11institutetext: 1 CEICO, Institute of Physics of the Czech Academy of Sciences, Na Slovance 2, Praha 8, Czech Republic
2 Department of Physics, Oxford University, Keble Road, Oxford OX1 3RH, UK
3 INAF-Osservatorio Astronomico di Roma, Via Frascati 33, 00078 Monteporzio Catone, Italy
4 INFN-Sezione di Roma, Piazzale Aldo Moro, 2 - c/o Dipartimento di Fisica, Edificio G. Marconi, 00185 Roma, Italy
5 Institut de Recherche en Astrophysique et Planétologie (IRAP), Université de Toulouse, CNRS, UPS, CNES, 14 Av. Edouard Belin, 31400 Toulouse, France
6 Université de Genève, Département de Physique Théorique and Centre for Astroparticle Physics, 24 quai Ernest-Ansermet, CH-1211 Genève 4, Switzerland
7 Dipartimento di Fisica, Universitá degli Studi di Torino, Via P. Giuria 1, 10125 Torino, Italy
8 INFN-Sezione di Torino, Via P. Giuria 1, 10125 Torino, Italy
9 INAF-Osservatorio Astrofisico di Torino, Via Osservatorio 20, 10025 Pino Torinese (TO), Italy
10 Université Paris-Saclay, CNRS, Institut d’astrophysique spatiale, 91405, Orsay, France
11 Institute of Cosmology and Gravitation, University of Portsmouth, Portsmouth PO1 3FX, UK
12 INAF-Osservatorio Astronomico di Brera, Via Brera 28, 20122 Milano, Italy
13 INAF-Osservatorio di Astrofisica e Scienza dello Spazio di Bologna, Via Piero Gobetti 93/3, 40129 Bologna, Italy
14 Dipartimento di Fisica e Astronomia, Universitá di Bologna, Via Gobetti 93/2, 40129 Bologna, Italy
15 INFN-Sezione di Bologna, Viale Berti Pichat 6/2, 40127 Bologna, Italy
16 Dipartimento di Fisica, Universitá di Genova, Via Dodecaneso 33, 16146, Genova, Italy
17 INFN-Sezione di Genova, Via Dodecaneso 33, 16146, Genova, Italy
18 Department of Physics ”E. Pancini”, University Federico II, Via Cinthia 6, 80126, Napoli, Italy
19 INAF-Osservatorio Astronomico di Capodimonte, Via Moiariello 16, 80131 Napoli, Italy
20 Instituto de Astrofísica e Ciências do Espaço, Universidade do Porto, CAUP, Rua das Estrelas, PT4150-762 Porto, Portugal
21 INAF-IASF Milano, Via Alfonso Corti 12, 20133 Milano, Italy
22 Institut de Física d’Altes Energies (IFAE), The Barcelona Institute of Science and Technology, Campus UAB, 08193 Bellaterra (Barcelona), Spain
23 Port d’Informació Científica, Campus UAB, C. Albareda s/n, 08193 Bellaterra (Barcelona), Spain
24 Institute for Theoretical Particle Physics and Cosmology (TTK), RWTH Aachen University, 52056 Aachen, Germany
25 INFN section of Naples, Via Cinthia 6, 80126, Napoli, Italy
26 Dipartimento di Fisica e Astronomia ”Augusto Righi” - Alma Mater Studiorum Universitá di Bologna, Viale Berti Pichat 6/2, 40127 Bologna, Italy
27 Centre National d’Etudes Spatiales – Centre spatial de Toulouse, 18 avenue Edouard Belin, 31401 Toulouse Cedex 9, France
28 Institut national de physique nucléaire et de physique des particules, 3 rue Michel-Ange, 75794 Paris Cédex 16, France
29 Institute for Astronomy, University of Edinburgh, Royal Observatory, Blackford Hill, Edinburgh EH9 3HJ, UK
30 European Space Agency/ESRIN, Largo Galileo Galilei 1, 00044 Frascati, Roma, Italy
31 ESAC/ESA, Camino Bajo del Castillo, s/n., Urb. Villafranca del Castillo, 28692 Villanueva de la Cañada, Madrid, Spain
32 University of Lyon, Univ Claude Bernard Lyon 1, CNRS/IN2P3, IP2I Lyon, UMR 5822, 69622 Villeurbanne, France
33 Institute of Physics, Laboratory of Astrophysics, Ecole Polytechnique Fédérale de Lausanne (EPFL), Observatoire de Sauverny, 1290 Versoix, Switzerland
34 UCB Lyon 1, CNRS/IN2P3, IUF, IP2I Lyon, 4 rue Enrico Fermi, 69622 Villeurbanne, France
35 Departamento de Física, Faculdade de Ciências, Universidade de Lisboa, Edifício C8, Campo Grande, PT1749-016 Lisboa, Portugal
36 Instituto de Astrofísica e Ciências do Espaço, Faculdade de Ciências, Universidade de Lisboa, Campo Grande, 1749-016 Lisboa, Portugal
37 Department of Astronomy, University of Geneva, ch. d’Ecogia 16, 1290 Versoix, Switzerland
38 INFN-Padova, Via Marzolo 8, 35131 Padova, Italy
39 INAF-Istituto di Astrofisica e Planetologia Spaziali, via del Fosso del Cavaliere, 100, 00100 Roma, Italy
40 Université Paris-Saclay, Université Paris Cité, CEA, CNRS, AIM, 91191, Gif-sur-Yvette, France
41 Institute of Space Sciences (ICE, CSIC), Campus UAB, Carrer de Can Magrans, s/n, 08193 Barcelona, Spain
42 Institut d’Estudis Espacials de Catalunya (IEEC), Carrer Gran Capitá 2-4, 08034 Barcelona, Spain
43 INAF-Osservatorio Astronomico di Trieste, Via G. B. Tiepolo 11, 34143 Trieste, Italy
44 Aix-Marseille Université, CNRS/IN2P3, CPPM, Marseille, France
45 INAF-Osservatorio Astronomico di Padova, Via dell’Osservatorio 5, 35122 Padova, Italy
46 Max Planck Institute for Extraterrestrial Physics, Giessenbachstr. 1, 85748 Garching, Germany
47 University Observatory, Faculty of Physics, Ludwig-Maximilians-Universität, Scheinerstr. 1, 81679 Munich, Germany
48 Dipartimento di Fisica ”Aldo Pontremoli”, Universitá degli Studi di Milano, Via Celoria 16, 20133 Milano, Italy
49 INFN-Sezione di Milano, Via Celoria 16, 20133 Milano, Italy
50 Institute of Theoretical Astrophysics, University of Oslo, P.O. Box 1029 Blindern, 0315 Oslo, Norway
51 Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA, 91109, USA
52 Department of Physics, Lancaster University, Lancaster, LA1 4YB, UK
53 Technical University of Denmark, Elektrovej 327, 2800 Kgs. Lyngby, Denmark
54 Cosmic Dawn Center (DAWN), Denmark
55 Max-Planck-Institut für Astronomie, Königstuhl 17, 69117 Heidelberg, Germany
56 Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, UK
57 Department of Physics and Helsinki Institute of Physics, Gustaf Hällströmin katu 2, 00014 University of Helsinki, Finland
58 Department of Physics, P.O. Box 64, 00014 University of Helsinki, Finland
59 Helsinki Institute of Physics, Gustaf Hällströmin katu 2, University of Helsinki, Helsinki, Finland
60 NOVA optical infrared instrumentation group at ASTRON, Oude Hoogeveensedijk 4, 7991PD, Dwingeloo, The Netherlands
61 Universität Bonn, Argelander-Institut für Astronomie, Auf dem Hügel 71, 53121 Bonn, Germany
62 Aix-Marseille Université, CNRS, CNES, LAM, Marseille, France
63 Dipartimento di Fisica e Astronomia ”Augusto Righi” - Alma Mater Studiorum Universitá di Bologna, via Piero Gobetti 93/2, 40129 Bologna, Italy
64 Department of Physics, Institute for Computational Cosmology, Durham University, South Road, DH1 3LE, UK
65 Université Côte d’Azur, Observatoire de la Côte d’Azur, CNRS, Laboratoire Lagrange, Bd de l’Observatoire, CS 34229, 06304 Nice cedex 4, France
66 Université Paris Cité, CNRS, Astroparticule et Cosmologie, 75013 Paris, France
67 European Space Agency/ESTEC, Keplerlaan 1, 2201 AZ Noordwijk, The Netherlands
68 Department of Physics and Astronomy, University of Aarhus, Ny Munkegade 120, DK-8000 Aarhus C, Denmark
69 Centre for Astrophysics, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada
70 Department of Physics and Astronomy, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada
71 Perimeter Institute for Theoretical Physics, Waterloo, Ontario N2L 2Y5, Canada
72 Université Paris-Saclay, Université Paris Cité, CEA, CNRS, Astrophysique, Instrumentation et Modélisation Paris-Saclay, 91191 Gif-sur-Yvette, France
73 Space Science Data Center, Italian Space Agency, via del Politecnico snc, 00133 Roma, Italy
74 CEA Saclay, DFR/IRFU, Service d’Astrophysique, Bat. 709, 91191 Gif-sur-Yvette, France
75 Institute of Space Science, Str. Atomistilor, nr. 409 Măgurele, Ilfov, 077125, Romania
76 Dipartimento di Fisica e Astronomia ”G. Galilei”, Universitá di Padova, Via Marzolo 8, 35131 Padova, Italy
77 Universitäts-Sternwarte München, Fakultät für Physik, Ludwig-Maximilians-Universität München, Scheinerstrasse 1, 81679 München, Germany
78 Departamento de Física, FCFM, Universidad de Chile, Blanco Encalada 2008, Santiago, Chile
79 Satlantis, University Science Park, Sede Bld 48940, Leioa-Bilbao, Spain
80 Centro de Investigaciones Energéticas, Medioambientales y Tecnológicas (CIEMAT), Avenida Complutense 40, 28040 Madrid, Spain
81 Instituto de Astrofísica e Ciências do Espaço, Faculdade de Ciências, Universidade de Lisboa, Tapada da Ajuda, 1349-018 Lisboa, Portugal
82 Universidad Politécnica de Cartagena, Departamento de Electrónica y Tecnología de Computadoras, Plaza del Hospital 1, 30202 Cartagena, Spain
83 Kapteyn Astronomical Institute, University of Groningen, PO Box 800, 9700 AV Groningen, The Netherlands
84 INFN-Bologna, Via Irnerio 46, 40126 Bologna, Italy
85 Infrared Processing and Analysis Center, California Institute of Technology, Pasadena, CA 91125, USA
86 IFPU, Institute for Fundamental Physics of the Universe, via Beirut 2, 34151 Trieste, Italy
87 Instituto de Astrofísica de Canarias, Calle Vía Láctea s/n, 38204, San Cristóbal de La Laguna, Tenerife, Spain
88 University of Applied Sciences and Arts of Northwestern Switzerland, School of Engineering, 5210 Windisch, Switzerland
89 Institut d’Astrophysique de Paris, 98bis Boulevard Arago, 75014, Paris, France
90 Junia, EPA department, 41 Bd Vauban, 59800 Lille, France
91 Instituto de Física Teórica UAM-CSIC, Campus de Cantoblanco, 28049 Madrid, Spain
92 CERCA/ISO, Department of Physics, Case Western Reserve University, 10900 Euclid Avenue, Cleveland, OH 44106, USA
93 Laboratoire de Physique de l’École Normale Supérieure, ENS, Université PSL, CNRS, Sorbonne Université, 75005 Paris, France
94 Observatoire de Paris, Université PSL, Sorbonne Université, LERMA, 750 Paris, France
95 Astrophysics Group, Blackett Laboratory, Imperial College London, London SW7 2AZ, UK
96 Scuola Normale Superiore, Piazza dei Cavalieri 7, 56126 Pisa, Italy
97 SISSA, International School for Advanced Studies, Via Bonomea 265, 34136 Trieste TS, Italy
98 INFN, Sezione di Trieste, Via Valerio 2, 34127 Trieste TS, Italy
99 Departamento de Astrofísica, Universidad de La Laguna, 38206, La Laguna, Tenerife, Spain
100 Dipartimento di Fisica e Scienze della Terra, Universitá degli Studi di Ferrara, Via Giuseppe Saragat 1, 44122 Ferrara, Italy
101 Istituto Nazionale di Fisica Nucleare, Sezione di Ferrara, Via Giuseppe Saragat 1, 44122 Ferrara, Italy
102 Institut de Physique Théorique, CEA, CNRS, Université Paris-Saclay 91191 Gif-sur-Yvette Cedex, France
103 Institut d’Astrophysique de Paris, UMR 7095, CNRS, and Sorbonne Université, 98 bis boulevard Arago, 75014 Paris, France
104 Dipartimento di Fisica - Sezione di Astronomia, Universitá di Trieste, Via Tiepolo 11, 34131 Trieste, Italy
105 NASA Ames Research Center, Moffett Field, CA 94035, USA
106 Kavli Institute for Particle Astrophysics & Cosmology (KIPAC), Stanford University, Stanford, CA 94305, USA
107 INAF, Istituto di Radioastronomia, Via Piero Gobetti 101, 40129 Bologna, Italy
108 Institute Lorentz, Leiden University, PO Box 9506, Leiden 2300 RA, The Netherlands
109 Institute for Astronomy, University of Hawaii, 2680 Woodlawn Drive, Honolulu, HI 96822, USA
110 Department of Physics & Astronomy, University of California Irvine, Irvine CA 92697, USA
111 Departamento Física Aplicada, Universidad Politécnica de Cartagena, Campus Muralla del Mar, 30202 Cartagena, Murcia, Spain
112 Department of Astronomy & Physics and Institute for Computational Astrophysics, Saint Mary’s University, 923 Robie Street, Halifax, Nova Scotia, B3H 3C3, Canada
113 Dipartimento di Fisica, Universitá degli studi di Genova, and INFN-Sezione di Genova, via Dodecaneso 33, 16146, Genova, Italy
114 Department of Computer Science, Aalto University, PO Box 15400, Espoo, FI-00 076, Finland
115 Ruhr University Bochum, Faculty of Physics and Astronomy, Astronomical Institute (AIRUB), German Centre for Cosmological Lensing (GCCL), 44780 Bochum, Germany
116 Université Paris-Saclay, CNRS/IN2P3, IJCLab, 91405 Orsay, France
117 Department of Physics and Astronomy, Vesilinnantie 5, 20014 University of Turku, Finland
118 Serco for European Space Agency (ESA), Camino bajo del Castillo, s/n, Urbanizacion Villafranca del Castillo, Villanueva de la Cañada, 28692 Madrid, Spain
119 AIM, CEA, CNRS, Université Paris-Saclay, Université de Paris, 91191 Gif-sur-Yvette, France
120 Oskar Klein Centre for Cosmoparticle Physics, Department of Physics, Stockholm University, Stockholm, SE-106 91, Sweden
121 Univ. Grenoble Alpes, CNRS, Grenoble INP, LPSC-IN2P3, 53, Avenue des Martyrs, 38000, Grenoble, France
122 Centre de Calcul de l’IN2P3/CNRS, 21 avenue Pierre de Coubertin 69627 Villeurbanne Cedex, France
123 Dipartimento di Fisica, Sapienza Università di Roma, Piazzale Aldo Moro 2, 00185 Roma, Italy
124 Centro de Astrofísica da Universidade do Porto, Rua das Estrelas, 4150-762 Porto, Portugal
125 Dipartimento di Fisica, Università di Roma Tor Vergata, Via della Ricerca Scientifica 1, Roma, Italy
126 INFN, Sezione di Roma 2, Via della Ricerca Scientifica 1, Roma, Italy
127 Department of Mathematics and Physics E. De Giorgi, University of Salento, Via per Arnesano, CP-I93, 73100, Lecce, Italy
128 INAF-Sezione di Lecce, c/o Dipartimento Matematica e Fisica, Via per Arnesano, 73100, Lecce, Italy
129 INFN, Sezione di Lecce, Via per Arnesano, CP-193, 73100, Lecce, Italy
130 Institute for Computational Science, University of Zurich, Winterthurerstrasse 190, 8057 Zurich, Switzerland
131 Institut für Theoretische Physik, University of Heidelberg, Philosophenweg 16, 69120 Heidelberg, Germany
132 Université St Joseph; Faculty of Sciences, Beirut, Lebanon
133 Mullard Space Science Laboratory, University College London, Holmbury St Mary, Dorking, Surrey RH5 6NT, UK
134 Department of Astrophysical Sciences, Peyton Hall, Princeton University, Princeton, NJ 08544, USA
135 Niels Bohr Institute, University of Copenhagen, Jagtvej 128, 2200 Copenhagen, Denmark
136 Cosmic Dawn Center (DAWN)
137 Universität Innsbruck, Institut für Astro- und Teilchenphysik, Technikerstr. 25/8, 6020 Innsbruck, Austria

Euclid preparation. XXXIV. The effect of linear redshift-space distortions in photometric galaxy clustering and its cross-correlation with cosmic shear

Euclid Collaboration: K. Tanidis1,2 [email protected]    V. F. Cardone3,4    M. Martinelli3,4    I. Tutusaus5,6    S. Camera7,8,9    N. Aghanim10    A. Amara11    S. Andreon12    N. Auricchio13    M. Baldi14,13,15    S. Bardelli13    E. Branchini16,17    M. Brescia18,19    J. Brinchmann20    V. Capobianco9    C. Carbone21    J. Carretero22,23    S. Casas24    M. Castellano3    S. Cavuoti19,25    A. Cimatti26    R. Cledassou27,28 Deceased    G. Congedo29    L. Conversi30,31    Y. Copin32    L. Corcione9    F. Courbin33    H. M. Courtois34    A. Da Silva35,36    H. Degaudenzi37    J. Dinis36,35    F. Dubath37    X. Dupac31    S. Dusini38    M. Farina39    S. Farrens40    S. Ferriol32    P. Fosalba41,42    M. Frailis43    E. Franceschi13    M. Fumana21    S. Galeotta43    B. Garilli21    W. Gillard44    B. Gillis29    C. Giocoli13,15    A. Grazian45    F. Grupp46,47    L. Guzzo48,12,49    S. V. H. Haugan50    W. Holmes51    I. Hook52    A. Hornstrup53,54    K. Jahnke55    B. Joachimi56    E. Keihanen57    S. Kermiche44    A. Kiessling51    M. Kunz6    H. Kurki-Suonio58,59    P. B. Lilje50    V. Lindholm58,59    I. Lloro60    E. Maiorano13    O. Mansutti43    O. Marggraf61    K. Markovic51    N. Martinet62    F. Marulli63,13,15    R. Massey64    S. Maurogordato65    E. Medinaceli13    S. Mei66    M. Meneghetti13,15    G. Meylan33    M. Moresco63,13    L. Moscardini63,13,15    E. Munari43    S.-M. Niemi67    C. Padilla22    S. Paltani37    F. Pasian43    K. Pedersen68    W. J. Percival69,70,71    V. Pettorino72    S. Pires40    G. Polenta73    J. E. Pollack74,66    M. Poncet27    L. A. Popa75    F. Raison46    A. Renzi76,38    J. Rhodes51    G. Riccio19    E. Romelli43    M. Roncarelli13    E. Rossetti14    R. Saglia77,46    D. Sapone78    B. Sartoris77,43    M. Schirmer55    P. Schneider61    A. Secroun44    G. Seidel55    S. Serrano42,41,79    C. Sirignano76,38    G. Sirri15    L. Stanco38    P. Tallada-Crespí80,23    A. N. Taylor29    I. Tereno35,81    R. Toledo-Moreo82    F. Torradeflot23,80    E. A. Valentijn83    L. Valenziano13,84    T. Vassallo77,43    A. Veropalumbo12    Y. Wang85    J. Weller77,46    G. Zamorani13    J. Zoubian44    E. Zucca13    A. Biviano43,86    A. Boucaud66    E. Bozzo37    C. Colodro-Conde87    D. Di Ferdinando15    R. Farinelli13    J. Graciá-Carpio46    S. Marcin88    N. Mauri26,15    V. Scottez89,90    M. Tenti84    A. Tramacere37    Y. Akrami91,92,93,94,95    V. Allevato19,96    C. Baccigalupi97,43,98    A. Balaguera-Antolínez87,99    M. Ballardini100,101,13    D. Benielli44    F. Bernardeau102,103    S. Borgani43,104,98,86    A. S. Borlaff105,106    C. Burigana107,84    R. Cabanac5    A. Cappi13,65    C. S. Carvalho81    G. Castignani63,13    T. Castro43,98,86    G. Cañas-Herrera67,108    K. C. Chambers109    A. R. Cooray110    J. Coupon37    A. Díaz-Sánchez111    S. Davini17    S. de la Torre62    G. De Lucia43    G. Desprez112    S. Di Domizio113    H. Dole10    J. A. Escartin Vigo46    S. Escoffier44    P. G. Ferreira2    I. Ferrero50    F. Finelli13,84    L. Gabarra76,38    J. García-Bellido91    E. Gaztanaga41,42,11    F. Giacomini15    G. Gozaliasl58,114    H. Hildebrandt115    S. Ilić116,27,5    J. J. E. Kajava117,118    V. Kansal119    C. C. Kirkpatrick57    L. Legrand6    A. Loureiro120,95    J. Macias-Perez121    M. Magliocchetti39    G. Mainetti122    R. Maoli123,3    C. J. A. P. Martins124,20    S. Matthew29    L. Maurin10    R. B. Metcalf63    M. Migliaccio125,126    P. Monaco104,43,98,86    G. Morgante13    S. Nadathur11    A. A. Nucita127,128,129    M. Pöntinen58    L. Patrizii15    A. Pezzotta46    V. Popa75    D. Potter130    A. G. Sánchez46    Z. Sakr131,5,132    J. A. Schewtschenko29    A. Schneider130    M. Sereno13,15    P. Simon61    A. Spurio Mancini133    J. Steinwagner46    M. Tewes61    R. Teyssier134    S. Toft54,135,136    J. Valiviita58,59    M. Viel86,43,97,98    L. Linke137
(May 1, 2024)
Abstract

Context. The cosmological surveys that are planned for the current decade will provide us with unparalleled observations of the distribution of galaxies on cosmic scales, by means of which we can probe the underlying large-scale structure (LSS) of the Universe. This will allow us to test the concordance cosmological model and its extensions. However, precision pushes us to high levels of accuracy in the theoretical modelling of the LSS observables, so that no biases are introduced into the estimation of the cosmological parameters. In particular, effects such as redshift-space distortions (RSD) can become relevant in the computation of harmonic-space power spectra even for the clustering of the photometrically selected galaxies, as has previously been shown in literature.

Aims. In this work, we investigate the contribution of linear RSD, as formulated in the Limber approximation by a previous work, in forecast cosmological analyses with the photometric galaxy sample of the Euclid survey. We aim to assess their impact and to quantify the bias on the measurement of cosmological parameters that would be caused if this effect were neglected.

Methods. We performed this task by producing mock power spectra for photometric galaxy clustering and weak lensing, as is expected to be obtained from the Euclid survey. We then used a Markov chain Monte Carlo approach to obtain the posterior distributions of cosmological parameters from these simulated observations.

Results. When the linear RSD is neglected, significant biases are caused when galaxy correlations are used alone and when they are combined with cosmic shear in the so-called 3×\times×2pt approach. These biases can be equivalent to as much as 5σ5𝜎5\,\sigma5 italic_σ when an underlying ΛΛ\Lambdaroman_ΛCDM cosmology is assumed. When the cosmological model is extended to include the equation-of-state parameters of dark energy, the extension parameters can be shifted by more than 1σ1𝜎1\,\sigma1 italic_σ.

Key Words.:
Cosmology: theory – large-scale structure of the Universe – cosmological parameters

1 Introduction

Within the current decade, several large-scale structure (LSS) surveys are expected to start their operations or to release their first results. They will provide a significant improvement to available cosmological data sets. These forthcoming LSS surveys will map the matter distribution in the Universe with exquisite precision. Some of the surveys will be ground-based, such as the Dark Energy Spectroscopic Instrument (DESI, DESI Collaboration: Aghamousa et al., 2016a, b), the Legacy Survey of Space and Time (LSST) at the Vera C. Rubin Observatory (LSST Science Collaboration: Abell et al., 2009; LSST Dark Energy Science Collaboration: Mandelbaum et al., 2018; Ivezić et al., 2019), and the Square Kilometre Array Observatory (SKAO; see, e.g., Abdalla et al., 2015; Santos et al., 2015; Brown et al., 2015; Bull et al., 2015; Camera et al., 2015b; Raccanelli et al., 2015; SKA Cosmology Science Working Group, 2020). Others will be space-borne, such as the Euclid satellite (Laureijs et al., 2011; Amendola et al., 2013, 2018; Euclid Collaboration: Blanchard et al., 2020), the Nancy Grace Roman Space Telescope (Spergel et al., 2015), and the Spectro-Photometer for the History of the Universe, Epoch of Reionization, and Ices Explorer (SPHEREx; see, e.g., Doré et al., 2014, 2018).

All these surveys rely on the observation of galaxy positions and shapes, with which summary statistics can be constructed that are customarily referred to in cosmology as galaxy clustering (GC) and weak lensing (WL). This can be done in multiple ways, using either the real-space 2pt correlation function, or the harmonic-space power spectrum that we study here, or even via other statistics such as COSEBIs (Schneider et al., 2010), or higher-order statistics (Euclid Collaboration: Ajani et al., 2023). The GC encodes information on the clustering of matter due to the effect of gravity, while the WL provides information on the projected matter distribution through its gravitational lensing effect.

In this work, we focus on Euclid and its surveys.111http://www.euclid-ec.org/. Euclid is a European Space Agency medium-class space mission whose launch took place on 1 July 2023. It will perform photometric and spectroscopic galaxy surveys over an area of 15000deg2similar-toabsent15000superscriptdegree2\sim 15000\,\deg^{2}∼ 15000 roman_deg start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the extragalactic sky (Laureijs et al., 2011), with the near-infrared instrument (Costille et al., 2018) and the visible imager (Cropper et al., 2018), which will be carried on board. The photometric survey will measure the positions and shapes of over a billion galaxies, enabling the analysis of photometric GC (GCph) and WL. Because the photometric measurements will provide relatively uncertain redshift measurements (compared to spectroscopic observations), the analyses of these observables will be performed via a tomographic approach by binning galaxies in redshift slices and considering the projected two-dimensional data sets. The precise radial measurements of the spectroscopic survey will instead allow us to perform a spectroscopic GC (GCsp) analysis, that is, a galaxy-clustering analysis in three dimensions.

In Euclid Collaboration: Blanchard et al. (2020, ‘Euclid preparation: VII’, hereafter EP:VII), the constraints expected from Euclid have been forecast for the individual GCsp, GCph, and WL probes and also for their combination. To obtain these results, EP:VII used several assumptions to simplify the theoretical computation of observables: the Limber approximation was used for all the photometric observables, and it was assumed for GCph that the only non-negligible contribution to the galaxy position correlation function comes from the anisotropies in the density field. It is known, however, that several other effects contribute to GCph, including lensing magnification, velocity, and relativistic effects (Yoo, 2010; Challinor & Lewis, 2011; Bonvin & Durrer, 2011).

These various contributions can be significant at very large scales, which we define as scales corresponding to a wavenumber smaller than that at which the matter power spectrum peaks at the matter-radiation equality scale. These scales are effectively within reach of wide surveys such as Euclid, and it has been shown that neglecting them could lead to inaccurate results. The cosmology that is recovered through parameter estimation pipelines might be significantly biased with respect to the true underlying cosmology (Camera et al., 2015a; Tanidis & Camera, 2019; Tanidis et al., 2019; Martinelli et al., 2022; Lepori et al., 2022).

In this work, we focus on one of the most important effects, namely redshift-space distortion (RSD). We also aim specifically to quantify its impact on the expected results of the Euclid wide survey. The effect of the linear RSD on the angular clustering is not new and has been thoroughly studied before (Scharf et al., 1994; Heavens & Taylor, 1995; Padmanabhan et al., 2007; Blake et al., 2007; Nock et al., 2010; Crocce et al., 2011; Balaguera-Antolínez et al., 2018; Tanidis & Camera, 2019). The effects have also been included in the Dark Energy Survey (DES, DES Collaboration: Abbott et al., 2005) 3×\times×2pt data analysis for Y3 in the configuration space (Abbott et al., 2022), although they were initially neglected in the Y1 analysis for the GC in the configuration and harmonic space (Elvin-Poole et al., 2018; Andrade-Oliveira et al., 2021). Here, we apply the approach of Tanidis & Camera (2019), where the linear RSD contribution to GC harmonic-space power spectra is obtained within the Limber approximation. We examine this contribution to the Euclid wide survey. This allows us to compute the theoretical prediction at a reasonable speed so that it can be used to estimate the parameters. Moreover, we only focus on GCph and do not discuss GCsp at all. We therefore always refer to the GCph probe simply as GC throughout.

The paper is organised as follows. The equations we used to compute the theoretical predictions for the observables of interest are reviewed in Sect. 2, where we also outline how RSD enters the calculations. In Sect. 2.4 we summarise the results of the Flagship simulation galaxy catalogue. Our analysis method is shown in Sect. 3 and the results we obtained are provided in Sect. 4. We finally summarise our conclusions in Sect. 5.

2 Photometric observables in Euclid

2.1 Harmonic-space power spectra

The harmonic-space power spectrum CAB(zi,zj)superscriptsubscript𝐶𝐴𝐵subscript𝑧𝑖subscript𝑧𝑗C_{\ell}^{AB}(z_{i},z_{j})italic_C start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT ( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) between an observable A𝐴Aitalic_A in the redshift bin i𝑖iitalic_i and an observable B𝐵Bitalic_B in the redshift bin j𝑗jitalic_j is defined as

Ai,mBj,m=CAB(zi,zj)δKδmmK,delimited-⟨⟩subscript𝐴𝑖𝑚subscriptsuperscript𝐵𝑗superscriptsuperscript𝑚subscriptsuperscript𝐶𝐴𝐵subscript𝑧𝑖subscript𝑧𝑗subscriptsuperscript𝛿Ksuperscriptsubscriptsuperscript𝛿K𝑚superscript𝑚\left\langle A_{i,\ell m}\,B^{\ast}_{j,\ell^{\prime}m^{\prime}}\right\rangle=C% ^{AB}_{\ell}(z_{i},z_{j})\,\delta^{\rm K}_{\ell\ell^{\prime}}\,\delta^{\rm K}_% {mm^{\prime}}\;,⟨ italic_A start_POSTSUBSCRIPT italic_i , roman_ℓ italic_m end_POSTSUBSCRIPT italic_B start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j , roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT ⟩ = italic_C start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) italic_δ start_POSTSUPERSCRIPT roman_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_ℓ roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_δ start_POSTSUPERSCRIPT roman_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_m start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (1)

where Xmsubscript𝑋𝑚X_{\ell m}italic_X start_POSTSUBSCRIPT roman_ℓ italic_m end_POSTSUBSCRIPT are the coefficients of the harmonic expansion of observable X𝑋Xitalic_X, and δKsuperscript𝛿K\delta^{\rm K}italic_δ start_POSTSUPERSCRIPT roman_K end_POSTSUPERSCRIPT denotes the Kronecker symbol. Here, the letters A𝐴Aitalic_A and B𝐵Bitalic_B stand for the observables of interest: galaxy number count fluctuations, ΔΔ\varDeltaroman_Δ, or galaxy ellipticities, ϵitalic-ϵ\epsilonitalic_ϵ. Both fields are discussed in more detail in Sect. 2.2 and Sect. 2.3, respectively.

Within the Limber approximation (Kaiser, 1992), which is valid at 1much-greater-than1\ell\gg 1roman_ℓ ≫ 1 and for broad redshift kernels, the harmonic-space (also called angular) power spectrum between two observables A𝐴Aitalic_A and B𝐵Bitalic_B is

CijAB()=drr2WiA(,r)WjB(,r)Pδδ(k=+1/2r,r),subscriptsuperscript𝐶𝐴𝐵𝑖𝑗d𝑟superscript𝑟2superscriptsubscript𝑊𝑖𝐴𝑟superscriptsubscript𝑊𝑗𝐵𝑟subscript𝑃𝛿𝛿𝑘12𝑟𝑟C^{AB}_{ij}(\ell)=\int\frac{\mathrm{d}r}{r^{2}}\,W_{i}^{A}(\ell,r)\,W_{j}^{B}(% \ell,r)\,P_{\delta\delta}\left(k=\frac{\ell+1/2}{r},r\right)\;,italic_C start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) = ∫ divide start_ARG roman_d italic_r end_ARG start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( roman_ℓ , italic_r ) italic_W start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( roman_ℓ , italic_r ) italic_P start_POSTSUBSCRIPT italic_δ italic_δ end_POSTSUBSCRIPT ( italic_k = divide start_ARG roman_ℓ + 1 / 2 end_ARG start_ARG italic_r end_ARG , italic_r ) , (2)

where Pδδsubscript𝑃𝛿𝛿P_{\delta\delta}italic_P start_POSTSUBSCRIPT italic_δ italic_δ end_POSTSUBSCRIPT is the (non-linear) matter power spectrum, k=|𝒌|𝑘𝒌k=|\boldsymbol{k}|italic_k = | bold_italic_k | is the wave number, which is the Fourier mode related to the comoving separation between pairs of galaxies in configuration space, and r(z)𝑟𝑧r(z)italic_r ( italic_z ) is the radial comoving distance to redshift z𝑧zitalic_z for a flat cosmology. The generic redshift-binned kernel WiA(,r)superscriptsubscript𝑊𝑖𝐴𝑟W_{i}^{A}(\ell,r)italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( roman_ℓ , italic_r ) takes different forms depending on the target observables, as we show in Sect. 2.2 and Sect. 2.3. We use the notation CijAB()subscriptsuperscript𝐶𝐴𝐵𝑖𝑗C^{AB}_{ij}(\ell)italic_C start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) as in EP:VII, as opposed to CAB(zi,zj)subscriptsuperscript𝐶𝐴𝐵subscript𝑧𝑖subscript𝑧𝑗C^{AB}_{\ell}(z_{i},z_{j})italic_C start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_z start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) of Eq. 1, to denote the fact that we refer to the Limber-approximated power spectrum.

2.2 Linear RSD in GCph

In GC, galaxies are biased tracers of the underlying matter field (Kaiser, 1987). At sufficiently large scales, the bias can be considered to only depend on redshift and not on scale (Abbott et al., 2018). On the other hand, when non-linear scales are added, the galaxy bias becomes non-local and a specific treatment is required to account for this effect (Sánchez et al., 2016; Desjacques et al., 2018). In addition, RSDs are additional observational effects due to the peculiar velocities of galaxies (Kaiser, 1987; Szalay et al., 1998). These are customarily split into linear RSD (also known as the Kaiser effect) and non-linear RSD (also known as the fingers of God, ‘FoG’ hereafter). The former causes the squashing of the galaxy 2pt correlation function on large scales in the direction perpendicular to the line of sight, while the latter enhances the clustering amplitude along the line-of-sight direction on small scales. In the current analysis we consider a simple model, that assumes that the galaxy bias is linear and scale independent, and we only account for linear RSD. The modelling of the non-linear galaxy bias and the FoG as well as their effects in the photometric observables for Euclid is left for future work. The kernel of Eq. 2 for the GC that includes up to linear RSD in the Limber approximation (for details, see Tanidis & Camera, 2019) takes the form

WiΔ(,r)=Widen(r)+WiRSD(,r),superscriptsubscript𝑊𝑖Δ𝑟superscriptsubscript𝑊𝑖den𝑟superscriptsubscript𝑊𝑖RSD𝑟W_{i}^{\varDelta}(\ell,r)=W_{i}^{\rm den}(r)+W_{i}^{\rm RSD}(\ell,r)\;,italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_Δ end_POSTSUPERSCRIPT ( roman_ℓ , italic_r ) = italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_den end_POSTSUPERSCRIPT ( italic_r ) + italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_RSD end_POSTSUPERSCRIPT ( roman_ℓ , italic_r ) , (3)

with the first term being due to fluctuations in the density field,

Widen(r)=ni(r)bi(r),superscriptsubscript𝑊𝑖den𝑟subscript𝑛𝑖𝑟subscript𝑏𝑖𝑟W_{i}^{\rm den}(r)=n_{i}(r)\,b_{i}(r)\;,italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_den end_POSTSUPERSCRIPT ( italic_r ) = italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) , (4)

with ni(r)drsubscript𝑛𝑖𝑟d𝑟n_{i}(r)\,\mathrm{d}ritalic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) roman_d italic_r the galaxy probability density in bin i𝑖iitalic_i between the comoving distance r𝑟ritalic_r and r+dr𝑟d𝑟r+\mathrm{d}ritalic_r + roman_d italic_r, and bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT being the corresponding linear galaxy bias, computed at rr(z)𝑟𝑟𝑧r\equiv r(z)italic_r ≡ italic_r ( italic_z ). We treat this as a constant within the redshift bin, and its actual amplitude is a nuisance parameter, over which we marginalised in our analysis. For the fiducial galaxy bias values in each bin (see Table 2), we used the fiducial model of Euclid Collaboration: Pocino et al. (2021, ‘Euclid Colaboration: XII’, hereafter EP:XII), who considered a magnitude cut at IE=24.5subscript𝐼E24.5I_{\scriptscriptstyle\rm E}=24.5italic_I start_POSTSUBSCRIPT roman_E end_POSTSUBSCRIPT = 24.5 for the Euclid imager, which will observe through an optical broad band. For the galaxy distributions ni(r)subscript𝑛𝑖𝑟n_{i}(r)italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ), we used the outcome of the Flagship galaxy simulation of the Euclid Consortium, for which we provide details in Sect. 2.4.

The second term in Eq. 3 is the RSD contribution (Tanidis & Camera, 2019),

WiRSD(,r)=n=11Ln()ni(2+1+4n2+1r)f(2+1+4n2+1r),superscriptsubscript𝑊𝑖RSD𝑟superscriptsubscript𝑛11subscript𝐿𝑛subscript𝑛𝑖214𝑛21𝑟𝑓214𝑛21𝑟W_{i}^{\rm RSD}(\ell,r)=\sum_{n=-1}^{1}L_{n}(\ell)\,n_{i}\left(\frac{2\,\ell+1% +4\,n}{2\,\ell+1}\,r\right)\,f\left(\frac{2\,\ell+1+4\,n}{2\,\ell+1}\,r\right)\;,italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_RSD end_POSTSUPERSCRIPT ( roman_ℓ , italic_r ) = ∑ start_POSTSUBSCRIPT italic_n = - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( roman_ℓ ) italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( divide start_ARG 2 roman_ℓ + 1 + 4 italic_n end_ARG start_ARG 2 roman_ℓ + 1 end_ARG italic_r ) italic_f ( divide start_ARG 2 roman_ℓ + 1 + 4 italic_n end_ARG start_ARG 2 roman_ℓ + 1 end_ARG italic_r ) , (5)

with

L0()subscript𝐿0\displaystyle L_{0}(\ell)italic_L start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( roman_ℓ ) =22+21(21)(2+3),absent2superscript2212123\displaystyle=\frac{2\,\ell^{2}+2\,\ell-1}{(2\,\ell-1)\,(2\,\ell+3)}\;,= divide start_ARG 2 roman_ℓ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 roman_ℓ - 1 end_ARG start_ARG ( 2 roman_ℓ - 1 ) ( 2 roman_ℓ + 3 ) end_ARG , (6)
L1()subscript𝐿1\displaystyle L_{-1}(\ell)italic_L start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ( roman_ℓ ) =(1)(21)(23)(2+1),absent1212321\displaystyle=-\frac{\ell\,(\ell-1)}{(2\,\ell-1)\sqrt{(2\,\ell-3)\,(2\,\ell+1)% }}\;,= - divide start_ARG roman_ℓ ( roman_ℓ - 1 ) end_ARG start_ARG ( 2 roman_ℓ - 1 ) square-root start_ARG ( 2 roman_ℓ - 3 ) ( 2 roman_ℓ + 1 ) end_ARG end_ARG , (7)
L+1()subscript𝐿1\displaystyle L_{+1}(\ell)italic_L start_POSTSUBSCRIPT + 1 end_POSTSUBSCRIPT ( roman_ℓ ) =(+1)(+2)(2+3)(2+1)(2+5).absent12232125\displaystyle=-\frac{(\ell+1)\,(\ell+2)}{(2\,\ell+3)\sqrt{(2\,\ell+1)\,(2\,% \ell+5)}}\;.= - divide start_ARG ( roman_ℓ + 1 ) ( roman_ℓ + 2 ) end_ARG start_ARG ( 2 roman_ℓ + 3 ) square-root start_ARG ( 2 roman_ℓ + 1 ) ( 2 roman_ℓ + 5 ) end_ARG end_ARG . (8)

In Eq. 5, f(1+z)dlnD/dz𝑓1𝑧d𝐷d𝑧f\coloneqq-(1+z)\,\mathrm{d}\ln D/\mathrm{d}zitalic_f ≔ - ( 1 + italic_z ) roman_d roman_ln italic_D / roman_d italic_z is the growth rate, with D𝐷Ditalic_D being the linear growth factor. In the top panel of Fig. 1, we quantify the signal loss when the linear RSD are not included in the photometric GC harmonic-space power spectrum. In particular, RSD are important at large scales, for approximately 100less-than-or-similar-to100\ell\lesssim 100roman_ℓ ≲ 100. Additionally, the contribution to the total signal increases with redshift for a given redshift bin width. For example, RSD contribute to the total signal from 2–3% at low redshift (solid purple curve, bin pair 1–1) and gradually increase with increasing redshift (dash-dotted yellow curve, bin pair 6–6) up to 40%similar-toabsentpercent40\sim 40\%∼ 40 % at the lowest available multipoles. However, we should note that this is not true for the highest redshift bin (dotted blue curve, bin pair 13–13). This bin contributes less to the full signal than the bin pairs 10–10 and 6–6, for example, Tanidis et al. (2019) showed that the linear RSD effect is gradually diluted when the width of the redshift bin increases. This is particularly the case for the highest-redshift bins due to the large photometric uncertainties at high redshifts.

Refer to caption
Figure 1: Ratio of the harmonic-space power spectrum for the density fluctuations alone with respect to that including RSD for some tomographic auto-bin correlations (i=j𝑖𝑗i=jitalic_i = italic_j). Top: GC. Bottom: XC.

Another important correction to the galaxy density field is the magnification bias. We wish to quantify the impact of linear RSD alone on Euclid photometric observables here and therefore neglected the magnification effect in our analysis. However, this effect has been thoroughly studied in Lepori et al. (2022) (in that study the linear RSD were neglected), and its inclusion was found to be crucial to avoid biases on the cosmological parameter estimation for Euclid. Similar studies of this effect have also been conducted for other future experiments (Tanidis et al., 2019).

In addition, there are also local and integrated contributions to the signal that are measurable at ultra-large scales, such as the Doppler terms, the Sachs-Wolfe and the integrated Sachs-Wolfe effects, and the time delay. We neglected these contributions in our analysis because for 1much-greater-than1\ell\gg 1roman_ℓ ≫ 1, where the Limber approximation holds, their effect is negligible (Yoo, 2010; Challinor & Lewis, 2011; Bonvin & Durrer, 2011; Martinelli et al., 2022).

2.3 Cosmic shear

The LSS of the Universe deflects the paths of photons that are emitted by distant sources. This distorts the source images. This distortion is decomposed into the convergence, κ𝜅\kappaitalic_κ, and shear, γ𝛾\gammaitalic_γ, which correspond to size magnification and shape distortion of the images, respectively, and are related linearly. Both signals contain useful cosmological information, but the former is more impossible to extract because it requires knowledge of the original source sizes (Heavens et al., 2013; Alsing et al., 2015). For this reason, shear is the usual focus of WL surveys of the LSS.

The harmonic-space power spectrum of the shear field is a probe of the growth of structures and the cosmological expansion. In addition to the cosmic shear, the correlation of the galaxy shapes also receives a contribution from intrinsic alignments (hereafter IA), which in the context of cosmological WL studies is regarded as a systematic effect. This accounts for the fact that, in addition to the random orientations of galaxies, which ideally would make the ellipticity an unbiased estimator of the shear field, there are also IA of the galaxies that are caused by the tidal interactions during galaxy formation, and also astrophysical effects that contaminate the WL analysis (Joachimi et al., 2015).

For the WL sample, the kernel of the observed ellipticity power spectrum including cosmic shear γ𝛾\gammaitalic_γ and IA, reads

Wiϵ(,r)=Wiγ(,r)+WiIA(r).subscriptsuperscript𝑊italic-ϵ𝑖𝑟subscriptsuperscript𝑊𝛾𝑖𝑟subscriptsuperscript𝑊IA𝑖𝑟W^{\epsilon}_{i}(\ell,r)=W^{\gamma}_{i}(\ell,r)+W^{\rm IA}_{i}(r)\;.italic_W start_POSTSUPERSCRIPT italic_ϵ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( roman_ℓ , italic_r ) = italic_W start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( roman_ℓ , italic_r ) + italic_W start_POSTSUPERSCRIPT roman_IA end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) . (9)

The shear contribution is

Wiγ(,r)=3Lγ()Ωm,0H022c2[1+z(r)]rqi(r),subscriptsuperscript𝑊𝛾𝑖𝑟3subscript𝐿𝛾subscriptΩm0superscriptsubscript𝐻022superscript𝑐2delimited-[]1𝑧𝑟𝑟subscript𝑞𝑖𝑟W^{\gamma}_{i}(\ell,r)=\frac{3\,L_{\gamma}(\ell)\,\varOmega_{{\rm m},0}\,H_{0}% ^{2}}{2c^{2}}\left[1+z(r)\right]\,r\,q_{i}(r)\;,italic_W start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( roman_ℓ , italic_r ) = divide start_ARG 3 italic_L start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ( roman_ℓ ) roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG [ 1 + italic_z ( italic_r ) ] italic_r italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) , (10)

where c𝑐citalic_c is the vacuum speed of light, qi(r)subscript𝑞𝑖𝑟q_{i}(r)italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) is the so-called lensing efficiency for a flat Universe,

qi(r)=rdrrrrni(r),subscript𝑞𝑖𝑟superscriptsubscript𝑟differential-dsuperscript𝑟superscript𝑟𝑟superscript𝑟subscript𝑛𝑖superscript𝑟q_{i}(r)=\int_{r}^{\infty}{\rm d}r^{\prime}\,\frac{r^{\prime}-r}{r^{\prime}}\,% n_{i}(r^{\prime})\;,italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) = ∫ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT roman_d italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT divide start_ARG italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT - italic_r end_ARG start_ARG italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , (11)

and the \ellroman_ℓ-dependent factor is given by

Lγ()=(+2)!(2)!(22+1)2.subscript𝐿𝛾22superscript2212L_{\gamma}(\ell)=\sqrt{\frac{(\ell+2)!}{(\ell-2)!}}\left(\frac{2}{2\,\ell+1}% \right)^{2}\;.italic_L start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ( roman_ℓ ) = square-root start_ARG divide start_ARG ( roman_ℓ + 2 ) ! end_ARG start_ARG ( roman_ℓ - 2 ) ! end_ARG end_ARG ( divide start_ARG 2 end_ARG start_ARG 2 roman_ℓ + 1 end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (12)

For WL we also used the ni(r)subscript𝑛𝑖𝑟n_{i}(r)italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) obtained from the Flagship simulation of Sect. 2.4. Because we restricted our analysis to multipoles, for which the Limber approximation applies (1much-greater-than1\ell\gg 1roman_ℓ ≫ 1), this factor can be considered to be Lγ()1subscript𝐿𝛾1L_{\gamma}(\ell)\approx 1italic_L start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT ( roman_ℓ ) ≈ 1.

The IA contribution can instead be modelled as in EP:VII, namely

WiIA(r)=𝒜IA𝒞IAΩm,0IA[z(r)]D[z(r)]ni(r),subscriptsuperscript𝑊IA𝑖𝑟subscript𝒜IAsubscript𝒞IAsubscriptΩm0subscriptIAdelimited-[]𝑧𝑟𝐷delimited-[]𝑧𝑟subscript𝑛𝑖𝑟W^{\rm IA}_{i}(r)=-\mathcal{A}_{\rm IA}\,\mathcal{C}_{\rm IA}\,\varOmega_{{\rm m% },0}\,\frac{\mathcal{F}_{\rm IA}\left[z(r)\right]}{D\left[z(r)\right]}\,n_{i}(% r)\;,italic_W start_POSTSUPERSCRIPT roman_IA end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) = - caligraphic_A start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT caligraphic_C start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT divide start_ARG caligraphic_F start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT [ italic_z ( italic_r ) ] end_ARG start_ARG italic_D [ italic_z ( italic_r ) ] end_ARG italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_r ) , (13)

where

IA(z)=(1+z)ηIA[L(z)L(z)]βIA.subscriptIA𝑧superscript1𝑧subscript𝜂IAsuperscriptdelimited-[]delimited-⟨⟩𝐿𝑧subscript𝐿𝑧subscript𝛽IA\mathcal{F}_{\rm IA}(z)=(1+z)^{\eta_{\rm IA}}\,\left[\frac{\langle L\rangle(z)% }{L_{\ast}(z)}\right]^{\beta_{\rm IA}}\;.caligraphic_F start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT ( italic_z ) = ( 1 + italic_z ) start_POSTSUPERSCRIPT italic_η start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT end_POSTSUPERSCRIPT [ divide start_ARG ⟨ italic_L ⟩ ( italic_z ) end_ARG start_ARG italic_L start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ( italic_z ) end_ARG ] start_POSTSUPERSCRIPT italic_β start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . (14)

The amplitude and shape of the IA signal is captured by the nuisance parameters 𝒜IAsubscript𝒜IA\mathcal{A}_{\rm IA}caligraphic_A start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT, βIAsubscript𝛽IA\beta_{\rm IA}italic_β start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT and ηIAsubscript𝜂IA\eta_{\rm IA}italic_η start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT (see Table 2 for their fiducial values), with CIAsubscript𝐶IAC_{\rm IA}italic_C start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT kept fixed at the value 0.01340.01340.01340.0134 because it is degenerate with 𝒜IAsubscript𝒜IA\mathcal{A}_{\rm IA}caligraphic_A start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT. The terms L(z)delimited-⟨⟩𝐿𝑧\langle L\rangle(z)⟨ italic_L ⟩ ( italic_z ) and L(z)subscript𝐿𝑧L_{\ast}(z)italic_L start_POSTSUBSCRIPT ∗ end_POSTSUBSCRIPT ( italic_z ) denote the mean and the characteristic luminosity of the source galaxies with respect to redshift.222We refer the reader to EP:VII for more details concerning the IA modelling. We neglected other sources of systematic effects for the WL probe such as the shear bias and the photometric redshift uncertainties (which are also present for the tomographic bins in photometric GC in principle).

Refer to caption
Figure 2: GC (orange) and the XC (green) spectrum for the bin correlation 6–6. The dashed lines correspond to spectra without RSD, and the solid lines show spectra with RSD.

In the bottom panel of Fig. 1 we show the signal loss when we neglect the linear RSD contribution in the cross power spectrum between GC and WL (hereafter XC). The picture is very similar compared to the one presented for the GC (top panel of Fig. 1). The main difference is that the RSD contribution to the total signal is lower than 1%similar-toabsentpercent1\sim 1\%∼ 1 % at the lowest redshift (bin pair 1–1), reaching a maximum of up to 70%similar-toabsentpercent70\sim 70\%∼ 70 % in bin pair 6–6 at the largest scales.

For some bin cases, the signal that is lost when the RSD is neglected with respect to the total signal in XC is more than the GC, which indeed seems to be counter-intuitive because the RSD is an additional term in the GC kernel Eq. 3. Although it is true that the RSD kernel given by Eq. 5 appears twice in Eq. 2 for all bin correlations due to Eq. 3 and only once in the XC, the GC spectra have more power than the XC spectra, as shown in Fig. 2. The WL kernel of Eq. 9 also appears once in the XC, and this kernel lowers the signal. For this reason, the XC spectra might be numerically more strongly affected when the RSD signal stronger, as in the bin correlations 6–6 and 10–10.

2.4 The Euclid Flagship simulation

In order to create realistic mock data vectors for our analysis, we used the simulated results from the Flagship galaxy simulation of the Euclid Consortium (Euclid Consortium, in preparation). The galaxy catalogue was produced using the N𝑁Nitalic_N-body Flagship dark matter simulation (Potter et al., 2017) with a ΛΛ\Lambdaroman_ΛCDM fiducial cosmology given by the total matter abundance, Ωm,0=0.319subscriptΩm00.319\varOmega_{{\rm m},0}=0.319roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT = 0.319; the baryon abundance, Ωb,0=0.049subscriptΩb00.049\varOmega_{{\rm b},0}=0.049roman_Ω start_POSTSUBSCRIPT roman_b , 0 end_POSTSUBSCRIPT = 0.049; the r.m.s. variance of the linear matter fluctuations at z=0𝑧0z=0italic_z = 0 in spheres with a radius of 8h1Mpc8superscript1Mpc8\,h^{-1}\,\mathrm{Mpc}8 italic_h start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Mpc, σ8=0.830subscript𝜎80.830\sigma_{8}=0.830italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT = 0.830; the spectral index of the primordial curvature power spectrum, ns=0.96subscript𝑛s0.96n_{\rm s}=0.96italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = 0.96; and the dimensionless Hubble constant, hH0/(100kms1Mpc1)=0.67subscript𝐻0100kmsuperscripts1superscriptMpc10.67h\equiv H_{0}/(100\,\mathrm{km\,s^{-1}\,Mpc^{-1}})=0.67italic_h ≡ italic_H start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / ( 100 roman_km roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Mpc start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) = 0.67. The N𝑁Nitalic_N-body simulation ran a box of 3.78h1Gpc3.78superscript1Gpc3.78\,h^{-1}\,\mathrm{Gpc}3.78 italic_h start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Gpc with a particle mass of 2.398×109h1M2.398superscript109superscript1subscript𝑀direct-product2.398\times 10^{9}\,h^{-1}\,\mathit{M}_{\odot}2.398 × 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT italic_h start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT.

The dark matter haloes were identified using ROCKSTAR (Behroozi et al., 2013) down to masses of 2.4×1010h1M2.4superscript1010superscript1subscript𝑀direct-product2.4\times 10^{10}\,h^{-1}\,\mathit{M}_{\odot}2.4 × 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT italic_h start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (corresponding to ten particles per halo). Then, the galaxies were assigned to the haloes using the halo-occupation distribution and the halo-abundance matching methods, following the recipe presented in Carretero et al. (2015). Several observational constraints were used to calibrate the galaxy mocks including the luminosity function (Blanton et al., 2003, 2005) applied for the faint galaxies, the measurements of galaxy clustering as a function of colour and luminosity (Zehavi et al., 2011), and the colour-magnitude diagram from Blanton et al. (2005). The final galaxy catalogue contains almost 3.43.43.43.4 billion galaxies over 500050005000\,5000deg2 and extends up to redshift 2.32.32.32.3.

We followed the analysis of EP:XII, where an optimisation of the galaxy sample for photometric GC analyses was performed using the Flagship simulation. EP:XII generated photometric redshift estimates with the directional-neighbourhood-fitting training-based algorithm (De Vicente et al., 2016) for all galaxies within a patch of 400deg2400superscriptdegree2400\,\deg^{2}400 roman_deg start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the Flagship simulation up to a magnitude limit of 25252525 in the VIS band. We considered the fiducial sample from EP:XII. This corresponds to a training of the algorithm with an incomplete spectroscopic training sample to mimic the lack of spectroscopic information at very faint magnitudes, and to optimistic magnitude limits for all photometric bands. There is an additional selection of objects with magnitudes brighter than 24.524.524.524.5 in the VIS band. By using the directional-neighbourhood-fitting algorithm, two different estimates for the photometric redshift are provided for each object. One estimate is the average of the redshifts from the neighbourhood, which we denote zmeansubscript𝑧meanz_{\rm mean}italic_z start_POSTSUBSCRIPT roman_mean end_POSTSUBSCRIPT. The second estimate is a Monte Carlo draw from the nearest neighbour, and we denote this zmcsubscript𝑧mcz_{\rm mc}italic_z start_POSTSUBSCRIPT roman_mc end_POSTSUBSCRIPT. We refer to EP:XII and De Vicente et al. (2016) for the similarities and differences between these two estimates. The final sample was then composed of 13131313 tomographic equispaced bins in zmeansubscript𝑧meanz_{\rm mean}italic_z start_POSTSUBSCRIPT roman_mean end_POSTSUBSCRIPT up to z=2𝑧2z=2italic_z = 2, that is, 13131313 bins with a constant redshift width in zmeansubscript𝑧meanz_{\rm mean}italic_z start_POSTSUBSCRIPT roman_mean end_POSTSUBSCRIPT, and therefore, with a different width in zmcsubscript𝑧mcz_{\rm mc}italic_z start_POSTSUBSCRIPT roman_mc end_POSTSUBSCRIPT. The normalised number densities of these bins as a function of redshift are shown in Fig. 3. In addition to the galaxy distributions, we also considered the linear galaxy biases for each of these distributions, which are provided in EP:XII.

Refer to caption
Figure 3: Normalised number densities of the 13131313 tomographic equispaced bins with redshift from the Euclid Flagship simulation (EP:XII).

3 Synthetic data and analysis method

In order to quantify the impact of linear RSD on the photometric observables of Euclid, we createed mock data that we compared with our theoretical predictions. For this purpose, we computed the power spectra for GC, WL, and XC following Eq. 2 in a fiducial cosmology. We chose this to be the same as EP:VII, namely a flat ΛΛ\Lambdaroman_ΛCDM model with one massive and two massless neutrinos fixed to mν=0.06subscript𝑚𝜈0.06\sum m_{\nu}=0.06∑ italic_m start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = 0.06eV. The fiducial values for the parameters are shown in Table 1 and Table 2.

Table 1: Fiducial values of the cosmological parameters.
Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT Ωb,0subscriptΩb0\varOmega_{{\rm b},0}roman_Ω start_POSTSUBSCRIPT roman_b , 0 end_POSTSUBSCRIPT hhitalic_h σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT
0.3200.3200.3200.320 0.0500.0500.0500.050 0.670.670.670.67 0.8160.8160.8160.816 0.960.960.960.96 1.01.0-1.0- 1.0 0.00.00.00.0
Table 2: Fiducial values of the nuisance parameters.
b1subscript𝑏1b_{1}italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT b2subscript𝑏2b_{2}italic_b start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT b3subscript𝑏3b_{3}italic_b start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT b4subscript𝑏4b_{4}italic_b start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT b5subscript𝑏5b_{5}italic_b start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT b6subscript𝑏6b_{6}italic_b start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT b7subscript𝑏7b_{7}italic_b start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT b8subscript𝑏8b_{8}italic_b start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT b9subscript𝑏9b_{9}italic_b start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT b10subscript𝑏10b_{10}italic_b start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT b11subscript𝑏11b_{11}italic_b start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT b12subscript𝑏12b_{12}italic_b start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT b13subscript𝑏13b_{13}italic_b start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT 𝒜IAsubscript𝒜IA\mathcal{A}_{\text{IA}}caligraphic_A start_POSTSUBSCRIPT IA end_POSTSUBSCRIPT βIAsubscript𝛽IA\beta_{\text{IA}}italic_β start_POSTSUBSCRIPT IA end_POSTSUBSCRIPT ηIAsubscript𝜂IA\eta_{\text{IA}}italic_η start_POSTSUBSCRIPT IA end_POSTSUBSCRIPT
1.0251.0251.0251.025 1.0371.0371.0371.037 1.0661.0661.0661.066 1.1101.1101.1101.110 1.1981.1981.1981.198 1.2931.2931.2931.293 1.4291.4291.4291.429 1.5591.5591.5591.559 1.7581.7581.7581.758 1.9471.9471.9471.947 2.2172.2172.2172.217 2.5372.5372.5372.537 2.7382.7382.7382.738 1.721.721.721.72 2.172.172.172.17 0.410.41-0.41- 0.41

Note: The linear galaxy bias parameter per redshift bin i𝑖iitalic_i for GC is denoted with bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and the WL parameters for the whole redshift range with 𝒜IAsubscript𝒜IA\mathcal{A}_{\text{IA}}caligraphic_A start_POSTSUBSCRIPT IA end_POSTSUBSCRIPT , βIAsubscript𝛽IA\beta_{\text{IA}}italic_β start_POSTSUBSCRIPT IA end_POSTSUBSCRIPT, and ηIAsubscript𝜂IA\eta_{\text{IA}}italic_η start_POSTSUBSCRIPT IA end_POSTSUBSCRIPT (for the resources of the IA modelling, see again EP:VII.). The centres of the redshift bins are the following: z¯1=0.14subscript¯𝑧10.14\bar{z}_{1}=0.14over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.14, z¯2=0.26subscript¯𝑧20.26\bar{z}_{2}=0.26over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.26, z¯3=0.39subscript¯𝑧30.39\bar{z}_{3}=0.39over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0.39, z¯4=0.53subscript¯𝑧40.53\bar{z}_{4}=0.53over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT = 0.53, z¯5=0.69subscript¯𝑧50.69\bar{z}_{5}=0.69over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT = 0.69, z¯6=0.84subscript¯𝑧60.84\bar{z}_{6}=0.84over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT = 0.84, z¯7=1.0subscript¯𝑧71.0\bar{z}_{7}=1.0over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT = 1.0, z¯8=1.14subscript¯𝑧81.14\bar{z}_{8}=1.14over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT = 1.14, z¯9=1.3subscript¯𝑧91.3\bar{z}_{9}=1.3over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT = 1.3, z¯10=1.44subscript¯𝑧101.44\bar{z}_{10}=1.44over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT = 1.44, z¯11=1.62subscript¯𝑧111.62\bar{z}_{11}=1.62over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT = 1.62, z¯12=1.78subscript¯𝑧121.78\bar{z}_{12}=1.78over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = 1.78 and z¯13=1.72subscript¯𝑧131.72\bar{z}_{13}=1.72over¯ start_ARG italic_z end_ARG start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT = 1.72.

We modelled the matter power spectrum, Pδδsubscript𝑃𝛿𝛿P_{\delta\delta}italic_P start_POSTSUBSCRIPT italic_δ italic_δ end_POSTSUBSCRIPT, on non-linear small scales using halofit (Smith et al., 2003), including corrections for both dark energy (Takahashi et al., 2012) and massive neutrinos (Bird et al., 2012) as in EP:VII. Using these assumptions, we can compute the harmonic-space power spectra of Eq. 2 for photometric GC, WL, and XC. These represent our synthetic data set, against which we can compare theoretical predictions from different models to constrain model parameters. In order to do this, we obtained the χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT for any set of parameters 𝜽𝜽\boldsymbol{\theta}bold_italic_θ including cosmological and nuisance, and assumed the Gaussian likelihood

χ2(𝜽)=[𝒅𝒕(𝜽)]𝖳𝖢1[𝒅𝒕(𝜽)].superscript𝜒2𝜽superscriptdelimited-[]𝒅𝒕𝜽𝖳superscript𝖢1delimited-[]𝒅𝒕𝜽\chi^{2}(\boldsymbol{\theta})=\left[\boldsymbol{d}-\boldsymbol{t}(\boldsymbol{% \theta})\right]^{\sf T}\mathsf{C}^{-1}\left[\boldsymbol{d}-\boldsymbol{t}(% \boldsymbol{\theta})\right]\;.italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( bold_italic_θ ) = [ bold_italic_d - bold_italic_t ( bold_italic_θ ) ] start_POSTSUPERSCRIPT sansserif_T end_POSTSUPERSCRIPT sansserif_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT [ bold_italic_d - bold_italic_t ( bold_italic_θ ) ] . (15)

The data vector 𝒅𝒅\boldsymbol{d}bold_italic_d, the vector of theoretical predictions 𝒕(𝜽)𝒕𝜽\boldsymbol{t}({\boldsymbol{\theta}})bold_italic_t ( bold_italic_θ ), and the covariance matrix 𝖢𝖢\mathsf{C}sansserif_C (which is assumed to be constant and independent of the cosmological parameters), were all stacked along the i𝑖iitalic_i, j𝑗jitalic_j, and \ellroman_ℓ indices.

The covariance matrix of the data, 𝖢𝖢\mathsf{C}sansserif_C, is the flattened version of the fourth-order Gaussian covariance (as in EP:VII), namely,

Cov[CijAB(),CmnAB()]=C~imAA()C~jnBB()+C~inAB()C~jmBA()(2+1)ΔfskyδK,Covsubscriptsuperscript𝐶𝐴𝐵𝑖𝑗subscriptsuperscript𝐶superscript𝐴superscript𝐵𝑚𝑛superscriptsubscriptsuperscript~𝐶𝐴superscript𝐴𝑖𝑚subscriptsuperscript~𝐶𝐵superscript𝐵𝑗𝑛subscriptsuperscript~𝐶𝐴superscript𝐵𝑖𝑛subscriptsuperscript~𝐶𝐵superscript𝐴𝑗𝑚21Δsubscript𝑓skysubscriptsuperscript𝛿Ksuperscript\text{Cov}\left[C^{AB}_{ij}(\ell),C^{A^{\prime}B^{\prime}}_{mn}(\ell^{\prime})% \right]=\frac{\widetilde{C}^{AA^{\prime}}_{im}(\ell)\,\widetilde{C}^{BB^{% \prime}}_{jn}(\ell)+\widetilde{C}^{AB^{\prime}}_{in}(\ell)\,\widetilde{C}^{BA^% {\prime}}_{jm}(\ell)}{(2\,\ell+1)\,\varDelta\ell\,f_{\rm sky}}\,\delta^{\rm K}% _{\ell\ell^{\prime}}\;,start_ROW start_CELL Cov [ italic_C start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) , italic_C start_POSTSUPERSCRIPT italic_A start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_m italic_n end_POSTSUBSCRIPT ( roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) ] = divide start_ARG over~ start_ARG italic_C end_ARG start_POSTSUPERSCRIPT italic_A italic_A start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT ( roman_ℓ ) over~ start_ARG italic_C end_ARG start_POSTSUPERSCRIPT italic_B italic_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j italic_n end_POSTSUBSCRIPT ( roman_ℓ ) + over~ start_ARG italic_C end_ARG start_POSTSUPERSCRIPT italic_A italic_B start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT ( roman_ℓ ) over~ start_ARG italic_C end_ARG start_POSTSUPERSCRIPT italic_B italic_A start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_j italic_m end_POSTSUBSCRIPT ( roman_ℓ ) end_ARG start_ARG ( 2 roman_ℓ + 1 ) roman_Δ roman_ℓ italic_f start_POSTSUBSCRIPT roman_sky end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUPERSCRIPT roman_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_ℓ roman_ℓ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , end_CELL end_ROW (16)

with C~ijAB()=CijAB()+NijAB()subscriptsuperscript~𝐶𝐴𝐵𝑖𝑗subscriptsuperscript𝐶𝐴𝐵𝑖𝑗subscriptsuperscript𝑁𝐴𝐵𝑖𝑗\widetilde{C}^{AB}_{ij}(\ell)=C^{AB}_{ij}(\ell)+N^{AB}_{ij}(\ell)over~ start_ARG italic_C end_ARG start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) = italic_C start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) + italic_N start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ), and NijAB()subscriptsuperscript𝑁𝐴𝐵𝑖𝑗N^{AB}_{ij}(\ell)italic_N start_POSTSUPERSCRIPT italic_A italic_B end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) being the noise contribution to the measurement, and fsky=0.3636subscript𝑓sky0.3636f_{\rm sky}=0.3636italic_f start_POSTSUBSCRIPT roman_sky end_POSTSUBSCRIPT = 0.3636 is the sky fraction observed by Euclid. Finally, we used

NijΔΔ()subscriptsuperscript𝑁ΔΔ𝑖𝑗\displaystyle N^{\varDelta\varDelta}_{ij}(\ell)italic_N start_POSTSUPERSCRIPT roman_Δ roman_Δ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) =1n¯iδijK,absent1subscript¯𝑛𝑖subscriptsuperscript𝛿K𝑖𝑗\displaystyle=\frac{1}{\bar{n}_{i}}\,\delta^{\rm K}_{ij}\;,= divide start_ARG 1 end_ARG start_ARG over¯ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUPERSCRIPT roman_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT , (17)
Nijϵϵ()subscriptsuperscript𝑁italic-ϵitalic-ϵ𝑖𝑗\displaystyle N^{\epsilon\epsilon}_{ij}(\ell)italic_N start_POSTSUPERSCRIPT italic_ϵ italic_ϵ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) =σϵ2n¯iδijK,absentsubscriptsuperscript𝜎2italic-ϵsubscript¯𝑛𝑖subscriptsuperscript𝛿K𝑖𝑗\displaystyle=\frac{\sigma^{2}_{\epsilon}}{\bar{n}_{i}}\,\delta^{\rm K}_{ij}\;,= divide start_ARG italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_δ start_POSTSUPERSCRIPT roman_K end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT , (18)
NijΔϵ()subscriptsuperscript𝑁Δitalic-ϵ𝑖𝑗\displaystyle N^{\varDelta\epsilon}_{ij}(\ell)italic_N start_POSTSUPERSCRIPT roman_Δ italic_ϵ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ( roman_ℓ ) =0,absent0\displaystyle=0\;,= 0 , (19)

with n¯isubscript¯𝑛𝑖\bar{n}_{i}over¯ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT being the number density of galaxies per steradian in each tomographic bin and σϵ2superscriptsubscript𝜎italic-ϵ2\sigma_{\epsilon}^{2}italic_σ start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT the intrinsic ellipticity variance, which we assumed to be σϵ=0.3subscript𝜎italic-ϵ0.3\sigma_{\epsilon}=0.3italic_σ start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = 0.3 (as in EP:VII, ).

We binned our data in multipoles considering N=20subscript𝑁20N_{\ell}=20italic_N start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT = 20 log-spaced multipole bins in the range {min,max}subscriptminsubscriptmax\{\ell_{\rm min},\ell_{\rm max}\}{ roman_ℓ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT , roman_ℓ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT }, with ΔΔ\varDelta\ellroman_Δ roman_ℓ the width of each bin. We considered the lowest multipole to be min=10subscriptmin10\ell_{\rm min}=10roman_ℓ start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 10 and considered an optimistic and a pessimistic scenario for the maximum multipole cut maxsubscriptmax\ell_{\rm max}roman_ℓ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT (see EP:VII, ), with

  • Optimistic — pessimistic GC: 3000 — 750,

  • Optimistic — pessimistic WL: 5000 — 1500.

For XC, we conservatively considered the smallest maxsubscriptmax\ell_{\rm max}roman_ℓ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT (corresponding to the GC values).

The assumptions we adopted from EP:VII for the purposes of this work in our modelling are not expected to affect the results strongly. For example, an increase in σϵsubscript𝜎italic-ϵ\sigma_{\epsilon}italic_σ start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT would have an effect on the WL part but not on the GC and XC. Then, a decrease in n¯isubscript¯𝑛𝑖{\bar{n}_{i}}over¯ start_ARG italic_n end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT would increase the shot noise, but the GC part of the analysis, which is what we are most interested in, with the linear RSD scales, should not be dominated by shot noise. Therefore, the 3×\times×2pt contours would increase through the WL contribution, but would not change our conclusions on RSD.

With this approach, we needed only one additional ingredient to obtain the χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of Eq. 15, which is the theory vector 𝒕(𝜽)𝒕𝜽\boldsymbol{t}(\boldsymbol{\theta})bold_italic_t ( bold_italic_θ ). We chose to focus on two models to be constrained with our mock data, which we refer to as ΛΛ\Lambdaroman_ΛCDM and w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM flat cosmologies following the same modelling as in EP:VII. The former has five free cosmological parameters, which are Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT, Ωb,0subscriptΩb0\varOmega_{{\rm b},0}roman_Ω start_POSTSUBSCRIPT roman_b , 0 end_POSTSUBSCRIPT, hhitalic_h, σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, and nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT. The latter also includes the equation-of-state parameters of dark energy, w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, which come from the CPL parametrisation (Chevallier & Polarski, 2001; Linder, 2003),

w(z)=w0+waz1+z.𝑤𝑧subscript𝑤0subscript𝑤𝑎𝑧1𝑧w(z)=w_{0}+w_{a}\frac{z}{1+z}\;.italic_w ( italic_z ) = italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT divide start_ARG italic_z end_ARG start_ARG 1 + italic_z end_ARG . (20)

As nuisance parameters, both models feature for the GC, the linear galaxy bias amplitudes bisubscript𝑏𝑖b_{i}italic_b start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (see Sect. 2.2), and for the WL, the IA modelling parameters 𝒜IAsubscript𝒜IA\mathcal{A}_{\rm IA}caligraphic_A start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT, βIAsubscript𝛽IA\beta_{\rm IA}italic_β start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT, and ηIAsubscript𝜂IA\eta_{\rm IA}italic_η start_POSTSUBSCRIPT roman_IA end_POSTSUBSCRIPT (see Sect. 2.3).

4 Results

In this section, we use the method presented in Sect. 3 to investigate the impact of the RSD contribution, modelled as in Sect. 2.2, on the final cosmological constraints. We explore the parameter space with a Markov chain Monte Carlo (MCMC) approach using emcee (Foreman-Mackey et al., 2013), and we obtain from CAMB (Lewis et al., 2000; Howlett et al., 2012) all the quantities needed to compute the theoretical predictions following Eq. 2 in a modified version of the code CosmoSIS (Zuntz et al., 2015). We adopt improper (flat) priors for all the parameters in Table 1 and Table 2.

4.1 Validation

Before we proceeded to set up the realistic and computationally expensive MCMC chains for all the different cosmologies and scenarios, we first ran a single case with an MCMC and compared it to a Fisher matrix forecast (see EP:VII, for details about Fisher forecasts and their derivatives accuracy) in order to validate the pipeline. In brief, entries of the Fisher matrix are constructed as

Fαβ=𝒕𝖳θα𝖢1𝒕θβ,subscript𝐹𝛼𝛽superscript𝒕𝖳subscript𝜃𝛼superscript𝖢1𝒕subscript𝜃𝛽F_{\alpha\beta}=\frac{\partial\boldsymbol{t}^{\sf T}}{\partial\theta_{\alpha}}% \mathsf{C}^{-1}\frac{\partial\boldsymbol{t}}{\partial\theta_{\beta}}\;,italic_F start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT = divide start_ARG ∂ bold_italic_t start_POSTSUPERSCRIPT sansserif_T end_POSTSUPERSCRIPT end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT end_ARG sansserif_C start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG ∂ bold_italic_t end_ARG start_ARG ∂ italic_θ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_ARG , (21)

where {θα}subscript𝜃𝛼\{\theta_{\alpha}\}{ italic_θ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT } are the elements of the parameter vector 𝜽𝜽\boldsymbol{\theta}bold_italic_θ (see Sect. 3).

In general, the Fisher forecasts, whose computation is usually faster than any Bayesian sampler, are used in order to investigate the likelihood curvature of the parameter hyperspace near its peak. When the posterior is very well described by a Gaussian, the Fisher forecasts are particularly accurate, and the smaller the uncertainties, the closer the peak of the posterior where the Gaussian approximation holds. This applies to forthcoming LSS experiments such as Euclid, which will provide us with an unprecedentedly large number of sources that will minimise the resulting errors. As expected, the precision will increase even further when we consider all the available signal. Therefore, we performed a Fisher forecast on the combination of GC, WL, and XC, which we labelled 3×\times×2pt. We show this by considering a w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM cosmological model and an optimistic scale cut, in order to investigate the potential Gaussianity of the posterior in an extended cosmology model with two additional parameters, {w0,wa}subscript𝑤0subscript𝑤𝑎\{w_{0},w_{a}\}{ italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT }. In Eq. 21 we calculated the theory vector 𝒕𝒕\boldsymbol{t}bold_italic_t for the fiducial cosmology model and the covariance matrix 𝖢𝖢\mathsf{C}sansserif_C following Sect. 3. We also included linear RSD as described in Sect. 2.2. With the same assumptions, we repeated the analysis with an MCMC approach, exploring the posterior and then finding the minimum χ2superscript𝜒2\chi^{2}italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over the parameter space 𝜽𝜽\boldsymbol{\theta}bold_italic_θ.

The comparison between the MCMC and Fisher approaches is shown in the top panel of Fig. 4, where the 68%percent6868\%68 % and 95%percent9595\%95 % credibility levels (C.L.) on cosmological parameters are presented, after the bias and IA parameters were marginalised over.333we note that 68%percent6868\%68 % and 95%percent9595\%95 % C.L. exactly correspond to one and two standard deviations in the Gaussian approximation of the Fisher matrix after marginalising over the remaining parameters.. The filled green contours correspond to the MCMC and the empty orange contours show the Fisher forecast. The agreement between the two is remarkable and highlights the Gaussianity of the posterior, with almost perfectly overlap** constraints and all the directions and widths of the Fisher ellipses recovered by the MCMC. In addition, the fiducial cosmology model values are presented with dotted black lines. They are located well within the 68%percent6868\%68 % C.L. contours.

Refer to caption
Refer to caption
Figure 4: 68%percent6868\%68 % and 95%percent9595\%95 % C.L. marginalised contours along with the one-dimensional posterior distributions on the cosmological parameters. Top: Constraints from the 3×\times×2pt analysis of the flat w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model for the optimistic scale cut considering RSD. The filled green contours are the MCMC constraints, and the open orange contours show the Fisher ellipses for the same modelling. The fiducial cosmology model is marked with dotted black lines. Bottom: Constraints from the GC alone of the flat w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model (pessimistic scenario), with (orange) and without (green) RSD. The fiducial cosmology model is marked with dotted black lines.

4.2 The contribution of RSD to constraining power

After we validated our MCMC pipeline, we investigated how the additional information encoded in RSD affects the constraining power on the model parameters. To do this, we considered GC alone and not the total 3×\times×2pt because as shown in Sect. 2.2, RSD is a correction for the GC signal alone, and therefore, a change in its constraining power would be easier to appreciate. We should mention at this point that the RSD also impacts the XC part of the 3×\times×2pt because the latter contains all the combinations of GC, WL, and XC. However, we did not perform the constraining power test due to the RSD in the 3×\times×2pt because the WL contribution which is not affected by RSD would make the RSD impact less evident. We again focused on the w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model and the pessimistic scale cut. This is reasonable because RSD mostly contributes on scales <100100\ell<100roman_ℓ < 100 (see Fig. 1), and therefore a higher maxsubscriptmax\ell_{\text{max}}roman_ℓ start_POSTSUBSCRIPT max end_POSTSUBSCRIPT cut would not affect the constraining power for this test.

In particular, we compared the following two models. For the first model, we constructed the synthetic data set and covariance matrix for the photometric GC spectra including RSD (GC with RSD), and we fit it against the predictions of the w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model, including RSD. The results of this analysis are shown by the orange contours in the bottom panel of Fig. 4. For the second model, we did exactly the same, but neither the synthetic data along with the covariance matrix nor the theory model included the RSD correction (GC without RSD); this corresponds to the green contours in the bottom panel of Fig. 4.

It is clear that the constraints of the two models agree very well, indicating that the RSD correction in our modelling does not add significant information on the projected cosmological parameters. This means that even though the RSD account for up to 40%percent4040\%40 % of the signal for some redshift bins on the largest scales (see again top panel of Fig. 1), the cosmological information does not come from the large scales because they are dominated by cosmic variance, but gradually include smaller scales, where the RSD signal contribution becomes progressively less important.

Table 3: Summary of the mean estimated values, θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT, along their corresponding 68%percent6868\%68 % C.L. intervals, σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT, and the relative bias, Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT (see Eq. 22) of a given parameter.
w/ RSD w/o RSD
Parameter Scale cut θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT
Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT pess. 0.31930.31930.31930.3193 0.00520.00520.00520.0052 0.134   0.30190.30190.30190.3019 0.00440.00440.00440.0044 4.11
opt. 0.31970.31970.31970.3197 0.00370.00370.00370.0037 0.08   0.30220.30220.30220.3022 0.00330.00330.00330.0033 5.39
Ωb,0subscriptΩb0\varOmega_{{\rm b},0}roman_Ω start_POSTSUBSCRIPT roman_b , 0 end_POSTSUBSCRIPT pess. 0.05030.05030.05030.0503 0.00240.00240.00240.0024 0.125   0.04990.04990.04990.0499 0.00220.00220.00220.0022 0.045
opt. 0.05020.05020.05020.0502 0.00180.00180.00180.0018 0.111   0.05060.05060.05060.0506 0.00180.00180.00180.0018 0.33
hhitalic_h pess. 0.6740.6740.6740.674 0.0230.0230.0230.023 0.174   0.6680.6680.6680.668 0.0210.0210.0210.021 0.095
opt. 0.6720.6720.6720.672 0.0150.0150.0150.015 0.133   0.6730.6730.6730.673 0.0160.0160.0160.016 0.18
σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT pess. 0.81540.81540.81540.8154 0.00590.00590.00590.0059 0.101   0.83270.83270.83270.8327 0.00620.00620.00620.0062 2.69
opt. 0.8160.8160.8160.816 0.00130.00130.00130.0013 0.0   0.82060.82060.82060.8206 0.00130.00130.00130.0013 3.53
nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT pess. 0.960.960.960.96 0.0120.0120.0120.012 0.0   0.9590.9590.9590.959 0.0110.0110.0110.011 0.091
opt. 0.95960.95960.95960.9596 0.00370.00370.00370.0037 0.108   0.96840.96840.96840.9684 0.00380.00380.00380.0038 2.21
S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT pess. 0.84120.84120.84120.8412 0.00950.00950.00950.0095 0.164   0.83530.83530.83530.8353 0.0090.0090.0090.009 0.829
opt. 0.84240.84240.84240.8424 0.00460.00460.00460.0046 0.078   0.82360.82360.82360.8236 0.00430.00430.00430.0043 4.45

Note: These are the results of the GC analysis for each ΛΛ\Lambdaroman_ΛCDM cosmological parameter for the complete (with RSD) and the incomplete (without RSD) model and for the pessimistic and optimistic scale cuts. We highlight the most biased cases (Bθ>1subscript𝐵𝜃1B_{\theta}>1italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT > 1) with bold.

Table 4: Same as Table 3 for GC but for the w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model.
w/ RSD w/o RSD
Parameter Scale cut θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT
Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT opt. 0.31900.31900.31900.3190 0.00690.00690.00690.0069 0.145   0.30920.30920.30920.3092 0.00560.00560.00560.0056 1.928
Ωb,0subscriptΩb0\varOmega_{{\rm b},0}roman_Ω start_POSTSUBSCRIPT roman_b , 0 end_POSTSUBSCRIPT opt. 0.04990.04990.04990.0499 0.00190.00190.00190.0019 0.052   0.05160.05160.05160.0516 0.00190.00190.00190.0019 0.84
hhitalic_h opt. 0.6720.6720.6720.672 0.0180.0180.0180.018 0.11   0.6640.6640.6640.664 0.0160.0160.0160.016 0.37
σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT opt. 0.81660.81660.81660.8166 0.0030.0030.0030.003 0.2   0.81580.81580.81580.8158 0.00270.00270.00270.0027 0.074
nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT opt. 0.960.960.960.96 0.00430.00430.00430.0043 0.0   0.96690.96690.96690.9669 0.00430.00430.00430.0043 1.60
w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT opt. 1.0011.001-1.001- 1.001 0.0420.0420.0420.042 0.0238   1.0071.007-1.007- 1.007 0.0390.0390.0390.039 0.17
wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT opt. 0.020.02-0.02- 0.02 0.180.180.180.18 0.111   0.210.210.210.21 0.170.170.170.17 1.235
S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT opt. 0.8420.8420.8420.842 0.0070.0070.0070.007 0.108   0.82810.82810.82810.8281 0.00580.00580.00580.0058 2.52

4.3 Ignoring RSD

Regardless of the absence of an additional constraining power on cosmological parameters encoded in linear RSD, we proceeded to investigate the effect of neglecting RSD in our modelling. This investigation was performed with an MCMC analysis in Tanidis & Camera (2019), assuming a Euclid-like survey of intermediate-width Gaussian bins (with a photometric redshift scatter 0.05(1+z)0.051𝑧0.05(1+z)0.05 ( 1 + italic_z )). We found that it is crucial to include RSD in the theoretical predictions to avoid biasing the cosmological parameters. This outcome of the importance of RSD in the modelling agreed with the decision in previous studies to include them in the analyses (see again the references in Sect. 1). Following the same approach, we created a mock data vector and covariance for the Euclid survey, with the same specifications as in EP:VII, including the RSD contribution (Sect. 2.2). The data were then analysed assuming an incorrect model without the RSD contribution. The impact of this approximation on the accuracy of the final constraints depends on the amplitude of the RSD signal, but also on the details of the experimental noise. We therefore extended the investigation of Tanidis & Camera (2019), which was only performed for GC in a ΛΛ\Lambdaroman_ΛCDM model and in the linear regime 444In Tanidis & Camera (2019) only the linear matter power spectrum was used and maxsubscriptmax\ell_{\text{max}}roman_ℓ start_POSTSUBSCRIPT max end_POSTSUBSCRIPT was defined by kmaxχ(z¯)subscript𝑘max𝜒¯𝑧k_{\text{max}}\,\chi(\bar{z})italic_k start_POSTSUBSCRIPT max end_POSTSUBSCRIPT italic_χ ( over¯ start_ARG italic_z end_ARG ) with kmax0.25hMpc1similar-to-or-equalssubscript𝑘max0.25superscriptMpc1k_{\text{max}}\simeq 0.25\,h\,\mathrm{Mpc}^{-1}italic_k start_POSTSUBSCRIPT max end_POSTSUBSCRIPT ≃ 0.25 italic_h roman_Mpc start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and z¯¯𝑧\bar{z}over¯ start_ARG italic_z end_ARG the mean redshift of the bin. for the density field, to more realistic settings. That is, we still considered the linear RSD, but now in the non-linear regime of the density field. In addition, we exploited photometric redshift bins from the Flagship simulation, including massive neutrinos in the analysis, and also investigated the effect on the w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model. Finally, we performed the analysis both for GC alone and for the full 3×\times×2pt, using the pessimistic and optimistic multipole cuts for both cases.

Finally, to gain better insight into the degeneracies that are present in the incorrect modelling, we recast all the final constraints on Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT in the derived parameter S8=σ8Ωm,0/0.3subscript𝑆8subscript𝜎8subscriptΩm00.3\smash{S_{8}=\sigma_{8}\,\sqrt{\varOmega_{{\rm m},0}/0.3}}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT = italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT square-root start_ARG roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT / 0.3 end_ARG. This parameter is particularly informative about the degenerate direction between Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT in the WL measurements that are included in the 3×\times×2pt. However, for the sake of completeness, we calculated it for the GC alone as well.

4.3.1 Biased parameter constraints in the photometric GC

We started by obtaining constraints for both the correct (den+RSD) and incorrect models (den-only), using photometric GC alone, both assuming ΛΛ\Lambdaroman_ΛCDM and its extension w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM. To assess the bias on the parameters of interest when an incorrect model is assumed, we used the relative bias which is valid for deterministic quantities assuming no stochasticity (e.g. a noiseless data vector such as we consider here),

Bθ=|θθfid|σθ,subscript𝐵𝜃superscript𝜃superscript𝜃fidsubscript𝜎𝜃B_{\theta}=\frac{|\theta^{\ast}-\theta^{\text{fid}}|}{\sigma_{\theta}}\;,italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = divide start_ARG | italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT - italic_θ start_POSTSUPERSCRIPT fid end_POSTSUPERSCRIPT | end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT end_ARG , (22)

with θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT being the mean estimated value on the marginalised posterior of the parameter θ𝜃\thetaitalic_θ, σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT the corresponding 68%percent6868\%68 % C.L., and θfidsuperscript𝜃fid\theta^{\rm\text{fid}}italic_θ start_POSTSUPERSCRIPT fid end_POSTSUPERSCRIPT the input fiducial value. It is defined as the offset we obtain from the input fiducial value (our benchmark) given the expected uncertainty on the parameter. Massey et al. (2012) suggested that after taking systematic effects into account, the values of Bθ0.3greater-than-or-equivalent-tosubscript𝐵𝜃0.3B_{\theta}\gtrsim 0.3italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ≳ 0.3 can be considered already as statistically significant 555Values of Bθ<0.3subscript𝐵𝜃0.3B_{\theta}<0.3italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT < 0.3 are considered to be permissible within the statistical fluctuation assuming that the same information for the synthetic data and the theory modelling was used..

All the ΛΛ\Lambdaroman_ΛCDM results in GC for the correct and incorrect model, as well as the pessimistic and optimistic scale cuts, are shown in Table 3. We opted to present the incorrect model constraints for the optimistic cases alone in the top panel of Fig. 5 because the resulting biases are larger than those in the pessimistic cut. We note that in the pessimistic case, neglecting RSD leads to a bias of 4.114.114.114.11 and 2.692.692.692.69 on the parameters Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, respectively; while with the correct modelling we always recover the fiducial cosmology with Bθ0.3less-than-or-similar-tosubscript𝐵𝜃0.3B_{\theta}\lesssim 0.3italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT ≲ 0.3 (see again Table 3). This bias is also imprinted on the increase in the best-fit value by Δχ284.61Δsuperscript𝜒284.61\varDelta\chi^{2}\approx 84.61roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≈ 84.61 with respect to the correct model.

The picture is similar for the optimistic scale cut, as the green contours in the top panel of Fig. 5 show. We note that the bias on the same parameters increases, and another mild bias now arises on nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT. In particular, the parameters Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT, σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, and nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT are biased with values 5.395.395.395.39, 3.533.533.533.53, and 2.212.212.212.21, respectively, and as expected there is increase in Δχ292.48Δsuperscript𝜒292.48\varDelta\chi^{2}\approx 92.48roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≈ 92.48 with respect to the correct model. This bias increase in the optimistic compared to the pessimistic scenario can be explained as follows: By extending the range to higher multipoles, we do not include more signal from the linear RSD which, as we saw in Sect. 2.2, contributes only on large scales (100less-than-or-similar-to100\ell\lesssim 100roman_ℓ ≲ 100). On the other hand, we increase the constraining power on the parameters, meaning that any existing bias is enhanced by lowering the projected errors. This shows that an inaccurate modelling becomes progressively more crucial with improving data quality.

Refer to caption
Refer to caption
Figure 5: 68%percent6868\%68 % and 95%percent9595\%95 % C.L. marginalised contours alongside the corresponding one-dimensional posterior distributions of the cosmological parameters for the incorrect modelling (without RSD). Top: Constraints from the GC for the ΛΛ\Lambdaroman_ΛCDM (green contours) and w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model (orange contours) for the optimistic scale cut. The fiducial cosmology is marked by dotted black lines. Bottom: Same as above, but for 3×\times×2pt.

We now turn our attention on the results obtained assuming the w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model. Again, we only considered the optimistic scales cut. The reason for this is that GC alone for the pessimistic scale cut is not very constraining in this model, and the bias on the parameters is expected to be small due to the smaller range of scales and the larger parameter set. The result of all this is that the w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT parameters vary over a wide range in the MCMC, also in regions where w0+wasubscript𝑤0subscript𝑤𝑎w_{0}+w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT can take non-negative values (see Cepa, J., 2004, for details on the high-redshift limit of w(z)𝑤𝑧w(z)italic_w ( italic_z ) in CPL). In this region of the parameter space, CAMB cannot obtain meaningful cosmological quantities, and these points are automatically rejected from the chain, thus introducing a cut in the parameter space that is physically motivated. Based on this restriction and because the modelling for this case is not very constraining, regardless of the assumed prior on the CPL parameters, we decided to focus our analysis on the optimistic case alone. The results are shown in Table 4 and in the top panel of Fig. 5, where we still have biased estimates of our free parameters, but with a lower significance than in the ΛΛ\Lambdaroman_ΛCDM analysis because the uncertainty brought by kee** the dark energy parameters w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT free is larger. The parameters that are shifted by more than 1σ1𝜎1\sigma1 italic_σ are Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT, nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, and wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, with their Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT values being 1.91.91.91.9, 1.61.61.61.6, and 1.21.21.21.2, respectively. Again, there is an increase in Δχ290.65Δsuperscript𝜒290.65\varDelta\chi^{2}\approx 90.65roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≈ 90.65 compared to the complete model. We summarise all the Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT values for the aforementioned scenarios in the left panel of Fig. 6 (see the caption for details), where the trend is clearer for larger biases in the optimistic compared to the pessimistic cases and for those of the ΛΛ\Lambdaroman_ΛCDM model against the w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM extension.

Table 5: Same as Table 3 but for the 3×\times×2pt.
w/ RSD w/o RSD
Parameter scale cut θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT
Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT pess. 0.320.320.320.32 0.00250.00250.00250.0025 0.0   0.31190.31190.31190.3119 0.00230.00230.00230.0023 3.52
opt. 0.320060.320060.320060.32006 0.000820.000820.000820.00082 0.07317   0.317860.317860.317860.31786 0.000820.000820.000820.00082 2.61
Ωb,0subscriptΩb0\varOmega_{{\rm b},0}roman_Ω start_POSTSUBSCRIPT roman_b , 0 end_POSTSUBSCRIPT pess. 0.050.050.050.05 0.00220.00220.00220.0022 0.0   0.04810.04810.04810.0481 0.0020.0020.0020.002 0.95
opt. 0.05010.05010.05010.0501 0.00180.00180.00180.0018 0.0555   0.04950.04950.04950.0495 0.00170.00170.00170.0017 0.29
hhitalic_h pess. 0.6710.6710.6710.671 0.0160.0160.0160.016 0.062   0.6250.6250.6250.625 0.0140.0140.0140.014 3.21
opt. 0.6710.6710.6710.671 0.0110.0110.0110.011 0.090   0.62830.62830.62830.6283 0.010150.010150.010150.01015 4.11
σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT pess. 0.8160.8160.8160.816 0.00320.00320.00320.0032 0.0   0.82790.82790.82790.8279 0.0030.0030.0030.003 3.96
opt. 0.8160.8160.8160.816 0.0010.0010.0010.001 0.0   0.820.820.820.82 0.0010.0010.0010.001 4.0
nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT pess. 0.95960.95960.95960.9596 0.00780.00780.00780.0078 0.0512   0.97470.97470.97470.9747 0.00720.00720.00720.0072 2.05
opt. 0.95980.95980.95980.9598 0.00310.00310.00310.0031 0.0645   0.97590.97590.97590.9759 0.0030.0030.0030.003 5.3
S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT pess. 0.84280.84280.84280.8428 0.000730.000730.000730.00073 0.0531   0.844220.844220.844220.84422 0.000720.000720.000720.00072 2.026
opt. 0.84280.84280.84280.8428 0.000540.000540.000540.00054 0.07189   0.844050.844050.844050.84405 0.000520.000520.000520.00052 2.478
Table 6: Same as Table 4 but for the 3×\times×2pt.
w/ RSD w/o RSD
Parameter scale cut θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT θsuperscript𝜃\theta^{\ast}italic_θ start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT σθsubscript𝜎𝜃\sigma_{\theta}italic_σ start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT
Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT pess. 0.32020.32020.32020.3202 0.00380.00380.00380.0038 0.0526   0.31590.31590.31590.3159 0.00350.00350.00350.0035 1.17
opt. 0.320.320.320.32 0.00120.00120.00120.0012 0.0   0.32020.32020.32020.3202 0.00120.00120.00120.0012 0.16
Ωb,0subscriptΩb0\varOmega_{{\rm b},0}roman_Ω start_POSTSUBSCRIPT roman_b , 0 end_POSTSUBSCRIPT pess. 0.05020.05020.05020.0502 0.00250.00250.00250.0025 0.0799   0.05200.05200.05200.0520 0.00240.00240.00240.0024 0.83
opt. 0.04990.04990.04990.0499 0.00190.00190.00190.0019 0.0526   0.05040.05040.05040.0504 0.00170.00170.00170.0017 0.23
hhitalic_h pess. 0.6710.6710.6710.671 0.0190.0190.0190.019 0.0526   0.6530.6530.6530.653 0.0170.0170.0170.017 1.0
opt. 0.670.670.670.67 0.010.010.010.01 0.0   0.63060.63060.63060.6306 0.00980.00980.00980.0098 4.02
σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT pess. 0.81580.81580.81580.8158 0.00470.00470.00470.0047 0.0425   0.82720.82720.82720.8272 0.00440.00440.00440.0044 2.54
opt. 0.8160.8160.8160.816 0.00120.00120.00120.0012 0.0   0.81860.81860.81860.8186 0.00110.00110.00110.0011 2.36
nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT pess. 0.95950.95950.95950.9595 0.00940.00940.00940.0094 0.0531   0.95570.95570.95570.9557 0.00870.00870.00870.0087 0.49
opt. 0.96020.96020.96020.9602 0.00350.00350.00350.0035 0.0571   0.97150.97150.97150.9715 0.00350.00350.00350.0035 3.28
w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT pess. 0.99960.9996-0.9996- 0.9996 0.0460.0460.0460.046 0.0086   0.9450.945-0.945- 0.945 0.0440.0440.0440.044 1.25
opt. 0.9990.999-0.999- 0.999 0.020.020.020.02 0.05   0.960.96-0.96- 0.96 0.020.020.020.02 2.0
wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT pess. 0.00.00.00.0 0.180.180.180.18 0.0   0.020.020.020.02 0.160.160.160.16 0.125
opt. 0.0050.005-0.005- 0.005 0.0750.0750.0750.075 0.066   0.0830.083-0.083- 0.083 0.0730.0730.0730.073 1.136
S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT pess. 0.84280.84280.84280.8428 0.00140.00140.00140.0014 0.0277   0.84880.84880.84880.8488 0.00140.00140.00140.0014 4.31
opt. 0.842730.842730.842730.84273 0.000860.000860.000860.00086 0.0362   0.845760.845760.845760.84576 0.000890.000890.000890.00089 3.37

4.3.2 Biased parameter constraints in the 3×\times×2pt

We repeated exactly the same analysis (different cosmologies and scale cuts) but now for the 3×\times×2pt (Table 5, Table 6, and the bottom panel of Fig. 5). Similarly to what we saw in Sect. 4.3.1 for the ΛΛ\Lambdaroman_ΛCDM model and the pessimistic scale cuts, the parameters Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT, hhitalic_h, nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT are now biased by 3.523.523.523.52, 3.213.213.213.21, 2.052.052.052.05, and 3.963.963.963.96 (see Table 5), while the overall Δχ2Δsuperscript𝜒2\varDelta\chi^{2}roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT compared to the correct model is increased by 96.27. It is interesting to note that while some of the parameters (e.g. nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT and hhitalic_h) exhibit an increased Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT value due to the increased constraining power of the 3×\times×2pt combination, for the other parameters the behaviour is less straightforward. On the one hand, WL strongly constrains S8subscript𝑆8S_{8}italic_S start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, which is a combination of Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, and because this probe is not biased by neglecting RSD, we would expect a lower bias for these parameters. On the other hand, however, the inclusion of XC lifts the degeneracy between σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT and the bias parameters present in the GC probe, thus making the latter more sensitive to this parameter. Because both GC and XC are biased when we neglect RSD, the overall effect is an increased Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT value on σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT with respect to the case of GC alone.

In the optimistic case, the balance between these effects changes because the increased number of scales available for WL makes this more relevant, thus reducing the bias with respect to the case of GC alone. This does not apply to nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT and hhitalic_h, which are still mostly constrained by GC. The Δχ2Δsuperscript𝜒2\varDelta\chi^{2}roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is increased by 106.78 with respect to the correct model.

Finally, the results of the 3×\times×2pt and the w0wasubscript𝑤0subscript𝑤𝑎w_{0}w_{a}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPTCDM model are shown in Table 6. In the pessimistic case, the biased estimate for σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT is 2.542.542.542.54, and the peaks for the parameters Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT, Ωb,0subscriptΩb0\varOmega_{{\rm b},0}roman_Ω start_POSTSUBSCRIPT roman_b , 0 end_POSTSUBSCRIPT, w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and hhitalic_h are misplaced with biases of 1.171.171.171.17, 0.830.830.830.83, 1.251.251.251.25, and 1.001.001.001.00, while for the optimistic scenario (see again Table 6 and the orange contours in the bottom panel of Fig. 5), there are biased estimates for w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT, hhitalic_h, nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT at 2.002.002.002.00, 1.141.141.141.14, 4.024.024.024.02, 3.283.283.283.28, and 2.362.362.362.36, respectively. Both models yield an increased Δχ2Δsuperscript𝜒2\varDelta\chi^{2}roman_Δ italic_χ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT compared to the complete model by 92.05 (pessimistic) and 103.6 (optimistic). When the latter case is compared to GC alone, the increased constraining power, brought by WL and XC, has the effect of increasing the significance of the bias on most of the parameters. Similarly to Sect. 4.3.1, we summarise the Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT values for all the cases of the 3×\times×2pt in the right panel of Fig. 6.

Refer to caption
Figure 6: Summary plot of the Bθsubscript𝐵𝜃B_{\theta}italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT values for the GC (left) and the 3×\times×2pt (right) for all the examined cases. The circles correspond to the ΛΛ\Lambdaroman_ΛCDM cosmology model and the stars show the w0wαsubscript𝑤0subscript𝑤𝛼w_{0}w_{\alpha}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_w start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPTCDM extension. The open and filled symbols show a theory modelling with and without RSD, respectively. The pessimistic and optimistic scale cuts are shown in brown and blue. The horizontal dashed line denotes the value Bθ=0.3subscript𝐵𝜃0.3B_{\theta}=0.3italic_B start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT = 0.3

5 Conclusions

In this work, we have aimed to quantify the contribution of linear RSD in photometric GC as is expected to be measured by Euclid, both as a stand-alone probe and in combination with cosmic shear (WL), in the so-called 3×\times×2pt approach. We followed Tanidis & Camera (2019) and included RSD in the angular power spectra of GC and its cross-correlation with WL (XC).

Using the galaxy distribution information coming from the Flagship simulation, and the Euclid specifications discussed in EP:VII, we generated synthetic data by generating angular power spectra and covariance matrix with a fiducial cosmology, for the photometric observations of Euclid and produced the posterior distributions for the free parameters of our model in an MCMC framework.

As a first step, we validated our results against a Fisher matrix approach and found that they agree very well. Then, we compared the constraints on the cosmological parameters that are obtained when the theoretical predictions are computed with the contribution of RSD to those obtained when RSD are neglected. When GC is used as a stand-alone probe and a ΛΛ\Lambdaroman_ΛCDM cosmology is assumed for the theoretical predictions, neglecting RSD can lead to significant inaccuracies on the reconstruction of cosmological parameters, in particular, in the most optimistic case when the constraining power of our experimental setup is maximum. We find that the parameters Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT, nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT are all significantly shifted from their (input) fiducial values with biases of 5.4σ5.4𝜎5.4\,\sigma5.4 italic_σ, 2.2σ2.2𝜎2.2\,\sigma2.2 italic_σ, and 3.5σ3.5𝜎3.5\,\sigma3.5 italic_σ, respectively. The statistical significance of these shifts is reduced when the cosmological model used to fit the data allows the w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and wasubscript𝑤𝑎w_{a}italic_w start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT parameters to take values different from ΛΛ\Lambdaroman_ΛCDM, however. The inclusion of these two additional parameters degrades the constraining power and leads to a less evident shift than in the fiducial model.

We included WL and XC in the analysis to perform a 3×\times×2pt analysis. In this case, we found a non-trivial effect on the significance of the RSD contribution. On the one hand, WL contributes to tighten the constraints that can be achieved with Euclid, thus potentially increasing the significance of the shifts that are obtained when RSD is neglected. On the other hand, the theoretical predictions for this probe are not biased by this approximation (although the XC is still biased), and the inclusion of WL can therefore drag the recovered posterior distribution towards the fiducial values even when RSD are neglected. We found that for parameters such as nssubscript𝑛sn_{\rm s}italic_n start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT and hhitalic_h, where galaxy GC dominate, WL simply improve the constraining power, and the bias on these parameters increases in significance. In contrast, the parameters that are mostly constrained by WL, such as Ωm,0subscriptΩm0\varOmega_{{\rm m},0}roman_Ω start_POSTSUBSCRIPT roman_m , 0 end_POSTSUBSCRIPT and σ8subscript𝜎8\sigma_{8}italic_σ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT, we found that their shifts decrease in the optimistic case, where the constraining power of WL dominates the constraining power from GC.

To summarise, we found that when the contribution of linear RSD is not included in the theoretical predictions for GC angular power spectra, it can significantly reduce the accuracy but not the precision of the constraints that can be achieved by Euclid. The reason is that the linear RSD contribute at scales <100100\ell<100roman_ℓ < 100, which is a cosmic-variance dominated regime that yields no gain in cosmological information. It is important to note that the necessity of including the effect of linear RSD in GC in order to avoid cosmology biases has been studied in depth (Scharf et al., 1994; Heavens & Taylor, 1995; Padmanabhan et al., 2007; Blake et al., 2007; Nock et al., 2010; Crocce et al., 2011; Balaguera-Antolínez et al., 2018; Abbott et al., 2022) and the findings of our work agree with this. However, we have demonstrated that the approach of Tanidis & Camera (2019), which is a fast and approximated way to account for the linear RSD correction, can easily be implemented and tested within a parameter estimation pipeline in the modelling of Euclid photometric observables. In addition to this, work is ongoing to improve the modelling and study the effect of the non-linear galaxy bias in the photometric GC, which becomes especially important for Stage-IV galaxy surveys such as Euclid.

Acknowledgements.
KT is supported by the STFC grant ST/W000903/1 and by the European Structural and Investment Fund. For most of the development of this project KT was supported by the Czech Ministry of Education, Youth and Sports (Project CoGraDS - CZ.02.1.01/0.0/0.0/15_003/0000437). V.C and M.M. acknowledge funding by the Agenzia Spaziale Italiana (asi) under agreement no. 2018-23-HH.0 and support from INFN/Euclid Sezione di Roma. IT acknowledges funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant agreement No. 863929; project title “Testing the law of gravity with novel large-scale structure observables”). SC acknowledges support from the ‘Departments of Excellence 2018-2022’ Grant (L. 232/2016) awarded by the Italian Ministry of University and Research (mur). The Euclid Consortium acknowledges the European Space Agency and a number of agencies and institutes that have supported the development of Euclid, in particular the Academy of Finland, the Agenzia Spaziale Italiana, the Belgian Science Policy, the Canadian Euclid Consortium, the French Centre National d’Etudes Spatiales, the Deutsches Zentrum für Luft- und Raumfahrt, the Danish Space Research Institute, the Fundação para a Ciência e a Tecnologia, the Ministerio de Ciencia e Innovación, the National Aeronautics and Space Administration, the National Astronomical Observatory of Japan, the Netherlandse Onderzoekschool Voor Astronomie, the Norwegian Space Agency, the Romanian Space Agency, the State Secretariat for Education, Research and Innovation (SERI) at the Swiss Space Office (SSO), and the United Kingdom Space Agency. A complete and detailed list is available on the Euclid web site (http://www.euclid-ec.org).

References

  • Abbott et al. (2018) Abbott, T. M. C., Abdalla, F. B., Annis, J., et al. 2018, MNRAS, 480, 3879
  • Abbott et al. (2022) Abbott, T. M. C., Aguena, M., Alarcon, A., et al. 2022, Phys. Rev. D, 105, 023520
  • Abdalla et al. (2015) Abdalla, F. B., Bull, P., Camera, S., et al. 2015, in Advancing Astrophysics with the Square Kilometre Array (AASKA14), 17
  • Alsing et al. (2015) Alsing, J., Kirk, D., Heavens, A., & Jaffe, A. H. 2015, MNRAS, 452, 1202
  • Amendola et al. (2018) Amendola, L., Appleby, S., Avgoustidis, A., et al. 2018, Living Rev. Rel., 21, 2
  • Amendola et al. (2013) Amendola, L., Appleby, S., Bacon, D., et al. 2013, Living Rev. Rel., 16, 6
  • Andrade-Oliveira et al. (2021) Andrade-Oliveira, F., Camacho, H., Faga, L., et al. 2021, MNRAS, 505, 5714
  • Balaguera-Antolínez et al. (2018) Balaguera-Antolínez, A., Bilicki, M., & Branchini, E. 2018, MNRAS., 476, 1050
  • Behroozi et al. (2013) Behroozi, P. S., Wechsler, R. H., & Wu, H.-Y. 2013, ApJ, 762, 109
  • Bird et al. (2012) Bird, S., Viel, M., & Haehnelt, M. G. 2012, MNRAS, 420, 2551
  • Blake et al. (2007) Blake, C., Collister, A., & Bridle, S. 2007, MRAS, 374, 1527
  • Blanton et al. (2003) Blanton, M. R., Hogg, D. W., Bahcall, N. A., et al. 2003, ApJ, 594, 186
  • Blanton et al. (2005) Blanton, M. R., Lupton, R. H., Schlegel, D. J., et al. 2005, ApJ, 631, 208
  • Blanton et al. (2005) Blanton, M. R., Schlegel, D. J., Strauss, M. A., et al. 2005, AJ, 129, 2562
  • Bonvin & Durrer (2011) Bonvin, C. & Durrer, R. 2011, Phys. Rev., D84, 063505
  • Brown et al. (2015) Brown, M., Bacon, D., Camera, S., et al. 2015, in Advancing Astrophysics with the Square Kilometre Array (AASKA14), 23
  • Bull et al. (2015) Bull, P., Camera, S., Raccanelli, A., et al. 2015, in Advancing Astrophysics with the Square Kilometre Array (AASKA14), 24
  • Camera et al. (2015a) Camera, S., Maartens, R., & Santos, M. G. 2015a, MNRAS, 451, L80
  • Camera et al. (2015b) Camera, S., Raccanelli, A., Bull, P., et al. 2015b, in Advancing Astrophysics with the Square Kilometre Array (AASKA14), 25
  • Carretero et al. (2015) Carretero, J., Castander, F. J., Gaztañaga, E., Crocce, M., & Fosalba, P. 2015, MNRAS, 447, 646
  • Cepa, J. (2004) Cepa, J. 2004, A&A, 422, 831
  • Challinor & Lewis (2011) Challinor, A. & Lewis, A. 2011, Phys. Rev., D84, 043516
  • Chevallier & Polarski (2001) Chevallier & Polarski. 2001, International Journal of Modern Physics D, 10, 213
  • Costille et al. (2018) Costille, A., Caillat, A., Rossin, C., et al. 2018, in Space Telescopes and Instrumentation 2018: Optical, Infrared, and Millimeter Wave, ed. M. Lystrup, H. A. MacEwen, G. G. Fazio, N. Batalha, N. Siegler, & E. C. Tong, Vol. 10698, International Society for Optics and Photonics (SPIE), 730
  • Crocce et al. (2011) Crocce, M., Cabré, A., & Gaztañaga, E. 2011, MNRAS, 414, 329
  • Cropper et al. (2018) Cropper, M., Pottinger, S., Azzollini, R., et al. 2018, in Space Telescopes and Instrumentation 2018: Optical, Infrared, and Millimeter Wave, ed. M. Lystrup, H. A. MacEwen, G. G. Fazio, N. Batalha, N. Siegler, & E. C. Tong, Vol. 10698, International Society for Optics and Photonics (SPIE), 709
  • De Vicente et al. (2016) De Vicente, J., Sánchez, E., & Sevilla-Noarbe, I. 2016, MNRAS, 459, 3078
  • DES Collaboration: Abbott et al. (2005) DES Collaboration: Abbott, T., Aldering, G., Annis, J., et al. 2005 [arXiv:astro-ph/0510346]
  • DESI Collaboration: Aghamousa et al. (2016a) DESI Collaboration: Aghamousa, A., Aguilar, J., Ahlen, S., et al. 2016a [arXiv:1611.00036]
  • DESI Collaboration: Aghamousa et al. (2016b) DESI Collaboration: Aghamousa, A., Aguilar, J., Ahlen, S., et al. 2016b [arXiv:1611.00037]
  • Desjacques et al. (2018) Desjacques, V., Jeong, D., & Schmidt, F. 2018, Physics Reports, 733, 1
  • Doré et al. (2014) Doré, O., Bock, J., Ashby, M., et al. 2014 [arXiv:1412.4872]
  • Doré et al. (2018) Doré, O., Werner, M. W., Ashby, M. L. N., et al. 2018 [arXiv:1805.05489]
  • Elvin-Poole et al. (2018) Elvin-Poole, J., Crocce, M., Ross, A. J., et al. 2018, Phys. Rev., D98, 042006
  • Euclid Collaboration: Ajani et al. (2023) Euclid Collaboration: Ajani, V., Baldi, M., Barthelemy, A., et al. 2023, A&A, 675, A120
  • Euclid Collaboration: Blanchard et al. (2020) Euclid Collaboration: Blanchard, A., Camera, S., Carbone, C., et al. 2020, A&A, 642, A191
  • Euclid Collaboration: Pocino et al. (2021) Euclid Collaboration: Pocino, A., Tutusaus, I., Castander, F. J., et al. 2021, A&A, 655, A44
  • Foreman-Mackey et al. (2013) Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, Publ. Astron. Soc. Pac., 125, 306
  • Heavens et al. (2013) Heavens, A., Alsing, J., & Jaffe, A. H. 2013, MNRAS, 433, L6
  • Heavens & Taylor (1995) Heavens, A. F. & Taylor, A. N. 1995, MNRAS, 275, 483
  • Howlett et al. (2012) Howlett, C., Lewis, A., Hall, A., & Challinor, A. 2012, JCAP, 2012, 027
  • Ivezić et al. (2019) Ivezić, v., Kahn, S. M., Tyson, A. J., et al. 2019, ApJ., 873, 111
  • Joachimi et al. (2015) Joachimi, B., Cacciato, M., Kitching, T. D., et al. 2015, Space Science Reviews, 193, 1
  • Kaiser (1987) Kaiser, N. 1987, MNRAS, 227, 1
  • Kaiser (1992) Kaiser, N. 1992, ApJ, 388, 272
  • Laureijs et al. (2011) Laureijs, R., Amiaux, J. Arduini, S., et al. 2011 [arXiv:1110.3193]
  • Lepori et al. (2022) Lepori, F., Tutusaus, I., Viglione, C., et al. 2022, A&A, 662, A93
  • Lewis et al. (2000) Lewis, A., Challinor, A., & Lasenby, A. 2000, ApJ, 538, 473
  • Linder (2003) Linder, E. V. 2003, Phys. Rev. Lett., 90, 091301
  • LSST Dark Energy Science Collaboration: Mandelbaum et al. (2018) LSST Dark Energy Science Collaboration: Mandelbaum, R., Eifler, T., Hložek, R., et al. 2018 [arXiv:1809.01669]
  • LSST Science Collaboration: Abell et al. (2009) LSST Science Collaboration: Abell, P., A., Allison, J., Anderson, S., F., et al. 2009 [arXiv:0912.0201]
  • Martinelli et al. (2022) Martinelli, M., Dalal, R., Majidi, F., et al. 2022, MNRAS, 510, 1964
  • Massey et al. (2012) Massey, R., Hoekstra, H., Kitching, T., et al. 2012, MNRAS, 429, 661
  • Nock et al. (2010) Nock, K., Percival, W. J., & Ross, A. J. 2010, MNRAS, 407, 520
  • Padmanabhan et al. (2007) Padmanabhan, N., Schlegel, D. J., Seljak, U., et al. 2007, MNRAS, 378, 852
  • Potter et al. (2017) Potter, D., Stadel, J., & Teyssier, R. 2017, Computational Astrophysics and Cosmology, 4, 2
  • Raccanelli et al. (2015) Raccanelli, A., Bull, P., Camera, S., et al. 2015, in Advancing Astrophysics with the Square Kilometre Array (AASKA14), 31
  • Santos et al. (2015) Santos, M., Bull, P., Alonso, D., et al. 2015, in Advancing Astrophysics with the Square Kilometre Array (AASKA14), 19
  • Scharf et al. (1994) Scharf, C. A., Fisher, K. B., & Lahav, O. 1994, MRAS, 266, 219
  • Schneider et al. (2010) Schneider, P., Eifler, T., & Krause, E. 2010, A&A, 520, A116
  • SKA Cosmology Science Working Group (2020) SKA Cosmology Science Working Group. 2020, PASA, 37, e007
  • Smith et al. (2003) Smith, R. E., Peacock, J. A., Jenkins, A., et al. 2003, MNRAS, 341, 1311
  • Spergel et al. (2015) Spergel, D., Gehrels, N., Baltay, C., et al. 2015 [arXiv:1503.03757]
  • Szalay et al. (1998) Szalay, A. S., Matsubara, T., & Landy, S. D. 1998, ApJ, 498, L1
  • Sánchez et al. (2016) Sánchez, A. G., Scoccimarro, R., Crocce, M., et al. 2016, MNRAS, 464, 1640
  • Takahashi et al. (2012) Takahashi, R., Sato, M., Nishimichi, T., Taruya, A., & Oguri, M. 2012, ApJ, 761, 152
  • Tanidis & Camera (2019) Tanidis, K. & Camera, S. 2019, MNRAS, 489, 3385
  • Tanidis et al. (2019) Tanidis, K., Camera, S., & Parkinson, D. 2019, MNRAS, 491, 4869
  • Yoo (2010) Yoo, J. 2010, Phys. Rev., D82, 083508
  • Zehavi et al. (2011) Zehavi, I., Zheng, Z., Weinberg, D. H., et al. 2011, ApJ, 736, 59
  • Zuntz et al. (2015) Zuntz, J., Paterno, M., Jennings, E., et al. 2015, Astron. Comput., 12, 45