Measuring the Binary Black Hole Mass Spectrum with an Astrophysically Motivated Parameterization

and

Published 2018 April 5 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Colm Talbot and Eric Thrane 2018 ApJ 856 173 DOI 10.3847/1538-4357/aab34c

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/856/2/173

Abstract

Gravitational-wave detections have revealed a previously unknown population of stellar mass black holes with masses above 20 M. These observations provide a new way to test models of stellar evolution for massive stars. By considering the astrophysical processes likely to determine the shape of the binary black hole mass spectrum, we construct a parameterized model to capture key spectral features that relate gravitational-wave data to theoretical stellar astrophysics. In particular, we model the signature of pulsational pair-instability supernovae, which are expected to cause all stars with initial mass 100 M ≲ M ≲ 150 M to form ∼40 M black holes. This would cause a cutoff in the black hole mass spectrum along with an excess of black holes near 40 M. We carry out a simulated data study to illustrate some of the stellar physics that can be inferred using gravitational-wave measurements of binary black holes and demonstrate several such inferences that might be made in the near future. First, we measure the minimum and maximum stellar black hole mass. Second, we infer the presence of a peak due to pair-instability supernovae. Third, we measure the distribution of black hole mass ratios. Finally, we show how inadequate models of the black hole mass spectrum lead to biased estimates of the merger rate and the amplitude of the stochastic gravitational-wave background.

Export citation and abstract BibTeX RIS

1. Introduction

The black holes observed by advanced gravitational-wave detectors such as the Laser Interferometer Gravitational-wave Observatory (LIGO) (Aasi et al. 2015) and Virgo (Acernese et al. 2015) are widely believed to be formed from massive stars with initial mass M ≳ 20 M (Heger et al. 2003). Gravitational-wave measurements constrain the mass and spin of merging binaries. These measurements, in turn, can be used to better understand the evolution of massive stars. In this paper, we study how gravitational-wave measurements of the black hole mass spectrum can be used to inform our understanding of stellar evolution.

Simulating the final stages of stellar binary evolution is computationally expensive. Additionally, there are significant theoretical uncertainties in key aspects of binary evolution, especially the common envelope phase (Ivanova et al. 2013) and the supernova mechanism. For these reasons, populations of compact objects are simulated using population synthesis models (e.g., Dominik et al. 2015; Belczynski et al. 2017; Stevenson et al. 2017b). These are phenomenological models calibrated against a small number of more detailed stellar simulations.

At the time of writing, there have been five confirmed detections of binary black hole mergers and one unconfirmed candidate event (Abbott et al. 2016a, 2016b, 2016c, 2017a, 2017b, 2017c). The 90% credible regions for the source masses of the black holes range from ∼5 M to ∼40 M. Of the observed events, only one (GW151226) provides unambiguous evidence of black hole spin (Abbott et al. 2016b), although two events (GW150914 and GW170104) show a weak preference for spins anti-aligned with respect to the orbital angular momentum vector (Abbott et al. 2016d, 2017a). The implications of these measurements are currently unclear. Current theories include: most binary black holes are formed dynamically (Rodriguez et al. 2016a, 2017), black holes are subject to large natal kicks (Rodriguez et al. 2016b; O'Shaughnessy et al. 2017), and/or that large black holes do not form with significant dimensionless spins (Belczynski et al. 2017; Wysocki et al. 2018).

There has been significant work using gravitational-wave data to infer the properties of black hole formation with ensembles of detections. These works range from comparing gravitational-wave data to specific, non-parameterized models (Mandel & O'Shaughnessy 2010; Dominik et al. 2015; Stevenson et al. 2015, 2017a; Belczynski et al. 2016, 2017; Barrett et al. 2017; Farr et al. 2017; Miyamoto et al. 2017; Wysocki et al. 2018; Zevin et al. 2017), to attempts to group the data by binning, clustering, or Gaussian mixture modeling (Farr et al. 2018; Mandel et al. 2017; Wysocki 2017), to fitting physically motivated phenomenological population (hyper)parameters (Fishbach & Holz 2017; Kovetz et al. 2017; Talbot & Thrane 2017). In this work, we take the last approach and demonstrate that it is possible to identify physical features in the black hole mass spectrum with an ensemble of detections using phenomenological models, building on work in Kovetz et al. (2017) and Fishbach & Holz (2017).

Previous attempts to determine the binary black hole mass spectrum have employed one or more of these three approaches. Clustering is applied to a binned mass distribution in Mandel et al. (2017) to demonstrate that a mass gap between neutron star and black hole masses can be identified after O(100) observations. Stevenson et al. (2015), Barrett et al. (2017), and Zevin et al. (2017) compare population synthesis models with different physical assumptions and show that the predicted mass distributions can be distinguished using O(10) observations.

Previous analyses by the LIGO and Virgo scientific collaborations fit a power-law model with variable spectral index, α. Fishbach and Holz point out that, given LIGO/Virgo's additional sensitivity to heavier binary systems, there is a possible dearth of black holes larger than ∼40 M. They suggest that this is due to the occurrence of pulsational pair-instability supernovae (Heger & Woosley 2002) and propose an extension of the current LIGO analysis where the maximum mass is a free parameter. Pulsational pair-instability supernovae occur in stars with initial masses 100 M ≲ M ≲ 150 M, causing all stars in that mass range to form black holes with mass ∼40 M. In addition to a cutoff in the black hole mass spectrum, we expect that there will be an excess of black holes around the cutoff mass.

Kovetz et al. model the lower mass limit of black holes and use a Fisher analysis to demonstrate that it should be possible to identify the presence of the neutron star–black hole (NS–BH) mass gap. They also model the distribution of mass ratios and propose a test for detecting primordial black holes. A method to simultaneously estimate the binary black hole mass spectrum and the merger rate is presented in Wysocki (2017). Wysocki also considers a Gaussian mixture model for fitting the distribution of compact binary parameters.

The rest of the paper is structured as follows. In Section 2, we introduce the statistical tools necessary to make statements about the black hole population. We then develop our model in Section 3 in terms of population (hyper)parameters by considering current observational constraints and predictions from theoretical astrophysics and population synthesis. In Section 4, we perform a Monte Carlo injection study. We consider how many detections will be necessary to identify different features using Bayesian parameter estimation and model selection. We show how the predicted mass distributions differ when using different (hyper)parameterizations. We also explore some of the consequences of using inadequate (hyper)parameterizations. In particular, we show that inadequate (hyper)parameterization can lead to significant bias in the estimate of the merger rate and the predicted amplitude of the stochastic gravitational-wave background (SGWB). Some closing thoughts are provided in Section 5.

2. Bayesian Inference

2.1. Gravitational-wave Detection

A binary black hole system is completely described by a set of 15 parameters, Θ. Recovering these parameters from the observed strain data requires the use of specialized Bayesian parameter inference software, e.g., LALInference (Veitch et al. 2015). The likelihood of a given set of binary parameters is computed by comparing the strain data to the signal predicted by general relativity. For the analysis presented here, the expected signal is calculated using phenomenological approximations to numerical relativity waveforms (Hannam et al. 2014; Schmidt et al. 2015; Smith et al. 2016). For a given set of strain data, hi, LALInference returns a set of ni samples, {Θ}, which are sampled from the posterior distribution, $p({\rm{\Theta }}| {h}_{i},H)$, of the binary parameters, along with the Bayesian evidence, ${ \mathcal Z }({h}_{i}| H)$, where H is the model being tested.

The distribution of binary black hole systems observed by current detectors is not representative of the astrophysical distribution of binary black holes. The observing volume of current gravitational-wave detectors is limited by the instruments' sensitivity. The observing volume of a gravitational-wave detector is primarily determined by the masses of the black holes, with spin entering as a higher order effect. More massive systems produce gravitational waves of greater amplitude. However, these more massive systems merge at a lower frequency and hence spend less time in the observing band of the detector. Additionally, distant sources undergo cosmological redshift and appear more massive than they actually are. Here, we will deal only with the unredshifted "source-frame" masses, not the "lab-frame" masses observed by gravitational-wave detectors. We note that the source/lab-frame distinction is about cosmological redshift and is not a statement about detectability and/or selection effects.

Accounting for these factors, we calculate Vobs(Θ), the sensitive volume for a binary with parameters Θ, following Abbott et al. (2016c), using semianalytic noise models corresponding to different sensitivities (Abbott et al. 2016e). The noise, and hence sensitivity, in real detectors is time-dependent, and so calculating this volume requires averaging over the observing time to obtain a mean sensitive volume $\langle {V}_{\mathrm{obs}}({\rm{\Theta }})\rangle $ (Abbott et al. 2016c).

2.2. Population Inference

We are interested in inferring population (hyper)parameters describing the distribution of source-frame black hole masses. The formalism to do this is briefly described below (see, e.g., Gelman et al. 2013, chapter 29 for a more detailed discussion of hierarchical Bayesian modeling and Mandel et al. 2014 for a discussion of selection biases).

Hierarchical inference of this kind can be cast as a post hoc method of changing from the prior distribution used in the single-event parameter estimation to a new prior, which depends on population (hyper)parameters, Λ. We marginalize over all of the binary parameters while reweighting the posterior samples by the ratio between our (hyper)parameterized model and the prior used to generate the posterior distribution. This marginalization integral is approximated by summing over the posterior samples for each event. The N events are then combined by multiplying together the new marginalized likelihoods for the individual events,

Equation (1)

Here $\pi ({\rm{\Theta }}| \mathrm{LAL})$ is the prior probability distribution used for single-event parameter estimation. The distribution $\pi ({{\rm{\Theta }}}_{j}^{i}| {\rm{\Lambda }},H)$ is the probability of a binary having parameters Θ in our model; see Section 3. We do not model any black hole parameters other than source-frame mass, and so Θ can be replaced by m = (m1, m2) in Equation (1) and all following equations. The prior distribution of source-frame masses in LALInference is a convolution of a prior that is uniform in component mass between limits determined by the reduced order model (Smith et al. 2016) and the redshift distribution that is uniform in luminosity distance extending out to 4 Gpc. This distance distribution is converted to redshift assuming ΛCDM cosmology using the results from the Planck 2015 data release (Planck Collaboration et al. 2016).

We combine this likelihood with $\pi ({\rm{\Lambda }}| H)$, the prior for the (hyper)parameters assuming a model H, and the Bayesian evidence for the data given H to obtain the posterior distribution for our (hyper)parameters,

Equation (2)

Equation (3)

To perform the (hyper)parameter estimation we use the Python implementation of MultiNest (Feroz et al. 2009; Buchner et al. 2014). Additionally, we calculate the posterior predictive distribution (PPD) of the binary parameters,

Equation (4)

where Λk are the nk (hyper)posterior samples. The PPD shows the probability that a subsequent detection will have parameters Θ given the previous data, {h}.

2.3. Model Selection

Model selection is performed in our Bayesian framework by considering Bayes factors,

Equation (5)

A large Bayes factor, ${\mathrm{BF}}_{\beta }^{\alpha }\gg 1$, indicates that Hα is strongly favored over Hβ. We adopt a conventional threshold of ln BF = 8 to distinguish between two models.

3. Phenomenology

In this section, we develop a parameterization of the black hole mass spectrum using predictions from astrophysics theory, population synthesis models, and electromagnetic observations. In this way, we can relate gravitational-wave measurements to stellar astrophysics. For low-mass systems, the parameter that most strongly affects the observable gravitational waveform is a combination of the component masses known as the chirp mass, ${ \mathcal M }={({m}_{1}{m}_{2})}^{3/5}/{({m}_{1}+{m}_{2})}^{1/5}$. For high-mass systems, the waveform is primarily determined by the total mass of the system. The mass ratio is more difficult to determine due to covariance between the mass ratio and the spin of the black holes.

The canonical assumed distribution of black hole masses is a power-law distribution in the primary mass between some maximum and minimum masses. This power-law distribution has three typical parameters: the spectral index α, the minimum mass mmin, and the maximum mass mmax. The distribution of secondary mass is typically taken to be flat between mmin and m1. We take this as the starting point for our parameterization.

3.1. High-mass Binaries

The observation of binary black hole mergers through the detection of gravitational waves revealed the presence of a previously unobserved population of black holes with mass ∼30 M (Abbott et al. 2016a, 2017a, 2017b). Since gravitational-wave detectors can observe more massive binaries at greater distances, binaries containing larger black holes are preferentially detected over less massive systems. Fishbach and Holz (2017) note that, given the observed rate of binaries with mass ∼30 M, it is somewhat surprising that we have not seen more massive black holes. They propose that this is due to a cutoff in the black hole mass spectrum around this mass. By comparing the Bayesian evidence for mmax = 41 M and mmax = 100 M using the first four events, they find tentative support for a cutoff. They further show that it will be possible to identify the presence of an "upper mass gap" with a Bayes factor of ≳150 (ln BF ≈ 5) using 10 detections, and the cutoff mass can be measured with ∼40 detections.

The theoretical motivation for such a cutoff is pulsational pair-instability supernovae (PPSNe) (Heger & Woosley 2002; Woosley & Heger 2015), whereby large amounts of matter are ejected prior to collapse to form a black hole. The expected result of this process is that all stars with initial mass 100 M ≲ M ≲ 150 M form black holes with masses ∼40 M. Stars with 150 M ≲ M ≲ 250 M are expected to undergo pair-instability supernovae (PISNe) and leave no remnant. Hence, we expect a gap in the black hole mass spectrum between ∼40 M and ∼250 M along with an excess of black holes at some mass mpp ∼ 40 M. The excess is due to the stars with 100 M ≲ M ≲ 150 M, which become PPSNe. The size, position, and shape of the peak are determined by the unknown details of PPSNe. While the observation of cutoff near 40 M could be interpreted as evidence for PPSNe, the additional observation of a peak would provide a smoking-gun signature that the highest mass stellar binaries are reduced in mass via PPSNe.

Thus, we extend our description of the upper end of the black hole mass spectrum to allow for the possibility of an excess due to PPSNe. We model this as a normal distribution with unknown mean mpp ∈ [25, 100]  M and variance σpp < 5 M. It is also necessary to introduce a mixing fraction parameter, λ, which describes the proportion of binaries that are drawn from the normal distribution.

We expect that any mass peak due to PPSNe should be near the high-mass cutoff, which corresponds to mmax ≈ mpp. Also, assuming that the power law is otherwise a good description of the black hole mass spectrum, we expect that the number of black holes in the PPSNe peak should be no more than the number of black holes that would have formed had the power law distribution continued to the upper limit of the upper mass gap, determined by the onset of pair-instability supernovae, mPI ∼ 150 M. We impose this condition by requiring that the extrapolated area that would be under the power-law curve is less than the area contained within the Gaussian. This amounts to a restriction on the allowed values of λ,

Equation (6)

Here, mmin is the upper limit of the NS–BH mass gap and mmax is the lower limit of the upper mass gap. The variable mPI is the mass above which stars become PISNe, leaving no remnant. Here, we assume α > 1 and mPI ≫ mmax.

To hone our intuition, we can plug in plausible values of α, mmin, mmax, and mPI to determine a typical value of λ. For example, if we set mmin = 5 M, mmax = 50 M, and α = 2, we find λ ∼ 0.1. If we measure a peak consistent with these values, we anticipate that the position and width of the peak can inform our physical understanding of this mechanism. If λ is measured to be inconsistent with this constraint it could indicate either that the extrapolation of the power law is not a valid assumption or that the peak is not entirely due to PPSNe. We note that the fraction of observed black holes that formed through PPSNe will be larger than λ since the more massive black holes are observable out to a greater distance.

3.2. Low-mass Binaries

The smaller sensitive volume for low-mass binaries means that it is more difficult to probe the low-mass end of the black hole mass spectrum with gravitational-wave detections. Previous analyses of the black hole mass spectrum from gravitational-wave detections have assumed that it has a sharp cutoff at some minimum mass mmin. However, this overestimates the number of low-mass black holes if the distribution of low-mass black holes in merging binaries is the same as that in low-mass X-ray binaries (Özel et al. 2010). Population synthesis models also generically predict that the primary mass distribution peaks above the minimum mass.

We replace the step function at the low-mass end of the fiducial black hole mass spectrum with a smoothing function, S(m, mmin, δm), which rises from unity at mmin to unity at mmin + δm,

Equation (7)

Equation (8)

We note that δm = 0 recovers the step function used in previous analyses. Since the mass distribution is expected to be an increasing function, the peak of p(m1) occurs below mmin + δm. We expect mmin ∼ 5 M and δm ≳ 3 given that the black hole mass spectrum inferred from electromagnetic observations peaks at 8 M, with no black holes less massive than 5 M.

Our model of the distribution of the primary mass can be summarized as

Equation (9)

where

encodes the power-law distribution with a smooth turn-on at low mass and

encodes the peak from PPSNe.

3.3. Mass Ratio

Previous analyses by the LIGO/Virgo scientific collaborations have assumed that the secondary mass is distributed uniformly between a lower limit set by mmin and an upper limit of m1. This is motivated by observations of the stellar initial binary population (e.g., Kroupa et al. 2013; Belloni et al. 2017). In contrast to this, population synthesis models typically predict that the distribution of mass ratios should prefer equal-mass binaries (e.g., Belczynski et al. 2017). We model the distribution of the mass ratio as a power law with spectral index β as in Fishbach & Holz (2017) and Kovetz et al. (2017). For a distribution of mass ratio peaked at equal masses, β > 0. We also impose the same smoothing at the lower limit as we apply to the primary mass. This allows us to write down the conditional probability distribution for secondary masses given a primary mass,

Equation (10)

3.4. Summary

A table listing the (hyper)parameters and their physical meaning is provided in Table 1. Including a factor of Vobs to account for selection biases, the probability of detecting a mass pair given our (hyper)parameters, Λ = {α, mmin, mmax, δm, λ, mpp, σpp, β} and under model H, is

Equation (11)

We consider six different models for our (hyper)prior distribution, each corresponding to decreasingly stringent physical assumptions. These different prior assumptions can be tested with our Bayesian framework using Bayes factors. In our first model, H0, we take the power-law distribution with maximum mass, mmax = 100 M. Since the injected data set considered in Section 4 has a minimum mass of 3 M, we allow mmin to vary, rather than fixing it at mmin = 5 M as in previous analyses. In H1 we introduce a uniform prior on mmax. In order to determine the relative importance of the different features, we switch to models that include all the effects described above except one. In H2, H3, and H4 we do not include the Gaussian component, mass ratio, and low-mass smoothing respectively. Finally, in H5 we include all of these effects. These prior choices are summarized in Table 2.

Table 1.  (Hyper)Parameters Describing the Black Hole Mass Spectrum

α Spectral index of m1 for the power-law distributed
  component as the mass spectrum.
mmax Maximum mass of the power-law distributed
  component as the mass spectrum.
λ Proportion of primary black holes formed via PPSNe.
mpp Mean mass of black holes formed via PPSNe.
σpp Standard deviation of masses of black holes formed
  via PPSNe.
mmin Minimum black hole mass.
δm Mass range over which the black hole mass spectrum
  turns on.
β Spectral index of m2.

Download table as:  ASCIITypeset image

Table 2.  Summary of Example Models

  α mmax λ mpp σpp β mmin δm
H0 [−3, 7] 100 0 N/A N/A 0 [2, 10] 0
H1 [−3, 7] [10, 100] 0 N/A N/A 0 [2, 10] 0
H2 [−3, 7] [10, 100] 0 N/A N/A [−5, 5] [2, 10] [0, 10]
H3 [−3, 7] [10, 100] [0, 1] [25, 100] (0, 5] 0 [2, 10] [0, 10]
H4 [−3, 7] [10, 100] [0, 1] [25, 100] (0, 5] [−5, 5] [2, 10] 0
H5 [−3, 7] [10, 100] [0, 1] [25, 100] (0, 5] [−5, 5] [2, 10] [0, 10]
MC 1.5 35 0.1 35 1 2 3 5

Note.  The prior ranges for our (hyper)parameters in each model are indicated. Each of these distributions is uniform over the stated range. The fixed parameters are in bold. "MC" refers to the values chosen for the simulated universe in Section 4. These values are chosen to be consistent with current observational data and theoretical predictions of pulsational pair-instability supernovae and population synthesis modeling.

Download table as:  ASCIITypeset image

3.5. Other Effects

As with any phenomenological model, our model has limitations. If the proposed mass gaps exist in the population of black holes formed as the endpoint of stellar evolution there may still be black holes found in these gaps. The remnant of the binary neutron star merger GW170817 has mass Mrem ≲2.8 M (Abbott et al. 2017d). It is not clear whether this object is a neutron star or a black hole. Similarly, the remnant from binary black hole mergers such as GW150914 is more massive than the suggested upper mass limit due to PPSNe, ${M}_{\mathrm{rem}}\,={62}_{-4}^{+4}\,{M}_{\odot }$ (Abbott et al. 2016f). Both of these objects lie within the proposed mass gaps. If either of these mergers happened in a dense environment such as a globular cluster, it is possible that such objects could merge with a new companion (Heggie 1975; Rodriguez et al. 2017). Similarly, primordial black holes (Hawking 1971) are not bound by the limitations of stellar evolution.

We do not expect either of these mechanisms to significantly affect the position and shape of an excess due to PPSNe, although they complicate the interpretation of the maximum black hole mass. It is possible that black holes formed through repeated mergers could be identified on a case-by-case basis. For example, black holes formed by a binary black hole merger event are expected to have large dimensionless spins, a ∼ 0.7 for equal-mass non-spinning pre-merger black holes (Scheel et al. 2009).

4. Monte Carlo Study

We verify that we are able to recover a distribution described by a particular set of (hyper)parameters using a Monte Carlo injection study. We create a simulated universe in which the black hole mass distribution follows our model with (hyper)parameters given in Table 2. For simplicity, we draw all of the extrinsic parameters from the geometrically determined prior distribution used by LALInference (Veitch et al. 2015) with luminosity distance extending to 4 Gpc. We draw the masses according to Equation (11) and draw black hole spins uniformly in spin magnitude (up to a maximum of a = 0.89) and isotropically in orientation.

Motivated by the prediction of pulsational pair-instability supernovae, our simulated universe includes a Gaussian component centered at the upper limit of the power-law component, mmax = 35 M. The Gaussian component has a width σpp = 1 M and the mixing fraction λ = 0.1. The inferred value of α is covariant with other parameters of the model; e.g., for the first four detections, decreasing the maximum black hole mass, decreases the inferred value of α (Fishbach & Holz 2017). We set α = 1.5 for our injection study, which is consistent with that analysis. We choose the spectral index of the secondary mass distribution to be β = 2 to reflect the preference of population synthesis models to produce binaries of near equal mass. We impose a lower mass cutoff of 3 M with a turn-on of δm = 5 M. These values are chosen to be consistent with current observational data. The distribution of primary masses with this choice of (hyper)parameters is shown in Figure 1. We can see that our model gives us a bimodal distribution with peaks at ≈7 M, due to the smooth turn-on, and 35 M, due to PPSNe.

Figure 1.

Figure 1. The astrophysical distribution of source-frame masses assuming our modeled distribution of m1 with (hyper)parameters as specified in Table 2. We identify the excess of black holes at 35 M due to PPSNe. We also see the smooth turn-on at low masses.

Standard image High-resolution image

To enforce selection effects, we keep only binaries with optimal matched filter signal-to-noise ratio ρ > 8 in a single Advanced LIGO detector operating at design sensitivity (Abbott et al. 2016e). We generate a set of 200 events for our simulated universe. Each signal is then injected into a three-detector LIGO/Virgo network with all detectors operating at their design sensitivities. Figure 2 shows the distribution of primary masses and mass ratio in our simulated universe before (dashed) and after (solid) accounting for observational bias. The blue histogram indicates the injected values.

Figure 2.

Figure 2. The distribution of source-frame primary mass and mass ratio (q ≡ m2/m1) for our simulated universe, see Table 2. The dashed and solid lines show the distribution before and after accounting for selection biases respectively. The blue histogram indicates the injected values.

Standard image High-resolution image

Using the recovered posterior distributions for the injected events, we employ the statistical methods described above for each of our models. The Bayes factors comparing H5 to the others are enumerated in Table 3. In Table 3 we also give an approximate number of events needed to reach our threshold ln BF = 8, assuming linear growth of ln BF with number of detections. We consider two cases. "Cosmic" assumes zero measurement error. All uncertainty comes from cosmic variance. "Design" uses posterior samples obtained through running LALInference for a three-detector network operating at design sensitivity. Including measurement errors reduces our resolving power between any pair of models by a significant factor for all the models. Unless otherwise specified we refer to the Design Bayes factors. Below, we consider the effect of each of the modifications on the mass distribution model.

Table 3.  Bayes factors comparing H5 to the others

  H0 H1 H2 H3 H4
Cosmic ${{\rm{ln}}\,{\rm{BF}}}_{i}^{5}$ 253.0 55.0 18.0 31.0 1.0
Design ${{\rm{ln}}\,{\rm{BF}}}_{i}^{5}$ 161.0 14.0 5.0 7.0 −1.0
Nexpected 10 100 300 250 ≫200

Note.  The table lists the log Bayes factor comparing each of the hypotheses summarized in Table 2 to the correct model, H5, given the 200 samples shown in Figure 2 with population (hyper)parameters as specified in Table 2. "Cosmic" indicates that the masses are used with no measurement error; this represents an upper limit on how well we can differentiate the two distributions. "Design" uses the output of LALInference for a three-detector advanced LIGO/Virgo network operating at design sensitivity. The bottom row gives an approximate number of events to reach our threshold of Design ln BF = 8. Measuring the shape of the low-mass cutoff will require many detections, because of the lower sensitivity at low masses.

Download table as:  ASCIITypeset image

4.1. Upper Mass Cutoff

After 200 events, our model without the variable upper mass cutoff, H0, is disfavored with a log Bayes factor of ∼160. We determine how many events are necessary to surpass the threshold of ln BF50 = 8 by considering subsets of our injection set. After 20 detections ${\mathrm{lnBF}}_{0}^{5}\sim N(\mu =13.3,\sigma =2.4$). Here, N(μ, σ) denotes a normal distribution with mean μ and variance σ2.

Given this, we expect to be able to identify an upper mass cutoff in the mass distribution after ≲20 events. We note that μ grows linearly with the number of events; this scaling is used in Table 3 to approximate the number of events to reach ln BF = 8. This is consistent with a similar study by Fishbach & Holz (2017).

4.2. PPSN Peak

After 200 events, the posterior distribution on λ, the fraction of black holes formed through PPSNe, is shown in Figure 4. We measure the maximum posterior probability point and 95% highest density confidence interval (HDI) to be $\lambda \,\sim \,{0.11}_{-0.04}^{+0.07}$ (all future confidence regions will be 95% HDI unless specified) and disfavor λ = 0 at ≳3σ. Correspondingly, ${\mathrm{lnBF}}_{2}^{5}=4.9$, which is moderate evidence for the existence of the PPSN peak, but below our threshold for a confident detection.

Figure 3 shows the one- and two-dimensional posterior distributions for the position and width of the PPSN peak. We measure ${m}_{{pp}}={34.4}_{-1.2}^{+1.0}{M}_{\odot }$, ${\sigma }_{{pp}}={1.2}_{-1.2}^{+0.9}{M}_{\odot }$. We note that the peak and width of the distribution are covariant, with smaller values of mpp requiring a larger σpp. This is unsurprising because the black holes with the highest mass, ∼40 M, must be accounted for. The posterior distribution on λ shows no significant correlation with mpp or σpp.

Figure 3.

Figure 3. The posterior on the mean and width of the PPSN peak using our 200 events injected into Gaussian noise. After 200 detections we can measure the position and width of the PPSN peak to within ∼1 M at 95% confidence. The dark and light shaded regions indicate the one-dimensional 68% and 95% confidence intervals.

Standard image High-resolution image

4.3. Mass Ratio

The posterior distribution on β is shown in Figure 4. After 200 events, the 1σ and 2σ confidence intervals on β span 1.1 and 2.2 respectively. We disfavor H3 with a log Bayes factor of ${\mathrm{lnBF}}_{3}^{5}\,=7.1$ after our 200 injections, just below our threshold of 8.

Figure 4.

Figure 4. The posterior on the fraction of black holes formed through PPSNe and the power-law index on the mass ratio using our 200 events injected into Gaussian noise. We can measure the fraction of black holes formed through PPSNe to be $\lambda \sim {0.11}_{-0.04}^{+0.07}$ at 95% confidence. We can determine the spectral index of the distribution of mass ratios to within ±1 at 95% confidence with 200 detections.

Standard image High-resolution image

4.4. Low Mass

The (hyper)parameters describing the low-mass end of the distribution are more difficult to measure than the high-mass (hyper)parameters because the observation bias favors high-mass systems. After 200 events, there is no evidence for or against the low-mass smoothing described by δm. This is unsurprising since only nine injected binaries have m1 ≲ 8 M, above which H4 and H5 are identical. Measuring the same events with improved strain sensitivity would not improve our sensitivity to δm. We are limited by cosmic variance: Cosmic ${\mathrm{lnBF}}_{4}^{5}=1$. The two-dimensional posterior distribution on mmin and δm, Figure 5, shows the correlation between low minimum masses and long turn-on lengths.

Figure 5.

Figure 5. The posterior on the (hyper)parameters describing the low-mass end of the black hole mass spectrum using our 200 events injected into Gaussian noise. These parameters are difficult to measure because only about 5% of events have m1 < 8 M. We can see the clear covariance between a low minimum mass and a long turn-on length.

Standard image High-resolution image

4.5. Recovery of the Mass Distribution

As a qualitative measure of the difference between the inferred mass distributions, we plot the posterior predictive distribution for the primary mass given the binaries in our injection study for our models in Figure 6. The dashed black line indicates the injected distribution. We can see the effect of the different (hyper)parameterizations.

Figure 6.

Figure 6. Posterior predictive distribution (see Equation (4)) for m1 and q for our models from Table 2 after 200 injected events. The dashed black line indicates the true distribution and the notches indicate the injected values. These distributions represent the observed distribution of source-frame masses.

Standard image High-resolution image

In order to accommodate the lack of black holes with m ≳ 40 M, α is overestimated in model H0. This leads to an overly steep inferred distribution and an overestimate of the total merger rate (see Section 4.6). Models H1 and H2, which do not include the Gaussian component, favor a mass spectrum that is less steep than the injected distribution. This is manifested as a positive gradient after accounting for observational bias. If we do not fit the power-law index of the mass ratio, we overestimate the number of heavy black holes. This bias is seen as an overestimate of λ in H3 and maximum mass in H1.

4.6. Impact on the Merger Rate

The majority of binary black hole mergers are not individually resolvable by Advanced LIGO/Virgo. Using a (hyper)parameterization that does not accurately describe the true distribution leads to a biased estimate of the fraction of mergers that are individually resolvable and hence the merger rate (Abadie et al. 2010; Abbott et al. 2017a). Compact binary coalescences are a Poisson process that can be described by a merger rate R(Λ). For a detector with time-independent sensitivity and a model of the distribution of binary black hole systems, the merger rate can be inferred from the number of observed events N, the sensitive volume of our detectors V(Λ), and the observation time T:

Equation (12)

where

Equation (13)

and Vobs(Θ) is the volume sensitive to a given binary as introduced in Section 2.

To illustrate the dependence of R on the mass distribution model, we calculate the posterior distribution for the inferred estimate of the merger rate for the models described in Table 2. For each model, we compute the posterior distribution for R. During the first observing run of Advanced LIGO, ∼3 binary black hole mergers were identified in ∼48 days joint observing time (Abbott et al. 2016c). We use these values, i.e., N = 3 and T = 48/365 yr, to normalize our rate estimates. We neglect the (currently large) Poisson uncertainty in the arrival rate since this will be small once 200 detections have been made. Figure 7 shows the posterior distribution for the merger rate for each of our models. We note that if we assume that the power-law mass distribution extends out to 100 M (the dashed–dotted line), we overestimate the merger rate by a factor of 2–3. This is due to the much larger α required to be consistent with the lack of detections at high masses. A similar result is obtained in Wysocki (2017) by simultaneously fitting the merger rate and power-law spectral index.

Figure 7.

Figure 7. Posterior distribution for the binary black hole merger rate after 200 simulated events. We assume a detection rate consistent with Advanced LIGO's first observing run, N/T ≈ 23 yr−1, and ignore Poisson uncertainties. The models are described in Table 2. The dashed black line indicates the rate for the injected distribution. If we do not fit the maximum mass (the dashed–dotted line) the rate is overestimated by a factor of 2–3. The inferred merger rate is not strongly sensitive to any of the other modifications to the mass function.

Standard image High-resolution image

4.7. Impact on the Stochastic Background

Unresolvable mergers are widely believed to make the dominant contribution to the SGWB (Abbott et al. 2016g, 2017e, 2017f). The SGWB is typically characterized by the ratio of the energy density of the Universe in gravitational waves to the energy required to close the Universe, ΩGW. The frequency of current detectors that is most sensitive to the SGWB is ∼30 Hz; this frequency corresponds to the inspiral phase of all binaries relevant to this work. The energy density due to binary black hole mergers depends on the distribution of chirp masses and the merger rate (Zhu et al. 2011),

Equation (14)

As seen above, cutting off the mass distribution around 40 M leads to a reduction in the merger rate; however, this is accompanied by an increase in ${{ \mathcal M }}^{5/3}$. Overall, this leads to a reduction of ∼10% in the expected SGWB as seen in Figure 8. We also observe that relaxing the assumption that the secondary mass is uniformly distributed leads to a further ∼10% reduction in ΩGW. This is because the chirp mass is maximized for equal-mass binaries for a given primary mass. These reductions are smaller than the current uncertainty on the amplitude of the background due to Poisson uncertainty in the observed merger rate. The current method of searching for this background is by cross-correlating the strain data from the two LIGO detectors; this method will take more time to resolve a weaker background.

Figure 8.

Figure 8. Posterior distribution for the ratio of the amplitude of the expected stochastic gravitational-wave background to the value using the injected distribution. The models are described in Table 2. The dashed lines indicate models in which the distribution of mass ratios is assumed to be uniform. The dashed–dotted line indicates the model in which the maximum black hole mass is fixed to be 100 M. Allowing the maximum mass of the power-law component to vary decreases the predicted amplitude of the stochastic background by ∼10%. Relaxing the assumption that the distribution of secondary masses is uniform between the minimum mass and the primary mass decreases the predicted amplitude by ∼10%.

Standard image High-resolution image

The cross-correlation method is expected to require years of observation before the background can be resolved. Recently, a method involving searching directly for the stochastic background due to binary black hole mergers has been introduced in Smith & Thrane (2017). This method is expected to be able to detect this component of the background using days of data. Since this method relies on the rate of binary black hole mergers rather than ΩGW it will be more sensitive to the black hole mass function than cross-correlation searches.

5. Discussion

The first gravitational-wave detections are revealing a previously unexplored population of black holes. While we are still in the regime of small-number statistics, the systems observed to date may be suggestive of a cutoff in the black hole mass spectrum at ∼40 M. This is consistent with the predicted black hole mass distribution if stars with initial masses M ≳ 100 M undergo pulsational pair-instability supernovae. We hypothesize that, if this is the cause of the cutoff, then there should be a corresponding excess of black holes at around the same mass. We construct a phenomenological model, which captures this behavior. In agreement with Fishbach & Holz (2017), we find that the presence of an upper mass cutoff can be identified at high significance with O(10) events.

We highlight several other interesting results that can be obtained using 200 detections at design sensitivity.

  • 1.  
    We will be able to identify the presence of an excess due to PPSNe at ∼3σ and constrain the fraction of black holes forming through PPSNe to within ∼0.05 at 95% confidence.
  • 2.  
    We can measure the position and width of the PPSN graveyard to within ∼1 M.
  • 3.  
    We will be able to measure the power-law index on the mass ratio to within ∼±1.

Detailed measurement of the low-mass end of the mass distribution will most likely require thousands of detections and may have to wait for future detectors, e.g., the proposed Einstein Telescope (Punturo et al. 2010) or Cosmic Explorer (Abbott et al. 2017g).

We demonstrate that neglecting the presence of either a cutoff or a mass peak can lead to a mis-recovery of the astrophysical distribution of black holes in merging binaries. For example, the higher sensitivity of current detectors to high-mass binaries means that in order to fit the upper mass range well, the low-mass distribution is biased. This leads to incorrect estimates of the total binary black hole merger rate and the predicted amplitude of the SGWB. The amplitude of the SGWB is also sensitive to the distribution of mass ratios.

Our analysis assumes that a clear distinction can be made between binary black hole systems and other compact binaries. In reality, if there is not a well-defined mass gap between neutron stars and black holes, it will be non-trivial to distinguish between binary black hole, neutron star–black hole, and binary neutron star systems (Yang et al. 2017), although differences in, e.g., the spins of the component objects may enable this distinction (Littenberg et al. 2015). Our framework can be naturally expanded to include these other classes of compact binaries.

This is LIGO document P1700449. We would like to thank Richard O'Shaughnessy for helpful comments. E.T. is supported by ARC CE170100004 and FT150100281.

Please wait… references are loading.
10.3847/1538-4357/aab34c