3D cosmic shear: numerical challenges, 3D lensing random fields generation and Minkowski Functionals for cosmological inference

A. Spurio Mancini Institut für Theoretische Physik, Universität Heidelberg, Philosophenweg 12, 69120 Heidelberg, Germany spuriomancini@thphys.uni-heidelberg.de    P. L. Taylor Mullard Space Science Laboratory, University College London, Holmbury St. Mary, Dorking, Surrey RH5 6NT, UK    R. Reischke Astronomisches Rechen-Institut, Zentrum für Astronomie der Universität Heidelberg, Philosophenweg 12, 69120 Heidelberg, Germany,
Institut für Kernphysik, Karlsruher Institut für Technologie, 76344 Eggenstein-Leopoldshafen, Germany
   T. Kitching Mullard Space Science Laboratory, University College London, Holmbury St. Mary, Dorking, Surrey RH5 6NT, UK    V. Pettorino AIM, CEA, CNRS, Université Paris-Saclay, Université Paris Diderot, Sorbonne Paris Cité, F-91191 Gif-sur-Yvette, France    B. M. Schäfer Astronomisches Rechen-Institut, Zentrum für Astronomie der Universität Heidelberg, Philosophenweg 12, 69120 Heidelberg, Germany    B. Zieser    Ph. M. Merkel Institut für Theoretische Astrophysik, Zentrum für Astronomie der Universität Heidelberg, Philosophenweg 12, 69120 Heidelberg, Germany
Abstract

Cosmic shear - the weak gravitational lensing effect generated by fluctuations of the gravitational tidal fields of the large-scale structure - is one of the most promising tools for current and future cosmological analyses. The spherical-Bessel decomposition of the cosmic shear field (“3D cosmic shear”) is one way to maximise the amount of redshift information in a lensing analysis and therefore provides a powerful tool to investigate in particular the growth of cosmic structure that is crucial for dark energy studies. However, the computation of simulated 3D cosmic shear covariance matrices presents numerical difficulties, due to the required integrations over highly oscillatory functions. We present and compare two numerical methods and relative implementations to perform these integrations. We then show how to generate 3D Gaussian random fields on the sky in spherical coordinates, starting from the 3D cosmic shear covariances. To validate our field-generation procedure, we calculate the Minkowski functionals associated with our random fields, compare them with the known expectation values for the Gaussian case and demonstrate parameter inference from Minkowski functionals from a cosmic shear survey. This is a first step towards producing fully 3D Minkowski functionals for a lognormal field in 3D to extract Gaussian and non-Gaussian information from the cosmic shear field, as well as towards the use of Minkowski functionals as a probe of cosmology beyond the commonly used two-point statistics.

I Introduction

The weak gravitational lensing effect is the distortion of light bundles due to the differential deflection of light rays coming from distant sources (see e.g. Bartelmann and Schneider, 2001; Hoekstra and Jain, 2008; Kilbinger, 2015, for reviews on the topic) caused by the gravitational tidal fields generated by the mass distribution between the source and the observer. This effect induces a small change in the ellipticity of a background galaxy’s image, which is referred to as ‘cosmic shear’ when it is caused by the gravitational fields of the large-scale structure of the Universe. Cosmic shear measurements are of a statistical nature, as the lensing effect is not associated with a particular intervening lens, but rather corresponds to small distortions (of the order of 1%percent\%) by all potential fluctuations along the line of sight; detecting the extremely faint cosmic shear signal requires averaging over many background galaxies. The statistical properties of the shear field reflect those of the underlying density field by virtue of the gravitational field equations. The cosmic shear field has zero mean; at the level of one-point statistics, cosmological information can be extracted from e.g. peak counts (Lin and Kilbinger, 2015; Peel et al., 2017; Fluri et al., 2018; Vallis et al., 2018), while for two-point statistics one looks in configuration space at the angular correlation function of the shear field, or its equivalent in Fourier space, the cosmic shear angular power spectrum. At higher order, cosmic shear can also break degeneracies between the dark sector and neutrinos (Peel et al., 2018).

Since the first detections in early 2000s (e.g. Bacon et al., 2000; Van Waerbeke et al., 2000; Brown et al., 2003), cosmic shear has flourished into a well-established theoretical framework and important cosmological constraints have already been derived from cosmic shear analyses of different surveys over the last decade, such as the Canada-France-Hawaii Telescope Lensing Survey (CFHTLenS; Heymans et al., 2013; Kitching et al., 2014; Joudaki et al., 2017) and the Kilo Degree Survey (KiDS; Köhlinger et al., 2017; Hildebrandt et al., 2017; van Uitert et al., 2018). Even stronger constraints are expected in the next decade from planned Stage IV surveys such as Euclid111https://www.euclid-ec.org/(Laureijs et al., 2011) and the Large Synoptic Survey Telescope222https://www.lsst.org/(LSST; Abell et al., 2009), which will use the weak gravitational lensing effect as one of the main cosmological probes to achieve their science goals, most importantly the investigation of the accelerated expansion of the Universe through its influence on the background expansion and cosmic structure growth. Studies to ascertain the nature of the dark energy component of the Universe, to which the acceleration can be ascribed, crucially depend on the sensitivity of the cosmological probes to the growth of cosmic structures. Cosmic shear appears therefore very promising for this purpose, because it is particularly sensitive to the growth.

Extracting information along the radial direction becomes then a necessary requirement for any cosmic shear analysis that aims at increasing the sensitivity to the growth of cosmic structure. For this reason, a standard approach (“tomography”, first introduced in Hu, 1999) for analysing a cosmic shear survey consists in calculating correlations of the lensing signal between different redshift bins, to which the observed galaxies are assigned according to their estimated (photometric) redshifts, in an attempt to reduce loss of information due to radial averaging.

Alternatively to this tomographic approach, a spherical-Bessel decomposition of the cosmic shear field (“3D cosmic shear”, first introduced in Heavens, 2003) allows for the inclusion in the analysis of the redshift information on each galaxy within the survey. This fully 3D approach has been recently studied in Spurio Mancini et al. (2018) and Pratten et al. (2016) in the context of modified gravity theories. In Spurio Mancini et al. (2018) and Taylor et al. (2018a) the cosmological sensivity of the 3D and tomographic approach have been compared. The code used to produce the results in Taylor et al. (2018a) has been released in Taylor et al. (2018b)333The code is available at https://github.com/astro-informatics/GLaSS. For questions on the code, please contact P. Taylor at peterllewelyntaylor@gmail.com , where a study of the flat universe approximation has been performed in view of future Stage IV surveys. Taylor et al. (2018a) also discussed alternative weights to the spherical-Bessel ones, which may be easier to compute while still preserving most of the radial information.

The 3D spherical-Bessel formalism presents challenging integrals to evaluate numerically: we highlight them while reviewing the 3D cosmic shear formalism in Sec. II. Subsequently we describe two different numerical techniques used to evaluate those integrals, namely the numerical recipes underlying the results presented in Spurio Mancini et al. (2018) and Taylor et al. (2018a, b): we present them in Sec. III.1 and III.2, respectively. In Sec. IV we present a comparison of 3D cosmic shear covariance matrices obtained with the two numerical methods. Having two completely independent numerical techniques to tackle the 3D cosmic shear integrations, producing results in excellent agreement between them, is useful for a number of future applications that employ the simulated 3D covariance matrices. These include e.g. a cross-correlation analysis of 3D cosmic shear and galaxy clustering (see Lanusse et al., 2015, for a spherical-Bessel analysis of a spectroscopic galaxy clustering survey), or the development of a Bayesian Hierarchical Model for 3D cosmic shear power spectra estimation (see Alsing et al., 2016, 2017, for a Bayesian Hierarchical Model for tomography). We show in Sec. V how to make use of the 3D covariance matrices to generate Gaussian random fields on the sky, showing in particular how to overcome the difficulties arising from the non-diagonality of the covariance matrices in the radial coordinate, which originates from the inhomogeneity of the lensing field along the line of sight. We test the validity of our field-generation procedure in Sec. VI, where we briefly describe and then calculate the Minkowski Functionals of our generated Gaussian random fields, and compare them with their expectation values, known analytically in the Gaussian case. We also demonstrate in a first simplified case how a likelihood analysis for cosmological inference can be carried out using the estimated Minkowski Functionals. Finally, we draw our conclusions in Sec. VII.

II 3D cosmic shear

We begin by reviewing the formalism for a fully 3D expansion of the shear field based on its spherical-Bessel decomposition, as first introduced in lensing studies by Heavens (2003). Here we follow the notation and conventions of Spurio Mancini et al. (2018) and Zieser and Merkel (2016) (see also Heavens et al., 2006; Kitching et al., 2007; Ayaita et al., 2012; Grassi and Schäfer, 2014; Kitching et al., 2015; Taylor et al., 2018a, b).

Information on the gravitational lensing effect is encoded in the lensing potential, a weighted projection of the gravitational potential along the line of sight. In a standard General Relativity context (see Spurio Mancini et al., 2018; Pratten et al., 2016, for an extended formalism valid also for modified gravity theories), the lensing potential ϕitalic-ϕ\phi is related to the gravitational potential ΦΦ\Phi by

ϕ(χ,𝐧^)=2c20χdχχχχχΦ(χ,𝐧^),italic-ϕ𝜒^𝐧2superscript𝑐2superscriptsubscript0𝜒differential-dsuperscript𝜒𝜒superscript𝜒𝜒superscript𝜒Φ𝜒^𝐧\displaystyle\phi(\chi,\hat{\mathbf{n}})=\frac{2}{c^{2}}\int_{0}^{\chi}\mathrm{d}\chi^{\prime}\frac{\chi-\chi^{\prime}}{\chi\chi^{\prime}}\Phi(\chi,\hat{\mathbf{n}})\;, (1)

where χ𝜒\chi is a comoving distance, and the normalized vector 𝐧^^𝐧\hat{\mathbf{n}} selects a direction on the sky. Here and throughout the paper spatial flatness will be assumed (for expressions for a non-flat Universe, see Taylor et al., 2018b), and the integration in Eq.1 is carried out in Born approximation, i.e. along the unperturbed light path. The shear tensor γ(χ,𝐧^)𝛾𝜒^𝐧\gamma(\chi,\hat{\mathbf{n}}) is defined as the second ∂̸not-partial-differential\not{\partial}-derivative (Newman and Penrose, 1962; Goldberg et al., 1967) of the lensing potential (Heavens, 2003; Castro et al., 2005)

γ(χ,𝐧^)=12ððϕ(χ,𝐧^).𝛾𝜒^𝐧12italic-ðitalic-ðitalic-ϕ𝜒^𝐧\displaystyle\gamma(\chi,\hat{\mathbf{n}})=\frac{1}{2}\eth\eth\phi(\chi,\hat{\mathbf{n}})\;. (2)

The ðitalic-ð\eth-derivative acts as a covariant differentiation operator on the celestial sphere and relates quantities of different spin, raising the spin s𝑠s of a field, a number which characterises its transformation properties under rotations. Acting twice on ϕitalic-ϕ\phi, the ðitalic-ð\eth operator relates the scalar (spin-0) lensing potential ϕitalic-ϕ\phi to the spin-2 shear field γ𝛾\gamma. The shear γ𝛾\gamma can be expanded with a choice of basis functions given by a combination of spherical Bessel functions j(z)subscript𝑗𝑧j_{\ell}(z) (Abramowitz et al., 1988) and spin 2-weighted spherical harmonics Ym2(𝐧^)subscriptsubscript𝑌𝑚2^𝐧{}_{2}Y_{\ell m}(\hat{\mathbf{n}})

γ(χ,𝐧^)=2πmk2dkγm(k)2Ym(𝐧^)j(kχ),𝛾𝜒^𝐧2𝜋subscript𝑚superscript𝑘2differential-d𝑘subscript𝛾𝑚subscript𝑘2subscript𝑌𝑚^𝐧subscript𝑗𝑘𝜒\displaystyle\gamma(\chi,\hat{\mathbf{n}})=\sqrt{\frac{2}{\pi}}\sum_{\ell m}\int k^{2}\,\mathrm{d}k\,\gamma_{\ell m}(k)\,_{2}Y_{\ell m}(\hat{\mathbf{n}})\,j_{\ell}(k\chi)\;, (3)

where the coefficients γm(k)subscript𝛾𝑚𝑘\gamma_{\ell m}(k) are given by

γm(k)=2πχ2dχdΩγ(χ,𝐧^)j(kχ)2Ym(𝐧^).subscript𝛾𝑚𝑘2𝜋superscript𝜒2differential-d𝜒differential-dΩ𝛾𝜒^𝐧subscript𝑗subscript𝑘𝜒2superscriptsubscript𝑌𝑚^𝐧\displaystyle\gamma_{\ell m}(k)=\sqrt{\frac{2}{\pi}}\int\chi^{2}\mathrm{d}\chi\int\mathrm{d}\Omega\,\gamma(\chi,\hat{\mathbf{n}})\,j_{\ell}(k\chi)\,_{2}Y_{\ell m}^{*}(\hat{\mathbf{n}})\;. (4)

Inserting Eqs.1 and 2 in 4, and applying a spherical-Bessel expansion to the gravitational potential ΦΦ\Phi, we can rewrite γ𝛾\gamma as

γ(χ,𝐧^)=2π1c20χdχχχχχ𝛾𝜒^𝐧2𝜋1superscript𝑐2superscriptsubscript0𝜒differential-dsuperscript𝜒𝜒superscript𝜒𝜒superscript𝜒\displaystyle\gamma(\chi,\hat{\mathbf{n}})=\sqrt{\frac{2}{\pi}}\frac{1}{c^{2}}\int_{0}^{\chi}\mathrm{d}\chi^{\prime}\,\frac{\chi-\chi^{\prime}}{\chi\chi^{\prime}} (5)
×k2dkm(+2)!(2)!Φm(k,χ)j(kχ)2Ym(𝐧^).\displaystyle\times\int k^{2}\mathrm{d}k\sum_{\ell m}\sqrt{\frac{(\ell+2)!}{(\ell-2)!}}\Phi_{\ell m}(k,\chi^{\prime})j_{\ell}(k\chi^{\prime})\,_{2}Y_{\ell m}(\hat{\mathbf{n}})\;.

Poisson’s equation can be used to link the coefficients in the spherical-Bessel decomposition of the lensing potential to those of the density contrast field δm(k,χ)subscript𝛿𝑚𝑘𝜒\delta_{\ell m}(k,\chi),

Φm(k,χ)c2=32Ωm(kχH)2δm(k,χ)a(χ),subscriptΦ𝑚𝑘𝜒superscript𝑐232subscriptΩ𝑚superscript𝑘subscript𝜒𝐻2subscript𝛿𝑚𝑘𝜒𝑎𝜒\displaystyle\frac{\Phi_{\ell m}(k,\chi)}{c^{2}}=-\frac{3}{2}\frac{\Omega_{m}}{(k\chi_{H})^{2}}\frac{\delta_{\ell m}(k,\chi)}{a(\chi)}\;, (6)

with the Hubble radius χHc/H0subscript𝜒𝐻𝑐subscript𝐻0\chi_{H}\equiv c/H_{0}. If fluctuations in ΦΦ\Phi are weak, the density field is statistically homogeneous and isotropic, characterised by a power spectrum Pδ(k,z,z)subscript𝑃𝛿𝑘𝑧superscript𝑧P_{\delta}(k,z,z^{\prime}) which is diagonal in harmonic space

δlm(k,z)δm(k,z)=Pδ(k,z,z)k2δD(kk)δKδmmK.delimited-⟨⟩subscript𝛿𝑙𝑚𝑘𝑧superscriptsubscript𝛿superscriptsuperscript𝑚superscript𝑘superscript𝑧subscript𝑃𝛿𝑘𝑧superscript𝑧superscript𝑘2superscript𝛿𝐷𝑘superscript𝑘superscriptsubscript𝛿superscript𝐾superscriptsubscript𝛿𝑚superscript𝑚𝐾\displaystyle\left\langle\delta_{lm}(k,z)\delta_{\ell^{\prime}m^{\prime}}^{*}(k^{\prime},z^{\prime})\right\rangle=\frac{P_{\delta}(k,z,z^{\prime})}{k^{2}}\delta^{D}(k-k^{\prime})\delta_{\ell\ell^{\prime}}^{K}\delta_{mm^{\prime}}^{K}\;. (7)

Using this, we can relate the covariance of shear modes to the matter power spectrum by

γ¯lm(k)γ¯m(k)delimited-⟨⟩subscript¯𝛾𝑙𝑚𝑘superscriptsubscript¯𝛾superscriptsuperscript𝑚superscript𝑘\displaystyle\left\langle\bar{\gamma}_{lm}(k)\bar{\gamma}_{\ell^{\prime}m^{\prime}}^{*}(k^{\prime})\right\rangle =9Ωm216π4χH4(+2)!(2)!absent9superscriptsubscriptΩ𝑚216superscript𝜋4superscriptsubscript𝜒𝐻422\displaystyle=\frac{9\Omega_{m}^{2}}{16\pi^{4}\chi_{H}^{4}}\frac{(\ell+2)!}{(\ell-2)!} (8)
×dk~k~2G(k,k~)G(k,k~)δKδmmK,\displaystyle\times\int\frac{\mathrm{d}\tilde{k}}{\tilde{k}^{2}}\,G_{\ell}(k,\tilde{k})\,G_{\ell}(k^{\prime},\tilde{k})\,\delta_{\ell\ell^{\prime}}^{K}\,\delta_{mm^{\prime}}^{K}\;,

where

G(k,k)subscript𝐺𝑘superscript𝑘\displaystyle G_{\ell}(k,k^{\prime}) =dznz(z)F(z,k)U(z,k),absentdifferential-d𝑧subscript𝑛𝑧𝑧subscript𝐹𝑧𝑘subscript𝑈𝑧superscript𝑘\displaystyle=\int\mathrm{d}z\,n_{z}(z)\,F_{\ell}(z,k)\,U_{\ell}(z,k^{\prime})\;, (9)
F(z,k)subscript𝐹𝑧𝑘\displaystyle F_{\ell}(z,k) =dzpp(zp|z)j[kχ0(zp)],absentdifferential-dsubscript𝑧𝑝𝑝conditionalsubscript𝑧𝑝𝑧subscript𝑗delimited-[]𝑘superscript𝜒0subscript𝑧𝑝\displaystyle=\int\mathrm{d}z_{p}\,p(z_{p}|z)\,j_{\ell}[k\chi^{0}(z_{p})]\;, (10)
U(z,k)subscript𝑈𝑧𝑘\displaystyle U_{\ell}(z,k) =0χ(z)dχa(χ)χχχχj(kχ)Pδ(k,z(χ)).absentsuperscriptsubscript0𝜒𝑧dsuperscript𝜒𝑎superscript𝜒𝜒superscript𝜒𝜒superscript𝜒subscript𝑗𝑘superscript𝜒subscript𝑃𝛿𝑘𝑧𝜒\displaystyle=\int_{0}^{\chi(z)}\frac{\mathrm{d}\chi^{\prime}}{a(\chi^{\prime})}\frac{\chi-\chi^{\prime}}{\chi\chi^{\prime}}j_{\ell}(k\chi^{\prime})\,\sqrt{P_{\delta}\left(k,z\left(\chi\right)\right)}\;. (11)

γ¯¯𝛾\bar{\gamma} are estimates of the shear modes that, in addition to the pure lensing effect, keep into account the redshift distribution of the lensed galaxies nz(z)subscript𝑛𝑧𝑧n_{z}(z) and the conditional probability p(zp|z)𝑝conditionalsubscript𝑧𝑝𝑧p(z_{p}|z) of estimating the redshift zpsubscript𝑧𝑝z_{p} given the true redshift z𝑧z. The contribution to the total signal coming from sources situated at different distances is governed by the source density nz(z)subscript𝑛𝑧𝑧n_{z}(z); through this term, the survey depth affects the strength of the overall signal. Angular variations are assumed to be negligible by considering a uniform source density: the number of sources per steradian and redshift interval is approximated by the mean nz(z)/(4π)subscript𝑛𝑧𝑧4𝜋n_{z}(z)/(4\pi) across the sky. The influence of incomplete sky coverage is ignored in this formalism: for applications to a Fisher matrix analysis, for example, the effect of partial sky coverage can be well compensated by a multiplying factor fskysubscript𝑓skyf_{\mathrm{sky}} denoting the fraction of sky spanned by the survey, prepended to the expression for the Fisher matrix (Heavens et al., 2006; Spurio Mancini et al., 2018). A finite field of view can be incorporated in the analysis by considering a suitable window function W(𝐧^)𝑊^𝐧W(\hat{\mathbf{n}}) that represents the angular distribution of the sources (e.g. a top hat filter corresponding to a rectangular field of view). For details on the extended formalism to include in the analysis this inhomogeneous sampling we refer the reader to e.g. Heavens (2003), while e.g. Leistedt et al. (2015) consider alternative methods, such as wavelets, to deal with survey geometry in 3D cosmic shear.

Statistical isotropy guarantees that the covariance in Eq. 8 does not depend on the multipole order m𝑚m, while the assumed full sky coverage also prevents mixing of different \ell modes. If the finite field of view is taken into account, statistical isotropy is broken (e.g. by the absence of data across parts of the sky) leading to a coupling of different \ell-modes; furthermore, if the field of view is not square, even for a fixed \ell there will be different results for different m𝑚m-modes. The lensing weight function, the redshift errors and the redshift-dependence of the source distribution, instead, always introduce correlations between the amplitudes of the signal on different scales; the covariance matrix then acquires off-diagonal terms, the calculation of which is numerically involved (see Kitching et al., 2014, for how to take these into account using a pseudo-Csubscript𝐶C_{\ell} approach in 3D).

The basis of spherical Bessel functions leads to integrals with rapidly oscillatory kernels, which in the inference process have to be solved for a large number of parameter combinations. The Pδ(k,z(χ))subscript𝑃𝛿𝑘𝑧𝜒\sqrt{P_{\delta}\left(k,z\left(\chi\right)\right)} term comes from an approximation, introduced and qualitatively justified in Castro et al. (2005), to calculate unequal-time correlators appearing in the comoving distance integrations by means of a geometric mean P(k,z,z)P(k,z)P(k,z)similar-to-or-equals𝑃𝑘𝑧superscript𝑧𝑃𝑘𝑧𝑃𝑘superscript𝑧P\left(k,z,z^{\prime}\right)\simeq\sqrt{P\left(k,z\right)P\left(k,z^{\prime}\right)} (see also Kitching and Heavens, 2017). This expression simplifies considerably in the linear regime of structure formation, retrieving the one presented in the seminal paper of Heavens (2003) where a product of the linear growth factors at different redshifts is present, acting on the matter power spectrum evaluated at the present time.

The noise term for the covariance matrix of the shear modes is given by the intrinsic ellipticity dispersion of source galaxies, as a result of the fact that the observed ellipticity ϵitalic-ϵ\epsilon is assumed to be the sum of the shear γ𝛾\gamma and the intrinsic ellipticity ϵSsubscriptitalic-ϵ𝑆\epsilon_{S}. The intrinsic ellipticity dispersion is given by ϵS2=σϵ2delimited-⟨⟩superscriptsubscriptitalic-ϵ𝑆2superscriptsubscript𝜎italic-ϵ2\left\langle\epsilon_{S}^{2}\right\rangle=\sigma_{\epsilon}^{2}. As given in Heavens et al. (2006) and Kitching et al. (2007), in the spherical-Bessel formalism (see Appendix A for a complete derivation), this leads to

γm(k)γm(k)SN=σϵ22π2dznz(z)j[kχ0(z)]j[kχ0(z)]δKδmmK,subscriptdelimited-⟨⟩subscript𝛾𝑚𝑘subscript𝛾superscriptsuperscript𝑚superscript𝑘SNsuperscriptsubscript𝜎italic-ϵ22superscript𝜋2differential-d𝑧subscript𝑛𝑧𝑧subscript𝑗delimited-[]𝑘subscript𝜒0𝑧subscript𝑗superscriptdelimited-[]superscript𝑘subscript𝜒0𝑧superscriptsubscript𝛿superscript𝐾superscriptsubscript𝛿𝑚superscript𝑚𝐾\displaystyle\left\langle\gamma_{\ell m}\left(k\right)\gamma_{\ell^{\prime}m^{\prime}}\!\left(k^{\prime}\right)\right\rangle_{\mathrm{SN}}\!=\!\frac{\sigma_{\epsilon}^{2}}{2\pi^{2}}\!\!\int\!\!\mathrm{d}z\,n_{z}(z)j_{\ell}\left[k\chi_{0}(z)\right]j_{\ell^{\prime}}\left[k^{\prime}\chi_{0}(z)\right]\delta_{\ell\ell^{\prime}}^{K}\delta_{mm^{\prime}}^{K}\;, (12)

where the subscript SN stands for “shot noise” and we set σϵ=0.3subscript𝜎italic-ϵ0.3\sigma_{\epsilon}=0.3. This expression for the noise holds only in absence of intrinsic alignments, i.e. assuming that the intrinsic ellipticities of galaxies are uncorrelated (see Merkel and Schäfer, 2013; Kitching et al., 2015, for a study of intrinsic alignments in 3D cosmic shear).

III Numerical Implementation

In this section we will briefly describe the two methods used to calculate the correlations from Eqs. (8) and (12). While one code implements in C++ the Levin collocation method (Levin, 1996, 1997) that makes use of the periodic oscillations of the Bessel functions and has been used to produce the results of Zieser and Merkel (2016) and Spurio Mancini et al. (2018), the other implements the integrations by matrix multiplications and appropriate use of the Limber approximation (Kaiser, 1992, 1998; Loverde and Afshordi, 2008) at high \ell. The second code is a Python module, implemented in the code GLaSS and used in Taylor et al. (2018a, b).

III.1 Levin integration

The method presented in (Levin, 1996, 1997) can be used for efficient evaluation of rapidly oscillatory integrals, once certain conditions are satisfied. The main idea behind the method is to transform the quadrature problem into the solution of a system of linear ordinary differential equations. These are then tackled by collocation, i.e. choosing candidate solutions (polynomials) and a number of points in the domain (called collocation points), and selecting that solution which satisfies the given equations at the collocation points.

As seen in Sec. II the 3D cosmic shear signal and noise (cf. Eqs. 8 and 12) present several integrals of the form

I1[h]=z1z2dzh(z)j(kχ(z))subscript𝐼1delimited-[]superscriptsubscriptsubscript𝑧1subscript𝑧2differential-d𝑧𝑧subscript𝑗𝑘𝜒𝑧\displaystyle I_{1}[h]=\int_{z_{1}}^{z_{2}}\mathrm{d}z\,h\left(z\right)j_{\ell}\left(k\chi\left(z\right)\right) (13)

or

I2[h]=z1z2dzh(z)j(k1χ(z))j(k2χ(z)).subscript𝐼2delimited-[]superscriptsubscriptsubscript𝑧1subscript𝑧2differential-d𝑧𝑧subscript𝑗subscript𝑘1𝜒𝑧subscript𝑗subscript𝑘2𝜒𝑧\displaystyle I_{2}[h]=\int_{z_{1}}^{z_{2}}\mathrm{d}z\,h\left(z\right)j_{\ell}\left(k_{1}\chi\left(z\right)\right)j_{\ell}\left(k_{2}\chi\left(z\right)\right). (14)

The comoving distance between two events at redshift z1subscript𝑧1z_{1} and z2subscript𝑧2z_{2} is given by

χ(z1,z2)=a(z2)a(z1)cdaa2H(a)=χHz1z2dzE(z),𝜒subscript𝑧1subscript𝑧2superscriptsubscript𝑎subscript𝑧2𝑎subscript𝑧1𝑐d𝑎superscript𝑎2𝐻𝑎subscript𝜒𝐻superscriptsubscriptsubscript𝑧1subscript𝑧2d𝑧𝐸𝑧\displaystyle\chi(z_{1},z_{2})=\int_{a(z_{2})}^{a(z_{1})}\frac{c\mathrm{d}a}{a^{2}H(a)}=\chi_{H}\int_{z_{1}}^{z_{2}}\frac{\mathrm{d}z}{E(z)}\;, (15)

where a𝑎a is the scale factor and a˙/a=H(a)=H0E(a)˙𝑎𝑎𝐻𝑎subscript𝐻0𝐸𝑎\dot{a}/a=H(a)=H_{0}E(a) is the Hubble function. Rather than redshift integrals, Eqs. 13 and 14 can be rewritten, using dz=dχE[z(χ)]d𝑧d𝜒𝐸delimited-[]𝑧𝜒\mathrm{d}z=\mathrm{d}\chi\,E[z(\chi)], with the comoving distance as the integration variable:

I1[h]subscript𝐼1delimited-[]\displaystyle I_{1}[h] =1χHχ(z1)χ(z2)dχE[z(χ)]h[z(χ)]j(kχ),absent1subscript𝜒𝐻superscriptsubscript𝜒subscript𝑧1𝜒subscript𝑧2differential-d𝜒𝐸delimited-[]𝑧𝜒delimited-[]𝑧𝜒subscript𝑗𝑘𝜒\displaystyle=\frac{1}{\chi_{H}}\int_{\chi(z_{1})}^{\chi(z_{2})}\mathrm{d}\chi\,E[z(\chi)]\,h[z(\chi)]\,j_{\ell}(k\chi)\;, (16)
I2[h]subscript𝐼2delimited-[]\displaystyle I_{2}[h] =1χHχ(z1)χ(z2)dχE[z(χ)]h[z(χ)]j(k1χ)j(k2χ).absent1subscript𝜒𝐻superscriptsubscript𝜒subscript𝑧1𝜒subscript𝑧2differential-d𝜒𝐸delimited-[]𝑧𝜒delimited-[]𝑧𝜒subscript𝑗subscript𝑘1𝜒subscript𝑗subscript𝑘2𝜒\displaystyle=\frac{1}{\chi_{H}}\int_{\chi(z_{1})}^{\chi(z_{2})}\mathrm{d}\chi\,E[z(\chi)]\,h[z(\chi)]\,j_{\ell}(k_{1}\chi)j_{\ell}(k_{2}\chi)\;. (17)

Due to the highly oscillatory nature of the spherical Bessel functions, especially at high \ell or k𝑘k, the numerical solution of these integrals by standard quadrature routines is extremely inaccurate when a large number of zero-crossings occurs in the interval [χ(z1),χ(z2)]𝜒subscript𝑧1𝜒subscript𝑧2[\chi(z_{1}),\chi(z_{2})], unless an enormous number of points is used to sample the integrand: however, the procedure becomes then exceedingly time-consuming, especially if many combinations of \ell and k𝑘k need to be considered.

Here we describe an alternative method, presented by Levin (1996, 1997), which we use to evaluate our integrals. It is applicable to integrals of the form

I[F]=abdxFT(x)w(x)=abdxF,w(x),𝐼delimited-[]𝐹superscriptsubscript𝑎𝑏differential-d𝑥superscript𝐹𝑇𝑥𝑤𝑥superscriptsubscript𝑎𝑏differential-d𝑥𝐹𝑤𝑥\displaystyle I[F]=\int_{a}^{b}\mathrm{d}x\,F^{T}(x)w(x)=\int_{a}^{b}\mathrm{d}x\left\langle F,w\right\rangle(x)\;, (18)

where F(x)=[F1(x),,Fd(x)]T𝐹𝑥superscriptsubscript𝐹1𝑥subscript𝐹𝑑𝑥𝑇F(x)=[F_{1}(x),\ldots,F_{d}(x)]^{T} and w(x)=[w1(x),,wd(x)]T𝑤𝑥superscriptsubscript𝑤1𝑥subscript𝑤𝑑𝑥𝑇w(x)=[w_{1}(x),\ldots,w_{d}(x)]^{T} are vectors of functions, for which the second equality of Eq. 18 defines a scalar product ,\left\langle,\right\rangle and the functions wi(x),i=1,2,,dformulae-sequencesubscript𝑤𝑖𝑥𝑖12𝑑w_{i}(x),i=1,2,...,d, but not Fi(x)subscript𝐹𝑖𝑥F_{i}(x), are rapidly oscillatory across the integration domain. A matrix of functions A(x)𝐴𝑥A(x) is defined, such that the derivatives of w(x)𝑤𝑥w(x), denoted by w(x)superscript𝑤𝑥w^{\prime}(x), fulfil

w(x)=A(x)w(x).superscript𝑤𝑥𝐴𝑥𝑤𝑥\displaystyle w^{\prime}(x)=A(x)w(x)\;. (19)

The components Aiq(x)subscript𝐴𝑖𝑞𝑥A_{iq}(x) should not be highly oscillatory. We show below an example of such a matrix for the particular cases given in Eqs. 16 and 17. In the Levin formalism a vector p(x)𝑝𝑥p(x) is constructed to approximate the integrand in Eq. 18 by

p,w=p+ATp,wF,w.superscript𝑝𝑤superscript𝑝superscript𝐴𝑇𝑝𝑤𝐹𝑤\displaystyle\left\langle p,w\right\rangle^{\prime}=\left\langle p^{\prime}+A^{T}p,w\right\rangle\approx\left\langle F,w\right\rangle\;. (20)

The first equality follows from applying the Leibniz rule for derivatives and Eq.19, with p,Aw=ATp,w𝑝𝐴𝑤superscript𝐴𝑇𝑝𝑤\left\langle p,Aw\right\rangle=\left\langle A^{T}p,w\right\rangle. If such a vector is found, then the integral in Eq.18 can be approximated by

I[F]abdxp,w(x)=p,w(b)p,w(a).𝐼delimited-[]𝐹superscriptsubscript𝑎𝑏differential-d𝑥superscript𝑝𝑤𝑥𝑝𝑤𝑏𝑝𝑤𝑎\displaystyle I[F]\approx\int_{a}^{b}\mathrm{d}x\left\langle p,w\right\rangle^{\prime}(x)=\left\langle p,w\right\rangle(b)-\left\langle p,w\right\rangle(a)\;. (21)

This can be achieved by demanding that both terms should be equal, p,w=F,wsuperscript𝑝𝑤𝐹𝑤\left\langle p,w\right\rangle^{\prime}=\left\langle F,w\right\rangle, at n𝑛n collocation points xj,j=1,2,,nformulae-sequencesubscript𝑥𝑗𝑗12𝑛x_{j},j=1,2,...,n. The requirement

p+ATpF,w(xj)=0,j=1,,nformulae-sequencesuperscript𝑝superscript𝐴𝑇𝑝𝐹𝑤subscript𝑥𝑗0𝑗1𝑛\displaystyle\left\langle p^{\prime}+A^{T}p-F,w\right\rangle(x_{j})=0,\quad j=1,...,n (22)

generally means that the vector p+ATpFdelimited-⟨⟩superscript𝑝superscript𝐴𝑇𝑝𝐹\left\langle p^{\prime}+A^{T}p-F\right\rangle must be orthogonal to w𝑤w at the points xjsubscript𝑥𝑗x_{j}, for example by demanding that it should be the null vector:

p(xj)+AT(xj)p(xj)=F(xj).superscript𝑝subscript𝑥𝑗superscript𝐴𝑇subscript𝑥𝑗𝑝subscript𝑥𝑗𝐹subscript𝑥𝑗\displaystyle p^{\prime}(x_{j})+A^{T}(x_{j})p(x_{j})=F(x_{j}). (23)

Finding a vector p𝑝p which has this property can be achieved by choosing a set of n𝑛n linearly independent and differentiable basis functions um(x)subscript𝑢𝑚𝑥u_{m}(x) and writing each component pi(x)subscript𝑝𝑖𝑥p_{i}(x) as a linear combination:

pi(x)=ci(m)um(x),i=1,,d;m=1,,n.formulae-sequencesubscript𝑝𝑖𝑥superscriptsubscript𝑐𝑖𝑚subscript𝑢𝑚𝑥formulae-sequence𝑖1𝑑𝑚1𝑛\displaystyle p_{i}(x)=c_{i}^{(m)}u_{m}(x),\quad i=1,...,d;\quad m=1,...,n\;. (24)

Equation 23 then leads to the following linear system of equations for the d×n𝑑𝑛d\times n coefficients ci(m)superscriptsubscript𝑐𝑖𝑚c_{i}^{(m)}:

ci(m)um(xj)+Aqicq(m)um(xj)=Fi(xj),i,q=1,,d;j,m=1,,n.formulae-sequencesuperscriptsubscript𝑐𝑖𝑚superscriptsubscript𝑢𝑚subscript𝑥𝑗subscript𝐴𝑞𝑖superscriptsubscript𝑐𝑞𝑚subscript𝑢𝑚subscript𝑥𝑗subscript𝐹𝑖subscript𝑥𝑗𝑖formulae-sequence𝑞1𝑑𝑗𝑚1𝑛\displaystyle c_{i}^{(m)}u_{m}^{\prime}(x_{j})+A_{qi}c_{q}^{(m)}u_{m}(x_{j})=F_{i}(x_{j}),\quad i,q=1,...,d;j,m=1,...,n\;. (25)

Levin (1996) showed how to concretely apply this algorithm to several cases of integrals with highly oscillatory kernels. The performance varies depending on the integrand, but accuracies below 106superscript10610^{-6} can often be achieved with less than 10 collocation points. As suggested by Levin (1996), in our implementation we use equidistant collocation points

xj=a+(j1)ban1,j=1,,nformulae-sequencesubscript𝑥𝑗𝑎𝑗1𝑏𝑎𝑛1𝑗1𝑛\displaystyle x_{j}=a+(j-1)\frac{b-a}{n-1},\quad j=1,...,n (26)

and choose the n𝑛n lowest-order polynomials as basis functions:

um(x)=(xa+b2ba)m1,m=1,,n.formulae-sequencesubscript𝑢𝑚𝑥superscript𝑥𝑎𝑏2𝑏𝑎𝑚1𝑚1𝑛\displaystyle u_{m}(x)=\left(\frac{x-\frac{a+b}{2}}{b-a}\right)^{m-1},\quad m=1,...,n\;. (27)

We note that the polynomials with m>1𝑚1m>1 and the derivatives with m>2𝑚2m>2 share the root x=(a+b)/2𝑥𝑎𝑏2x=(a+b)/2: to prevent the linear system of equations from becoming singular, that root should not be used as a collocation point. The factor 1/(ba)1𝑏𝑎1/(b-a) is included for numerical reasons: if b1much-greater-than𝑏1b\gg 1 or b1much-less-than𝑏1b\ll 1, the values of polynomials of different order may differ by several orders of magnitude; the normalising factor guarantees that |um(x)|1subscript𝑢𝑚𝑥1|u_{m}(x)|\leq 1 across the integration domain, in order to regulate the range of the coefficients of the linear system of equations in Eq. 25 and thus the condition of the corresponding matrix. Suitable vectors w𝑤w for the integrals in Eqs. 16 and 17 can be identified by considering the following recurrence relations for the spherical Bessel functions (Abramowitz et al., 1988):

ddxj(x)dd𝑥subscript𝑗𝑥\displaystyle\frac{\mathrm{d}}{\mathrm{d}x}j_{\ell}(x) =j1(x)+1xj(x),absentsubscript𝑗1𝑥1𝑥subscript𝑗𝑥\displaystyle=j_{\ell-1}(x)-\frac{\ell+1}{x}j_{\ell}(x)\;, (28)
ddxj1(x)dd𝑥subscript𝑗1𝑥\displaystyle\frac{\mathrm{d}}{\mathrm{d}x}j_{\ell-1}(x) =j(x)+1xj1(x).absentsubscript𝑗𝑥1𝑥subscript𝑗1𝑥\displaystyle=-j_{\ell}(x)+\frac{\ell-1}{x}j_{\ell-1}(x)\;. (29)

Rewriting these relations in the form w=Awsuperscript𝑤𝐴𝑤w^{\prime}=Aw, one finds that

w(χ)=(j(kχ)j1(kχ)),A(χ)=(+1χkk1χ)formulae-sequence𝑤𝜒matrixsubscript𝑗𝑘𝜒subscript𝑗1𝑘𝜒𝐴𝜒matrix1𝜒𝑘𝑘1𝜒\displaystyle w(\chi)=\left(\begin{matrix}j_{\ell}(k\chi)\\ j_{\ell-1}(k\chi)\end{matrix}\right),\quad A(\chi)=\left(\begin{matrix}-\frac{\ell+1}{\chi}&k\\ -k&\frac{\ell-1}{\chi}\end{matrix}\right) (30)

is a suitable choice for the integral in Eq. 16, with F(χ)={E[z(χ)]h[z(χ)],0}T𝐹𝜒superscript𝐸delimited-[]𝑧𝜒delimited-[]𝑧𝜒0𝑇F(\chi)=\{E[z(\chi)]h[z(\chi)],0\}^{T}. It is easy to verify that neither the entries of the matrix A𝐴A nor the integral kernels F𝐹F are rapidly oscillatory. For integrals of the type in Eq. 17, four-dimensional vectors are needed:

w(χ)=(j(k1χ)j(k2χ)j1(k1χ)j(k2χ)j(k1χ)j1(k2χ)j1(k1χ)j1(k2χ)),A(χ)=(2(l+1)χk1k20k12χ0k2k203χk10k2k12(1)χ).formulae-sequence𝑤𝜒matrixsubscript𝑗subscript𝑘1𝜒subscript𝑗subscript𝑘2𝜒subscript𝑗1subscript𝑘1𝜒subscript𝑗subscript𝑘2𝜒subscript𝑗subscript𝑘1𝜒subscript𝑗1subscript𝑘2𝜒subscript𝑗1subscript𝑘1𝜒subscript𝑗1subscript𝑘2𝜒𝐴𝜒matrix2𝑙1𝜒subscript𝑘1subscript𝑘20subscript𝑘12𝜒0subscript𝑘2subscript𝑘203𝜒subscript𝑘10subscript𝑘2subscript𝑘121𝜒\displaystyle w(\chi)=\left(\begin{matrix}j_{\ell}(k_{1}\chi)j_{\ell}(k_{2}\chi)\\ j_{\ell-1}(k_{1}\chi)j_{\ell}(k_{2}\chi)\\ j_{\ell}(k_{1}\chi)j_{\ell-1}(k_{2}\chi)\\ j_{\ell-1}(k_{1}\chi)j_{\ell-1}(k_{2}\chi)\end{matrix}\right),\quad A(\chi)=\left(\begin{matrix}-\frac{2(l+1)}{\chi}\!\!\!\!\!\!&k_{1}\!\!\!\!\!\!&k_{2}\!\!\!\!\!\!&\!\!\!\!\!\!0\\ -k_{1}&-\frac{2}{\chi}&0&k_{2}\\ -k_{2}&0&-\frac{3}{\chi}&k_{1}\\ 0&-k_{2}&-k_{1}&\frac{2(\ell-1)}{\chi}\end{matrix}\right). (31)

Similarly, F(χ)={E[z(χ)]h[z(χ)],0,0,0}T𝐹𝜒superscript𝐸delimited-[]𝑧𝜒delimited-[]𝑧𝜒000𝑇F(\chi)=\{E[z(\chi)]h[z(\chi)],0,0,0\}^{T}.

III.2 GLaSS

Refer to caption
Figure 1: Comparison of the differential signal-to-noise curve (eq.37) as a function of the angular multipole. The two curves have been obtained from the signal and noise parts of the covariance matrices produced with GLaSS (red) and the Levin method (blue).
Refer to caption
Figure 2: Relative difference of the signal-to-noise curve calculated with the GLaSS and Levin method, as a function of the multipole \ell.

The Generalised Lensing and Shear Spectra (GLaSS) code is written in Python and integrated into the modular cosmological package Cosmosis (Zuntz et al., 2015). Cosmological information can be read from an external source as in this work, or directly from the Cosmosis pipeline. More information can be found in  Taylor et al. (2018b).

GLaSS is written to compute the lensing spectra for an arbitrary weight function W[kχ0(zp)]subscript𝑊delimited-[]𝑘superscript𝜒0subscript𝑧𝑝W_{\ell}[k\chi^{0}(z_{p})] which takes the place of the Bessel functions in Eq. 10; see Taylor et al. (2018a) for more details about this generalized spherical-transform. Nevertheless, 3D cosmic shear comes as an in-built run-mode option.

All nested integrals in Eqs. 8-11 are computed as matrix multiplications because this is one of the few operations that releases the Global Interpreter Lock in Python allowing parallelisation. For example,

U(z,k)χA(χ(z),χ)B(χ,k),subscript𝑈𝑧𝑘subscriptsuperscript𝜒𝐴𝜒𝑧superscript𝜒𝐵superscript𝜒𝑘U_{\ell}\left(z,k\right)\approx\sum_{\chi^{\prime}}A\left(\chi\left(z\right),\chi^{\prime}\right)B\left(\chi^{\prime},k\right), (32)

A(χ,χ)ΔχFK(χ,χ)a(χ)𝐴𝜒superscript𝜒Δsuperscript𝜒subscript𝐹𝐾𝜒superscript𝜒𝑎superscript𝜒A\left(\chi,\chi^{\prime}\right)\equiv\Delta\chi^{\prime}\frac{F_{K}\left(\chi,\chi^{\prime}\right)}{a\left(\chi^{\prime}\right)}, where ΔχΔsuperscript𝜒\Delta\chi^{\prime} is the spacing between the sampled points in χsuperscript𝜒\chi^{\prime} and B(χ,χ)j(kχ)P(k;χ)𝐵𝜒superscript𝜒subscript𝑗𝑘superscript𝜒𝑃𝑘superscript𝜒B\left(\chi,\chi^{\prime}\right)\equiv j_{\ell}\left(k\chi^{\prime}\right)\sqrt{P\left(k;\chi^{\prime}\right)} .

To further speed up computations all Bessel functions data is pre-computed in GLaSS. To save memory, values of the Bessel functions j(x)subscript𝑗𝑥j_{\ell}\left(x\right) are stored in a 2D look up table in \ell and x𝑥x and the j(kχ)subscript𝑗𝑘𝜒j_{\ell}\left(k\chi\right) are found as needed. This procedure was first described in Seljak and Zaldarriaga (1996); Kosowsky (1998).

While computing the lensing spectra in terms of nested matrix multiplications allows for easy parallelisation, this procedure does not efficiently sample the z𝑧z-k𝑘k space as efficiently as the Levin integration. At high-\ell where the Bessel functions oscillate quickly this means the lensing spectra must be evaluated at very high resolutions.

To reduce the resolution at which the lensing spectra must be evaluated, GLaSS takes the extended Limber approximation  (LoVerde and Afshordi, 2008) above >100100\ell>100. This was shown to have negligible impact for stage IV surveys (Kitching et al., 2017). Taking the Limber approximation, equation 11 can be rewritten as:

U(χ,k)=Fk(χ,ν(k))ka(ν(k))π2(+1/2)P(k,ν(k)),subscript𝑈𝜒𝑘subscript𝐹𝑘𝜒𝜈𝑘𝑘𝑎𝜈𝑘𝜋212𝑃𝑘𝜈𝑘U_{\ell}\left(\chi,k\right)=\frac{F_{k}\left(\chi,\nu\left(k\right)\right)}{ka\left(\nu\left(k\right)\right)}\sqrt{\frac{\pi}{2\left({\ell}+1/2\right)}}\sqrt{P\left(k,\nu\left(k\right)\right)}\;, (33)

where ν(k)+1/2k𝜈𝑘12𝑘\nu\left(k\right)\equiv\frac{{\ell}+1/2}{k}. Meanwhile at low-\ell the Bessel functions oscillate slowly and the nested integrals can be evaluated at lower resolution.

IV Code comparison

In the following we compare the predictions for the 3D cosmic shear covariance matrices produced with the Levin method and with the algorithm implemented in the GLaSS code. For the code comparison we fix the fiducial cosmological model to a flat cosmology with parameters given in Tab. 1. The source distribution and the redshift error probability need to be the same for the two codes. For the source distribution we follow Amendola et al. (2016) and choose

nz(z)n0(2zm)3z2exp[(2zzm)3/2],proportional-tosubscript𝑛𝑧𝑧subscript𝑛0superscript2subscript𝑧𝑚3superscript𝑧2superscript2𝑧subscript𝑧𝑚32\displaystyle n_{z}(z)\propto n_{0}\left(\frac{\sqrt{2}}{z_{m}}\right)^{3}z^{2}\exp\left[-\left(\frac{\sqrt{2}z}{z_{m}}\right)^{3/2}\right], (34)

where zmsubscript𝑧𝑚z_{m} is the median redshift of the survey and n0subscript𝑛0n_{0} is the observed redshift-integrated source density. We set zm=0.9,n0=30arcmin2formulae-sequencesubscript𝑧𝑚0.9subscript𝑛030superscriptarcmin2z_{m}=0.9,n_{0}=30\,\mathrm{arcmin}^{-2}. We take the redshift error distribution to be a Gaussian

p(zp|z)=12πσ(z)exp[(zpz)22σ2(z)],𝑝conditionalsubscript𝑧𝑝𝑧12𝜋𝜎𝑧superscriptsubscript𝑧𝑝𝑧22superscript𝜎2𝑧\displaystyle p(z_{p}|z)=\frac{1}{\sqrt{2\pi}\sigma(z)}\exp\left[-\frac{(z_{p}-z)^{2}}{2\sigma^{2}(z)}\right], (35)

with a redshift-dependent dispersion

σ(z)=σz(1+z).𝜎𝑧subscript𝜎𝑧1𝑧\displaystyle\sigma(z)=\sigma_{z}(1+z)\;. (36)
ΩmsubscriptΩm\Omega_{\rm m} ΩbsubscriptΩb\Omega_{\rm b} ΩrsubscriptΩr\Omega_{\rm r} ΩksubscriptΩ𝑘\Omega_{k} w0subscript𝑤0w_{0} wasubscript𝑤𝑎w_{a} σ8subscript𝜎8\sigma_{8} nssubscript𝑛sn_{\rm s} hh
0.315 0.0486 9.187×1059.187superscript1059.187\times 10^{-5} 0.0 -1.0 0.0 0.834 0.962 0.674
Table 1: Values of the cosmological parameters in the fiducial model assumed for the code comparison.

We first compare the signal-to-noise curve. The cumulative signal-to-noise ratio, summed over the contributions at different multipoles up to a maximum multipole \ell, is defined as

Σ2()=fsky=min2+12Tr[𝑪1𝑺𝑪1𝑺]=minΔΣ2(),annotatedsuperscriptΣ2absentsubscript𝑓skysuperscriptsubscriptsuperscriptsubscriptmin2superscript12Trdelimited-[]subscriptsuperscript𝑪1superscriptsubscript𝑺superscriptsuperscriptsubscript𝑪superscript1subscript𝑺superscriptsuperscriptsubscriptsuperscriptsubscriptminΔsuperscriptΣ2superscript\Sigma^{2}(\leq\ell)=f_{\mathrm{sky}}\sum_{\ell^{\prime}=\ell_{\mathrm{min}}}^{\ell}\frac{2\ell^{\prime}+1}{2}\mathrm{Tr}\left[\boldsymbol{C}^{-1}_{\ell^{\prime}}\boldsymbol{S}_{\ell^{\prime}}\boldsymbol{C}_{\ell^{\prime}}^{-1}\boldsymbol{S}_{\ell^{\prime}}\right]\equiv\sum_{\ell^{\prime}=\ell_{\mathrm{min}}}^{\ell}\Delta\Sigma^{2}(\ell^{\prime})\;, (37)

where 𝑺𝑺\boldsymbol{S} is the signal covariance (Eq. 8) only, while 𝑪𝑪\boldsymbol{C} refers to the sum of signal and shot noise, i.e. Eqs. 8++12. The signal-to-noise curves produced by both codes are shown in Fig.1, depicting the differential contributions to the signal-to-noise coming from the different multipoles. The Levin and GLaSS predictions for the signal-to-noise curves show agreement with each other for the multipoles considered, reaching differences below 4%percent\%, as evidenced by Fig. 2 where the relative difference between the predictions of the two codes are shown. GLaSS has slightly lower signal-to-noise at high \ell. This is because GLaSS is not designed specifically for 3D cosmic shear and the signal-to-noise converges as the resolution of the computation grid is increased. This problem will be exacerbated as one includes higher and higher \ell (e.g. =30003000\ell=3000 where there is still useful signal to add) because the Bessel functions oscillate more quickly. For our comparison we used 2000 k𝑘k-modes linearly spaced between k=0.005h/Mpc𝑘0.005Mpck=0.005\,h/\mathrm{Mpc} and k=2.0h/Mpc𝑘2.0Mpck=2.0\,h/\mathrm{Mpc}, and restricted the comparison to multipoles 10001000\ell\leq 1000; see Taylor et al. (2018b) for details on how much information is captured by GLaSS at different k𝑘k-resolutions and Taylor et al. (2018a) for a discussion of the run time at different resolutions.

As a second diagnostic for our comparison, we consider individually the signal and noise contributions to the covariance matrices (Eqs. 8 and 12, respectively) for two different multipoles, =100100\ell=100 and =500500\ell=500. For both signal and noise we compare the elements on the diagonal C(k,k)subscript𝐶𝑘𝑘C_{\ell}(k,k), and plot them respectively in Fig. 3 and Fig. 4. In the noise case we also multiply the curves by k2superscript𝑘2k^{2}, to check that they effectively become flat as expected. The predictions show good agreement, with differences of at most a few percent (in the lower k𝑘k range for the signal, and over the entire k𝑘k range for the noise), as visible also from Fig. 5, where we show the differences between the codes, normalised to the sum of their predictions. The disagreement in the signal plot towards the higher end of the k𝑘k range is due to the numerical noise present in the GLaSS computations; however, this discrepancy can be disregarded because the contributions from those k𝑘k-regimes (k0.2h/Mpcgreater-than-or-equivalent-to𝑘0.2Mpck\gtrsim 0.2h/\mathrm{Mpc} for =100100\ell=100, k0.4h/Mpcgreater-than-or-equivalent-to𝑘0.4Mpck\gtrsim 0.4h/\mathrm{Mpc} for =500500\ell=500) are many orders of magnitude smaller than the main contributions around the peak of the curves, and also much smaller than contributions from the noise (cf. Fig 4). For =100100\ell=100, the Levin and GLaSS predictions coincide until approximately k0.2h/Mpcsimilar-to-or-equals𝑘0.2Mpck\simeq 0.2h/\mathrm{Mpc}: at this point the behaviour of the curve for GLaSS starts being dominated by numerical noise, while the Levin signal decreases in a smoother way. The same happens for =500500\ell=500, but the disagreement starts at approximately k0.4h/Mpcsimilar-to-or-equals𝑘0.4Mpck\simeq 0.4\,h/\mathrm{Mpc}. In both cases however, the signal predictions in those k𝑘k-regimes are at least 3-4 orders of magnitude smaller than the contributions around the peak of the curves, just before and after approximately k0.1h/Mpcsimilar-to-or-equals𝑘0.1Mpck\simeq 0.1h/\mathrm{Mpc}, respectively. Importantly, the values of the signal curves for those k𝑘k-regimes are even smaller than the contributions from the noise, which dominates in that regime by many orders of magnitude. This means that for practical purposes we can safely ignore the contributions from those k𝑘k-regimes where the codes are apparently in disagreement in their signal predictions. In Figs. 3,4, and the left panels of Fig. 5 we demonstrate this point by shading the regions where the signal contribution represents a fraction 1/1000absent11000\leq 1/1000 of the noise contribution at the same k𝑘k. These regions turn out to be the same where the signal predictions of the two codes disagree, thus demonstrating that this discrepancy can be safely disregarded. In the bottom panel of Fig. 3 we plot the same comparison between the signal predictions produced by both two codes, with a linear scale on the y𝑦y-axis instead of the logarithmic one used in the top panel; this is another way to appreciate how subdominant the contributions coming from the higher end of the k𝑘k-range are with respect to the signal coming from the lower k𝑘k-range.

It is interesting to note that the disagreement is practically only evident in the signal predictions, while the noise part is much less affected. This may be due to the increased number of matrix multiplications that need to be performed in the calculation of the signal with respect to the noise (cf. Eqs. 8 and 12). The fact that the number of integrations to carry out for the noise is higher means, in the GLaSS implementation, that more matrix multiplications are required and these are sensitive to the resolution in k𝑘k. Additionally, since in the noise part of the covariance matrix there are no multiplications by Bessel functions, this may suggest that the spikes at high-k𝑘k in the signal may also be due to the Bessel function resolution breaking down (as explained in Sec. III.2, in GLaSS the Bessel functions j(x)subscript𝑗𝑥j_{\ell}(x) are precomputed in a look up table in \ell and x𝑥x).

Refer to caption
Refer to caption
Figure 3: Comparison of the diagonal elements of the signal part of the covariance matrices (eq.8) for two multipoles =100100\ell=100 and =500500\ell=500, produced with GLaSS (solid lines, cyan and red for =100100\ell=100 and =500500\ell=500, respectively) and the Levin method (dashed lines, blue and black for =100100\ell=100 and =500500\ell=500, respectively). All curves have been plotted without performing any interpolation. We show the same curves using a linear (upper panel) and a logarithmic (bottom panel) scale on the y𝑦y-axis. The differences at higher k𝑘k (k0.2h/Mpcgreater-than-or-equivalent-to𝑘0.2Mpck\gtrsim 0.2\,h/\mathrm{Mpc} for =100100\ell=100, k0.4h/Mpcgreater-than-or-equivalent-to𝑘0.4Mpck\gtrsim 0.4\,h/\mathrm{Mpc} for =500500\ell=500) arise from the higher numerical noise present in the GLaSS computations in that k𝑘k regime. However, these contributions are many orders of magnitude smaller than the main contributions around the peaks of the curves, and much smaller than the contributions from the noise (cf. Fig. 4), therefore can be safely neglected. We demonstrate this point in the upper panel by indicating the shaded region for each multipole \ell where the signal represents a fraction 1/1000absent11000\leq 1/1000 of the noise: these regions correspond to the k𝑘k-ranges where the GLaSS and Levin predictions for the signal are in apparent disagreement (cf. also Fig. 5). In both panels the curves for =500500\ell=500 have been multiplied by a factor 1000 for easier visualisation.
Refer to caption
Figure 4: Comparison of the diagonal elements of the noise part of the covariance matrices (eq.12) for the same two multipoles =100100\ell=100 and =500500\ell=500 of Fig. 3, produced with GLaSS (solid, cyan for =100100\ell=100 and red for =500500\ell=500) and the Levin method (dashed, blue for =100100\ell=100 and black for =500500\ell=500). All curves have been plotted without performing any interpolation and multiplied by a factor k2superscript𝑘2k^{2}. In the case =500500\ell=500 the curves produced by both methods have also been multiplied by a factor 20 for easier visualisation.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5: Differences between the predictions for the signal (left panels) and noise (right panels) contributions to the covariance matrices for multipoles =100100\ell=100 (top panels) and =500500\ell=500 (bottom panels), normalised to their sum. We stress here again that the discrepancies at high k𝑘k should not be a concern because the k𝑘k-regimes where they originate produce contributions very much subdominant with respect to the peaks of the signal curves, and also with respect to the relevant contributions from the noise (cf. Figs. 3, 4). In the signal plots we shade the regions where the signal is a fraction 1/1000absent11000\leq 1/1000 of the noise (cf. Fig. 3): these regions correspond tho the k𝑘k values where the differences between the two codes are bigger, however since the signal contributions from these regions are negligeable, this discrepancy can be safely ignored.

The code implementing the Levin method sources the matter power spectrum from the Einstein-Boltzmann solver Cosmic Linear Anisotropy Solving System (CLASS, Blas et al., 2011), while for this code comparison the matter power spectrum used by GLaSS has been sourced from the Code for Anisotropies in the Microwave Background (CAMB, Lewis et al., 2000). The CLASS and CAMB codes have been compared in their predictions (Lesgourgues, 2011). Therefore, in comparing the Levin and GLaSS methods, the matter power spectrum has been ruled out as a possible source of discrepancy.

We conclude this section with a note on the performance of the two codes. The code implementing the Levin integration has been developed explicitly for the production of precise 3D cosmic shear cosmological forecasts and has been recently used to this purpose in Spurio Mancini et al. (2018). As one can see from Figs. 3 and 4, the curves produced with the Levin method are very smooth, showing the high precision achieved by the method. This compensates for the relatively low speed of the code, necessary to achieve that precision. GLaSS on the other hand, has not been developed for 3D cosmic shear only; in Taylor et al. (2018a, b) it is introduced as a means to compute lensing spectra for arbitrary weighting functions and, importantly, for integration within the cosmological module Cosmosis. This means that speed has been a crucial goal in developing the code and the method used for the matrix multiplications indeed allows for greater speed than the one achieved with the Levin method. However, numerical noise remains higher: to overcome this issue, one would need to increase the resolution at which the matrix multiplications are performed, but this would inevitably imply a slower performance of the code. We conclude that the use of the Levin or the GLaSS method depends on the task to perform: if a high level of precision is required, the Levin method should be preferred, while if speed is a crucial requirement, GLaSS can be a better option. For our purposes in this paper, i.e. the demonstration of a method for generating 3D lensing random fields on the sky and the calculation of Minkowski Functionals associated to these fields, both methods are equally valid for the computation of the 3D cosmic shear covariance matrices, which represent the starting point of the algorithms described in the following sections.

V Generation of spin-2 random fields on the sky

In this section we show how to generate random fields on the sky starting from the full 3D cosmic shear covariance matrix. As shown in Sec. II, the full covariance matrix can be decomposed in C(k,k)subscript𝐶𝑘superscript𝑘C_{\ell}(k,k^{\prime}) for each multipole \ell, given that the assumed isotropy of the shear field implies multipole independence γm(k)γm(k)=C(k,k)δδmmdelimited-⟨⟩subscript𝛾𝑚𝑘subscript𝛾superscriptsuperscript𝑚superscript𝑘subscript𝐶𝑘superscript𝑘subscript𝛿superscriptsubscript𝛿𝑚superscript𝑚\left\langle\gamma_{\ell m}(k)\gamma_{\ell^{\prime}m^{\prime}}(k^{\prime})\right\rangle=C_{\ell}(k,k^{\prime})\delta_{\ell\ell^{\prime}}\delta_{mm^{\prime}}. We detail our procedure considering for simplicity the convergence κ𝜅\kappa, as it is a scalar field and therefore easier to analyse. The convergence shares essentially the same covariance matrix with the shear field, each \ell-block only being rescaled by a prefactor (+1)(+2)(1)121\frac{\ell(\ell+1)}{(\ell+2)(\ell-1)} (Castro et al., 2005) that plays a role only for the very largest angular scales. The generalisation to the spin-2 case for the shear field simply requires starting from the original shear covariance matrix and replacing the transforms from Fourier coefficients to configuration space with their spin-2 extensions. To demonstrate our field generation procedure we use the covariance matrices produced with the Levin integration.

Our aim is to generate modes of the convergence field in Fourier space κmsubscript𝜅𝑚\kappa_{\ell m} and to transform them back into configuration space using the HEALPix (Górski et al., 2005) in-built function alm2map, in its scalar version for the convergence case (for the shear, one simply needs to activate the option pol = True that allows the user to deal with spin-2 fields). This way we can obtain samples of the convergence field in configuration space κ(r,θ,ϕ)𝜅𝑟𝜃italic-ϕ\kappa(r,\theta,\phi), on spherical shells corresponding to different values of the radius; on each shell, the field can be discretised on a HEALPix map (an example of the final result is given in Fig. 6). The procedure described in the following is similar to the one used in the code FLASK (Xavier et al., 2016) to generate samples of the density, convergence and shear fields on redshift slices, starting from tomographic weak lensing covariance matrices Cij()subscript𝐶𝑖𝑗C_{ij}(\ell), where the indices i𝑖i and j𝑗j run over the redshift slices and the type of field (density, convergence or shear). In FLASK, the problem of generating correlated random fields across different redshift slices is dealt with by means of a Cholesky decomposition of the correlated covariance matrices Cij()subscript𝐶𝑖𝑗{C}_{ij}(\ell). The Cholesky decomposition rewrites the covariance as the product of an upper and lower triangular matrix. Here, the situation is similar in that we also have correlated multipoles belonging to the different radial slices, however the correlation is in terms of the wavector k𝑘k rather than the tomographic/field index i𝑖i. This difference originates from the fact that we start from the 3D cosmic shear covariance matrices C(k,k)subscript𝐶𝑘superscript𝑘C_{\ell}(k,k^{\prime}), as opposed to the tomographic Cij()subscript𝐶𝑖𝑗C_{ij}(\ell) matrices in FLASK. Additionally, in FLASK correlations between density, convergence and shear fields can be considered if the user desires, while here we concentrate singularly on the generation of convergence or shear fields and do not consider their cross-correlations. The fact that the random fields at different wavectors are correlated is ultimately due to the fact that the lensing field is not homogenenous along the line of sight, due to the mode-coupling effect of the lensing kernel, the source redshift distributions and the redshift error probability (cf. Eqs. 8-11).

The assumption of statistical isotropy implies that modes κmsubscript𝜅𝑚\kappa_{\ell m} of the convergence field at different multipoles \ell and m𝑚m can be generated independently. The number of \ell multipoles is in principle infinite, however practically there will be a maxsubscriptmax\ell_{\mathrm{max}} which sets the maximum resolution. We use max=3Nsidesubscriptmax3subscript𝑁side\ell_{\mathrm{max}}=3N_{\mathrm{side}}, where Nsidesubscript𝑁sideN_{\mathrm{side}} is a HEALPix parameter describing the resolution of the HEALPix grid (Górski et al., 2005). The choice max=3Nsidesubscriptmax3subscript𝑁side\ell_{\mathrm{max}}=3N_{\mathrm{side}} is the same made by Lim and Simon (2012) in their CMB analysis and guarantees that the grid size is comparable to the smallest angular scale considered, corresponding to maxsubscriptmax\ell_{\mathrm{max}}. For each \ell value, m𝑚m ranges from -\ell to ++\ell, so that there are 2+1212\ell+1 m𝑚m values for each multipole \ell. However, due to the hermiticity of the convergence field, we actually consider only +11\ell+1 modes from 0 to \ell. We employ a Cholesky decomposition of the covariance matrices to deal with the fact that modes corresponding to different k𝑘k values are correlated:

C(k,k)=pT(k,p)T(p,k),subscript𝐶𝑘superscript𝑘subscript𝑝subscript𝑇𝑘𝑝subscript𝑇𝑝superscript𝑘\displaystyle C_{\ell}(k,k^{\prime})=\sum_{p}T_{\ell}(k,p)T_{\ell}(p,k^{\prime}), (38)

where 𝐓()𝐓\mathbf{T}(\ell) are (lower) triangular matrices, which we can later use to generate correlated random variables κm(k)subscript𝜅𝑚𝑘\kappa_{\ell m}(k), e.g. Gaussian distributed,

κm(k)=pT(k,p)nm(k),subscript𝜅𝑚𝑘subscript𝑝subscript𝑇𝑘𝑝subscript𝑛𝑚𝑘\displaystyle\kappa_{\ell m}(k)=\sum_{p}T_{\ell}(k,p)\,n_{\ell m}(k), (39)

where nm(k)subscript𝑛𝑚𝑘n_{\ell m}(k) are independent, Gaussian distributed random variables with zero mean and unit variance. To obtain our convergence field samples in configuration space, we transform back from Fourier space, first by multiplying by a spherical Bessel function and k2superscript𝑘2k^{2} and integrating over k𝑘k, as indicated by Eq. 3, and then acting with the HEALPix routine alm2map to obtain the field samples on a discretised grid in the angular coordinates. This way we are essentially performing in two steps an inverse spherical-Bessel transform.

We summarise schematically our procedure in Algorithm 1. We implemented it in a Python routine, leveraging parallelisation on multiple cores with joblib. The problem is embarassingly parallel, since the correlation of the fields on different radii is preserved by the starting cosmic shear covariance matrix, while different realisations of the random fields are completely independent. The fact that the covariance does not depend on m𝑚m, but only on the multipole \ell, can be used to speed up calculations, as one needs to perform the Cholesky decomposition only once per each multipole \ell, and can then use the decomposition for all m𝑚m’s pertaining to that \ell mode.

input : Covariance matrix (e.g. for the convergence) κmκmdelimited-⟨⟩subscript𝜅𝑚subscript𝜅superscriptsuperscript𝑚\left\langle\kappa_{\ell m}\kappa_{\ell^{\prime}m^{\prime}}\right\rangle = C(k,k)δδmmsubscript𝐶𝑘superscript𝑘subscript𝛿superscriptsubscript𝛿𝑚superscript𝑚C_{\ell}(k,k^{\prime})\delta_{\ell\ell^{\prime}}\delta_{mm^{\prime}}
output : κ(r,θ,ϕ)𝜅𝑟𝜃italic-ϕ\kappa(r,\theta,\phi). For each fixed radius r𝑟r in r1,rNχsubscript𝑟1subscript𝑟subscript𝑁𝜒r_{1},\dots r_{N_{\chi}}, create a HEALPix map on discretised θ𝜃\theta and ϕitalic-ϕ\phi
method :  r[r1,rNχ]for-all𝑟subscript𝑟1subscript𝑟subscript𝑁𝜒\forall r\in[r_{1},\dots r_{N_{\chi}}]:
[0,max]for-all0subscriptmax\forall\ell\in[0,\ell_{\mathrm{max}}]:
  Cholesky decompose 𝑪=𝑻𝑻Tsubscript𝑪subscript𝑻superscriptsubscript𝑻𝑇\boldsymbol{C}_{\ell}=\boldsymbol{T}_{\ell}\boldsymbol{T}_{\ell}^{T};
  m[0,]for-all𝑚0\forall m\in[0,\ell]:
  sample 𝒛N(0,I)similar-to𝒛N0𝐼\boldsymbol{z}~{}\sim~{}\operatorname{N}({0},I);
  κm(k)=𝑻𝒛subscript𝜅𝑚𝑘subscript𝑻𝒛\kappa_{\ell m}(k)=\boldsymbol{T}_{\ell}\boldsymbol{z};
  κm(k)dkk2j(kr)κm(r)subscript𝜅𝑚𝑘differential-d𝑘superscript𝑘2subscript𝑗𝑘𝑟subscript𝜅𝑚𝑟\kappa_{\ell m}(k)\rightarrow\int\mathrm{d}k\,k^{2}\,j_{\ell}(kr)\rightarrow\kappa_{\ell m}(r);
  κm(r)subscript𝜅𝑚𝑟absent\kappa_{\ell m}(r)\rightarrow HEALPix alm2map κ(r,θ,ϕ)absent𝜅𝑟𝜃italic-ϕ\rightarrow\kappa(r,\theta,\phi)
Algorithm 1 Algorithm for generation of lensing Gaussian random fields on spherical shells

We show an example of 3D reconstruction on 3 slices in redshift (or equivalently, comoving distance) in Fig. 6.

VI Minkowski Functionals of scalar fields on the sphere

In this section we briefly introduce Minkowski Functionals (MFs) and apply them to the generated random fields introduced in Sec. V. For Gaussian random fields MFs can be calculated analytically. We will compare this theoretical prediction with the MFs calculated directly from the HEALPix maps as a proof of concept. In particular we will calculate the MFs on spheres of different radii (compare Fig. 6) and estimate the covariance between the different MFs at those radii. Repeating this whole procedure for different starting lensing covariance matrices (e.g varying each time one parameter), we can then produce a likelihood function dependent on the underlying cosmology.

Refer to caption
Figure 6: Convergence field sampled at three different values of the radius χ𝜒\chi, with the observer situated in the centre. A section of the outer and middle sphere has been removed to facilitate visualisation. The lensing covariance matrix which we used for sampling the random field is given by Eq.8. We consider only contributions from the signal part of the covariance matrix, and use 30 \ell modes ranging between 10 and 1000. We use a linear matter power spectrum for the calculation of the covariance.

VI.1 General definition

Here we define the MFs, concentrating on the aspects that are more interesting for cosmological applications and referring the reader to e.g. Mecke et al. (1994) for further mathematical details. In our definitions we follow the notation of Schmalzing and Gorski (1998) and Lim and Simon (2012).

MFs are quantities that characterize the morphology of sets in an n𝑛n dimensional space \mathcal{M}. To be considered morphological, a quantity needs to be invariant under translation and rotations; while one could think that there are many such quantities, Hadwiger’s theorem (Hadwiger, 1957) states that in an n𝑛n dimensional space there exist only n+1𝑛1n+1 linearly independent morphological functionals, from which all the others can be derived. These are the so–called Minkowski Functionals. On the 2-dimensional sphere, 𝕊2subscript𝕊2\mathbb{S}_{2}, there are 2+1=32132+1=3 MFs, carrying clear geometrical interpretations and in particular representing the area, circumference and integrated geodesic curvature of an excursion set, i.e. a region where the field exceeds some threshold level. Given a threshold ν𝜈\nu and a smooth scalar field u𝑢u, the excursion set Qνsubscript𝑄𝜈Q_{\nu} is mathematically defined as

Qν={x|u(x)>ν},subscript𝑄𝜈conditional-set𝑥𝑢𝑥𝜈\displaystyle Q_{\nu}=\big{\{}x\in\mathcal{M}\,|\,u(x)>\nu\big{\}}\;, (40)

while its boundary Qνsubscript𝑄𝜈\partial Q_{\nu} is given by

Qν={x|u(x)=ν}.subscript𝑄𝜈conditional-set𝑥𝑢𝑥𝜈\displaystyle\partial Q_{\nu}=\big{\{}x\in\mathcal{M}\,|\,u(x)=\nu\big{\}}\;. (41)

When considering the 2-sphere 𝕊2subscript𝕊2\mathbb{S}_{2}, the first MF V0(ν)subscript𝑉0𝜈V_{0}(\nu) can be interpreted as the area of Qνsubscript𝑄𝜈Q_{\nu},

V0(ν):=14π𝕊2dΩΘ(uν),assignsubscript𝑉0𝜈14𝜋subscriptsubscript𝕊2differential-dΩΘ𝑢𝜈\displaystyle V_{0}(\nu):=\frac{1}{4\pi}\int_{\mathbb{S}_{2}}\mathrm{d}\Omega\,\Theta(u-\nu)\;, (42)

where ΘΘ\Theta is the Heaviside function. The total length of the boundary of Qνsubscript𝑄𝜈Q_{\nu} gives the second MF

V1(ν):=116π𝕊2dl=116π𝕊2dΩδ(uν)|u|.assignsubscript𝑉1𝜈116𝜋subscriptsubscript𝕊2differential-d𝑙116𝜋subscriptsubscript𝕊2differential-dΩ𝛿𝑢𝜈𝑢\displaystyle V_{1}(\nu):=\frac{1}{16\pi}\int_{\partial\mathbb{S}_{2}}\mathrm{d}l=\frac{1}{16\pi}\int_{\mathbb{S}_{2}}\mathrm{d}\Omega\,\delta(u-\nu)\,|\nabla u|\;. (43)

Here δ𝛿\delta is the delta distribution and |u|𝑢|\nabla u| is the norm of the gradient of u𝑢u. Finally, the third MF is the integral of the quantity κ𝜅\kappa along the boundary

V2(ν):=18π2𝕊2dlκ=18π2𝕊2dΩδ(uν)|u|κ,assignsubscript𝑉2𝜈18superscript𝜋2subscriptsubscript𝕊2differential-d𝑙𝜅18superscript𝜋2subscriptsubscript𝕊2differential-dΩ𝛿𝑢𝜈𝑢𝜅\displaystyle V_{2}(\nu):=\frac{1}{8\pi^{2}}\int_{\partial\mathbb{S}_{2}}\mathrm{d}l\,\kappa=\frac{1}{8\pi^{2}}\int_{\mathbb{S}_{2}}\mathrm{d}\Omega\,\delta(u-\nu)\,|\nabla u|\,\kappa\;, (44)

where κ𝜅\kappa is the geodesic curvature: this describes how much a curve γ𝛾\gamma is different from a straight line, i.e. from a geodetic. For a normalised tangent, i.e. |γ˙|=1˙𝛾1|\dot{\gamma}|=1, the curvature is defined through

κ:=|γ˙γ˙|,assign𝜅subscript˙𝛾˙𝛾\displaystyle\kappa:=|\nabla_{\dot{\gamma}}\dot{\gamma}|\;, (45)

where γ˙subscript˙𝛾\nabla_{\dot{\gamma}} represents the covariant derivative along the tangent vector γ˙˙𝛾\dot{\gamma} of the curve. Schmalzing and Gorski (1998) show how to calculate κ𝜅\kappa on a generic manifold, which in the case of 𝕊2subscript𝕊2\mathbb{S}_{2} reads

κ=2u;θu;ϕu;θϕu;θ2u;ϕϕu;ϕ2uθθu;θ2+u;ϕ2,\displaystyle\kappa=\frac{2u_{;\theta}u_{;\phi}u_{;\theta\phi}-u_{;\theta}^{2}u_{;\phi\phi}-u_{;\phi}^{2}u_{\theta\theta}}{u_{;\theta}^{2}+u_{;\phi}^{2}}, (46)

with the semicolon indicating a covariant derivative.

Refer to caption
Figure 7: Numerical estimations of the first MF V0Gsuperscriptsubscript𝑉0𝐺V_{0}^{G} (dots), calculated on our generated Gaussian fields at different values of the radius (represented by different colours), compared with the theoretical predictions given by Eq. 54 (joined by lines), as a function of the threshold ν𝜈\nu. The range of the thresholds always varies between 4σ4𝜎-4\sqrt{\sigma} and +4σ4𝜎+4\sqrt{\sigma}, where σ𝜎\sigma is the (average) variance of the lensing field at a fixed radius.

VI.2 Numerical calculation of MFs

Numerical estimates of the MFs calculated on our realisations of the lensing fields can be obtained using the software HEALPix (Górski et al., 2005), as we explain in the following. We first generate full sky maps of e.g. the convergence field, on concentric spherical shells at different radii, starting from the 3D covariance matrices; to this purpose we follow the procedure described in sec. V. We then calculate numerically the MFs by directly implementing the integrals in Eqs. 42-44; our algorithm closely follows the one used in Schmalzing and Gorski (1998) and Lim and Simon (2012) and is reported in the following.

Refer to caption
Figure 8: Numerical estimations of the second MF V1Gsuperscriptsubscript𝑉1𝐺V_{1}^{G} (dots), calculated on our generated Gaussian fields at different values of the radius (represented by different colours), compared with the theoretical predictions. The colour scheme is the same as in Fig. 7.

Given the values u(xi)𝑢subscript𝑥𝑖u(x_{i}) of a field on a pixelated map, HEALPix provides useful routines that allow for the calculation of first and second partial derivatives at each pixel in (,m)𝑚(\ell,m) spherical harmonic space. We use this to obtain the three numerical MFs for 𝕊2subscript𝕊2\mathbb{S}_{2}, which we label Vi(i=0,1,2)subscript𝑉𝑖𝑖012V_{i}(i=0,1,2), via a sum over all pixels

Vi(ν):=1Npixj=1Npixi(ν,xj)assignsubscript𝑉𝑖𝜈1subscript𝑁pixsuperscriptsubscript𝑗1subscript𝑁pixsubscript𝑖𝜈subscript𝑥𝑗\displaystyle V_{i}(\nu):=\frac{1}{N_{\mathrm{pix}}}\sum_{j=1}^{N_{\mathrm{pix}}}\mathcal{I}_{i}(\nu,x_{j})\; (47)

of the respective integrands

0(ν,xj)subscript0𝜈subscript𝑥𝑗\displaystyle\mathcal{I}_{0}(\nu,x_{j}) :=Θ(uν),assignabsentΘ𝑢𝜈\displaystyle:=\Theta(u-\nu)\;, (48)
1(ν,xj)subscript1𝜈subscript𝑥𝑗\displaystyle\mathcal{I}_{1}(\nu,x_{j}) :=14δ(uν)u;θ2+u;ϕ2,\displaystyle:=\frac{1}{4}\delta(u-\nu)\sqrt{u_{;\theta}^{2}+u_{;\phi}^{2}}\;, (49)
2(ν,xj)subscript2𝜈subscript𝑥𝑗\displaystyle\mathcal{I}_{2}(\nu,x_{j}) :=12πδ(uν)2u;θu;ϕu;θϕu;θ2u;ϕϕu;ϕ2uθθu;θ2+u;ϕ2,\displaystyle:=\frac{1}{2\pi}\delta(u-\nu)\frac{2u_{;\theta}u_{;\phi}u_{;\theta\phi}-u_{;\theta}^{2}u_{;\phi\phi}-u_{;\phi}^{2}u_{\theta\theta}}{u_{;\theta}^{2}+u_{;\phi}^{2}}\;, (50)

where the semicolon indicates again a covariant derivative,

u;θ\displaystyle u_{;\theta} :=θu,u;ϕ:=1sinθϕu,u;ϕϕ:=1sin2θϕ2u+cosθsinθθu,\displaystyle:=\partial_{\theta}u,\quad u_{;\phi}:=\frac{1}{\sin\theta}\partial_{\phi}u,\quad u_{;\phi\phi}:=\frac{1}{\sin^{2}\theta}\partial_{\phi}^{2}u+\frac{\cos\theta}{\sin\theta}\partial_{\theta}u\;, (51)
u;θθ\displaystyle u_{;\theta\theta} :=θ2u,u;θϕ:=1sinθθϕucosθsin2θϕu,u;θ:=θ.\displaystyle:=\partial_{\theta}^{2}u,\quad u_{;\theta\phi}:=\frac{1}{\sin\theta}\partial_{\theta}\partial_{\phi}u-\frac{\cos\theta}{\sin^{2}\theta}\partial_{\phi}u,\quad u_{;\theta}:=\frac{\partial}{\partial\theta}\;. (52)

The integrands I1subscript𝐼1I_{1} and I2subscript𝐼2I_{2} involve the delta function: to approximate this numerically, Schmalzing and Gorski (1998) and Lim and Simon (2012) use the Heaviside function

δN(x):=(Δν)1[Θ(x+Δν/2)Θ(xΔν/2)].assignsubscript𝛿𝑁𝑥superscriptΔ𝜈1delimited-[]Θ𝑥Δ𝜈2Θ𝑥Δ𝜈2\displaystyle\delta_{N}(x):=(\Delta\nu)^{-1}[\Theta(x+\Delta\nu/2)-\Theta(x-\Delta\nu/2)]\;. (53)

This approximation of the delta function produces some numerical noise, which Lim and Simon (2012) demonstrate to be due to the delta function discretization rather than some random noise which should disappear averaging over nRsubscript𝑛𝑅n_{R} realisations. For our purposes, we do not consider the corrections proposed by Lim and Simon (2012) to remove this discretisation effect and simply average over many realisations of the field. This is enough for our purposes, as our main goal is to test the field generation procedure rather than using the MFs to study e.g. non-Gaussianity as in Lim and Simon (2012) (in which case these corrections should be taken into account).

For Gaussian fields, as the ones we are considering here, the expectation values for the MFs are known analytically and equal to

V¯0G(ν)superscriptsubscript¯𝑉0𝐺𝜈\displaystyle\bar{V}_{0}^{G}(\nu) :=V0G(ν)=12(1erf(νμ2σ)),assignabsentdelimited-⟨⟩superscriptsubscript𝑉0𝐺𝜈121erf𝜈𝜇2𝜎\displaystyle:=\left\langle{V}_{0}^{G}(\nu)\right\rangle=\frac{1}{2}\left(1-\mathrm{erf}\left(\frac{\nu-\mu}{\sqrt{2\sigma}}\right)\right)\;, (54)
V¯1G(ν)superscriptsubscript¯𝑉1𝐺𝜈\displaystyle\bar{V}_{1}^{G}(\nu) :=V1G(ν)=18τσexp((νμ)22σ),assignabsentdelimited-⟨⟩superscriptsubscript𝑉1𝐺𝜈18𝜏𝜎expsuperscript𝜈𝜇22𝜎\displaystyle:=\left\langle{V}_{1}^{G}(\nu)\right\rangle=\frac{1}{8}\sqrt{\frac{\tau}{\sigma}}\mathrm{exp}\left(-\frac{(\nu-\mu)^{2}}{2\sigma}\right)\;, (55)
V¯2G(ν)superscriptsubscript¯𝑉2𝐺𝜈\displaystyle\bar{V}_{2}^{G}(\nu) :=V2G(ν)=1(2π)3/2τσνμσexp((νμ)22σ).assignabsentdelimited-⟨⟩superscriptsubscript𝑉2𝐺𝜈1superscript2𝜋32𝜏𝜎𝜈𝜇𝜎expsuperscript𝜈𝜇22𝜎\displaystyle:=\left\langle{V}_{2}^{G}(\nu)\right\rangle=\frac{1}{(2\pi)^{3/2}}\frac{\tau}{\sigma}\frac{\nu-\mu}{\sqrt{\sigma}}\mathrm{exp}\left(-\frac{(\nu-\mu)^{2}}{2\sigma}\right)\;. (56)

Therefore we can compare our numerical estimates with the theoretical expectation values as a check for the validity of our field generation procedure. We perform this comparison in Figs. 789, where we overplot our numerical estimates and their expectation values. We consider all three MFs and show the comparison for five values of the radii, corresponding to five concentric shells over which we generate our lensing field. We calculate our MFs over a set of thresholds that always ranges between 4σ4𝜎-4\sqrt{\sigma} and +4σ4𝜎+4\sqrt{\sigma}, where σ𝜎\sigma is the variance of the lensing field at a certain radius.

The error bars associated to our numerical estimates of the MFs are taken as the square root of the diagonal elements of the covariance matrix of the MFs, computed as

CovijsubscriptCov𝑖𝑗\displaystyle\mathrm{Cov}_{ij} =1nR1m=1nR(VimVi)(VjmVj),absent1subscript𝑛𝑅1superscriptsubscript𝑚1subscript𝑛𝑅superscriptsubscript𝑉𝑖𝑚delimited-⟨⟩subscript𝑉𝑖superscriptsubscript𝑉𝑗𝑚delimited-⟨⟩subscript𝑉𝑗\displaystyle=\frac{1}{n_{R}-1}\sum_{m=1}^{n_{R}}\Big{(}V_{i}^{m}-\left\langle V_{i}\right\rangle\Big{)}\left(V_{j}^{m}-\langle V_{j}\rangle\right),
i,j=0,3nχnνformulae-sequence𝑖𝑗03subscript𝑛𝜒subscript𝑛𝜈\displaystyle i,j=0,\cdots 3\cdot n_{\chi}\cdot n_{\nu} (57)

where the indices i,j𝑖𝑗i,j run over the type of Minkowski Functional (the three MFs V0,V1,V2subscript𝑉0subscript𝑉1subscript𝑉2V_{0},V_{1},V_{2}), the number of radii nχsubscript𝑛𝜒n_{\chi} and the number of thresholds nνsubscript𝑛𝜈n_{\nu}. Videlimited-⟨⟩subscript𝑉𝑖\left\langle V_{i}\right\rangle denotes the mean of the MFs over all realisations nRsubscript𝑛𝑅n_{R}, Vi=1nRm=1nRVimdelimited-⟨⟩subscript𝑉𝑖1subscript𝑛𝑅superscriptsubscript𝑚1subscript𝑛𝑅superscriptsubscript𝑉𝑖𝑚\left\langle V_{i}\right\rangle=\frac{1}{n_{R}}\sum_{m=1}^{n_{R}}V_{i}^{m}.

Refer to caption
Figure 9: Numerical estimations of the third MF V2Gsuperscriptsubscript𝑉2𝐺V_{2}^{G} (dots), calculated on our generated Gaussian fields at different values of the radius (represented by different colours), compared with the theoretical predictions. The colour scheme is the same as in Fig. 7.

An example of this matrix is presented in Fig. 10. We consider the covariance between all three MFs (V0subscript𝑉0V_{0}, V1subscript𝑉1V_{1}, V2subscript𝑉2V_{2}, each of them as a function of the threshold ν𝜈\nu, ranging from ν1subscript𝜈1\nu_{1} to νmaxsubscript𝜈max\nu_{\mathrm{max}}), and we include the correlations between MFs belonging to each different radius (labelled by different χ𝜒\chi value, from χ1subscript𝜒1\chi_{1} to χmaxsubscript𝜒max\chi_{\mathrm{max}}). The values of the radii are the same used for Figs. 789. We stress here that the error bars depicted in Figs. 789, associated to the MFs calculated for each value of the threshold are not independent. Also, since the realisations of the random fields on different radial shells are not statistically independent, the MFs on different radii are not independent either. To give a flavour of the correlations between the MFs, in Fig. 11 we show the elements of the correlation matrix, i.e. the Pearson correlation coefficients rijsubscript𝑟𝑖𝑗r_{ij} calculated as

rij=CovijCoviiCovjj,subscript𝑟𝑖𝑗subscriptCov𝑖𝑗subscriptCov𝑖𝑖subscriptCov𝑗𝑗\displaystyle r_{ij}=\frac{\mathrm{Cov}_{ij}}{\sqrt{\mathrm{Cov}_{ii}}\sqrt{\mathrm{Cov}_{jj}}},\quad i,j=0,3nχnνformulae-sequence𝑖𝑗03subscript𝑛𝜒subscript𝑛𝜈\displaystyle i,j=0,\cdots 3\cdot n_{\chi}\cdot n_{\nu} (58)

from the covariance matrix CovijsubscriptCov𝑖𝑗\mathrm{Cov}_{ij} of our numerical estimates of the MFs, Eq. VI.2. It follows from the Cauchy-Schwarz inequality that 1rij11subscript𝑟𝑖𝑗1-1\leq r_{ij}\leq 1. A value of rij=1subscript𝑟𝑖𝑗1r_{ij}=1 indicates a perfect linear correlation between the two variables i𝑖i and j𝑗j; a common interpretation is that in this case all data points in a sample lie on a straight line. This is also true if rij=1subscript𝑟𝑖𝑗1r_{ij}=-1, but the slope of the line is negative. A vanishing correlation coefficient implies that there is no linear correlation. If the correlation coefficient is positive, deviations of both variables from the mean tend to have the same sign, whereas opposite signs lead to a negative correlation coefficient.

As expected, we notice in particular a strong anti-correlation for V0subscript𝑉0V_{0} centered around ν=0𝜈0\nu=0, as was expected by looking at Fig. 7. The same is true for V2subscript𝑉2V_{2} (cf. Fig. 9), while V1subscript𝑉1V_{1} is strongly positively correlated (cf. Fig 8). This high amount of (anti)correlation suggests that in the Gaussian case analysed here it is not necessary to consider a very high number of threshold values; however, this may not be true in the non-Gaussian case, where a higher resolution in the threshold values may be important to identify non-Gaussian features. Crucially important is, in all cases, a sufficient resolution in the HEALPix maps used at the beginning for the generation of the random fields, and later for the calculation of the MFs (in our estimates, we used the HEALPix parameters Nside=256subscript𝑁side256N_{\mathrm{side}}=256 and max=3Nside=768subscriptmax3subscript𝑁side768\ell_{\mathrm{max}}=3N_{\mathrm{side}}=768). This affects considerably the speed of the numerical implementation of these computations, however as mentioned earlier in Sec. V the generation of random fields and, separately, the calculation of the MFs (both happening at each realisation and at each radius) are embarrassingly parallel problems; this can be leveraged in practical implementations by employing parallelisation across multiple cores and nodes, without the need to worry about inter-process communication.

VI.3 Inference from Minkowski functionals of Gaussian fields

Introduced in cosmology by Mecke et al. (1994), the main applications of MFs so far have been as probes of primordial non-Gaussianities (Schmalzing and Buchert, 1997; Winitzki and Kosowsky, 1998; Schmalzing and Gorski, 1998), widely used in two and three dimensions, for instance on WMAP CMB data (Hikage et al., 2008), Planck CMB data (Ducout et al., 2013; Planck Collaboration et al., 2014, 2016; Novaes et al., 2016; Buchert et al., 2017) and on the SDSS galaxy catalogue (Park et al., 2005; Hikage et al., 2006). In the CMB case, MFs are suboptimal estimators of primordial non-Gaussianity parameters, while it has been shown that polyspectra provide minimum error bars for weak levels of non Gaussianity (Babich, 2005). Nevertheless, MFs constitute an attractive alternative to an analysis with polyspectra for a number of reasons: firstly, contrary to the bispectrum, MFs are defined in configuration rather than in Fourier space, so that a robust implementation for MFs becomes in practice easier to achieve; secondly, MFs are sensitive to the full hierarchy of higher order correlations, instead of third order only, and can provide additional information on all the non-linear coupling parameters fNLsubscript𝑓𝑁𝐿f_{NL}, gNLsubscript𝑔𝑁𝐿g_{NL}, … which appear in the perturbative development of the primordial curvature perturbation (Komatsu and Spergel, 2001; Okamoto and Hu, 2002); additionally, MFs can be analytically determined for Gaussian random fields; lastly, they are additive, which makes accounting for complicated survey geometries much easier compared to estimators of polyspectra.

In this work we propose (for the first time, as to our knowledge) MFs as an alternative probe of Gaussianity, in addition to non Gaussianity, in the sense specified in the following. We show how, assuming our MFs to be Gaussian distributed, we can use the MFs to probe the cosmology dependence of the fields realisations. This can be leveraged in future work to develop a full cosmological inference process based on the MFs calculated on lensing fields, of which we provide a first example here.

From a Bayesian perspective, assuming that our likelihood L(Vi|Ω)𝐿conditionalsubscript𝑉𝑖ΩL(V_{i}|\Omega) (the probability of having MFs Visubscript𝑉𝑖V_{i} given the cosmological parameters ΩΩ\Omega) is Gaussian is equivalent, considering a flat prior p(Ω)𝑝Ωp(\Omega) on the cosmological parameters ΩΩ\Omega, to having a Gaussian posterior p(Ω|Vi)𝑝conditionalΩsubscript𝑉𝑖p(\Omega|V_{i}), since by virtue of Bayes theorem

p(Vi|Ω)L(Ω|Vi)p(Ω).proportional-to𝑝conditionalsubscript𝑉𝑖Ω𝐿conditionalΩsubscript𝑉𝑖𝑝Ω\displaystyle p(V_{i}|\Omega)\propto L(\Omega|V_{i})p(\Omega). (59)

It follows that we are allowed to consider the likelihood and the posterior equivalently. In the Gaussian case, defining =lnLln𝐿\mathcal{L}=-\mathrm{ln}L and ignoring additive constants, we have that 2=χ22superscript𝜒2-2\mathcal{L}=\chi^{2}, where the χ2superscript𝜒2\chi^{2} can be evaluated as

χ2(θ)=i,j=1nνnχ(Vi(θ)Vi(θ0))covij1(θ0)(Vj(θ)Vj(θ0)),superscript𝜒2𝜃superscriptsubscript𝑖𝑗1subscript𝑛𝜈subscript𝑛𝜒delimited-⟨⟩subscript𝑉𝑖𝜃delimited-⟨⟩subscript𝑉𝑖subscript𝜃0subscriptsuperscriptcov1𝑖𝑗subscript𝜃0delimited-⟨⟩subscript𝑉𝑗𝜃delimited-⟨⟩subscript𝑉𝑗subscript𝜃0\displaystyle\chi^{2}(\theta)=\sum_{i,j=1}^{n_{\nu}\cdot n_{\chi}}\Bigg{(}\langle V_{i}(\theta)\rangle-\langle V_{i}(\theta_{0})\rangle\Bigg{)}\;\mathrm{cov}^{-1}_{ij}\left(\theta_{0}\right)\Bigg{(}\langle V_{j}(\theta)\rangle-\langle V_{j}(\theta_{0})\rangle\Bigg{)}, (60)

where the averages are performed over the number of realisations nRsubscript𝑛𝑅n_{R}, while the indices i,j𝑖𝑗i,j run over the length of our data vector, i.e. we consider the MFs evaluated at all the nνsubscript𝑛𝜈n_{\nu} thresholds and all the nχsubscript𝑛𝜒n_{\chi} radii. The MFs depend on the cosmological parameters and so does the covariance matrix; for the calculation of the chi-square, we use the inverse evaluated at the fiducial model θ0subscript𝜃0\theta_{0}.

We calculate the χ2superscript𝜒2\chi^{2} statistics with MFs obtained from the realisations of the lensing random fields (we consider the convergence in this example) at different values of one cosmological parameter, for simplicity. We consider 11 values of ΩmsubscriptΩm\Omega_{\mathrm{m}}, ranging from 0.25 to 0.35 in equidistant intervals of 0.01 centered on the fiducial value of 0.3. For each of the ΩmsubscriptΩm\Omega_{\mathrm{m}} values we produce our 3D cosmic shear covariance matrix following the equations in Sec. II, with either the Levin or the GLaSS method. Once the full lensing covariance matrix is available, we use it to generate, according to the procedure described in Sec. V, nRsubscript𝑛𝑅n_{R} realisations of the convergence field at rNχsubscript𝑟subscript𝑁𝜒r_{N_{\chi}} values of the radius in configuration space. On each shell and for each realisation we also calculate the associated MFs, and store them in memory. Subsequently we use them to build the full covariance matrix, exactly as the one shown in the previous subsection, however this time we will have one covariance matrix of the MFs for each starting value of ΩmsubscriptΩm\Omega_{\mathrm{m}}. Inverting the covariance corresponding to our fiducial value Ωm=0.3subscriptΩm0.3\Omega_{\mathrm{m}}=0.3, we can then use it to calculate the χ2superscript𝜒2\chi^{2} following Eq. 60.

The calculation of this inverse covariance matrix poses a numerical problem, in that its entries are very small and standard methods such as Gaussian elimination fail in producing a sensible inverse. We use therefore a Moore-Penrose pseudo-inverse matrix (Dresden, 1920; Penrose, 1955), after checking that it effectively produces an inverse covariance matrix that, multiplied by the covariance, gives back the identity matrix to within numerical precision.

We calculate the χ2superscript𝜒2\chi^{2} isolating the different MFs in our data vector. This implies isolating from the full covariance matrix the relevant sub-blocks for the auto-correlation of V0,V1subscript𝑉0subscript𝑉1V_{0},V_{1} and V2subscript𝑉2V_{2} (which we will in the following schematically indicate with V0,V0,V1,V1,V2,V2subscript𝑉0subscript𝑉0subscript𝑉1subscript𝑉1subscript𝑉2subscript𝑉2\left\langle V_{0},V_{0}\right\rangle,\left\langle V_{1},V_{1}\right\rangle,\left\langle V_{2},V_{2}\right\rangle, or Cov(V0,V0),Cov(V1,V1),Cov(V2,V2)Covsubscript𝑉0subscript𝑉0Covsubscript𝑉1subscript𝑉1Covsubscript𝑉2subscript𝑉2\mathrm{Cov}(V_{0},V_{0}),\,\mathrm{Cov}(V_{1},V_{1}),\,\mathrm{Cov}(V_{2},V_{2})). These sub-blocks can be visualised by looking at the corresponding sub-blocks in the covariance matrix plotted in Fig.10 (e.g. the correlation V0,V0subscript𝑉0subscript𝑉0\left\langle V_{0},V_{0}\right\rangle isolates the top left corner block); in Fig. 12 we plot the χ2superscript𝜒2\chi^{2} curves obtained with the three MFs. We notice how the χ2superscript𝜒2\chi^{2} increases going from V0subscript𝑉0V_{0} to V2subscript𝑉2V_{2}.

VII Conclusions

Refer to caption
Figure 10: Covariance matrix between different MFs at different radii. We consider correlations between all three MFs V0subscript𝑉0V_{0}, V1subscript𝑉1V_{1}, V2subscript𝑉2V_{2}, all functions of the threshold ν𝜈\nu (ranging from ν1subscript𝜈1\nu_{1} to νmaxsubscript𝜈max\nu_{\mathrm{max}}), as calculated at different radii (labelled by different χ𝜒\chi values, ranging from χ1subscript𝜒1\chi_{1} to χmaxsubscript𝜒max\chi_{\mathrm{max}} and specifically equal to 1000, 2000, 3000, 4000 and 5000 Mpc/hh, as in Figs. 7, 8, 9). In the matrix we indicate the block sub-matrices that represent the covariance between the three MFs. We used a logarithmic scale for both positive and negative values to highlight the many orders of magnitude spanned by the entries of the matrix and the different contibutions given by the three MFs.
Refer to caption
Figure 11: Correlation matrix between different MFs at different radii; the matrix entries represent the Pearson correlation coefficient, obtained from the covariance matrix entries (the same plotted in in Fig. 10) following Eq. 58. We consider correlations between all three MFs V0subscript𝑉0V_{0}, V1subscript𝑉1V_{1}, V2subscript𝑉2V_{2}, all functions of the threshold ν𝜈\nu (ranging from ν1subscript𝜈1\nu_{1} to νmaxsubscript𝜈max\nu_{\mathrm{max}}), as calculated at different radii (labelled by different χ𝜒\chi values, ranging from χ1subscript𝜒1\chi_{1} to χmaxsubscript𝜒max\chi_{\mathrm{max}} and specifically equal to 1000, 2000, 3000, 4000 and 5000 Mpc/hh, as in Figs. 7, 8, 9). In the matrix we indicate the block sub-matrices that represent the correlation between the three MFs.
Refer to caption
Figure 12: χ2superscript𝜒2\chi^{2} obtained considering the covariance of different combinations of MFs, i.e. considering the three MFs singularly (V0,V0subscript𝑉0subscript𝑉0\left\langle V_{0},V_{0}\right\rangle (blue), V1,V1subscript𝑉1subscript𝑉1\left\langle V_{1},V_{1}\right\rangle (green) and V2,V2subscript𝑉2subscript𝑉2\left\langle V_{2},V_{2}\right\rangle (red)). Our fiducial model is represented by the choice Ωm=0.3subscriptΩm0.3\Omega_{\mathrm{m}}=0.3.

3D cosmic shear constitutes an alternative to a traditional tomographic analysis of a cosmic shear survey. The spherical-Bessel expansion of the shear field at the core of its formalism maximises the amount of redshift information; however, the calculation of the covariance matrices presents numerical difficulties due to the numerous integrations over highly oscillatory functions.

In this paper we described and compared two methods for the calculation of simulated 3D cosmic shear covariance matrices. While the first method implements the Levin technique for integration of the periodic oscillations of the Bessel functions, the second method, implemented in the code GLaSS, tackles the integrations by matrix multiplications and appropriate use of the Limber approximation.

We first compared the predictions of the two codes in terms of covariance matrices and found excellent agreement. We compared the output of the codes both in terms of the total signal-to-noise ratio and the single contributions to the covariance matrices Csubscript𝐶C_{\ell}, for two different values of the multipole \ell, for both the signal and noise parts.

Once tested the accuracy of the predictions for the covariance matrices, we used the simulated matrices to generate Gaussian lensing fields on the sky. The procedure we described, based on a Cholesky decomposition of the Csubscript𝐶C_{\ell} matrices, allowed us to generate correlated Gaussian fields at different slices in comoving distance. We remark here that in our formalism for 3D cosmic shear we ignored complications arising from survey masks, assuming full sky coverage. Masked data can be readily accounted for by applying a mixing matrix/pseudo-Csubscript𝐶C_{\ell} like formalism. For 3D cosmic shear this is described in Kitching et al. (2014), where it was applied to data. Both methods for the computation of 3D cosmic shear covariance matrices studied in this paper can be trivially modified to include such an effect. In our work we did not investigate masked data and the ability of a pseudo-Csubscript𝐶C_{\ell} method to produce power spectra that account for a mask; however, thanks to our method to generate 3D cosmic shear data, we could now test this pseudo-Csubscript𝐶C_{\ell} approach by masking the simulated data and we will investigate this in a future paper.

The generation of normal and lognormal fields (the latter being easily obtainable from the former, by exponentiation of the Gaussian maps) can be used in future work to compute a realistic covariance matrix for a full 3D cosmic shear likelihood analysis. This should improve upon e.g. the CFHTLenS analysis for 3D cosmic shear (Kitching et al., 2014), where a covariance implementation similar to GLaSS was used. Kitching et al. (2014) constructed a likelihood, in which the parameter dependency was in the covariance rather than the mean shear transform coefficients. This could be improved by having a likelihood in which the covariance is used as the mean and the 4-point covariance of the covariance used.

The generation of normal and lognormal random fields, starting from the 3D cosmic shear covariance matrices, also constitutes the first step for the development of a Bayesian Hierarchical Model for 3D cosmic shear power spectra estimation (following e.g. the work of Alsing et al., 2016, 2017, and extending it to a spherical-Bessel formalism), which can be investigated in future work.

In this work we tested our random field generation procedure by calculating Minkowski Functionals associated to our Gaussian random fields and comparing them with their known expectation values. We found good agreement between our numerical estimates and their theoretical expectation values. We calculated our Minkowski Functionals separately on each spherical shell, however we stress here that the realisations of the random fields on different radial shells are not statistically independent, as one can appreciate from the correlation matrix presented in Fig. 11. Future work should concentrate on estimating the full correlation between the Minkowski Functionals at different values of the radii, implementing a fully three-dimensional approach for their calculation (see e.g. Hikage et al., 2003; Gleser et al., 2006; Yoshiura et al., 2017; Appleby et al., 2018, for examples of Minkowski Functionals in 3D). Producing fully 3D Minkowski functionals for a lognormal field in 3D can be used in particular to extract non-Gaussian information from the shear field.

Finally, we showed how Minkowski Functionals can also be used to extract Gaussian information by means of a likelihood analysis. We show an example of this in Fig. 12, where we plot the χ2superscript𝜒2\chi^{2} obtained from the covariance of the different Minkowski Functionals as a function of the varying cosmological parameter ΩmsubscriptΩ𝑚\Omega_{m}. This is a first example of a full cosmological inference process, making use of the Minkowski Functionals, that we plan to develop in future work.

Acknowledgements

ASM and RR acknowledge financial support from the graduate college Astrophysics of cosmological probes of gravity by Landesgraduiertenakademie Baden-Württemberg. PLT is supported by the UK Science and Technology Facilities Council. TK is supported by a Royal Society University Research Fellowship. The authors thank the German DFG Excellence Initiative for funding a mobility program, through the grant Gravity on the largest scales and the cosmic large-scale structure, within which this work originated.

References

  • Abell et al. [2009] P. A. Abell, J. Allison, S. F. Anderson, J. R. Andrew, J. R. P. Angel, L. Armus, D. Arnett, S. J. Asztalos, T. S. Axelrod, and et al. LSST Science Book, Version 2.0. ArXiv e-prints, December 2009.
  • Abramowitz et al. [1988] M. Abramowitz, I. A. Stegun, and R. H. Romer. Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. American Journal of Physics, 56:958–958, October 1988. doi: 10.1119/1.15378.
  • Alsing et al. [2016] J. Alsing, A. Heavens, A. H. Jaffe, A. Kiessling, B. Wandelt, and T. Hoffmann. Hierarchical cosmic shear power spectrum inference. MNRAS, 455:4452–4466, February 2016. doi: 10.1093/mnras/stv2501.
  • Alsing et al. [2017] J. Alsing, A. Heavens, and A. H. Jaffe. Cosmological parameters, shear maps and power spectra from CFHTLenS using Bayesian hierarchical inference. MNRAS, 466:3272–3292, April 2017. doi: 10.1093/mnras/stw3161.
  • Amendola et al. [2016] L. Amendola, S. Appleby, A. Avgoustidis, D. Bacon, T. Baker, M. Baldi, N. Bartolo, A. Blanchard, C. Bonvin, S. Borgani, E. Branchini, C. Burrage, S. Camera, C. Carbone, L. Casarini, M. Cropper, C. de Rham, J. P. Dietrich, C. Di Porto, R. Durrer, A. Ealet, P. G. Ferreira, F. Finelli, J. Garcia-Bellido, T. Giannantonio, L. Guzzo, A. Heavens, L. Heisenberg, C. Heymans, H. Hoekstra, L. Hollenstein, R. Holmes, O. Horst, Z. Hwang, K. Jahnke, T. D. Kitching, T. Koivisto, M. Kunz, G. La Vacca, E. Linder, M. March, V. Marra, C. Martins, E. Majerotto, D. Markovic, D. Marsh, F. Marulli, R. Massey, Y. Mellier, F. Montanari, D. F. Mota, N. J. Nunes, W. Percival, V. Pettorino, C. Porciani, C. Quercellini, J. Read, M. Rinaldi, D. Sapone, I. Sawicki, R. Scaramella, C. Skordis, F. Simpson, A. Taylor, S. Thomas, R. Trotta, L. Verde, F. Vernizzi, A. Vollmer, Y. Wang, J. Weller, and T. Zlosnik. Cosmology and Fundamental Physics with the Euclid Satellite. ArXiv e-prints, June 2016.
  • Appleby et al. [2018] S. Appleby, P. Chingangbam, C. Park, K. P. Yogendran, and P. K. Joby. Minkowski tensors in three dimensions: Probing the anisotropy generated by redshift space distortion. The Astrophysical Journal, 863(2):200, 2018. URL http://stacks.iop.org/0004-637X/863/i=2/a=200.
  • Ayaita et al. [2012] Y. Ayaita, B. M. Schäfer, and M. Weber. Investigating clustering dark energy with 3D weak cosmic shear. MNRAS, 422:3056–3066, June 2012. doi: 10.1111/j.1365-2966.2012.20822.x.
  • Babich [2005] D. Babich. Optimal estimation of non-Gaussianity. Phys. Rev. D, 72(4):043003, August 2005. doi: 10.1103/PhysRevD.72.043003.
  • Bacon et al. [2000] D. J. Bacon, A. R. Refregier, and R. S. Ellis. Detection of weak gravitational lensing by large-scale structure. MNRAS, 318:625–640, October 2000. doi: 10.1046/j.1365-8711.2000.03851.x.
  • Bartelmann and Schneider [2001] M. Bartelmann and P. Schneider. Weak gravitational lensing. Phys. Rep., 340:291–472, January 2001. doi: 10.1016/S0370-1573(00)00082-X.
  • Blas et al. [2011] D. Blas, J. Lesgourgues, and T. Tram. The Cosmic Linear Anisotropy Solving System (CLASS). Part II: Approximation schemes. JCAP, 7:034, July 2011. doi: 10.1088/1475-7516/2011/07/034.
  • Brown et al. [2003] M. L. Brown, A. N. Taylor, D. J. Bacon, M. E. Gray, S. Dye, K. Meisenheimer, and C. Wolf. The shear power spectrum from the COMBO-17 survey. MNRAS, 341:100–118, May 2003. doi: 10.1046/j.1365-8711.2003.06237.x.
  • Buchert et al. [2017] T. Buchert, M. J. France, and F. Steiner. Model-independent analyses of non-Gaussianity in Planck CMB maps using Minkowski functionals. Classical and Quantum Gravity, 34(9):094002, May 2017. doi: 10.1088/1361-6382/aa5ce2.
  • Castro et al. [2005] P. G. Castro, A. F. Heavens, and T. D. Kitching. Weak lensing analysis in three dimensions. Phys. Rev. D, 72(2):023516, July 2005. doi: 10.1103/PhysRevD.72.023516.
  • Dresden [1920] Arnold Dresden. The fourteenth western meeting of the american mathematical society. Bull. Amer. Math. Soc., 26(9):385–396, 06 1920. URL https://projecteuclid.org:443/euclid.bams/1183425340.
  • Ducout et al. [2013] A. Ducout, F. R. Bouchet, S. Colombi, D. Pogosyan, and S. Prunet. Non-Gaussianity and Minkowski functionals: forecasts for Planck. MNRAS, 429:2104–2126, March 2013. doi: 10.1093/mnras/sts483.
  • Fluri et al. [2018] J. Fluri, T. Kacprzak, R. Sgier, A. Refregier, and A. Amara. Weak lensing peak statistics in the era of large scale cosmological surveys. JCAP, 10:051, October 2018. doi: 10.1088/1475-7516/2018/10/051.
  • Gleser et al. [2006] L. Gleser, A. Nusser, B. Ciardi, and V. Desjacques. The morphology of cosmological reionization by means of Minkowski functionals. MNRAS, 370:1329–1338, August 2006. doi: 10.1111/j.1365-2966.2006.10556.x.
  • Goldberg et al. [1967] J. N. Goldberg, A. J. Macfarlane, E. T. Newman, F. Rohrlich, and E. C. G. Sudarshan. Spins𝑠-s spherical harmonics and ðitalic-ð\eth. Journal of Mathematical Physics, 8(11):2155–2161, 1967. doi: 10.1063/1.1705135. URL http://dx.doi.org/10.1063/1.1705135.
  • Górski et al. [2005] K. M. Górski, E. Hivon, A. J. Banday, B. D. Wandelt, F. K. Hansen, M. Reinecke, and M. Bartelmann. HEALPix: A Framework for High-Resolution Discretization and Fast Analysis of Data Distributed on the Sphere. ApJ, 622:759–771, April 2005. doi: 10.1086/427976.
  • Grassi and Schäfer [2014] A. Grassi and B. M. Schäfer. Detecting baryon acoustic oscillations by 3d weak lensing. MNRAS, 437:2632–2641, January 2014. doi: 10.1093/mnras/stt2075.
  • Hadwiger [1957] H. Hadwiger. Vorlesungen über Inhalt, Oberfläche und Isoperimetrie. Springer, 1957.
  • Heavens [2003] A. Heavens. 3D weak lensing. MNRAS, 343:1327–1334, August 2003. doi: 10.1046/j.1365-8711.2003.06780.x.
  • Heavens et al. [2006] A. F. Heavens, T. D. Kitching, and A. N. Taylor. Measuring dark energy properties with 3D cosmic shear. MNRAS, 373:105–120, November 2006. doi: 10.1111/j.1365-2966.2006.11006.x.
  • Heymans et al. [2013] C. Heymans, E. Grocutt, A. Heavens, M. Kilbinger, T. D. Kitching, F. Simpson, J. Benjamin, T. Erben, H. Hildebrandt, H. Hoekstra, Y. Mellier, L. Miller, L. Van Waerbeke, M. L. Brown, J. Coupon, L. Fu, J. Harnois-Déraps, M. J. Hudson, K. Kuijken, B. Rowe, T. Schrabback, E. Semboloni, S. Vafaei, and M. Velander. CFHTLenS tomographic weak lensing cosmological parameter constraints: Mitigating the impact of intrinsic galaxy alignments. MNRAS, 432:2433–2453, July 2013. doi: 10.1093/mnras/stt601.
  • Hikage et al. [2003] C. Hikage, J. Schmalzing, T. Buchert, Y. Suto, I. Kayo, A. Taruya, M. S. Vogeley, F. Hoyle, J. R. Gott, III, and J. Brinkmann. Minkowski Functionals of SDSS Galaxies I : Analysis of Excursion Sets. PASJ, 55:911–931, October 2003. doi: 10.1093/pasj/55.5.911.
  • Hikage et al. [2006] C. Hikage, E. Komatsu, and T. Matsubara. Primordial Non-Gaussianity and Analytical Formula for Minkowski Functionals of the Cosmic Microwave Background and Large-Scale Structure. ApJ, 653:11–26, December 2006. doi: 10.1086/508653.
  • Hikage et al. [2008] C. Hikage, T. Matsubara, P. Coles, M. Liguori, F. K. Hansen, and S. Matarrese. Limits on primordial non-gaussianity from minkowski functionals of the wmap temperature anisotropies. Monthly Notices of the Royal Astronomical Society, 389(3):1439–1446, 2008. doi: 10.1111/j.1365-2966.2008.13674.x. URL http://dx.doi.org/10.1111/j.1365-2966.2008.13674.x.
  • Hildebrandt et al. [2017] H. Hildebrandt, M. Viola, C. Heymans, S. Joudaki, K. Kuijken, C. Blake, T. Erben, B. Joachimi, D. Klaes, L. Miller, C. B. Morrison, R. Nakajima, G. Verdoes Kleijn, A. Amon, A. Choi, G. Covone, J. T. A. de Jong, A. Dvornik, I. Fenech Conti, A. Grado, J. Harnois-Déraps, R. Herbonnet, H. Hoekstra, F. Köhlinger, J. McFarland, A. Mead, J. Merten, N. Napolitano, J. A. Peacock, M. Radovich, P. Schneider, P. Simon, E. A. Valentijn, J. L. van den Busch, E. van Uitert, and L. Van Waerbeke. KiDS-450: cosmological parameter constraints from tomographic weak gravitational lensing. MNRAS, 465:1454–1498, February 2017. doi: 10.1093/mnras/stw2805.
  • Hoekstra and Jain [2008] H. Hoekstra and B. Jain. Weak Gravitational Lensing and Its Cosmological Applications. Annual Review of Nuclear and Particle Science, 58:99–123, November 2008. doi: 10.1146/annurev.nucl.58.110707.171151.
  • Hu [1999] W. Hu. Power Spectrum Tomography with Weak Lensing. ApJ, 522:L21–L24, September 1999. doi: 10.1086/312210.
  • Joudaki et al. [2017] S. Joudaki, C. Blake, C. Heymans, A. Choi, J. Harnois-Deraps, H. Hildebrandt, B. Joachimi, A. Johnson, A. Mead, D. Parkinson, M. Viola, and L. van Waerbeke. CFHTLenS revisited: assessing concordance with Planck including astrophysical systematics. MNRAS, 465:2033–2052, February 2017. doi: 10.1093/mnras/stw2665.
  • Kaiser [1992] N. Kaiser. Weak gravitational lensing of distant galaxies. ApJ, 388:272–286, April 1992. doi: 10.1086/171151.
  • Kaiser [1998] N. Kaiser. Weak Lensing and Cosmology. ApJ, 498:26–42, May 1998. doi: 10.1086/305515.
  • Kilbinger [2015] M. Kilbinger. Cosmology with cosmic shear observations: a review. Reports on Progress in Physics, 78(8):086901, July 2015. doi: 10.1088/0034-4885/78/8/086901.
  • Kitching and Heavens [2017] T. D. Kitching and A. F. Heavens. Unequal-time correlators for cosmology. Phys. Rev. D, 95(6):063522, March 2017. doi: 10.1103/PhysRevD.95.063522.
  • Kitching et al. [2007] T. D. Kitching, A. F. Heavens, A. N. Taylor, M. L. Brown, K. Meisenheimer, C. Wolf, M. E. Gray, and D. J. Bacon. Cosmological constraints from COMBO-17 using 3D weak lensing. MNRAS, 376:771–778, April 2007. doi: 10.1111/j.1365-2966.2007.11473.x.
  • Kitching et al. [2014] T. D. Kitching, A. F. Heavens, J. Alsing, T. Erben, C. Heymans, H. Hildebrandt, H. Hoekstra, A. Jaffe, A. Kiessling, Y. Mellier, L. Miller, L. van Waerbeke, J. Benjamin, J. Coupon, L. Fu, M. J. Hudson, M. Kilbinger, K. Kuijken, B. T. P. Rowe, T. Schrabback, E. Semboloni, and M. Velander. 3D cosmic shear: cosmology from CFHTLenS. MNRAS, 442:1326–1349, August 2014. doi: 10.1093/mnras/stu934.
  • Kitching et al. [2015] T. D. Kitching, A. F. Heavens, and S. Das. 3D weak gravitational lensing of the CMB and galaxies. MNRAS, 449:2205–2214, May 2015. doi: 10.1093/mnras/stv193.
  • Kitching et al. [2017] T. D. Kitching, J. Alsing, A. F. Heavens, R. Jimenez, J. D. McEwen, and L. Verde. The limits of cosmic shear. MNRAS, 469:2737–2749, August 2017. doi: 10.1093/mnras/stx1039.
  • Köhlinger et al. [2017] F. Köhlinger, M. Viola, B. Joachimi, H. Hoekstra, E. van Uitert, H. Hildebrandt, A. Choi, T. Erben, C. Heymans, S. Joudaki, D. Klaes, K. Kuijken, J. Merten, L. Miller, P. Schneider, and E. A. Valentijn. KiDS-450: the tomographic weak lensing power spectrum and constraints on cosmological parameters. MNRAS, 471:4412–4435, November 2017. doi: 10.1093/mnras/stx1820.
  • Komatsu and Spergel [2001] E. Komatsu and D. N. Spergel. Acoustic signatures in the primary microwave background bispectrum. Phys. Rev. D, 63(6):063002, March 2001. doi: 10.1103/PhysRevD.63.063002.
  • Kosowsky [1998] Arthur Kosowsky. Efficient computation of hyperspherical bessel functions. arXiv preprint astro-ph/9805173, 1998.
  • Lanusse et al. [2015] F. Lanusse, A. Rassat, and J.-L. Starck. 3D galaxy clustering with future wide-field surveys: Advantages of a spherical Fourier-Bessel analysis. A&A, 578:A10, June 2015. doi: 10.1051/0004-6361/201424456.
  • Laureijs et al. [2011] R. Laureijs, J. Amiaux, S. Arduini, J. . Auguères, J. Brinchmann, R. Cole, M. Cropper, C. Dabin, L. Duvet, A. Ealet, and et al. Euclid Definition Study Report. ArXiv e-prints, October 2011.
  • Leistedt et al. [2015] B. Leistedt, J. D. McEwen, T. D. Kitching, and H. V. Peiris. 3D weak lensing with spin wavelets on the ball. Phys. Rev. D, 92(12):123010, December 2015. doi: 10.1103/PhysRevD.92.123010.
  • Lesgourgues [2011] J. Lesgourgues. The Cosmic Linear Anisotropy Solving System (CLASS) III: Comparision with CAMB for LambdaCDM. ArXiv e-prints, April 2011.
  • Levin [1996] David Levin. Fast integration of rapidly oscillatory functions. Journal of Computational and Applied Mathematics, 67(1):95 – 101, 1996. ISSN 0377-0427. doi: https://doi.org/10.1016/0377-0427(94)00118-9. URL http://www.sciencedirect.com/science/article/pii/0377042794001189.
  • Levin [1997] David Levin. Analysis of a collocation method for integrating rapidly oscillatory functions. Journal of Computational and Applied Mathematics, 78(1):131 – 138, 1997. ISSN 0377-0427. doi: https://doi.org/10.1016/S0377-0427(96)00137-9. URL http://www.sciencedirect.com/science/article/pii/S0377042796001379.
  • Lewis et al. [2000] Antony Lewis, Anthony Challinor, and Anthony Lasenby. Efficient computation of CMB anisotropies in closed FRW models. Astrophys. J., 538:473–476, 2000. doi: 10.1086/309179.
  • Lim and Simon [2012] E. A. Lim and D. Simon. Can we detect hot/cold spots in the CMB with Minkowski Functionals? JCAP, 1:048, January 2012. doi: 10.1088/1475-7516/2012/01/048.
  • Lin and Kilbinger [2015] C.-A. Lin and M. Kilbinger. A new model to predict weak-lensing peak counts. I. Comparison with N-body simulations. A&A, 576:A24, April 2015. doi: 10.1051/0004-6361/201425188.
  • Loverde and Afshordi [2008] M. Loverde and N. Afshordi. Extended Limber approximation. Phys. Rev. D, 78(12):123506, December 2008. doi: 10.1103/PhysRevD.78.123506.
  • LoVerde and Afshordi [2008] Marilena LoVerde and Niayesh Afshordi. Extended limber approximation. Physical Review D, 78(12):123506, 2008.
  • Mecke et al. [1994] K. R. Mecke, T. Buchert, and H. Wagner. Robust morphological measures for large-scale structure in the Universe. A&A, 288:697–704, August 1994.
  • Merkel and Schäfer [2013] P. M. Merkel and B. M. Schäfer. Intrinsic alignments and 3d weak gravitational lensing. MNRAS, 434:1808–1820, September 2013. doi: 10.1093/mnras/stt1151.
  • Newman and Penrose [1962] Ezra Newman and Roger Penrose. An approach to gravitational radiation by a method of spin coefficients. Journal of Mathematical Physics, 3(3):566–578, 1962.
  • Novaes et al. [2016] C. P. Novaes, A. Bernui, G. A. Marques, and I. S. Ferreira. Local analyses of Planck maps with Minkowski functionals. MNRAS, 461:1363–1373, September 2016. doi: 10.1093/mnras/stw1427.
  • Okamoto and Hu [2002] T. Okamoto and W. Hu. Angular trispectra of CMB temperature and polarization. Phys. Rev. D, 66(6):063008, September 2002. doi: 10.1103/PhysRevD.66.063008.
  • Park et al. [2005] Changbom Park, Yun-Young Choi, Michael S. Vogeley, J. Richard Gott III, Juhan Kim, Chiaki Hikage, Takahiko Matsubara, Myeong-Gu Park, Yasushi Suto, David H. Weinberg, and The SDSS Collaboration. Topology analysis of the sloan digital sky survey. i. scale and luminosity dependence. The Astrophysical Journal, 633(1):11, 2005. URL http://stacks.iop.org/0004-637X/633/i=1/a=11.
  • Peebles [1980] P. J. E. Peebles. The large-scale structure of the universe. 1980.
  • Peel et al. [2017] A. Peel, C.-A. Lin, F. Lanusse, A. Leonard, J.-L. Starck, and M. Kilbinger. Cosmological constraints with weak-lensing peak counts and second-order statistics in a large-field survey. A&A, 599:A79, March 2017. doi: 10.1051/0004-6361/201629928.
  • Peel et al. [2018] Austin Peel, Valeria Pettorino, Carlo Giocoli, Jean-Luc Starck, and Marco Baldi. Breaking degeneracies in modified gravity with higher (than 2nd) order weak-lensing statistics. A&A, 619:A38, November 2018. doi: 10.1051/0004-6361/201833481.
  • Penrose [1955] R. Penrose. A generalized inverse for matrices. Mathematical Proceedings of the Cambridge Philosophical Society, 51(3):406–413, 1955. doi: 10.1017/S0305004100030401.
  • Planck Collaboration et al. [2014] Planck Collaboration, P. A. R. Ade, N. Aghanim, C. Armitage-Caplan, M. Arnaud, M. Ashdown, F. Atrio-Barandela, J. Aumont, C. Baccigalupi, A. J. Banday, R. B. Barreiro, J. G. Bartlett, N. Bartolo, E. Battaner, K. Benabed, A. Benoît, A. Benoit-Lévy, J. P. Bernard, M. Bersanelli, P. Bielewicz, J. Bobin, J. J. Bock, A. Bonaldi, L. Bonavera, J. R. Bond, J. Borrill, F. R. Bouchet, M. Bridges, M. Bucher, C. Burigana, R. C. Butler, J. F. Cardoso, A. Catalano, A. Challinor, A. Chamballu, H. C. Chiang, L. Y. Chiang, P. R. Christensen, S. Church, D. L. Clements, S. Colombi, L. P. L. Colombo, F. Couchot, A. Coulais, B. P. Crill, A. Curto, F. Cuttaia, L. Danese, R. D. Davies, R. J. Davis, P. de Bernardis, A. de Rosa, G. de Zotti, J. Delabrouille, J. M. Delouis, F. X. Désert, J. M. Diego, H. Dole, S. Donzelli, O. Doré, M. Douspis, A. Ducout, J. Dunkley, X. Dupac, G. Efstathiou, F. Elsner, T. A. Enßlin, H. K. Eriksen, J. Fergusson, F. Finelli, O. Forni, M. Frailis, E. Franceschi, S. Galeotta, K. Ganga, M. Giard, Y. Giraud-Héraud, J. González-Nuevo, K. M. Górski, S. Gratton, A. Gregorio, A. Gruppuso, F. K. Hansen, D. Hanson, D. Harrison, A. Heavens, S. Henrot-Versillé, C. Hernández-Monteagudo, D. Herranz, S. R. Hildebrandt, E. Hivon, M. Hobson, W. A. Holmes, A. Hornstrup, W. Hovest, K. M. Huffenberger, A. H. Jaffe, T. R. Jaffe, W. C. Jones, M. Juvela, E. Keihänen, R. Keskitalo, T. S. Kisner, J. Knoche, L. Knox, M. Kunz, H. Kurki-Suonio, F. Lacasa, G. Lagache, A. Lähteenmäki, J. M. Lamarre, A. Lasenby, R. J. Laureijs, C. R. Lawrence, J. P. Leahy, R. Leonardi, J. Lesgourgues, A. Lewis, M. Liguori, P. B. Lilje, M. Linden-Vørnle, M. López-Caniego, P. M. Lubin, J. F. Macías-Pérez, B. Maffei, D. Maino, N. Mandolesi, A. Mangilli, D. Marinucci, M. Maris, D. J. Marshall, P. G. Martin, E. Martínez-González, S. Masi, M. Massardi, S. Matarrese, F. Matthai, P. Mazzotta, P. R. Meinhold, A. Melchiorri, L. Mendes, A. Mennella, M. Migliaccio, S. Mitra, M. A. Miville-Deschênes, A. Moneti, L. Montier, G. Morgante, D. Mortlock, A. Moss, D. Munshi, J. A. Murphy, P. Naselsky, P. Natoli, C. B. Netterfield, H. U. Nørgaard- Nielsen, F. Noviello, D. Novikov, I. Novikov, S. Osborne, C. A. Oxborrow, F. Paci, L. Pagano, F. Pajot, D. Paoletti, F. Pasian, G. Patanchon, H. V. Peiris, O. Perdereau, L. Perotto, F. Perrotta, F. Piacentini, M. Piat, E. Pierpaoli, D. Pietrobon, S. Plaszczynski, E. Pointecouteau, G. Polenta, N. Ponthieu, L. Popa, T. Poutanen, G. W. Pratt, G. Prézeau, S. Prunet, J. L. Puget, J. P. Rachen, B. Racine, R. Rebolo, M. Reinecke, M. Remazeilles, C. Renault, A. Renzi, S. Ricciardi, T. Riller, I. Ristorcelli, G. Rocha, C. Rosset, G. Roudier, J. A. Rubiño-Martín, B. Rusholme, M. Sandri, D. Santos, G. Savini, D. Scott, M. D. Seiffert, E. P. S. Shellard, K. Smith, L. D. Spencer, J. L. Starck, V. Stolyarov, R. Stompor, R. Sudiwala, R. Sunyaev, F. Sureau, P. Sutter, D. Sutton, A. S. Suur-Uski, J. F. Sygnet, J. A. Tauber, D. Tavagnacco, L. Terenzi, L. Toffolatti, M. Tomasi, M. Tristram, M. Tucci, J. Tuovinen, L. Valenziano, J. Valiviita, B. Van Tent, J. Varis, P. Vielva, F. Villa, N. Vittorio, L. A. Wade, B. D. Wandelt, M. White, S. D. M. White, D. Yvon, A. Zacchei, and A. Zonca. Planck 2013 results. XXIV. Constraints on primordial non-Gaussianity. A&A, 571:A24, November 2014. doi: 10.1051/0004-6361/201321554.
  • Planck Collaboration et al. [2016] Planck Collaboration, P. A. R. Ade, N. Aghanim, M. Arnaud, F. Arroja, M. Ashdown, J. Aumont, C. Baccigalupi, M. Ballardini, A. J. Banday, R. B. Barreiro, N. Bartolo, S. Basak, E. Battaner, K. Benabed, A. Benoît, A. Benoit-Lévy, J. P. Bernard, M. Bersanelli, P. Bielewicz, J. J. Bock, A. Bonaldi, L. Bonavera, J. R. Bond, J. Borrill, F. R. Bouchet, F. Boulanger, M. Bucher, C. Burigana, R. C. Butler, E. Calabrese, J. F. Cardoso, A. Catalano, A. Challinor, A. Chamballu, H. C. Chiang, P. R. Christensen, S. Church, D. L. Clements, S. Colombi, L. P. L. Colombo, C. Combet, F. Couchot, A. Coulais, B. P. Crill, A. Curto, F. Cuttaia, L. Danese, R. D. Davies, R. J. Davis, P. de Bernardis, A. de Rosa, G. de Zotti, J. Delabrouille, F. X. Désert, J. M. Diego, H. Dole, S. Donzelli, O. Doré, M. Douspis, A. Ducout, X. Dupac, G. Efstathiou, F. Elsner, T. A. Enßlin, H. K. Eriksen, J. Fergusson, F. Finelli, O. Forni, M. Frailis, A. A. Fraisse, E. Franceschi, A. Frejsel, S. Galeotta, S. Galli, K. Ganga, C. Gauthier, T. Ghosh, M. Giard, Y. Giraud-Héraud, E. Gjerløw, J. González- Nuevo, K. M. Górski, S. Gratton, A. Gregorio, A. Gruppuso, J. E. Gudmundsson, J. Hamann, F. K. Hansen, D. Hanson, D. L. Harrison, A. Heavens, G. Helou, S. Henrot- Versillé, C. Hernández-Monteagudo, D. Herranz, S. R. Hildebrandt, E. Hivon, M. Hobson, W. A. Holmes, A. Hornstrup, W. Hovest, Z. Huang, K. M. Huffenberger, G. Hurier, A. H. Jaffe, T. R. Jaffe, W. C. Jones, M. Juvela, E. Keihänen, R. Keskitalo, J. Kim, T. S. Kisner, J. Knoche, M. Kunz, H. Kurki-Suonio, F. Lacasa, G. Lagache, A. Lähteenmäki, J. M. Lamarre, A. Lasenby, M. Lattanzi, C. R. Lawrence, R. Leonardi, J. Lesgourgues, F. Levrier, A. Lewis, M. Liguori, P. B. Lilje, M. Linden-Vørnle, M. López-Caniego, P. M. Lubin, J. F. Macías-Pérez, G. Maggio, D. Maino, N. Mandolesi, A. Mangilli, D. Marinucci, M. Maris, P. G. Martin, E. Martínez-González, S. Masi, S. Matarrese, P. McGehee, P. R. Meinhold, A. Melchiorri, L. Mendes, A. Mennella, M. Migliaccio, S. Mitra, M. A. Miville-Deschênes, A. Moneti, L. Montier, G. Morgante, D. Mortlock, A. Moss, M. Münchmeyer, D. Munshi, J. A. Murphy, P. Naselsky, F. Nati, P. Natoli, C. B. Netterfield, H. U. Nørgaard-Nielsen, F. Noviello, D. Novikov, I. Novikov, C. A. Oxborrow, F. Paci, L. Pagano, F. Pajot, D. Paoletti, F. Pasian, G. Patanchon, H. V. Peiris, O. Perdereau, L. Perotto, F. Perrotta, V. Pettorino, F. Piacentini, M. Piat, E. Pierpaoli, D. Pietrobon, S. Plaszczynski, E. Pointecouteau, G. Polenta, L. Popa, G. W. Pratt, G. Prézeau, S. Prunet, J. L. Puget, J. P. Rachen, B. Racine, R. Rebolo, M. Reinecke, M. Remazeilles, C. Renault, A. Renzi, I. Ristorcelli, G. Rocha, C. Rosset, M. Rossetti, G. Roudier, J. A. Rubiño-Martín, B. Rusholme, M. Sandri, D. Santos, M. Savelainen, G. Savini, D. Scott, M. D. Seiffert, E. P. S. Shellard, M. Shiraishi, K. Smith, L. D. Spencer, V. Stolyarov, R. Stompor, R. Sudiwala, R. Sunyaev, P. Sutter, D. Sutton, A. S. Suur-Uski, J. F. Sygnet, J. A. Tauber, L. Terenzi, L. Toffolatti, M. Tomasi, M. Tristram, A. Troja, M. Tucci, J. Tuovinen, L. Valenziano, J. Valiviita, B. Van Tent, P. Vielva, F. Villa, L. A. Wade, B. D. Wandelt, I. K. Wehus, D. Yvon, A. Zacchei, and A. Zonca. Planck 2015 results. XVII. Constraints on primordial non-Gaussianity. A&A, 594:A17, September 2016. doi: 10.1051/0004-6361/201525836.
  • Pratten et al. [2016] G. Pratten, D. Munshi, P. Valageas, and P. Brax. 3D weak lensing: Modified theories of gravity. Phys. Rev. D, 93(10):103524, May 2016. doi: 10.1103/PhysRevD.93.103524.
  • Schmalzing and Buchert [1997] J. Schmalzing and T. Buchert. Beyond Genus Statistics: A Unifying Approach to the Morphology of Cosmic Structure. ApJ, 482:L1–L4, June 1997. doi: 10.1086/310680.
  • Schmalzing and Gorski [1998] J. Schmalzing and K. M. Gorski. Minkowski functionals used in the morphological analysis of cosmic microwave background anisotropy maps. MNRAS, 297:355–365, June 1998. doi: 10.1046/j.1365-8711.1998.01467.x.
  • Seljak and Zaldarriaga [1996] U. Seljak and M. Zaldarriaga. A Line-of-Sight Integration Approach to Cosmic Microwave Background Anisotropies. ApJ, 469:437, October 1996. doi: 10.1086/177793.
  • Spurio Mancini et al. [2018] A. Spurio Mancini, R. Reischke, V. Pettorino, B. M. Schäfer, and M. Zumalacárregui. Testing (modified) gravity with 3D and tomographic cosmic shear. MNRAS, 480:3725–3738, November 2018. doi: 10.1093/mnras/sty2092.
  • Taylor et al. [2018a] Peter L. Taylor, Thomas D. Kitching, and Jason D. McEwen. Preparing for the cosmic shear data flood: Optimal data extraction and simulation requirements for stage iv dark energy experiments. Phys. Rev. D, 98:043532, Aug 2018a. doi: 10.1103/PhysRevD.98.043532. URL https://link.aps.org/doi/10.1103/PhysRevD.98.043532.
  • Taylor et al. [2018b] Peter L. Taylor, Thomas D. Kitching, Jason D. McEwen, and Thomas Tram. Testing the cosmic shear spatially-flat universe approximation with generalized lensing and shear spectra. Phys. Rev. D, 98:023522, Jul 2018b. doi: 10.1103/PhysRevD.98.023522. URL https://link.aps.org/doi/10.1103/PhysRevD.98.023522.
  • Vallis et al. [2018] Z.M. Vallis, C.G.R. Wallis, and T.D. Kitching. On the effect of projections on convergence peak counts and minkowski functionals. Astronomy and Computing, 24:84 – 96, 2018. ISSN 2213-1337. doi: https://doi.org/10.1016/j.ascom.2018.06.004. URL http://www.sciencedirect.com/science/article/pii/S221313371730149X.
  • van Uitert et al. [2018] E. van Uitert, B. Joachimi, S. Joudaki, A. Amon, C. Heymans, F. Köhlinger, M. Asgari, C. Blake, A. Choi, T. Erben, D. J. Farrow, J. Harnois-Déraps, H. Hildebrandt, H. Hoekstra, T. D. Kitching, D. Klaes, K. Kuijken, J. Merten, L. Miller, R. Nakajima, P. Schneider, E. Valentijn, and M. Viola. KiDS+GAMA: cosmology constraints from a joint analysis of cosmic shear, galaxy-galaxy lensing, and angular clustering. MNRAS, 476:4662–4689, June 2018. doi: 10.1093/mnras/sty551.
  • Van Waerbeke et al. [2000] L. Van Waerbeke, Y. Mellier, T. Erben, J. C. Cuillandre, F. Bernardeau, R. Maoli, E. Bertin, H. J. McCracken, O. Le Fèvre, B. Fort, M. Dantel-Fort, B. Jain, and P. Schneider. Detection of correlated galaxy ellipticities from CFHT data: first evidence for gravitational lensing by large-scale structures. A&A, 358:30–44, June 2000.
  • Winitzki and Kosowsky [1998] S. Winitzki and A. Kosowsky. Minkowski functional description of microwave background Gaussianity. , 3:75–99, March 1998. doi: 10.1016/S1384-1076(97)00046-8.
  • Xavier et al. [2016] H. S. Xavier, F. B. Abdalla, and B. Joachimi. Improving lognormal models for cosmological fields. MNRAS, 459:3693–3710, July 2016. doi: 10.1093/mnras/stw874.
  • Yoshiura et al. [2017] S. Yoshiura, H. Shimabukuro, K. Takahashi, and T. Matsubara. Studying topological structure of 21-cm line fluctuations with 3D Minkowski functionals before reionization. MNRAS, 465:394–402, February 2017. doi: 10.1093/mnras/stw2701.
  • Zieser and Merkel [2016] B. Zieser and P. M. Merkel. The cross-correlation between 3D cosmic shear and the integrated Sachs-Wolfe effect. MNRAS, 459:1586–1595, June 2016. doi: 10.1093/mnras/stw665.
  • Zuntz et al. [2015] Joe Zuntz, Marc Paterno, Elise Jennings, Douglas Rudd, Alessandro Manzotti, Scott Dodelson, Sarah Bridle, Saba Sehrish, and James Kowalkowski. Cosmosis: modular cosmological parameter estimation. Astronomy and Computing, 12:45–59, 2015.

Appendix A Shot noise in 3D cosmic shear

Here we derive explicitly the expression for the shot noise contribution to the 3D cosmic shear covariance matrix, as given by Eq. 12. The noise contribution is present in all the literature on 3D cosmic shear [see e.g. the seminal papers Heavens, 2003, Heavens et al., 2006], however in the context of our code comparison for 3D cosmic shear it may arise difficulties due to the different conventions used for the spherical-Bessel formalism. We refer the reader also to Appendix C in Lanusse et al. [2015], where a derivation of the shot noise term for 3D galaxy clustering is presented.

Shot noise arises by discretising the survey in cells that either contain one or zero galaxies [Peebles, 1980]. We will keep the discussion more general here, for a random field f(x)𝑓𝑥f(\vec{x}) that is discretised on our series of cells labelled by index i𝑖i. We will later specialise to our intrinsic ellipticity field. nisubscript𝑛𝑖n_{i} represents the occupation number of the cell and fisubscript𝑓𝑖f_{i} the value of the field in cell i𝑖i:

f(x)=iδ(xxi)nifi.𝑓𝑥subscript𝑖𝛿𝑥subscript𝑥𝑖subscript𝑛𝑖subscript𝑓𝑖\displaystyle f(\vec{x})=\sum_{i}\delta(\vec{x}-\vec{x}_{i})\,n_{i}\,f_{i}. (61)

We calculate the correlation (where V𝑉V is the “volume” factor for our field):

f(x)f(x)delimited-⟨⟩𝑓𝑥𝑓𝑥\displaystyle\left\langle f(\vec{x})f(\vec{x})\right\rangle =i,jδD(xxi)δD(xxi)ninjfifj1V2absentsubscript𝑖𝑗delimited-⟨⟩superscript𝛿𝐷𝑥subscript𝑥𝑖superscript𝛿𝐷𝑥subscript𝑥𝑖subscript𝑛𝑖subscript𝑛𝑗subscript𝑓𝑖subscript𝑓𝑗1superscript𝑉2\displaystyle=\sum_{i,j}\left\langle\delta^{D}(\vec{x}-\vec{x}_{i})\delta^{D}(\vec{x}-\vec{x}_{i})\,n_{i}\,n_{j}\,f_{i}\,f_{j}\right\rangle\frac{1}{V^{2}} (62)
=i,jδD(xixj)ni2fi21V2absentsubscript𝑖𝑗superscript𝛿𝐷subscript𝑥𝑖subscript𝑥𝑗delimited-⟨⟩superscriptsubscript𝑛𝑖2delimited-⟨⟩superscriptsubscript𝑓𝑖21superscript𝑉2\displaystyle=\sum_{i,j}\delta^{D}(\vec{x}_{i}-\vec{x}_{j})\left\langle n_{i}^{2}\right\rangle\left\langle f_{i}^{2}\right\rangle\frac{1}{V^{2}} (63)
=iδD(0)nifi21V2absentsubscript𝑖superscript𝛿𝐷0delimited-⟨⟩subscript𝑛𝑖delimited-⟨⟩superscriptsubscript𝑓𝑖21superscript𝑉2\displaystyle=\sum_{i}\delta^{D}(0)\left\langle n_{i}\right\rangle\left\langle f_{i}^{2}\right\rangle\frac{1}{V^{2}} (64)
=inifi2Vabsentsubscript𝑖delimited-⟨⟩subscript𝑛𝑖delimited-⟨⟩superscriptsubscript𝑓𝑖2𝑉\displaystyle=\sum_{i}\langle n_{i}\rangle\frac{\langle f_{i}^{2}\rangle}{V} (65)

where we used the fact that nisubscript𝑛𝑖n_{i} and fisubscript𝑓𝑖f_{i} are uncorrelated, ni2=nidelimited-⟨⟩superscriptsubscript𝑛𝑖2delimited-⟨⟩subscript𝑛𝑖\langle n_{i}^{2}\rangle=\langle n_{i}\rangle due to Poisson sampling and we assumed that only equal cells are correlated. In the last step we used that δD(0)=V=4πsuperscript𝛿𝐷0𝑉4𝜋\delta^{D}(0)=V=4\pi. In our case, the random field we consider is the intrinsic ellipticity of the galaxies ϵSsubscriptitalic-ϵ𝑆\epsilon_{S}. This is because, as already mentioned in Sec. II, we assume the observed ellipticity ϵitalic-ϵ\epsilon to be the sum of the shear γ𝛾\gamma and the intrinsic ellipticity ϵSsubscriptitalic-ϵ𝑆\epsilon_{S}, and neglect correlations between γ𝛾\gamma and ϵSsubscriptitalic-ϵ𝑆\epsilon_{S} as given by intrinsic alignments. We denote the intrinsic ellipticity dispersion as σϵsubscript𝜎italic-ϵ\sigma_{\epsilon} (with a typical value σϵ0.3similar-to-or-equalssubscript𝜎italic-ϵ0.3\sigma_{\epsilon}\simeq 0.3). Expressing the field f𝑓f in a spherical basis as

fm(k)=2πinifij(kχi)Ym(𝐧^𝐢),subscript𝑓𝑚𝑘2𝜋subscript𝑖subscript𝑛𝑖subscript𝑓𝑖subscript𝑗𝑘subscript𝜒𝑖subscript𝑌𝑚subscript^𝐧𝐢\displaystyle f_{\ell m}(k)=\sqrt{\frac{2}{\pi}}\sum_{i}\,n_{i}\,f_{i}\,j_{\ell}(k\chi_{i})\,Y_{\ell m}\,(\mathbf{\hat{n}_{i}}), (66)

and taking into account the redshift distribution of galaxies, one arrives at Eq. 12.