Nonparametric analysis of correlations in the binary black hole population with LIGO–Virgo–KAGRA data

Jack Heinzel\orcidlink0000-0002-5794-821X [email protected] LIGO, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA Kavli Institute for Astrophysics and Space Research and Department of Physics, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA    Matthew Mould\orcidlink0000-0001-5460-2910 LIGO, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA Kavli Institute for Astrophysics and Space Research and Department of Physics, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA    Salvatore Vitale\orcidlink0000-0003-2700-0767 LIGO, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA Kavli Institute for Astrophysics and Space Research and Department of Physics, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA
(June 24, 2024)
Abstract

Formation channels of merging compact binaries imprint themselves on the distributions and correlations of their source parameters. But current understanding of this population observed in gravitational waves is hindered by simplified parametric models. We overcome these limitations using PixelPop [Heinzel et al. (2024)]—our multidimensional Bayesian nonparametric population model. We analyze data from the first three LIGO–Virgo–KAGRA observing runs and make high resolution, minimally modeled measurements of the pairwise distributions of binary black hole masses, redshifts, and spins. There is no evidence that the mass spectrum evolves over redshift and we show that such measurements are fundamentally limited by the detector horizon. We find support for correlations of the spin distribution with binary mass ratio and redshift, but at reduced significance compared to overly constraining parametric models. Confident data-driven conclusions about population-level correlations using very flexible models like PixelPop will require more informative gravitational-wave catalogs.

Introduction.—From the first gravitational-wave (GW) detection of a merging black hole (BH) binary [1] to the catalog observed by the end of the third observing run (O3) [2, 3, 4, 5] of the LIGO–Virgo–KAGRA collaboration (LVK) [6, 7, 8], our understanding of the underlying population has improved significantly [9, 10, 11]: the merger rate increases with redshift, consistent with cosmic star-formation history [12, 13, 14, 15]; there is a continuum of masses with peaks at around 10M10subscript𝑀direct-product10M_{\odot}10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and 35M35subscript𝑀direct-product35M_{\odot}35 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, with most binaries having nearly equal masses [16, 17, 18, 19, 20]; and there are few BHs with large spins [21, 22, 23]. Potential astrophysical correlations between source parameters have also been identified [24, 25, 26, 27, 28]. Despite this, the physical origins of these features are unclear, though there are a plethora of potential formation channels (e.g., Refs. [29, 30, 31, 32, 33, 34, 35, 36, 37, 38, 39, 40, 41, 42, 43, 44]) with distinct GW signatures (e.g., Refs [45, 46, 47, 48, 49, 50, 51, 52, 53, 54, 55]). Interpreting GW data in light of these predictions could yield key astrophysical insights.

The above population constraints are made using a model for the merger rate density as a function of the binary source parameters, for which the combined catalog of GW events can be used to infer a Bayesian posterior [56, 57, 58, 59]. Due to large theoretical uncertainties [60], a major difficulty is choosing appropriate models. Perhaps the most common approach due to its simplicity is to parametrize the rate density with basic functions (such as power laws and normal distributions) and infer their parameters (e.g., Refs. [11, 13, 16, 21, 22, 61, 62]). However, such choices can be overly constraining and lead to missed features in the inferred population or, conversely, features driven by the strong model assumptions rather than data [63, 64, 65]. An alternative approach is to use nonparametric models in order to impose any assumptions as weakly as possible [66, 11, 67, 68, 69, 70, 71, 72, 73, 74]. Such flexible models are better able to fit complicated structures in the source distributions that may arise from astrophysical formation processes, though typically at the cost of larger uncertainties, higher computational cost, and lack of interpretability. Moreover, it is challenging to extend these methods to simultaneously probe multiple parameter dimensions and their correlations with high fidelity [11, 75, 73, 76, 77, 65].

In Ref. [78] we developed PixelPop—a Bayesian nonparametric multidimensional population model for inference on correlated parameter distributions, such as those that may result from the astrophysical formation of compact binaries. PixelPop works by densely binning the joint space of binary source parameters and inferring the comoving merger rate density in each bin. The only assumption made is a weak smoothing prior that couples each bin to its nearest neighbors—the conditional autoregressive (CAR) prior. The computational efficiency of the CAR model allows us to dramatically increase its resolution compared to similar nonparametric methods [66, 11, 73]—using 𝒪(104)𝒪superscript104\mathcal{O}(10^{4})caligraphic_O ( 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) bins compared to 𝒪(102)𝒪superscript102\mathcal{O}(10^{2})caligraphic_O ( 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT )—and thus provide a much more complete view of GW populations while making fewer assumptions about them.

In this Letter, we analyze the LVK catalog of binary BH mergers with PixelPop, including the 69 binary BH events with false-alarm rates <1yr1absent1superscriptyr1<1\,\mathrm{yr}^{-1}< 1 roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT in the third GW transient catalog (GWTC-3). We specifically target the joint distributions of: (1) the heavier (primary) BH mass m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and redshift z𝑧zitalic_z, (2) binary mass ratio 0<q10𝑞10<q\leq 10 < italic_q ≤ 1 and effective aligned spin χeff(1,1)subscript𝜒eff11\chi_{\mathrm{eff}}\in(-1,1)italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ∈ ( - 1 , 1 ) [79], and (3) χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and z𝑧zitalic_z (when not inferred with PixelPop we fit the mass, redshift, and spin distributions with the parametric Power Law + Peak mass model, Power Law redshift model, and Default spin model from Ref. [11]). We use the publicly available [80, 81] parameter-estimation samples [82, 83, 84, 85] (Overall, PrecessingSpinIMRHM, and Mixed for events from GWTC-1, 2, and 3, respectively) and sensitivity estimates [86] to compute the GW population likelihood, following a Monte Carlo approximation [57, 58, 11] and penalizing the likelihood in regions of high approximation uncertainty [87, 88, 89]. Full details of PixelPop and its analysis methods are available in Heinzel et al. [78].

Primary mass and redshift.—The observed primary mass distribution spans roughly two orders of magnitude, so we place 100 bins uniformly over both lnm1subscript𝑚1\ln m_{1}roman_ln italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT from m1=2Msubscript𝑚12subscript𝑀direct-productm_{1}=2M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 2 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT to m1=100Msubscript𝑚1100subscript𝑀direct-productm_{1}=100M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 100 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and redshift from z=0𝑧0z=0italic_z = 0 to z=1.9𝑧1.9z=1.9italic_z = 1.9 (104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT bins total). In Fig. 1 we show the comoving merger rate density (m1;z)subscript𝑚1𝑧\mathcal{R}(m_{1};z)caligraphic_R ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ; italic_z ) inferred from GWTC-3 over m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and z𝑧zitalic_z; note that (m1;z)subscript𝑚1𝑧\mathcal{R}(m_{1};z)caligraphic_R ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ; italic_z ) is a function of redshift z𝑧zitalic_z through the comoving volume element, but is not a density in z𝑧zitalic_z. A representative average over the posterior uncertainty is given by the median rate bins in the central panel. One dimensional slices of the joint distribution and their associated uncertainties are displayed in the upper and right-hand panels.

Refer to caption
Figure 1: Comoving merger rate density (lnm1;z)=m1(m1;z)subscript𝑚1𝑧subscript𝑚1subscript𝑚1𝑧\mathcal{R}(\ln m_{1};z)=m_{1}\mathcal{R}(m_{1};z)caligraphic_R ( roman_ln italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ; italic_z ) = italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT caligraphic_R ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ; italic_z ) as a function of primary mass m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and redshift z𝑧zitalic_z, inferred from GWTC-3. The central panel shows the posterior median, with lower to higher rates colored black to yellow, respectively. The white curve defines the detection horizon above which a binary BH merger is not detectable. The upper horizontal panel shows \mathcal{R}caligraphic_R as a function of m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT at two fixed redshifts, z=0.1𝑧0.1z=0.1italic_z = 0.1 (blue) and z=0.75𝑧0.75z=0.75italic_z = 0.75 (orange). The solid line is the posterior median while the shaded region encloses the 90% symmetric credible region. Similarly, the right-hand vertical panel shows \mathcal{R}caligraphic_R as a function of z𝑧zitalic_z at the two peaks in the merger rate of m1=10Msubscript𝑚110subscript𝑀direct-productm_{1}=10M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (blue) and m1=35Msubscript𝑚135subscript𝑀direct-productm_{1}=35M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 35 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (orange).

We observe peaks at m110Msubscript𝑚110subscript𝑀direct-productm_{1}\approx 10M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and m135Msubscript𝑚135subscript𝑀direct-productm_{1}\approx 35M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 35 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, consistent with previous parametric and nonparametric results [16, 11, 90, 67, 91, 20, 72]. For z[0,0.2]𝑧00.2z\in[0,0.2]italic_z ∈ [ 0 , 0.2 ], the merger rate density integrated over m1/M[8,15]subscript𝑚1subscript𝑀direct-product815m_{1}/M_{\odot}\in[8,15]italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ∈ [ 8 , 15 ] is higher than when integrated over m1/M[8,100]subscript𝑚1subscript𝑀direct-product8100m_{1}/M_{\odot}\in[8,100]italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ∈ [ 8 , 100 ] at almost 100% posterior credibility. The secondary peak has somewhat lower significance in the same redshift interval; we quantify this with the local integrated rate density over m1/M[25,40]subscript𝑚1subscript𝑀direct-product2540m_{1}/M_{\odot}\in[25,40]italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ∈ [ 25 , 40 ] and m1/M[15,50]subscript𝑚1subscript𝑀direct-product1550m_{1}/M_{\odot}\in[15,50]italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ∈ [ 15 , 50 ], where the former is higher than the latter at 72% credibility.

Within the posterior uncertainty we find no evidence for structure beyond these two peaks. But unlike for parametric mass models, the inferred merger rate does not decrease at small masses and the uncertainty grows as there is insufficient information in the catalog to constrain the merger rate with the flexibility of PixelPop. This is to be expected because at low masses the detector horizon—the redshift at which a source is not detectable, given by the white curve111We compute this as the maximum redshift at which a source has optimal signal-to-noise ratio >8absent8>8> 8 in a LIGO–Virgo detector network with O3-like sensitivity [92]. in Fig. 1—excludes most of the mass range. In the absence of GW events, inference in this region of parameter space is dominated by the CAR prior, similar to other nonparametric models [72, 93]. In contrast, the merger rate decreases at large masses and low redshifts where sources should be detectable; having not detected such sources implies a low astrophysical merger rate.

There is a mild preference for an increasing merger rate at small redshifts z0.2less-than-or-similar-to𝑧0.2z\lesssim 0.2italic_z ≲ 0.2, consistent with parametric models [13, 14, 11, 15]. For masses m1/M[8,100]subscript𝑚1subscript𝑀direct-product8100m_{1}/M_{\odot}\in[8,100]italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ∈ [ 8 , 100 ] beyond the prior-dominated regime, the integrated comoving rate density is higher over z[0.1,0.2]𝑧0.10.2z\in[0.1,0.2]italic_z ∈ [ 0.1 , 0.2 ] than z[0,0.1]𝑧00.1z\in[0,0.1]italic_z ∈ [ 0 , 0.1 ] at 67% credibility. At m1=10Msubscript𝑚110subscript𝑀direct-productm_{1}=10M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, the comoving merger rate increases from z=0𝑧0z=0italic_z = 0 to z=0.2𝑧0.2z=0.2italic_z = 0.2 at 73% credibility. With monotonic parametric models [11, 13], a positive slope at low redshifts enforces the same at higher redshifts, whereas PixelPop does not make this requirement. There appears to be another redshift mode at z0.7𝑧0.7z\approx 0.7italic_z ≈ 0.7 for masses within the detection horizon, coincident with the onset of a plateau in the merger rate inferred in Ref. [72]. At m1=35Msubscript𝑚135subscript𝑀direct-productm_{1}=35M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 35 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, the comoving merger rate increases from z=0𝑧0z=0italic_z = 0 to z=0.7𝑧0.7z=0.7italic_z = 0.7 at 79% credibility. Above this redshift the posterior uncertainty increases as the majority of the parameter space lies beyond the detector horizon where the CAR model prefers broad distributions.

Beyond this, we do not confidently identify any correlations in the mass–redshift parameter space, in agreement with Ref. [73, 94]. While the m135Msubscript𝑚135subscript𝑀direct-productm_{1}\approx 35M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 35 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT peak extends out to the z0.7𝑧0.7z\approx 0.7italic_z ≈ 0.7 mode and beyond, we cannot conclude that the mass spectrum evolves over redshift as suggested in Ref. [95] because the m110Msubscript𝑚110subscript𝑀direct-productm_{1}\approx 10M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT mode is limited by the detector horizon. We cannot exclude the possibility that the lower-mass peak has the same redshift extent as the higher-mass peak.

Mass ratio and effective spin.—The effective spin χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT is the mass-weighted average of the two BH spin components along the binary orbital angular momentum; it is large and positive (negative) for binaries with spinning BHs aligned (antialigned) with the orbital angular momentum, and closer to zero for small or misaligned BH spins. Callister et al. [24] originally discovered evidence for an anticorrelation between mass ratios q𝑞qitalic_q and effective spins χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT—i.e., binaries with larger χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT tend to have more unequal masses—with a strongly parameterized model. This result was corroborated by Refs. [96, 97, 70, 11] with additional models, though with reduced confidence in regions of uninformative data [70]. However, mechanisms that could produce this correlation are not well understood [49, 45, 48, 98, 99, 100, 101, 100] and the models used make strong assumptions about the form of the correlation. Instead, we quantify the evidence for this correlation and constrain its form with minimal assumptions by fitting the merger rate density (χeff,q;z)subscript𝜒eff𝑞𝑧\mathcal{R}(\chi_{\mathrm{eff}},q;z)caligraphic_R ( italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT , italic_q ; italic_z ) with 100 bins over both χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and q𝑞qitalic_q using PixelPop.

In Fig. 2, we show the inferred comoving merger rate density evaluated at a redshift of z=0.2𝑧0.2z=0.2italic_z = 0.2, corresponding to the peak in Fig. 1. The merger rate is highest for near-equal masses and χeff0subscript𝜒eff0\chi_{\mathrm{eff}}\approx 0italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ≈ 0 and decreases towards unequal masses and |χeff|>0subscript𝜒eff0|\chi_{\mathrm{eff}}|>0| italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT | > 0, though the data do not require the global maximum to be q=1𝑞1q=1italic_q = 1 and χeff=0subscript𝜒eff0\chi_{\mathrm{eff}}=0italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0. PixelPop does not enforce that the merger rate vanishes as q0𝑞0q\to 0italic_q → 0 (unlike common parametric models [11]) and there is also support for negative χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT, though the tails of the distributions again become dominated by the CAR prior (cf. Ref. [72]). The bulk of the population has q>0.6𝑞0.6q>0.6italic_q > 0.6 and in this region the integrated merger rate is higher for χeff[0,0.5]subscript𝜒eff00.5\chi_{\mathrm{eff}}\in[0,0.5]italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ∈ [ 0 , 0.5 ] than χeff[0.5,0]subscript𝜒eff0.50\chi_{\mathrm{eff}}\in[-0.5,0]italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ∈ [ - 0.5 , 0 ] at 91.7% posterior credibility, indicating that component BH spins tend to prefer alignment with the orbital angular momentum [11]. Specifically, the merger rate density at z=0.2𝑧0.2z=0.2italic_z = 0.2 is 11.33.8+5.8Gpc3yr1superscriptsubscript11.33.85.8superscriptGpc3superscriptyr111.3_{-3.8}^{+5.8}\;\mathrm{Gpc}^{-3}\mathrm{yr}^{-1}11.3 start_POSTSUBSCRIPT - 3.8 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 5.8 end_POSTSUPERSCRIPT roman_Gpc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT for positive spins χeff[0,0.5]subscript𝜒eff00.5\chi_{\mathrm{eff}}\in[0,0.5]italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ∈ [ 0 , 0.5 ] (median and 90% credible interval) but 7.43.1+4.8Gpc3yr1superscriptsubscript7.43.14.8superscriptGpc3superscriptyr17.4_{-3.1}^{+4.8}\;\mathrm{Gpc}^{-3}\mathrm{yr}^{-1}7.4 start_POSTSUBSCRIPT - 3.1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 4.8 end_POSTSUPERSCRIPT roman_Gpc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT for negative χeff[0.5,0]subscript𝜒eff0.50\chi_{\mathrm{eff}}\in[-0.5,0]italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ∈ [ - 0.5 , 0 ], in both cases taking q[0.6,1]𝑞0.61q\in[0.6,1]italic_q ∈ [ 0.6 , 1 ].

Refer to caption
Figure 2: Joint comoving merger rate density (χeff,q;z=0.2)subscript𝜒eff𝑞𝑧0.2\mathcal{R}(\chi_{\mathrm{eff}},q;z=0.2)caligraphic_R ( italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT , italic_q ; italic_z = 0.2 ) as a function of effective spin χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and binary mass ratio q𝑞qitalic_q, evaluated at a fixed redshift z=0.2𝑧0.2z=0.2italic_z = 0.2. The central panel displays the joint two-dimensional posterior median. The merger rate density evaluated at χeff=0subscript𝜒eff0\chi_{\mathrm{eff}}=0italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0 (blue) and χeff=0.25subscript𝜒eff0.25\chi_{\mathrm{eff}}=0.25italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.25 (orange) is shown as a function of q𝑞qitalic_q in the upper panel, and evaluated at q=1𝑞1q=1italic_q = 1 (blue) and q=0.4𝑞0.4q=0.4italic_q = 0.4 (orange) as a function of χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT in the right-hand panel. For the one-dimensional slices the bold line gives the posterior median while the shaded region encloses the 90% symmetric credible region.

The two-dimensional posterior median in Fig. 2 suggests a tail in the distribution toward (q,χeff)(0.25,0.25)𝑞subscript𝜒eff0.250.25(q,\chi_{\mathrm{eff}})\approx(0.25,0.25)( italic_q , italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ) ≈ ( 0.25 , 0.25 ). In the region q[0.2,0.4]𝑞0.20.4q\in[0.2,0.4]italic_q ∈ [ 0.2 , 0.4 ], we find a higher rate density for χeff[0,0.5]subscript𝜒eff00.5\chi_{\mathrm{eff}}\in[0,0.5]italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ∈ [ 0 , 0.5 ] than χeff[0.5,0]subscript𝜒eff0.50\chi_{\mathrm{eff}}\in[-0.5,0]italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ∈ [ - 0.5 , 0 ] at 66% credibility when accounting for the posterior uncertainty. To further test for the aforementioned q𝑞qitalic_qχeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT correlation, we compute the Spearman rank correlation coefficient ρ𝜌\rhoitalic_ρ [102, 103]—which is positive (negative) when there is an increasing (decreasing) monotonic correlation—for (χeff,q)subscript𝜒eff𝑞\mathcal{R}(\chi_{\mathrm{eff}},q)caligraphic_R ( italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT , italic_q ), restricting to q[0.2,1]𝑞0.21q\in[0.2,1]italic_q ∈ [ 0.2 , 1 ] and χeff[0.5,0.5]subscript𝜒eff0.50.5\chi_{\mathrm{eff}}\in[-0.5,0.5]italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ∈ [ - 0.5 , 0.5 ] to avoid prior-dominated regions. Whereas Ref. [11] identify a negative correlation between q𝑞qitalic_q and χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT at 97.5% significance, we find ρ=0.0230.270+0.281𝜌superscriptsubscript0.0230.2700.281\rho=0.023_{-0.270}^{+0.281}italic_ρ = 0.023 start_POSTSUBSCRIPT - 0.270 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.281 end_POSTSUPERSCRIPT and no significant evidence for a correlation. We stress, however, that the PixelPop posterior does not exclude such a correlation and that this result is not at odds with the results of strongly parametrized models [24, 11, 96, 70] when used to analyze the same GW catalog—if one mandates a functional form that allows a global monotonic trend then the data prefer an anticorrelation, but this is neither required nor excluded by the more flexible PixelPop model that allows a much broader class of nontrivial correlations.

Redshift and effective spin.Biscoveanu et al. [25] found evidence for an astrophysical broadening of the χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT distribution with increasing redshift, assuming a linear relationship. This was confirmed by Heinzel et al. [70] with a more flexible model for the correlation, but in both cases a simple Gaussian parametrization was made for the underlying population. We remove these assumptions by using PixelPop to infer the comoving merger rate density (χeff;z)subscript𝜒eff𝑧\mathcal{R}(\chi_{\mathrm{eff}};z)caligraphic_R ( italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ; italic_z ) jointly over 100 bins in each of χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and z𝑧zitalic_z.

We show the PixelPop posterior in Fig. 3. The merger rate may evolve differently over redshift depending on the effective spin: for χeff=0subscript𝜒eff0\chi_{\mathrm{eff}}=0italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0 the median rate density is initially higher at lower redshifts than χeff=0.25subscript𝜒eff0.25\chi_{\mathrm{eff}}=0.25italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.25, increases from z=0𝑧0z=0italic_z = 0 to z0.25𝑧0.25z\approx 0.25italic_z ≈ 0.25, then drops at z0.5𝑧0.5z\approx 0.5italic_z ≈ 0.5 before increasing again at z0.7𝑧0.7z\approx 0.7italic_z ≈ 0.7; for χeff=0.25subscript𝜒eff0.25\chi_{\mathrm{eff}}=0.25italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.25 the median rate density increases monotonically between z=0𝑧0z=0italic_z = 0 and z0.7𝑧0.7z\approx 0.7italic_z ≈ 0.7; in both cases the rate density plateaus in the prior-dominated region z1greater-than-or-equivalent-to𝑧1z\gtrsim 1italic_z ≳ 1. While these results hold for the median inferred rate, this marginalizes over large uncertainties, within which a lack of redshift peaks is also possible.

Refer to caption
Figure 3: Comoving merger rate density (χeff;z)subscript𝜒eff𝑧\mathcal{R}(\chi_{\mathrm{eff}};z)caligraphic_R ( italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ; italic_z ) as a function of effective spin χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and redshift z𝑧zitalic_z. The central panel displays the joint two-dimensional posterior median. The merger rate density evaluated at χeff=0subscript𝜒eff0\chi_{\mathrm{eff}}=0italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0 (blue) and χeff=0.25subscript𝜒eff0.25\chi_{\mathrm{eff}}=0.25italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.25 (orange) is shown as a function of z𝑧zitalic_z in the upper panel, and evaluated at z=0.1𝑧0.1z=0.1italic_z = 0.1 (blue) and z=0.75𝑧0.75z=0.75italic_z = 0.75 (orange) as a function of χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT in the right-hand panel. For the one-dimensional slices the bold line gives the posterior median while the shaded region encloses the 90% symmetric credible region.

We can confidently conclude that the merger rate significantly decreases for |χeff|0.5greater-than-or-equivalent-tosubscript𝜒eff0.5|\chi_{\mathrm{eff}}|\gtrsim 0.5| italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT | ≳ 0.5 at low redshifts where sources would be detectable but have not been detected, implying that the astrophysical merger rate must be low. This corresponds to the broadening of the χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT spin distribution visible in both the two-dimensional median posterior rate density and the one-dimensional slices in Fig. 3. There is a more pronounced peak just above χeff0subscript𝜒eff0\chi_{\mathrm{eff}}\approx 0italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ≈ 0 at z=0.1𝑧0.1z=0.1italic_z = 0.1 compared to z=0.75𝑧0.75z=0.75italic_z = 0.75 where the overall rate is higher. We quantify this trend by computing the Spearman rank correlation coefficient ρ𝜌\rhoitalic_ρ between the redshift and the width of the effective spin distribution—if ρ𝜌\rhoitalic_ρ is positive, the χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT distribution broadens at larger redshifts. However, we must be careful not to mistake the natural broadening of the CAR model in uninformative regions of parameter space—such as higher redshifts—for an astrophysical correlation. For z<0.5𝑧0.5z<0.5italic_z < 0.5, we find a weak preference for broadening, with ρ=0.0960.240+0.231𝜌superscriptsubscript0.0960.2400.231\rho=0.096_{-0.240}^{+0.231}italic_ρ = 0.096 start_POSTSUBSCRIPT - 0.240 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.231 end_POSTSUPERSCRIPT and ρ>0𝜌0\rho>0italic_ρ > 0 at 75% credibility; cf. 98% in Refs. [25, 70]. Limiting to z<1𝑧1z<1italic_z < 1 instead yields ρ=0.1640.236+0.267𝜌superscriptsubscript0.1640.2360.267\rho=0.164_{-0.236}^{+0.267}italic_ρ = 0.164 start_POSTSUBSCRIPT - 0.236 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + 0.267 end_POSTSUPERSCRIPT and ρ>0𝜌0\rho>0italic_ρ > 0 at a larger but still mild significance of 84%. The broadening of χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT over redshift appears to be a genuine astrophysical feature in the data, but with PixelPop we cannot make a definitive statement about the underlying nature of the population without more informative data.

Conclusions.—As the catalog of binary BH mergers observed with GW detectors grows, increasingly more information can be obtained about the underlying population. In particular, nontrivial structures and population-level parameter correlations may be identified and interpreted in terms of their astrophysical origins and formation channels, but only by using a suitable population model. Indeed, correlations have been identified in the population of binary BHs with simplified parametric models [24, 25, 70, 96, 97, 26, 28], but these risk misspecification and overconfident inference.

PixelPop [78] is a nonparametric multidimensional population model that imposes minimal assumptions on the form of the underlying merger rate and is computationally efficient. We used PixelPop to analyze bivariate source distributions with high resolution—primary BH masses and redshifts, binary mass ratios and effective spins, and effective spins and redshifts—using LVK binary BH detections [2, 3, 4, 5]. We showed that: there is no evidence for an evolving BH mass spectrum over redshift and this measurement is currently prohibited by the detector horizon; and the GW data are consistent with a distribution of effective spins that is correlated with binary mass ratio and redshift, but that we cannot be confident in this hypothesis without strong model assumptions.

Flexible population inference methods such as Bayesian nonparametrics will be increasingly useful in the future as GW catalogs become more informative. They offer a valuable consistency check that strongly parameterized models are not overfitting to or extrapolating from observational data, as the shapes of distributions as well as correlations in their joint space may be complicated and impossible to model with simple parameterizations. Flexible models can also be used to account for unmodeled contributions to the GW population that would otherwise lead to systematically biased inference [104], or for cosmological constrains with astrophysics-agnostic assumptions about the source population [105, 93]. The ability of PixelPop to capture arbitrary correlations between source parameters make it a standout model for such GW population analyses.

Acknowledgements.—We thank Sofía Álvarez-López, Jacob Golomb, Cailin Plunkett, Noah Wolfe, and the Rates and Populations LIGO working group for useful discussions and helpful comments. J.H is supported by the NSF Graduate Research Fellowship under Grant No. DGE1122374. M.M. is supported by LIGO Laboratory through the National Science Foundation award PHY-1764464. J.H. and S.V. are partially supported by the NSF grant PHY-2045740. This material is based upon work supported by NSF’s LIGO Laboratory which is a major facility fully funded by the National Science Foundation. The authors are grateful for computational resources provided by the LIGO Laboratory and supported by National Science Foundation Grants PHY-0757058 and PHY-0823459.

References

I Supplemental Material

We include the analysis of additional pairwise correlations in LIGO–Virgo–KAGRA (LVK) data from the first three gravitational-wave (GW) catalogs using PixelPop. Below, we infer the joint distributions of: (1) primary binary black hole (BH) mass m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and effective spin χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT; (2) m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and binary mass ratio q𝑞qitalic_q; and (3) BH spin magnitudes a𝑎aitalic_a and orbital misalignment angles θ𝜃\thetaitalic_θ.

Primary mass and effective spin.Ray et al. [76] find evidence that the χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT distribution is different for merging binary BHs with masses in the range [30M,40M]30subscript𝑀direct-product40subscript𝑀direct-product[30M_{\odot},40M_{\odot}][ 30 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT , 40 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ], using a three-dimensional binned Gaussian process prior to model primary mass, secondary mass, and effective spin. We use PixelPop to infer the merger rate density (m1,χeff;z)subscript𝑚1subscript𝜒eff𝑧\mathcal{R}(m_{1},\chi_{\mathrm{eff}};z)caligraphic_R ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ; italic_z ) with 100 uniform bins in both lnm1subscript𝑚1\ln m_{1}roman_ln italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT between m1=2Msubscript𝑚12subscript𝑀direct-productm_{1}=2M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 2 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and m1=100Msubscript𝑚1100subscript𝑀direct-productm_{1}=100M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 100 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT and χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT over (1,1)11(-1,1)( - 1 , 1 ). We plot the inferred merger-rate density evaluated at z=0.2𝑧0.2z=0.2italic_z = 0.2 in Fig. 4.

Refer to caption
Figure 4: Comoving merger rate density (lnm1,χeff;z)=m1(m1,χeff;z)subscript𝑚1subscript𝜒eff𝑧subscript𝑚1subscript𝑚1subscript𝜒eff𝑧\mathcal{R}(\ln m_{1},\chi_{\rm eff};z)=m_{1}\mathcal{R}(m_{1},\chi_{\rm eff};z)caligraphic_R ( roman_ln italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ; italic_z ) = italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT caligraphic_R ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ; italic_z ) as a function of primary mass m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and effective spin χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT, evaluated at a fixed redshift z=0.2𝑧0.2z=0.2italic_z = 0.2. The central panel displays the joint two-dimensional posterior median. The merger rate density evaluated at χeff=0subscript𝜒eff0\chi_{\mathrm{eff}}=0italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0 (blue) and χeff=0.25subscript𝜒eff0.25\chi_{\mathrm{eff}}=0.25italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0.25 (orange) is shown as a function of m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT in the upper panel, and evaluated at m1=10Msubscript𝑚110subscript𝑀direct-productm_{1}=10M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (blue) and m1=35Msubscript𝑚135subscript𝑀direct-productm_{1}=35M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 35 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (orange) as a function of χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT in the right-hand panel. For the one-dimensional slices the bold line gives the posterior median while the shaded region encloses the 90% symmetric credible region.

We find no evidence for different effective-spin distributions at different primary masses. Due to the different nonparametric approaches, this is not necessarily at odds with the findings of Ray et al. [76]. Since we do not jointly model the secondary masses in our analysis, this may suggest that when modeled independently, the two component BH mass distributions [77, 65] have different χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT distributions. However, no evidence for this has previously been found with parametric population models [106]. Interestingly, when modeling the joint m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPTχeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT distribution an additional mode between 10<m1/M<3010subscript𝑚1subscript𝑀direct-product3010<m_{1}/M_{\odot}<3010 < italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT < 30 appears at χeff=0subscript𝜒eff0\chi_{\mathrm{eff}}=0italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT = 0, but at low significance.

Primary mass and mass ratio.—To assess differences between the component mass distributions, we run PixelPop to infer the merger-rate density (m1,q;z)subscript𝑚1𝑞𝑧\mathcal{R}(m_{1},q;z)caligraphic_R ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_q ; italic_z ) jointly over primary mass m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and mass ratio q𝑞qitalic_q with the same bin placement as previously. As PixelPop allows for any correlations over the parameter space, this is entirely equivalent to modeling the joint distribution of component masses with a reparametrization, as in Refs. [11, 75, 73, 76, 77, 65]. We show the PixelPop results, evaluated at z=0.2𝑧0.2z=0.2italic_z = 0.2, in Fig. 5.

Refer to caption
Figure 5: Comoving merger rate density (lnm1,q;z)=m1(m1,q;z)subscript𝑚1𝑞𝑧subscript𝑚1subscript𝑚1𝑞𝑧\mathcal{R}(\ln m_{1},q;z)=m_{1}\mathcal{R}(m_{1},q;z)caligraphic_R ( roman_ln italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_q ; italic_z ) = italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT caligraphic_R ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_q ; italic_z ) as a function of primary mass m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and mass ratio q𝑞qitalic_q, evaluated at a fixed redshift z=0.2𝑧0.2z=0.2italic_z = 0.2. The central panel displays the joint two-dimensional posterior median. The merger rate density evaluated at q=0.7𝑞0.7q=0.7italic_q = 0.7 (blue) and q=1𝑞1q=1italic_q = 1 (orange) is shown as a function of m1subscript𝑚1m_{1}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT in the upper panel, and evaluated at m1=10Msubscript𝑚110subscript𝑀direct-productm_{1}=10M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (blue) and m1=35Msubscript𝑚135subscript𝑀direct-productm_{1}=35M_{\odot}italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 35 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (orange) as a function of q𝑞qitalic_q in the right-hand panel. For the one-dimensional slices the bold line gives the posterior median while the shaded region encloses the 90% symmetric credible region.

There is a potential trend in the primary mass distribution shifting to larger masses for mass ratios further from unity, and the marginal mass-ratio distribution may be flatter at larger primary masses. However, within the large PixelPop posterior uncertainties the merger rate may also be uncorrelated and even flat over component masses.

Component spins.—We also perform a nonparametric inference on the joint distribution of the component spins using PixelPop. These are specified by the dimensionless spin magnitudes a1subscript𝑎1a_{1}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, a2subscript𝑎2a_{2}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and the polar angles θ1subscript𝜃1\theta_{1}italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, θ2subscript𝜃2\theta_{2}italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT between the component BH spin vectors and the binary orbital angular momentum for the primary (1) and secondary (2) BHs, respectively. We assume that the spins are independent and identically (IID) distributed. Phrased in terms of the comoving differential merger rate density (a,cosθ)𝑎𝜃\mathcal{R}(a,\cos\theta)caligraphic_R ( italic_a , roman_cos italic_θ ), this means that

(a1,cosθ1,a2,cosθ2)=(a1,cosθ1)(a2,cosθ2)dadcosθ(a,cosθ).subscript𝑎1subscript𝜃1subscript𝑎2subscript𝜃2subscript𝑎1subscript𝜃1subscript𝑎2subscript𝜃2𝑎cos𝜃𝑎𝜃\displaystyle\mathcal{R}(a_{1},\cos\theta_{1},a_{2},\cos\theta_{2})=\frac{% \mathcal{R}(a_{1},\cos\theta_{1})\mathcal{R}(a_{2},\cos\theta_{2})}{\int% \differential a\,\differential\mathrm{cos}\theta\ \mathcal{R}(a,\cos\theta)}\,.caligraphic_R ( italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = divide start_ARG caligraphic_R ( italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , roman_cos italic_θ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) caligraphic_R ( italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , roman_cos italic_θ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG start_ARG ∫ start_DIFFOP roman_d end_DIFFOP italic_a start_DIFFOP roman_d end_DIFFOP roman_cos italic_θ caligraphic_R ( italic_a , roman_cos italic_θ ) end_ARG . (1)

This enforces the IID probability density for the component BH spins without double counting the overall merger rate.

As the merger rate model contributes twice due to the IID assumption, the uncertainty in the Monte Carlo likelihood estimator mentioned in the Introduction and described in detail in Ref. [78] is effectively doubled. We therefore replace the parameter-estimation samples for a subset of GW events (GW150914, GW170608, GW170729, GW170818, GW170823, GW200129_065458, and GW200112_155838) with the samples from the IMRPhenomXP [107] analyses. This increases the minimum sample count over all events from 3,194 to 14,802, thereby reducing Monte Carlo uncertainty in estimating the likelihood for each event. Additionally, we create a custom set of 109absentsuperscript109\approx 10^{9}≈ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT simulated GW signals using the IMRPhenomXP waveform approximant [107], added to Gaussian noise colored by representative power spectral densities from the first, second, and third LVK observing runs, selected according to their respective observing durations [92]. Sources are drawn from a fixed population using the Power Law + Peak mass model and Power Law redshift model, with parameters chosen to match the distributions inferred in Ref. [11]. The spin magnitudes are drawn uniformly over [0,1)01[0,1)[ 0 , 1 ) and their directions are drawn isotropically. The remaining parameters are drawn from the default uniform parameter-estimation priors [5]. We approximate the detection criterion by selecting signals with a network matched-filter signal-to-noise ratio >9absent9>9> 9 [108], which leaves 1.4×106absent1.4superscript106\approx 1.4\times 10^{6}≈ 1.4 × 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT found signals, compared to 4×104absent4superscript104\approx 4\times 10^{4}≈ 4 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT in the public sensitvity estimates [86], thereby significantly reducing the Monte Carlo uncertainty when estimating selection effects.

In Fig. 6, we show the inferred joint merger rate density over spin magnitudes and tilts, evaluated at a fixed redshift z=0.2𝑧0.2z=0.2italic_z = 0.2. There is a preference for more slowly spinning BHs aligned with the binary orbital angular momentum, as noted in our analyses of the effective spin χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT distribution. However, there are large uncertainties in the spin distributions and they are also consistent with being uniform. Information about spins in the binary BH population mostly comes from effective spin degrees of freedom [109], such as χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and the effective precessing spin parameter χpsubscript𝜒p\chi_{\mathrm{p}}italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT [110, 111]. As these represent all six spin degrees of freedom (the two BH spin vectors) with reduced dimensionality, if there is no additional information in the data beyond the effective spins then any component-spin distributions that are consistent with the inferred effective-spin distributions are allowed.

Refer to caption
Figure 6: Comoving merger rate density (a,cosθ;z)𝑎𝜃𝑧\mathcal{R}(a,\cos\theta;z)caligraphic_R ( italic_a , roman_cos italic_θ ; italic_z ) as a function of BH spin magnitude a𝑎aitalic_a and polar misalignment angle θ𝜃\thetaitalic_θ between the BH spin vector and binary orbital angular momentum, evaluated at a fixed redshift of z=0.2𝑧0.2z=0.2italic_z = 0.2. The central panel displays the joint two-dimensional posterior median. The merger rate density evaluated at cosθ=1𝜃1\cos\theta=1roman_cos italic_θ = 1 (blue) and cosθ=1𝜃1\cos\theta=-1roman_cos italic_θ = - 1 (orange) is shown as a function of a𝑎aitalic_a in the upper panel, and evaluated at a=0.3𝑎0.3a=0.3italic_a = 0.3 (blue) and a=0.75𝑎0.75a=0.75italic_a = 0.75 (orange) as a function of cosθ𝜃\cos\thetaroman_cos italic_θ in the right-hand panel. For the one-dimensional slices the bold line gives the posterior median while the shaded region encloses the 90% symmetric credible region.

We illustrate this in Fig. 7 by plotting the marginal spin magnitude and tilt distributions inferred by PixelPop and the implied marginal distributions on the effective spins χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and χpsubscript𝜒p\chi_{\mathrm{p}}italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT, compared to results of the Default spin parametric model [11]. Despite PixelPop having much larger posterior uncertainties compared to the parametric model, particular in the distribution of cosθ𝜃\cos\thetaroman_cos italic_θ, the uncertainties in the effective spin distributions, particular χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT, are much smaller and more similar to the parametric analysis.

Refer to caption
Figure 7: Marginal distributions of component and effective BH spins inferred with GWTC-3. BH spin magnitudes a𝑎aitalic_a and BH spin–orbit misalignment angles θ𝜃\thetaitalic_θ are shown in the top and bottom left panels, respectively. Effective aligned and precessing spins, χeffsubscript𝜒eff\chi_{\mathrm{eff}}italic_χ start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT and χpsubscript𝜒p\chi_{\mathrm{p}}italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT, are shown in the top and bottom right panels, respectively. The result of our nonparametric PixelPop analysis are shown in black and are compared to the parametric Default spin analysis of Ref. [11]. The bold lines denote posterior medians while the shaded regions enclose the 90% symmetric credible regions.