of GR quantiles for the PN testing param-eters corresponding to the same simulations as for Fig. 8. Thoughbased on analyses of only 21 stretches of data, the results areconsistent with the quantiles being uniformly distributed on theinterval [0,1].

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-12

Page 13: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

for the 21 stretches of data.We note how the deviations in thePN parameters tend to alternate in sign, due to the fact thatthere is some correlation between them, and that the φithemselves have alternating signs. Next, in Fig. 9, we showthe distribution of theQi, which despite the small sample sizeis indeed suggestive of uniformity on [0,1].Needless to say, a full investigation for systems like

GW150914 would require performing the parametrizedtests for a much larger sample of data stretches than the 21used here and it would be of interest to repeat the study forother choices of masses and spins; due to computationalrestrictions, this was not practicable. Nevertheless, theoutcome is indicative of the expected behavior.

V. MEASUREMENT SENSITIVITIES

Next, we want to assess the power of our parametrizedtests in constraining GR violations and their sensitivity toselected GR violations, by adding simulated signals tostationary Gaussian detector noise with the power spectraldensity of the Advanced LIGO detectors at design sensi-tivity [73] and performing parameter estimation as in theprevious section.As far as GR violations are concerned, ideally one would

like to do this using specific alternative theories of gravity.However, in most cases, the effects of particular theorieshave only been calculated for the inspiral and then only toleading PN order [13,36,51,79]; to our knowledge, fullinspiral-merger-ringdown waveform models with reason-able inclusion of all relevant physical effects so far onlyexist for GR itself. Hence, we confine ourselves toinjections that have a deviation δpi in a particular coef-ficient pi or in several of the pi at the same time, startingfrom some PN order. However, in the template waveformsused for the measurements, we still only vary a single oneof the δpi at a time. As we shall see, if the injections havedeviations in multiple coefficients, then single-parametertests will still pick this up. In fact, even parameters that arenot associated with the deviations in the signal must showdeviations. Such effects had already been observed in[52,53,55], and should not come as a surprise: template

waveform models will use whatever additional freedomthey have to accommodate anomalies in the signals. At thesame time, only varying one testing parameter leads to ahigher measurement accuracy than for multiple parametersbeing varied at the same time. A drawback is that posteriordensities for testing parameters can not be straightfor-wardly mapped to statements about whatever additionalcharges, coupling constants, or energy scales may bepresent in some particular alternative theories. For this tobe possible, accurate and complete inspiral-merger-ring-down waveforms for alternative theories would be required,but these are not currently available. However, the purposeof the parametrized tests is not to place bounds onparameters characterizing other theories, but rather to testthe theory of general relativity itself, with as high anaccuracy as possible.

A. Bounding GR violations

First, we illustrate the ability of the parametrized tests inputting bounds on GR violations, which will get increas-ingly sharper as information from multiple events iscombined. The posterior density functions for each ofthe δpi obtained from the simulated signals in Sec. IVAlead to combined posterior densities according to theprescription of Eq. (9). As shown in Fig. 10, after a fewtens of detections, these will be sharply peaked near thevalue of zero. In these examples, after 50 (100) detections,the 1 − σ accuracies on δφ3, δβ2, and δα2 are, respectively,0.013 (0.008), 0.020 (0.013), and 0.054 (0.032).

B. Simulated signals with deviations inparticular coefficients

We now consider injections that have a deviation in aparticular coefficient pi. The GR parameters are picked tobe the means of the posterior density distributions forGW150914 [2]. We focus on this type of source so as tohave some amount of sensitivity to each of the inspiral,intermediate, and merger-ringdown regimes. The injectionsare done in stationary Gaussian noise with the predicted

FIG. 10. Sharper constraints on deviation from GR can be obtained by combining posterior density functions for the δpi from allavailable detections. This is illustrated for δφ3 (left), δβ2 (middle), and δα2 (right). The black curve shows the median of the jointdistribution, and the darker and lighter shadings show the 68% and 95% confidence intervals, respectively.

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-13

Page 14: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

power spectral density at design sensitivity for the twoAdvanced LIGO detectors [73]. For the deviations, weconsider in turn two representative parameters from each ofthe inspiral, intermediate, and merger-ringdown regimesand give the corresponding δpi a magnitude that roughlycorresponds to 5 times the standard deviation observedfor GW150914, with both positive and negative signs.In particular, δφ3 ¼ �0.4, δφ4 ¼ �3.3, δβ2 ¼ �0.7, δβ3 ¼�0.8, δα2 ¼ �1.3, and δα4 ¼ �1.6.Figure 11 shows posterior densities for the cases where

the injection has either nonzero δφ3 or nonzero δφ4, and inthe measurements, all of the δpi are allowed to vary in turn.A few things can be noted:(1) In each case, the posterior density for the testing

parameter where the deviation in the signal resideshas no support at the GR value of zero, but thesupport does contain the injected value.

(2) The posterior densities of all of the other PN testingparameters, with the exception of δφ1, show strongoffsets away from zero.

(3) On the other hand, the intermediate-regime andmerger-ringdown testing parameters show much lessof a response to a deviation in a PN parameter.

(4) The deviations in the PN parameters tend to alternatein sign. This reflects the fact that there is someamount of correlation between these parameters andthat the φi themselves have alternating signs.

The posteriors in Figs. 12 and 13, where either anintermediate-regime parameter or a merger-ringdown

parameter in the signal has a deviation, show analogousbehavior: for the parameter where the deviation resides,posteriors have no support at zero, but this is also the casefor at least one other parameter, usually one in the sameregime.

C. Simulated signals with deviations in multiplecoefficients

Next, we consider injections in which all the δpi arenonzero starting from some PN order. Two scenarios areconsidered:(1) All testing parameters starting from 1.5 PN have the

same fractional shifts δpi ¼ 0.5. This includes thesets δφ3;4;5l;6;6l;7, δβ2;3, and δα2;3;4.

(2) All testing parameters starting from1.5PNhave shiftswhose sign alternates from one parameter to the next,according to the way they are correlated: δφ3;5l;6l;7 ¼−0.4 and δφ4;6 ¼ þ0.4. For the intermediate-regimeand merger-ringdown parameters, we choose δβ2 ¼δβ3 ¼ −0.4 and δα2 ¼ δα3 ¼ δα4 ¼ 0.4.

The results are shown in Fig. 14, and can be summarizedas follows:(1) Again strong deviations are picked up even by testing

parameters that are not associated with the violationsin the signal; both for the same- and alternating-signviolations, all of the testing parameters return aposterior density function whose support does notcontain the GR value of zero.

FIG. 11. (Top) Posterior densities for testing parameters for an injection with δφ3 ¼ þ0.4 (orange) and δφ3 ¼ −0.4 (blue). (Bottom)posteriors for an injection with δφ4 ¼ þ3.3 (orange) and δφ4 ¼ −3.3 (blue). Note how all the PN testing parameters indicate a deviationfrom GR, not just the ones that deviate from zero in the signal.

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-14

Page 15: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

FIG. 12. (Top) Posterior densities for testing parameters for an injection with δβ2 ¼ þ0.7 (orange) and δβ2 ¼ −0.7 (blue). (Bottom)Posteriors for an injection with δβ3 ¼ þ0.8 (orange) and δβ3 ¼ −0.8 (blue). In each case, the GR violation is also picked up by theother δβi.

FIG. 13. (Top) Posterior densities for testing parameters for an injection with δα2 ¼ þ1.3 (orange) and δα2 ¼ −1.3 (blue). (Bottom)Posteriors for an injection with δα4 ¼ þ1.6 (orange) and δα4 ¼ −1.6 (blue). Here too, in each case, the other δαi also pick up the GRviolation.

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-15

Page 16: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

(ii) Even in the case where the signs of all the deviationsare the same, we see alternation in the offsets of theposterior densities for PN parameters, following theway they are correlated.

(iii) For PN parameters from 1.5 PN onwards, themeasured GR violation is larger than the injecteddeviation; individual parameters respond to thecollective change in the waveform induced by theshifts in all of the testing parameters together.

Hence, measuring the δpi one by one can enable thediscovery of GR violations also when the signal hasmultiple pi that deviate from their GR values.

VI. SUMMARY AND CONCLUSIONS

In [4,5,12], the detected binary black hole signals wereanalyzed using template waveforms that allow for para-metrized deviations from GR, so as to test the strong-fielddynamics of the theory. In this work, we have introducedreduced-order quadratures that speed up likelihood calcu-lations by factors of a few to more than a hundred, whichwill significantly ease the computational burden in apply-ing the method to future events. Our chosen waveformmodel is IMRPhenomPv2, though we note that the methodused in this paper can in principle also be applied toreduced-order models for other frequency domain wave-forms with parametrized deviations added, such as the onesin [61,62]. We also established the method’s robustnessthrough p-p plots for simulated signals in syntheticGaussian noise and by examining the results for a numeri-cal relativity injection in different stretches of real data fromthe S6 data set, recolored to the Advanced LIGO finaldesign sensitivity. Finally, the sensitivity of the method wasevaluated using both GR injections and injections with GRviolations in various parameters.A range of alternative theories of gravity have been

considered, which are often characterized by additional

charges or coupling constants. The tests presented here donot easily map to statements about such parameters; puttingconstraints on particular alternative theories would requirefull inspiral-merger-ringdown waveforms of similar qualityas the ones we have for GR. The regular observation ofbinary black hole coalescences will be an incentive fortheorists to develop such models. However, the main aim ofthe parametrized tests is to perform stringent tests of GRitself, and as we have demonstrated, our method provides areliable and accurate way of doing this.Recently a binary neutron star merger was also discov-

ered [44]. Here too the parametrized tests can be applied,although care should be taken so that the effects of theneutron stars’ tidal deformation are not confused with aviolation of GR. This can be done by analyzing the signalup to frequencies of only a few hundred hertz so that tidaleffects can be neglected [52–54] or by including tidaldeformabilities in the signal. The latter approach has theadvantage that the entire signal can be used, but there willalso be some loss of sensitivity due to the increaseddimensionality of parameter space; which approach willbe the most efficient is yet to be determined. Especially forthese kinds of events, which involve longer signals than forbinary black holes, it would be beneficial to constructreduced-order quadratures; this too is left for future work.

ACKNOWLEDGMENTS

The authors have benefited from discussions withmany LIGO Scientific Collaboration (LSC) and VirgoCollaboration members. J. M., K.W. T., A. G., P. S., andC. V. D. B. are supported by the research program of theNetherlands Organisation for Scientific Research (NWO).M. A. acknowledges NWO-Rubicon Grant No. RG86688.T. L. was partially supported by a grant from the ResearchGrants Council of the Chinese University of Hong Kong(Project No. CUHK 24304317) and the Direct Grant for

FIG. 14. Results for injections where all of the signal’s testing parameters starting from 1.5 PN have a fractional shift jδpij ¼ 0.4, inone case, all with positive sign (orange), in another case, with a sign that alternates from one parameter to the next (blue); see the maintext for details. In both cases, the offsets of the posterior densities follow the way successive PN coefficients are correlated. Also notehow in both cases all of the δpi clearly indicate a GR violation in the signal. In fact, from 1.5 PN order onwards, the measured violationsin PN parameters is larger than the injected deviation at a given order: individual testing parameters will try to accommodate thecollective change in the signal resulting from the shifts in all the parameters together.

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-16

Page 17: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

Research from the Research Committee of the ChineseUniversity of Hong Kong. J. V. is supported by U.K.Science and Technology Facilities Council (STFC) GrantNo. ST/K005014/1. K. B. and S. V. acknowledge thesupport of the National Science Foundation and theLIGO Laboratory. LIGO was constructed by theCalifornia Institute of Technology and Massachusetts

Institute of Technology with funding from the NationalScience Foundation and operates under cooperative agree-ment PHY-0757058. S. E. F. acknowledges support fromthe National Science Foundation through Grant No. PHY-1606654, the Sherman Fairchild Foundation, and helpfuldiscussions with Saul Teukolsky. P. S. acknowledges NWOVeni Grant No. 680-47-460.

[1] J. Aasi et al., Advanced LIGO, Classical Quantum Gravity32, 074001 (2015).

[2] B. P. Abbott et al., Observation of Gravitational Waves froma Binary Black Hole Merger, Phys. Rev. Lett. 116, 061102(2016).

[3] B. P. Abbott et al., GW151226: Observation of Gravita-tional Waves from a 22-Solar-Mass Binary Black HoleCoalescence, Phys. Rev. Lett. 116, 241103 (2016).

[4] B. P. Abbott et al., Binary Black Hole Mergers in the FirstAdvanced LIGO Observing Run, Phys. Rev. X 6, 041015(2016).

[5] B. P. Abbott et al., GW170104: Observation of a 50-Solar-Mass Binary Black Hole Coalescence at Redshift 0.2, Phys.Rev. Lett. 118, 221101 (2017).

[6] B. P. Abbott et al., GW170608: Observation of a 19-solar-mass binary black hole coalescence, Astrophys. J., Lett.851, L35 (2017).

[7] F. Acernese et al., Advanced Virgo: A second-generationinterferometric gravitational wave detector, Classical Quan-tum Gravity 32, 024001 (2015).

[8] B. P. Abbott et al., GW170814: A Three-Detector Obser-vation of Gravitational Waves from a Binary Black HoleCoalescence, Phys. Rev. Lett. 119, 141101 (2017).

[9] Y. Aso, Y. Michimura, K. Somiya, M. Ando, O. Miyakawa,T. Sekiguchi, D. Tatsumi, and H. Yamamoto, Interferometerdesign of the KAGRA gravitational wave detector, Phys.Rev. D 88, 043007 (2013).

[10] B. R. Iyer et al., Report No. LIGO-M1100296-v2, 2011.[11] E. Berti et al., Testing general relativity with present and

future astrophysical observations, Classical Quantum Grav-ity 32, 243001 (2015).

[12] B. P. Abbott et al., Tests of General Relativity withGW150914, Phys. Rev. Lett. 116, 221101 (2016).

[13] N. Yunes, K. Yagi, and F. Pretorius, Theoretical physicsimplications of the binary black-hole mergers GW150914and GW151226, Phys. Rev. D 94, 084002 (2016).

[14] P. C. Peters and J. Mathews, Gravitational radiation frompoint masses in a Keplerian orbit, Phys. Rev. 131, 435(1963).

[15] F. Pretorius, Evolution of Binary Black Hole Spacetimes.,Phys. Rev. Lett. 95, 121101 (2005).

[16] M. Campanelli, C. O. Lousto, P. Marronetti, and Y.Zlochower, Accurate Evolutions of Orbiting Black-HoleBinaries without Excision, Phys. Rev. Lett. 96, 111101(2006).

[17] J. G. Baker, J. R. van Meter, S. T. McWilliams, J. Centrella,and B. J. Kelly, Consistency of Post-Newtonian Waveformswith Numerical Relativity, Phys. Rev. Lett. 99, 181101(2007).

[18] A. Buonanno and T. Damour, Effective one-body approachto general relativistic two-body dynamics, Phys. Rev. D 59,084006 (1999).

[19] A. Buonanno and T. Damour, Transition from inspiral toplunge in binary black hole coalescences, Phys. Rev. D 62,064015 (2000).

[20] T. Damour, P. Jaranowski, and G. Schaefer, Effective onebody approach to the dynamics of two spinning black holeswith next-to-leading order spin-orbit coupling, Phys. Rev. D78, 024009 (2008).

[21] T. Damour and A. Nagar, An Improved analytical descrip-tion of inspiralling and coalescing black-hole binaries, Phys.Rev. D 79, 081503 (2009).

[22] E. Barausse and A. Buonanno, An Improved effective-one-body Hamiltonian for spinning black-hole binaries, Phys.Rev. D 81, 084024 (2010).

[23] L. Blanchet, Gravitational radiation from post-Newtoniansources and inspiralling compact binaries, Living Rev.Relativity 5, 3 (2002).

[24] C. V. Vishveshwara, Scattering of gravitational radiation bya Schwarzschild black-hole, Nature (London) 227, 936(1970).

[25] W. H. Press, Long wave trains of gravitational waves from avibrating black hole, Astrophys. J. 170, L105 (1971).

[26] S. Chandrasekhar and S. Detweiler, The quasi-normalmodes of the Schwarzschild black hole, Proc. R. Soc. A344, 441 (1975).

[27] A. Taracchini et al., Effective-one-body model for black-hole binaries with generic mass ratios and spins, Phys. Rev.D 89, 061502 (2014).

[28] P. Ajith et al., A template bank for gravitational waveformsfrom coalescing binary black holes. I. Non-spinningbinaries, Phys. Rev. D 77, 104017 (2008); Erratum, Phys.Rev. D 79, 129901 (2009).

[29] P. Ajith et al., Inspiral-Merger-Ringdown Waveforms forBlack-Hole Binaries with Non-Precessing Spins, Phys. Rev.Lett. 106, 241101 (2011).

[30] L. Santamaría, F. Ohme, P. Ajith, B. Brügmann, N. Dorband,M. Hannam, S. Husa, P. Moesta, D. Pollney, C. Reisswig,E. L. Robinson, J. Seiler, and B. Krishnan, Matching post-Newtonian and numerical relativity waveforms: Systematic

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-17

Page 18: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

errors and a new phenomenologicalmodel for non-precessingblack hole binaries, Phys. Rev. D 82, 064016 (2010).

[31] S. Husa, S. Khan, M. Hannam, M. Pürrer, F. Ohme, X. J.Forteza, and A. Bohe, Frequency-domain gravitationalwaves from non-precessing black-hole binaries. I. Newnumerical waveforms and anatomy of the signal, Phys.Rev. D 93, 044006 (2016).

[32] S. Khan, S. Husa, M. Hannam, F. Ohme, M. Pürrer, X. J.Forteza, and A. Bohe, Frequency-domain gravitationalwaves from non-precessing black-hole binaries. II. Aphenomenological model for the advanced detector era,Phys. Rev. D 93, 044007 (2016).

[33] M. Hannam, P. Schmidt, A. Bohe, L. Haegel, S. Husa, F.Ohme, G. Pratten, and M. Pürrer, Simple Model ofComplete Precessing Black-Hole-Binary GravitationalWaveforms, Phys. Rev. Lett. 113, 151101 (2014).

[34] G. F. Giudice, M. McCullough, and A. Urbano, Hunting fordark particles with gravitational waves, J. Cosmol. Astro-part. Phys. 10 (2016) 001.

[35] V. Cardoso, E. Franzin, A. Maselli, P. Pani, and G. Raposo,Testing strong-field gravity with tidal Love numbers, Phys.Rev. D 95, 084014 (2017); Publisher’s Note, Phys. Rev. D95, 089901 (2017).

[36] N. Yunes and F. Pretorius, Fundamental theoretical bias ingravitational wave astrophysics and the parametrized post-Einsteinian framework, Phys. Rev. D 80, 122003 (2009).

[37] V. Cardoso, S. Hopper, C. F. B. Macedo, C. Palenzuela, andP. Pani, Gravitational-wave signatures of exotic compactobjects and of quantum corrections at the horizon scale,Phys. Rev. D 94, 084031 (2016).

[38] T. Littenberg and N. J. Cornish, A Bayesian approach to thedetection problem in gravitational wave astronomy, Phys.Rev. D 80, 063007 (2009).

[39] N. J. Cornish and T. B. Littenberg, BayesWave: Bayesianinference for gravitational wave bursts and instrumentglitches, Classical Quantum Gravity 32, 135012 (2015).

[40] A. Ghosh et al., Testing general relativity using goldenblack-hole binaries, Phys. Rev. D 94, 021101 (2016).

[41] A. Ghosh, N. K. Johnson-Mcdaniel, A. Ghosh, C. K. Mishra,P. Ajith, W. Del Pozzo, C. P. L. Berry, A. B. Nielsen, andL. London, Testing general relativity using gravitational wavesignals from the inspiral,merger and ringdownof binaryblackholes, Classical and Quantum Gravity 35, 014002 (2017).

[42] C. M. Will, Bounding the mass of the graviton usinggravitational-wave observations of inspiralling compactbinaries, Phys. Rev. D 57, 2061 (1998).

[43] S. Mirshekari, N. Yunes, and C. M. Will, Constraininggeneric lorentz violation and the speed of the graviton withgravitational waves, Phys. Rev. D 85, 024041 (2012).

[44] B. P. Abbott et al., GW170817: Observation of Gravita-tional Waves from a Binary Neutron Star Inspiral, Phys.Rev. Lett. 119, 161101 (2017).

[45] V. Alan Kosteleck and M. Mewes, Testing local Lorentzinvariance with gravitational waves, Phys. Lett. B 757, 510(2016).

[46] B. P. Abbott et al., Gravitational waves and gamma-raysfrom a binary neutron star merger: GW170817 and GRB170817A, Astrophys. J. 848, L13 (2017).

[47] C. M.Will, The confrontation between general relativity andexperiment, Living Rev. Relativity 9, 3 (2006).

[48] L. Blanchet and B. S. Sathyaprakash, Signal analysis ofgravitational wave tails, Classical Quantum Gravity 11,2807 (1994).

[49] L. Blanchet and B. S. Sathyaprakash, Detecting the TailEffect in Gravitational Wave Experiments, Phys. Rev. Lett.74, 1067 (1995).

[50] C. K. Mishra, K. G. Arun, B. R. Iyer, and B. S.Sathyaprakash, Parametrized tests of post-Newtoniantheory using Advanced LIGO and Einstein telescope, Phys.Rev. D 82, 064010 (2010).

[51] N. Cornish, L. Sampson, N. Yunes, and F. Pretorius,Gravitational wave tests of general relativity with theparameterized post-Einsteinian framework, Phys. Rev. D84, 062003 (2011).

[52] T. G. F. Li, W. Del Pozzo, S. Vitale, C. Van Den Broeck, M.Agathos, J. Veitch, K. Grover, T. Sidery, R. Sturani, and A.Vecchio, Towards a generic test of the strong field dynamicsof general relativity using compact binary coalescence,Phys. Rev. D 85, 082003 (2012).

[53] T. G. F. Li, W. Del Pozzo, S. Vitale, C. Van Den Broeck, M.Agathos, J. Veitch, K. Grover, T. Sidery, R. Sturani, and A.Vecchio, Towards a generic test of the strong field dynamicsof general relativity using compact binary coalescence:Further investigations, J. Phys. Conf. Ser. 363, 012028(2012).

[54] M. Agathos, W. Del Pozzo, T. G. F. Li, C. Van Den Broeck,J. Veitch, and S. Vitale, TIGER: A data analysis pipeline fortesting the strong-field dynamics of general relativity withgravitational wave signals from coalescing compact bina-ries, Phys. Rev. D 89, 082001 (2014).

[55] L. Sampson, N. Cornish, and N. Yunes, Gravitational wavetests of strong field general relativity with binary inspirals:Realistic injections and optimal model selection, Phys. Rev.D 87, 102001 (2013).

[56] H. Antil, S. E. Field, F. Herrmann, R. H. Nochetto, and M.Tiglio, Two-step greedy algorithm for reduced order quad-ratures, J. Sci. Comput. 57, 604 (2013).

[57] P. Canizares, S. E. Field, J. R. Gair, and M. Tiglio, Gravi-tational wave parameter estimation with compressed like-lihood evaluations, Phys. Rev. D 87, 124005 (2013).

[58] R. Smith, S. E. Field, K. Blackburn, C.-J. Haster, M. Prrer,V. Raymond, and P. Schmidt, Fast and accurate inference ongravitational waves from precessing compact binaries, Phys.Rev. D 94, 044031 (2016).

[59] S. E. Field, C. R. Galley, F. Herrmann, J. S. Hesthaven,E. Ochsner, and M. Tiglio, Reduced Basis Catalogs forGravitational Wave Templates, Phys. Rev. Lett. 106, 221102(2011).

[60] S. E. Field, C. R. Galley, J. S. Hesthaven, J. Kaye, and M.Tiglio, Fast Prediction and Evaluation of GravitationalWaveforms Using Surrogate Models, Phys. Rev. X 4,031006 (2014).

[61] M. Pürrer, Frequency domain reduced order models forgravitational waves from aligned-spin compact binaries,Classical Quantum Gravity 31, 195010 (2014).

[62] M. Pürrer, Frequency domain reduced order model ofaligned-spin effective-one-body waveforms with genericmass-ratios and spins, Phys. Rev. D 93, 064041 (2016).

[63] J. Blackman, S. E. Field, C. R. Galley, B. Szilgyi, M. A.Scheel, M. Tiglio, and D. A. Hemberger, Fast and Accurate

JEROEN MEIDAM et al. PHYS. REV. D 97, 044033 (2018)

044033-18

Page 19: PHYSICAL REVIEW D 97, 044033 (2018)orca.cf.ac.uk/110510/1/PhysRevD.97.044033.pdf · Jeroen Meidam,1 Ka Wa Tsang,1 Janna Goldstein,2 Michalis Agathos,3 Archisman Ghosh,1 Carl-Johan

Prediction of Numerical Relativity Waveforms from BinaryBlack Hole Coalescences Using Surrogate Models, Phys.Rev. Lett. 115, 121102 (2015).

[64] J. Blackman, S. E. Field, M. A. Scheel, C. R. Galley, D. A.Hemberger, P. Schmidt, and R. Smith, A surrogate model ofgravitational waveforms from numerical relativity simula-tions of precessing binary black hole mergers, Phys. Rev. D95, 104023 (2017).

[65] R. O’Shaughnessy, J. Blackman, and S. E. Field, Anarchitecture for efficient gravitational wave parameter esti-mation with multimodal linear surrogate models, ClassicalQuantum Gravity 34, 144002 (2017).

[66] J. Blackman, S. E. Field, M. A. Scheel, C. R. Galley, C. D.Ott, M. Boyle, L. E. Kidder, H. P. Pfeiffer, and B. Szilgyi,Numerical relativity waveform surrogate model for generi-cally precessing binary black hole mergers, Phys. Rev. D 96,024058 (2017).

[67] P. Schmidt, M. Hannam, and S. Husa, Towards modelsof gravitational waveforms from generic binaries: A simpleapproximate mapping between precessing and non-precessing inspiral signals, Phys. Rev. D 86, 104063 (2012).

[68] P. Schmidt, F. Ohme, and M. Hannam, Towards models ofgravitational waveforms from generic binaries II: Modellingprecession effects with a single effective precession param-eter, Phys. Rev. D 91, 024043 (2015).

[69] S. Khan (private communication).

[70] M. Campanelli, C. O. Lousto, Y. Zlochower, B. Krishnan,and D. Merritt, Spin flips and precession in black-hole-binary mergers, Phys. Rev. D 75, 064030 (2007).

[71] J. Veitch and A. Vecchio, Bayesian coherent analysis of in-spiral gravitational wave signals with a detector network,Phys. Rev. D 81, 062003 (2010).

[72] J. Veitch et al., Parameter estimation for compact binarieswith ground-based gravitational-wave observations usingthe LALInference software library, Phys. Rev. D 91, 042003(2015).

[73] B. P. Abbott et al., Report No. LIGO-T0900288-v3, 2015.[74] B. P. Abbott et al., Prospects for observing and localizing

gravitational-wave transients with Advanced LIGO, Ad-vanced Virgo and KAGRA, Living Rev. Relativity 19, 1(2016).

[75] https://data.black-holes.org/waveforms/index.html.[76] C. R. Galley and P. Schmidt, Report No. LIGO-P1600064,

2016.[77] P. Schmidt, I. W. Harry, and H. P. Pfeiffer, Report

No. LIGO-T1500606, 2017.[78] B. P. Abbott et al., Effects of waveform model systematics

on the interpretation of GW150914, Classical QuantumGravity 34, 104002 (2017).

[79] N. Yunes and X. Siemens, Gravitational-wave tests ofgeneral relativity with ground-based detectors and pulsartiming-arrays, Living Rev. Relativity 16, 9 (2013).

PARAMETRIZED TESTS OF THE STRONG-FIELD … PHYS. REV. D 97, 044033 (2018)

044033-19