Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Forecasting the sensitivity of Pulsar Timing Arrays to gravitational wave backgrounds

Stanislav Babak stas@apc.in2p3.fr Astroparticule et Cosmologie, CNRS, Université Paris Cité, F-75013 Paris, France    Mikel Falxa falxa@apc.in2p3.fr LPC2E, CNRS, 3 Av. de la Recherche Scientifique, 45071 Orléans, France Astroparticule et Cosmologie, CNRS, Université Paris Cité, F-75013 Paris, France    Gabriele Franciolini gabriele.franciolini@cern.ch CERN, Theoretical Physics Department, Esplanade des Particules 1, Geneva 1211, Switzerland    Mauro Pieroni mauro.pieroni@cern.ch CERN, Theoretical Physics Department, Esplanade des Particules 1, Geneva 1211, Switzerland
Abstract

Pulsar Timing Array (PTA) observations hinted towards the existence of a stochastic gravitational wave background (SGWB) in the nHz frequency band. Still, the nature of the SGWB signal cannot be confidently inferred from current data, and the leading explanation invokes mergers of supermassive black holes. If confirmed, such discovery would not only represent a turning point in our understanding of astrophysics, but it may severely limit the capability of searching for additional cosmological sources in the nHz frequency range. In this work, we build a simple framework to forecast the sensitivity of future PTA configurations and assess the parameter estimation of SGWB, which could consist of several contributions. We release the python code fastPTA implementing this framework and ready to use.

preprint: CERN-TH-2024-039

I Introduction

Pulsar timing array (PTAs) experiments are unique probes to study the gravitational-wave (GW) spectrum in the nano-Hertz (nHz) frequency range. Recently, the results reported by the NANOGrav Agazie et al. (2023a, b), EPTA (in combination with InPTA) Antoniadis et al. (2023a, b, c), PPTA Reardon et al. (2023a); Zic et al. (2023); Reardon et al. (2023b) and CPTA Xu et al. (2023) collaborations hinted towards the existence of a common spectrum of a stochastic nature, with around (2÷4)σ24𝜎(2\div 4)\sigma( 2 ÷ 4 ) italic_σ significance for a Hellings-Down (HD) angular correlation, consistent with the quadrupolar nature of GWs in general relativity Hellings and Downs (1983).

All collaborations reported the signal could be explained by a stochastic GW background (SGWB) with a preference for blue-tilted spectrum of energy density ΩGW(f)fnTproportional-tosubscriptΩGW𝑓superscript𝑓subscript𝑛𝑇\Omega_{\rm GW}(f)\propto f^{n_{T}}roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT ( italic_f ) ∝ italic_f start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, with nT2similar-to-or-equalssubscript𝑛𝑇2n_{T}\simeq 2italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ≃ 2, and large uncertainties varying among different collaborations. The astrophysical explanation of this signal would rely on the incoherent superposition of GW signals from a population of inspiralling supermassive black hole (SMBH) binaries on circular orbits, whose frequency spectrum is characterised by a scaling law ΩGWf2/3proportional-tosubscriptΩGWsuperscript𝑓23\Omega_{\rm GW}\propto f^{2/3}roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT ∝ italic_f start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT Phinney (2001). Nevertheless, orbital eccentricity and stellar environment could affect the binary evolution which is coupled to the statistical fluctuation in the distribution of SMBHBs may play an important role and can lead to modified spectra (see, e.g., Sesana et al. (2008); Kocsis and Sesana (2011); Kelley et al. (2017); Perrodin and Sesana (2018); Ellis et al. (2023); Agazie et al. (2023c); Afzal et al. (2023); Ghoshal and Strumia (2024); Bonetti et al. (2018)). On the other hand, with current data, it is still impossible to rule out a cosmological explanation for the observed signal. The most studied scenarios capable of producing a SGWB in the nHz frequency range would be: first-order phase transitions Arzoumanian et al. (2021); Xue et al. (2021); Nakai et al. (2021); Di Bari et al. (2021); Sakharov et al. (2021); Li et al. (2021); Ashoorioon et al. (2022); Benetti et al. (2022); Barir et al. (2023); Hindmarsh and Kume (2023); Gouttenoire and Volansky (2023); Baldes et al. (2023); Li and Xie (2023); cosmic strings and domain walls Ellis and Lewicki (2021); Datta et al. (2021); Samanta and Datta (2021); Buchmuller et al. (2020); Blasi et al. (2021); Ramazanov et al. (2022); Babichev et al. (2022); Gorghetto et al. (2021); Buchmuller et al. (2021); Blanco-Pillado et al. (2021); Ferreira et al. (2023); An and Yang (2023); Qiu and Yu (2023); Zeng et al. (2023); King et al. (2024); Babichev et al. (2023); Kitajima et al. (2024); Barman et al. (2023); scalar-induced GWs generated from primordial fluctuations Vaskonen and Veermäe (2021); De Luca et al. (2021); Bhaumik and Jain (2021); Inomata et al. (2021); Kohri and Terada (2021); Domènech and Pi (2022); Namba and Suzuki (2020); Sugiyama et al. (2021); Zhou et al. (2020); Lin et al. (2023); Rezazadeh et al. (2022); Kawasaki and Nakatsuka (2021); Ahmed et al. (2022); Yi and Fei (2023); Yi (2023); Dandoy et al. (2023); Zhao et al. (2023); Ferrante et al. (2023); Cai et al. (2023); Franciolini et al. (2023); Balaji et al. (2023); Liu et al. (2024); tensor perturbations generated during inflation (see, e.g., Vagnozzi (2023)) and many others (see also Franciolini et al. (2024); Madge et al. (2023); Figueroa et al. (2023); Garcia-Bellido et al. (2024); Murai and Yin (2023); Konoplya and Zhidenko (2023); Smarra et al. (2023)).

Future observations will narrow down the range of models compatible with the data. For example, the forecasted sensitivity of future PTA experiments may improve by various orders of magnitude compared to the ones achieved by current experiments. This will allow us to better understand the nature of the dominant source of GWB in the nHz while constraining signals from new physics appearing with amplitude below current sensitivity. Crucially, the presence of the currently observed foreground will greatly limit the constraining power of future observations on subdominant SGWB. In other words, regardless of the signal origin, currently hinted SGWB will constitute an unavoidable source of foreground noise when searching for subdominant contributions.

In this work, we aim to build a simplified framework capable of reliably estimating future PTA sensitivity without the need for expensive data simulation and analyses. We will remain agnostic on the nature of the signal, which we assume to be an isotropic, stationary SGWB characterized by the maximum likelihood power-law model resulting from the analysis of the recent PTA datasets Antoniadis et al. (2023b), and we will model noises based on the currently observed pulsars. We then forecast future PTA uncertainties and describe how they will shrink both with a longer observation time and a larger number of pulsars. Finally, we will forecast the sensitivity to subdominant contribution to the SGWB, considering simple models as a proof of concept. Detailed analysis considering motivated cosmological SGWB, coming from new physics would be presented elsewhere.

The paper is organized as follows. In Sec. II, we discuss how we build the PTA noise sensitivity, following Ref. Hazboun et al. (2019) and with some additional simplifying assumptions. In Sec. III we describe the PTA data simulations we adopt to validate our proposed methodology against state-of-the-art pipelines for the PTA data analysis. In sec. IV we report the results while presenting conclusions and outlook in Sec. V.

At the repository linked in Ref. cod , we release the code fastPTA readily usable to estimate the sensitivities and the measurement uncertainties with future PTA configuration, with longer observation times and a simulated larger set of pulsars, which also features the possibility of including additional subdominant GWB contributions.

II Building the Pulsar timing array sensitivity

In this section, we describe how, following Ref. Anholm et al. (2009); Chamberlin et al. (2015); Rosado et al. (2015); Hazboun et al. (2019), we build the PTA sensitivity curves based on a few simplifying assumptions. In the following sections, we show that this framework reproduces measurement uncertainties estimated using current data (EPTA DR2, in particular, Antoniadis et al. (2023a)) and forecasted from mock data sets generated with state-of-the-art techniques.

As a starting point, we assume the combined PTA data consists of residuals of time of arrivals (TOAs) dI(t)subscript𝑑𝐼𝑡d_{I}(t)italic_d start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_t ), where the index I𝐼Iitalic_I runs up the total number of pulsars observed 1,,Np1subscript𝑁𝑝1,\dots,N_{p}1 , … , italic_N start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT. Moreover, we will assume the GW signal and noise to be stationary111 In reality, we only need stationarity in the frequency band relevant to the targeted stochastic GW signal. See conclusions for a more comprehensive discussion of this point and define the Fourier domain data as

d~ik=Tobs/2Tobs/2dI(t)ei2πfktdt.superscriptsubscript~𝑑𝑖𝑘superscriptsubscriptsubscript𝑇obs2subscript𝑇obs2subscript𝑑𝐼𝑡superscript𝑒𝑖2𝜋subscript𝑓𝑘𝑡differential-d𝑡\tilde{d}_{i}^{k}=\int_{-T_{\rm obs}/2}^{T_{\rm obs}/2}d_{I}(t)e^{-i2\pi f_{k}% t}{\rm d}t.over~ start_ARG italic_d end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT = ∫ start_POSTSUBSCRIPT - italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT / 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT / 2 end_POSTSUPERSCRIPT italic_d start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_t ) italic_e start_POSTSUPERSCRIPT - italic_i 2 italic_π italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_t end_POSTSUPERSCRIPT roman_d italic_t . (1)

Given the finite duration of the data stream, we adopt a finite set of frequencies for the Fourier basis, fk=k/Tobssubscript𝑓𝑘𝑘subscript𝑇obsf_{k}=k/T_{\rm obs}italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_k / italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT, starting from the inverse of the observation time Tobssubscript𝑇obsT_{\rm obs}italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT and going up to the Nyquist frequency. We assume that the data contains the stochastic GW signal s~Iksuperscriptsubscript~𝑠𝐼𝑘\tilde{s}_{I}^{k}over~ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT and noise n~Iksuperscriptsubscript~𝑛𝐼𝑘\tilde{n}_{I}^{k}over~ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT, as d~Ik=s~Ik+n~Iksuperscriptsubscript~𝑑𝐼𝑘superscriptsubscript~𝑠𝐼𝑘superscriptsubscript~𝑛𝐼𝑘\tilde{d}_{I}^{k}=\tilde{s}_{I}^{k}+\tilde{n}_{I}^{k}over~ start_ARG italic_d end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT = over~ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT + over~ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT, which we will further assume to be Gaussian with zero mean s~Ik=n~Ik=0delimited-⟨⟩superscriptsubscript~𝑠𝐼𝑘delimited-⟨⟩superscriptsubscript~𝑛𝐼𝑘0\langle\tilde{s}_{I}^{k}\rangle=\langle\tilde{n}_{I}^{k}\rangle=0⟨ over~ start_ARG italic_s end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ⟩ = ⟨ over~ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ⟩ = 0.

Given the aforementioned assumptions, one can write down the full covariance matrix, including contributions from both intrinsic pulsar noise, measurement process, and GWB as

CIJ=Cn,IJ+Ch,IJ.subscript𝐶𝐼𝐽subscript𝐶𝑛𝐼𝐽subscript𝐶𝐼𝐽C_{IJ}=C_{n,IJ}+C_{h,IJ}\,.italic_C start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT italic_n , italic_I italic_J end_POSTSUBSCRIPT + italic_C start_POSTSUBSCRIPT italic_h , italic_I italic_J end_POSTSUBSCRIPT . (2)

In the following, we will assume the noise contribution to the covariant matrix to be dominated by the diagonal terms in the pulsar indices, i.e., to be uncorrelated among different pulsars. This allows us to approximate Cn,IJ=δIJPn,Isubscript𝐶𝑛𝐼𝐽subscript𝛿𝐼𝐽subscript𝑃𝑛𝐼C_{n,IJ}=\delta_{IJ}P_{n,I}italic_C start_POSTSUBSCRIPT italic_n , italic_I italic_J end_POSTSUBSCRIPT = italic_δ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_n , italic_I end_POSTSUBSCRIPT. Additionally, the GW contribution presents the unique Hellings and Downs (HD) correlation pattern to be discussed in Sec. II.2.

Transmission function.

The signal is extracted from timing residuals built by subtracting the expected time of arrival computed using the timing model. Fitting for the parameters entering the timing model (i.e., describing each pulsar intrinsic rotation frequency, its derivative, proper motion, etc), results in a polynomial suppression of sensitivity at low frequencies.222Sensitivity to the GWs with frequencies below 1/Tobs1subscript𝑇obs1/T_{\text{\tiny obs}}1 / italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT can be retained in delays compared to higher order spin-down terms of the pulsar timing model (see e.g., DeRocco and Dror (2024, 2023)). Crucially, the drastic decrease only happens around the frequency 1/Tobs1subscript𝑇obs1/T_{\rm obs}1 / italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT. This suppression of signal can effectively be described by a transmission function scaling as 1/f61superscript𝑓61/f^{6}1 / italic_f start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT (for a quadratic spin-down model) below the frequency of 1/Tobs1subscript𝑇obs1/T_{\rm obs}1 / italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT Hazboun et al. (2019). We model this transfer function as

𝒯(f)[1+1/(fTobs)]6.similar-to-or-equals𝒯𝑓superscriptdelimited-[]11𝑓subscript𝑇obs6{\cal T}(f)\simeq\left[1+1/\left(fT_{\text{\tiny obs}}\right)\right]^{-6}.caligraphic_T ( italic_f ) ≃ [ 1 + 1 / ( italic_f italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT ) ] start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT . (3)

An additional loss of the sensitivity is induced around f=1/yr𝑓1yrf=1/{\rm yr}italic_f = 1 / roman_yr from fitting the sky position and the proper motion and around f=2/yr𝑓2yrf=2/{\rm yr}italic_f = 2 / roman_yr due to parallax. This generates large spikes in the sensitivity curve. Furthermore, if the pulsar is in a binary system with a period that falls within the frequency range, an additional dip in the transmission function will appear due to the fitting of the orbital parameters. As these losses typically appear at relatively high frequencies, we will neglect them in our estimates.

II.1 Noise model

Following the discussion in Ref. Hazboun et al. (2019), we isolate a few key features of the noise budget. For each pulsar, the noise contribution comes schematically from

Noise budget=WN + RN + DM + SV,Noise budgetWN + RN + DM + SV\text{Noise budget}=\text{WN + RN + DM + SV},Noise budget = WN + RN + DM + SV , (4)

defined as follows:

White noise (WN).

The time of arrival of the signal shows an overall white noise due to the finite signal-to-noise ratio (SNR) of the match filtering process adopted to extract them. It appears as a flat contribution to the power spectra of timing residuals. We characterize the white noise contribution for each pulsar as PIWNsuperscriptsubscript𝑃𝐼WNP_{I}^{\rm WN}italic_P start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_WN end_POSTSUPERSCRIPT, where I𝐼Iitalic_I is the index associated with each pulsar. This is typically controlled by the parameters EQUAD, ECORR, and EFAC in PTA analyses Antoniadis et al. (2023b); Agazie et al. (2023b). This can be described by a time delay power spectrum

PIWN=2σ2Δt,superscriptsubscript𝑃𝐼WN2superscript𝜎2Δ𝑡P_{I}^{\rm WN}=2\sigma^{2}\Delta t,italic_P start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_WN end_POSTSUPERSCRIPT = 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Δ italic_t , (5)

where ΔtΔ𝑡\Delta troman_Δ italic_t is the inverse of the observing cadence and σ𝜎\sigmaitalic_σ is the rms timing uncertainty.

Red noise (RN).

The stochasticity in the pulsar’s rotation causes an “achromatic" red noise, which we model with a power-law power spectral density (PSD). This noise is independent of the frequency of the radio observation (hence “achromatic”) and often dominates at low frequencies:

PIRN=AIRN(ffr)γIRN;superscriptsubscript𝑃𝐼RNsubscriptsuperscript𝐴RN𝐼superscript𝑓subscript𝑓𝑟subscriptsuperscript𝛾RN𝐼P_{I}^{\rm RN}=A^{\rm RN}_{I}\left(\frac{f}{f_{r}}\right)^{\gamma^{\rm RN}_{I}};italic_P start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_RN end_POSTSUPERSCRIPT = italic_A start_POSTSUPERSCRIPT roman_RN end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( divide start_ARG italic_f end_ARG start_ARG italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_γ start_POSTSUPERSCRIPT roman_RN end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ; (6)

where frsubscript𝑓𝑟f_{r}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is a reference frequency often assumed to be 1/yr, and the amplitude AIRNsubscriptsuperscript𝐴RN𝐼A^{\rm RN}_{I}italic_A start_POSTSUPERSCRIPT roman_RN end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT and the spectral index γIRNsubscriptsuperscript𝛾RN𝐼\gamma^{\rm RN}_{I}italic_γ start_POSTSUPERSCRIPT roman_RN end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT is intrinsic to each pulsar I𝐼Iitalic_I.

Chromatic noise components (DM, SV).

Other important noise sources are temporal variations in dispersion measure (DM) and scattering variations (SV). Those components are chromatic and add time delays to the ToAs as ν2proportional-toabsentsuperscript𝜈2\propto\nu^{-2}∝ italic_ν start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT (DM) and ν4proportional-toabsentsuperscript𝜈4\propto\nu^{-4}∝ italic_ν start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT (SV), where ν𝜈\nuitalic_ν is the observed radio frequency, and both are caused by the time-varying electron column density in the interstellar medium along the line of sight (see Chalumeau et al. (2021) and references therein). The PSD for both noises is usually described by power-law similar to Eq. (6) but with added chromaticity.

All in all, we can approximate the noise model as a white noise contribution with the addition of a low-frequency red power law (see e.g., Ref. Caballero et al. (2016)) of the form

Pn,I=PIWN+AIRN(ffr)γIRN+PIDM,SV,subscript𝑃𝑛𝐼superscriptsubscript𝑃𝐼WNsubscriptsuperscript𝐴RN𝐼superscript𝑓subscript𝑓𝑟superscriptsubscript𝛾𝐼RNsuperscriptsubscript𝑃𝐼DMSV\displaystyle P_{n,I}=P_{I}^{\rm WN}+A^{\rm RN}_{I}\left(\frac{f}{f_{r}}\right% )^{\gamma_{I}^{\rm RN}}+P_{I}^{\rm DM,SV},italic_P start_POSTSUBSCRIPT italic_n , italic_I end_POSTSUBSCRIPT = italic_P start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_WN end_POSTSUPERSCRIPT + italic_A start_POSTSUPERSCRIPT roman_RN end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( divide start_ARG italic_f end_ARG start_ARG italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_γ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_RN end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT + italic_P start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_DM , roman_SV end_POSTSUPERSCRIPT , (7)

where we arbitrarily fix the reference frequency at fr=fyrsubscript𝑓𝑟subscript𝑓yrf_{r}=f_{\rm yr}italic_f start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT roman_yr end_POSTSUBSCRIPT and γIRN<0superscriptsubscript𝛾𝐼RN0\gamma_{I}^{\rm RN}<0italic_γ start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_RN end_POSTSUPERSCRIPT < 0. PIDM,SVsuperscriptsubscript𝑃𝐼DMSVP_{I}^{\rm DM,\,SV}italic_P start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_DM , roman_SV end_POSTSUPERSCRIPT in the previous equation indicate the possible presence of DM and SV terms, modeled as power-laws analogously to eq. (6), only present for some of the pulsars. In this work, we have neglected the chromaticity and treated DM and SV on an equal footing as achromatic noise. The main reason is that there is still a significant coupling between chromatic and achromatic noise components in the currently observed data. We will re-discuss this point in the concluding section.

As we will see, the simplified modeling summarised above provides a sufficiently accurate description of current sensitivity, allowing the reproduction of the uncertainties obtained in current analyses. After validating this construction, we will use this framework to forecast future detector performance. In particular, we will investigate the limit in which the currently observed SGWB dominates over the experimental noise, in what is referred to as the signal-dominated regime. This further motivates the simplifying assumptions we made to describe the noise model.

II.2 Stochastic GW background model

The spin-2 perturbations of the metric can be expressed in terms of plane GWs with frequency f𝑓fitalic_f, polarizations {+,×}\{+,\times\}{ + , × }, and propagation directions k^^𝑘\hat{k}over^ start_ARG italic_k end_ARG as

hab(t,x)=d2Ωk^df[h~+(f,k^)eab+(k^)+h~×(f,k^)eab×(k^)]ei2πf(tk^x/c),subscript𝑎𝑏𝑡𝑥superscriptd2subscriptΩ^𝑘superscriptsubscriptd𝑓delimited-[]subscript~𝑓^𝑘subscriptsuperscript𝑒𝑎𝑏^𝑘subscript~𝑓^𝑘subscriptsuperscript𝑒𝑎𝑏^𝑘superscript𝑒𝑖2𝜋𝑓𝑡^𝑘𝑥𝑐h_{ab}(t,\vec{x})=\int{\rm d}^{2}\Omega_{\hat{k}}\>\int_{-\infty}^{\infty}{\rm d% }f\>\left[\tilde{h}_{+}(f,\hat{k})e^{+}_{ab}(\hat{k})\right.\\ \left.+\tilde{h}_{\times}(f,\hat{k})e^{\times}_{ab}(\hat{k})\right]e^{i2\pi f(% t-\hat{k}\cdot\vec{x}/c)}\,,start_ROW start_CELL italic_h start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( italic_t , over→ start_ARG italic_x end_ARG ) = ∫ roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT over^ start_ARG italic_k end_ARG end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_f [ over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT + end_POSTSUBSCRIPT ( italic_f , over^ start_ARG italic_k end_ARG ) italic_e start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( over^ start_ARG italic_k end_ARG ) end_CELL end_ROW start_ROW start_CELL + over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT × end_POSTSUBSCRIPT ( italic_f , over^ start_ARG italic_k end_ARG ) italic_e start_POSTSUPERSCRIPT × end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( over^ start_ARG italic_k end_ARG ) ] italic_e start_POSTSUPERSCRIPT italic_i 2 italic_π italic_f ( italic_t - over^ start_ARG italic_k end_ARG ⋅ over→ start_ARG italic_x end_ARG / italic_c ) end_POSTSUPERSCRIPT , end_CELL end_ROW (8)

where we introduced the polarization tensors eab+,×(k^)subscriptsuperscript𝑒𝑎𝑏^𝑘e^{+,\times}_{ab}(\hat{k})italic_e start_POSTSUPERSCRIPT + , × end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( over^ start_ARG italic_k end_ARG ), with the two polarizations denoted P={+,×}𝑃P=\{+,\times\}italic_P = { + , × } in the following. We assume that the SGWB is stationary, unpolarized, and isotropic, which means

h~P(f,k^)h~P(f,k^)=116πSh(f)δ(ff)δPPδ2(k^,k^),delimited-⟨⟩subscript~𝑃𝑓^𝑘superscriptsubscript~superscript𝑃superscript𝑓superscript^𝑘116𝜋subscript𝑆𝑓𝛿𝑓superscript𝑓subscript𝛿𝑃superscript𝑃superscript𝛿2^𝑘superscript^𝑘\langle\tilde{h}_{P}(f,\hat{k})\tilde{h}_{P^{\prime}}^{*}(f^{\prime},\hat{k}^{% \prime})\rangle=\frac{1}{16\pi}S_{h}(f)\delta(f-f^{\prime})\delta_{PP^{\prime}% }\delta^{2}(\hat{k},\hat{k}^{\prime})\,,⟨ over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( italic_f , over^ start_ARG italic_k end_ARG ) over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , over^ start_ARG italic_k end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ⟩ = divide start_ARG 1 end_ARG start_ARG 16 italic_π end_ARG italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_f ) italic_δ ( italic_f - italic_f start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_δ start_POSTSUBSCRIPT italic_P italic_P start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_δ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( over^ start_ARG italic_k end_ARG , over^ start_ARG italic_k end_ARG start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , (9)

where Sh(f)subscript𝑆𝑓S_{h}(f)italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_f ) is the (one-sided) strain power spectral density of the GWB. We define the SGWB energy density (per logarithmic frequency interval) as (see e.g., Caprini and Figueroa (2018))

ΩGWh2=h2ρcdρGWdlogf2π2f33H02/h2Sh,subscriptΩGWsuperscript2superscript2subscript𝜌𝑐𝑑subscript𝜌GW𝑑𝑓2superscript𝜋2superscript𝑓33superscriptsubscript𝐻02superscript2subscript𝑆\Omega_{\rm GW}h^{2}=\frac{h^{2}}{\rho_{c}}\frac{d\rho_{\rm GW}}{d\log f}% \equiv\frac{2\pi^{2}f^{3}}{3H_{0}^{2}/h^{2}}S_{h},roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG divide start_ARG italic_d italic_ρ start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT end_ARG start_ARG italic_d roman_log italic_f end_ARG ≡ divide start_ARG 2 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG 3 italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT , (10)

where ρc/h2=3(H0/h)2/(8πG)subscript𝜌𝑐superscript23superscriptsubscript𝐻028𝜋𝐺\rho_{c}/h^{2}=3(H_{0}/h)^{2}/(8\pi G)italic_ρ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 3 ( italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_h ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 8 italic_π italic_G ) is the Universe critical energy density and H0/h=1/(9.78Gyr)subscript𝐻019.78GyrH_{0}/h=1/(9.78{\rm Gyr})italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_h = 1 / ( 9.78 roman_Gyr ) is the Hubble parameter today. In the last equality, we have associated ΩGWh2subscriptΩGWsuperscript2\Omega_{\rm GW}h^{2}roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to the strain power spectral density Shsubscript𝑆S_{h}italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT.

The timing residual response of the signal coming from a pulsar I𝐼Iitalic_I to the SGWB can be expressed in Fourier space as Hazboun et al. (2019)

h~I(f)=d2Ωk^PRIP(f,k^)h~P(f;k^),subscript~𝐼𝑓superscriptd2subscriptΩ^𝑘subscript𝑃superscriptsubscript𝑅𝐼𝑃𝑓^𝑘subscript~𝑃𝑓^𝑘\displaystyle\tilde{h}_{I}(f)=\int{\rm d}^{2}\Omega_{\hat{k}}\>\sum_{P}R_{I}^{% P}(f,\hat{k})\tilde{h}_{P}(f;\hat{k}),over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_f ) = ∫ roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT over^ start_ARG italic_k end_ARG end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_P end_POSTSUPERSCRIPT ( italic_f , over^ start_ARG italic_k end_ARG ) over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( italic_f ; over^ start_ARG italic_k end_ARG ) , (11)

which is integrated over all propagation directions k^^𝑘\hat{k}over^ start_ARG italic_k end_ARG. The response function for a pulsar located at a distance D𝐷Ditalic_D along the direction p^Isubscript^𝑝𝐼\hat{p}_{I}over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT is

RIP(f,k^)ϵabP(k^)i4πfpI^apI^b1+pI^k^(1ei2πfD(1+k^p^I)/c).superscriptsubscript𝑅𝐼𝑃𝑓^𝑘subscriptsuperscriptitalic-ϵ𝑃𝑎𝑏^𝑘𝑖4𝜋𝑓superscript^subscript𝑝𝐼𝑎superscript^subscript𝑝𝐼𝑏1^subscript𝑝𝐼^𝑘1superscript𝑒𝑖2𝜋𝑓𝐷1^𝑘subscript^𝑝𝐼𝑐\displaystyle R_{I}^{P}(f,\hat{k})\equiv\frac{\epsilon^{P}_{ab}(\hat{k})}{i4% \pi f}\frac{\hat{p_{I}}^{a}\hat{p_{I}}^{b}}{1+\hat{p_{I}}\cdot\hat{k}}\left(1-% e^{-i2\pi fD(1+\hat{k}\cdot\hat{p}_{I})/c}\right)\,.italic_R start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_P end_POSTSUPERSCRIPT ( italic_f , over^ start_ARG italic_k end_ARG ) ≡ divide start_ARG italic_ϵ start_POSTSUPERSCRIPT italic_P end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_a italic_b end_POSTSUBSCRIPT ( over^ start_ARG italic_k end_ARG ) end_ARG start_ARG italic_i 4 italic_π italic_f end_ARG divide start_ARG over^ start_ARG italic_p start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG start_POSTSUPERSCRIPT italic_a end_POSTSUPERSCRIPT over^ start_ARG italic_p start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG start_POSTSUPERSCRIPT italic_b end_POSTSUPERSCRIPT end_ARG start_ARG 1 + over^ start_ARG italic_p start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ⋅ over^ start_ARG italic_k end_ARG end_ARG ( 1 - italic_e start_POSTSUPERSCRIPT - italic_i 2 italic_π italic_f italic_D ( 1 + over^ start_ARG italic_k end_ARG ⋅ over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ) / italic_c end_POSTSUPERSCRIPT ) . (12)

When dealing with PTA observations, it is easy to see that the characteristic frequency f=(2πD/c)1subscript𝑓superscript2𝜋𝐷𝑐1f_{*}=(2\pi D/c)^{-1}italic_f start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = ( 2 italic_π italic_D / italic_c ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT associated to a distance D𝐷Ditalic_D (which is of order kpc) turns out to be f2×1012similar-to-or-equalssubscript𝑓2superscript1012f_{*}\simeq 2\times 10^{-12}italic_f start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≃ 2 × 10 start_POSTSUPERSCRIPT - 12 end_POSTSUPERSCRIPTHz. On the other hand, the minimum frequency accessible in our analysis is limited by the observation time f1/Tobssimilar-to𝑓1subscript𝑇obssimilar-toabsentf\sim 1/T_{\rm obs}\simitalic_f ∼ 1 / italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT ∼nHz. Therefore, one finds that for PTA experiments fD1much-greater-than𝑓𝐷1fD\gg 1italic_f italic_D ≫ 1, and the frequency-dependent term in the response function (12) is well approximated by RP(f)1/fproportional-tosuperscript𝑅𝑃𝑓1𝑓R^{P}(f)\propto 1/fitalic_R start_POSTSUPERSCRIPT italic_P end_POSTSUPERSCRIPT ( italic_f ) ∝ 1 / italic_f, while the rapidly oscillating piece is negligible when averaged over the directions k^^𝑘\hat{k}over^ start_ARG italic_k end_ARG in (11).

It follows that the SGWB signal covariance matrix is

Ch,IJh~Ih~JT=RIJSh(f),subscript𝐶𝐼𝐽delimited-⟨⟩subscript~𝐼superscriptsubscript~𝐽absent𝑇subscript𝑅𝐼𝐽subscript𝑆𝑓C_{h,IJ}\equiv\langle\tilde{h}_{I}\tilde{h}_{J}^{*T}\rangle=R_{IJ}S_{h}(f),italic_C start_POSTSUBSCRIPT italic_h , italic_I italic_J end_POSTSUBSCRIPT ≡ ⟨ over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT over~ start_ARG italic_h end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ italic_T end_POSTSUPERSCRIPT ⟩ = italic_R start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ( italic_f ) , (13)

where we introduced Hazboun et al. (2019)

RIJ=χIJ(f)[𝒯I(f)𝒯J(f)TIJ/Tobs]1/2,subscript𝑅𝐼𝐽subscript𝜒𝐼𝐽𝑓superscriptdelimited-[]subscript𝒯𝐼𝑓subscript𝒯𝐽𝑓subscript𝑇𝐼𝐽subscript𝑇obs12R_{IJ}=\chi_{IJ}\cdot{\cal R}(f)\left[{\cal T}_{I}(f){\cal T}_{J}(f){T_{IJ}}/{% T_{\text{obs}}}\right]^{1/2},italic_R start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT = italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT ⋅ caligraphic_R ( italic_f ) [ caligraphic_T start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_f ) caligraphic_T start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ( italic_f ) italic_T start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT , (14)

and we conveniently defined the frequency dependent, sky-averaged, quadratic response function (f)1/12π2f2𝑓112superscript𝜋2superscript𝑓2{\cal R}(f)\equiv 1/12\pi^{2}f^{2}caligraphic_R ( italic_f ) ≡ 1 / 12 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_f start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The first geometrical factor is the well-known HD correlation pattern as a function only of the angular separation between pulsars as required by symmetries (see e.g. Kehagias and Riotto (2024)), discussed below. In contrast, the time-dependent factor is introduced to account for each pulsar transmission function as well as the individual observation time of each pulsar TIsubscript𝑇𝐼T_{I}italic_T start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT. This is because we can only include off-diagonal components correlating signals between different pulsars for an effective overlapping time defined as TIJ=min[TI,TJ]subscript𝑇𝐼𝐽minsubscript𝑇𝐼subscript𝑇𝐽T_{IJ}={\rm min}[T_{I},T_{J}]italic_T start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT = roman_min [ italic_T start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT , italic_T start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ].

II.2.1 The Hellings-Downs correlation

We introduced the HD function Hellings and Downs (1983) for a pair of pulsars separated by an angle ζIJarccos(p^Ip^J)subscript𝜁𝐼𝐽arccosinesubscript^𝑝𝐼subscript^𝑝𝐽\zeta_{IJ}\equiv\arccos(\hat{p}_{I}\cdot\hat{p}_{J})italic_ζ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT ≡ roman_arccos ( start_ARG over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ⋅ over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT end_ARG ), which takes the analytic expression

χIJ=12+32ξIJ[lnξIJ16]+12δIJ,subscript𝜒𝐼𝐽1232subscript𝜉𝐼𝐽delimited-[]subscript𝜉𝐼𝐽1612subscript𝛿𝐼𝐽\chi_{IJ}=\frac{1}{2}+\frac{3}{2}\xi_{IJ}\left[\ln\xi_{IJ}-\frac{1}{6}\right]+% \frac{1}{2}\,\delta_{IJ}\,,italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG + divide start_ARG 3 end_ARG start_ARG 2 end_ARG italic_ξ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT [ roman_ln italic_ξ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT - divide start_ARG 1 end_ARG start_ARG 6 end_ARG ] + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_δ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT , (15)

where ξIJ(1cos(ζIJ))/2subscript𝜉𝐼𝐽1subscript𝜁𝐼𝐽2\xi_{IJ}\equiv\left({1-\cos(\zeta_{IJ})}\right)/{2}italic_ξ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT ≡ ( 1 - roman_cos ( start_ARG italic_ζ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT end_ARG ) ) / 2.

To estimate the sensitivity to the HD angular correlation function, it is convenient to decompose it on a suitable basis. Following Ref. Antoniadis et al. (2023a), we adopt the following templates:

  • Binned HD function: we model the correlation function as a step-wise constant over n𝑛nitalic_n bins in the angular variable ξIJsubscript𝜉𝐼𝐽\xi_{IJ}italic_ξ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT. This takes the form

    χIJ=i=0n1biΘ(ζIJζi)Θ(ζi+1ζIJ),subscript𝜒𝐼𝐽superscriptsubscript𝑖0𝑛1subscript𝑏𝑖Θsubscript𝜁𝐼𝐽superscript𝜁𝑖Θsuperscript𝜁𝑖1subscript𝜁𝐼𝐽\chi_{IJ}=\sum_{i=0}^{n-1}b_{i}\Theta(\zeta_{IJ}-\zeta^{i})\Theta(\zeta^{i+1}-% \zeta_{IJ}),italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Θ ( italic_ζ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT - italic_ζ start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ) roman_Θ ( italic_ζ start_POSTSUPERSCRIPT italic_i + 1 end_POSTSUPERSCRIPT - italic_ζ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT ) , (16)

    where the bins in the angular variables are equally spaced ζi=iπ/nsuperscript𝜁𝑖𝑖𝜋𝑛\zeta^{i}=i\pi/nitalic_ζ start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT = italic_i italic_π / italic_n.333Notice that for a small number of pulsars, a sensible choice would be to divide the angular range unevenly and choose bins containing an equal number of pulsar pairs Antoniadis et al. (2023a). We also introduced the Heaviside theta function ΘΘ\Thetaroman_Θ. We impose Gaussian priors on the coefficients bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT with width σbi=1subscript𝜎subscript𝑏𝑖1\sigma_{b_{i}}=1italic_σ start_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 1.

  • HD expansion in Legendre polynomials: following Gair et al. (2014); Antoniadis et al. (2023a), we expand the HD function as

    χIJ==0naP(cosζIJ).subscript𝜒𝐼𝐽superscriptsubscript0𝑛subscript𝑎subscript𝑃subscript𝜁𝐼𝐽\chi_{IJ}=\sum_{\ell=0}^{n}a_{\ell}P_{\ell}(\cos\zeta_{IJ}).italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT roman_ℓ = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ( roman_cos italic_ζ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT ) . (17)

    Using the standard normalization of the Legendre polynomials, the coefficients are found by Gair et al. (2014); Romano and Allen (2023)

    asubscript𝑎\displaystyle a_{\ell}italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT =2+1211χHD(x)P(x)dxabsent212superscriptsubscript11subscript𝜒HD𝑥subscript𝑃𝑥d𝑥\displaystyle=\frac{2\ell+1}{2}\int_{-1}^{1}\chi_{\text{\tiny HD}}(x)P_{\ell}(% x)\textrm{d}x= divide start_ARG 2 roman_ℓ + 1 end_ARG start_ARG 2 end_ARG ∫ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT HD end_POSTSUBSCRIPT ( italic_x ) italic_P start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ( italic_x ) d italic_x
    =32(2+1)(+2)(+1)(1).absent3221211\displaystyle=\frac{3}{2}\frac{(2\ell+1)}{(\ell+2)(\ell+1)\ell(\ell-1)}.= divide start_ARG 3 end_ARG start_ARG 2 end_ARG divide start_ARG ( 2 roman_ℓ + 1 ) end_ARG start_ARG ( roman_ℓ + 2 ) ( roman_ℓ + 1 ) roman_ℓ ( roman_ℓ - 1 ) end_ARG . (18)

    for 22\ell\geq 2roman_ℓ ≥ 2, and a0=a1=0subscript𝑎0subscript𝑎10a_{0}=a_{1}=0italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0. We will estimate the sensitivity to the HD curve by constraining the expansion parameters asubscript𝑎a_{\ell}italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT. We impose priors on asubscript𝑎a_{\ell}italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT to be Gaussian with width σa=1subscript𝜎subscript𝑎1\sigma_{a_{\ell}}=1italic_σ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 1.

II.2.2 Effective sensitivity

We compute the SNR by

SNR2fk[CIJ1CKL1RJKRLI]Sh2,superscriptSNR2subscriptsubscript𝑓𝑘delimited-[]subscriptsuperscript𝐶1𝐼𝐽subscriptsuperscript𝐶1𝐾𝐿subscript𝑅𝐽𝐾subscript𝑅𝐿𝐼superscriptsubscript𝑆2{\rm SNR}^{2}\equiv\sum_{f_{k}}\left[C^{-1}_{IJ}C^{-1}_{KL}R_{JK}R_{LI}\right]% S_{h}^{2},roman_SNR start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≡ ∑ start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT [ italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_K italic_L end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_J italic_K end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_L italic_I end_POSTSUBSCRIPT ] italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (19)

where, here and throughout the work, we implicitly assume Einstein’s summation convention on the pulsar indices in capital letters. It is instructive to define an effective sensitivity, as Hazboun et al. (2019)

Seff(f)=(CIJ1CKL1RJKRLI)1/2,subscript𝑆eff𝑓superscriptsubscriptsuperscript𝐶1𝐼𝐽subscriptsuperscript𝐶1𝐾𝐿subscript𝑅𝐽𝐾subscript𝑅𝐿𝐼12S_{\rm eff}(f)=\left(C^{-1}_{IJ}C^{-1}_{KL}R_{JK}R_{LI}\right)^{-1/2},italic_S start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_f ) = ( italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_K italic_L end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_J italic_K end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_L italic_I end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT , (20)

in such a way that the SNR computation takes the familiar form SNR2=Tobsk(Sh/Seff)2superscriptSNR2subscript𝑇obssubscript𝑘superscriptsubscript𝑆subscript𝑆eff2{\rm SNR}^{2}=T_{\text{obs}}\sum_{k}(S_{h}/S_{\text{eff}})^{2}roman_SNR start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_S start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT.

In the weak signal limit (which is in practice the limit in which ShPnmuch-less-thansubscript𝑆subscript𝑃𝑛{\cal R}S_{h}\ll P_{n}caligraphic_R italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ≪ italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT), one can expand the correlation matrix above to find

Seff(f)=(IJTIJTobs𝒯I(f)𝒯J(f)χIJ2Pn,I(f)Pn,J(f)/2(f))1/2,subscript𝑆eff𝑓superscriptsubscript𝐼𝐽subscript𝑇𝐼𝐽subscript𝑇obssubscript𝒯𝐼𝑓subscript𝒯𝐽𝑓subscriptsuperscript𝜒2𝐼𝐽subscript𝑃𝑛𝐼𝑓subscript𝑃𝑛𝐽𝑓superscript2𝑓12S_{\rm eff}(f)=\left(\sum_{I\neq J}\frac{T_{IJ}}{T_{\text{obs}}}\frac{\mathcal% {T}_{I}(f)\mathcal{T}_{J}(f)\chi^{2}_{IJ}}{P_{n,I}(f)P_{n,J}(f)/\mathcal{R}^{2% }(f)}\right)^{-1/2}\,,italic_S start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_f ) = ( ∑ start_POSTSUBSCRIPT italic_I ≠ italic_J end_POSTSUBSCRIPT divide start_ARG italic_T start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT end_ARG divide start_ARG caligraphic_T start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_f ) caligraphic_T start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ( italic_f ) italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT end_ARG start_ARG italic_P start_POSTSUBSCRIPT italic_n , italic_I end_POSTSUBSCRIPT ( italic_f ) italic_P start_POSTSUBSCRIPT italic_n , italic_J end_POSTSUBSCRIPT ( italic_f ) / caligraphic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_f ) end_ARG ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT , (21)

which includes contributions from the Hellings and Downs factors χIJsubscript𝜒𝐼𝐽\chi_{IJ}italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT and the individual pulsar strain-noise power spectral densities. This corresponds to the result of Ref. Hazboun et al. (2019) and shown in Refs. Agazie et al. (2023d, e) to report current PTA sensitivities. As we will see, when considering future PTA datasets (including longer observation times and a larger number of pulsars), it is important to adopt the full expression for the effective sensitivity to go beyond the weak signal approximation. We report examples of future effective sensitivity, in the presence of the currently observed GW background in Fig. 2 in the next section.

II.3 Likelihood function and parameter estimation

The log-likelihood for Gaussian and zero mean data d~ksuperscript~𝑑𝑘\tilde{d}^{k}over~ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT, with k𝑘kitalic_k running over frequencies fksubscript𝑓𝑘f_{k}italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, described only by their variance, dubbed CIJ(fk,θ)subscript𝐶𝐼𝐽subscript𝑓𝑘𝜃C_{IJ}(f_{k},{\theta})italic_C start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT ( italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ ), can be written as Contaldi et al. (2020); Bond et al. (1998)

ln(d~|θ)k,IJln[CIJ(fk,θ)]+d~IkCIJ1(fk,θ)d~Jk,proportional-toconditional~𝑑𝜃subscript𝑘𝐼𝐽subscript𝐶𝐼𝐽subscript𝑓𝑘𝜃superscriptsubscript~𝑑𝐼𝑘superscriptsubscript𝐶𝐼𝐽1subscript𝑓𝑘𝜃superscriptsubscript~𝑑𝐽𝑘-\ln\mathcal{L}(\tilde{d}|{\theta})\propto\sum_{k,IJ}\ln\left[C_{IJ}(f_{k},{% \theta})\right]+\tilde{d}_{I}^{k}C_{IJ}^{-1}(f_{k},{\theta})\tilde{d}_{J}^{k*},- roman_ln caligraphic_L ( over~ start_ARG italic_d end_ARG | italic_θ ) ∝ ∑ start_POSTSUBSCRIPT italic_k , italic_I italic_J end_POSTSUBSCRIPT roman_ln [ italic_C start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT ( italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ ) ] + over~ start_ARG italic_d end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_C start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ ) over~ start_ARG italic_d end_ARG start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k ∗ end_POSTSUPERSCRIPT , (22)

which is also known as Whittle likelihood. This likelihood function for θ𝜃{\theta}italic_θ corresponds to the probability of the data d~~𝑑\tilde{d}over~ start_ARG italic_d end_ARG given θ𝜃{\theta}italic_θ, the signal model parameters.

Fisher information matrix estimates.

Here we briefly summarise the main ingredients of the Fisher Information Matrix (FIM) formalism, which is often used in the limit of large SNR to assess parameter uncertainties.

The FIM Fαβsubscript𝐹𝛼𝛽F_{\alpha\beta}italic_F start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT is defined as

Fαβ2logθαθβ|θ=θ0=kTr[C1CθαC1Cθβ]θ=θ0,subscript𝐹𝛼𝛽evaluated-atsuperscript2superscript𝜃𝛼superscript𝜃𝛽𝜃subscript𝜃0subscript𝑘Trsubscriptdelimited-[]superscript𝐶1𝐶superscript𝜃𝛼superscript𝐶1𝐶superscript𝜃𝛽𝜃subscript𝜃0F_{\alpha\beta}\equiv-\left.\frac{\partial^{2}\log\mathcal{L}}{\partial{\theta% }^{\alpha}\partial{\theta}^{\beta}}\right|_{{\theta}={\theta}_{0}}=\sum_{k}% \textrm{Tr}\left[C^{-1}\frac{\partial C}{\partial{\theta}^{\alpha}}C^{-1}\frac% {\partial C}{\partial{\theta}^{\beta}}\right]_{{\theta}={\theta}_{0}}\;,italic_F start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT ≡ - divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_log caligraphic_L end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT ∂ italic_θ start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT end_ARG | start_POSTSUBSCRIPT italic_θ = italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT Tr [ italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_C end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_C end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT end_ARG ] start_POSTSUBSCRIPT italic_θ = italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT , (23)

where, θ0subscript𝜃0{\theta}_{0}italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT represents the maximum likelihood estimator of model parameters, determined by imposing

logθα|θ=θ0kCθα[C1C1d~kd~kC1]=0,proportional-toevaluated-atsuperscript𝜃𝛼𝜃subscript𝜃0subscript𝑘𝐶superscript𝜃𝛼delimited-[]superscript𝐶1superscript𝐶1superscript~𝑑𝑘superscript~𝑑𝑘superscript𝐶10\left.\frac{\partial\log\mathcal{L}}{\partial{\theta}^{\alpha}}\right|_{{% \theta}={\theta}_{0}}\propto\sum_{k}\frac{\partial C}{\partial{\theta}^{\alpha% }}\left[C^{-1}-C^{-1}\;\tilde{d}^{k}\tilde{d}^{k*}\;C^{-1}\right]=0\;,divide start_ARG ∂ roman_log caligraphic_L end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG | start_POSTSUBSCRIPT italic_θ = italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ∝ ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT divide start_ARG ∂ italic_C end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG [ italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT - italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over~ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT over~ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT italic_k ∗ end_POSTSUPERSCRIPT italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] = 0 , (24)

with C(fk,θ0)=d~kd~k𝐶subscript𝑓𝑘subscript𝜃0superscript~𝑑𝑘superscript~𝑑𝑘C(f_{k},{\theta}_{0})=\tilde{d}^{k}\tilde{d}^{k*}italic_C ( italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = over~ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT over~ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT italic_k ∗ end_POSTSUPERSCRIPT. In practice, the discrete sum over finite frequencies can be replaced with a continuous integral over the frequency range as

FαβfklogCθαlogCθβ.subscript𝐹𝛼𝛽subscriptsubscript𝑓𝑘𝐶superscript𝜃𝛼𝐶superscript𝜃𝛽F_{\alpha\beta}\equiv\sum_{f_{k}}\frac{\partial\log C}{\partial{\theta}^{% \alpha}}\frac{\partial\log C}{\partial{\theta}^{\beta}}\,.italic_F start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT ≡ ∑ start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT divide start_ARG ∂ roman_log italic_C end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG divide start_ARG ∂ roman_log italic_C end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT end_ARG . (25)

Keeping all indices fully expressed, one obtains

FαβfkCIJ1CKL1(RJKSh)θα(RLISh)θβ.subscript𝐹𝛼𝛽subscriptsubscript𝑓𝑘subscriptsuperscript𝐶1𝐼𝐽subscriptsuperscript𝐶1𝐾𝐿subscript𝑅𝐽𝐾subscript𝑆superscript𝜃𝛼subscript𝑅𝐿𝐼subscript𝑆superscript𝜃𝛽F_{\alpha\beta}\equiv\sum_{f_{k}}C^{-1}_{IJ}C^{-1}_{KL}\frac{\partial(R_{JK}S_% {h})}{\partial{\theta}^{\alpha}}\frac{\partial(R_{LI}S_{h})}{\partial{\theta}^% {\beta}}.italic_F start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT ≡ ∑ start_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT italic_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_K italic_L end_POSTSUBSCRIPT divide start_ARG ∂ ( italic_R start_POSTSUBSCRIPT italic_J italic_K end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG divide start_ARG ∂ ( italic_R start_POSTSUBSCRIPT italic_L italic_I end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_θ start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT end_ARG . (26)

Finally, the covariance matrix Cαβsubscript𝐶𝛼𝛽C_{\alpha\beta}italic_C start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT, is obtained by inverting the FIM, from which one can estimate uncertainties as σαFαα1subscript𝜎𝛼superscriptsubscript𝐹𝛼𝛼1\sigma_{\alpha}\equiv\sqrt{F_{\alpha\alpha}^{-1}}italic_σ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ≡ square-root start_ARG italic_F start_POSTSUBSCRIPT italic_α italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_ARG.

MCMC analyses.

Given that the FIM is an unreliable estimator in the low SNR limit, where the validity of the Gaussian approximation is expected to degrade, we will test the FIM results with full-fledged Bayesian parameter estimation. For this purpose, we simulate frequency domain data d~kd~(fk)superscript~𝑑𝑘~𝑑subscript𝑓𝑘\tilde{d}^{k}\equiv\tilde{d}(f_{k})over~ start_ARG italic_d end_ARG start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ≡ over~ start_ARG italic_d end_ARG ( italic_f start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ), where k𝑘kitalic_k indexes the frequencies within the detector sensitivity range. We generate Gaussian realizations for the signal and all noise components, with zero mean and variances defined by their respective power spectral densities. The parameter estimation is then performed by running a Monte Carlo Markov Chain (MCMC) using emcee Foreman-Mackey et al. (2013) with uninformative priors on signal parameters.

III Simulating future pulsar timing array networks

To test our method, we compare to currently available datasets, while we also propose to simulate future generation datasets. In the upcoming years, the Square Kilometer Array (SKA) will be a central protagonist among PTA collaborations Janssen et al. (2015). At the moment, PTA datasets are limited by radiometer noise. The SKA will provide high-precision pulsar timing measurement with uncertainties below similar-to\sim100ns Janssen et al. (2015), making it roughly 10 times better than current generation telescopes Lazio (2013). With a reduction of the white noise at relatively high frequencies, it also comes with a better determination of the RN, relevant at low frequencies. For our purposes, we will simulate two mock datasets:

  • An EPTA-like with 50 pulsars and 20 yr observations. We will denote this as EPTA20 in the rest of the paper.

  • A SKA-like with 50 pulsars and 10 yr observations. We will denote this configuration as SKA10.

III.1 Fake PTA data and time-domain Bayesian analysis

III.1.1 Time domain representation

As mentioned in Sec. II, we are dealing with the timing residuals obtained by fitting and subtracting the timing model from observed pulses’ ToAs. The residuals from all pulsars are concatenated in a single array forming PTA data set d=[d1(t),d2(t),,dNpulsars(t)]𝑑subscript𝑑1𝑡subscript𝑑2𝑡subscript𝑑subscript𝑁pulsars𝑡\vec{d}=[d_{1}(t),d_{2}(t),...,d_{N_{\rm pulsars}}(t)]over→ start_ARG italic_d end_ARG = [ italic_d start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) , italic_d start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) , … , italic_d start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT end_POSTSUBSCRIPT ( italic_t ) ]. We expect that the residual errors in fitting the timing model parameters are small and could be approximated by a linear model

dIdIkαkMIk,superscriptsubscript𝑑𝐼subscript𝑑𝐼subscript𝑘subscript𝛼𝑘subscript𝑀𝐼𝑘d_{I}^{\prime}\rightarrow d_{I}-\sum_{k}\alpha_{k}\vec{M}_{Ik},italic_d start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT → italic_d start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT - ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT over→ start_ARG italic_M end_ARG start_POSTSUBSCRIPT italic_I italic_k end_POSTSUBSCRIPT , (27)

where MIksubscript𝑀𝐼𝑘M_{Ik}italic_M start_POSTSUBSCRIPT italic_I italic_k end_POSTSUBSCRIPT is a design matrix (representing the derivative of the timing model w.r.t timing model parameters), and αksubscript𝛼𝑘\alpha_{k}italic_α start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT are residual timing model errors, which we consider as random variables. We have already used the design matrix in the frequency domain while building the transmission function.

For the description of the noise components, we use Gaussian process representation (see van Haasteren and Vallisneri (2014) for a detailed description) with a discrete Fourier basis fi=i/TIsubscript𝑓𝑖𝑖subscript𝑇𝐼f_{i}=i/T_{I}italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_i / italic_T start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT with i=[1,,Nf]𝑖1subscript𝑁𝑓i=[1,...,N_{f}]italic_i = [ 1 , … , italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ] as the eigenfunctions:

nI(t)=iXi,Icos(2πfit)+Yi,Isin(2πfit),subscript𝑛𝐼𝑡subscript𝑖subscript𝑋𝑖𝐼2𝜋subscript𝑓𝑖𝑡subscript𝑌𝑖𝐼2𝜋subscript𝑓𝑖𝑡n_{I}(t)=\sum_{i}X_{i,I}\cos(2\pi f_{i}t)+Y_{i,I}\sin(2\pi f_{i}t),italic_n start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_X start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT roman_cos ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_t end_ARG ) + italic_Y start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT roman_sin ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_t end_ARG ) , (28)

where Xi,I,Yi,I𝒩(0,S(fi)Δfi)similar-tosubscript𝑋𝑖𝐼subscript𝑌𝑖𝐼𝒩0𝑆subscript𝑓𝑖Δsubscript𝑓𝑖X_{i,I},Y_{i,I}\sim\mathcal{N}(0,S(f_{i})\Delta f_{i})italic_X start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT , italic_Y start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT ∼ caligraphic_N ( 0 , italic_S ( italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) roman_Δ italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ), S(f)𝑆𝑓S(f)italic_S ( italic_f ) the one-sided PSD of the noise and Δfi=fi+1fiΔsubscript𝑓𝑖subscript𝑓𝑖1subscript𝑓𝑖\Delta f_{i}=f_{i+1}-f_{i}roman_Δ italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT - italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The weights Xi,I,Yi,Isubscript𝑋𝑖𝐼subscript𝑌𝑖𝐼X_{i,I},Y_{i,I}italic_X start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT , italic_Y start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT for the pulsar intrinsic noise components are uncorrelated among the pulsars. For GWB signals, they are spatially correlated between pulsars I𝐼Iitalic_I and J𝐽Jitalic_J and are distributed as a Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT dimensional multivariate normal distribution Xi,Yi𝒩(0,χIJS(fi)Δfi)similar-tosubscript𝑋𝑖subscript𝑌𝑖𝒩0subscript𝜒𝐼𝐽𝑆subscript𝑓𝑖Δsubscript𝑓𝑖\vec{X}_{i},\vec{Y}_{i}\sim\mathcal{N}(0,\chi_{IJ}S(f_{i})\Delta f_{i})over→ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , over→ start_ARG italic_Y end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ caligraphic_N ( 0 , italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT italic_S ( italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) roman_Δ italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) where χIJsubscript𝜒𝐼𝐽\chi_{IJ}italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT are the introduced above HD correlation coefficients. In general, the noise covariance matrix is given as

CInsuperscriptsubscript𝐶𝐼𝑛\displaystyle C_{I}^{n}italic_C start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT =\displaystyle== nI(t)nI(t)delimited-⟨⟩subscript𝑛𝐼𝑡subscript𝑛𝐼𝑡\displaystyle\langle n_{I}(t)n_{I}(t)\rangle⟨ italic_n start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_t ) italic_n start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_t ) ⟩
=\displaystyle== [cos(2πf1tI)sin(2πf1tI)cos(2πfNftI)sin(2πfNftI)][X1,I2Y1,I2XNf,I2YNf,I2][cos(2πf1tI)sin(2πf1tI)cos(2πfNftI)sin(2πfNftI)],matrix2𝜋subscript𝑓1subscript𝑡𝐼2𝜋subscript𝑓1subscript𝑡𝐼2𝜋subscript𝑓subscript𝑁𝑓subscript𝑡𝐼2𝜋subscript𝑓subscript𝑁𝑓subscript𝑡𝐼matrixdelimited-⟨⟩superscriptsubscript𝑋1𝐼2delimited-⟨⟩superscriptsubscript𝑌1𝐼2delimited-⟨⟩superscriptsubscript𝑋subscript𝑁𝑓𝐼2delimited-⟨⟩superscriptsubscript𝑌subscript𝑁𝑓𝐼2matrix2𝜋subscript𝑓1subscript𝑡𝐼2𝜋subscript𝑓1subscript𝑡𝐼2𝜋subscript𝑓subscript𝑁𝑓subscript𝑡𝐼2𝜋subscript𝑓subscript𝑁𝑓subscript𝑡𝐼\displaystyle\begin{bmatrix}\cos(2\pi f_{1}\vec{t}_{I})\\ \sin(2\pi f_{1}\vec{t}_{I})\\ \vdots\\ \cos(2\pi f_{N_{f}}\vec{t}_{I})\\ \sin(2\pi f_{N_{f}}\vec{t}_{I})\end{bmatrix}\cdot\begin{bmatrix}\langle X_{1,I% }^{2}\rangle\\ \langle Y_{1,I}^{2}\rangle\\ \vdots\\ \langle X_{N_{f},I}^{2}\rangle\\ \langle Y_{N_{f},I}^{2}\rangle\end{bmatrix}\cdot\begin{bmatrix}\cos(2\pi f_{1}% \vec{t}_{I})\\ \sin(2\pi f_{1}\vec{t}_{I})\\ \vdots\\ \cos(2\pi f_{N_{f}}\vec{t}_{I})\\ \sin(2\pi f_{N_{f}}\vec{t}_{I})\end{bmatrix},[ start_ARG start_ROW start_CELL roman_cos ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ) end_CELL end_ROW start_ROW start_CELL roman_sin ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ) end_CELL end_ROW start_ROW start_CELL ⋮ end_CELL end_ROW start_ROW start_CELL roman_cos ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ) end_CELL end_ROW start_ROW start_CELL roman_sin ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ) end_CELL end_ROW end_ARG ] ⋅ [ start_ARG start_ROW start_CELL ⟨ italic_X start_POSTSUBSCRIPT 1 , italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_CELL end_ROW start_ROW start_CELL ⟨ italic_Y start_POSTSUBSCRIPT 1 , italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_CELL end_ROW start_ROW start_CELL ⋮ end_CELL end_ROW start_ROW start_CELL ⟨ italic_X start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_CELL end_ROW start_ROW start_CELL ⟨ italic_Y start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT , italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ end_CELL end_ROW end_ARG ] ⋅ [ start_ARG start_ROW start_CELL roman_cos ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ) end_CELL end_ROW start_ROW start_CELL roman_sin ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ) end_CELL end_ROW start_ROW start_CELL ⋮ end_CELL end_ROW start_ROW start_CELL roman_cos ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ) end_CELL end_ROW start_ROW start_CELL roman_sin ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT italic_N start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ) end_CELL end_ROW end_ARG ] ,

where X1,I2delimited-⟨⟩superscriptsubscript𝑋1𝐼2\langle X_{1,I}^{2}\rangle⟨ italic_X start_POSTSUBSCRIPT 1 , italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ being the variance of the Xi,I,Yi,Isubscript𝑋𝑖𝐼subscript𝑌𝑖𝐼X_{i,I},Y_{i,I}italic_X start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT , italic_Y start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT as defined by their Gaussian distribution.

III.1.2 Generating fake data

We simulate the data in the time domain444The time-domain simulations were made using https://github.com/mfalxa/fakepta.. We start with choosing the pulsar array and epochs of observation, which we populate with the white noise with dispersion σ𝜎\sigmaitalic_σ. We simulated the red (spin) noise of each pulsar using Gaussian Process according to (28) where the weights Xi,Isubscript𝑋𝑖𝐼\vec{X}_{i,I}over→ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT and Yi,Isubscript𝑌𝑖𝐼\vec{Y}_{i,I}over→ start_ARG italic_Y end_ARG start_POSTSUBSCRIPT italic_i , italic_I end_POSTSUBSCRIPT are drawn from their respective normal probability distributions with variance defined by PSD, S(f)𝑆𝑓S(f)italic_S ( italic_f ). The simulated timing model includes only five components with the following design matrix:

  • Offset : M01proportional-tosubscript𝑀01\vec{M}_{0}\propto\vec{1}over→ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∝ over→ start_ARG 1 end_ARG,

  • Spin rate : M1tIproportional-tosubscript𝑀1subscript𝑡𝐼\vec{M}_{1}\propto\vec{t}_{I}over→ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∝ over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT,

  • Quadratic spin-down rate : M2tI2proportional-tosubscript𝑀2superscriptsubscript𝑡𝐼2\vec{M}_{2}\propto\vec{t}_{I}^{2}over→ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ∝ over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT,

  • RA : M3sin(2πfyrtI)proportional-tosubscript𝑀32𝜋subscript𝑓𝑦𝑟subscript𝑡𝐼\vec{M}_{3}\propto\sin(2\pi f_{yr}\vec{t}_{I})over→ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ∝ roman_sin ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT italic_y italic_r end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ),

  • DEC : M4cos(2πfyrtI)proportional-tosubscript𝑀42𝜋subscript𝑓𝑦𝑟subscript𝑡𝐼\vec{M}_{4}\propto\cos(2\pi f_{yr}\vec{t}_{I})over→ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT ∝ roman_cos ( start_ARG 2 italic_π italic_f start_POSTSUBSCRIPT italic_y italic_r end_POSTSUBSCRIPT over→ start_ARG italic_t end_ARG start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT end_ARG ).

The overall offset, spin rate, and quadratic spindown rate correspond to the inaccuracy in the pulsar rotation. The offset in the pulsar sky position is modeled by RA and DEC terms, giving annual modulation due to the Earth’s orbital motion. During the Bayesian PTA analysis, we marginalize over the coefficients αksubscript𝛼𝑘\alpha_{k}italic_α start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT of the linear timing model, assuming non-informative improper prior. This marginalization is responsible for the transmission function described in (3Hazboun et al. (2019).

III.2 Mock datasets

We simulated two PTA mock datasets:

  • EPTA20 : built from the 25 pulsars of the EPTA DR2new 10.4 yr dataset with the same noise properties (as in Antoniadis et al. (2023d)) but with a doubled observation time, to which we append 25 additional pulsars with the same noise properties and observation time as EPTA DR2new but randomized positions drawn uniformly in the sky. The noises are re-injected according to their maximum likelihood values in Antoniadis et al. (2023d).

  • SKA10 : SKA-like PTA with 50 pulsars and 10 yr observations. Every pulsar is simulated with a 2-week cadence of observation, σ=𝜎absent\sigma=italic_σ =100ns timing uncertainty, and random position uniformly drawn in the sky. Time-correlated noises with power-law spectra are included (red noise and dispersion measure noise) with log-amplitude log10Asubscript10𝐴\log_{10}Aroman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_A and spectral index γ𝛾\gammaitalic_γ drawn uniformly between log10A=[17,13]subscript10𝐴1713\log_{10}A=[-17,-13]roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_A = [ - 17 , - 13 ] and γ=[1,5]𝛾15\gamma=[1,5]italic_γ = [ 1 , 5 ].

These properties are summarised in Tab. 1.

EPTA20 SKA10
Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT 25 + 25 50
ν𝜈\nuitalic_ν [GHz] Antoniadis et al. (2023d) 1.4 - 3
Tobssubscript𝑇obsT_{\rm obs}italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT [yr] 20.8 + 10.4 10
σ𝜎\sigmaitalic_σ [s] 106similar-toabsentsuperscript106\sim 10^{-6}∼ 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT 107superscript10710^{-7}10 start_POSTSUPERSCRIPT - 7 end_POSTSUPERSCRIPT
ΔtΔ𝑡\Delta troman_Δ italic_t [days] 3similar-toabsent3\sim 3∼ 3 14
TM TM5 + Antoniadis et al. (2023d) TM5
log10ARNsubscript10subscript𝐴RN\log_{10}A_{\rm RN}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_RN end_POSTSUBSCRIPT Antoniadis et al. (2023d) [-17, -13]
γRNsubscript𝛾RN\gamma_{\rm RN}italic_γ start_POSTSUBSCRIPT roman_RN end_POSTSUBSCRIPT Antoniadis et al. (2023d) [1, 5]
log10ADMsubscript10subscript𝐴DM\log_{10}A_{\rm DM}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_DM end_POSTSUBSCRIPT Antoniadis et al. (2023d) [-17, -13]
γDMsubscript𝛾DM\gamma_{\rm DM}italic_γ start_POSTSUBSCRIPT roman_DM end_POSTSUBSCRIPT Antoniadis et al. (2023d) [1, 5]
log10ASVsubscript10subscript𝐴SV\log_{10}A_{\rm SV}roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_A start_POSTSUBSCRIPT roman_SV end_POSTSUBSCRIPT Antoniadis et al. (2023d) -
γSVsubscript𝛾SV\gamma_{\rm SV}italic_γ start_POSTSUBSCRIPT roman_SV end_POSTSUBSCRIPT Antoniadis et al. (2023d) -
Table 1: Summary of mock datasets properties. Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT is the number of pulsars, ν𝜈\nuitalic_ν are the observing frequency sub-bands, Tobssubscript𝑇obsT_{\rm obs}italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT is the time of observation, σ𝜎\sigmaitalic_σ is the level of white noise, ΔtΔ𝑡\Delta troman_Δ italic_t is the cadence, TM is the timing model where TM5 is the five-component model described in III.1.2, log10Asubscript10𝐴\log_{10}Aroman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_A and γ𝛾\gammaitalic_γ give the range of the uniform distribution used to draw the red noise parameters. The entries marked by Ref. Antoniadis et al. (2023d) correspond to the choice of parameters consistent with ERPTA DR2new analysis.
Refer to captionRefer to captionRefer to caption
Figure 1: Comparison between the measurement uncertainties estimated using FIM (blue) and time-domain Bayesian parameter estimation (red, see Sec. III.1 for more details). For visualization purposes, the FIM result is shown by building samples drawn from a multivariate Gaussian distribution with covariance FIM1𝐹𝐼superscript𝑀1FIM^{-1}italic_F italic_I italic_M start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. For both parameters, we remove the mean observed values to compare with the FIM which does not provide information on it. In the two-dimensional panels, we indicate 68%, 95%, and 99% C.I. Left panel: current posterior obtained with EPTADR2new datasets (EPTA10). Center panel: forecasted EPTA 20 yr dataset. Right panel: forecasted SKA 10 yr dataset.

In both datasets, we inject a stochastic GW signal with HD correlations and a power-law spectrum with log10A=14subscript10𝐴14\log_{10}A=-14roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_A = - 14 and γ=3𝛾3\gamma=3italic_γ = 3 to reproduce the signal observed in Antoniadis et al. (2023a); Agazie et al. (2023a), potentially originating from the population of SMBHB in the Universe. This signal will act as foreground, and we will also inject additional stochastic GW signals into the datasets to test our ability to detect them in the presence of a foreground (see Sec. IV.4).

III.3 Time-domain Bayesian analysis

The timing model marginalized Gaussian likelihood used in time-domain PTA analysis van Haasteren and Vallisneri (2014) can be expressed as (analogously to the previously defined likelihood (22))

ln(d|θ)det{CIJ}+IJdI(t)CIJ1(θ)dJ(t),proportional-toconditional𝑑𝜃detsubscript𝐶𝐼𝐽subscript𝐼𝐽subscript𝑑𝐼𝑡superscriptsubscript𝐶𝐼𝐽1𝜃superscriptsubscript𝑑𝐽𝑡-\ln\mathcal{L}(d|{\theta})\propto\textrm{det}\{C_{IJ}\}+\sum_{IJ}d_{I}(t)C_{% IJ}^{-1}({\theta})d_{J}^{\intercal}(t),- roman_ln caligraphic_L ( italic_d | italic_θ ) ∝ det { italic_C start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT } + ∑ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT italic_d start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT ( italic_t ) italic_C start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( italic_θ ) italic_d start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ⊺ end_POSTSUPERSCRIPT ( italic_t ) , (29)

where CIJ=CIδIJ+Ch,IJsubscript𝐶𝐼𝐽subscript𝐶𝐼subscript𝛿𝐼𝐽subscript𝐶𝐼𝐽C_{IJ}=C_{I}\delta_{IJ}+C_{h,IJ}italic_C start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT + italic_C start_POSTSUBSCRIPT italic_h , italic_I italic_J end_POSTSUBSCRIPT has a block structure of size Npulsars×Npulsarssubscript𝑁pulsarssubscript𝑁pulsarsN_{\rm pulsars}\times N_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT × italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT in which the GW signal Ch,IJ=χIJChsubscript𝐶𝐼𝐽subscript𝜒𝐼𝐽subscript𝐶C_{h,IJ}=\chi_{IJ}C_{h}italic_C start_POSTSUBSCRIPT italic_h , italic_I italic_J end_POSTSUBSCRIPT = italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT italic_C start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT (IJ𝐼𝐽I\neq Jitalic_I ≠ italic_J) appears off diagonal. The individual pulsar’ noise components form the diagonal I=J𝐼𝐽I=Jitalic_I = italic_J where CI=CITM+CIRN+CIDM+CISVsubscript𝐶𝐼superscriptsubscript𝐶𝐼𝑇𝑀superscriptsubscript𝐶𝐼𝑅𝑁superscriptsubscript𝐶𝐼𝐷𝑀superscriptsubscript𝐶𝐼𝑆𝑉C_{I}=C_{I}^{TM}+C_{I}^{RN}+C_{I}^{DM}+C_{I}^{SV}italic_C start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T italic_M end_POSTSUPERSCRIPT + italic_C start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_R italic_N end_POSTSUPERSCRIPT + italic_C start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_D italic_M end_POSTSUPERSCRIPT + italic_C start_POSTSUBSCRIPT italic_I end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_S italic_V end_POSTSUPERSCRIPT (corresponding to the timing model, the red noise, the dispersion measure noise, and the chromatic noise respectively).

We perform Bayesian analysis using the software ENTERPRISE Ellis et al. (2020); Taylor et al. (2021) and sampler PTMCMC Ellis and van Haasteren (2017). ENTERPRISE provides computation of the likelihood and prior based on the chosen model. All noise components are modelled as Gaussian Processes with incomplete Fourier basis as eigenfunctions van Haasteren and Vallisneri (2014). We are using PTMCMC without a parallel tempering option, which is sufficient for our purpose. The resulting posterior distributions for both EPTA20 and SKA10 are shown in Fig. 1 (central and right panels, red color).

IV Results

Throughout this section, we will assume that the SGWB observed by the various collaborations is described by a power-law (PL)

h2ΩGWPL(f)=AGWB(ffyr)nTsuperscript2superscriptsubscriptΩGWPL𝑓subscript𝐴GWBsuperscript𝑓subscript𝑓yrsubscript𝑛𝑇h^{2}\Omega_{\rm GW}^{\rm PL}(f)=A_{\rm GWB}\left(\frac{f}{f_{\rm yr}}\right)^% {n_{T}}italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_PL end_POSTSUPERSCRIPT ( italic_f ) = italic_A start_POSTSUBSCRIPT roman_GWB end_POSTSUBSCRIPT ( divide start_ARG italic_f end_ARG start_ARG italic_f start_POSTSUBSCRIPT roman_yr end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_POSTSUPERSCRIPT (30)

with AGWB6.3×108similar-to-or-equalssubscript𝐴GWB6.3superscript108A_{\rm GWB}\simeq 6.3\times 10^{-8}italic_A start_POSTSUBSCRIPT roman_GWB end_POSTSUBSCRIPT ≃ 6.3 × 10 start_POSTSUPERSCRIPT - 8 end_POSTSUPERSCRIPT and nT2similar-to-or-equalssubscript𝑛𝑇2n_{T}\simeq 2italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ≃ 2. For definiteness, these parameters are fixed to the maximum likelihood values obtained by the EPTA Antoniadis et al. (2023a), which is in good agreement with the other observations Agazie et al. (2023e). To make contact with other quantities used in the literature, this corresponds to a strain amplitude (hcfShsubscript𝑐𝑓subscript𝑆h_{c}\equiv\sqrt{fS_{h}}italic_h start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≡ square-root start_ARG italic_f italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT end_ARG) at f=fyr𝑓subscript𝑓yrf=f_{\text{yr}}italic_f = italic_f start_POSTSUBSCRIPT yr end_POSTSUBSCRIPT of Ahc1014similar-to-or-equalssubscript𝐴subscript𝑐superscript1014A_{h_{c}}\simeq 10^{-14}italic_A start_POSTSUBSCRIPT italic_h start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUBSCRIPT ≃ 10 start_POSTSUPERSCRIPT - 14 end_POSTSUPERSCRIPT and pulsar timing residuals spectral index γ=5nT=3𝛾5subscript𝑛𝑇3\gamma=5-n_{T}=3italic_γ = 5 - italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 3. When performing parameter estimation, it is convenient to define the log of the amplitude αlog10(AGWB)subscript𝛼subscript10subscript𝐴GWB\alpha_{*}\equiv\log_{10}(A_{\rm GWB})italic_α start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ≡ roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_A start_POSTSUBSCRIPT roman_GWB end_POSTSUBSCRIPT ).

Assuming this SGWB, we obtain SNR = 4.5 with the current EPTA (DRNew) pulsar configuration, which is compatible with the observations Antoniadis et al. (2023a). On the other hand, we obtain SNR=184SNR184{\rm SNR}=184roman_SNR = 184 for EPTA20 and SNR=292SNR292{\rm SNR}=292roman_SNR = 292 for SKA10.

To validate our framework based on the approximated likelihood introduced in section II.3, we compare our results to the ones obtained with state-of-the-art techniques described in III.3. In Fig. 1, we compare the uncertainties obtained with our simplified FIM approach to actual PTA parameter estimation. We remove the mean from the posterior distribution of both amplitude and tilt showing the deviations from the mean values. In the left panel, we compare results for the current EPTA DR2new configuration, assuming Tobs=10.33subscript𝑇obs10.33T_{\text{obs}}=10.33italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT = 10.33yrs of data with the 25 pulsars adopted by EPTA in their analysis Antoniadis et al. (2023b). In the center and right panels, we compare our forecasted sensitivity to signal parameters with the simulations for EPTA20 and SKA10, with Tobs=20subscript𝑇obs20T_{\text{obs}}=20italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT = 20yrs and Tobs=10subscript𝑇obs10T_{\text{obs}}=10italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT = 10yrs, respectively and Npulsars=50subscript𝑁pulsars50N_{\rm pulsars}=50italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT = 50, built as described in Sec. III. In all cases, we find very good agreement between both results, showing the validity of our assumptions in Sec. II for our purposes.

IV.1 Future effective sensitivity

Using the metrics defined in sec. II.2.2, we can forecast the effective sensitivity achieved by current and future experiments. In particular, we plot Seff(f)subscript𝑆eff𝑓S_{\rm eff}(f)italic_S start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ( italic_f ) Fig. 2 for different detector designs, i.e., EPTA10 and SKA10, translated in the GW energy density using Eq. (10). Notice we plot the effective sensitivity as a function of frequencies also below f1/(10yrs)similar-to𝑓110yrsf\sim 1/(10{\rm yrs})italic_f ∼ 1 / ( 10 roman_y roman_r roman_s ) for presentation purposes. In the same plot, we show the maximum likelihood power-law SGWB for comparison. These results assume the HD function to be fixed and described by Eq. (15) (i.e., we do not infer it from the data).

In both cases, the dashed lines report the putative sensitivity obtained without injecting an SGWB. The effective sensitivity reaches a minimum of around f3similar-to𝑓3f\sim 3italic_f ∼ 3nHz, with a steep high-frequency slope mostly induced by the large WN components. Indeed, in the high-frequency part of the sensitivity curve, we observe ΩGW(f)f5similar-tosubscriptΩGW𝑓superscript𝑓5\Omega_{\rm GW}(f)\sim f^{5}roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT ( italic_f ) ∼ italic_f start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT, as expected from a WN-dominated regime, including frequency-dependent factors coming from the definition of energy density and the response function. Instead, at low frequencies, the sensitivity is degraded with a tilt steeper than 1/fproportional-toabsent1𝑓\propto 1/f∝ 1 / italic_f, expected from the behavior of transmission functions alone, showing the relevance of the red-noise components included in the pulsar noise budget. SKA10 sensitivity greatly surpasses EPTA10 due to the much smaller expected WN (with an improvement that is more than an order of magnitude at high frequencies).

The solid lines report the effective sensitivity obtained including the SGWB background as in Eq. (30). While including the SGWB does not significantly affect the sensitivity of EPTA10, showing that one still falls in the weak signal regime, the SKA10 effective sensitivity strongly degrades, driven by the large SGWB foreground compared to the pulsar noises.

In practice, the take-home message is that searches for subdominant contributions to the SGWB below the solid lines would be inconclusive, at odds with the conclusion drawn considering the effective sensitivity with neglected foregrounds. This is one of the main results of this paper. We will show examples of multi-SGWB analyses in the following sections, corroborating the qualitative conclusions shown here.

Refer to caption
Figure 2: Current sensitivity of the EPTADR2new dataset, denoted as EPTA10, (blue lines), and forecasted future sensitivity achieved by the SKA 10 yr (yellow lines). Both scenarios are reported assuming the presence (absence) of an SGWB in a solid (dashed) line. The black solid line corresponds to the maximum likelihood power-law signal (30).
Refer to caption
Figure 3: Left panel: scaling of the SNR as a function of the observation time Tobssubscript𝑇obsT_{\text{\tiny obs}}italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT assuming a PTA network composed of 30 pulsars. The vertical error bars indicate the standard deviation of the result obtained by varying over a random realization of the PTA array (both for pulsar positions and noises). The black dashed lines indicate the two limiting scaling regimes expected for the SNR. The low-Tobssubscript𝑇obsT_{\rm obs}italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT regime is dictated by our choice of the spectrum with nT=2subscript𝑛𝑇2n_{T}=2italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 2, while the high-Tobssubscript𝑇obsT_{\rm obs}italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT regime is universal. Right panel: scaling of the uncertainties on the amplitude and tilt of the SGWB as a function of observation time.
Refer to caption
Figure 4: Scaling of the SNR (left panel) and uncertainties (right panel) as a function of the number of pulsars, assuming fixed Tobs=20subscript𝑇obs20T_{\text{\tiny obs}}=20italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT = 20yrs. The error band indicates the standard deviation of both results induced by producing multiple realizations of the pulsar sample.

IV.2 Scaling with time and number of pulsars

With the proposed framework, we can forecast the evolution with time and the number of pulsars of both the SNR and measurement precision.

In Fig. 3, we report the results of different random realizations of 30303030 pulsars with variable observation time Tobssubscript𝑇obsT_{\text{obs}}italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT. We indicate with the uncertainty bands the corresponding spread obtained from randomly selecting pulsars with different sky localization (assuming isotropic distribution) and noises.555For both Figs. 3 and 4, we generated 30303030 different realizations for each value of Tobssubscript𝑇obsT_{\text{obs}}italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT and Npulsarssubscript𝑁pulsarsN_{\text{pulsars}}italic_N start_POSTSUBSCRIPT pulsars end_POSTSUBSCRIPT, to estimate the scatter induced by randomly choosing noise parameter for each pulsar. The parameters determining the noise of each pulsar are sampled from distributions built out of the currently observed ones Antoniadis et al. (2023b).

In these plots, we aim to test the scaling of the relevant quantities. For this reason, we show FIM results also for low-SNR values where FIM predictions are unreliable. It should be kept in mind that results obtained in this regime only serve the purpose of testing the scalings and should not be used to deduce the detectability of the signal. Also, in this case, these results assume the HD function to be fixed and described by Eq. (15) (i.e., we do not infer it from the data). We do not expect this assumption to modify the scaling obtained.

Refer to captionRefer to caption
Figure 5: Uncertainties on the HD curve were obtained using the Fisher matrix analysis and assuming a power-law signal. Left panel: constraints on the coefficients of the binned HD function. We divide the angular direction into 7 equally spaced bins. Uncertainties are comparable with the one observed found in EPTADR2 analyses Antoniadis et al. (2023a). By construction, within the FIM formalism, the central value corresponds to the injection. Right panel: Posterior predictive distribution of the HD function obtained from the uncertainties on the coefficients of the Legendre polynomial expansion in Eq. (17).
Refer to captionRefer to caption
Figure 6: Same as Fig. 5, but assuming the SKA10 future configuration.

Focusing on the evolution of SNR, we observe that at small observation times, it grows rapidly as SNRTobs3similar-toabsentsuperscriptsubscript𝑇obs3\sim T_{\text{\tiny obs}}^{3}∼ italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. This can be explained as follows (see analytical considerations reported in Ref. Siemens et al. (2013)). Assuming WN dominates at the initially relatively high frequencies fmin1/Tobssimilar-tosubscript𝑓min1subscript𝑇obsf_{\text{min}}\sim 1/T_{\text{\tiny obs}}italic_f start_POSTSUBSCRIPT min end_POSTSUBSCRIPT ∼ 1 / italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT, and that γ>0𝛾0\gamma>0italic_γ > 0, the integral for the SNR is dominated by the contribution close to fminsubscript𝑓minf_{\text{min}}italic_f start_POSTSUBSCRIPT min end_POSTSUBSCRIPT. Longer observation times allow to probe lower frequencies, enhancing the ratio Sh/SefffminγTobsγsimilar-tosubscript𝑆subscript𝑆effsuperscriptsubscript𝑓min𝛾similar-tosuperscriptsubscript𝑇obs𝛾S_{h}/S_{\text{eff}}\sim f_{\text{min}}^{-\gamma}\sim T_{\text{obs}}^{\gamma}italic_S start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT / italic_S start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT ∼ italic_f start_POSTSUBSCRIPT min end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - italic_γ end_POSTSUPERSCRIPT ∼ italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT. Including the additional pre-factor and frequency summation, this gives SNRTobsγproportional-toabsentsuperscriptsubscript𝑇obs𝛾\propto T_{\text{obs}}^{\gamma}∝ italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT. When the effective sensitivity falls below the GW amplitude at the minimum frequency, one transitions to the intermediate regime, and the above scaling is broken. Eventually, the SNR growth with time converges to Tobsproportional-toabsentsubscript𝑇obs\propto\sqrt{T_{\text{obs}}}∝ square-root start_ARG italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT end_ARG as predicted in the strong signal regime, which we observe on the right side of the plot.

An analogous behavior is observed in the evolution of uncertainties on signal parameters αsubscript𝛼\alpha_{*}italic_α start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT and nTsubscript𝑛𝑇n_{T}italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT. By construction, they are expected to scale inversely with the SNR, in the high SNR regime. We observe that both feature an initially steep decline that converges towards a 1/Tobs1subscript𝑇obs1/\sqrt{T_{\rm obs}}1 / square-root start_ARG italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT end_ARG in the far future. We stress that current PTA experiments are currently probing the intermediate regime between the two asymptotic scalings, seen in the plot around Tobs10similar-tosubscript𝑇obs10T_{\rm obs}\sim 10italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT ∼ 10yrs (see also Pol et al. (2021)). Additionally, we see that the uncertainty on the tilt reaches the scaling regime more slowly because more frequency modes need to be observed to better pin down the spectral tilt compared to the overall amplitude.

In Fig. 4 we show the corresponding scaling as a function of the number of pulsars. In the left panel, we see that the SNR grows as N𝑁\sqrt{N}square-root start_ARG italic_N end_ARG as a function of number of pulsars. This scaling originates from including auto-correlation terms in our analysis (i.e., diagonal terms in the pulsar indices of the covariant matrix). Correspondingly, the uncertainties on SGWB spectral amplitude and tilt decrease as 1/Npulsars1subscript𝑁pulsars1/\sqrt{N_{\rm pulsars}}1 / square-root start_ARG italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT end_ARG. Had we only included the off-diagonal components, featuring the HD correlation, we would have obtained a scaling SNR(IJχIJ2)1/2similar-toSNRsuperscriptsubscript𝐼𝐽superscriptsubscript𝜒𝐼𝐽212{\rm SNR}\sim\left(\sum_{I\neq J}\chi_{IJ}^{2}\right)^{1/2}roman_SNR ∼ ( ∑ start_POSTSUBSCRIPT italic_I ≠ italic_J end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT which behaves as SNR[Npulsars(Npulsars1)]1/2Npulsarssimilar-toSNRsuperscriptdelimited-[]subscript𝑁pulsarssubscript𝑁pulsars112similar-tosubscript𝑁pulsars{\rm SNR}\sim\left[N_{\rm pulsars}(N_{\rm pulsars}-1)\right]^{1/2}\sim N_{\rm pulsars}roman_SNR ∼ [ italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT ( italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT - 1 ) ] start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ∼ italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT.

Refer to captionRefer to caption
Figure 7: Uncertainties on the fifth-order expansion coefficients of the HD curve in Legendre’s polynomials asubscript𝑎a_{\ell}italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT as a function of the number of pulsars Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT. We adopt Nmodes=30subscript𝑁modes30N_{\rm modes}=30italic_N start_POSTSUBSCRIPT roman_modes end_POSTSUBSCRIPT = 30. Left panel: zero noise limit. The 0thsuperscript0th0^{\rm th}0 start_POSTSUPERSCRIPT roman_th end_POSTSUPERSCRIPT and 1stsuperscript1st1^{\rm st}1 start_POSTSUPERSCRIPT roman_st end_POSTSUPERSCRIPT orders are zero, and the corresponding uncertainty scales as 1/Npulsarsproportional-toabsent1subscript𝑁pulsars\propto 1/N_{\rm pulsars}∝ 1 / italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT (indicated by the dashed black line). The higher-order pieces reach the plateau predicted by cosmic variance in Eq. (31). See the discussion in the main text. Right panel: Same as the left panel, but includes noise. The plateau at small values of Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT is reached when information is small enough that our prior on σa=1subscript𝜎subscript𝑎1\sigma_{a_{\ell}}=1italic_σ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 1 becomes dominant.

IV.3 Sensitivity to the HD correlation pattern

In this section, we show how the formalism developed in the current work can also forecast the sensitivity to the HD correlation. This is a crucial observable to identify GW signals and could be used to test modified theories of gravity (e.g., Liang et al. (2023)). For this purpose, we allow χIJsubscript𝜒𝐼𝐽\chi_{IJ}italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT to vary, and, adopting the HD curve parameterizations described in Sec. II.2.1, we assess the ability of different PTA configurations to constrain the expansion coefficients.

First of all, we consider the case of the EPTA10 configuration. In Fig. 5 (left panel), we report the uncertainties on the coefficient bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT of the binned parameterization (16) found using FIM. For this analysis, we considered 7 bins as well as the signal amplitude and slope as free parameters. As one can notice, the uncertainties are large both at narrow and wide angular separations, where fewer pulsar pairs are present, while reaching down to values around Δbi0.3similar-toΔsubscript𝑏𝑖0.3\Delta b_{i}\sim 0.3roman_Δ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ 0.3 at 68%percent6868\%68 % C.L. in the central bins. These results are in good agreement with the posterior distributions reported by the EPTA collaboration Antoniadis et al. (2023a), although they choose unequally spaced bins to evenly distribute the number of pulsar pairs as a function of angular separation.

In the right panel of Fig. 5, we show the posterior predictive distribution of χIJsubscript𝜒𝐼𝐽\chi_{IJ}italic_χ start_POSTSUBSCRIPT italic_I italic_J end_POSTSUBSCRIPT obtained using the uncertainties estimated with FIM for the current EPTA10 configuration. In this case, the FIM analysis adopts the coefficients of the Legendre polynomial expansion (17) up to the sixth order. As expected, relatively tight bounds are obtained in the central portion of the angular separation variable (which contains a larger number of pulsar pairs), with constraints that degrade at the edges of the plot.

Having checked the validity of our framework to estimate the sensitivity to HD correlations, we forecast the corresponding uncertainties obtained with the future SKA10 dataset. In Fig. 6, we show the corresponding posterior on binned expansion parameters (left panel) as well as posterior predictive distribution reconstructed using Legendre expansion (right panel). As expected, the improved SNR ratio in the SKA10 forecast leads to significantly narrower uncertainties on the HD correlation function. In the central bins, one finds uncertainties around Δbi0.04similar-toΔsubscript𝑏𝑖0.04\Delta b_{i}\sim 0.04roman_Δ italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ 0.04 at 68%percent6868\%68 % C.L.

Cosmic variance.

For a Gaussian background, the multipole moments are drawn from a Gaussian distribution with variance asubscript𝑎a_{\ell}italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT defined in Eq. (17). Even neglecting noise, and assuming perfect measurement precision, the estimate of the variance of the distribution is limited by the number of modes observed for each \ellroman_ℓ. As described in Refs. Roebber and Holder (2017); Allen (2023), assuming single mode observation and no noise, the uncertainty on asubscript𝑎a_{\ell}italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT should plateau towards the value

Δa=a2+1.Δsubscript𝑎subscript𝑎21\Delta a_{\ell}=\frac{a_{\ell}}{\sqrt{2\ell+1}}\;.roman_Δ italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT = divide start_ARG italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG 2 roman_ℓ + 1 end_ARG end_ARG . (31)

In reality, multiple frequency modes are observed by PTA, and one expects the limit on the uncertainties due to cosmic variance to scale as Δa(Nmodes)1/2similar-toΔsubscript𝑎superscriptsubscript𝑁modes12\Delta a_{\ell}\sim\left(N_{\rm modes}\right)^{-1/2}roman_Δ italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ∼ ( italic_N start_POSTSUBSCRIPT roman_modes end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT.

In Fig. 7, we show how the uncertainty on asubscript𝑎a_{\ell}italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT scales as a function of the number of pulsars. To better show the qualitative behavior, we first compute the results assuming no noise and 30303030 observed frequencies (left plot) and Tobs=10subscript𝑇obs10T_{\rm obs}=10italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT = 10yr, only including terms up to =66\ell=6roman_ℓ = 6 in the FIM. As one can see, the uncertainty for =0,101\ell=0,1roman_ℓ = 0 , 1 scales as 1/Npulsars1subscript𝑁pulsars1/N_{\rm pulsars}1 / italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT, without showing any significant departure at high Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT. This is because, a0,1=0subscript𝑎010a_{0,1}=0italic_a start_POSTSUBSCRIPT 0 , 1 end_POSTSUBSCRIPT = 0 due to the quadruple nature of GR, and their inference does not suffer from self-noise. On the other hand, the uncertainty on high multiples converges towards a plateau, which is dictated by cosmic variance (31). The plateau is reached at an increasing number of pulsars for higher 22\ell\geq 2roman_ℓ ≥ 2 due to the increasingly larger Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT needed to cover the sky positions with sufficient finer angular resolution to saturate the information available.

In the right panel of Fig. 7, we show analogous results assuming the noise to be present. In this case, we simulate 30 pulsars observed for Tobs=10subscript𝑇obs10T_{\text{obs}}=10italic_T start_POSTSUBSCRIPT obs end_POSTSUBSCRIPT = 10yrs. We see that uncertainties on asubscript𝑎a_{\ell}italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT feature a plateau at low values of Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT, where their posterior distribution is dominated by the prior (assumed to be Gaussian with σa=1subscript𝜎subscript𝑎1\sigma_{a_{\ell}}=1italic_σ start_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 1). When Npulsars10greater-than-or-equivalent-tosubscript𝑁pulsars10N_{\rm pulsars}\gtrsim 10italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT ≳ 10 (incidentally close to the currently analyzed sample of pulsars by EPTA), uncertainties start to decrease, especially at lower multipoles. Qualitatively, we observe the same behavior obtained in the left panel for the idealized case of no noise. The =0,101\ell=0,1roman_ℓ = 0 , 1 uncertainties keep the 1/Npulsars1subscript𝑁pulsars1/N_{\rm pulsars}1 / italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT scaling, while starting from the lowest orders in \ellroman_ℓ (especially =2,323\ell=2,3roman_ℓ = 2 , 3) we notice a flattening of the curve already from Npulsars𝒪(few102)similar-tosubscript𝑁pulsars𝒪fewsuperscript102N_{\rm pulsars}\sim{\cal O}({\rm few}\cdot 10^{2})italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT ∼ caligraphic_O ( roman_few ⋅ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). We do not push the computations to even larger Npulsarssubscript𝑁pulsarsN_{\rm pulsars}italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT for two main reasons: 1) this configuration becomes increasingly optimistic for any future scenario, and 2) to retain a reasonable computational cost.

IV.4 Sensitivity to subdominant SGWBs

In this section, we forecast future sensitivity to subdominant contribution to the nHz SGWB. For presentation purposes, we consider the scenario in which the additional subdominant GWB contribution is described by a lognormal (LN) shape

h2ΩCGW(f)=10αLNe[ln2(f/(10γLNHz))/(2102βLN)].superscript2subscriptΩCGW𝑓superscript10subscript𝛼LNsuperscript𝑒delimited-[]superscript2𝑓superscript10subscript𝛾LNHz2superscript102subscript𝛽LNh^{2}\Omega_{\rm CGW}(f)=10^{\alpha_{\rm LN}}e^{\left[-\ln^{2}(f/(10^{\gamma_{% \rm LN}}{\rm Hz}))/(2\cdot 10^{2\beta_{\rm LN}})\right]}.italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT roman_CGW end_POSTSUBSCRIPT ( italic_f ) = 10 start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT [ - roman_ln start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_f / ( 10 start_POSTSUPERSCRIPT italic_γ start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_Hz ) ) / ( 2 ⋅ 10 start_POSTSUPERSCRIPT 2 italic_β start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) ] end_POSTSUPERSCRIPT . (32)

This should serve as a toy model we adopt to test our framework. Typical new physics scenarios of interest for PTA observations predict the existence of a blue tilted spectrum which reaches a peak close to the nHz frequencies, also to comply with ΔNeffΔsubscript𝑁eff\Delta N_{\rm eff}roman_Δ italic_N start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT upper bounds on cosmological (i.e., from the early universe) GWs, forcing the total primordial GW abundance ΩGWh2<1.6106(ΔNeff/0.28)subscriptΩGWsuperscript21.6superscript106Δsubscript𝑁eff0.28\Omega_{\rm GW}h^{2}<1.6\cdot 10^{-6}\left(\Delta N_{\rm eff}/0.28\right)roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT < 1.6 ⋅ 10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT ( roman_Δ italic_N start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT / 0.28 ) at 95% C.L. Aghanim et al. (2020). Nonetheless, flatter spectra in the nHz range are predicted by cosmic string scenarios, as well as second order induced GWs from enhanced and flat curvature power spectra. In-depth analyses which assume specific spectral shapes motivated by new physics cosmological scenarios (see e.g., Afzal et al. (2023); Antoniadis et al. (2023c); Madge et al. (2023); Figueroa et al. (2023); Ellis et al. (2024)) will be presented elsewhere et al. (2024).

Again, to validate our framework in the case of multiple SGWB contributions, we generate a synthetic dataset as described in Sec. III.2 with the SKA10 configuration (50 pulsars and Tobs=10subscript𝑇obs10T_{\rm obs}=10italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT = 10yrs) of observation of an SGWB which takes contributions from two sources

h2ΩGW=h2ΩGWPL+h2ΩGWLN.superscript2subscriptΩGWsuperscript2superscriptsubscriptΩGWPLsuperscript2superscriptsubscriptΩGWLNh^{2}\Omega_{\rm GW}=h^{2}\Omega_{\rm GW}^{\text{PL}}+h^{2}\Omega_{\rm GW}^{% \text{LN}}.italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT = italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT start_POSTSUPERSCRIPT PL end_POSTSUPERSCRIPT + italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT start_POSTSUPERSCRIPT LN end_POSTSUPERSCRIPT . (33)

Collectively, the injected signal features five parameters, which are fixed to {α,nT,αLN,βLN,γLN}={7.2,2,7,1,7.5}subscript𝛼subscript𝑛𝑇subscript𝛼LNsubscript𝛽LNsubscript𝛾LN7.22717.5\{\alpha_{*},n_{T},\alpha_{\rm LN},\beta_{\rm LN},\gamma_{\rm LN}\}=\{-7.2,2,-% 7,-1,-7.5\}{ italic_α start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT , italic_α start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT , italic_β start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT , italic_γ start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT } = { - 7.2 , 2 , - 7 , - 1 , - 7.5 }. In Fig. 8, we compare the FIM estimates to the state-of-the-art multi-components injection and inference. As in Fig. 1, one observes excellent agreement between both results, validating our framework even in the presence of multiple contributions to the SGWB.

We further explore the sensitivity to an LN subdominant contribution to the SGWB in Fig. 9. We simulate different scenarios in which the amplitude of the LN signal is varied, keeping fixed remaining parameters {α,nT,βLN,γLN}={7.2,2,1,7.5}subscript𝛼subscript𝑛𝑇subscript𝛽LNsubscript𝛾LN7.2217.5\{\alpha_{*},n_{T},\beta_{\rm LN},\gamma_{\rm LN}\}=\{-7.2,2,-1,-7.5\}{ italic_α start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT , italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT , italic_β start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT , italic_γ start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT } = { - 7.2 , 2 , - 1 , - 7.5 }. Uncertainties on PL parameters remain practically constant. This is because most of the injections feature a negligible contribution from the LN bump. Only large values of αLNsubscript𝛼LN\alpha_{\rm LN}italic_α start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT the uncertainty grow slightly, due to the partial contamination around LN peak frequencies. On the other hand, we observe increasing precision on the LN parameters with larger αLNsubscript𝛼LN\alpha_{\rm LN}italic_α start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT. Interestingly, the relative uncertainty on the signal amplitude falls below 30%similar-toabsentpercent30\sim 30\%∼ 30 % around αLN7.4similar-to-or-equalssubscript𝛼LN7.4\alpha_{\rm LN}\simeq-7.4italic_α start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT ≃ - 7.4, indicating detection capability, when the LN amplitude crosses the effective sensitivity.

Refer to caption
Figure 8: Same as Fig. 1, but injecting a SGWB from multiple sources: i) the dominant GW signal parametrized as a PL with α=7.2subscript𝛼7.2\alpha_{*}=-7.2italic_α start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT = - 7.2 and nT=2subscript𝑛𝑇2n_{T}=2italic_n start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT = 2 and ii) subdominant LN bump controlled by αLN=7subscript𝛼LN7\alpha_{\rm LN}=-7italic_α start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT = - 7, βLN=1subscript𝛽LN1\beta_{\rm LN}=-1italic_β start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT = - 1 and γLN=7.5subscript𝛾LN7.5\gamma_{\rm LN}=-7.5italic_γ start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT = - 7.5.
Refer to caption
Figure 9: Left panel: Injected SGWB including PL (purple, Eq. (30)) and LN (cyan, Eq. (32)). The grey solid lines show the effective sensitivity obtained with the future SKA10 configuration (with Npulsars=100subscript𝑁pulsars100N_{\rm pulsars}=100italic_N start_POSTSUBSCRIPT roman_pulsars end_POSTSUBSCRIPT = 100 and Tobs=10subscript𝑇obs10T_{\rm obs}=10italic_T start_POSTSUBSCRIPT roman_obs end_POSTSUBSCRIPT = 10yrs). Different values of the amplitude are shown, varying αLN[9,7]subscript𝛼LN97\alpha_{\rm LN}\in[-9,-7]italic_α start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT ∈ [ - 9 , - 7 ] with regular spacing. Right panel: Relative uncertainties on the model parameters, injecting the same values shown in the left panel. While uncertainties on PL parameters remain practically constant, we see improved accuracy on the LN parameters with larger αLNsubscript𝛼LN\alpha_{\rm LN}italic_α start_POSTSUBSCRIPT roman_LN end_POSTSUBSCRIPT.

V Conclusions

Pulsar timing array experiments provide an unprecedented opportunity to search for the existence of GWs in the nHZ frequency range. This will allow to constrain both astrophysical and cosmological sources of such signatures, with increasing precision. Current data report growing evidence for the existence of a signal with an amplitude ΩGWh2109similar-tosubscriptΩGWsuperscript2superscript109\Omega_{\rm GW}h^{2}\sim 10^{-9}roman_Ω start_POSTSUBSCRIPT roman_GW end_POSTSUBSCRIPT italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT at the best constrained frequencies f1/(10yrs)similar-to𝑓110yrsf\sim 1/(10{\rm yrs})italic_f ∼ 1 / ( 10 roman_y roman_r roman_s ). In the future, longer observation time and the growing number of pulsars available will drastically expand PTA sensitivity.

In this work, we have demonstrated how to achieve fast estimation of detectability and parameter estimation for a stochastic GW signal which one could expect for a particular Pulsar Timing Array. Despite the simplistic assumptions adopted, we find that, compared to a full Bayesian analysis using the software ENTERPRISE on the EPTA DR2new and simulated data, this approach gives reliable results. Moreover, we have demonstrated that this method recovers an expected scaling of the PTA sensitivity with observation time and the number of pulsars. We can also study the spatial correlation in the pulsar’s pairs either using a binned consideration or Legendre decomposition. Finally, the formalism allows us to investigate multi-component stochastic GW signals of different strengths, going beyond the weak signal approximation so far adopted in the literature, providing fast and reliable forecasts on the constraining power of PTA configurations on both astrophysical and cosmological SGWBs. We release the code fastPTA which implements this framework.

If we neglect a pulsar term, extending this work to include the resolvable GW signals from individual sources (e.g., continuous GW (CGW) from inspiralling SMBHBs) would be quite simple. In this case, assuming that the CGW candidates are identified, we can include CGW in the transmission function, which will appear as a notch at the CGW frequency. To demonstrate this, we can write the CGW in the form s=kak,hk(t,fCGW,sky)𝑠subscript𝑘subscript𝑎𝑘subscript𝑘𝑡subscript𝑓CGWskys=\sum_{k}a_{k},h_{k}(t,f_{\rm{CGW}},\rm{sky})italic_s = ∑ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT , italic_h start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_t , italic_f start_POSTSUBSCRIPT roman_CGW end_POSTSUBSCRIPT , roman_sky ) (see Babak and Sesana (2012); Ellis et al. (2012)) which is similar to the timing (linearised) model and will give a contribution to the G𝐺Gitalic_G matrix Hazboun et al. (2019) and, therefore, to the transmission 𝒯𝒯{\cal T}caligraphic_T. Extension of this approach to CGWs with pulsar terms is somewhat more complicated and will be discussed elsewhere.

As promised, we want to say a few words about mimicking the chromatic noises in our approach. We can restore the dependence on the radio frequency, as it will become important for the simultaneous ultra-broad band observations. One possibility on how it can be done is presented in Hazboun et al. (2019) using the DMX model. In the case of Gaussian Process representation, we need to translate it to DMX-compatible representation. The DM variations through the DMX model appear in the transmission function.

We conclude by mentioning possible further future directions:

  • Astrophysical models predict the SGWB comes from the superposition of unresolved signals. In this case, spectral fluctuations due to finite source numbers within frequency bins may appear in the observable band at a relatively "high" frequency. It would be interesting to estimate the future sensitivity as well as frequency resolutions to such spikes, which are not expected in cosmological scenarios.

  • Throughout our work, we have assumed the dominant GWB to be stationary and isotropic. In particular, the effective sensitivity reported in our work assumes this background cannot be reduced. If the signal is of astrophysical origin, with sufficient sensitivity one should be able to resolve individual SMBH mergers and remove some CGW. This could improve the effective sensitivity to subdominant GWBs.

  • With our code, one could also test the sensitivity to different correlation patterns beyond HD (e.g. motivated by modified theories of gravity).

  • Finally, it would be interesting to extend this framework to include SGWB anisotropies (see e.g., Ali-Haïmoud et al. (2020, 2021); Cruz et al. (2024) for work in this direction), going beyond the weak signal approximation.

Acknowledgements.
We thank Bruce Allen, Paul Frederik Depta, and Valerie Domcke for valuable discussions on the HD cosmic variance. MP thanks Rutger van Haasteren for the useful discussion on PTA data analysis. SB acknowledges support from funding from the French National Research Agency (grant ANR21-CE31-0026, project MBH waves), the European Union’s Horizon 2020 research and innovation program under the Marie Skłodowska-Curie grant agreement No. 101007855. MP acknowledges the hospitality of Imperial College London, which provided office space during some parts of this project.

References