Detecting Gravitational Wave Anisotropies from Supermassive Black Hole Binaries

Anna-Malin Lemke \XeTeXLinkBox anna-malin.lemke@desy.de II. Institute of Theoretical Physics, Universität Hamburg, Luruper Chaussee 149, 22761, Hamburg, Germany    Andrea Mitridate \XeTeXLinkBox andrea.mitridate@desy.de Deutsches Elektronen-Synchrotron DESY, Notkestr. 85, 22607 Hamburg, Germany    Kyle A. Gersbach \XeTeXLinkBox kyle.gersbach@nanograv.org Department of Physics and Astronomy, Vanderbilt University, 2301 Vanderbilt Place, Nashville, Tennessee 37235, USA
Abstract

Several Pulsar Timing Array (PTA) collaborations have recently found evidence for a gravitational wave background (GWB) permeating our galaxy. The origin of this background is still unknown. Indeed, while the gravitational wave emission from inspiraling supermassive black hole binaries (SMBHB) is the primary candidate for its origin, several cosmological sources have also been proposed. One distinctive feature of SMBHB-generated backgrounds is the presence of GWB anisotropies stemming from the binaries distribution in the local Universe. However, none of the anisotropy searches performed to date reported a detection. In this work, we show that the lack of anisotropy detection is not currently in tension with a SMBHB origin of the background. We accomplish this by calculating the anisotropy detection probability of present and future PTAs. We find that a PTA with the noise characteristics of the NANOGrav 15-year data set had only a 2%11%percent2percent11~{}2\%-11\%2 % - 11 % probability of detecting SMBHB-generated anisotropies, depending on the properties of the SMBHB population. However, we estimate that for the IPTA DR3 data set these probabilities will increase to 4%28%percent4percent28~{}4\%-28\%4 % - 28 %, putting more pressure on the SMBHB interpretation in case of a null detection. We also identify SMBHB populations that are more likely to produce detectable levels of anisotropies. This information could be used together with the spectral properties of the GWB to characterize the SMBHB population.

preprint: DESY-24-097

I Introduction

Several Pulsar Timing Array (PTA) collaborations have recently found strong evidence for the presence of a gravitational wave background (GWB) in the nHz frequency band [1, 2, 3]. The most plausible source for such a background is a population of supermassive black hole binaries (SMBHB) formed in galaxy mergers through cosmic history. However, cosmological sources (including cosmic inflation, scalar-induced GW, phase transitions, cosmic strings, and domain walls) have also been proposed as a possible source of the background (see, for example, Ref. [4]). Discriminating between these two possible interpretations is one of the main near-term goals of PTA experiments.

Searches for anisotropies in the GWB provide a promising path to establishing the origin of the GWB. Indeed, if the GWB is produced by a population of inspiraling SMBHB, clustering of host galaxies and Poisson fluctuations in binaries properties are expected to induce a detectable level of anisotropies in the GWB [5]. On the other hand, the level of anisotropies produced by most cosmological sources is well below the sensitivity of present and future PTAs [6, 7]. Therefore, detecting anisotropies in the GWB would provide strong evidence in favor of an astrophysical origin of the signal. At the same time, we might use the lack of GWB anisotropies to constrain the properties of the SMBHB population and eventually rule out the astrophysical interpretation in favor of a cosmological one.

The recent search for anisotropies in NANOGrav 15-year data set has reported null evidence, and placed an upper limit on the level of broadband anisotropies [8]. In view of these results a few questions arise: Is the lack of evidence for GWB anisotropies in tension with an SMBHB origin of the background? What can we learn about the SMBHB population given this null detection? Are the properties of the SMBHB population needed to reproduce the intensity of the observed signal compatible with a null anisotropy detection?

In this work, we address these questions by estimating the anisotropy detection probability for GWBs sourced by SMBHB populations whose demographic and evolution are consistent with the background spectral properties. We do this by running realistic anisotropy searches on mock PTA data in which we injected the GW signal generated from synthetic SMBHB populations produced using holodeck [9]. Specifically, we proceed as follows: for a given set of SMBHB population parameters, we generate several mock SMBHB populations and their associated PTA signal, specifically the pulsar-pair correlations that such populations would induce (see Sec. III). We then run a frequentist anisotropy search [10] on these mock pulsar-pairs correlations, and derive the anisotropy detection probability for a given set of population parameters (see Sec. II and Sec. IV for details). Finally, we discuss our results (see Sec. V), which consist in a mapping between sets of SMBHB population parameters and anisotropies detection probabilities. This mapping gives us the tools to interpret the results of present and future searches for anisotropies. Specifically, it allows us to assess if a null anisotropy detection can be considered in tension with an SMBHB origin of the background, and use the results of anisotropy searches to better characterize the properties of the SMBHB population.

II How to Search for Anisotropies

In this section, we provide a summary of how GWB anisotropies appear in PTA data and discuss some of the data analysis techniques commonly adopted to extract these signals.

II.1 Anisotropies and pulsar-pair correlations

PTAs track the arrival times of radio pulses emitted by a collection of galactic millisecond pulsars. The presence of a GWB will appear in PTA data as correlated perturbations, δt𝛿𝑡\delta titalic_δ italic_t, in the arrival times of these pulses. Specifically, let’s consider a GWB whose plane-wave expansion is given by

hij(t,𝒙)=A𝑑fS2𝑑Ω^hA(f,Ω^)ei2πf(tΩ^𝒙)eijA(Ω^),subscript𝑖𝑗𝑡𝒙subscript𝐴differential-d𝑓subscriptsuperscript𝑆2differential-d^Ωsubscript𝐴𝑓^Ωsuperscript𝑒𝑖2𝜋𝑓𝑡^Ω𝒙superscriptsubscript𝑒𝑖𝑗𝐴^Ωh_{ij}(t,\bm{x})=\sum_{A}\int df\int_{S^{2}}d\hat{\Omega}\;h_{A}(f,\hat{\Omega% })e^{i2\pi f(t-\hat{\Omega}\cdot\bm{x})}e_{ij}^{A}(\hat{\Omega})\,,italic_h start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( italic_t , bold_italic_x ) = ∑ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ∫ italic_d italic_f ∫ start_POSTSUBSCRIPT italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_d over^ start_ARG roman_Ω end_ARG italic_h start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_f , over^ start_ARG roman_Ω end_ARG ) italic_e start_POSTSUPERSCRIPT italic_i 2 italic_π italic_f ( italic_t - over^ start_ARG roman_Ω end_ARG ⋅ bold_italic_x ) end_POSTSUPERSCRIPT italic_e start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( over^ start_ARG roman_Ω end_ARG ) , (1)

where f𝑓fitalic_f is the GW frequency, A=+,×𝐴A=+,\timesitalic_A = + , × labels the two GW polarizations, eijAsuperscriptsubscript𝑒𝑖𝑗𝐴e_{ij}^{A}italic_e start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT are the GW polarization tensors, and Ω^^Ω\hat{\Omega}over^ start_ARG roman_Ω end_ARG identifies the propagation direction for a single GW plane wave. For a stationary, Gaussian, and unpolarized background the polarization amplitudes, hA(f,Ω^)subscript𝐴𝑓^Ωh_{A}(f,\hat{\Omega})italic_h start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_f , over^ start_ARG roman_Ω end_ARG ), are complex random variables with vanishing expectation values and a two-point correlator given by:

hA(f,Ω^)hA(f,Ω^)=116πδAAδ(ff)δ(Ω^,Ω^)Sh(Ω^,f).delimited-⟨⟩superscriptsubscript𝐴𝑓^Ωsubscriptsuperscript𝐴superscript𝑓superscript^Ω116𝜋subscript𝛿𝐴superscript𝐴𝛿𝑓superscript𝑓𝛿^Ωsuperscript^Ωsubscript𝑆^Ω𝑓\langle h_{A}^{*}(f,\hat{\Omega})h_{A^{\prime}}(f^{\prime},\hat{\Omega}^{% \prime})\rangle=\frac{1}{16\pi}\delta_{AA^{\prime}}\delta(f-f^{\prime})\delta(% \hat{\Omega},\hat{\Omega}^{\prime})S_{h}(\hat{\Omega},f)\,.⟨ italic_h start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_f , over^ start_ARG roman_Ω end_ARG ) italic_h start_POSTSUBSCRIPT italic_A start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ( italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , over^ start_ARG roman_Ω end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ⟩ = divide start_ARG 1 end_ARG start_ARG 16 italic_π end_ARG italic_δ start_POSTSUBSCRIPT italic_A italic_A start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_δ ( italic_f - italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_δ ( over^ start_ARG roman_Ω end_ARG , over^ start_ARG roman_Ω end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( over^ start_ARG roman_Ω end_ARG , italic_f ) . (2)

The GWB power spectrum can be factorized as Sh(Ω^,f)=Sh(f)P(Ω^,f)subscript𝑆^Ω𝑓subscript𝑆𝑓𝑃^Ω𝑓S_{h}(\hat{\Omega},f)=S_{h}(f)P(\hat{\Omega},f)italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( over^ start_ARG roman_Ω end_ARG , italic_f ) = italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_f ) italic_P ( over^ start_ARG roman_Ω end_ARG , italic_f ), where the function Sh(f)subscript𝑆𝑓S_{h}(f)italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_f ) describes the spectral content of the GWB, and P(Ω^,f)𝑃^Ω𝑓P(\hat{\Omega},f)italic_P ( over^ start_ARG roman_Ω end_ARG , italic_f ) describes the distribution of the GWB power in the sky.

The expected correlations between the perturbations in the pulse time of arrivals, usually referred to as timing residuals, for pulsar a𝑎aitalic_a at time tisubscript𝑡𝑖t_{i}italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and the timing residials for pulsar b𝑏bitalic_b at time tjsubscript𝑡𝑗t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT induced by such a GWB are given by [11]:

δta(ti)δtb(tj)=df24π2f2e2πif(titj)Sh(f)Γab(f).delimited-⟨⟩𝛿subscript𝑡𝑎subscript𝑡𝑖𝛿subscript𝑡𝑏subscript𝑡𝑗𝑑𝑓24superscript𝜋2superscript𝑓2superscript𝑒2𝜋𝑖𝑓subscript𝑡𝑖subscript𝑡𝑗subscript𝑆𝑓subscriptΓ𝑎𝑏𝑓\langle\delta t_{a}(t_{i})\delta t_{b}(t_{j})\rangle=\int\frac{df}{24\pi^{2}f^% {2}}\,e^{2\pi if(t_{i}-t_{j})}\,S_{h}(f)\,\Gamma_{ab}(f)\,.⟨ italic_δ italic_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_δ italic_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) ⟩ = ∫ divide start_ARG italic_d italic_f end_ARG start_ARG 24 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT 2 italic_π italic_i italic_f ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_f ) roman_Γ start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_f ) . (3)

From this equation is clear that Sh(f)subscript𝑆𝑓S_{h}(f)italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_f ) encodes the time correlation of the residuals. The spatial correlations of the residuals are instead encoded in the overlap reduction function, ΓabsubscriptΓ𝑎𝑏\Gamma_{ab}roman_Γ start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT, which is related to the sky-distribution of the GWB power as:111Here we are neglecting the pulsar term contribution to the overlap reduction function. While this is the standard assumption in all anisotropy searches, an in-depth analysis of how this assumption affects anisotropy searches will be presented in a follow-up work [12].

Γab(f)=32S2dΩ^4πP(Ω^,f)A=+,×A(p^a,Ω^)A(p^b,Ω^)subscriptΓ𝑎𝑏𝑓32subscriptsuperscript𝑆2𝑑^Ω4𝜋𝑃^Ω𝑓subscript𝐴superscript𝐴subscript^𝑝𝑎^Ωsuperscript𝐴subscript^𝑝𝑏^Ω\Gamma_{ab}(f)=\frac{3}{2}\int_{S^{2}}\frac{d\hat{\Omega}}{4\pi}\,P(\hat{% \Omega},f)\sum_{A=+,\times}\mathcal{F}^{A}(\hat{p}_{a},\hat{\Omega})\mathcal{F% }^{A}(\hat{p}_{b},\hat{\Omega})roman_Γ start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_f ) = divide start_ARG 3 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT divide start_ARG italic_d over^ start_ARG roman_Ω end_ARG end_ARG start_ARG 4 italic_π end_ARG italic_P ( over^ start_ARG roman_Ω end_ARG , italic_f ) ∑ start_POSTSUBSCRIPT italic_A = + , × end_POSTSUBSCRIPT caligraphic_F start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT , over^ start_ARG roman_Ω end_ARG ) caligraphic_F start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , over^ start_ARG roman_Ω end_ARG ) (4)

where A(p^,Ω^)superscript𝐴^𝑝^Ω\mathcal{F}^{A}(\hat{p},\hat{\Omega})caligraphic_F start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( over^ start_ARG italic_p end_ARG , over^ start_ARG roman_Ω end_ARG ) is the antenna response function for the A𝐴Aitalic_A-th GW polarization:

A(p^,Ω^)=p^ip^j2(1Ω^p^)eijA(Ω^).superscript𝐴^𝑝^Ωsuperscript^𝑝𝑖superscript^𝑝𝑗21^Ω^𝑝superscriptsubscript𝑒𝑖𝑗𝐴^Ω\mathcal{F}^{A}(\hat{p},\hat{\Omega})=\frac{\hat{p}^{i}\hat{p}^{j}}{2(1-\hat{% \Omega}\cdot\hat{p})}e_{ij}^{A}(\hat{\Omega})\,.caligraphic_F start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( over^ start_ARG italic_p end_ARG , over^ start_ARG roman_Ω end_ARG ) = divide start_ARG over^ start_ARG italic_p end_ARG start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT over^ start_ARG italic_p end_ARG start_POSTSUPERSCRIPT italic_j end_POSTSUPERSCRIPT end_ARG start_ARG 2 ( 1 - over^ start_ARG roman_Ω end_ARG ⋅ over^ start_ARG italic_p end_ARG ) end_ARG italic_e start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( over^ start_ARG roman_Ω end_ARG ) . (5)

Working with a discrete set of frequencies, indexed by n𝑛nitalic_n, and rewriting the integral in Eq. (4) as a sum over equal-areal pixels, indexed by k𝑘kitalic_k, we get:

Γab,n=32kPn,k[a,k+b,k++a,k×b,k×]ΔΩ^k4π.subscriptΓ𝑎𝑏𝑛32subscript𝑘subscript𝑃𝑛𝑘delimited-[]superscriptsubscript𝑎𝑘superscriptsubscript𝑏𝑘superscriptsubscript𝑎𝑘superscriptsubscript𝑏𝑘Δsubscript^Ω𝑘4𝜋\Gamma_{ab,n}=\frac{3}{2}\sum_{k}P_{n,k}\left[\mathcal{F}_{a,k}^{+}\mathcal{F}% _{b,k}^{+}+\mathcal{F}_{a,k}^{\times}\mathcal{F}_{b,k}^{\times}\right]\frac{% \Delta\hat{\Omega}_{k}}{4\pi}\,.roman_Γ start_POSTSUBSCRIPT italic_a italic_b , italic_n end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 2 end_ARG ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT [ caligraphic_F start_POSTSUBSCRIPT italic_a , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT caligraphic_F start_POSTSUBSCRIPT italic_b , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + caligraphic_F start_POSTSUBSCRIPT italic_a , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT × end_POSTSUPERSCRIPT caligraphic_F start_POSTSUBSCRIPT italic_b , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT × end_POSTSUPERSCRIPT ] divide start_ARG roman_Δ over^ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π end_ARG . (6)

In this work, we will normalize the GWB power such that 𝑑Ω^P(Ω^,fn)=kPn,kΔΩ^k=4πdifferential-d^Ω𝑃^Ωsubscript𝑓𝑛subscript𝑘subscript𝑃𝑛𝑘Δsubscript^Ω𝑘4𝜋\int d\hat{\Omega}\,P(\hat{\Omega},f_{n})=\sum_{k}P_{n,k}\Delta\hat{\Omega}_{k% }=4\pi∫ italic_d over^ start_ARG roman_Ω end_ARG italic_P ( over^ start_ARG roman_Ω end_ARG , italic_f start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT roman_Δ over^ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 4 italic_π. With this normalization, a perfectly isotropic GWB is given by Pn,k=1subscript𝑃𝑛𝑘1P_{n,k}=1italic_P start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT = 1, and when plugged into Eq. (6) it returns the familiar Hellings-Downs correlation pattern [13].

II.2 Anisotropy searches

Given a set of timing residuals, we can estimate the cross-correlation coefficients using the frequentist estimator developed in Refs. [14, 15, 16, 17]:

ρab=δ𝐭aT𝐍a1𝐒^ab𝐍b1δ𝐭bTtr[𝐍a1𝐒^ab𝐍b1𝐒^ba],subscript𝜌𝑎𝑏𝛿superscriptsubscript𝐭𝑎𝑇superscriptsubscript𝐍𝑎1subscript^𝐒𝑎𝑏superscriptsubscript𝐍𝑏1𝛿superscriptsubscript𝐭𝑏𝑇trdelimited-[]superscriptsubscript𝐍𝑎1subscript^𝐒𝑎𝑏superscriptsubscript𝐍𝑏1subscript^𝐒𝑏𝑎\rho_{ab}=\frac{\delta\mathbf{t}_{a}^{T}\mathbf{N}_{a}^{-1}\hat{\mathbf{S}}_{% ab}\mathbf{N}_{b}^{-1}\delta\mathbf{t}_{b}^{T}}{\,{\rm tr}\left[\mathbf{N}_{a}% ^{-1}\hat{\mathbf{S}}_{ab}\mathbf{N}_{b}^{-1}\hat{\mathbf{S}}_{ba}\right]},italic_ρ start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT = divide start_ARG italic_δ bold_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_N start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT bold_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_δ bold_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG roman_tr [ bold_N start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT bold_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_b italic_a end_POSTSUBSCRIPT ] end_ARG , (7)

where ρabsubscript𝜌𝑎𝑏\rho_{ab}italic_ρ start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT are the best estimates for the cross-correlation coefficients, δ𝐭a𝛿subscript𝐭𝑎\delta\mathbf{t}_{a}italic_δ bold_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT contains the timing residuals for pulsar a𝑎aitalic_a, 𝐍a=δ𝐭aδ𝐭aTsubscript𝐍𝑎delimited-⟨⟩𝛿subscript𝐭𝑎𝛿superscriptsubscript𝐭𝑎𝑇\mathbf{N}_{a}=\langle\delta\mathbf{t}_{a}\delta\mathbf{t}_{a}^{T}\ranglebold_N start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = ⟨ italic_δ bold_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_δ bold_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ⟩ is the a𝑎aitalic_a-th pulsar’s autocovariance matrix, and 𝐒^absubscript^𝐒𝑎𝑏\hat{\mathbf{S}}_{ab}over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT is the template-scaled covariance matrix. This template-scaled covariance matrix encodes information about the spectral shape of the GWB but is agnostic about its overall amplitude and cross-correlation signature. It is related to the full covariance matrix, 𝐒abδ𝐭aδ𝐭bTsubscript𝐒𝑎𝑏delimited-⟨⟩𝛿subscript𝐭𝑎𝛿superscriptsubscript𝐭𝑏𝑇\mathbf{S}_{ab}\equiv\langle\delta\mathbf{t}_{a}\delta\mathbf{t}_{b}^{T}\ranglebold_S start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ≡ ⟨ italic_δ bold_t start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT italic_δ bold_t start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ⟩, as 𝐒ab=A2Γab𝐒^absubscript𝐒𝑎𝑏superscript𝐴2subscriptΓ𝑎𝑏subscript^𝐒𝑎𝑏\mathbf{S}_{ab}=A^{2}\Gamma_{ab}\hat{\mathbf{S}}_{ab}bold_S start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT = italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Γ start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT over^ start_ARG bold_S end_ARG start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT, where A𝐴Aitalic_A is the GWB amplitude for a given spectral shape template. The values of 𝐍asubscript𝐍𝑎\mathbf{N}_{a}bold_N start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT and 𝐒^^𝐒\hat{\mathbf{S}}over^ start_ARG bold_S end_ARG are not known a priori, and need to be extracted from the data, usually using Bayesian inference techniques.

Equation (7) allows us to estimate broad-band correlations for the timing residuals. However, the GWB sky sourced by realistic SMBHB populations will not be constant across frequencies, and therefore induce correlations which are not constant in frequency space. These frequency-resolved correlations can be estimated using the recently developed per-frequency Optimal Statistic (PFOS) methods described in Ref. [18]. Assuming stationary Gaussian noise for the cross-correlation uncertainties, the likelihood function for these frequency-resolved correlations, 𝝆nsubscript𝝆𝑛\bm{\rho}_{n}bold_italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, given a GWB sky, 𝐏nsubscript𝐏𝑛\mathbf{P}_{n}bold_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, can be written as:222Generally, cosmic variance and deviations from isotropy will source off-diagonal elements in the noise covariance matrix [19]. However, since the scope of this work is to closely replicate current anisotropy searches, we ignore these off-diagonal elements following the same convention adopted in Ref. [8].

p(𝝆n|𝐏n)=exp[12(𝝆n𝐑𝐏n)𝚺n1(𝝆𝐑𝐏n)]det(2π𝚺n),𝑝conditionalsubscript𝝆𝑛subscript𝐏𝑛12subscript𝝆𝑛subscript𝐑𝐏𝑛subscriptsuperscript𝚺1𝑛𝝆subscript𝐑𝐏𝑛2𝜋subscript𝚺𝑛p(\bm{\rho}_{n}|\mathbf{P}_{n})=\frac{\exp[-\frac{1}{2}(\bm{\rho}_{n}-\mathbf{% R}\mathbf{P}_{n})\mathbf{\Sigma}^{-1}_{n}(\bm{\rho}-\mathbf{RP}_{n})]}{\sqrt{% \det(2\pi\mathbf{\Sigma}_{n})}},italic_p ( bold_italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT | bold_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) = divide start_ARG roman_exp [ - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( bold_italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - bold_RP start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( bold_italic_ρ - bold_RP start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ] end_ARG start_ARG square-root start_ARG roman_det ( 2 italic_π bold_Σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) end_ARG end_ARG , (8)

where 𝚺nsubscript𝚺𝑛\bm{\Sigma}_{n}bold_Σ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is the Ncc×Nccsubscript𝑁ccsubscript𝑁ccN_{\text{cc}}\times N_{\text{cc}}italic_N start_POSTSUBSCRIPT cc end_POSTSUBSCRIPT × italic_N start_POSTSUBSCRIPT cc end_POSTSUBSCRIPT diagonal noise covariance matrix for the cross-correlation measurements in the n𝑛nitalic_n-th frequency bin, and we have defined the PTA overlap response matrix

Rab,k=32Npix[a,k+b,k++a,k×a,k×].subscript𝑅𝑎𝑏𝑘32subscript𝑁𝑝𝑖𝑥delimited-[]superscriptsubscript𝑎𝑘superscriptsubscript𝑏𝑘superscriptsubscript𝑎𝑘superscriptsubscript𝑎𝑘R_{ab,k}=\frac{3}{2N_{pix}}[\mathcal{F}_{a,k}^{+}\mathcal{F}_{b,k}^{+}+% \mathcal{F}_{a,k}^{\times}\mathcal{F}_{a,k}^{\times}].italic_R start_POSTSUBSCRIPT italic_a italic_b , italic_k end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 2 italic_N start_POSTSUBSCRIPT italic_p italic_i italic_x end_POSTSUBSCRIPT end_ARG [ caligraphic_F start_POSTSUBSCRIPT italic_a , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT caligraphic_F start_POSTSUBSCRIPT italic_b , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + caligraphic_F start_POSTSUBSCRIPT italic_a , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT × end_POSTSUPERSCRIPT caligraphic_F start_POSTSUBSCRIPT italic_a , italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT × end_POSTSUPERSCRIPT ] . (9)

Therefore, given a set of measured cross-correlations, we can maximize this likelihood to obtain an estimate of the GWB sky, 𝐏^nsubscript^𝐏𝑛\hat{\mathbf{P}}_{n}over^ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. In this work, we choose to work in the pixel basis [20], in which the GWB sky is divided into independent equal-area pixels. In the pixel basis, reconstructing the GWB sky corresponds to reconstructing the GWB power in each of these equal area pixels. Other typical basis choices include decomposition of the sky into linear [21, 22, 23, 24] and square root spherical harmonics [25] or eigenmaps of the PTA Fisher matrix [26].

In the rest of this section, we discuss the three techniques used in this work to maximize the likelihood function in Eq. (8). Examples of sky maps reconstructed by using some of these techniques are shown in Fig. 1.

II.2.1 Analytical Solution

In the pixel basis, there exists an analytical maximum solution for the likelihood in Eq. (8), given by [27, 28, 29]:

𝐏^n=𝐌n1𝐗n,subscript^𝐏𝑛superscriptsubscript𝐌𝑛1subscript𝐗𝑛\mathbf{\hat{P}}_{n}=\mathbf{M}_{n}^{-1}\mathbf{X}_{n},over^ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = bold_M start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_X start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT , (10)

where we have defined the Fisher information matrix, 𝐌n=𝐑T𝚺n1𝐑subscript𝐌𝑛superscript𝐑𝑇subscriptsuperscript𝚺1𝑛𝐑\mathbf{M}_{n}=\mathbf{R}^{T}\mathbf{\Sigma}^{-1}_{n}\mathbf{R}bold_M start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = bold_R start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT bold_R, and the “dirty map”, 𝐗n=𝐑T𝚺n1𝝆nsubscript𝐗𝑛superscript𝐑𝑇subscriptsuperscript𝚺1𝑛subscript𝝆𝑛\mathbf{X}_{n}=\mathbf{R}^{T}\mathbf{\Sigma}^{-1}_{n}\bm{\rho}_{n}bold_X start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = bold_R start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT bold_italic_ρ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. The diagonal elements of the Fisher matrix provide an estimate of the uncertainty on the recovered pixel values, while the dirty map provides an inverse noise-weighted representation of the power as seen by the PTA.

The inversion of the Fisher matrix typically limits the resolution of our maximum likelihood sky maps beyond the fundamental PTA limit. The spatial resolution of a PTA is approximately set by requiring NpixNccsubscript𝑁pixsubscript𝑁ccN_{\text{pix}}\leq N_{\text{cc}}italic_N start_POSTSUBSCRIPT pix end_POSTSUBSCRIPT ≤ italic_N start_POSTSUBSCRIPT cc end_POSTSUBSCRIPT [28], where Nccsubscript𝑁𝑐𝑐N_{cc}italic_N start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT is the number of cross-correlations in the data set, and Npixsubscript𝑁pixN_{\rm pix}italic_N start_POSTSUBSCRIPT roman_pix end_POSTSUBSCRIPT is the number of pixels in the sky-map tessellation. For a HEALpix [30] sky map, Npixsubscript𝑁pixN_{\rm pix}italic_N start_POSTSUBSCRIPT roman_pix end_POSTSUBSCRIPT is related to the Nsidesubscript𝑁sideN_{\text{side}}italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT parameter as Npix=12Nsidesubscript𝑁pix12subscript𝑁sideN_{\text{pix}}=12\cdot N_{\text{side}}italic_N start_POSTSUBSCRIPT pix end_POSTSUBSCRIPT = 12 ⋅ italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT. Given the NANOGrav 15yr configuration with 67 pulsars and Ncc=2211subscript𝑁cc2211N_{\text{cc}}=2211italic_N start_POSTSUBSCRIPT cc end_POSTSUBSCRIPT = 2211, this translates to Nside8subscript𝑁side8N_{\text{side}}\leq 8italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT ≤ 8. However, we find the inverse of the Fisher matrix to be numerically ill-defined for Nside=8subscript𝑁side8N_{\text{side}}=8italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT = 8, which limits the resolution of the NANOGrav 15-year data set to Nside4subscript𝑁side4N_{\textrm{side}}\leq 4italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT ≤ 4, corresponding to 12 (Nside=1)N_{\textrm{side}}=1)italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT = 1 ), 48 (Nside=2subscript𝑁side2N_{\text{side}}=2italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT = 2) and 192 (Nside=4subscript𝑁side4N_{\text{side}}=4italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT = 4) pixels for the sky maps.

Another downside of this map-reconstruction method is that it usually produces maps with negative power in some regions of the sky. These negative-power regions are –of course– unphysical and prevent us from normalizing the recovered maps.

II.2.2 Radiometer Search

If we assume that most of the GW power is coming from a single bright pixel, instead of trying to reconstruct the entire GW sky, we can focus on reconstructing the power in each pixel independently, neglecting correlations between neighbouring pixels. The power in each pixel can be derived by replacing the Fisher matrix with its diagonal components in Eq. (10).

This reconstruction method, usually known as radiometer [31, 32], is best suited for GWB skies that contain a single, very bright hot spot instead of a collection of sources with similar signal amplitudes. While some SMBHB populations might generate GW skies with very bright hot spots, it is not clear a priori if a radiometer search is best suited to recover generic SMBHB skies. As part of our analysis, we will assess how well the radiometer search performs on realistic SMBHB skies.

II.2.3 Numerical Solution

Finally, we can maximize the cross-correlation likelihood numerically. Specifically, we can reconstruct the GW sky by numerically solving the following equation333We use the CVXPY package [33, 34] to obtain the least-squares solution to Eq. (11) while imposing positivity and the map normalization kPn,kΔΩ^k=4πsubscript𝑘subscript𝑃𝑛𝑘Δsubscript^Ω𝑘4𝜋\sum_{k}P_{n,k}\Delta\hat{\Omega}_{k}=4\pi∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT roman_Δ over^ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 4 italic_π as boundary conditions.

𝐌n𝐏^n=𝐗n.subscript𝐌𝑛subscript^𝐏𝑛subscript𝐗𝑛\mathbf{M}_{n}\mathbf{\hat{P}}_{n}=\mathbf{X}_{n}\,.bold_M start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT over^ start_ARG bold_P end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = bold_X start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT . (11)

The advantage of this method is that it allows us to impose positivity on the recovered sky maps and therefore interpret each pixel value as physical power. Additionally, this approach does not rely on our ability to calculate the inverse of 𝐌nsubscript𝐌𝑛\mathbf{M}_{n}bold_M start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT.

III Mock Data

In this section we describe the procedure we used to generate mock PTA data. While, in principle, we could produce a set of mock timing residuals for each synthetic SMBHB population and then derive the corresponding pulsar cross-correlations, we instead directly generate the cross-correlation coefficients, ρab,nsubscript𝜌𝑎𝑏𝑛\rho_{ab,n}italic_ρ start_POSTSUBSCRIPT italic_a italic_b , italic_n end_POSTSUBSCRIPT, to maximize the computational efficiency of our procedure. Specifically, we proceed by following these three main steps:

Population Synthesis (Sec. III.1) –

We explore the parameter space describing formation and evolution of SMBHBs. For each point in this space, we generate several realizations of the cosmic SMBHB population using the holodeck package [9].

Sky Creation (Sec. III.2) –

For each SMBHB population generated in the previous step, we create a GWB sky, Pn,ksubscript𝑃𝑛𝑘P_{n,k}italic_P start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT, by randomly placing the population binaries in the sky.

Correlation Creation (Sec. III.3) –

For each GWB sky, we generate theoretical pulsar-pair correlation according to Eq. (6). To simulate measurement noise, on top of these theoretical correlations, we add zero-mean Gaussian noise whose standard deviation is given by the cross-correlations uncertainty measured from the NANOGrav 15-year data set [8].

III.1 Population Synthesis with holodeck

To generate SMBHB populations, we used the same procedure as in Ref. [35], which we briefly summarize in this section. For more details, we refer the reader to Refs. [35, 9].

Model Component Parameter Fiducial value
GSMF [36] ψzsubscript𝜓𝑧\psi_{z}italic_ψ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT -0.60
mψ,zsubscript𝑚𝜓𝑧m_{\psi,z}italic_m start_POSTSUBSCRIPT italic_ψ , italic_z end_POSTSUBSCRIPT +0.11
αψ,0subscript𝛼𝜓0\alpha_{\psi,0}italic_α start_POSTSUBSCRIPT italic_ψ , 0 end_POSTSUBSCRIPT -1.21
αψ,zsubscript𝛼𝜓𝑧\alpha_{\psi,z}italic_α start_POSTSUBSCRIPT italic_ψ , italic_z end_POSTSUBSCRIPT -0.03
MBHMbulgesubscript𝑀BHsubscript𝑀bulgeM_{\text{BH}}-M_{\text{bulge}}italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT - italic_M start_POSTSUBSCRIPT bulge end_POSTSUBSCRIPT [37, 38, 39, 40, 41] αμsubscript𝛼𝜇\alpha_{\mu}italic_α start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT 1.10
f,bulgesubscript𝑓bulgef_{\star,\text{bulge}}italic_f start_POSTSUBSCRIPT ⋆ , bulge end_POSTSUBSCRIPT 0.615
dadt𝑑𝑎𝑑𝑡\dfrac{da}{dt}divide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG [42, 35] ac/pcsubscript𝑎𝑐pca_{c}/{\rm pc}italic_a start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / roman_pc 102superscript10210^{2}10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT
ainit/pcsubscript𝑎initpca_{\text{init}}/{\rm pc}italic_a start_POSTSUBSCRIPT init end_POSTSUBSCRIPT / roman_pc 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT
νoutersubscript𝜈outer\nu_{\text{outer}}italic_ν start_POSTSUBSCRIPT outer end_POSTSUBSCRIPT 2.5
Table 1: Fiducial model parameters as presented in table B1 of Ref. [35].

III.1.1 Galaxy Merger Rate

As SMBHB are formed in galaxy mergers, one of the key ingredients needed to model SMBHB populations is the rate at which such mergers take place. The comoving number density of galaxies mergers, ηgalgalsubscript𝜂galgal\eta_{\rm gal-gal}italic_η start_POSTSUBSCRIPT roman_gal - roman_gal end_POSTSUBSCRIPT, as function of the stellar mass of the primary galaxy, m1subscript𝑚1m_{1\star}italic_m start_POSTSUBSCRIPT 1 ⋆ end_POSTSUBSCRIPT, the stellar mass ratio, qm2/m1subscript𝑞subscript𝑚2subscript𝑚1q_{\star}\equiv m_{2\star}/m_{1\star}italic_q start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ≡ italic_m start_POSTSUBSCRIPT 2 ⋆ end_POSTSUBSCRIPT / italic_m start_POSTSUBSCRIPT 1 ⋆ end_POSTSUBSCRIPT, and redshift, z𝑧zitalic_z, can be written as [36]:

3ηgal-galm1qz=Ψ(m1,z)m1ln(10)P(m1,q,z)Tgal-gal(m1,q,z)tz,superscript3subscript𝜂gal-galsubscript𝑚absent1subscript𝑞𝑧Ψsubscript𝑚absent1superscript𝑧subscript𝑚absent110𝑃subscript𝑚absent1subscript𝑞superscript𝑧subscript𝑇gal-galsubscript𝑚absent1subscript𝑞superscript𝑧𝑡superscript𝑧\frac{\partial^{3}\eta_{\text{gal-gal}}}{\partial m_{\star 1}\partial q_{\star% }\partial z}=\frac{\Psi(m_{\star 1},z^{\prime})}{m_{\star 1}\ln(10)}\frac{P(m_% {\star 1},q_{\star},z^{\prime})}{T_{\text{gal-gal}}(m_{\star 1},q_{\star},z^{% \prime})}\frac{\partial t}{\partial z^{\prime}},divide start_ARG ∂ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_η start_POSTSUBSCRIPT gal-gal end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT ∂ italic_q start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ∂ italic_z end_ARG = divide start_ARG roman_Ψ ( italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT , italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT roman_ln ( 10 ) end_ARG divide start_ARG italic_P ( italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT , italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_T start_POSTSUBSCRIPT gal-gal end_POSTSUBSCRIPT ( italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT , italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_ARG divide start_ARG ∂ italic_t end_ARG start_ARG ∂ italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG , (12)

where ΨΨ\Psiroman_Ψ denotes the galaxy stellar-mass function (GSMF), P𝑃Pitalic_P the galaxy pair fraction and Tgal-galsubscript𝑇gal-galT_{\text{gal-gal}}italic_T start_POSTSUBSCRIPT gal-gal end_POSTSUBSCRIPT the galaxy merger time. Since the galaxy merger spans a finite timescale (Tgalgalsubscript𝑇galgalT_{\rm gal-gal}italic_T start_POSTSUBSCRIPT roman_gal - roman_gal end_POSTSUBSCRIPT), in Eq. (12), we distinguish between the redshift at which the galaxy pair forms (z=z[t]superscript𝑧superscript𝑧delimited-[]𝑡z^{\prime}=z^{\prime}[t]italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT [ italic_t ] at some initial time t𝑡titalic_t) and the redshift at which the system becomes a post-merger galaxy remnant (z=z[t+Tgalgal]𝑧𝑧delimited-[]𝑡subscript𝑇galgalz=z[t+T_{\rm gal-gal}]italic_z = italic_z [ italic_t + italic_T start_POSTSUBSCRIPT roman_gal - roman_gal end_POSTSUBSCRIPT ]).

The GSMF provides the differential number-density of galaxies per decade of stellar mass. Following Ref. [35], we parametrize the GSMF as a single Schechter function [43]:

Ψ(m1,z)=ln(10)Ψ0(m1Mψ)αψexp(m1Mψ),Ψsubscript𝑚absent1superscript𝑧10subscriptΨ0superscriptsubscript𝑚absent1subscript𝑀𝜓subscript𝛼𝜓subscript𝑚absent1subscript𝑀𝜓\Psi(m_{\star 1},z^{\prime})=\ln(10)\Psi_{0}\cdot\bigg{(}\frac{m_{\star 1}}{M_% {\psi}}\bigg{)}^{\alpha_{\psi}}\exp\bigg{(}-\frac{m_{\star 1}}{M_{\psi}}\bigg{% )},roman_Ψ ( italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT , italic_z start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) = roman_ln ( 10 ) roman_Ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⋅ ( divide start_ARG italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_exp ( - divide start_ARG italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT end_ARG ) , (13)

where the phenomenological parameters Ψ0subscriptΨ0\Psi_{0}roman_Ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, Mψsubscript𝑀𝜓M_{\psi}italic_M start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT and αψsubscript𝛼𝜓\alpha_{\psi}italic_α start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT follow the redshift parametrization introduced in [36]:

log10(Ψ0/Mpc3)=ψ0+ψzzlog10(Mψ/M)=mψ,0+mψ,zzαψ=1+αψ,0+αψ,zzsubscript10subscriptΨ0superscriptMpc3subscript𝜓0subscript𝜓𝑧𝑧subscript10subscript𝑀𝜓subscript𝑀direct-productsubscript𝑚𝜓0subscript𝑚𝜓𝑧𝑧subscript𝛼𝜓1subscript𝛼𝜓0subscript𝛼𝜓𝑧𝑧\begin{split}\log_{10}(\Psi_{0}/\,{\rm Mpc}^{-3})&=\psi_{0}+\psi_{z}\cdot z\\ \log_{10}(M_{\psi}/M_{\odot})&=m_{\psi,0}+m_{\psi,z}\cdot z\\ \alpha_{\psi}&=1+\alpha_{\psi,0}+\alpha_{\psi,z}\cdot z\end{split}start_ROW start_CELL roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( roman_Ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / roman_Mpc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) end_CELL start_CELL = italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_ψ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ⋅ italic_z end_CELL end_ROW start_ROW start_CELL roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) end_CELL start_CELL = italic_m start_POSTSUBSCRIPT italic_ψ , 0 end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT italic_ψ , italic_z end_POSTSUBSCRIPT ⋅ italic_z end_CELL end_ROW start_ROW start_CELL italic_α start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT end_CELL start_CELL = 1 + italic_α start_POSTSUBSCRIPT italic_ψ , 0 end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_ψ , italic_z end_POSTSUBSCRIPT ⋅ italic_z end_CELL end_ROW (14)

This parametrization introduces six free parameters for the GSMF. Of these six parameters, the mass normalization ψ0subscript𝜓0\psi_{0}italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and the characteristic mass mψ,0subscript𝑚𝜓0m_{\psi,0}italic_m start_POSTSUBSCRIPT italic_ψ , 0 end_POSTSUBSCRIPT are allowed to vary in our analysis, while the remaining four are fixed to the fiducial values provided in Table 1.

The galaxy pair fraction and merger time are kept fixed to fiducial power-law functions of the galaxy stellar mass, galaxy mass ratio, and initial redshift. For a full description of these functions we direct the reader to Ref. [35].

III.1.2 SMBH-Host Relation

By populating galaxies with supermassive black holes (SMBHs), we can relate the galaxies merger rate in Eq. (12) with the SMBHs merger rate. We populate the galaxies assuming a correspondence between the host galaxy and the SMBH at its center. Specifically, we assume that the SMBH mass, MBHsubscript𝑀BHM_{\rm BH}italic_M start_POSTSUBSCRIPT roman_BH end_POSTSUBSCRIPT, is related to the bulge mass of the host galaxy, Mbulgesubscript𝑀bulgeM_{\rm bulge}italic_M start_POSTSUBSCRIPT roman_bulge end_POSTSUBSCRIPT, according to the relation [44]:

log10(MBH/M)=μ+αμlog10(Mbulge1011M)+𝒩(0,ϵμ)subscript10subscript𝑀BHsubscript𝑀direct-product𝜇subscript𝛼𝜇subscript10subscript𝑀bulgesuperscript1011subscript𝑀direct-product𝒩0subscriptitalic-ϵ𝜇\log_{10}(M_{\text{BH}}/M_{\odot})=\mu+\alpha_{\mu}\log_{10}\bigg{(}\frac{M_{% \text{bulge}}}{10^{11}M_{\odot}}\bigg{)}+\mathcal{N}(0,\epsilon_{\mu})\,roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT BH end_POSTSUBSCRIPT / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) = italic_μ + italic_α start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT bulge end_POSTSUBSCRIPT end_ARG start_ARG 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT end_ARG ) + caligraphic_N ( 0 , italic_ϵ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ) (15)

where 𝒩(0,ϵμ)𝒩0subscriptitalic-ϵ𝜇\mathcal{N}(0,\epsilon_{\mu})caligraphic_N ( 0 , italic_ϵ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT ) is a normally distributed random scatter with zero mean and standard deviation ϵμsubscriptitalic-ϵ𝜇\epsilon_{\mu}italic_ϵ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT, and Mbulge=f,bulgemsubscript𝑀bulgesubscript𝑓bulgesubscript𝑚M_{\rm bulge}=f_{\star,{\rm bulge}}\cdot m_{\star}italic_M start_POSTSUBSCRIPT roman_bulge end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT ⋆ , roman_bulge end_POSTSUBSCRIPT ⋅ italic_m start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT is the fraction of the galaxy stellar mass contained in the bulge. In our analysis, we vary the mass normalization μ𝜇\muitalic_μ and the scattering width ϵμsubscriptitalic-ϵ𝜇\epsilon_{\mu}italic_ϵ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT, while keeping the other parameters fixed to the fiducial values reported in Table 1.

Using the MBHMbulgesubscript𝑀BHsubscript𝑀bulgeM_{\rm BH}-M_{\rm bulge}italic_M start_POSTSUBSCRIPT roman_BH end_POSTSUBSCRIPT - italic_M start_POSTSUBSCRIPT roman_bulge end_POSTSUBSCRIPT relation we can then translate the number density of galaxy mergers into a number density for SMBHB mergers, η𝜂\etaitalic_η, as:

3ηMqz=3ηgal-galm1qzm1Mqq,superscript3𝜂𝑀𝑞𝑧superscript3subscript𝜂gal-galsubscript𝑚absent1subscript𝑞𝑧subscript𝑚absent1𝑀subscript𝑞𝑞\frac{\partial^{3}\eta}{\partial M\partial q\partial z}=\frac{\partial^{3}\eta% _{\text{gal-gal}}}{\partial m_{\star 1}\partial q_{\star}\partial z}\frac{% \partial m_{\star 1}}{\partial M}\frac{\partial q_{\star}}{\partial q}\,,divide start_ARG ∂ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_η end_ARG start_ARG ∂ italic_M ∂ italic_q ∂ italic_z end_ARG = divide start_ARG ∂ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_η start_POSTSUBSCRIPT gal-gal end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT ∂ italic_q start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT ∂ italic_z end_ARG divide start_ARG ∂ italic_m start_POSTSUBSCRIPT ⋆ 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_M end_ARG divide start_ARG ∂ italic_q start_POSTSUBSCRIPT ⋆ end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_q end_ARG , (16)

where M=m1+m2𝑀subscript𝑚1subscript𝑚2M=m_{1}+m_{2}italic_M = italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is the total binary mass, qm2/m1<1𝑞subscript𝑚2subscript𝑚11q\equiv m_{2}/m_{1}<1italic_q ≡ italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < 1 is the binary mass ratio, and m1,2subscript𝑚12m_{1,2}italic_m start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT are the masses of the SMBHs in the binary system.

Binary property Range Number of bins
M/M𝑀subscript𝑀direct-productM/M_{\odot}italic_M / italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT [104,1012]superscript104superscript1012[10^{4},10^{12}][ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ] 91
q𝑞qitalic_q [103,1]superscript1031[10^{-3},1][ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT , 1 ] 81
z𝑧zitalic_z [103,10]superscript10310[10^{-3},10][ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT , 10 ] 101
f/nHz𝑓nHzf/\rm{nHz}italic_f / roman_nHz [1.85,5.93]1.855.93[1.85,5.93][ 1.85 , 5.93 ] 3
Table 2: Details of the SMBHB parameter space binning. Frequency bins are linearly spaced, while the bins for all the other dimensions are logarithmically spaced.

III.1.3 Binary Evolution

For binaries on circular orbits, the observer-frame GW frequency, f𝑓fitalic_f, is related to the binary’s rest frame orbital frequency, fpsubscript𝑓𝑝f_{p}italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, as f=2fp/(1+z)𝑓2subscript𝑓𝑝1𝑧f=2f_{p}/(1+z)italic_f = 2 italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / ( 1 + italic_z ). The orbital frequency, in turn, is uniquely related to the binary orbital separation, a𝑎aitalic_a, according to fp2=4πGM/a3superscriptsubscript𝑓𝑝24𝜋𝐺𝑀superscript𝑎3f_{p}^{2}=4\pi GM/a^{3}italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 4 italic_π italic_G italic_M / italic_a start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT.

The evolution of the binary system’s orbital separation, a𝑎aitalic_a, is driven by a combination of GW emission and interactions with the astrophysical environment. The GW contribution to the binary hardening is given by [45]:

dadt|GW=64G35c5m1m2Ma3,evaluated-at𝑑𝑎𝑑𝑡GW64superscript𝐺35superscript𝑐5subscript𝑚1subscript𝑚2𝑀superscript𝑎3\frac{da}{dt}\bigg{|}_{\text{GW}}=-\frac{64G^{3}}{5c^{5}}\frac{m_{1}m_{2}M}{a^% {3}}\,,divide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG | start_POSTSUBSCRIPT GW end_POSTSUBSCRIPT = - divide start_ARG 64 italic_G start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG 5 italic_c start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT end_ARG divide start_ARG italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_M end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG , (17)

where G𝐺Gitalic_G is the gravitational constant. We describe environmental effects using a phenomenological model which aims to capture the overall effects of more complicated explicit binary evolution models, while introducing a small number of free parameters:

dadt|phenom=Ha(aac)1νinner(1+aac)νinnerνouter.evaluated-at𝑑𝑎𝑑𝑡phenomsubscript𝐻𝑎superscript𝑎subscript𝑎𝑐1subscript𝜈innersuperscript1𝑎subscript𝑎𝑐subscript𝜈innersubscript𝜈outer\frac{da}{dt}\bigg{|}_{\text{phenom}}=H_{a}\bigg{(}\frac{a}{a_{c}}\bigg{)}^{1-% \nu_{\text{inner}}}\bigg{(}1+\frac{a}{a_{c}}\bigg{)}^{\nu_{\text{inner}}-\nu_{% \text{outer}}}.divide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG | start_POSTSUBSCRIPT phenom end_POSTSUBSCRIPT = italic_H start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ( divide start_ARG italic_a end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 1 - italic_ν start_POSTSUBSCRIPT inner end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( 1 + divide start_ARG italic_a end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_ν start_POSTSUBSCRIPT inner end_POSTSUBSCRIPT - italic_ν start_POSTSUBSCRIPT outer end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . (18)

Here, the power law indices νinnersubscript𝜈inner\nu_{\text{inner}}italic_ν start_POSTSUBSCRIPT inner end_POSTSUBSCRIPT and νoutersubscript𝜈outer\nu_{\text{outer}}italic_ν start_POSTSUBSCRIPT outer end_POSTSUBSCRIPT are introduced, dividing the hardening process into a small-separation and a large-separation regime with different power-law behaviours and a turnover at a critical separation acsubscript𝑎𝑐a_{c}italic_a start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. While νinnersubscript𝜈inner\nu_{\text{inner}}italic_ν start_POSTSUBSCRIPT inner end_POSTSUBSCRIPT is allowed to vary in our model, νoutersubscript𝜈outer\nu_{\text{outer}}italic_ν start_POSTSUBSCRIPT outer end_POSTSUBSCRIPT is fixed to a standard literature value of +2.52.5+2.5+ 2.5 [42]. The total rate of binary evolution is obtained by summing the two contributions:

dadt=dadt|GW+dadt|phenom.𝑑𝑎𝑑𝑡evaluated-at𝑑𝑎𝑑𝑡GWevaluated-at𝑑𝑎𝑑𝑡phenom\frac{da}{dt}=\frac{da}{dt}\bigg{|}_{\text{GW}}+\frac{da}{dt}\bigg{|}_{\text{% phenom}}\,.divide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG = divide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG | start_POSTSUBSCRIPT GW end_POSTSUBSCRIPT + divide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG | start_POSTSUBSCRIPT phenom end_POSTSUBSCRIPT . (19)

The normalization of the phenomenological hardening rate, Hasubscript𝐻𝑎H_{a}italic_H start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, is determined by requiring that the binary lifetime, from the initial separation ainitsubscript𝑎inita_{\rm init}italic_a start_POSTSUBSCRIPT roman_init end_POSTSUBSCRIPT to the innermost stable circular orbit aisco=6GM/c2subscript𝑎isco6𝐺𝑀superscript𝑐2a_{\rm isco}=6GM/c^{2}italic_a start_POSTSUBSCRIPT roman_isco end_POSTSUBSCRIPT = 6 italic_G italic_M / italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, matches a target value given by:

τf=ainitaisco(dadt)1𝑑a.subscript𝜏𝑓superscriptsubscriptsubscript𝑎initsubscript𝑎iscosuperscript𝑑𝑎𝑑𝑡1differential-d𝑎\tau_{f}=\int_{a_{\text{init}}}^{a_{\text{isco}}}\bigg{(}\frac{da}{dt}\bigg{)}% ^{-1}da\,.italic_τ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = ∫ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT init end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT isco end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ( divide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_d italic_a . (20)

The value of τfsubscript𝜏𝑓\tau_{f}italic_τ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT is a free parameter that we allow to vary in our analysis.

Given a model for the binary orbital evolution, we can write the number of SMBHB per given mass, mass-ratio, redshift, and log orbital frequency, as [46, 47, 48]:

4NMqzlnfp=3ηMqzτ(fp)4πc(1+z)dc2,superscript4𝑁𝑀𝑞𝑧subscript𝑓𝑝superscript3𝜂𝑀𝑞𝑧𝜏subscript𝑓𝑝4𝜋𝑐1𝑧superscriptsubscript𝑑𝑐2\frac{\partial^{4}N}{\partial M\partial q\partial z\partial\ln f_{p}}=\frac{% \partial^{3}\eta}{\partial M\partial q\partial z}\cdot\tau(f_{p})\cdot 4\pi c(% 1+z)d_{c}^{2}\,,divide start_ARG ∂ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_N end_ARG start_ARG ∂ italic_M ∂ italic_q ∂ italic_z ∂ roman_ln italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG = divide start_ARG ∂ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_η end_ARG start_ARG ∂ italic_M ∂ italic_q ∂ italic_z end_ARG ⋅ italic_τ ( italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) ⋅ 4 italic_π italic_c ( 1 + italic_z ) italic_d start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (21)

where dcsubscript𝑑𝑐d_{c}italic_d start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is the comoving distance of a source at redshift z𝑧zitalic_z, and τ(fp)𝜏subscript𝑓𝑝\tau(f_{p})italic_τ ( italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) is the binary hardening timescale, which characterizes the amount of time a SMBHB spends in a given frequency bin:

τ(fp)fp/(dfp/dt)=2a3dadt.𝜏subscript𝑓𝑝subscript𝑓𝑝𝑑subscript𝑓𝑝𝑑𝑡2𝑎3𝑑𝑎𝑑𝑡\tau(f_{p})\equiv f_{p}/(df_{p}/dt)=-\frac{2a}{3}\frac{da}{dt}\,.italic_τ ( italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) ≡ italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / ( italic_d italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_d italic_t ) = - divide start_ARG 2 italic_a end_ARG start_ARG 3 end_ARG divide start_ARG italic_d italic_a end_ARG start_ARG italic_d italic_t end_ARG . (22)

III.1.4 Population Synthesis

Given a set of astrophysical parameters, 𝜽={ψ0,mψ,0,μ,ϵμ,τf,νinner}𝜽subscript𝜓0subscript𝑚𝜓0𝜇subscriptitalic-ϵ𝜇subscript𝜏𝑓subscript𝜈inner\bm{\theta}=\{\psi_{0},\,m_{\psi,0},\,\mu,\,\epsilon_{\mu},\,\tau_{f},\,\nu_{% \rm inner}\}bold_italic_θ = { italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_m start_POSTSUBSCRIPT italic_ψ , 0 end_POSTSUBSCRIPT , italic_μ , italic_ϵ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT , italic_τ start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , italic_ν start_POSTSUBSCRIPT roman_inner end_POSTSUBSCRIPT }, we generate random SMBHB populations by drawing the number of binaries in a given (M,q,z,f)𝑀𝑞𝑧𝑓(M,\,q,\,z,\,f)( italic_M , italic_q , italic_z , italic_f ) bin from a Poisson distribution centered around the expectation value

N(M,q,z,fp)=4NMqzlnfpΔMΔqΔzΔlnfp,𝑁𝑀𝑞𝑧subscript𝑓𝑝superscript4𝑁𝑀𝑞𝑧subscript𝑓𝑝Δ𝑀Δ𝑞Δ𝑧Δsubscript𝑓𝑝N(M,q,z,f_{p})=\frac{\partial^{4}N}{\partial M\partial q\partial z\partial\ln f% _{p}}\Delta M\,\Delta q\,\Delta z\,\Delta\ln f_{p},italic_N ( italic_M , italic_q , italic_z , italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) = divide start_ARG ∂ start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT italic_N end_ARG start_ARG ∂ italic_M ∂ italic_q ∂ italic_z ∂ roman_ln italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG roman_Δ italic_M roman_Δ italic_q roman_Δ italic_z roman_Δ roman_ln italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT , (23)

where (ΔM,Δq,Δz,Δfp)Δ𝑀Δ𝑞Δ𝑧Δsubscript𝑓𝑝(\Delta M,\,\Delta q,\,\Delta z,\,\Delta f_{p})( roman_Δ italic_M , roman_Δ italic_q , roman_Δ italic_z , roman_Δ italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) are the bin sizes for each SMBHB parameter. Details about the binning of the SMBHB parameter space are reported in Tab. 2.

To ensure that the SMBHB populations considered in our analysis are consistent with the spectral properties of the observed GWB, we sample the astrophysical parameters from the posterior distributions derived by fitting the SMBHB signal to the observed GWB spectrum. Specifically, we sample from the Monte Carlo chains obtained for the Phenom+Astro model in Ref. [35]. For each 𝜽𝜽\bm{\theta}bold_italic_θ in these chains, we then generate 50 independent SMBHB populations.

III.2 GWB Sky Maps

Given a mock SMBHB population, the next step consists in generating an associated GW sky. We start by assuming that each of the SMBHB is on a circular orbit, and assigning them the M𝑀Mitalic_M, q𝑞qitalic_q, z𝑧zitalic_z, and f𝑓fitalic_f values corresponding to their bin centers. We can then derive their sky and polarization-averaged GW amplitude as [49]:

hs2(f)=325c8(G)10/3dc2(2πfp)4/3,superscriptsubscript𝑠2𝑓325superscript𝑐8superscript𝐺103superscriptsubscript𝑑𝑐2superscript2𝜋subscript𝑓𝑝43h_{s}^{2}(f)=\frac{32}{5c^{8}}\frac{(G\mathcal{M})^{10/3}}{d_{c}^{2}}(2\pi f_{% p})^{4/3}\,,italic_h start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_f ) = divide start_ARG 32 end_ARG start_ARG 5 italic_c start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT end_ARG divide start_ARG ( italic_G caligraphic_M ) start_POSTSUPERSCRIPT 10 / 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_d start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( 2 italic_π italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 4 / 3 end_POSTSUPERSCRIPT , (24)

where =Mq3/5/(1+q)6/5𝑀superscript𝑞35superscript1𝑞65\mathcal{M}=Mq^{3/5}/(1+q)^{6/5}caligraphic_M = italic_M italic_q start_POSTSUPERSCRIPT 3 / 5 end_POSTSUPERSCRIPT / ( 1 + italic_q ) start_POSTSUPERSCRIPT 6 / 5 end_POSTSUPERSCRIPT, and fp=f(1+z)/2subscript𝑓𝑝𝑓1𝑧2f_{p}=f(1+z)/2italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = italic_f ( 1 + italic_z ) / 2. This amplitude can then be translated into a characteristic strain as [50]:

hc2(f)=hs2(f)fΔf,superscriptsubscript𝑐2𝑓superscriptsubscript𝑠2𝑓𝑓Δ𝑓h_{c}^{2}(f)=h_{s}^{2}(f)\frac{f}{\Delta f}\,,italic_h start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_f ) = italic_h start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_f ) divide start_ARG italic_f end_ARG start_ARG roman_Δ italic_f end_ARG , (25)

where the frequency bin width is given by Δf=1/TobsΔ𝑓1subscript𝑇obs\Delta f=1/T_{\text{obs}}roman_Δ italic_f = 1 / italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT.

For each frequency bin, we then place the loudest 20 sources on random pixels of a sky-tessellation derived using HEALpix [30], and add their squared characteristic strain to the corresponding pixels. The remaining binaries are added to the sky map as an isotropic background (equally divided among all pixels) with a characteristic strain given by

hc,BG2(f)=M,q,z,fpN(M,q,z,fp)hc2(f),subscriptsuperscript2𝑐BG𝑓subscript𝑀𝑞𝑧subscript𝑓𝑝𝑁𝑀𝑞𝑧subscript𝑓𝑝superscriptsubscript𝑐2𝑓h^{2}_{c,\scriptscriptstyle{\rm BG}}(f)=\sum_{M,q,z,f_{p}}N(M,q,z,f_{p})h_{c}^% {2}(f)\,,italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_c , roman_BG end_POSTSUBSCRIPT ( italic_f ) = ∑ start_POSTSUBSCRIPT italic_M , italic_q , italic_z , italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_N ( italic_M , italic_q , italic_z , italic_f start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) italic_h start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_f ) , (26)

where the sum runs over all binaries expect the loudest 20.444To generate the mock sky maps we use Nside=16subscript𝑁side16N_{\text{side}}=16italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT = 16, resulting in a resolution of 12Nside2=307212superscriptsubscript𝑁side2307212\cdot N_{\text{side}}^{2}=307212 ⋅ italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 3072 pixels. We have checked that the finite resolution effects on the induced pulsar-pair correlations are negligible for Nside16subscript𝑁side16N_{\text{side}}\geq 16italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT ≥ 16. The anisotropy of the GWB is mainly captured by the loudest 10 sources in terms of spherical harmonic coefficients [51], which we have confirmed also applies for our pixel basis. The maps are then normalized such that kPn,kΔΩ^k=4πsubscript𝑘subscript𝑃𝑛𝑘Δsubscript^Ω𝑘4𝜋\sum_{k}P_{n,k}\Delta\hat{\Omega}_{k}=4\pi∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n , italic_k end_POSTSUBSCRIPT roman_Δ over^ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 4 italic_π, such that Pn=1subscript𝑃𝑛1P_{n}=1italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT = 1 describes an isotropic sky and a theoretical upper limit of Pn,k,max=4π/ΔΩ^=Npixsubscript𝑃𝑛𝑘max4𝜋Δ^Ωsubscript𝑁𝑝𝑖𝑥P_{n,k,\rm max}=4\pi/\Delta\hat{\Omega}=N_{pix}italic_P start_POSTSUBSCRIPT italic_n , italic_k , roman_max end_POSTSUBSCRIPT = 4 italic_π / roman_Δ over^ start_ARG roman_Ω end_ARG = italic_N start_POSTSUBSCRIPT italic_p italic_i italic_x end_POSTSUBSCRIPT is imposed for the power in each pixel, which corresponds to a singular GW source being present.

To marginalize over the sky locations of the loudest sources, for each SMBHB population, we generate 30 independent sky maps per frequency bin. Given that we generate 50 SMBHB populations for each set of astrophysical parameters (see the previous section for details), we produce a total of 1500 sky maps per frequency bin for each set of astrophysical parameters. Examples of mock sky maps are provided in the upper panels of Fig. 1.

Refer to caption
Figure 1: Examples of GWB power distributions induced by SMBHB populations (upper panel), and the reconstructed sky-maps obtained using the analytical (middle panel) and numerical (lower panel) reconstruction methods. The recovered pixel values are rescaled by their corresponding uncertainties derived from a null distribution of recovered sky maps from isotropic GW skies, σPsubscript𝜎𝑃\sigma_{P}italic_σ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT, indicating the significance of deviation from isotropy. The left panels illustrate the reconstruction of a GW sky with a hot spot that leads to an anisotropy detection in our analysis. The right panels show a GW sky that does not produce a detection.
Refer to caption
Figure 2: Pixel upper limits derived for fGW=3.95subscript𝑓GW3.95f_{\rm GW}=3.95italic_f start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT = 3.95 nHz and Nside=2subscript𝑁side2N_{\text{side}}=2italic_N start_POSTSUBSCRIPT side end_POSTSUBSCRIPT = 2 using the analytical (left panel) and numerical (right panel) reconstruction methods. The red stars indicate the position of the pulsars contained in the NANOGrav 15-year data set. The higher pixel values for the analytical upper limits reflect our inability to normalize the reconstructed sky maps due to the presence of negative pixel values.

III.3 Mock Correlations

The final step of the procedure consists of generating mock cross-correlations for each of the mock GWB skies. For a given sky map, Pi,ksubscript𝑃𝑖𝑘P_{i,k}italic_P start_POSTSUBSCRIPT italic_i , italic_k end_POSTSUBSCRIPT, the expected cross-correlations values are given by Eq. (6). However, any realistic measurement will not return these theoretical values due to finite experimental precision. Therefore, to simulate a realistic anisotropy search, we generate mock correlation values by sampling from a Gaussian distribution centered around the expected value given by Eq. (6), and with a standard deviation given by the cross-correlation uncertainties measured from the NANOGrav 15-year data set using the PFOS [18].

We use this procedure to generate two mock PTA data sets. The first data set consists of the cross-correlations values for the 67676767 pulsars monitored for the NANOGrav 15-year data release. For the second data set, we try to mimic the capabilities of a future IPTA and generate mock cross-correlations for the 115 pulsars currently observed by NANOGrav [52] together with EPTA+InPTA [53], PPTA [54], and MeerKat [55]. Not having a realistic noise estimate for this data set, we use the cross-correlation uncertainties derived from the NANOGrav 15-year data set for the subset of pulsar pairs already included in this data set, and draw the noise values randomly from this set of cross-correlation uncertainties for pulsar pairs not included in the current NANOGrav observations.

It is important to note that, even for a PTA with infinite resolution, cosmic variance, polarization of the GW emission, and contributions from the pulsar term will induce deviations from the expected cross-correlations in Eq. (6) (see, for example, Ref. [56]). While these deviations are below the current precision at which we can reconstruct the cross-correlations, their impact on our ability to reconstruct the underlying GWB sky has not yet been investigated. In this work, we will follow standard conventions of anisotropy searches, and ignore the impact of these effects. We leave a dedicated study of their impact to a follow-up analysis [12].

IV Detection Statistics

In the final step of our procedure, we take the mock cross-correlations generated for each SMBHB population and try to reconstruct the underlying GW sky by using the techniques discussed in Sec. II. For each of these reconstructed skies, we want to establish if they show any evidence for anisotropies. We do this by constructing an appropriate detection statistic that we then calibrate against a null distribution.

When searching for anisotropies, we assume isotropy as the null hypothesis. Therefore, we calibrate all our detection statistics against a null distribution derived by re-running our pipeline on mock cross-correlations generated from perfectly isotropic skies (instead of the SMBHB populated skies used in Sec. III). Specifically, we generate null distributions by proceeding as follows:

  • For each pulsar pair, we generate a cross-correlation value drawing from a normal distribution with a mean given by the Hellings & Downs (HD) value and a standard deviation given by the cross-correlation uncertainties measured from the NANOGrav 15-year data set using the PFOS estimator [18]. We note that, even for isotropic skies, the cosmic variance effect already discussed in the previous section will induce deviations from HD correlations. In our analysis, following standard conventions (see for example Refs. [21, 24, 8]), we ignore this effect. The impact of this assumption will be investigated in a follow-up analysis [12].

  • By using the methods discusses in Sec. II, we recover the GWB sky associated to these mock correlations.

  • We compute the value of the detection statistic on the recovered GWB sky.

  • We repeat this procedure 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT times to derive a null distribution for the detection statistic.

By comparing the detection statistic values of a reconstructed sky with this null distribution we can assess the significance of the recovered anisotropies.

In the rest of this section, we discuss three detection statistics used in this work to define anisotropy detection.

IV.1 Pixel upper limits

An intuitive detection statistic is provided by the reconstructed GWB power in each pixel. By comparing the reconstructed pixel power with the pixel-specific null distribution, we can claim that a reconstructed sky map provides a detection with a 3σsimilar-toabsent3𝜎\sim 3\sigma∼ 3 italic_σ global significance if any of the reconstructed pixel power has a p𝑝pitalic_p-value smaller than 3×103/Npix3superscript103subscript𝑁pix3\times 10^{-3}/N_{\rm pix}3 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT / italic_N start_POSTSUBSCRIPT roman_pix end_POSTSUBSCRIPT. Here, the trial factor 1/Npix1subscript𝑁pix1/N_{\rm pix}1 / italic_N start_POSTSUBSCRIPT roman_pix end_POSTSUBSCRIPT converts the local significance of each pixel to a global significance.555Here we are assuming that the recovered pixel values are independent. Indeed, for N𝑁Nitalic_N independent trials, and in the limit of small false detection probability, the probability for a false signal classification scales as Pfalse,NNPfalse,1proportional-tosubscript𝑃false𝑁𝑁subscript𝑃false1P_{\text{false},N}\propto N\cdot P_{\text{false},1}italic_P start_POSTSUBSCRIPT false , italic_N end_POSTSUBSCRIPT ∝ italic_N ⋅ italic_P start_POSTSUBSCRIPT false , 1 end_POSTSUBSCRIPT. However, recovered pixel values are generally correlated. Therefore assuming Npixsubscript𝑁pixN_{\text{pix}}italic_N start_POSTSUBSCRIPT pix end_POSTSUBSCRIPT independent trials tends to underestimate the global significance of our detections.

The pixel threshold values for maps with Nside=2subscript𝑁side2N_{\rm side}=2italic_N start_POSTSUBSCRIPT roman_side end_POSTSUBSCRIPT = 2, and for a GW frequency of fGW=3.95nHzsubscript𝑓GW3.95nHzf_{\rm GW}=3.95\,{\rm nHz}italic_f start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT = 3.95 roman_nHz are reported in Fig. 2. As expected, the lowest threshold values are found in the region of the sky with the highest density of pulsars.

IV.2 Anisotropic signal-to-noise ratio

Instead of defining a detection statistic based on the individual pixels, we can also use one that takes the full sky map into account. One such statistic is the anisotropic signal-to-noise ratio (SNR) [10], which is given by the maximum likelihood ratio between an anisotropic and an isotropic sky:

SNR=2ln[p(𝝆|𝐏^)p(𝝆|𝐏iso)],SNR2𝑝conditional𝝆^𝐏𝑝conditional𝝆subscript𝐏iso{\rm SNR}=\sqrt{2\ln\bigg{[}\frac{p(\bm{\rho}|\mathbf{\hat{P}})}{p(\bm{\rho}|% \mathbf{P}_{\text{iso}})}\bigg{]}}\,,roman_SNR = square-root start_ARG 2 roman_ln [ divide start_ARG italic_p ( bold_italic_ρ | over^ start_ARG bold_P end_ARG ) end_ARG start_ARG italic_p ( bold_italic_ρ | bold_P start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT ) end_ARG ] end_ARG , (27)

where p𝑝pitalic_p is the likelihood function given in Eq. (8), 𝐏^^𝐏\mathbf{\hat{P}}over^ start_ARG bold_P end_ARG is the recovered anisotropic sky map, and 𝐏isosubscript𝐏iso\mathbf{P}_{\text{iso}}bold_P start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT is the perfectly isotropic sky map. This detection statistics takes advantage of the full reconstructed sky, such that sources located in different regions of the sky can contribute to a classification as a detection (even if below the individual pixel thresholds discussed in the previous section).

Due to its global nature, this detection statistic is not suitable for assessing the significance of GW skies reconstructed using the radiometer approach. Indeed, the radiometer method only reconstructs the power of individual pixels while assuming that all the others do not contribute to the observed correlation pattern. Because of this, we will not use the SNR detection statistic in conjunction with the radiometer reconstruction method.

As for the pixel-based statistic, the SNR values are calibrated against null distributions to identify sky maps that give us a detection with a p𝑝pitalic_p-value p<3×103𝑝3superscript103p<3\times 10^{-3}italic_p < 3 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (corresponding to a 3σsimilar-toabsent3𝜎\sim 3\sigma∼ 3 italic_σ Gaussian-equivalent significance).

IV.3 χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-Statistic

Finally, we consider a third test statistic based on the “distance” between the reconstructed sky map and the “typical” recovered map for an underlying isotropic sky. Specifically, assuming that the uncertainties on the pulsar-pair correlations are normally distributed, the null distribution of the recovered pixel values also follows a multivariate normal distribution, 𝒩(μ,𝚺)𝒩𝜇𝚺\mathcal{N}(\mathbf{\mu},\mathbf{\Sigma})caligraphic_N ( italic_μ , bold_Σ ), where μ𝜇\mathbf{\mu}italic_μ and 𝚺𝚺\mathbf{\Sigma}bold_Σ are the mean and covariance of the recovered pixel values for the sky maps of the null distribution. We can than calculate a probability distance estimator, the Mahalanobis distance [57], for a specific sky map, 𝐏𝐏\mathbf{P}bold_P, as

D2=(𝐏μ)T𝚺1(𝐏μ).superscript𝐷2superscript𝐏𝜇𝑇superscript𝚺1𝐏𝜇D^{2}=(\mathbf{P}-\mu)^{T}\mathbf{\Sigma}^{-1}(\mathbf{P}-\mu).italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( bold_P - italic_μ ) start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_Σ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_P - italic_μ ) . (28)

As for the previous test statistics, we compare the Mahalanobis distance for any recovered map with its null distribution. In this case the null distribution can be derived analytically, it follows a χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT distribution with the number of degrees of freedom given by the number of pixels. Recovered skies that produce a Mahalanobis distance with a p𝑝pitalic_p-value smaller than p<3×103𝑝3superscript103p<3\times 10^{-3}italic_p < 3 × 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT are considered as detections.

Our application of the Mahalanobis distance as a detection statistic assumes that the null distribution is given by a multivariate normal distribution. While this is the case for both the radiometer pixel basis and the analytical map reconstruction, the positivity constraint that we impose during the numerical reconstruction changes the shape of the null distribution such that we can not use the Mahalanobis distance as an estimator for anisotropy detection in combination with the numerical approach.

V Results

The main result of our analysis consists of an estimate of the anisotropy detection probability for different combinations of SMBHB population parameters. For each set of these parameters, 𝜽𝜽\bm{\theta}bold_italic_θ, the anisotropy detection probability, p𝜽subscript𝑝𝜽p_{\bm{\theta}}italic_p start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT, is estimated as

p𝜽=N𝜽,detNtot,subscript𝑝𝜽subscript𝑁𝜽detsubscript𝑁totp_{\bm{\theta}}=\frac{N_{\bm{\theta},\rm det}}{N_{\rm tot}}\,,italic_p start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT = divide start_ARG italic_N start_POSTSUBSCRIPT bold_italic_θ , roman_det end_POSTSUBSCRIPT end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_tot end_POSTSUBSCRIPT end_ARG , (29)

where Ntot=1500subscript𝑁tot1500N_{\rm tot}=1500italic_N start_POSTSUBSCRIPT roman_tot end_POSTSUBSCRIPT = 1500 is the total number of mock GWB skies generated for each set of SMBHB population parameters, and N𝜽,detsubscript𝑁𝜽detN_{\bm{\theta},\rm det}italic_N start_POSTSUBSCRIPT bold_italic_θ , roman_det end_POSTSUBSCRIPT is the fraction of those skies that gave us an anisotropy detection.

In Fig. 3, we report the posterior distributions of the SMBHB parameters used in this work. These distributions were derived in Ref. [35] by fitting the spectrum of SMBHB-sourced GWBs against the NANOGrav 15-year data. In this figure, the individual samples are colored according to their anisotropy detection probability, p𝜽subscript𝑝𝜽p_{\bm{\theta}}italic_p start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT, for a NANOGrav-like PTA at fGW=3.95nHzsubscript𝑓GW3.95nHzf_{\rm GW}=3.95\,{\rm nHz}italic_f start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT = 3.95 roman_nHz (the most sensitive frequency bin). From this plot, we can identify some general trends in the detection probability values. The parameter that mostly impacts the detection probability is νinnersubscript𝜈inner\nu_{\rm inner}italic_ν start_POSTSUBSCRIPT roman_inner end_POSTSUBSCRIPT, with larger values of this parameter leading to larger anisotropy detection probabilities (up to p𝜽0.45similar-tosubscript𝑝𝜽0.45p_{\bm{\theta}}\sim 0.45italic_p start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT ∼ 0.45 for νinner0)\nu_{\rm inner}\sim 0)italic_ν start_POSTSUBSCRIPT roman_inner end_POSTSUBSCRIPT ∼ 0 ). This is consistent with the results of Ref. [51], where it was shown that larger νinnersubscript𝜈inner\nu_{\rm inner}italic_ν start_POSTSUBSCRIPT roman_inner end_POSTSUBSCRIPT lead to a faster evolution of the binaries at small separation and fewer sources contributing to the background in the lowest frequency bins. Averaging over the SMBHB population parameters, we find an average detection probability of p¯𝜽0.07similar-to-or-equalssubscript¯𝑝𝜽0.07\bar{p}_{\bm{\theta}}\simeq 0.07over¯ start_ARG italic_p end_ARG start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT ≃ 0.07. This suggests that the null-detection for anisotropies reported in Ref. [8] is still perfectly compatible with an SMBHB interpretation of the GWB.

Refer to caption
Figure 3: The off-diagonal panels show the 2D posterior distribution for the SMBHB population parameters derived in Ref. [35] by fitting the SMBHB signal to the observed GWB spectrum. The black contours represent the 68%, 95% and 99.7% Bayesian credible regions for these posterior distributions, while the dots show the MCMC samples from which the posterior distributions are derived. The individual MCMC samples are colored according to their estimated anisotropy detection probability, p𝜽subscript𝑝𝜽p_{\bm{\theta}}italic_p start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT. The diagonal panels show the marginalized 1D posterior distribution, the lines are colored according to the marginalized anisotropy detection probability.

The observed trends in detection probabilities suggest that anisotropy detection (or lack thereof) provides additional information about the SMBHB properties in addition to the ones already encoded in the background spectral properties. This extra information could be used in future analyses to break the degeneracy between SMBHB populations leading to similar GWB spectra. Specifically, we envision an analysis in which the information provided by a positive (or negative) anisotropy detection is included in the likelihood used in Monte Carlo explorations of the SMBHB parameter space.

As expected, we find that most of the GW skies leading to a detection present a hot spot in the region of the sky in which we have the highest density of pulsars, and consequently a higher sensitivity to GWB anisotropies. This fact is clearly illustrated by Fig. 4, which reports the average of the detected GW skies, and shows how the typical detectable sky has more power in the region of the sky that is more densely populated by pulsars included in the data set. This fact also suggests that searches for anisotropies will greatly benefit from the more uniform sky coverage that will be provided by the upcoming IPTA DR3. Estimates for anisotropy detection probabilities in an IPTA-like data set are reported in Fig. 5. We find that, for this PTA, detection probabilities can reach up to pθ0.85similar-tosubscript𝑝𝜃0.85p_{\theta}\sim 0.85italic_p start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ∼ 0.85 for νinner0similar-tosubscript𝜈inner0\nu_{\rm inner}\sim 0italic_ν start_POSTSUBSCRIPT roman_inner end_POSTSUBSCRIPT ∼ 0, while the average detection probability grows to p¯𝜽0.16similar-tosubscript¯𝑝𝜽0.16\bar{p}_{\bm{\theta}}\sim 0.16over¯ start_ARG italic_p end_ARG start_POSTSUBSCRIPT bold_italic_θ end_POSTSUBSCRIPT ∼ 0.16. This result illustrates how future anisotropy searches could provide the first evidence in support of SMBHB as the origin of the background, or –in case of a null detection– put this interpretation under pressure.

Refer to caption
Figure 4: Average of the GWB skies (at fGW=3.95Hzsubscript𝑓GW3.95Hzf_{\rm GW}=3.95\,{\rm Hz}italic_f start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT = 3.95 roman_Hz) that would lead to a detection of anisotropies in a PTA dataset with the noise properties of NANOGrav 15-year data set (using numerical map reconstruction and SNR-based detection statistic). The red stars indicate the position of the pulsars contained in the NANOGrav 15-year data set.
Refer to caption
Figure 5: Same for Fig. 3 but for a PTA containing all the pulsars currently observed by EPTA+InPTA, MeerKAT, NANOGrav, and PPTA.
Refer to caption
Figure 6: The left panel shows the average (histograms) and 68%percent6868\%68 % intervals (bars) for anisotropy detection probability (for Nside=2subscript𝑁side2N_{\rm side}=2italic_N start_POSTSUBSCRIPT roman_side end_POSTSUBSCRIPT = 2) in the first three frequency bins and for different combinations of the map reconstruction method (numerical N, analytical A, or radiometer R) and detection statistic (SNR S,pixel-based P, or χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT-based statistic χ𝜒\chiitalic_χ). The right panel shows the same quantities for the second frequency bin and different values of Nsidesubscript𝑁sideN_{\rm side}italic_N start_POSTSUBSCRIPT roman_side end_POSTSUBSCRIPT.

While in Fig. 3 and 5 we only report detection probabilities for the second frequency bin, we have also derived analogous results for the remaining lowest five frequency bins. We find that detection probabilities at other frequencies show similar trends in the SMBHB parameter space, even if with different overall normalizations. The left panel of Figure 6 reports the average detection probabilities in the first three frequency bins, and shows how detection probabilities are the largest in the second (we have also checked detection probabilities at higher frequencies, finding that sensitivity to anisotropies keeps degrading going to higher frequencies). The second frequency bin results the most likely to detect GW anisotropies from SMBHB due to the balancing of competing effects: the level of anisotropies from SMBHBs is expected to increase at higher frequencies due to decreasing numbers of binaries contributing to the signal at those frequencies [51]; however, the red-tilt of the GWB spectrum causes the cross-correlation uncertainties to strongly increase at higher frequencies, limiting the sensitivity to anisotropies to the lower frequency range. The inclusion of pulsars with a timing baseline significantly shorter than 15 years introduces additional uncertainties in the first frequency bin, which leaves the second bin as the one most sensitive to searches for anisotropies.

While in Fig. 3 and Fig. 5 we only report the results derived with a combination of numerical map reconstruction and SNR-based detection statistic, we also derived these detection probabilities using all possible combinations of sky reconstruction techniques (discussed in Sec. II) and detection statistics (discussed in Sec. IV). We find that for all these possible map reconstruction-detection statistic combinations we observe the same detection probability trends in the SMBHB parameter space, but with different overall normalization. Specifically, we find that numerical map reconstruction combined with the SNR-based detection statistics leads to the best detection prospects when coupled with a map resolution of Nside=2subscript𝑁side2N_{\rm side}=2italic_N start_POSTSUBSCRIPT roman_side end_POSTSUBSCRIPT = 2 or Nside=4subscript𝑁side4N_{\rm side}=4italic_N start_POSTSUBSCRIPT roman_side end_POSTSUBSCRIPT = 4 (see the right panel of Fig. 6). Except for the lower resolution maps (Nside=1)N_{\rm side}=1)italic_N start_POSTSUBSCRIPT roman_side end_POSTSUBSCRIPT = 1 ), the analytical reconstruction tends to underperform the numerical reconstruction methods, while the radiometer basis delivers similar results. This shows how, despite realistic SMBHB skies presenting more than a single bright pixel, the radiometer basis is still able to resolve them.

Since most of the GW skies that lead to a positive detection for anisotropies are characterized by a loud binary, it is interesting to consider the interplay with continuous wave (CW) searches. By comparing the CW upper limits provided in Ref. [58] with the characteristic strain of all binaries in our mock GW skies, we find that roughly 35%percent3535\%35 % of the anisotropy detections in the second frequency bin would be accompanied by a positive detection in CW searches at the 2σ2𝜎2\sigma2 italic_σ-level (see the left panel of Fig. 7). Similarly, we find that around 40%percent4040\%40 % of the skies with a CW-detectable binary in the second frequency bin would also lead to a positive anisotropy detection (see the right panel of Fig. 7). Of course, these results are only rough estimates based on marginalized CW upper limits. A more detailed analysis would require performing a full CW search on the SMBHB populations used to derive the mock GW skies. We leave this analysis to future studies [59].

Refer to caption
Figure 7: Left: Fraction of anisotropy detections above the CW upper limits. Right: Fraction of CW detectable skies that also led to an anisotropy detection.

Finally, we want to emphasize that in deriving our results we ignored the impact that interference between the binaries signals can have on the pulsars cross correlations [56, 19]. This effect –which is a combination of what is commonly known as pulsar variance and cosmic variance– can induce significant deviations from the expected cross-correlation values given in Eq. (4) and make anisotropy detection more challenging. The precise impact of these effects, together with possible mitigation strategies, will be discussed in future work [12]. However, because of this approximation, the detection probabilities reported herein should be regarded as optimistic estimates, contingent upon improvements in detection methods to account for such interference effects.

VI Discussion

In this work, we developed a pipeline to perform realistic anisotropy searches on GWBs produced by synthetic SMBHB populations. We then used this framework to study the prospects of anisotropy detection by current and future PTAs, under the assumption that the observed GWB is produced by a population of SMBHB. The main results of our analysis are the following:

  1. 1.

    Based on our analysis, the lack of evidence for GWB anisotropies in a PTA with the noise characteristics of the NANOGrav 15-year data set is not in tension with assuming SMBHBs as a source of the observed background. Indeed, we find that for a SMBHB-generated GWB there was only a 7%similar-toabsentpercent7\sim 7\%∼ 7 % probability of detecting anisotropies in such a data set.

  2. 2.

    We estimate that the probability of detecting GWB anisotropies in a SMBHB-generated background will increase to 16%similar-toabsentpercent16\sim 16\%∼ 16 % for the upcoming IPTA DR3 data set. This indicates that, as we keep improving the sensitivity of our observations, a null anisotropy detection will start to put pressure on an SMBHB interpretation of the background.

  3. 3.

    We find that the anisotropy detection probability strongly depends on the properties of the SMBHB population. For example, while the average detection probability is around 7%similar-toabsentpercent7\sim 7\%∼ 7 % for a PTA with the noise properties of NANOGrav 15-year data set, probabilities can reach up to 45%similar-toabsentpercent45\sim 45\%∼ 45 % for SMBHB populations where the phenomenological hardening processes speed up at smaller separations (i.e., when νinner0)\nu_{\rm inner}\sim 0)italic_ν start_POSTSUBSCRIPT roman_inner end_POSTSUBSCRIPT ∼ 0 ). This indicate that an anisotropy detection (or lack of thereof) can convey information about the SMBHB population properties, and break the degeneracy between SMBHB populations with similar GWB spectral properties.

  4. 4.

    Among the considered reconstruction methods and detection statistics, the combination that performs the best uses a numerical map reconstruction method and an SNR-based detection statistic.

  5. 5.

    We find that only 40%similar-toabsentpercent40\sim 40\%∼ 40 % of the GWB skies leading to detection of anisotropies contain a binary with a CW signal above current CW upper limits. This emphasizes the complementarity between CW and anisotropy searches in the effort to detect the first direct evidence for SMBHB being the source of the GWB.

Acknowledgments

We would like to thank Emiko C. Gardiner, Nihan Pol, Luke Zoltan Kelley, Tristan Smith, Stephen R. Taylor and the entire NANOGrav collaboration for useful discussions and comments on a preliminary form of this work. The work of AM and AL was supported by the Deutsche Forschungsgemeinschaft under Germany’s Excellence Strategy - EXC 2121 Quantum Universe - 390833306. KAG acknowledges support from an NSF CAREER #2146016. This work used the Maxwell computational resources operated at Deutsches Elektronen-Synchrotron DESY, Hamburg (Germany).

References