License: CC BY 4.0
arXiv:2402.13912v1 [astro-ph.CO] 21 Feb 2024

Comparing Mass Mapping Reconstruction Methods with Minkowski Functionals

Nisha Grewal1{}^{1}\aststart_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT ∗    Joe Zuntz11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT    Tilman Tröster22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT 11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPTInstitute for Astronomy, University of Edinburgh, Royal Observatory, Blackford Hill, Edinburgh, EH9 3HJ, UK
22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPTEidgenoessische Technische Hochschule Zürich, Rämistrasse 101, 8092 Zürich, Switzerland
\ast nisha.grewal@ed.ac.uk
(February 21, 2024)
Abstract

Using higher-order statistics to capture cosmological information from weak lensing surveys often requires a transformation of observed shear to a measurement of the convergence signal. This inverse problem is complicated by noise and boundary effects, and various reconstruction methods have been developed to implement the process. Here we evaluate the retention of signal information of four such methods: Kaiser-Squires, Wiener filter, DarkMappy, and DeepMass. We use the higher order statistics Minkowski functionals to determine which method best reconstructs the original convergence with efficiency and precision. We find DeepMass produces the tightest constraints on cosmological parameters, while Kaiser-Squires, Wiener filter, and DarkMappy are similar at a smoothing scale of 3.5 arcmin. We also study the MF inaccuracy caused by inappropriate training sets in the DeepMass method and find it to be large compared to the errors, underlining the importance of selecting appropriate training cosmologies.

1 Introduction

Weak gravitational lensing uses the distortion of the shapes and/or size of galaxies as a probe of the matter distribution along the line of sight. Maps quantifying lensing can be generated by measuring the observed shapes of galaxies, utilizing this shear effect from gravitational lensing (Hikage et al., 2019; Hamana et al., 2020; Asgari et al., 2021; Amon et al., 2022; Secco et al., 2022). Statistics measured from weak gravitational lensing are effective probes of the evolution of large-scale structure. Galaxy surveys, such as the Dark Energy Survey111https://www.darkenergysurvey.org/ (DES; Flaugher, 2005), Hyper Supreme-Cam222https://www.naoj.org/Projects/HSC/ (HSC; Aihara et al., 2017) survey, and Kilo-Degree Survey333http://kids.strw.leidenuniv.nl/ (KiDS; Kuijken et al., 2019), constructed detailed galaxy catalogues of tens to hundreds of millions of observed galaxies (Gatti et al., 2021a; Giblin et al., 2021; Aihara et al., 2022). The Legacy Survey Space and Time at the Rubin Observatory444https://www.lsst.org/ (LSST; LSST Science Collaboration et al., 2009; LSST Dark Energy Science Collaboration, 2012) and the space telescopes Euclid555https://www.euclid-ec.org/ (Scaramella et al., 2021) and Nancy Grace Roman666https://roman.gsfc.nasa.gov/ (Spergel et al., 2015) will go further in depth and area, and so improve our understanding of dark matter and dark energy.

Weak lensing causes source galaxies shapes to be magnified by the convergence field and distorted by the shear field. While only the shear signal can be directly inferred from galaxy shape observations, convergence can conveniently be represented as an integral of the projected matter density along the line of sight and can be derived from shear. Without noise or boundary effects, the observed shear field can be used to measure the convergence field trivially, but this is not the case for real data. We need to solve an inverse problem, working backwards from observational shear measurements to infer the underlying convergence in the presence of these effects in real data, which becomes complex.

Various non-parametric mass mapping reconstruction methods have been developed to address this problem and take into account noise and masks in the data to solve the inverse problem. Linear approaches like Kaiser Squires (KS) (Kaiser & Squires, 1993), Wiener filter (WF) (Lahav et al., 1994), Simon et al. (2009), and VanderPlas et al. (2011) and non-linear approaches like GLIMPSE (Leonard et al., 2014), Glimpse2D (Lanusse et al., 2016), KS+ (Pires et al., 2020), DeepMass (Jeffrey et al., 2020), MCAlens (Starck et al., 2021), DarkMappy (Price et al., 2021), KaRMMa (Fiedorowicz et al., 2022), DeepPosterior (Remy et al., 2022), SKS+ (Kansal, 2023), and others (Marshall et al., 2002; Simon et al., 2011b; Leonard et al., 2012; Alsing et al., 2015; Schneider et al., 2017; Horowitz et al., 2019; Porqueres et al., 2021) have all been shown to be effective. The original method, Kaiser-Squires (KS) is a linear inversion of the shear map into a convergence map, but it cannot take noise or boundary effects into account (Kaiser & Squires, 1993). Wiener filtering is also a linear filter, but unlike KS, it can transform noisy data well (Lahav et al., 1994). Sparsity methods like GLIPMSE and DarkMappy are non-linear and able to reconstruct information in the non-Gaussian regime (Leonard et al., 2014; Price et al., 2021). Some methods like DarkMappy and DeepPosterior are able to approximate the uncertainty that comes from reconstruction. Other methods including DeepPosterior have employed machine learning as a useful tool in denoising convergence signal (Jeffrey et al., 2020; Shirasaki et al., 2021; Remy et al., 2022). In this paper we explore four reconstruction methods: Kaiser Squires, Wiener filter, DarkMappy, and DeepMass.

Measuring statistics from the best reconstructed convergence maps is expected to lead to tighter constraints on cosmological parameters. Two-point statistics, which describe the correlation between pairs of data points, can constrain these parameters with high precision (Heymans et al., 2021; Miyatake et al., 2021; Abbott et al., 2022). However, statistics beyond two-point are required to fully probe non-Gaussian structure on small scales. Higher order statistics including N-point such as the one-point probability distribution function (Liu & Madhavacheril, 2019; Barthelemy et al., 2020; Thiele et al., 2020; Boyle et al., 2021) and three-point correlation functions (Schneider & Lombardi, 2003; Takada & Jain, 2004; Gong et al., 2023), moments (Van Waerbeke et al., 2013; Petri et al., 2015; Vicinanza et al., 2016; Chang et al., 2018; Peel et al., 2018; Vicinanza et al., 2018; Gatti et al., 2020, 2021b; Porth & Smith, 2021), topological descriptors such as Betti numbers (Feldbrugge et al., 2019; Heydenreich et al., 2021; Parroni et al., 2021) and peak counts (Marian et al., 2009; Kratochvil et al., 2010; Liu et al., 2015; Kacprzak et al., 2016; Shan et al., 2017; Martinet et al., 2018; Peel et al., 2018; Ajani et al., 2020; Harnois-Déraps et al., 2021; Zürcher et al., 2022), scattering transform coefficients (Cheng et al., 2020; Cheng & Ménard, 2021; Gatti et al., 2023), and field level inference (Porqueres et al., 2021; Boruah et al., 2022) can measure information in the non-Gaussian regime in weak lensing maps.

Another higher-order statistic, Minkowski functionals (MFs) are morphological descriptors that describe the topology of a continuous field (Minkowski, 1903). MFs are unbiased, have low variance, and can have additional resilience to some systematic uncertainties, making them effective probes of the underlying dark matter distribution (Zürcher et al., 2021). MFs have been applied to Cosmic Microwave Background data (Schmalzing & Gorski, 1998; Eriksen et al., 2004; Hikage et al., 2006; Buchert et al., 2017; Collaboration et al., 2020; Hamann & Kang, 2023; Chingangbam & Rahman, 2023), 2D convergence maps (Mecke et al., 1993; Kratochvil et al., 2012; Petri et al., 2013; Vicinanza et al., 2019; Parroni et al., 2020; Grewal et al., 2022; Euclid Collaboration et al., 2023), 2D density fields (Grewal et al., 2022), and 3D density fields from spectroscopic data (Hikage et al., 2003; Wiegand & Eisenstein, 2017; Sullivan et al., 2019; Appleby et al., 2022; Liu et al., 2022). Like other higher-order statistics, MFs produce tighter constraints on cosmological parameters because they are sensitive to small-scale structure.

In this paper we use MFs as a test of different convergence reconstruction methods, using them to quantify the information (particularly on small scales) that the methods can reconstruct in maps. We compare the four different mass mapping methods KS, WF, DarkMappy, and DeepMass, applying them to simulated images and using MFs to constrain w𝑤witalic_wCDM parameters and thus computing figures of merit to describe method power.

In Section 2 we review weak lensing formalism and the mass mapping reconstruction methods, in Section 3 we describe the simulations, observables, and evaluation metrics, in Section 4 we compare the cosmological constraints from the four reconstruction methods, and we conclude in Section 5.

2 Formalism

2.1 Convergence and Shear

The magnification and distortion of a lensed galaxy image is represented to first order by the Jacobian matrix

𝒜(𝜽)=(δij2ψ(𝜽)θiθj)=(1κγ1γ2γ21κ+γ1),𝒜𝜽subscript𝛿𝑖𝑗superscript2𝜓𝜽subscript𝜃𝑖subscript𝜃𝑗matrix1𝜅subscript𝛾1subscript𝛾2subscript𝛾21𝜅subscript𝛾1\mathcal{A}(\boldsymbol{\theta})=\left(\delta_{ij}-\frac{\partial^{2}\psi(% \boldsymbol{\theta})}{\partial\theta_{i}\partial\theta_{j}}\right)=\begin{% pmatrix}1-\kappa-\gamma_{1}&-\gamma_{2}\\ -\gamma_{2}&1-\kappa+\gamma_{1}\end{pmatrix},caligraphic_A ( bold_italic_θ ) = ( italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT - divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ ( bold_italic_θ ) end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∂ italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG ) = ( start_ARG start_ROW start_CELL 1 - italic_κ - italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL - italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL - italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_CELL start_CELL 1 - italic_κ + italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) , (1)

where θ𝜃\thetaitalic_θ is the observed angular position, ψ𝜓\psiitalic_ψ is the lensing potential, the shear γγ1+iγ2𝛾subscript𝛾1𝑖subscript𝛾2\gamma\equiv\gamma_{1}+i\gamma_{2}italic_γ ≡ italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_i italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and convergence is κ𝜅\kappaitalic_κ (Bartelmann & Schneider, 2001). Convergence is a measure of the matter distribution along the line of sight. It is given by a projection of the overdensity δ𝛿\deltaitalic_δ:

κ(θ^)=3H02Ωm2c20χlim𝑑χq(χ)a(χ)fK(χ)δ(fK(χ)θ^,χ),𝜅^𝜃3superscriptsubscriptH02subscriptΩm2superscript𝑐2superscriptsubscript0subscript𝜒limdifferential-d𝜒𝑞𝜒𝑎𝜒subscript𝑓𝐾𝜒𝛿subscript𝑓𝐾𝜒^𝜃𝜒\kappa(\mathbf{\hat{\theta}})=\frac{3\mathrm{H}_{0}^{2}\Omega_{\mathrm{m}}}{2c% ^{2}}\int_{0}^{\chi_{\mathrm{lim}}}d\chi\frac{q(\chi)}{a(\chi)}f_{K}(\chi)% \delta(f_{K}(\chi)\mathbf{\hat{\theta}},\chi),italic_κ ( over^ start_ARG italic_θ end_ARG ) = divide start_ARG 3 roman_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT roman_lim end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_d italic_χ divide start_ARG italic_q ( italic_χ ) end_ARG start_ARG italic_a ( italic_χ ) end_ARG italic_f start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( italic_χ ) italic_δ ( italic_f start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( italic_χ ) over^ start_ARG italic_θ end_ARG , italic_χ ) , (2)

where H0subscriptH0\mathrm{H}_{0}roman_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the Hubble constant, ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT is the matter density, c𝑐citalic_c is the speed of light, χ𝜒\chiitalic_χ is comoving distance, χlimsubscript𝜒lim\chi_{\mathrm{lim}}italic_χ start_POSTSUBSCRIPT roman_lim end_POSTSUBSCRIPT is the horizon distance, a𝑎aitalic_a is the scale factor, and the lens efficiency q𝑞qitalic_q is defined as

q(χ)=0χlimn(χ)𝑑χfK(χχ)fK(χ).𝑞𝜒superscriptsubscript0subscript𝜒lim𝑛superscript𝜒differential-dsuperscript𝜒subscript𝑓𝐾superscript𝜒𝜒subscript𝑓𝐾superscript𝜒q(\chi)=\int_{0}^{\chi_{\mathrm{lim}}}n(\chi^{\prime})d\chi^{\prime}\frac{f_{K% }(\chi^{\prime}-\chi)}{f_{K}(\chi^{\prime})}.italic_q ( italic_χ ) = ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_χ start_POSTSUBSCRIPT roman_lim end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_n ( italic_χ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) italic_d italic_χ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT divide start_ARG italic_f start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( italic_χ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_χ ) end_ARG start_ARG italic_f start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT ( italic_χ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) end_ARG . (3)

n(χ)dχ𝑛𝜒d𝜒n(\chi)\mathrm{d}\chiitalic_n ( italic_χ ) roman_d italic_χ is the source galaxy distribution and fKsubscript𝑓𝐾f_{K}italic_f start_POSTSUBSCRIPT italic_K end_POSTSUBSCRIPT is the comoving angular distance (Kilbinger, 2015).

While convergence is harder to observe directly, shear can be measured from galaxy shape observations. The convergence κ𝜅\kappaitalic_κ is related to the shear γ𝛾\gammaitalic_γ via the lensing potential ψ𝜓\psiitalic_ψ. Rearranging the following equations for κ𝜅\kappaitalic_κ and γ𝛾\gammaitalic_γ and solving for ψ𝜓\psiitalic_ψ,

κ=12Δψ,𝜅12Δ𝜓\kappa=\frac{1}{2}\Delta\psi,italic_κ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG roman_Δ italic_ψ , (4)
γ1=12(12ψ22ψ);γ2=12ψ,formulae-sequencesubscript𝛾112superscriptsubscript12𝜓superscriptsubscript22𝜓subscript𝛾2subscript1subscript2𝜓\gamma_{1}=\frac{1}{2}\left(\partial_{1}^{2}\psi-\partial_{2}^{2}\psi\right);% \gamma_{2}=\partial_{1}\partial_{2}\psi,italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( ∂ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ - ∂ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ ) ; italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = ∂ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∂ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_ψ , (5)

and then performing a Fourier transform in the flat sky regime, the convergence can be expressed as

κ~=k12k22k2γ1~+2k1k2k2γ2~.~𝜅superscriptsubscript𝑘12superscriptsubscript𝑘22superscript𝑘2~subscript𝛾12subscript𝑘1subscript𝑘2superscript𝑘2~subscript𝛾2\tilde{\kappa}=\frac{k_{1}^{2}-k_{2}^{2}}{k^{2}}\tilde{\gamma_{1}}+\frac{2k_{1% }k_{2}}{k^{2}}\tilde{\gamma_{2}}.over~ start_ARG italic_κ end_ARG = divide start_ARG italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG over~ start_ARG italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG + divide start_ARG 2 italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG over~ start_ARG italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG . (6)

k1subscript𝑘1k_{1}italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are spatial frequencies in Fourier space (k2=k12+k22superscript𝑘2superscriptsubscript𝑘12superscriptsubscript𝑘22k^{2}=k_{1}^{2}+k_{2}^{2}italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), and γ1~~subscript𝛾1\tilde{\gamma_{1}}over~ start_ARG italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG and γ2~~subscript𝛾2\tilde{\gamma_{2}}over~ start_ARG italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG are the two orthogonal components of shear distortion (Remy et al., 2022). As the expectation value of matter overdensity δ𝛿\deltaitalic_δ is zero, the first moment of shear and convergence will also be zero. Known as mass sheet degeneracy, this feature means that at k𝑘kitalic_k = 0, the relationship between convergence and shear is undefined. It is important to note that this transformation is only optimal when performed on pure signal in the absence of noise and missing data.

In the presence of noise and gaps from masking, we must go beyond this simple treatment. In this more realistic regime, various algorithms attempt the inversion from shear to convergence in different ways. We review several of these methods here.

2.2 Kaiser-Squires

Kaiser-Squires uses a direct linear inversion of the shear in Fourier space to obtain the reconstructed convergence field (Kaiser & Squires, 1993). Scalar perturbations in matter density are captured by the E mode gradient-like patterns, and the tensor perturbation counterparts are captured by the curl-like B mode. The shear-convergence relation can be expanded to give both the E and B modes:

κ~=κE~+iκB~=(k12k22k2+i2k1k2k2)(γ1~+iγ2~).~𝜅~subscript𝜅𝐸𝑖~subscript𝜅𝐵superscriptsubscript𝑘12superscriptsubscript𝑘22superscript𝑘2𝑖2subscript𝑘1subscript𝑘2superscript𝑘2~subscript𝛾1𝑖~subscript𝛾2\tilde{\kappa}=\tilde{\kappa_{E}}+i\tilde{\kappa_{B}}=\left(\frac{k_{1}^{2}-k_% {2}^{2}}{k^{2}}+i\frac{2k_{1}k_{2}}{k^{2}}\right)(\tilde{\gamma_{1}}+i\tilde{% \gamma_{2}}).over~ start_ARG italic_κ end_ARG = over~ start_ARG italic_κ start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT end_ARG + italic_i over~ start_ARG italic_κ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG = ( divide start_ARG italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_i divide start_ARG 2 italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) ( over~ start_ARG italic_γ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG + italic_i over~ start_ARG italic_γ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) . (7)

We can define the forward model in Fourier space 𝐏𝐏\mathbf{P}bold_P as

𝐏=(k12k22k2+i2k1k2k2)𝐏superscriptsubscript𝑘12superscriptsubscript𝑘22superscript𝑘2𝑖2subscript𝑘1subscript𝑘2superscript𝑘2\textbf{{P}}=\left(\frac{k_{1}^{2}-k_{2}^{2}}{k^{2}}+i\frac{2k_{1}k_{2}}{k^{2}% }\right)P = ( divide start_ARG italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_i divide start_ARG 2 italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) (8)

(Kaiser & Squires, 1993). We note that we do not expect weak lensing to generate B modes, which come from systematics and higher-order lensing effects.

KS disregards noise and boundary effects, which means it requires continuous fields to reconstruct maps with masks (Van Waerbeke et al., 2013). In common with other approaches, we use a Gaussian kernel to fill in the gaps and smooth the field, though this leads to loss of structure on small scales and suppresses peaks in the convergence (Jeffrey et al., 2018). This also leads to leakage of the E and B modes of the reconstructed convergence field (Remy et al., 2022).

2.3 Wiener filter

A Wiener filter (WF) is a linear inversion of signal and noise together, aiming to suppress the latter. The WF method uses the expected power spectrum of the convergence field at a fiducial cosmology and of the noise (Simon et al., 2011a). In Fourier space:

κ~=𝐒𝐏[𝐏𝐒𝐏+𝐍]1γ~,~𝜅superscript𝐒𝐏superscriptdelimited-[]superscript𝐏𝐒𝐏𝐍1~𝛾\tilde{\kappa}=\textbf{SP}^{\dagger}\left[\textbf{{PSP}}^{\dagger}+\textbf{{N}% }\right]^{-1}\tilde{\gamma},over~ start_ARG italic_κ end_ARG = SP start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT [ PSP start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT + N ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT over~ start_ARG italic_γ end_ARG , (9)

where S is the covariance matrix of power spectrum signal and N is the noise covariance matrix (Lahav et al., 1994; Zaroubi et al., 1995). However, in filtering out the noise, WF smooths out small-scales features in the map more generally. While WF can reconstruct the field with precision on large scales, it does not typically retain small-scale information in convergence maps as well (Simon et al., 2011a; Horowitz et al., 2019).

2.4 DarkMappy

DarkMappy was primarily designed for cluster analysis; here we apply the method to a wide-field convergence. DarkMappy models the convergence field using wavelets (Price et al., 2021). It uses a sparsity approach to minimise the number of wavelets used and varies their coefficients in a hierarchical Bayesian approach, which enables an approximation of the signal uncertainty that comes from reconstruction. As it is a non-linear approach, we expect it to measure small-scale effects smoothed out by KS and WF.

A wavelet dictionary is a collection of wavelets, which are oscillatory functions with zero mean that are localised in Fourier and real space (Mallat, 2012). A wavelet dictionary 𝚿𝚿\mathbf{\Psi}bold_Ψ can sparsely represent the signal κ𝜅\kappaitalic_κ:

κ=𝚿α,𝜅𝚿𝛼\kappa=\mathbf{\Psi}\alpha,italic_κ = bold_Ψ italic_α , (10)

where only a subset of non-zero coefficients α𝛼\alphaitalic_α is needed (Leonard et al., 2014).

Unlike the previous methods, DarkMappy fits the data in real space. The forward model operation in Eq. (8) is transformed into a measurement operator 𝚽𝚽\mathbf{\Phi}bold_Φ using a Fourier transform 𝐅𝐅\mathbf{F}bold_F:

𝚽=𝐅1𝐏𝐅.𝚽superscript𝐅1𝐏𝐅\mathbf{\Phi}=\mathbf{F}^{-1}\mathbf{PF}.bold_Φ = bold_F start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT bold_PF . (11)

𝚽𝚽\mathbf{\Phi}bold_Φ maps convergence onto shear in the measurement equation, where the total shear is then given by:

γ=𝚽κ+n,𝛾𝚽𝜅𝑛\gamma=\mathbf{\Phi}\kappa+n,italic_γ = bold_Φ italic_κ + italic_n , (12)

where n𝑛nitalic_n is the noise contaminating the signal. Using a wavelet-based, sparsity-promoting, l1subscript𝑙1l_{1}italic_l start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT-norm prior, the DarkMappy model for the reconstructed map is then given by the maximum a posteriori solution:

κmap=argmin𝜅{μ𝚿κ1+12Σ12(𝚽κγ)22}.superscript𝜅map𝜅argminconditional-set𝜇evaluated-atsuperscript𝚿𝜅112superscriptsubscriptnormsuperscriptΣ12𝚽𝜅𝛾22\kappa^{\mathrm{map}}=\underset{\kappa}{\mathrm{argmin}}\biggl{\{}\mu\parallel% \mathbf{\Psi}^{\dagger}\kappa\parallel_{1}+\frac{1}{2}\parallel\Sigma^{-\frac{% 1}{2}}(\mathbf{\Phi}\kappa-\gamma)\parallel_{2}^{2}\biggr{\}}.italic_κ start_POSTSUPERSCRIPT roman_map end_POSTSUPERSCRIPT = underitalic_κ start_ARG roman_argmin end_ARG { italic_μ ∥ bold_Ψ start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT italic_κ ∥ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∥ roman_Σ start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( bold_Φ italic_κ - italic_γ ) ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT } . (13)

The first term is the prior, which imposes sparsity, and the second term is the likelihood of the data. μ𝜇\muitalic_μ regularises the weighting between the two terms and ΣΣ\Sigmaroman_Σ is the covariance of the measured shear (Price et al., 2021).

The search process for the MAP uses knock-out hypothesis testing of the posterior to differentiate between original signal and reconstruction effects (Cai et al., 2018). This algorithm tests whether features (e.g. peaks) are significant enough to be included in the MAP model using a statistical test that compares κ𝜅\kappaitalic_κ models that include them to those using κ𝜅\kappaitalic_κ smoothed by surrounding pixels.

2.5 DeepMass

DeepMass is a convolutional neural network (CNN) that aims to denoise noisy maps. The network learns parameters ΘΘ\Thetaroman_Θ in convolution layers to reconstruct convergence maps from noisy shear maps:

κ=Θ(γ),𝜅subscriptΘ𝛾\kappa=\mathcal{F}_{\Theta}(\gamma),italic_κ = caligraphic_F start_POSTSUBSCRIPT roman_Θ end_POSTSUBSCRIPT ( italic_γ ) , (14)

where Θ(γ)subscriptΘ𝛾\mathcal{F}_{\Theta}(\gamma)caligraphic_F start_POSTSUBSCRIPT roman_Θ end_POSTSUBSCRIPT ( italic_γ ) is the posterior estimate of convergence (Jeffrey et al., 2020). Convergence and shear training data is used to minimise a mean-square-error (MSE) cost function J𝐽Jitalic_J as the network finds ΘΘ\Thetaroman_Θ:

J(Θ)=Θ(γ)κ22.𝐽ΘsuperscriptsubscriptnormsubscriptΘ𝛾𝜅22J(\Theta)=\parallel\mathcal{F}_{\Theta}(\gamma)-\kappa\parallel_{2}^{2}.italic_J ( roman_Θ ) = ∥ caligraphic_F start_POSTSUBSCRIPT roman_Θ end_POSTSUBSCRIPT ( italic_γ ) - italic_κ ∥ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (15)

Conceptually, the ‘true’ output noise-free convergence maps 𝜿𝜿\boldsymbol{\kappa}bold_italic_κ are drawn from a prior distribution P(κ)P𝜅\mathrm{P}(\kappa)roman_P ( italic_κ ), and the corresponding input noisy shear maps are drawn from the likelihood P(γ|κ)Pconditional𝛾𝜅\mathrm{P}(\gamma|\kappa)roman_P ( italic_γ | italic_κ ) (Jeffrey et al., 2020). The reconstructed convergence κ𝜅\kappaitalic_κ can then be approximated as

κ=Θ(𝜸)=κP(κ|γ)dκ.𝜅subscriptΘ𝜸𝜅Pconditional𝜅𝛾d𝜅\kappa=\mathcal{F}_{\Theta}(\boldsymbol{\gamma})=\int\kappa\mathrm{P}(\kappa|% \gamma)\textrm{d}\kappa.italic_κ = caligraphic_F start_POSTSUBSCRIPT roman_Θ end_POSTSUBSCRIPT ( bold_italic_γ ) = ∫ italic_κ roman_P ( italic_κ | italic_γ ) d italic_κ . (16)

In practice, the network is trained on pairs of truth convergence maps and noisy WF convergence maps, which are reconstructed from shear maps (Jeffrey et al., 2020). The model is two-dimensional, so that multiple models need to be built for data with more than one redshift bin. DeepMass utilises a U-Net architecture with a contracting path (encoder), reducing the dimensions of the input data while extracting high-level features, and an expanding path (decoder), which complements the contracting path by restoring the full spatial dimension and generating a high-resolution output (Ronneberger et al., 2015). It also uses average pooling, where each region of the input data is replaced with its mean value. As the resolution decreases, the model captures more comprehensive physical features in the convergence map by considering a wider area and reducing the dimensionality of the data (Géron, 2019). While DeepMass can measure small-scale structure in convergence maps, the model does not retrieve uncertainties in the reconstruction (Jeffrey et al., 2020; Remy et al., 2022). As with all CNNs, DeepMass risks overfitting to map structure and/or creating fake artefacts in the reconstructed signal.

2.6 Minkowski functionals

Minkowski functionals are field integrals that characterise the topological properties of continuous fields (Minkowski, 1903; Zürcher et al., 2021). In this work we measure these mathematical descriptors from a convergence field. MFs measure the properties of excursion sets of this field, which are given by the region of the field above a given threshold (Parroni et al., 2020). The first three functionals quantify area, perimeter, and mean curvature of an excursion set and are defined as

V0(t)=1AAΞ(κ(𝐱)t)d2𝐱,subscript𝑉0𝑡1𝐴subscript𝐴Ξ𝜅𝐱𝑡superscriptd2𝐱V_{0}(t)=\frac{1}{A}\int\limits_{A}\Xi(\kappa(\mathbf{x})-t)\mathrm{d}^{2}% \mathbf{x},italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG 1 end_ARG start_ARG italic_A end_ARG ∫ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT roman_Ξ ( italic_κ ( bold_x ) - italic_t ) roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_x , (17)
V1(t)=14AAδ(κ(𝐱)t)κx2+κy2d2𝐱,subscript𝑉1𝑡14𝐴subscript𝐴𝛿𝜅𝐱𝑡superscriptsubscript𝜅𝑥2superscriptsubscript𝜅𝑦2superscriptd2𝐱V_{1}(t)=\frac{1}{4A}\int\limits_{A}\delta(\kappa(\mathbf{x})-t)\sqrt{\kappa_{% x}^{2}+\kappa_{y}^{2}}\mathrm{d}^{2}\mathbf{x},italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG 1 end_ARG start_ARG 4 italic_A end_ARG ∫ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT italic_δ ( italic_κ ( bold_x ) - italic_t ) square-root start_ARG italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_x , (18)
V2(t)=12πAAδ(κ(𝐱)t)(2κxκyκxyκx2κyyκy2κxxκx2+κy2)d2𝐱,subscript𝑉2𝑡12𝜋𝐴subscript𝐴𝛿𝜅𝐱𝑡2subscript𝜅𝑥subscript𝜅𝑦subscript𝜅𝑥𝑦superscriptsubscript𝜅𝑥2subscript𝜅𝑦𝑦superscriptsubscript𝜅𝑦2subscript𝜅𝑥𝑥superscriptsubscript𝜅𝑥2superscriptsubscript𝜅𝑦2superscriptd2𝐱V_{2}(t)=\frac{1}{2\pi A}\int\limits_{A}\delta(\kappa(\mathbf{x})-t)\\ \left(\frac{2\kappa_{x}\kappa_{y}\kappa_{xy}-\kappa_{x}^{2}\kappa_{yy}-\kappa_% {y}^{2}\kappa_{xx}}{\kappa_{x}^{2}+\kappa_{y}^{2}}\right)\mathrm{d}^{2}\mathbf% {x},start_ROW start_CELL italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_A end_ARG ∫ start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT italic_δ ( italic_κ ( bold_x ) - italic_t ) end_CELL end_ROW start_ROW start_CELL ( divide start_ARG 2 italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT - italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT - italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT end_ARG start_ARG italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT bold_x , end_CELL end_ROW (19)

where 𝐱=(x,y)𝐱𝑥𝑦\mathbf{x}=(x,y)bold_x = ( italic_x , italic_y ) is the location in the field, t𝑡titalic_t is a chosen threshold, A𝐴Aitalic_A is the total area of the map, x𝑥xitalic_x and y𝑦yitalic_y are flat sky coordinates, κ(𝐱)𝜅𝐱\kappa(\textbf{x})italic_κ ( x ) is the field value in two dimensions, and κxsubscript𝜅𝑥\kappa_{x}italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT, κysubscript𝜅𝑦\kappa_{y}italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT, κxysubscript𝜅𝑥𝑦\kappa_{xy}italic_κ start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT, κyysubscript𝜅𝑦𝑦\kappa_{yy}italic_κ start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT, and κxxsubscript𝜅𝑥𝑥\kappa_{xx}italic_κ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT are derivatives of the field (Minkowski, 1903; Petri et al., 2013). The derivatives are also evaluated at position 𝐱𝐱\mathbf{x}bold_x; this is dropped in Eq. (18) and (19) for brevity.

The Heaviside function ΞΞ\Xiroman_Ξ in Eq. (17) identifies the area of the field region above the threshold; similarly, the Dirac delta functions in Eq. (18) and (19) select the regions where the height matches the threshold to measure the perimeter and connectivity of the field as the curvature of the boundary, respectively (Minkowski, 1903).

3 Methodology

3.1 Simulations

To generate our convergence maps we use CosmoGridV1, which comprises a suite of full-sky lightcone simulations with nside = 2048 and shells that allow for probing of multiple redshifts, up to z <<< 3.5 (Kacprzak et al., 2022). The simulations were run with PKDGRAV3, a high performance self-gravitating astrophysical N-body treecode (Potter et al., 2016). CosmoGridV1 varies the cosmological parameters ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT, σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, nssubscript𝑛sn_{\mathrm{s}}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, ΩbsubscriptΩb\Omega_{\mathrm{b}}roman_Ω start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT, and H0subscriptH0\mathrm{H}_{0}roman_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and baryonic parameters Mc0subscriptsuperscript𝑀0cM^{0}_{\mathrm{c}}italic_M start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT and ν𝜈\nuitalic_ν, but uses fixed neutrino masses, each with mvsubscript𝑚𝑣m_{v}italic_m start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT = 0.02eV. To make realistic lensing maps, baryonic effects are included using a shell-based baryon correction model at the map level. Particle count maps are adjusted by a 2D displacement function measured from halo catalogues (Kacprzak et al., 2022). Following Schneider et al. (2019), the mass dependence of the gas profile Mcsubscript𝑀cM_{\mathrm{c}}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is broken down into the amplitude Mc0superscriptsubscript𝑀c0M_{\mathrm{c}}^{0}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and redshift dependence parameter ν𝜈\nuitalic_ν:

Mc=Mc0(1+z)ν.subscript𝑀csuperscriptsubscript𝑀c0superscript1𝑧𝜈M_{\mathrm{c}}=M_{\mathrm{c}}^{0}(1+z)^{\nu}.italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ( 1 + italic_z ) start_POSTSUPERSCRIPT italic_ν end_POSTSUPERSCRIPT . (20)

CosmoGridV1 provides 200 simulations at a fiducial cosmology with different seeds for the initial conditions, accompanied by maps with paired matching seeds with two steps, one above and one below for each of the cosmological parameters, shown in Table 1. For each cosmology, we cut out 50,000 15 degree ×\times× 15 degree patches with random positions and rotations from the 200 simulations, removing values at the poles to avoid distortion. The resulting convergence maps have resolution 256 ×\times× 256 pixels with pixel size 3.5 arcmin.

The CosmoGridV1 convergence maps have only positive values; here we subtract the mean after reconstruction before measuring observables. The mass-sheet degeneracy means the zero-point is arbitrary when going from shear back to convergence.

Fiducial ΔΔ\Deltaroman_Δ Fiducial
ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT 0.26 ±plus-or-minus\pm± 0.01
σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT 0.84 ±plus-or-minus\pm± 0.015
w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT -1 ±plus-or-minus\pm± 0.05
nssubscript𝑛sn_{\mathrm{s}}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT 0.9649 ±plus-or-minus\pm± 0.02
ΩbsubscriptΩb\Omega_{\mathrm{b}}roman_Ω start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT 0.0493 ±plus-or-minus\pm± 0.001
H0subscriptH0\mathrm{H}_{0}roman_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 67.3 ±plus-or-minus\pm± 2.0
Mc0superscriptsubscript𝑀c0M_{\mathrm{c}}^{0}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT 13.82 ±plus-or-minus\pm± 0.1
ν𝜈\nuitalic_ν 0.0 ±plus-or-minus\pm± 0.1
Table 1: CosmoGridV1 Parameter Shift Values. H0subscriptH0\mathrm{H}_{0}roman_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT has units kms1Mpc1superscriptkms1superscriptMpc1\mathrm{kms^{-1}Mpc^{-1}}roman_kms start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Mpc start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

In this work, we do our analysis with 5,000 map patches, a balance between efficiency and precision found by requiring convergence of the KS contours described below, where there is less than a 5% shift in contours. For DeepMass training and testing, we use 50,000 random patches from 200 maps to avoid overfitting at the given cosmology. However, all observables are only measured from 5,000 maps.

3.1.1 Redshift

In our analysis, we use approximately DES Y3-like tomographic redshifts from the CosmoGridV1 shells, modelled as a Smail-type distribution:

n(z)=zαexp[(zz0)β],𝑛𝑧superscript𝑧𝛼expdelimited-[]superscript𝑧subscript𝑧0𝛽n(z)=z^{\alpha}\mathrm{exp}\left[-\left(\frac{z}{z_{0}}\right)^{\beta}\right],italic_n ( italic_z ) = italic_z start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT roman_exp [ - ( divide start_ARG italic_z end_ARG start_ARG italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT ] , (21)

(Smail et al., 1995; Kacprzak et al., 2022). The survey parameter values can be seen in Table 2 and the final distributions can be seen in Figure 1.

Bin α𝛼\alphaitalic_α β𝛽\betaitalic_β z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT
1 1.99 1.44 0.20
2 3.46 2.34 0.39
3 6.03 3.60 0.66
4 3.53 4.49 1.03
Table 2: Smail-type distribution parameters for DES Y3-like redshift bins (Fischbacher et al., 2022)
Refer to caption
Figure 1: The DES Y3-like nzsubscript𝑛𝑧n_{z}italic_n start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT redshift distributions from CosmoGridV1 used in this work (Kacprzak et al., 2022).

3.1.2 Noise and Mask

At our chosen resolution our maps are in the low noise regime. We only take shape noise into account, disregarding any other observational effects. To obtain the noise standard deviation per pixel, we use

σ=σengalnbinp2,𝜎subscript𝜎𝑒subscript𝑛galsubscript𝑛binsuperscript𝑝2\sigma=\frac{\sigma_{e}}{\sqrt{\frac{n_{\mathrm{gal}}}{n_{\mathrm{bin}}}p^{2}}},italic_σ = divide start_ARG italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT end_ARG start_ARG square-root start_ARG divide start_ARG italic_n start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT end_ARG start_ARG italic_n start_POSTSUBSCRIPT roman_bin end_POSTSUBSCRIPT end_ARG italic_p start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG , (22)

where the shape dispersion σesubscript𝜎𝑒\sigma_{e}italic_σ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT is 0.26 per galaxy, ngalsubscript𝑛galn_{\mathrm{gal}}italic_n start_POSTSUBSCRIPT roman_gal end_POSTSUBSCRIPT is a DES Y3-like number density (10 galaxies/arcmin22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT), nbinsubscript𝑛binn_{\mathrm{bin}}italic_n start_POSTSUBSCRIPT roman_bin end_POSTSUBSCRIPT is the four DES Y3-like redshift bins, and p𝑝pitalic_p is the pixel size in arcmin. With 25622{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT total pixels and 15 ×\times× 15 square degree maps, the pixel size p𝑝pitalic_p is 3.5 arcmin and the noise level σ𝜎\sigmaitalic_σ is 0.047 per pixel. The map noise is computed as an uncorrelated random Gaussian field with mean 0 and σ𝜎\sigmaitalic_σ from Eq. (22).

Refer to caption
Figure 2: Mask applied to a 15 ×\times× 15 degree flat sky convergence map of redshift bin 0 at fiducial cosmology. The colourbar shows convergence of each pixel. We note that in practice the map is applied to shear map, only visualised here on convergence data.

We apply a mask to our maps by hand to approximate the structure of real masks, covering approximately 2.8% of the map, as seen in Figure 2.

The noise and mask are added at the shear step. A Gaussian kernel with a standard deviation of one pixel has to be applied to the masked shear map before reconstruction. This smoothing fills in gaps and decreases the resolution of the maps. This is a crucial step for Kaiser Squires, which only works for continuous fields.

3.1.3 Applications of Reconstruction Methods

For the Wiener filter reconstruction calculation of the signal matrix 𝐒𝐒\mathbf{S}bold_S in Eq. (9), we compute the 2D flat-sky power spectrum P(k)𝑃𝑘P(k)italic_P ( italic_k ) with CCL (Chisari et al., 2019) using the cosmology of the fiducial convergence map and incorporating baryonic effects using HMCode (Mead et al., 2015; McCarthy et al., 2022).

Splitting k𝑘kitalic_k into its components k1subscript𝑘1k_{1}italic_k start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and k2subscript𝑘2k_{2}italic_k start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in Eq. (6), P(k)𝑃𝑘P(k)italic_P ( italic_k ) can be found in 2D space. We relate the flat-sky 2D wavenumber k𝑘kitalic_k to the angular multipole \ellroman_ℓ as

=2πbsk,2𝜋𝑏𝑠𝑘\ell=\frac{2\pi}{bs}k,roman_ℓ = divide start_ARG 2 italic_π end_ARG start_ARG italic_b italic_s end_ARG italic_k , (23)

where b𝑏bitalic_b is the box width in pixels and s𝑠sitalic_s is the pixel size in radians. Using CCL, we then compute the angular power spectrum Csubscript𝐶C_{\ell}italic_C start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT at the corresponding cosmology and perform a scalar transformation using

P(k)=bsC.𝑃𝑘𝑏𝑠subscript𝐶P(k)=\frac{b}{s}C_{\ell}.italic_P ( italic_k ) = divide start_ARG italic_b end_ARG start_ARG italic_s end_ARG italic_C start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT . (24)

The noise matrix 𝐍𝐍\mathbf{N}bold_N in Eq. (9) is a uniform matrix with value σ𝜎\sigmaitalic_σ from Eq. (22). It describes the statistics of the noise in the convergence map.

We train and test the DeepMass model on realisations from a single (fiducial) cosmology and apply it to the nearby non-fiducial cosmologies. To test potential systematic uncertainties from this approach, we will also compare it to models trained directly on the non-fiducial data.

DeepMass does not include a padding functionality in the model, so the output map does not have the same dimensions as the input map. As such, we pad input data to reduce edge effects on the signal. To account for any correlation in signal between different redshifts, we have extended DeepMass to a 3D model across all four redshift bins. We note sporadic issues where DeepMass yields NaN values. We then rerun the model at the same training cosmology.

3.2 Observables

The functional integrals from Eq. (17), (18), and (19) are converted to sums over the number of pixels N𝑁Nitalic_N, so we can calculate MFs from the 2D reconstructed CosmoGridV1 convergence fields:

V0(tj)=1NiΘ(κ(𝐱i)t),subscript𝑉0subscript𝑡𝑗1𝑁subscript𝑖Θ𝜅subscript𝐱𝑖𝑡V_{0}(t_{j})=\frac{1}{N}\sum_{i}\Theta(\kappa(\textbf{x}_{i})-t),italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Θ ( italic_κ ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - italic_t ) , (25)
V1(tj)=14NiΔ(κ(𝐱i)t)κx2+κy2,subscript𝑉1subscript𝑡𝑗14𝑁subscript𝑖Δ𝜅subscript𝐱𝑖𝑡superscriptsubscript𝜅𝑥2superscriptsubscript𝜅𝑦2V_{1}(t_{j})=\frac{1}{4N}\sum_{i}\Delta(\kappa(\textbf{x}_{i})-t)\sqrt{\kappa_% {x}^{2}+\kappa_{y}^{2}},italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG 4 italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Δ ( italic_κ ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - italic_t ) square-root start_ARG italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (26)
V2(tj)=12πNiΔ(κ(𝐱i)t)(2κxκyκxyκx2κyyκy2κxxκx2+κy2),subscript𝑉2subscript𝑡𝑗12𝜋𝑁subscript𝑖Δ𝜅subscript𝐱𝑖𝑡2subscript𝜅𝑥subscript𝜅𝑦subscript𝜅𝑥𝑦superscriptsubscript𝜅𝑥2subscript𝜅𝑦𝑦superscriptsubscript𝜅𝑦2subscript𝜅𝑥𝑥superscriptsubscript𝜅𝑥2superscriptsubscript𝜅𝑦2V_{2}(t_{j})=\frac{1}{2\pi N}\sum_{i}\Delta(\kappa(\textbf{x}_{i})-t)\\ \left(\frac{2\kappa_{x}\kappa_{y}\kappa_{xy}-\kappa_{x}^{2}\kappa_{yy}-\kappa_% {y}^{2}\kappa_{xx}}{\kappa_{x}^{2}+\kappa_{y}^{2}}\right),start_ROW start_CELL italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG 2 italic_π italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Δ ( italic_κ ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - italic_t ) end_CELL end_ROW start_ROW start_CELL ( divide start_ARG 2 italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT italic_x italic_y end_POSTSUBSCRIPT - italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT italic_y italic_y end_POSTSUBSCRIPT - italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT italic_x italic_x end_POSTSUBSCRIPT end_ARG start_ARG italic_κ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_κ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , end_CELL end_ROW (27)

where ΔΔ\Deltaroman_Δ is 1 when κ(𝐱i)𝜅subscript𝐱𝑖\kappa(\textbf{x}_{i})italic_κ ( x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) is between the thresholds tjsubscript𝑡𝑗t_{j}italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and tj+1subscript𝑡𝑗1t_{j+1}italic_t start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT and 0 outside the range. Following Grewal et al. (2022), we evenly space the thresholds from μ3σ𝜇3𝜎\mu-3\sigmaitalic_μ - 3 italic_σ to μ+3σ𝜇3𝜎\mu+3\sigmaitalic_μ + 3 italic_σ, where μ𝜇\muitalic_μ is 0 and σ𝜎\sigmaitalic_σ is the field value standard deviation for each redshift in each cosmology. We differ slightly here, as in that work we chose thresholds for each individual map realisation. We note the thresholds are dynamic across reconstruction methods; we expect this to slightly improve our posterior contours compared to other approaches.

3.3 Evaluating the Posterior

We measure MFs from the pixels in the reconstructed posterior distributions of each method, then build a Fisher matrix FF\mathrm{F}roman_F to evaluate the constraining power on cosmological parameters:

Fij=mnXmθiCmn1Xnθj,subscriptF𝑖𝑗subscript𝑚𝑛subscript𝑋𝑚subscript𝜃𝑖subscriptsuperscriptC1𝑚𝑛subscript𝑋𝑛subscript𝜃𝑗\mathrm{F}_{ij}=\sum_{mn}\frac{\partial X_{m}}{\partial\theta_{i}}\mathrm{C}^{% -1}_{mn}\frac{\partial X_{n}}{\partial\theta_{j}},roman_F start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT divide start_ARG ∂ italic_X start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG roman_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT divide start_ARG ∂ italic_X start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG , (28)

where Xmsubscript𝑋𝑚X_{m}italic_X start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT is the measured MF, θisubscript𝜃𝑖\theta_{i}italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the cosmological parameter, and the covariance matrix Cmnsubscript𝐶𝑚𝑛C_{mn}italic_C start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT is (XmX¯m)(XnX¯n)delimited-⟨⟩subscript𝑋𝑚subscript¯𝑋𝑚subscript𝑋𝑛subscript¯𝑋𝑛\langle(X_{m}-\overline{X}_{m})(X_{n}-\overline{X}_{n})\rangle⟨ ( italic_X start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT - over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT ) ( italic_X start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT - over¯ start_ARG italic_X end_ARG start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) ⟩. The covariance is obtained from simulations at the fiducial cosmology. The Fisher derivatives are taken from the slope of the line of best fit through the three cosmologies in Table 1: minus parameter step size, fiducial, and plus parameter step size. The paired initial condition seeds decrease the noise in this calculation. To match DES Y3 constraining power, we divide the covariance matrix by the square root of the ratio of DES sky area (5000 sq deg) to our map size (15×\times×15 sq deg); this scale factor is 4.714 (Gatti et al., 2021b).

The correlation matrix derived from the covariance matrix is shown in Figure 3. While V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT has some correlation with itself, it is anti-correlated with V1subscript𝑉1V_{1}italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and V2subscript𝑉2V_{2}italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, which are highly correlated with each other. There is also strong correlation across redshift bins. We note the difference between this matrix and our previous MF correlation matrix in Grewal et al. (2022) can be attributed to the thresholding methodology.

Refer to caption
Figure 3: Correlation coefficient matrix of all MFs, which have been concatenated from all four redshift bins. There is strong correlation between the MFs across redshift. The boxes represent one functional, where there are three per bin (denoted by a thicker black line). The MFs have been measured from a true convergence map at fiducial cosmology.

With the center set at the fiducial cosmology, we follow the process in Coe (2009) to generate the ellipse parameters and build the contours in Section 4. We marginalise over the other parameters by inverting the Fisher matrix before removing them.

3.4 Figures of Merit

We use two figures of merit (FOM) to measure the constraining power of different reconstruction methods: the lensing amplitude parameter S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT and the inverse of the area of the ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT contour ellipse. S8σ8(Ωm0.3)αsubscript𝑆8subscript𝜎8superscriptsubscriptΩm0.3𝛼S_{8}\equiv\sigma_{8}\left(\frac{\Omega_{\mathrm{m}}}{0.3}\right)^{\alpha}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT ≡ italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT ( divide start_ARG roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT end_ARG start_ARG 0.3 end_ARG ) start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT (where α𝛼\alphaitalic_α=0.5) is perpendicular to the lensing degeneracy direction for ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT for typical survey redshifts. It is thus the minor axis of the ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT contours, and is usually the best constrained parameter combination.

Following Euclid Collaboration et al. (2023), we can transform the Fisher matrix in ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT into the matrix for ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT with

F(Ωm,S8)=MTF(Ωm,σ8)M,FsubscriptΩmsubscriptS8superscriptMTFsubscriptΩmsubscript𝜎8M\mathrm{F}(\Omega_{\mathrm{m}},\mathrm{S}_{8})=\mathrm{M^{T}}\mathrm{F}(\Omega% _{\mathrm{m}},\sigma_{8})\mathrm{M},roman_F ( roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT , roman_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT ) = roman_M start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT roman_F ( roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT ) roman_M , (29)

where

M=[ΩmΩmΩmS8σ8Ωmσ8S8]=[100.5σ8Ωm(0.3Ωm)0.5].MmatrixsubscriptΩmsubscriptΩmsubscriptΩmsubscriptS8subscript𝜎8subscriptΩmsubscript𝜎8subscriptS8matrix100.5subscript𝜎8subscriptΩmsuperscript0.3subscriptΩm0.5\mathrm{M}=\begin{bmatrix}\dfrac{\partial\Omega_{\mathrm{m}}}{\partial\Omega_{% \mathrm{m}}}&\dfrac{\partial\Omega_{\mathrm{m}}}{\partial\mathrm{S}_{8}}\\[10.% 0pt] \dfrac{\partial\sigma_{8}}{\partial\Omega_{\mathrm{m}}}&\dfrac{\partial\sigma_% {8}}{\partial\mathrm{S}_{8}}\end{bmatrix}=\begin{bmatrix}1&0\\ -0.5\dfrac{\sigma_{8}}{\Omega_{\mathrm{m}}}&\left(\dfrac{0.3}{\Omega_{\mathrm{% m}}}\right)^{0.5}\end{bmatrix}.roman_M = [ start_ARG start_ROW start_CELL divide start_ARG ∂ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT end_ARG start_ARG ∂ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT end_ARG end_CELL start_CELL divide start_ARG ∂ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT end_ARG start_ARG ∂ roman_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL divide start_ARG ∂ italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT end_ARG start_ARG ∂ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT end_ARG end_CELL start_CELL divide start_ARG ∂ italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT end_ARG start_ARG ∂ roman_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT end_ARG end_CELL end_ROW end_ARG ] = [ start_ARG start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL - 0.5 divide start_ARG italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT end_ARG start_ARG roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT end_ARG end_CELL start_CELL ( divide start_ARG 0.3 end_ARG start_ARG roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT 0.5 end_POSTSUPERSCRIPT end_CELL end_ROW end_ARG ] . (30)

In this work, we also aim to study the uncertainty in the DeepMass reconstruction. We quantify it with χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT per pixel:

χ2=1Npixels(XreconXtrueσ)2,superscript𝜒21subscript𝑁pixelssuperscriptsubscript𝑋reconsubscript𝑋true𝜎2\chi^{2}=\frac{1}{N_{\mathrm{pixels}}}\sum\left(\frac{X_{\mathrm{recon}}-X_{% \mathrm{true}}}{\sigma}\right)^{2},italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_pixels end_POSTSUBSCRIPT end_ARG ∑ ( divide start_ARG italic_X start_POSTSUBSCRIPT roman_recon end_POSTSUBSCRIPT - italic_X start_POSTSUBSCRIPT roman_true end_POSTSUBSCRIPT end_ARG start_ARG italic_σ end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (31)

using the σ𝜎\sigmaitalic_σ defined in Eq. (22). Xtruesubscript𝑋trueX_{\mathrm{true}}italic_X start_POSTSUBSCRIPT roman_true end_POSTSUBSCRIPT is the true MF observable and Xreconsubscript𝑋reconX_{\mathrm{recon}}italic_X start_POSTSUBSCRIPT roman_recon end_POSTSUBSCRIPT is the reconstructed MF observable, fiducial or non-fiducial.

4 Results

Refer to caption
Figure 4: Visualising the output for true maps, true smoothed maps, and the four different reconstruction methods. 15 ×\times× 15 degree flat sky convergence maps of redshift bin 3 at fiducial cosmology. The shared colourbar shows convergence of each pixel. A Gaussian kernel with a standard deviation of one arcmin has been applied to the true map to achieve the smoothed map.

Figure 4 shows the true, true smoothed, and reconstructed convergence maps for the four methods: Kaiser Squires, Wiener filter, DarkMappy, and DeepMass. As expected, peaks in the KS and WF maps have been smoothed out because small scales are noise-dominated; however, DarkMappy looks similarly noise-dominated despite it being a non-linear method. DeepMass appears most like the truth convergence map, which is evidence that it has denoised the convergence signal; this is consistent with predictions. While DeepMass takes more time to train the model than the other methods do to run, once the model is generated, DeepMass reconstructs convergence maps very quickly.

Since MFs are not linear, their total value does not come from summing the MFs from noise and signal. Here we have a separate analysis with signal and noise to study the effects of the latter. The Minkowski functional outputs for true, true-plus-noise, KS, WF, DarkMappy, and DeepMass are compared in Figure 5. The MF curves have the same shape, which is the typical shape for a convergence field. The distributions are not expected to be exactly the same, and here we see differing amplitudes for different methods corresponding to the noise level in the field. We do not expect KS to reduce the noise in any way, but some of the noise in KS has been smoothed out. As such, KS is not as noisy as true-plus-noise. Figure 5 shows KS, WF, and DarkMappy are comparable, while DeepMass has the lowest variation in amplitude due to noise suppression after smoothing.

Refer to caption
Figure 5: The first three MFs, V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, V1subscript𝑉1V_{1}italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and V2subscript𝑉2V_{2}italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, calculated from fiducial convergence maps. The curves shown are for true signal and true-plus-noise, along with the four reconstructed signals visualised in Figure 4: Kaiser Squires, Wiener filter, DarkMappy, and DeepMass. The t𝑡titalic_t values on the curve correspond to an excursion set, and the three MFs on the y𝑦yitalic_y-axis describe the perimeter, area, and curvature of the excursion set of the convergence.

We marginalise over all cosmological and baryonic parameters in Table 1 in our analysis except for the two used in our Fisher matrix. Figure 6 shows the contour comparison in the reconstruction methods for ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, and ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The contour directions are typical for convergence constraints. Figures of merit are shown in Table 3 with the standard deviation of S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT in the first column and the inverse of the ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT contour area in the second column.

Method σS8subscript𝜎subscript𝑆8\sigma_{S_{8}}italic_σ start_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT end_POSTSUBSCRIPT 1/EAsubscript𝐸𝐴E_{A}italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT
True 0.00392 28328.11
True + Mask 0.00387 19445.44
True + Noise 0.00819 3378.91
Kaiser Squires 0.00848 3484.66
Wiener Filter 0.00700 3994.15
DarkMappy 0.00940 2967.18
Deepmass 0.00409 6606.95
Table 3: Figures of merit for different reconstruction methods. EAsubscript𝐸𝐴E_{A}italic_E start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT is the ellipse area of the ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT contour.
Refer to caption
Refer to caption Refer to caption
Figure 6: w𝑤witalic_wCDM Fisher constraints for true signal, true-plus-mask, true-plus-noise, KS reconstruction, WF reconstructions, DarkMappy reconstruction, and DeepMass reconstruction. The top left plot shows constraints on ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT vs w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the bottom left plot shows constraints on ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT vs σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, and the bottom right plot shows constraints on ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT vs S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT. KS, WF, and DarkMappy are comparable at these scales, while DeepMass shows significantly tighter constraints, implying the model has effectively denoised the signal.

DeepMass outperforms the other reconstruction methods by roughly 50%percent\%% and has constraints closest to the true contour. KS, WF, and DarkMappy have similar constraints within a 30% interval of each other, and are comparable to true-plus-noise, which shows adding noise doubles the size of the contour. We note the choice of threshold range can affect the size of the contours. As mentioned previously, KS is not expected to perform as well as true-plus-noise, but some of the noise has been smoothed out. We can see in Figure 6 that the step of masking the data has a much smaller impact than the step of adding noise to the fields, the latter doubling the size of the parameter constraints. We use standard value 0.5 for α𝛼\alphaitalic_α in the σS8subscript𝜎subscript𝑆8\sigma_{S_{8}}italic_σ start_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT end_POSTSUBSCRIPT calculation, which corresponds to the width perpendicular to the direction of the true-plus-mask contour. We also include the second column for inverse ellipse area to more completely quantify constraining power of cosmological parameters, and we see that the true map has the largest value, followed by true-plus-mask. On the whole, ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT is better constrained than other cosmological parameters. We can see the same trends in the ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT and ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT - w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT planes. We investigate baryonic parameters in Appendix A.

Using Eq. (31), we compare χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT values per pixel to measure the effect of training DeepMass at a (slightly) wrong cosmology. Figure 7 shows a comparison of DeepMass outputs for the fiducial and non-fiducial models (-ΔΩmΔsubscriptΩm\Delta\Omega_{\mathrm{m}}roman_Δ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT) applied to non-fiducial data (-ΔΩmΔsubscriptΩm\Delta\Omega_{\mathrm{m}}roman_Δ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT). Each histogram displays the χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT values across 5000 output maps. The incorrect model histogram illustrates the distinction between the fiducial model and true data with -ΔΩmΔsubscriptΩm\Delta\Omega_{\mathrm{m}}roman_Δ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT, while the correct model histogram depicts differences for the non-fiducial model on the same input data. As expected, the histogram for the correct model has a lower average χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT value. The quantification of uncertainty from the reconstruction is determined as the difference between the means of the two histograms: 3.8%. While this seems relatively small, it leads to changes in the MFs comparable to cosmologically-induced changes, as seen in Figure 8. As this can only worsen the constraining power of DeepMass, which is already close to the ideal constraints from truth data, it cannot have had a significant effect on our results here, as we focus on the reconstruction precision (not accuracy). However, this demonstrates that the selection of training model cosmologies is critical for future analysis accuracy. Tests in Jeffrey et al. (2020) used the truth cosmologies for optimal training, but analyses of real data will need thoughtful and perhaps iteratively-chosen training samples and careful emulators to interpolate between them.

Refer to caption
Figure 7: Normalised χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT/pixel comparison of DeepMass reconstruction using the correct model versus the incorrect model. The correct model (purple) gives a lower χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT value than the incorrect model (orange), indicating that choice of model is important. We measure a 3.8% uncertainty in the cosmology choice in DeepMass training, which has been calculated as the percent change in the difference in the means of the two χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT histograms.
Refer to caption
Figure 8: MFs for the two DeepMass outputs in Figure 7. Errorbars are shown on the output from the correct model.

5 Conclusion

We want to use the most precise reconstructed convergence maps in order to best capture the original signal. In this way, measuring statistics from such detailed reconstructed maps enables improved measurement of cosmological parameters. In this paper we investigate four non-parametric mass mapping methods developed to reconstruct convergence from shear: Kaiser Squires, Wiener filter, DarkMappy, and DeepMass. We smooth masked and noisy shear maps and reconstruct convergence maps using the different methods. KS requires smoothing because it can only take a continuous field as input. We then measure Minkowski functionals from the reconstructed convergence maps and perform a Fisher analysis to make contours. We compare constraints on ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT, σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, and w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (see Appendix A for baryonic parameter analysis).

The contour plots for different combinations of cosmological parameters all show the same results for different reconstruction methods. DeepMass outperforms the rest of the methods by a significant degree, demonstrating the model is a successful denoiser and an effective tool for reconstructing convergence from shear. Beyond a 50%percent\%% improvement in constraining power, the model is also computationally efficient.

We also measure the uncertainty from using different cosmologies in the DeepMass model training. By comparing χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT values of true data MFs and MFs from data trained on the correct model or the same data trained on an incorrect model, we can quantify the impact of cosmological choices on model performance with the difference in average χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT values. This is about 3.8%, indicating the choice of training model cosmology has an impact on reconstruction precision.

This difference aligns closely with observed variations in average MF DeepMass outputs from data trained on the correct or incorrect model. Both effects are comparable with the change in cosmology for ΩmsubscriptΩ𝑚\Omega_{m}roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT. These findings demonstrate a strong dependence of DeepMass on the cosmology of the training dataset, emphasising the need for the consideration of input cosmology in model training to ensure accurate and reliable results.

The remaining methods, KS, WF, and DarkMappy are comparable at the scales we use here. DarkMappy was expected to outperform KS and WF. One reason it did not could be that it was designed for cluster analysis and may be tuned for such scenarios rather than our wide-field tests; alternative choices of its parameters and wavelet dictionary could improve its performance. Another could be the resolution of the field. In this paper we have used CosmoGridV1 simulations with DES Y3-like redshift bins and noise, which have led to a more limited resolution. We expect an improvement in resolution with less noisy survey conditions like for LSST or in future LSST-like forecasts, as well as with newer simulations.

Additionally, the choice of threshold levels can influence contour sizes. Here we use dynamic thresholds, where the range is set by the mean and standard deviation of the field for each cosmology, redshift, and reconstruction method. Using fixed thresholds more carefully tuned to each case has less variability. This can lead to smaller contours, but the thresholds are less consistently useful for different inputs like redshift bins, reconstruction methods, and simulation conditions like noise level.

We use Minkowski functionals to evaluate the different reconstruction methods, but there are many other statistics that could be used. Other higher order statistics may yield different results and show more distinction between the contours. Likewise, we have chosen four reconstruction methods here, representative of the different approaches to mass mapping. Other methods may be as or more precise.

In future analysis, using different (or multiple statistics) could lead to better information about how the different methods perform. Additionally, using lower noise conditions like LSST and other reconstruction methods could lead to stronger differentiation among methods.

We finally note that mass mapping for higher-order statistics is an ideal problem for a community challenge like the ones presented in Zuntz et al. (2021) and Leonard et al. (2023). This would both focus authors of reconstruction methods on useful metrics for upcoming surveys and incentivise them to optimise parameters and choices for a specific scenario.

6 Acknowledgements

We thank Tomasz Kacprzak for help with the CosmoGridV1 simulations. NG thanks Harry Rendell-Bhatti for the useful discussions. TT acknowledges funding from the Swiss National Science Foundation under the Ambizione project PZ00P2_193352. Results in this paper made use of many software packages, including Numpy, Scipy, and CCL (Harris et al., 2020; Virtanen et al., 2020; Chisari et al., 2019).

References

  • Abbott et al. (2022) Abbott, T. M. C., Aguena, M., Alarcon, A., et al. 2022, Phys. Rev. D, 105, 023520, doi: 10.1103/PhysRevD.105.023520
  • Aihara et al. (2017) Aihara, H., Arimoto, N., Armstrong, R., et al. 2017, Publications of the Astronomical Society of Japan, 70, doi: 10.1093/pasj/psx066
  • Aihara et al. (2022) Aihara, H., AlSayyad, Y., Ando, M., et al. 2022, Publications of the Astronomical Society of Japan, 74, 247, doi: 10.1093/pasj/psab122
  • Ajani et al. (2020) Ajani, V., Peel, A., Pettorino, V., et al. 2020, Physical Review D, 102, doi: 10.1103/PhysRevD.102.103531
  • Alsing et al. (2015) Alsing, J., Heavens, A., Jaffe, A. H., et al. 2015, Monthly Notices of the Royal Astronomical Society, 455, 4452–4466, doi: 10.1093/mnras/stv2501
  • Amon et al. (2022) Amon, A., Gruen, D., Troxel, M. A., et al. 2022, Physical Review D, 105, doi: 10.1103/physrevd.105.023514
  • Appleby et al. (2022) Appleby, S., Park, C., Pranav, P., et al. 2022, The Astrophysical Journal, 928, 108, doi: 10.3847/1538-4357/ac562a
  • Asgari et al. (2021) Asgari, M., Lin, C.-A., Joachimi, B., et al. 2021, Astronomy & Astrophysics, 645, A104, doi: 10.1051/0004-6361/202039070
  • Bartelmann & Schneider (2001) Bartelmann, M., & Schneider, P. 2001, Physics Reports, 340, 291–472, doi: 10.1016/s0370-1573(00)00082-x
  • Barthelemy et al. (2020) Barthelemy, A., Codis, S., Uhlemann, C., Bernardeau, F., & Gavazzi, R. 2020, Monthly Notices of the Royal Astronomical Society, 492, 3420
  • Boruah et al. (2022) Boruah, S. S., Rozo, E., & Fiedorowicz, P. 2022, Monthly Notices of the Royal Astronomical Society, 516, 4111
  • Boyle et al. (2021) Boyle, A., Uhlemann, C., Friedrich, O., et al. 2021, Monthly Notices of the Royal Astronomical Society, 505, 2886
  • Buchert et al. (2017) Buchert, T., France, M. J., & Steiner, F. 2017, Classical and Quantum Gravity, 34, 094002, doi: 10.1088/1361-6382/aa5ce2
  • Cai et al. (2018) Cai, X., Pereyra, M., & McEwen, J. D. 2018, Monthly Notices of the Royal Astronomical Society, 480, 4170, doi: 10.1093/mnras/sty2015
  • Chang et al. (2018) Chang, C., Pujol, A., Mawdsley, B., et al. 2018, MNRAS, 475, 3165, doi: 10.1093/mnras/stx3363
  • Cheng & Ménard (2021) Cheng, S., & Ménard, B. 2021, Monthly Notices of the Royal Astronomical Society, 507, 1012
  • Cheng et al. (2020) Cheng, S., Ting, Y.-S., Ménard, B., & Bruna, J. 2020, Monthly Notices of the Royal Astronomical Society, 499, 5902
  • Chingangbam & Rahman (2023) Chingangbam, P., & Rahman, F. 2023, Minkowski Functionals for composite smooth random fields. https://arxiv.org/abs/2311.12571
  • Chisari et al. (2019) Chisari, N. E., Alonso, D., Krause, E., et al. 2019, The Astrophysical Journal Supplement Series, 242, 2, doi: 10.3847/1538-4365/ab1658
  • Coe (2009) Coe, D. 2009, Fisher Matrices and Confidence Ellipses: A Quick-Start Guide and Software. https://arxiv.org/abs/0906.4123
  • Collaboration et al. (2020) Collaboration, P., Akrami, Y., Ashdown, M., et al. 2020, Astronomy & Astrophysics, 641, A7, doi: 10.1051/0004-6361/201935201
  • Eriksen et al. (2004) Eriksen, H. K., Novikov, D. I., Lilje, P. B., Banday, A. J., & Gorski, K. M. 2004, The Astrophysical Journal, 612, 64, doi: 10.1086/422570
  • Euclid Collaboration et al. (2023) Euclid Collaboration, Ajani, V., Baldi, M., et al. 2023, A&A, 675, A120, doi: 10.1051/0004-6361/202346017
  • Feldbrugge et al. (2019) Feldbrugge, J., Van Engelen, M., van de Weygaert, R., Pranav, P., & Vegter, G. 2019, Journal of Cosmology and Astroparticle Physics, 2019, 052
  • Fiedorowicz et al. (2022) Fiedorowicz, P., Rozo, E., Boruah, S. S., Chang, C., & Gatti, M. 2022, Monthly Notices of the Royal Astronomical Society, 512, 73–85, doi: 10.1093/mnras/stac468
  • Fischbacher et al. (2022) Fischbacher, S., Kacprzak, T., Blazek, J., & Refregier, A. 2022, Redshift requirements for cosmic shear with intrinsic alignment, arXiv, doi: 10.48550/ARXIV.2207.01627
  • Flaugher (2005) Flaugher, B. 2005, International Journal of Modern Physics A, 20, 3121, doi: 10.1142/S0217751X05025917
  • Gatti et al. (2023) Gatti, M., Jeffrey, N., Whiteway, L., et al. 2023, Dark Energy Survey Year 3 results: simulation-based cosmological inference with wavelet harmonics, scattering transforms, and moments of weak lensing mass maps I: validation on simulations. https://arxiv.org/abs/2310.17557
  • Gatti et al. (2020) Gatti, M., Chang, C., Friedrich, O., et al. 2020, MNRAS, 498, 4060, doi: 10.1093/mnras/staa2680
  • Gatti et al. (2021a) Gatti, M., Sheldon, E., Amon, A., et al. 2021a, Monthly Notices of the Royal Astronomical Society, 504, 4312–4336, doi: 10.1093/mnras/stab918
  • Gatti et al. (2021b) Gatti, M., Jain, B., Chang, C., et al. 2021b, doi: 10.48550/ARXIV.2110.10141
  • Géron (2019) Géron, A. 2019, Hands-On Machine Learning with Scikit-Learn, Keras, and TensorFlow, 2nd Edition (O’Reilly Media, Inc.)
  • Giblin et al. (2021) Giblin, B., Heymans, C., Asgari, M., et al. 2021, Astronomy & Astrophysics, 645, A105, doi: 10.1051/0004-6361/202038850
  • Gong et al. (2023) Gong, Z., Halder, A., Barreira, A., Seitz, S., & Friedrich, O. 2023, Journal of Cosmology and Astroparticle Physics, 2023, 040, doi: 10.1088/1475-7516/2023/07/040
  • Grewal et al. (2022) Grewal, N., Zuntz, J., Tröster, T., & Amon, A. 2022, The Open Journal of Astrophysics, 5, doi: 10.21105/astro.2206.03877
  • Hamana et al. (2020) Hamana, T., Shirasaki, M., Miyazaki, S., et al. 2020, Publications of the Astronomical Society of Japan, 72, doi: 10.1093/pasj/psz138
  • Hamann & Kang (2023) Hamann, J., & Kang, Y. 2023. https://arxiv.org/abs/2310.14618
  • Harnois-Déraps et al. (2021) Harnois-Déraps, J., Martinet, N., Castro, T., et al. 2021, Monthly Notices of the Royal Astronomical Society, 506, 1623
  • Harris et al. (2020) Harris, C. R., Millman, K. J., van der Walt, S. J., et al. 2020, Nature, 585, 357, doi: 10.1038/s41586-020-2649-2
  • Heydenreich et al. (2021) Heydenreich, S., Brück, B., & Harnois-Déraps, J. 2021, Astronomy & Astrophysics, 648, A74
  • Heymans et al. (2021) Heymans, C., Tröster, T., Asgari, M., et al. 2021, Astronomy & Astrophysics, 646, A140, doi: 10.1051/0004-6361/202039063
  • Hikage et al. (2006) Hikage, C., Komatsu, E., & Matsubara, T. 2006, The Astrophysical Journal, 653, 11, doi: 10.1086/508653
  • Hikage et al. (2003) Hikage, C., Schmalzing, J., Buchert, T., et al. 2003, Publications of the Astronomical Society of Japan, 55, 911, doi: 10.1093/pasj/55.5.911
  • Hikage et al. (2019) Hikage, C., Oguri, M., Hamana, T., et al. 2019, Publications of the Astronomical Society of Japan, 71, doi: 10.1093/pasj/psz010
  • Horowitz et al. (2019) Horowitz, B., Seljak, U., & Aslanyan, G. 2019, Journal of Cosmology and Astroparticle Physics, 2019, 035–035, doi: 10.1088/1475-7516/2019/10/035
  • Jeffrey et al. (2020) Jeffrey, N., Lanusse, F., Lahav, O., & Starck, J.-L. 2020, Monthly Notices of the Royal Astronomical Society, 492, 5023–5029, doi: 10.1093/mnras/staa127
  • Jeffrey et al. (2018) Jeffrey, N., Abdalla, F. B., Lahav, O., et al. 2018, Monthly Notices of the Royal Astronomical Society, 479, 2871, doi: 10.1093/mnras/sty1252
  • Kacprzak et al. (2022) Kacprzak, T., Fluri, J., Schneider, A., Refregier, A., & Stadel, J. 2022, doi: 10.48550/ARXIV.2209.04662
  • Kacprzak et al. (2016) Kacprzak, T., Kirk, D., Friedrich, O., et al. 2016, Monthly Notices of the Royal Astronomical Society, 463, 3653, doi: 10.1093/mnras/stw2070
  • Kaiser & Squires (1993) Kaiser, N., & Squires, G. K. 1993, The Astrophysical Journal, 404, 441
  • Kansal (2023) Kansal, V. 2023, Astronomy & Astrophysics, 670, A34, doi: 10.1051/0004-6361/202245198
  • Kilbinger (2015) Kilbinger, M. 2015, Reports on Progress in Physics, 78, 086901, doi: 10.1088/0034-4885/78/8/086901
  • Kratochvil et al. (2010) Kratochvil, J. M., Haiman, Z., & May, M. 2010, Physical Review D, 81, doi: 10.1103/physrevd.81.043519
  • Kratochvil et al. (2012) Kratochvil, J. M., Lim, E. A., Wang, S., et al. 2012, Physical Review D, 85, doi: 10.1103/physrevd.85.103513
  • Kuijken et al. (2019) Kuijken, K., Heymans, C., Dvornik, A., et al. 2019, A&A, 625, A2, doi: 10.1051/0004-6361/201834918
  • Lahav et al. (1994) Lahav, O., Fisher, K. B., Hoffman, Y., Scharf, C. A., & Zaroubi, S. 1994, The Astrophysical Journal, 423, L93, doi: 10.1086/187244
  • Lanusse et al. (2016) Lanusse, F., Starck, J.-L., Leonard, A., & Pires, S. 2016, Astronomy & Astrophysics, 591, A2, doi: 10.1051/0004-6361/201628278
  • Leonard et al. (2012) Leonard, A., Dupé, F.-X., & Starck, J.-L. 2012, Astronomy & Astrophysics, 539, A85, doi: 10.1051/0004-6361/201117642
  • Leonard et al. (2014) Leonard, A., Lanusse, F., & Starck, J.-L. 2014, Monthly Notices of the Royal Astronomical Society, 440, 1281, doi: 10.1093/mnras/stu273
  • Leonard et al. (2023) Leonard, C. D., et al. 2023, Open Journal of Astrophysics, 6, 1, doi: 10.21105/astro.2212.04291
  • Liu & Madhavacheril (2019) Liu, J., & Madhavacheril, M. S. 2019, Physical Review D, 99, 083508
  • Liu et al. (2015) Liu, J., Petri, A., Haiman, Z., et al. 2015, Phys. Rev. D, 91, 063507, doi: 10.1103/PhysRevD.91.063507
  • Liu et al. (2022) Liu, W., Jiang, A., & Fang, W. 2022, Journal of Cosmology and Astroparticle Physics, 2022, 045, doi: 10.1088/1475-7516/2022/07/045
  • LSST Dark Energy Science Collaboration (2012) LSST Dark Energy Science Collaboration. 2012, doi: 10.48550/ARXIV.1211.0310
  • LSST Science Collaboration et al. (2009) LSST Science Collaboration, Abell, P. A., Allison, J., et al. 2009, doi: 10.48550/ARXIV.0912.0201
  • Mallat (2012) Mallat, S. 2012, Communications on Pure and Applied Mathematics, 65, doi: 10.1002/cpa.21413
  • Marian et al. (2009) Marian, L., Smith, R. E., & Bernstein, G. M. 2009, The Astrophysical Journal, 698, L33
  • Marshall et al. (2002) Marshall, P. J., Hobson, M. P., Gull, S. F., & Bridle, S. L. 2002, Monthly Notices of the Royal Astronomical Society, 335, 1037–1048, doi: 10.1046/j.1365-8711.2002.05685.x
  • Martinet et al. (2018) Martinet, N., Schneider, P., Hildebrandt, H., et al. 2018, MNRAS, 474, 712, doi: 10.1093/mnras/stx2793
  • McCarthy et al. (2022) McCarthy, F., Hill, J. C., & Madhavacheril, M. S. 2022, Physical Review D, 105, doi: 10.1103/physrevd.105.023517
  • Mead et al. (2015) Mead, A. J., Peacock, J. A., Heymans, C., Joudaki, S., & Heavens, A. F. 2015, Monthly Notices of the Royal Astronomical Society, 454, 1958–1975, doi: 10.1093/mnras/stv2036
  • Mecke et al. (1993) Mecke, K. R., Buchert, T., & Wagner, H. 1993, doi: 10.48550/ARXIV.ASTRO-PH/9312028
  • Minkowski (1903) Minkowski, H. 1903, Mathematische Annalen, Vol. Volumen und Oberfläche, 447–495
  • Miyatake et al. (2021) Miyatake, H., et al. 2021, Publications of the Astronomical Society of Japan, doi: 10.48550/ARXIV.2111.02419
  • Parroni et al. (2020) Parroni, C., Cardone, V. F., Maoli, R., & Scaramella, R. 2020, Astronomy & Astrophysics, 633, A71, doi: 10.1051/0004-6361/201935988
  • Parroni et al. (2021) Parroni, C., Tollet, É., Cardone, V. F., Maoli, R., & Scaramella, R. 2021, Astronomy & Astrophysics, 645, A123
  • Peel et al. (2018) Peel, A., Pettorino, V., Giocoli, C., Starck, J.-L., & Baldi, M. 2018, A&A, 619, A38, doi: 10.1051/0004-6361/201833481
  • Petri et al. (2013) Petri, A., Haiman, Z., Hui, L., May, M., & Kratochvil, J. M. 2013, Physical Review D, 88, doi: 10.1103/physrevd.88.123002
  • Petri et al. (2015) Petri, A., Liu, J., Haiman, Z., et al. 2015, Phys. Rev. D, 91, 103511, doi: 10.1103/PhysRevD.91.103511
  • Pires et al. (2020) Pires, S., Vandenbussche, V., Kansal, V., et al. 2020, Astronomy & Astrophysics, 638, A141, doi: 10.1051/0004-6361/201936865
  • Porqueres et al. (2021) Porqueres, N., Heavens, A., Mortlock, D., & Lavaux, G. 2021, Monthly Notices of the Royal Astronomical Society, 509, 3194, doi: 10.1093/mnras/stab3234
  • Porth & Smith (2021) Porth, L., & Smith, R. E. 2021, Monthly Notices of the Royal Astronomical Society, 508, 3474, doi: 10.1093/mnras/stab2819
  • Potter et al. (2016) Potter, D., Stadel, J., & Teyssier, R. 2016, PKDGRAV3: Beyond Trillion Particle Cosmological Simulations for the Next Era of Galaxy Surveys. https://arxiv.org/abs/1609.08621
  • Price et al. (2021) Price, M. A., McEwen, J. D., Cai, X., Kitching, T. D., & and, C. G. R. W. 2021, Monthly Notices of the Royal Astronomical Society, 506, 3678, doi: 10.1093/mnras/stab1983
  • Remy et al. (2022) Remy, B., Lanusse, F., Jeffrey, N., et al. 2022, Probabilistic Mass Mapping with Neural Score Estimation. https://arxiv.org/abs/2201.05561
  • Ronneberger et al. (2015) Ronneberger, O., Fischer, P., & Brox, T. 2015, U-Net: Convolutional Networks for Biomedical Image Segmentation. https://arxiv.org/abs/1505.04597
  • Scaramella et al. (2021) Scaramella, R., Amiaux, J., Mellier, Y., et al. 2021, doi: 10.48550/ARXIV.2108.01201
  • Schmalzing & Gorski (1998) Schmalzing, J., & Gorski, K. M. 1998, Monthly Notices of the Royal Astronomical Society, 297, 355, doi: 10.1046/j.1365-8711.1998.01467.x
  • Schneider et al. (2019) Schneider, A., Teyssier, R., Stadel, J., et al. 2019, Journal of Cosmology and Astroparticle Physics, 2019, 020, doi: 10.1088/1475-7516/2019/03/020
  • Schneider et al. (2017) Schneider, M. D., Ng, K. Y., Dawson, W. A., et al. 2017, The Astrophysical Journal, 839, 25, doi: 10.3847/1538-4357/839/1/25
  • Schneider & Lombardi (2003) Schneider, P., & Lombardi, M. 2003, Astronomy & Astrophysics, 397, 809, doi: 10.1051/0004-6361:20021541
  • Secco et al. (2022) Secco, L. F., Samuroff, S., Krause, E., et al. 2022, Physical Review D, 105, doi: 10.1103/physrevd.105.023515
  • Shan et al. (2017) Shan, H., Liu, X., Hildebrandt, H., et al. 2017, Monthly Notices of the Royal Astronomical Society, 474, 1116–1134, doi: 10.1093/mnras/stx2837
  • Shirasaki et al. (2021) Shirasaki, M., Moriwaki, K., Oogi, T., et al. 2021, Monthly Notices of the Royal Astronomical Society, 504, 1825, doi: 10.1093/mnras/stab982
  • Simon et al. (2011a) Simon, P., Heymans, C., Schrabback, T., et al. 2011a, Monthly Notices of the Royal Astronomical Society, 419, 998–1016, doi: 10.1111/j.1365-2966.2011.19760.x
  • Simon et al. (2009) Simon, P., Taylor, A. N., & Hartlap, J. 2009, Monthly Notices of the Royal Astronomical Society, 399, 48–68, doi: 10.1111/j.1365-2966.2009.15246.x
  • Simon et al. (2011b) Simon, P., Heymans, C., Schrabback, T., et al. 2011b, Monthly Notices of the Royal Astronomical Society, 419, 998–1016, doi: 10.1111/j.1365-2966.2011.19760.x
  • Smail et al. (1995) Smail, I., Hogg, D. W., Yan, L., & Cohen, J. G. 1995, The Astrophysical Journal, 449, doi: 10.1086/309647
  • Spergel et al. (2015) Spergel, D., Gehrels, N., Baltay, C., et al. 2015, doi: 10.48550/ARXIV.1503.03757
  • Starck et al. (2021) Starck, J.-L., Themelis, K. E., Jeffrey, N., Peel, A., & Lanusse, F. 2021, Astronomy & Astrophysics, 649, A99, doi: 10.1051/0004-6361/202039451
  • Sullivan et al. (2019) Sullivan, J. M., Wiegand, A., & Eisenstein, D. J. 2019, Monthly Notices of the Royal Astronomical Society, 485, 1708, doi: 10.1093/mnras/stz498
  • Takada & Jain (2004) Takada, M., & Jain, B. 2004, Monthly Notices of the Royal Astronomical Society, 348, 897, doi: 10.1111/j.1365-2966.2004.07410.x
  • Thiele et al. (2020) Thiele, L., Hill, J. C., & Smith, K. M. 2020, Physical Review D, 102, 123545
  • Van Waerbeke et al. (2013) Van Waerbeke, L., Benjamin, J., Erben, T., et al. 2013, MNRAS, 433, 3373, doi: 10.1093/mnras/stt971
  • VanderPlas et al. (2011) VanderPlas, J. T., Connolly, A. J., Jain, B., & Jarvis, M. 2011, The Astrophysical Journal, 727, 118, doi: 10.1088/0004-637x/727/2/118
  • Vicinanza et al. (2016) Vicinanza, M., Cardone, V. F., Maoli, R., Scaramella, R., & Er, X. 2016, doi: 10.48550/ARXIV.1606.03892
  • Vicinanza et al. (2018) Vicinanza, M., Cardone, V. F., Maoli, R., Scaramella, R., & Er, X. 2018, Phys. Rev. D, 97, 023519, doi: 10.1103/PhysRevD.97.023519
  • Vicinanza et al. (2019) Vicinanza, M., Cardone, V. F., Maoli, R., et al. 2019, Physical Review D, 99, doi: 10.1103/physrevd.99.043534
  • Virtanen et al. (2020) Virtanen, P., Gommers, R., Oliphant, T. E., et al. 2020, Nature Methods, 17, 261, doi: 10.1038/s41592-019-0686-2
  • Wiegand & Eisenstein (2017) Wiegand, A., & Eisenstein, D. J. 2017, Monthly Notices of the Royal Astronomical Society, 467, 3361, doi: 10.1093/mnras/stx292
  • Zaroubi et al. (1995) Zaroubi, S., Hoffman, Y., Fisher, K. B., & Lahav, O. 1995, The Astrophysical Journal, 449, 446, doi: 10.1086/176070
  • Zuntz et al. (2021) Zuntz, J., Lanusse, F., Malz, A. I., et al. 2021, The Open Journal of Astrophysics, 4, doi: 10.21105/astro.2108.13418
  • Zürcher et al. (2021) Zürcher, D., Fluri, J., Sgier, R., Kacprzak, T., & Refregier, A. 2021, Journal of Cosmology and Astroparticle Physics, 2021, 028, doi: 10.1088/1475-7516/2021/01/028
  • Zürcher et al. (2022) Zürcher, D., Fluri, J., Sgier, R., et al. 2022, Monthly Notices of the Royal Astronomical Society, 511, 2075, doi: 10.1093/mnras/stac078

Appendix A Effects of Baryons

A.1 Comparing Fisher calculations for Baryonic Parameter Constraints

Here we explore the effects of baryons in our simulations and analysis. Figure 9 shows contour plots for baryonic parameters. The left plot shows the Mc0νsubscriptsuperscript𝑀0c𝜈M^{0}_{\mathrm{c}}-\nuitalic_M start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT - italic_ν plane, and the right plot shows the σ8Mc0subscript𝜎8subscriptsuperscript𝑀0c\sigma_{8}-M^{0}_{\mathrm{c}}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT - italic_M start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT plane. The results are unexpected, as the true maps should have the tightest constraints. We would expect smoothing and noise in the reconstruction process to lead to a loss of information, especially for small-scales where we expect to see baryonic effects. These contours are not reliable. To investigate this discrepancy, we study the baryonic effects at each step of the analysis.

Refer to caption
Refer to caption
Figure 9: Contours for different baryonic parameter combinations. The same four reconstruction methods are compared, which produce similar constraints, with DeepMass performing the best.

A.2 Baryonic non-fiducial Convergence Maps

We first compare the true baryonic and fiducial convergence maps to visualise the effect of a change in the baryonic parameters Mc0superscriptsubscript𝑀c0M_{\mathrm{c}}^{0}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and ν𝜈\nuitalic_ν. Figure 10 shows the difference between the non-fiducial ±ΔMc0plus-or-minusΔsuperscriptsubscript𝑀c0\pm\Delta M_{\mathrm{c}}^{0}± roman_Δ italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and ±Δνplus-or-minusΔ𝜈\pm\Delta\nu± roman_Δ italic_ν baryonic convergence maps and the fiducial convergence maps. We can see that the change in signal is very small compared to the original convergence value. This is roughly two orders of magnitude smaller than the change for nonbaryonic parameters.

Refer to caption
Figure 10: The difference in the non-fiducial and fiducial convergence maps for baryon parameters Mc0superscriptsubscript𝑀c0M_{\mathrm{c}}^{0}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and ν𝜈\nuitalic_ν.

A.3 Minkowski functionals for Baryonic Parameters

We next compare the Minkowski functional observables measured from the baryonic and fiducial convergence maps. In Figures 11 and 12, the functionals at different baryon cosmologies are shown to be nearly identical. This is not the case for the cosmological parameters we constrain in this paper; an example of variation in MF observables for ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT is shown in Figure 13. Here, any difference in the MFs falls within the noise regime. This may be because the ΔΔ\Deltaroman_Δ parameter step size is too small or the pixel size is too large to visualise or measure baryonic effects. These changes in step size are small enough that numerical noise dominates true Fisher matrix contours.

Refer to caption
Refer to caption
Figure 11: MFs at different values of ν𝜈\nuitalic_ν (top row). Difference in MFs of ΔνΔ𝜈\Delta\nuroman_Δ italic_ν and fiducial maps (bottom row).
Refer to caption
Refer to caption
Figure 12: MFs at different values of Mc0superscriptsubscript𝑀c0M_{\mathrm{c}}^{0}italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT (top row). Difference in MFs of ΔMc0Δsuperscriptsubscript𝑀c0\Delta M_{\mathrm{c}}^{0}roman_Δ italic_M start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT and fiducial maps (bottom row).

Appendix B Minkowski functionals for Non-Baryonic Parameters

For non-baryonic parameters, we see the expected change in the Minkowski functionals that corresponds to a change in cosmology. Figure 13 shows MFs for ±ΔΩmplus-or-minusΔsubscriptΩm\pm\Delta\Omega_{\mathrm{m}}± roman_Δ roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT as well as the fiducial value for ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT.

Refer to caption
Figure 13: MFs at different values of ΩmsubscriptΩm\Omega_{\mathrm{m}}roman_Ω start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT; the change is comparable to the variation from training cosmology in Figure 8.