Sky marginalization in black hole spectroscopy and tests of the area theorem

Alex Correia University of Massachusetts Dartmouth, 285 Old Westport Rd, North Dartmouth, MA 02747    Collin D. Capano [email protected] Department of Physics, Syracuse University, Syracuse, NY 13244 University of Massachusetts Dartmouth, 285 Old Westport Rd, North Dartmouth, MA 02747
Abstract

Direct observation of gravitational waves from binary black hole (BBH) mergers has made it possible to test the laws of black-hole thermodynamics using real astrophysical sources. These tests rely on accurate and unbiased parameter estimates from the pre- and post-merger portions of a signal. Due to numerical complications, previous analyses have fixed the sky location and coalescence time when independently estimating the parameters of the pre- and post-merger signal. Here we overcome the numerical complications and present a novel method of marginalizing over sky location and coalescence time. Doing so, we find that it is not possible to model only the pre- or post-merger portions of the signal while marginalizing over timing uncertainty. We surmount this problem by simultaneously yet independently modelling the pre- and post-merger signal, with only the sky location and coalescence time being shared between the models. This allows us to marginalize over all parameters. We use our method to measure the change in area ΔAmeasured=AfAiΔsubscript𝐴measuredsubscript𝐴𝑓subscript𝐴𝑖\Delta A_{\rm measured}=A_{f}-A_{i}roman_Δ italic_A start_POSTSUBSCRIPT roman_measured end_POSTSUBSCRIPT = italic_A start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT between the final and initial black holes in the BBH merger GW150914. To measure the final black hole’s area Afsubscript𝐴𝑓A_{f}italic_A start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT we do an analysis using quasi-normal modes (QNMs) to model the post-merger signal, and another analysis using the post-merger portion of an inspiral-merger-ringdown (IMR) template. We find excellent agreement with expectations from General Relativity. The Hawking area theorem (which states that AfAisubscript𝐴𝑓subscript𝐴𝑖A_{f}\geq A_{i}italic_A start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ≥ italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT) is confirmed to 95.4%percent95.495.4\%95.4 % and 99.5%percent99.599.5\%99.5 % confidence using the QNM and IMR post-merger models, respectively. Both models yield ΔAmeasured/ΔAexpected1similar-toΔsubscript𝐴measuredΔsubscript𝐴expected1\Delta A_{\rm measured}/\Delta A_{\rm expected}\sim 1roman_Δ italic_A start_POSTSUBSCRIPT roman_measured end_POSTSUBSCRIPT / roman_Δ italic_A start_POSTSUBSCRIPT roman_expected end_POSTSUBSCRIPT ∼ 1, where ΔAexpectedΔsubscript𝐴expected\Delta A_{\rm expected}roman_Δ italic_A start_POSTSUBSCRIPT roman_expected end_POSTSUBSCRIPT is the expected change in area derived from fits to numerical relativity simulations.

I Introduction

The direct detection of gravitational waves (GWs) throughout the past decade has been one of the most significant advancements in observational relativity. These observations not only confirmed the existence of GWs, but also opened the door to direct tests of General Relativity (GR) in the strong-field regime. One such prediction that can be verified is Hawking’s area theorem, which states that the remnant from a binary-black hole (BBH) merger must have an event horizon surface area greater than the sum of the progenitor horizon areas [1].

Tests of the area theorem using GWs have been carried out previously [2, 3, 4]. Generally, they involve measuring the initial two black holes’ mass and spin (from which the total initial area Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is derived) from the pre-merger part of the signal, during which the two black holes inspiral into each other. The area of the final black hole is independently measured using the GW that is emitted during the post-merger, or “ringdown” phase. In both cases Bayesian inference is used to produce a “posterior” probability density on the black holes’ parameters. Ideally, all parameters that describe the BBH should be allowed to vary in the analysis, in order to fully account for all statistical uncertainties. However, previous tests of the area theorem have fixed the sky position and coalescence time tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT of the events to nominal values when doing their analysis [2, 4]. Fixing the values in this way may lead to biases in the resultant parameter estimates, obfuscating the true nature of the system [3, 5]. At the very least, it may cause an underestimate of the statistical uncertainty of measured parameters, yielding constraints on deviations from GR that are misleadingly strong.

The sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT have been fixed in earlier studies due to technical hurdles in calculating the likelihood function. In order to independently analyze the pre- and post-merger portion of the signal it is necessary to excise the post- and pre-merger data, respectively, from the analysis. Excising the pre-merger data is also necessary when just analyzing the post-merger signal for tests of the no-hair theorem using quasi-normal modes (known as black hole spectroscopy). In either case, the excision (or “gate”) results in a modified likelihood function that cannot be solved using conventional, frequency-domain means. Existing pipelines for doing such analyses in the time domain, such as pyring [6, 7] and ringdown [8], instead calculate the likelihood by numerically inverting the noise covariance matrix for the data. This calculation can be numerically costly [9, 5]. If the sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT are not fixed, the gate will vary in time, and the full likelihood will need to be recalculated for every unique gate time, significantly increasing computational costs. However, if the sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT are fixed, the gate is static, and the most computationally demanding part of the calculation need only be done once.

Finch & Moore [10] devised a method to overcome this issue and vary the sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In their method the inspiral and merger are modeled with a wavelet series stitched to the beginning of the ringdown. This allows them to use the traditional frequency-domain likelihood, and vary the start time of the ringdown. However, this does not allow for an area theorem test, since the parameters of the initial black holes cannot be estimated from the wavelet signal model. Furthermore, the traditional frequency-domain likelihood causes the start of the ringdown to be coupled to information from just before the merger due to convolution with the whitening filter. This means that recovered pre- and post-merger parameters will not be independent measurements.

This paper presents a novel method of calculating the likelihood function using the parameter estimation code PyCBC Inference [11]. This code already utilizes “gating and in-painting” [12] to excise data from the analysis, which allows for cheaper likelihood evaluation under certain conditions as compared to the methods used by pyring and ringdown [9, 5]. However, an additional normalization factor has traditionally been omitted from this calculation, as this also required computationally expensive numerical methods to evaluate. This paper presents a method of linear interpolation that calculates this normalization factor with good approximation. Together, these methods allow for fast likelihood calculations regardless of gate position, allowing for marginalization over tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and sky location and a full accounting of parameter uncertainties.

This paper is structured as follows. Section II describes the modifications to the likelihood calculation in PyCBC in detail. Full waveform analyses can be conducted with these modifications to marginalize over sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. However, the method cannot be used in partial waveform analyses; Section III documents the issues that arise in these models. Using this method, a test of the Hawking area theorem is conducted using data from GW150914 [13]. Section IV describes the configuration of the analyses used in this test, and Section V describes and analyzes the results. Section VI summarizes the findings and the implications of the marginalization method for future GW analyses.

II Determinant approximation

PyCBC Inference [11] utilizes Bayesian inference to conduct parameter estimation on GW events. Bayes’ theorem is used to extract information about the parameter space ϑbold-italic-ϑ\bm{\vartheta}bold_italic_ϑ from a data set 𝒔𝒔\bm{s}bold_italic_s [14]. Assuming the data set is composed of a signal hhitalic_h and noise n𝑛nitalic_n such that 𝒔=𝒏+𝒉𝒔𝒏𝒉\bm{s}=\bm{n}+\bm{h}bold_italic_s = bold_italic_n + bold_italic_h, the probability of observing a specific ϑbold-italic-ϑ\bm{\vartheta}bold_italic_ϑ is given by

p(ϑ|𝒔,h)=p(𝒔|ϑ,h)p(ϑ|h)p(𝒔|h).𝑝conditionalbold-italic-ϑ𝒔𝑝conditional𝒔bold-italic-ϑ𝑝conditionalbold-italic-ϑ𝑝conditional𝒔p(\bm{\vartheta}|\bm{s},h)=\frac{p(\bm{s}|\bm{\vartheta},h)p(\bm{\vartheta}|h)% }{p(\bm{s}|h)}.italic_p ( bold_italic_ϑ | bold_italic_s , italic_h ) = divide start_ARG italic_p ( bold_italic_s | bold_italic_ϑ , italic_h ) italic_p ( bold_italic_ϑ | italic_h ) end_ARG start_ARG italic_p ( bold_italic_s | italic_h ) end_ARG . (1)

The term p(𝒔|ϑ,h)𝑝conditional𝒔bold-italic-ϑp(\bm{s}|\bm{\vartheta},h)italic_p ( bold_italic_s | bold_italic_ϑ , italic_h ) is the likelihood, which describes the probability of observing a signal 𝒔𝒔\bm{s}bold_italic_s assuming the event has a parameter space ϑbold-italic-ϑ\bm{\vartheta}bold_italic_ϑ. The term p(ϑ|h)𝑝conditionalbold-italic-ϑp(\bm{\vartheta}|h)italic_p ( bold_italic_ϑ | italic_h ) is the prior, a distribution that describes the a priori probability of observing ϑbold-italic-ϑ\bm{\vartheta}bold_italic_ϑ given a signal model hhitalic_h. The denominator p(𝒔|h)𝑝conditional𝒔p(\bm{s}|h)italic_p ( bold_italic_s | italic_h ) is the evidence, a normalization factor used to compare analyses using different models for hhitalic_h. The resultant probability distribution on the left-hand side of the equation is known as the posterior.

The prior is chosen at the discretion of the analyst based on assumed plausible values for ϑbold-italic-ϑ\bm{\vartheta}bold_italic_ϑ, whereas the likelihood is calculated directly from the data. For a system of K𝐾Kitalic_K detectors each taking N𝑁Nitalic_N time series samples of an event with a stochastic Gaussian noise background, the likelihood can be written as [15]

p(𝒔net|n)=exp[12d=1K𝒔dT𝚺d1𝒔d][(2π)NKd=1Kdet𝚺d]1/2,𝑝conditionalsubscript𝒔𝑛𝑒𝑡𝑛12superscriptsubscript𝑑1𝐾subscriptsuperscript𝒔𝑇𝑑subscriptsuperscript𝚺1𝑑subscript𝒔𝑑superscriptdelimited-[]superscript2𝜋𝑁𝐾superscriptsubscriptproduct𝑑1𝐾subscript𝚺𝑑12p(\bm{s}_{net}|n)=\frac{\exp{[-\frac{1}{2}\sum_{d=1}^{K}\bm{s}^{T}_{d}\bm{% \Sigma}^{-1}_{d}\bm{s}_{d}]}}{[(2\pi)^{NK}\prod_{d=1}^{K}\det\bm{\Sigma}_{d}]^% {1/2}},italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | italic_n ) = divide start_ARG roman_exp [ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT bold_italic_s start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ] end_ARG start_ARG [ ( 2 italic_π ) start_POSTSUPERSCRIPT italic_N italic_K end_POSTSUPERSCRIPT ∏ start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT roman_det bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG , (2)

where 𝚺dsubscript𝚺𝑑\bm{\Sigma}_{d}bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is the covariance matrix associated with n𝑛nitalic_n in detector d𝑑ditalic_d.

To evaluate the likelihood, further assumptions must be made to calculate 𝚺d1superscriptsubscript𝚺𝑑1\bm{\Sigma}_{d}^{-1}bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The result of these assumptions is:

p(𝒔net|n)exp[12d=1K𝒔d,𝒔d],proportional-to𝑝conditionalsubscript𝒔𝑛𝑒𝑡𝑛12subscriptsuperscript𝐾𝑑1subscript𝒔𝑑subscript𝒔𝑑p(\bm{s}_{net}|n)\propto\exp\bigg{[}-\frac{1}{2}\sum^{K}_{d=1}\langle\bm{s}_{d% },\bm{s}_{d}\rangle\bigg{]},italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | italic_n ) ∝ roman_exp [ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT ⟨ bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT , bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ⟩ ] , (3)

where the inner product is defined by Eq. (22). (The full derivation of this likelihood function is given in Appendix A). This gives the likelihood function for a signal that is assumed to be entirely noise. To get the likelihood function with respect to the signal model hhitalic_h evaluated for a parameter space ϑbold-italic-ϑ\bm{\vartheta}bold_italic_ϑ, Eq. (3) can be rewritten by substituting 𝒔d𝒔d𝒉d(ϑ)subscript𝒔𝑑subscript𝒔𝑑subscript𝒉𝑑bold-italic-ϑ\bm{s}_{d}\rightarrow\bm{s}_{d}-\bm{h}_{d}(\bm{\vartheta})bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT → bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - bold_italic_h start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( bold_italic_ϑ ) on the right-hand side, yielding

logp(𝒔net|ϑ,h)=𝑝conditionalsubscript𝒔𝑛𝑒𝑡bold-italic-ϑabsent\displaystyle\log p(\bm{s}_{net}|\bm{\vartheta},h)=roman_log italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | bold_italic_ϑ , italic_h ) = 12d=1K𝒔d𝒉d(ϑ),𝒔d𝒉d(ϑ)12subscriptsuperscript𝐾𝑑1subscript𝒔𝑑subscript𝒉𝑑bold-italic-ϑsubscript𝒔𝑑subscript𝒉𝑑bold-italic-ϑ\displaystyle-\frac{1}{2}\sum^{K}_{d=1}\langle\bm{s}_{d}-\bm{h}_{d}(\bm{% \vartheta}),\bm{s}_{d}-\bm{h}_{d}(\bm{\vartheta})\rangle- divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT ⟨ bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - bold_italic_h start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( bold_italic_ϑ ) , bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - bold_italic_h start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( bold_italic_ϑ ) ⟩
12[NKlog2π+d=1Klogdet𝚺d].12delimited-[]𝑁𝐾2𝜋subscriptsuperscript𝐾𝑑1subscript𝚺𝑑\displaystyle-\frac{1}{2}\bigg{[}NK\log 2\pi+\sum^{K}_{d=1}\log\det\bm{\Sigma}% _{d}\bigg{]}.- divide start_ARG 1 end_ARG start_ARG 2 end_ARG [ italic_N italic_K roman_log 2 italic_π + ∑ start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT roman_log roman_det bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ] . (4)

(Here and throughout, we use log\logroman_log to refer to the natural logarithm.) For a GW analysis that examines the entirety of 𝒔𝒔\bm{s}bold_italic_s, this expression is easily evaluated using the approximated eigenvalues of 𝚺dsubscript𝚺𝑑\bm{\Sigma}_{d}bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT. However, many analyses only examine a portion of the time series (generally either the pre- or post-merger signal). Therefore, a region of 𝒔𝒔\bm{s}bold_italic_s is omitted, or “gated”, to reduce numerical biases due to the merger. Doing this also requires excising rows and columns from 𝚺dsubscript𝚺𝑑\bm{\Sigma}_{d}bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, thereby breaking the Toeplitz form of the matrix and the corresponding approximations. The inverse of the truncated covariance matrix 𝚺d,trsubscript𝚺𝑑𝑡𝑟\bm{\Sigma}_{d,tr}bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT is evaluated in PyCBC using “gating and in-painting” (explained in detail in [12] and Appendix B), allowing for easy calculation of the first term of Eq. (4).

The second term, however, requires the calculation of logdet𝚺d,trsubscript𝚺𝑑𝑡𝑟\log\det\bm{\Sigma}_{d,tr}roman_log roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT, which is a notoriously expensive numerical problem. The most widely-used matrix decomposition methods are O(N3)similar-toabsent𝑂superscript𝑁3\sim\!O(N^{3})∼ italic_O ( italic_N start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) [16], which become impractical for GW time series such as GW150914, where N104similar-to𝑁superscript104N\!\sim\!10^{4}italic_N ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT. This problem is amplified when varying sky position and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in a gated analysis. The start and end times of the gate must be converted between the geocentric and individual detector frames. The time conversions directly depend on the time of merger tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and the sky location of the event relative to the detectors. Therefore, when varying sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the gate times in the detector frames will also vary. Subsequently, the elements of 𝚺d,trsubscript𝚺𝑑𝑡𝑟\bm{\Sigma}_{d,tr}bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT will vary with each unique sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT value. Analyses that marginalize over these parameters require recalculating logdet𝚺d,trsubscript𝚺𝑑𝑡𝑟\log\det\bm{\Sigma}_{d,tr}roman_log roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT on multiple steps of the sampler, which would be impractical using numerical decomposition methods. This is not an issue if the sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT are fixed, since in that case det𝚺d,trsubscript𝚺𝑑𝑡𝑟\det\bm{\Sigma}_{d,tr}roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT will not change throughout the analysis (and in fact can be ignored, as it amounts to a constant normalization term).

In order to marginalize over sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT it is necessary to have a fast method for calculating det𝚺d,trsubscript𝚺𝑑𝑡𝑟\det\bm{\Sigma}_{d,tr}roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT. We find that logdet𝚺d,trsubscript𝚺𝑑𝑡𝑟\log\det\bm{\Sigma}_{d,tr}roman_log roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT is strongly linearly correlated to the length of the power spectral density (PSD) of the data, which is equivalent the number of rows and columns in 𝚺d,trsubscript𝚺𝑑𝑡𝑟\bm{\Sigma}_{d,tr}bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT. Fig. 1 shows this relationship using logdet𝚺d,trsubscript𝚺𝑑𝑡𝑟\log\det\bm{\Sigma}_{d,tr}roman_log roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT values calculated for various gate sizes applied to GW150914 data. Using a least squares fit to the points, the correlation coefficient was calculated as 0.999<R2<10.999superscript𝑅210.999<R^{2}<10.999 < italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT < 1, which for a sample of 8 points implies a highly significant correlation [17].

Refer to caption
Figure 1: Plot of logdet𝚺d,trsubscript𝚺𝑑𝑡𝑟\log\det\bm{\Sigma}_{d,tr}roman_log roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT versus PSD length for GW150914 in the Hanford detector. Blue points represent determinants calculated exactly using SciPy functions [18], while the black dashed line is a least squares linear fit. The correlation coefficient of 0.999<R2<10.999superscript𝑅210.999<R^{2}<10.999 < italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT < 1 indicates a highly significant linear correlation between logdet𝚺d,trsubscript𝚺𝑑𝑡𝑟\log\det\bm{\Sigma}_{d,tr}roman_log roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT and the size of 𝚺d,trsubscript𝚺𝑑𝑡𝑟\bm{\Sigma}_{d,tr}bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT.

We also find that the position of the gate in the time series has no significant effect on the value of logdet𝚺d,trsubscript𝚺𝑑𝑡𝑟\log\det\bm{\Sigma}_{d,tr}roman_log roman_det bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT. The maximum range over which the determinant values varied was 106similar-toabsentsuperscript106\sim\!10^{-6}∼ 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT. However, since these values were generally of order 106similar-toabsentsuperscript106\sim\!10^{6}∼ 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT, this represents a maximum fractional change of 1012similar-toabsentsuperscript1012\sim\!10^{-12}∼ 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPT over the entire time domain. While gate position does have a minor effect on determinant value, the relative effect is so small that the value is well-approximated as a constant for a static gate length.

Using these two facts, we calculate the normalization term in Eq. (4) using a linear interpolation based solely on the size of 𝚺d,trsubscript𝚺𝑑𝑡𝑟\bm{\Sigma}_{d,tr}bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT. Specifically, a least squares linear fit is generated using the determinant values for the full matrix and three differently-sized truncated matrices. The full matrix determinant is calculated using its approximated eigenvalues, while the truncated determinants are calculated numerically using SciPy [18]. The determinant can then be calculated on each step through linear interpolation based on the size of the truncated covariance matrix.

III Need for Joint Analyses

Refer to caption
Refer to caption
Figure 2: Comparison of waveform templates fitting to a simulated signal similar to GW150914 in the Hanford detector. The simulated signal (“injection”) is generated using the maximum likelihood values from a 4-OGC analysis of GW150914 [19] and added to zero noise. The ungated zero-noise injection is plotted with a gray line on each set of axes. The plots on the left (right) contain maximum likelihood pre-merger (post-merger) waveform templates plotted with blue (red) lines. In each analysis, a 1-second gate is applied to the data, represented by a shaded gray region and a dashed green line representing the start (end) of the gate. The data with this gate applied is shown with a black line. The horizontal axes represent the time in seconds relative to the geocentric gate start (end) time. The top plots show the waveform templates for analyses where the sky location was fixed to the same values as the injection. The center plots show the templates for analyses where the sky location was allowed to vary and the pre- and post-merger signals were modeled separately. Notice that the pre-merger (post-merger) gate is significantly earlier (later) than the injected values shown in the corresponding fixed sky location templates. The bottom plots show the templates from an analysis with variable sky location where the pre- and post-merger signals were modeled simultaneously. Besides sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the parameter distributions were independent of each other. The resultant gate positions are similar to those observed in the fixed sky location analyses.

The strategy outlined in Sec. II fixes the technical hurdle of calculating the likelihood when the gate length and position are varied. Even with that, however, we find that it is not possible to only model a portion of a signal — whether it be the post-merger or the pre-merger — if tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT or sky location are uncertain. This is because a larger likelihood can generally be obtained if the gate is shifted such that it removes as much of a signal as possible.

From Eq. (4) it is evident that the likelihood is maximized as 𝒔d𝒉d(ϑ)0subscript𝒔𝑑subscript𝒉𝑑bold-italic-ϑ0\bm{s}_{d}-\bm{h}_{d}(\bm{\vartheta})\rightarrow 0bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - bold_italic_h start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( bold_italic_ϑ ) → 0. When analyzing the entire time series this can only occur if the template 𝒉d(ϑ)subscript𝒉𝑑bold-italic-ϑ\bm{h}_{d}(\bm{\vartheta})bold_italic_h start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ( bold_italic_ϑ ) is similar to the signal that exists in the data (as desired). However, if a variable gate is involved, then a large likelihood can also be obtained if the gate is shifted such that it excises most (or all) of the signal in the data. The template then only needs to match the remaining noise. Since the noise manifold is larger than the signal manifold, a larger volume of prior space will generally be able to match the noise than the volume that matches the signal. For example, when using a QNM model, if the gate is shifted such that it excises the entire signal then one only needs to reduce the amplitude of the QNM template below noise level to get a relatively large likelihood. The same likelihood would then be obtained regardless of what the other parameters describing the template are set to. The end result is that the posterior probability will favor excising the entire signal even if the template that is being used can match the signal.

Specifically, in pre-merger-only analyses, the gate will be shifted to as early a time that is possible while in post-merger-only analyses, the gate will be shifted to as late as possible. These phenomena are illustrated in the middle row of Fig. 2. Shown are the results from separate pre-merger and post-merger analyses of a simulated signal in which the gate time is varied due to time and sky-location marginalization. The maximum likelihood template in the pre-merger-only analysis has a tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roughly 25ms25ms25\text{ms}25 ms earlier than what was injected. This significantly shortens the pre-merger signal, leading to inaccurate measurements in parameters such as total spin. Meanwhile, the maximum likelihood template for the post-merger-only analysis had a ringdown start time 10Msimilar-toabsent10subscript𝑀direct-product\sim\!10M_{\odot}∼ 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT later than the injected value, well into the regime where the signal is expected to be noise-dominated. This leads to a “prior-in prior-out” posterior – a roughly uniform distribution across parameter space – as the signal model can fit any arbitrary model to late-time noise.

Fundamentally, this problem is due to the gated signal model not being the appropriate model for the observed data. Excising data from the analysis is mathematically equivalent to assuming that the excised data is Gaussian noise and marginalizing over all possible realizations of it. A property of multivariate Gaussian distributions is that marginalizing over a subset of dimensions yields another multivariate Gaussian distribution with the marginalized dimensions excised from the covariance matrix. This is exactly the same form as Eq. (3); in our case, each time sample is a dimension in the multivariate Gaussian. The problem is the excised times are not Gaussian noise. They contain a signal, albeit a portion of the signal that we want to ignore. Consequently, a gated signal model is not representative of the data.

This issue could be mitigated by modifying the prior to exclude the portion of parameter space that matches the noise while keeping the portion that matches the signal. With a QNM template this would mean setting a lower bound on the amplitude such that it is above noise level. However, modifying the prior in this manner is challenging to do in an unbiased way, as it would require a priori knowledge of both the noise and signal properties. It also does not solve the fundamental problem that the gated signal model is not representative of the data.

We resolve this issue by simultaneously yet independently modeling both the pre- and post-merger signals. Each domain is treated as a separate gated analysis. The signal parameters used in each domain are independent of each other, except for a set of common parameters ϑcomsubscriptbold-italic-ϑcom\bm{\vartheta}_{\rm com}bold_italic_ϑ start_POSTSUBSCRIPT roman_com end_POSTSUBSCRIPT. In our case the common parameters are the right ascension α𝛼\alphaitalic_α, declination δ𝛿\deltaitalic_δ, and geocentric coalescence time tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. These parameters determine the coalescence time tcdsubscriptsuperscript𝑡𝑑𝑐t^{d}_{c}italic_t start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in each detector d𝑑ditalic_d, which the gate time depends on. The likelihood is calculated in each domain, then combined to get the overall likelihood.

Explicitly, our algorithm to calculate the likelihood for a given set of parameter values ϑ={ϑinsp,ϑrd,ϑcom}bold-italic-ϑsubscriptbold-italic-ϑinspsubscriptbold-italic-ϑrdsubscriptbold-italic-ϑcom\bm{\vartheta}=\{\bm{\vartheta}_{\rm insp},\bm{\vartheta}_{\rm rd},\bm{% \vartheta}_{\rm com}\}bold_italic_ϑ = { bold_italic_ϑ start_POSTSUBSCRIPT roman_insp end_POSTSUBSCRIPT , bold_italic_ϑ start_POSTSUBSCRIPT roman_rd end_POSTSUBSCRIPT , bold_italic_ϑ start_POSTSUBSCRIPT roman_com end_POSTSUBSCRIPT } is:

  1. 1.

    Generate the pre-merger (“inspiral”) template hinspsuperscriptinsph^{\rm insp}italic_h start_POSTSUPERSCRIPT roman_insp end_POSTSUPERSCRIPT using parameters ϑinspsubscriptbold-italic-ϑinsp\bm{\vartheta}_{\rm insp}bold_italic_ϑ start_POSTSUBSCRIPT roman_insp end_POSTSUBSCRIPT.

  2. 2.

    Project the template into each detector’s frame using ϑcom={α,δ,tc}subscriptbold-italic-ϑcom𝛼𝛿subscript𝑡𝑐\bm{\vartheta}_{\rm com}=\{\alpha,\delta,t_{c}\}bold_italic_ϑ start_POSTSUBSCRIPT roman_com end_POSTSUBSCRIPT = { italic_α , italic_δ , italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT }.

  3. 3.

    Excise times after tcdetsubscriptsuperscript𝑡det𝑐t^{\rm det}_{c}italic_t start_POSTSUPERSCRIPT roman_det end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT by gating and in-painting the residual sdhdinsp(ϑinsp,ϑcom)subscript𝑠𝑑superscriptsubscript𝑑inspsubscriptbold-italic-ϑinspsubscriptbold-italic-ϑcoms_{d}-h_{d}^{\rm insp}(\bm{\vartheta}_{\rm insp},\bm{\vartheta}_{\rm com})italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT - italic_h start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_insp end_POSTSUPERSCRIPT ( bold_italic_ϑ start_POSTSUBSCRIPT roman_insp end_POSTSUBSCRIPT , bold_italic_ϑ start_POSTSUBSCRIPT roman_com end_POSTSUBSCRIPT ) using a 111\,1s gate that spans [tcd,tcd+1)subscriptsuperscript𝑡𝑑𝑐subscriptsuperscript𝑡𝑑𝑐1[t^{d}_{c},t^{d}_{c}+1)[ italic_t start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_t start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT + 1 ).

  4. 4.

    Calculate the pre-merger likelihood p(𝒔net|ϑinsp,ϑcom,hinsp)𝑝conditionalsubscript𝒔𝑛𝑒𝑡subscriptbold-italic-ϑinspsubscriptbold-italic-ϑcomsuperscriptinspp(\bm{s}_{net}|\bm{\vartheta}_{\rm insp},\bm{\vartheta}_{\rm com},h^{\rm insp})italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | bold_italic_ϑ start_POSTSUBSCRIPT roman_insp end_POSTSUBSCRIPT , bold_italic_ϑ start_POSTSUBSCRIPT roman_com end_POSTSUBSCRIPT , italic_h start_POSTSUPERSCRIPT roman_insp end_POSTSUPERSCRIPT ) via Eq. (4).

  5. 5.

    Repeat steps 1-4 for the post-merger (“ringdown”) template hrdsuperscriptrdh^{\rm rd}italic_h start_POSTSUPERSCRIPT roman_rd end_POSTSUPERSCRIPT to get the post-merger likelihood p(𝒔net|ϑrd,ϑcom,hrd)𝑝conditionalsubscript𝒔𝑛𝑒𝑡subscriptbold-italic-ϑrdsubscriptbold-italic-ϑcomsuperscriptrdp(\bm{s}_{net}|\bm{\vartheta}_{\rm rd},\bm{\vartheta}_{\rm com},h^{\rm rd})italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | bold_italic_ϑ start_POSTSUBSCRIPT roman_rd end_POSTSUBSCRIPT , bold_italic_ϑ start_POSTSUBSCRIPT roman_com end_POSTSUBSCRIPT , italic_h start_POSTSUPERSCRIPT roman_rd end_POSTSUPERSCRIPT ). However, for the ringdown the gate spans [tcd1,tcd)subscriptsuperscript𝑡𝑑𝑐1subscriptsuperscript𝑡𝑑𝑐[t^{d}_{c}-1,t^{d}_{c})[ italic_t start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - 1 , italic_t start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ); i.e., it ends at tcdsubscriptsuperscript𝑡𝑑𝑐t^{d}_{c}italic_t start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT whereas it starts at tcdsubscriptsuperscript𝑡𝑑𝑐t^{d}_{c}italic_t start_POSTSUPERSCRIPT italic_d end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for the pre-merger. The data used in each domain is thereby (nearly) mutually exclusive (see Appendix B for more details).

  6. 6.

    The total likelihood is then

    p(𝒔net|ϑ,h)𝑝conditionalsubscript𝒔𝑛𝑒𝑡bold-italic-ϑ\displaystyle p(\bm{s}_{net}|\bm{\vartheta},h)italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | bold_italic_ϑ , italic_h ) =\displaystyle==
    p(𝒔net\displaystyle p(\bm{s}_{net}italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT |ϑinsp,ϑcom,hinsp)p(𝒔net|ϑrd,ϑcom,hrd)\displaystyle|\bm{\vartheta}_{\rm insp},\bm{\vartheta}_{\rm com},h^{\rm insp})% p(\bm{s}_{net}|\bm{\vartheta}_{\rm rd},\bm{\vartheta}_{\rm com},h^{\rm rd})| bold_italic_ϑ start_POSTSUBSCRIPT roman_insp end_POSTSUBSCRIPT , bold_italic_ϑ start_POSTSUBSCRIPT roman_com end_POSTSUBSCRIPT , italic_h start_POSTSUPERSCRIPT roman_insp end_POSTSUPERSCRIPT ) italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | bold_italic_ϑ start_POSTSUBSCRIPT roman_rd end_POSTSUBSCRIPT , bold_italic_ϑ start_POSTSUBSCRIPT roman_com end_POSTSUBSCRIPT , italic_h start_POSTSUPERSCRIPT roman_rd end_POSTSUPERSCRIPT ) (5)

We fix the issue of the gate trying to excise the signal by using this hierarchical likelihood. The two domains offset each other: in order for the post-merger gate to shift to later times the pre-merger template must match more of the signal, and vice versa. This also addresses the fundamental issue with the gated signal model highlighted above: our global model for the entire data set now contains a non-Gaussian element (the other domain’s signal model) in each domain’s excised region. Note also that no coupling occurs across the domain boundaries due to the whitening filter.

The bottom plots in Fig. 2 show the maximum likelihood waveform templates from an analysis with this configuration. Besides sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, both models generated independent parameter measurements. The resultant waveform templates closely match those obtained by fixing sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to the injected values (top row of Fig.  2); the erroneous gate motion observed in the pre- and post-merger-only analyses is no longer present.

IV Methods

We performed eight analyses in total. Each analysis was differentiated by the waveform used, post-merger approximant, and whether or not sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT were marginalized over. All other aspects of the analyses were kept constant. Both the pre- and post-merger signals in all models utilized the PyCBC model GatedGaussianMargPol, which inherits the normalization protocols described in Section II. The dynesty sampler [20] was used to generate posterior distributions. The samplers in these analyses used 4000 live points to ensure the convergence of each model (see Appendix C).

Half of the analyses used the original data of GW150914 with a sample rate of 2048 Hz obtained from the Gravitational Wave Open Science Center [21]. To validate these results, we repeat each run on a simulated signal in zero noise. The simulated signal was generated using the IMRPhenomXPHM waveform approximant [22] with the maximum likelihood parameters for GW150914 from Ref. [19].

All template models utilized IMRPhenomXPHM (abbreviated here on as IMR) to model the pre-merger signal of the waveform. Half of the models used this IMR approximant to model the post-merger signal. The other half utilized a quasinormal mode (QNM) approximant to model the post-merger section of the waveform. Table 1 lists the sampled parameters and priors used in the pre-merger models, and Table 2 lists the same for the post-merger models. The QNM approximant was configured such that the ringdown was composed of a dominant (2,2,0)220(2,2,0)( 2 , 2 , 0 ) mode and a subdominant (2,2,1)221(2,2,1)( 2 , 2 , 1 ) mode as proposed by [23]. All models apply the ringdown model starting at merger time tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. The priors of the QNM post-merger models were restricted such that the (2,2,0)220(2,2,0)( 2 , 2 , 0 ) mode contribution to the post-merger signal-to-noise ratio (SNR) was at least 2. This condition was imposed to ensure that the dominant mode was present in the model, preventing possible “label-switching” that may occur due to the (2,2,1)221(2,2,1)( 2 , 2 , 1 ) mode erroneously matching to the (2,2,0)220(2,2,0)( 2 , 2 , 0 ) mode in the signal.

Only half of the models allowed for sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to vary. The other half fixed these parameters to nominal values for GW150914 to replicate previous works. Specifically, the fixed parameter analyses set α=1.95𝛼1.95\alpha=1.95italic_α = 1.95, δ=1.27𝛿1.27\delta=-1.27italic_δ = - 1.27, and tc=1126259462.408subscript𝑡𝑐1126259462.408t_{c}=1126259462.408italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 1126259462.408, in accordance with the maximum likelihood values in [23].

Parameter Description Prior Dist. Prior Range
tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT Coalescence time uniform 1126259462.43+[0.05,0.05]1126259462.430.050.051126259462.43+[-0.05,0.05]1126259462.43 + [ - 0.05 , 0.05 ] s
α𝛼\alphaitalic_α Right ascension uniform [0,2π]02𝜋[0,2\pi][ 0 , 2 italic_π ]
δ𝛿\deltaitalic_δ Declination sine angle [π/2,π/2]𝜋2𝜋2[-\pi/2,\pi/2][ - italic_π / 2 , italic_π / 2 ]
ι𝜄\iotaitalic_ι Inclination sine angle [0,π]0𝜋[0,\pi][ 0 , italic_π ]
Mchirpsubscript𝑀𝑐𝑖𝑟𝑝M_{chirp}italic_M start_POSTSUBSCRIPT italic_c italic_h italic_i italic_r italic_p end_POSTSUBSCRIPT Source frame chirp mass M1subscript𝑀1M_{1}italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, M2subscript𝑀2M_{2}italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT [23,42]M2342subscript𝑀direct-product[23,42]M_{\odot}[ 23 , 42 ] italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT
q𝑞qitalic_q Mass ratio M1/M2subscript𝑀1subscript𝑀2M_{1}/M_{2}italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT M1subscript𝑀1M_{1}italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, M2subscript𝑀2M_{2}italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT [1,4]14[1,4][ 1 , 4 ]
χa,(1/2)subscript𝜒𝑎12\chi_{a,(1/2)}italic_χ start_POSTSUBSCRIPT italic_a , ( 1 / 2 ) end_POSTSUBSCRIPT Spin magnitude uniform [0,0.99]00.99[0,0.99][ 0 , 0.99 ]
χθ,(1/2)subscript𝜒𝜃12\chi_{\theta,(1/2)}italic_χ start_POSTSUBSCRIPT italic_θ , ( 1 / 2 ) end_POSTSUBSCRIPT Spin polar angle solid angle [0,2π]02𝜋[0,2\pi][ 0 , 2 italic_π ]
χϕ,(1/2)subscript𝜒italic-ϕ12\chi_{\phi,(1/2)}italic_χ start_POSTSUBSCRIPT italic_ϕ , ( 1 / 2 ) end_POSTSUBSCRIPT Spin azimuthal angle solid angle [0,π]0𝜋[0,\pi][ 0 , italic_π ]
ϕcsubscriptitalic-ϕ𝑐\phi_{c}italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT Reference phase uniform [0,2π]02𝜋[0,2\pi][ 0 , 2 italic_π ]
VCsubscript𝑉𝐶V_{C}italic_V start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT Comoving volume uniform [5000,92918664351]500092918664351[5000,92918664351][ 5000 , 92918664351 ] Mpc3
Table 1: Varied parameters in IMR models and their associated prior distributions. The subscript (1/2)12(1/2)( 1 / 2 ) indicates that the same prior was used for the primary and secondary masses. The third column indicates the sampling method for each prior. Parameters listed in this column indicate uniform sampling over those parameters rather than what is listed in column 1. (For example, Mchirpsubscript𝑀𝑐𝑖𝑟𝑝M_{chirp}italic_M start_POSTSUBSCRIPT italic_c italic_h italic_i italic_r italic_p end_POSTSUBSCRIPT is sampled using uniform priors for M1subscript𝑀1M_{1}italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and M2subscript𝑀2M_{2}italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.)
Parameter Description Prior Dist. Prior Range
Mfsubscript𝑀𝑓M_{f}italic_M start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT Source frame final mass uniform [10,200]M10200subscript𝑀direct-product[10,200]M_{\odot}[ 10 , 200 ] italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT
χfsubscript𝜒𝑓\chi_{f}italic_χ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT Final spin uniform [0.99,0.99]0.990.99[-0.99,0.99][ - 0.99 , 0.99 ]
A220subscript𝐴220A_{220}italic_A start_POSTSUBSCRIPT 220 end_POSTSUBSCRIPT Initial (2,2,0)220(2,2,0)( 2 , 2 , 0 ) mode amplitude log10subscript10\log_{10}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT [1025,8×1017]superscript10258superscript1017[10^{-25},8\times 10^{-17}][ 10 start_POSTSUPERSCRIPT - 25 end_POSTSUPERSCRIPT , 8 × 10 start_POSTSUPERSCRIPT - 17 end_POSTSUPERSCRIPT ]
ϕ220subscriptitalic-ϕ220\phi_{220}italic_ϕ start_POSTSUBSCRIPT 220 end_POSTSUBSCRIPT (2,2,0)220(2,2,0)( 2 , 2 , 0 ) mode phase uniform [0,2π]02𝜋[0,2\pi][ 0 , 2 italic_π ]
A221/A220subscript𝐴221subscript𝐴220A_{221}/A_{220}italic_A start_POSTSUBSCRIPT 221 end_POSTSUBSCRIPT / italic_A start_POSTSUBSCRIPT 220 end_POSTSUBSCRIPT Initial (2,2,1)221(2,2,1)( 2 , 2 , 1 ) mode amplitude (as ratio of A220subscript𝐴220A_{220}italic_A start_POSTSUBSCRIPT 220 end_POSTSUBSCRIPT) uniform [0,5]05[0,5][ 0 , 5 ]
ϕ221subscriptitalic-ϕ221\phi_{221}italic_ϕ start_POSTSUBSCRIPT 221 end_POSTSUBSCRIPT (2,2,1)221(2,2,1)( 2 , 2 , 1 ) mode phase uniform [0,2π]02𝜋[0,2\pi][ 0 , 2 italic_π ]
Table 2: Varied parameters in ringdown models and their associated priors. The tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, ι𝜄\iotaitalic_ι, and sky location priors used in ringdown analyses were identical to those shown in Table 1. The amplitude priors indicate the amplitudes of the corresponding quasinormal modes at the start of the ringdown model (i.e. at the merger). The third column indicates the sampling method for each prior. Here, log10subscript10\log_{10}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT indicates a uniform distribution over the base 10 logarithm of the given parameter.

As a preliminary test, the sky location posteriors of the analyses with variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT parameters on real data were plotted. As seen in Fig. 3, the analyses were able to recover most of the posterior for the full IMR analysis conducted in [19].

Refer to caption
Figure 3: Sky position posteriors for GW150914 analyses. The colored contours represent analyses done using the normalization methods described in Section II The blue contour represents the analysis with a QNM post-merger model, and the red contour depicts the analysis with an IMR post-merger model. The black contour represents the posterior from a full IMR analysis conducted in [19]. All contours mark the 90th percentile of each distribution.

V Area theorem

The simplest area theorem test is to compare the progenitor horizon areas A1subscript𝐴1A_{1}italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and A2subscript𝐴2A_{2}italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT to the remnant horizon area Afsubscript𝐴𝑓A_{f}italic_A start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT to check that

A1+A2AiAf.subscript𝐴1subscript𝐴2subscript𝐴𝑖subscript𝐴𝑓A_{1}+A_{2}\equiv A_{i}\leq A_{f}.italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≡ italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≤ italic_A start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT . (6)

In natural units (G=c=1𝐺𝑐1G=c=1italic_G = italic_c = 1) the area of each black hole is given by [1]:

A=8πM2(1+1χ2),𝐴8𝜋superscript𝑀211superscript𝜒2A=8\pi M^{2}(1+\sqrt{1-\chi^{2}}),italic_A = 8 italic_π italic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 + square-root start_ARG 1 - italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (7)

where M𝑀Mitalic_M the black hole’s mass and χ=J/M2𝜒𝐽superscript𝑀2\chi=J/M^{2}italic_χ = italic_J / italic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is its dimensionless spin.

A more robust test can be performed by comparing the measured change in area to the expected change in area [2]

H=Af,measuredAiAf,expectedAi.𝐻subscript𝐴𝑓measuredsubscript𝐴𝑖subscript𝐴𝑓expectedsubscript𝐴𝑖H=\frac{A_{f,\rm measured}-A_{i}}{A_{f,\rm expected}-A_{i}}.italic_H = divide start_ARG italic_A start_POSTSUBSCRIPT italic_f , roman_measured end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_f , roman_expected end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG . (8)

Here, Af,measuredsubscript𝐴𝑓measuredA_{f,\rm measured}italic_A start_POSTSUBSCRIPT italic_f , roman_measured end_POSTSUBSCRIPT is the area of the final black hole inferred from the post-merger analysis and Aisubscript𝐴𝑖A_{i}italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the sum of the initial areas inferred from the pre-merger analysis. The expected area of the final black hole Af,expectedsubscript𝐴𝑓expectedA_{f,\rm expected}italic_A start_POSTSUBSCRIPT italic_f , roman_expected end_POSTSUBSCRIPT is derived from the initial masses and spins measured from the pre-merger analysis. These are converted into an estimated final mass and spin using fits from numerical relativity, which is then converted into an area via Eq. (7).

Since Af,expectedsubscript𝐴𝑓𝑒𝑥𝑝𝑒𝑐𝑡𝑒𝑑A_{f,expected}italic_A start_POSTSUBSCRIPT italic_f , italic_e italic_x italic_p italic_e italic_c italic_t italic_e italic_d end_POSTSUBSCRIPT is evaluated with numerical relativity using A1subscript𝐴1A_{1}italic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and A2subscript𝐴2A_{2}italic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, the denominator of Eq. (8) is positive definite. Therefore, if H<0𝐻0H<0italic_H < 0, then Af,measured<Aisubscript𝐴𝑓𝑚𝑒𝑎𝑠𝑢𝑟𝑒𝑑subscript𝐴𝑖A_{f,measured}<A_{i}italic_A start_POSTSUBSCRIPT italic_f , italic_m italic_e italic_a italic_s italic_u italic_r italic_e italic_d end_POSTSUBSCRIPT < italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and the area theorem is violated. The confidence interval in favor of the area theorem is the fraction of the posterior for which H>0𝐻0H>0italic_H > 0. More robustly, H=1𝐻1H=1italic_H = 1 indicates that signal is consistent with GR specifically, not just the more generic class of theories that satisfy the area increase law.

Refer to caption
Figure 4: Final mass and final spin posteriors of GW150914 analyses using a QNM post-merger approximant and real waveform data. All masses are in the detector frame. Contours represent the 90th percentile of each distribution. The darker colored lines correspond to the model with variable sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, and the lighter lines represent the model where these parameters were fixed. Blue lines correspond to post-merger results, while red lines correspond to pre-merger results, obtained using numerical relativity with the IMR parameters. The innermost black contour represents the 4-OGC posterior from [19].

Figure 4 shows the posteriors of the final mass Mfsubscript𝑀𝑓M_{f}italic_M start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and final spin χfsubscript𝜒𝑓\chi_{f}italic_χ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT of the QNM post-merger models on real data. Specifically, the Mfsubscript𝑀𝑓M_{f}italic_M start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and χfsubscript𝜒𝑓\chi_{f}italic_χ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT posteriors from the analysis with variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT parameters are compared with the corresponding posteriors from fixed gate analysis. All posteriors contain within them the distribution from a full IMR analysis from 4-OGC [19]. Furthermore, both posteriors from the variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT run overlap almost completely with their corresponding fixed parameter posteriors.

Refer to caption
Figure 5: Final mass and final spin posteriors of GW150914 analyses using an IMR post-merger approximant and real waveform data. All masses are in the detector frame. Contours represent the 90th percentile of each distribution. The darker colored lines correspond to the model with variable sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, and the lighter lines represent the model where these parameters were fixed. Blue lines correspond to post-merger results, while red lines correspond to pre-merger results, obtained using numerical relativity with the IMR parameters. The innermost black contour represents the 4-OGC posterior from [19].

A similar plot is shown in Fig. 5, except with the analyses that used an IMR post-merger model on real data. Again, the pre-merger and post-merger posteriors from the analysis with variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT parameters are compared to their fixed counterparts as well as the full IMR posterior from [19]. Notably, while still in very close agreement, the variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT posteriors have a greater discrepancy from their fixed counterparts than seen in Fig. 4. Specifically, the post-merger posterior tends towards slightly higher masses and spins, while the pre-merger posterior includes lower spin values.

Directly comparing the pre-merger posteriors in Figs.4 and 5 shows a slight discrepancy in final mass and spin estimates. Namely, the analysis with a QNM post-merger model tends towards higher final mass and spin estimates than the IMR post-merger model. This indicates minor coupling between the pre- and post-merger models caused by their shared sky and time parameters. However, as made evident by these two plots, this is not a major effect, and both pre-merger models maintain strong consistency with the full IMR posterior.

Figure 6 compares the H𝐻Hitalic_H posteriors between the four analyses of the real GW150914 data. All posteriors agree with the expected value H=1𝐻1H=1italic_H = 1, with each distribution peaking near this value. While the QNM post-merger posteriors are almost in exact agreement with each other, there is a slight discrepancy between the IMR post-merger posteriors. Specifically, the fixed sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT posterior has a much sharper peak at a lower value than the variable gate counterpart. This may be explained by Fig. 5, where the fixed IMR post-merger posterior was observed to contain lower masses and spins than the variable gate model. This could lead to the slight bias to lower H𝐻Hitalic_H values seen here.

Refer to caption
Figure 6: Posteriors of H=(Af,measuredAi)/(Af,expectedAi)𝐻subscript𝐴𝑓𝑚𝑒𝑎𝑠𝑢𝑟𝑒𝑑subscript𝐴𝑖subscript𝐴𝑓𝑒𝑥𝑝𝑒𝑐𝑡𝑒𝑑subscript𝐴𝑖H=(A_{f,measured}-A_{i})/(A_{f,expected}-A_{i})italic_H = ( italic_A start_POSTSUBSCRIPT italic_f , italic_m italic_e italic_a italic_s italic_u italic_r italic_e italic_d end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) / ( italic_A start_POSTSUBSCRIPT italic_f , italic_e italic_x italic_p italic_e italic_c italic_t italic_e italic_d end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) for the GW150914 analyses with real waveform data. Blue histograms correspond to QNM post-merger models, while red lines correspond to IMR post-merger models. Darker histograms correspond to models with variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT parameters, while lighter histograms correspond to models where these parameters are fixed to the maximum likelihood values from [19]. The expected value H=1𝐻1H=1italic_H = 1 is shown with a vertical black dotted line. Points in the shaded region correspond to H<0𝐻0H<0italic_H < 0 and therefore disagree with the area theorem.

Table 3 summarizes the H𝐻Hitalic_H credible intervals and area theorem confidence intervals for all eight analyses. All credible intervals are in agreement with the expected value H=1𝐻1H=1italic_H = 1. Additionally, all analyses of the real GW150915 signal have H>0𝐻0H>0italic_H > 0 at greater than 95% confidence, indicating very high agreement with the area theorem. Results from the analysis of the simulated signal are largely consistent with these results (see Appendix C for equivalent figures).

The area theorem confidence interval of any given analysis appears to be uncorrelated to whether or not the sky location and coalescence time were allowed to vary. However, IMR post-merger models tended to have slightly higher agreement with the area theorem than corresponding QNM post-merger models. This pattern may be explained by trends in the pre- and post-merger mass and spin results. The IMR post-merger models in Fig. 5 had post-merger measurements that were either in agreement with or greater than the corresponding pre-merger estimates. This corresponds to a very high concentration of H𝐻Hitalic_H measurements approximating or exceeding 1, with very few points skewed to negative H𝐻Hitalic_H values. Conversely, the post-merger posteriors shown in Fig. 4 have significant amounts of points with higher and lower mass and spin values than their pre-merger counterparts. This leads to a wider distribution of H𝐻Hitalic_H values to both higher and lower values, allowing for a higher concentration of negative measurements.

GW150914 analysis Credible interval P(H>0)𝑃𝐻0P(H>0)italic_P ( italic_H > 0 )
Variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, QNM post-merger, real waveform 1.01.0+1.4subscriptsuperscript1.01.41.01.0^{+1.4}_{-1.0}1.0 start_POSTSUPERSCRIPT + 1.4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1.0 end_POSTSUBSCRIPT 95.4%
Variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, IMR post-merger, real waveform 1.10.6+0.6subscriptsuperscript1.10.60.61.1^{+0.6}_{-0.6}1.1 start_POSTSUPERSCRIPT + 0.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.6 end_POSTSUBSCRIPT 99.5%
Fixed sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, QNM post-merger, real waveform 1.21.0+1.2subscriptsuperscript1.21.21.01.2^{+1.2}_{-1.0}1.2 start_POSTSUPERSCRIPT + 1.2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1.0 end_POSTSUBSCRIPT 97.4%
Fixed sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, IMR post-merger, real waveform 1.00.6+0.6subscriptsuperscript1.00.60.61.0^{+0.6}_{-0.6}1.0 start_POSTSUPERSCRIPT + 0.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.6 end_POSTSUBSCRIPT 98.4%
Variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, QNM post-merger, zero-noise injection 0.91.3+1.5subscriptsuperscript0.91.51.30.9^{+1.5}_{-1.3}0.9 start_POSTSUPERSCRIPT + 1.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1.3 end_POSTSUBSCRIPT 89.1%
Variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, IMR post-merger, zero-noise injection 0.90.6+0.6subscriptsuperscript0.90.60.60.9^{+0.6}_{-0.6}0.9 start_POSTSUPERSCRIPT + 0.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.6 end_POSTSUBSCRIPT 98.4%
Fixed sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, QNM post-merger, zero-noise injection 1.10.9+0.9subscriptsuperscript1.10.90.91.1^{+0.9}_{-0.9}1.1 start_POSTSUPERSCRIPT + 0.9 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.9 end_POSTSUBSCRIPT 97.3%
Fixed sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, IMR post-merger, zero-noise injection 1.00.5+0.5subscriptsuperscript1.00.50.51.0^{+0.5}_{-0.5}1.0 start_POSTSUPERSCRIPT + 0.5 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.5 end_POSTSUBSCRIPT 98.8%
Table 3: Summary of posteriors for H=(Af,measuredAi)/(Af,expectedAi)𝐻subscript𝐴𝑓𝑚𝑒𝑎𝑠𝑢𝑟𝑒𝑑subscript𝐴𝑖subscript𝐴𝑓𝑒𝑥𝑝𝑒𝑐𝑡𝑒𝑑subscript𝐴𝑖H=(A_{f,measured}-A_{i})/(A_{f,expected}-A_{i})italic_H = ( italic_A start_POSTSUBSCRIPT italic_f , italic_m italic_e italic_a italic_s italic_u italic_r italic_e italic_d end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) / ( italic_A start_POSTSUBSCRIPT italic_f , italic_e italic_x italic_p italic_e italic_c italic_t italic_e italic_d end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) from various GW150914 analyses. The first column lists the properties of each model, namely whether or not sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT were allowed to vary, the post-merger model used, and the waveform data used. The approximant IMRPhenomXPHM [22] was used to model the post-merger signals of analyses labeled “IMR post-merger,” as were all pre-merger signals. Analyses with injected waveforms used a zero-noise injection of the GW150914 waveform as input data. The second column lists the median and the 90% credible interval for each H𝐻Hitalic_H posterior. The third column lists the percentage of posterior points for which H>0𝐻0H>0italic_H > 0, thereby agreeing with the area theorem.

VI Conclusions

This paper presented a novel method of marginalizing over sky location and coalescence time when performing a pre- or post-merger analysis, which allows for a full accounting of uncertainties in parameter estimates. Tests of the area theorem were conducted using data from GW150914 using this method. It was found that the data for this event agrees very well with the area theorem regardless of whether sky position and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT were allowed to vary. The only noticeable changes in agreement in the area theorem were caused by the post-merger approximant used, though in general this was a very minor change.

While we focus on tests of the area theorem here, our method for marginalizing over sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT applies to any analysis in which only a portion of the signal is modelled. In particular, black hole spectroscopy involves analyzing the post-merger signal using a QNM template in order to perform a test of the no-hair theorem [24, 25, 26]. As with previous tests of the area theorem, previous black hole spectroscopy studies have fixed the sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for the reasons discussed in Sec. II [23, 27, 28]. Our finding that a signal model is needed for the entire observable signal applies equally well to black hole spectroscopy, even though the pre-merger signal is purposely excluded in such analyses. The hierarchical method we develop in Sec. III is therefore equally relevant for marginalizing over sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT when doing black hole spectroscopy. In Ref. [29] we use our QNM analysis of GW150914 to investigate the evidence for the presence of the (2,2,1)221(2,2,1)( 2 , 2 , 1 ) mode, which has been hotly contested in the literature [23, 27, 30, 5].

In all our analyses here we end the pre-merger analysis when the post-merger analysis beings. This is necessitated by our finding that some model must exist for the entire observable signal when the gate time is allowed to vary. Since there was no gap between our pre- and post-merger template, we obtained agreement with the area theorem in excess of 95%percent9595\%95 %. A more rigorous test of the area theorem is to excise the merger from the data, since any biases introduced by including the merger as part of the pre-merger model would be omitted. Refs. [4, 2] did this additional test with fixed sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. However, this is not possible when marginalizing over sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for the reasons highlighted in Sec. III. Introducing a gap between the pre- and post-merger will once again favor points that excise as much of the signal as possible, even when both the pre- and post-merger were modeled simultaneously.

Introducing a gap between the pre- and post-merger models could be achieved by using three sub-domains rather than two: one each for the inspiral, merger, and ringdown. The gate for one domain would start/end at the end/start of the next. For the merger domain, an arbitrary signal model using wavelets could be used, similar to what Finch and Moore used in Ref. [10]. This would ensure that pre- and post-merger parameters are fully unbiased by the merger while ensuring that the merger itself is not arbitrarily gated out. The initial black hole areas could then be estimated using an inspiral model and the final area using a QNM model. This would also be useful in black hole spectroscopy studies involving fundamental angular QNMs. These modes are not expected to become relevant until 10Msimilar-toabsent10𝑀\sim 10M∼ 10 italic_M after merger, necessitating a gap between the merger and the start of the QNM model. We plan to investigate this in a future study.

This research was conducted using PyCBC [31]. Our data is available at [32].

VII Acknowledgements

A.C. was supported by funds from the Massachusetts Space Grant Consortium. C.C. acknowledges support from NSF award PHY-2309356. All computations were performed on Unity, a collaborative, multi-institutional high-performance computing cluster managed by UMass Amherst Research Computing and Data.

This research has made use of data or software obtained from the Gravitational Wave Open Science Center (gwosc.org), a service of the LIGO Scientific Collaboration, the Virgo Collaboration, and KAGRA. This material is based upon work supported by NSF’s LIGO Laboratory which is a major facility fully funded by the National Science Foundation, as well as the Science and Technology Facilities Council (STFC) of the United Kingdom, the Max-Planck-Society (MPS), and the State of Niedersachsen/Germany for support of the construction of Advanced LIGO and construction and operation of the GEO600 detector. Additional support for Advanced LIGO was provided by the Australian Research Council. Virgo is funded, through the European Gravitational Observatory (EGO), by the French Centre National de Recherche Scientifique (CNRS), the Italian Istituto Nazionale di Fisica Nucleare (INFN) and the Dutch Nikhef, with contributions by institutions from Belgium, Germany, Greece, Hungary, Ireland, Japan, Monaco, Poland, Portugal, Spain. KAGRA is supported by Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan Society for the Promotion of Science (JSPS) in Japan; National Research Foundation (NRF) and Ministry of Science and ICT (MSIT) in Korea; Academia Sinica (AS) and National Science and Technology Council (NSTC) in Taiwan.

Appendix A Likelihood Function Derivation

To calculate the posterior from Bayes’ theorem (Eq. (1)), one requires a model for both the signal hhitalic_h and the noise n𝑛nitalic_n. While hhitalic_h can be determined from the linearized EFE wave solution, n𝑛nitalic_n requires more statistical considerations.

To start, consider a network of K𝐾Kitalic_K gravitational wave detectors that each sample at a rate ΔtΔ𝑡\Delta troman_Δ italic_t over a total time length T𝑇Titalic_T. Defining the total number of samples N=T/Δt𝑁𝑇Δ𝑡N=T/\Delta titalic_N = italic_T / roman_Δ italic_t, the full data series 𝒔netsubscript𝒔𝑛𝑒𝑡\bm{s}_{net}bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT can be expressed as a series of K𝐾Kitalic_K N𝑁Nitalic_N-dimensional vectors 𝒔Ksuperscript𝒔𝐾\bm{s}^{K}bold_italic_s start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT such that 𝒔K={s0K,s1K,,sNK}superscript𝒔𝐾subscriptsuperscript𝑠𝐾0subscriptsuperscript𝑠𝐾1subscriptsuperscript𝑠𝐾𝑁\bm{s}^{K}=\{s^{K}_{0},s^{K}_{1},...,s^{K}_{N}\}bold_italic_s start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT = { italic_s start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_s start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_s start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT } and 𝒔net={𝒔1,𝒔2,,𝒔K}subscript𝒔𝑛𝑒𝑡superscript𝒔1superscript𝒔2superscript𝒔𝐾\bm{s}_{net}=\{\bm{s}^{1},\bm{s}^{2},...,\bm{s}^{K}\}bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT = { bold_italic_s start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT , bold_italic_s start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , … , bold_italic_s start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT }. To simplify calculations, assume that the signal is zero such that 𝒔net=𝒏subscript𝒔𝑛𝑒𝑡𝒏\bm{s}_{net}=\bm{n}bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT = bold_italic_n. Additionally, assume that noise model is a stochastic Gaussian distribution and is uncorrelated between detectors. Under these assumptions, the noise likelihood function is:

p(𝒔net|n)=exp[12d=1K𝒔dT𝚺d1𝒔d][(2π)NKd=1Kdet𝚺d]1/2,𝑝conditionalsubscript𝒔𝑛𝑒𝑡𝑛12superscriptsubscript𝑑1𝐾subscriptsuperscript𝒔𝑇𝑑subscriptsuperscript𝚺1𝑑subscript𝒔𝑑superscriptdelimited-[]superscript2𝜋𝑁𝐾superscriptsubscriptproduct𝑑1𝐾subscript𝚺𝑑12p(\bm{s}_{net}|n)=\frac{\exp{[-\frac{1}{2}\sum_{d=1}^{K}\bm{s}^{T}_{d}\bm{% \Sigma}^{-1}_{d}\bm{s}_{d}]}}{[(2\pi)^{NK}\prod_{d=1}^{K}\det\bm{\Sigma}_{d}]^% {1/2}},italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | italic_n ) = divide start_ARG roman_exp [ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT bold_italic_s start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ] end_ARG start_ARG [ ( 2 italic_π ) start_POSTSUPERSCRIPT italic_N italic_K end_POSTSUPERSCRIPT ∏ start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT roman_det bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG , (9)

where 𝚺dsubscript𝚺𝑑\bm{\Sigma}_{d}bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT is the covariance matrix of the noise model for detector d𝑑ditalic_d, defined using the ensemble average:

𝚺d[j,k]=𝒔d[j]𝒔d[k].subscript𝚺𝑑𝑗𝑘delimited-⟨⟩subscript𝒔𝑑delimited-[]𝑗subscript𝒔𝑑delimited-[]𝑘\bm{\Sigma}_{d}[j,k]=\langle\bm{s}_{d}[j]\bm{s}_{d}[k]\rangle.bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT [ italic_j , italic_k ] = ⟨ bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT [ italic_j ] bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT [ italic_k ] ⟩ . (10)

(By dropping the d𝑑ditalic_d subscripts, one obtains an equivalent expression for the full covariance matrix and data. We do so in the following steps for brevity.) This is the exact expression of the likelihood function for noise. However, this function is infeasible to calculate analytically due to the inverse covariance matrix in the numerator.

To do this, one may expand the covariance matrix definition in Eq. (10) as follows, defining Δkj=kjsubscriptΔ𝑘𝑗𝑘𝑗\Delta_{kj}=k-jroman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT = italic_k - italic_j:

𝚺[j,k]𝚺𝑗𝑘\displaystyle\bm{\Sigma}[j,k]bold_Σ [ italic_j , italic_k ] =𝒔[j]𝒔[k]absentdelimited-⟨⟩𝒔delimited-[]𝑗𝒔delimited-[]𝑘\displaystyle=\langle\bm{s}[j]\bm{s}[k]\rangle= ⟨ bold_italic_s [ italic_j ] bold_italic_s [ italic_k ] ⟩
=𝒔[j]𝒔[Δkj+j]absentdelimited-⟨⟩𝒔delimited-[]𝑗𝒔delimited-[]subscriptΔ𝑘𝑗𝑗\displaystyle=\langle\bm{s}[j]\bm{s}[\Delta_{kj}+j]\rangle= ⟨ bold_italic_s [ italic_j ] bold_italic_s [ roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT + italic_j ] ⟩
=limn1nl=0n1sl[j]sl[Δkj+j],absentsubscript𝑛1𝑛subscriptsuperscript𝑛1𝑙0superscript𝑠𝑙delimited-[]𝑗superscript𝑠𝑙delimited-[]subscriptΔ𝑘𝑗𝑗\displaystyle=\lim_{n\to\infty}\frac{1}{n}\sum^{n-1}_{l=0}s^{l}[j]s^{l}[\Delta% _{kj}+j],= roman_lim start_POSTSUBSCRIPT italic_n → ∞ end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_n end_ARG ∑ start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT [ italic_j ] italic_s start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT [ roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT + italic_j ] , (11)

where in the last step the ensemble average is written out fully.

In general, this expression is dependent on time tj=jΔtsubscript𝑡𝑗𝑗Δ𝑡t_{j}=j\Delta titalic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT = italic_j roman_Δ italic_t and displacement τkj=ΔkjΔtsubscript𝜏𝑘𝑗subscriptΔ𝑘𝑗Δ𝑡\tau_{kj}=\Delta_{kj}\Delta titalic_τ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT = roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT roman_Δ italic_t. However, one can make the assumption that the noise is wide sense stationary, where the mean and variance are both constant in time. Under this assumption, any constant can be added to the indices in Eq. (10) to obtain the same result. This makes 𝚺𝚺\bm{\Sigma}bold_Σ symmetric (since the factors in Eq. (11) commute) and Toeplitz (since the elements along the diagonals are equal) [33]. Additionally, since 𝚺[0,Δkj]=𝚺[Δkj,0]=𝚺[0,Δkj]𝚺0subscriptΔ𝑘𝑗𝚺subscriptΔ𝑘𝑗0𝚺0subscriptΔ𝑘𝑗\bm{\Sigma}[0,\Delta_{kj}]=\bm{\Sigma}[-\Delta_{kj},0]=\bm{\Sigma}[0,-\Delta_{% kj}]bold_Σ [ 0 , roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT ] = bold_Σ [ - roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT , 0 ] = bold_Σ [ 0 , - roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT ], the elements of 𝚺𝚺\bm{\Sigma}bold_Σ are even functions of ΔkjsubscriptΔ𝑘𝑗\Delta_{kj}roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT.

Additionally, one can assume that the data is ergodic, meaning that new realizations of 𝒔𝒔\bm{s}bold_italic_s are obtained via time. Under this assumption and the properties of the elements of 𝚺𝚺\bm{\Sigma}bold_Σ, the ensemble averages in Eq. (11) can be replaced with time averages:

𝚺[j,k]𝚺𝑗𝑘\displaystyle\bm{\Sigma}[j,k]bold_Σ [ italic_j , italic_k ] =limn1nl=0n1sl[0]sl[Δkj]absentsubscript𝑛1𝑛subscriptsuperscript𝑛1𝑙0superscript𝑠𝑙delimited-[]0superscript𝑠𝑙delimited-[]subscriptΔ𝑘𝑗\displaystyle=\lim_{n\to\infty}\frac{1}{n}\sum^{n-1}_{l=0}s^{l}[0]s^{l}[\Delta% _{kj}]= roman_lim start_POSTSUBSCRIPT italic_n → ∞ end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_n end_ARG ∑ start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT [ 0 ] italic_s start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT [ roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT ]
=limn1nl=0n1sl[l]sl[Δkj+l]absentsubscript𝑛1𝑛subscriptsuperscript𝑛1𝑙0superscript𝑠𝑙delimited-[]𝑙superscript𝑠𝑙delimited-[]subscriptΔ𝑘𝑗𝑙\displaystyle=\lim_{n\to\infty}\frac{1}{n}\sum^{n-1}_{l=0}s^{l}[l]s^{l}[\Delta% _{kj}+l]= roman_lim start_POSTSUBSCRIPT italic_n → ∞ end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_n end_ARG ∑ start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT [ italic_l ] italic_s start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT [ roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT + italic_l ]
=limn12nl=nn1sl[l]sl[Δkj+l]absentsubscript𝑛12𝑛subscriptsuperscript𝑛1𝑙𝑛superscript𝑠𝑙delimited-[]𝑙superscript𝑠𝑙delimited-[]subscriptΔ𝑘𝑗𝑙\displaystyle=\lim_{n\to\infty}\frac{1}{2n}\sum^{n-1}_{l=-n}s^{l}[l]s^{l}[% \Delta_{kj}+l]= roman_lim start_POSTSUBSCRIPT italic_n → ∞ end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG 2 italic_n end_ARG ∑ start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l = - italic_n end_POSTSUBSCRIPT italic_s start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT [ italic_l ] italic_s start_POSTSUPERSCRIPT italic_l end_POSTSUPERSCRIPT [ roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT + italic_l ]
=12Rss((kj)Δt).absent12subscript𝑅𝑠𝑠𝑘𝑗Δ𝑡\displaystyle=\frac{1}{2}R_{ss}((k-j)\Delta t).= divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_R start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( ( italic_k - italic_j ) roman_Δ italic_t ) . (12)

The last step defines the autocorrelation function Rss(τ)subscript𝑅𝑠𝑠𝜏R_{ss}(\tau)italic_R start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_τ ), which describes the correlation between points in the time series 𝒔𝒔\bm{s}bold_italic_s. If Rss(τ)subscript𝑅𝑠𝑠𝜏R_{ss}(\tau)italic_R start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_τ ) goes to zero in some finite time τmaxsubscript𝜏𝑚𝑎𝑥\tau_{max}italic_τ start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT, then all diagonals with |Δkj|>floor(τmax/Δt)=ΔmaxsubscriptΔ𝑘𝑗𝑓𝑙𝑜𝑜𝑟subscript𝜏𝑚𝑎𝑥Δ𝑡subscriptΔ𝑚𝑎𝑥|\Delta_{kj}|>floor(\tau_{max}/\Delta t)=\Delta_{max}| roman_Δ start_POSTSUBSCRIPT italic_k italic_j end_POSTSUBSCRIPT | > italic_f italic_l italic_o italic_o italic_r ( italic_τ start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT / roman_Δ italic_t ) = roman_Δ start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT will equal zero. This is similar to the form of a circulant matrix 𝑪𝑪\bm{C}bold_italic_C, a special case of a Toeplitz matrix where each row is a right-cycle permutation of the same vector (in this case, 𝒔𝒔\bm{s}bold_italic_s). The eigenvectors of circulant matrices are well-known [33]:

𝒖p[k]=1Nexp(2πikp/N)subscript𝒖𝑝delimited-[]𝑘1𝑁2𝜋𝑖𝑘𝑝𝑁\bm{u}_{p}[k]=\frac{1}{\sqrt{N}}\exp{(-2\pi ikp/N)}bold_italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT [ italic_k ] = divide start_ARG 1 end_ARG start_ARG square-root start_ARG italic_N end_ARG end_ARG roman_exp ( - 2 italic_π italic_i italic_k italic_p / italic_N ) (13)

This generally is not true for Toeplitz matrices, but one may take advantage of Eq. (13) by recognizing that the matrix described by Eq. (12) asymptotes to a circulant matrix for large N:

limN|𝚺𝑪|=𝟎subscript𝑁𝚺𝑪0\lim_{N\to\infty}|\bm{\Sigma}-\bm{C}|=\bm{0}roman_lim start_POSTSUBSCRIPT italic_N → ∞ end_POSTSUBSCRIPT | bold_Σ - bold_italic_C | = bold_0 (14)

Therefore, the eigenvalues λpsubscript𝜆𝑝\lambda_{p}italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT of 𝚺𝚺\bm{\Sigma}bold_Σ can be evaluated using the usual eigenvalue equation as long as Δmax<<N/2much-less-thansubscriptΔ𝑚𝑎𝑥𝑁2\Delta_{max}<<N/2roman_Δ start_POSTSUBSCRIPT italic_m italic_a italic_x end_POSTSUBSCRIPT < < italic_N / 2:

𝚺𝒖pλp𝒖p𝚺subscript𝒖𝑝subscript𝜆𝑝subscript𝒖𝑝\bm{\Sigma}\bm{u}_{p}\approx\lambda_{p}\bm{u}_{p}bold_Σ bold_italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≈ italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT bold_italic_u start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT (15)

Therefore, using the fact that 𝚺𝚺\bm{\Sigma}bold_Σ is symmetric and Rss(l)subscript𝑅𝑠𝑠𝑙R_{ss}(l)italic_R start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_l ) is even:

λpsubscript𝜆𝑝\displaystyle\lambda_{p}italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT =12Re{l=N/2N/2Rss(l)exp(2πipl/N)}absent12𝑅𝑒subscriptsuperscript𝑁2𝑙𝑁2subscript𝑅𝑠𝑠𝑙2𝜋𝑖𝑝𝑙𝑁\displaystyle=\frac{1}{2}Re\bigg{\{}\sum^{N/2}_{l=-N/2}R_{ss}(l)\exp(-2\pi ipl% /N)\bigg{\}}= divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_R italic_e { ∑ start_POSTSUPERSCRIPT italic_N / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l = - italic_N / 2 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_l ) roman_exp ( - 2 italic_π italic_i italic_p italic_l / italic_N ) }
=12Re{l=0N1Rss(l)exp(2πipl/N)}absent12𝑅𝑒subscriptsuperscript𝑁1𝑙0subscript𝑅𝑠𝑠𝑙2𝜋𝑖𝑝𝑙𝑁\displaystyle=\frac{1}{2}Re\bigg{\{}\sum^{N-1}_{l=0}R_{ss}(l)\exp(-2\pi ipl/N)% \bigg{\}}= divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_R italic_e { ∑ start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_l = 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_l ) roman_exp ( - 2 italic_π italic_i italic_p italic_l / italic_N ) }
=12Re{R~ss(p)/Δt},absent12𝑅𝑒subscript~𝑅𝑠𝑠𝑝Δ𝑡\displaystyle=\frac{1}{2}Re\{\tilde{R}_{ss}(p)/\Delta t\},= divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_R italic_e { over~ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_p ) / roman_Δ italic_t } , (16)

where R~ss(p)subscript~𝑅𝑠𝑠𝑝\tilde{R}_{ss}(p)over~ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_p ) is the discrete Fourier transform of the autocorrelation function:

R~ss(p)=Δtk=0N1Rss(k)exp(2πipk/N)subscript~𝑅𝑠𝑠𝑝Δ𝑡subscriptsuperscript𝑁1𝑘0subscript𝑅𝑠𝑠𝑘2𝜋𝑖𝑝𝑘𝑁\tilde{R}_{ss}(p)=\Delta t\sum^{N-1}_{k=0}{R}_{ss}(k)\exp(-2\pi ipk/N)over~ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_p ) = roman_Δ italic_t ∑ start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT ( italic_k ) roman_exp ( - 2 italic_π italic_i italic_p italic_k / italic_N ) (17)

To simplify Eq. (16), one may impose the Wiener-Khinchin Theorem [34], which defines the power spectral density Snsubscript𝑆𝑛S_{n}italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT as the Fourier transform of Rsssubscript𝑅𝑠𝑠R_{ss}italic_R start_POSTSUBSCRIPT italic_s italic_s end_POSTSUBSCRIPT for a wide-sense stationary stochastic process:

λp=Sn[p]2Δtsubscript𝜆𝑝subscript𝑆𝑛delimited-[]𝑝2Δ𝑡\lambda_{p}=\frac{S_{n}[p]}{2\Delta t}italic_λ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = divide start_ARG italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT [ italic_p ] end_ARG start_ARG 2 roman_Δ italic_t end_ARG (18)

One can then construct the inverse of 𝚺𝚺\bm{\Sigma}bold_Σ using the eigenvalues from Eq. (18) and the eigenvectors of Eq. (13):

𝚺1[j,k]superscript𝚺1𝑗𝑘\displaystyle\bm{\Sigma}^{-1}[j,k]bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ italic_j , italic_k ] =2ΔtNp=0N1exp[2πip(jk)/N]Sn(p)absent2Δ𝑡𝑁subscriptsuperscript𝑁1𝑝02𝜋𝑖𝑝𝑗𝑘𝑁subscript𝑆𝑛𝑝\displaystyle=\frac{2\Delta t}{N}\sum^{N-1}_{p=0}\frac{\exp[-2\pi ip(j-k)/N]}{% S_{n}(p)}= divide start_ARG 2 roman_Δ italic_t end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p = 0 end_POSTSUBSCRIPT divide start_ARG roman_exp [ - 2 italic_π italic_i italic_p ( italic_j - italic_k ) / italic_N ] end_ARG start_ARG italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_p ) end_ARG
=2Δf(Δt)2p=0N/21e2πip(jk)/N+e2πip(jk)/NSn(p),absent2Δ𝑓superscriptΔ𝑡2subscriptsuperscript𝑁21𝑝0superscript𝑒2𝜋𝑖𝑝𝑗𝑘𝑁superscript𝑒2𝜋𝑖𝑝𝑗𝑘𝑁subscript𝑆𝑛𝑝\displaystyle={2\Delta f(\Delta t)^{2}}\sum^{N/2-1}_{p=0}\frac{e^{-2\pi ip(j-k% )/N}+e^{2\pi ip(j-k)/N}}{S_{n}(p)},= 2 roman_Δ italic_f ( roman_Δ italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∑ start_POSTSUPERSCRIPT italic_N / 2 - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p = 0 end_POSTSUBSCRIPT divide start_ARG italic_e start_POSTSUPERSCRIPT - 2 italic_π italic_i italic_p ( italic_j - italic_k ) / italic_N end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT 2 italic_π italic_i italic_p ( italic_j - italic_k ) / italic_N end_POSTSUPERSCRIPT end_ARG start_ARG italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_p ) end_ARG , (19)

using the fact that Snsubscript𝑆𝑛S_{n}italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is symmetric about N/2𝑁2N/2italic_N / 2 and defining the sample frequency Δf=1/T=1/NΔtΔ𝑓1𝑇1𝑁Δ𝑡\Delta f=1/T=1/N\Delta troman_Δ italic_f = 1 / italic_T = 1 / italic_N roman_Δ italic_t. Since 𝒔𝒔\bm{s}bold_italic_s is real and

(Δt)2j,k=0N1𝒔[j]𝒔[k](e2πip(jk)/N+e2πip(jk)/N)=2|𝒔~|2[p],superscriptΔ𝑡2subscriptsuperscript𝑁1𝑗𝑘0𝒔delimited-[]𝑗𝒔delimited-[]𝑘superscript𝑒2𝜋𝑖𝑝𝑗𝑘𝑁superscript𝑒2𝜋𝑖𝑝𝑗𝑘𝑁2superscript~𝒔2delimited-[]𝑝(\Delta t)^{2}\sum^{N-1}_{j,k=0}\bm{s}[j]\bm{s}[k](e^{-2\pi ip(j-k)/N}+e^{2\pi ip% (j-k)/N})=2|\tilde{\bm{s}}|^{2}[p],( roman_Δ italic_t ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∑ start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , italic_k = 0 end_POSTSUBSCRIPT bold_italic_s [ italic_j ] bold_italic_s [ italic_k ] ( italic_e start_POSTSUPERSCRIPT - 2 italic_π italic_i italic_p ( italic_j - italic_k ) / italic_N end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT 2 italic_π italic_i italic_p ( italic_j - italic_k ) / italic_N end_POSTSUPERSCRIPT ) = 2 | over~ start_ARG bold_italic_s end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [ italic_p ] , (20)

one can write:

𝒔T𝚺1𝒔=4Δfp=0N/212|𝒔~|2[p]Sn(p).superscript𝒔𝑇superscript𝚺1𝒔4Δ𝑓subscriptsuperscript𝑁21𝑝02superscript~𝒔2delimited-[]𝑝subscript𝑆𝑛𝑝\bm{s}^{T}\bm{\Sigma}^{-1}\bm{s}=4\Delta f\sum^{N/2-1}_{p=0}\frac{2|\tilde{\bm% {s}}|^{2}[p]}{S_{n}(p)}.bold_italic_s start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_italic_s = 4 roman_Δ italic_f ∑ start_POSTSUPERSCRIPT italic_N / 2 - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p = 0 end_POSTSUBSCRIPT divide start_ARG 2 | over~ start_ARG bold_italic_s end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [ italic_p ] end_ARG start_ARG italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_p ) end_ARG . (21)

Finally, if the inner product between two arbitrary vectors 𝒂𝒂\bm{a}bold_italic_a and 𝒃𝒃\bm{b}bold_italic_b is defined as:

𝒂,𝒃=4Re{Δfp=0N/21𝒂~[p]𝒃~[p]Sn(p)},𝒂𝒃4𝑅𝑒Δ𝑓subscriptsuperscript𝑁21𝑝0superscript~𝒂delimited-[]𝑝~𝒃delimited-[]𝑝subscript𝑆𝑛𝑝\langle\bm{a},\bm{b}\rangle=4Re\bigg{\{}\Delta f\sum^{N/2-1}_{p=0}\frac{\tilde% {\bm{a}}^{\*}[p]\tilde{\bm{b}}[p]}{S_{n}(p)}\bigg{\}},⟨ bold_italic_a , bold_italic_b ⟩ = 4 italic_R italic_e { roman_Δ italic_f ∑ start_POSTSUPERSCRIPT italic_N / 2 - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_p = 0 end_POSTSUBSCRIPT divide start_ARG over~ start_ARG bold_italic_a end_ARG start_POSTSUPERSCRIPT ⁢ end_POSTSUPERSCRIPT [ italic_p ] over~ start_ARG bold_italic_b end_ARG [ italic_p ] end_ARG start_ARG italic_S start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_p ) end_ARG } , (22)

Eq. (21) can be written as an inner product:

𝒔T𝚺1𝒔=𝒔,𝒔,superscript𝒔𝑇superscript𝚺1𝒔𝒔𝒔\bm{s}^{T}\bm{\Sigma}^{-1}\bm{s}=\langle\bm{s},\bm{s}\rangle,bold_italic_s start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_italic_s = ⟨ bold_italic_s , bold_italic_s ⟩ , (23)

and the likelihood function can be expressed as:

p(𝒔net|n)=exp[12d=1K𝒔d,𝒔d][(2π)NKd=1Kdet𝚺d]1/2𝑝conditionalsubscript𝒔𝑛𝑒𝑡𝑛12superscriptsubscript𝑑1𝐾subscript𝒔𝑑subscript𝒔𝑑superscriptdelimited-[]superscript2𝜋𝑁𝐾superscriptsubscriptproduct𝑑1𝐾subscript𝚺𝑑12p(\bm{s}_{net}|n)=\frac{\exp{[-\frac{1}{2}\sum_{d=1}^{K}\langle\bm{s}_{d},\bm{% s}_{d}\rangle]}}{[(2\pi)^{NK}\prod_{d=1}^{K}\det\bm{\Sigma}_{d}]^{1/2}}italic_p ( bold_italic_s start_POSTSUBSCRIPT italic_n italic_e italic_t end_POSTSUBSCRIPT | italic_n ) = divide start_ARG roman_exp [ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT ⟨ bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT , bold_italic_s start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ⟩ ] end_ARG start_ARG [ ( 2 italic_π ) start_POSTSUPERSCRIPT italic_N italic_K end_POSTSUPERSCRIPT ∏ start_POSTSUBSCRIPT italic_d = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_K end_POSTSUPERSCRIPT roman_det bold_Σ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT end_ARG (24)

Appendix B Gating and In-Painting

Generally, BBH models do not require the entire waveform to be analyzed at once. For example, the analyses throughout this paper independently examine the pre- and post-merger portions of the GW150914 waveform. To maintain independence between the models, any points not corresponding to the respective model (i.e. after tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for the pre-merger model, or before tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for the post-merger model) were excised, or “gated,” from the data. Here, a gate of length M𝑀Mitalic_M applied starting at a sample a𝑎aitalic_a will excise all samples within the range [a,a+M]𝑎𝑎𝑀[a,a+M][ italic_a , italic_a + italic_M ], corresponding to a gate of time length t=MΔt𝑡𝑀Δ𝑡t=M\Delta titalic_t = italic_M roman_Δ italic_t. The gated time series will therefore take the form 𝒔d,tr={s0,s1,sa,sa+M,sN1,sN}subscript𝒔𝑑𝑡𝑟subscript𝑠0subscript𝑠1subscript𝑠𝑎subscript𝑠𝑎𝑀subscript𝑠𝑁1subscript𝑠𝑁\bm{s}_{d,tr}=\{s_{0},s_{1},...s_{a},s_{a+M},...s_{N-1},s_{N}\}bold_italic_s start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT = { italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_s start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … italic_s start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , italic_s start_POSTSUBSCRIPT italic_a + italic_M end_POSTSUBSCRIPT , … italic_s start_POSTSUBSCRIPT italic_N - 1 end_POSTSUBSCRIPT , italic_s start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT }.

By doing this, the simplifications made to derive the likelihood function are no longer valid, since 𝚺d,trsubscript𝚺𝑑𝑡𝑟\bm{\Sigma}_{d,tr}bold_Σ start_POSTSUBSCRIPT italic_d , italic_t italic_r end_POSTSUBSCRIPT is no longer Toeplitz (and, subsequently, no longer approximately circulant for large N𝑁Nitalic_N). There are numerical methods to calculate the matrix inverse directly, but in general they can be unstable and time-intensive. Therefore, a method known as “gating and in-painting” is employed for gated waveform analyses [12].

First, the method assumes that the noise time series 𝒏𝒏\bm{n}bold_italic_n is the sum of 𝒏gsubscript𝒏𝑔\bm{n}_{g}bold_italic_n start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, the noise series with the gated times zeroed out, and 𝒙𝒙\bm{x}bold_italic_x, a time series containing only the gated samples in 𝒏𝒏\bm{n}bold_italic_n. The goal of in-painting is to solve the following equation in the gated region:

𝚺1(𝒏g+𝒙)=𝟎superscript𝚺1subscript𝒏𝑔𝒙0\bm{\Sigma}^{-1}(\bm{n}_{g}+\bm{x})=\bm{0}bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_italic_n start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT + bold_italic_x ) = bold_0 (25)

If the nonzero elements of 𝒙𝒙\bm{x}bold_italic_x are such that (𝚺1𝒏)[k]=𝟎superscript𝚺1𝒏delimited-[]𝑘0(\bm{\Sigma}^{-1}\bm{n})[k]=\bm{0}( bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_italic_n ) [ italic_k ] = bold_0 for all samples k𝑘kitalic_k in the gate [a,a+M]𝑎𝑎𝑀[a,a+M][ italic_a , italic_a + italic_M ], then the inner product 𝒏T𝚺𝒏superscript𝒏𝑇𝚺𝒏\bm{n}^{T}\bm{\Sigma}\bm{n}bold_italic_n start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_Σ bold_italic_n will be equal for the truncated and raw data set. Since 𝒙𝒙\bm{x}bold_italic_x is zero outside of the gate, 𝚺1𝒙superscript𝚺1𝒙\bm{\Sigma}^{-1}\bm{x}bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_italic_x will form an M×M𝑀𝑀M\times Mitalic_M × italic_M Toeplitz matrix containing the [a,a+M]𝑎𝑎𝑀[a,a+M][ italic_a , italic_a + italic_M ] rows and columns of 𝚺1superscript𝚺1\bm{\Sigma}^{-1}bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Therefore, Eq. (25) can be rewritten within the gated region as

𝚺1𝒙=𝚺1𝒏g,superscript𝚺1𝒙superscript𝚺1subscript𝒏𝑔\bm{\Sigma}^{-1}\bm{x}=-\bm{\Sigma}^{-1}\bm{n}_{g},bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_italic_x = - bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_italic_n start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT , (26)

and adding 𝒙𝒙\bm{x}bold_italic_x to 𝒏gsubscript𝒏𝑔\bm{n}_{g}bold_italic_n start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT will give the same result as truncating 𝒏𝒏\bm{n}bold_italic_n and 𝚺𝚺\bm{\Sigma}bold_Σ. Unlike trying to solve for the inverse directly, this solution is readily found using a Toeplitz solver [18, 35]. Given gated data 𝒔gsubscript𝒔𝑔\bm{s}_{g}bold_italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT containing some gated signal 𝒉gsubscript𝒉𝑔\bm{h}_{g}bold_italic_h start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, Eq. (26) can be evaluated using 𝒏g=𝒔g𝒉gsubscript𝒏𝑔subscript𝒔𝑔subscript𝒉𝑔\bm{n}_{g}=\bm{s}_{g}-\bm{h}_{g}bold_italic_n start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = bold_italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT - bold_italic_h start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, and the value 𝒙+𝒔g𝒉g𝒙subscript𝒔𝑔subscript𝒉𝑔\bm{x}+\bm{s}_{g}-\bm{h}_{g}bold_italic_x + bold_italic_s start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT - bold_italic_h start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT can be used to calculate the likelihood.

The analyses conducted in this paper utilize gating and in-painting to apply a gate starting/ending at tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to the waveform template. In theory, the gates should extend to the edges of the analysis segment (i.e., the pre-merger gate starts at the segment start time, and the post-merger gate ends at segment end). However, the in-painting algorithm, which is the dominant cost in our analysis, is O(M2)similar-toabsent𝑂superscript𝑀2\sim\!O(M^{2})∼ italic_O ( italic_M start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) operations for a gate of M𝑀Mitalic_M samples. Although the observable signal is only 0.2similar-toabsent0.2\sim 0.2\,∼ 0.2s long [13], we use an analysis segment that is 444\,4s in duration in order to resolve line artifacts in the PSD. We also use a sample rate of 204820482048\,2048Hz, in order to fully capture all observable signal power. The in-painting would therefore need to span M4096similar-to𝑀4096M\sim 4096italic_M ∼ 4096 samples if the gates were to extend to the beginning/end of the analysis segment. This is computationally expensive: some of our analyses would take 1similar-toabsent1\sim 1∼ 1 month to complete (utilizing 64 CPU cores).

To reduce the computational cost, we instead use a 1-second gate starting (ending) at tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and ensuring that the remainder of the pre-merger (post-merger) template 𝒉𝒉\bm{h}bold_italic_h after (before) the gate is zero. This is equivalent to applying a full gate to the template when a whitening filter is applied, since the early- and late-time template will be identically zero in both cases.

We do not zero the data 𝒔𝒔\bm{s}bold_italic_s outside of the gates, however, since doing so would introduce additional ringing artifacts at the beginning/end of the analysis segment not accounted for by the in-painting. This means that the pre- and post-merger models will share data outside of the gates (namely, greater than 1 second before and after tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT). However, the data is expected to be noise-dominated in this region, yielding an insignificant contribution to the likelihood. This can be seen in Appendix C, which portrays the remnant posteriors for analyses using zero-noise waveform injections. Since the data in these analyses contain no noise, the shared data has identically zero contribution to the likelihood. These results are in strong agreement with Figs. 4, 5, and 6, indicating that any effects due to early- and late-time shared noise are indeed negligible.

Appendix C Simulation results

This section contains additional plots showing results from the analysis of the GW150914-like simulated signal in zero noise. Figure 7 shows the Mfsubscript𝑀𝑓M_{f}italic_M start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and χfsubscript𝜒𝑓\chi_{f}italic_χ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT posteriors for the QNM post-merger models, and Fig. 8 shows the same for the IMR post-merger models. Figure 9 shows the H𝐻Hitalic_H posteriors for the zero-noise injection models.

Refer to caption
Figure 7: Final mass and final spin posteriors of GW150914 analyses using a QNM post-merger approximant and a zero-noise waveform injection. All masses are in the detector frame. Contours represent the 90th percentile of each distribution. The darker colored lines correspond to the model with variable sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, and the lighter lines represent the model where these parameters were fixed. Blue lines correspond to post-merger results, while red lines correspond to pre-merger results, obtained using numerical relativity with the IMR parameters. The innermost black contour represents the 4-OGC posterior from [19].
Refer to caption
Figure 8: Final mass and final spin posteriors of GW150914 analyses using an IMR post-merger approximant and a zero-noise waveform injection. All masses are in the detector frame. Contours represent the 90th percentile of each distribution. The darker colored lines correspond to the model with variable sky location and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, and the lighter lines represent the model where these parameters were fixed. Blue lines correspond to post-merger results, while red lines correspond to pre-merger results, obtained using numerical relativity with the IMR parameters. The innermost black contour represents the 4-OGC posterior from [19].
Refer to caption
Figure 9: Posteriors of H=(Af,measuredAi)/(Af,expectedAi)𝐻subscript𝐴𝑓𝑚𝑒𝑎𝑠𝑢𝑟𝑒𝑑subscript𝐴𝑖subscript𝐴𝑓𝑒𝑥𝑝𝑒𝑐𝑡𝑒𝑑subscript𝐴𝑖H=(A_{f,measured}-A_{i})/(A_{f,expected}-A_{i})italic_H = ( italic_A start_POSTSUBSCRIPT italic_f , italic_m italic_e italic_a italic_s italic_u italic_r italic_e italic_d end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) / ( italic_A start_POSTSUBSCRIPT italic_f , italic_e italic_x italic_p italic_e italic_c italic_t italic_e italic_d end_POSTSUBSCRIPT - italic_A start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) for the GW150914 analyses with zero-noise injected waveform data. Blue histograms correspond to QNM post-merger models, while red lines correspond to IMR post-merger models. Darker histograms correspond to models with variable sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT parameters, while lighter histograms correspond to models where these parameters are fixed to the maximum likelihood values from [19]. The expected value H=1𝐻1H=1italic_H = 1 is shown with a vertical black dotted line. Points in the shaded region correspond to H<0𝐻0H<0italic_H < 0 and therefore disagree with the area theorem.
Refer to caption
Figure 10: Final mass and spin posteriors for analyses of GW150914 data with a QNM post-merger model and a variable ringdown start time. The blue contour indicates the 90th percentile of an analysis where the dynesty sampler used 2000 live points. The orange countour represents the 90th percentile of an analysis with 4000 live points. Besides the number of sampler points used, both models were identical.
Refer to caption
Figure 11: Final mass and spin posteriors for analyses of GW150914 data with a IMR post-merger model and a variable ringdown start time. The blue contour indicates the 90th percentile of an analysis where the dynesty sampler used 2000 live points. The orange countour represents the 90th percentile of an analysis with 4000 live points. Besides the number of sampler points used, both models were identical.

Appendix D Sampler convergence

To test the convergence of the sampler we repeated the analyses with 2000 and 4000 live points. Figure 10 shows the Mfsubscript𝑀𝑓M_{f}italic_M start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT and χfsubscript𝜒𝑓\chi_{f}italic_χ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT posteriors for QNM post-merger models on real data with 2000 and 4000 live points, while Fig. 11 depicts the same for the IMR post-merger models. Both the QNM and IMR models were able to converge to similar distributions regardless of the number of live points used. All sky/tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT results are reported using 4000 live points.

References

  • Bardeen et al. [1973] J. M. Bardeen, B. Carter, and S. W. Hawking, The four laws of black hole mechanics, Communications in Mathematical Physics 31, 161 (1973).
  • Kastha et al. [2022] S. Kastha, C. D. Capano, J. Westerweck, M. Cabero, B. Krishnan, and A. B. Nielsen, Model systematics in time domain tests of binary black hole evolution, Physical Review D 10510.1103/physrevd.105.064042 (2022).
  • Cabero et al. [2018] M. Cabero, C. D. Capano, O. Fischer-Birnholtz, B. Krishnan, A. B. Nielsen, A. H. Nitz, and C. M. Biwer, Observational tests of the black hole area increase law, Physical Review D 9710.1103/physrevd.97.124069 (2018).
  • Isi et al. [2021] M. Isi, W. M. Farr, M. Giesler, M. A. Scheel, and S. A. Teukolsky, Testing the black-hole area law with GW150914, Physical Review Letters 12710.1103/physrevlett.127.011103 (2021).
  • Wang et al. [2023] Y.-F. Wang, C. D. Capano, J. Abedi, S. Kastha, B. Krishnan, A. B. Nielsen, A. H. Nitz, and J. Westerweck, A frequency-domain perspective on GW150914 ringdown overtone,   (2023), arXiv:2310.19645 [gr-qc] .
  • Carullo et al. [2019] G. Carullo, W. D. Pozzo, and J. Veitch, Observational black hole spectroscopy: A time-domain multimode analysis of GW150914, Physical Review D 9910.1103/physrevd.99.123029 (2019).
  • Carullo et al. [2023] G. Carullo, W. Del Pozzo, and J. Veitch, pyRing: a time-domain ringdown analysis python packagegit.ligo.org/lscsoft/pyring (2023).
  • Isi and Farr [2021a] M. Isi and W. M. Farr, maxisi/ringdown: Initial ringdown release (2021a).
  • Isi and Farr [2021b] M. Isi and W. M. Farr, Analyzing black-hole ringdowns,   (2021b), arXiv:2107.05609 [gr-qc] .
  • Finch and Moore [2022] E. Finch and C. J. Moore, Searching for a ringdown overtone in GW150914, Physical Review D 10610.1103/physrevd.106.043005 (2022).
  • Biwer et al. [2019] C. M. Biwer, C. D. Capano, S. De, M. Cabero, D. A. Brown, A. H. Nitz, and V. Raymond, PyCBC Inference: A Python-based parameter estimation toolkit for compact binary coalescence signals, Publications of the Astronomical Society of the Pacific 131, 024503 (2019).
  • Zackay et al. [2021] B. Zackay, T. Venumadhav, J. Roulet, L. Dai, and M. Zaldarriaga, Detecting gravitational waves in data with non-stationary and non-gaussian noise, Physical Review D 10410.1103/physrevd.104.063034 (2021).
  • Abbott et al. [2016] B. P. Abbott et al. (LIGO Scientific Collaboration and Virgo Collaboration), Observation of gravitational waves from a binary black hole merger, Physical Review Letters 116, 061102 (2016).
  • Sivia and Skilling [2006] D. Sivia and J. Skilling, Data analysis: a Bayesian tutorial (Oxford University Press, Oxford, 2006).
  • Finn [1992] L. S. Finn, Detection, measurement, and gravitational radiation, Physical Review D 46, 5236 (1992).
  • Pan et al. [1997] V. Pan, Y. Yu, and C. Stewart, Algebraic and numerical techniques for the computation of matrix determinants, Computers & Mathematics with Applications 34, 43 (1997).
  • Taylor [1996] J. R. Taylor, An Introduction to Error Analysis: The Study of Uncertainties in Physical Measurements, 2nd ed. (University Science Books, 1996).
  • et al. [2020a] P. V. et al., SciPy 1.0: fundamental algorithms for scientific computing in Python, Nature Methods 17, 261 (2020a).
  • Nitz et al. [2022] A. H. Nitz, S. Kumar, Y. Wang, S. Kastha, S. Wu, M. Schäfer, R. Dhurkunde, and C. D. Capano, 4-OGC: Catalog of gravitational waves from compact-binary mergers (2022), arXiv:2112.06878 [astro-ph.HE] .
  • Speagle [2020] J. S. Speagle, dynesty: a dynamic nested sampling package for estimating Bayesian posteriors and evidences, Monthly Notices of the Royal Astronomical Society 493, 3132 (2020).
  • et al. [2021] R. A. et al., Open data from the first and second observing runs of Advanced LIGO and Advanced Virgo, SoftwareX 1310.1016/j.softx.2021.100658 (2021).
  • Pratten et al. [2021] G. Pratten, C. García-Quirós, M. Colleoni, A. Ramos-Buades, H. Estellés, M. Mateu-Lucena, R. Jaume, M. Haney, D. Keitel, J. E. Thompson, and S. Husa, Computationally efficient models for the dominant and subdominant harmonic modes of precessing binary black holes, Physical Review D 10310.1103/physrevd.103.104056 (2021).
  • Isi et al. [2019] M. Isi, M. Giesler, W. M. Farr, M. A. Scheel, and S. A. Teukolsky, Testing the no-hair theorem with GW150914, Physical Review Letters 12310.1103/physrevlett.123.111102 (2019).
  • Vishveshwara [1970] C. V. Vishveshwara, Scattering of Gravitational Radiation by a Schwarzschild Black-hole, Nature 227, 936 (1970).
  • Chandrasekhar and Detweiler [1975] S. Chandrasekhar and S. L. Detweiler, The quasi-normal modes of the Schwarzschild black hole, Proc. Roy. Soc. Lond. A 344, 441 (1975).
  • Dreyer et al. [2004] O. Dreyer, B. J. Kelly, B. Krishnan, L. S. Finn, D. Garrison, and R. Lopez-Aleman, Black hole spectroscopy: Testing general relativity through gravitational wave observations, Class. Quant. Grav. 21, 787 (2004)arXiv:gr-qc/0309007 .
  • Cotesta et al. [2022] R. Cotesta, G. Carullo, E. Berti, and V. Cardoso, Analysis of ringdown overtones in GW150914, Physical Review Letters 12910.1103/physrevlett.129.111102 (2022).
  • Abbott et al. [2021] R. Abbott et al. (LIGO Scientific, VIRGO, KAGRA), Tests of General Relativity with GWTC-3,   (2021), arXiv:2112.06861 [gr-qc] .
  • Correia et al. [2023] A. Correia, Y.-F. Wang, and C. D. Capano, Low evidence for ringdown overtone in GW150914 when marginalizing over time and sky location uncertainty,   (2023), arXiv:2312.14118 [gr-qc] .
  • Isi and Farr [2022] M. Isi and W. M. Farr, Revisiting the ringdown of GW150914 (2022), arXiv:2202.02941 [gr-qc] .
  • Nitz et al. [2024] A. Nitz, I. Harry, D. Brown, C. M. Biwer, J. Willis, T. D. Canton, C. Capano, T. Dent, L. Pekowsky, G. S. C. Davies, S. De, M. Cabero, S. Wu, A. R. Williamson, B. Machenschalk, D. Macleod, F. Pannarale, P. Kumar, S. Reyes, dfinstad, S. Kumar, M. Tápai, L. Singer, P. Kumar, veronica villa, maxtrevor, B. U. V. Gadre, S. Khan, S. Fairhurst, and A. Tolley, gwastro/pycbc: v2.3.3 release of pycbc (2024).
  • Correia and Capano [2024] A. Correia and C. Capano, Sky marginalization in black hole spectroscopy (2024).
  • Gray [2001] R. Gray, Toeplitz and circulant matrices: a review, Foundations and Trends in Communications and Information Theory 2 (2001).
  • Chatfield [1996] C. Chatfield, The analysis of time series: an introduction, 6th ed. (CRC Press, Florida, US, 1996).
  • et al. [2020b] C. R. H. et al., Array programming with NumPy, Nature 585, 357 (2020b).