This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\mciteErrorOnUnknownfalse

The impact of large-scale galaxy clustering on the variance of the Hellings-Downs correlation: theoretical framework

Nastassia Grimm nastassia.grimm@unige.ch Département de Physique Théorique and Center for Astroparticle Physics, Université de Genève, Quai E. Ansermet 24, CH-1211 Geneve 4, Switzerland    Martin Pijnenburg martin.pijnenburg@unige.ch Département de Physique Théorique and Center for Astroparticle Physics, Université de Genève, Quai E. Ansermet 24, CH-1211 Geneve 4, Switzerland    Giulia Cusin giulia.cusin@iap.fr Institut d’Astrophysique de Paris, UMR-7095 du CNRS, Paris, France Département de Physique Théorique and Center for Astroparticle Physics, Université de Genève, Quai E. Ansermet 24, CH-1211 Geneve 4, Switzerland    Camille Bonvin camille.bonvin@unige.ch Département de Physique Théorique and Center for Astroparticle Physics, Université de Genève, Quai E. Ansermet 24, CH-1211 Geneve 4, Switzerland
(August 2, 2025)
Abstract

While pulsar timing array experiments have recently found evidence for the existence of a stochastic gravitational wave (GW) background, its origin is still unclear. If this background is of astrophysical origin, we expect the distribution of GW sources to follow the one of galaxies. Since galaxies are not perfectly isotropically distributed at large scales, but follow the cosmological large-scale structure, this would lead to an intrinsic anisotropy in the distribution of GW sources. In this work, we develop a formalism to account for this anisotropy, by considering a Gaussian ensemble of sources in each realization of the universe and then taking ensemble averages over all such realizations. We find that large-scale galaxy clustering has no impact on the expectation value of pulsar timing residual correlations, described by the Hellings-Downs curve. However, it introduces a new contribution to the variance of the Hellings-Downs correlation. Hence, the anisotropic distribution of sources contributes to the amount by which the measurements of pulsar timing residual correlations, in our single realization of the universe, may differ from the Hellings-Downs curve.

I Introduction

Recently, pulsar timing arrays (PTAs) were used to deliver evidence for the existence of a stochastic gravitational wave (GW) background NANOGrav:2023gor ; Reardon:2023gzh ; EPTA:2023sfo ; Xu:2023wog . This detection utilizes the angular correlation of timing residuals, i.e. small perturbations in the arrival times of radio pulses caused by the passing of GWs. The origin of the GW background is still unclear, and it might be of both astrophysical or cosmological nature (see e.g. Refs. Shannon:2015ect ; Chen:2019xse ; NANOGrav:2021flc ; NANOGrav:2023hvm ; Vagnozzi:2023lwo ). One possibility to distinguish between these scenarios is to evaluate whether galaxy clustering has an impact on the angular correlation of timing residuals: if sources contributing to the GW background are of astrophysical origin, we expect their distribution to follow the distribution of galaxies. Galaxies are however not perfectly isotropically distributed on the sky: they are grouped in clusters, that are themselves arranged into a large-scale structure. Anisotropies in the galaxy distribution have been observed by various surveys, leading to correlations between galaxy positions up to very large cosmological scales, see e.g. Ref. Blake_2011 ; Howlett:2014opa ; Pezzotta:2016gbo ; Alam2016:1607.03155v1 ; eBOSS:2020yzd . Therefore, any GW background of astrophysical origin is expected to be as well anisotropic: an intrinsic source of anisotropy is due to clustering (the distribution of GW sources follows the one of the large-scale structure), a secondary one is due to the fact that once a GW signal is emitted, it feels the effect of the gravitational potential of structures Cusin:2017fwz ; Cusin:2017mjm ; Cusin:2018rsq . These secondary anisotropies are expected to be subdominant on all scales Pitrou:2019rjz , hence we will focus here on the effect of large-scale galaxy clustering.

Usually, the correlation of pulsar timing residuals due to the passing of GWs, as a function of the pulsar pair separation on the sky, is assumed to be described by the Hellings and Downs (HD) curve Hellings:1983fr . Recent literature has investigated to which extent this curve is precise in the description of correlations of timing residuals Allen:2022dzg ; Allen:2022ksj ; Romano:2023zhb ; Allen:2024rqk ; Bernardo:2022xzl . It has been found that, in practice, PTA observations are not expected to exactly recover the HD curve in any realistic models of the GW background. For instance, the interference of GW sources would cause a departure from this idealized curve Allen:2022bjz ; Allen:2022dzg . In all these studies, the fact that sources are clustered is systematically neglected.111The possibility of anisotropies in the distribution of GW sources in the PTA band has been pointed out in previous work, however either not in the context of the HD correlation Allen:1996gp ; Cusin:2017fwz ; Cusin:2017mjm ; Cusin:2018rsq ; Cusin:2018avf ; Cusin:2019jpv ; Cusin:2019jhg ; Pitrou:2019rjz ; Jenkins:2019nks ; Alonso:2020mva ; renzini2022 , or without targeting the impact on its variance Mingarelli:2013dsa ; Taylor:2013esa ; Gair:2014rwa ; Ali-Haimoud:2016mbv .

In this work, we present, for the first time, a framework to account for the impact of the cosmological large-scale structure on the HD correlation.222Note that in this manuscript, we will use the notion HD correlation to denote the observed correlation of pulsar timing delays, while HD curve denotes the theoretical prediction made by Hellings & Downs in 1983 Hellings:1983fr . This method, novel in the context of PTAs, is based on a two-step averaging approach: first, we describe a Gaussian ensemble of GW sources in a realization of the universe with a given distribution of galaxies. Second, to determine the impact of the galaxy distribution on the HD correlation, we study how it varies not only from ensemble to ensemble of GW sources, but also from realization to realization of the galaxy distribution. Mathematically, the latter is done by taking a second ensemble average over all such realizations. We find that the mean of the HD correlation itself is not affected, which means that anisotropies in the distribution of sources do not generate a systematic bias in the HD correlation. In other words, if we were able to measure the correlations of pulsar timing residuals in many realizations of the universe, we would in average recover the HD curve predicted in Ref. Hellings:1983fr for the case of isotropically distributed sources. However, in practice we have only access to one such realization. To account for this, we calculate the variance of the HD correlation due to the large-scale structure of the Universe, which tells us by how much the measurements in our specific realization of the universe can differ from the mean. In this paper we find that, indeed, large-scale galaxy clustering leads to a new contribution to the HD variance, which needs to be carefully evaluated in light of future observations. Numerical results for specific population models of supermassive black holes are presented in a companion paper Grimm:2024hgi .

In addition to the variance due the clustering of sources, there is as well a variance due to shot noise, i.e. due to the fact that GW sources are not a continuous field, but rather have a discrete Poisson distribution (see e.g. Ref. Jenkins:2019nks ). In this paper we do not take the shot noise contribution into account (see Ref. Allen:2024mtn for a derivation of this contribution for PTAs), and we concentrate only on the modeling of the clustering variance. Consequently, we refer to an isotropic distribution of sources when no clustering is present, and an anisotropic distribution of sources when clustering is accounted for. This use of terminology, although being common in literature, is not fully accurate due to shot noise: since sources have a discrete nature, they will never have a perfectly isotropic distribution, even in the absence of clustering.

The remainder of this work is structured as follows: In section II, we first revise basic notations in the case of an isotropic distribution of GW sources, and then introduce galaxy clustering and discuss the respective ensemble averaging procedure. We present our calculation of the variance of the HD correlation due to large-scale clustering in section III, with our main result given in Eq. (31). We conclude in section IV, and comment on the positivity of this novel contribution to the variance in Appendix A.

II Ensemble averages of GW signals

Before calculating the impact of large-scale structure on the HD correlation, we first determine how it impacts the two-point function of the GW amplitude. In the following, we first discuss the different layers of stochasticity present in our analyses, then discuss the GW signal for an isotropic distribution of sources, and subsequently introduce the impact of galaxy clustering.

II.1 Two different layers of stochasticity

The GW signal is a stochastic quantity, first due to the fact that it arises from an (unknown) superposition of waves, and second because we do not know the spatial distribution of sources. As a consequence, the HD correlation is also a statistical quantity. The aim of this paper is to compute the mean and variance of the HD correlation accounting for both sources of stochasticity. We introduce therefore two types of averages, laying the foundation for the subsequent sections:

  1. (i)

    Ensemble average over GW realizations. The GW signal is treated as a stochastic gravitational wave background (SGWB). It arises indeed from the superposition of a large number of sources, emitting incoherently at distinct frequencies. The total strain in a given direction is stochastic since we do not know the details of the distribution of sources emitting GWs during the observing time TT (e.g. we do not know at which stage of the inspiral a given source is, and this reflects in a stochasticity of both amplitude and phase; orbital planes are also randomly oriented, hence the GW polarization is random; see also Ref. Maggiore:1900zz for a description of the stochastic background). We can therefore not predict the value of the strain, but we can predict its 2-point function (since we know the statistical properties of the sources), determined by the spectral density.

  2. (ii)

    Ensemble average over stochastic initial conditions of cosmological variables, in particular the galaxy density field. The stochasticity of these variables is inherited from their quantum origin during inflation. Since the distribution of GW sources follows the distribution of matter, the total strain obtained by summing the contribution from different sources along the line of sight inherits the stochasticity of the matter distribution. We note that this average (ii) is the usual ensemble average used in cosmology to compute correlation functions and angular power spectra of cosmological observables. We also emphasize that if a GW observable is a stochastic quantity, then the average (i) over GW realizations alone still maintains the stochasticity of the matter distribution.

We emphasize again that, in practice, we only have observational access to one realization of the Universe, i.e. only one realization of the SGWB and of the galaxy density field. Therefore, for any observable, it is important to know its variance in addition to its theoretical expectation value: the variance provides a measure by how much we expect the observation and the theoretical expectation value to differ.

II.2 Isotropic distribution of sources

We start by summarizing the basic concepts and notations (see e.g. Ref. Maggiore:1900zz for a standard reference) necessary to describe the GW signal in absence of anisotropies in the galaxy density (i.e., the average (ii) described above plays no role yet). We work in units where the speed of light is given by c=1c=1 for convenience. We consider a homogeneous and isotropic universe without galaxy clustering and a stochastic GW background, causing a perturbation to the space-time metric:

ds2=dt2+d𝐱2+hij(t,𝐱)dxidxj.\mathrm{d}s^{2}=-\mathrm{d}t^{2}+\mathrm{d}\mathbf{x}^{2}+h_{ij}(t,\mathbf{x})\mathrm{d}x^{i}\mathrm{d}x^{j}\,. (1)

We work in the transverse-traceless synchronous gauge, where the quantity hijh_{ij} can be written, using the plane-wave expansion, as

hij(t,𝐱)=A=+,×dfd𝐧hA(f,𝐧)e2πif(t𝒏𝐱)eijA(𝐧),h_{ij}(t,\mathbf{x})=\sum_{A=+,\times}\int\mathrm{d}f\int\mathrm{d}\mathbf{n}\,h_{A}(f,\mathbf{n}){\rm e}^{2\pi if(t-\boldsymbol{n}\cdot\mathbf{x})}e^{A}_{ij}(\mathbf{n})\,, (2)

where eijA(𝐧)e_{ij}^{A}(\mathbf{n}) is the polarization tensor and hA(f,𝐧)h_{A}(f,\mathbf{n}) denotes the amplitude of the polarization state AA at frequency ff and propagating in direction 𝐧\mathbf{n} (hence originating from direction 𝐧-\mathbf{n}). The amplitudes are complex functions satisfying hA(f,𝐧)=hA(f,𝐧)h_{A}(-f,\mathbf{n})=h_{A}^{\ast}(f,\mathbf{n}). Moreover, by definition of a stochastic background the amplitudes have zero mean: hA(f,𝐧)=0\left\langle h_{A}(f,\mathbf{n})\right\rangle=0. The expectation value of the product of two hA(f,𝐧)h_{A}(f,\mathbf{n}) is however non-zero and given by

hA(f,𝐧)hB(f,𝐧)=12δABδ(ff)S¯h(f)δ2(𝐧,𝐧)4π.\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle=\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})\bar{S}_{h}(f)\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}\,. (3)

The delta functions account for the fact that only GWs with the same polarization state, at the same frequency and coming from the same direction are correlated Cusin:2017mjm . In other words, at each unique polarization state and unique value of ff and 𝐧\mathbf{n}, the quantity hA(f,𝐧)h_{A}(f,\mathbf{n}) is a Gaussian random variable with variance determined by the spectral density S¯h(f)\bar{S}_{h}(f).333This spectral density is related to the isotropic GW energy parameter Ω¯GW\bar{\Omega}_{\text{GW}} via Ω¯GW=πf3S¯h/(2Gρc)\bar{\Omega}_{\text{GW}}=\pi f^{3}\bar{S}_{h}/(2G\rho_{c}), with ρc\rho_{c} the critical energy density of the Universe Maggiore:1900zz . For a stochastic background due to astrophysical sources, the spectral density S¯h(f)\bar{S}_{h}(f) appearing in Eq. (3) is related to the number of sources as well as to their characteristics. It depends on frequency, but by construction it does not depend on directions 𝐧\mathbf{n} given that we assume an isotropic distribution of GW sources. Thus, while hA(f,𝐧)h_{A}(f,\mathbf{n}) generally depends on direction 𝐧\mathbf{n} due to the stochasticity of the GW background, its variance does not unless other forms of anisotropy are present in the Universe.

The expectation value given in Eq. (3) occurs in standard calculations leading to the HD curve, i.e. in the calculation of the mean of timing residual correlations, for a Gaussian ensemble of GW sources. As a next step, we want to extend Eq. (3) to account for the fact that GW sources follow the large-scale structure of the Universe, hence have an anisotropic distribution. This will require taking the ensemble averaging on the left-hand side of Eq. (3) in two steps, in order to distinguish the two different layers of stochasticity described in section II.1.

II.3 Galaxy clustering

In an inhomogeneous universe, the galaxy density distribution can be treated as a random field (see average (ii) in section II.1), directly linked to the primordial fluctuations generated by inflation. We denote by 𝔭g(𝐧,z)\mathfrak{p}_{g}(\mathbf{n},z) this random field at redshift zz and spatial coordinates 𝐱=rz𝐧\mathbf{x}=-r_{z}\mathbf{n}, with rzr_{z} being the comoving distance444The comoving distance is related to redshift as rz=0zdzH(z)1r_{z}=\int_{0}^{z}{\mathrm{d}z^{\prime}}\,{H(z^{\prime})^{-1}}, where H(z)H(z) is the Hubble parameter. The negative sign in the spatial coordinates 𝐱=rz𝐧\mathbf{x}=-r_{z}\mathbf{n} is introduced for notational convenience in subsequent sections, where galaxy density will be related to GWs propagating in direction 𝐧\mathbf{n}, i.e. originating from direction 𝐧-\mathbf{n}.. Our Universe is a particular realization of this random field, that we denote by ρg(𝐧,z)\rho_{g}(\mathbf{n},z). Assuming that inflation generates statistically isotropic fluctuations (i.e. no statistically preferred direction is present in the Universe, which means that statistical expectation values are preserved under rotation on the sky, see e.g. Refs. Armendariz-Picon:2005lfa ; Planck:2015igc ; Hajian:2003qq ), the mean of 𝔭g(𝐧,z)\mathfrak{p}_{g}(\mathbf{n},z) over all possible realizations does not depend on direction:

𝔭g(𝐧,z)=ρ¯g(z).\left\langle\mathfrak{p}_{g}(\mathbf{n},z)\right\rangle=\bar{\rho}_{g}(z)\,. (4)

Due to the ergodic theorem, this mean also represents the average distribution of galaxies over all directions 𝐧\mathbf{n} in one realization ρg(𝐧,z)\rho_{g}(\mathbf{n},z).555In the context of cosmological large-scale structure, the ergodic theorem usually refers to replacing ensemble averages over different realizations of the universe by spatial averages over different regions of the sky, see e.g. Ref. 2012reco.book…..E . The covariance of 𝔭g(𝐧,z)\mathfrak{p}_{g}(\mathbf{n},z) is given by

(𝔭g(𝐧,z)ρ¯g(z))(𝔭g(𝐧,z)ρ¯g(z))=ρ¯g(z)ρ¯g(z)ξg(𝐧𝐧,z,z),\left\langle\left(\mathfrak{p}_{g}(\mathbf{n},z)-\bar{\rho}_{g}(z)\right)\left(\mathfrak{p}_{g}(\mathbf{n}^{\prime},z^{\prime})-\bar{\rho}_{g}(z^{\prime})\right)\right\rangle=\bar{\rho}_{g}(z)\bar{\rho}_{g}(z^{\prime})\xi_{g}(\mathbf{n}\cdot\mathbf{n}^{\prime},z,z^{\prime})\,, (5)

where ξg(𝐧𝐧,z,z)\xi_{g}(\mathbf{n}\cdot\mathbf{n}^{\prime},z,z^{\prime}) is the galaxy 2-point correlation function. Note that, due to statistical isotropy, ξg(𝐧𝐧,z,z)\xi_{g}(\mathbf{n}\cdot\mathbf{n}^{\prime},z,z^{\prime}) depends only on the angle between 𝐧\mathbf{n} and 𝐧\mathbf{n}^{\prime} and on the two redshifts zz, zz^{\prime}. Our aim is to determine how the clustering of galaxies described by the correlation function ξg(𝐧𝐧,z,z)\xi_{g}(\mathbf{n}\cdot\mathbf{n}^{\prime},z,z^{\prime}) impacts the HD correlation. In a given cosmological model, this correlation function can be easily calculated using Boltzmann codes such as CLASS Lesgourgues:2011re . Moreover, it has been measured by various surveys, such as WiggleZ Blake_2011 , SDSS Howlett:2014opa , VIPERS Pezzotta:2016gbo , BOSS Alam2016:1607.03155v1 and eBOSS eBOSS:2020yzd .

II.4 Ensemble average in one specific realization: conditional averaging

The case discussed in section II.2 describes a situation where the distribution of sources is homogeneous and isotropic. This corresponds to an average over all possible density realizations of the universe, which eliminates the directional dependence assuming that the universe is statistically isotropic, i.e. does not have any preferred direction.666Indeed, the average on the left-hand side of Eq. (3) is eliminating all stochasticity, hence the right-hand side is as well non-stochastic. The only way to obtain a direction-dependent spectral density after eliminating stochasticity is by breaking statistical isotropy, i.e. by assuming that there would exist some special direction in our Universe. In practice, anisotropies are typically not specifically mentioned, and the average over all realizations can be treated as an average done at a special realization, 𝔭g(𝐧,z)=ρ¯g(z)\mathfrak{p}_{g}(\mathbf{n},z)=\bar{\rho}_{g}(z), where sources are isotropically distributed. In this spirit, Eq. (3) can be written as

hA(f,𝐧)hB(f,𝐧)|𝔭g(𝐧,z)=ρ¯g(z)=12δABδ(ff)S¯h(f)δ2(𝐧,𝐧)4π.\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\,\big{|}\,\mathfrak{p}_{g}(\mathbf{n},z)=\bar{\rho}_{g}(z)\right\rangle=\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})\bar{S}_{h}(f)\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}\,. (6)

The expression on the left is a conditional ensemble average: it represents the average over all realizations of hAh_{A} (i.e. the expectation value) of the product hA(f,𝐧)hB(f,𝐧)h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}}), assuming a fixed realization of the density field 𝔭g(𝐧,z)=ρ¯g(z)\mathfrak{p}_{g}(\mathbf{n},z)=\bar{\rho}_{g}(z). On the right-hand side, S¯h(f)\bar{S}_{h}(f) determines the variance (over all possible realizations of hAh_{A}) of the field hAh_{A}, in a universe where the galaxy density distribution is perfectly homogeneous. Since the density distribution does not depend on direction in this case, S¯h(f)\bar{S}_{h}(f) is also independent on 𝐧\mathbf{n}. Moreover, for a fixed realization of the density, hAh_{A} is a Gaussian random field.

Let us now consider a different realization that is not perfectly homogeneous and isotropic: 𝔭g(𝐧,z)=ρg(𝐧,z)=ρ¯g(z)(1+δg(𝐧,z))\mathfrak{p}_{g}(\mathbf{n},z)=\rho_{g}(\mathbf{n},z)=\bar{\rho}_{g}(z)(1+\delta_{g}(\mathbf{n},z)). Here, δg(𝐧,z)\delta_{g}(\mathbf{n},z) is the galaxy over- or underdensity, which can thus be positive or negative. In such a realization, Eq. (3) changes to

hA(f,𝐧)hB(f,𝐧)|𝔭g(𝐧,z)=ρg(𝐧,z)=12δABδ(ff)Sh(f,𝐧)δ2(𝐧,𝐧)4π.\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\,\big{|}\,\mathfrak{p}_{g}(\mathbf{n},z)=\rho_{g}(\mathbf{n},z)\right\rangle=\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})S_{h}(f,\mathbf{n})\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}\,. (7)

As before, hA(f,𝐧)h_{A}(f,\mathbf{n}) is a Gaussian random field given the specific realization ρg(𝐧,z)\rho_{g}(\mathbf{n},z), with the variance at each point determined by the spectral density Sh(f,𝐧)S_{h}(f,\mathbf{n}) and no correlations between distinct angles, frequencies or polarization states. The only difference to before is that the spectral density Sh(f,𝐧)S_{h}(f,\mathbf{n}) now depends on direction 𝐧\mathbf{n}, since the distribution of galaxies (and therefore the distribution of GW sources) is not isotropic in this realization of the universe. We can decompose the spectral density Sh(f,𝐧)S_{h}(f,\mathbf{n}) into an isotropic part S¯h(f)\bar{S}_{h}(f), denoting the spectral density in an isotropic realization, and δSh(f,𝐧)\delta S_{h}(f,\mathbf{n}), encoding the departure from isotropy in the realization ρg(𝐧,z)\rho_{g}(\mathbf{n},z),

Sh(f,𝐧)=S¯h(f)+δSh(f,𝐧)=S¯h(f)[1+dzbGW(f,z)δg(𝐧,z)].S_{h}(f,\mathbf{n})=\bar{S}_{h}(f)+\delta S_{h}(f,\mathbf{n})=\bar{S}_{h}(f)\left[1+\int\mathrm{d}z\,b_{\rm GW}(f,z)\delta_{g}(\mathbf{n},z)\right]\,. (8)

Here, bGW(f,z)b_{\rm GW}(f,z) determines the relation between the galaxy overdensity δg(𝐧,z)\delta_{g}(\mathbf{n},z) and the GW spectral density. Since we assume that the distribution of GW sources follows the distribution of galaxies, bGW(f,z)b_{\rm GW}(f,z) does not depend on direction 𝐧\mathbf{n}. Its value, as well as its dependence on redshift and on frequency, is related to the astrophysical model describing the formation of GW sources and will need to be specified once numerical calculations are performed. In this work, to develop a theoretical framework broadly applicable to many astrophysical models, we keep bGW(f,z)b_{\rm GW}(f,z) as a generic quantity.

II.5 Total ensemble averages

Since we aim to determine the variance of the HD correlation due to galaxy clustering, we need to calculate the ensemble average of hAh_{A} over all realizations of the galaxy density field 𝔭g(𝐧,z)\mathfrak{p}_{g}(\mathbf{n},z). For ease of notation, we first introduce a more concise notation for the ensemble average over hh in a fixed realization ρg\rho_{g},

hA(f,𝐧)hB(f,𝐧)h|ρghA(f,𝐧)hB(f,𝐧)|𝔭g(𝐧,z)=ρg(𝐧,z).\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle_{h|\rho_{g}}\equiv\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\,\big{|}\,\mathfrak{p}_{g}(\mathbf{n},z)=\rho_{g}(\mathbf{n},z)\right\rangle\,. (9)

The full ensemble average, i.e. the average over all realizations of hh and all realizations of the density is then given by777Similar two-step averaging procedures have been applied in the context of anisotropies in the energy density of a stochastic GW background, see Refs. Jenkins:2018lvb ; Jenkins:2019nks ; renzini2022 . Here, we apply this procedure, for the first time, to compute the impact of galaxy clustering on the variance of the HD correlation.

hA(f,𝐧)hB(f,𝐧)=hA(f,𝐧)hB(f,𝐧)h|ρg𝔭g.\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle=\left\langle\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle_{h|\rho_{g}}\right\rangle_{\mathfrak{p}_{g}}\,. (10)

Here, we first compute the conditional ensemble average, and then integrate over the probability distribution of 𝔭g(𝐧,z)\mathfrak{p}_{g}(\mathbf{n},z). This way, we obtain

hA(f,𝐧)hB(f,𝐧)=12δABδ(ff)Sh(f,𝐧)δ2(𝐧,𝐧)4π𝔭g=12δABδ(ff)S¯h(f)δ2(𝐧,𝐧)4π,\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle=\left\langle\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})S_{h}(f,\mathbf{n})\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}\right\rangle_{\mathfrak{p}_{g}}=\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})\bar{S}_{h}(f)\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}\,, (11)

where we used Eq. (8) and the fact that δg(𝐧,z)𝔭g=0\left\langle\delta_{g}(\mathbf{n},z)\right\rangle_{\mathfrak{p}_{g}}=0 (since 𝔭g(𝐧,z)=ρ¯g(z)\left\langle\mathfrak{p}_{g}(\mathbf{n},z)\right\rangle=\bar{\rho}_{g}(z)). Hence, the variance of hAh_{A} over all realizations of the density field is indeed the same as the variance of hAh_{A} in a perfectly isotropic realization. This is not surprising since the mean of the density field over all realizations, ρ¯g(z)\bar{\rho}_{g}(z), is itself homogeneous and isotropic.

With the tools developed in this section, we can now determine how the galaxy density field impacts the HD correlations and their variance. For clarity, we provide a summary of our notations for the different kinds of ensemble averages in Table 1.

\langle\dots\rangle Total ensemble average over all realizations of the universe, i.e. over all GW and density field realizations
|𝔭g(𝐧,z)=ρg(𝐧,z)\left\langle\dots\big{|}\mathfrak{p}_{g}(\mathbf{n},z)=\rho_{g}(\mathbf{n},z)\right\rangle Conditional ensemble average over all GW realizations, for a fixed galaxy density realization
h|ρg\left\langle\dots\right\rangle_{h|\rho_{g}} A shorter notation for the one in the line above
𝔭g\langle\dots\rangle_{\mathfrak{p}_{g}} Ensemble average over all density field realizations of the universe
Table 1: We list our notations for the different kinds of ensemble averages appearing in this work.

III Variance of the HD correlation due to galaxy clustering

The HD correlation describes the correlation of timing residuals between pulsars, due to the passing of GWs. Following Ref. Allen:2022dzg , we calculate the redshift, Za(t)Z_{a}(t), induced by the GW on a pulse emitted by pulsar aa located in direction 𝐧a{\bf n}_{a} and observed at time tt on Earth:

Za(t)Adfd𝐧FaA(𝐧)hA(f,𝐧)e2πift(1e2πifτa(1+𝐧𝐧a)).Z_{a}(t)\equiv\sum_{A}\int{\rm d}f\int{\rm d}\mathbf{n}\,F_{a}^{A}(\mathbf{n})h_{A}(f,\mathbf{n}){\rm e}^{2\pi ift}\left(1-{\rm e}^{-2\pi if\tau_{a}(1+\mathbf{n}\cdot{\bf n}_{a})}\right)\,. (12)

Here, τa\tau_{a} is the light travel time between emission and observation, and the antenna pattern functions FaA(𝐧)F_{a}^{A}(\mathbf{n}) are defined as

FaA(𝐧)nainajeijA(𝐧)2(1+𝐧𝐧a).F_{a}^{A}(\mathbf{n})\equiv\frac{n_{a}^{i}n_{a}^{j}e_{ij}^{A}(\mathbf{n})}{2(1+\mathbf{n}\cdot\mathbf{n}_{a})}\,. (13)

Note that the property hA(f,𝐧)=hA(f,𝐧)h_{A}(-f,\mathbf{n})=h_{A}^{\ast}(f,\mathbf{n}) implies that Za(t)Z_{a}(t)\in\mathbb{R}. Since the redshift of a pulsar is related to its timing residual by time differentiation, the goal of PTA observations can be formulated as the search for correlations between the redshifts of pulsars. We define therefore ρab\rho_{ab} as the product of the redshifts of two pulsar signals averaged over the observation time TT,888In this paper, ρab\rho_{ab} will always be used as defined in Eq. (14), and should not be confused with a density.

ρab\displaystyle\rho_{ab} ZaZb¯=ZaZb¯=1TT/2T/2dtZaZb\displaystyle\equiv\overline{Z_{a}Z_{b}}=\overline{Z_{a}Z_{b}^{\ast}}=\frac{1}{T}\int_{-T/2}^{T/2}{\mathrm{d}}t\ Z_{a}Z_{b}^{\ast}
=A,Adfdfd𝐧d𝐧RaA(f,𝐧)RbA(f,𝐧)hA(f,𝐧)hA(f,𝐧)sinc(π(ff)T),\displaystyle=\sum_{A,A^{\prime}}\int\mathrm{d}f\int\mathrm{d}f^{\prime}\int\mathrm{d}\mathbf{n}\int\mathrm{d}\mathbf{n}^{\prime}R^{A}_{a}(f,\mathbf{n})^{\ast}R_{b}^{A^{\prime}}(f^{\prime},\mathbf{n}^{\prime})h_{A}^{\ast}(f,\mathbf{n})h_{A^{\prime}}(f^{\prime},\mathbf{n}^{\prime})\mbox{sinc}\left(\pi(f-f^{\prime})T\right)\,, (14)

where we followed Ref. Allen:2022dzg in the use of the shorthand notation,

RaA(f,𝐧)(1e2πifτa(1+𝐧𝐧a))FaA(𝐧).R^{A}_{a}(f,\mathbf{n})\equiv\left(1-{\rm e}^{-2\pi if\tau_{a}(1+\mathbf{n}\cdot\mathbf{n}_{a})}\right)F^{A}_{a}(\mathbf{n})\,. (15)

The function sinc(x)sin(x)/x\mbox{sinc}(x)\equiv\sin(x)/x in Eq. (14) arises from the time averaging process:

1TT/2T/2dte2πifte2πift=sin(π(ff)T)π(ff)T.\frac{1}{T}\int_{-T/2}^{T/2}{\rm d}t\ {\rm e}^{-2\pi ift}{\rm e}^{2\pi if^{\prime}t}=\frac{\sin\left(\pi(f-f^{\prime})T\right)}{\pi(f-f^{\prime})T}\,. (16)

In the standard case of isotropically distributed sources, the expectation value ρab\left\langle\rho_{ab}\right\rangle over field realizations of hAh_{A} leads to the standard HD angular correlation, as is done e.g. in Ref. Allen:2022dzg . When sources are anisotropically distributed, however, the expectation value of ρab\rho_{ab} depends on the specific realization through the spectral density Sh(f,𝐧)S_{h}(f,\mathbf{n}), as we see from Eq. (7). We can compute the ensemble average of ρab\rho_{ab} over all realizations of the galaxy density, and using Eq. (11), we see immediately that we recover the standard HD curve predicted in Ref. Hellings:1983fr . This is not surprising, since the ensemble average of the galaxy density is itself homogeneous and isotropic. For future reference, we write this property of ρab\rho_{ab} as

ρab=ρabst..\left\langle\rho_{ab}\right\rangle=\left\langle\rho_{ab}\right\rangle^{\text{st.}}\,. (17)

In other words, even when taking anisotropies into account, ρab\left\langle\rho_{ab}\right\rangle takes the “standard” form ρabst.\left\langle\rho_{ab}\right\rangle^{\text{st.}}, namely the one that it has in computations assuming isotropy. The relevant question is to determine to which extent the observed HD correlation in our specific realization of the universe may differ from the expectation value. To assess this, we need to compute the variance of ρab\rho_{ab}. Before proceeding with the calculation, we note that ρab\rho_{ab} is defined involving only one pair of pulsars. In practice, PTA surveys track many pulsars forming a large number of pairs, which needs to be taken into account when evaluating the variance for a realistic scenario Allen:2022ksj ; Allen:2022dzg ; Romano:2023zhb . Indeed, the concepts presented in the previous sections are valid independently of how many pulsars are considered. Here, we demonstrate how our formalism is applied in the case of a single pulsar pair. In our companion paper, Ref. Grimm:2024hgi , we present results for the observational limits of a single as well as infinitely many pulsar pairs.

From Eq. (14) we have

ρab2=A,A,A′′,A′′′dfdfdf′′df′′′d𝐧d𝐧d𝐧′′d𝐧′′′sinc(π(ff)T)sinc(π(f′′f′′′)T)\displaystyle\rho^{2}_{ab}=\sum_{A,A^{\prime},A^{\prime\prime},A^{\prime\prime\prime}}\iiiint\mathrm{d}f\mathrm{d}f^{\prime}\mathrm{d}f^{\prime\prime}\mathrm{d}f^{\prime\prime\prime}\iiiint\mathrm{d}\mathbf{n}\,\mathrm{d}\mathbf{n}^{\prime}\mathrm{d}\mathbf{n}^{\prime\prime}\mathrm{d}\mathbf{n}^{\prime\prime\prime}\mbox{sinc}\left(\pi(f-f^{\prime})T\right)\mbox{sinc}\left(\pi(f^{\prime\prime}-f^{\prime\prime\prime})T\right)
×RaA(f,𝐧)RbA(f,𝐧)RaA′′(f′′,𝐧′′)RbA′′′(f′′′,𝐧′′′)hA(f,𝐧)hA(f,𝐧)hA′′(f′′,𝐧′′)hA′′′(f′′′,𝐧′′′).\displaystyle\times R^{A}_{a}(f,\mathbf{n})^{\ast}R^{A^{\prime}}_{b}(f^{\prime},\mathbf{n}^{\prime})R^{A^{\prime\prime}}_{a}(f^{\prime\prime},\mathbf{n}^{\prime\prime})R^{A^{\prime\prime\prime}}_{b}(f^{\prime\prime\prime},\mathbf{n}^{\prime\prime\prime})^{\ast}h_{A}^{\ast}(f,\mathbf{n})h_{A^{\prime}}(f^{\prime},\mathbf{n}^{\prime})h_{A^{\prime\prime}}(f^{\prime\prime},\mathbf{n}^{\prime\prime})h_{A^{\prime\prime\prime}}^{\ast}(f^{\prime\prime\prime},\mathbf{n}^{\prime\prime\prime})\,. (18)

The variance is given by

ρab2ρab2=ρab2h|ρg𝔭gρabh|ρg𝔭g2.\left\langle\rho^{2}_{ab}\right\rangle-\left\langle\rho_{ab}\right\rangle^{2}=\left\langle\left\langle\rho^{2}_{ab}\right\rangle_{h|\rho_{g}}\right\rangle_{\mathfrak{p}_{g}}-\left\langle\left\langle\rho_{ab}\right\rangle_{h|\rho_{g}}\right\rangle^{2}_{\mathfrak{p}_{g}}\,. (19)

For the inner ensemble average we use Eq. (7), and the fact that the field hAh_{A} is Gaussian given any fixed realization of the density field, as discussed in section II.4. We can thus apply Isserlis’ theorem (also known as Wick’s theorem) to the conditioned field. This theorem implies that for a Gaussian field, correlations of four fields can be expressed as sums of products of correlations of two fields. Our calculation is following Appendix C of Ref. Allen:2022dzg with the extension that now, we need to take the anisotropic term of Sh(f,𝐧)S_{h}(f,\mathbf{n}) in Eq. (8) into account. More specifically, Isserlis’ theorem reads

hA(f,𝐧)hA(f,𝐧)hA′′(f′′,𝐧′′)hA′′′(f′′′,𝐧′′′)=h|ρg\displaystyle\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{A^{\prime}}(f^{\prime},\mathbf{n}^{\prime})h_{A^{\prime\prime}}(f^{\prime\prime},\mathbf{n}^{\prime\prime})h_{A^{\prime\prime\prime}}^{\ast}(f^{\prime\prime\prime},\mathbf{n}^{\prime\prime\prime})\right\rangle{{}_{h|\rho_{g}}}=
hA(f,𝐧)hA(f,𝐧)hA′′(f′′,𝐧′′)hA′′′(f′′′,𝐧′′′)h|ρgh|ρg\displaystyle\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{A^{\prime}}(f^{\prime},\mathbf{n}^{\prime})\right\rangle{{}_{h|\rho_{g}}}\left\langle h_{A^{\prime\prime}}^{\ast}(-f^{\prime\prime},\mathbf{n}^{\prime\prime})h_{A^{\prime\prime\prime}}(-f^{\prime\prime\prime},\mathbf{n}^{\prime\prime\prime})\right\rangle{{}_{h|\rho_{g}}}
+hA(f,𝐧)hA′′(f′′,𝐧′′)hA(f,𝐧)hA′′′(f′′′,𝐧′′′)h|ρgh|ρg\displaystyle+\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{A^{\prime\prime}}(f^{\prime\prime},\mathbf{n}^{\prime\prime})\right\rangle{{}_{h|\rho_{g}}}\left\langle h_{A^{\prime}}^{\ast}(-f^{\prime},\mathbf{n}^{\prime})h_{A^{\prime\prime\prime}}(-f^{\prime\prime\prime},\mathbf{n}^{\prime\prime\prime})\right\rangle{{}_{h|\rho_{g}}}
+hA(f,𝐧)hA′′′(f′′′,𝐧′′′)hA(f,𝐧)hA′′(f′′,𝐧′′)h|ρg,h|ρg\displaystyle+\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{A^{\prime\prime\prime}}(-f^{\prime\prime\prime},\mathbf{n}^{\prime\prime\prime})\right\rangle{{}_{h|\rho_{g}}}\left\langle h_{A^{\prime}}^{\ast}(-f^{\prime},\mathbf{n}^{\prime})h_{A^{\prime\prime}}(f^{\prime\prime},\mathbf{n}^{\prime\prime})\right\rangle{{}_{h|\rho_{g}}}\,, (20)

where negative signs in the arguments appear making use of the relation hA(f,𝐧)=hA(f,𝐧)h_{A}(f,\mathbf{n})=h_{A}^{\ast}(-f,\mathbf{n}). Applying Eq. (7), we find that

hA(f,𝐧)hA(f,𝐧)hA′′(f′′,𝐧′′)hA′′′(f′′′,𝐧′′′)=h|ρg\displaystyle\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{A^{\prime}}(f^{\prime},\mathbf{n}^{\prime})h_{A^{\prime\prime}}(f^{\prime\prime},\mathbf{n}^{\prime\prime})h_{A^{\prime\prime\prime}}^{\ast}(f^{\prime\prime\prime},\mathbf{n}^{\prime\prime\prime})\right\rangle{{}_{h|\rho_{g}}}=
14δAAδA′′A′′′δ2(𝐧,𝐧)4πδ2(𝐧′′,𝐧′′′)4πδ(ff)δ(f′′f′′′)Sh(f,𝐧)Sh(f′′,𝐧′′)\displaystyle\frac{1}{4}\delta_{AA^{\prime}}\delta_{A^{\prime\prime}A^{\prime\prime\prime}}\frac{\delta^{2}(\mathbf{n},\mathbf{n}^{\prime})}{4\pi}\frac{\delta^{2}(\mathbf{n}^{\prime\prime},\mathbf{n}^{\prime\prime\prime})}{4\pi}\delta(f-f^{\prime})\delta(f^{\prime\prime}-f^{\prime\prime\prime})S_{h}(f,\mathbf{n})S_{h}(f^{\prime\prime},\mathbf{n}^{\prime\prime})
+14δAA′′δAA′′′δ2(𝐧,𝐧′′)4πδ2(𝐧,𝐧′′′)4πδ(ff′′)δ(ff′′′)Sh(f,𝐧)Sh(f,𝐧)\displaystyle+\frac{1}{4}\delta_{AA^{\prime\prime}}\delta_{A^{\prime}A^{\prime\prime\prime}}\frac{\delta^{2}(\mathbf{n},\mathbf{n}^{\prime\prime})}{4\pi}\frac{\delta^{2}(\mathbf{n}^{\prime},\mathbf{n}^{\prime\prime\prime})}{4\pi}\delta(f-f^{\prime\prime})\delta(f^{\prime}-f^{\prime\prime\prime})S_{h}(f,\mathbf{n})S_{h}(f^{\prime},\mathbf{n}^{\prime})
+14δAA′′′δAA′′δ2(𝐧,𝐧′′′)4πδ2(𝐧,𝐧′′)4πδ(f+f′′′)δ(f+f′′)Sh(f,𝐧)Sh(f,𝐧).\displaystyle+\frac{1}{4}\delta_{AA^{\prime\prime\prime}}\delta_{A^{\prime}A^{\prime\prime}}\frac{\delta^{2}(\mathbf{n},\mathbf{n}^{\prime\prime\prime})}{4\pi}\frac{\delta^{2}(\mathbf{n}^{\prime},\mathbf{n}^{\prime\prime})}{4\pi}\delta(f+f^{\prime\prime\prime})\delta(f^{\prime}+f^{\prime\prime})S_{h}(f,\mathbf{n})S_{h}(f^{\prime},\mathbf{n}^{\prime})\,. (21)

Whenever products of the type RaARbAR^{A\ast}_{a}R^{A}_{b} are considered in integrals, it is common to use the approximation that τa,bf1\tau_{a,b}f\gg 1 for situations of practical interest, and thus functions with unequal and highly oscillating phases average out:

FaAFbA[1e2πifτa(1+𝐧𝐧a)][1e2πifτb(1+𝐧𝐧b)]FaAFbA(1+δab).F^{A}_{a}F^{A}_{b}\ \left[1-{\rm e}^{2\pi if\tau_{a}(1+\mathbf{n}\cdot\mathbf{n}_{a})}\right]\left[1-{\rm e}^{-2\pi if\tau_{b}(1+\mathbf{n}\cdot\mathbf{n}_{b})}\right]\simeq F^{A}_{a}F^{A}_{b}\ (1+\delta_{ab})\,. (22)

This approximation is equivalent to considering only the correlation of the Earth-term for two distinct pulsars. The contribution of the pulsar term is only relevant for the auto-correlation, i.e. with identical pulsars a=ba=b, and falls off fast for non-zero angular or radial separations (see e.g. Ref. Mingarelli:2014xfa or Appendix C.2 of Ref. Allen:2022dzg for the contributions of pulsar terms). Thus, this term can be neglected when aba\neq b. The auto-correlation (a=ba=b) term, on the other hand, carries contributions from the Earth and pulsar terms, and therefore the corresponding terms are twice as large as those for aba\neq b.

With this further approximation, substituting Eq. (21) in Eq. (18) leads to

ρab2=h|ρg\displaystyle\left\langle\rho_{ab}^{2}\right\rangle{{}_{h|\rho_{g}}}= 14dfdfd𝐧4πSh(f,𝐧)χab(𝐧)d𝐧4πSh(f,𝐧)χab(𝐧)\displaystyle\frac{1}{4}\iint\mathrm{d}f\,\mathrm{d}f^{\prime}\int\frac{\mathrm{d}\mathbf{n}}{{4\pi}}\,S_{h}(f,\mathbf{n})\chi_{ab}(\mathbf{n})\int\frac{\mathrm{d}\mathbf{n}^{\prime}}{{4\pi}}\,S_{h}(f^{\prime},\mathbf{n}^{\prime})\chi_{ab}(\mathbf{n}^{\prime})
+dfdfsinc2(π(ff)T)d𝐧4πSh(f,𝐧)χaa(𝐧)d𝐧4πSh(f,𝐧)χbb(𝐧)\displaystyle+\iint\mathrm{d}f\,\mathrm{d}f^{\prime}\mathrm{sinc}^{2}\left(\pi(f-f^{\prime})T\right)\int\frac{\mathrm{d}\mathbf{n}}{{4\pi}}\,S_{h}(f,\mathbf{n})\chi_{aa}(\mathbf{n})\int\frac{\mathrm{d}\mathbf{n}^{\prime}}{{4\pi}}\,S_{h}(f^{\prime},\mathbf{n}^{\prime})\chi_{bb}(\mathbf{n}^{\prime})
+14dfdfsinc2(π(ff)T)d𝐧4πSh(f,𝐧)χab(𝐧)d𝐧4πSh(f,𝐧)χab(𝐧),\displaystyle+\frac{1}{4}\iint\mathrm{d}f\,\mathrm{d}f^{\prime}\mathrm{sinc}^{2}\left(\pi(f-f^{\prime})T\right)\int\frac{\mathrm{d}\mathbf{n}}{{4\pi}}\,S_{h}(f,\mathbf{n})\chi_{ab}(\mathbf{n})\int\frac{\mathrm{d}\mathbf{n}^{\prime}}{{4\pi}}\,S_{h}(f^{\prime},\mathbf{n}^{\prime})\chi_{ab}(\mathbf{n}^{\prime})\,, (23)

where

χab(𝐧)=A=+,×FaA(𝐧)FbA(𝐧),\chi_{ab}(\mathbf{n})=\sum_{A=+,\times}F_{a}^{A}(\mathbf{n})F_{b}^{A}(\mathbf{n})\,, (24)

and we focused on non-vanishing pulsar pair separations only, i.e. ρab\rho_{ab} with aba\neq b (note that terms χaa,χbb\propto\chi_{aa},\chi_{bb} may nevertheless appear, for which we consider contributions from both the pulsar and Earth terms as discussed above).

The GW spectral density, Sh(f,𝐧)S_{h}(f,\mathbf{n}), appearing in Eq. (23) can be expressed as the sum of an isotropic and anisotropic contribution, see Eq. (8). We therefore split the expression for ρab2h|ρg\left\langle\rho^{2}_{ab}\right\rangle{{}_{h|\rho_{g}}} into a “standard” and a “clustering” part,

ρab2=h|ρgρab2h|ρgst.+ρab2h|ρgclust..\displaystyle\left\langle\rho_{ab}^{2}\right\rangle{{}_{h|\rho_{g}}}=\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{st.}}_{h|\rho_{g}}+\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{clust.}}_{h|\rho_{g}}\,. (25)

The standard terms,

ρab2h|ρgst.=\displaystyle\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{st.}}_{h|\rho_{g}}= 14dfdfd𝐧4πS¯h(f)χab(𝐧)d𝐧4πS¯h(f)χab(𝐧)\displaystyle\frac{1}{4}\iint\mathrm{d}f\,\mathrm{d}f^{\prime}\int\frac{\mathrm{d}\mathbf{n}}{{4\pi}}\,\bar{S}_{h}(f)\chi_{ab}(\mathbf{n})\int\frac{\mathrm{d}\mathbf{n}^{\prime}}{{4\pi}}\,\bar{S}_{h}(f^{\prime})\chi_{ab}(\mathbf{n}^{\prime})
+dfdfsinc2(π(ff)T)d𝐧4πS¯h(f)χaa(𝐧)d𝐧4πS¯h(f)χbb(𝐧)\displaystyle+\iint\mathrm{d}f\,\mathrm{d}f^{\prime}\mathrm{sinc}^{2}\left(\pi(f-f^{\prime})T\right)\int\frac{\mathrm{d}\mathbf{n}}{{4\pi}}\,\bar{S}_{h}(f)\chi_{aa}(\mathbf{n})\int\frac{\mathrm{d}\mathbf{n}^{\prime}}{{4\pi}}\,\bar{S}_{h}(f^{\prime})\chi_{bb}(\mathbf{n}^{\prime})
+14dfdfsinc2(π(ff)T)d𝐧4πS¯h(f)χab(𝐧)d𝐧4πS¯h(f)χab(𝐧),\displaystyle+\frac{1}{4}\iint\mathrm{d}f\,\mathrm{d}f^{\prime}\mathrm{sinc}^{2}\left(\pi(f-f^{\prime})T\right)\int\frac{\mathrm{d}\mathbf{n}}{{4\pi}}\,\bar{S}_{h}(f)\chi_{ab}(\mathbf{n})\int\frac{\mathrm{d}\mathbf{n}^{\prime}}{{4\pi}}\,\bar{S}_{h}(f^{\prime})\chi_{ab}(\mathbf{n}^{\prime})\,, (26)

are exactly those specified in Eq. (C25) of Ref. Allen:2022dzg . The clustering terms, on the other hand, arise from the anisotropic part of Sh(f,𝐧)S_{h}(f,\mathbf{n}) and are given by

ρab2h|ρgclust.=\displaystyle{\left\langle\rho^{2}_{ab}\right\rangle^{\mathrm{clust.}}_{h|\rho_{g}}}= dzdz[G(z)G(z)+Γ(z,z)]d𝐧4πd𝐧4πδg(𝐧,z)δg(𝐧,z)χab(𝐧)χab(𝐧)\displaystyle\iint\mathrm{d}z\,\mathrm{d}z^{\prime}\left[G(z)G(z^{\prime})+\Gamma(z,z^{\prime})\right]\iint\frac{\mathrm{d}\mathbf{n}}{4\pi}\frac{\mathrm{d}\mathbf{n}^{\prime}}{4\pi}\delta_{g}(\mathbf{n},z)\delta_{g}(\mathbf{n}^{\prime},z^{\prime})\chi_{ab}(\mathbf{n})\chi_{ab}(\mathbf{n}^{\prime})
+4dzdzΓ(z,z)d𝐧4πd𝐧4πδg(𝐧,z)δg(𝐧,z)χaa(𝐧)χbb(𝐧)\displaystyle+4\iint\mathrm{d}z\,\mathrm{d}z^{\prime}\,\Gamma(z,z^{\prime})\iint\frac{\mathrm{d}\mathbf{n}}{4\pi}\frac{\mathrm{d}\mathbf{n}^{\prime}}{4\pi}\delta_{g}(\mathbf{n},z)\delta_{g}(\mathbf{n}^{\prime},z^{\prime})\chi_{aa}(\mathbf{n})\chi_{bb}(\mathbf{n}^{\prime})
+terms linear in δg(𝐧,z),\displaystyle+\mbox{terms linear in $\delta_{g}(\mathbf{n},z)$}\,, (27)

where

G(z)\displaystyle G(z) 0dfS¯h(f)bGW(f,z),\displaystyle\equiv\int_{0}^{\infty}\mathrm{d}f\,\bar{S}_{h}(f)b_{\rm GW}(f,z)\,,
Γ(z,z)\displaystyle\Gamma(z,z^{\prime}) 14dfdfsinc2(π(ff)T)S¯h(f)S¯h(f)bGW(f,z)bGW(f,z),\displaystyle\equiv\frac{1}{4}\iint\mathrm{d}f\,\mathrm{d}f^{\prime}\mbox{sinc}^{2}\left(\pi(f-f^{\prime})T\right)\bar{S}_{h}(f)\bar{S}_{h}(f^{\prime})b_{\rm GW}(f,z)b_{\rm GW}({f^{\prime}},z^{\prime})\,, (28)

are measures of the strain weighted by the quantity bGW(f,z)b_{\rm GW}(f,z) relating galaxy overdensity and GW intensity.999We note that, in the case where bGW(f,z)=1b_{\rm GW}(f,z)=1, GG and Γ\Gamma respectively correspond to the quantities h2h^{2} and 𝒽4\mathscr{h}^{4} defined in Ref. Allen:2022dzg . We note that, while terms linear in δg\delta_{g} appear in Eq. (31), they vanish in the full ensemble average since δg(𝐧,z)𝔭g=0\left\langle\delta_{g}(\mathbf{n},z)\right\rangle_{\mathfrak{p}_{g}}=0.

Taking the ensemble average of Eqs. (26) and (27) over all realizations of the galaxy density field leads to

ρab2=ρab2h|ρg𝔭g=ρab2st.+ρab2clust.,\displaystyle\left\langle\rho_{ab}^{2}\right\rangle=\left\langle\left\langle\rho_{ab}^{2}\right\rangle{{}_{h|\rho_{g}}}\right\rangle_{\mathfrak{p}_{g}}=\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{st.}}+\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{clust.}}\,, (29)

where, for the standard terms which are independent of the density field,

ρab2st.=ρab2h|ρgst.\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{st.}}=\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{st.}}_{h|\rho_{g}} (30)

and, for the clustering terms,

ρab2clust.=\displaystyle\left\langle\rho^{2}_{ab}\right\rangle^{\mathrm{clust.}}= dzdz[G(z)G(z)+Γ(z,z)]d𝐧4πd𝐧4πξg(𝐧𝐧,z,z)χab(𝐧)χab(𝐧)\displaystyle\iint\mathrm{d}z\,\mathrm{d}z^{\prime}\left[G(z)G(z^{\prime})+\Gamma(z,z^{\prime})\right]\iint\frac{\mathrm{d}\mathbf{n}}{4\pi}\frac{\mathrm{d}\mathbf{n}^{\prime}}{4\pi}\xi_{g}(\mathbf{n}\cdot\mathbf{n}^{\prime},z,z^{\prime})\chi_{ab}(\mathbf{n})\chi_{ab}(\mathbf{n}^{\prime})
+4dzdzΓ(z,z)d𝐧4πd𝐧4πξg(𝐧𝐧,z,z)χaa(𝐧)χbb(𝐧).\displaystyle+4\iint\mathrm{d}z\,\mathrm{d}z^{\prime}\Gamma(z,z^{\prime})\iint\frac{\mathrm{d}\mathbf{n}}{4\pi}\frac{\mathrm{d}\mathbf{n}^{\prime}}{4\pi}\xi_{g}(\mathbf{n}\cdot\mathbf{n}^{\prime},z,z^{\prime})\chi_{aa}(\mathbf{n})\chi_{bb}(\mathbf{n}^{\prime})\,. (31)

Here, we have used Eq. (5) in the form δg(𝐧,z)δg(𝐧,z)=ξg(𝐧𝐧,z,z)\left\langle\delta_{g}(\mathbf{n},z)\delta_{g}(\mathbf{n}^{\prime},z^{\prime})\right\rangle=\xi_{g}(\mathbf{n}\cdot\mathbf{n}^{\prime},z,z^{\prime}). Given Eq. (17), which states that ρab\left\langle\rho_{ab}\right\rangle is unaffected by clustering, we have

ρab2ρab2=ρab2st.(ρabst.)2+ρab2clust..\left\langle\rho^{2}_{ab}\right\rangle-\left\langle\rho_{ab}\right\rangle^{2}=\left\langle\rho^{2}_{ab}\right\rangle^{\text{st.}}-\left(\left\langle\rho_{ab}\right\rangle^{\text{st.}}\right)^{2}+\left\langle\rho^{2}_{ab}\right\rangle^{\mathrm{clust.}}\,. (32)

Eq. (31) presents the main result of our work: an analytical expression encoding the effect of galaxy clustering (described by the galaxy correlation function ξg\xi_{g}) on the HD variance. In other words, this provides a measure of how much we expect the measured correlation of timing residuals to differ, due to galaxy clustering, from the idealized HD curve in a real measurement. We comment on the positivity of the contribution ρab2clust.\left\langle\rho^{2}_{ab}\right\rangle^{\mathrm{clust.}} to the total HD variance in Appendix A.

We close this section by noting that, instead of the galaxy correlation function, an alternative measure of galaxy clustering is provided by the angular power spectrum Cl(z,z)C_{l}(z,z^{\prime}),

ξg(𝐧𝐧,z,z)=14πl(2l+1)Cl(z,z)𝒫l(𝐧𝐧),\xi_{g}(\mathbf{n}\cdot\mathbf{n}^{\prime},z,z^{\prime})=\frac{1}{4\pi}\sum_{l}(2l+1)C_{l}(z,z^{\prime})\mathcal{P}_{l}(\mathbf{n}\cdot\mathbf{n}^{\prime})\,, (33)

with 𝒫l\mathcal{P}_{l} being the Legendre polynomials. This leads to

ρab2clust.=\displaystyle\left\langle\rho^{2}_{ab}\right\rangle^{\mathrm{clust.}}= dzdz[G(z)G(z)+Γ(z,z)]l2l+14πCl(z,z)d𝐧4πd𝐧4π𝒫l(𝐧𝐧)χab(𝐧)χab(𝐧)\displaystyle\iint\mathrm{d}z\,\mathrm{d}z^{\prime}\left[G(z)G(z^{\prime})+\Gamma(z,z^{\prime})\right]\sum_{l}\frac{2l+1}{4\pi}C_{l}(z,z^{\prime})\iint\frac{\mathrm{d}\mathbf{n}}{4\pi}\frac{\mathrm{d}\mathbf{n}^{\prime}}{4\pi}\,\mathcal{P}_{l}(\mathbf{n}\cdot\mathbf{n}^{\prime})\chi_{ab}(\mathbf{n})\chi_{ab}(\mathbf{n}^{\prime})
+4dzdzΓ(z,z)l2l+14πCl(z,z)d𝐧4πd𝐧4π𝒫l(𝐧𝐧)χaa(𝐧)χbb(𝐧).\displaystyle+4\iint\mathrm{d}z\,\mathrm{d}z^{\prime}\,\Gamma(z,z^{\prime})\sum_{l}\frac{2l+1}{4\pi}C_{l}(z,z^{\prime})\iint\frac{\mathrm{d}\mathbf{n}}{4\pi}\frac{\mathrm{d}\mathbf{n}^{\prime}}{4\pi}\mathcal{P}_{l}(\mathbf{n}\cdot\mathbf{n}^{\prime})\chi_{aa}(\mathbf{n})\chi_{bb}(\mathbf{n}^{\prime})\,. (34)

As a further alternative, one can also choose to work with the galaxy power spectrum Pg(k,z,z)P_{g}(k,z,z^{\prime}),

δg(𝐤,z)δg(𝐤,z)=(2π)3δD(𝐤𝐤)Pg(k,z,z).\left\langle\delta_{g}(\mathbf{k},z)\delta_{g}^{\ast}(\mathbf{k}^{\prime},z^{\prime})\right\rangle=(2\pi)^{3}\delta_{D}(\mathbf{k}-\mathbf{k^{\prime}})P_{g}(k,z,z^{\prime})\,. (35)

Here, δg(𝐤,z)\delta_{g}(\mathbf{k},z) is the 3D Fourier transform of the galaxy overdensity δg(𝐧,z)\delta_{g}(\mathbf{n},z). Using the galaxy power spectrum, Eq. (31) can be simplified and expressed as one instead of two integrals over redshift by using the Limber approximation limber1953analysis , a commonly used approximation in large-scale structure observables. In particular, following Ref. Bartelmann:1999yn , we obtain

ρab2clust.=\displaystyle\left\langle\rho^{2}_{ab}\right\rangle^{\mathrm{clust.}}= dz[G2(z)+Γ(z,z)]H(z)kdk2πPg(k,z)d𝐧4πd𝐧4πχab(𝐧)χab(𝐧)J0(krz|𝐧𝐧|)\displaystyle\int\mathrm{d}z\left[{G^{2}(z)}+\Gamma(z,z)\right]{H(z)}\int\frac{k\,\mathrm{d}k}{2\pi}P_{g}(k,z)\iint\frac{\mathrm{d}\mathbf{n}}{4\pi}\frac{\mathrm{d}\mathbf{n}^{\prime}}{4\pi}\chi_{ab}(\mathbf{n})\chi_{ab}(\mathbf{n}^{\prime})J_{0}\big{(}kr_{z}\,{|\mathbf{n}-\mathbf{n}^{\prime}|}\big{)}
+4dzΓ(z,z)H(z)kdk2πPg(k,z)d𝐧4πd𝐧4πχaa(𝐧)χbb(𝐧)J0(krz|𝐧𝐧|),\displaystyle+4\int\mathrm{d}z\,\Gamma(z,z){H(z)}\int\frac{k\,\mathrm{d}k}{2\pi}P_{g}(k,z)\iint\frac{\mathrm{d}\mathbf{n}}{{4\pi}}\frac{\mathrm{d}\mathbf{n}^{\prime}}{{4\pi}}\chi_{aa}(\mathbf{n})\chi_{bb}(\mathbf{n}^{\prime})J_{0}\big{(}kr_{z}\,{|\mathbf{n}-\mathbf{n}^{\prime}|}\big{)}\,, (36)

with J0J_{0} denoting the 0th order Bessel function and H(z)H(z) being the Hubble parameter.101010The factor H(z)H(z) appears due to the relation dz=H(z)dr\mathrm{d}z=H(z)\mathrm{d}r, i.e. the change of integration variable from redshift zz to comoving distance rr, which is needed to apply the Limber approximation (cf. Eq. (2.83) in Ref. Bartelmann:1999yn ).

IV Conclusion

This work introduces a rigorous theoretical framework to assess how the HD correlation, describing pulsar pair correlations caused by a stochastic GW background of astrophysical origin, is affected by anisotropies in the distribution of sources due to large-scale galaxy clustering. The presence of such anisotropies is naturally expected, since the GW sources (e.g. supermassive black holes at the center of galaxies) are expected to follow the distribution of matter in the Universe, which is itself anisotropically distributed.

We showed that to properly account for such anisotropies, it is necessary to consider not only ensemble averages over realizations of GW sources, but also ensemble averages over realizations of the galaxy density field. Modeling fluctuations around an isotropic universe by Gaussian random fields and performing ensemble averaging taking two layers of stochasticity into account, we find that the expectation value of the HD correlation is unchanged with respect to that of an isotropic universe. However, we find that anisotropies lead to a new contribution to the variance of the correlation, which has not been considered in previous work Allen:2022dzg ; Allen:2022ksj ; Romano:2023zhb ; Allen:2024rqk . Thereby, large-scale galaxy clustering impacts how much one expects observed timing residual correlations to deviate from the idealized HD curve.

This contribution is important to assess in view of future observations. While the present-day PTA data has error bars sufficiently large to fully overlap with the HD curve NANOGrav:2023gor , the precision of observations is expected to improve babak2024forecasting . Correctly modeling the variance of the HD correlation is therefore important in view of future observations. Indeed, a GW background of cosmological origin is only marginally affected by clustering (i.e. only via relativistic line-of-sight effects, which give very small contributions that we neglect in this work). As such, the clustering induced contribution to the variance evaluated in this work, which is only present in the case of an astrophysical origin, is a new interesting observable which may provide us with precious information on the nature of the stochastic GW background. Moreover, since various work on PTA observations has investigated the possibility to search for new physics (see e.g. Refs. liang2023test ; NANOGrav:2023hvm ; omiya2023hellings ), it is crucial to properly evaluate timing residual correlations accounting for all fundamental contributions to their variance, to enable an accurate comparison between future high-precision data and models.

In a companion paper Grimm:2024hgi , we apply our theoretical formalism to specific population models of supermassive black holes, considering as well the scenario of many pulsar pairs. In particular, in order to precisely quantify the size of the clustering induced variance, we need to determine how the distribution of sources is linked to the distribution of galaxies. This depends closely on the details of the source formation and evolution. We expect that different formation scenarios, leading to different source distributions, may result in different magnitudes of the HD variance, thus influencing how much we expect observations of the HD correlation to differ from the idealized HD curve. This work, in combination with its companion paper Grimm:2024hgi , provides a new framework to study the link between two important observational probes of the properties of the Universe: the large-scale clustering of galaxies and the stochastic GW background observed via PTAs.

Acknowledgements

We acknowledge discussions with Bruce Allen. N.G. and C.B. acknowledge funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (Grant agreement No. 863929; project title “Testing the law of gravity with novel large-scale structure observables”). C.B. is also supported by the Swiss National Science Foundation. The work of G.C. is supported by CNRS and G.C. and M.P. acknowledge support from the Swiss National Science Foundation (Ambizione grant, “Gravitational wave propagation in the clustered universe”).

References

Appendix A Positivity of the clustering contribution to the HD variance

Intuitively, we expect the fact that we can observe only one universe, i.e. one realization of the galaxy density field, to increase the deviation from the HD curve that one measures in a real observation. Indeed, the positivity of ρab2clust.\left\langle\rho^{2}_{ab}\right\rangle^{\mathrm{clust.}} can be shown by writing the wave random field hAh_{A} as a superposition of two independent random fields: the background field hAst.h_{A}^{\mathrm{st.}} corresponding to the absence of anisotropies, and a perturbation δhA\delta h_{A} due to the presence of anisotropies in the galaxy density. More precisely, we introduce

hA(f,𝐧)=hAst.(f,𝐧)+δhA(f,𝐧),h_{A}(f,\mathbf{n})=h_{A}^{\mathrm{st.}}(f,\mathbf{n})+\delta h_{A}(f,\mathbf{n})\,, (37)

and the ensemble averages over realizations of hAh_{A} are treated as joint averages over realizations of both hAst.h^{\mathrm{st.}}_{A} and δhA\delta h_{A}. For any realization ρg\rho_{g} of the density field, the means of background and perturbation wave fields are given by

hAst.h|ρg=hAst.(hst.,δh)|ρg=0,δhAh|ρg=δhA(hst.,δh)|ρg=0,\left\langle h_{A}^{\mathrm{st.}}\right\rangle_{h|\rho_{g}}=\left\langle h_{A}^{\mathrm{st.}}\right\rangle_{(h^{\mathrm{st.}},\delta h)|\rho_{g}}=0\,,\qquad\left\langle\delta h_{A}\right\rangle_{h|\rho_{g}}=\left\langle\delta h_{A}\right\rangle_{(h^{\mathrm{st.}},\delta h)|\rho_{g}}=0\,, (38)

and the correlations by

hAst.(f,𝐧)hBst.(f,𝐧)h|ρg=12δABδ(ff)S¯h(f)δ2(𝐧,𝐧)4π,\left\langle h_{A}^{{\mathrm{st.}}\ast}(f,\mathbf{n})h^{\mathrm{st.}}_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle_{h|\rho_{g}}=\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})\bar{S}_{h}(f)\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}\,, (39)

and

δhA(f,𝐧)δhB(f,𝐧)h|ρg\displaystyle\left\langle\delta h_{A}^{\ast}(f,\mathbf{n})\delta h_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle_{h|\rho_{g}} =12δABδ(ff)δSh(f,𝐧)δ2(𝐧,𝐧)4π\displaystyle=\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})\delta S_{h}(f,\mathbf{n})\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}
=12δABδ(ff)S¯h(f)dzbGW(f,z)δg(𝐧,z)δ2(𝐧,𝐧)4π.\displaystyle=\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})\bar{S}_{h}(f)\int\mathrm{d}z\,b_{\rm GW}(f,z)\delta_{g}(\mathbf{n},z)\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}\,. (40)

Moreover, the independence of (hAst.,δhA)(h_{A}^{\mathrm{st.}},\delta h_{A}) makes any average over their product vanish,

δhA(f,𝐧)hBst.(f,𝐧)h|ρg=δhA(f,𝐧)h|ρghBst.(f,𝐧)h|ρg=0.\left\langle\delta h_{A}(f,\mathbf{n})h^{\mathrm{st.}}_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle_{h|\rho_{g}}=\left\langle\delta h_{A}(f,\mathbf{n})\right\rangle_{h|\rho_{g}}\left\langle h^{\mathrm{st.}}_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle_{h|\rho_{g}}=0\,. (41)

Each of the field hAst.h_{A}^{\rm st.} and δhA\delta h_{A} is, for a given density realization ρg\rho_{g} of the universe, considered to be Gaussian, i.e. fully specified by Eqs. (38)–(40). With these properties, the full field hAh_{A} is as well Gaussian for each density field realization ρg\rho_{g}, and obeys

hA(f,𝐧)h|ρg=0,hA(f,𝐧)hB(f,𝐧)h|ρg=12δABδ(ff)Sh(f,𝐧)δ2(𝐧,𝐧)4π.\left\langle h_{A}(f,\mathbf{n})\right\rangle_{h|\rho_{g}}=0,\qquad\left\langle h_{A}^{\ast}(f,\mathbf{n})h_{B}(f^{\prime},\mathbf{n^{\prime}})\right\rangle_{h|\rho_{g}}=\frac{1}{2}\delta_{AB}\,\delta(f-f^{\prime})S_{h}(f,\mathbf{n})\frac{\delta^{2}(\mathbf{n},\mathbf{n^{\prime}})}{4\pi}\,. (42)

Thus, the expectation value ρab=ρabh|ρg𝔭g\left\langle\rho_{ab}\right\rangle=\big{\langle}\left\langle\rho_{ab}\right\rangle_{h|\rho_{g}}\big{\rangle}_{\mathfrak{p}_{g}} remains unchanged and equals the HD curve predicted in Ref. Hellings:1983fr .

Let us now consider ρab2\rho_{ab}^{2}, the quantity needed to calculate the variance of the HD correlation. Splitting the wave field as h=hst.+δhh=h^{\mathrm{st.}}+\delta h leads to ρab2\rho_{ab}^{2} containing the following types of terms (with arguments suppressed for conciseness):

ρab2\displaystyle\rho_{ab}^{2}\sim\ 𝒪(hAst.hAst.hA′′st.hA′′′st.)+𝒪(δhAδhAδhA′′δhA′′′)\displaystyle\mathcal{O}(h_{A}^{{\mathrm{st.}}\ast}h^{\mathrm{st.}}_{A^{\prime}}h^{\mathrm{st.}}_{A^{\prime\prime}}h_{A^{\prime\prime\prime}}^{{\mathrm{st.}}\ast})+\mathcal{O}(\delta h_{A}^{\ast}\delta h_{A^{\prime}}\delta h_{A^{\prime\prime}}\delta h_{A^{\prime\prime\prime}}^{\ast})
+Terms(hst.δhδhδh)+Terms(δhhst.hst.hst.)+Terms(δhδhhst.hst.).\displaystyle+\text{Terms$(h^{\mathrm{st.}}\cdot\delta h\cdot\delta h\cdot\delta h)$}+\text{Terms$(\delta h\cdot h^{\mathrm{st.}}\cdot h^{\mathrm{st.}}\cdot h^{\mathrm{st.}})$}+\text{Terms$(\delta h\cdot\delta h\cdot h^{\mathrm{st.}}\cdot h^{\mathrm{st.}})$}\,. (43)

When taking the h|𝔭g\left\langle\cdot\right\rangle_{h|\mathfrak{p}_{g}} ensemble average, the Terms(hst.δhδhδh)(h^{\mathrm{st.}}\cdot\delta h\cdot\delta h\cdot\delta h) and Terms(δhhst.hst.hst.)(\delta h\cdot h^{\mathrm{st.}}\cdot h^{\mathrm{st.}}\cdot h^{\mathrm{st.}}) vanish in virtue of Eq. (41), as they are linear in either hAh_{A} or δhA\delta h_{A} which are zero-mean random fields. The Terms(δhδhhst.hst.)(\delta h\cdot\delta h\cdot h^{\mathrm{st.}}\cdot h^{\mathrm{st.}}) do not vanish immediately when taking the h|𝔭g\left\langle\cdot\right\rangle_{h|\mathfrak{p}_{g}} ensemble average, but given Eq. (40) it leads to an expression linear in δg\delta_{g}, which vanishes once averaging over density field realizations to obtain the full ensemble average.

Finally, one recovers the results of section III for ρab2st.\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{st.}} and ρab2clust.\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{clust.}}, respectively, when taking the full ensemble average over the terms in the first line of Eq. (43). We thus still have ρab2=ρab2st.+ρab2clust.\left\langle\rho_{ab}^{2}\right\rangle=\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{st.}}+\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{clust.}}. However, splitting hAh_{A} into the fields hAst.h_{A}^{\mathrm{st.}} and δhA\delta h_{A} allows to define the random variable

ρ~ab=A,Adfdfd𝐧d𝐧RaA(f,𝐧)RbA(f,𝐧)δhA(f,𝐧)δhA(f,𝐧)sinc(π(ff)T),\tilde{\rho}_{ab}=\sum_{A,A^{\prime}}\int\mathrm{d}f\int\mathrm{d}f^{\prime}\int\mathrm{d}\mathbf{n}\int\mathrm{d}\mathbf{n}^{\prime}R^{A}_{a}(f,\mathbf{n})^{\ast}R_{b}^{A^{\prime}}(f^{\prime},\mathbf{n}^{\prime})\delta h_{A}^{\ast}(f,\mathbf{n})\delta h_{A^{\prime}}(f^{\prime},\mathbf{n}^{\prime})\mbox{sinc}\left(\pi(f-f^{\prime})T\right)\,, (44)

i.e. the same as the definition of ρab\rho_{ab} given in Eq. (14), but with the perturbation δhA\delta h_{A} instead of the full hAh_{A} (beware that ρabρabst.+ρ~ab\rho_{ab}\neq\rho^{\mathrm{st.}}_{ab}+\tilde{\rho}_{ab}, since ρab\rho_{ab} is quadratic in the strain). This variable satisfies ρ~ab=0\left\langle\tilde{\rho}_{ab}\right\rangle=0, again due to the fact that the ensemble average h|𝔭g\left\langle\cdot\right\rangle_{h|\mathfrak{p}_{g}} leads to an expression linear in δg\delta_{g}. Therefore, Var[ρ~ab]=ρ~ab20\text{Var}[\tilde{\rho}_{ab}]=\left\langle\tilde{\rho}^{2}_{ab}\right\rangle\geq 0. The evaluation of ρ~ab2\left\langle\tilde{\rho}^{2}_{ab}\right\rangle follows the exact same steps as that of ρab2\left\langle\rho_{ab}^{2}\right\rangle in the absence of clustering, substituting hAst.h_{A}^{\mathrm{st.}} by δhA\delta h_{A}. In particular, we can still use Isserlis’ theorem since δhA\delta h_{A} is Gaussian for a given density realization of the universe. We immediately obtain

Var[ρ~ab]=ρ~ab2=ρab2clust.0.\text{Var}[\tilde{\rho}_{ab}]=\left\langle\tilde{\rho}^{2}_{ab}\right\rangle=\left\langle\rho_{ab}^{2}\right\rangle^{\mathrm{clust.}}\geq 0\,. (45)