[go: up one dir, main page]

0% found this document useful (0 votes)
39 views416 pages

The Cosmoverse White Paper: Addressing Observational Tensions in Cosmology With Systematics and Fundamental Physics

The CosmoVerse White Paper discusses the standard model of cosmology and its observational tensions, particularly regarding the measurement of core cosmological parameters. It highlights the need for addressing these discrepancies through systematic investigations and potential new physics, as well as the importance of novel data analysis methods. The document concludes with recommendations for future research directions in observational cosmology and fundamental physics over the next decade.

Uploaded by

Teddy Filho
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
39 views416 pages

The Cosmoverse White Paper: Addressing Observational Tensions in Cosmology With Systematics and Fundamental Physics

The CosmoVerse White Paper discusses the standard model of cosmology and its observational tensions, particularly regarding the measurement of core cosmological parameters. It highlights the need for addressing these discrepancies through systematic investigations and potential new physics, as well as the importance of novel data analysis methods. The document concludes with recommendations for future research directions in observational cosmology and fundamental physics over the next decade.

Uploaded by

Teddy Filho
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 416

The CosmoVerse White Paper: Addressing observational tensions in cosmology with

systematics and fundamental physics

The CosmoVerse Network


(April 3, 2025)
The standard model of cosmology has provided a good phenomenological description of a wide
range of observations both at astrophysical and cosmological scales for several decades. This concor-
dance model is constructed by a universal cosmological constant and supported by a matter sector
described by the standard model of particle physics and a cold dark matter contribution, as well
as very early-time inflationary physics, and underpinned by gravitation through general relativity.
There have always been open questions about the soundness of the foundations of the standard
model. However, recent years have shown that there may also be questions from the observational
sector with the emergence of differences between certain cosmological probes. In this White Paper,
we identify the key objectives that need to be addressed over the coming decade together with the
core science projects that aim to meet these challenges. These discordances primarily rest on the
divergence in the measurement of core cosmological parameters with varying levels of statistical
confidence. These possible statistical tensions may be partially accounted for by systematics in
various measurements or cosmological probes but there is also a growing indication of potential new
arXiv:2504.01669v1 [astro-ph.CO] 2 Apr 2025

physics beyond the standard model. After reviewing the principal probes used in the measurement
of cosmological parameters, as well as potential systematics, we discuss the most promising array of
potential new physics that may be observable in upcoming surveys. We also discuss the growing set
of novel data analysis approaches that go beyond traditional methods to test physical models. These
new methods will become increasingly important in the coming years as the volume of survey data
continues to increase, and as the degeneracy between predictions of different physical models grows.
There are several perspectives on the divergences between the values of cosmological parameters,
such as the model-independent probes in the late Universe and model-dependent measurements in
the early Universe, which we cover at length. The White Paper closes with a number of recom-
mendations for the community to focus on for the upcoming decade of observational cosmology,
statistical data analysis, and fundamental physics developments.
2

Editors1 : Eleonora Di Valentino2 , Jackson Levi Said3

Forward Writers: Adam Riess, Agnieszka Pollo, and Vivian Poulin

Section Coordinators: Adrià Gómez-Valent, Amanda Weltman, Antonella Palmese, Caroline D. Huang, Carsten
van de Bruck, Chandra Shekhar Saraf, Cheng-Yu Kuo, Cora Uhlemann, Daniela Grandón, Dante Paz, Dominique
Eckert, Elsa M. Teixeira, Emmanuel N. Saridakis, Eoin Ó Colgáin, Florian Beutler, Florian Niedermann, Francesco
Bajardi, Gabriela Barenboim, Giulia Gubitosi, Ilaria Musella, Indranil Banik, Istvan Szapudi, Jack Singal, Jaume
Haro Cases, Jens Chluba, Jesús Torrado, Jurgen Mifsud, Karsten Jedamzik, Khaled Said, Konstantinos Dialektopou-
los, Laura Herold, Leandros Perivolaropoulos, Lei Zu, Lluís Galbany, Louise Breuval, Luca Visinelli, Luis A. Escamilla,
Luis A. Anchordoqui, M.M. Sheikh-Jabbari, Margherita Lembo, Maria Giovanna Dainotti, Maria Vincenzi, Marika
Asgari, Martina Gerbino, Matteo Forconi, Michele Cantiello, Michele Moresco, Micol Benetti, Nils Schöneberg, Özgür
Akarsu, Rafael C. Nunes, Reginald Christian Bernardo, Ricardo Chávez, Richard I. Anderson, Richard Watkins,
Salvatore Capozziello, Siyang Li, Sunny Vagnozzi, Supriya Pan, Tommaso Treu, Vid Irsic, Will Handley, William
Giarè, and Yukei Murakami

Section Contributors: Adèle Poudou, Alan Heavens, Alan Kogut, Alba Domi, Aleksander Łukasz Lenart, Alessan-
dro Melchiorri, Alessandro Vadalà, Alexandra Amon, Alexander Bonilla, Alexander Reeves, Alexander Zhuk, Alfio
Bonanno, Ali Övgün, Alice Pisani, Alireza Talebian, Amare Abebe, Amin Aboubrahim, Ana Luisa González Morán,
András Kovács, Andreas Papatriantafyllou, Andrew R. Liddle, Andronikos Paliathanasis, Andrzej Borowiec, Anil
Kumar Yadav, Anita Yadav, Anjan Ananda Sen, Anjitha John William Mini Latha, Anne Christine Davis, Anowar
J. Shajib, Anthony Walters, Anto Idicherian Lonappan, Anton Chudaykin, Antonio Capodagli, Antonio da Silva,
Antonio De Felice, Antonio Racioppi, Araceli Soler Oficial, Ariadna Montiel, Arianna Favale, Armando Bernui,
Arrianne Crystal Velasco, Asta Heinesen, Athanasios Bakopoulos, Athanasios Chatzistavrakidis, Bahman Khanpour,
Bangalore S. Sathyaprakash, Bartek Zgirski, Benjamin L’Huillier, Benoit Famaey, Bhuvnesh Jain, Biesiada Marek,
Bing Zhang, Biswajit Karmakar, Branko Dragovich, Brooks Thomas, Carlos Correa, Carlos G. Boiza, Catarina
Marques, Celia Escamilla-Rivera, Charalampos Tzerefos, Chi Zhang, Chiara De Leo, Christian Pfeifer, Christine
Lee, Christo Venter, Cláudio Gomes, Clecio Roque De bom, Cristian Moreno-Pulido, Damianos Iosifidis, Dan Grin,
Daniel Blixt, Dan Scolnic, Daniele Oriti, Daria Dobrycheva, Dario Bettoni, David Benisty, David Fernández-Arenas,
David L. Wiltshire, David Sanchez Cid, David Tamayo, David Valls-Gabaud, Davide Pedrotti, Deng Wang, Denitsa
Staicova, Despoina Totolou, Diego Rubiera-Garcia, Dinko Milaković, Dom Pesce, Dominique Sluse, Duško Borka,
Ebrahim Yusofi, Elena Giusarma, Elena Terlevich, Elena Tomasetti, Elias C. Vagenas, Elisa Fazzari, Elisa G. M.
Ferreira, Elvis Barakovic, Emanuela Dimastrogiovanni, Emil Brinch Holm, Emil Mottola, Emre Özülker, Enrico
Specogna, Enzo Brocato, Erik Jensko, Erika Antonette Enriquez, Esha Bhatia, Fabio Bresolin, Felipe Avila, Filippo
Bouchè, Flavio Bombacigno, Fotios K. Anagnostopoulos, Francesco Pace, Francesco Sorrenti, Francisco S. N. Lobo,
Frédéric Courbin, Frode K. Hansen, Greg Sloan, Gabriel Farrugia, Gabriel Lynch, Gabriela Garcia-Arroyo, Gabriella
Raimondo, Gaetano Lambiase, Gagandeep S. Anand, Gaspard Poulot, Genly Leon, Gerasimos Kouniatalis, Ger-
mano Nardini, Géza Csörnyei, Giacomo Galloni, Giada Bargiacchi, Giannis Papagiannopoulos, Giovanni Montani,
Giovanni Otalora, Giulia De Somma, Giuliana Fiorentino, Giuseppe Fanizza, Giuseppe Gaetano Luciano, Giuseppe
Sarracino, Gonzalo J. Olmo, Goran S. Djordjević, Guadalupe Cañas-Herrera, Hanyu Cheng, Harry Desmond, Hassan
Abdalla, Houzun Chen, Hsu-Wen Chiang, Hume A. Feldman, Hussain Gohar, Ido Ben-Dayan, Ignacio Sevilla-Noarbe,
Ignatios Antoniadis, Ilim Cimdiker, Inês S. Albuquerque, Ioannis D. Gialamas, Ippocratis Saltas, Iryna Vavilova,
Isidro Gómez-Vargas, Ismael Ayuso, Ismailov Nariman Zeynalabdi, Ivan De Martino, Ivonne Zavala Carrasco, J.
Alberto Vázquez, Jacobo Asorey, Janusz Gluza, Javier Rubio, Jenny G. Sorce, Jenny Wagner, Jeremy Sakstein,
Jessica Santiago, Jim Braatz, Joan Solà Peracaula, John Blakeslee, John Webb, Jose A. R. Cembranos, José Pedro
Mimoso, Joseph Jensen, Juan García-Bellido, Judit Prat, Kathleen Sammut, Kay Lehnert, Keith R. Dienes, Kishan
Deka, Konrad Kuijken, Krishna Naidoo, László Árpád Gergely, Laur Järv, Laura Mersini-Houghton, Leila L. Graef,
Léo Vacher, Levon Pogosian, Lilia Anguelova, Lindita Hamolli, Lu Yin, Luca Caloni, Luca Izzo, Lucas Macri , Luis
E. Padilla, Luz Ángela García, Maciej Bilicki, Mahdi Najafi, Manolis Plionis, Manuel Gonzalez-Espinoza, Manuel
Hohmann, Marcel A. van der Westhuizen, Marcella Marconi, Marcin Postolak, Marco de Cesare, Marco Regis, Marek
Biesiada, Maret Einasto, Margus Saal, Maria Caruana, Maria Petronikolou, Mariam Bouhmadi-López, Mariana Melo,
Mariaveronica De Angelis, Marie-Noëlle Célérier, Marina Cortês, Mark Reid, Markus Michael Rau, Martin S. Sloth,
Martti Raidal, Masahiro Takada, Masoume Reyhani, Massimiliano Romanello, Massimo Marengo, Mathias Garny,
Matías Leizerovich, Matteo Martinelli, Matteo Tagliazucchi, Mehmet Demirci, Miguel A. S. Pinto, Miguel A. Sabogal,

1 Detailed author information at the end of the document and lists alphabetized by first name
2 Email: e.divalentino@sheffield.ac.uk
3 Email: jackson.said@um.edu.mt
3

Miguel A. García-Aspeitia , Milan Milošević, Mina Ghodsi, Mustapha Ishak, Nelson J. Nunes, Nick Samaras, Nico
Hamaus, Nico Schuster, Nicola Borghi, Nicola Deiosso, Nicola Tamanini, Nicolao Fornengo, Nihan Katırcı, Nikolaos E.
Mavromatos, Nikolaos Petropoulos, Nikolina Šarčević, Nils A. Nilsson, Noemi Frusciante, Octavian Postavaru, Oem
Trivedi, Oleksii Sokoliuk, Olga Mena, Paloma Morilla, Paolo Campeti, Paolo Salucci, Paula Boubel, Paweł Bielewicz,
Pekka Heinämäki, Petar Suman, Petros Asimakis, Pierros Ntelis, Pilar Ruiz-Lapuente, Pran Nath, Predrag Jovanović,
Purba Mukherjee, Radosław Wojtak, Rafaela Gsponer, Rafid H. Dejrah, Rahul Shah, Rasmi Hajjar, Rebecca Briffa,
Rebecca Habas, Reggie C. Pantig, Renier Mendoza, Riccardo Della Monica, Richard Stiskalek, Rishav Roshan, Rita
B. Neves, Roberto Molinaro, Roberto Terlevich, Rocco D’Agostino, Rodrigo Sandoval-Orozco, Ronaldo C. Batista,
Ruchika Kaushik, Ruth Lazkoz, Saeed Rastgoo, Sahar Mohammadi, Salvatore Samuele Sirletti, Sandeep Haridasu,
Sanjay Mandal, Saurya Das, Sebastian Bahamonde, Sebastian Grandis, Sebastian Trojanowski, Sergei D. Odintsov,
Sergij Mazurenko, Shahab Joudaki, Sherry H. Suyu, Shouvik Roy Choudhury, Shruti Bhatporia, Shun-Sheng Li,
Simeon Bird, Simon Birrer, Simone Paradiso, Simony Santos da Costa, Sofia Contarini, Sophie Henrot-Versillé, Spy-
ros Basilakos, Stefano Casertano, Stefano Gariazzo, Stylianos A. Tsilioukas, Surajit Kalita, Suresh Kumar, Susana J.
Landau, Sveva Castello, Swayamtrupta Panda, Tanja Petrushevska, Thanasis Karakasis, Thejs Brinckmann, Tiago
B. Gonçalves, Tiziano Schiavone, Tom Abel, Tomi Koivisto, Torsten Bringmann, Umut Demirbozan, Utkarsh Kumar,
Valerio Marra, Maurice H.P.M. van Putten, Vasileios Kalaitzidis, Vasiliki A. Mitsou, Vasilios Zarikas, Vedad Pasic,
Venus Keus , Verónica Motta, Vesna Borka Jovanović, Víctor H. Cárdenas, Vincenzo Ripepi, Vincenzo Salzano,
Violetta Impellizzeri, Vitor da Fonseca, Vittorio Ghirardini, Vladas Vansevičius, Weiqiang Yang, Wojciech Hellwing,
Xin Ren, Yu-Min Hu, and Yuejia Zhai

Endorsers: Abdul Malik Sultan, Adrienn Pataki, Alessandro Santoni, Aliya Batool, Aneta Wojnar, Arman Tursunov,
Avik De, Ayush Hazarika, Baojiu Li, Benjamin Bose, Bivudutta Mishra, Bobomurat Ahmedov, Chandra Shekhar
Saraf, Claudia Scóccola, Crescenzo Tortora, D’Arcy Kenworthy, Daniel E. Holz, David F. Mota, David S. Pereira,
Devon M. Williams, Dillon Brout, Dong Ha Lee, Eduardo Guendelman, Edward Olex, Emanuelly Silva, Emre Onur
Kahya, Enzo Brocato, Eva-Maria Mueller, Felipe Andrade-Oliveira, Feven Markos Hunde, F. R. Joaquim, Florian
Pacaud, Francis-Yan Cyr-Racine, Pozo Nuñez, F, Gábor Rácz, Gene Carlo Belinario, Geraint F. Lewis, Gergely Dálya,
Giorgio Laverda, Guido Risaliti, Guillermo Franco-Abellán, Hayden Zammit, Hayley Camilleri, Helene M. Courtois,
Hooman Moradpour, Igor de Oliveira Cardoso Pedreira, Ilídio Lopes, István Csabai, James W. Rohlf, J. Bogdanoska,
Javier de Cruz Pérez, Joan Bachs-Esteban, Joseph Sultana , Julien Lesgourgues, Jun-Qian Jiang, Karem Peñaló
Castillo, Lavinia Heisenberg, Laxmipriya Pati, Léon V.E. Koopmans, Lokesh kumar Duchaniya, Lucas Lombriser,
María Pérez Garrote, Mariano Domínguez, Marine Samsonyan, Mark Pace, Martin Krššák, Masroor C. Pookkillath,
Matteo Peronaci, Matteo Piani, Matthildi Raftogianni , Meet J. Vyas, Melina Michalopoulou, Merab Gogberashvili,
Michael Klasen, Michele Cicoli, Michele Moresco, Miguel Quartin, Miguel Zumalacárregui, Milan S. Dimitrijević, Milos
Dordevic, Mindaugas Karčiauskas, Morgan Le Delliou, Nastassia Grimm, Nicolás Augusto Kozameh, Nicoleta Voicu,
Nicolina Pop, Nikos Chatzifotis, Oliver Fabio Piattella, Paula Boubel, Pedro da Silveira Ferreira, Péter Raffai, Peter
Schupp, Pierros Ntelis, Pradyumn Kumar Sahoo, Roberto V. Maluf, Ruth Durrer, S. A. Kadam, Sabino Matarrese,
Samuel Brieden, Santiago González-Gaitán, Santosh V. Lohakare, Scott Watson, Shao-Jiang Wang, Simão Marques
Nunes, Soumya Chakrabarti, Suvodip Mukherjee, Tajron Jurić, Tessa Baker, Theodoros Nakas, Tiago Barreiro, Upala
Mukhopadhyay, Veljko Vujčić, Violetta Sagun, Vladimir A. Srećković, Wangzheng Zhang, Yo Toda, Yun-Song Piao,
and Zahra Davari
4

Contents

Executive Summary 7

1. Introduction 8
1.1. State of the art status of cosmological tensions 8
1.2. Early vs local measurements and efforts for a solution 12
1.3. Additional curiosities and anomalies 17
1.4. Data analysis: How to tackle the problem 19

2. Observational cosmology and systematics 22


2.1. H0 Tension: Measurements and systematics 22
2.1.1. The distance ladder 22
2.1.2. Cepheid variables as standard candles 24
2.1.3. Maser driven constraints 27
2.1.4. On the tip of the red giant branch method 28
2.1.5. The surface brightness fluctuations method 33
2.1.6. Mira variables 35
2.1.7. Type Ia supernovae 37
2.1.8. J-regions of the asymptotic giant branch methods 40
2.1.9. The Hubble constant from Type II supernovae 41
2.1.10. HII galaxy distance indicators 42
2.1.11. The baryonic Tully-Fisher relation approach 45
2.1.12. The Hubble tension in our Backyard: DESI and the nearness of the Coma cluster 47
2.1.13. Cosmic chronometers 48
2.1.14. Strong lensing and time delay measurements 49
2.1.15. Extended QSO cosmologies 50
2.1.16. GRBs as cosmological standard candles 53
2.1.17. Gravitational wave constraints 55
2.1.18. The cosmic microwave background radiation 56
2.1.19. Baryonic acoustic oscillations 62
2.2. S8 Tension: measurements and systematics 67
2.2.1. Weak lensing 67
2.2.2. Galaxy cluster counts 74
2.2.3. Galaxy Clustering – Other probes 81
2.3. Other challenges 83
2.3.1. Systematics and the Alens parameter 83
2.3.2. Evidence for a nonvanishing Ωk 84
2.3.3. Anisotropic anomalies in the cosmic microwave background radiation 85
2.3.4. Hints of dynamical dark energy in DESI baryon acoustic oscillations and beyond 88
2.3.5. Neutrino tensions 90
2.3.6. Cosmic dipoles 92
2.3.7. Big bang nucleosynthesis 96
2.3.8. Anomalies with Lyman-α measurements 97
2.3.9. Cosmic superstructures and the ISW anomalies 99
2.3.10. JWST anomalies 102
2.3.11. Cosmic voids 104
2.3.12. Fast radio burst probes of cosmic tensions 105
2.3.13. Radio background excess 108
2.3.14. Tension between the large scale bulk flow and the standard cosmological model 110
2.3.15. Ultra long period cepheids 110

3. Data analysis in cosmology 113


3.1. Cosmology simulators and Markov chain Monte Carlo approaches 114
3.1.1. Markov chain Monte Carlo sampling algorithms 114
3.1.2. Hamiltonian Monte Carlo 116
3.1.3. Nested sampling 116
3.1.4. Practical considerations for choosing an MC sampler 117
5

3.1.5. Cosmological inference frameworks 118


3.1.6. Modeling problems: Prior dependence and projection effects 118
3.1.7. Model comparison criteria 119
3.1.8. Tension metrics 120
3.2. Machine learning based inference techniques 122
3.2.1. Artificial neural networks 123
3.2.2. Convolutional neural networks 124
3.2.3. Bayesian neural networks 126
3.2.4. Deep learning 129
3.2.5. Conclusion 131
3.3. Reconstruction techniques 131
3.3.1. Parametric reconstructions 132
3.3.2. Phenomenological parameterizations 132
3.3.3. Physically motivated parameterizations 133
3.3.4. Model-independent parameterizations 134
3.3.5. Cosmography 136
3.3.6. Non-parametric methods 136
3.4. Bio-inspired algorithms in model selection 138
3.4.1. Genetic algorithm 139
3.4.2. Machine learning and GA variants 140
3.4.3. GA for cosmology 141
3.4.4. GA for cosmological tensions 142
3.5. Inference from cosmological simulations 144
3.5.1. Cosmological simulation 144
3.5.2. Nonlinear effects 144
3.5.3. Cosmological simulations beyond ΛCDM 145
3.5.4. Cosmological simulations in nonlinear observation and cosmic tension 147
3.6. Profile likelihoods in cosmology 149
3.6.1. Motivation 149
3.6.2. Confidence intervals from Profile Likelihoods 149
3.6.3. Calculation of profile likelihoods 150
3.6.4. Applications of profile likelihood in cosmology 152

4. Fundamental physics 154


4.1. Early-time proposals 155
4.1.1. Early dark energy and variants 155
4.1.2. New early dark energy and variants 159
4.1.3. Extra relativistic degrees of freedom 162
4.2. Late-time proposals 165
4.2.1. Late dark energy 165
4.2.2. Dark energy models exhibiting a rapid density transition from negative to positive values in the
late Universe 171
4.2.3. Interacting dark energy 174
4.3. Modified gravity 178
4.3.1. Modified gravity in light of cosmic tensions 178
4.3.2. Early modifications to gravity 184
4.4. Matter sector solutions 188
4.4.1. Cold dark matter 188
4.4.2. Warm dark matter 192
4.4.3. Interacting and decaying dark matter 194
4.5. Other solutions 197
4.5.1. Estimates based on a possible local void 197
4.5.2. Primordial magnetic fields 199
4.5.3. Feasibility of inflationary models to ameliorate the cosmic tensions 201
4.6. The cosmological principle 204
4.6.1. FLRW spacetime 205
4.6.2. Moving beyond FLRW 207
4.6.3. Mc-Vittie spacetime 208
6

4.6.4. What needs to happen going forward 209


4.7. Quantum gravity phenomenology 210
4.7.1. QG modified gravitational dynamics 210
4.7.2. Quantum gravity effects on the physics of particles and fields 211
4.8. Varying fundamental constants and their role in the Hubble tension 212
4.9. Local New physics solutions to the Hubble and growth tensions 214

5. Discussion and future opportunities 217


5.1. Future survey prospects 217
5.1.1. Cosmic microwave background 217
5.1.2. Baryon acoustic oscillations 220
5.1.3. Weak lensing experiments 222
5.1.4. Gravitational waves as probes of cosmological tensions 224
5.1.5. 21 cm Cosmology 226
5.1.6. Type Ia supernovae and distance ladder 228
5.1.7. Time delay cosmography 230
5.1.8. Fast Radio Bursts (FRBs) 231
5.1.9. Line-intensity mapping 233
5.1.10. Potential new probes 235
5.2. Recommendations 236
5.3. Conclusions 245

6. Conventions 248

7. List of Acronyms 249

References 255
7

Executive Summary
The CosmoVerse network is born out of the CosmoVerse COST Action (formally CA21136 - Addressing observational
tensions in cosmology with systematics and fundamental physics [1, 2]), which traces its origins to the SNOWMASS
2021 effort in the cosmic tensions sector [3–7]. This will be one of the key deliverables of the COST Action and one
of its lasting legacies. The aim of this network is to establish synergy among researchers working across the disparate
themes of observational cosmology, novel techniques of data analysis, and fundamental physics. This White Paper
demonstrates how this effort has been successful, while also laying out a roadmap to sustain those efforts and expand
the interdisciplinary nature of the topic across different areas of the community. Another core aspect of this effort
is to harmonize approaches between the constituent subcommunities and to create a common language in which to
discuss the topic of cosmic tensions.
In the White Paper, the accomplishment of these goals is demonstrated through the interwoven connections between
the key communities of the network and the granular topics. This was achieved through a diverse set of actions,
including the CosmoVerse Seminar Series led by Eleonora Di Valentino [8], the CosmoVerse conferences and the
discussions and presentations involved in these events [9], the CosmoVerse School held in Corfu in 2024 [10], the
CosmoVerse Training Series, which involved a significant effort to bridge [11], as well as the CosmoVerse Journal
Club, led by Enrico Specogna and Mahdi Najafi [12].
The CosmoVerse White Paper, edited by Jackson Levi Said and Eleonora Di Valentino, is a collective effort from the
community to review the state of the art and identify opportunities to address tensions in cosmology over the coming
years. This includes progress in observational cosmology and the exhaustive treatment of potential systematics, the
development of new data analysis approaches for upcoming surveys and potential new physics models, as well as
the construction of cosmological models that appropriately address the areas where the concordance model exhibits
tensions. These topics are organized as follows:

• Section 1 - Introduction and state-of-the-art.


• Section 2 - Observational cosmology. Coordinators: Adam Riess, Amanda Weltman, Antonella Palmese, Caro-
line D. Huang, Chandra Shekhar Saraf, Cheng-Yu Kuo, Cora Uhlemann, Dan Scolnic, Daniela Grandón, Dante
Paz, Dominique Eckert, Florian Beutler, Gabriela Barenboim, Ilaria Musella, Istvan Szapudi, Jack Singal,
Khaled Said, Leandros Perivolaropoulos, Lluís Galbany, Louise Breuval, Louise Breuval, Maria Giovanna Dain-
otti, Maria Vincenzi, Marika Asgari, Martina Gerbino, Margherita Lembo, Matteo Forconi, Michele Cantiello,
Michele Moresco, Nils Schöneberg, Ricardo Chávez Murillo, Richard I. Anderson, Rick Watkins, Shahin Sheikh-
Jabbari, Siyang Li, Tommaso Treu, Vid Iršič, Will Handley, and William Giarè.
• Section 3 - Novel data analysis techniques. Coordinators: Agnieszka Pollo, Adrià Gómez-Valent, Daniela
Grandón, Jesus Torrado, Jurgen Mifsud, Lei Zu, Luis Escamilla, and Reginald Christian Bernardo.
• Section 4 - Fundamental physics. Coordinators: Vivian Poulin, Carsten van de Bruck, Elsa Teixeira, Em-
manuel N. Saridakis, Eoin O Colgain, Florian Niedermann, Francesco Bajardi, Giulia Gubitosi, Indranil Banik,
Jaime Haro Cases, Jens Chluba, Karsten Jedamzik, Konstantinos F. Dialektopoulos, Laura Herold, Lean-
dros Perivolaropoulos, Luca Visinelli, Luis Anchordoqui, Micol Benetti, Özgür Akarsu, Rafael Nunes, Sunny
Vagnozzi, Supriya Pan.
• Section 5 - Discussion, White Paper recommendations, and future prospects.

In every section, the coordinators and contributing writers are identified at the beginning of each contribution. The
preparation of the CosmoVerse White Paper involved a substantial number of people, with ∼ 65 coordinators, ∼ 350
contributing writers, and ∼ 130 endorsers. The project was reviewed by Alba Domi, Alexandra Amon, Anjitha John
William Mini Latha, Anton Chudaykin, Bivudutta Mishra, Emil Mottola, Emmanuel N. Saridakis, Florian Pacaud,
Germano Nardini, Marcin Postolak, Mariano Domínguez, Miguel A. García-Aspeitia, Nelson J. Nunes, Oem Trivedi,
Oliver Fabio Piattella, Özgür Akarsu, Paolo Salucci, Pilar Ruiz-Lapuente, Rafid H. Dejrah, and Supriya Pan, and
the bibliographic information was organized by Luca Visinelli. The notational conventions of the work are defined in
Sec. 6. This is followed by a list of the glossary abbreviations in Sec. 7. The CosmoVerse Action and White Paper
were aided by the administrative contributions of Hiba Wazaz.
8

1. Introduction
Cosmology has entered an era of unprecedented precision in and volume of observational measurements, with large-
scale surveys providing high-quality data across multiple redshifts and cosmic epochs. This wealth of observational
data allows for a deeper understanding of the Universe’s composition, dynamics, and structure formation processes
of the Universe. However, these advancements have also revealed significant tensions between early- and late-time
cosmological measurements, challenging the standard cosmological model. These discrepancies not only question the
consistency of this framework but also suggest the possibility of unrecognized systematic errors or the need for new
physics beyond the standard model of cosmology.
The standard model of cosmology, or ΛCDM describes the Universe using a cosmological constant (Λ) and Cold
Dark Matter (CDM), and gravity through Einstein’s General relativity (GR). It provides an excellent fit for a range
of cosmological datasets, including the Cosmic Microwave Background Radiation (CMB) and large-scale structure
surveys. Nevertheless, the emergence of tensions in key cosmological parameters has become increasingly difficult to
ignore. Among these, the most prominent include discrepancies in the measurements of the Hubble constant (H0 ), the
amplitude of matter fluctuations (S8 ), and the sound horizon at the epoch of Baryon Acoustic Oscillations (BAO).
These tensions raise profound questions about our understanding of the Universe’s expansion history, structure
formation, and the fundamental nature of Dark Matter (DM) and Dark Energy (DE). If not due to systematic errors,
such tensions may indicate the need for modifications of the standard model or additional components in the cosmic
inventory. Addressing these issues requires a careful consideration of both observational and theoretical aspects, as
well as a comprehensive approach that combines multiple cosmological probes.

1.1. State of the art status of cosmological tensions


The most statistically significant tension in cosmology is the H0 tension, which refers to a significant and persistent
discrepancy between measurements of the Hubble constant obtained from early- and late-time cosmological probes,
challenging the completeness of the standard ΛCDM model and suggesting the possible need for new physics. Early
Universe constraints, primarily from the Planck satellite (Sec. 2.1.18), which maps CMB anisotropies, provide a
precise estimate of H0 = 67.4 ± 0.5 km s−1 Mpc−1 . This result relies on the angular scale of the acoustic peaks in the
CMB power spectrum, under the assumptions of a flat ΛCDM model with standard radiation content. Consistency
with Planck’s result has also been demonstrated by ground-based experiments such as Atacama Cosmology Telescope
(ACT) and South Pole Telescope (SPT), both of which measure the damping tail and lensing-induced smoothing of
the CMB power spectrum with high precision, reinforcing the lower H0 value.
BAO (see Sec. 2.1.19) provide an additional, independent probe of the expansion history by measuring the char-
acteristic scale imprinted by sound waves in the early Universe, observable in the large-scale distribution of galaxies.
This standard ruler, linked to the sound horizon at the epoch of baryon drag, enables precise distance measurements
at various redshifts. Surveys such as Baryon Oscillation Spectroscopic Survey (BOSS), Extended Baryon Oscillation
Spectroscopic Survey (eBOSS), and The Dark Energy Spectroscopic Instrument (DESI) have reported H0 values
consistent with CMB constraints, further supporting the lower early Universe based estimate. Notably, DESI’s latest
BAO results, based on over six million galaxies across multiple redshift bins, when calibrated by the Planck + ΛCDM
constraint on the sound horizon measured H0 = 68.5 ± 0.6 km s−1 Mpc−1 .
Late-time measurements of the Hubble constant H0 using the distance ladder approach (Sec. 2.1.1) suggest a
higher value than early Universe constraints, contributing to the persistent Hubble tension. The most precise distance
ladder method involves three steps: (1) geometric distance measurements to calibrate the luminosities of Cepheid
variables using Gaia parallaxes, detached eclipsing binaries, and water masers; (2) using these calibrated Cepheids to
standardize Type Ia supernovae (SNIa) observed with Hubble Space Telescope (HST); and (3) measuring H0 from
SNIa distances in the Hubble flow, where cosmic expansion dominates.
Cepheid variables (Sec. 2.1.2) serve as primary standard candles due to their well-defined Period-Luminosity
(P-L) relation, where the pulsation period correlates with intrinsic brightness. Systematic uncertainties, such as
metallicity effects, crowding, and dust extinction, are mitigated through near-infrared photometry and consistent
datasets from HST and James Webb Space Telescope (JWST). The SH0ES collaboration recently measured H0 =
73.17 ± 0.86 km s−1 Mpc−1 using this approach, indicating a 5 − 6σ tension with Planck. JWST’s independent obser-
vations, particularly in Cepheid-rich fields, have validated and refined these results by reducing crowding biases and
confirming the reliability of the Cepheid calibration, strengthening the significance of the Hubble tension.
SNIa (Sec. 2.1.7) serves as the most common, far-field rung of the distance ladder, extending measurements into the
Hubble flow. The Pantheon+ dataset provides precise H0 measurements consistent with local results but in tension
with early Universe estimates.
While the above tools remain central to the Hubble tension, offering the strongest leverage, alternative standard
candles such as the Tip of the Red Giant Branch (TRGB), Mira variables, J-region Asymptotic Giant Branch (JAGB)
stars, Type II supernovae (SNII), Surface Brightness Fluctuations (SBF), and the Baryonic Tully-Fisher Relation
9

(BTFR) provide valuable cross-checks. These independent methods, while varying in calibration techniques and
stellar populations, consistently yield higher values of H0 than early Universe constraints from the CMB, emphasizing
the robustness of the tension and fueling interest in theoretical refinements.
The TRGB method (Sec. 2.1.4) measures distances using the sharp luminosity cutoff where QSO (QSO) stars ignite
helium as a standard candle. Calibrated with galaxies like the Magellanic Clouds and NGC 4258, TRGB provides
precise distance estimates. For state-of-the-art measurements, it yields consistent distance measures to SNIa hosts as
Cepheids. TRGB remains a valuable cross-check for the Hubble tension, with minimal sensitivity to metallicity (if
measured in the I-band) and dust.
Mira variables (Sec 2.1.6) are long-period pulsating stars used as standard candles for measuring H0 . Their period-
luminosity relation, especially in the near-infrared, provides reliable distances. Calibrated using galaxies like the Large
Magellanic Clouds (LMC) and NGC 4258, Miras offer an independent cross-check on H0 with minimal sensitivity to
metallicity and crowding.
JAGB stars (Sec. 2.1.8) are standard candles used for measuring H0 based on their narrow luminosity range in the
near-infrared. They are carbon-rich Asymptotic Giant Branch (AGB) stars in an advanced stellar phase, providing
distance estimates at long range. Calibrated using galaxies like NGC 4258, JAGB measurements offer a route to
calibrate SNIa. Though newer and less tested than Cepheids or TRGB, they contribute to cross-validation efforts.
SNII (Sec. 2.1.9) measure H0 using the correlation between their luminosity and decline rate during the plateau
phase. They offer an alternative to SNIa, aiding cross-checks in the Hubble tension.
The SBF method (Sec. 2.1.5) measures H0 using pixel-to-pixel luminosity variations in elliptical galaxies. Calibrated
with nearby galaxies, SBF provides H0 estimates which are independent of SNIa.
The BTFR (Sec. 2.1.11) estimates H0 using the correlation between a galaxy’s rotational velocity and baryonic
mass. Recent calibrations, particularly addressing zero-point uncertainties between different standard candles like
Cepheids and TRGB, have refined the measurement to H0 = 76.3 ± 2.1(stat) ±1.5(sys) km s−1 Mpc−1 , yielding
another SNIa-independent measurement and further emphasizing the tension with early Universe estimates.
As an alternative to the distance ladder measurements, masers offer independent insights into the Hubble constant
H0 . The maser technique (Sec. 2.1.3) uses 22 GHz H2 O maser emissions from rotating disks around supermassive
black holes to directly measure distances. Each of 5 masers (leaving NGC 4258 aside as its often used to calibrate
the distance ladder) yields a geometric distance and H0 = 73.9 ± 3.0 km s−1 Mpc−1 . The Megamaser Cosmology
Project (MCP) has extended the precision of this method, producing results consistent with late-time measurements.
Although rare, water masers provide a model-independent check, reinforcing the higher H0 values obtained from local
probes.
Strong gravitational lensing with time-delay measurements (Sec. 2.1.14) offers another independent method for
estimating H0 . Multiple images of a background source, produced by a massive foreground lens, create time delays
due to differences in the light paths. These delays depend on the lensing geometry and the Universe’s expansion rate.
Collaborations such as H0LiCOW and TDCOSMO have applied this technique, yielding H0 ≈ 74.2±2.6 km s−1 Mpc−1 ,
consistent with other late-time estimates.
Gravitational Wave (GW) events with electromagnetic counterparts, known as standard sirens (Sec. 2.1.17), provide
an independent estimate of H0 by measuring the luminosity distance from GWs emitted during compact object
mergers. When an electromagnetic counterpart identifies the host galaxy, the redshift can be measured directly.
Recent standard siren measurements yield H0 ≈ 70.0 ± 3.0 km s−1 Mpc−1 , offering a model-independent constraint
on cosmic expansion and contributing to the Hubble tension debate. However, this approach is still severely limited
by the rarity of GW events with EM counterparts, with only a single event and which is too close for a reliable
measurement of H0 .
Cosmic chronometers (CC) (Sec. 2.1.13) estimate H0 using the differential ages of passively evolving galaxies. By
dating stellar populations, this method infers the expansion rate without relying on standard candles. However, unlike
the use of distances, this measure is not empirical, depending on astrophysical modeling of aging stellar populations,
their metallicities, star formation histories, and initial mass functions, with an uncertainty dominated by the modeling
effort. Recent measurements suggest H0 ≈ 70.6 ± 6.7 km s−1 Mpc−1 , still consistent with both early and most late-
time measurements. As a distance-ladder-independent probe, CC offer a different dimension of study of the Hubble
tension.
HII galaxies (Sec. 2.1.10) measure H0 using the correlation between their luminosity and emission line flux from
ionized gas. The strong emission lines, primarily from young massive stars, allow for distance estimates based on the
luminosity-size relation. Calibrated using nearby galaxies, HII galaxy measurements provide independent constraints
on H0 consistent with other local probes.
DESI has provided a novel avenue for measuring H0 through the Fundamental Plane (FP) relation of early-type
galaxies, with a calibration tied to the distance to the Coma cluster (Sec. 2.1.12). By leveraging DESI’s extensive
sample of over 4,000 early-type galaxies, the FP relation was calibrated to provide precise distance estimates. For a dis-
tance to the Coma cluster of 98.5±2.2 Mpc based on SNIa, DESI yields a local value of H0 = 76.5±2.2 km s−1 Mpc−1 .
10

This result is in significant tension with the value inferred from Planck measurements where ΛCDM is assumed, and
which predicts a distance to Coma exceeding 110 Mpc. Historically, the distance to the Coma cluster has ranged from
90-100 Mpc, highlighting the discrepancy. By extending the Hubble diagram with FP-calibrated distances, DESI
highlights a persistent conflict between local measurements and early Universe predictions.
Extended Red Giant Branch (RGB) cosmologies (Sec. 2.1.15) constrain H0 using correlations between RGB lumi-
nosity and variability timescales. Their ability to probe higher redshifts than standard candles makes them valuable
for testing cosmic expansion over extended epochs. Similarly, Gamma-ray burst (GRB) observations (see Sec. 2.1.16)
use correlations between their luminosity and spectral properties to estimate H0 , serving as additional high-redshift
probes relevant to the Hubble tension.
Despite significant methodological diversity and substantial precision improvements, the tension between early- and
late-time measurements of the Hubble constant H0 persists at a statistically significant level, exceeding 5.9σ (Planck
2018 vs. SH0ES, alone, but higher when combining measures). This discrepancy challenges the completeness of the
standard ΛCDM model and raises the possibility of new physics.
Related to the H0 tension there is the sound horizon tension, which refers to a discrepancy in the inferred comoving
sound horizon scale at the end of the baryon drag epoch, rsdrag , derived from early- and late-time cosmological
measurements. This standard ruler, crucial for calibrating cosmological distances, is determined by the physics of
the early Universe, specifically the acoustic oscillations in the photon-baryon plasma before recombination. It is
primarily constrained by measurements of the CMB power spectrum, where Planck data assuming ΛCDM suggest
rsdrag ≈ 147.09 ± 0.26 Mpc. However, late-time measurements, such as BAO data combined with local distance ladder
determinations of H0 , suggest a lower sound horizon, with BAO-based estimates yielding approximately 137 Mpc, a
difference of about 7% and a tension of 2.6σ significance.
This discrepancy arises because rsdrag is directly connected to the expansion rate around recombination, meaning
changes to the early Universe physics could shift its value. Solutions involving early-time modifications often require
an enhanced expansion rate before recombination, which can reduce rsdrag . Examples include models introducing
additional relativistic species or Early Dark Energy (EDE) components. However, these scenarios face tight con-
straints from CMB observations, as the acoustic peak structure is highly sensitive to changes in photon diffusion and
gravitational driving effects.
Late-time measurements rely on the BAO feature imprinted at the drag epoch, calibrated through local H0 mea-
surements, such as the SH0ES collaboration results. Since rsdrag anchors the BAO scale, discrepancies in H0 esti-
mates propagate into inferred sound horizon measurements. This tension therefore highlights the need for consistent
cross-calibration between early- and late-time probes and motivates further investigation into both systematics and
extensions to the standard model of cosmology.
The combination of the estimates resulting from these, and other, approaches to estimating the value of H0 points
to a significant tension in the value of this parameter. This tension in the expansion rate of the Universe at current
times is further detailed in Sec. 2.1, with the most prominent estimates shown in Fig. 1.
Another interesting tension is that related to S8 , which highlights a persistent discrepancy between measurements
of the amplitude of matter fluctuations
p on cosmological scales inferred from early and late Universe observations. The
parameter S8 , defined as S8 ≡ σ8 Ωm /0.3, combines the clustering amplitude σ8 (the root-mean-square of matter
fluctuations on scales of 8 h−1 Mpc) with the present-day matter density parameter Ωm . It provides a key measure
of the growth of cosmic structures, making it a crucial probe of the standard cosmological model.
Current measurements from early-time data, such as the Planck 2018 results, which constrain the CMB temperature
and polarization anisotropies (Sec 2.1.18.f), yield a high precision estimate of S8 = 0.834 ± 0.016 assuming the
ΛCDM model. However, late-time measurements derived from weak gravitational lensing (Sec. 2.2.1) and galaxy
clustering surveys (Sec. 2.2.2), including Kilo-Degree Survey (KiDS), Dark Energy Survey (DES), and Hyper Suprime-
Cam (HSC), consistently report lower values of S8 . For example, DES Year 3 results obtained from combined galaxy
clustering and Weak Lensing (WL) measurements suggest S8 = 0.772 ± 0.017, while KiDS-1000 and HSC report
similarly low values around S8 ≈ 0.76 with comparable uncertainties. This tension, at the 2 − 3σ level, persists
across multiple independent data sets, indicating a possible systematic deviation between early and late-time probes
of structure formation.
The origin of this tension remains under debate. On the one hand, it could be driven by systematic uncertainties
in the analysis of WL data. These systematics include shear calibration biases, uncertainties in photometric redshift
estimates, and baryonic feedback effects that suppress the small-scale power spectrum due to processes like Active
Galactic Nucleus (AGN) feedback and gas ejection from galaxies. DES and KiDS collaborations have both extensively
explored the role of such systematics, yet the tension persists even after conservative scale cuts and improved modeling
of non-linear clustering.
On the other hand, the S8 tension might hint at the need for extensions to the standard ΛCDM model. One
potential modification involves evolving DE models, where a time-dependent equation of state parameter w(z) could
alter structure growth rates. Another possibility is modifications to the theory of gravity itself, such as f (R) models
11

CMB 2018 Planck


CMB 2025 (ACT-DR6)
CMB 2024 (SPT-3G+lensing+tauprior)
BBN+DESIBAO 2024
BBN+eBOSS 2022 H0 [km s 1 Mpc 1]
BBN+BAO+Shapefit eBOSS 2022 Cosmological Model Dependent

HST SH0ES 2024 (4 anchors) Direct


JWST SH0ES 2024 (1 anchor) (D vs z)
Cepheids 2022 (2 rungs, no SNIa)
Masers 2019 (no rungs)
TRGB CCHP + SNIa CSP 2025
TRGB EDD + SNIa CSP 2021
TRGB CATs + SNIa PanthP 2023
TRGB JWST + SBF 2025
TRGB HST + SBF 2021
Cepheids HST + SBF 2021
Miras + SNIa 2023
JAGB JWST SH0ES set + SNIa 2024
JAGB JWST CCHP set + SNIa 2024
JAGB JWST all + SNIa 2025
SN II (no rungs)
HII 2024
Tully-Fisher 2024
Tully-Fisher 2022 (baryonic)
Tully-Fisher 2020 (baryonic)
DESI Fundamental Plane + COMA 2024

Modeled Phenomena
Strong lensing 2020 (7 lensed QSO asser)
FRBs 2023 (18 local)
FRBs 2024 (64 local)

66 68 70 72 74 76 78 80

FIG. 1: Summary of H0 estimates from different cosmological probes with error bars smaller than
3.5 km s−1 Mpc−1 .

or scalar-tensor theories, which could modify the relationship between the matter distribution and lensing potential.
Additionally, the presence of massive neutrinos, which suppress structure formation at late-times due to their free-
streaming behavior, could also contribute to lowering the observed S8 value if the neutrino mass is larger than currently
assumed in the base ΛCDM model.
Notably, cross-correlation measurements between WL and other probes, such as CMB lensing (e.g., from Planck,
ACT, and SPT) and galaxy clustering from BAO surveys, have shown consistency with both early and late-time
datasets, complicating the overall picture. Furthermore, analyses that vary only the normalization of the power
spectrum, such as the σ8 parameter itself, have not fully resolved the tension, suggesting a more complex interplay
between cosmic structure formation and expansion history.
The tension in the S8 parameter may be milder but there is a growing effort to improve the constraints in obser-
vational estimates of the amplitude of matter fluctuations. These are detailed in Sec. 2.2, with the most prominent
estimates also shown in Fig. 4.
12

CMB 2018 Planck


CMB 2025 (ACT-DR6)
CMB 2024 (SPT-3G+lensing+tauprior)
BBN+DESIBAO 2024
BBN+eBOSS 2022 H0 [kmModel
Cosmological
s 1 Mpc 1]
Dependent
BBN+BAO+Shapefit eBOSS 2022

HST SH0ES 2024 (4 anchors) Direct


JWST SH0ES 2024 (1 anchor)
Cepheids 2022 (2 rungs, no SNIa)
(D vs z)
Masers 2019 (no rungs)
TRGB CCHP + SNIa CSP 2025
TRGB EDD + SNIa CSP 2021
TRGB CATs + SNIa PanthP 2023
TRGB JWST + SBF 2025
TRGB HST + SBF 2021
Cepheids HST + SBF 2021
Miras + SNIa 2023
JAGB JWST SH0ES set + SNIa 2024
JAGB JWST CCHP set + SNIa 2024
JAGB JWST all + SNIa 2025
SN II EPM method 2024
SN II SCM method 2022
SN II (no rungs)
HII 2024
Tully-Fisher 2024
Tully-Fisher 2022 (baryonic)
Tully-Fisher 2020 (baryonic)
DESI Fundamental Plane + COMA 2024

CC (GP) 2022
Modeled Phenomena
Strong lensing 2020 (7 lensed QSO cons)
Strong lensing 2020 (7 lensed QSO asser)
Strong lensing 2023 (SN Refsdal)
Strong lensing 2024 (SN H0pe)
GWs 2024 (1 bright standard siren)
GWs 2024 (dark standard sirens)
GWs 2023 (spectral+1 standard siren)
Cosmic Voids 2022
FRBs 2022 (16 unlocal + 60 local)
FRBs 2023 (18 local)
FRBs 2024 (64 local)

55 60 65 70 75 80 85

FIG. 2: Summary of H0 estimates from all the cosmological probes in this White Paper following the order of the
different sections.

1.2. Early vs local measurements and efforts for a solution


The persistence of the H0 tension across independent datasets suggests either unresolved systematic errors or the
need for extensions to the ΛCDM framework. Numerous theoretical solutions have been proposed, aiming to reconcile
the observed discrepancy while preserving the success of the standard cosmological model in describing early and late-
time observables. The most prominent of these efforts for a solution are shown in Fig. 6, which are also highlighted
below, and detailed in Sec. 4:

• Early Dark Energy (EDE): EDE models (Sec. 4.1.1) propose a transient DE component that briefly domi-
nates the cosmic energy budget before recombination, reducing the size of the sound horizon and allowing for
a higher inferred H0 from CMB data. This mechanism involves a scalar field with a potential that activates
temporarily and dilutes before significantly affecting late-time cosmology. Several variants of EDE have been
explored in the literature. Oscillatory EDE models, often motivated by axion-like fields, involve a scalar field
oscillating around the recombination epoch, injecting energy into the expansion rate. New Early Dark En-
ergy (NEDE) (see Sec. 4.1.2) refines this idea by introducing a phase transition where the scalar field rapidly
decays, creating a sharper and more controlled impact on the expansion rate. Adiabatic Fluctuation EDE
models, meanwhile, adjust the evolution of the scalar potential to balance early- and late-time cosmological
13

Planck+BAO+SNIa+SH0ES Planck+BAO+BBN+SNIa+SH0ES Planck+DESI+SNIa+SH0ES


Planck+DESI+SNIa PP+SH0ES Planck+BAO Planck+B2K+BAO+SNIa
Planck+BAO+SH0ES Planck Planck+DESI+SNIa
Planck+BAO+SNIa+S8 Planck+BAOtr+KiDS1000+SNIa+SH0ES Planck+BAO+RSD+SNIa
Planck+BAO+S8 Planck+BAOtr+DESY5 SN+CC+f 12+SH0ES Planck+BAO+CC+SNIa+SH0ES
Planck+BAO+SNIa Planck+DESI SNIa
Planck+BAO+SNIa+BBN

EDE (Poulin+ 2018)


EDE (Poulin+ 2024)
EDE freq (Herold+ 2022)
EDE (Hill+ 2020)
EDE freq (Herold+ 2022) H0 [km s 1 Mpc 1]
Cold NEDE (Cruz+ 2023)
Cold NEDE (Cruz+ 2023)
Cold NEDE freq (Cruz+ 2023)
Cold NEDE freq (Cruz+ 2023)
Hot NEDE (Garny+ 2024)
Hot NEDE (Garny+ 2024)
Majoron (Escudero & Witte 2021)
Non-thermal DM+phantom DE (da Costa+ 2023)
Wess Zumino DR (Schoneberg & Abellan 2022)
Vacuum Metamorphosis (Di Valentino+ 2020)
Emergent DE (Banihashemi+ 2020)
LsCDM (Akarsu+ 2023)
LsCDM+string model (Anchordoqui+ 2024)
wXCDM (Gomez-Valent+ 2024)
IDE (Giare+ 2024)
IDE (Zhai+2023)
IDE (Pan+2020)
Generalized Cubic Galileon (Frusciante+ 2020)
String Inspired Chern-Simons (Gomez-Valent+ 2023)
Ultralight DM-DE interaction (Aboubrahim & Nath 2024)
DM-photons Interaction (Becker+ 2021)
Decaying DM (Simon+ 2024)
DM-DE Interaction (Teixeira+ 2024)
KBC void galaxy counts (Haslbauer+ 2020)
PMF (Jedamzik+ 2025)
Higgs Inflation (Rodriguez+ 2023)
Electron mass + Omegak (Schoneberg & Vacher 2024)
Electron mass (Schoneberg & Vacher 2024)
Modified Recombination (Lynch+ 2024)
Modified Recombination MODREC (Lynch+ 2024)
LMT (Alestas+ 2021)
LwMT (Alestas+ 2021)
AvERA-625k (Pataki+ 2025)

68 70 72 74 76

FIG. 3: Summary of models proposed to solve the H0 tension in this White Paper following the order of the differ-
ent sections.

constraints, though these often introduce parameter degeneracies. While EDE models have shown promise in
alleviating the H0 tension by increasing the inferred value from early-time data, they also introduce challenges.
The models can exacerbate tensions in other cosmological parameters, particularly the amplitude of matter
fluctuations S8 and the matter density Ωm . Depending on the specific data set, EDE can reduce the H0 tension
to 2 − 3σ, however, recent CMB and Large Scale Structure (LSS) data sets show no evidence for this model and
disfavour large fractions of EDE.

• Late Dark Energy (LDE): Late Dark Energy (LDE) models (Sec. 4.2.1) attempt to address the Hubble
tension by modifying the expansion history at low redshifts (z ≲ 1). These models alter the equation of state
parameter of DE, w(z), away from the cosmological constant value w = −1. Some LDE scenarios involve a rapid
transition in w(z), where DE density evolves from a negative to a positive contribution, effectively accelerating
the late-time expansion rate. Although LDE models can slightly raise H0 without modifying the early Universe
physics, they often face challenges in matching BAO and Type Ia supernova data simultaneously, as the shift
14

CMB 2018 Planck S8


CMB 2023 (ACT-DR4+BAO)
CMB 2024 (SPT-3G+lensing+tauprior) CMB
Low-z
3x2pt KiDS-1000 + BOSS 2021
3x2pt DESY3 2022
3x2pt HSCY3 + SDSS BOSS DR11 2023
Peak Counts KiDS-1000 + DESY1 2024
Peak Counts DESY3 2022
Peak Counts HSC-DR1 2024
Cosmic Shear KiDS-Legacy 2025
Blue Cosmic Shear DESY3 2024
Cosmic Shear DESY3 + KiDS-1000 2023
Cosmic Shear HSCY3 2023
Cosmic Shear DECADE 2025
Cluster Counts eROSITA eRASS 2024
Cluster Counts SPTPol 2024
Cluster Counts DES 2025
Cluster Counts Subaru/HSC-SSP 2024
BOSS Galaxy Bispectrum 2024
Cross-Correlation unWISE-ACT DR6 2023
Cross-Correlation Gaia-Planck DR4 2024
Cross-Correlation DESI-Planck-ACT 2024
Stacked void-galaxy cross correlation 2022
UNIONS cluster cosmology 2025

0.70 0.75 0.80 0.85 0.90

FIG. 4: Summary of S8 estimates from different cosmological probes.

in the expansion rate can introduce tension with large-scale structure constraints.

• Rapid transitions in the late-Universe: Rapid transitions in the DE from negative to positive values in the
late Universe (Sec. 4.2.2) exhibit a sign-switching action in the vacuum energy while leaving the DE magnitude
unchanged. These models allow for the possibility of simultaneously increasing the value of H0 while also
suppressing changes in S8 . The transition point is established primarily by BAO data sets, while also being
supported by other contributions. Most models express an abrupt transition with a discontinuity in the vacuum
energy density.

• Interacting Dark Matter and Dark Energy (IDE) Models: Interacting Dark Matter and Interacting
Dark Energy (IDE) models (Sec. 4.2.3) propose a non-trivial energy exchange between DM and DE, modifying
both the cosmic expansion rate and structure growth. This interaction, described by a coupling function Q,
allows energy to transfer from one component to the other. Depending on the direction and strength of the
coupling, IDE can either slow down or accelerate the expansion rate. IDE models can increase H0 estimates
while modifying the growth of structure, offering a way to reduce the S8 tension simultaneously. However,
strong constraints from CMB lensing and BAO measurements limit the parameter space of IDE models, and
they often require fine-tuning to remain consistent with multiple datasets.
15

Planck+BAO+SNIa+SH0ES Planck+BAO+SNIa Planck+SNIa+SH0ES


Planck+DESI+SNIa PP+SH0ES Planck+BAO+BBN+SNIa+SH0ES Planck+BAOtr+DESY5 SN+CC+f 12+SH0ES
Planck+BAO+SH0ES Planck+BAO Planck+BAO+RSD+SNIa
Planck+BAO+SNIa+S8 Planck Planck+BAO+CC+SNIa+SH0ES
Planck+BAO+S8 Planck+BAOtr+KiDS1000+SNIa+SH0ES Planck+BAO+DES

EDE (Poulin+ 2018)


EDE (Poulin+ 2024)
EDE (Hill+ 2020)
Cold NEDE (Cruz+ 2023) S8
Cold NEDE (Cruz+ 2023)
Non-thermal DM+phantom DE (da Costa+ 2023)
Wess Zumino DR (Schoneberg & Abellan 2022)
LsCDM (Akarsu+ 2023)
LsCDM (Akarsu+ 2024)
wXCDM (Gomez-Valent+ 2024)
IDE (Sabogal+ 2025)
IDE (Zhai+2023)
Generalized Cubic Galileon (Frusciante+ 2020)
String Inspired Chern-Simons (Gomez-Valent+ 2023)
DM-baryon Interaction (He+ 2023)
Ultralight DM-DE Interaction (Aboubrahim & Nath 2024)
DM-photons Interaction (Becker+ 2021)
Decaying DM (Simon+ 2024)
DM-DE Interaction (Teixeira+ 2024)
PMF (Jedamzik+ 2025)
LMT (Alestas+ 2021)

0.750 0.775 0.800 0.825 0.850 0.875 0.900

FIG. 5: Summary of models proposed to solve the S8 tension in this White Paper following the order of the differ-
ent sections.

• Modified Gravity Theories: Modified Gravity (MG) models (Sec. 4.3.1 and Sec. 4.3.2) propose extensions
to GR by altering the Einstein-Hilbert action, introducing additional scalar degrees of freedom, altering the
underlying geometry itself, or changing the spacetime dimensionality, among other alternatives, which can
affect both the cosmic expansion rate and the growth of large-scale structure. These modifications can impact
the lensing potential and structure formation, leading to changes in the inferred values of H0 and S8 . Several
frameworks have been explored to address the Hubble tension within MG. f (R) gravity, for example, generalizes
GR by replacing the Ricci scalar with a nonlinear function of R, introducing an additional scalar mode that
modifies both the background expansion and structure growth. Horndeski theories, which include non-minimally
coupled scalar fields with derivative interactions, allow modifications to both the expansion history and lensing
effects but often face constraints from CMB lensing and large-scale structure data. Other models, such as
massive gravity and bimetric gravity, alter the graviton’s properties, leading to modified cosmic acceleration
and structure formation patterns. While some variants can reduce both the H0 and S8 tensions, they often
require fine-tuning or introduce instabilities in the late Universe. Teleparallel gravity approaches, including
f (T ) and f (Q) theories, replace curvature with torsion or non-metricity as the primary geometrical property
16

driving cosmic evolution. These models have been shown to affect the late-time expansion rate but remain
under scrutiny for consistency with both early- and late-time datasets. Emerging curvature theories, such as
AvERA [13, 14] modify the late expansion history compared to the concordance model (Sec. 2.3.9) to solve
the H0 and ISW puzzles. While MG models offer a rich theoretical landscape to explore, many face difficulties
in simultaneously resolving the H0 and S8 tensions while remaining consistent with solar system tests, CMB
lensing, and BAO constraints.

• Exotic Scenarios and Non-Standard Dark Matter Models: Exotic scenarios (Sec. 4.4.1, Sec. 4.4.2, and
Sec. 4.4.3) explore extensions to the standard cosmological model involving non-standard physics in both the
DM and DE sectors, which can influence the cosmic expansion history and structure formation. Among these
are decaying dark matter (DDM) models, where DM particles decay into lighter states. If this decay occurs
before or during the recombination epoch, it can alter the expansion rate and reduce the sound horizon, leading
to a higher inferred H0 . However, constraints from CMB lensing and large-scale structure limit the viability
of DDM as a complete solution. Ultralight scalar fields, such as axion-like particles or fuzzy dark matter, have
also been proposed. These fields can modify the standard expansion rate through their contributions to the
cosmic energy density, potentially mimicking an additional relativistic species or altering structure formation on
small scales. Nevertheless, precise measurements of the CMB power spectrum and the Lyman-alpha forest have
placed stringent bounds on their properties. Primordial black holes and non-CDM models represent further
exotic avenues, with the latter deviating from the standard cold, collisionless DM paradigm, often leading to a
suppression of small-scale structure growth. While these models offer novel mechanisms to address cosmological
tensions, most current models are tightly constrained by precision data from Planck, DESI, and the Pantheon+
supernova sample. However, they remain of theoretical interest, especially with upcoming data from CMB
Stage-4 and JWST capable of probing extreme scenarios further.

• Extra Relativistic Species and Neutrino Physics: Extra relativistic species (Sec. 4.1.3) represent another
class of proposed extensions aimed at resolving the Hubble tension. An increased number of relativistic degrees
of freedom, commonly parametrized by Neff , leads to a higher early Universe expansion rate and a reduced sound
horizon, potentially increasing the inferred H0 . This scenario can arise from new light particles, such as sterile
neutrinos or dark radiation. Sterile neutrinos, in particular, can act as an additional relativistic species if they
decouple before standard neutrinos while self-interacting or secret neutrino interactions can modify the thermal
history of the Universe and delay neutrino decoupling, effectively altering Neff . However, current constraints
from Planck, BAO, and CMB lensing data have limited viable deviations to ∆Neff ≈ 0.2. While insufficient
to fully resolve the Hubble tension, such scenarios remain relevant in the context of early Universe physics and
potential beyond-the-Standard-Model extensions.

• Local Void Hypothesis: The local void hypothesis (Sec. 4.5.1) posits that the Milky Way galaxy (MW)
resides in a large, underdense region of the Universe, potentially biasing measurements of the Hubble constant
due to a locally faster expansion rate compared to the cosmic average. The idea suggests that a void could lead
to a larger measured H0 locally while being lower on cosmological scales. However, the size and depth of such
a void required to explain the entire Hubble tension would be inconsistent with current observations. Studies
using CosmicFlows-4 and Pantheon+ supernova data, as well as analyses of bulk flows, have not identified an
underdensity significant enough to fully explain the observed discrepancy assuming ΛCDM.

• Primordial Magnetic Fields (PMFs): Primordial Magnetic Field (PMF) (Sec. 4.5.2) contributions are
generated before recombination could influence the early Universe’s expansion history and the formation of large-
scale structures. PMF contributions can modify the photon-baryon plasma dynamics, alter the acoustic peaks
in the CMB power spectrum, and shift the inferred sound horizon, all of which can affect the measurement of
H0 . However, current constraints from the CMB, including non-detections of Faraday rotation and the damping
tail suppression, limit the strength of PMFs, making it unlikely that they can fully account for the observed
tension.

• Inflationary Models: Certain inflationary scenarios (Sec. 4.5.3.b) propose modifications to the early Universe’s
physics that could indirectly affect the H0 measurement. Models involving non-standard reheating phases,
features in the inflationary potential, or a modified primordial power spectrum have been investigated in this
context. These scenarios can alter the early expansion history or the acoustic peak structure, impacting the
inferred value of H0 from CMB measurements. However, most inflationary modifications struggle to generate a
significant enough shift in H0 while remaining consistent with current CMB and large-scale structure data.

• Varying Fundamental Constants: The idea of varying fundamental constants (Sec. 4.8) explores the pos-
sibility that parameters such as the fine-structure constant α, the electron mass me , or the proton-to-electron
17

mass ratio µ could change over cosmic time or with spatial position. If fundamental constants were to evolve,
they could impact key cosmological observables, such as the sound horizon and the physics of recombination,
altering the inferred values of H0 . Scalar fields coupled to the electromagnetic sector, as in models like varying-α
theories, could drive such changes. However, stringent constraints from CMB, Big Bang Nucleosynthesis (BBN),
and RGB absorption spectra have significantly limited the variation of some of these constants and their ability
to resolve the Hubble tension, while the others face issues of finding a coherent theoretical framework that can
explain their variation.
• Local Physics Solutions: Local physics solutions (Sec. 4.9) suggest that the H0 tension could arise from
new physical effects specific to the local Universe rather than requiring modifications to the global cosmological
model. One possibility is a local modification to the gravitational constant, Geff , which could alter the calibration
of standard candles like Cepheid variables or SNIa, leading to a biased measurement of H0 . Another proposed
local effect involves variations in the transparency of the Interstellar Medium (ISM) or modified extinction laws
that could influence supernova observations. While some of these scenarios can explain a portion of the tension,
they generally fail to provide a complete resolution and are often tightly constrained by measurements from
cosmic bulk flows, galaxy clustering, and large-scale structure surveys.
• Systematic Uncertainties and Calibration Issues: A significant portion of the community continues to
explore whether systematic errors could account for the observed tension. Potential sources include Cepheid
calibration errors, host galaxy dust corrections, and selection effects in both early and late-time measurements.
However, rigorous cross-calibrations between HST, JWST, and ground-based surveys like DESI and SH0ES have
yet to identify a significant bias capable of resolving the tension completely.

1.3. Additional curiosities and anomalies


Other cosmological anomalies beyond the primary H0 and S8 tensions have emerged in recent years, revealing
potential challenges to the standard ΛCDM model. These discrepancies span a wide range of physical scales and
cosmological probes, motivating deeper investigations into both systematic effects and new physics. Key anomalies
include:

• The Alens anomaly: The Alens parameter, introduced as a phenomenological extension to the standard CMB
analysis, quantifies the amplitude of lensing-induced smoothing of acoustic peaks in the CMB power spectrum.
Planck data suggest a value for Alens slightly higher than expected under ΛCDM, with a significance of 2 − 3σ,
potentially pointing to unmodeled effects in CMB lensing or physics beyond standard cosmology (Sec. 2.3.1).
• Evidence for nonzero curvature: While the standard model assumes a flat universe, Planck’s CMB data,
when analyzed without external priors, have shown mild evidence for a closed geometry with Ωk < 0. However,
this claim remains controversial, as BAO and lensing measurements are consistent with a flat universe within
1σ (Sec. 2.3.2).
• CMB anisotropic anomalies: Several statistically significant features have been identified in CMB data,
including the hemispherical power asymmetry, alignments in low multipole moments, and the cold spot anomaly.
While most of these could be attributed to cosmic variance, their persistence in multiple datasets (e.g., Planck,
Wilkinson Microwave Anisotropy Probe (WMAP)) raises the question of whether they hint at new physics or
residual systematics (Sec. 2.3.3).
• The w0 wa tension: Discrepancies have emerged in constraints on the DE equation of state parameters, w0 and
wa , particularly when comparing results from CMB data with late-time probes such as Supernovae (SN) and
DESI BAO. Some results suggest a deviation from the cosmological constant value of w0 = −1, which could
indicate evolving DE or unresolved systematics (Sec. 2.3.4).
• Neutrino Tension: The sum of neutrino masses, mν , inferred from cosmological data, such as CMB and
P
BAO observations, has shown tension with the lower boundsP set by terrestrial neutrino oscillation experiments,
which require mν ≳ 0.06 eV (normal ordering, NO) and mν ≳ 0.1 eV (inverted ordering, IO).P Recent DESI
P
BAO data combined with Planck and ACT CMB constraints yield tight upper limits, such as mν < 0.05
eV (2σ), challenging the inverted hierarchy and preferring the NO with a Bayes factor exceeding 46.5 in some
datasets. Interestingly, some analyses allowing for extended models have reported a preference for negative
neutrino masses, linked to enhanced CMB lensing signals in Planck and ACT data. This tension could stem
from unaccounted systematics in lensing measurements or may indicate new physics beyond minimal ΛCDM,
such as modified neutrino properties or evolving DE models (Sec. 2.3.5).
18

Quantum
gravity
Warm dark cold dark effects on
matter matter particles Quantum
and fields gravity
modified
Interacting
dynamics
and
decaying
dark
matter Quantum
Matter sector gravity phe- Non-
solutions nomenology Riemanian
Local voids geometries
Higher di-
mensional
models
Primordial
magnetic
fields

Other Modified Non-local


solutions gravity models

Inflationary
models
Higher
Fundamental order
Physics Solutions models

Early dark
energy
Deviations
Early-time from the
solutions Cosmological
Principle
New early
dark
energy

Extra
degrees of
freedom
Varying
(neutrino Late-time
fundamental
physics) solutions
constants
Interacting
dark
energy

Dark
Late dark energy
energy transition
models

FIG. 6: Fundamental physics provides a number of potential solutions to address the challenge of tensions in cos-
mology.

• Cosmic dipole anomalies: Tensions exist between the inferred cosmic dipole from the CMB and measurements
from radio galaxies and RGBs. Separately, statistically significant directional variations in H0 have been reported
in both CMB and the local Universe, providing an alternative perspective on CMB anisotropic anomalies and
bulk flow anomalies. These discrepancies could point to non-standard cosmic expansion or incomplete modeling
of large-scale structure effects (Sec. 2.3.6).

• Big Bang Nucleosynthesis (BBN) anomalies: Discrepancies exist between deuterium and helium abun-
dances inferred from BBN and those predicted by CMB-based baryon density estimates, suggesting either
measurement systematics or small deviations from standard physics during the early Universe (Sec. 2.3.7).
19

• Lyman-α anomalies: High-redshift Lyman-α forest measurements show mild tension with low-redshift con-
straints on the matter power spectrum, raising questions about the evolution of structure and thermal history
of the Intergalactic Medium (IGM) (Sec. 2.3.8).
• Integrated Sachs-Wolfe (ISW) and cosmic superstructures: Measurements of the Integrated Sachs–Wolfe
(ISW) effect, which probes late-time structure growth through correlations between CMB and large-scale struc-
ture data, have shown excess correlations on large scales, potentially linked to cosmic superstructures (Sec. 2.3.9).
• Cosmic void anomalies: Observations of cosmic voids have revealed unexpected properties in their size
distribution and their contribution to the ISW effect, potentially suggesting deviations from the standard growth
of structure (Sec. 2.3.11).
• Fast Radio Burst (FRB) probes: Fast Radio Burst (FRB) observations have recently emerged as a powerful
probe of cosmic tensions, offering a unique way to constrain the expansion history and large-scale bulk flows.
Preliminary results suggest some inconsistencies with ΛCDM predictions (Sec. 2.3.12).
• Radio background excess: Several observations of the diffuse radio sky have reported an unexplained excess
in both the surface brightness and the anisotropy power of the radio background, which could be linked to new
diffuse or low flux radio sources, PMFs, exotic decays, or residual systematics across a range of radio observations
(Sec. 2.3.13).
• Bulk flow anomalies: Peculiar velocity measurements and bulk flows inferred from galaxy catalogs show a
higher amplitude than predicted by ΛCDM, potentially pointing to the presence of large-scale structures beyond
the standard model (Sec. 2.3.14).
• Ultra Long Period Cepheids: These variable stars have been proposed as potential standard candles, but
their properties appear inconsistent with standard stellar evolution models, raising questions about distance
ladder calibrations (Sec. 2.3.15).

Ongoing and upcoming surveys such as Euclid, The Rubin Observatory’s Legacy Survey of Space and Time (LSST),
and the Roman Space Telescope will play a pivotal role in clarifying these anomalies. Additionally, methodological
advances, including cross-correlations between multiple probes and the application of Machine Learning (ML) tech-
niques (Sec. 3, Sec. 3.2–3.5), will help assess whether these tensions are genuine physical discoveries or the result of
residual systematics. A deeper understanding of these anomalies is essential to determin whether extensions to the
standard ΛCDM framework are required and to uncover the fundamental nature of DE, DM, and cosmic structure
formation.

1.4. Data analysis: How to tackle the problem


The analysis of cosmological data has seen remarkable advancements, leveraging a wide range of statistical, compu-
tational, and interdisciplinary approaches. These methodologies have not only refined parameter constraints but also
opened new avenues to tackle cosmological tensions such as the H0 and S8 discrepancies. These tensions challenge
the completeness of the ΛCDM model, suggesting the possibility of new physics or unresolved systematics.
Cosmology simulators and Markov chain Monte Carlo (MCMC) techniques (Sec 3.1) are foundational tools for
exploring high-dimensional parameter spaces. They allow for robust testing of cosmological models against observa-
tional data while accounting for degeneracies and systematics. Advances in MCMC methods, such as nested sampling
and Hamiltonian Monte Carlo, have improved the efficiency and accuracy of parameter estimation, particularly when
combining data from the CMB, BAO, and SN datasets.
ML approaches (Sec. 3.2) have emerged as powerful tools in cosmological analysis. Neural networks, Gaussian
processes (GP), and decision trees have demonstrated their utility in extracting patterns, accelerating data analysis,
and refining model predictions. For example, ML techniques have been used to process large-scale structure data
and improve constraints on DE models. Hybrid approaches, combining ML with traditional statistical methods, have
further enhanced the reliability of these analyses by integrating the strengths of both paradigms.
Reconstruction methods (Sec. 3.3) have also played a critical role, enabling the inference of the Universe’s expansion
history from observational data. Techniques such as GP regression have been applied to reconstruct the Hubble
parameter H(z) and the growth of cosmic structures without assuming a specific cosmological model. These methods
reduce systematic biases and provide model-independent insights into cosmological tensions.
Bio-inspired algorithms, particularly Genetic Algorithm (GA) (Sec. 3.4), have introduced novel ways to optimize
model selection and parameter estimation in cosmology. These algorithms, inspired by evolutionary processes in
nature, use mechanisms such as selection, mutation, and crossover to identify optimal solutions in complex and high-
dimensional parameter spaces. GAs are particularly well-suited for exploring cosmological models where traditional
20

gradient-based methods may struggle, such as those involving nonlinear dynamics or multimodal likelihood functions.
By iteratively refining a population of candidate solutions, GAs can efficiently locate regions of interest in the pa-
rameter space, even in the presence of degeneracies or non-Gaussian distributions. Applications of GAs in cosmology
have been diverse. They have been employed to address key tensions, such as the H0 and S8 discrepancies, by testing
alternative models including EDE, Interacting Dark Matter (IDM)-DE scenarios, and MG theories. For instance,
GAs have been used to explore the parameter space of IDE models, identifying regions that minimize tensions with
both early- and late-time observations. Their flexibility allows for the incorporation of priors from ML or other
inference techniques, further enhancing their robustness. Moreover, hybrid approaches that combine GAs with ML
have shown promise in improving the convergence and accuracy of solutions. These methods leverage the pattern
recognition capabilities of neural networks or decision trees to guide the evolutionary search process, thereby reducing
computational overhead while maintaining accuracy. In addition to model testing, GAs have proven effective in en-
hancing observational strategies. By simulating survey configurations and optimizing the allocation of observational
resources, they help maximize the scientific output of upcoming missions. This includes optimizing survey strategies
for WL, galaxy clustering, and other probes of large-scale structure. The adaptability of GAs also extends to their
role in reconstructing the initial conditions of the Universe from observed data. By evolving populations of initial
conditions, these algorithms can identify those that best reproduce observed structures, offering insights into the
fundamental physics of the early Universe. Overall, bio-inspired algorithms like GAs represent a powerful addition to
the cosmological toolkit, enabling the exploration of complex models and the refinement of observational strategies.
Their continued integration with other advanced methods, such as ML and high-performance simulations, holds great
potential for addressing the most pressing tensions in cosmology and uncovering new physics.
Cosmological simulations (Sec. 3.5) have become indispensable for interpreting observational data and testing
theoretical models. High-resolution N -body simulations, hydrodynamic models, and semi-analytic methods provide
a detailed understanding of the large-scale structure formation and the interplay of DM and DE. These simulations
enable the calibration of observables such as WL signals, galaxy clustering, and BAO measurements, thereby improving
the accuracy of cosmological parameter constraints.
Statistical tools, including profile likelihoods, have gained prominence in cosmology for estimating confidence inter-
vals and assessing parameter significance. Unlike traditional Bayesian methods, profile likelihoods provide a frequentist
alternative (Sec. 3.6) that is less sensitive to prior assumptions. This makes them particularly useful in scenarios where
prior information is limited or where strong assumptions may bias the results. Profile likelihoods excel in analyzing
models with complex parameter spaces, including those with non-Gaussian distributions or significant degeneracies
between parameters. Such complexities often arise in cosmological contexts, for instance, when exploring EDE models,
IDE scenarios, or the inclusion of additional relativistic species. These models frequently introduce additional degrees
of freedom, leading to correlations that can obscure parameter constraints when traditional Bayesian approaches are
employed. A key advantage of the profile likelihood approach is its ability to disentangle degeneracies and provide
robust confidence intervals without relying on the full posterior distribution. This is achieved by profiling the nuisance
parameters, effectively marginalizing them without requiring explicit integration. As a result, the profile likelihood
method can highlight the true parameter dependencies and offer more transparent interpretations of the data. Appli-
cations of profile likelihoods in cosmology have included constraining the equation of state of DE, testing deviations
from the standard ΛCDM model, and evaluating the statistical significance of cosmological tensions, such as those in-
volving H0 and S8 . For example, in the context of EDE, profile likelihoods have been used to assess whether parameter
shifts can resolve these tensions or whether they indicate systematic effects in the data. The computational efficiency
of profile likelihoods also makes them a valuable tool in high-dimensional analyses, particularly for next-generation
cosmological surveys. Their ability to isolate relevant subspaces of parameter space enables efficient exploration of
hypotheses while minimizing the computational cost associated with evaluating full posterior distributions. Further-
more, recent advances in computational methods have improved the applicability of profile likelihoods. Techniques
such as adaptive mesh refinement and ML-assisted likelihood evaluations have reduced the computational demands
of high-precision analyses. These innovations make profile likelihoods increasingly attractive for analyzing the large
datasets expected from upcoming missions like the Simons Observatory, Euclid, and the Roman Space Telescope.
These methodologies, often used in combination, reflect a collaborative effort across theoretical, computational, and
observational domains. For example, the integration of MCMC methods with ML-enhanced simulations and recon-
struction techniques has provided a comprehensive framework for addressing persistent cosmological tensions. This
synergy has also enabled the exploration of exotic scenarios, such as decaying DM, PMFs, and varying fundamental
constants, which aim to explain anomalies in the current cosmological framework.
The ongoing advancements in data analysis tools and methodologies are pivotal for the next generation of cosmolog-
ical surveys, including the Simons Observatory, CMB—Stage 4 (CMB-S4), Euclid, and the Roman Space Telescope.
These surveys will generate unprecedented datasets that require state-of-the-art techniques to extract meaningful in-
sights. A schematic is shown in Fig. 7 of these methods. By embracing these innovations, the cosmological community
continues to push the boundaries of our understanding of the Universe, addressing existing tensions and paving the
21

ML-
Boosted Tradi-
tional
Sam-
plers
Monte- CosmoSIS
Python

Cobaya
Nested
Bayes samplers
Factor No-U-
Inference Turn:
Frame- JAX +
works numpyro
Information
Cri-
teria Model Hamiltonian
selection Monte
criteria Carlo

Based
Bayesian Ensemble
on inference and Se-
Bayes methods in Markov quen-
Evi- Tension Cosmology Chain tial
dence metrics Monte MC Nonlinear
Carlo scale
Based ob-
on Pos- Affine- ser-
terior invariant va-
distri- tions
bution N-body
simu-
lations
Binned
MCMC
Models
beyond
LCDM
Dedicated
opti-
mization Inference
methods Frequentist from cos-
Approaches mological
simulations

Memetic
Cosmological Algorithm
Data Analysis
Artificial Techniques
Neural Grammatical
Networks Evolution

Machine Bio-inspired
Learning algorithms
Concurrent techniques in model
Neural Differential
Networks selection
Evolution

Bayesian Reconstruction Approximate


Neural Techniques Genetic Bayesian
Networks Algorithm Compu-
Recurrent
Neural Particle tation
Networks Swarm Op-
timization

Gaussian Parametric
Processes Methods
Artificial
Neural
Networks

FIG. 7: Schematic of the different data analysis approaches used in the cosmology community.

way for new discoveries in fundamental physics.


22

2. Observational cosmology and systematics


2.1. H0 Tension: Measurements and systematics
Introduction: Adam Riess

It is cliché to say we live in the era of precision cosmology—but it is true. Over the past several decades, the
type, scope, and precision of cosmological measurements have grown enormously. Two of the most powerful tools,
SNIa, and BAOs, joined the cadre of first-rank indicators during this time. Other observables—such as the CMB,
gravitational lensing, primordial abundances, and components of the distance ladder—have been refined and matured,
sharpening our view of the Universe. It is an exhilarating time to be a cosmologist. The surprising and non-intuitive
composition of the Universe demands the full use of our observational toolkit. Whether the correct cosmological
model is the simplest form of ΛCDM or one with additional complexities will ultimately be determined by the quality
of our measurements. In the following sections, you will read how the astronomical zoo of objects and features has
been employed to produce precision tests of the Universe, with a level of control over systematic errors once known
only to particle physics. To be fair, not every tool and technique has reached the same level of robustness (and some
may never do so), but this is something you can judge for yourself as you explore the state of each art. Perhaps the
biggest question for observational cosmology in the 2020s is what to make of the growing evidence of tensions. The
first appearance of any tension or anomaly is usually attributed to experimental error or systematics—an assumption
that makes sense when playing the odds. However, signals that may herald new physics will be treated the same
unless critical thinking and extensive analysis follow. At present, one or more tensions have surpassed thresholds
of statistical significance, reproducibility, and independent cross-checks, earning them the continued attention of the
field. In the end, as Einstein once said, the Universe (or Lord) is subtle but not malicious. Our goal is to measure
these subtleties.
2.1.1. The distance ladder
Coordinator: Louise Breuval
Contributors: Adam Riess, Giulia De Somma, Leandros Perivolaropoulos, Lluís Galbany, Lucas Macri, Richard I.
Anderson, Siyang Li, and Vladas Vansevicius

The Hubble constant measures the present expansion rate of the Universe. This cosmological parameter represents
the slope of the redshift-distance relation, cz = H0 D, in the limit of z ∼ 0. Only distant galaxies are sensitive to
the Universe’s expansion, unlike nearby systems which are dominated by local gravitational interactions. The best
method to reach galaxies in this regime, called the Hubble Flow (0.02 < z < 0.20), is to build a distance ladder based
on a succession of distance indicators. The most widely used and best calibrated distance ladder is based on three
rungs. First, geometric distances are used to calibrate the Period-Luminosity relation of Cepheid variables in nearby
galaxies. This law is, in turn, adopted to calibrate SNIa on the second rung, which is limited to the volume where
both Cepheids and SNIa are observable. Finally, on the third rung, distances and redshifts of SNIa in the Hubble
Flow directly measure H0 . Although Cepheids provide the most homogeneous and reliable set of distances to many
nearby galaxies, they can be substituted by alternative standard candles such as TRGB stars [15, 16], JAGB [17, 18],
or Mira variables [19, 20]. On the second rung, SNIa can be replaced by SBF [21, 22], SNII [23, 24] or the Tully-Fisher
relation [25]. These methods provide valuable independent checks of the distance ladder.
A Cepheid-SNIa distance ladder was included in the HST Key Project on the Extragalactic Distance Scale [27] result
of H0 = 72 ± 8 km s−1 Mpc−1 (which calibrated multiple secondary distance indicators using Cepheids). Since then,
this measurement was significantly improved using state-of-the-art data from later HST instruments, homogeneous
calibration, and near-IR Cepheid observations (Fig. 8). The first rung of the distance ladder is currently supported
by four geometric calibrators, each providing a direct absolute determination of the Cepheid luminosity: Gaia DR3
parallaxes of Cepheids and host clusters in the MW [28], detached eclipsing binaries distances in the Large and Small
Magellanic Clouds (SMC) [29, 30], and the water maser distance to NGC 4258 [31]. The second rung now comprises a
total of 42 SNIa in 37 galaxies, where all hosts are observed with the same instrument (Wide Field Camera 3 (WFC3))
on the same HST, an investment of more than 1000 orbits of observing time. On the last rung, the Pantheon+ team
provides ∼ 300 SNIa in the Hubble Flow with the highest quality data and calibration, standardized across many
surveys [32, 33]. These recent developments provided the most precise H0 measurement from a simultaneous fit of
the three rungs with 73.17 ± 0.86 km s−1 Mpc−1 [34, 35]. This HST-based measurement has recently been confirmed
with observations of a subsample of SNIa host galaxies with JWST, which show excellent consistency between both
telescopes [26] (Fig. 9). Regardless of the method, essential elements for a precise H0 measurement in the late Universe
include the use of near-infrared photometry to minimize the impact of dust [36] and consistent data to cancel flux
calibration errors between rungs.
23

Type Ia Supernovae → redshift(z)


σ=0.135 mag
42

µ (z,H0=73.0,q0,j0)
40

38

36

34
Cepheids → Type Ia Supernovae
σ=0.130 mag, N=42 0.4

∆ mag
34
0.0

SN Ia: m-M (mag)


33
-0.4
32 34 36 38 40 42
SN Ia: m-M (mag)
31

30

Geometry → Cepheids 29
30
N4258 0.4

∆ mag
Cepheid: m-M (mag)

25
0.0
-0.4
SMC 29 30 31 32 33 34
20 LMC
Cepheid: m-M (mag)
Milky Way
15

10

0.4 0.4
∆ mag

0.2
0.0 0.0
-0.2
-0.4 -0.4
10 15 20 25 30
Geometry: 5 log D [Mpc] + 25

FIG. 8: The Cepheid-SNIa distance ladder [26]. The left-hand panel (first rung) shows the four anchor galaxies
with geometric distances which are adopted to calibrate Cepheids. In the middle (second rung), 42 SNIa are cal-
ibrated with Cepheids. On the right-hand panel (third rung), redshifts and distances of SNIa in the Hubble Flow
directly measure the expansion rate of the Universe.

FIG. 9: Comparisons of H0 between HST Cepheids (baseline, green circles) and JWST Cepheids (squares) for dif-
ferent SNIa host subsamples. The CCHP (blue) and SH0ES (red) subsamples selected for JWST observations pro-
duce a difference of 3 − 4 km s−1 Mpc−1 in H0 owing to selection. The HST and JWST distance measurements them-
selves are in good agreement [26].
24

2.1.2. Cepheid variables as standard candles


Coordinator: Louise Breuval
Contributors: Adam Riess, Giulia De Somma, Leandros Perivolaropoulos, Lluís Galbany, Lucas Macri, Richard I.
Anderson, Siyang Li, and Vladas Vansevicius

OGLE-SMC-CEP-4697
2.1.2.a. Cepheid variables as standard candles Classical
Cepheids are luminous evolved stars located in the instability [4.5] -0.8
strip region of the Hertzsprung–Russell diagram. These stars
11.5
undergo regular radial pulsations driven by two mechanisms: [3.6] -0.5
the κ (opacity) and γ (adiabatic exponent, Γ3 − 1) mecha-
nisms, which operate in the partial ionization zones of abun- 12.0 K-0.2
dant elements such as H, He, and He+ [40, 41]. Cepheids
are regarded as the “gold standard” for distance measure-
12.5 J-0.1
ments due to their tight Period-Luminosity (P-L) relation

mag
[42], expressed as M = a log P + b, where M is the abso-
lute magnitude, P is the pulsation period, and a and b are 13.0
constants. By measuring a Cepheid’s P , its M can be de- I
termined, and its distance D (in Mpc) can be inferred using 13.5
the distance modulus: m − M = 5 log10 (D) + 25, where m is
the apparent magnitude. From a theoretical perspective, the
P-L relation arises from fundamental principles of stellar as- 14.0
trophysics [43]. The pulsation period is related to the mean V
density ρ of the star by P ∝ ρ−1/2 , where ρ ∝ M/R3 (M
and R are the stellar mass and radius, respectively). The
0.2 0.6 1.0 1.4 1.8
phase
Stefan-Boltzmann law, L ∝ R Teff , and the mass-luminosity
2 4

relation, L ∝ M3.5 , link the star’s luminosity (L) and effec- FIG. 10: Multi-band light curves for Cepheid
tive temperature (Teff ) to its pulsation period, providing a OGLE-SMC-CEP-4697 based on a compilation of
direct foundation for the P-L relation. data from Refs. [37–39].
Cepheids play a key role in the distance ladder [34]. Their
P-L relation is locally calibrated with geometric distances and is used, in turn, to calibrate SNIa on the second
rung. Cepheids are easy to identify due to their periodicity and are bright enough to be observed up to about 50
Mpc with the current generation of space telescopes. They are generally identified at optical wavelengths [44], where
their pulsation amplitudes peak (∼ 1 mag in the V band), allowing precise determination of their pulsation periods
(Fig. 10). Follow-up observations in the near-infrared minimize the impact of interstellar dust on their magnitudes.
P-L relations are typically constructed using intensity-averaged magnitudes over the pulsation cycle. These averages
are derived either through extensive photometric sampling or by fitting random-phase measurements with template
light curves [45, 46]. Using mean magnitudes significantly reduces the P-L dispersion by a factor of ∼2 compared to
random-phase measurements.
However, systematic uncertainties in Cepheid distances may arise due to differences between nearby calibrators and
distant Cepheids observed in SNIa host galaxies. These include chemical composition, period range, crowding, dust
properties, and potential nonlinearity of the P-L relation (see Ref. [47] for a comprehensive review). In the following
sections, we outline the most significant systematics affecting Cepheid distances, and we propose solutions to mitigate
them.
2.1.2.b. Improvements in reducing the systematics
Photometric systems Combining multiple photometric systems to observe Cepheids in the first and second
rung introduces a 1.4 − 1.8% systematic error in distance measurements, as described in Ref. [48]. To mitigate this, it
is advantageous to use the same telescope and instrument, thereby minimizing flux-calibration errors. Ground-based
telescopes, while offering dedicated facilities with high availability [38], are impacted by atmospheric absorption
and scattering. In contrast, space-based observatories like the HST and JWST provide superior image quality, are
unaffected by weather, exhibit excellent long-term stability, and are not constrained by the day-night cycle. Prior
to the launch of JWST, WFC3 on HST was the only instrument capable of resolving Cepheids in the most distant
SNIa host galaxies at optical and near-infrared wavelengths [34]. While WFC3 is well-suited for observing both
faint and bright targets, it suffers from a slight non-linearity in the infrared. Specifically, photons from faint stars
(low count rates) are measured less efficiently than those from bright stars (high count rates), introducing a Count
Rate Non-Linearity (CRNL). Though small, this bias is critical for achieving percent-level precision and amounts to
0.0077 ± 0.0006 mag/dex [49]. A CRNL correction is typically applied to bright Cepheids in anchor galaxies to ensure
consistent calibration [34, 35].
25

Crowding Separating Cepheids from their surrounding stellar populations is one of the most significant challenges
in measuring Cepheids in SNIa host galaxies [15]. Contamination from redder stars, such as QSO and AGB stars along
the line of sight, limits the precision of Cepheid measurements, especially in the near-infrared. A common approach
to address crowding is to add artificial stars of known brightness at random positions near Cepheids and recover their
magnitudes using the same photometric methods applied to real stars [44]. This allows for a statistical correction for
crowding (although such corrections remove the bias, but not the added scatter or crowding noise). With the advent
of JWST, the ability to separate Cepheids from background stars has improved significantly, resulting in a substantial
reduction in crowding noise. Recent JWST observations of Cepheids in SNIa host galaxies including the anchor galaxy
NGC 4258 [50, 51] have demonstrated excellent agreement with prior HST measurements, achieving a mean difference
of only 0.01 mag. Furthermore, JWST data reduced the scatter in the P-L relation by a factor of 2.5 (Fig. 11). This
result decisively rules out Cepheid crowding from HST photometry as the cause of the Hubble tension, with a 8.2σ CL.

Metallicity The luminosity of Cepheids at a given period is known to correlate with their chemical composition,
but the sign and magnitude of this dependence (γ) have historically been difficult to constrain [52–54]. Accounting
for the metallicity dependence is essential in the distance ladder, particularly when calibrating Cepheids in the Mag-
ellanic Clouds, which are more metal-poor than typical SNIa host galaxies [35, 48]. However, Cepheids in NGC 4258
and the MW resemble those in large spiral SNIa hosts, making metallicity differences negligible for determining
H0 . Recent calibrations of the metallicity effect, leveraging accurate distances and expanded Cepheid samples, have
converged on a consensus. These calibrations include studies using Gaia parallaxes and individual spectroscopic
abundances of MW Cepheids [55–57] and combined analyses of MW and Magellanic Cloud Cepheids with geometric
distances [35, 58]. The derived metallicity dependence (γ) lies between −0.2 and −0.3 mag/dex, with the negative
sign indicating that metal-rich Cepheids are intrinsically brighter than metal-poor ones. Recent theoretical stud-
ies have corroborated both the sign and magnitude of this effect [59, 60], particularly for the Wesenheit index (a
dereddened magnitude combining multiple filters [61]) employed in the SH0ES distance ladder [34]. For nearby
galaxies, direct [Fe/H] abundances can be determined from high-resolution spectroscopy of Cepheids [56, 62]. In more
distant galaxies, metallicities are inferred from the [O/H] gradient measured via HII region spectroscopy [63–65] or
optical Integral Field Spectroscopy (IFS) of calibrator galaxies [66]. These methods are consistent in the MW, where
Cepheid spectroscopy agrees with HII region gradients to within 1σ ∼ 0.05 dex [34]. Because the metallicities of
Cepheids in SNIa host galaxies closely match those in the anchor galaxies, the Hubble constant is only weakly affected
by the metallicity effect. Neglecting the correction would change H0 by just 0.5 km s−1 Mpc−1 (in the higher H0
direction) [34]. Although the impact on H0 is minimal, the metallicity correction plays a crucial role in ensuring con-
sistency among anchor galaxies while independently matching the radial dependence with Gaia parallaxes in the MW.

Binaries While crowding can be mitigated using artificial star measurements, unresolved companion stars physi-
cally associated with Cepheids can bias their measured flux,thereby affecting inferred distances. A significant fraction
of Cepheids reside in binary or multiple systems [67], and unresolved companions typically increase the measured
brightness of Cepheids. However, because unresolved binarity is equally present in both SNIa hosts and anchor
galaxies, its effect largely cancels out in calibrated distances. In the Wesenheit index adopted in the SH0ES distance
ladder, [68] used synthetic populations of binary Cepheids to estimate the impact of binarity, finding it to be a minor
effect, contributing only 0.004 mag in distance modulus, or a 0.1% change in H0 . Similarly, [69] determined that the
primary source of flux contamination arises from Cepheids located in open clusters, leading to an overestimate of H0
by approximately 0.23%. However, the fraction of Cepheids in open clusters is relatively small (see also Refs. [46, 70]).
Thus, while binarity and clustering introduce minor biases, their overall impact on H0 remains negligible at the cur-
rent level of precision.

2.1.2.c. Perspectives
Dust laws Dust has long been a critical systematic uncertainty in the distance ladder [71]. To correct for dust,
one can either subtract Rλ × E(V − I) to Cepheid apparent magnitudes or adopt a reddening-free “Wesenheit”
magnitude (see Ref. [34, 61] for details). Additionally, to the uncertainties in the reddening E(V − I) itself, the value
of the RV parameter is not independently known in each SNIa host galaxy and at each Cepheid position, but rather
is assumed to match the MW [34, 72]. However, the use of near-infrared observations greatly limits the impact of
reddening, reducing the size of Rλ to ∼ 0.4 and variations due to these different values of RV to ±0.03. Varying the
reddening law across different hosts first requires subtracting the intrinsic color of Cepheids in order to separate the
component of the color that results from dust reddening (see Sect. 6.3 and Appendix D in Ref. [34]). A complete
sampling of the reddening curve at long and short wavelengths in each host as well as an improved characterization
26

of the dependence of RV with host properties such as mass, star formation rate, and color [73] can further reduce this
source of uncertainty in Cepheid measurements.

Cepheid parallaxes Early measurements of trigonometric parallaxes were obtained with the Fine Guidance
Sensor (FGS) on HST [74] and later by spatial scanning with HST/WFC3 [75, 76] for ∼10 MW Cepheids. The
European Space Agency (ESA) Gaia mission now provides parallaxes for thousands of Cepheids with individual
precision at the ∼ 5% level [28]. The main systematic uncertainty in this method arises from the parallax zero-point,
a small corrective term estimated with distant RGBs, LMC stars and binaries. The Early Data Release 3 constrained
this term and characterized its dependence on magnitude, color, and position [77]. However, MW Cepheids are bright,
resulting in limited sampling of the offset term for these stars. Fortunately, the parallax zero-point offset is an additive
term, while the distance scale derived from parallaxes is multiplicative. This distinction allows for separation of the
two terms, provided a sufficient range of Cepheid distances is available. The analysis of 75 MW Cepheids with high
S/N and HST/WFC3 photometry in Ref. [78] yielded a residual parallax offset of −14 µas for Cepheid-like bright
stars, which aligns well with the majority of independent studies for stars in this magnitude range [79]. MW Cepheids
thus remain the strongest anchor of the distance ladder.
Greater individual precision can be achieved with Cepheids in open clusters by using the average parallax of cluster
members [80, 81]. Cluster members, being fainter than Cepheids, allow for a more accurate calibration of the parallax
offset. In this magnitude range, studies by Refs. [82, 83] find a zero-point consistent with zero, demonstrating good
agreement between P-L relations derived from individual Cepheids and cluster members. These results further confirm
the robust accuracy of Gaia parallaxes across a broad range of magnitudes.

New anchors for the cepheid distance ladder The distance ladder is currently supported by geometric
distances in four galaxies: Gaia parallaxes in the MW [78, 82], detached eclipsing binary (DEB) distances in the
Large [29, 48] and Small [30, 35] Magellanic Clouds, and the water maser distance to NGC 4258 [31, 84]. Adding new
anchors to the Cepheid calibration would reduce uncertainties and strengthen the first rung of the distance ladder.
A galaxy can serve as an anchor if it meets two conditions: (1) having Cepheid photometry measured in the same
system (e.g., HST/WFC3) as the rest of the distance ladder and (2) having a sufficiently precise geometric distance.
The two nearby galaxies, M31 and M33, are excellent candidates for this role, as they contain large Cepheid samples
observed with HST/WFC3 [46, Letters,
The Astrophysical Journal 85]. However, they
962:L17 (13pp), 2024 remain
February 10 beyond the reach of late-type DEBs, which
Riess et al. are too faint
at these distances, or suffer from imprecise geometric distances obtained with other methods [86]. On the other
hand, early-type DEBs are brighter and could enable geometric distances measurements for M31 and M33 [87]. This
method relies on model atmosphere theory, in contrast to the 1% precision surface brightness color relation used for
late-type DEBs in the Magellanic Clouds. Significant advancements in these methods are anticipated [88, 89], and
they should soon provide additional robust anchors for the Cepheid distance ladder.

Future prospects Ongoing surveys (e.g., Zwicky Transient Facility (ZTF)) and the next generation of ground-
based (Rubin, Extremely Large Telescope (ELT)) and space telescopes (Roman) will provide new insights into the
Cepheid distance ladder [90, 91]. These advancements will enable the discovery of new Cepheids through deep
optical time-series observations of distant galaxies, extending the volume of the local Universe where Cepheids can
be studied. High-precision photometry for bright stars in the MW and Magellanic Clouds will be measured by the
PhotSat mission in optical bandpasses, which will also help characterize the extinction law in different environments.
Direct abundance measurements obtained with ground-based ELT will complement current datasets and extend the
range of distances over which different metallicity tracers – such as Cepheids, blue supergiants, or HII regions – can
The Astrophysical Journal Letters, 962:L17 (13pp), 2024 February 10 Riess et al.

Figure 7. Comparison between the standard (SH0ES: R22) dereddened magnitude mWH period–magnitude relation used to measure distances to SN Ia hosts. The red
FIG. 11: Cepheid
pointsP-L relations
use JWST F150W (λ =in 1.50the
eff μm) andWesenheit
the gray points areindex
from HST obtained
0.036[(V − I) − 1.0] to account for the differences in these passbands.
from
F160W (λ = 1.53 HSTF
μm), including
eff a small
160W (grey)
transformation and JWSTF
F150W–F160W = 0.033 + 150W (red),

showing the 2.5 reduction in scatter in NGC 4258 (left) and NGC 5584 (right), figure taken from Ref. [51].
found and which reduces the sample size), and split the sample bias in either measurements or errors (see further discussion
into the more and less crowded halves (where the DOLPHOT in R23).
crowding parameter is above or below the sample median). We also determined the results for a single epoch (the first);
This latter test is meant to mimic the subselection of Cepheids this provides useful information on the dispersion of the P–L
by those that appear visually least crowded with JWST relation when no phase information is available. The single-
resolution, as proposed in Freedman & Madore (2023) for epoch solution yields standard deviation (SD) = 0.224 mag
host NGC 7250 (see also the Appendix). This produces a versus SD = 0.178 for two epochs and phase corrected (plus an
significant difference in sample mean dispersion, namely additional 100 Cepheids with only a single epoch), a 25%
∼0.17 mag for the less crowded half and ∼0.20 mag for the reduction in dispersion or equivalent weight to a 70% increase
more crowded half, but only a 0.015 ± 0.053 mag difference in in the sample size (58% from the scatter and 12% more
H0 between them. Although the difference is small, this objects). With perfect measurements, a single random phase at
selection has the potential to bias the remaining sample because this wavelength will produce ∼0.15 mag scatter, two random
it is enacted on the appearance of the source, which may phases reduces this to ∼0.11 mag, and two phase-corrected
include unresolved blending rather than the statistical proper- magnitudes to ∼0.075 mag. We note that most of these variants
ties of the scene. For example, a coincidental superposition of a are not statistically independent of each other, as they use the
27

FIG. 12: The distribution and Keplerian rotation curve [103, 104] of the 22 GHz H2 O masers in NGC 4258. The
red, blue, and green dots in the plot represent the redshifted, blueshifted, and systemic maser components, respec-
tively.

be compared [62, 63, 92, 93]. Expanding Cepheid samples in both nearby and distant galaxies will allow further
tests of the linearity and universality of the P-L relation, a subject of ongoing debate [94–96]. In SNIa host galaxies,
long-period Cepheids dominate as they are the brightest, while in nearby anchors, they may saturate or have invalid
parallaxes. Conversely, P-L relations in nearby galaxies often include mostly short-period Cepheids. This period
disparity could influence the inferred distances in the event of a P-L break. Recent studies [34, 97, 98] show no
evidence for a non-linear P-L relation and demonstrate that allowing for different slopes does not improve the P-L
fit. [34] also finds the P-L slope consistent with a single value within 1σ across the current sample of host galaxies.
The significance of this effect will be confirmed with larger samples at greater distances. Additionally, the metallicity
dependence of the P-L slope [55, 59] will be clarified through observations in different environments.

2.1.3. Maser driven constraints


Coordinator: Cheng-Yu Kuo
Contributors: Dom Pesce, Jim Braatz, Mark Reid, and Violetta Impellizzeri

For the purpose of measuring H0 using accurate distances external galaxies, the so-called H2 O maser technique,
e.g., see Ref. [99, 100] provides a novel tool that allows one to by-pass the traditional extragalactic distance ladders
and measure the angular-diameter distance to a galaxy in a single step without relying on the CMB. This method
involves sub-milliarcsecond resolution imaging and single-dish monitoring of the 22 GHz H2 O maser emission from
sub-parsec circumnuclear disks at the center of active galaxies. These disk maser emissions, which arise from the
JK− K+ = 616 −523 transition4 of the ortho-H2 O molecule, usually have extremely high surface brightnesses, permitting
mapping with Very Long Baseline interferometry (VLBI) that allows for a unique probe of the gas distribution and
kinematics on sub-parsec scales at the centers of distant galaxies [102].
As revealed in the prototypical maser galaxy NGC 4258, e.g., see Ref. [103, 105], the masing gas in a disk maser
system often resides in a nearly edge-on thin disk around the central supermassive black hole (BH). The rotation
curve of the gas can be nearly perfectly traced by Keplerian rotation, enabling black hole (BH) mass measurements
to percent-level accuracy, e.g., see Ref. [106, 107]. Given the assumption that the gas dynamics is dominated by
the gravity of the central BH, the simplicity of disk geometry and kinematics allow one to use the orbital radii
2
rsys = vsys /asys of the systemic maser components, see Fig. 12, e.g., see Ref. [106], as a standard ruler for measuring
the angular-diameter distance DA of the galaxy, e.g., see Ref. [105], where vsys and asys standing for the orbital
velocity and centripetal acceleration of a systemic maser component. By using rsys as the standard ruler, one can
easily show that the angular-diameter distance of a maser galaxy can be expressed as DA = v02 sin(i)/a∆θ, where i

4 Here, J is the total angular momentum of the H2 O molecule, with K− and K+ representing the projections of J on two molecular axes,
e.g., see Ref. [101].
28

FIG. 13: Left panel: Hubble diagram for the six maser galaxies considered in the MCP H0 measurement [110]. As-
suming a fixed velocity uncertainty of 250 km s−1 associated with peculiar motions, the MCP constrains the Hub-
ble constant to be H0 = (73.9 ± 3.0) km s−1 Mpc−1 , independent of distance ladders and the CMB. Right panel:
posterior distributions for H0 from the five different peculiar velocity treatments considered in Ref. [110], with the
fiducial treatment plotted in black.

and ∆θ indicate the maser disk inclination and the angular radius of the systemic masers, respectively. As long as
the masing components of the gas follow circular orbits, one can determine an accurate galaxy distance by measuring
the four disk parameters including asys , vsys , i, and ∆θ, where a0 can be measured from single-dish monitoring of the
maser lines, e.g., see Ref. [100] and the rest of the three parameters can be obtained by modeling the maser disk in
three dimensions, e.g., see Ref. [108, 109].
To apply this maser technique to galaxies in the local Universe, one has to first search for maser disks similar to
NGC 4258 that allow for robust disk modeling. To this end, the Megamaser Cosmology Project (MCP; [111, 112]) has
carried out an extensive survey of H2 O megamaser emission from >4800 AGNs [113, 114] within redshift z ≲ 0.05,
resulting in the detection of ≳30 candidate disk masers [115]. The follow-up imaging of these candidates has increased
the number of H2 O maser disks with high precision VLBI maps by a factor of ≳4 over the past decade. The resultant
Hubble constant (see Fig. 13) from this effort is H0 = 73.9 ± 3.0 km s−1 Mpc−1 [110], a ∼4% uncertain H0 estimate
determined based on six “clean” disk maser systems that have the required spectral qualities for reliable disk modeling.
This maser-based H0 measurement is well consistent with the majority of direct, late Universe measurements of
Hubble constant. Its uncertainty is currently dominated by by measurement errors in the maser position obtained
from VLBI observation and in the determination of the maser acceleration with the single-dish monitoring. The
systematic uncertainties, which could result from non-circular orbits of the masing gas, e.g., see Ref. [108] or from the
impacts of non-gravitational forces such as shocks in the maser disk, e.g., see Ref. [115–117], are currently negligible
in comparison with measurement errors. It is expected that the accuracy of the maser-based H0 measurement can
be further improved to ∼1%, the long-term goal of the observational cosmology community, by measuring distances
to ≳50 maser galaxies, with ∼7% accuracy per measurement [118]. It is promising that this goal can be realized
after the full array operation of the next-generation Very Large Array (ngVLA), which will bring about an order
of magnitude improvement in sensitivity, permitting efficient detection of disk maser systems in a ∼30 times larger
volume compared with the MCP [118] as well as significant improvement in VLBI maser position measurement.

2.1.4. On the tip of the red giant branch method


Coordinator: Richard I. Anderson
Contributors: Adam Riess, Gagandeep S. Anand, Giulia de Somma, Ippocratis Saltas, Louise Breuval, Siyang Li,
and Vladas Vansevicius

The TRGB method provides the most precise stellar alternative to classical Cepheids (Sec. 2.1.2) on the first and
second rungs of the extragalactic distance ladder used to measure the Hubble constant [119]. Comparison between
TRGB and Cepheid distances in supernova hosts shows very good agreement between both methods, Fig. 11 in
Refs. [16] or [26]. Fig. 14 illustrates two paths to H0 in which the TRGB calibrates either SBF (Sec. 2.1.5) [22, 120]
measured in elliptical galaxies or SNIa (Sec. 2.1.7), e.g., see Ref. [15, 121, 122] measured in any type of galaxy for tracing
the Universe’s isotropic expansion in the Hubble flow. In both cases, the TRGB serves as an intermediary to translate
the relative apparent magnitude differences of SBF or SNIa as a function of redshift to an absolute scale anchored to
geometrically measured distances. As part of the past decade’s unfurling Hubble tension, the TRGB method has seen
major improvements and inspired deep investigations into systematics of late-Universe H0 measurements. Here, we
briefly review the astrophysical basis of the TRGB as a standard candle to determine luminosity distances, present
29

methodological considerations relevant for determining TRGB distances to better than ∼ 5%, and consider likely
future developments relevant for the H0 tension in light of catalysts provided by new observational facilities and
upcoming space missions. [123] provides further background and in-depth discussion.

Surface Brightness Fluctuations Type-Ia supernovae (SNeIa)

TRGB

TRGB
s
eid
ph
Ce

SNIa
J+21 F+11 R+22

𝐻! ≈ 𝑐𝑧/10!.%('(%))
𝜇 " calibrates SBF or SNeIa
to trace Hubble flow
𝜇 " = 𝑚 " − 𝑚 ",! + 𝜇!
𝑚!

𝑚 !,#
W+23

𝜇! from geometry
e.g., Gaia, MCs, NGC4258 A+24

FIG. 14: Illustrations of extragalactic distance ladders built on the TRGB method. The scales considered increase
upwards. Bottom right: The TRGB feature is calibrated based on color-magnitude diagrams of stars whose dis-
tances (µ0 ) are known from geometric measurements, such as Gaia parallaxes, the Magellanic Clouds, or NGC4258.
The choice and systematics of (the combination of) geometric anchors directly impact the Hubble constant by set-
ting µ0 . Center left: The apparent magnitude of the TRGB feature is determined in other galaxies. Differences
in the absolute magnitude between TRGB populations, either compared to the calibrating set or as field-to-field
variations, must be mitigated by standardization. Upper part, right: The measured TRGB distances calibrate the
fiducial absolute magnitude of SNIa in any type of galaxy. Cross-checks between multiple stellar standard can-
dles can primarily be obtained in galaxies hosting young, intermediate-age, and old stellar populations at the same
time. Note that TRGB fields should be placed on galaxy halos populated by old, metal-poor stars, whereas classi-
cal Cepheid fields target young stellar populations in supernova host galaxies. Upper part, left: Alternatively, the
TRGB is used to calibrate the absolute scale of the SBF method for distance determination to elliptical galaxies
that trace the Hubble flow. SBF exploits the fact that the variance of SBF decreases as d2 .
Figure credit: R.I. Anderson based on images as labeled from: J+21 [124]; F+11 B. J. Fulton, Las Cumbres Obser-
vatory; R+22 [34]; W+23 [125], A+24 [126].
30

-6 M F 1 5 0 W

-4
M I

M
-2
M V

-6 M F 1 5 0 W

M F 0 9 0 W
M I
-4

M V
M

-2

(V - I)+ 1 (V - I)+ 2 (V - I)+ 3


0

1 2 3 4
V - I
h

FIG. 15: Dependence of MT on chemical composition and age based on PARSEC v.1.2S isochrones [131, 132]. Top:
8 Gyr isochrones for increasing metallicity (from left to right: [M/H]= −2.2, −1.5, −0.7, −0.4, 0.0) in Johnson V
(dark blue), Cousins I (cyan), and JWST/NIRCAM F150W (red) passbands. Bottom: [M/H]= −1.5 isochrones
for increasing ages (left to right) of 2 Gyr (blue), 4 Gyr (cyan), 8 Gyr (green), and 14 Gyr (red) in the Johnson V ,
Cousins I, JWST/NIRCAM F090W and F150W passbands, successively offset in color for visibility.

2.1.4.a. The astrophysical basis The TRGB as a standard candle is a robust empirical concept that is usefully
supported by a solid understanding of the evolution of QSO stars. Empirically, the magnitude mT is measured as
the inflection point of a QSO Luminosity Function (LF) [126, 127]. Measuring mT therefore requires a large number
of stars to avoid stochastic effects and, in particular, cannot be done for each star separately. This characteristic is
shared with the J-region AGB method [128, 129] (Sec. 2.1.8) and distinguishes the TRGB as a statistical standard
candle from pulsating stars, such as classical Cepheids, whose luminosity can be measured and calibrated individually
[47, 130].
The TRGB feature is astrophysically explained by the very rapid thermonuclear ignition of degenerate helium cores
of low-mass first-ascent QSO stars, the so-called helium flash (HeF), see e.g., [133–137]. In particular, old QSO with
masses ≲ 1.6 M⊙ develop helium cores of nearly equal mass (∼ 0.5 M⊙ ), resulting in a nearly equal luminosity of
the HeF that can be exploited as a standard candle. Younger, higher-mass giants reach lower luminosity due to
lower electron degeneracy at higher core temperature and lower density. This mass limit corresponds to an age-
effect: giants older than ∼ 4 Gyr have nearly constant luminosity in I−band. Differences in chemical composition
(metallicity) significantly affect TRGB luminosity and temperature due to line blanketing: the bolometric magnitude
behaves as MT,Bol ∝ −0.19[Fe/H] [136]. At the same time, higher metallicity results in cooler effective temperature.
Hence, more metal-poor giants are brighter and bluer. Thankfully, the dependence on metallicity is partially mitigated
in Cousins I−band where the bolometric correction (MI = MBol − BCI ) nearly compensates the change in luminosity
for low-metallicity ([Fe/H] ∈ [−0.7, −2.0]) giants. Color-based metallicity calibrations can further mitigate this issue,
e.g., see Ref. [138, 139]. Hence, the TRGB is a theoretically well-supported standard candle when low-metallicity
31

([Fe/H] < −0.7) old (> 4 Gyr) QSO stars are observed in I−band. However, observations at infrared wavelengths
require caution and additional study with respect to the aforementioned effects, in particular if the observed QSO
populations exhibit diversity in age and/or chemical composition, see Fig. 15.
Theoretical predictions of the TRGB luminosity rely on stellar evolution models that require various assumptions
and simplified treatment of stellar physics, including the modelling of opacities, diffusion, convection (mixing-length
theory) and overshooting, mass-loss, electron screening, neutrino losses [140], nuclear reaction rates, among others
[137, 141]. In particular, the dominant uncertainties related to radiative opacity introduce systematics of up to 1.6%.
Simultaneously varying radiative and conductive opacities, nuclear reaction rates (e.g., triple-alpha reaction), and
neutrino losses affects the bolometric Tip luminosity by δL/L ∼ 10−3 , depending on mass and metallicity [142].
Comparisons with observations are subject to larger uncertainties, notably related to the translation of bolometric
luminosity to magnitude, which require stellar atmosphere models that are usually considered separately from the
evolutionary models [137].
2.1.4.b. Methodological considerations and HST-based results
The absolute calibration of the TRGB method requires measuring mT in QSO populations of known
distance to obtain MT = mT − µ0 . The best available, geometrically measured, distances (in ascending order) are:
trigonometric parallaxes from the ESA Gaia mission [143], the distances to the LMC and SMC measured by detached
eclipsing binary stars [29, 30], and the megamaser distance to NGC 4258 (M106) [31]. Importantly, the distance
moduli µ0,i of the LMC, SMC, and NGC 4258 are frequently used in distance ladders calibrated using stellar standard
candles and hence their systematics are often shared among different H0 measurements, e.g., see Ref. [34, 35, 144, 145].
Comparing distances to other galaxies measured using an absolute calibration based on the same anchor, such as
NGC 4258, thus yields the strongest test of distance systematics [e.g., 26, 146]. As implied in Fig. 14, uncertainties in
µ0 propagate directly into H0 measurements, as they set the absolute scale of the secondary tracers (SNIa, SBF) via
the primary standard candle (here: TRGB). Basing an absolute TRGB calibration on the broadest possible set of
“anchors” reduces possible bias and allows one to determine appropriate standardization procedures, e.g., to account
for metallicity differences [121, 138]. At present, the most accurate TRGB calibrations to date are obtained in the
Magellanic Clouds [126, 147] and NGC 4258 [16] using passbands similar to I−band, such as HST’s ACS/F814W
or JWST’s NIRCAM/F090W. TRGB calibration based on Gaia parallaxes [148] has been achieved based on EDR3
parallaxes, which require correction for complex systematics, e.g., see Ref. [77, 78, 83, 149–152]. Furthermore, a
parallax-based TRGB calibration has been obtained for ω Centauri [153]. The period-color relation of stars near the
TRGB may provide a useful avenue to deal with (differential) reddening when calibrating the TRGB based on field
RGs [147].
A critical element of determining accurate distances using standard candles is to ensure consistency between the
absolute calibration and the standardized application in the target environments. However, there are measurable
differences even between the I−band absolute magnitudes in the LMC and the SMC, and they reflect noticeably
in the variability periods of the small-amplitude QSO stars that make up the QSO population at the Tip [126].
These ubiquitous small amplitude pulsations near the QSO Tip allow to probe the intrinsic diversity (e.g., in age
and metallicity) ofQSO populations that is not typically known a priori and thus constitute a difficult-to-control
astrophysical systematic at the level of a couple of percent [147]. Additionally, a common feature of QSO populations
in other galaxies is contamination by other stars, notably at higher luminosity (typically AGB stars). Unfortunately,
the specifics and degree of such contamination differ from environment to environment (cf. CATs below) and are not
reliably known a priori. The exception to this rule are globular clusters (GC), which rarely contain stars brighter than
the TRGB, e.g., see Ref. [154], so that measuring mT conceptually corresponds to searching for the brightest cluster
star without knowing how much it differs from the HeF luminosity. As a result, GC-based values of mT represent
lower limits to the desired measurement of mT and are furthermore also subject to variations in metallicity and age,
depending on the photometric band. Thus, GC-based mT values differ conceptually from the values of mT measured
in QSO populations of mixed age and metallicity. Further study is needed to assess to what degree such effects impact
MT , and hence, H0 .
Extinction corrections are routinely applied for TRGB calibration in the Magellanic Clouds [155]. However,
dust corrections to TRGB measurements in other galaxies typically, e.g., see Ref. [144] rely on all-sky dust extinction
maps [156, 157] that account only for MW foreground dust and whose accuracy is limited in the vicinity of resolved
galaxies. Recently, it has been pointed out [158] that small, albeit non-zero, extinction in galaxy halos leads to
underestimated H0 values and that dust extinction estimates based on background RGBs [159] provide a possibility
for correcting this one-sided systematic. Statistical corrections for local extinction have since been applied [16, 122].
However, further study of extinction effects and their variation across circumgalactic media would be useful to provide
improved corrections.
The TRGB magnitude (mT ) is measured either via an edge detection algorithm [119, 127], such as a Sobel
filter, or via a maximum likelihood fit of the LF [79, 160, 161]. When Sobel filters are used for edge detection, LFs
32

are typically smoothed to some degree to reduce noise [127], and different weighting schemes have been considered
in the literature [125, 162]. Recently, it has been pointed out that both the LF smoothing and the Sobel response
weighting introduce biases that depend on the shape of the observed LFs [126]. In particular, Sobel response weighting
was shown to introduce the tip-contrast relation determined independently using observations [125]. Both issues can
easily bias the measured value of mT by 0.06 mag (∼ 3% in distance), and it is therefore crucial to apply the same
TRGB measurement method in all target QSO populations to avoid bias. Similarly, observer choices, such as color
cuts (notably near the Tip), can affect the measured mT and should be decided based on objective criteria [122].
Field-to-field variations of mT have been recently reported within the same galaxies. This led to the develop-
ment of an unsupervised TRGB detection algorithm, called CATs (Comparative Analysis of TRGBs), which sought to
reduce the impact of subjective observer choices on the measurement of mT [122, 125]. In the process, a tip-contrast
relation was determined, which has since been shown to result from the use of weighted Sobel response curves [126].
Bias of TRGB distances can thus be avoided if either a) unweighted Sobel filters are used to measure mT [126], or b)
if an appropriate tip-contrast relation is used to standardize mT a posteriori when weighted Sobel response curves are
used [122]. Given the simpler algorithm and the non-uniqueness of the tip contrast, which depends on the measured
mT , it appears prudent to avoid Sobel filter weighting. First results from an I−band (F090W) TRGB calibration
using the JWST also support this approach [16].
In observations based on the Hubble Space Telescope or JWST, the placement of TRGB fields in target galaxies is
crucial to ensure that the targeted QSO population is similar or standardizeable to the QSO population that provides
the absolute calibration [163], for an example, e.g., see Ref. [146]. Thanks to very large fields-of-view, observations
with the ESA Euclid mission [164] or the future Nancy Grace Roman telescope 5 will allow us to integrate field-to-field
variation analysis into TRGB algorithms, rendering them much more robust. However, the existence of field-to-field
variations within a single galaxy highlights the need for TRGB standardization to account for differences between
QSO populations in different galaxies.
The choice of photometric system The TRGB has been most commonly measured in the I−band where
it is nearly flat and relatively insensitive to metallicity and age effects [119, 145], cf. Fig. 15. There has been
growing interest in NIR TRGB measurements due to the capabilities of the JWST and the intrinsically higher
(1 − 2 mag) luminosity in the NIR [165, 166], and several I−band and NIR observing programs are currently being
analyzed with JWST (e.g., GO-1685, 1995, 2875, 3055). However, the (color) slope of the TRGB at NIR wavelengths
requires additional consideration. Both empirical (see Refs. [165–172]) and theoretical approaches [137, 173] have been
considered for the time being, and further study is required to determine competitive and accurate TRGB distances
at infrared wavelengths. Furthermore, K−corrections are expected to exceed 1% in distance for single-band JWST
TRGB observations at distances above ∼ 70 Mpc for the F150W passband and shorter wavelengths [158].
2.1.4.c. Implications for H0 and the Hubble constant tension Several studies have reported H0 measurements
involving the TRGB as a standard candle, with rather significant differences. To understand these differences, it
is crucial to consider the differing assumptions and methods underlying each analysis. In particular, Refs. [144]
and [121] applied different absolute TRGB calibrations and TRGB measurement methodologies to the same set
of HST I−band TRGB observations in conjunction with the CSP SNIa dataset in the Hubble flow and measured
H0 = 69.8±0.6±1.6 km s−1 Mpc−1 and 71.5±1.8 km s−1 Mpc−1 , respectively. In an update to Ref. [144], Ref. [174](v3)
employed JWST I−band (F090W) TRGB measurements together with CSP SNIa and obtained H0 = 70.4 ± 1.2 ±
1.3 ± 0.7 km s−1 Mpc−1 . Ref. [122] corrected the tip-contrast TRGB systematic and replaced the CSP SNIa with
Pantheon+ to ensure a consistent calibration between SNIa in host galaxies and the Hubble flow and found 72.94 ±
1.98 km s−1 Mpc−1 . Replacing HST entirely with JWST I−band TRGB observations of 17 unique SNIa yields 72.1 ±
2.2 ± 1.2 km s−1 Mpc−1 [26]. Recently, Ref. [120] replaced SNIa entirely with SBF calibrated by JWST I−band TRGB
observations [16] and found 73.8 ± 0.7 ± 2.3 km s−1 Mpc−1 , confirming the conclusion from Ref. [122] that at least half
the difference between the H0 values reported by Ref. [144] and Ref. [82] can be attributed to SNIa alone.
The prospects for further improvements of the TRGB method and, in turn, to increase the accuracy in H0 are very
promising. New wide-field space-based telescopes (Euclid, Roman) will allow us to comprehensively measure mT in a
very large number of galaxies, while JWST will allow to push the limits of TRGB distances. Potentially, Roman could
allow to directly calibrate mT in the LMC and SMC, circumventing differences among photometric systems. However,
the target magnitudes are close to saturation (∼ 6s at AB 15 mag in F0876 ). An SBF calibration based on the TRGB
method will yield an accurate H0 measurement fully independent of the Cepheid-SNIa distance ladder [22, 120], if
different geometric distances are used to calibrate the two types of standard candles. Future improvements in the
Gaia astrometric solution will play an important role in this endeavor, as will synthetic Gaia photometry based on

5 https://roman.gsfc.nasa.gov
6 https://roman.gsfc.nasa.gov/science/apttables2021/table-timetosaturation.html
33

Bp and Rp spectra [175]. Pushing the TRGB to the limit using JWST may motivate infrared observations owing to
the combination of QSO stars being brighter in the near-IR and JWST providing optimal sensitivity there, although
it has been already shown that I-band JWST TRGB measurements are feasible out to at least 50 Mpc [22, 176].
In order to ensure that stellar standard candles provide the clearest and most accurate picture of cosmic tensions,
notably the Hubble constant tension, it is crucial to pursue the most direct and simple assessments of systematics.
To this end, direct and detailed comparisons of distances measured using the TRGB method, classical Cepheids, and
the JAGB method are preferred, cf. [26]. In contrast, trying to understand TRGB or Cepheid-based systematics
from published values of H0 is complicated by several other possible differences between distance ladder set ups, e.g.,
with respect to the treatment of peculiar motions and supernova standardization [122]. Cross-checks based on nearby
galaxies, e.g., M31 [85], M33 [46, 177], and other nearby galaxies (e.g., HST GO-17520), will be particularly insightful
to this end as they provide the greatest possible precision for each of the standard candles and an optimal ability to
investigate causes of systematic differences, e.g., using spectroscopy on future 30m-class telescopes, such as the ELT.

2.1.5. The surface brightness fluctuations method


Coordinator: Michele Cantiello
Contributors: Enzo Brocato, Gabriella Raimondo, John Blakeslee, Joseph Jensen, and Rebecca Habas

2.1.5.a. Introduction & state of the art The SBF method is a powerful technique used to measure distances of
elliptical galaxies out to ∼ 150 Mpc, and may reach distances as far as ∼ 300 Mpc with telescopes like the JWST
[22]. Introduced in the late 1980s, the SBF method relies on the fact that the surface brightness of a galaxy exhibits
small-scale variations due to the statistical distribution of stars [178–182]. When imaging a galaxy, the number of stars
that fall within each detector pixel varies, causing pixel-to-pixel variations in the surface brightness. The amplitude of
these fluctuations is inversely proportional to the square of the distance to the galaxy [183], allowing one to estimate
distances from imaging alone.
In principle, the SBF method is straightforward. The only difficulty lies in isolating the signal fluctuations of the
underlying Population II stars from contaminants such as globular clusters, star forming regions, and background
galaxies. Once this is done, the absolute SBF magnitude can be calibrated in a given passband, M ξ , using galaxies
at well-known distances [184] or through stellar populations numerical modeling [185]. The distance modulus µ0 ≡
(m−M ) = 5 log 10(D) + 25 can then be inferred as usual: µ0 = mξ − M ξ where mξ is the ξ-band observed fluctuation
magnitude and D is the distance in Mpc.
In practical terms, the SBF distance measurement involves several steps, summarized briefly here and illustrated in
Fig. 16: i) determining and subtracting the sky background; ii) modeling and subtracting the galaxy (upper middle
panel in Fig. 16); iii) masking all potential sources of contamination to the fluctuation signal (e.g., globular clusters
within the galaxy, background galaxies, dust patches, etc.; upper right panel); iv) modeling the LF of sources in
the frame to estimate the spurious fluctuation contribution from unexcised sources (lower left panel); and v) power
spectrum analysis of the residual masked frame, to measure the amplitude of the SBF signal (lower middle and right
panels). Additional details on this procedure are available in the references cited in this review.
In recent years, the application of the SBF method has seen significant advancements, particularly with the con-
tributions from the high-resolution space-based data from the HST. Recently, Refs. [124] and [21] used HST/WFC3
data to measure highly accurate distances for a sample of ∼ 60 elliptical galaxies, taking particular care to analyze
the M F 110W calibration, and succeeded in obtaining distances with median statistical uncertainties ≤ 4%.
From the ground, the Next Generation Virgo Cluster Survey (NGVS), a project based on 104 degrees2 deep optical
imaging from the CFHT [187], has provided a dataset ideal for SBF measurements in the Virgo Cluster. The NGVS
has enabled detailed studies of galaxy distances in this cornerstone galaxy cluster [186], and represents a crucial
survey, serving as a precursor for future large ground- and space-based surveys.
Most of the focus for the use of SBF has been on bright galaxies, where the method provides the lowest uncertainty,
reaching levels ≤ 5% on single targets in well-designed observations. The advent of large area deep surveys has,
however, led to the discovery of an increasingly large number of dwarf galaxies which play a fundamental role in
Cosmology [188]. Several authors [189–191] have applied the SBF technique to smaller and fainter systems. These
studies have demonstrated that m can be used to estimate distances to dwarf galaxies, although with relatively large
errors compared to massive galaxies, providing a typical uncertainty of δD/D ∼ 15%. Despite the relatively larger
error, these elusive galaxies are often challenging targets for distance measurement due to their low surface brightness
and complex stellar populations. This has opened new avenues for using SBF to probe the Local Universe and improve
the cosmic distance ladder.
2.1.5.b. The Hubble constant tension In addition to providing distance estimates for individual galaxies, SBF
can also be used to probe tensions in the measurement of the Hubble constant that have arisen between the CMB
observations [192] and the value obtained from the late-Universe [34].
34

FIG. 16: SBF analysis for the galaxy VCC 1146. Top row: i-band image, residual image, and masked residual im-
age (left to right). Black annuli in the second and third panels mark the inner and outer radii used for SBF mea-
surements. Bottom left: Fitted LF for external sources. Green circles are observational data with a faint-end
downturn from incompleteness. The solid green line shows the best fit, corrected for incompleteness, with blue dot-
ted and red dashed curves representing background galaxy and GC LFs. Bottom middle: Azimuthal average of
the residual power spectrum (gray circles) and the fit (solid black line). Bottom right: Fitted P0 values vs. kstart ,
with kend = 450. Vertical lines at kstart = 80 and 160 show the range over which the fit is stable. The median P0 in
this range (dotted horizontal line) is adopted. Figure adapted from Ref. [186].

The SBF method, with its ability to provide distance measurements to galaxies beyond the 50 Mpc range, offers
a unique perspective on this tension, independent of the classical Cepheid/SNIa route. This topic has been explored
by Ref. [21] for a sample of bright galaxies between 20-100 Mpc, with M F110W calibrated using both Cepheids and
TRGB distances. For both calibrations, the authors derived a value of H0 that is fully consistent with the direct
distance estimates in the local Universe. The work by Ref. [21] was based on a small sample, and this result should
be corroborated with larger studies. That study, as well as some of the newly programmed ones [22], used SBF in
near-IR passbands where the amplitude of fluctuations is “enhanced” because the brightest stars contributing to the
signal are brighter compared to the optical bands. This enhancement is key for the recent results on H0 and will be
crucial for future observations, as it allows for reliable measurements for galaxies at larger distances.
A complementary approach to the H0 tension uses SBF indirectly, as a calibration for SNIa distances. Rather than
the Cepheid-SNIa route to Cosmology, one can use a Cepheid-SBF-SNIa route. This method has larger systematic
and statistical uncertainties from the extra rung in the distance ladder, but offers the advantage of having a larger
calibration sample of SNIa from the local Universe because of the larger overlap in galaxies that host SNIa and have
SBF measurements. This approach was adopted by Ref. [193], who used a heterogeneous collection of SNIa and
SBF measurements from existing literature, obtaining H0 = 71.1 ± 2.4(stat) ± 3.4(sys) km s−1 Mpc−1 . More recently,
Ref. [194] used a very homogeneous sample of SBF measurements from Ref. [124], and found that the Hubble-
Lemaître parameter derived from the revised SALT2 [32] parameter fit and calibrated with near-IR SBF distances
is H0 = 73.3 ± 1.0(stat) ± 2.7(sys) km s−1 Mpc−1 . Note that the SNIa-independent estimate from Ref. [21], based
on Cepheid zero-point calibration, yielded H0 = 73.3 ± 0.7(stat) ± 2.4(sys) km s−1 Mpc−1 , whereas the most recent
recalibration, which relies on TRGB distances from JWST observations of Virgo and Fornax galaxies by Ref. [120],
provided H0 = 73.8 ± 0.7(stat) ± 2.3(sys) km s−1 Mpc−1 .
2.1.5.c. Future prospects and projects The SBF method is expected to become even more relevant in the future,
taking advantage of the fact that most of the forthcoming facilities will operate in the near-IR regime, where the
35

SBF signal is much stronger. The JWST, with its superior resolution and sensitivity, is expected to enable detailed
SBF studies of distant galaxies, potentially reaching ∼ 300 Mpc [22]. The large sky surveys from the Vera Rubin
Observatory’s LSST and from the Euclid mission, will further enhance the capabilities of SBF by providing an
enormous dataset for SBF, largely dominated by the dwarfs. This will allow astronomers to apply the SBF method
to a much larger and more diverse sample of galaxies, improving the statistical robustness of the 3D mapping of the
Universe within 40-70 Mpc.
On longer time scales, the Nancy Grace Roman Space Telescope (to be launched in the mid-late 2020s), will also
play a crucial role in the advancement of SBF, due to its combination of wide-field, high-resolution, and infrared
wavelength coverage. Roman will enable high-precision SBF measurements over large areas of the sky, potentially
reaching the depth of the JWST, but over a wider area. The future 30-40 m class ground based telescopes, with their
adaptive optics (AO), may also be more useful for SBF than the JWST. Correcting for atmospheric distortions using
the AO systems could provide better FWHM resolutions than the JWST in certain passbands (e.g., K-band) which,
combined with the huge telescope collecting area, would allow SBF to be applied to more and more distant systems.
Although time-consuming, because of the specific requests of AO observations of extended objects and the needs for
AO activation loops, observing a handful of well-chosen targets with one of these massive telescopes could allow one
to discriminate between different cosmological models, independently from SNIa.
Beyond improved telescopes, SBF observations also benefit from improved numerical stellar population studies.
Accurate models of stellar populations are essential for interpreting SBF measurements, reducing systematic uncer-
tainties and deriving reliable k-corrections which cannot be ignored beyond ∼ 150 Mpc. Additionally, the application
of SBF to blue, low-mass galaxies requires the calibration of M across a broader range of stellar populations and
environments. Stellar population modeling will offer a reference to test and prove the reliability of the SBF method
over a wide interval of population properties.
Theoretical efforts over the past 20 years have focused on incorporating stars in peculiar evolutionary stages, such
as thermally pulsing AGB stars and hot HB stars in old-intermediate populations (e.g., see Ref. [185, 195]). These
efforts have also included the impact of α-elements and He-abundance enhancements. Advancements in this area
will benefit from the insights gained by applying theories of late stellar evolution —dominated by poorly understood
physics such as mass loss processes— to the observations in mid-infrared instruments like MIRI on the JWST. This
will result in more refined stellar population models for distance estimates and a better understanding of these crucial
stages of stellar evolution.
Future SBF measurements can also benefit from the adoption of ML techniques. ML methods require training on
large samples which are not yet available, but when they are, ML algorithms can optimize the measurement of SBF,
which is currently slow and heavily reliant on human intervention. ML can further identify patterns and correlations
in large datasets, improving the efficiency of SBF measurements and reducing measurement time. Sufficient training
datasets will be available after the first years of Euclid or LSST public data releases.
In conclusion, the SBF method currently stands as a robust and versatile tool in extragalactic distance measure-
ments. To date, the method’s range of usefulness allows it to be adopted for tests on the H0 tension, providing results
that agree with the late-time/direct Hubble constant estimates. With advancements in telescope technology, data
processing, and collaborative efforts, the SBF method is well-positioned to make significant contributions to resolving
cosmic tensions, possibly even beyond the H0 parameter, and enhancing our understanding of the Universe. By
implementing the proposed future developments, the astronomical community can further refine the SBF technique
and extend its applicability to new frontiers in extragalactic astronomy.

2.1.6. Mira variables


Coordinator: Caroline Huang
Contributors: Antonio Capodagli, Lucas Macri, and Massimo Marengo

Mira variables are fundamental-mode, thermally-pulsing AGB stars with periods ranging from ∼ 100 − 1000 days or
longer. They fall into a broader category of variables known as Long-Period Variables (LPVs), which includes semi-
regular variables and OGLE small-amplitude QSO stars, which are overtone or irregular pulsators. The pulsation of
Miras is likely driven by a κ-mechanism similar to that found in Cepheids. As highly-evolved stars, they contribute
to the chemical enrichment of the ISM through stellar mass loss, have low effective temperatures (Teff < 3500 K)
with spectral intensity peaking between 1-2 µm, and are often amongst the brightest stars in an intermediate-to-old
population (L ∼ 104 L⊙ ). They are also ubiquitous as nearly all stars (0.8M⊙ < M < 8M⊙ ) will experience a Mira
phase in evolution.
Like other AGB stars, they may be classified into two main spectral types based on their photospheric carbon-to-
oxygen ratio (C/O ratio). The exact boundaries are somewhat fluid, but typical C/O values are: C/O ratio ≳ 1 are
classified as C-rich (C-type), C/O ratio ≲ 0.5 are O-rich (M-type), and C/O ratio ∼ 0.5 − 1.0 are intermediate, or
S-type. While the identification of carbonaceous or silicate molecular features in stellar spectra is the gold standard for
36

classification, in practice, spectra are difficult or expensive to obtain and colors, or color-color diagrams, are typically
used as a proxy for spectral type.

µM 101 = 29.10 mag


21 σ = 0.24 mag
F160W Magnitude

22

23

2.40 2.45 2.50 2.55 2.60


Log Period (Days)

FIG. 17: Period-Luminosity Relation for short-period, presumed O-rich Miras in the SNIa host galaxy M101 from
Ref. [20].

The first Period-Luminosity Relations (PLRs) derived for Miras used bolometric magnitudes [196, 197]. Over the
past several decades, Miras have been observed in bands ranging from optical to far-infrared have been used to
measure distances to many Local Group galaxies (see reviews by Refs. [198–200] and references therein) and map
the structure of the MW [201]. Studies of nearby dwarf galaxies have suggested that metallicity does not appear to
have an observable effect on the pulsational properties of these stars [202, 203], although possible metallicity effects
have been suggested by theoretical models. Recently, nonlinear stellar pulsational models have shown significant
improvements in matching theoretical predictions of long-period variable PLRs with observations [204] which may
shed light on this apparent discrepancy in the future.
2.1.6.a. Application For the purposes of this review, we will focus on the O-rich Miras, which are more commonly
used as distance indicators because they follow PLRs with lower scatter in near-infrared wavelengths (∼ 0.12 mag)
compared to their C-rich counterparts. In order to create the PLRs, time-series observations — often with either an
uneven, power-law spacing or monthly sampling — are also required in order to determine periods.
The first rung of Mira distance ladder uses Hubble Space Telescope observations of Miras in the water megamaser
host galaxy NGC 4258 and ground-based near-infrared observations of Miras in the LMC [205, 206] as “anchor”
galaxies to obtain an absolute magnitude calibration [19]. In order to tie the ground-based observations to the HST
photometric system, Ref. [207] used a photometric transformation derived from O-rich Mira spectra. Unlike the
Cepheid distance ladder, there is currently no precise, parallax-based absolute magnitude calibration. This is due
to the fact that Miras (as well as AGB stars in general, and other evolved stars such as red supergiants) exist at
the intersection of many difficulties for Gaia — they are highly luminous, red, have angular sizes larger than their
parallaxes, and are known to have photocenter variations, all of which can contribute to underestimated uncertainties
and bias in the inversion of their parallaxes [208–211].
The second rung of the distance ladder is currently built on observations of Miras in two nearby Type Ia Supernova
host galaxies — NGC 1559 and M101 (hosts of SN 2005df and SN 2011fe respectively, with PL relation for M101 shown
in Fig. 17) [19, 20]. Thus far, these are the only two SNIa host galaxies with Mira-based distances. The statistical
uncertainty in the SNIa calibration is also the dominant source of error in the Mira-H0 measurement (even after
standardization, the uncertainty in SNIa peak magnitude is ∼ 0.1 mag). Thus the greatest reduction in uncertainty
will be obtained from obtaining more observations of Miras in SNIa host galaxies.
37

2.1.6.b. Implications for Hubble constant and Hubble constant tension The current most precise Mira-based mea-
surement yields H0 = 72.37 ± 2.97 km s−1 Mpc−1 (∼ 4% total uncertainty, including both statistical and systematic
components) [20]. While it is not yet as precise as more established distance indicators, as an independent measure-
ment, this result does support the hypothesis that the local measurement of H0 is greater than the early-Universe
measurement with ∼ 95% confidence. In addition, it provides an important cross-check to distances made with other
more established distance indicators such as Cepheids and TRGB.
2.1.6.c. Prospects The next decade should be particularly exciting for using Miras and other LPVs as distance
indicators. In the NIR/MIR bands accessible with HST, JWST and Roman, Miras have PLRs with smaller scatter
than in the optical, nearly-sinusoidal light curves, and are relatively easy to identify as some of the brightest stars in
resolved stellar populations. The Vera C. Rubin Observatory’s LSST is expected to have sufficient depth and number
of epochs to enable the detection of Miras at optical wavelengths (griz) [212] in galaxies out to D ∼ 10 Mpc [213].

2.1.7. Type Ia supernovae


Coordinator: Maria Vincenzi
Contributors: Jenny Wagner, Lluís Galbany, Luca Izzo, M. Pilar Ruiz Lapuente, and Sanjay Mandal

SNIa are thermonuclear explosions of carbon oxygen white dwarfs in close binary systems. Their peak brightness after
various empirical corrections is highly homogeneous (r.m.s. <0.15 mag in B-band), hence their usage in cosmology
as standardizable candles. SNIa have played and still play a crucial role in measurements of the Hubble constant,
and they are typically used in two out of three rungs in the typical distance ladder. In the second rung, their
luminosities are calibrated with stellar distance indicators, e.g., Cepheids or TRGB or JAGB, in the same galaxy. In
the third rung, their brightness values calibrate the Hubble-Lemaître relation into what is now deemed the ‘Hubble
flow’ (∼100 to 600 Mpc, or z < 0.15), where the cosmological expansion dominates over the peculiar motions of
galaxies. Ideally, one could remove the intermediate step (second rung), and go straight from geometric calibration
to SN. Unfortunately, the rate of SN in the local Universe is not nearly frequent enough to provide multiple SN
in the few galaxies used for geometric anchors (the MW, the LMC, or the mega-maser NGC 4258). Even with
Cepheids/TRGB/JAGB as a go-between, the low rate of SN in the nearby Universe (roughly one per galaxy per 100
years) is the limiting component of the precision of H0 measurements, for instance Ref. [34] utilizes every SNIa that
pass cosmological quality requirements and Cepheid suitability within 40 Mpc. The most comprehensive three-rung
ladder H0 measurements using Cepheids is shown in Fig. 8.
2.1.7.a. SN Ia standardization and derivation of H0 We review here the formalism for deriving the Hubble
constant with SNIa in the local distance ladder, as presented by Ref. [34]. Using a set of Cepheids or TRGB distance
moduli (µ0 ) calibrated with geometric measurements such as parallax or megamasers (first rung in the distance
ladder) and comparing these distances to brightnesses mX of SNIa exploding in the same galaxies, we can estimate
the single offset MB , which describes the absolute magnitude of an SNIa (second rung). Following the SALT modeling
framework [214, 215],7 SNIa standardized brightness, mX ,8 is generally measured with the Tripp formula [218] such
that:
mX = mB + αx1 − βc − δBias + δHost , (2.1)
where mB , x1 and c are all independent properties of each light curve derived using the SALT light-curve model, α
and β are correlation coefficients that help standardize the brightness, δBias is a correction due to selection effects and
other biases as predicted by simulations, and δHost is a final correction due to residual correlations with host galaxy
properties.
For a SNIa in the i-th Cepheid host,
mX,i = µ0,i +MB , (2.2)
where MB is the fiducial SNIa absolute magnitude (assumed to be the same across the whole sample), and µ0,i is the
distance modulus derived from Cepheid measurements for the ith galaxy. The ladder is completed with a set of SNIa
that measure the expansion rate quantified as the intercept, aB , of the distance (or magnitude)–redshift relation. For
an arbitrary expansion history and for z > 0 as
 
1 1
1 − q0 − 3q02 + j0 z 2 + O(z 3 ) − 0.2mHF (2.3)

aB = log cz 1 + [1 − q0 ] z − X ,
2 6

7 The SALT2 and SALT3 frameworks has been used in all most recent cosmological analyses. Other frameworks are also available, i.e.,
SnooPy [216], BayeSN [217].
8 We call the standardized brightness mX instead of mB as in Ref. [34] to be clear that the brightness is standardized.
38

SurveyRedshift uncertainties
Selection Efficiency
SN Astrophysics
(different dust and intrinsic scatter models)

Intra-survey Calibration

stat + syst

stat + syst
LightCurve model training
Absolute Calibration

syst

syst
H0

H0

H0

H0
Intra-survey photometric Calibration
Baseline
0.6 0.4 0.2 0.0 0.2 0.4 0.6
H0 [km s Mpc 1 1]

FIG. 18: Adapted from Ref. [33], the impact on recovery of H0 of the various systematic uncertainties tabulated.
The units of these measurements are km s−1 Mpc−1 . The dashed lines are given at ∆H0 of 0.7, which is the entire
contribution of the uncertainty in Ref. [34] from SN measurements. We labeled the different categories of system-
atic uncertainties.

measured from a set of SNIa (z, mHF


X ) in the Hubble flow, where z is the redshift due to expansion, q0 is the deceleration
parameter, and j0 is the jerk parameter. Typically, for ΛCDM, j0 is set to 1. The determination of H0 follows from

log H0 = 0.2MB +aB +5 . (2.4)

Covariances between rungs are taken into account following the approach of Refs. [219] and [220]. Finally, q0 (and
j0 ) can only be constrained from SNIa data, without the requirement of any additional information.
2.1.7.b. Systematics on the path to H0 with SN Ia H0 measurements using SNIa are affected by various sources
of systematic uncertainties.

TABLE I: A summary of the various cross-checks and systematics on the supernova component of the distance lad-
der. If two uncertainties are given, the first one is the statistical uncertainty and the second one is the systematic
uncertainty.

Reference Notes on the specific systematics check implemented Result


( km s−1 Mpc−1 )
Optical SNIa:
[221] Uses spectral feature twinning process to improve standardization; checks dust 73.01 ± 0.92
modeling, intrinsic scatter modeling;
[222] Checks different models of peculiar velocities / bulk flows σH0 < 0.2
[223] Checks SNIa Calibration by allowing individual SN survey offsets σH0 < 0.2
[224] Checks impact of mass step, global vs. local correlations σH0 < 0.15
[225] Checks light-curve fitting method; also does NIR fits 73 ± 2
[194] Uses 4-rung distance ladder, checks SNIa host demographic systematic 74.6 ± 0.9 ± 2.7
[226] Uses ZTF data alone, check on SNIa calibration 76.94 ± 6.4
NIR SNIa:
[227] Uses literature NIR SN (restframe J) and peak fitting 72.8 ± 2.8
[36] Uses literature NIR SN (restframe J and H band) 72.3 ± 1.4 ± 1.4
[228] Uses RAISIN\+literature NIR SN (restframe Y band) and SNooPy fitting, check 75.9 ± 2.2
on dust
[229] Uses literature NIR SN and BayesN fitting 74.82 ± 0.97 ± 0.84

Ref. [33] gives a comprehensive overview of many of these systematics and how they may affect measurements of
H0 . In Fig. 18, adapted from Ref. [33], we show the impact on H0 of applying 1σ shift of each systematic. We group
the systematics into various categories: Redshifts, SN Physics, Selection, Calibration, and MW Dust. Before delving
into each category of systematics in detail, it’s crucial to first understand which systematics H0 is most sensitive to,
and conversely, which have minimal impact due to being mitigated. The distance ladder is structured so that the same
probe is used in two of its three rungs (e.g., Cepheids in the first and second rungs, and SNIa in the second and third
rungs). If a systematic error disproportionately affects SNIa in the second and third rungs, or introduces significant
39

differences in the SNIa populations between these rungs, it will have a substantial impact on H0 . Conversely, if a
systematic introduces a consistent offset that affects all SNIa uniformly, the impact of that systematic will be negligible
due to the cancellation effect inherent in this formalism. In other words, consistent offsets will cancel out, mitigating
the effect of the systematic error.
Calibration: Since many of the same surveys measure SN in both the second and third rungs of the distance
ladder, the effects of calibration systematics on H0 is mitigated by the effect discussed above. As illustrated in
Ref. [223] in the context of the latest SH0ES analysis, per-survey “gray” calibration offsets affect the uncertainty by
0.2 km s−1 Mpc−1 at most. However, when different surveys are used for the second and third rungs, as shown in
Ref. [15, 225], this cancellation does not occur and effects of calibration can be larger.
Redshifts: H0 measurements are more sensitive to systematics related to redshift since redshift information is used
only in the third rung of the distance ladder, with no systematic mitigation. Ref. [222] explore various bulk flow
models in the nearby Universe, finding that changes in H0 could be up to 0.2 km s−1 Mpc−1 . They also note that
including or excluding peculiar velocity corrections could lead to ∆H0 ∼ 0.5. Additionally, Refs. [230] and [231] find
redshift measurement biases causing uncertainties around 0.1 km s−1 Mpc−1 .
SN Physics and Selection: SN intrinsic astrophysics and the role of SN dust remain one of the most poorly
understood aspects of SNIa cosmology and can affect H0 measurements [221, 232]. Most analyses have found that
these effects have a small impact on H0 because the differences between SN sub-populations selected in the second
and third rung are not expected to be significantly different. For example, Ref. [233] showed evidence for a correlation
between standardized brightness and the age of the host galaxy (quantified estimating the specific star-formation at
the SN location). In earlier measurements like Ref. [234], the third rung of SN had no galaxy-based selection applied.
Only the second rung had this selection, tied to Cepheid discovery, which favored star-forming host galaxies. This
would potentially lead to a bias in the recovery of H0 . The size of the bias would depend on the relative differential
fraction of host-galaxy demographics between the second and third rung multiplied by the size of the correlation.
Subsequent analyses [224] showed that this effect would likely be insufficient to explain the Hubble tension. Still, in
the most recent SH0ES analysis [34], the selection of SN in the third rung of the distance ladder was done to be as
similar as possible as the second rung. Only SN found in star-forming galaxies are selected, which thereby removed
the sensitivity to this systematic. Yet, the impact from this change was less than the statistical uncertainty from the
supernova component of the distance ladder. Similarly, significant differences in dust extinction and/or color-related
effects between SN in the second and third rung could potentially bias H0 measurements [235]. SN dust extinction
and color-dependent corrections are encapsulated in the nuisance parameter β (see Eq. (2.1)). As a cross-check, the β
parameter was fitted separately in SN used in the second and third rung, and it was found to be consistent [33]. This
test, together with various NIR SN H0 measurements, see Table I [36, 227–229], suggests that dust or color-dependent
effects are not expected to significantly bias H0 , or to be a significantly underestimated systematic in current H0
measurements.
Other systematics: Additional systematic tests, such as changing the light-curve fitter or adding spectroscopic
information, have shown consistent results, within ≈ 0.3 km s−1 Mpc−1 , e.g., see Ref. [221, 225]. Even when isolating
data to a single survey, as in Ref. [229], the results remain consistent, though with larger uncertainties due to the
smaller sample size.
2.1.7.c. Variants on the path to H0 with Supernovae Ia Some of the main cross-checks on the SNIa used for these
analyses is varying the wavelength regime in which light curves are measured (i.e., optical to NIR) or the dataset used
(i.e., the survey used to measure the light curves). As H0 constraints are limited by the number of SN found within
40 Mpc, there is considerable overlap in the data between these various studies. The most popular path to check and
improve the distance ladder with SNIa is by measuring NIR light curves. We list these papers in Table I. Overall, even
though the rest-frame band in which light curves are measured varies between these analyses, and the fitting method
varies between these methods, there is generally very good agreement in the recovered values of H0 . One challenge
multiple of these studies have found (e.g., [227, 228]) is larger calibration offsets between samples than those found
for optical studies. A benefit of this type of study is the possibility of improved precision of distance measurements
from NIR data, but the quality of older light curves has not typically been good enough to evaluate this possibility.
An additional path is creating a “4-rung distance ladder”, as done in Ref. [194]. SNIa used in the SH0ES distance
ladder are those found in late-type galaxies. To avoid this specific sub-sample, one can add another rung in the
distance ladder from SBF between TRGB/Cepheids and SN. The analysis of Ref. [194] improves on that of Ref. [193]
because the latter follows a similar method, but uses an inhomogeneous set of SBF measurements. The inhomogeneous
set significantly increases the scatter of the tie between SBF and SN measurements and appears to bias H0 to lower
values. Ref. [194] find a value of H0 = 74.6 ± 2.8 km s−1 Mpc−1 , in good agreement with the SH0ES value.
2.1.7.d. Inverse distance ladder to H0 with supernova The same set of SNIa used in the SH0ES distance ladder
can also serve as uncalibrated relative distance indicators to constrain H0 when combined with other probes like
BAO. BAO constrain the expansion history H(z) and extrapolate H0 . But this approach assumes the sound horizon
40

from CMB constraints and is model-dependent, relying on ΛCDM to infer H(z = 0) from BAO data at z ∼ 0.5.
However, SNIa can address the latter issue. Instead of calibrating SNIa to the distance ladder, they can be calibrated
to the BAO distance scale at typical BAO redshifts (z ∼ 0.5), allowing SNIa to constrain the expansion history at
later times (z < 0.1) without assuming ΛCDM. Studies such as Refs. [236–238] using this technique found H0 =
68.57 ± 0.9 km s−1 Mpc−1 , consistent with CMB under ΛCDM. Since BAOs obtain the physical scale from the sound
speed in the early Universe, these low H0 values have prompted discussions about the impact of the sound horizon
value.
2.1.7.e. Improving measurements of H0 in the future Improving constraints on H0 using SNIa and the distance
ladder is challenging. The limiting factor in the past decade has been the fact that HST can discover Cepheids/TRGB
within a radius of 40 Mpc, and there is one to three SNIa per year exploding within this volume. However, new
telescopes like the JWST (or future instruments like the Nancy Grace Roman Space Telescope and the ELT) are
already showing impressive improvements (both in quality and depth) in Cepheids, TRGB and JAGB measurements9
as shown by Refs. [26, 174], with exciting hints of shifts to lower H0 values presented by Ref. [26], even though the
statistics is still too low to draw firm conclusions.
The other path towards improving the constraint from SNIa is to improve the precision of the measurements. This
is the path followed by papers like Refs. [221, 239] which tried using spectral features to improve the standardization,
or the large number of papers that measure SNIa in the NIR as better standard candles. The main challenge is that
these new standardisation approaches depend on the amount of data available for its application in past literature
measurements. New types of measurements can be made for nearby SN in the future, but we can not re-measure past
SNIa.

2.1.8. J-regions of the asymptotic giant branch methods


Coordinator: Siyang Li
Contributors: Adam Riess, Bartek Zgirski, Caroline Huang, Dan Scolnic, Gagandeep S. Anand, Greg Sloan, Louise
Breuval, Richard I. Anderson, and Stefano Casertano

The JAGB refers to a group of stars in a NIR color magnitude diagram (see Fig. 19) that contains thermally pulsating
carbon-rich AGB stars with photospheric C/O greater than 1. The potential for using carbon stars (CS) as distance
indicators was first realized in the 1980s [240–244], and the method was later revived in the 2000s [129, 177, 245–248].
These pioneering studies, among others, have proposed that the mean, median, mode, or model fit of a near-infrared
J-region LF can be used as a standard candle to measure extragalactic distances. The basis for doing so originates
from the expectation that only oxygen-rich AGB stars that have masses falling within a relatively narrow range can
evolve into carbon-rich AGB stars, constrained by hot bottom burning and the efficiency of the 3rd dredge-up events.
2.1.8.a. Application The JAGB method relies on a measurement of the CS infrared magnitude distribution in
a color range between weakly pulsating CS that produce little dust and strongly pulsating CS (i.e., Mira variables)
that produce significant amounts of amorphous carbon dust. The JAGB is typically measured in the “outer disk” of
a galaxy where the stellar population is young enough to contain a substantial amount of carbon AGB stars and far
enough from the center to minimize the effects of crowding and internal reddening [249]. The region is also typically
chosen to be close enough to avoid sparser fields which increases the statistical uncertainty. In addition, the red
background of galaxies can be similar to that of CS and contaminate the CS sample [248]. Consistent selection of
fields is also important in maintaining consistent stellar populations.
Several methods have been proposed to measure the JAGB reference magnitude in the outer disk; these are typically
designed to measure some variation of the peak of the J-region LF. For instance, Refs. [129, 247] use the mean and
median, Ref. [177] smooth a binned J-region LF and take the mode, while Refs. [246, 251] introduce maximum
likelihood and binned versions of medians and calibrate their JAGB measurements using the skew parameter in a
Lorentzian fit. Ref. [248] fit a Gaussian+Quadratic model to the LF. There is an important need for standardization
of the technique to ensure accuracy.
2.1.8.b. Prospects The JAGB method has the potential to independently support or refute local measurements
of the Hubble constant from the second rung of the distance ladder. Recent studies have found JAGB-based Hubble
constants of 74.7 ± 2.1 (stat) ± 2.3 (sys) km s−1 Mpc−1 , with a full range spanning 71 to 78 km s−1 Mpc−1 depending
on the measurement method used [18], and 67.96 ± 1.85 (stat) ± 1.90 (sys) km s−1 Mpc−1 [174, 252], noting that the
lower value here originates from galaxy subsample selection, see Ref. [26]. However, it is important to be aware that

9 Even a 25% increase in the Cepheid distance would allow a doubling in the number of usable SNIa in the second rung of the distance
ladder (the volume of discovered SN will increase with distance cubed).
41

this method is still much less mature than other standard candles, such as Cepheids. For this method to be robust,
several aspects will need to be better understood, such as the empirical effects of metallicity, molecular atmospheric
diversity effects on photometry, and the shift of CS from weak to strong pulsations in the context of population
diversity.
In addition, past literature has found evidence of non-uniform
asymmetry in the J-region LF [79, 246, 251, 253]. This can produce
methodological variations (i.e., the mean will be different from the
median, mode, and model fit). Non-uniformity of the asymmetry
can also result in a mismatch in the degree of methodological vari-
ations across the distance ladder, thus increasing uncertainties of
the method. It will be important to standardize this effect.
The JAGB method introduces observational flexibility for mea-
suring H0 . JAGB only requires a single epoch while Cepheids
need multiple epochs to measure their periods and mean magni-
tudes. CS are ubiquitous in most galaxies, unlike Cepheids which
are generally observed in face-on spiral galaxies. In the near in- FIG. 19: J vs. J - K color-magnitude diagram
frared, JAGB stars are as bright as long-period Cepheids, making of stars in the LMC using the sample from
them competitive for reaching large distances. On the other hand, Ref. [250].
the metallicity dependence of Cepheids is very well calibrated [58], which is not yet the case for CS. The JAGB feature
is also brighter than TRGB. However, JAGB stars are an intermediate-age (300 Myr - 1 Gyr [129]) population, and
are therefore not as ubiquitous as TRGB stars, which populate essentially every galaxy.
The JAGB is primarily measured in the NIR, typically in the J- and H-bands (or space-based equivalents: HST
F110W [249], JWST, F115W, [17] & F150W [18]). It is important to understand whether either of these bands
necessarily offers more consistency in color dependence over the other and whether either is sloped as a function of
color. Multiple studies have found evidence of asymmetric LF and non-flat color dependence in both the J- and H-
bands [79, 246, 251, 253].
Increasing the SNIa sample size decreases fluctuations in H0 arising from cosmic variance and facilitates closer
reversion to the mean (see Ref. [26]). Recently, Ref. [254] augmented the JAGB sample used to measure H0 by
combining all available JAGB distances to SNIa host galaxies in the literature, as well as measuring new distances to
galaxies from JWST Cycles 1 & 2 for a total of 15 galaxies hosting 18 SNIa. As in Ref. [18], they find methodical
variations consequent of non-uniform asymmetry in the J-region LF; taking the middle measurement variant (as
described in Ref. [18]) yields H0 = 73.3 ± 1.4(stat) ± 2.0(sys) km s−1 Mpc−1 .
A powerful way to characterize, and potentially better standardize, variations in the JAGB method is to conduct
a field-to-field comparison of the JAGB. This approach has been implemented for the TRGB by the Comparative
Analysis for TRGBs (CATs) team [122, 125, 255]. They compared multiple fields in a given galaxy and developed a
standardization procedure via a contrast ratio. Future missions, such as Roman, and more observations with JWST
can obtain a larger field coverage that can be used to measure the JAGB and make a similar analysis possible.

2.1.9. The Hubble constant from Type II supernovae


Coordinator: Lluís Galbany
Contributors: Anil Kumar Yadav, Anto Idicherian Lonappan, David Benisty, Géza Csörnyei, Ismailov Nariman
Zeynalabdi, and Vladas Vansevičius

With the currently ongoing tension between local and distant measurements of the Hubble constant, it is crucial
to test and employ independent methods that adhere to a separate set of systematic uncertainties. SNII offer such
independent routes with the sufficient accuracy. SNII are the explosions of red supergiant stars, with multiple direct
progenitor detections to date (e.g., see Refs. [256, 257]). Given the well constrained progenitor type along with their
relatively simple composition (with the red supergiant retaining its hydrogen envelope and being made up mostly of H
and He before the explosion) and the recent advancements in the understanding of the explosion mechanism, multiple
theoretically well-founded distance estimation techniques exist for SNII. The spectra of SNII are characterized by the
presence of broad P-Cygni profiles, which allow constraining the photospheric properties. To date, three techniques
have been used to measure the Hubble constant: the expanding photosphere method (EPM, [258]), the spectral
modelling based techniques (either the spectral expanding atmosphere method SEAM [259] or the tailored EPM
[260]) and the standardizeable candle method (SCM, [261]). Of the three, SCM is an empirical technique, which
employs the relation present between the luminosity and the expansion velocity of the supernova. On the other
hand, EPM and SEAM both aim to constrain the physical parameters and the luminosity based on the spectra
directly, providing a single step measurement tool. Both Refs. [262] and [23] demonstrated that SCM and EPM
42

distance measurements up to redshifts of z ≈ 0.34 are feasible, providing important alternative paths for cosmology,
as depicted in Fig. 20.
2.1.9.a. Methods and approaches to measure H0 using SNe II The use of EPM for SNII, which essentially constits
of estimating the size of the photosphere, then comparing it to the observed flux by assuming a blackbody spectrum,
was first demonstrated by Ref. [258]. This initial exploration yielded H0 = 65 ± 15 km s−1 Mpc−1 . However, it was
highlighted later by Ref. [263], SNII radiate with a much smaller surface flux than a blackbody of the same colour
temperature (the photosphere flux appears diluted), owing to the scattering dominated atmosphere of SNII and the
fact that the thermalization layer from which the blackbody radiation is generated is deeper than the photosphere. To
take this into account, the EPM was refined incorporating a multiplicative ξ dilution factor, which relates the position
of the thermalization layer to the photosphere, which were shown to depend largely only on the color temperature
(see Refs. [264–266]). Systematic discrepancies between the different sets of dilution factors can explain differences
on the scale of 20% in the inferred distance [267, 268]. The EPM is to date one of the most frequently used distance
−1
estimation technique for SNII, with the latest H0 estimate being 72.9+5.7 −4.3 km s Mpc−1 based on 12 SNII following
Ref. [269], using the dilution factors from Ref. [265]. To rule out the possible systematic effects introduced by dilution
factors, one has to carry out radiative transfer based spectral modeling of observations, which yield self-consistent
results for the physical parameters.
This is achieved in both the customized EPM and in the SEAM, which incorporate radiative transfer modeling of
spectra into the distance estimation process. In both techniques, the modeling of the complete spectral time series is
carried out, which yields the relevant physical parameters at each epoch in a self-consistent way, avoiding the use of
dilution factors. However, estimating a single radiative transfer model takes hours to days, hence estimating a distance
this way is a time-consuming process. Due to this, until recent years, this method was only used for a handful of SN
[259, 260, 270]. Recent advances in ML allowed for the faster estimation of radiative transfer models, hence significantly
speeding up this modeling process, as described in Ref. [271]. This method was shown to provide internally consistent
−1
results [24] at a competitive precision [146] and yielded a Hubble constant of H0 = 74.9+1.9 −1.9 km s Mpc−1 based on
the analysis of literature SN [272]. However, this method is currently largely limited to local redshifts: to date only
a few SNII were observed far enough out into the Hubble flow in sufficient detail for the application of either tailored
EPM or SEAM.
To date, the most widely used method for estimating H0 using SNII is the SCM [23, 261, 273]. The method is based
on an empirically found correlation between the luminosity and the photospheric expansion velocity of the objects
[274, 275], and in terms of philosophy is similar to the method employed for SNIa (e.g., see Ref. [34]). In contrast to
the Ia’s, however, to standardize SNII, spectroscopic information is required in the form of line velocity measurements.
Over the years, the standardization has been further extended with colour term accounting for the varying level of
extinction present for different SN [276]. Recently, Ref. [23] applied this method with nine calibrator SN to estimate
−1
the Hubble constant by using SNII as the final rung of the distance ladder, obtaining H0 = 75.4+3.8 −3.7 km s Mpc−1 .
The method is still under further development, with the aim of increasing the currently available set of calibrators
with high precision cases, and also with attempts to investigate the need for additional standardization terms based
on spectroscopic information.
These results show that SNII indeed provides important and independent distance measures that can serve as the
tools for a sanity check for multiple rungs of the distance ladder. With the steady increase in the number of objects
with detailed observations, the coming years will see multiple SNII-based Hubble constant estimates with competitive
precision, which will allow for an alternative look at the Hubble tension.

2.1.10. HII galaxy distance indicators


Coordinator: Ricardo Chávez
Contributors: Ana Luisa González Morán, David Fernández-Arenas, David Valls-Gabaud, Elena Terlevich, Fabio
Bresolin, Iryna Vavilova, Ismailov Nariman Zeynalabdi, Manolis Plionis, Roberto Terlevich, Rodrigo Sandoval-Orozco,
Spyros Basilakos, and Vladas Vansevičius

HII galaxies (HIIGs) are characterized by intense, compact episodes of star formation predominantly occurring
within dwarf irregular galaxies, significantly enhancing their luminosity. These galaxies are spectroscopically identified
based on the prominent equivalent width of their Balmer emission lines, specifically EW (Hβ) > 50 Å, which indicates
their young age (less than 5 Myr). Similarly, Giant Extragalactic HII Regions (GEHRs) exhibit vigorous star formation
but are typically found in the outer disks of late-type galaxies. The rest-frame optical spectra of both HIIGs and
GEHRs display pronounced emission lines, indicative of gas ionization by massive Young Stellar Clusters (YSC) or
Super Star Clusters (SSC), leading to similar spectral features [279–283].
Numerous studies have confirmed a consistent correlation, found by Ref. [281], between the luminosity of Balmer
lines, such as L(Hβ), and the ionized gas velocity dispersion, σ, measured through these emission lines, in both HIIGs
43

Calibrated SNe II in the Hubble flow


42

[mag] w. H0 = 75.4 km / s / Mpc


40
42 EPM distances for cosmology
38 EPM - B14
40 EPM - E96
36 EPM - G18
EPM - J09
38 EPM - D24
34
34 Cepheid + TRGB calibration of SNe II tailored EPM - V24

[mag]
1 tailored EPM - C23
36
33
0

EPM
SN II: m M [mag]

32 34

SN II:
1
31 34 36 38 40 42
SN II: m M [mag] 32
30
29 30
28
1 28
28 30 32 34 36 38 40 42
[mag] w. H0 = 75.4 km / s / Mpc
0
1
28 30 32 34
Cepheid / TRGB m M [mag]

FIG. 20: SNII for Hubble parameter estimation. Left: The distance ladder formalism applied on SNII, as presented
by Ref. [23]. The bottom panels with the blue dots depict the intermediate rung of the distance ladder, where the
SNII brightnesses are calibrated through Cepheids and TRGB, along with the residuals after the standardization.
The top panels with the red dots show the final rung of the distance ladder, where this standardization is applied
on SNII in the Hubble flow which allow for a Hubble constant estimation, while the residuals are shown on the
panel below. Right: SNII Hubble diagram using EPM distances, as updated from Ref. [277]. EPM distances do
not require external calibration, hence they provide a single step measurement of H0 . The empty circles denote the
EPM distances that employ dilution factors, while the filled squares mark the tailored EPM estimates that make
use of radiative transfer modelling. The references for the distance estimates are as follows: B14 – [278], E96 –
[264], G18 – [262], J09 – [267], D24 – [269], V24 – [272] and C23 – [24].

and GEHRs. This relationship, known as the L − σ relation [283–285], is recognized as a potent cosmological distance
indicator [286–289], where GEHRs and nearby HIIGs are used as the “anchor” sample because their distances can be
independently estimated from Cepheid variables or TRGB measurements. As a result, the L − σ relation provides
an exceptional method for utilizing this distance metric to investigate the Hubble flow over an extensive range of
redshifts (z).
The characteristic strong emission lines within the rest-frame optical spectra of GEHRs and HIIGs render them
effective tools for exploring nascent star formation at high redshifts. Using instruments such as NIRSpec [290] on
board the JWST [291], it is possible to study these regions up to z ∼ 6.5 through the Hα emission line, or even up
to z ∼ 9 using the Hβ and [OIII]λλ4959, 5007 Å emission lines. This capability enables the observation of luminous
HIIGs that date back to the Epoch of Reionization (EoR) as shown in Fig. 21.
Using a dataset of 231 GEHRs and HIIGs, some observed with NIRSpec on the JWST up to z ∼ 7.5, a study by
Ref. [289] employs the MultiNest Bayesian inference algorithm [292–294] to derive constraints for various cosmological
models. They used uniform, non-informative priors across all parameters for unbiased estimations. The derived
constraints are detailed in Table II, showing marginalized best-fit values and 1σ uncertainties for each parameter,
with some parameters held constant during the analysis. The study explores a generalized parameter space θ =
{α, β, h, Ωm , w0 , wa }, where θn = {α, β} represents nuisance parameters of the L − σ relation for GEHRs and HIIGs
with α as the intercept and β as the slope. For the flat ΛCDM model, θc = {h, Ωm , −1, 0} sets h as the reduced Hubble
constant and Ωm as the total matter density, fixing the first two DE equation of state (DE Equation of State (EoS))
parameters at w0 = −1 and wa = 0, corresponding to a cosmological constant (Λ). Adjusting w0 allows for models with
evolving DE EoS, akin to quintessence [295, 296], while including wa aligns with the Chevallier-Polarski-Linder (CPL)
model [297–299].
The constraints derived on cosmological parameters, specifically {h, Ωm , w0 } = {0.731±0.039, 0.302+0.12 +0.52
−0.069 , −1.01−0.29 }
(stat) from GHIIR and HIIG data, align closely with recent Pantheon+ results from 1550 SNIa, which produces
{h, Ωm , w0 } = {0.735 ± 0.011, 0.334 ± 0.018, −0.90 ± 0.14} [33]. This agreement emphasizes the validity and impor-
44

TABLE II: Marginalized best-fit parameter values and associated 1σ uncertainties for the HIIGs and anchor sam-
ples. Values enclosed in parentheses indicate parameters that were held constant during the analysis.

Data Set α β h Ωm w0 wa N
HIIG — (5.022 ± 0.058) — 0.282+0.037
−0.045 (-1.0) (0.0) 195
HIIG — (5.022 ± 0.058) — 0.278+0.092
−0.051 −1.21+0.45
−0.40 (0.0) 195
HIIG (33.268 ± 0.083) (5.022 ± 0.058) 0.715 ± 0.018 0.267+0.038
−0.048 (-1.0) (0.0) 195
HIIG (33.268 ± 0.083) (5.022 ± 0.058) 0.718 ± 0.020 0.278+0.091 +0.46
−0.050 −1.22−0.40 (0.0) 195
Anchor+HIIG 33.28 ± 0.11 4.997 ± 0.089 0.730 ± 0.038 (0.3) (-1.0) (0.0) 231
Anchor+HIIG 33.28 ± 0.14 5.00 ± 0.11 0.730 ± 0.040 0.335+0.044
−0.055 (-1.0) (0.0) 231
Anchor+HIIG 33.29 ± 0.14 4.99 ± 0.11 0.731 ± 0.039 0.302+0.12 +0.52
−0.069 −1.01−0.29 (0.0) 231
Anchor+HIIG 33.29 ± 0.14 4.99 ± 0.11 0.730 ± 0.039 0.321+0.10 +0.55 +0.65
−0.063 −0.91−0.33 −0.71−1.2 231

Age of the Universe (Gyr)


10 6 4 2 1.5 1 0.8 0.66
55 wCDM (h, m, w0 = 0.731, 0.302, 1.01)
CDM (h, m = 0.731, 1.0)
50
45
40 40 0.4

35 0.3
35
pdf

30 0.2
30
25 0.1
25 20 0.0
0.000 0.025 0.050 0.075 0.100 0.125 0.150 2.5 0.0 2.5
20 z /
0 1 2 3 4 5 6 7 8
z
FIG. 21: Hubble diagram for GEHRs and HIIGs, here z is the redshift and µ is the distance modulus. In cyan we
present the “anchor” sample of 36 GEHRs which have been analyzed in Ref. [300], in black we present the full sam-
ple of 181 HIIGs which have been analyzed in Ref. [289], while in blue we present the 9 new HIIGs from Ref. [301]
and in green 5 new HIIGs studied with JWST by Ref. [302]. The black line is the cosmological model that best fits
the data with the red shaded area representing the 1σ uncertainties to the model, while the grey dashed line is a
flat cosmological model without DE. The inset at the left shows a close-up of the Hubble diagram for z ≤ 0.15. The
inset at the right presents the pulls Probability Density function (PDF) of the entire sample of GEHRs and HIIGs
and the red line shows the best Gaussian fit to the PDF. Adapted from Ref. [289].

tance of HIIGs as distance indicators within the context of current cosmological studies.
The enduring validity of the L − σ relation at high redshifts (z > 3), reaching into the EoR, suggests remarkable
uniformity in HIIGs properties over vast cosmic timescales. This not only confirms the reliability of the L − σ relation
as a cosmological tool but also illuminates the fundamental processes underlying the formation and evolution of early
Universe galaxies.
In our efforts to improve the accuracy of cosmological parameters using HIIGs, the addition of data from JWST,
45

extending to z ∼ 9 and beyond, is proving crucial. The JWST’s exceptional sensitivity and resolution enable detailed
observations of HIIGs at these higher redshifts, offering a unique view of the early Universe. This data range is
essential for analyzing the dynamics of the Universe’s expansion across various cosmological scenarios.

2.1.11. The baryonic Tully-Fisher relation approach


Coordinator: Khaled Said
Contributors: Benoit Famaey, Cláudio Gomes, Duško Borka, Esha Bhatia, Jenny G. Sorce, Maurice H.P.M. van
Putten, Milan Milošević, Paolo Salucci, Paula Boubel, Predrag Jovanović, and Vesna Borka Jovanović

The BTFR ([303–312]) extends the classic Tully-Fisher (TF; [25, 313–335]) relation by incorporating both the stellar
and gas masses of galaxies, correlating a galaxy’s baryonic mass (Mb ) with its rotational velocity (Vrot ). This relation
provides critical insights into galaxy formation and dynamics and is crucial in testing and constraining models of
galaxy evolution within both the ΛCDM framework and alternative theories [336–349].
The BTFR is defined as: Mb ∝ Vrot x
, where Mb is the sum of stellar and gas masses, Mb = M∗ + Mgas and x is
typically around 4, suggesting a deep connection between the visible matter and the dynamics of galaxies [303, 350–
352].
2.1.11.a. Historical context and development The original TF relation, proposed by Ref. [313], established a
link between the intrinsic brightness of spiral galaxies and their maximum rotational velocities. This relation has
been pivotal in determining distances to galaxies and measuring the Hubble constant [353]. For example, Ref. [313]
applied their TF relation to derive distances to the Virgo and Ursa Major clusters, estimating a Hubble constant
of approximately H0 ≈ 84 km s−1 Mpc−1 for Virgo and H0 ≈ 75 km s−1 Mpc−1 for Ursa Major. The extension to the
BTFR was motivated by the need to include galaxies with significant gaseous components, particularly dwarf galaxies
[303]. Over the years, numerous studies have validated and expanded the BTFR, demonstrating its robustness and
utility in various cosmological and astrophysical applications [312, 347, 354]. A crucial step towards the understanding
of the physics of the TF and BTFR relationships has been the finding, in disk systems, of a family of local Tully-
Fisher-like relationships [355–357], holding at various specific galactocentric radii and thus dubbed as the Radial Tully
Fisher (RTF: for a review see Ref. [358]).
2.1.11.b. Theoretical models and their implications
The standard ΛCDM model In the ΛCDM, the TF and BTFR are seen as consequences of the gravitational
dynamics dictated by DM halos and baryonic matter. The rotational velocities of galaxies are thought to be influenced
by both visible and DM, with the BTFR providing a means to probe these interactions [347]. However, as illustrated
in Fig. 22, the observed slope and intrinsic scatter of the BTFR often differ from ΛCDM predictions, which typically
forecast a lower slope and higher scatter [359]. This discrepancy poses challenges to the ΛCDM model and prompts
further investigation into galaxy dynamics and mass distribution.
Modified gravity models Some studies propose that the BTFR can be explained without DM through modified
theories like f (R) theories of gravity [360–363]. Alternative gravitational theories, such as Modified Newtonian
dynamics (MOND) and non-minimal matter-curvature coupling models, offer different perspectives on the BTFR.
MOND, proposed by [364], adjusts Newtonian dynamics at low accelerations, predicting a BTFR with a slope exactly
equal to 4, consistent with many observations [347, 354]. MOND has made several successful predictions regarding
the detailed shapes of rotation curves, galaxy dynamics, and galaxy lensing [365–367].
Another approach, non-minimal matter-curvature coupling model, generalizes the pure gravity sector by introducing
a generic function of the curvature scalar, f1 (R), and a non-minimal coupling between the matter Lagrangian and
another generic function of the curvature scalar, f2 (R) [368],
ˆ

S = d4 x −g (f1 (R) + f2 (R)L) . (2.5)

This model leads to an extra force term in the geodesics for a perfect fluid. In three dimensions and in the Newtonian
limit, assuming that the total acceleration ⃗a is collinear with the 3-force, f , and in the limit
 of very small gravitational
f2
accelerations, the Newtonian acceleration, ⃗aN , is ⃗aN ≈ aaE ⃗a, where a−1
E = (2f ) −1
1 − a2 , which is remarkably similar

to MOND’s result. Furthermore, a ≈ aE GM /r = v 2 /r, hence a Tully-Fisher relation appears with a luminosity of
the form L ∼ v 4 with v 4 = aE GM . Thus aE = 10−10 m s−2 as in MOND or aE = (8.5 ± 1.3) × 10−10 ms−2 [368].
This model has been shown to be consistent with several observations ranging from the cosmic version of the virial
theorem at Abell 586 cluster [369], to Jeans instability in Bok globules [370, 371], or to SN distance data and the
BAO data allowing for an attempt to solve the Hubble tension [372].
46

1012

1011 ΛCDM

Mb (M⊙)
1010

109

Observed:
108 s = 3.75 ± 0.11
log A = 2.18 ± 0.23
107
2 × 101 1 × 102 5 × 102
Vf (km s−1)
FIG. 22: Comparison between the BTFR obtained from error-weighted fits of galaxies with accurate distances from
Ref. [359] to the BTFR predicted by ΛCDM cosmology. The observed BTFR (blue solid line) includes a light blue
band representing the intrinsic scatter of 0.1 dex. The ΛCDM prediction is shown as a black dashed line. Both
relations are plotted using equations from Ref. [359]: Eq. (6) for the observed BTFR and Eq. (8) for the ΛCDM
prediction. [Reprinted under CC BY 4.0. from: V. Borka Jovanović, D. Borka, and P. Jovanović, Contrib. Astron.
Obs. Skalnate Pleso 55/2, 24 - 33 (2025)].

A finite sensitivity of weak gravitation in galaxy dynamics to background cosmology may be detectable below the de
Sitter acceleration scale [373] using SPARC [374] and large galaxy surveys such as MaNGA [375]. Redshift dependence
would point to a potentially unified picture, linking BTFR to JWST observations of ultra-high redshift galaxies at
cosmic dawn [376, 377]. This approach may also place novel observational constraints on cosmological parameters,
notably H0 and the deceleration parameter q0 [375].
2.1.11.c. Observations and cosmic tensions Various studies have extensively explored the BTFR using data
from large galaxy surveys. The Spitzer Photometry and Accurate Rotation Curves (SPARC), for example, includes
measurements of 175 rotationally supported galaxies in the near-IR, minimizing the effect of star-halo degeneracy
and providing precise baryonic mass and rotational velocity data [378, 379]. Such robust datasets enable accurate fits
to the BTFR and facilitate comparisons between observed galaxy properties and theoretical predictions. The BTFR
for the SPARC sample aligns with a scale of x = 4. Research by Ref. [380] using the BTFR with a sample of 95
independent galaxies also challenges the Hubble constant value H0 < 70 km s−1 Mpc−1 with 95% confidence level. A
More recent study by Ref. [25] demonstrated that using the BTFR allowed them to determine a Hubble constant of
H0 = 75.5 ± 2.5 km s−1 Mpc−1 , which again challenges the standard cosmological model that suggests lower H0 values
[192]. Additionally, studies have shown variations in BTFR parameters across different distance bins, which could
contribute to the understanding of the Hubble tension [381–386].
The RTF relationship, applied to a large sample of 843 local galaxies extending out to z = 0.03, has recently been
used as a distance indicator to determine the maximum allowed variance of the H(z)/H0 parameter [356]. They found
that the maximum allowed local ‘radial’ variation is 1%, which is not enough to resolve the H0 tension. In the near
future, with the PROBES sample [387] (3000 objects) we will be able to use the RTF to test the hypothesis of a Giant
Local Void at greater distances and investigate the anisotropy of the expansion of the Universe (see Ref. [356] for the
present situation). Finally, the RTF can be applied at z ∼ 1 − 2 to investigate a major part of the expansion history
of the Universe [388].
The original TF relation is also useful for tracing the distance-redshift relation, provided that the relation is carefully
calibrated. Unlike the BTFR, the TF relation is not universally linear, so studies typically introduce more complex
models to account for the curvature and varying intrinsic scatter [333, 389]. The Cosmicflows-4 (CF4) Tully-Fisher
catalog [25] of ∼10,000 spiral galaxies is currently the largest TF dataset, combining HI line widths with photometric
magnitudes. There have been several published H0 measurements using this dataset. For the full CF4 TF catalog,
Ref. [25] derived H0 = 75.1 ± 0.2 km s−1 Mpc−1 with an estimated systematic error of ± 3 km s−1 Mpc−1 . Using an
improved measurement methodology [389] and updated primary distance calibration [390], this was re-measured
47

as H0 = 76.3 ± 2.1 (stat) ± 1.5 (sys) km s−1 Mpc−1 , where the statistical error reflects the relatively small number of
primary calibrators in the sample (∼ 50 objects with CPLR and/or TRGB distances). These are in agreement with
other recent H0 measurements using the TF relation as the final rung of the distance ladder, which consistently return
H0 > 70 km s−1 Mpc−1 (see Ref. [353] for a review). The next generation of Tully-Fisher datasets resulting from large
surveys such as WALLABY [391], DESI [392], and FAST [393] will increase the precision of TF-derived measurements
of the Hubble constant.

2.1.12. The Hubble tension in our Backyard: DESI and the nearness of the Coma cluster
Coordinator: Yukei Murakami
Contributors: Daniel Scolnic

The recent work by DESI [394] analyzed their first samples of early-type galaxies in the DESI peculiar velocity
survey to study the fundamental plane (FP). FP is an empirical three-dimensional relation that links elliptical galaxies’
radii, surface brightness, and the velocity dispersions [395, 396]10 . This tight correlation serves as a standardization
to measure distances to the galaxies over a large redshift range once the absolute distance scale is anchored to an
independent, well-measured distance. In the DESI work [394], the Hubble Constant H0 is measured with FP using
an SBF measurement to the Coma cluster to calibrate the distance scale (i.e., distance ladder). Alternatively, recent
study [397] showed that one can calibrate the FP in the early universe using Planck+ΛCDM inference of H0 to
measure the distance to the Coma cluster (inverse distance ladder). The inverse distance ladder with Planck+ΛCDM-
-calibration of FP places Coma at ∼ 10% further than the direct measurement of distance by SBF. In addition, the
new SNIa distance, as well as other independent distance measurements to Coma collected over the past few decades,
are all consistent with the SBF, creating a tension against Planck+ΛCDM-prediction. This discrepancy provides a
unique opportunity to shed light on the implications of the Hubble Tension in our local universe.
2.1.12.a. DESI FP and inverse distance ladder DESI has provided an extensive FP measurement and a Hubble
diagram over the redshift range of 0.01 < z < 0.1 using a sample of 4191 early-type galaxies in the Hubble flow and
226 galaxies in Coma [394]. The FP relation, which correlates a galaxy’s velocity dispersion, surface brightness, and
physical size, serves as a secondary distance indicator when calibrated with independent distance anchors. When
the FP relation is calibrated using a near-infrared SBF distance of DComa = 99.1 ± 5.8 Mpc [124], DESI derives
−1 −1
a value of H0 = 76.05  s Mpc . This result relates the Hubble constant and the Coma distance as
 ± 1.3 km
H0 = (76.05 ± 1.3) × 99.1 Mpc
DComa km s−1 Mpc−1 . This relation can be used to obtain the distance to Coma predicted
by Planck+ΛCDM, using the Planck-inferred value of the Hubble constant H0 = 67.4 km s−1 Mpc−1 . The implied
distance to Coma shifts to DComa = 111.8 ± 1.8 Mpc, a distance larger than any of the historical measurements that
we discuss later.
2.1.12.b. Latest Coma distance with ATLAS SNe Ia A recent analysis [397] compiled a sample of 10 spectro-
scopically confirmed SNIa within the Coma Cluster observed by the ATLAS survey [398] and YSE [399], as well
as two additional SN previously included in the Pantheon+ dataset[32]. Applying the standardization and cali-
bration methods established in the Pantheon+ analysis, the study derived a mean standardized peak brightness of
m0B = 15.712 ± 0.041 mag for SN in Coma cluster. Using the absolute magnitude calibration of SNIa from the HST
Cepheids [34], this results in a distance modulus of µ = 34.97 ± 0.05 mag, corresponding to DComa = 98.5 ± 2.2
Mpc. This measurement is in excellent agreement with the SBF distance used to calibrate DESI, with a significantly
reduced uncertainty. The measured distance is at 4.6σ tension from the Planck-ΛCDMprediction.
2.1.12.c. New & historical measurements in tension with ΛCDM Historical measurements, as well as additional
distance indicators and calibration methods further support a distance to Coma around 100 Mpc (see Fig. 23). A
compilation of historical measurements summarized in [400], including I-band Tully-Fisher [319], K-band SBF [401],
I-band SBF [402], Dn − σ [403], FP [404] and Globular Cluster LF (GCLF) [405], produce an average of DComa =
95.1 ± 3.1 Mpc. The HST Key Project (KP) calibrated the FP relation using Cepheid variable stars, and its results,
accounted for the recent updates, measure DComa = 85 ± 8 Mpc [406]. A more recent JWST measurement of TRGB
and SBF, allows to re-calibrate the FP relation by HST-KP, which places Coma at 90 ± 9 Mpc. When combined with
the SBF distance used in the DESI work and the new SNIa distance, these independent approaches yield a consensus
distance of 98.0 ± 2.0 Mpc.
The tension between local and Planck-calibrated distances to Coma reflects a broader challenge in reconciling the
local and early-universe distance scales. If Coma were truly at DComa > 110 Mpc as ΛCDM-Planck calibration of

10 See Sec. 2.1.11 for the Tully-Fisher relation, a late-type galaxy counterpart
48

FIG. 23: Historical (1990 onward) distance modulus measurements of the Coma cluster (as reviewed in [406]). Only
distance measurements that do not depend on redshift and H0 are included. Figure taken from [397], and the right-
most point represents the SNIa distance measured in their work.

DESI FP predicts, it would be inconsistent with decades of direct distance measurements based on a wide range of
methodologies. The upcoming year 1 data release for DESI FP is expected to be significantly larger than the current
study, and the targeted JWST observation of Coma galaxies (PI Jensen, GO 5989) are underway. When combined,
these data will significantly reduce the statistical uncertainties in the FP-based H0 or distance measurements, and the
increased sky coverage in DESI data will allow FP to be calibrated with more than one galaxy clusters. Additionally,
a few SNIa are being discovered in Coma every year, and the SNIa–based distance measurements are expected to be
more robust in the next few years. Coma distance serves as an alternative or additional perspective on the issues
surrounding the cosmological distance scales, and it showcases a new type of distance ladder analysis that is expected
to develop rapidly in the coming years.

2.1.13. Cosmic chronometers


Coordinator: Michele Moresco
Contributors: Adrià Gómez-Valent, Anto Idicherian Lonappan, Arianna Favale, David Benisty, David Valls-
Gabaud, Dinko Milakovic, Elena Tomasetti, Marek Biesiada, Rodrigo Sandoval-Orozco, Ruth Lazkoz, Sanjay Mandal,
Swayamtrupta Panda, and Vavilova Iryna

CC have emerged in the last few years [182] as one of the most promising cosmological probes that can provide
direct measurements of the expansion history of the Universe without the need to rely on any assumed cosmological
model. The method [408], which is only based on the validity of the Cosmological Principle and on the assumption
that General Relativity and standard physics hold in the environment of the galactic stars, proposes to derive the
Hubble parameter H(z) directly from the differential age evolution of the Universe (dt) in a redshift interval (dz), as:

1 dz
H(z) = − . (2.6)
(1 + z) dt

To map the age evolution of the Universe, the best tracers that have been found are very massive (log10 (M/M⊙ ) > 11)
and passively evolving galaxies since they have been proven to be extremely homogeneous systems, with synchronized
formation, representing at each redshift the oldest population of galaxies [182]. Several authors have applied this
method with different approaches to derive the differential age dt [409–417]. It currently counts more than 30 H(z)
measurements in the range 0 < z < 2.1, with accuracy from ∼10-15% at the higher redshifts down to ∼5% at the
lower redshifts (see Fig. 24). See Refs. [182, 418] for extensive reviews on the topic.
Since it provides a measurement of the Hubble parameter without assuming any cosmological model, recently,
CCs have been widely employed in discussing the Hubble tension. To fully take advantage of the potential of these
data, several different cosmology-independent techniques have been explored, from GP [419–431], to non-parametric
smoothing [432, 433], Weighted Polynomial Regression method [422], Padé approximations [427, 434], and Artificial
Neural Network (AGB) architectures [431]; all these analysis provide a reconstruction of H(z) independent of any
cosmological model that can, in turn, be extrapolated to z = 0 for a determination of the Hubble constant H0 . The
49

250 0.6
current CC data f CDM (forecast)
CC forecast f CDM
0.5 o CDM
200 fwCDM
H(z) [km/s/Mpc]

150 0.4

m
100 0.3

50 0.2
0.0 0.5 1.0 1.5 2.0 50 60 70 80 90
redshift H0
FIG. 24: Left plot. H(z) measurements obtained with the CC method from the literature (white points) and from
future surveys (green points, from Refs. [182, 407]). Right plot. Cosmological constraints in the Ωm -H0 plane; the
different colors show the 68% CL results (also reported in the left panel) obtained by fitting current data with a flat
ΛCDM (in blue), open ΛCDM (in yellow), and flat wCDM (in red) cosmological model, and future CC data with a
flat ΛCDM model (in grey).

other alternative is, instead, to assume a cosmological model and derive cosmological constraints from the parametric
fit of the data [182, 435, 436]. Both these approaches, considering all the sources of statistical and systematic errors,
give comparable results, yielding a constraint on the Hubble constant of H0 = 70.7 ± 6.7 km s−1 Mpc−1 [429]. Future
measurements, however, leveraging on much larger statistics (like from the ESA mission Euclid [437] or from future
spectrographs like WST [407]) and better modeling of the data, promise to significantly improve these constraints,
improving the accuracy from ∼8% to ∼2%. An example is given in Fig. 24, where current and future CC data are
fitted with different cosmological models (flat ΛCDM, open ΛCDM, and wCDM) to provide constraints in the ΩM -H0
plane (for more details, see Refs. [182, 407]).
A difference in H0 can be reinterpreted also as a tension in the calibrators of the cosmic distance ladders, i.e., the
absolute magnitude M of standard candles such as SNIa and the standard ruler represented by the comoving sound
horizon at the baryon-drag epoch, rd . Assuming CC as reliable CCs, it is also possible to use them to measure these
distance calibrators independently from the CMB and the first rungs of the direct distance ladder. Thus, it is clear
why CC plays an important role in the discussion. So far, the uncertainties for M and rd derived by applying these
calibrations are still too large to arbitrate the tension [424, 429, 438], but also this is bound to improve in the near
future with the advent of upcoming surveys and data. Euclid [437], for instance, is expected to increase by 2 orders
of magnitude the currently available statistics [182]. CCs have been used to calibrate baryon acoustic oscillations and
SNIa [422–424, 429, 439–441] and also other standardizable objects like ultra-compact radio RGBs [420], GRBs or
RGBs [430, 438, 442, 443] which can be exploited to extend the ladder beyond the SNIa redshifts, z > 2.
Beyond the H0 tension, CCs are also the ideal data to test and compare with different cosmological models, from
the simplest to the more exotic ones. Among the litmus tests of the ΛCDM model, the Om(z) diagnostic and its
two-point version Om(z1 , z2 ) have a distinguishing feature [444] that depends only upon the expansion rate H(z).
Hence, CCs have been used [445, 446] to perform tests of scenarios assuming either a cosmological constant or an
evolving DE equation of state.

2.1.14. Strong lensing and time delay measurements


Coordinator: Tommaso Treu
Contributors: Alba Domi, Alexander Bonilla Rivera, Ali Övgün, Anowar J. Shajib, Clecio Roque De bom, David
Valls-Gabaud, Dominique Sluse, Eoin O Colgain, Frédéric Courbin, Jenny Wagner, Lindita Hamolli, Predrag Jo-
vanović, Sherry H. Suyu, Simon Birrer, Tanja Petrushevska, and Veronica Motta

Time variable sources that are multiply imaged – i.e., strongly lensed – provide an opportunity to measure absolute
distances, and therefore cosmological parameters including the Hubble constant H0 . The main advantages of this
so-called time delay cosmography (TDC) with respect to other cosmological probes described in this white paper
are: i) it is a direct measurement of distances well into the Hubble flow; ii) it is independent of all other types of
distance measurements; iii) it relies on well understood fundamental physics such as general relativity. In this section
50

we provide a brief summary of TDC, the current state of the art, and future prospects. The reader is referred to
dedicated reviews for more extensive discussion, e.g., see Ref. [447–451].
The method was originally proposed for lensed SNIa by Refsdal in Ref. [452], well before the discovery of any actual
strong gravitational lens. Since the mid-eighties it has been applied to lensed RGBs [447], and only recently – with
the discovery of multiply imaged SN [453, 454] – it has been possible to fulfill Refsdal’s dream [455, 456].
TDC measures angular diameter distances, typically between z = 0.5 − 3, where the majority of deflector galaxies
and lensed sources lie. More precisely, it measures the so-called time delay distance D∆t = (1 + zd ) DDdds Ds
, where
D indicates angular diameter distance, and the subscripts s and d stand for source and deflector, respectively. In
combination with stellar kinematics, it also measures Dd [457].
TDC is thus primarily sensitive to H0 . It is, however, also sensitive to other cosmological parameters, such as
curvature or the DE equation of state, in a complementary way to probes such as the CMB and SNIa, breaking some
of the degeneracies of those probes [458–461].
Several representative measurements using lensed RGBs prior to 2022 are summarized in Fig. 25. Already in 2020,
just 7 lenses yielded H0 at 2% precision [462, 463], under standard assumptions about the mass density profile of the
deflector. That measurement is in agreement with, and completely independent, of late Universe probes. If those
assumptions are relaxed, and a parametrization of the mass density profile that is maximally degenerate with the
Hubble constant via the mass-sheet degeneracy [464, 465] is considered, the uncertainty increases to 8% for the same
7 lenses [466]. In 2023, the first lensed supernova [455] yielded an H0 measurement with ∼7% precision. Whereas this
precision is very encouraging for a single system, especially given the complexity of modeling a cluster of galaxies, see
also Ref. [456], it is not sufficient to distinguish between the early and late-time measurements and resolve the “Hubble
tension”. As more lensed SN are being discovered and analyzed, e.g., see Ref. [467–474], the precision is expected to
improve rapidly, especially once the Roman and Rubin telescopes come on line [475].
In order to increase the precision on H0 to the desired level of 1% under these weaker assumptions, three comple-
mentary approaches can be pursued [476]. The first is to increase the sample size. The Euclid, Roman, and Rubin
Telescopes are expected to discover thousands of strong lenses, including lensed SN [477–480]. There is little doubt
that large numbers of lenses will be discovered. The challenge will be to transform them into cosmological probes
by obtaining the necessary additional data and computing the necessary cosmography grade models. Additional
data include high-resolution images, time delays, deflector and source redshift, deflector kinematics, and supernova
classification if relevant. Some of these data will be provided by the surveys themselves, e.g., high-resolution images
by Roman and Euclid [475, 481] and some time delays by Rubin [482], but others will require concerted follow-up
effort. A coordinated follow-up for measuring precise time delays is critical. RGB time delays require high precision
lightcurves minimally affected by microlensing [483, 484], while SN require dedicated multi-color or spectroscopic
follow-up [485, 486]. The effort will be substantial, and it will require, e.g., repurposing a 4m class telescope for
time delays or building a dedicated satellite. The advent of ELTs will simplify matters, especially for high-resolution
imaging and kinematics.
Modeling hundreds of lenses with precision sufficient for cosmography hundreds of lenses would be prohibitive at
the moment, both in terms of human power and computing power. Fortunately, efforts are underway to develop
automated pipelines [487–490], in some cases with the help of ML algorithms [491–493], to speed things up. The
first results are promising, in the sense that models from independent automated pipelines are found to agree in their
predicted time delays within the uncertainties, when the data contain sufficient information models that agree to
within the uncertainties in predicted time delay can be found [494].
The second approach consists of increasing the precision per lens. In order to break the mass sheet degeneracy, non-
lensing information is needed, for example stellar kinematics of the deflector [461, 495–497], the absolute luminosity
of a lensed supernova Ia [498], or WL [499]. If the quality of the kinematics is sufficient, in principle, each lens can
deliver distances to 2-4% precision, limited by systematic errors in stellar velocity dispersion measurements [500].
The third approach consists of utilizing the much more abundant non-time delay lenses (a.k.a. “external samples”)
to learn about the internal structure of time delay lenses. This method has been shown to increase the precision of
TDC significantly [466]. The key to using external samples without introducing biases is to understand the selection
function of both the time delay and non-time delay lenses sufficiently well. At the moment, this is not a limiting
factor [501] - statistical errors are larger - but the selection function will have to be properly modeled and understood
in order to achieve 1% precision and accuracy on H0 , sufficient to solve the Hubble tension.

2.1.15. Extended QSO cosmologies


Coordinator: Maria Giovanna Dainotti
Contributors: Aleksander Łukasz Lenart, Catarina Marques, Celia Escamilla-Rivera, Giada Bargiacchi, Giovanni
Montani, Rodrigo Sandoval-Orozco, and Swayamtrupta Panda
51

FIG. 25: The measurements are categorized according to i) assumptions on the mass distribution of the main de-
flector, “assertive” and “conservative” for simply parametrized models or “free form” for pixellated models; ii) to
the amount of information used per lens, “low info” utilizes RGB positions and time delays, "high info" adds the
extended surface brightness distribution of the multiple images of the RGB host galaxy (or “Einstein Ring”), stel-
lar kinematics of the main deflector, and number counts or WL to estimate the line-of-sight convergence. For each
measurement, the reference and the number of time delay lenses are given. The measurements shown in red were
performed with blinding to prevent experimenter bias. The figure is taken from Ref. [449], see therein for details.

RGB are one of the most luminous non-transient energy sources, whose use at cosmological scales allows us to
study the Universe at higher redshifts, e.g., up to z ∼ 7 [502, 503]. At this redshift range, it is possible to reveal
some interesting aspects and deviations from different cosmological models that can be indistinguishable at low z.
In this path, some examples of the RGB cosmological use are: (i) the reverberation mapping technique [504, 505],
and (ii) the relation between variability in X-ray amplitude and black hole mass scatter [506]. However, there is a
lack of a clear definition of RGBs to order the diversity of AGN objects [507], although recent studies have begun
to streamline the various RGB categories [508, 509] with particular emphasis on sources accreting at or above the
Eddington limit [510]. There is a need to homogenize the modeling of broadband RGB spectra and examine and assess
the evolution of RGB properties (Black Hole mass, Eddington ratio, viewing angle to the source) over a broad range
of redshifts [511, 512]. Under such techniques, several RGB catalogs have been used to find cosmological constraints
under certain theoretical assumptions, e.g., consistencies or inconsistencies between time delays from reverberation
mapping of prominent emission lines extending beyond the cosmic noon [513] and impact of the inherent heterogeneity
in the data set on cosmological parameter constraints [514], constraints using GRBs with RGB cosmology analysis
for future landscapes [515].
Furthermore, studies including RGB catalogs combined with GRB observations achieve small uncertainties on
cosmological parameters [516, 517]. Moreover, combining the Risaliti-Lusso relation for RGB with SNIa can allow
us to obtain corrections on the redshift evolution as a function of cosmology [518]. Recent studies have been going
through the Risaliti-Lusso relation, which has been validated in Ref. [512], where Ωm,0 value seems larger when only
it is considered RGB baselines [518–522], and some systematics in the RGBs measurements should be expected in
52

0.25 2.0 0.4


norm logistic t
0.20 1.5 norm norm
0.3
0.15
N 1.0 0.2

N
0.10
0.05 0.5 0.1
0.00 0.0 0.0
4 2 0 2 4 1.0 0.5 0.0 0.5 1.0 6 4 2 0 2 4
(GRBs) log10LX (QSOs) (Pantheon + )
FIG. 26: Normalized residuals ∆µ histogram for GRBs (left panel), RGBs (middle panel), and SNIa from Pantheon
+ samples (right panel). The red curve is the best-fit distribution, while the orange one is the Gaussian distribu-
tion. In the case of GRBs, the two coincide since the Gaussian is the best-fit distribution.

future new cosmological probes [523]. Despite the systematics, this relation is a promising cosmological probe since
its observed dispersion on average is ∼0.20-0.25 dex, and becomes significantly smaller when only the sources with
the highest quality X-ray observations are considered [524], suggesting that intrinsically the X-ray to UV relation has
a scatter lower than 0.1 dex, pointing to a tight, universal physical process regulating the energy flow between the
accretion disk and the X-ray emitting region in RGBs. In this scheme, several studies related to extended theories of
gravity have been performed and cosmologically constrained using RGB baselines, e.g., f (T ) cosmologies have been
constrained through cosmography by considering non-flat and flat geometries [525] with GRB observables, using RGB
objects [526] detected through high-quality UV and X-ray fluxes up to z ∼ 5.1 [527]. In view of possible discrepancies
among measurement of the matter critical parameter via sources with different z-ranges, it has been proposed a
promising model in which Ω0,m becomes a dynamical quantity. The proposed scenario relies on a metric f (R)-gravity
in the so-called Jordan frame and dominant linear contribution is retained in the potential term of the non-minimally
coupled scalar field (a small deviation being included too). It is such a dominant contribution that is responsible for a
non-standard critical matter parameter [528]. From the cosmographic point of view, other analyses have highlighted a
significant tension between the standard cosmological model and the best-fit cosmographic model, when using RGBs,
SNIa, and GRBs [527], RGBs and SNIa [529], and RGBs combined with GRBs, BAO, and SNIa [530]. Also, some
studies use RGBs as standard rulers [526] by their angular size–luminosity using very-long-baseline interferometry
[531]. In addition, the joint analysis of spectroastrometry and reverberation mapping can measure AGN distances
and provide a new way to measure H0 [532]. Furthermore, recent results have considered RGB physics from UV,
X-ray, and optical plane techniques behind the local observations as cosmological probes to study the H0 tension
[533]. Another application of RGBs in cosmology is the Sandage test of the cosmological redshift drift, i.e., a small
dynamic change in the redshift of objects following the Hubble flow [534], which requires not only high instrumental
resolution but also very bright targets. One of the main objectives of the QUBRICS (QUasars as BRIght beacons for
Cosmology in the Southern Hemisphere) survey [535] is to provide a sample of bright targets for the Sandage test.
The two brightest RGBs in this Golden Sample have started to be analyzed in the pilot program “An ESPRESSO
Redshift Drift Experiment”, which will be a complete end-to-end proof of concept for this experiment at ANDES in
ELT, where the redshift drift signal is expected to be detected [536]. Although the Sandage test is a direct and real-
time model-independent mapping of the expansion rate of the Universe, recent theoretical studies such as Refs. [537]
and [538] show synergies with other cosmological probes, in particular regarding the characterization of the physical
properties of DE.
Regarding the efforts to reach a higher precision in the determination of cosmological parameters, also a purely
statistical approach has been employed. This is based on the fact that the normalized residuals of RGB logarithmic lu-
minosities of the Risaliti-Lusso relation are not normally distributed. Hence, the Gaussian is not the most appropriate
distribution to be used as cosmological likelihood. Indeed, Ref. [516] prove that the RGB have the logistic distribution
as a best-fit. This is shown in the right upper panel of Fig. 26 along with the same investigation for GRBs (left upper
panel) and SNIa (lower panels). Remarkably, the employment of this proper likelihood for RGBs, compared to the
standard Gaussian one, reduces the uncertainty on H0 up to 35%, on Ωm,0 up to 27%, on Ωk up to 32%, and on w up
to 31%, when non-flat and wCDM models are investigated by combining RGBs with GRBs, BAOs, and SNIa, where
every single probe is fitted with its own best-fit cosmological likelihood [517]. Another approach aiming at inferring
cosmological parameters with lower uncertainties consists of identifying a RGB sub-sample composed only of sources
that better follow the Risaliti-Lusso relation, i.e., are closer to the best-fit relation. Indeed, this method allows us
to reduce the intrinsic dispersion of the correlation, thus yielding a better precision of the cosmological parameters.
Different techniques have been applied to select such a sub-sample [528, 539, 540] which have promoted RGBs as
standalone cosmological probes, even with the same precision as SNIa.
53

FIG. 27: The distance luminosity vs redshift when a calibration with CC is considered. The figure has been taken
by Ref. [430].

2.1.16. GRBs as cosmological standard candles


Coordinator: Maria Giovanna Dainotti
Contributors: Aleksander Łukasz Lenart, Arianna Favale, Denitsa Staicova, Giada Bargiacchi, Hassan Abdalla,
and Radosław Wojtak

Long-duration GRBs (LGRBs) could be used as standard candles, extending the capabilities of the Hubble diagram
to measure distances further than currently possible with SNIa, thereby helping to constrain cosmological parameters.
Phenomenological relations derived from spectral modeling, e.g., the Amati relation [541–543] can be used for this
purpose. This relation connects the source-frame energy (Ei,p ), where the gamma-ray spectral energy peaks, with the
isotropic-equivalent bolometric energy (Eiso ) released during the prompt phase [544]. Another significant empirical
correlation in the prompt emission is the Yonetoku relation [545] between Epeak with the intrinsic peak luminosity Liso .
This correlation has been validated for both LGRBs and short-duration GRBs (SGRBs) [545, 546]. Given the large
variability in the feature of the prompt emission, a more promising approach is obtained with the use of correlations
involving the plateau emission [547], a relatively flat segment in the light curve (LC). The Dainotti relation relates
the rest-frame time at the end of the plateau (Ta∗ ) and its corresponding luminosity (La ). This relation succeeds
in standardizing GRBs by utilizing the shape of their LCs. Indeed, around 500 GRBs observed by Swift exhibit a
distinct, plateau phase in X-rays. The great advantage of this relation, compared to the ones in the prompt emission,
is that the plateau phase can be attributed more straightforwardly to various theoretical models, including magnetar
spin-down [548–551] and forward/reverse shock mechanisms [552–554].
This correlation was later updated in Refs. [555–557], and subsequently extended to a 3D correlation by incorpo-
rating the peak luminosity (Lpeak ) [555, 558–560], the so-called 3D Dainotti relation. [557, 558] demonstrated that
this correlation is devoid of biases and it can be corrected for the redshift evolution.
Further studies by Refs. [559, 561] confirmed the robustness of this correlation for high-quality LGRBs, which show
well-defined plateau properties that obey the fundamental plane relation and constitute the so-called Gold sample.
An improvement of this sample has been classified as the Platinum sample [562]. As a result, this correlation has been
successfully applied as a cosmological tool [517, 563–566], providing insights into the high-redshift Universe. GRBs,
being at high-z, could be one of the important tools to differentiate between competing cosmological models. Indeed,
Ref. [516] pinpointed traces of rising tension between flat ΛCDM and the equivalent model with Ωk ̸= 0. There is a
trend that reveals Ωk < 0, marking the importance of further studies in this direction at high-z.
The future cosmological role of GRBs has also been investigated in Ref. [565] where it has been determined how
many GRBs are needed as stand-alone probes to achieve a comparable precision on Ωm,0 to the one obtained by SNIa
only. They obtained the same error measurements derived using SNIa in 2011 and 2014 with 142 and 284 simulated
optical GRBs, respectively, considering the error bars on the variables halved. These error limits will be reached in
2038 and in 2047, respectively. Using a doubled sample, obtained by the current ML approaches [567, 568] allowing
a LC reconstruction [569] and the estimates of GRB redshifts, with error bars halved, the same precision as SNIa in
2011 and 2014, is reached now and in 2026, respectively. If we need to reach the current precision of SNIa, it would
require 18 years from now, but this estimate is very conservative since it does not consider the new redshifts from
Euclid and JWST.
Indeed, in light of the existing cosmological tensions, such as the one on the Hubble constant, H0 , it has become
fundamental to obtain unbiased cosmological distances via intermediate redshift probes. Thus, the calibration proce-
54

dure of the GRB correlations must not be subject to strong model-dependent assumptions. Works have shown that
this can be done by employing low-z probes, such as SNIa from the Pantheon Plus sample, a collection of 1701 SNIa,
discussed already in the previous section, in combination with advanced statistical techniques such as GP [570] or
cosmographical analyses [571, 572]. On the other hand, data on the H(z) obtained from CC can play the role of
calibrating the ladders [429] since these data only rely on General Relativity and the validity of the standard physics
in the environment of the galaxy stars. Ref. [430] found that current data on CC can identify a subset of GRBs in the
Platinum sample in 0.553 ≤ z ≤ 1.96 that reveal a tight fundamental plane relation, with one of the lowest intrinsic
+0.03
scatter observed to date, σint = 0.20−0.05 . This result is obtained considering a relation corrected for evolutionary
effects and with parameters compatible with the magnetar model [549–551, 573]. The GRB sub-sample pinpointed
by this model-independent calibration can represent a valuable set of standardizable candles that is used to extend
the cosmic distance ladder by providing unbiased luminosity distances up to z = 5. For RGB, calibrating with
SNIa, varying only H0 in the cases without and with correction for evolution, H0 = 73.76 ± 2.18 km s−1 Mpc−1 and
H0 = 69.82 ± 2.27 km s−1 Mpc−1 , respectively [518]. For GRBs, using the Gaussian priors, varying only H0 in the
cases without and with correction for evolution, H0 = 73.23 ± 3.31 km s−1 Mpc−1 and H0 = 72.87 ± 2.92 km s−1 Mpc−1 ,
respectively [566].
Additionally, the 3D X-ray Dainotti relation has been also recently employed in cosmographic analyses to investigate
the reliability of the standard cosmological model in a cosmology-independent way. In this concern, as mentioned in
the previous section, GRBs combined with SNIa, RGBs, and BAO point toward a statistically significant discrepancy
between the prediction of the concordance model and the observational data Ref. [530]. From another point of
view, [574] constrained the transition epoch between a matter-dominated and a DE-dominated universe by using
two cosmographic approaches and the combination of GRBs, SNIa, and BAO obtaining results compatible with the
concordance model. Concerning instead the statistical analyses already mentioned for RGBs, the residuals of the
distance moduli of GRBs remarkably prove to be effectively Gaussian, as reported in the upper left panel of Fig. 26,
differently from the cases of RGBs and SNIa.
Given the importance of the 3D Dainotti relation, it is crucial to explore the existence of the same correlations but
with different associated classes (not only the Platinum sample). The LGRBs associated with Supernovae Type Ib/c
(GRB-SN) are interesting since they have been discussed in the literature as possible standard candles [562, 575].
Ref. [576] investigate the existence of probable correlations among SN parameters and GRB prompt and afterglow
features considering the largest compilation of GRB-v possible associations observed from 1997 up to 2021 and find a
possible correlation between the GRB optical luminosity at the end of the plateau (log10 La,opt ) and the log10 of SN
rest-frame peak time (log10 t∗p ). The correlation can be expressed as log10 La,opt = (9.43±1.47) log10 t∗p −(13.60±11.89).
The uncertainties on the fitting parameters (16%-87%) are too high to allow using this correlation for standardizing
GRBs, thus new observations are needed to validate it. If confirmed, then it will represent a crucial support for
cosmological analysis.
Another interesting application of GRB physics in cosmology comes from the search for quantum gravity models.
Some such models predict an energy-dependent speed of light, which can be observed as simultaneously emitted high
and low-energy photons arriving at different times. Such a tiny quantum effect is expected to be amplified by very high
energies of the photons and cosmological distances, making GRBs a promising probe. In Refs. [577, 578], the authors
explored the use of different GRB time-delay datasets in combination with the Pantheon/Pantheon+ SNIa datasets,
BAOs, and the CMB distance priors under different approximations for the intrinsic lag to constrain cosmological
parameters and investigate the impact on the H0 tension. The analysis revealed that the inclusion of such time-delay
datasets still leads to a deviation in the parameter c/H0 rd , where c is the speed of light and rd is the comoving
sound horizon at the drag epoch. Such deviation can be interpreted as an artifact of the H0 tension on the remaining
quantities, demonstrating again that the tension is spread to the whole H0 − rd − Ωm,0 plane. Such a dataset is not
directly affected by the Ep − Eiso correlation, but relies critically on the assumption of Lorentz Invariance Violation
(LIV), which so far has not been detected. In the case where LIV is not zero, GRB time delays would be a new
independent cosmological probe.
In addition, the constraints on the cosmological parameters can be obtained with the interaction of the gamma rays
with their surrounding medium or other radiations. Gamma rays with energies above around 10 GeV are attenuated
via interactions with the extragalactic background light (EBL) photons, resulting in electron-positron pair production.
The attenuation effect was systematically measured in spectra of blazars and GRBs, using a wide range of instruments
and techniques, e.g., see Ref. [579, 580]. Since the attenuation scales with the comoving distance, these measurements
can be used to constrain cosmological parameters, in particular H0 , given a model of the EBL based on integrating
light from galaxies in deep cosmological surveys [581]. Recent joint modeling of the EBL and γ-ray attenuation yields
−1
H0 = 65.1+6.0
−4.9 km s Mpc−1 , and H0 = 66.5−2.1+2.2
km s−1 Mpc−1 when combined with the BAO observations and the
BBN prior [582]. The H0 measurements are independent of any external distance calibration, and the best-fit values
are in close agreement with the CMB-based Planck value [192].
55

2.1.17. Gravitational wave constraints


Coordinator: Antonella Palmese
Contributors: Bangalore S. Sathyaprakash, Ivan de Martino, Matteo Tagliazucchi, Nicola Borghi, and Nicola
Tamanini

The field of GW astronomy has recently facilitated novel measurements of the Universe’s expansion rate by exploiting
unique properties of compact binary mergers. By providing a distinct perspective on the Universe, in contrast to
traditional astronomical and cosmological probes (such as electromagnetic radiation, cosmic rays, and neutrinos),
GWs offer a complementary approach to addressing cosmological tensions.
Since 2015, a ground-based GW detector network, including the Laser Interferometer Gravitational Wave Obser-
vatory (LIGO) [583] and later Virgo [584] and KAGRA [585], has been detecting coalescences of compact object
binaries, which so far include binary neutron star (BNS) [586], neutron star-black hole (NSBH) [587], and binary
black hole (BBH) [588] mergers. It is anticipated that LIGO-Virgo-KAGRA (LVK) will complete their fourth and
fifth observing runs (O4 and O5) by the early 2030s [589], after which a significantly upgraded detector network,
called A# [590, 591] is expected to become operational. This will be followed by the deployment of next-generation
(XG) observatories, such as the Cosmic Explorer [592–594] and Einstein Telescope (ET) [595]. Additionally, the Laser
Interferometer Space Antenna (LISA) [596] is scheduled for launch in the next decade and will be sensitive to GWs
from the inspiral and merger of binaries within the 0.1 mHz to 1 Hz frequency range, including massive BBHs and
extreme mass ratio inspirals (EMRIs). LISA and XG observatories will observe mergers throughout the cosmos from
an epoch before the first stars formed.
GW events can be used as “standard sirens” (StS; [597, 598]), as they can act as absolute distance indicators. The
GW signal of a compact binary merger is directly sensitive to the luminosity distance of the source and can therefore
be combined with redshift information to act as a probe of the expansion of the Universe. Depending on where the
redshift information is derived from, StSs are typically divided into different classes, with bright and dark StSs being
the two main classes. In what follows, we briefly describe each method and the state of the field, including latest
measurements and systematics studies, and end with prospects for future measurements, focusing on H0 constraints
as a pathway to shed light on the Hubble tension.
2.1.17.a. Bright standard sirens For bright StSs, an electromagnetic (EM) counterpart is identified, and the
redshift is derived from its host galaxy, assuming that both are unique. So far, there exists only one EM counterpart
that is confidently associated with a GW event, GW170817 [599]. This association has been used to derive the first
−1
StS measurement [600], finding H0 = 70+12 −8 km s Mpc−1 , a ∼ 14% precision measurement. Subsequent analyses
have taken a more in-depth approach to estimate the peculiar velocity of the host galaxy [601–603]. Moreover, various
works attempted to take advantage of the EM observations to constrain the viewing angle of the binary and break the
distance-inclination angle degeneracy which exists when estimating these parameters from the GW data [604–607].
The latest of these measurements reached a ∼ 7% precision and they are all consistent with both early and late-time
measurements of H0 .
It is worth noting some of the major systematics that may affect future bright StS measurements. A StS analysis
only using the GW data and the host galaxy spectroscopic redshift of a high-confidence EM counterpart provides an
H0 measurement which is not expected to first order to be affected by major systematic uncertainties, assuming that
the peculiar velocities measurements are well constrained (also considering that their uncertainties will become less
important as we move towards larger GW detector distance horizons [608]) and detector calibration errors are kept
under control (≲ 2% [609]). Although GW selection effects are well understood and can be more easily modeled,
selection effects from EM searches also need to be taken into account and may be coupled to binary parameters such
as source luminosity distance and viewing angle [610]. Beyond selection effects, if including EM estimates of the
viewing angle, H0 measurements may also be affected by modeling uncertainties of the jet structure [611], kilonova
geometry [612], or even a possible jet misalignment with the binary angular momentum [613].
Future bright StSs are most likely expected to arise from binary neutron star (BNS) and neutron star black hole
(NSBH; which would prove valuable StSs [614]) mergers, but candidate EM counterparts to BBH mergers also exist
[615–617]. Although the BBH EM counterparts association is currently uncertain [618, 619], they may in the future
be used to infer the Hubble constant along with other cosmological parameters [620–624].
2.1.17.b. Dark standard sirens Most mergers detected by the LVK (currently in the order of ∼ 200 when including
significant detection candidates from the ongoing observing run) do not have an associated EM counterpart and are
mostly comprised of BBH mergers. In that case, the redshift information necessary for an StS analysis may be taken
from galaxy catalogs (e.g., see Refs. [625, 626]), from the redshifted mass measured with the GW data [627, 628], or
a combination of the two [629–631].
In the first case, the redshift distribution of galaxies along the line of sight of a GW event’s localization can be
used to infer cosmological parameters, including H0 , from the distance-redshift relation. Various works have produced
56

such measurements so far [632–640], in general finding a precision on H0 down to ∼ 20% following the third LVK
observing run, again in agreement with both Planck and SH0ES constraints. Even when combined with the bright
StS available, the precision reaches ∼ 10% (when ignoring EM viewing angle constraints) and does not yet allow us
to distinguish between the two leading H0 measurements.
Golden dark sirens are a subclass of dark sirens when the catalog contains, or follow-up observations reveal, a single
galaxy in the 3D localization volume. Although rare, they may provide exquisite few percent-level measurements
of H0 [641]. It is also possible to cross-correlate GW events with galaxy catalogs to infer cosmological parameters
(e.g., see Ref. [642]). In all cases, the GW localization precision and the availability of extended spectroscopic galaxy
catalogs [631] will be crucial to establish dark StS as a cosmological probe competitive with bright StSs.
In the second case, the source redshift is reconstructed at the statistical level by taking advantage of the presence
of features in the distributions of the BBH population properties, such as the mass gap that may be explained by the
theory of pair-instability supernova [643, 644], or the overdensity peaks in the mass distribution observed in LVK data
[645, 646]. This “spectral sirens” approach has allowed constraints on H0 at the ∼ 60% level [635] and on modified
GW propagation [647, 648]. These constraints are expected to significantly improve with XG detectors, as all BBH
events can be used as spectral sirens. However, this poses a computational challenge, since the time required for a
full cosmological and population analysis scales linearly with the number of events.
The major sources of systematics for dark sirens methods are expected from incorrect assumptions on the mass
function [635]; while inferring cosmological parameters in conjunction may mitigate this effect [629], it is still crucial
to correctly model the mass function including its potential evolution [628]. Other systematics may arise when
using photometric redshifts [649, 650], and host galaxy weighting or galaxy assumptions that do not match the true
underlying distribution of merger hosts [651, 652].
2.1.17.c. Prospects for ground-based detectors Fig. 28 in Ref. [653] summarizes the prospect of terrestrial GW
observatories in resolving the Hubble tension. The x-axis lists GW detector networks that are expected to become
operational or newly built over the next decade or more. For each network, the y-axis shows the precision with which
the Hubble constant would be measured, for four classes of binary merger events: BBH, NSBH and BNS as golden
dark sirens, as well as BNS mergers with associated EM counterpart. For each network the figure also shows the
number of events expected to be detected in each source class. A target precision of 2% could be achieved by the HLV
network with dark sirens but not guaranteed since the number of events expected is O(1). If the current median BNS
merger rate holds, then the HLV network might observe ∼ 130 bright sirens and accomplish the 2% target. Assuming
one ET detector and its synergy with Transient High Energy Sources and Early Universe Surveyor (THESEUS) [654]
for bright StSs, one may achieve an accuracy on H0 of only 0.40 km s−1 Mpc−1 in the case of a non-flat ΛCDM model
after five years of observations (which corresponds to ∼ 166 bright StSs detections with redshift below z ∼ 4.3) [655].
ET alone, or a network of only two CE detectors, will not be able to accomplish the 2% target with dark sirens,
although a network of two CE detectors could do so with bright sirens. A network consisting of an ET and at least
one CE will determine H0 to sub-percent precision with dark sirens of all source classes, as well as with hundreds
of bright sirens. In the case of DE models, the accuracy on the Hubble constant may achieve 1% after 10 years of
observations due to the correlations between H0 and the DE parameters of the specific model under consideration
[656]. It is also worth noting that bright sirens with XG will reach a per-cent level precision in distance in the local
Universe, thus enabling precision measurements of σ8 [657] which will inform our understanding of the S8 tension,
should it persist. With XG observatories we will enter the era of precision cosmology with GW observations.
2.1.17.d. LISA perspective The LISA [658] will push GW cosmology at high redshift (z ≳ 3), mapping the
expansion of the Universe in a still poorly charted cosmic epoch [659]. LISA will observe both bright sirens, in the
form of massive black hole binary mergers with an identified EM counterpart [660, 661], and dark sirens, in the form
of extreme mass ratio inspirals (EMRIs) [662]. Independently these two populations of standard sirens are expected
to deliver percent constraints on the Hubble constant [663–668]. Combined together however they should provide a
one percent, or even better, constraint on H0 [669], making LISA a rightful competitor to solve the Hubble tension.
The large redshift of LISA standard sirens will furthermore provide original tests of ΛCDM and of general relativity,
delivering new potential insights on the nature of DE [664, 670–674].

2.1.18. The cosmic microwave background radiation


Coordinator: Margherita Lembo and Martina Gerbino
Contributors: Anto Idicherian Lonappan, Chandra Shekhar Saraf, Enrico Specogna, Luca Caloni, Özgür Akarsu,
Nils A. Nilsson, Paweł Bielewicz, Simone Paradiso, and Venus Keus

2.1.18.a. Review of CMB basics and recent observations The CMB plays a crucial role in shaping our under-
standing of modern physics and remains a powerful tool for advancing knowledge in cosmology and particle physics.
57

Gravitational − wave Standard Sirens


BBH (Dark) NSBH (Dark) BNS (Dark) BNS (Bright)
0
10

−0 events

−0 events

−0 events

−0 events

−0 events

−0 events

−0 events
H0 prior

20 events
10−1

−1 events
0+3

0+1

0+1

0+1

0+1

0+3

0+1
∆H0 /H0

−1 events

196 events

1+2
2%
1+3

131 events

−3 events
−2
10

−2 events
−3 events

346 events

−7 events

347 events
3+5

−3 events

−4 events
2+4
5+3

11+8
7+2

5+5
10−3

10−4
HLA ET CE4020 CE40ET CE4020ET
Networks

FIG. 28: Precision with which H0 would be constrained by standard sirens with the HLV network upgraded to A#
sensitivity and networks containing one or more XG observatories (see Sec. 2.1.17.c text for details). Figure taken
from Ref. [653].

The CMB dates back to roughly 380000 years after the Big Bang, at a redshift of z∗ ≃ 1100, when CMB photons
decoupled from electrons, close to the moment in the thermal history of the Universe when electrons and protons
combined into neutral hydrogen atoms. Since then, CMB photons free stream with a nearby blackbody distribution
with a temperature today of approximately T ≃ 2.7 K [675]. The CMB spectrum peaks around 100 GHz, within the
microwave frequency range of the electromagnetic spectrum. Primordial density perturbations at the last scattering
surface gave rise to temperature anisotropies in the CMB field of the order of ∆T /T ∼ 10−5 [676]. The CMB radi-
ation, because of its quadrupolar anisotropy, is indeed also partially linearly polarized due to Thompson scattering,
with an amplitude of ∼ 10% of temperature anisotropies.
The polarization pattern is usually decomposed into so-called (curl-free) E-modes and (divergence-free) B-modes.
On small angular scales, both the temperature and the polarization fields undergo a tiny distortion when they pass
close to large distributions of matter. The weak gravitational lensing effect (CMB lensing) is one of the most important
mechanisms that can generate secondary anisotropies in the CMB. In particular, it smears out anisotropies at small
angular scales in temperature and E-mode polarization, and it gives rise to an additional B-mode contribution sourced
by the lensed E-mode pattern.
The first full sky mapping of the CMB anisotropies was made by the COBE satellite [677], refined by WMAP [678]
and finally by Planck [192], that delivered the state-of-the-art for full-sky CMB measurements both in temperature
and polarization. At smaller (arcmin) angular scales in temperature and E-mode polarization, observations have been
dominated in sensitivity by the ground-based experiments ACT and SPT; the search for B modes has been led so far
by ground-based telescopes observing degree angular scales, i.e., BICEP/KECK, POLARBEAR/Simons Array.
Next-generation CMB experiments, both space-borne (e.g., LiteBIRD [679]) and ground-based (e.g., Simons Ob-
servatory [680] and CMB-S4 [681]), aim to achieve precise measurements of CMB polarization. A clear science goal
is to detect the imprint on CMB pattern of primordial GWs, which are the smoking gun of an inflationary phase in
the early Universe [679, 682, 683]. Moreover, improved measurements of CMB polarization would provide insights on
potential new physics beyond the standard model [684–692]. CMB lensing is one of the most relevant observable of
near-future CMB experiments, as it provides an invaluable tool to reconstruct the integrated distribution of matter
in the Universe and to place stringent constraints on physical properties and effects mostly related to the late-time
phases of the Universe, such as the growth of structures and neutrino masses (see e.g., Planck [693], ACT [694],
SPT [695]).
2.1.18.b. CMB-driven constraints in the context of cosmological tensions CMB observations allowed us to test
the predictions of the standard cosmological model and to shape our knowledge of the Universe, its history and
58

composition. Despite its phenomenological success, the ΛCDM model is still incomplete in that, for example, fails in
addressing the fundamental nature of the most abundant dark components. In addition, it has been facing challenges
due to statistical tensions between results obtained with recent cosmological and astrophysical surveys of increased
accuracy. Most notably, there is a severe (> 5σ) discrepancy in the value of the Hubble constant H0 estimated with
CMB observations with respect to the result obtained with local distance ladder measurements [34, 192]. Furthermore,
CMB and LSS measurements show a less severe albeit still intriguing disagreement on the amplitude of matter
perturbations quantified via the σ8 parameter [192, 696–698]. These discrepancies and their implications will be
explored in more detail in the following subsections from the perspective of CMB observations.
2.1.18.c. Constraints on H0 from CMB measurements The Hubble tension is one of the major unresolved issues
in modern cosmology. This tension refers to the discrepancy between local measurements of the Hubble constant, H0 ,
and the value inferred from early Universe observations assuming the ΛCDM model. Early Universe estimates of H0
are driven by CMB observations made by the Planck satellite, which yields the value H0 = 67.27±0.60 km s−1 Mpc−1 at
68% CL, using a combination of TT,TE,EE+lowE data [192] (see also Refs. [699, 700] for re-analyses of Planck data,
with no significant deviations of H0 from the Planck result). On the other hand, the most sensitive local measurement
of H0 is obtained by the SH0 ES collaboration, which finds H0 = 73.04 ± 1.04 km s−1 Mpc−1 exploiting type SNIa
calibrated with Cepheids [34]. These two values are in 5σ tension.
Alternative CMB datasets, such as WMAP [678], ACT [701, 702], and SPT [703], or CMB-independent probes,
such as combinations of BBN with BAO measurements [704], yield H0 values that align with those obtained from
Planck. The analysis of ACT-DR6 data [702] yields a 68% CL value of H0 = 66.11 ± 0.79 km s−1 Mpc−1 (ACT
TTTEEE+Planck-Sroll2 EE at large scales to constrain tau) consistent with measurements obtained by Planck satel-
lite and lower than local measurements. Similarly, the combination of ACT-DR6 and WMAP data results in a value
of H0 = 66.78 ± 0.68 km s−1 Mpc−1 at 68% CL [702, 705]. Additionally, the combination of ACT and WMAP data
results in a value of H0 = 67.6 ± 1.1 km s−1 Mpc−1 at 68% CL [701]. The same collaboration also reported an es-
timate of H0 = 68.3 ± 1.1 km s−1 Mpc−1 from measurements of CMB lensing in combination with BAO data (6dF
and SDSS) [694]. The SPT collaboration finds a compatible result, with a value of H0 = 68.3 ± 1.5 km s−1 Mpc−1 at
68% CL [703], improved to 66.81 ± 0.81 km s−1 Mpc−1 with the use of the latest unlensed EE and CMB lensing data
in combination with a prior on the optical depth τ [706]. It is relevant to emphasize that all the results above are
independent of Planck data. On the other hand, late-time measurements different from SH0 ES, including additional
methods for calibrating SNIa at large distances (see Sec. 2.1.7) or observations independent of SNIa (see e.g., Sec. 2.1.9,
Sec. 2.1.13, Sec. 2.1.10, Sec. 2.1.17), are all suggesting higher values of the Hubble parameter.
Depending on the combination of measurements, the tension ranges roughly from 4σ to 6σ. Fig. 29 shows a
comprehensive list of CMB measurements of the Hubble parameter. It is important to emphasize that these results
were derived under the assumption of the ΛCDM model. For example, as shown in Ref. [707], the uncertainty on H0
increases when considering extensions to the ΛCDM framework.
Assuming a different cosmological model may lead to different estimates of H0 , either in the mean value or in the
width (i.e., in the errorbar) of the PDF (or both) [3, 708, 709].
Since the value of H0 inferred from CMB observations depends on the underlying cosmological model, this dis-
crepancy with the local measurements of the Hubble parameter might be an indication of physics beyond the ΛCDM
model. An alternative possibility, which will be discussed in Sec. 2.1.18.e, is that this tension is due to unknown
systematics that affects either early-time or late-time observations (or both). However, given the vast array of inde-
pendent probes of H0 , it seems quite unlikely that systematic effects may be the only responsible for this discrepancy.
This further motivates the search for an explanation in terms of new physics. The landscape of theoretical scenarios
to address the H0 tension is broadly discussed in Sec. 4. Here, we just recall that the easiest extensions to ΛCDM are
unable to solve the tension. More exotic scenarios, like those involving new BSM interactions between the components
of the stress-energy tensor, the existence of new components or variation of fundamental constants, usually do not
properly alleviate the tension, but rather yield estimates of H0 with a larger uncertainty than the ΛCDM-based infer-
ence. Some models which gained visibility in the past, like those with strongly self-interacting neutrinos, require new
physics which is at odds with complementary constraints from laboratory probes [710]. The search for new physics is
still ongoing.
In order to understand the possible physical origin of this tension, we need first to discuss how CMB observations
constrain H0 (see Refs. [711, 712]). The key quantity from which the Hubble constant is inferred is the angular scale
of the sound horizon at recombination, which we assume to happen instantaneously at redshift z∗ ≃ 1100. This is
defined as
rs (z∗ )
θs ≡ , (2.7)
DA (z∗ )

where rs (z∗ ) and DA (z∗ ) are the comoving sound horizon at recombination and the comoving angular diameter
59

Indirect measurements of H0 , assuming ΛCDM

CMB only
Planck18, Aghanim et al. (2020)

Planck18+CMB lensing, Aghanim et al. (2020)

Planck PR4, Tristram et al. (2024)

WMAP9, Hinshaw et al. (2013)

ACT DR4, Aiola et al. (2020)

ACT DR4+WMAP9, Aiola et al. (2020)


SH0 ES
SPT-3G, Dutcher et al. (2021) Riess et al. (2022)

Planck18+SPT-3G+ACT DR4, Balkenhol et al. (2021)

CMB + other cosmological probes


DESI BAO+Planck18+CMB lensing, Adame et al. (2024)

eBOSS DR16+Planck18 ωm , Pogosian et al. (2020)

WMAP9+BAO, Zhang and Huang (2019)

SPT-3G+BAO, Dutcher et al. (2021)

ACT CMB lensing+BAO, Madhavacheril et al. (2024)

65 67 69 71 73 75
−1 −1
H0 [km s Mpc ]

FIG. 29: Whisker plot with 68% CL constraints on the Hubble constant H0 through different indirect measure-
ments performed over the years. The constraints shown in this plot have been derived assuming ΛCDM model. If
not specified, “BAO” refers to a combination of BAO data prior to DESI (see the corresponding paper for more de-
tails). The green vertical band corresponds to the H0 value from SH0 ES team [34], while the red one corresponds to
the value from Planck 2018 [192]. This figure has been adapted from Fig. 1 of Ref. [708].

distance to recombination, respectively:


ˆ ∞  ˆ z∗ 
cs (z) 1
rs (z∗ ) = Sk dz , DA (z∗ ) = Sk dz , (2.8)
z∗ H(z) 0 H(z)

with

for k < 0
 1 p
 √|k| sinh( |k|x) ,


Sk (x) = x , for k = 0 . (2.9)
 √1 sin( |k|x) , for k > 0

 p
|k|

Here, cs (z) = 1/ 3[1 + R(z)] is the sound speed of the acoustic waves in the baryon-photon plasma, where R(z) =
p

3ρb (z)/(4ργ (z)) = 3ωb /(4ωγ )(1 + z)−1 denotes the ratio between the baryon and photon energy densities, and k
is the spatial curvature of the Universe. Because of the so called “geometric degeneracy” [713], CMB anisotropy
measurements alone do not constrain the curvature of the Universe. To large extent the degeneracy can be broken
using CMB lensing effect. Tighter constraints can be imposed by analyzing CMB data in combination with other
cosmological probes (see Sec. 2.1.18.d). The Hubble parameter, H(z), is given in terms of the physical densities of
DE, matter, photons, neutrinos and curvature as
q
H(z) = (100 km s−1 Mpc−1 ) ωDE (z) + ωm,0 (1 + z)3 + ωγ (1 + z)4 + ων (z) + ωk (1 + z)2 . (2.10)
60

Within the ΛCDM model, ωDE (z) = ωΛ and ωk = 0. The only unknown quantities in Eqs. (2.8-2.10) are ωb , ωm,0 and
ωΛ . The radiation density ωγ is fixed by the CMB temperature obtained from measurements of the CMB blackbody
spectrum [675] and the neutrino density can be obtained at any time by properly integrating the neutrino distribution
function11 .
Then, H0 is inferred from CMB data as follows (see Ref. [712] for more details). First, we determine the baryon
and matter densities from their impact on CMB power spectra [714]. The baryon density affects the relative heights
of the acoustic peaks and the diffusion damping in the CMB power spectrum, while the matter density is mostly
determined via the “potential envelope” effect, i.e., the increase in power for modes that re-entered the horizon before
matter-radiation equality. The characteristic scale of this power boost corresponds to the comoving size of the horizon
at matter-radiation equality projected on the last scattering surface, which depends on the ratio between the matter
and radiation densities. Since ωΛ is negligible at redshifts z ≥ z∗ , we have everything we need to determine rs (z∗ ) via
Eq. (2.8).
Then, we obtain the acoustic scale θs from the spacing between the acoustic peaks, and, using Eq. (2.7) we derive
DA (z∗ ). As a last step, we adjust ωΛ in order to match the value of DA (z∗ ) calculated via Eq. (2.8) with the one
inferred as described above. Having determined the physical density of each component, we can finally reconstruct
H(z) via Eq. (2.10). For z = 0, this provides us with the value of H0 .
Note that θs is very tightly constrained by observations. In particular, it is measured to 0.03% accuracy by Planck
data, which gives 100θs = 1.04109 ± 0.00030 using a combination of TT,TE,EE+lowE data [192]. Other CMB probes
measure θs to be consistent with Planck, such as 104 θs = 104.056 ± 0.031 from ACT-DR6 (TTTEEE+Planck-Sroll2
EE [702]) and 100θs = 1.04016 ± 0.00067 (SPT+WMAP [703]).
2.1.18.d. CMB in combination with other cosmological probes The same acoustic oscillations that we observe in
the CMB are also left imprinted in the galaxy power spectra in the form of BAO. The characteristic scale of BAO is
the sound horizon at the drag epoch, zd ≃ 1059 [192], i.e., the time when baryons were released from the drag of CMB
photons. This characteristic scale provides us with a cosmological standard ruler that serves as an independent way
to measure the expansion rate of the Universe and improve our bounds on the cosmological parameters. Notably, the
inclusion of BAO data allows us to better constrain the spatial curvature of the Universe and break the aforementioned
geometric degeneracy [713, 715]. We refer to Sec. 2.1.19 for a more detailed discussion of BAO measurements of H0 .
In what follows, we report an incomplete list of the most recent estimates of H0 obtained with the combination of
CMB and LSS data, to give a sample of the complementarity and constraining power of these combinations.
A recent analysis of DESI BAO data in combination with Planck PR4 primary anisotropies (using small, commander
and CamSpec likelihoods), Planck and ACT-DR6 CMB lensing, leads to H0 = 68.17 ± 0.28 km s−1 Mpc−1 [716]. A
combined analysis of Planck CMB lensing, BOSS DR12 galaxy power spectra, and the PANTHEON+supernova
−1
constraints yields H0 = 64.8+2.2 −2.5 km s Mpc−1 [717], by imposing a BBN prior on physical baryon density and
assuming ΛCDM cosmology. A multi-tracer full-shape analysis of luminous red galaxy (LRG) and emission line
galaxy (ELG) samples from the eBOSS DR16 measures H0 = 70.0 ± 2.3 km s−1 Mpc−1 in the ΛCDM framework when
combining the BBN prior and fixing the spectral tilt ns to Planck value [718]. Another analysis of the BOSS DR12
data by fixing ns and the baryon-to-DM ratio Ωb /ΩDM to Planck value gives H0 = 68.5 ± 2.2 km s−1 Mpc−1 [719].
Adopting a forward modeling approach for the BOSS DR12 bispectrum monopole, with BBN prior on Ωb h2 , [720]
−1
found H0 = 67.6+2.2 −1.8 km s Mpc−1 . Another study of the BOSS galaxy power spectrum with bispectrum monopole
and quadrupole estimated H0 = 69.2±1.1 km s−1 Mpc−1 [721] within the ΛCDM model, using BBN and fixing spectral
tilt to Planck value.
It is crucial to highlight that these results assume a ΛCDM cosmology. A combination of Planck TT,TE,EE+lowE,
Planck lensing, BAO measurements from BOSS and eBOSS, and Pantheon+ SNIa data, assuming a non-flat ΛCDM
model yielded H0 = 68.24 ± 0.54 km s−1 Mpc−1 with Ωk = 0.0004 ± 0.0017 [722]. A similar constraint of H0 =
68.53 ± 0.56 km s−1 Mpc−1 was found when treating Alens as a free parameter (for more details, see the discussion
on the Alens anomaly in Sec. 2.1.18.e). Extending the analysis to non-flat wCDM model resulted in H0 = 67.95 ±
0.66 km s−1 Mpc−1 and Ωk = 0.0016±0.0019 (see Ref. [722] and references therein). To conclude this list of constraints,
the take-home message is that the combination of CMB and LSS measurements provide estimates of H0 which are in
agreement with the (lower) value inferred from Planck alone. This is true also in case of combinations which do not
include Planck anisotropies - or do not include Planck data at all - in their fit.
2.1.18.e. The role of potential systematic effects in CMB-based estimates of H0 Unresolved systematic effects
(hereafter systematics) in CMB measurements could potentially bias the cosmological constraints. This possibility
has motivated a collective effort to examine potential sources of systematics within the CMB dataset.

11 The parameter of the neutrino equation of state is not constant. Therefore, it is not possible to generally parametrize ων via a simple
scaling of the redshift, unless one considers the limiting cases of ultra-relativistic and non-relativistic neutrinos.
61

Features in the Planck spectra (especially temperature and at small scales) have been also interpreted as hints to
unsolved systematics. In this context, the Alens anomaly in Planck data is noteworthy. First introduced in Ref. [723],
Alens is an phenomenological (somehow “unphysical”) parameter used to rescale the effects of gravitational lensing on
the CMB angular power spectra. Therefore, the expected value is Alens = 1. Interestingly, Planck data (primary
anisotropies) show a preference for Alens > 1 at more than 2σ, increasing to more than 3σ when combined with
BAO data [192]. Theoretical explanations for such values are challenging, as it would require either a closed universe,
posing conflicts with other datasets and simple inflationary models, or more exotic solutions such as modifications to
General Relativity. Furthermore, the lensing anomaly does not appear in Planck trispectrum data (i.e., in the lensing
power spectrum), which offers an independent and complementary measurement of CMB lensing. If not indicative of
new physics or not interpreted as a statistical fluke, the Alens anomaly might arise from an undetected systematics
error in Planck data and this systematic could potentially bias the measurement of the H0 parameter. However,
when the effect of Alens is marginalized over in the analyses, the Planck and Planck + BAO constraints on H0 shift
only slightly towards higher values: H0 = 68.28 ± 0.72 km s−1 Mpc−1 and H0 = 68.23 ± 0.49 km s−1 Mpc−1 at 68%
CL, respectively [192], not enough to provide a satisfying solution to the H0 tension. If the Alens anomaly is due to
systematics which may impact the final H0 constraints, it raises questions whether the same systematics can be fully
captured by Alens alone or if further modeling is needed. Moreover, recent analyses by ACT and SPT, which also
find a lower value of H0 than local measurements, find no deviation from the standard lensing effect predicted by
the ΛCDM model [701, 702, 724], supporting the idea that the Alens anomaly might not be the right solution to the
tension.
The consistency of cosmological parameters estimated from different multipole ranges of Planck data has also been
investigated extensively in the past years [725–727], sometimes claiming evidence for internal inconsistencies which
may impact the H0 estimates. However, it has been showed [714] that the shifts in cosmological parameters recovered
from different ranges - mostly due to the combined effects in temperature of non-lensing-related residuals at high
multipoles and power deficit at large scales (ℓ < 30) - are not very significant and consistent with sample variance.
In the context of potential systematics in CMB data, it is also worth mentioning that the mild discrepancy seen
between ACT-DR4 and both WMAP and Planck has been solved with the final data release of ACT. Indeed, ACT-
DR6 finds very good agreement with both Planck Legacy data (PR3) and Planck PR4 (NPIPE) already at the power
spectrum level [702]. In terms of cosmological parameters, the difference between the best fit values in ΛCDMobtained
with ACT-DR6 and Planck is estimated to be within 1.6σ (for Planck Legacy) and 2.5σ (for Planck PR4), see [702, 705]
for details.
As far as SPT results are concerned, good agreement has been found [703] between SPT and Planck, both at
the power spectrum level (in TT over angular scales that are signal-dominated in both experiments) and at the
cosmological parameter level (assuming ΛCDM). As noted by the SPT collaboration, this excellent agreement between
two effectively independent experiments (given the negligible overlap between observed sky fractions and different
sensitivity to angular scales) is a strong argument in favour of the robustness of the results and of the consistency of
ΛCDM across different scales and spectra. A good agreement has been found also between SPT and WMAP. The
agreement between SPT and ACT is acceptable [703], with the largest shift in recovered ΛCDM parameters being on
θ (2σ larger in ACT than in SPT). However, a multi-dimensional statistical test results in no significant deviations
to be noted and agree with an explanation in terms of statistical fluctuations.
Another crucial aspect of the CMB data analysis to be explored in the context of H0 estimation is the possible
impact of instrumental systematic effects. In CMB data analysis, particular care is devoted to the quantification of
residual systematics, which are corrected for through the pipeline and/or propagated in the likelihood analysis either
by incorporating it in the noise model or modelling specific residuals in the data vector. In this context, the use
of end-to-end simulations is key. Among the different sources of systematics, beam characterization and instrument
calibration (especially in polarization) are particularly worrisome. As an example, the determination of polarization
efficiencies is considered one of the main limitations of the Planck data products [728]. Nevertheless, the use of large
suite of null tests and cross-checks on several data splits, and the comparison with realistic end-to-end simulations
allow to show, for all the data products released by the main CMB experiments, that possible biases of cosmological
parameters - included H0 - induced by residual systematics are small-to-negligible [701, 703, 706, 728–731]. Similar
analyses on dedicated simulations for next generation CMB experiments have recently shown that, even in the presence
of unaccounted-for systematics (i.e., effects which have not been corrected for in the analysis pipeline) due to, e.g.,
beam chromaticity, bandpass mismatch, polarization angle miscalibration and incorrect calibration in polarization,
the bias on H0 is at the level of a small fraction of σ [732, 733]. Therefore, we conclude that instrumental systematic
effects cannot be the (only) source of discrepancy between the CMB-based estimate of H0 and local measurements.
Finally, the interplay between instrumental systematics and foreground contamination can have non-linear effects on
the final results, potentially biasing cosmological parameter estimates, including the Hubble constant. The presence
of galactic and extragalactic foregrounds, such as dust emission and radio sources, complicates the separation of the
CMB signal from other astrophysical contributions. This complexity underscores the need for precise modeling and
62

mitigation strategies to ensure accurate parameter estimation, but it is unlikely the source of the H0 tension.
2.1.18.f. S8 tension The amplitude of the CMB power spectrum, especially CMB lensing, places stringent con-
straints on the matter density and hence on σ8 . Here, we briefly address the tension related to the amplitude of matter
clustering in the late Universe, described by the parameter S8 ≡ σ8 (Ωm,0 /0.3)0.5 . For a more detailed discussion,
refer to Sec. 2.2.
CMB-based estimates of S8 are consistently larger than those derived from galaxy-based measurements [192, 694,
699–701, 703, 706]. Observations of weak gravitational lensing at low redshifts (z ≲ 0.5 − 1) indeed suggest weaker
matter clustering than predicted by the ΛCDM model with parameters inferred from CMB data. Simply put, the
distribution of galaxies and matter in the late Universe appears smoother than expected based on the evolution of the
fluctuations seen in the CMB. The CMB estimate of S8 is model-dependent and can shift with other parameters that
affect the growth and amplitude of matter fluctuations. Notably, the optical depth to reionization (τ ), which carries
the largest uncertainty among the ΛCDM parameters, and the sum of neutrino masses ( mν ), which is a derived
P
parameter and is usually fixed. Uncertainties in these parameters affect the derived S8 constraints. Additionally,
Planck ’s excess lensing anomaly, as discussed in Ref. [734], can mimic a larger S8 . However, the CMB-LSS tension
persists even without Planck ’s lensing excess, as shown by the latest ACT+WMAP analysis [702, 705], which reports
a high value of S8 = 0.857±0.020 without an anomalous lensing amplitude, and South Pole Telescope - 3rd Generation
(SPT-3G) analysis [706], which is independent of temperature data and reports S8 = 0.850 ± 0.017.
On the theoretical side, efforts to resolve the S8 tension involve modifying the matter or gravity sectors of the
ΛCDM model, leading to a range of alternative cosmological scenarios (see e.g., Sec. 4.1 and Sec. 4.2 for more details).
Although some models alleviate the S8 tension, most fail to provide consistent solutions when all cosmological probes
are considered. In particular, due to the correlation between H0 and S8 , models that resolve the S8 tension often
worsen the H0 tension, and vice versa [735–737]. For example, late-time dark sector transitions, which prefer a
higher H0 value, often result in lower Ωm,0 to preserve the measured value of Ωm,0 h2 , leading to changes in structure
growth and CMB anisotropies and typically yielding a higher σ8 than ΛCDM due to extended matter domination.
Similarly, EDE models that address the H0 tension tend to increase σ8 because they require a higher initial curvature
perturbation amplitude to counterbalance the unclustered component. It is therefore crucial to perform joint analyses
fitting a full array of multiple datasets, without fixing any of the parameters of the model. Simultaneously any solution
to the S8 tension must be consistent with the growth history (typically studied through the parameter f σ8 (z)), probed
by e.g., BAO, galaxy power spectrum and void measurements.

2.1.19. Baryonic acoustic oscillations


Coordinator: Florian Beutler
Contributors: Armando Bernui, David Benisty, Denitsa Staicova, Felipe Avila, Ignacio Sevilla-Noarbe, Özgür
Akarsu, Maret Einasto, Mustapha Ishak, Nicola Deiosso, Rafael C. Nunes, Ruchika Kaushik, Samuel Brieden, and
Sveva Castello

BAO represent a characteristic scale in the distribution of galaxies, or more generally of any tracer of the matter
density field [738, 739]. This scale, known as the sound horizon rd , originated from sound waves travelling through
the early (z ≳ 1060) Universe before baryons and photons decoupled (see Refs. [740–742] for reviews of the subject).
We can employ this scale as a standard ruler, which allows us to measure the expansion history of the Universe. In
practice, the BAO features can be extracted from the two-point correlation function of galaxies, where it appears
as a peak [743, 744], or from the Fourier-space equivalent, the power spectrum, where it is manifested as a series of
oscillations [745–747]. In the following, we will review the different methods to extract the BAO scale from the galaxy
distribution. We will then discuss the most recent measurements from DESI.
2.1.19.a. Methods The 3D position of a galaxy can be determined by measuring its location on the sky (in right
ascension and declination) and its redshift. For this reason, measuring the BAO scale in the perpendicular direction to
the line-of-sight and along the line-of-sight provides constraints on slightly different quantities. The angular component
constrains the comoving angular diameter distance DM /rd = 1/∆θ, where ∆θ is the angular separation of the pair of
galaxies, while the line-of-sight component constrains the comoving Hubble distance DH /rd = 1/∆z, where ∆z is the
redshift separation of the galaxy pair. In the Friedmann-Lemaître-Robertson-Walker (FLRW) metric, the comoving
angular diameter distance DM , the related DA and the Hubble distance DH are given by
ˆ z
dz ′
 
c DM c
DM = p sinn |ΩK | 1/2
′)
, DA = , DH = , (2.11)
H0 |ΩK | 0 E(z 1 + z H 0 E(z)

where sinn(x) ≡ sin(x), x, sinh(x) for ΩK < 0, ΩK = 0, ΩK > 0 respectively, and E(z) is the normalized Hubble
function. Rather than separating DM and DH , often measurements are reported as a combination of the two given
63
1/3
by DV ≡ zDM (z)2 DH (z) . BAO constraints from photometric galaxy surveys usually have significant redshift

uncertainties that only allow for constraints on DM . One important point to note from the equations above is that
BAO measurements always constrain distances relative to the sound horizon scale. This introduces degeneracies
between cosmological parameters and the sound horizon scale rd . In ΛCDM for example, BAO can only constrain
H0 rd , and breaking this degeneracy is crucial to obtain measurements of H0 and provide key information in light of
the Hubble tension. We will discuss this further in Sec. 2.1.19.c.
The density field sources a velocity field that will displace galaxies away from their initial position, effectively
leading to a redshift-dependent smoothing of the localized BAO feature. This effect can reduce the signal-to-noise by
up to a factor of two at low redshift (e.g., see Ref [748]) and also results in a sub-percent level systematic biasing of
the BAO scale [749, 750]. Given that the velocity field is sourced by the density, one can estimate the displacement of
galaxies assuming standard gravity and correct for this effect [750, 751]. Such techniques are known as density field
reconstruction and are routinely applied to most spectroscopic BAO measurements.
The standard BAO analysis relies on converting the angular position and the redshift into Cartesian coordinates,
from which the 3D clustering statistics can be calculated. This step requires adopting a fiducial cosmology, which
typically is taken to be a flat ΛCDM model with parameters based on the CMB. The BAO analysis accounts for
these assumptions by including geometric scaling parameters and so far studies with alternative models (e.g., EDE
or MG) have not found that the fiducial cosmology assumptions inflict any significant bias in the standard BAO
analysis [752–755]. Through density field reconstruction, the standard analysis also makes assumptions about the
connection between the measured galaxy density field, the matter density field, and the large-scale velocity field.
However, just like with the fiducial cosmology assumptions, the impact on the measurements has been shown to be
negligible (e.g., see Refs. [756, 757]). Assumptions about the number of relativistic particles, Neff at high redshift have
been shown to impact the BAO phase and could bias BAO measurements [758, 759]. The standard analysis usually
fixes this parameter to the standard model value. While a Planck prior on Neff (within ΛCDM) reduces any potential
bias to well below the uncertainties of current measurements, one still should be aware of these prior assumptions.
Given the priors implicit in the standard BAO analysis, alternative analysis methods have been developed that
can (at least partly) avoid the assumption of a fiducial cosmology [760]. This technique often runs under the name
transverse BAO [761, 762], which should not be confused with the many angular BAO measurements in the literature
which follow the standard analysis technique described above (see e.g., DES Y6 in Fig. 30). The basic idea of
such methods is to bin the data into redshift bins and measure the angular BAO signature in those redshift bins
through two-point angular-clustering statistics, avoiding the need to convert redshifts into distances. However, to
account for the redshift evolution within the redshift bin one still needs to use a fiducial cosmology, and density
field reconstruction is generally not applied in such cases. For this reason, such methods usually have much larger
measurement uncertainties (see e.g., see Refs. [763–769]).
2.1.19.b. BAO measurements The BAO feature has been measured both with photometric and spectroscopic
surveys. Photometric surveys allow for a direct identification of a huge number of galaxies, but generally only provide
poor photometric redshift information. Even for measurements of the angular BAO scale, the redshift information is
necessary to split the sample into redshift bins, to avoid the smearing of the BAO signal due to its redshift evolution.
Although the photometric redshift quality can be improved with an adequate subselection, this usually also reduces
the number of galaxies in the sample [770]. To date, the best photometric BAO measurement comes from DES year-6
analysis using about 16 000 000 galaxies and yielding a 2.1% constraint on DM /rd [771]. Observational systematics
for such measurements are dominated by uncertainties in the redshift distribution, which are nevertheless still below
the statistical error budget.
Compared to photometric surveys, spectroscopic surveys have the advantage of providing precise redshift mea-
surements and detailed spectral information. This is achieved through the identification of specific spectral lines or
features, such as the Hα, Hβ, OII lines or the 4000 Å break for galaxies, and broad emission lines for RGBs. The
Lyman-α (Lyα) forest represents a series of absorption features in the spectra of distant RGBs that can be used to
map the distribution of intergalactic hydrogen gas and provides additional BAO measurements at higher redshift [772].
The first convincing detections of the BAO signal in the distribution of galaxies came from the 2-degree Field Galaxy
Redshift Survey (2dFGRS) and the Sloan Digital Sky Survey (SDSS) in the early and mid-2000s [743, 745, 773], while
the first detection in the Lyα forest was made by the BOSS [774]. There have been many more subsequent detections
in other galaxy surveys [747, 748, 775–781]. The best constraints to date come from DESI, which measured the BAO
signal in 8 independent redshift bins with a detection significance ranging from 3.3 to 9.1σ [744, 782]. A comparison
between these DESI measurements and previous BAO measurements is shown in Fig. 30.
Observational and instrumental systematics, such as fiber collisions [783], distortions due to peculiar velocities
(Redshift-Space Distortions (RSD)), non-linear matter and galaxy clustering, atmospheric dispersion, and spectro-
graph calibration are critical factors in any analysis based on spectroscopic surveys datasets. Since the BAO signal is
located on very large scales (rd ∼ 150 Mpc) and represents a distinguishable feature unlikely to be mimicked by any
instrumental or physical process, BAO measurements have proven to be very robust (see e.g., see Ref. [757]). The
64

27.5
DESI
6dFGS
25.0 SDSS
DES Y6
22.5 WiggleZ
Carvalho’20 (SDSS-tr)
Menote’22 (SDSS-tr)
20.0
Dx(z)/rd z

Planck ±1σ

DESI best-fit
17.5 √
DM (z)/rd z

DV (z)/rd z

15.0 zDH (z)/rd z

12.5

10.0 QSO
ELG Lya
BGS LRG
(DV /rd)/(DV /rd)best

1.1

1.0

0.1 0.2 0.5 1.0 2.0


Redshift z

FIG. 30: DESI BAO measurements (filled circles) compared to prior BAO measurements (empty symbols) from
spectroscopic (6dFGS, WiggleZ, SDSS) and photometric (DES Y6) surveys, where SDSS includes the MGS, BOSS
DR12 and eBOSS DR16 galaxy and Lyman-α Samples. Additionally, we show the transverse BAO measured from
SDSS catalogs (SDSS-tr) in thin redshift shells, neglecting the longitudinal modes. Note that these measurements
are not official SDSS products, but were instead obtained independently in Refs. [762, 784–787] (small grey squares)
and Ref.[761] (small pink squares). Horizontal arrows and coloured√regions indicate the redshift range spanned by
each DESI tracer type. Solid lines show BAO distances (scaled by z for improved visibility) as a function of red-
shift z for the standard flat ΛCDM DESI best-fit, and dashed lines indicate the Planck best-fit with shaded region
denoting the ±1σ regime. Upper panel: Comparison between transverse (pink), radial (blue), and isotropic (or-
ange) BAO measurements, where the latter is only displayed for data that does not exhibit an anisotropic measure-
ment due to low signal-to-noise. Lower panel: Isotropic BAO distance residuals where the anisotropic BAO mea-
surements from the upper panel (apart from DES-Y6 and SDSS-tr, which do not measure radial BAO) have been
combined into an isotropic measurement (not shown in the upper panel to avoid redundancy).

galaxy clustering analysis of DESI identified the galaxy-halo connection (HOD) as the dominant systematic for their
BAO measurement. However, all combined systematics still only added 5% to the final statistical error budget even
for the highest precision BAO measurement (see table 13 and 15 in Ref. [782]).
The BAO signal is most commonly measured as a mixture of radial and transversal modes. Alternatively, the BAO
signature can be extracted in a thin redshift bin using the two-point angular-clustering statistics (where the angular
separation of pairs instead of the comoving distances are computed, avoiding converting redshifts into distances using a
fiducial cosmology), providing a measurement of the transverse BAO, ∆θBAO (z). The finer the redshift bin, the purer
this transverse contribution. If rd is known, ∆θBAO (z) determines DM (z), as explained in Sec. 2.1.19.a. Fig. 30 shows
such transverse BAO measurements from SDSS data (labelled SDSS-tr) derived in Refs. [761] and [762, 784–787].
However, for the remainder of this section, we will focus on the most recent DESI analysis.
2.1.19.c. BAO and the H0 tension One of the challenges in the use of BAO measurements in cosmology is the
degeneracy between the sound horizon rd and Hubble constant H0 . As mentioned in Sec. 2.1.19.a, this arises because
all measurable quantities in the BAO data depend on the product c/H0 rd . The only way to decouple them is by using
additional information beyond the BAO scale [712, 717, 788–795] or imposing an external prior on rd from the CMB
or BBN [698, 712, 796]. The 2024 DESI BAO results combined with a BBN prior from Ref. [797] yield a constraint
of H0 = 68.53 ± 0.80 km s−1 Mpc−1 [782]. This measurement is independent of the CMB and solely based on DESI
BAO + BBN. While this value of the Hubble constant is slightly higher than the one preferred by Planck [192] as well
as the one reported in previous data from SDSS [796], it is still in 3.4σ tension with the SH0ES result of Ref. [78].
Fig. 31 shows the values for H0 and Ωm,0 inferred from the DESI data, along with the SH0ES results on H0 as a
shaded band. One sees that only the w0 CDM model comes close to resolving the Hubble tension for the DESI+CMB
65

DESI+CMB+DESY5
DESI+CMB+Union3
DESI+CMB+Panth. w0waCDM+ k
DESI+CMB
DESI+BBN+
DESI+CMB+DESY5
DESI+CMB+Union3
DESI+CMB+Panth w0waCDM
DESI+CMB
DESI+BBN+
DESI+CMB+DESY5
DESI+CMB+Union3
DESI+CMB+Panth. w0CDM
DESI+CMB
DESI+BBN+
kCDM
DESI+CMB
DESI+BBN+
DESI+CMB
DESI+BBN+
DESI+BBN
SDSS IV + BBN CDM
CMB
62 64 66 68 70 72 74 0.28 0.30 0.32 0.34 0.36 0.38
H0(kms 1Mpc 1) m

FIG. 31: 68% credible-interval constraints on the Hubble constant H0 (left panel) and the matter density Ωm,0
(right panel), assuming different cosmological models and different combinations of datasets (on the plot, DESI,
Planck 2018, Pantheon Plus, DESY5, Union3) and the BBN and θ∗ priors. The grey band indicates the SH0ES
measurement for H0 .

dataset to the price of much lower matter density. A further look of the DESI results is presented in Fig. 32, where
we show the 1σ posteriors for some of the models presented in Fig. 31. In this figure, the degeneracy between H0
and Ωm,0 manifests as the covariance of the inferred data, which is particularly visible on the right panel of Fig. 32
zooming in on the DESI+BBN models. This leads to the known issue that, across different models, increasing H0
generally decreases Ωm,0 , and vice versa. On the plots, rd is missing since it is not a free parameter for all tested
models.
BAO measurements rely on a model-dependent extrapolation to redshift zero when constraining the Hubble con-
stant. However, combining BAO with SNIa datasets and the CMB does constrain the redshift evolution and makes
it difficult to find extensions to ΛCDM that could resolve the Hubble tension [238, 788, 798–801].
Ref. [802] has found that marginalising over rd can lead to a degeneracy between H0 and the primordial power
spectrum and to a model-dependent H0 measurement. A similar observation has been published in Ref. [803] showing
that models that only reduce rd cannot resolve the Hubble tension while at the same time remaining consistent with
other cosmological datasets. A possible solution to this consists of working with the sound horizon as a free parameter
(thus not making any assumption concerning the recombination physics) and using other quantities to break the
degeneracy [804–806]. In these works, it was shown that while this approach can produce interesting results when
examining ΛCDM alternatives, it does not solve the Hubble tension. Another approach consists in marginalizing
over both rd and H0 [807], which integrates the quantity c/H0 rd out of the likelihood (χ2 ) and thus removes entirely
the dependence on these parameters from the theoretical predictions. The results obtained in this way show a slight
preference for Dynamical Dark Energy (DDE) models but they do not solve the Hubble tension.
While early BOSS and eBOSS Ly-α BAO results showed a 1.5 to 2.5σ tension with Planck-ΛCDM [788, 796], the
most recent Ly-α analysis in DESI [744] shows no tension anymore. However, DE density reconstructions using DESI
BAO still suggest the possibility of zero or negative DE densities for z ≳ 1.5 − 2 in some dataset combinations [808].
Additionally, a shift of the Ly-α BAO peak at the 2.2σ and 3.5σ levels in real and redshift space, respectively, was
recently suggested in Ref. [809]. The reasons behind these findings are yet to be understood, warranting further
investigation. Since the comoving angular diameter distance to last scattering, DM (z∗ ), is strictly constrained by
66

0.33
ΛCDM CMB (ωm prior)
ΛCDM + ΩK SDSS+BBN
0.40
wCDM DESI+BBN
w0 wa CDM
0.32 DESI+BBN+θ∗
w0 waCDM + ΩK DESI+CMB+SN
0.35

Ωm
Ωm

0.31

0.30

0.30
0.25

0.29
55 60 65 70 75 80 66 67 68 69 70
−1 −1 −1 −1
H0 [km s Mpc ] H0 [km s Mpc ]
FIG. 32: Illustrations of the degeneracy between the Hubble constant and the matter density parameter (Ωm,0 )
constrained by different data combinations (indicated by line style) for different models (indicated by colour) at
1σ. “CMB” refers to Planck + ACT lensing data and “SN” to the Pantheon+ SNIa sample. The parameter region
constrained by CMB alone is shown as a prior on ωm (black dotted contours) that is independent of the late-time
model extensions considered here. The most constraining data combination (DESI+CMB+SN) is shown on the
right as a zoom into the black rectangle on the left for better visibility. For conciseness, the SDSS BAO + BBN
constraint (dash-dot-dotted) is shown for ΛCDM (orange) only. Note that the legends shown in the left and right
panels apply to both simultaneously.

CMB data almost model-independently, a lower H(z) for z ≳ 1.5 − 2 due to vanishing or negative DE density
should be compensated by a higher H(z) for z ≲ 1.5 − 2, provided that the pre-recombination Universe is not
modified [800, 801, 810–812]. This results in a higher H0 and correspondingly smaller Ωm , which can reduce the S8
value and address the most discussed tensions in the ΛCDM model [800, 801, 810–812].
The most recent DESI data release, DESI DR2, improves on the DR1 findings, analyzing over 14 million galaxies
and RGBs from three years of observations (surpassing SDSS in the effective volume of all tracers). Compared to
DR1, the precision of BAO measurements has increased by approximately a factor of two, with uncertainties reduced
to about 0.24% for galaxy/RGB measurements and 0.65% at high redshift from the Lyman-α forest (which is now
twice the DR1 sample). The DESI DR2 combined with a BBN prior yield H0 = 68.51 ± 0.58 km s−1 Mpc−1 - 28% more
precise than DR1. The cosmological analysis shows stronger evidence for evolving DE, with statistical significance
amounting to 2.8σ, 3.8σ and 4.2σ depending on which datasets are combined with DESI (Pantheon +, Union3 or
DES Y5) and 3.1σ for DESI+CMB. The results show a preference for a DE equation of state with w0 > −1 and
wa < 0 and a possible phantom crossing in the past. While the degeneracy direction for the constraints in the w0 − wa
plane approximately points towards the ΛCDM solution, the statistical evidence consistently challenges the standard
ΛCDM across multiple analysis approaches and dataset combinations [716].

2.1.19.d. Outlook The DESI experiment is expected to publish its Y3 BAO analysis in 2025, while the final Y5
analysis should be made public in 2026/27 [813]. The final dataset will be three times larger than the current Y1
catalogue, and proposals for a DESI extension until 2029 and a second phase (DESI-II) well into the 2030s are under
consideration. At the same time, the Euclid mission has also started, with the first cosmology analysis expected in
2025/26 [164], and the Vera Rubin Observatory will see first light in 2025.
On a longer timescale, there have been several proposals for stage-V spectroscopic experiments with a start in the
early to mid-2030s. These experiments increase the telescope aperture and the spectroscopic multiplexing capabilities
with the aim of collecting hundreds of millions of galaxy redshifts (e.g., see Refs. [407, 814–816]). After the DESI
mission is completed, the BAO signal in the low redshift Universe (z ≲ 1) will have been measured up to the sample
variance limit over almost the entire sky available for cosmological measurements (∼ 15 000 to 18 000 deg2 ), see figure
16 in Ref. [817]. Future experiments therefore focus on the high redshift regime, where significant gains are still
possible.
67

2.2. S8 Tension: measurements and systematics


2.2.1. Weak lensing
Coordinator: Marika Asgari, Daniela Grandón, Cora Uhlemann
Contributors: David Sanchez Cid, Ignacio Sevilla, Judit Prat, Konrad Kuijken, Maciej Bilicki, Markus Michael
Rau, Masahiro Takada, Nikolina Šarčević, and Shun-Sheng Li

Cosmic shear is one of the key probes of the LSS. A cosmic shear analysis measures the weak gravitational lensing
signal [818–823] imprinted in the observed galaxy shapes by LSS (see Ref. [824–826] for a review and references
therein). The strength of this signal depends directly on the total matter distribution between the source galaxies
and the observer. Therefore, cosmic shear provides a precise mapping of the projected matter density field, capturing
its statistical properties and enabling constraints on cosmological parameters.
To perform a cosmic shear analysis, we need a galaxy catalogue that comprises two measurements: galaxy shapes
(Sec. 2.2.1.a) and the redshift information (Sec. 2.2.1.b). The redshifts are used to divide source galaxies into tomo-
graphic bins. The redshift distribution in each bin determines the strength of the lensing effect and serves as an input
for the theoretical predictions of the cosmic shear signal, which require prescriptions for baryonic effects (Sec. 2.2.1.c)
and the intrinsic alignment of galaxies (Sec. 2.2.1.d). In addition, the tomographic analysis allows us to extract
information about the evolution of structures and to disentangle the astrophysical systematics from the cosmological
information.
The most commonly used estimator for the cosmic shear signal is the two-point statistic. In real space, the two-point
shear correlation functions are measured as the average correlation between the shapes of galaxy pairs in parallel and
orthogonal directions as a function of their angular separation [827]. Both of those correlations can be related to
one angular convergence power spectrum, which is a line-of-sight projection of the matter power spectrum, and thus
sensitive to nonlinear effects on small scales and low redshifts. However, because the lensing field is a non-Gaussian
random field, cosmic shear surveys have also implemented estimators that capture information beyond the traditional
two-point statistics [828]. These higher-order or non-Gaussian statistics, such as the three-point shear correlation
function [829–832], the bispectrum [833, 834], the lensing convergence PDF [835–838] or its moments [839], and peak
statistics [840–843], among others [844–847], have proven effective in enhancing the constraining power and promise
to be valuable in Stage IV analysis [828, 848].
Cosmic shear analysis is most sensitive to the S8 parameter and the strength of the intrinsic alignment signal
of galaxies. In fact, with the available data, S8 is the only cosmological parameter that cosmic shear analysis can
robustly constrain from two-point statistics. A joint analysis of two-point function and higher-order statistics however
can improve constraints in Ωm,0 , as well as self-calibrate systematics and break degeneracies between cosmological
and nuisance parameters.
The current generation of Stage III cosmic shear surveys—such as the KiDS ([849, 850]), DES ([851, 852]), the Sub-
aru HSC Survey ([853–855]) and the Dark Energy Camera All Data Everywhere (DECADE; [856]) —have conducted
multiple cosmic shear analyses from at least two separate data releases. These surveys consistently infer S8 values
that are 2 − 3σ lower than what is expected from the CMB anisotropies and CMB lensing measurements by Planck
and ACT [854, 855, 857–863]. A summary of these results is presented in Fig. 33. This disagreement in the S8 value
between early- and late-time probes, often referred to as the S8 tension, was first observed between the tomographic
analysis of the Canada France Hawaii Telescope Lensing Survey (CFHTLenS) [864] and the first Planck cosmology
results [865], at the level of 2.3σ. More recently, the DES and KiDS teams collaborated on a unified analysis pipeline
which resulted in a joint constraint with a slightly higher value of S8 , translating to a 1.7σ difference with Planck
[866]. The most recent KiDS results (KiDS-Legacy) shows consistent results with Planck at the level of 1σ.
The “3×2pt” analysis has emerged in the past decade as a mature cosmological probe that extends cosmic shear
by incorporating two additional two-point correlations: galaxy-galaxy lensing (the cross-correlation of lens galaxy
positions and source galaxy shapes) and photometric galaxy clustering (the autocorrelation of lens galaxy positions).
The primary advantage of this combination is its ability to break the degeneracy between Ωm,0 and σ8 , enabling robust
constraints on Ωm,0 in addition to S8 , as well as improving the self-calibration of the nuisance parameters, especially
of the intrinsic alignment and redshift parameters. The three major current-generation surveys have conducted this
kind of analysis, with their results presented in Fig. 33. Generally, the S8 values measured by 3×2pt have been found
to be ∼ 1 − 2σ below the Planck value predicted assuming ΛCDM cosmology, thus with a slightly smaller tension
with respect to the cosmic shear results.
The constraints on S8 and other cosmological parameters can also be affected by the adopted statistical inference
pipeline. Most cosmic shear analyses use a Gaussian likelihood, which is defined using the data, the theoretical
predictions, and the data covariance matrix (for exceptions employing simulation-based inference (SBI), e.g., see
Ref. [875–877]). Studies have shown that an accurate estimation of the covariance matrix is crucial to avoid biases in
cosmological constraints [878, 879]. Additionally, properly accounting for likelihood asymmetry in two-point functions
68

Planck TT,TE,EE+lowE
ACT CMB lensing+BAO
SPT CMB lensing+BAO

HSC Y3 C
DES Y3 C
KiDS-1000 C

HSC Y3 ±
DES Y3 ±
KiDS-1000 ±
DES Y3+KiDS-1000 ±

HSC Y3 3x2
DES Y3 3x2
KiDS-1000 3x2

0.70 0.75 0.80 0.85 0.90 0.95


S8 = 8( m/0.3)0.5
FIG. 33: The S8 constraints derived from the Stage III WL surveys DES Y3 [860, 861, 867, 868], HSC Y3 [863,
869–871] and KiDS-1000 [859, 872, 873] from two-point shear statistics in harmonic space using Cl or in angular
space using ξ± , and 3x2-point statistics including galaxy clustering. Also shown are the joint constraint from DES
Y3+KiDS-1000 [866]. We include the CMB measurements from the Planck satellite [693], ACT [694] and SPT [874]
on top for comparison. The error bars denote 1-σ uncertainties.

is essential for robust parameter inference [880].


In this section, we focus on weak gravitational lensing and the associated systematics that influence cosmic shear
cosmological constraints. These systematics can be broadly categorized as observational or theoretical. Observational
systematics affect either the accuracy of shape measurements or the redshift distributions. On the theoretical side,
achieving percent-level precision in predicting the matter power spectrum is required across a wide range of scales [881],
determined by the survey depth and analysis choices. Since these scales extend into the nonlinear regime, an accurate
nonlinear power spectrum prescription is necessary. Additionally, two key astrophysical systematics—intrinsic galaxy
alignments and baryonic feedback—remain among the least understood aspects of theoretical modeling for this signal.
Various modeling approaches have been developed to address these issues, which can generally be categorized into
emulators based on N -body simulations [882, 883], and the more widely used halo model-based approaches, see
Ref. [884] for a recent review, such as halofit [885, 886] and HMcode [887, 888]. While emulators may offer greater
accuracy within the specific parameter range they are trained on, they lack the flexibility of halo model approaches,
which are capable of extrapolating beyond this range. As a result, most cosmic shear analyses to date, including all
of the flagship publications by current surveys have relied on the latter. To limit the contributions from small scales,
69

suitable linear combinations of the signal can achieve a nulling of the effective WL kernel [889]. This can be used to
perform scale cuts in physical rather than angular space and render predictions more robust to small-scale physics,
e.g., see Ref. [890, 891]. While this discussion focuses on current cosmic shear surveys, the next generation of cosmic
shear observations (Stage IV surveys) is approaching [437, 892, 893]. As constraints tighten, more precise modeling
and calibration will be required. Consequently, the interplay of these systematic effects will become increasingly
important in future analyses.
2.2.1.a. Observational effects: Shape measurement and calibration Extracting WL signals from observed galaxy
images is a non-trivial task. The cosmic shear signal is typically orders of magnitude smaller than the intrinsic galaxy
ellipticities, necessitating robust statistical measurements of a large number of galaxies. Furthermore, systematic
biases introduced by various instrumental and measurement effects further complicate the process. While the statis-
tical limitations are readily overcome with the ever-growing galaxy samples in ongoing and upcoming WL surveys;
addressing the systematic biases remains a challenging task. This challenge is further exacerbated by increasingly
stringent requirements driven by the growing statistical power of these surveys. For upcoming Stage IV surveys, the
overall residual shear biases must be reduced to the order of 10−3 , compared to the current percent-level requirements
of Stage III surveys, e.g., see Ref. [894]. Achieving this stringent requirement demands continuous development of
shear measurement and calibration techniques, with particular attention to subtle effects arising from object detection,
blending, selection, and redshift estimation.
Shape measurement and data-based calibration Given the critical importance of accurate shear measure-
ment, the WL community has dedicated significant early efforts to developing robust shear measurement algorithms.
A number of methods have now achieved an overall bias at the percent level. Broadly, these shear measurement
methods can be categorized into two main classes: moment-based methods, e.g., see Ref. [895] and model-fitting
methods, e.g., see Ref. [896, 897]. Moment-based methods estimate galaxy ellipticities from the second moments of
observed galaxy images after correcting for the point-spread function (PSF) caused by instrumental and observa-
tional effects. Conversely, model-fitting methods employ forward modelling with PSF-convolved parametric galaxy
profiles. These techniques are susceptible to specific biases. Moment-based methods are affected by missing pixels,
and contamination by light from detected and undetected neighbouring galaxies, whereas model-fitting approaches
suffer from “model bias”. Additionally, both are affected by common biases such as “noise bias” and selection effects,
see Ref. [898] for a recent review. These biases need to be calibrated and corrected for shear measurement methods
to meet the requirement of modern WL surveys.
A recent significant advancement in shear measurement algorithms is the introduction of self-calibration or meta-
calibration procedures that serve as a first-order correction to the initial shear measurement, e.g., see Ref. [899, 900].
These techniques, which can rely on model profiles, observed galaxy images, or priors from deep observations, have
demonstrated the ability to control the residual shear bias to a percent or sub-percent level, nearly meeting the
requirements of current Stage III surveys, e.g., see Ref. [901–903]. However, these data-based calibrations cannot
correct biases introduced at the detection stage as required for the accuracy of Stage IV surveys, e.g., see Ref. [904].
Detection biases and simulation-based calibration Detection biases arise primarily from blending, where
the light of neighbouring galaxies overlaps in the image plane. Pixel noise and PSF convolution further exacerbate
these effects, introducing shear-dependent detection biases, e.g., see Ref. [905–908]. Because these biases arise before
shear measurement and are already imprinted in the data, they are difficult to calibrate using observations alone.
Meta-calibration methods attempt to correct detection biases by introducing an additional detection step, known as
metadetection, but this becomes challenging when accounting for redshift estimation [909]. Furthermore, the shear
of galaxies varies with their environment and redshift, resulting in redshift-shear interplay effects that cannot be
calibrated with observational data where the true shear is unknown, e.g., see Ref. [901, 903]. Thus, simulation-
based calibration is essential for shear measurement methods to meet the stringent requirements of Stage IV surveys.
This approach originated from a series of community-wide blind challenges, where mock images with realistic galaxy
properties and real data features were used to assess shear measurement algorithms [910–914]. With simulated images,
where the ground truth is known, shear biases can be directly measured for any algorithm.
If the simulated images accurately replicate a survey’s properties (e.g., resolution, PSF, signal-to-noise SNR, size
and ellipticity distributions, source density, and clustering), the measured shear biases can be used to calibrate residual
biases in real observations. This approach has been adopted by all Stage III surveys, e.g., see Ref. [901–903] and will
remain vital for the Stage IV surveys.
Impact on S8 tension and future direction Decades of dedicated work on shear measurement and calibration
have significantly improved the accuracy of shear measurement. The latest KiDS cosmic shear analysis demonstrates
that residual shear biases, primarily stemming from uncertainties in galaxy profiles, introduce only sub-percent uncer-
tainties in S8 estimates [915]. Shear measurement uncertainties are thus unlikely to be the main driver of the current
S8 tension.
70

Since shear measurement is closely tied to redshift estimation in real observations, developing a consistent, joint
calibration of shear and redshift estimates using multi-band image simulations remains essential and will be a key
focus for future simulation-based calibration, e.g., see Ref. [901, 903]. Furthermore, due to the inherent limitations of
image simulations—especially in capturing galaxy profile details—advancing shear measurement algorithms to reduce
sensitivity to these intricacies remains a crucial goal.
2.2.1.b. Observational effects: Redshift Measurement and Calibration WL measurements rely on two-dimensional
(projected) quantities, meaning individual source redshifts are not strictly required. However, redshift estimates are
essential for cosmic shear tomography, which enhances the signal and probes the time evolution of the growth of
structure. Historically, the requirements for photometric redshifts of individual lensing sources have been relatively
lenient, but upcoming surveys such as Euclid [164] and LSST [893] demand higher precision, e.g., see Ref. [916] for a
review.
Photometric redshift estimation Photometric redshift estimation is crucial for extracting cosmological in-
formation in optical surveys like LSST and Euclid. In contrast to spectroscopic redshift measurement, photometric
redshift inference uses image, or photometric information. Broad band optical surveys like LSST and Euclid take
images in 100 − 200 nanometer wide optical filter bands, which does not allow the recovery of atomic lines. This
limitation means that redshift estimation must rely on the overall shape and characteristic features of a galaxy’s
spectral energy distribution (SED), for a review see Ref. [916, 917]. Two primary approaches exist: ML and template
fitting. ML methods [918–924] construct a direct mapping from observed photometry of individual galaxies to their
redshift based on training data. While this training data does not need to spatially overlap, it has to be representative
and complete in the color-redshift mapping being learned. Template fitting methods, e.g., see Ref. [925–931] match
observed photometry to pre-existing SED models. It relies on accurate SED modelling and requires validation using
similar calibration datasets. Although complementary, both techniques face challenges such as epistemic uncertainty
in the SED and selection function models and incompleteness of accurate reference data at the faint end of the
color-magnitude space, e.g., see Ref. [932–936].
Redshift distribution calibration for cosmic shear Beyond individual redshift estimates for cosmic shear
sources, accurately calibrating their redshift distributions is crucial for interpreting observed signals with underlying
cosmological models. For a given set of sources – usually selected in tomographic redshift bins – it is essential to
recover their underlying true redshift distribution as closely as possible [937]. This is necessary to relate the observed
signal to theoretical predictions, as the the redshift distribution of sources determines the observed shape correlations
via the matter power spectrum. Calibrating the redshift distribution in current WL surveys is challenging due to the
faint nature of most cosmic shear sources and the limited availability of spectroscopic measurements. Consequently, a
range of calibration methods have been developed, and efforts are ongoing to extend spectroscopic samples to better
match the needs of cosmic shear surveys [933, 938].
One of the earliest approaches for redshift distribution calibration, employed by e.g., the Deep Lens Survey (DLS)
[939] or CFHTLenS [940], involved stacking photo-z posterior PDFs, derived with Bayesian Photometric Redshifts
(BPZ) [941], to estimate the population distribution dN/dz. As Stage III surveys increased in statistical power,
more reliable calibration methods were developed. These methods generally rely on either mapping the relation
between colors and redshifts, or on the cross-correlation (clustering) approach. The former employs spectroscopic
or narrow-filter multiband photometric calibration data to re-weight dN/dzphot , either directly in magnitude space
[857, 858, 942], by incorporating additional properties, e.g., see Ref. [943], or via self-organizing maps (SOM) [944].
Two key ingredients are required for these calibration methods: spectroscopic data that closely match the cosmic shear
sample (in magnitude, color and redshift range) and a sufficient number of multi-band filters to break degeneracies in
the color-redshift mapping. This is most readily achievable in KiDS, which covers nine bands and includes multiple
deep calibration fields [945, 946]. The DES redshift inference and calibration utilizes a mapping from deep multi-band
to wide four-band data that is combined with spatial cross-correlations [947–949]. This choice is motivated by studies
[936] that found that the photometric redshift biases from direct calibration strategies induced by the incompleteness
in spectroscopic calibration samples is unacceptably high for DES. Similar to DES, the HSC calibration [950] relies
on a combination of spectroscopic and narrow filter multiband calibration data as well spatial cross-correlation mea-
surements described in the following section. We note that the choice of photometric redshift calibration method and
the incorporation of systematic effects is dictated by the survey specification and cannot be generalized across survey
designs. However it can be concluded that the treatment of selection functions of calibration sources poses a major
challenge for DES, HSC and KiDS.
The cross-correlation method, also known as clustering redshifts, leverages the spatial correlation of galaxy positions
to estimate dN/dz for a photometric sample by measuring its angular cross-correlation with a reference sample that
has known redshifts, e.g., see Ref. [951, 952]. Either spectroscopic or high-fidelity photometric redshifts can serve
as reference samples in this method. A major limitation of the clustering redshift approach is the availability of
wide-angle calibration samples that extend to sufficiently high redshifts. As a result, its application in current cosmic
71

shear surveys has been restricted primarily to the calibration of low-redshift bins [945, 947, 950].
Impact on S8 tension and future direction Ref. [953] showed that an improved redshift calibration of the
full sample or calibrations using different subsets change the S8 obtained from the fiducial KiDS-1000 analysis at
most at the level of 0.5σ. For upcoming surveys, photometric redshift estimation based on ML methods could benefit
from training sample augmentation with simulated galaxies possessing otherwise unrepresented features [954]. The
complex selection function of spectroscopic surveys presents a major challenge in the calibration of photometric
redshifts, especially at the faint end of the color-magnitude space [936].
Computationally efficient approaches for the marginalisation over photometric redshift uncertainties are crucial to
tackle the challenge of running inference in high-dimensional parameter spaces. Strategies can involve simplified n(z)
parametrisations, such as in terms of shifts of a mean redshift ∆z per bin, or a resampling approach using multiple
MCMC chains at a fixed n(z) sampled from the uncertainties. Ref. [955] showed that for the forecasted precision
for HSC Y3, these methods recover statistically consistent error bars. However, when the constraining power of
the full HSC survey (or other surveys of comparable or greater statistical power) is considered, the choice of the
marginalisation method may modify the 1σ uncertainties on Ωm,0 − S8 constraints by a few percent.
2.2.1.c. Astrophysical effects: Baryonic feedback With the increasing precision of modern cosmological surveys,
small-scale astrophysical processes—commonly referred to as baryonic feedback—have become increasingly relevant
in WL analyses. Baryonic feedback, driven by galaxy formation, supernova explosions, and AGN, suppresses the total
matter power spectrum on scales of k ∼ 1 − 10 h Mpc−1 by redistributing gas within and beyond halos and influencing
through back-reaction [956]. Unlike gravity, which acts as a long-range force, baryonic effects follow an inside-out
pattern: halos, stars, galaxies, and central supermassive black holes form on small scales through hierarchical structure
formation, while supernova and AGN feedback modify the baryon distribution by, for example, expelling intrahalo gas
into the IGM [957–961]. In particular, hydrodynamical simulations [962–966] indicate that AGN feedback is one of the
most significant effects in suppressing the matter power spectrum at small scales. However, the spatial extent of the
AGN ejected gas remains an open question—whether it affects a few Mpc or extends to tens of Mpc [965, 967]. Below
this maximum scale, the total matter distribution remains unchanged due to conservation of mass and momentum.
On smaller scales of k > 10 h Mpc−1 , baryonic effects such as gas cooling and star formation become efficient, leading
to an upturn in the matter power spectrum [881]. Since these effects alter the total matter distribution, cosmic
shear summary statistics—such as two-point correlation functions — are significantly affected compared to a purely
collisionless scenario.
Mitigation strategies Despite the importance of baryonic feedback, modeling it from first principles remains
challenging due to the complex and nonlinear nature of these processes. This introduces significant theoretical
uncertainties in WL analyses. A common mitigation strategy is to apply scale cuts to the summary statistics,
excluding small scale data where hydrodynamical simulations suggest feedback effects dominate [859–861, 863, 869].
While this method preserves the robustness of cosmological constraints, it comes at the cost of discarding high SNR
scales that are rich in cosmological information, thereby limiting the constraining power of WL surveys. To fully
exploit the potential of upcoming Stage IV surveys, it is crucial to incorporate baryonic effects into cosmological
analyses rather than simply removing affected scales.
Several modeling strategies have been developed to address this issue by incorporating baryonic feedback into the
total matter power spectrum. One widely used approach is the baryon correction model (BCM), which modifies the
position of particles in N -body simulations using parametric prescriptions that approximate the impact of feedback
on the DM halo profiles [883, 958, 968, 972].
Another method integrates baryonic effects within the halo model framework by modifying density profiles and gas
expulsion mechanisms [887, 888, 973]. Some of these models and their impact on the matter power spectrum are
summarized in Fig. 34. Principal Component Analysis (PCA) and alternative basis function approaches provide a
more data-driven way to capture a broad range of feedback scenarios [974]. The most detailed recipe comes from
cosmological hydrodynamical simulations, which explicitly simulate baryonic physics using sub-grid prescriptions for
unresolved processes such as supernova and AGN feedback [881]. However, discrepancies arise due to differences in
sub-grid implementations, calibration strategies, and simulation methods. These variations lead to disagreements in
both the amplitude and scale dependence of the suppression of the total matter power spectrum [967]. Consequently,
an effective model for WL must be flexible enough to encompass the full range of plausible baryonic effects. Moreover,
other nonlinear effects—including intrinsic alignments, nonlinear clustering, and reduced shear corrections—must be
accounted for to ensure robust cosmological constraints. The aforementioned strategies provide a way to marginalize
over baryonic feedback uncertainties while preserving small-scale cosmological information. However, degeneracies
between baryonic feedback and cosmological parameters may degrade the constraints on S8 , highlighting the need of
estimators beyond the two-point functions that can disentangle these effects while maximizing the constraining power
of next-generation surveys.
Higher-order statistics of the WL field, such as bispectrum, peak counts [848, 975] and the one-point lensing PDF
72

1.15
1.10
1.05
1.00
0.95
F(k)

0.90 P(k)Schneider15
0.85
P(k)VanDaalen19
P(k)Mead20
0.80 P(k)Aricò21
0.75 P(k)Moran22
10 3 10 2 10 1 100 101
k [Mpc 1]

FIG. 34: Suppression of the total matter power spectrum on small scales due to baryonic effects for various baryon
models [883, 887, 956, 968, 969]. All spectra were computed for a flat ΛCDM cosmology with Ωc = 0.25, Ωb = 0.05,
Ωk = 0.0, σ8 = 0.81, ns = 0.96, and h = 0.67, assuming no massive neutrinos. The y-axis shows the ratio F (k) of
each model including modeling of baryons, to the nonlinear matter power spectrum without baryonic effects. The
nonlinear power spectrum was computed using pyccl.nonlin_matter_power method, while baryonic suppression
was modeled with the baryonic modules in pyCCL [970], all evaluated at a = 1 (z = 0) over the same k-range.
(Adapted from Ref. [971].)

[835], are more sensitive to non-Gaussian and smaller-scale structures than two-point correlations [846, 976–978]. Thus,
combining the two-point correlations with the higher-order statistics would allow us to obtain tighter constraints on
cosmological parameters [841, 845], while self-calibrating the baryonic effects [979].
Finally, multi-probe approaches—combining WL with X-ray, measurements of the Sunyaev–Zeldovich (SZ) effect
(thermal tSZ and kinetic kSZ), and FRB observations—offer complementary constraints on the distribution of baryons
in large-scale structures [980–984]. These techniques collectively help preserve cosmological information while mini-
mizing biases from astrophysical processes.
Impact on cosmic tension and S8 To explore the impact of baryons on the inferred S8 , Stage III surveys
have conducted parameter inference with two-point functions down to small scales. They found that moderate to
small baryonic effects are present in the data [972, 985–987]. In particular, Ref. [985] performed a joint cosmic
shear analysis of the Stage III surveys DES-Y3, KiDS-1000 and HSC-DR1 using small scale data in harmonic space.
This reanalysis implements a single pipeline that extends the cosmic shear angular power spectrum to scales up to
ℓ < 4500, using BACCOemu to model the non-linear regime and baryonic effects in the total matter power spectrum.
Their resulting S8 constraint is 1.8σ lower than Planck, however [985] found a Ωm,0 tension between both data sets
of 3.5σ. When analyzing the parameter space in terms of variables without an implicit dependence on the Hubble
constant H0 , (S12 , wm ), the authors find no tension with Planck. On the other hand, Ref. [986] explore signatures
of baryonic effects in the HSC Y3 two-point correlation functions ξ± down to small scales of 0.28 arcminutes (up to
k ∼ 20h Mpc−1 ). The theoretical modeling is implemented by means of a non-linear DM-only matter power spectrum
emulator (similar to DarkEmulator [988]). The authors found no significant shift in the inferred S8 when including
the small scales. Moreover, the DM-only theory fits the observations within the statistical error of the survey, meaning
baryons are not playing an important role in these measurements. Similar results were obtained for HSC Y1 data
using multiple higher-order statistics at small scales [846]. These studies of cosmic shear data, together with other
analyses from the Sunyaev-Zel’dovich observations, indicate that strong feedback scenarios are needed to reconcile
the S8 inferred from cosmic shear surveys with the value derived from Planck ΛCDM [989]. This translates into
a large suppression of the matter power spectrum, of the order of ∼25% at k ∼ 1h Mpc−1 . The latter, however,
is larger than the suppression found from hydrodynamical simulations, and contradicts the constraints imposed by
73

X-ray observations of the baryon mass fraction for cluster-scale halos [986, 987]. Therefore, it is possible that baryons
in combination with other non-linear effects and unmodeled systematics may explain the discrepancy between cosmic
shear results and Planck ΛCDM cosmology, or extensions of ΛCDM and new physics are required.
2.2.1.d. Astrophysical effects: Intrinsic alignments of galaxies Intrinsic alignments (IA) of galaxies represent a
significant astrophysical effect that can systematically affect the measurement of cosmic shear and, consequently,
the inference of cosmological parameters (see Ref. [990] for a comprehensive review). These alignments arise due to
gravitational interactions and the tidal forces exerted by the large-scale structure of the Universe, causing the shapes
of nearby galaxies to correlate with each other and with the surrounding matter density field.
For cosmic shear two-point correlation functions, intrinsic alignments can be categorized primarily into two types:
intrinsic-intrinsic (II) alignments and gravitational-intrinsic (GI) and intrinsic-gravitational (IG) alignments. II align-
ments occur when the shapes of physically close galaxies are directly influenced by the same gravitational tidal field,
leading them to align with each other. GI and IG alignments describe the correlation between intrinsic galaxy shapes
and the shear induced by gravitational lensing. GI refers to the correlation between the intrinsic shape of a fore-
ground galaxy and the shear of a background galaxy, while IG represents the reverse case [991]. These terms are
related through symmetry properties of the shear field, with IG often being treated as the transpose of GI in theoretical
and numerical treatments.
Galaxy-galaxy lensing (GGL) probe, or the correlation between galaxy shapes and galaxy positions, is also sensitive
to intrinsic alignments when the shape and the position catalog overlap in redshift. GGL is sensitive only to GI types
of alignments and through a different kernel than for cosmic shear (see equations in e.g., Ref. [992]). Because of this,
3×2pt analyses significantly help to constrain the intrinsic alignment parameters. Recently, GGL shear-ratios have
been used to fold in prior IA information to the cosmic shear analysis [860, 861, 993]. Recently, inverse galaxy-galaxy
lensing has also been proposed as an additional probe to constrain IA parameters [994].
Modeling intrinsic alignments Intrinsic alignment (IA) models vary in complexity and scale-dependent effec-
tiveness. The Nonlinear Alignment Model (NLA) [995] has been the standard approach to model red (elliptical) galax-
ies. This model assumes a linear coupling between the nonlinear tidal field and galaxy shapes, typically parametrized
by an amplitude parameter alongside an optional redshift evolution parameter. In this framework, blue (spiral) galax-
ies are assumed to have negligible intrinsic alignments. However, the linear tidal shear assumption breaks down on
small scales, necessitating more sophisticated models.
The Tidal Alignment and Tidal Torquing (TATT) model [996] represents an important extension, introducing
additional terms to capture the alignments of spiral galaxies where angular momentum plays a critical role. This
model incorporates a separate amplitude parameter for the tidal torquing effect, along with its own redshift scaling
parameter. In total, the basic TATT model typically employs four parameters: A1 (tidal alignment amplitude), A2
(tidal torquing amplitude), η1 (alignment redshift scaling), and η2 (torquing redshift scaling). A fifth parameter, bTA
(the linear bias of source galaxies), is frequently included to account for the clustered distribution of source galaxies in
observational data. Many studies such as Refs. [997–999] demonstrate that while NLA fits measurements well above
6−8 Mpc/h, TATT extends this range down to 1−2 Mpc/h. These findings emerge from both simulation-based tests
(N -body with semi-analytic IA components or hydrodynamic simulations) and direct fits to observational data.
For modeling at even smaller scales, the halo model describes alignments using one-halo and two-halo terms, based
on galaxy correlations within and between DM halos [990, 1000] while the Effective Field Theory (EFT) framework
treats alignments as a tensor field and incorporates small-scale physics through free parameters. Ref. [1001] shows
that EFT describes DM halo intrinsic alignments up to kmax = 0.30 h/Mpc at z = 0, significantly outperforming both
NLA and TATT which they find to only reach kmax = 0.05 h/Mpc.
Despite these advances in model complexity, practical applications raise important questions about necessary so-
phistication. The DES Y3 cosmic shear analysis [860] found that NLA was sufficient to fit their data, even though
TATT had been chosen for fiducial results in a blinded fashion. Similarly, Ref. [997] found no significant evidence for
non-zero tidal torquing amplitude (A2 ) in Illustris TNG hydrodynamic simulations, challenging the need for TATT’s
full 5-parameter framework in typical applications. Moreover, it is important to note that increasing model complexity
does not necessarily reduce biases. The interplay between IA and other uncertainties (e.g., from photometric redshift
errors [1002]) can introduce additional complications. Recent work [1003] shows that jointly modeling IA and source
redshift distributions via a shared LF can help mitigate such biases while maintaining consistency in WL analyses.
Careful calibration and validation remain crucial to ensure these models accurately represent physical processes with-
out introducing unintended distortions in cosmological analyses. To navigate these modeling challenges, researchers
have developed strategic approaches for IA model selection. Ref. [1004] recently introduced a data-driven methodol-
ogy to objectively determine which IA model best suits a particular analysis, providing a more systematic framework
for model choice. In parallel, many cosmological studies have adopted a multi-model approach, presenting results
across several IA formulations to demonstrate robustness and identify potential model-dependent biases. Below, we
examine how these modeling decisions have shaped recent cosmological parameter constraints and what they reveal
74

about the relative importance of IA modeling in contemporary WL analyses.


Modeling IA for higher order statistics While IA modeling has primarily focused on two-point statistics, recent
efforts have extended these frameworks to beyond 2-point statistics. Analytical advances have provided foundations
for modeling IA in bispectra [1005, 1006] and the lensing PDF [1007], while Ref. [1008] demonstrated that high
signal-to-noise peaks in aperture mass maps can experience deviations up to 30% for Euclid -like surveys due to IA
effects. Many recent analyses employ simulation-based inference, which relies on map-level IA modeling and presents
unique challenges. For example, Ref. [1009] showed that source clustering impacts non-Gaussian statistics significantly
more than two-point functions [1010]. Current approaches address this either by implementing additional clustering
terms in analytical predictions [1007] or by directly incorporating clustering effects in simulations [875]. The latter
approach extends beyond traditional NLA implementation by using simulation-based clustering rather than tree-
level perturbation theory, showing that modeling IA at the map level can also offer advantages for various summary
statistics.
Impact on cosmic tension and S8 Misestimations in intrinsic alignments can lead to significant biases in the
structure growth parameter S8 , thereby affecting our understanding of cosmic tensions. Recent cosmic shear analyses,
as illustrated in Fig. 35, demonstrate that among various analysis choices, the selection of IA model often produces
shifts in the S8 constraints, which will become substantial for Stage IV surveys [1011]. The TATT model (pink
points) tends to yield somewhat lower S8 values compared to the NLA model (blue points) across different analysis
combinations. This difference, while modest, may be relevant when considering potential tensions between early and
late-Universe probes, as indicated by the comparison with the reference line in the figure. The largest change comes
from the joint analysis of data from the DES and KiDS, which revealed that transitioning from the NLA model to
the more complex TATT model can lead to a heightened tension in S8 measurements. Quantitatively, the difference
between the redshift-dependent NLA and TATT analysis results was 0.9σ in the S8 marginalized posterior, but
reached 1.3σ when comparing the maximum a posteriori (MAP) values in the full parameter space. This distinction
is particularly significant as it indicates that the discrepancy arises from genuine differences in model physics rather
than merely from differences in prior volumes or projection effects. These results highlight the importance of carefully
considering IA modeling choices when interpreting cosmological constraints from WL surveys. 3x2pt analyses in DES
Y3 were less susceptible to the IA model compared to statistics based on cosmic shear alone.
Future directions As WL analyses continue to advance, refining IA models and developing more realistic cosmo-
logical simulations will be essential for providing robust validation tests. These improvements will work synergistically
with upcoming surveys such as LSST by the Vera C. Rubin Observatory, which will deliver high-quality imaging across
unprecedented sky areas, enabling studies of IA with greater statistical power and tightening constraints on cosmic
shear measurements.
Exciting new approaches are also already emerging from recent observations. Measurements have consistently
found IA amplitudes lower than theoretical expectations, with blue galaxies showing negligible alignment signals
[1012]. Building on these findings, [1013] proposed an innovative strategy of conducting cosmic shear analysis exclu-
sively with blue galaxies – potentially eliminating the need for complex IA modeling entirely. While this approach
sacrifices approximately half the galaxy sample, reducing statistical power, it could significantly mitigate systematic
uncertainties. The viability of this or other compromises will become clearer as theoretical understanding advances
and next-generation survey data becomes available.
2.2.1.e. Confirmation bias and blinding Modern cosmological analyses in surveys like DES, KiDS, and HSC are
often highly complex and require the seamless interaction of large teams responsible for different parts of the analysis.
Sources of epistemic error are difficult to control, which can lead to a dependence of cosmological parameter constraints
on analyses choices. At the same time, complementary cosmological constraints like from the CMB might induce a
subconscious bias in a select parameter subspace, like σ8 and Ωm,0 . In cases where the value of the cosmological
parameters are revealed during the analysis this might lead to ‘confirmation bias’, where analysis assumptions are
subconsciously tuned towards a certain scientific narrative. The complexity of cosmological analysis combined with
the significant dependency on analysis choices, therefore necessitates strict strategies to avoid this confirmation bias.
The blinding strategies differ across surveys but generally consist of a catalog-level blinding where a set of synthetic
and encrypted multiplicative shear values are artificially added to obfuscate the true shear field and an analysis-level
blinding where the inferred parameter posteriors are altered to avoid a readily comparison with prior work [1014].

2.2.2. Galaxy cluster counts


Coordinator: Dominique Eckert
Contributors: Antonio da Silva, Filippo Bouchè, Francesco Pace, Iryna Vavilova, Jenny G. Sorce, Massimiliano
Romanello, Radosław Wojtak, Sebastian Grandis, Shahab Joudaki, and Vittorio Ghirardini
75

DES Y3 + KiDS (Cosmic Shear) TATT

DES Y3 + KiDS (Cosmic Shear) NLA

HSC Y3 (Cosmic Shear) TATT

HSC Y3 (Cosmic Shear) NLA

DES Y3 (Cosmic Shear + SR) TATT

DES Y3 (Cosmic Shear + SR) NLA

DES Y3 (3×2pt) TATT

DES Y3 (3×2pt) NLA

0.750 0.775 0.800 0.825 0.850


S8 = σ8 (Ωm /0.3)0.5

FIG. 35: Impact of IA model choice to S8 in recent WL analyses assuming the ΛCDM model. The DES Y3 +
KiDS results are from Ref. [866], HSC Y3 from Ref. [863], DES Y3 cosmic shear (COSEBIS and ξ± )+ SR (shear-
ratio) from Ref. [860, 861] and DES Y3 3×2pt from Ref. [867]. The vertical dashed line is the Planck value with its
corresponding uncertainty shaded in orange. All error bars correspond to 1-σ uncertainties.

2.2.2.a. Methodology Galaxy clusters are the most massive gravitationally bound structures in the present Uni-
verse and originate from the largest fluctuations of the primordial matter distribution. Their mass and number
density strongly depends on the underlying cosmological parameters [1015], such that galaxy cluster surveys are
powerful probes of cosmological parameters. In this section, we describe the main steps that are needed to extract
cosmological information from galaxy cluster surveys. We identify the major bottlenecks that need to be addressed
to advance our cosmological knowledge with the use of this technique, and summarize the main recent results in the
field.
The halo mass function and its cosmological dependency In the hierarchical structure formation scenario,
small structures formed at high redshifts progressively merge under the influence of gravity to form the massive
systems we see today. As such, the number density of halos and its time evolution trace the growth of structures in
the Universe, which depends on the underlying cosmological model. The halo mass function describes the number of
collapsed halos at a given time per unit volume and halo mass. Its shape, normalization, and redshift evolution are
highly sensitive to cosmological parameters.
In mathematical terms, the halo mass function can be expressed in the following way:

dn ρm,0 d ln σ
= f (σ) , (2.12)
d ln M M d ln M

with ρm,0 the mean matter density of the Universe at present day, σ the mass variance on a mass scale M and f (σ)
a function characterizing different fitting (or theoretical) expressions. Often, f (σ) is written in as νf (ν), where the
peak height ν = δc /σ is a function of mass and time. Here, δc is the linearly extrapolated overdensity and can be
evaluated within the framework of the spherical collapse model. Many different expressions exist for f (σ) [1016–1021].
In Fig. 37 we present the dependence of the Tinker halo mass function [1019] on the matter power spectrum
normalization σ8 (left panel) and on the matter density Ωm,0 . We see that the effect on the halo mass function is that
of increasing the number of structures. A higher value of Ωm,0 implies also a higher σ8 value. The two parameters are
highly degenerate at the high-mass end, as shown by Fig. 36 where a 50 deg2 mock cluster counts survey is compared
76

FIG. 36: Ωm,0 − σ8 degeneracy in typical cosmic shear and cluster counts experiments. The blue curve and shaded
area show the degeneracy of the expected Ωm,0 − σ8 contours in a mock 50 deg2 X-ray cluster count survey, whereas
the green contours show the same degeneracy in the combined DES Y3 + KiDS-1000 cosmic shear estimate. We
−0.3
can see that the degeneracy is shallower in cluster counts experiments (σ8 ∝ Ωm,0 ) compared to cosmic shear (σ8 ∝
−0.5
Ωm,0 ), such that any definition of S8 that is optimal for either experiment will not perfectly follow the degeneracy
of the other. For comparison, we add the Planck CMB Ωm,0 − σ8 contours in orange.

with DES+KiDS WL and the latest Planck collaboration data. The degeneracy between Ω, and σ8 for galaxy cluster
counts is clearly visible.
It is important to note that halo masses are usually defined within a spherical overdensity ∆ with respect to
either the critical density ρc (z) or the mean density of the Universe ρm (z). For a given overdensity ∆c or ∆m , the
corresponding mass M∆ and radius R∆ are defined as the mass and radius within which the condition

M∆
3 = ∆c/m ρc/m (z) , (2.13)
4/3πR∆

is satisfied. The choice of the halo mass function and of the observational mass definition must therefore be adapted
to the chosen spherical overdensity. Common choices for the definition of the overdensity are 500c, 200c, and 200m.
Mass-observable scaling relations and the scatter around these relations While the halo mass function
is highly sensitive to cosmological parameters, constraining it from galaxy cluster surveys is a challenging task. The
main underlying issue is that halo mass is not a directly observable quantity. Samples of galaxy clusters are extracted
from surveys of the baryonic component that fills these structures, whether it is in the form of hot gas (X-ray,
Sunyaev-Zel’dovich effect) or stars (optical/IR). Galaxy cluster detection techniques and their primary observables
are described in more detail in Sec. 2.2.2.b. In all cases, the primary observable of the survey is not directly the halo
mass but rather a quantity that depends on it. The relation between survey observable and the halo mass, usually
referred as a scaling relation, needs to be properly understood to turn the detected cluster samples into constraints
on the halo mass function.
If galaxy clusters grow and evolve in a self-similar way [1022], the relations between observable quantities O and
spherical overdensity halo mass M∆ are expected to be power laws. While astrophysical effects such as mergers,
77

10−2 10−2

10−4
10−4
−6
10
dn/d ln M [h3 Mpc−3 ]

dn/d ln M [h3 Mpc−3 ]


10−6
10−8

10−10 10−8

10−12
10−10
10−14
Ωm = 0.20 σ8 = 0.75
−16 Ωm = 0.25 10−12 σ8 = 0.80
10
Ωm = 0.30 σ8 = 0.85
10−18 10−14
1012 1013 1014 1015 1016 1012 1013 1014 1015 1016
M [M h−1 ] M [M h−1 ]

FIG. 37: Dependence of the Tinker mass function on matter density Ωm (left panel) and on the matter spectrum
normalization σ8 (right panel) at z = 0.

cooling, and AGN feedback can cause deviations from the simple self-similar scenario, evidence suggests [1023–1026]
that these relations can be well described by power laws with a log-normal intrinsic scatter σlog O ,
 
M∆
log O = AOM + BOM + γOM E(z) ± σlog O , (2.14)
Mpivot

with E(z) = H(z)/H0 . Here the parameters AOM and BOM are the normalization and the slope of the mass-
observable scaling relation and γOM governs its redshift evolution. The scaling relation allows us to construct the
probability of detecting a system with a measured observable O given the source’s mass and redshift, P (O|M, z). A
precise knowledge of these parameters is required to make accurate cosmological inferences from galaxy surveys. We
discuss the associated issues in more detail in Sec. 2.2.2.c.
Selection function, forward modeling, likelihood function The cosmological analysis of galaxy cluster
catalogs follows a Bayesian inference approach. A generative model for the data given unknown model parameters
has to be postulated, and the PDF of the actual data as a function of the model parameters, the likelihood, has to
be computed. The generative model assumed for a cluster catalog with observable Ô and redshift z at a sky position
θ is a Poisson realization of the density field
dñ dn
= (1 + beff δ(θ, z)) , (2.15)
dÔ dÔ
with beff the effective bias, δ(θ, z) the matter density contrast, and
ˆ ˆ
dn dn
= dO dM P (O|M, z)P (Ô|O, z, θ)P (sel|Ô, O, z, θ) , (2.16)
dÔ dM

where d ln
dn
M is the halo mass function (Eq. (2.12)), P (O|M, z) the observable mass relation and its scatter (Eq. (2.14)),
P (Ô|O, z, θ) the effect of instrumental noise of the intrinsic observable O, and P (sel|Ô, O, z, θ) the selection function.
The latter takes values between 0 and 1 based on the probability of including the cluster with properties (Ô, O, z, θ).
The second term in Eq. (2.15) can be ignored for sufficiently high-mass and wide survey area samples, yielding
the Poisson likelihood used by Refs. [735, 1027–1033]. For smaller area surveys and lower mass surveys, the second
term contributes significantly via the variance of the matter field. In such cases, a composite Gaussian-Poisson
likelihood has to be used [1034–1041]. The resulting likelihood depends on the cosmological parameters via the halo
mass function and the cosmological volume. It also depends on the nuisance parameters describing the observable
mass relation and, potentially, the selection function. Marginalizing over sufficiently flexible observable mass relation
parameterizations is crucial to accurately represent the cluster population, especially with regard to the halo mass.
2.2.2.b. Galaxy cluster surveys
X-ray The vast majority of the baryonic content of galaxy clusters is in the form of a hot (> 107 [K]), highly ionized
plasma that fills the system’s gravitational potential well. Given the high temperatures involved, the intracluster
78

medium (ICM) shines predominantly in the X-ray range through thermal bremsstrahlung and line emission. As such,
galaxy clusters are luminous (LX = 1044 − 1045 [erg/s]) X-ray sources spanning a diameter of several arcmin. The
overwhelming majority of extended X-ray sources in the extragalactic sky are galaxy clusters, such that X-ray surveys
are powerful cluster detection machines. In the early 1990s, Röntgensatellit (ROSAT) [1042–1044] performed an all-
sky X-ray survey and several deeper but small area surveys, e.g., ROSAT Deep Cluster Survey (RDCS) [1045] which
allowed for the detection of several thousand clusters. This spurred the build-up of several sub-samples, either volume-
limited (REXCESS) [1046], or flux-limited (HIFLUGCS) [1047, 1048] with a well-understood selection function.
Deep follow-up observations of ROSAT-selected clusters with XMM-Newton and Chandra allowed to constrain cos-
mological parameters with cluster number counts by measuring the mass of these systems with better precision [1049].
XMM-Newton, thanks to its larger field-of-view, enabled to scan deeply (30-40 ks) several tens of square degrees, en-
abling the serendipitous detection of the fainter galaxy group population, extending the mass range used in the
cosmological analysis. The XXL survey [1050] detected ∼ 450 clusters over an area of 50 deg2 [1034, 1038]. More
recently, eROSITA undertook several all-sky surveys, with the ultimate goal of detecting about 100, 000 galaxy clus-
ters. The results from the first of these all-sky surveys were recently released [1033] and provided the detection of
∼ 12, 000 clusters over half the sky.
Sunyaev-Zel’dovich effect The thermal SZ effect [tSZ; 1051] describes the inverse-Compton scattering of CMB
photons off the hot electrons of the ICM. Photons crossing the path of a hot electron cloud occasionally encounter
hot electrons and gain energy, which induces a modification of the spectral shape of the CMB, see Ref. [1052] for a
review. At the peak frequency of the CMB, the tSZ effect takes the form of a decrement in the temperature of the
CMB, as a fraction of the photons are upscattered to higher frequencies. At first order, the spectral signal of the tSZ
depends only on the Compton parameter, y, which describes the average energy gain over the line of sight. The total
integrated signal over the area of a cluster, YSZ , is proportional to the integrated thermal energy of the ICM [1053].
The advantage of the tSZ effect as a cluster detection technique is twofold. First, the signal is essentially independent
of redshift, which makes it very efficient to detect high-redshift systems. Second, the primary tSZ observable, YSZ ,
is believed to be an excellent proxy of cluster mass [1054–1057]. In the past decade, the Planck survey yielded the
detection of > 2, 000 galaxy clusters with this technique extending out to z ∼ 1 [736, 1058]. While Planck was very
sensitive to the tSZ thanks to its wide frequency range and all-sky coverage, the detection efficiency of high-redshift
sources was hampered by the poor Planck beam, such that the bulk of the detected systems are at redshift z < 0.5.
Conversely, ground-based experiments have larger collecting areas and better angular resolution. SPT [1059–1061]
and ACT [1062, 1063] were now able to gather catalogues of several thousand galaxy clusters out to z ∼ 2.
Optical/near-IR photometry While in the past couple of decades, a great deal of attention was devoted to
the selection of galaxy clusters through the ICM, there is now renewed interest in the detection of galaxy clusters
as overdensities of galaxies thanks to the new generation of optical and near-IR surveys like Euclid and Rubin. As
their name suggests, galaxy clusters were originally defined as such based on the detection of overdensities of galaxies
in the same region of the sky. Based on visual inspection of photometric plates, Abell (1989) [1064] constructed
a catalogue of more than 4,000 clusters that has been used since more than two decades as reference of the local
large-scale structure. Since that time, sophisticated galaxy cluster detection algorithms have been developed which
allowed to extract large samples of clusters from galaxy surveys. Modern detection algorithms are mainly split into
two classes: red sequence and photometric redshifts. Red sequence cluster finders [1065–1067] take advantage of the
predominance of passive red galaxies in clusters with respect to the field to search for overdensities of red galaxies. This
method reduces spurious detections thanks to the use of the red sequence, which is known to trace rich environments
[1068, 1069]. Alternatively, methods based on photometric redshifts [719, 1070] make use of galaxy SEDs to search for
galaxy overdensities in three-dimensional space. These techniques rely on the availability of high-quality photometric
redshifts but make no assumption on the existence of a red sequence. For a detailed comparison of various cluster
finding algorithm we refer to Ref. [1071].
The main advantage of optical cluster catalogues is their high sensitivity, with the deepest current catalogues
containing more than 100,000 clusters (e.g., see Refs. [1072–1076]). Modern algorithms also provide an estimate of
the cluster redshift as a direct byproduct. The primary observable of optical cluster finders is the number of member
galaxies, or richness N . A standardized richness estimate is computed iteratively to include cluster members above a
given absolute magnitude. Richness is known to correlate with halo mass [1077, 1078], although with a large intrinsic
scatter induced by projection effects and miscentering [1079, 1080].
Shear selection The techniques discussed above all rely on detection through the baryonic content of clusters,
whether it is in the form of gas or galaxies, whereas the vast majority of the mass is made of DM. In the past
decade, direct detection of galaxy clusters as peaks in projected WL mass maps has become possible [902, 1081–1084].
This technique allows to detect clusters directly through their total mass content rather than through a proxy that
is related to the halo mass in a complex way. Thanks to the depth of the Subaru/HSC-SSP survey (∼ 30 galaxies
per square arcmin), Ref. [1082] presented a sample of 65 shear-selected clusters within a ∼ 160 deg2 area with a S/N
79

greater than 4.7. All the most significant peaks were associated with galaxy overdensities or X-ray signal [1085]. More
recently, the first cosmological constraints from shear-selected cluster samples were extracted from an area of ∼ 500
deg2 [1083, 1086]. While at the present day shear-selected cluster catalogs are much smaller than catalogs obtained
from other techniques, upcoming lensing experiments like Euclid will yield much larger samples directly extracted
from the projected mass distribution. The primary issue associated with shear selection is projection associated with
correlated large-scale structure and halo orientation, such that the masses of the detected systems are expected to
be over-estimated by ∼ 50% [1087]. Future works will need to model projection effects in detail to make accurate
cosmological inference from large samples of shear-selected systems.
2.2.2.c. Mass calibration and systematics
Mass calibration techniques The number of galaxy clusters as an observable displays a direct degeneracy
between the mean mass scale of the sample and the cosmological parameters, which dictate the number density of
halos via the mass function and the cosmological volume. Number counts of galaxy clusters can thus only constrain
cosmological parameters if the relation between halo mass and cluster observables is determined, a problem called
mass calibration. Our knowledge of subgrid astrophysical processes limits the direct calibration of this relation on
hydrodynamical simulation. It is thus very challenging to derive the degree of uncertainty on the mass calibration
from pure simulation-based studies. The latter is, however, the main contributor to the final cosmological constraining
power of cluster number counts experiments and should, therefore, be assessed in a reliable, empirical fashion.
In the context of a CDM cosmology, most of the halo mass is composed of DM. Empirical mass calibration can
thus only proceed via measurements of the cluster’s gravitational potential. Three techniques have been established
to observe the cluster gravitational potential: i) hydrostatic equilibrium; ii) cluster galaxy dynamics; iii) WL. Owing
to its independence on the dynamical state, WL has become in recent years the method of choice to calibrate galaxy
cluster mass, see Ref. [1088] for a review.
While WL is expected to be close to unbiased on average, the signal-to-noise of individual cluster mass measurement
is usually very low. Mass calibration is thus integrated into the likelihood framework discussed above by considering
the marginal PDF of the observable directly linked to halos mass ÔM conditional on the observable one seeks to
calibrate Ô [1028, 1029, 1031, 1032], reading
−1
dn dn

p(OM |Ô, z, θ) = , (2.17)
dÔ dÔdÔM
with
ˆ ˆ ˆ
dn dn
= dO dOM dM P (O, OM |M, z)P (Ô|O, z, θ)P (ÔM |OM , z)P (sel|Ô, O, z, θ) , (2.18)
dÔdÔM dM

where P (O, OM |M, z) is the joint scaling relation between mass calibration observable and primary observable, and
P (ÔM |OM , z) parameterized the measurement uncertainty on the mass calibration observable. Note that the cor-
relation between the observable mass relation is crucial to account for physical selection effects [1089, 1090]. The
marginal relation P (OM |M, z) needs to be anchored, and its parameters tightly constraint within prior, such that the
overall likelihood of the mass calibration depends only on the parameters of the primary observable mass relation,
thus constraining them.
Mass calibration uncertainties For the sake of brevity, we shall focus on the systematic uncertainty affecting
WL mass calibration. Given the advent of wide and deep photometric surveys, WL mass calibration has emerged as
the principal way to calibrate cluster number count. Three sources of statistical uncertainty impact the WL signal
of galaxy clusters: the shape noise due to the intrinsic ellipticity of source galaxies, the intrinsic heterogeneity of
cluster mass profiles at the same mass (due to scatter around the concentration mass relation, varying degrees of
substructure, orientation, triaxiality, etc.), and the statistical fluctuations in the line-of-sight integrated matter field
[1091, 1092]. Deeper and wider WL surveys can only reduce the first of these three noise sources.
WL mass calibration shares the systematics of the shear measurement and photometric redshift estimation inherent
in the source catalogs from wide and deep photometric surveys. Furthermore, the displacement between the observed
and true cluster position washes out the WL signal (mis-centering). As clusters are significant overdensities in the
galaxy field, depending on the source background selection, a fraction of unlensed cluster members might contaminate
the source sample and dilute the signal (cluster member contamination). Finally, baryonic effects do not only alter
the total halo mass but also the halo mass profile that sources the lensing. The combined impact of these effects is
summarized by establishing the relation between the WLL mass (that results from fitting the WL signal) and the
halo mass on realistic WL simulations [1093–1095]. WL masses are biased low by 5-10% depending on the fitting
formula used and scatter around the true halo mass [1096–1098]. The systematic uncertainties of the aforementioned
effects are distilled into an uncertainty on the bias of the WL mass [1031, 1090, 1099, 1100]. Current understanding
80

of baryon feedback puts a 2% lower limit on this uncertainty [1095]. Depending on the WL survey, photometric
source redshift uncertainties increase this to 10% at high cluster redshifts. The other effects have a quantifiably
smaller impact, as, for instance, mis-centering and cluster member contamination are partially controlled empirically
[1031, 1078, 1090, 1101].
Halo mass function The halo mass function described above only considers structures made by CDM. However,
halos contain a certain amount of baryons. Neglecting this contribution leads to a systematic bias in determining the
halo mass function. This can be easily understood by considering that baryons will change the total mass of the halo.
Hydrodynamical simulations found that the effects of baryons are different whether, in the simulations, radiative
effects are taken into account or gas heating is only due to gravitational effects. For instance, models with AGN
feedback decrease the abundance of clusters compared to DM-only simulations [1028]. The impact of baryons appears
to depend on the halo mass. Considering an overdensity of 1500 times the critical density, baryons contribute up to
6%–7% to the total mass, while for 200 times the critical density, their contribution reduces to only about 1%.
In terms of number counts, [1102, 1103] found that the number density of halos decreases up to 15% at low masses
and redshifts, while being in broad agreement with that of DM-only simulations in the other cases. In addition,
to better match X-Ray and SZ observational studies, [1103] proposed fits to estimate the baryon contribution when
moving from a halo defined as 200 times the critical density to one found using 500 times this density. The retrieved
halo masses in the presence of baryons can also be corrected to match the expected halo masses in gravity-only
simulations [1095, 1104].
2.2.2.d. Results
Cosmological parameters measurements and tensions Cosmological results on the mean matter density at
present time, Ωm , from cluster number counts agree with CMB results that it should be around 0.3. Recent results
from eROSITA (0.29 ± 0.01 [1033]), SPT (0.29 ± 0.03 [1032]), WtG (0.26 ± 0.03 [1027]), XXL (0.31 ± 0.03 [1038]),
+0.08
Planck SZ (0.33 ± 0.03 [735] and DES (0.32−0.07 [1105]) are all consistent with the results from Planck CMB [192] of
0.315 ± 0.007.
Conversely, there is currently some disagreement between the values of σ8 from cluster number counts obtained in
various studies. While some studies find σ8 in agreement with the CMB value of 0.811 ± 0.006 (eROSITA 0.88 ± 0.02
[1033]; SPT 0.82 ± 0.03 [1032]; WtG 0.83 ± 0.04 [1027]; XXL 0.84 ± 0.04 [1038]), Planck SZ found a value of 0.76 ± 0.03
[735] which is in tension with the CMB at the level of ∼ 2.5σ.
Similar to the case of cosmic shear, and as highlighted in Fig. 37, consistency with CMB should be assessed
by considering theentire multi-dimensional posterior distribution. For this reason, it is customary to consider the
α
Ωm,0
quantity, S8 = σ8 0.3 , with α = 0.5. We note, however, that the degeneracy between Ωm,0 and σ8 in cluster
count experiments follows a slightly different slope from cosmic shear, with an index in the range 0.2 (e.g., SPT [1032])
to 0.4 (e.g., eROSITA, [1033]).
Considering only experiments including an internally calibrated WL mass-observable scaling relation, i.e., eROSITA,
SPT, WtG, and XXL, the tension between Ωm and CMB measurements is at most 1.2-σ, thus showing a high level
of consistency between experiments. The same exercise applied to σ8 indicates a slightly higher level of tension (1.7
σ), such that potential systematics still remain at this level of precision.
Constraints on dark energy Cluster number counts are useful in providing constraints on the DE equation of
state w. DE affects the growth of structures in the Universe at redshifts below ∼ 1, thereby reducing the growth of
massive halos. Most current studies on the DE equation of state from cluster counts assume a constant equation of
state w and a flat universe. All the results published thus far are consistent with w = −1 (Chandra w = −1.14 ± 0.21
[1049]; WtG w = −0.98 ± 0.15 [1027]; SPT w = −1.45 ± 0.31 [1032]; eROSITA w = −1.12 ± 0.12 [1033]). Similar
results were obtained constraining the gas mass fraction to be constant with time (w = −1.13+0.17
−0.20 [1106]).
Alternative models The halo mass function, as discussed above, is a great tool to investigate the underlying
cosmological model. For this reason, several authors studied its behaviour in MG models, such as the normal branch
of Dvali-Gabadadze-Porrati theory (nDGP) [1107] and f (R) [1108] simulations. To model the HMF in MG scenarios,
N -body simulations are required to calibrate the difference between the HMF in the MG simulations with respect
to ΛCDM. Ref. [1107] studied the evolution of perturbations for the nDGP model, which is characterized by a time-
varying gravitational constant. The numerical nDGP mass function predicts up to three times more objects at very
high masses, for strong deviations from ΛCDM, explained by the fact that the halo mass function is very sensitive to
σ8 .
A similar approach, but which combines the benefits of the previous ones is that used by Refs. [1108, 1109] to
study the halo mass function in f (R) models. The authors parameterize the modifications induced by f (R) models
both in the spherical collapse model and in the HMF, calibrating these modifications on N -body simulations. These
modifications are a function of mass, redshift. Finally, Ref. [1110] showed that the halo mass function is universal
when expressed in terms of ln (σ −1 ) rather than the halo mass M .
81

2.2.2.e. Other cosmological probes from galaxy clusters: cluster clustering The spatial distribution of galaxy
clusters represents a valuable cosmological probe, depending on the composition of the Universe and on the nature of
gravitational interaction. In particular, cluster clustering traces the growth rate of density perturbations, therefore it
has been widely employed for inferring fundamental cosmological parameters, such as the matter density, Ωm,0 , and
the amplitude of mass fluctuations on scales of 8 h−1 Mpc, σ8 [1111–1114], as well as the neutrino properties [1115]
and deviations from the ΛCDM model, possibly described by the DDE parameters, i.e., w0 and wa [1116]. Moreover,
it can be used to significantly improve the precision on the constraints from cluster weak gravitational lensing [1111]
and number counts [1116], when analyzed in combination with them.

Despite the limit in the exploitation of the clustering properties due to the difficulty in collecting large homo-
geneous cluster samples, which leads to poorer statistics with respect to galaxy samples, cosmology with galaxy
clusters presents a series of relevant advantages. First, galaxy clusters, being placed at the nodes of the filamentary
structure of the cosmic web, are hosted by the latest and most massive virialised haloes formed by the hierarchical
growth of cosmic structures. Therefore they are highly biased tracers, more clustered than galaxies [1015, 1117, 1118].
While, in general, the galaxy bias is difficult to model properly, especially on small physical scales, and it is usually
seen as a nuisance parameter in cosmological analyses based on galaxy statistics, the effective bias of galaxy cluster
samples can be theoretically predicted [1113, 1114, 1119, 1120], depending on the mass and the redshift of the hosting
haloes [1121]. In turn, the cluster mass, mainly consisting of DM, can be linked to directly observable mass proxies,
through the so-called mass-observable scaling relation [1015, 1101, 1122, 1123]. However, the uncertainties on scaling
relation parameters and their redshift evolution affect our ability to recover the true cluster mass and represent one
of the limits of cosmological inference from galaxy clusters. This highlights the importance of having a robust WL
mass calibration, capable of assessing the systematics caused by halo orientation, selection, and projection effects
[1037, 1078]. A second advantage is that galaxy clusters present lower peculiar velocities with respect to their host
galaxies, reducing the small-scale distortions in the clustering signal, namely the Fingers of God effect, caused by
nonlinear dynamics and incoherent motion within virialised structures [1124, 1125].

Cluster clustering is traditionally investigated through the analysis of the two-point correlation function, ξ(r), which
measures the excess probability of finding a pair of objects in the volume elements δV1 and δV2 , at the comoving
separation r, relative to that expected from a random distribution. Increasing attention is being paid to higher-order
statistics, and in particular to the three-point correlation function, ζ(r12 , r13 , r23 ), which provides the probability of
finding triplets of objects at comoving separations r12 , r13 , and r23 , and has been successfully applied for the detection
of the BAO peak of galaxy clusters [1126]. These probes require the assumption of a fiducial cosmology to convert
the observed angular and redshift coordinates into distances, which might be different from the true one, resulting in
geometric distortions [1127], which have to be taken into account in the modeling of the clustering signal [1128, 1129].
In recent years, the two-point redshift-space correlation function of an homogeneous sample of X-ray selected galaxy
clusters from the XXL survey, carried out by the XMM-Newton satellite, has been used to measure a value of Ωm,0
in full agreement with the ΛCDM model [1130]. On the other hand, optically selected clusters from the Sloan Digital
Sky Survey (SDSS) have been employed to get an estimate of f σ8 , which was found consistent to General Relativity
predictions [1112], to determine the distance–redshift relation, Ωm,0 and H0 , from the BAO peak [1124, 1131] and, in
combination to stacked gravitational lensing, to constrain σ8 [1111]. Moreover, from the 3D clustering of the Planck
SZ selected galaxy clusters, the Planck mass bias, bSZ , and Ωm,0 , have been derived, while σ8 was not constrained
[1132]. The tomographic cluster autocorrelation from the Constrain Dark Energy with X-ray (CODEX) sample, after
redshift-dependent richness selection, has been employed to explore the Ωm,0 − σ8 degeneracy, putting constraints
on the structure growth parameter, S8 [1133]. Since cluster clustering is particularly sensitive to these parameters,
independent constraints on S8 have been put by measuring the two-point correlation function of the photometric
clusters detected in the third data release of KiDS-DR3 [1113]. An equivalent approach consists in the study of the
angular correlation function, w(θ), and its spherical harmonic counterpart, the angular power spectrum, Cℓ [1134].
In particular, a first measurement of the auto-correlation cluster power spectrum and of the galaxy-galaxy cluster
cross-spectrum has been performed on the SDSS-DR8 photometric sample [1120]. Moreover, from the tomographic
study of w(θ) and Cℓ of the same KiDS-DR3 catalogue, competitive constraints on S8 have been found [1114]. These
results, being based on angular positions alone, are not affected by geometric distortions, highlighting the potential
of angular cluster clustering as a cosmological probe in future photometric redshift surveys.

2.2.3. Galaxy Clustering – Other probes


Coordinator: Chandra Shekhar Saraf
Contributors: Kishan Deka, and Paweł Bielewicz
82

2.2.3.a. Cross-correlations The S8 parameter can be measured from the redshift space clustering of galaxies by
estimating the two- and three-point correlation functions or their Fourier counterparts, the galaxy power spectrum
and bispectrum. These statistics can be used to gain insights into the growth rate of cosmic structures, f σ8 (z), from
RSD at the effective redshift of the galaxy sample [1135–1137]. The galaxy clustering data can be combined with
CMB surveys to yield constrains on σ8 and S8 . Several studies [1138–1141] have reported values of σ8 consistent with
the CMB-only analyses. An illustrative schematic of estimating σ8 parameter from cross-correlation measurements
between Planck CMB lensing convergence map and DESI’s Legacy Imaging Survey (DESI-LIS; [1142]) is shown in
Fig. 38 (see Ref. [1143] for details on cross-correlation analysis). A cross-correlation between unWISE galaxies and
ACT DR6 CMB lensing measurements found S8 = 0.813±0.021, and S8 = 0.810±0.015 with a combination of Planck
and ACT cross-correlations [1140]. These measurements are fully consistent with the predictions from primary CMB
measurements within our standard cosmology. Another study correlating Gaia–unWISE RGB catalog with Planck
PR4 CMB lensing found σ8 = 0.7766 ± 0.034 and Ωm,0 = 0.343+0.017 −0.019 , which translates to S8 = 0.819 ± 0.058 [1139].
However, many cross-correlation analyses between CMB datasets and galaxy samples divided into narrow redshift
bins [1144–1151] find low values of σ8 and S8 parameter (when Ωm,0 is generally assumed to be consistent with the
well constrained values from SN and CMB analyses), resulting in mild tension with Planck analysis. A tomographic
cross-correlation of DES-Y3 data and CMB lensing from SPT and Planck estimated S8 = 0.734+0.035 −0.028 [1145]. Another
cross-correlation of DESI Legacy Imaging survey LRGs and CMB lensing from Planck found S8 = 0.765 ± 0.023 and
S8 = 0.790+0.024
−0.027 with ACT DR6 [1152].

Planck lensing convergence map

window functions estimated parameters

DESI-LIS galaxy density maps

(a) (b) (c)

FIG. 38: An illustration of estimating σ8 parameter from tomographic cross-correlation measurements between
DESI-LIS galaxy survey and Planck lensing convergence map, assuming ΛCDM model [1153]. (a) The Planck lens-
ing convergence map and DESI-LIS galaxy density maps in four tomographic bins, smoothed with a Gaussian beam
of 60′ FWHM (only to better show the large scale distribution). (b) The galaxy redshift distribution for the four
redshift bins (solid blue lines) and the CMB lensing kernel (solid red line). The vertical dashed orange lines mark
the boundaries of redshift bins. (c) The σ8 parameter estimated from cross-correlation measurements of four tomo-
graphic bins. The red line is the evolution of σ8 parameter as estimated from cross-correlation data. The dashed
black line is the evolution of σ8 parameter measured from the Planck CMB data alone.

2.2.3.b. Full shape analyses The σ8 parameter can also be measured by fitting the full shape of the observed
power spectrum from spectroscopic surveys as well as the galaxy bispectrum [720, 1154–1158]. Furthermore, the σ8
parameter can be translated into S8 parameter with additional measurements of Ωm,0 (from power spectrum shape,
SN or CMB). The measurements of S8 parameters employing full shape analysis of the galaxy power spectrum
[1154, 1158] are in tension with Planck measurements [192], but with typically less than 3σ statistical significance.
Recent constraints on the S8 parameter from the BOSS galaxy bispectrum multipoles found S8 = 0.774+0.056 −0.053 [720]
and S8 = 0.77 ± 0.04 [1157], which are statistically consistent with both CMB and WL measurements.
83

2.2.3.c. 3×2pt analyses The combined analysis of two-point statistics corresponding to galaxy clustering, galaxy-
galaxy lensing and cosmic shear is commonly known as the “3 × 2pt” analysis. The joint analyses provide us with an
improved control on the systematic uncertainties such as galaxy bias and intrinsic alignments. The first 3×2pt analysis
were published with the datasets from KiDS-450 × {2dFLenS + BOSS} [1159], KiDS-450 × GAMA [1160] and DES-Y1
[1161]. The DES-Y1 analysis used photometric galaxies as both lens and sources, finding S8 = 0.773+0.026
−0.020 . The KiDS
collaboration, on the other hand, used overlapping imaging and spectroscopic surveys, resulting in S8 = 0.742 ± 0.035
when combined with 2dFLenS + BOSS, and S8 = 0.800+0.029 −0.027 with GAMA survey.
The updated constraints on S8 with 3 × 2pt analyses combining KiDS-1000 with {2dFLenS + BOSS} datasets gives
S8 = 0.766+0.020
−0.014 [872], at 3.1σ tension with Planck constraints within the ΛCDM model. The recent measurements
from the DES-Y3 datasets resulted in S8 = 0.776 ± 0.017 in the ΛCDM cosmology and S8 = 0.775+0.026 −0.024 assuming
a wCDM model [867]. However, a joint analysis of DES-Y3 3 × 2pt data, Planck CMB anisotropy data (without
lensing), eBOSS BAO and RSD measurements, and DES SNIa data provides S8 = 0.812 ± 0.008 in the ΛCDM model.
A joint cosmic shear analysis of KiDS-1000 and DES Y3 data has obtained stringent constrain of S8 = 0.790+0.018 −0.014
[866]. Another 3 × 2pt analyses performed with the HSC-Y3 imaging data and SDSS BOSS DR11 spectroscopic
+0.043
galaxies yielded S8 = 0.775−0.038 for the ΛCDM model [870]. The HSC-Y3 analysis is highly consistent with both
KiDS-1000 and DES-Y3 measurements of S8 within one standard deviation.
2.2.3.d. Systematics The measurements of cosmological parameters from large-scale structure evolution at low
redshifts require state-of-the-art analysis pipelines to correct for a large number of systematic effects. The observation
sector can include effects such as atmospheric and extragalactic extinction, stellar contamination, airmass and blurring
[1162], fibre collisions [1163], angular and radial modes systematics [1164].
For the cross-correlation between galaxy clustering and CMB, galaxy survey systematics such as photometric
calibration errors, foreground contamination and catastrophic errors need to be modelled and mitigated for unbiased
inferences [1148, 1151, 1165–1167]. In order to allow tomographic cross-correlations with photometric galaxy surveys,
both precise individual galaxy photometric redshifts and accurate redshift distributions for the ensemble, require
techniques for redshift determination. Typically, a subset of spectroscopically observed sources are used to train,
validate and calibrate photometric redshifts and redshift distributions. The accuracy of the redshift distribution
directly impacts the accuracy of estimated cosmological parameters. Biases in the mean redshifts of the tomographic
bins will cause a bias in S8 . Recently, [1143, 1153] have shown the importance of precise modelling of redshift error
distributions and redshift bin mismatch of objects, when estimating σ8 and S8 parameters.
In the theory sector, the largest uncertainty comes from the accuracy of the models for non-linear clustering of
galaxies. The cross-correlation measurements, full shape analyses and the bispectrum multipoles, all depend on
the accurate modelling of the non-linear regime in the redshift-space galaxy clustering. This uncertainty has been
improved after the significant progress from the effective field theory of the large-scale structure which provides
accurate description of galaxy clustering at large scales [718, 721, 1168–1176].

2.3. Other challenges


2.3.1. Systematics and the Alens parameter
Coordinator: William Giarè
Contributors: Alessandro Melchiorri, Anto Idicherian Lonappan, Deng Wang, Elsa M. Teixeira, Enrico Specogna,
Giulia Gubitosi, Ido Ben-Dayan, Matteo Forconi, Özgür Akarsu, and Rita B. Neves

On scales smaller than ten arcminutes, the interaction between CMB photons and the Universe’s large-scale struc-
ture becomes significant, giving rise to second-order anisotropies. One of the most important contributions is the
gravitational deflection, or lensing, experienced by CMB photons along their paths [1177]. At arcminute scales, lens-
ing deflections distort the observed image of the CMB fluctuations, imprinting a distinctive non-Gaussian four-point
correlation function (or trispectrum) in both the temperature and polarization anisotropies [1177]. This signal can be
extracted with a high signal-to-noise ratio by correlating power in different directions on the sky, leading to a direct
reconstruction of the power spectrum of the CMB lensing field, now measured at ∼ 40σ [693, 694, 1178, 1179].
Although gravitational lensing does not alter the overall distribution of primary CMB anisotropies, it leaves dis-
tinctive signatures in the spectra of temperature and polarization anisotropies. The most relevant effects are the
conversion of E-mode polarization to B-modes, the transfer of power to the damping tail, and the lensing-induced
smoothing of the acoustic peaks and troughs in the TT, TE, and EE spectra. The degree of smoothing is directly tied
to the amplitude of the power spectrum of the lensing potential, which is derived from the six ΛCDM parameters. As
a result, interesting consistency tests have been proposed to verify whether the amplitude of lensing inferred through
the high-ℓ smoothing of acoustic peaks matches the theoretical predictions of the standard cosmological model.
In Ref. [723], the phenomenological parameter Alens was introduced to encompass various physical mechanisms
that might affect the lensing amplitude. This parameter scales the amplitude of the lensing trispectrum, effectively
84

accounting for the associated smoothing effects in the CMB temperature and polarization spectra. Alens is defined
such that Alens = 1 matches the standard ΛCDM prediction, whereas Alens = 0 corresponds to a scenario where CMB
lensing is disregarded altogether. By allowing Alens to remain a free parameter in theoretical models, its value can be
directly constrained by data, potentially confirming or deviating from the ΛCDM predictions.
Focusing on the Planck-2018 data release (PR3), the analysis of the Planck plik likelihood for the TT, TE, and
EE spectra at ℓ > 30, combined with the Commander likelihood for the TT spectrum at 2 ≤ ℓ ≤ 30 and the SimAll
likelihood for the EE spectrum at 2 ≤ ℓ ≤ 30, yields Alens = 1.180 ± 0.065 [192]. This finding suggests an excess
lensing signal at approximately 2.8σ, resulting from a significant improvement in χ2 of approximately ∆χ2 ∼ 9.7. This
improvement primarily originates from high-ℓ temperature and polarization data, particularly within the multipole
range 600 < ℓ < 1500. As shown in Fig. 24 of Ref. [192], there is a visible preference for increased lensing smoothing
in the oscillatory residuals in the TT spectrum at 1100 < ℓ < 2000.
Since 2018, the Planck data has undergone substantial reanalyses. The new Planck PR4 (NPIPE) CMB maps
incorporate significant improvements, such as increased coverage of sky area at high frequencies, improved processing
of time-ordered data, and approximately 8% more data accounted for in the lensing trispectrum reconstruction.
Updated likelihoods for temperature and polarization spectra have been released following these developments. The
latest versions of CamSpec [699] and HiLLiPoP [700] – two likelihoods already employed in various studies by the
Planck collaboration – are now based on the Planck PR4 (NPIPE) maps, reducing small-scale noise compared to plik
and enhancing the constraints on cosmological parameters by up to 10%. Both likelihoods indicate a shift towards
Alens = 1. For CamSpec, the constraints on the lensing amplitude inferred from high-ℓ smoothing of acoustic peaks
are summarized in Table 6 of Ref. [699]. Temperature and polarization data reduce the preference for Alens > 1 to
less than 1.7σ while focusing solely on the TT spectrum increases the lensing anomaly to 2.3σ. Constraints on Alens
resulting from the new HiLLiPoP likelihood are summarized in Table 6 of Ref. [700]. They consistently agree with
ΛCDM at 1σ.
These new reanalyses suggest that the preference for excess power smoothing has decreased in the NPIPE maps,
lending weight to interpreting the lensing anomaly as systematics in the Planck PR3 data. The latter interpretation
finds support in CMB experiments other than Planck. ACT and the SPT have released precise small-scale measure-
ments of the spectra of temperature and polarization anisotropies [703, 724, 731], as well as precise reconstructions of
the lensing trispectrum [694, 1179, 1180]. The Data Release 4 of the ACT constrains Alens = 1.01 ± 0.11 in remarkable
agreement with the baseline value (see Fig. 16 of Ref. [701]). Similarly, the SPT TT, TE, and EE spectra analysis
gives Alens = 0.87 ± 0.11, consistent with ΛCDM at 1.2σ. In this case, we refer to Table V of Ref. [703] and Table VII
of Ref. [724] for earlier analyses involving only the EE and TE spectra (yielding Alens = 0.98 ± 0.12).
Fig. 39 summarizes the 68% CL intervals for the parameter Alens , inferred over the years from the analysis of
lensing-induced smoothing of the acoustic peaks and troughs in the TT, TE, and EE spectra measured by various
experiments and likelihoods discussed so far.

2.3.2. Evidence for a nonvanishing Ωk


Coordinator: Will Handley

The spatial curvature of the Universe, parameterized by Ωk , is a fundamental parameter in cosmology, intimately tied
to the geometry and fate of the Universe. While the inflationary paradigm predicts a value extremely close to zero,
it is important to observationally test this prediction. In the standard cosmological model, Ωk is degenerate with
other parameters when using CMB data alone, such that CMB data do not provide strong constraints on curvature.
However, the addition of external data sets, such as BAO measurements, can break this degeneracy.
The Planck 2018 data release presented an intriguing hint of a possible departure from a flat universe. Analyzing
the TT,TE,EE+lowE data, the Planck team found a preference for negative values of Ωk in the range −0.095 < Ωk <
−0.007 at 99% confidence level, with ∆χ2eff = −11 compared to the baseline ΛCDM model [192]. This preference
was attributed to a combination of volume effects and better fits to both the high-ℓ and low-ℓ data in closed models.
Notably, closed models predict a higher lensing amplitude (AL ), which can explain the preference for AL > 1 in Planck
data, and also a better fit to the low-ℓ temperature power spectrum. Adding Planck lensing data to TT,TE,EE+lowE,
the constraint on Ωk shifts to Ωk = −0.0106±0.0065 at 68% confidence level, reducing the preference for closed models
to less than 2σ. Finally, including BAO data results in Ωk = 0.0007±0.0019, consistent with a flat universe at 1σ [192].
However, this conclusion has been challenged in the literature. Ref. [1181] and Ref. [1182] show that there is Bayesian
evidence for a closed universe, albeit not quantified in the Planck paper, and that the Planck lensing reconstruction
is in tension with TT,TE,EE+lowE data at 2.5σ. In particular, the tension is driven by the preference for a higher
value of AL in closed models, which is not supported by the Planck lensing reconstruction. They argue that BAO
data are also in tension with TT,TE,EE+lowE data, and that combining these datasets is not justified. Ref. [1183]
conversely shows that the preference for Ωk < 0 is reduced when using the CamSpec likelihood, which uses more data
than the Planck baseline likelihood, and argue that BAO data strongly prefer flat universes.
85

SPT-3G

ACT-DR4

Planck-PR4 (HiLLiPoP)

Planck-PR4 (CamSpec)

Planck-PR3 (CamSpec)

Planck-PR3 (Plik)

0.7 0.8 0.9 1.0 1.1 1.2 1.3


Alens
FIG. 39: The whisker plot summarizes the 68% CL intervals for the parameter Alens , inferred over the years from
the analysis of lensing-induced smoothing of the acoustic peaks and troughs in the TT, TE, and EE spectra mea-
sured by various experiments with different likelihoods. Constraints based on the Planck-PR3 measurements indi-
cate an excess of lensing, which is reduced in the recent Planck-PR4 updated likelihoods. Small-scale temperature
and polarization spectra from ACT-DR4 and SPT-3G are in overall agreement with Alens = 1, consistently recov-
ered within one standard deviation (or slightly more).

More recently, Ref. [1184] showed that the BAO scale measurements implicitly assume a fiducial ΛCDM cosmology,
and that the preference for a flat universe can be weakened when relaxing these assumptions. They find that using
the full shape of the galaxy power spectra, the tension between BAO and Planck data in the Ωk parameter is reduced
to ∼ 1.5σ.
The tension between Planck CMB data and BAO measurements in the context of non-vanishing curvature has
motivated numerous works that explore its possible origin and its implications for cosmology. For example, Ref. [1185]
used CC as an alternative dataset to break the geometric degeneracy, finding Ωk = −0.0054 ± 0.0055, consistent with
a flat universe. However, the tension between different H0 measurements persists even in the context of non-flat
models [1186–1188].
The curvature tension therefore remains an open problem in cosmology. Future observations, such as those from
DESI, Euclid, and the Vera Rubin Observatory will provide more precise BAO measurements, which will allow us to
test the robustness of the flat Universe assumption and potentially shed light on the origin of this discrepancy.

2.3.3. Anisotropic anomalies in the cosmic microwave background radiation


Coordinator: Leandros Perivolaropoulos
Contributors: András Kovács, Anto Idicherian Lonappan, Emanuela Dimastrogiovanni, Eoin Ó Colgáin, Frode K.
Hansen, Giulia Gubitosi, Laura Mersini-Houghton, Marina Cortês, Nils A. Nilsson, Shahin Sheikh-Jabbari, and Venus
Keus

The CMB, a relic from the early Universe, provides crucial insights into cosmology. Observations by COBE, WMAP,
and Planck have revealed several anisotropic anomalies that may challenge the prevailing ΛCDM cosmological model
and the cosmological principle, prompting a reevaluation of our understanding of the early Universe’s structure and
the inflationary paradigm [1189, 1190].
One such anomaly is the unusually low quadrupole moment (ℓ = 2) in the CMB power spectrum, which deviates from
the predictions of cosmic variance and the ΛCDM model [725, 1191]. This anomaly has implications for inflationary
cosmology and may require theoretical adjustments. The low quadrupole moment was first observed by COBE and
later confirmed by WMAP and Planck, with increasing precision. The statistical significance of this anomaly has
been thoroughly investigated, and its persistence across multiple observations challenges our understanding of the
primordial power spectrum and the inflationary paradigm. Theoretical efforts to explain the low quadrupole moment
have included modifications to the inflation model, such as a running spectral index or a cutoff in the primordial
power spectrum at large scales [1192, 1193].
Another striking feature is the alignment of the quadrupole and octopole moments, which appears to be oriented
86

with the ecliptic plane and the motion of the Solar System [1194, 1195]. This alignment violates statistical isotropy
and suggests potential cosmological or local explanations. The quadrupole-octopole alignment was first reported
using WMAP data and later confirmed by Planck, with both missions providing strong evidence for its existence. The
alignment has been studied extensively, with some proposing that it could be a signature of cosmic topology [1196],
while others have investigated the possibility of local foreground contamination or systematic effects [1197, 1198]. The
statistical significance of this alignment and its potential origins remain active areas of research.

FIG. 40: Upper Panel: Derived quadrupole moment from the Planck SMICA map, showing the preferred axis of the
quadrupole. Lower Panel: Derived octopole moment, indicating the alignment with the quadrupole. These align-
ments suggest a significant anomaly in the CMB, challenging the assumption of statistical isotropy. The plus and
star symbols indicate the axes of the quadrupole and octopole, respectively, around which the angular momentum
dispersion is maximized. The diamond symbols correspond to the quadrupole axes after correction for the kine-
matic quadrupole. From Ref. [1199].

The upper and lower panels of Fig. 40 illustrate the derived quadrupole and octopole moments, respectively. The
alignment of these moments with the ecliptic plane highlights one of the key anisotropic anomalies in the CMB. This
alignment suggests a possible violation of statistical isotropy and indicates that there may be underlying cosmological
or local causes for this observed pattern. Understanding the significance and origins of these alignments is crucial for
interpreting the implications for the standard cosmological model and exploring potential new physics.
The CMB also exhibits a hemispherical power asymmetry, known as the dipolar power distribution, where one
half of the celestial sphere has significantly more temperature fluctuations than the other [1200, 1201]. The scale
and statistical significance of this asymmetry raise questions about the density variations in the early Universe. The
hemispherical power asymmetry was first detected in WMAP data and later confirmed by Planck. The hemispherical
asymmetry can be considered as two different anomalies as described in detail in Ref. [1201]. For larger scales,
ℓ < 100, the asymmetry can be modeled as a dipolar modulation of an isotropic field. The p-value this modulation
asymmetry has been reported as 0.1−1%, but considerably less significant when correcting for the choice of ℓmax . The
second hemispherical asymmetry, the angular clustering asymmetry, extends to much smaller scales. By estimating
the CMB power spectrum locally in different parts of the sky, maps of the spatial distribution of the power spectrum
can be estimated for different multipole ranges. When estimating a dipole for these local power spectrum maps at
different scales, these dipoles are found to cluster. This clustering is found to persist at least to ℓ = 1000 and only
1/1000 simulations show similar clustering. The clustering direction is close to the direction of the large scale dipolar
asymmetry. The origin of this asymmetry remains unclear, with some proposing that it could be a signature of non-
Gaussianity in the primordial perturbations [1202–1204], while others have explored the possibility of a super-horizon
scale mode modulating the primordial power spectrum [1205].
Additionally, there is a parity asymmetry in the CMB, with a preference for odd-parity modes over even-parity
87

modes, challenging the scale invariance of primordial fluctuations [1206, 1207]. The parity asymmetry was first
reported using WMAP data and later confirmed by Planck, with both missions providing evidence for a statistically
significant difference between the power in odd and even multipoles. This asymmetry has been studied in the context
of the primordial tensor-to-scalar ratio, with some suggesting that it could be a signature of chiral gravity [1208, 1209].
The physical origin of this asymmetry and its implications for the inflationary model remain open questions.
The Cold Spot is another intriguing anomaly – a large, unusually cold region in the CMB, surrounded by a hot ring
[1210–1212]. Both the WMAP and Planck missions identified it as a significant deviation from the Gaussian random
field expectation (∼ 3σ), and its region has been studied extensively. To explain it, novel cosmological mechanisms
[1213, 1214], a cosmic texture [1215], a large void in the line-of-sight [1216], and a systematic effect [1210, 1217] have
all been investigated. Importantly, the Eridanus supervoid (R ≈ 200h−1 Mpc, δ ≈ −0.2) was detected at z < 0.3
aligned with the Cold Spot, based on galaxy counts from multiple surveys [1218, 1219]. This evidence is supported
by reconstructions of the cosmic velocity field [1220] and WL analyses by the DES team [1221], suggesting a causal
relation between these individually rare objects in the CMB and in the cosmic web. Yet, the statistical significance
of the Cold Spot and its potential origins remain active areas of research.
Another very significant CMB anomaly, which potentially could give rise to many of the above mentioned anoma-
lies, is the cooling of CMB photons around nearby galaxies. First discovered in Ref. [1222], the stacking of CMB
temperatures in areas around the halos in nearby z < 0.015 late type spiral galaxies showed lower CMB temperatures
compared to elsewhere on the sky. This was followed up in Ref. [1223] where it was shown that an attempt at mod-
elling the discovered cooling of CMB photons around galactic halos gives rise to several of the observed anomalies.
In particular, the hemispherical asymmetries and the cold spot arise naturally in such a scenario with the correct
directions on the sky. Correlations are found between the largest scales of the CMB and the nearby galaxy distribution
obtained from the 2MRS galaxy catalogue [1224]. In Ref. [1217] a detailed study of the galaxies in the cold spot areas
showed that the shape and size of the cold spot to a large degree could be explained by cooling of the galaxies in
this area. In particular, the nearby large Eridanus group of galaxies is located at this position. In Ref. [1225], the
cooling was found to correlate to the nearby cosmic density field. In Ref. [1226], the cooling is detected at the 5.7σ
significance level when looking at galactic halos in the most massive nearby cosmic filaments. Even when correcting
for the look-elsewhere-effect by allowing for different choices of galaxy properties, the detection is stronger than in any
of 10.000 simulated CMB skies. Finally Ref. [1227] shows that by masking the affected area on the sky, the estimate
of the cosmological parameters are not significantly altered. The origin of the cooling of CMB photons in galactic
halos is unknown, but in Ref. [1226], it is speculated whether DM could be involved.
The statistical independence and potential interconnections between these anomalies are crucial areas of investiga-
tion. For example, the relationship between the low quadrupole moment and the lack of large-angle correlations in
the CMB is of particular interest [1228, 1229]. The lack of large-angle correlations was first reported using COBE
data and later confirmed by WMAP and Planck, with all three missions providing evidence for a significant deficit of
correlations on angular scales greater than ∼ 60◦ . This anomaly has been studied in the context of cosmic topology
[1230, 1231], with some proposing that it could be a signature of a non-trivial topology of the Universe. The rela-
tionship between the lack of large-angle correlations and the low quadrupole moment has also been investigated, with
some suggesting that they may share a common origin [1232].
Collectively, these anomalies pose a significant challenge to the standard model of cosmology. While some of
these anomalies may be the result of foreground contamination or systematic effects, their persistence across multiple
observations and their statistical significance suggests that they may have a cosmological origin. The standard ΛCDM
model, which assumes a flat, homogeneous, and isotropic Universe, struggles to explain these anomalies, and their
existence may require modifications to the model or the development of new theoretical frameworks.
Upcoming observational missions, such as the Euclid space telescope and next-generation ground-based CMB exper-
iments, have the potential to shed new light on these anomalies [437, 681]. These missions will provide unprecedented
sensitivity and angular resolution, allowing for a more detailed study of the CMB and its anomalies. In particular,
the Euclid mission will provide detailed measurements of the large-scale structure of the Universe, which can be used
to test models of the early Universe and the origin of the CMB anomalies. Next-generation CMB experiments, such
as CMB-S4, will provide a significant improvement in sensitivity and angular resolution, allowing for a more detailed
study of the CMB power spectrum and its anomalies.
Several theoretical frameworks and models have been proposed to explain these anomalies, such as the role of
topology in large-angle CMB correlations [1231], the impact of a cosmological GW background [1233], and the effects
of superhorizon isocurvature DE [1205]. Other models, including anisotropic k-essence [1234], loop quantum cosmology
[1235], non-canonical anisotropic inflation [1236], and unexpected topology of temperature fluctuations [1230], have
also been explored. These models aim to provide a theoretical framework that can account for the observed anomalies
while maintaining consistency with other cosmological observations.
The development of new theoretical models and advanced simulations will also be crucial in understanding the early
Universe’s complexities. Theoretical efforts to explain the CMB anomalies will require a deep understanding of the
88

physics of the early Universe, including the dynamics of inflation, the generation of primordial perturbations, and the
evolution of the Universe in the presence of DE and DM. Advanced simulations, such as those performed with the
Planck Sky Model [1237], will be essential in understanding the impact of foreground contamination and systematic
effects on the observed CMB anomalies.
In the early 2000’s in the program of investigation of the selection of the initial conditions of our Universe [1238–
1241], authors proposed to allow the wave-function of the Universe to propagate on the landscape of string theory
and used quantum cosmology to derive the probability of our origin. This was the first work where the answer
was derived from an underlying fundamental theory. It showed that in contrast to previous beliefs, the most likely
Universe to spontaneously come into existence is the one that starts at very high energy. The authors proposed to
use the quantum entanglement between branches of the wave function as a way to test the theory and, for the first
time to have a handle to test the existence of the quantum multiverse [1214, 1242]. Traces of earlier entanglement
are imprinted as anomalies in our sky. The authors calculated and predicted a series of seven anomalies, including
the Cold Spot, power asymmetry, alignments, and suppression of power in lowest multipoles, etc., all of which are by
now observed. Status of the predicted anomalies against observations showed perfect agreement [1243–1246].
As cosmology continues to evolve, the anisotropic anomalies in the CMB serve as a reminder of the importance
of critically examining our models and assumptions, driving us to deepen our understanding of the cosmos. The
existence of these anomalies suggests that our current understanding of the early Universe and the origin of cosmic
structure may be incomplete, and that new theoretical frameworks and observational efforts may be necessary to
fully explain the observed features of the CMB. As we continue to study these anomalies and their implications for
cosmology, we may uncover new insights into the fundamental nature of the Universe and the physical processes that
shaped its evolution.

2.3.4. Hints of dynamical dark energy in DESI baryon acoustic oscillations and beyond
Coordinator: William Giarè
Contributors: Anton Chudaykin

DESI has recently released data from its first (DR1)[698, 744, 782, 808, 1247] and second (DR2)[716, 1248–1250] year
of BAO measurements, based on observations of tens of millions of extragalactic objects, including galaxies, RGBs,
and Lyman-α forest tracers. These DESI BAO observations provide precise constraints on the transverse comoving
distance, the Hubble rate, and their combination (all relative to the sound horizon at the drag epoch) across seven
redshift bins in the range 0.1 < z < 4.2.
One of the most notable results from DESI BAO observations concerns the nature of DE. As initially highlighted by
the DESI collaboration’s DR1 results [698, 1247] and recently confirmed by DR2 results [716, 1251], the combination of
DESI BAO with CMB data from the Planck satellite and SNIa distance moduli measurements from three independent
SNIa samples (i.e., the Pantheon-plus catalog [32, 33], the Union3 compilation [1252], and five-year observations from
DESy5 [1253–1255] ) provides moderate to strong evidence for a time-evolving DE component, commonly referred to
as DDE. One important aspect of this preference is that the data indicates a phantom crossing scenario, in which
the dark energy equation of state crosses the w = −1 [1256]. While this behavior presents theoretical challenges
for most simple scalar-field models, it can be realized in multi-field scenarios, alternative models of gravity [1257–
1259]. The DESI preference for an evolving DE has been explored in the context of general scalar-tensor Horndeski
theories [1260–1263] and in modified gravity models with non-minimal coupling [1261, 1264].
Given the potential impact of these results on our understanding of the Universe, caution is essential, and thoroughly
testing the robustness of these findings is of paramount importance. Unsurprisingly, a significant portion of the
cosmology and high-energy physics community has actively engaged with this issue, clarifying and/or bringing up
several aspects and concerns surrounding this preference towards DDE. In the following, we review some of the key
results that have emerged from the consistency checks conducted over the past few months, highlighting both the
strengths and weaknesses of the observed preference, while stressing aspects that warrant further clarification.

• Parameterization of the Equation of State – Barring, for the moment, any potential systematic issues
in the different datasets involved in the analyses, a first key aspect that has undergone significant cross-
checking [808, 1264–1276] is the parameterization used to describe the DE EoS. Originally, the DESI col-
laboration parameterized the time evolution of the EoS using the linear CPL form, w(a) = w0 + wa (1 − a),
arguing that various combinations of data indicate a consistent preference for a present-day quintessence-like
EoS (w0 > −1) that crosses the phantom barrier (wa < 0). Specifically, within this parameterization, DESI
DR1 BAO measurements, when combined with Planck CMB and DESy5 SNIa, yield a preference for DDE
at a significance level of ∼ 3.9σ. This preference decreases to ∼ 2.5σ (∼ 3.5σ) when replacing DESy5 with
Pantheon-plus (Union3) [698]. Notably, within the same CPL parameterization, these results remain stable
when substituting DESI DR1 BAO with the latest DR2 DESI measurements. The latter not only confirm the
89

preference for DDE but also consistently strengthen it by approximately ∼ 0.3σ across all dataset combinations
with BAO and SNIa. Although the CPL parameterization has been shown to match the background evolution
of distances arising from the exact DE equations of motion with about 0.1% accuracy for viable cosmologies
across a broad range of physics (including scalar fields, MG, and phase transitions see, e.g., Refs. [298, 1277])
alternative parameterizations that deviate from CPL at both z ≪ 1 and z ≳ 1 remain consistent with current
observations. In Ref. [1278], it was argued that assuming the CPL parameterization is not the primary driver of
this preference. When combining CMB, DESI BAO, and SNIa measurements within various EoS parameteriza-
tions, w0 consistently remains in the quintessence regime, while the constraints on wa indicate a preference for
a dynamical evolution crossing into the phantom regime (see also Ref. [1275]). This result holds for both DESI
DR1 and the latest DESI DR2 BAO measurements, as confirmed by the DESI collaboration in Ref. [1251]. In
this sense, the preference is to be considered robust against different models.

FIG. 41: Observational constraints in the w0 -wa plane (at 68% and 95% CL) assuming a CPL parametrization for
DE across different datasets. The left panel shows constraints from a combination of the Planck Plik likelihood for
high-ℓ TT, TE, and EE PR3 spectra, the Commander and SimAll likelihoods for low-ℓ temperature and polarization
(TT and EE), and Planck PR4 NPIPE combined with ACT-DR6 lensing likelihoods, along with DESI DR1 BAO
and Pantheon-plus / DESy5 SNIa data. The middle panel examines the impact of replacing DESI BAO with SDSS
BAO, while the right panel explores how the constraints shift when Planck is replaced with ACT-DR4 for tempera-
ture and polarization spectra and ACT-DR6 for lensing.

• DESI BAO measurements (and beyond)– Since the preference for DDE was first highlighted by the
DESI collaboration, the impact of DESI BAO DR1 data on this preference has been carefully scrutinized, with
systematic effects in DESI BAO measurements remaining a topic of debate, see, e.g., Refs. [1260, 1279–1282].
Notably, the DESI BAO measurement at z = 0.71 exhibits a ∼ 3σ tension with predictions from the Planck
best-fit ΛCDM cosmology, making it a key driver of several hints for new physics, including (part of) the
preference for DDE. However, it is crucial to emphasize that (i) the latest DESI DR2 BAO measurements not
only confirm the preference for DDE but also consistently strengthen it by approximately ∼ 0.3σ across all
dataset combinations with BAO and SNIa. While this increase in preference from DESI DR1 to DESI DR2 is
not highly statistically significant, it definitively reinforces the signal, suggesting that the initial preference for
DDE observed in DESI DR1 may not be merely a statistical fluctuation resulting from a preliminary one-year
observation. (ii) In combined analyses of CMB, BAO, and SNIa data, the preference for DDE is not exclusively
driven by DESI BAO, whether from DR1 or DR2. In fact, as reported in Ref. [1283], a similar shift toward
DDE is observed when replacing DESI BAO measurements with SDSS BAO. Planck CMB combined with
SDSS BAO and DESy5 SNIa indicates a preference for DDE at a statistical significance exceeding 2.5σ, see
also Fig. 41. Likewise, Planck CMB combined with SDSS BAO and Union 3 points to a DDE component at
≳ 2σ significance. The only combination where this preference is notably weakened below 2σ is when SDSS
BAO and Pantheon-plus SNIa are analyzed together with Planck CMB. Importantly, across all independent
datasets – each showing hints of deviation from a cosmological constant at varying levels of statistical significance
– a consistent shift in parameter space is observed in the same direction, namely towards a quintessence-like
equation of state in the present epoch that transitions into a phantom-like regime in the past. Overall, while
DESI BAO data undoubtedly strengthen the preference for DDE, the other datasets involved in the analysis –
most prominently SNIa distance modulus measurements – play an equally significant role in shaping this result.
90

For further discussions on BAO data, we refer to Sec. 2.1.19.


• SNIa– It quickly became clear that SNIa data play a major role in driving the preference for DDE. Excluding
DESI BAO and considering data combinations involving Planck CMB and either DESy5 or Union3 (i.e., without
including any BAO surveys) already leads to a preference for DDE at ∼ 2 − 2.5σ (see, e.g., Refs. [1283, 1284]).
Given the crucial role of SNIa measurements in this trend, the potential impact of systematic effects has been
widely investigated [522, 1285–1289], particularly for the DESy5 sample, which exhibits the strongest shift
toward DDE. For instance, Ref. [1290] highlighted the influence of SNIa measurements, reporting a cross-
correlation between the PantheonPlus and DESy5 supernova samples that suggested a calibration difference
of ∼ 0.04 mag between low and high redshifts. Correcting for this offset brings the DESy5 sample into closer
agreement with Planck’s ΛCDM cosmology. Since the parameter space favored by the uncorrected DESy5
sample diverges from many other cosmological datasets, it has been suggested that the apparent evidence for
DDE might primarily arise from systematic effects in DESy5 SNIa. In response to these concerns, the DES
collaboration in Ref. [1291] showed that the debated ∼ 0.04 mag offset between Pantheon+ and DESy5 at low
and high redshift is partly due to improvements in the modeling of supernova intrinsic scatter and host galaxy
properties in DESy5 (which account for up to ∼ 43% of the offset) and partly (∼ 38%) due to a misleading
comparison. The latter arises because different selection functions characterize the DES subsets included in
Pantheon+ and DES-SN5YR, leading to differences in individual supernova distance measurements due to
distinct bias corrections. Therefore, while SNIa data do play a pivotal role in shaping the preference for DDE,
these findings might warrant additional tests to carefully assess potential systematics before drawing definitive
conclusions. For further discussions on SNIa data, we refer to Sec. 2.1.7.
• Planck CMB data (and beyond): The final dataset considered in these analyses is the CMB angular
power spectra, measured by the Planck Collaboration. While a DDE component primarily affects the late-time
evolution of the Universe, it also has a non-negligible impact on the CMB spectra. The most significant effect
of DE dynamics on the CMB angular power spectrum is observed in the amplitude of the ISW plateau at
very large angular scales. This amplitude is mainly determined by primordial inflationary parameters (As and
ns ) and contributions from the late-time ISW effect, which is sensitive to DE dynamics (see also Sec. 2.3.9).
Specifically, a phantom (quintessence) DE component suppresses (enhances) the decay of gravitational potentials
that source the late ISW effect. Therefore, assessing the impact of Planck measurements on the preference for
DDE is certainly worthwhile. As discussed in Ref. [1292], Planck data – particularly the large-scale temperature
and E-mode polarization measurements – play a crucial role in reinforcing the preference for DDE reported by
the DESI collaboration. Several studies in the literature [192, 1293–1298] have highlighted that temperature
and polarization data at large angular scales (ℓ ≤ 30) drive a number of mild anomalies, including a latent
preference for a phantom-like DE component [192, 1293, 1294]. As discussed in Ref. [1292], temperature and
E-mode polarization anisotropy measurements at ℓ ≲ 30 are precisely the subsets of Planck data strengthening
the shift toward DDE, as well. Notably, when Planck’s large-scale data are excluded from the analysis or when
considering alternative CMB experiments that are independent of Planck, the preference for DDE diminishes
significantly. In these cases, no strong preference for DDE is observed when combining DESI with Pantheon-plus
SNIa data, and the ΛCDM model remains consistent within (or close to) the 95% confidence level results. For
instance, no convincing preference for DDE is found in any combination involving SPT data, whether using DESI
BAO together with Pantheon-plus or DESy5, or even when including Planck/WMAP large-scale temperature
and polarization measurements. Similarly, in analyses involving ACT data, a preference for DDE is observed
only when combining ACT with DESI BAO and DESy5 SNIa, while it gets diluted when replacing DESy5 with
Pantheon-plus, see also Fig. 41. However, when Planck large-scale temperature and polarization measurements
are combined with ACT small-scale temperature and polarization data, a preference for DDE (re-)emerges.
Long story short, in the CMB front of the analysis, the shift toward DDE is primarily strengthened by Planck’s
large-scale data, while other CMB experiments generally weaken the case for DDE. Taking a conservative stance,
this adds to broader concerns about potential systematic issues raised in other datasets discussed earlier.

2.3.5. Neutrino tensions


Coordinator: Gabriela Barenboim
Contributors: Adèle Poudou, Janusz Gluza, Mariana Melo, Matteo Forconi, Olga Mena, Rasmi Hajjar, Rishav
Roshan, and Stefano Gariazzo

Neutrino physics is in its Golden Age era. Neutrinos possess unique characteristics that set them apart. Firstly, they
are significantly lighter than other fermions by several orders of magnitude. Remarkably, no direct mass measurement
has yet provided evidence for a non-zero neutrino mass; for all detected neutrinos, E = pc within errors. Secondly,
91

they have their own cosmological background. Neutrinos are often described as “elusive” due to their nature, yet
they are incredibly abundant throughout the Universe. Their near absence of interaction is why we seldom detect
their presence. This relic neutrino background has never been detected directly. Nevertheless, neutrinos are hot
thermal relics, and their masses, even if tiny, affect the Universe’s evolution, having its largest impact on the growth
of structure at small scales due to the neutrino free steaming nature. Therefore, albeit indirectly, we can set an upper
bound on them by cosmological observations. Although official Planck CMB results reportP i < 0.24 eV [192], the
P
m
addition of LSS data in the form of BAO allows reaching mi < 0.09 eV [1299] or even mi < 0.072 eV from the
P
very recent DESI BAO survey observations [698, 1300], all at 95% CL. Similar bounds are obtained when different
likelihood approaches are adopted for analyzing Planck data [700], while the limits are relaxed when high-multipole
CMB data are considered instead of Planck [1301].
Nevertheless, the tightest limits to date are those reported in Ref. [1302] after the addition to Planck and DESI
observations of other background probes, such as CC, galaxy clusters angular diameter distances, and GRBs distance
moduli. For instance, the combination of CMB with GRBs and DESI BAO provides mν < 0.049 eV. The most
P
constraining bound is obtained when CMB, SNIa luminosity distances, DESI BAO, and all background probes are
combined: this limit is 0.043 eV at 95% CL. Background probes are therefore able to provide strong bounds on the
neutrino mass due to both the preferred higher mean value of the Hubble constant and the smaller errors on both H0
and the matter mass-energy density Ωm . Consequently, cosmological limits currently possess the highest constraining
power on neutrino masses if the standard scenario is assumed for neutrino decoupling and if neutrinos are massive
stable particles.
In the past, neutrinos were supposed to provide a possible explanation to the H0 and σ8 tensions, see e.g., [1303].
More recently, however, it has been shown that an increase in H0 due to neutrino physics is normally related to an
increase also in σ8 , see e.g., Fig. 34 in Ref. [192]. A solution to both tensions thanks to neutrinos seems therefore
unlikely.
On the other hand, neutrino oscillations measured at terrestrial experiments while being insensitive to the overall
neutrino mass scale, determine the value of two squared mass differences, the atmospheric |∆m231 | ≈ 2.55 · 10−3 eV2
and the solar ∆m221 ≈ 7.5 · 10−5 eV2 splittings [1304, 1305]. Since the sign of |∆m231 | is unknown, two mass orderings
are possible, the normal (NO) and the inverted (IO) orderings:

NO: m2ν + pm2ν + ∆m221 + p m2ν + ∆m231 ≥ 0.058 eV ,


 p p
X
mi = (2.19)
IO: m2ν + |∆m231 | + m2ν + |∆m231 | + ∆m221 + m2ν ≥ 0.101 eV .

Notice that not only some of the above-mentioned current cosmological limits are below the minimum sum of the
neutrino masses allowed in the inverted hierarchical scenario, but also some of them are below the minimum mass re-
quired from neutrino oscillation probes. Very interestingly, several of the possible data combinations imply a neutrino
mass limit at 2σ smaller than the minimum expected from oscillation experiments, i.e., mν ≲ 0.06 eV, pointing
P
towards a clear tension between cosmological and oscillation neutrino mass limits, see Ref. [1302]. Therefore, despite
the fact that such tension has so far not been present [1306], not only one but several sets of cosmological measure-
ments indicate a clear problem between these cosmological and terrestrial searches of neutrino masses. Cosmological
constraints, therefore, at present, reflect some tension regarding neutrino masses:

• Through cosmological observations, we are about to rule out the minimum value of mi allowed by neutrino
P
oscillations. It is important to remember that cosmology does not directly measure neutrino masses but rather
constrains the neutrino energy density, which is proportional to neutrino masses in the standard scenario.
This relation could be altered by the presence of new physics: current cosmological bounds may point to the
existence of very exotic cosmological scenarios (possibly related to dark sector physics) and/or non-standard
neutrino physics since these bounds are extremely robust within simple extensions of the ΛCDM model; see,
e.g., [1307]. Possible scenarios to relax the cosmological neutrino mass bounds include time-varying neutrino
masses [1308], decaying neutrinos [1309–1312], strongly-interacting neutrinos [1313, 1314], long-range neutrino
forces [1315], among others [1316–1320].
• Using Planck 2018, BOSS Lyman-alpha, andPDESI data, if negative P neutrino masses are allowed in the fitting
pipeline, a slightly better fit is obtained with mi < 0 than with mi ≥ 0 [1319]. It should be noted, however,
that the absolute sign of the mass of a fermion is a phase (i.e., is unobservable) and therefore this preference
for negative masses should be taken as an indication of an incomplete model and rather than an actual physical
measurement.

These tensions will need to be addressed in the coming years, and upcoming cosmological measurements, such as
those from e.g., future observations by DESI or CMB-S4, will be able to test all the current tensions (and possibly
additional ones; see below) while sharpening the cosmological neutrino mass limits.
92

FIG. 42: The sum of the neutrino masses as a function of the lightest neutrino mass for normal (red) and inverted
(blue) orderings. Cosmological constraints from CMB, CMB plus DESI, and including measurements from CC,
distance moduli from GRBs, and angular diameter distances from galaxy clusters are also depicted by the shaded
green and yellow regions.

Last but not least, the Karlsruhe Tritium Neutrino (KATRIN) experiment, a tritium beta decay experiment, may
directly discover the electron neutrino mass, finding a positive signal.12 Such a putative neutrino mass detection by
KATRIN will show an additional clear tension with current cosmological mass limits. As previously stated, possible
inconsistencies among laboratory and cosmological searches would definitely point to a much richer neutrino sector, to
a departure from the standard model of cosmology, or to a combination of both. Such a situation would offer exciting
possibilities to discover new physics. Neutrino physics has brought the first departure from Standard Model Physics
and may also be the first one to falsify the current standard cosmological scenarios where they are hot thermal stable
relics interacting exclusively via weak interactions.

2.3.6. Cosmic dipoles


Coordinator: M.M. Sheikh-Jabbari
Contributors: Dinko Milaković, Eoin Ó Colgáin, Francesco Sorrenti, Iryna Vavilova, Jenny Wagner, José Pedro
Mimoso, Laura Mersini-Houghton, Leandros Perivolaropoulos, Lu Yin, Maciej Bilicki, and Manolis Plionis

Cosmological principle, the assumption that the Universe around us at cosmological scales is (statistically or on the
average) homogeneous and isotropic, is the cornerstone of the modern cosmology, see Ref. [1323] for a recent review.
CMB temperature fluctuations [192] has been regarded as a very precise verification of cosmic isotropy at the surface
of last scattering at around z ∼ 1100. However, the existence of a very small deviation from isotropy in the CMB is
not ruled out theoretically [1324] and references therein or observationally [1323, 1325–1330] and references therein.
Such small anisotropies, if they exist and are not due to systematics, in principle can grow in time through the large
structure formation, especially when we enter nonlinear growth regime. It is therefore, prudent to thoroughly explore
evidence for traces of anisotropy especially in late cosmology.
Any physical observable may be expanded in terms of spherical harmonics on the celestial sphere. The lowest mode
is ℓ = 0 which preserves isotropy. The next mode is ℓ = 1, corresponding to dipoles of the physical observable in
question. Given that deviations from spherical symmetry are expected to be small, one expects higher multipoles to
have smaller contributions and hence harder to detect. On the other hand, the dipole component, which is the largest
mode beyond isotropy, is observer dependent and can be a kinematical effect. For example, it is well established
that the dipole in the CMB, which corresponds to ∼ 10−4 − 10−3 fluctuation in the CMB temperature (compared

12 The current limit is mβ < 0.45 eV (90% CL) [1321], and the expected sensitivity is 0.2 eV (90% CL) [1322]
93

to usual 10−5 fluctuations), can be attributed to peculiar motion of the observer compared to the “CMB frame”
that moves with the Hubble flow. So, exploring the dipoles in cosmic observables one should always note this frame
dependence. The very local Universe is obviously not isotropic around us. One can ask how far one should look to
be confidently in the Hubble/CMB frame. It is established that this distances is larger than ∼ 100 Mpc, as indicated
also by the convergence scale of the acceleration dipole due to matter fluctuations traced by galaxies and clusters of
galaxies distributions, first at Ref. [1331–1333] and later at Ref. [1334–1337]. The “bulk flow” in the Local Universe
is a collective phenomenon due to the peculiar motions of matter structures [1338] that instead of moving in random
directions, appear to follow an approximate dipole velocity flow, see Ref. [1339] and references therein. Existence of
bulk flows that are too large for ΛCDM expectations may be taken as a sign of departure from isotropy that shows
itself as a dipole in various cosmological observables.
2.3.6.a. Non-comoving cosmology, tilted cosmology Cosmological dipole is a frame dependent notion. To distin-
guish kinematical and a non-kinematical (not removable by the choice of frame) one needs to study usual FLRW
cosmological models in a non-comoving frame, see Ref. [1340–1344]. A related, but conceptually different, notion
is the “tilt” introduced by King and Ellis [1345]. Suppose we choose a comoving frame, a frame in which metric
does not have time-space off-diagonal elements, gti = 0. In this frame cosmic fluids may exhibit a momentum flow,
i.e., the time-space component of energy momentum tensor Tit may be nonzero. Tilt may be different for different
components of cosmic fluid, e.g., (pressureless) matter or radiation may have different tilts and hence there does not
exist a Hubble or CMB frame. Tilt is a dynamical variable in cosmology, like the shear or Hubble parameter, e.g.,
see Ref. [1324, 1346–1350] and references therein. This is in contrast with the non-comoving cosmology in which the
dipole components are not dynamical ones. The existence of tilts yields dipole anisotropy in various cosmological
observables.
2.3.6.b. Observational hints for cosmic dipoles We start by recalling that typical cosmological observables are
redshift (as a measure of distance), look-back time, Hubble expansion rate, deceleration/acceleration parameter,
luminosity distance, angular diameter distance, and number counts. In an anisotropic cosmology, these quantities
besides the redshift z also depend on line-of-sight. Redshift, the ratio of the frequency of the photon emitted by a
source and the one observed by the observer minus one, receives a Doppler shift if there is a bulk flow of velocity v:

(2.20)
p
1 + z = (1 + z0 )∆ , ∆ := (1 − v/c cos θLoS )/ 1 − v 2 /c2 ,

where 1 + z0 = 1/a is redshift when the source and observer have zero relative velocity and θLoS denotes the line of
sight angle. For more formal and detailed derivations and analysis see Refs. [1349, 1351–1356].
2.3.6.c. Ellis-Baldwin test A standard technique for searching for a dipole component in the distribution of
sources in the sky is the seminal Ellis-Baldwin test [1357]. The number N per solid angle of a local source in the sky
(say radio galaxies or RGB) in two frames moving with a relative velocity v and angle cos θLoS is given by

dN dN
= ∆2+k , (2.21)
dΩ obs dΩ rest

where k is a constant parameterizing astrophysical characteristics of the source population.


2.3.6.d. Cosmic microwave background dipole A dipole component in the CMB is customarily treated as purely
kinematical, due to the motion of the observer. This may be viewed as the definition of “CMB frame”, a frame in
which the CMB has no dipole component. While one can always find such a frame, only in an isotropic universe all
other cosmological observables are necessarily dipole free in the CMB frame. The heliocentric frame, in which our
cosmological observations are made, is moving with respect to the CMB frame with velocity v = (369.82±0.11) kms−1
along the direction (in degrees) (l, b) = (264.02±0.0085, 48.253±0.004) [192]. While the CMB dipole may be treated as
kinematical, there have been persistent CMB quadrupole and octopole components which may be hints of deviations
from the cosmological principle, see Refs. [1323, 1325, 1358] for reviews.
The Planck CMB data in the CMB frame may still show traces of a dipole once one makes a hemispherical split
analysis (HSA) of the data. This method involves splitting the celestial sphere into two hemispheres and fitting
ΛCDM to CMB data in each hemisphere. One then reads the values of H0 and Ωm,0 in each hemisphere, rotates
the hemisphere separation plane in the sky, and records the ratio of the difference between the two values of H0 in
each hemisphere, ∆H0 , over the mean H0 value over the whole sky, H̄0 . This analysis has yielded a maximal value of
∆H0 /H̄0 ≃ 10%, roughly along 45◦ to the direction of the CMB kinematic dipole [1329], a direction close to the axis
of the hemispherical power asymmetry, a well-recognized CMB anomaly [1325]. Intriguingly, this is the same level as
the discrepancy in Hubble tension; however see Ref. [1359].
2.3.6.e. Maximum Temperature Asymmetry (MTA) is another dipolar anomaly in the CMB which is identified
by examining the temperature differences between opposite pixels in the CMB sky map after subtracting the dipole
94

component. MTA selects a preferred axis by maximizing the temperature difference between these opposite pixels
and indicates a preferred direction in the CMB temperature fluctuations, signature of a dipole. See Ref. [1360] for
MTA analysis in the WMAP 7-year data. The proximity of the MTA axis to the α dipole, DE dipole, and dark flow
directions (see below) further strengthens the case for a potential underlying physical cause for these anisotropies.
2.3.6.f. Fine structure constant dipole Initially reported by Ref. [1361], was identified through the analysis of RGB
absorption spectra using the Very Large Telescope (VLT) and the Keck Observatory. The dipole suggests a spatial
variation in the fine structure constant α across the sky, with a dipole axis pointing towards (l, b) = (331◦ , −14◦ ), and
an amplitude A = (0.97 ± 0.21) × 10−5 . However, recent studies have reduced the significance of the α dipole to less
than 4σ [1358, 1362].
2.3.6.g. Type Ia supernovae dipole SNIa are widely accepted as standard candles with a fairly good distribution
over the sky [JLA, Pantheon(+), Union3, DES-5Y, ZTF]. The SNIa observations are of course made in the heliocentric
frame, correcting for the presumed (peculiar) motion w.r.t. the CMB frame, assuming that a Hubble flow frame exists
and that the far enough (farther than ∼ 100 Mpc SNIa) are already in the Hubble flow frame w.r.t us. This may
obscure possible dipole components in the data in the heliocentric frame [1363]. SNIa data may be explored for a
possible cosmological dipole component in 3 different kind of analyses:
(I) Ellis-Baldwin test for SNIa We still do not have enough observed SNIa to carry out this test. It may be feasible
in some years.
(II) HSA using a cosmographic expansion, dipole fitting method. For relatively low redshift data, z ≲ 1, one can
expand Hubble parameter and luminosity distance in powers of z,
 
cz 1
H = H0 1 + (1 + q0 )z + O(z 2 ) , 1 + (1 − q0 )z + O(z 2 ) . (2.22)
 
DL (z) =
H0 2
One may now consider a dipole in H0 , i.e., H0 (θ) = H0 (1+DH cos θ) or similarly in acceleration, q0 (θ) = q0 (1+Dq cos θ)
and perform HSA. This test has been carried out for JLA [1364, 1365] and Pantheon(+) [1343, 1363, 1366–1368]
datasets claiming for a small dipole in H0 and a notable dipole for q0 roughly along the CMB dipole in the heliocentric
frame. See also, [1339, 1369]. There have been analyses that claim not seeing a dipole see e.g., [33, 1370–1377]. We
need to wait for more SNIa data, e.g, the DES data and ZTF to confirm or refute these claims.
(III) HSA and dipole fitting of ΛCDM into SNIa data. This analysis started in 2010 in Ref. [1378] with the
Union2 SNIa dataset and identified a significant anisotropy in the accelerating expansion rate in direction (l, b) =
◦ ◦
(309◦ +23 ◦ +11
−3◦ , 18 −10◦ ). The anisotropy level ∆Ω0m,max = 0.43 ± 0.06 was consistent with statistical isotropy in about
30% of simulations. The alignment with other observations like bulk velocity flows and CMB low multipole moments,
fine structure constant, and DE dipole obtained through MTA method [1379] suggested a potential underlying physical
cause.
Similar recent
´ z approaches use the luminosity distance and distance ladder relation µ = 5 log DL (z) + 25 with
DL (z) = Hc0 dz ′ / 1 − Ωm,0 + (1 + z ′ )3 Ωm,0 and perform HSA. This has been carried out yielding a small variation
p

∆H0 ≲ few km s−1 Mpc−1 over the sky [1380], see also Refs. [1369, 1381–1385] and Refs. [1367, 1369, 1386] for further
analysis. It would be instructive to repeat a similar analysis with other datasets like DES 5Y [1369, 1384] or ZTF data
release. Alternatively, one may repeat a similar analysis with absolute magnitudes of SNIa (instead of H0 and Ωm,0 )
[1387]. Analyzing the latest ZTF data release, particularly at low redshifts, may yield further valuable information.
Ultimately, the combination of these different approaches and datasets will enhance our understanding of cosmic
anisotropies and contribute to a more comprehensive picture of the Universe’s large-scale structure.
2.3.6.h. Quasar (QSO) dipoles RGB catalogs contain about 1000 times more data points compared to SNIa
datasets, nonetheless, unlike the SNIa their standardizability [182, 1388, 1389] is still a matter of debate, see Refs. [523,
1390–1392] and references therein. The 3 anisotropy/dipole searches mentioned above using SNIa data, can be repeated
for the RGB samples. Ellis-Baldwin test for RGB has been carried out yielding a significant (up to 4.9σ) departure
from isotropy [1393]: RGB of redshifts z ≳ 1 are not in Hubble/CMB frame, while in the heliocentric frame they
move in the same direction as the CMB dipole, their peculiar velocity is few orders of magnitude bigger than the
CMB velocity 370 kms−1 , see also Refs. [1394, 1395]. This result is still under debate e.g., see Refs. [1396, 1397] and
references therein.
The HSA for RGB data with the dipole fitting method has also been carried out, in a cosmographic expansion
and/or using Risaliti-Lusso [1389] X-ray/UV flux relation. This has also led to a 2 − 3σ level result for a dipole
component in H0 [1398].
2.3.6.i. Radio Galaxy dipole Radio galaxies are typically located at cosmological distances and hence within the
standard cosmological principle paradigm, they are expected to be moving with the Hubble flow; they should not
exhibit a flow in the CMB frame and CMB frame should be their rest frame. This was first put to the test in
2002 [1399], confirming the expectation. However, the same observation has been repeated with better accuracy,
95

refuting the earlier result in about 3σ level [1400], finding radio galaxies moving with speed 4 times bigger than the
CMB velocity in the heliocentric frame, and roughly in the same direction. Similar results have been reported in
Refs. [1401–1408], see however Refs. [1409–1414].
2.3.6.j. Gamma-ray burst dipole GRBs may also be used to perform similar dipole tests as in RGB or SNIa.
However, a relative low number of data points in GRB samples and their still debated standardizability [541, 1415–
1417] (see Ref. [515] for a recent review) puts limits on using GRBs as reliable cosmological observables. See Refs. [1398,
1418] for the search of a dipole in H0 through an HSA within ΛCDM and GRB data, reporting 2 − 3σ deviation from
isotropy.
2.3.6.k. Cosmicflow-4 (CF4) bulk flow A measure for a possible (non)kinematic dipole may be inferred from
peculiar velocities of large catalogs of distant galaxies with known distances (which typically means cosmologically
close ≲ 500 Mpc). Such large distance, non-random and coherent in direction velocity fields, if they exist, are called
bulk flows. Bulk flows may be compared to the expected Hubble flow. CF4 (superseded CF2 and CF3) [1419]
complies ∼ 56000 galaxies to this end, while reporting H0 = 74.6 ± 0.8 ± 3(Sys) km s−1 Mpc−1 also reports their
peculiar velocities.13 The amplitude and alignment of the inferred velocity field from the CF4 data is at ∼ 2 − 3σ
discrepancy with respect to the ΛCDM model [1420]. While closer galaxies at around 50 − 100 Mpc seem to show
no excess or deficit in their velocity fields compared to expected Hubble flow, farther ones show sizable excesses (by
more than 3 times) [1421, 1422] and in a direction compatible with the CMB dipole, confirming similar anomalies
seen in CF3 [1423]. Similar bulk flows have also been reported from the low redshift part of Pantheon+ compilation
(0.015 ≤ z ≤ 0.06); these flows are toward the Shapley supercluster [1339] (see also Ref. [1369]).
A specific kind of bulk flow (as defined above) has been given a different name, dark flows. Dark flows are associated
with peculiar velocities of galaxy clusters that can be measured through fluctuations in CMB generated by the CMB
photons off X-ray emitting gas inside clusters [1424]. Dark flows are found in scales larger than 300 Mpc and the
dark flow dipole is a dipole found at the position of galaxy clusters in filtered maps of CMB temperature anisotropies
[1425] see also Refs. [1426–1428], see however, Ref. [1429]. It is hard to accommodate dark flow dipoles in the ΛCDM
cosmology.
2.3.6.l. Cosmological models accommodating the presence of cosmic dipoles As discussed, dipoles can appear in
various cosmological observables, such as the distribution of RGB and SNIa, distances, acceleration parameter, H0 ,
the fine-structure constant α. One can construct models that accommodate dipoles in some of these observables
and/or introduce mechanisms to yield large-scale inhomogeneities or anisotropies in the Universe. Below, we discuss
some of the prominent theoretical models that could explain the observed cosmic dipoles.
Off-Center observers in spherical cosmological scale inhomogeneities. Off-center observers situated in a spherically
symmetric inhomogeneous universe can see a dipolar anisotropy due to the observer’s position relative to the center
of the inhomogeneity. This idea is explored in models where large-scale structures or voids introduce anisotropies due
to the observer’s peculiar location within the cosmic structure [1430, 1431].
Primordial dipolar horizon scale perturbations. Primordial perturbations generated during the inflationary epoch
can also have dipole anisotropy that can be imprinted on the largest scales and re-enter the horizon at later times,
leading to observable dipolar anisotropies in the CMB and other cosmological observables. Such mechanisms can be
related to specific inflationary models that predict large-scale anisotropies [1202, 1432].
Nontrivial topology of the cosmos. Spatial topology of the Universe may play a role in creating cosmic dipoles.
Nontrivial topologies, such as those involving multiply connected spaces, can introduce preferred directions and
anisotropies on large scales which may be detectable in the distribution of galaxies, the CMB, and other cosmological
datasets [1230, 1433]. (Non-trivial topology may induce anisotropy through parity-violation without involving a dipole
[1327, 1434].)
Hubble scale topological defects. Hubble scale topological defect, such as a global monopole created by the DE scalar
field at recent cosmological times, known as topological quintessence, can be another source of dipolar anisotropy in
various observable (from the off-center observer viewpoint) [1358, 1379, 1435, 1436]. This scenario may also explain
the alignment of various observed dipoles, such as the RGB dipole, the SNIa dipole, and the α dipole.
Other physical mechanisms. Other physical mechanisms that can induce cosmic dipoles include:
- Anisotropic Dark Energy: Models with an anisotropic equation of state for DE can lead to directional dependencies
in the expansion rate of the Universe, thereby creating dipolar anisotropies [1437–1439].
- Large-Scale magnetic fields can introduce anisotropies in the CMB and other observables, potentially explaining
some of the observed dipoles [1440, 1441].
- Vector field models: Inflationary models involving vector fields can naturally generate anisotropic perturbations
[1442], leading to observable dipoles in various cosmological datasets [1443, 1444].

13 It is notable that this value for H0 has independent systematics from those based on cosmic distance ladder.
96

- Tilted models: Tilted cosmological models, where observers have peculiar velocities relative to the cosmic rest frame,
can lead to locally observed acceleration and dipole-like anisotropies in the deceleration parameter [1342, 1347–
1351, 1364, 1396, 1445–1448].
These theoretical models provide a rich framework for understanding the observed cosmic dipoles. Future observa-
tions and more refined data will be crucial in testing these models and determining the true nature of the large-scale
anisotropies in our Universe.
2.3.6.m. Conclusion and future directions We briefly discussed dipoles in various cosmological observables, from
the dipole in the distribution of astrophysical sources like SNIa, RGB, GRB and radio galaxies over the sky, to dipole
in the coherent flows of matter (clusters and superclusters), to dipoles in the cosmological model observables like
H0 , Ωm,0 . Besides the dipoles we discussed here, there have been reports on intrinsic (non-kinematical) dipole in the
CMB [1449–1453], the dipole in the fine structure constant [1361, 1379, 1454] and in distribution of baryon matter in
the Universe using FRB [1455, 1456]. Moreover, if the CMB dipole is interpreted as our departure from the Hubble
flow it will induce ‘time dilation dipole’, a directionally-dependent time dilation over the sky. Detection of time
dilation dipole can provide a new assessment of the cosmological principle [1457, 1458].
These dipoles, if of a cosmological origin and if not kinematical, hint to a breakdown of cosmological principle, the
very pillar of modern cosmology. Most of the dipoles discussed here may be individually of not a large statistical
significance, nonetheless, the synergy between them suggests that they may not be dismissed. Future data with better
sky coverage will be instrumental in establishing/refuting true cosmological dipoles. If established, one should take
more serious steps in accommodating a cosmological dipole in the cosmological models, first such steps have been
taken in [1347, 1349].

2.3.7. Big bang nucleosynthesis


Coordinator: Nils Schöneberg
Contributors: Dinko Milaković, Ismailov Nariman Zeynalabdi, John Webb, Luca Izzo, and Venus Keus

BBN describes the process of generating the atomic nuclei of light elements shortly after the big bang as a result of
the Universe cooling down beyond the point at which photo-disintegration of these nuclei becomes inefficient. Since
the heavier elements need to be built up from lighter components,14 the process of the nucleosynthesis only begins
when the Deuterium photo-disintegration becomes inefficient, which occurs when the photon temperature is around
T ∼ 100keV (corresponding to redshift z ∼ 5 · 108 ). For reviews see Refs. [1459–1462]. Not all stable elements are
readily produced from BBN due to the instability of isobars with 5 or 8 baryons. In particular, the stable elements
produced in significant proportion from BBN are hydrogen (1 H and 2 H), Helium (3 He and 4 He), and Lithium (6 Li
and 7 Li), with the remaining elements produced only in tiny proportions compared to later stellar and cosmic-ray
induced generation. As such, the primordial abundances of these heavier elements are almost impossible to measure,
given that already for the small abundance of Lithium the post-BBN generation presents a significant systematic,
see Sec. 2.3.7.a. Despite generally showing an excellent agreement between predicted and observed abundances, we
discuss hints at possible tensions for these different elements in Sec. 2.3.7.a–Sec. 2.3.7.c. Furthermore, BBN can play
a critical role in other tensions, see Sec. 2.3.7.d.
2.3.7.a. Lithium tension The two stable isotopes of Lithium (6 Li and 7 Li) are produced only in very subdominant
amounts, with 7 Li at a level 10−9 times below that of hydrogen (1 H). At face value, the prediction of the 7 Li abundance
from standard BBN is about 2-4 times higher than that measured in the atmospheres of metal-poor stars in the galactic
halo, for example see Ref. [1463] computes 7 Li/H=(4.94 ± 0.72) · 10−10 (for a CMB-motivated baryon density, see
also Refs. [1464–1468] for other such high computations) while Ref. [1469] measured 7 Li/H=(1.58+0.35 −0.28 ) · 10
−10
(see
Ref. [1470] for a higher measurement in the SMC, possibly related to nova enrichment [1471]). See Ref. [1472] for a
review. Naïvely this appears to be a large issue for standard BBN. However, there are a few caveats to take into
account.
The observations of the 7 Li abundances are extrapolated to zero metallicity (iron abundance), but the plateau
originally observed (the so-called Spite plateau) Ref. [1473, 1474] appears to not persist at lower metallicities (increased
scatter falling on average below the plateau), see for example Refs. [1472, 1475, 1476]. The issue is that Lithium can
be burnt up in stellar environments, possibly even in metal-poor stars. In particular, convective mixing in metal-poor
stars could bring the Lithium down to layers with a temperature above 3 · 106 K sufficient for further burning, see for
example Ref. [1477]. This stellar depletion argument is still under investigation, for a recent discussion see Ref. [1478]

14 The reason for this is that the direct n-particle fusion is suppressed by a power of ηbn where ηb ∼ 6 · 10−10 is the tiny baryon-to-photon
ratio. As such, subsequent progressive 2-nuclei fusion processes are more relevant for the fusion of heaver elements, which makes the
isobar stability gaps crucially important.
97

– In this case 6 Li would be burnt up and thus its abundance is used as a useful argument on whether such burning is
expected to have taken place. However, the observation of 6 Li is even more difficult due to its even smaller abundance
(and its ready generation through cosmic rays). Initial observations indicating a plateau at 6 Li/7 Li∼0.05 (BBN
predictions are at ∼10−4 Refs. [1460, 1464]) would indicate that stellar burning could not have taken place [1479–
1485], but newer observations currently only put upper bounds [1486–1489], overall making stellar depletion a likely
explanation. See Ref. [1490] for another argument towards stellar processes and Ref. [1491] for possible observations
of a truly primordial plateau.
Another argument could be made on the aspect of the nuclear reaction rates, though the rates appear to be
consistent, see Ref. [1478] for a discussion. As such, while nuclear rates are unlikely to be the cause of the Lithium
tension, stellar depletion could very well be an entirely sufficient explanation. Given this unclear footing of the
Lithium tension, any claims of a “required” modification of standard BBN should be seen with great caution. Despite
this, a range of non-standard BBN models have been proposed as possible solutions to the Lithium tension, such as
for example Refs. [1476, 1492–1502], and Ref. [1503] for a review.
2.3.7.b. Helium anomaly Primordial 3 He is remarkably difficult to measure [1462] (see also Ref. [1504]) since 3 He
is generally readily produced and depleted in stars. It is thus rarely used to constrain BBN. Conversely, 4 He is a
crucial part of the BBN predictions that can be readily measured in ionized metal poor extragalactic HII regions. Since
4
He is produced in stellar cores, the primordial Helium mass fraction YP is usually found by extrapolating the ratio of
4
He/1 H from the measured systems as a function of metallicity to zero metallicity, which is presumed to represent the
primordial value. These determinations from 214 systems [1505–1510] (which roughly average to YP = 0.245 ± 0.003
[1462]) generally agree extremely well with the value predicted from baryon abundance and neutrino abundance
determinations from the Planck satellite under the standard BBN assumptions (giving YP = 0.241 ± 0.025 [192]). The
recent determination from the EMPRESS survey on the Subaru telescope by Ref. [1511] returns a much lower value
(YP = 0.2370±0.0033) compared to all previous results, including that of Ref. [1509] (which gets YP = 0.2436±0.0040).
This is interesting because Ref. [1511] uses the entire sample 1 of Ref. [1509] and the only difference is the addition of
five additional extremely metal-poor galaxies (J1631+4426, J0133+1342, J0825+3532, J0125+0759, and J0935-0115).
Four out of five of these galaxies lie significantly below the expected Helium mass fraction (from the fit of Ref. [1509])
given their metallicity, see Fig. 43. Taking the measurements at face value, one would conclude a non-zero lepton
asymmetry [1512, 1513] and a slight preference for additional relativistic degrees of freedom [1511, 1512]. However,
such an analysis is yet to be independently confirmed.We also note that Ref. [1514] finds a higher-than-mean value.
2.3.7.c. Deuterium – A new tension? The final element that is produced in decent quantity from BBN is Deu-
terium (2 H, alternatively D). While the Deuterium measurements are largely consistent, see Fig. 43, its abundance can
also be seen to be in tension with the baryon abundance determined from CMB data, depending on which modeling
of the nuclear rates is performed [1466, 1515, 1516], thus requiring new measurements of nuclear rates to clear up the
posited tension. In particular, the crucial rates for Deuterium are
2
H + 2 H → 3 H + p+ (ddp) , 2
H + 2 H → 3 He + n (ddn) . (2.23)

Conversely, to probe the consistency with the CMB one can also check what nuclear rates would be preferred given
the baryon abundance and neutrino number from the CMB (see Sec. 2.1.18) [1515, 1517].
2.3.7.d. Impact of BBN on other tensions While BBN is extremely self-consistent, it can also be used to sup-
port other tensions. For example, the combination of BAO+BBN data provides one of the tightest non-Planck
determinations of a low H0 value, see for example Refs. [698, 704, 788, 1518–1522], recently reaching a precision of
H0 = 68.52 ± 0.62 km s−1 Mpc−1 [698], which puts a significant strain on solutions of the Hubble tension that focus
solely on the CMB. See also in particular Sec. 2.1.19. BBN data also aids in constraining dark radiation or varying
constant solutions, see Sec. 4.1.3 and Sec. 4.8 (although the latter is not tightly enough constrained to exclude it as
a solution to the Hubble tension, see Ref. [1523]).

2.3.8. Anomalies with Lyman-α measurements


Coordinator: Vid Iršič
Contributors: Simeon Bird

Similarly to how galaxies trace the matter distribution at lower redshifts, the IGM has been used to trace the matter
distribution at the time when first galaxies are forming (z > 2) (e.g., Ref. [1525]). The most well-studied observable
of the IGM is the Lyman-α (Lyα) forest – a series of absorption lines in the spectra of RGBs, that arise due to the
scattering of RGB light on the intervening neutral hydrogen atoms in the ground state. In the last two decades the
Lyα forest has been measured in large spectroscopic surveys to study the precise position of the BAO [744, 1526],
full-shape RSD [1527, 1528], as well as small-scale 1D clustering statistics [1529, 1530]. The latter relies on estimation
98

35 Planck (Fields+2019)
Abundance ratio He/H (number ratio) 0.10
0.28 PDG value
Measurements (PDG)
30

Ratio He/H (mass ratio)


0.09
0.26

25

D/H ·106
0.08 0.24

S1E (bestfit, this work) 20


0.07 S2E (bestfit, this work) 0.22
S1 (mean, Hsyu+2020)
15
S1E (mean, Matsumoto+2022) 0.20
0.06
EMPRESS
Hsyu+2020 (Sample 1) 0.18 10

0.05 Hsyu+2020 (Sample 2)


0.16
0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6
5 zabs
Abundance ratio 10 O/H

FIG. 43: Left: A comparison of the Helium abundance measurements from Ref. [1509] (Hysu+2020) to some recent
fits of Ref. [1511] (Mastumoto+2022). For the fits, S1 is sample 1, S2 is sample 2, and E denotes the EMPRESS
data. Right: A comparison of Deuterium abundance measurements from Ref. [1462](PDG) compared to the corre-
sponding summary value and the Planck prediction according to Ref. [1524] (Fields+2019).

of the 1D power spectrum along each RGB spectrum and then averaged over all the RGB spectra in the data sample
[1531, 1532]. This methodology has also been applied to smaller samples, with 10-100’s of high-resolution and high
signal-to-noise RGB spectra [1533–1536]. The highest precision constraints from the Lyα forest on the amplitude and
shape of the matter power spectrum come from the measurements of the 1D flux power spectra.
The cosmological parameter inference from the 1D flux power spectra relies on state-of-the-art cosmological sim-
ulations [1537–1540] as well as a data compression technique [1541]. This method assumes that at 2 < z < 4 and
small scales (< 10 Mpc/h) the Lyα forest is constraining only two effective parameters – the amplitude and slope of
the linear matter power spectrum at a pivot wavenumber and redshift, with the exact pivot point depending on the
survey. This data compression has been shown to be valid also for some of the extensions of the ΛCDM model, such
as cosmology with massive neutrinos or running of the spectral index [1542].
However, this methodology is typically valid when limited to scales > 1 Mpc/h or cosmological models that do
not drastically change the behaviour of the small-scale linear matter power spectrum (such as light DM models). In
those cases, at least one more parameter can be extracted from the data corresponding, for example, to the scale of
suppression of the matter power spectrum due to free-streaming nature of DM [1543]. This has been successfully used
to produce bounds on various types of alternative DM models [1544–1548].
While the ongoing DESI survey will provide updated constraints, the current most precise measurements of the 1D
flux power spectrum at 0.1 − 1.0 h Mpc−1 come from the SDSS-IV/eBOSS spectroscopic survey [1549]. A subsequent
cosmological inference analysis [1546] reported a tension with the CMB inferred cosmological parameters, suggesting
a hint of new physics in the form of a running of the spectral index. No non-statistical effects in the data analysis were
identified that could explain the signal and significant effort was invested towards mitigation strategies of systematic
effects. A later re-analysis of the eBOSS DR14 measurements used improved cosmological simulations and sampling
techniques [1550] and pointed to possible internal tension in the data between lower redshift (z < 2.5) and higher
redshift (z > 2.5) Lyα data. The analysis also proposed that the previous tension driving the constraints on the
running of the spectral index can be completely alleviated at the expense of the increased tension in the amplitude of
the matter clustering σ8 . The S8 tension between the amplitude of structure expected by the CMB and the amplitude
of structure detected by WL at lower redshifts has also been detected at z = 2 by the Lyα analysis of Ref. [1550],
with agreement between the measurements of the Lyα forest and WL.
Similarly, several analyses of high-resolution data pointed to a small tension between the CMB and Lyα forest
from the high-resolution observations of RGB spectra [1551, 1552]. A 2σ tension was reported between Planck
CMB analysis and high redshift (3 < z < 5.5) Lyman-α forest data from XQ-100/MIKE/HIRES in Ref. [1553].
The study argued that several systematic effects in the data, such as RGB continuum mis-estimation, residual high
column density systems not masked in the data selection, and astrophysical uncertainties in the mean transmission
of the high redshift Lyman-α scattering could potentially alleviate the tension. A further study combining both high
redshift XQ-100/MIKE/HIRES and low redshift eBOSS DR14 data [1554] showcased that the internal tension is
strongest in the compressed parameter space of amplitude and slope of the linear matter power spectrum at pivot
scale (kp ∼ 1 h Mpc−1 ) and redshift (z ∼ 2.5) as probed by the Lyman-α forest data. Both studies [1553, 1554]
99

further proposed that systematics in the data would pull in the direction of higher slope and lower amplitude (or
vice-versa). Study of Ref. [1554] then argued that for that reason EDE models are severely penalized in this part of
the parameter space and can thus be robustly excluded. A recent study by Ref. [1555] summarized these findings
and proposed a new physics model of mixed ultra-light axions to solve the tension between eBOSS and Planck CMB
data. While successful in explaining the results of the eBOSS survey, such a model would induce heavy suppression
of matter power spectrum on smaller scales < 1 h−1 Mpc, that would be detected high redshift data [1534, 1535].
These data sets access smaller scales and put stringent constraints on such on DM suppression, including mixed DM
models [1556, 1557].
The dependence of the Lyα 1D analysis on the cosmological simulations makes independent tests costly. Moreover,
the cosmological parameter inference is not immune to astrophysical effects. While certain physical properties of the
IGM are becoming well understood, and can be measured independently, such as the thermal history evolution [1558];
the impact of other effects, such as inhomogeneous helium reionization, on the 1D flux power spectrum is not yet fully
understood [1559, 1560]. Aside from the uncertainty of the astrophysics of the IGM, several observational systematic
effects can potentially explain the differences between the current measurements.
Nevertheless, these anomalies in the Lyα forest data could offer a fascinating hint of new physics that can be tested
and explored in a unique range of redshifts and clustering scales probed.

2.3.9. Cosmic superstructures and the ISW anomalies


Coordinator: Istvan Szapudi
Contributors: András Kovács, Christine Lee, Deng Wang, Maret Einasto, Mina Ghodsi, and Pekka Heinämäki

2.3.9.a. Galaxy superclusters The largest structures in the cosmic web are superclusters of galaxies - overdensity
regions which embed richest galaxy clusters connected by filaments of poor clusters and galaxy groups [1561–1563].
Galaxies and galaxy systems form due to initial density perturbations of different scales. Perturbations of a scale of
about 100h−1 Mpc give rise to the largest superclusters, the largest coherent systems in the Universe [1564]. The sizes
of the richest and largest superclusters are up to almost 200h−1 Mpc [1563, 1565, 1566]. With their galaxies, groups,
clusters, filaments and gas, superclusters can be considered as miniature versions of universes. Therefore, they are
ideal laboratories to study the properties and evolution of various elements of the cosmic web [1567, 1568].
Superclusters as the connected overdensity regions in the cosmic web have been defined based on individual objects
(typically optical or X-ray groups and clusters of galaxies) or luminosity-density or velocity fields [1563, 1565, 1569–
1571].
The extreme cases of observed objects usually provide the most stringent tests for theories; this motivates the need
for a detailed understanding of various properties of the richest superclusters [1572]. Various properties of superclusters
can be used to test cosmological models, such as their abundance, masses, sizes, shapes, morphology, the properties
of their high-density cores and so on [1567, 1573]. Recent observational results (e.g., DESI Collaboration [698]) have
shown a tension in the Hubble value based on different observational probes, challenging the concordance model of
the DE equation of state. Superclusters, representing the grand finale of hierarchical merging events, evolve at the
forefront under the influence of two competing forces driven by DE and DM. They may provide a useful probe through
their superior mass and extent. For example, the abundance of rich superclusters at a given epoch and the stacked
signal through the ISW effect generated by large superclusters can be useful to constrain the DE equation of state
[1574–1578]. In what follows, we focus on supercluster imprint on the CMB, and on the properties of supercluster
planes, shells, and on the regularity in the supercluster distribution.
2.3.9.b. The ISW puzzle When photons cross the decaying potential well of a supercluster, they become hotter;
in voids, they become colder. This phenomenon is the ISW effect [1579]. The ISW puzzle consists of anomalous signal
levels detected when stacking the CMB aligned with supervoids and superclusters. The effect has persisted since 2008
[1576].
The linear contribution to the late-time ISW CMB TISW is integrated over the line-of-sight from the present time
(z = 0) to the surface of last scattering (zLS ) [1579],
ˆ zLS ˆ zLS
∆TISW a
(n̂) = 2 Φ̇ (n̂, χ(z)) dz = −2 a (1 − f (z)) Φ (n̂, z) dz , (2.24)
T 0 H(z) 0

where a is the scale factor, χ is the comoving distance, Φ is the gravitational potential, and H(z) is the Hubble
parameter. The second equality reveals the connection with the logarithmic growth function f (τ ) ≡ d ln D/d ln a [14].
While the analogous non-linear Rees-Sciama effect [1580] happens in all models, the ISW effect gauges the divergence
of growth history from the Einstein-de Sitter (EdS) model where it is zero.
The CMB stacked towards superclusters and supervoids creates hot and cold spots, respectively (c.f., Fig. 44).
The ISW amplitude relative to concordance expectations, AISW ≡ ∆Tobs /∆TΛCDM quantifies any anomalies when
100

AISW ̸= 1. The first measurement to stack superstructures [1576] found AISW ≃ 4 with the correct signs for supervoids
and superclusters.

FIG. 44: The first stacking measurement of WMAP towards SDSS LRGs [1576] from 2008. The significance ex-
ceeded 4.4σ with A ≃ 4. The anomalous signal has been reproduced with Planck CMB maps and in most wide-
angle LSS maps over different areas of the sky.

2.3.9.c. Summary of observations Later stacking measurements, mainly based on voids, are consistent with
AISW ≈ 4 − 5 [1575, 1581–1584]. Cross-correlating the CMB with large-scale structure is typically consistent with the
concordance model [1149, 1585], although with less statistical power from a given data set. Stacking superstructures
focuses on the highest signal-to-noise data in contrast with the unweighted average of the two-point statistics.
The BOSS redshift survey updated the void stacking measurements of the SDSS. Void finding techniques solving the
“void-in-void” problem by combining voids into larger structures confirmed the earlier excess [1586, 1587]. Algorithms
with no merging identified no excess [1588]. Stacking [1589, 1590] DES Year-1 and Year-3 photometric redshift data
sets further corroborated earlier results by Ref. [1576]. Since DES covers a different part of the sky than the original
SDSS, it lessens the likelihood of a statistical fluke. The combined DES and BOSS data sets [1587] yield the ultimate
result AISW ≈ 5.2 ± 1.6 in the redshift range of 0.2 < z < 0.9. Fig. 45 summarizes the principal observations of the
ISW tension.
2.3.9.d. The cold spot The stacking measurements draw attention to the largest anomaly of the CMB, the Cold
Spot (CS), an exceptionally cold of approximately 70 µK area centred on Galactic coordinates (l, b) ≃ (209◦ , −57◦ )
[1189, 1210, 1591]. Could it be due to the ISW effect? The Eridanus void, found by Refs. [1218, 1592] targeting an
extended underdensity aligned with the CS, is the likely cause. While the full CS signal with substructures taken
into account predicted by the concordance model AISW = 1 [1219] is not enough to explain observations, the excess
AISW ≈ 4 − 5 would suffice. Bayesian hypothesis testing overwhelmingly favours the hypothesis of the Eridanus void
causing the CS over chance alignment [1218].
2.3.9.e. AvERA and the ISW and Hubble puzzles The Average Expansion Rate Approximation (AvERA) model
[13, 14] simultaneously explains the excess ISW signal and Hubble constant anomaly, see Ref. [1593], where they
+1.05
fit H0 = 71.99−1.03 km s−1 Mpc−1 to SNIa. AvERA approximates emerging curvature models [1594–1596] with a
modified N -body simulation. AvERA reverses the order of averaging and solving the Friedman equation; a minor but
surprisingly consequential modification of the standard N -body algorithm. Its motivation is the Separate Universe
Hypothesis (SUH), stating that an under- or overdense region evolves like a universe with modified cosmological
parameters. Recent general relativistic simulations [1597] confirmed that emerging curvature in voids dominates the
late cosmic expansion history as predicted by AvERA.
The final Hubble constant in AvERA depends slightly on the coarse-graining parameter. Between the extreme
scales of the box size (no effect) and resolution (SUH breaks down), the final results are only mildly sensitive to
coarse-graining: an order of magnitude change will alter the results by a few percent. The best fit, expressed as a
mass scale of ≃ 1 − 2 × 1011 M⊙ , reproduces the locally observed higher Hubble constant [13].
The AvERA expansion history is similar to concordance expectations with minor differences in detail. By definition,
it starts on the same trajectory as a ΛCDM or EdS model at high z. Around z ≃ 4, the initial collapse of high-
density regions accelerates growth compared to the concordance model. At lower redshifts, z ≃ 1.5, the voids start
dominating. Their expansion stunts growth more effectively than Λ, thus the “void effect” explains the local higher
Hubble constant. The higher derivative also produces an excess ISW effect [14] comparable to observations. Thus,
the AvERA explains simultaneously the Hubble and ISW puzzles. In AvERA, the coarse-graining scale drives the local
Hubble constant. While boosting the coarse-graining scale by a factor of two induces only a few per cent change in H0 ,
101

a reasonable choice of a few times 1011 M⊙ (Lagrangian) achieves perfect consistency with the local measurements.
The late complexity of AvERA expansion history affects S8 as well, although this has not been investigated in detail.
AvERA has a robust and surprising prediction: a sign change of the ISW effect above z ≃ 1.5 [14]. In contrast, the
concordance DE models predict an unmeasurably small ISW signal at the same redshift.

1.75 eBOSS DR16 QSO


eBOSS data 0.8<z<2.2
5 1.50

QSO
CDM 1.25

f(z) dlnD /dlna


0
TISW = 0
1.00
AISW

LRG
eBOSS 1.9 < z < 2.2
5 eBOSS 1.5 < z < 1.9 0.75 CDM
eBOSS 1.2 < z < 1.5 QSO voids* AvERA
eBOSS 0.8 < z < 1.2 0.50 eBOSS ISW
10 DES+BOSS 0.2 < z < 0.9
ELG DES+BOSS
CMB Cold Spot z 0.15 CMASS ISW
5 0.25 FastSound*
Tension [ ]

2 0.00 eBOSS DR16*


SDSS Ly
2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
0.2 0.5 0.8 1.1 1.4 1.7 2.0 z
z

FIG. 45: These figures from Ref. [1598] sum up the ISW tension as of today. Left: Observed ISW amplitudes in
different redshift bins. The excess ISW signal transitions around z ≲ 1.5 and changes sign. The bottom panel dis-
plays the significance of the deviations compared to ΛCDM predictions (1σ and 2σ correspond to shaded bands).
Right: Solid and dashed lines show the cosmic growth history in Planck 2018 ΛCDM cosmology and in AvERA
as used by Ref. [1149]. Points with an asterisk display f σ8 (z) constraints divided by the Planck σ8 (z) value. The
DES, BOSS, and eBOSS ISW anomalies have a consistent trend in terms of the re-scaled ΛCDM growth rate values
(AISW (z) × [1 − f ΛCDM (z)]). The CMASS [1588] and eBOSS LRGs ISW amplitude is not significantly anomalous.
Nevertheless, the eBOSS ELG, RGB, and high-z constraints from the FastSound and SDSS Lyα lean toward the
late complexity trend predicted by the AvERA.

2.3.9.f. Opposite sign ISW effect The eBOSS DR16 RGBs [1599] enabled the first look for ISW sign reversal
[1598]. The data cover a crucial redshift range 0.8 < z < 2.2 where ΛCDM and AvERA diverge: the first tends to
zero while the latter predicts a sign change around z ≃ 1.5.
The concordance model signal was estimated with a RGB mock catalogue from the Millennium-XXL for comparison
with stacking 800 supervoids of the eBOSS DR16 RGB catalogue. The excess signal AISW ≈ 3.6 ± 2.1 in the redshift
range of 0.8 < z < 1.2 is comparable with earlier observations. At 1.5 < z < 2.2, the AvERA-predicted opposite-
sign ISW signal emerged in 2.7σ tension with the ΛCDM. Despite the moderate significance of these measurements,
taken together with the excess at low redshift and the CMB CS, it suggests a more complex growth history than any
variations of the concordance model. The observed late complexity (c.f., Fig. 45) consistent with AvERA implies that
alternative models, such as emerging curvature, will explain the late expansion history of the Universe.
2.3.9.g. Supercluster complexes and planes In several cases the richest superclusters form complexes in which
several very rich superclusters are almost connected. These are, for example, the Sloan Great Wall at redshift
z = 0.08 and the BOSS Great Wall at redshift z = 0.47 with a length of over 250h−1 Mpc [1566, 1600–1602]. It has
been questioned whether the presence of several such rich complexes in the local Universe is in agreement with the
standard cosmological model [1576, 1603–1605].
Also, in the local Universe galaxy superclusters are arranged on huge planes which span many hundreds of Mpc in
space. Among these are the Local Supercluster Plane, discovered by Ref. [1606] in the distribution of superclusters.
Rich optical and X-ray clusters, as well as luminous galaxies and radio galaxies in the local Universe are located on
this plane [1565, 1607–1609]. Perpendicular to the Local supercluster plane, [1610] discovered in the distribution of
rich superclusters another plane - the Dominant supercluster plane. It is not yet clear whether the presence of such
planes are in agreement with our standard cosmological model. For example, using constrained simulations of the
local Universe in the 500h−1 Mpc box [1611] found that a plane with the extent of approximately 100h−1 Mpc can
102

be recovered in simulations. However, 0.28% of random realizations of simulations only matched both underdense
and overdense regions of the local Universe, and their simulation box was not large enough to test the presence of
a supercluster plane hundreds of Mpc long. Also, they did not analyse the probability of finding two perpendicular
supercluster planes. Refs. [1609, 1612] speculate that such planes may be a signature of long, nearly straight strings.
2.3.9.h. Regularity in the distribution of rich superclusters and supercluster shells Ref. [1565] analyzed for the first
time the spatial distribution of superclusters based on the all-sky catalogue of superclusters using various methods.
In their study [1565] discovered that local rich superclusters are arranged in an almost regular pattern with the
characteristic distance between superclusters 120 − 140h−1 Mpc, see also Ref. [1613] on the regularity periodogram of
the local structures. The power spectrum of rich clusters has a bump at this scale, and the correlation function has a
series of wiggles [1614–1616]. Rich superclusters are separated by giant voids or supervoids [1617, 1618]. Superclusters
are not arranged randomly in the cosmic web, and this makes projections smaller. As a result, supervoids are in some
cases connected and form very large voids, as the Eridanus supervoid. Such huge structures can leave their signature
to the CMB. These signatures are discussed in Sec. 2.3.9.b. In addition, Ref. [1619] detected a hint of the regularity
with the scale of about 400h−1 Mpc in the distribution of RGB systems at redshift range 1 < z < 1.8.
The Sloan Digital Sky Survey covers a part of the regular patterns of superclusters. Ref. [1620] showed that rich
superclusters in the sky region, covered by the SDSS, are arranged in six intertwined shell-like structures with the same
characteristic radius as found for the whole pattern of superclusters, approximately 120 − 140 h−1 Mpc [1620]. The
centres and walls of these shells are marked by rich galaxy clusters in superclusters such as the Bootes supercluster,
the Sloan Great Wall, the Corona Borealis supercluster, the Ursa Major supercluster, and other superclusters, see
also Ref. [1621]. The most prominent of these shells was recently determined also using the CosmicFlows 4 data, and
named “Ho’oleilana” [1622].
The huge sizes of these shells tell us that the origin of these shells, and regular pattern in the supercluster distribution
should come from the very early Universe. Ref. [1620] concluded that the process behind these shells is still unknown.
Ref. [1622] interpreted the shell which they named as Ho´oleilana as BAO shell. However, in the BAO shells the central
mass is always much higher than the mass in shell walls [1623]. Analysis of the luminosities and masses of superclusters
have shown that rich superclusters typically have mass-to-light ratios M/L ≈ 300 [1601, 1624, 1625]. Therefore, we can
estimate masses of superclusters in the Ho´oleilana using their luminosities provided by Ref. [1620] and this M/L value.
The mass of the central cluster in the Bootes supercluster, Abell cluster A1795 is M ≈ 6.6−11.2×1014 M⊙ [1626, 1627].
The mass of the Bootes supercluster itself is the lowest among superclusters in this shell, MBoo ≈ 0.1 × 1016 M⊙ .
Masses of superclusters in the walls of this structure are more than a hundred times higher than the mass of the
central cluster, A1795, of order of Mscl ≈ 0.2 − 3.5 × 1016 M⊙ , and their total mass is at least Mtot ≈ 25 × 1016 M⊙ .
Therefore, masses of superclusters contradict the interpretation of Ho’oleilana as BAO shell, as proposed in
Ref. [1622]. Ref. [1620] already provided arguments against BAO interpretation. Namely, the radius of the shell
is larger than ≈ 109 h−1 Mpc, the BAO scale, and shell is wide, being in the interval of 120 − 140 h−1 Mpc.
Ref. [1622] proposed that for the interpretation that this structure represents a BAO shell, the value of the Hubble
constant should be approximately 77 km s−1 Mpc−1 , see Ref. [1622] for details. However, so high value of H0 is very
unlikely. Therefore it is still, even thirty years after discovery, an open question whether the presence of such struc-
tures can be explained within the ΛCDM cosmological model, and which processes in the very early Universe are
behind this regularity.

2.3.10. JWST anomalies


Coordinator: Matteo Forconi
Contributors: William Giarè

One of the major challenges in testing the ΛCDM model is the lack of direct observations at high redshift (z ≳ 10).
Current probes, such as BAO and SNIa, do not reach this regime, making it difficult to study the accuracy of the
ΛCDM model around these epochs; this is particularly important considering that the large-scale structures we observe
today largely emerged in the early stages of the Universe’s evolution. The JWST, however, has been able to achieve
this goal, offering a glimpse into these distant epochs of galaxy evolution history. Because of the expansion of our
Universe, photons emitted by distant sources experience a non-negligible redshifting, such that UV light falls into
the infrared spectrum today. Moreover, the farther away the source (i.e., the earlier in cosmic time), the fainter the
observed signal. To address these observational challenges, JWST Near Infrared Camera (NIRCam) [1628] covers
a wavelength range of [0.6 − 5]µm, improving the coverage of its predecessor, HST, that was optimized for visible
wavelengths. Coupled with a primary mirror which is nearly three times larger than that of HST, JWST significantly
increases our ability to detect and resolve faint and very distant targets.
Recent JWST observations have identified a large population of photometric galaxy candidates [1629–1638]. Many of
these galaxies are very bright [1630, 1632, 1639], with UV LFs that differ from previous theoretical and observational
103

predictions [1634, 1640–1644]. They are found at remarkably high redshifts (z ≥ 12) [1634, 1645–1648] and some
exhibit extremely large stellar masses ( M ≥ 1010.5 M⊙ ) [1649–1653]. The unveiling of these extreme objects is
difficult to reconcile within the standard galaxy formation models and, by extension, the assumed underlying cosmology
(ΛCDM) [1654–1656]. Moreover, JWST’s ability to discover new SN candidates [1657–1662] introduces additional
prospects for tackling tension in cosmology [1663, 1664].
It is important to emphasize that these preliminary findings are from photometric data which, alone, are not enough
to draw definitive conclusions. A major source of uncertainty is the difficulty in differentiating early star-forming
galaxies from quiescent galaxies at lower redshifts [1665], as well as the spectral energy distribution templates used
to interpret photometric bands may be unsuitable for very massive galaxies in the early Universe - though updated
templates have been suggested, e.g., see Refs. [1666, 1667]. Nonetheless, comparisons of photometric and spectroscopic
redshifts in overlapping samples of galaxies with both measurements further validate these observations [1668–1675].
A way to alleviate this emerging tension is to reconsider the physics and assumptions behind the structure evolution
and star formation process at high redshift [1676–1687]. Because measured stellar masses strongly depend on the
Initial Mass Function (IMF), adopting a top-heavy IMF [1684, 1688, 1689] instead of the typical standard assumption
of a Salpeter IMF [1690] could shift the inferred stellar masses. Additionally, increasing the gas temperatures can
lead to a greater contribution to the brightness of early galaxies and consequently lowering the stellar mass estimates,
although feedback effects might compensate this mechanism [1691]. Another critical factor is the star formation
history and star formation efficiency ϵ. The results are highly sensitive to the model adopted [1692–1695]; a higher ϵ
can explain large early galaxies by accelerating the reionization process [1696–1700] though this must be reconciled
with constraints from lower-redshift data. Indeed, extremely high ϵ with minimal feedback is required to explain
bright z ∼ 16 JWST candidates [1701, 1702].
These possible explanations deal with the complex physics relating star formation to the evolution of DM halos at
high redshifts. However, emerging evidence suggests the issue might lie within our cosmological model. While it is
premature to draw any definitive conclusions from these preliminary observations, if neither of the aforementioned
possibilities can reconcile theory with the JWST data, it may be necessary to rethink the assumptions of the underlying
cosmological model itself.
A common way to quantify discrepancies with JWST data is through the Cumulative Stellar Mass Density (e.g.,
see Refs. [1629, 1649, 1650, 1703]), defined as
ˆ z2 ˆ ∞
dnh dV
ρ⋆ (M̄ ) = ϵfb M dM , (2.25)
z1 M̄ dM V (z1 , z2 )

where fb = Ωb /Ωm is the cosmic baryon fraction, ϵ is the efficiency of converting baryons into stars, V is the comoving
volume of the Universe between redshift z2 and z1 and dnh /dM is the mass function. For computing the mass function,
the halo multiplicity [1016] is needed and one of the most used is the Sheth-Tormen alternative [1017, 1118] due to its
theoretical motivations for collapsed halos [1704, 1705] and multiple successful N-body tests [1021, 1706] (even though
other phenomenological fitting mass functions exist [1707, 1708]). A further assumption often made is to use a Top-
Hat window function for smoothing the density field. In focusing on the cosmological framework, one typically sets ϵ
to a fixed value (for a conservative approach usually ϵ = 0.2 is enough). In principle, star formation efficiency can be
a function of the halo mass [1709] and further adjustments to star formation physics might be needed for more precise
computations [1653]. Nevertheless, for massive halos (e.g., CEERS observations [1649]), one can approximate ϵ as a
smoothly varying power law. If the mass range is short, this approximation does not affect greatly the assumptions
for the model. Lastly, the cosmic baryon fraction is usually considered, instead of computing the baryon evolution
in different halos [1710–1716]. If fb is not a function of the halo mass, it plays a role of a multiplicative factor and
a greater fb can only push the theoretical predictions towards the observed data points. All these methodological
choices and simplifications are widely used in the literature and allow to present conservative results that can be
directly compared with similar works following the same approach.
One strategy for reconciling JWST data is to boost the halo mass function at the tail, thereby producing enough
massive galaxies at early-times. This can be achieved with massive primordial black holes [1717, 1718] with a fractional
contribution of fPBH ∼ 10−3 , consistent with current constraints [1719, 1720]. Super massive primordial black holes
not only can accelerate the structure formation process but they could also explain the observed super massive
black holes that, otherwise, would have insufficient time to form from Pop III stellar remnants [1721–1729]. Other
approaches involve modifying the primordial fluctuations through, for instance, rare density fluctuations from the
inflaton field [1730], a blue tilted [1731, 1732] or modified [1733] primordial power spectrum, or also introducing
non-Gaussianities in the initial condition, affecting the abundance of DM halos [1734]. Barring self-interactions, the
properties of DM halos around galaxies imply that ma > 10−21 eV i.e. that this particle is indistinguishable from a
cold one on galactic scale [1735].
Axion-related solutions offer a further possibility for understanding JWST’s detection of unusually massive galax-
ies [1736]. For example, axion miniclusters can boost early galaxy formation [1737] though their required masses
104

exceeded super-radiance constraints. Alternatively, axions with masses between 10−22 eV ≲ ma ≲ 10−19 eV with
delayed oscillations, can produce efficient axion field fragmentation driving the formation of more massive galax-
ies [1738]. Introducing a fuzzy DM component, usually modeled with axions, in the structure formation process can
lead to suppression of small halos and galaxies, thus potentially alleviating tension if paired with altered star formation
efficiency [1739]. Yet another axion-like solution is found in axion quark nuggets [1740].
These JWST observations can also be interpreted by invoking PMF [1741] or cosmic string loops with a tension
below the limit imposed by CMB observations but above constraints from astrophysical probes (that can be avoided
by adjusting the effective low radius cut-off) [1742–1744]. Observed massive galaxies also allow us to revisit the
proprieties of Warm Dark Matter (WDM) particles [1745–1747]. Explored in the context of EDE, the fit to JWST
observations is improved with respect to the one in the minimal ΛCDM framework. In fact, EDE leads to an increased
ns and simultaneously raises the fraction of EDE, that allows for a larger number of halos and massive galaxies, and,
at the same time, achieving a non-negligible improvement toward easing the Hubble tension [1748–1751]. Another
possibility to increase the predicted cumulative stellar mass density in agreement with JWST observations, is to employ
alternatives in the DE sector such as phantom crossing in the DE equation of state [1752], a negative cosmological
constant [1753–1755] a sign-switching model [1756] or a simple CPL parametrization [1757]. Conversely, interactive
dark energy appears to exacerbate the tension [1748].

2.3.11. Cosmic voids


Coordinator: Dante Paz
Contributors: Alice Pisani, Carlos Correa, Cora Uhlemann, Maret Einasto, Mina Ghodsi, Nico Hamaus, Nico
Schuster, Sofia Contarini, and Umut Demirbozan

2.3.11.a. Properties and identification of cosmic voids Cosmic voids are vast underdense regions of the Universe.
Since their discovery [1758–1761], they have been recognized as powerful cosmological laboratories [1762–1765]. Hence,
they encode key information about the expansion history and geometry of the Universe [1765, 1766]. The statistical
properties of voids depend on two factors: (i) the matter tracers used to map the large-scale structure, namely, galaxies,
and (ii) the method used to identify them from the spatial distribution of these tracers. Regarding the first factor,
the non-trivial biased relation between galaxies and matter distribution in low-density environments has been studied
intensively in recent years [1617, 1767–1770]. Regarding (ii), there are different classes of void finders; for a comparison
see Ref. [1771]. Some methods exploit the watershed transform technique, see for instance Refs. [1772, 1773], others
are based on dynamical properties [1774], and some methods identify low integrated density in spherical or free-shape
regions, see for instance Ref. [1775]. Despite the intrinsic differences between all the methods, there is a consensus
on the basic statistical properties of voids. Roughly speaking, voids are underdense regions with densities as low as
10 − 20% of the average density of the Universe, their properties described here are almost exclusively governed by
gravity [1597, 1776–1778], and the motion of tracers around them is consistent with linear dynamics [1779, 1780].
They can have diameters starting from few Mpc, spanning several tens of Mpc and the sizes of the largest voids may
even exceed hundreds of Mpc [1565, 1605, 1781]. Such huge voids are delineated by the richest structures - galaxy
superclusters, connected by a hierarchical pattern of filaments within these large voids [1605, 1617].
2.3.11.b. Void size function as cosmological test The evolution of underdense regions in the Universe, particularly
isolated spherically symmetric density perturbations, involves several key stages. Initially, these regions expand until
internal matter overtakes the outer shells, transitioning from linear to non-linear growth. As voids expand, their
density decreases, forming spherical shapes with reverse top-hat density profiles, leading to super-Hubble bubbles
with suppressed structure growth and boundary ridges. There are two modes of this void evolution: the void-in-
void mode, where regions expand at all scales, and the void-in-cloud mode, where voids expand but surrounding
matter density leads to contraction and eventual collapse [1779, 1782]. The spherical expansion model, combined
with excursion set theory, can model the void size function (VSF)—the number density of voids as a function of their
comoving size [1782–1784]. The VSF depends on the power spectrum of fluctuations, the growth rate of structures,
and the critical density threshold when expanding shells cross. Using the cosmological model, we can predict the
abundance of voids of a given comoving size. By comparing this modeled function with observations of abundances
at different redshift bins, it is possible to conduct cosmological tests.
2.3.11.c. The void-galaxy cross-correlation function The cross-correlation function for voids and galaxies can also
be used to perform an Alcock-Paczynski test, similar to the galaxy correlation function at BAO scales. The one void
term can be interpreted as the stacked mean void profile and serves as a standard sphere at different redshifts to test
the Universe’s geometry and the relations between the angular diameter and the comoving distances. This profile, in
combination with a model for RSD, can be used to test cosmological parameters and gravity models [1785–1788].
105

FIG. 46: Confidence contours from void counts (light blue, [1791]) and void shape distortions (green, [1792]) of the
BOSS DR12 voids, and their combination (black) as independent constraints on Ωm , H0 and σ8 . Adapted from
[1791], licensed under Creative Commons Attribution 4.0 International (CC BY 4.0)

2.3.11.d. Redshift-space distortions in voids The assumption of spherical symmetry for the galaxy distribution
around voids breaks down in observations due to RSD acting along the line of sight. The general characteristics of RSD
include an elongation of inner correlation contours towards larger scales (due to void expansion) and a flattening of
the correlation levels at large scales due to the infall and lower expansion rates at the void outskirts. These anisotropic
patterns encode valuable information about void dynamics and the galaxy velocity field. The first model for RSD in
voids is the linear model, equivalent to the Kaiser factor in the galaxy autocorrelation [1787]. The pairwise velocity
distribution can also be described with the Gaussian streaming model [1779]. For the impact of RSD in VSF and
void-galaxy cross correlation see for instance [1789, 1790].
2.3.11.e. An independent probe in the landscape of cosmic tensions Recent studies have highlighted the potential
to test the standard cosmological model using voids in upcoming redshift surveys, such as Euclid [1793–1795]. In a
recent analysis presented in Ref. [1791], the void size function in the BOSS DR12 galaxies were examined. Through a
combined analysis of void abundances and void-galaxy cross-correlations, these findings proved to be both competitive
and compatible with cosmological constraints derived from other probes (see Fig. 46). Furthermore, recent studies
of the CMB lensing imprints of cosmic voids have become a valuable test for the cosmological model. In particular,
the BOSS survey reported up to 5.3σ [1796], DES reached up to 5.9σ [1797], the WISE–PanSTARRS combination
obtained 13.3σ [1798], and the DESI Legacy Survey DR9 achieved up to 17σ [1799], all in excellent agreement with
ΛCDM predictions. However, the achieved precision on the parameters does not yet allow for a decisive position in
the context of current cosmological tensions. Nevertheless, these results emphasize the significance of cosmic voids as
they provide an independent cosmological probe that comes at no extra costs in modern cosmological surveys.

2.3.12. Fast radio burst probes of cosmic tensions

Coordinator: Amanda Weltman


Contributors: Anthony Walters, Bing Zhang, Christo Venter, Shruti Bhatporia, and Surajit Kalita
106

FIG. 47: Illustrative diagram indicating key components of contributing to measured DM of FRBs.

2.3.12.a. Introduction to fast radio burst cosmology FRB observations are relatively recently discovered bright
transient events (typically observed for a few ms duration) detected in the radio spectrum [1800]. Of particular use
to us in cosmology, is the observation that they originate at cosmological distance, apparently from host galaxies at
these distances [1801], and thus they provide us with potentially very powerful probes of cosmology [1802, 1803] as
well as the IGM [1804, 1805].
To date, there have been approximately 800 observed FRBs with information available in the public domain. The
majority of these have been detected by the Canadian Hydrogen Intensity Mapping Experiment (CHIME).15 The
primary open problem of FRB science is understanding their origins and the progenitor mechanism at work [1806].
Compelling evidence has accumulated for a magnetar engine, as exemplified by the only Galactic FRB 20200428D
detected from a magnetar in the MW. On the other hand, some other engines for cosmological FRBs are still possible
(see Refs. [1806–1808] for detailed reviews on FRB progenitor mechanisms). Perhaps the greatest open puzzle is the
curious observation that while some FRBs appear to repeat, not all of them do, and there does not appear to be
any pattern or periodicity to so-called FRB repeaters. While there may be some selection effects at work, it is likely
that there are simply different classes of bursts and thus even more to discover about FRBs and with FRBs. Hence,
multi-wavelength observations or detections of GWs [1809] and neutrinos from FRB sites will be important to better
understand their progenitor mechanisms.
Recent searches for persistent radio emission from one-off and repeating FRBs identified a nearly unresolved source
for FRB 20190714A, with a peak brightness of 53 µJy/beam, marking the first detection of persistent radio emission
potentially linked to a non-repeating FRB. Multi-wavelength follow-ups in ultraviolet, optical, X-ray, and gamma-
ray bands set upper limits [1810], aiding the distinction between repeating and one-off FRBs. Persistent emission
modeling can constrain central source energetics and test magnetar-driven emission scenarios. Follow-ups also refine
host galaxy properties, localization, and redshift-dependent FRB characteristics, enhancing their use in cosmology.
2.3.12.b. Fundamental parameter inference from Fast Radio Bursts One of the key parameters in understanding
FRBs is the dispersion measure (DM), which gives information about the ionized plasma along the path of the light
ray. Broadly it gets contributions from each of the 4 environments along the line of sight, our Galaxy, its halo, the
IGM, and the host galaxy as illustrated in Fig. 47. The (1 +zS ) factor in the denominator of the last term accounts for
the observed value of redshifted DMHost . In general, DMIGM contains the information of the underlying cosmology.
The average DMIGM is given by the following equation (now known as the Macquart relation [1805])
ˆ
3cΩb H02 zS fIGM (z)χ(z)(1 + z)
⟨DMIGM (zS )⟩ = dz, (2.26)
8πGmp 0 H(z)

15 https://www.chime-frb.ca/catalog
107

where Ωb is the baryonic matter density, mp is the proton mass, fIGM is the baryon mass fraction in the IGM, and
χ(z) is the ionisation fraction along the line of sight. As ⟨DMIGM ⟩ contains cosmological parameters, this equation
can be used to constrain their values. For instance, utilizing a set of mock FRBs, DE equation of state for wCDM
cosmology was constrained [1811, 1812] and predicted that with O(104 ) localized FRBs, this constraint can be tighter
than the same obtained from BAO measurements.
FRBs have been observed over a relatively narrow redshift range to date, 100 MHz – 8 GHz [1813, 1814], and
partly this is most likely a result of observation bias, where we can more easily observe the bursts that are closest to
us. This has limited their potential in constraining cosmological parameters directly [1802], though using cosmology
as a prior allows for IGM constraints, specifically a potential solution to the missing baryon problem [1804, 1805].
Moreover, using a specific case of FRB 150418, earlier studies established a limit on the photon mass of mγ < 1.8 ×
10−14 eV c−2 [1815] and this bound has been continuously strengthened with the inclusion of additional FRBs [1816,
1817]. Furthermore, localized FRBs have been used to constrain the parameterized post-Newtonian (PPN) parameter
[1818] and other fundamental constants like the fine-structure constant and the proton-to-electron mass ratio [1819,
1820].
2.3.12.c. Hubble constant estimations using Fast Radio Bursts Several recent studies have leveraged FRBs with
measured redshifts to constrain H0 within a Bayesian framework. This approach compares the measured DMIGM
of localized FRBs with their theoretical predictions, yielding constraints on H0 and other cosmological parameters.
Ref. [1802] was among the first to investigate the utility of FRBs as cosmological probes through simulations. Their
mock FRB catalogue highlighted the challenge posed by DM variations due to IGM inhomogeneities. Notably, the
most significant improvement was observed in the Ωb,0 h2 parameter when combining FRBs with CMB, BAO, SN,
and H0 data. Additionally, they found that FRBs offered limited constraints on the DE equation of state, while the
curvature parameter (Ωk ) showed some improvement when combined with other data. Moving forward, Ref. [1805]
−1
leveraged 8 localised FRBs to constrain Ωb,0 = 0.051+0.021
−0.025 h70 where h70 = H0 /(70 km s
−1
Mpc−1 ), consistent with
values derived from CMB and BBN data. Another interesting work using strong lensing effect with 10 FRBs found
H0 ≈ 70 km s−1 Mpc−1 [1821].
More recently, as more FRB data are available, these constraints have been revised to a great extent. Ref. [1822] an-
alyzed 9 localized FRBs and found H0 = 62.3 ± 9.1 km s−1 Mpc−1 , assuming a homogeneous host galaxy contribution,
which may limit the complete validation of the result. Subsequently, Ref. [1823] classified 18 localized FRBs based on
host galaxy morphology and employed the IllustrisTNG simulation to estimate individual host contributions. This
−1
approach yielded H0 = 68.81+4.99
−4.33 km s Mpc−1 . However, their model-based host DM values were lower than the ac-
tual reported values for some FRBs. Moreover, Ref. [1824] included 16 localised and 60 unlocalized FRBs detected by
−1
ASKAP, estimating H0 = 73+12 −8 km s Mpc−1 . Their larger error bars, despite a substantial number of FRBs, likely
reflect the inclusion of systematic uncertainties. Moving beyond, Ref. [1825] incorporated a scatter in the Macquart
−1
relation with 78 FRBs (21 localized) and obtained H0 = 85.3+9.4 −8.1 km s Mpc−1 . Ref. [1826] employed a model-
independent method with 18 localised FRBs and found H0 = 71 ± 3 km s−1 Mpc−1 . Conversely, Ref. [1827] estimated
−1
H0 = 95.8+7.8
−9.2 km s Mpc−1 using 24 localised FRBs with a wide flat prior on DMHalo . Furthermore, Ref. [1828] com-
−1
bined 12 unlocalized FRBs with BBN data, resulting in H0 = 80.4+24.1 −19.4 km s Mpc−1 . Additionally, Ref. [1829] used
−1
a combination of 18 localised FRBs and the Pantheon dataset to estimate H0 = 65.5+6.4 −5.4 km s Mpc−1 . Ref. [1830]
propose a novel method using AGB architectures for 23 localized FRBs to estimate H0 = 67.3 ± 6.6 km s−1 Mpc−1 .
Furthermore, utilizing Bayesian analysis with different likelihood functions and distinct host distributions for 64 local-
ized FRBs, Ref. [1831] obtained H0 values well above 70 km s−1 Mpc−1 aligning with the late-Universe H0 values with
1σ error bars no longer overlap with those obtained from early-Universe measurements. More recently, accounting 98
localized FRBs, Ref. [1832] showed multiple estimates of H0 utilizing different methodologies.
As we see, FRBs do offer preliminary constraints on cosmological parameters, although they currently lack the
power to definitively resolve existing tensions in cosmology. While not as stringent as constraints derived from
established probes like CMB and SNIa, FRB-derived constraints are significantly tighter than those obtained from
GW observations. A primary contributor to the discrepancies in H0 values reported by different researchers lies in
the choice of models for IGM and host DM contributions, which remain largely unconstrained from observations.
Future surveys aiming to detect a larger number of FRBs, with a substantial fraction localized to their host galaxies,
hold promise for significantly improved constraints. These advancements could potentially contribute to resolving the
current Hubble tension.
2.3.12.d. Understanding nature of Dark Matter using Fast Radio Bursts Beyond constraining the Hubble con-
stant and other cosmological parameters related to the evolution of the Universe, FRBs hold promise for constraining
the fraction of primordial mass black holes made up of DM (fPBH ). In this context, strong gravitational lensing
of FRBs plays a crucial role. When light rays from FRBs pass near a massive object with mass ML , they can be
deflected, potentially producing multiple, time-delayed copies of the original burst. The time delays and magnification
108

(µ) of these lensed images provide an estimate of the source’s optical depth (τ ), which is linked to fPBH as follows
ˆ zS
3 H02 DL DLS 2 2 2
(2.27)

τ (ML , zS ) = fPBH Ωc dzL (1 + zL ) ymax (µ) − ymin (ML , zL ) ,
2 0 cH(zL ) DS
where DL and DS are respectively angular diameter distances to the lens object and the source from the observer,
DLS is the same between lens and source, ymin and ymax are minimum and maximum impact parameters, respectively.
Ref. [1833] pioneered this method, setting an initial constraint of fPBH < 0.08 assuming FRB lensing by black holes
exceeding mass M > 20 M⊙ . Subsequent studies incorporated FRB microstructure [1834] and extended mass functions
[1835] to refine these bounds. Notably, Ref. [1836] visualised using a null detection of lensed FRBs within a sample
of 110 real FRB observations. Further investigations explored the impact of intervening plasma, which mimics the
lensing effects. Ref. [1837] demonstrated this using 172 CHIME FRBs, highlighting the need to account for such
decoherence or scattering screens. More recently, Ref. [1838] analysed 636 FRBs from CHIME and the absence of
confirmed lensing events suggesting that MG might introduce a screening effect akin to the plasma scenario. In
summary, FRBs offer constraints on fPBH in the mass range of approximately 10−4 − 104 M⊙ and these constraints
are comparable or even surpassing existing experiments like OGLE, EROS, Icarus, and MACHO.
The landscape of FRB-based constraints is constantly evolving as new and more powerful radio telescopes come
online. Arrays such as Hydrogen Intensity and Real-time Analysis eXperiment (HIRAX) [1839], Deep Synoptic Ar-
ray (DSA)-2000 [1840], Bustling Universe Radio Survey Telescope in Taiwan (BURSTT) [1841] and Square Kilometer
Array Observatory (SKAO) [1842] promise to detect a significantly larger number of FRBs, including potentially
localised ones. This will undoubtedly lead to further refinements in our understanding of the Universe.

2.3.13. Radio background excess


Coordinator: Jack Singal
Contributors: Alan Kogut, Marco Regis, and Nicolao Fornengo

Throughout most of the electromagnetic spectrum the observed level of surface brightness and anisotropy of the sky
background radiation is roughly consistent with that expected from known emission mechanisms from astrophysical
and cosmological sources. This is the case for the backgrounds at infrared, e.g., see Ref. [1843], microwave, e.g.,
see Ref. [1429], optical/UV, e.g., see Ref. [1844], X-ray, e.g., see Ref. [1845], and gamma-ray, e.g., see Ref. [1846]
wavelengths. However, in the radio region of the electromagnetic spectrum, both the observed surface brightness level
of the photon background and the level of its anisotropy angular power are seemingly much larger than that which
can be produced by known source classes, resulting in a tension that is, as of now, an anomaly.
As recently summarized in Ref. [1847], with the known structure resulting from Galactic diffuse emission subtracted,
the surface brightness of the level of diffuse background radiation on the sky, from at least ∼20 MHz to 3 GHz, as a
function of frequency ν is, in radiometric temperature units
 ν −2.66±0.04
TBGND (ν) = 30.4 ± 2.6K + TCMB , (2.28)
310 MHz
where TCMB is the frequency-independent contribution of 2.725 K due to the CMB. This background level has been
measured by the Absolute Radiometer for Cosmology, Astrophysics, and Diffuse Emission 2 (ARCADE 2) [1848, 1849]
at the high frequency end, and several radio maps at lower frequencies from which an absolute zero level calibration
can be determined, including recently the Long Wavelength Array (LWA) [1850]. Because the spectral index of −2.6
is consistent with synchrotron radiation, this has been referred to as the radio synchrotron background (RSB). The
RSB surface brightness level is shown in radiometric temperature units and spectral energy density units in Fig. 48.
This level of surface brightness is several times higher than can be produced by known classes of radio sources in the
Universe. Studies of radio source counts show that the total emission from the known discrete extragalactic sources,
particularly AGN and star-forming galaxies, is around a factor of five lower than the surface brightness level of the
RSB [1856–1858]. Producing the surface brightness of the RSB with discrete extragalactic sources would require an
enormous number of a new class of low-flux sources [1856].
For lines of sight far away from the Galactic plane, the RSB is considerably brighter than the expected contribution
from models of Galactic diffuse emission [1859]. The possibility that the RSB originates from a larger spherical halo
of radio emission surrounding our Galaxy is highly disfavored by several considerations, e.g., see Ref. [1847, 1859]
including that it would make our Galaxy highly atypical [1860]. Observational constraints on the polarization structure
of the diffuse radio emission rule out the Local Bubble as an origin scenario [1861].
In a parallel situation to the surface brightness, the anisotropy level of the RSB, as measured at 140 MHz, is
also higher than that which could result from models of known source classes [1862, 1863]. The measured angular
power spectrum of the diffuse radio emission is a power law (in K 2 or “(∆T )2ℓ ” units) with increasing ℓ with index
109

FIG. 48: Left: The radio sky zero level in radiometric temperature units, reproduced from Ref. [1850], as measured
by several different instruments or surveys reporting an absolute zero-level calibration. Results are shown for AR-
CADE 2 at 3–10 GHz [1848, 1849], Reich & Reich at 1.4 GHz, [1851], Ref. [1852] at 408 MHz , Ref. [1853] at 45
MHz, and Ref. [1854] at 22 MHz, as well as several points reported by Ref. [1850]. Right: The photon backgrounds
in the Universe in units of spectral energy surface brightness density. Reproduced from Ref. [1855].

β = 2.17 ± 0, 08 from 700 ≲ ℓ ≲ 4000 (angular scales from ∼2’ to ∼20’), indicative of unclustered point sources,
diffuse sources, or a large number of faint, clustered point sources [1863]. However, its level of anisotropy power rules
out the former, given that the density of high flux point sources is constrained by source counts observations such as
e.g., [1857]. This leaves very numerous but highly clustered point sources or diffuse sources as the source classes that
can seemingly reproduce the level of the observed angular power spectrum of the RSB.
There are other observational constraints on possible origin scenarios for the RSB. Any possible source class and
emission mechanism must not:
• overproduce the observed level of the background radiation at far-infrared wavelengths, given the known correlation
between radio and far-infrared emission in galaxies, e.g., see Ref. [1864]. Any origin scenario involving galaxies or
their components would need this correlation to evolve with redshift [1865, 1866].
• overproduce the observed level of the X-ray background through inverse-Compton upscattering of the CMB and/or
optical/UV background [1865]. If the RSB indeed originates from a synchrotron process, this puts a lower limit on
the strength of the magnetic field in the sources from which the RSB originates of around 1µG at redshift z = 0. At
higher redshifts this lower limit increases by a factor of up to (1 + z)4 due to the increase in the surface brightness
level of the CMB at higher redshifts.
• overproduce the observed limits on the 21-cm absorption trough, due to the possible presence of the RSB at
redshifts z ∼ 10 increasing the temperature of the background relative to the temperature of the 21-cm transition,
e.g., see Ref. [1867].
• overproduce the observed cross-correlation angular power spectrum between the RSB and optical source catalogs,
indicating that the vast majority of its sources, if discrete, must be at z ≳ 0.5 [1866].
Some RSB origin scenarios that have been suggested include SN of massive population III stars [1868], emission
from Alfvén reacceleration in merging galaxy clusters [1869], annihilating DM in halos or filaments [1870–1873] or
ultracompact halos [1874], “dark” stars in the early Universe [1875, 1876], dense nuggets of quarks [1877], accretion
onto primordial black holes [1878–1880], decays to dark photons [1881], dark photon decays [1882, 1883], and other
injections of photons in the early Universe from processes [1884] including superconducting cosmic strings [1885] and
decays of relic neutrinos [1886]. Some of these models likely violate one or more of the observational constraints
discussed here, while others contain parameter spaces where they do not.
Upcoming observatories, and observations possible with current facilities, have the potential to further constrain
and test RSB origin scenarios. A measurement, such as with the space-based LuSSE-Night [1887] to determine if the
spectrum hardens below 10 MHz would be useful in this regard. The radio SZ effect [1888, 1889] when combined
with the well-known CMB SZ effect, would result in observed cluster radio emission having a null frequency at
700 ≤ ν ≤ 800 MHz, below which there would be an increment in the observed surface brightness in the direction of
a cluster and above which there would be a decrement, each on the order of ≲ 1 mk (both on top of the contribution
from the background RSB). A detection or lack thereof at the required sensitivity in clusters of various redshifts would
constrain the redshift(s) of the origin of the RSB. The radio background being far in excess, in both surface brightness
and anisotropy angular power, of that which is expected to be produced by known source classes and processes in the
110

Universe is an intriguing issue.

2.3.14. Tension between the large scale bulk flow and the standard cosmological model
Coordinator: Richard Watkins
Contributors: Hume A. Feldman

The Cosmological Principle requires that the motions of galaxies averaged over a sphere of radius R (bulk flow) should
go to zero as R becomes cosmologically large. Thus the determination of the Large-Scale Bulk Flow is an important
probe of this principle and of the Standard Cosmological Model more generally. Estimating the bulk flow requires
large catalogs of galaxy peculiar velocity measurements; in the last decade the quantity and quality of this type of
data has improved to where bulk flows can provide an important test of our understanding of the Universe.
2.3.14.a. Analyzing the bulk flow using the Cosmic Flows 4 catalog The Cosmic Flows 4 (CF4) peculiar velocity
catalog [1419] contains velocities and their uncertainties for 38,057 groups and individual galaxies; this represents
most of the peculiar velocity data in existence. Analyzing the bulk flow using this data presents several significant
challenges. First, we can only measure the radial component of the peculiar velocity. Second, peculiar velocities
typically have large uncertainties; objects in the CF4 catalog typically have uncertainties that are around 15% of the
redshift. This means that since distant galaxies have much larger absolute uncertainties, most of the information in
the catalog is concentrated nearby. Thus care must be taken to ensure that the bulk flow estimate is not dominated
by objects near the center of the volume. Finally, the CF4 survey has a very irregular distribution both radially and
on the sky. Thus analyzing the bulk flow using the CF4 requires a method that can “even out” the distribution of
the information in the volume of interest both so that the volume is uniformly sampled and so that radial flows do
not make spurious contributions. The latter is particularly important given the current tension in the value of the
Hubble constant. The use of an incorrect value of the Hubble constant in calculating peculiar velocities can result
in phantom radial flows that can contribute to the bulk flow if the volume of interest is not sampled in an isotropic
manner.
Ref. [1421] deals with these challenges by analyzing the CF4 catalog using the Minimum Variance (MV) method
[1423, 1890, 1891]. The MV method generates estimates of the bulk flow components, in a volume of radius R, that
are as close as possible to those that would be measured from an ideal (isotropic and well sampled) survey; thus it
balances the information in the catalog so that the volume is evenly sampled. In addition, a constraint is put on the
MV bulk flow estimates so that they are completely independent of the value of the Hubble constant. It should be
noted that even though the analysis uses only radial velocity measurements, if it is assumed that the velocity field is
irrotational (as it should be if it were generated via gravitational instability), then the MV method estimates reflect
averages of the full three-dimensional velocities.
2.3.14.b. The bulk flow tension Ref. [1421] determined that the largest radius for which the bulk flow could be
estimated with the CF4 catalog is around 200h−1 Mpc. Their bulk flow estimates as a function of radius are shown in
Fig. 49 and displayed in Table III. The figure shows both the estimates and uncertainties for the components of the
bulk flow (in galactic coordinates) as well as the magnitude. Included in the figure (in red) is the expected magnitude
of the bulk flow as a function of radius calculated using the Cosmological Standard Model. Contrary to expectations,
they find that the bulk flow magnitude increases with increasing radius. In addition, they find that the bulk flow has a
much larger magnitude than expected in the Cosmological Standard Model [192]. Indeed, they found the probability
of obtaining a bulk flow as large or larger in the Standard Model for R = 200h−1 Mpc to be 1.5 × 10−6 , equivalent to
about 4.8σ. Ref. [1422] find a similar magnitude for the bulk flow (also using the MV method) but estimate larger
uncertainties using cosmological simulation data. Ref. [1420] also analyze the bulk flow from the CF4 catalog using
a very different method and obtain a result that is consistent with the Standard model. However, it is difficult to
discern the sensitivity of their bulk flow to different scale motions. Thus the large-scale bulk flow potentially poses
a significant challenge to the standard model, but determining the strength of this challenge will require resolving
differences in analysis methods.

2.3.15. Ultra long period cepheids


Coordinator: Ilaria Musella
Contributors: Giuliana Fiorentino, Marcella Marconi, Roberto Molinaro, and Vincenzo Ripepi

The Ultra Long Period Cepheids are pulsating stars characterised by periods longer than ∼ 80 days with a mean
absolute magnitude in the I band −9 < MI < −7 mag. Thanks to their significant brightness they are observable at
very long distances, up to cosmologically interesting scales. Thus they could represent competitive standard candles,
not requiring the combination with secondary distance indicators [1892–1898] to reach the Hubble flow. On this
111

TABLE III: Summary of Bulk Flows for R = 150h−1 Mpc and R = 200h−1 Mpc. The uncertainties include both
the theoretical difference between the estimate and the bulk flow from an ideal survey and the measurement noise.

R = 150h−1 Mpc R = 200h−1 Mpc


Expectation (km/s) 139 120
Bulk Flow (km/s) 395 ± 29 427 ± 37
Direction l = 297◦ ± 4◦ l = 298◦ ± 5◦
◦ ◦
b = −4 ± 3 b = −7◦ ± 4◦
χ2 with 3 d.o.f. 20.19 29.84
Probability 1.54 × 10−4 1.49 × 10−6

400
ux uy
300
km/s

200
100

uz |u|
400
300
km/s

200
100

100 150 200 250 100 150 200 250


h 1Mpc h 1Mpc

FIG. 49: The green points with error bars show the bulk flow components and magnitude estimated from the CF4
catalog as a function of radius R. The error bars indicate the uncertainty in the estimates due to measurement
noise. The dotted blue lines show the theoretical standard deviation of the expected differences between the bulk
flow estimates and the bulk flow from an ideal survey calculated using the cosmological standard model (not includ-
ing measurement noise). The red dashed line indicates the theoretical expectation for the magnitude of the bulk
flow calculated using the Cosmological Standard Model. The figure is reproduced with permission from Ref. [1421].

basis, using ULPs might reduce the possible effect of systematic errors on the calibration of the extragalactic distance
scale and, in turn, on the local determination of H0 . Despite their expected important role in distance scale studies,
their application as standard candles is still challenging due to the relatively small sample of ULPs with well-sampled
light curves covering more than one cycle and some inconsistencies between theoretical evolutionary and pulsational
predictions and the observed properties. Indeed, the number of known ULPs is 73 observed in different galaxies. In
particular, only one has recently been discovered in our Galaxy in the Gaia DR3 catalogue. Their position in the
period-luminosity planes and the colour-magnitude diagrams suggest that these objects represent the extension at
higher luminosity and mass of the Classical Cepheids, but their mass-luminosity relation could be different according
to their evolutionary stage. As recently suggested in Ref. [1898], the consistency between ULP and Classical Cepheids
relations increases when the photometry accuracy improves, thus supporting the hypothesis that they are the same
type of pulsating variables but with different mass and period ranges.
However, at such high luminosity and mass levels, stellar evolution does not predict the occurrence of the blue loop
typical of the central helium-burning phase of intermediate-mass stars, so the population of the Instability Strip might
correspond to the first crossing toward the QSO. Moreover, the observed pulsation period provides a strong constraint
on the stellar mass at fixed temperature and luminosity. Current evolutionary tracks matching the observed position
of the ULPs do not always provide a combination of mass, luminosity and effective temperature consistent with the
observed periodicity.
The agreement between the ULP and Classical Cepheid Wesenheit relations tends to confirm the Hubble constant
112

value obtained based on Classical Cepheids, even if the remaining uncertainties on these objects’ evolutionary and
pulsational properties need to be solved and the dispersion of their Wesenheit relation has to be reduced to contribute
to the understanding of the Hubble tension. To this aim, future theoretical and observational investigations will
include different steps: to better understand the evolutionary status of the ULPs through a detailed comparison with
updated sets of evolutionary tracks and isochrones, to produce an extended grid of nonlinear convective pulsation
models covering the high mass and luminosity regime expected for these variable start to increase the number of
ULPs and improve their photometric accuracy. In this respect, exploring the Gaia database to find new ULPs will
be fundamental. Recently, the first MW ULP was found among the variables classified as Long Period Variables in
Gaia DR3 and there are likely other misclassified ULPs. In addition, the forthcoming Rubin-LSST survey will also
represent a fundamental opportunity to improve the photometry of the known Local Group ULPs and/or increase
the sample.
113

3. Data analysis in cosmology


Coordinator: Agnieszka Pollo

The volume and complexity of data available for observational cosmology has increased greatly during the last
years, and in the coming years the amount of data generated by different ground-based and space observatories and
experiments is expected to grow by orders of magnitude.
For example, the Early Data Release of DESI comprising 2 million spectra of extragalactic sources amounts to 80
TB, and it is only 2% of the expected full final DESI catalog [1899–1901]. Imaging data are even more voluminous: a
single exposure of the Dark Energy Camera (DECam) produces an image of a size of a gigabyte [1902]; and the latest
DES Data Release 2 [1903] was built upon 76,217 such single-epoch images. Similarly, one night of observations with
the Subaru HSC camera results in several hundred gigabytes of data [1904], which translated into ∼ 70 TB database
already with the first data release of the Hyper Suprime-Cam Subaru Strategic Program [1905], and several hundreds
of terabytes after the third data release [1906].
All these numbers are expected to fade soon with the expected arrival of Euclid [164], Vera Rubin Observatory
[893], or SKAO data.16 In particular, Vera Rubin Observatory is expected to deliver more than 500 petabytes of
imaging data per year, while for SKAOthe data volume is expected to exceed 700 petabytes per year.17
In addition to the expected multiwavelength and multimessenger observational or experimental data, there exists
a whole additional realm of data useful for cosmology, i.e., simulations. Simulated catalogues are being used to test
different cosmological scenarios, to probe mechanisms of baryonic interactions in the dark-matter-dominated Universe,
especially at small, non-linear scales; and finally, to create mock realizations of real catalogues which can be used
both for physical predictions and for tests of specific measurement techniques against a variety of biases introduced
by observational strategy and technique. Creation of simulated catalogs is usually a multi-level process, requiring
a lot of computational power and creating amounts of data much larger than the corresponding observational data
themselves.
Using all types of data expected in the near future, both observed and simulated, to provide reliable scientific
results, requires developing new approaches to data handling and analysis at different levels. Firstly, all these data
need to be properly stored and transferred. Ideally, they should also be made publicly accessible in formats that are
easy to handle for users from different scientific communities. Datasets observed at different wavelengths or coming
from different experiments need to be cross-correlated to facilitate multi-messenger studies. Finally, scientific analysis
of these data results in creation of derived data sets, in amounts often exceeding the original data. Storage, transfer,
and subsequent analysis of large amounts of data is not only costly but also requires a well thought technological and
strategic approach. ASTRONET Roadmap for years 2022-203518 indicates near-future big data as one of the biggest
challenges that astronomy in Europe will face in the coming decade and recommends the development of a “Tiered”
approach for data infrastructure, to optimize the use of available resources.
With new big data, unavoidably, come new methods, ideally fully automated and requiring little human intervention
(e.g., see Ref. [1907]). It is therefore not surprising that a majority of sections in this chapter deal with different
aspects of ML-based approaches to the data analysis, in order to speed up and optimize the science output, and to
probe all possible parameter spaces. However, with the growing popularity of these methods, new challenges arise.
An old saying popular in ML community says “rubbish in, rubbish out”; it concisely summarizes the fact that
the errors or biases that exist in the original data, also those unknown to the users, are unavoidably transferred to
the results. Informed application of any machine-learning methods requires in-depth understanding of the properties
and limitations of the data. A related problem is a common “black box” approach to ML applications. Insufficient
understanding of algorithms used may result in users not being aware of their limitations and applicability to a
given scientific problem and given data. The next big challenges are related to interpretability and reproducibility of
the results obtained with the aid of machine-learning-based methods. In particular, reproducibility strengthens the
importance of keeping codes and data open to the scientific community so that all the scientific results obtained can
be independently verified. However, it cannot be forgotten that storage - especially long term storage - and computing
power come for a cost, both monetary, and environmental. The latter aspect draws more and more attention especially
now, when commercial ML and artificial intelligence projects fuel an increasing demand for resources.
Observational cosmology has already become an extremely data-intensive endeavour, and the future large, complex
and interdependent data sets that will be generated by observatories, space missions, real and mock experiments,
theoretical model simulations and different types of numerical simulations will require new tools and approaches, and
close co-operation between different fields of science. It is safe to say that cosmology and extragalactic astrophysics
is now only at the beginning of this road.

16 https://www.skao.int/
17 https://www.skao.int/sites/default/files/documents/SKA-TEL-SKO-0001818-01_DataProdSummary-signed_0.pdf
18 https://www.astronet-eu.org/wp-content/uploads/2023/05/Astronet_RoadMap2022-2035_Interactive.pdf
114

3.1. Cosmology simulators and Markov chain Monte Carlo approaches


Coordinator: Jesús Torrado
Contributors: Alessandro Vadalà, Benjamin l’Huillier, Denitsa Staicova, Guadalupe Cañas-Herrera, Jenny G. Sorce,
Laura Herold, Matías Leizerovich, Matteo Martinelli, Ruth Lazkoz, and Susana J. Landau

Modern cosmology has entered an era of precision measurements, where extracting robust constraints on theoretical
models from increasingly complex datasets demands sophisticated statistical approaches. Monte Carlo (MC) methods
have emerged as the cornerstone of parameter inference in cosmology, allowing researchers to efficiently explore high-
dimensional parameter spaces and quantify uncertainties in a fully Bayesian framework. This section presents an
overview of the most commonly used MC techniques in cosmological analyses, their strengths and limitations, and
practical considerations for their implementation.
The goal of Bayesian parameter inference is to determine the posterior p(θ|d, M), which is the probability of the
parameters of model M taking some value θ, given the data d. The posterior can be related to the likelihood of the
data d being a realization of the model with given parameter values, i.e., L(d|θ, M), via the Bayes theorem
L(d|θ, M) π(θ, M)
p(θ|d, M) = , (3.1)
Z(d, M)
where π(θ, M) is the prior and Z(d, M) the model evidence. Since one typically cares about the constraints imposed
on the model parameters, which are independent of the overall normalization of the posterior, the model evidence is
often neglected as it contributes only a constant factor. However, as we will discuss below, this evidence becomes
important for model selection. In standard parameter estimation contexts, we usually drop the explicit conditioning on
the model M. When required, lower-dimensional constraints, e.g., Bayesian credible intervals, are easily obtained by
marginalization, i.e., integration of the posterior over the remaining
´ parameters. By construction, the model evidence
is the marginal likelihood over the full parameter space, Z(d) = L(d|θ)π(θ) dθ, determining the probability of the
data having been realized under the given model, whatever the values of its parameters (hence their usefulness for
model comparison).
There are various Monte Carlo (MC) techniques aimed at determining the posterior p(θ|d) by means of obtaining
a fair sample from it. Below we offer a short review of the most popular MC algorithms used by the astrophysical and
cosmological community, which is illustrated in Fig. 50. These approaches have diverse strengths, but also limitations,
which we refer to as “sampling problems” below.
Moreover, the choice of prior π(θ) can lead to unwanted (and sometimes unknown) effects like prior dependence or
projection/prior volume effects, which we refer to as “modeling problems” and will be discussed below in the context
of cosmological tensions.19
3.1.1. Markov chain Monte Carlo sampling algorithms
MCMC algorithms explore the parameter space through a proposal-acceptance mechanism, building a chain of
states, or points in parameter space, from which a (decorrelated) fair sample from the posterior can be extracted.
Robust convergence tests (usually based on the stability of a single Markov chain or similarity between parallel ones),
are crucial to avoid biased inferences and inaccurate error estimates, guaranteeing that the MCMC samples reliably
represent the target posterior distribution.
The family of MCMC algorithms encompasses a large number of different approaches. They can be classified
under different criteria: the strategy followed to propose new points, the acceptance-rejection test performed on the
proposals, the distribution from which they effectively sample (not always the target distribution), whether they
pre-condition the parameter space and how.
In the simplest approach, at each step θ, a new state θ ′ is drawn from a proposal density q(θ ′ |θ) and accepted with
probability
ϕ(θ ′ )q(θ|θ ′ )
 
α = min 1, , (3.2)
ϕ(θ)q(θ ′ |θ)
where ϕ(θ) is the target distribution, i.e., the posterior in a Bayesian framework. This test is called the Metropolis-
Hastings (MH) criterion. Rejected proposals increase the weight of the current state within the chain. A typical
choice for the proposal distribution is a multivariate normal θ ′ ∼ N (θ, ΣT ), with a variable dispersion covariance ΣT
tailored to parameter values at each step.

19 Of course, the modeling of the likelihood L(d|θ) is also often based on approximations and can lead to uncertainties. Since the likelihood
is the basis of all parameter inference, not only MC approaches, we will not discuss those here.
115

FIG. 50: A flowchart summarizing the main differences between different samplers.

While unsophisticated, this method can produce accurate inference for simple distributions in large dimensions,
but its efficiency depends critically on the choice of an appropriate proposal distribution, and otherwise deteriorates
quickly for highly correlated or complex distributions.
Efficiency can be increased by using the current chain (or parallel chains) to compute an affine transformation
that would decorrelate the posterior locally (or globally if sufficiently Gaussian). This affine transformation can be
chosen so that proposals involve the re-calculation of only parts of the likelihood pipeline [1908] (as implemented
in MontePython [1909, 1910] and Cobaya [1911]). Parallelization, together with these strategies, can make MCMC
well-suited for computationally expensive likelihood evaluations.
For more complicated posterior structures, such as curving degeneracies that would need a local transformation
to be decorrelated, ensemble methods such as emcee20 [1912] can use the information on the position of multiple
walkers to automatically adapt proposals. More recently, the authors of pocoMC21 [1913] have proposed the use of
Sequential Monte Carlo (i.e., sequentially sampling from a family of distributions that interpolate between the prior
and the posterior) while learning a generalized transformation that decorrelates arbitrary parameter spaces, using
Normalizing Flows [1914]. pocoMC can compute model evidences along MC samples.
Slice sampling [1915] is an alternative adaptive approach to propose new Markov chain states, that automatically
adjusts step sizes to the local shape of the distribution. For each parameter update, it samples uniformly from a

20 https://github.com/dfm/emcee
21 https://github.com/minaskar/pocomc
116

randomly-oriented “slice” under the PDF curve, Sy = {θ : y < ϕ(θ)}, where y is uniformly drawn from [0, ϕ(θ current )].
A proposal θ ′ is then uniformly sampled from this slice through a stepping-out and shrinking procedure. This adaptiv-
ity makes slice sampling particularly robust, requiring minimal tuning while maintaining good mixing properties. The
algorithm naturally accommodates different scales in different regions of the parameter space, though its efficiency
can decrease in higher dimensions. A modern implementation combining slice sampling with an ensemble approach,
with applications in cosmology and astronomy, is zeus22 [1916]. Slice sampling has also been used to increase the
efficiency and robustness of nested samplers (see below).
3.1.2. Hamiltonian Monte Carlo
Hamiltonian Monte Carlo (HMC), also known as hybrid Monte Carlo, is an MCMC approach that incorporates
gradient information through an augmented parameter space with momentum variables. The system evolves according
to Hamiltonian dynamics
1
H(θ, p) = − ln ϕ(θ) + pT M −1 p , (3.3)
2
where M is a mass matrix. In this approach, proposals for accepted/rejected new states are obtained by letting the
current state evolve according to an energy-conserving trajectory for a set amount of time – in practice some number of
steps of corresponding of small but finite duration. By following continuous trajectories determined by local dynamics,
rather than diffusive random steps, HMC significantly reduces random-walk behavior, improving sampling efficiency
for posterior with complicated geometries.
A key challenge in HMC is selecting an appropriate integration time for the dynamic evolution of the steps, i.e.,
the number and duration of finite steps. The No-U-Turn Sampler (NUTS) [1917] refines this approach by adaptively
determining the trajectory length, ensuring efficient exploration without manual tuning. By dynamically terminating
paths upon noticing signs of retracing, NUTS avoids redundant computations while preserving detailed balance, thus
enhancing the efficiency, particularly in high-dimensional spaces.
In order to exploit the advantages of HMC, cosmological codes and likelihoods need to be able to compute derivatives
with respect to input parameters and intermediate observables, respectively. The most common approach is to build
these codes using automatically-differentiable numerical frameworks that are able to produce gradients for arbitrary
inputs. The most capable and popular one at the time of writing this paper is the Python library JAX [1918]. There
is a growing body of JAX-based cosmological and astronomical libraries (JAX-COSMO23 [1919], DISCO-DJ24 [1920]) and
likelihoods (candl25 [1921]), as well as machine-learning cosmological emulators (CosmoPower-JAX26 [1922]). Another
popular programming framework that allows for automatic-differentiation is julia,27 that has been employed for
example in LimberJack.jl28 [1923].
An implementation of NUTS that can be interfaced with JAX-based codes is numpyro29 [1924]. Using numpyro and
diverse JAX-based cosmological libraries and emulators, recent works [1925, 1926] have reported efficiency gains of up
to orders of magnitude, especially for parameter spaces of large dimensionality that are expected when performing
inference on data from next-generation surveys.
3.1.3. Nested sampling
Nested sampling [1927] transforms the evidence calculation into a one-dimensional integral over prior mass X
accumulated up to a given likelihood value
ˆ 1 ˆ
Z= L(X) dX with X(λ) = π(θ) dθ , (3.4)
0 L(θ)>λ

where the likelihood is evaluated such


Pthat L(X(λ)) = λ for some likelihood-cutoff value λ. Numerically, this integral
m
is computed as a weighted sum Z ∼ i=1 wi Li , with wi ∼ ∆X. The algorithm maintains a set of live points updated
at every step: the one with the lowest likelihood is dropped (and referred to as a dead point), and a new one is
sampled restricted to having a likelihood larger than that of the recently-dead point. The prior volume shrinkage
∆X is computed probabilistically and assigned the likelihood value of the dead point. In addition to the evidence
calculation, the pool of dead points may be used to construct a fair sample from the posterior.
The procedure to sample a point from the prior subject to a minimum likelihood, i.e., within an iso-likelihood
contour, is the main hurdle of the algorithm. Some implementations, such as MultiNest30 [293], construct a minimal

22 https://github.com/minaskar/zeus
23 https://github.com/DifferentiableUniverseInitiative/jax_cosmo
24 https://github.com/ohahn/DISCO-EB
25 https://github.com/Lbalkenhol/candl
26 https://github.com/dpiras/cosmopower-jax
27 https://github.com/JuliaLang/julia
28 https://github.com/jaimerzp/LimberJack.jl
29 https://github.com/pyro-ppl/numpyro
30 https://github.com/farhanferoz/MultiNest
117

ellipsoid containing the set of live points, and sample uniformly from an enlarged version of it. UltraNest31 [1928],
implements a different sampling approach based on ellipsoids centred around the live points themselves, and includes
more robust uncertainty estimation for the integrated evidence. Others, like PolyChord32 [1929], propose new points by
running a short affine-invariant MCMC chain (in particular using slice-sampling) from one of the current live points.
MCMC-based approaches have better dimensionality scaling, but tend to be slower for simpler problems. Other
modern implementations like dynesty33 [1930] dynamically vary the number of live points to improve efficiency. All
these approaches allow for efficient sampling of multi-modal distributions by using clustering algorithms to separate
and evolve different subsets of live points in parallel.
In general, nested samplers are preferred when the geometry of the posterior distribution is complicated or exhibits
multiple peaks. Due to its ability to compute the Bayesian evidence, nested sampling is also particularly valuable for
model selection problems. Nested sampling can however be computationally very expensive for highly dimensional
parameter spaces, but it parallelizes very efficiently up to the number of live points.
There have been in recent years various attempts at accelerating nested sampling using ML, focusing on efficiently
sampling within iso-likelihood contours. nautilus34 [1931] and neuralike35 [1932] train a neural network on the
spatial dependence of the likelihood in the vicinity of the current live set (only on the live set for neuralike) to
construct an estimate of the iso-likelihood contour, and reject points predicted to be outside it before the true
likelihood is evaluated. nessai36 [1933] takes a similar approach, training a normalizing flow on the set of live points
to find a transformation of the iso-likelihood contour into a much simpler hyper-sphere.
3.1.4. Practical considerations for choosing an MC sampler
The sampling methods discussed above differ in both their theoretical foundations and practical applications, each
offering distinct advantages and limitations in cosmological parameter inference. √
Traditional Metropolis-Hastings MCMC uses local random-walk proposals, leading to O( d) scaling in dimen-
sionality d. While simple and reliable, these proposals often require many likelihood evaluations, making them
computationally expensive, especially for high-dimensional parameter spaces or when the likelihood poorly constrains
the parameters. Non-Gaussian posteriors, particularly multi-modal distributions, can further challenge efficient pa-
rameter space exploration. A good dimensionality scaling can be retained for simple distributions using decorrelating
affine-invariant transforms. For more complicated, highly non-Gaussian and multi-modal distributions, ensemble and
sequential MC approaches can help, especially when in combination with generalized decorrelating transforms such
as NFs.
Slice sampling takes a geometric approach by sampling from level sets of the PDF, automatically adjusting to the
local structure of the posterior. This self-adaptive behavior results in O(d) scaling without requiring manual tuning,
making it effective for problems with widely varying parameter scales, such as fitting galaxy LFs where parameters
span multiple orders of magnitude. While computationally inexpensive in low dimensions, its efficiency can degrade
in high-dimensional spaces.
HMC improves upon this by using gradient information to make physics-inspired proposals, achieving better O(d1/4 )
scaling. This makes HMC particularly effective for problems with strong parameter correlations, such as hierarchical
models in galaxy clustering. However, it requires differentiable likelihoods and careful tuning of integration parameters.
Nested sampling transforms the problem into a one-dimensional integration over nested likelihood contours. Despite
its O(d3 ) scaling due to constrained sampling requirements when using simple acceptance/rejection sampling, it offers
important advantages: direct computation of the evidence Z (essential for model comparison in DE studies) and
natural handling of multi-modal distributions. This makes it particularly valuable in cosmology, GW physics, and
gravitational lensing analyses. A better dimensionality scaling can be achieved by using slice-sampling for constrained
sampling, or performing acceptance/rejection sampling on a surrogate model.
The choice of sampler often depends on specific requirements like parameter space dimensionality, likelihood char-
acteristics, and the need for evidence calculations [1934–1937]. For problems in which the likelihood is extremely fast,
O(< 10−2 s), and the dimensionality is of orders or a few 10s, simple approaches should suffice, such as affine-invariant
MCMC, or a non-boosted nested sampler if a model evidence estimate is needed. For harder problems (very high
dimensionality, strong non-Gaussianity, multi-modality), the use of the more advanced methods discussed above, such
as preconditioned Sequential Monte Carlo and ML-boosted nested samplers, can prove optimal. For extremely slow
likelihoods, O(> 10 s), the approaches discussed in Sec. 3.2 such as surrogate posterior methods or simulation-based
inference may be the only viable option, at the cost of some statistical robustness.

31 https://github.com/JohannesBuchner/UltraNest
32 https://github.com/PolyChord/PolyChordLite
33 https://github.com/joshspeagle/dynesty
34 https://github.com/johannesulf/nautilus
35 https://github.com/igomezv/neuralike
36 https://github.com/mj-will/nessai
118

Runtime vs Dimensions 1.0


Traditional Accuracy Metric 1.0
Wasserstein Distribution Accuracy
102

0.8 0.8

Wasserstein Accuracy
Traditional Accuracy
Runtime (seconds)

(higher is better)

(higher is better)
0.6 0.6
101

0.4 0.4

0.2 0.2
100

2 3 4 5 6 7 8 9 10 2 3 4 5 6 7 8 9 10 2 3 4 5 6 7 8 9 10
Dimensions Dimensions Dimensions

MH-MCMC (PyMC) (gaussian) Emcee (emcee) (gaussian) HMC (NumPyro) (gaussian) Nested (dynesty) (gaussian) Slice (custom) (gaussian) PolyChord (gaussian)
MH-MCMC (PyMC) (rosenbrock) Emcee (emcee) (rosenbrock) HMC (NumPyro) (rosenbrock) Nested (dynesty) (rosenbrock) Slice (custom) (rosenbrock) PolyChord (rosenbrock)

FIG. 51: Illustration of how different sampling methods scale with dimensionality and target distribution complex-
ity. The left panel shows runtime scaling with increasing dimensions, while the middle and right panels display
sampler accuracy using two different metrics. Results are shown for both a simple multivariate Gaussian distribu-
tion (solid lines) and the challenging banana-shaped Rosenbrock distribution (dashed lines). The traditional accu-
racy metric combines deviations from mean and function evaluation errors. The Wasserstein accuracy metric uses
optimal transport theory to measure distributional similarity between samples and theoretical distributions.

An illustration of the problem of dimensionality and the complexity of the distribution can be seen Fig. 51. On it
we show how the runtime and the accuracy scale with dimensions for a multi-dimensional Gaussian distribution and
the Rosenbrock distribution. For the accuracy test, we use the Wasserstein distance metric [1938], which measures
the true distributional similarity between sampler outputs and reference distributions. One can see that all samplers
show excellent performance on Gaussian problems across dimensions, but exhibit a dramatic accuracy drop-off for
Rosenbrock problems beyond 6-8 dimensions, reflecting the inherent challenge of sampling correctly from complex,
highly correlated distributions in higher dimensions.
3.1.5. Cosmological inference frameworks
There exist a number of public codes facilitating the integration of MC samplers such as the ones mentioned above
with cosmological inference pipelines: numerical codes to calculate cosmological observables (Boltzmann codes such as
CAMB [1939], CLASS [1940], and PyCosmo [1941] and their extensions, as well as emulators) and experimental likelihoods
for cosmological and astrophysical surveys and data.
The most used ones in the last years are CosmoSIS37 [1942] MontePython38 [1909, 1910] and Cobaya39 [1911] (which
uses GetDist40 [1943] for analyzing MC results). The choice for a particular one will be determined by the scientific
problem at hand, since each provides a different combination of features and integrates different theoretical codes and
experimental likelihoods.
Their use is thoroughly recommended since they facilitate not only the interfacing between the different parts of
the inference pipeline, but also help in specifying the models and priors as well as the configuration of samplers and
theoretical codes, and provide sane defaults for each. They usually either include in their source or automate the
installation of cosmological codes and likelihoods, as well as MC samplers.
3.1.6. Modeling problems: Prior dependence and projection effects
Analyses of present-day data require a significant modeling effort in order to achieve accurate theoretical predictions
that can be compared to data without biasing the results. This includes the modeling of complicated physics and
systematic or non-linear effects, which require introducing additional parameters (nuisance parameters) that need
to be kept free during the sampling of the corresponding posterior distribution. If the data is not sensitive to all
parameters in such a high-dimensional spaces, this can lead to a flat likelihood surface L(D|θ) in Eq. (3.1), which in
turn can result in a sensitivity to the choice of prior π(θ). If the prior distribution is not informed by past experiments

37 https://github.com/joezuntz/cosmosis?tab=readme-ov-file
38 https://github.com/brinckmann/montepython_public
39 https://github.com/CobayaSampler/cobaya
40 https://github.com/cmbant/getdist
119

or theory, the dependence of the results on π(θ) may be unwanted and needs to be explored in a sensitivity analysis.
A second – more subtle – dependence of the one-dimensional posteriors on the prior is introduced at marginalization.
Since the marginalization procedure takes into account the posterior volume, for very flat or very non-Gaussian
likelihood surfaces, there may be an (unwanted) up-weighting of regions with large (prior) volume. These so-called
projection or prior volume effects can lead to shifts of the marginalized posterior away from the maximum-likelihood
estimate or “best fit” point (see also Sec. 3.6), and may bias the constraints on the physics of the problem if the choice
of prior was not sufficiently informed.
An example of this in the context of the σ8 tension is the Effective Field Theory of Large Scale Structure (EFTofLSS)
(e.g., see Refs. [1173, 1174, 1944–1946]), a perturbation-theory based technique to describe the power spectrum of
biased tracers up to mildly non-linear scales, that is applied to LSS measurements. In such an approach, several new
parameters are introduced, as they enter the perturbative expansion of the galaxy density field. These parameters
will be free parameters in the context of an MCMC analysis and, therefore, require a choice of prior. However, the
arbitrary nature of the choice of such priors can become a problem for the analysis, since the inferred cosmological
parameters can depend on the particular choice of prior. It has been shown how different prior choices, e.g., extending
the range of allowed values to extreme intervals, can significantly affect the outcome of the analysis, with projection
effects shifting some parameters, such as the primordial amplitude As and the amplitude of matter fluctuations,
σ8 , with respect to their true values [1947, 1948]. It was further demonstrated that different – albeit theoretically
equivalent – choices of the EFT parametrization lead to discrepant inferred cosmological parameters [1948, 1949].
Ref. [1950] confirm a significant impact of priors on the nuisance parameters on the inferred cosmological parameter
using frequentist profile likelihoods. Approaches to choose well motivated priors on the EFT parameters include the
use of a Jeffreys prior [718, 1951] or simulation-inferred priors [1952, 1953]. Such an effect hinders the robustness of
the analysis results, as strong assumptions, only partially motivated by theoretical considerations, need to be done
on the choice of priors for the extra parameters.
A similar projection effect, shifting the values of parameters away from their true value, can be introduced also
when considering priors on the baryon abundance Ωb,0 h2 . Several nuisance effect modeled in LSS observables require
the value of this parameter, which is however not strongly constrained by LSS observations. For such a reason the
choice of prior for the abundance of baryons can be significant for the analysis, as switching from a uniform prior to
a Gaussian constrain as provided by BBN [797] can change the outcome in a significant way [1948].
In the context of the Hubble tension, an impact of prior effects on the inferred value of H0 has been reported and
explored for several extended models, for example: EDE (see Sec. 4.1.1) [1954–1960], NEDE (see Sec. 4.1.2) [1961,
1962], number of relativistic species (see Sec. 4.1.3) [1963–1965], Brans-Dicke model (see Sec. 4.3.1) [1966] decaying
DM (see Sec. 4.4.3) [1967] and more.
3.1.7. Model comparison criteria
The increasing complexity of cosmological models and datasets necessitates robust statistical frameworks for model
comparison and consistency checks. Information criteria and Bayesian methods provide quantitative tools for selecting
between competing models while accounting for model complexity and data support.
There exist several information criteria that can be used to compare models [1968]. Common criteria that bal-
ance goodness-of-fit of a model against its complexity are the Akaike Information Criterion (AIC) and the Bayesian
Information Criterion

AIC = −2 ln(Lmax ) + 2k , BIC = −2 ln(Lmax ) + k ln(N ) , (3.5)

where Lmax is the maximum likelihood, k is the number of free parameters, and N is the effective number of data
points on which the likelihood is defined. When comparing a test model versus a baseline one (e.g., ΛCDM) using
ICs (AIC and BIC), we compute the difference in criterion values as

∆ICtest = ICbaseline − ICtest .

The model with the lowest IC is preferred [1969], meaning, in this sign convention, ∆IC > 0 favors the test model,
while ∆IC < 0 favors the baseline model. The value of |∆IC| indicates the strength of the preference: |∆IC| ≥ 2
(weak), |∆IC| ≥ 6 (medium), |∆IC| ≥ 10 (strong).
The Deviance Information Criterion (DIC) extends AIC to hierarchical models by accounting for parameter uncer-
tainty:
h i
DIC = −2 ln(L(θ̂)) + 2pD , with pD = 2 ln(L(θ̂)) − ⟨ln(L(θ))⟩ , (3.6)

where θ̂ is the posterior mean of the free parameters, pD the effective number of them, and ⟨ln(L(θ))⟩ the mean
log-likelihood over the posterior samples. Unlike AIC and BIC, DIC uses the entire posterior distribution rather than
120

just the maximum likelihood estimate, making it particularly suitable for Bayesian hierarchical models where the
effective number of parameters may be less than the actual number due to prior constraints [1970].
The Bayes Factor (BF) provides a fully Bayesian approach to model comparison
´
Z1 L1 (θ 1 )π1 (θ 1 ) dθ 1
BF12 = =´ . (3.7)
Z2 L2 (θ 2 )π2 (θ 2 ) dθ 2
Jeffreys’ scale is commonly used to interpret ln(BF12 ) values: 0–1 indicates inconclusive evidence, 1–3 suggests
positive evidence, 3–5 shows strong evidence, and values larger than 5 represent very strong evidence in favor of
model 1 [1969, 1971]. While AIC and BIC are easily computed from MCMC chains, Bayes factors require accurate
evidence estimation, typically obtained through nested sampling or thermodynamic integration methods. As a fully
Bayesian approach, the value of the Bayes Factor depends strongly on the choice of prior density, unlike for information
criteria, so a prior sensitivity analysis is advisable.
3.1.8. Tension metrics
Tension metrics are statistical tools that are used to quantify the agreement (or lack thereof) between the estimation
of cosmological parameters obtained with different data sets, beyond the widely applied “rule of thumb”. According to
this rule the distancepbetween two posterior 1-d distributions can be quantified in 1-dimensional standard deviations
by Nσ = (µA − µB )/ σA 2 + σ 2 where µ
B A/B and σA/B refer to the means and variances obtained with each data set.
Most studied tension metrics can be grouped into two families: i) those based on Bayesian evidence [1972, 1973] and
ii) those based on the posterior distribution [1974–1976]. While group (i) answers the question about what hypothesis
is preferred by the data under the assumed model, group (ii) intends to establish the statistical significance between
the posteriors of data sets A and B within the parameter space of both experiments. For each metric, an estimator
is computed and a corresponding probability P is defined that quantifies the agreement or√disagreement between
data sets in terms of its corresponding equivalent 1-dimensional Nσ defined as: P = Erf(Nσ / 2). In most cases the
probability P is related to the probability-to-exceed (PTE) of an estimator Q that follows a χ2d distribution, defined as
ˆ ∞
PTE = χ2d (x) dx . (3.8)
Q

i) Metrics defined on Bayesian evidence (Z). The Bayesian ratio statistic R corresponding to two datasets A
and B, and the combination AB of both of them, can be written as [1972]
ZAB
R= . (3.9)
Z A ZB
R ≫ 1 is interpreted as both datasets being consistent, while R ≪ 1 means that the datasets are inconsistent.
However, it has been shown that the R estimator depends on the choice of the prior and can therefore underestimate
inconsistencies. This estimator has two primary contributions, one from the unlikeliness of two datasets being in
agreement, the information ratio I, and another from their disagreement, the suspiciousness S [1973]
ln I = DA + DB − DAB , ln S = ln R − ln I . (3.10)
Here D refers to the Kullback-Leibler divergence between prior and posterior, quantifying the information gain/compression
produced by the given data set. The suspiciousness estimator S remains unaltered under a change of the prior widths
as long as this change does not significantly alter the posterior. In addition, if the posterior is a d-dimensional
Gaussian, the quantity d − 2 ln S follows a χ2d distribution. Under that approximation, the probability of two datasets
being discordant by chance can be computed as the PTE of d − 2 ln S.
ii) Metrics based on the posterior distribution. First, we discuss some metrics based on quadratic estimators
that require Gaussian posteriors. For these, the probability P of the data sets agreeing is defined as P = 1 − PTE
where PTE is the probability-to-exceed defined above. Let us call θ̂ A the mean of the parameters inferred with
dataset A and θ̂ B the one obtained with dataset B, while Σ̂A and Σ̂B refer to their respective covariance matrices.
The method called parameter differences in standard form relies on the quadratic estimator [1974]

QDM = (θ̂ B − θ̂ A )T (Σ̂B + Σ̂A )−1 (θ̂ B − θ̂ A ) , (3.11)

This estimator follows a χ2ν distribution with ν = Rank[Σ̂B + Σ̂A ] and can be regarded as a generalization of the rule
of thumb. The method called parameter differences in updated form measures how one data set updates the other
and is based on [1974]

QUDM = (θ̂ AB − θ̂ A )T (Σ̂AB − Σ̂A )−1 (θ̂ AB − θ̂ A ) , (3.12)


121

Here θ̂ AB are the inferred parameters obtained with the joint datasets A and B, while Σ̂AB is the respective covariance
matrix. This estimator follows a χ2ν distribution with ν = rank[(Σ̂AB − Σ̂A )] degrees of freedom. Next, we describe the
Goodness-of-fit loss which evaluates the likelihood function at the maximum of the posterior of the joint distribution
and of each distribution separately [1974]
QDMAP = 2 ln LA (θ p,A ) + 2 ln LB (θ p,B ) − 2 ln LAB (θ p,AB ) , (3.13)
Here θ p,A/B are the Maximum a posteriori (MAP) parameters considering the dataset A/B. The estimator QDMAP
follows a χ2 distribution with ∆ν = ν A + ν B − ν AB degrees of freedom, where for a given data set/combination
ν = N − tr[Σ−1 π Σp ], with N being the effective number of data points defining the likelihood, and Σπ , Σp the
covariance matrices of the prior and the posterior, respectively. This metric is effective at evaluating how well the
theoretical prediction fits the data, but does not evaluate directly the disagreement between inferred parameters.

PPS (1) PPS (1)


Planck18 (2) CC (2)
PPS+Planck18 (1, 2) PPS+CC (1, 2)

DM Nσ = 6.420 DM Nσ = 2.772
UDM 1 → ©1, 2ª Nσ = 4.620 UDM 1 → ©1, 2ª Nσ = 4.778
UDM 2 → ©1, 2ª Nσ = 2.316 UDM 2 → ©1, 2ª Nσ = 2.471
DMAP Nσ = 4.218 DMAP Nσ = 1.704
EPS Nσ = ∞ EPS Nσ = 2.820
NΩm = 0.942 NΩm = 0.0171
NH0 = 6.093 NH0 = 1.899
76
75
74
70
72
H0

H0

70 65

68 60
66
0.30 0.34 0.38 66 70 74 0.2 0.3 0.4 0.5 60 65 70 75
Ωm H0 Ωm H0
FIG. 52: Tensions metrics applied to a) Planck 2018 vs Pantheon Plus + SH0ES (PPS), b) CC vs Pantheon Plus
+ SH0ES. In all cases, blue contours correspond to PPS and gray contours to the joint analysis. Nσ refers to the
1-d standard deviations that correspond to a 1-d PDF associated to each metric estimator that quantifies the dis-
agreement between different data sets (see the beginning of this subsection). QUDM quantifies how the parameters
inferred from one dataset are updated when another is incorporated into the analysis. QDMAP is related to the dif-
ference in goodness of fit to the different data of the underlying model. QDM and EPS quantify the difference in the
value of the inferred parameters. NΩm and NH0 quantify the tension in each inferred parameter applying the 1-d
rule of thumb, adapted from Ref. [1977]†† .

The next method, called exact parameter shift (EPS) does not require Gaussian posteriors and is based on the
parameter difference probability [1974, 1976]
ˆ
P (∆θ) = PA (θ)PB (θ − ∆θ) dθ , (3.14)

where ∆θ = θ̂ A − θ̂ B is the difference between the means of the posterior parameters that correspond to datasets A/B.
The probability to identify the tension is calculated by summing over all the values of P (∆θ) over the iso-contour
corresponding to no difference ∆θ = 0
ˆ
∆= P (∆θ) d∆θ . (3.15)
P (∆θ)>P (0)

†† Reprinted from Physics Letters B, Volume 855, id.138844 Leizerovich, M., Landau, S.J., Scóccola, C.G. Tensions in cosmology: A
discussion of statistical tools to determine inconsistencies Copyright (2024), with permission from Elsevier.
122

Usually, this integral is computed numerically from the posterior samples. In cases of strong disagreement, this
estimator may be difficult to compute.42
Many authors use the tools presented here to discuss tensions in the inferred value of cosmological parameters with
recent data sets [698, 1977, 1978]. It has been discussed [1977] that the different metrics answer different questions.
For example, QUDM quantifies how a given data set updates the values of the inferred parameters of another data
set, while QDMAP quantifies the difference of the goodness of fit between the individual and joint data sets to the
underlying theoretical model. On the other hand, to quantify the tension between the value of inferred parameters,
the appropriate metrics are QDM and EPS, emphasizing that the former requires posteriors to be sufficiently Gaussian,
while the latter does not. Fig. 52 shows two examples of 2-d contours and posteriors that correspond to different
data sets and the corresponding Nσ obtained with each metric, together with the ones computed with the 1-d rule
of thumb for each parameter. If EPS can be computed, usually the equivalent Nσ that corresponds to DM and EPS
indicate similar values, while there is a difference with the ones obtained using the rule of thumb [1977]. Note that
the rule of thumb only quantifies the tension in a given parameter, while the metrics presented above quantify the
disagreement in the whole parameter space.

3.2. Machine learning based inference techniques


Coordinator: Jurgen Mifsud
Contributors: Alba Domi, Anto Idicherian Lonappan, Benjamin L’Huillier, Celia Escamilla-Rivera, Clecio Roque De
bom, Daniela Grandón, Daria Dobrycheva, David Valls-Gabaud, Denitsa Staicova, Elena Giusarma, Filippo Bouchè,
Germano Nardini, Ippocratis Saltas, Iryna Vavilova, J Alberto Vazquez, Jacobo Asorey, Jenny G. Sorce, Judit Prat,
Konstantinos Dialektopoulos, Leandros Perivolaropoulos, Luca Izzo, Luis A. Escamilla, Matteo Martinelli, Purba
Mukherjee, Rahul Shah, and Ruth Lazkoz

In traditional approaches, analytical likelihood functions are employed to characterize the PDFs of observed data,
enabling parameter estimation through posterior inference. However, certain datasets exhibit errors that cannot be
adequately described by simple analytical distributions, such as multivariate Gaussian distributions. Consequently,
the true likelihood in such cases is often complex and lacks an explicit analytical form, making traditional parameter
inference methods challenging to implement. To address this issue, various likelihood-free techniques have been
developed to circumvent the direct computation of likelihood functions.
In standard Bayesian inference, posterior distributions are typically explored using methods such as MCMC sam-
pling, variational inference, or other Bayesian computational techniques. These conventional approaches generally
rely on evaluating the likelihood function for the models and parameters under consideration. However, for highly
complex and computationally intensive models, the likelihood function may become intractable, and simulations can
demand significant time and resources, rendering parameter inference impractical. Therefore, parameter estima-
tion techniques that mitigate or overcome these challenges are highly beneficial, particularly in cosmological studies.
Likelihood-free inference has emerged as a promising paradigm for Bayesian inference in scenarios involving complex
generative models, leveraging only forward simulations to perform analysis.
With the current generation of large-scale structure observational programs, such as the European Space Agency’s
Euclid mission [437] and LSST [893], sub-percent-level precision measurements are anticipated. The upcoming decade
is thus expected to witness an unprecedented increase in the quantity, diversity, and quality of multi-wavelength
astronomical observations of the large-scale structure. This surge in data will necessitate the development of highly
sophisticated computational tools. At this stage, the limitations may shift from data quality or availability to the
capabilities of statistical and data-driven methodologies. ML techniques have demonstrated substantial potential in
addressing the computational challenges associated with traditional statistical methods, positioning them as valuable
tools for advancing cosmological analyses.
For instance, current ML techniques were adopted to the scatter in cluster mass estimates [1979] and to distinguish
between standard and MG theories from statistically similar WL maps [1980]. Such techniques have also been found
to be useful for the next generation CMB experiments [1981], N-body simulations [1982], cosmological parameters
inference [1983], DE model comparison [1984], supernova classification [1985] and strong lensing probes [1986].
Cosmology is experiencing an unprecedented growth in the volume and complexity of astronomical datasets, driven
by large-scale surveys like Euclid, LSST, and DESI. These datasets are used to construct data vectors for cosmological
parameter inference, with which we can shed light on current parameter tensions within ΛCDM. To fully leverage the
wealth of cosmological information encoded in these datasets, ML methods have emerged as a powerful tool.
In this section, we will discuss a number of ML based inference techniques, with an emphasis on those techniques
which were adapted to address cosmological tensions. Neural networks are becoming ever more widely employed in

42 We point out that in the case of strong disagreement, care must be taken with the sampling of the tails of the distributions.
123

FIG. 53: The network structure used for parameter estimation in Ref. [1992]. The structure of the left panel is for
one dataset, while the multi-branch network of the right panel is for multiple datasets {D1 , D2 , D3 , ..., Dn } that are
from different experiments.

physics, including in the field of cosmology. We refer the interested reader to a review of the core concepts surrounding
the use of neural networks in Refs. [1987, 1988]. In this section we discuss the cosmological applications of artificial
neural networks (Sec. 3.2.1), convolutional neural networks (Sec. 3.2.2), Bayesian neural networks (Sec. 3.2.3), and
deep learning (Sec. 3.2.4).
3.2.1. Artificial neural networks
The method by which AGB architectures [1654] are used to learn to mimic the iterative process by which an
MCMC analysis takes place is described in this section together with some details of the adopted methodology to
optimize the AGB structure for comparative analyses. AGBs are non-parametric techniques in the sense that they do
not contain the cosmological parameters themselves, unlike the likelihood functions. This may extend cosmological
analyses to otherwise overly complex systems, or reduce the computational requirements of regular cosmological
model analyses making the approach more accessible. Inspired by biological neural networks, an AGB is a setup
composed of a collection of neurons that are organized into layers [1989–1991]. This family of algorithms offers a
powerful base on which to apply complex numerical tasks. AGBs are optimized to work in computer systems with a
large number of threads which has become a feasible prospect in recent years with the development of very powerful
graphical processing units (GPUs), thus making the process of training AGBs competitive with traditional techniques.
In recent years, AGBs have become very useful in meeting the accuracy and computational efficiency required for
analyses in cosmology. AGBs are constructed with input and output layers to represent the MCMC parameter set
that is being sampled and output likelihood respectively. These layers are connected to a series of internal, or hidden,
layers that are structured in a way to optimize how closely the AGB can imitate the MCMC iterations. Each layer in
this network is composed of neurons, each of which is connected to the neurons of the layers preceding and after it,
as illustrated in the left panel of Fig. 53 for a single data set.
For the task of inferring cosmological parameter constraints, the observational data is fed to the input layer, then
the information of observational data passes through each hidden layer, and finally the cosmological parameters are
outputted from the output layer. Specifically, each layer accepts a vector, the elements of which are called neurons,
from the former layer as input, then applies a linear transformation and a nonlinear activation on the input, and
finally propagates the current result to the next layer.
AGBs have been widely implemented as emulators of cosmological observables, such as the lensing power spectrum
[1993, 1994], CMB source functions [1995, 1996], the CMB temperature anisotropies power spectrum and the matter
power spectrum [1994, 1997]. The main purpose of these applications is to accelerate the statistical analysis, when
expensive Boltzmann codes and big volumes of data prevent us from efficiently sampling the posterior distributions.
In particular, CosmicNet I aims to remove bottlenecks from Einstein-Boltzmann solvers (such as CLASS or CAMB)
by training neural network emulator that learn the mapping from four ΛCDM parameters to source functions of
CMB anisotropies. The resulting trained AGB model is then injected into the CLASS code for public use. This work
was later extended to include extensions of a flat ΛCDM, such as DDE, spatial curvature. Emulators are crucial
for summary statistics of cosmological probes that lack an analytical prescription that models their dependence on
cosmology and other parameters. This is the case of higher-order statistics of cosmic fields, such as the peak counts (a
124

FIG. 54: The left panel depicts the schematic diagram of the likelihood-free architecture adopted in Ref. [1992]. In
the right panel, a visualization of a mapping between the data space of measurements and the cosmological parame-
ter space is illustrated.

distribution of local maxima in a field) and the one-point PDF. As emulators are only approximations of the target
observables, recent works have explored how to propagate the uncertainty derived from such approximation, in order
to safeguard parameter inference against biases [1998].
In recent years, likelihood-free analyses relied on ensembles of neural networks as density estimators of the sampling
distribution of the data. In Refs. [1992, 1999], a cosmological likelihood-free inference technique has been developed.
The left panel of Fig. 54 depicts the schematic diagram of the likelihood-free architecture adopted in Ref. [1992] con-
taining the main processes of training and parameter estimation. First, a class object for the cosmological model that
contains the simulation method of the measurements is developed, where the simulation method here is used to gener-
ate the training set. Then, the initial parameters are set without any constraints in their selection. Subsequently, the
training set will be simulated automatically, where two sampling methods can be considered to generate cosmological
parameters: sampling uniformly in a hypercube or a hyper-ellipsoid. It should be noted that the posterior distribu-
tion is unknown before the first estimation using the AGB model. Thus, the cosmological parameters here cannot be
generated in a hyper-ellipsoid. Therefore, the cosmological parameters are uniformly generated in a hypercube using
the set of initial parameters. Consequently, an AGB model will be constructed automatically based on the respective
training set. At the same time, the training set is then preprocessed before training the network, while noise will be
automatically generated based on the observation errors and added to the training set. The training set will then be
normalized, and the training set will be fed to the AGB model, consisting of thousands of epochs.
Cosmological parameters are then estimated using the well-trained AGB model. In simple terms, the AGB model
actually learns a mapping between the data space of the measurements and the parameter space of cosmological
parameters, as depicted in the right panel of Fig. 54 via a mapping between the data space of measurements and the
cosmological parameter space. Therefore, in order to infer the respective posterior distribution, the corresponding
distribution of the measurements is fed to the AGB model. Indeed, a large number of data-like samples of the
measurements using the observational data are generated and fed to the AGB model to obtain the corresponding
AGB chain, similar to the concept of an MCMC chain.
The AGB method in Refs. [1992, 1999] was shown to be very successful in estimating cosmological parameters
with one data set, which was then extended to multiple data sets by utilizing a multi-branch network The latter
multi-branch network is illustrated in the right panel of Fig. 53, in which it was successfully tested with CMB, SNIa,
and BAO datasets in Refs. [1992, 1999], as illustrated in the derived results of Fig. 55 with the Planck 2015 data set.
Thus, the AGB method was shown to be capable of estimating parameters using data sets of multiple experiments
and is a direct competitor of the customary MCMC technique.
3.2.2. Convolutional neural networks
With the advent of precision cosmology, we are reaching an era in which it is extremely important to avoid any
potential biases in the cosmological analysis when extracting cosmological parameters. For example, most of the
time large-scale structure analyses are done using summary statistical observables like 2-point correlation functions,
after assuming a cosmology to go from angular coordinates to Cartesian ones. This type of analysis can bias the
cosmological parameters and a tomographic analysis can avoid this [2000].
One alternative to the use of summary statistics is the use of Convolutional Neural Network (CNN) architectures.
They can be used to make direct inferences on cosmological parameters from astrophysical images. For instance,
Ref. [2001] used CNNs to constrain axions as DM from 21cm line intensity mapping probes. In Fig. 56 we show a
schematic of the adopted CNN model architecture. It took as input a 64 × 64 × 128 lightcone data cube where each
125

FIG. 55: One-dimensional and two-dimensional marginalized distributions [1992] constrained from Planck 2015
power spectrum. The blue solid lines are the results of the AGB method, the red dashed lines represent those of
the MCMC method, while the gray circles are the fiducial values of the cosmological parameters.

FIG. 56: The architecture of the CNN adopted in Ref. [2001], with input data cube of 64 × 64 × 128 voxels where
each voxel has dimension 4.68cMpc × 4.68cMpc × 1.48Mhz (cMpc means co-moving Mpc).

voxel has dimension 4.68cMpc × 4.68cMpc × 1.48Mhz (cMpc means co-moving Mpc).
If DM were described by axion-like particles (ALPs), the small scale structure would change the value of Ωm,0
leading to a change in H0 and σ8 , potentially solving the potential tensions. To make a direct inference from CNNs,
we need a training set of simulations of the given astrophysical image for different sets of parameters (e.g., axion mass
or H0 ). The input layer is basically the simulation with the angular maps and the frequency maps while the final
output layer is the vector with the cosmological parameters inferred (in this case, the axion mass). The input layer
is filtered and pooled down before being inputted as a vector to the connected neural network layers. The data is
passed through the network a number of epochs updating the corresponding weights each time.
The inferred results, as depicted in Fig. 57, show that the CNN successfully recovered the true values of the axion
126

FIG. 57: The distribution of the CNN inferred predictions [2001] on 100 noiseless simulations fixed axion mass of
10−21 eV, along with the vertical dashed lines showing the 1-σ (68%) credible interval. We should remark that the
CNN could predict the axion mass because the CNN is trained only with the data including axions.

mass in the testing data with a precision of ∼ 20% across a broad range of masses. To evaluate the uncertainty in
the recovered values, 100 additional independent simulations were generated for both an axion model and a standard
CDM model. For an axion mass of MX = 10−21 eV, a mass uncertainty of +5.94 −4.80 × 10
−21
eV was inferred. It was
further concluded that this method could potentially enable the detection of axion DM using SKAO1-Low with 68%
confidence if the axion mass is MX < 1.86 × 10−20 eV. However, these findings depend on the Planck 2015 cosmological
parameters and the specific design parameters of future SKAO1-Low configurations.
A CNN was also adopted to perform the typically computationally expensive task of estimating the parameters
of GW events. The considered CNN in Ref. [2002] is able to produce posterior distributions that in all cases are
compatible with the already published results. The schematic of the respective CNN is depicted in Fig. 58, whereas
the inferred results for the event GW200224_222234 are illustrated in Fig. 59. In the case of event GW200224_222234,
all the estimated parameters were found to be in accordance with those published in the GWTC-3 catalog. Having
said that, the effective spin was the worst performing parameter. Furthermore, the uncertainty yielded by the CNN on
the sky position was found to be too large to accurately pinpoint the event, but it still is unbiased and is expected to
produce a first alert for the instruments that look for electromagnetic counterparts to start pointing their instruments
to a given patch in the sky. The latter would be more feasible with the upcoming more accurate pipelines which are
expected to improve the precision of the sky localization. Overall, the GW event results indicate a good response
from CNNs. One of the most important advantages is its computational efficiency, in which the CNN model could
provide several posterior samples in a fraction of the time required by Bayesian inference methods, thus facilitating
multi-messenger astronomy.
3.2.3. Bayesian neural networks
A Bayesian Neural Network (BNN) that outputs Gaussian PDFs was adopted in Ref. [2003]. The BNN has five
input nodes for the five-band grizy photometry. A parameter grid search was performed in order to optimize for free
parameters, including the number of epochs, number of layers, number of nodes per layer, learning rate, loss function,
activation function, and optimizer. The BNN has a final output node that produces a mean and standard deviation
127

FIG. 58: Architecture of the CNN used in Ref. [2002]. The first set of layers are 1-dimensional convolutions and 1-
dimensional maxpooling ones, followed by a set of fully connected dense layers. Between each of the dense layers, a
dropout of 20% is applied to prevent overfitting during the training stage.

FIG. 59: Full posterior distribution inferred [2002] from the CNN and MCMC approaches for the
GW200224_222234 event.

assuming a Gaussian distribution for each photo-z prediction.


A key attribute of the BNN model is the production of photo-z uncertainties, which are needed for using photo-z
results in cosmological analyses. It was also found that the BNN produces accurate uncertainties. In Ref. [2003], the
HSC Public Data Release 2 [2004], which is designed to reach similar depths as LSST but over a smaller portion of
the sky was adopted for the BNN. In total, the data consists of 286,401 galaxies with broadband grizy photometry
and known spectroscopic redshifts. The considered galaxy sample covers a redshift of 0.01 < z < 4, however the
128

FIG. 60: Visualization of the neural network (top left) and BNN (top right) performance compared to the BNNN
with outlier removal criteria σz < 0.5 (bottom left) and σz < 0.3 (bottom right) [2003].

majority of the sample lies between 0.01 < z < 2.5. In the BNN analysis of Ref. [2003], 80% of the galaxies were used
for training, 10% for validation, and 10% for testing. The performance of the BNN technique is clearly illustrated in
Fig. 60, in which the BNN was reported to have superior photo-z estimations with respect to other competing models.
The capability of BNNs was confronted in Refs. [2005, 2006] on different avenues. The first is how effective BNNs can
be in recognizing the distinct features in the power spectrum for a particular modification to the concordance model
of cosmology, such as f (R) or Dvali–Gabadadze–Porrati (DGP) [2007]. The second avenue which was exploited in
Refs. [2005, 2006] is their ability to detect a deviation from ΛCDM in the power spectrum irrespective of the particular
modification. Two BNNs with the same architecture were trained in Refs. [2005, 2006], the first for five labels divided
between ΛCDM and the considered four extensions, whilst the second trained to distinguish between the two labels of
ΛCDM and non-ΛCDM. Due to the fact that only four redshift bins were considered in these works, it was beneficial
to treat the data as four separate time series and use one-dimensional convolutional layers.
The architecture of the network used to train both the five-label and two-label networks is depicted in the left
panel of Fig. 61. Initially, the adopted structure consisted of three 1D convolutional flip out layers with 8, 16 and 32
filters, kernel sizes of 10, 5 and 2 with strides of 2, 2 and 1, respectively. Each of the first two 1D convolutional layers
were followed by a max pooling layer with a pool size of 2 and a pooling stride of 2 for the first max pooling layer
and a pooling stride of 1 for the second max pooling layer. After both of these max pooling layers there is a batch
normalization layer. Following the final convolutional layer there is a global average pooling layer to reduce the filter
size to one in order to pass it to a dense layer with 32 nodes. Finally, after a further batch normalization there is a
softmax layer consisting of five or two neurons for either the five- or two-label networks, respectively.
It was found that a five-label network trained to classify between ΛCDM, f (R) gravity, DGP gravity, wCDM and
a “random” class provided more reliable predictions than a two-label network trained to distinguish simply between
ΛCDM and non-ΛCDM. While generally being less sensitive to variations in the noise distribution, it can also determine
whether a power spectrum does not belong to any class included in the training set. Since the selection of the correct
model is crucial when performing conventional statistical analyses such as with MCMCs, this ability could prove
beneficial in indicating prospective models to consider. However, the network used in this work is currently limited
to classification tasks while the notion of model selection on firm statistical grounds in the context of BNNs remains
129

FIG. 61: The right panel illustrates the BNN workflow as adopted in Refs. [2005, 2006], whilst the left panel depicts
the corresponding BNN architecture employed for both the five-label and two-label classification tasks. The height
of each block illustrates the dimension size for each layer, while the number of blocks per layer corresponds to the
number of filters. The dense blocks embedded in the first three transparent layers indicate the kernels for the first
three one-dimensional convolutional layers scaled by their respective size.

an open problem.
It could be concluded that BNNs may provide a powerful new means to search for hints of new physics in cosmological
datasets. In particular, it is anticipated that they will serve as a powerful “filter”, allowing us to narrow down the
theory space before moving on to constrain model parameters with MCMCs while perhaps even signaling the presence
of new physics that does not belong to any known model.
3.2.4. Deep learning
Neural networks have lately become ubiquitous in cosmology [1991, 1999, 2008–2021]. In this regard, the LADDER
- Learning Algorithm for Deep Distance Estimation and Reconstruction - suite is a novel deep learning algorithm
designed to learn the cosmic distance ladder in a model-independent, non-parametric manner [2022]. The schematic
overview of the training algorithm is depicted in Fig. 62, whilst the corresponding algorithm is illustrated in Fig. 63.
Trained on late-time datasets like the Pantheon SNIa dataset [2023], LADDER incorporates associated errors and
complete covariance information. It interpolates from the joint distribution of a randomly chosen subset of the
dataset to estimate target variables and errors simultaneously, effectively handling correlations and the sequential
nature of the data. This leads to robust predictions resilient to input noise and outliers, even in data-sparse regions.
Optimized by physically motivated metrics such as monotonicity and smoothness, LADDER leverages the Gaussian
nature of the observations by employing the Kullback–Leibler (KL) divergence as the loss function during training.
The LSTM network, found to perform best with the Pantheon dataset, is included in the LADDER suite, available
publicly on GitHub (https://github.com/rahulshah1397/LADDER) under a MIT License and version 1.0 is archived
in Zenodo [2022].
Having learned the cosmic distance ladder, LADDER is a powerful tool for cosmological applications. It can inde-
pendently verify the consistency of SNIa datasets like Pantheon+ [32] and DES [1253], and serve as a pathology
test for qualitatively different datasets, such as 2D vs. 3D BAO. Additionally, it acts as a model-independent
calibrator for high-redshift datasets, including GRBs and RGBs. Moreover, previous ML approaches in cosmology
struggled with stable error prediction at higher redshifts, rendering them unsuitable for precision cosmological tests.
LADDER overcomes these challenges and shows promise in extrapolating beyond available data, useful for simulating
intermediate-redshift data or augmenting current data to higher redshifts.
In Ref. [2024] LADDER is employed to recalibrate SDSS BAO and DESI BAO in a model-independent manner, which
helps address the longstanding H0 and S8 tensions simultaneously. Traditionally, BAO distances are inferred using
the sound horizon at the drag epoch (rd ), which is calibrated based on CMB data under the assumption of the ΛCDM
model. In contrast, LADDER, trained on the Pantheon dataset, derives rd purely from observational data without relying
on cosmological model assumptions, instead incorporating merely an astrophysical prior on the absolute magnitude of
SNIa (MB ). When combined with Planck 2018 data, this recalibrated BAO framework, along with Pantheon, yields
joint constraints on ΛCDM that significantly alleviate the H0 and S8 tensions, due to an “in-plane shift", despite
130

FIG. 62: Schematic overview of the deep learning algorithm adopted in Ref. [2022].

FIG. 63: The deep learning algorithm adopted in Ref. [2022].

their negative correlation. These results highlight the potential biases introduced by CMB-based BAO calibration
and demonstrate the robustness of data-driven approaches, offering a promising new avenue for precision cosmology.
Beyond conventional deep learning architectures, more advanced generative models have been developed to tackle
complex challenges in cosmological inference. These methods, such as normalizing flows [2025] and diffusion mod-
els [2026, 2027], extend deep learning capabilities by efficiently learning high-dimensional PDFs and accelerating
likelihood-free inference. Unlike traditional MCMC approaches, which often struggle with computational inefficiency,
generative models can rapidly approximate complex posteriors while maintaining high accuracy.
Normalizing flows have shown promise in addressing cosmological tensions, such as discrepancies in the Hubble
constant and matter fluctuation amplitude (S8 ). By modeling joint PDFs across datasets, they enhance uncertainty
quantification and provide a robust statistical framework for analyzing these tensions [2028, 2029]. One example is
the EMUFLOW framework [2030], which utilizes normalizing flows to model joint posterior distributions from multiple
datasets, enabling efficient constraint combination without increasing parameter space dimensionality. Additionally,
methods employing normalizing flows have been developed for non-Gaussian tension estimation, improving the quan-
tification of agreement levels between different cosmological experiments and offering insights into discrepancies such
as the H0 tension [2031]. Further developments in normalizing flow architectures have introduced approaches that
incorporate symmetries of the universe, such as translation and rotation equivariance, optimizing parameter inference
131

for cosmological fields [2032]. Other extensions [2033] leverage hierarchical wavelet-based decompositions to improve
robustness across spatial scales, while iterative flow-based techniques facilitate the transformation of complex PDFs
into tractable forms, improving efficiency in cosmological data analysis.
Diffusion models, leveraging stochastic iterative denoising processes, have also been applied successfully to cosmo-
logical parameter inference. These models excel in uncertainty quantification and high-fidelity generative sampling
of cosmological fields, particularly in large-scale structure and WL studies. A recent application integrates diffusion
models with Hamiltonian Monte Carlo methods [2034, 2035], enabling accurate posterior sampling and accelerated
convergence in extracting key cosmological parameters such as Ωm,0 and S8 . The ability of diffusion models to re-
construct high-fidelity statistical distributions while remaining robust to observational noise makes them valuable
tools for precision cosmology. As ML techniques continue to evolve, integrating generative models into cosmological
inference is expected to enhance the precision and reliability of parameter estimation, especially for upcoming surveys
like LSST and Euclid. These advancements underscore the growing role of deep learning-driven methodologies in
shaping the future of cosmological analysis.
3.2.5. Conclusion
We reviewed the application of ML methods for the inference of cosmological parameters, emphasizing their ver-
satility and capacity to address the inherent complexity of astronomical datasets. By integrating ML techniques
into cosmological inference, we enable efficient and accurate predictions, facilitating faster convergence of traditional
algorithms such as MCMC and advancing implicit likelihood or likelihood-free inference using deep learning. These
innovations are particularly valuable in handling the high-dimensional, noisy, and often incomplete nature of cosmo-
logical data.
Across the subsections of this review, it is evident that the ML techniques considered provide powerful new tools
to uncover potential signals of new physics within cosmological datasets. These methods act as a "filter," allowing
researchers to effectively narrow down the theory space and refine subsequent analyses with standard techniques like
MCMC. Furthermore, ML based approaches may even reveal evidence of new physics that does not fit within the
framework of existing theoretical models, opening new avenues for exploration.
While alternative methodologies exist, we emphasize the transformative potential of advanced learning algorithms
in optimizing the extraction of information from cosmological data. By demonstrating their utility, we aim to showcase
how these novel approaches can complement or enhance traditional inference techniques. The innovative integration of
ML into the cosmological framework promises to address current challenges, such as improving parameter constraints
and identifying hidden structures or patterns in data, thereby paving the way for more accurate reconstructions of
cosmological phenomena.
We also highlight the possibility of yet unexplored ML algorithms that may surpass the performance of those
outlined in this review. This underscores the importance of continued exploration and innovation within the ML and
cosmology communities. By encouraging the adoption and further development of ML based inference techniques, we
aim to inspire new efforts to tackle unresolved questions in cosmology from fresh perspectives. Such endeavors would
significantly improve the precision and reliability of cosmological analyses, contributing to a deeper understanding of
the Universe’s fundamental nature and addressing ongoing tensions in cosmological observations.
In conclusion, the integration of ML methods into cosmological research represents a paradigm shift, offering not only
computational efficiency but also novel insights into the underlying physics of the Universe. We urge the community
to actively engage in the exploration of these advanced tools, as they hold the potential to revolutionize the study of
cosmology in the years to come.

3.3. Reconstruction techniques


Coordinator: Luis Escamilla, Daniela Grandón
Contributors: Adrià Gómez-Valent, Anil Kumar Yadav, Anjan Ananda Sen, Anto Idicherian Lonappan, Ariadna
Montiel, Arianna Favale, Benjamin L’Huillier, Biesiada Marek, Celia Escamilla-Rivera, David Benisty, David Valls-
Gabaud, Eoin Ó Colgáin, Filippo Bouchè, Iryna Vavilova, Isidro Gómez-Vargas, J. Alberto Vázquez, Jenny G. Sorce,
Jenny Wagner, Jurgen Mifsud, Konstantinos Dialektopoulos, Luis E. Padilla, Matteo Martinelli, Miguel A. Sabogal,
Purba Mukherjee, Rafael C. Nunes, Rahul Shah, Rocco D’Agostino, Ruth Lazkoz, Víctor H. Cárdenas, and Wojciech
Hellwing

Reconstruction techniques have become crucial for understanding the phenomenology of DE, the expansion history,
and the growth of large-scale structures in our Universe. They aim to capture specific features or general trends in
the data, test a predefined physical model against observations, or explore a flexible, theory-agnostic framework to
elucidate the phenomenology of cosmological functions. In general, reconstructions can be categorized into parametric
and non-parametric approaches. They mainly differ in their underlying assumptions about a cosmological model
and the functional form of the reconstructed quantity. The former imposes an explicit form on the function being
132

reconstructed, thereby reducing uncertainties in the result. However, the a priori assumption of a specific model may
limit the flexibility of the reconstruction and introduce biases regarding the properties of the cosmological function of
interest. To mitigate this, some studies instead perform a non-parametric (“free-form”) reconstruction within a theory-
agnostic framework. Non-parametric reconstructions are particularly useful for gaining insights into the dynamics of
cosmological functions when the true underlying function is not directly observable but can be inferred from observed
data.
Many works on the parametric and non-parametric approaches investigate the impact of H0 priors, to see whether
the H0 tension is also imprinted in the derived cosmological reconstructions. Hence, these techniques can shed
light on the cosmological tensions and advance our understanding of their effects on a broader context. Moreover, a
joint analysis of parametric and non-parametric reconstructions using cosmological observations is valuable to robustly
identify departures from the ΛCDM model that emerge despite the differences of the methods used (e.g., see Ref. [2036–
2038] for examples). In summary, these methods serve as data-driven frameworks to decode the phenomenology of
cosmological functions from complex datasets, and offer an alternative approach to distinguish between competing
cosmological models.
In what follows we present some examples of reconstructions and how they can be classified according to their
methodology.
3.3.1. Parametric reconstructions
In a parametric reconstruction, the desired physical quantity is modeled using an analytical function with a set of
free parameters that describe its behavior across cosmic time or spatial scales. The free parameters of the postulated
function, and the cosmological parameters, are hence jointly inferred in light of the observational data. Hence,
parameterizations are a crucial tool for modeling unknown physical phenomena in a mathematically tractable way.
As an example, in the absence of a fundamental and well-defined theory of DE, several functions, i.e., w(z), ρ(z), have
been parameterized in a broad number of different ways. These parametrizations track the dynamics of a DE back in
time simply extending the cosmological model with extra parameters. While parametric methods are computationally
efficient and easy to interpret, they are inherently limited by the assumptions built into the chosen parameterization.
A poorly chosen model may lead to biased results that fail to capture the true complexity of the underlying physics. As
such, model selection criteria for various parametric candidates are critical in assessing the robustness of the results.
3.3.2. Phenomenological parameterizations
Phenomenological parameterizations focus on capturing specific behaviors, often without direct derivation from
fundamental physics. Their primary goal is to test whether certain trends (such as evolving equations of state,
deviations from General Relativity, or modified growth of structure) are compatible with observational data. These
parameterizations provide a practical way to explore a broad range of possibilities without requiring a detailed
theoretical foundation. Some well-known examples include:
• CPL parameterization [297, 298] for the DE equation of state
w(z) = w0 + wa z/(1 + z) , (3.16)
which is the general go-to parameterization used to test if any dynamic behaviour can be present in DE (this
parameterization is further explained in Sec. 2.3.4). The DESI collaboration recently employed this parame-
terization in their analysis of year-one BAO data [698] and found evidence suggesting a preference for a DDE
model with a significance exceeding 2σ.
• The Jassal-Bagla-Padmanabhan (JBP) parameterization [2039], also for the DE equation of state
w(z) = w0 + wa z/(1 + z)2 , (3.17)
which serves as a modification of the CPL designed to exhibit different behaviours at large redshifts.
• Interaction term between DE and DM [2040–2043]
Q = ξHρDM , (3.18)
with the parameter ξ governing the strength of the interaction. It is proposed as a way to represent the interaction
as a function of time (H) and energy density (ρDM ). Given its phenomenological nature, this approach can be
extended further, as demonstrated in Ref. [2044], where the interaction term was modified to depend not only on
DM but also on DE. In this case, the interaction takes the form Q = ξH(ρDM − αρDE ) with the new parameter
α regulating the strength of the dependence on DE. Moreover, the authors in Ref. [2042] discuss the interaction
between DM and DE in the context of the second law of thermodynamics and its consequences on DE evolution.
For a more complete discussion on DE-DM interactions please refer to Sec. 4.4.3.
133

• A parametric form for the DE density X(z) was developed by Refs. [2045–2047] and further tested in Refs. [2036,
2048]. The quadratic parametrization corresponds to
   2
ρDE (z) z z
X(z) ≡ = 1 + (4x1 − x2 − 3) − 2 (2x1 − x2 − 1) , (3.19)
ρDE (0) zm zm
where zm is the maximum redshift in the data set, x1 and x2 are the free parameters of the model to be
constrained in light of the data. When x1 = x2 = 1, the parametrization reduces to ΛCDM where the DE
density is constant X(z) = 1. This approach aims to study evidence for DE evolution by allowing its energy
density to change with redshift. Reconstructing X(z), instead of w(z), is advantageous since this quantity is
more directly traced by the luminosity distance. A cubic parametrization was also explored in Ref. [2048] to
allow for an extra degree of freedom and to investigate whether further transitions respect to X(z) = 1 are
found in the reconstruction.
• The growth rate parameterization [2049]

f (z) = Ωm,0 (z)γ , (3.20)

where γ is named the growth index, which for ΛCDM has a value of ∼ 0.55. A recent application of this
parameterization can be found in Ref. [2050], where an analysis using f σ8 and Planck CMB revealed a tension
of 4.2σ with the concordance model’s expected value of γ = 0.55, suggesting a potential need for modifications
to the way the standard model predicts structure growth.
• The gravitational slip [2051]
Φ
η= , (3.21)
Ψ
where Φ and Ψ are the metric potentials in the perturbed Einstein equations. If η deviates from 1 there may be
an indication for MG. In Ref. [2052] four different models for η were analyzed using forecasts from WL, galaxy
clustering, and supernova data to evaluate how effectively future surveys could constrain the slip parameter.
Furthermore, in Ref. [2053] a deviation from GR was found using GW data along with the slip parameter.

As we can see, phenomenological parameterizations allow for a broad exploration of deviations from the standard
model without committing to a specific underlying theory.
3.3.3. Physically motivated parameterizations
Physically motivated parameterizations, in contrast, are derived from fundamental theories and ensure consistency
with known physics. These models are often constructed from Lagrangian formulations, field equations, or extensions
of Einstein’s General Relativity, ensuring that theoretical principles like energy conditions and stability constraints
are satisfied. Some of these parameterizations which can be found in the literature are:

• Scalar fields for DE. A well-known one being the Quintessence [295], where a scalar field ϕ with a potential
V (ϕ) leads to an equation of state

ϕ̇/2 − V
w(z) = , (3.22)
ϕ̇/2 + V

having the peculiarity of behaving as, in which the kinetic term has the wrong sign, ϕ̇2 /2 → −ϕ̇2 /2, leading to
w(z) > −1. Complementary to it we have the Phantom scalar field [2054] where w(z) < −1. The combination
of them, resulting in the interacting-two-scalar-field model named Quintom [2055–2057] has also been studied
due to its peculiarity of being able to cross the Phantom-divide line (w = −1) [2058, 2059]. K-essence scalar
field [2060], which can be seen as an extension of Quintessence where non-canonical kinetic terms are included
in the Lagrangian.
• Horndeski theories of MG [664, 2061–2063], which provide the most general scalar-tensor theories leading to
second-order field equations.
• f (R) theories [2064] which modify the Ricci scalar in the Einstein-Hilbert action
ˆ
4 √
 
1
S = d x −g f (R) + Lm , (3.23)
16πG
134

where the f (R) can take many forms. In Ref. [2065] two different parameterizations for f (R) were studied, them
being f (R) = R − 2Λ(1 − e−R/Λb ) and f (R) = R − 2Λ 1 − 1+(R/Λb)1
n , finding concordance with the standard


model when using BAO, SN and CC data. Similarly, in Ref. [2066] a preference for the standard model is
recovered when the two models f (R) = R − β/Rn and f (R) = R − α ln R − β are used in tandem with simulated
GW data.
• For DM models, we encounter WDM [2067–2071], where a nonzero velocity dispersion modifies the matter
power spectrum. Another example is Self-Interacting DM [2072–2075], in which a cross-section σ/m accounts
for self-interactions, affecting structure formation at small scales while preserving the CDM behavior on larger
scales.
Both phenomenological and physically motivated parameterizations aim to describe the evolution of cosmic com-
ponents, test deviations from ΛCDM, and assess the viability of new models against observations. Phenomenological
parameterizations provide a flexible way to analyze trends, while the physically motivated ones ensure consistency
with fundamental physics. The choice between the two depends on whether the focus is on empirical data fitting or
theoretical consistency.
3.3.4. Model-independent parameterizations
While both phenomenological and physically motivated parameterizations help constrain the behavior of cosmolog-
ical functions, they inherently require assumptions about the functional form. However, these assumptions can limit
the flexibility of the analysis, as the chosen parameterization may not fully capture the true underlying features of
the function being reconstructed. To overcome this limitation, one can employ more sophisticated parameterization
methods that minimize the number of underlying assumptions. Some of these approaches, referred to as model-
independent methods, remain parametric but offer significantly greater flexibility, allowing them to capture a wider
range of possible features in the data.
3.3.4.a. Basis representation Basis-representation parameterizations provide a flexible and systematic way to
reconstruct cosmological functions by expanding them in terms of a chosen set of basis functions. Common examples
found in the literature include:
• Given its simplicity and ease of use, the Taylor expansion has been applied to many different cosmological
functions. As an example, in Ref. [2076] the DE equation of state, through its pressure, was expressed as a
truncated Taylor series
N −1
X 1 dn p
p = p0 + (ρ − ρ0 ) + O[(ρ − ρ0 )N ] . (3.24)
n=1
n! dρn 0

This idea can be extended even further. By performing a Taylor expansion of the Hubble parameter around
z = 0, one arrives at the cosmographic approach [2077–2080], which will be explored in more detail in a later
subsection.
• In a manner similar to the Taylor series expansion, one can instead adopt a different basis, such as a Fourier
series. In Ref. [2081], this approach was applied to the DE equation of state, revealing a preference for an
oscillatory behavior at late-times. The equation of state was found to cross the phantom divide multiple times,
suggesting that single scalar field models are disfavored by the analyzed datasets, as they fail to reproduce this
behavior.
• The Padé approximation could be seen as an extension to the Taylor series, being itself a rational of two polyno-
mial series. It has the advantage of a faster convergence rate, but potentially using more new terms/parameters.
Its uses are well varied in the literature, going from the luminosity distance [434, 2082–2085], the Hubble expan-
sion rate [2085], the DE EoS [2084] and even f (R) [2086]. In Ref. [2082] the luminosity distance was expressed
as dL = 1+b1az+b
1z
2z
2 , which is usually called the (1, 2) Padé approximant for dL , finding some agreement with

Planck’s CMB data at small redshifts but allowing DE to be dynamical.


• Wavelet expansion. Wavelets work as a basis and have the peculiarity of being localized in frequency and
configuration space, which is useful for capturing both global trends and local fluctuations. In Ref. [2087], a
wavelet-basis was used for the DE EoS in the next manner
X
w(zj ) + 1 = Pi ψi (zj ) , (3.25)

with zj being the redshift points at which w is calculated, Pi the coefficients of the expansion and ψi the wavelet
function. In this particular work the wavelets used were the Haar ones (where ψi (x) = 1 for 0 ≤ x ≤ 1/2,
135

ψi (x) = −1 for 1/2 < x ≤ 1 and ψi (x) = 0 otherwise) and, using SNIa, WMAP and BAO data, a hint for
dynamical DE was found.
• Reconstructions using an orthonormal basis set of functions (ONB) has been used to reconstruct the normalized
cosmic expansion function, H(z)/H0 , from the Pantheon dataset [2088], without assuming a specific cosmology.
In this approach, the luminosity distance function DL (a, c) was expanded into orthonormal basis functions ϕα (a)
as
B −1
NX
DL (a, c) = cα ϕα (a) = c ◦ Φ , (3.26)
α=0

where a is the scale factor. The cα elements of c ∈ RNB denote the weights of the NB basis functions. These
terms are summed up in the short-hand notation of the right hand side. The choice of the ONB can be adapted
to the problem at hand, as has been done in Ref. [2088], but other general choices like Chebycheff polynomials
can be used. Then the values of cα are obtained by minimizing the χ2 function [2089]. Using a special ONB
adapted to approximate a ΛCDM universe with its first few basis functions, [2088] showed that the reconstructed
cosmic expansion function from the Pantheon set of SNIa still yields very broad confidence bounds, such that
ΛCDM and other competitive models of DE are still statistically consistent with the dataset.

3.3.4.b. Interpolation methods Another model-independent approach to parameterization, similar to the basis
representation method, involves interpolation techniques. In this approach, a set of “positions” is defined and then
connected in a specific manner, with the chosen connection method determining the type of interpolation used.
These positions act as free parameters, whose optimal values are determined through Bayesian parameter inference
algorithms, such as MCMC or Nested Sampling.
An example of such a class of methods consists in binning the function one wants to reconstruct and treat each
one of its binned values as a free parameter of the analysis, e.g., one could aim at constraining the value of w(z) at a
given number of redshift zi . This completely removes assumptions done on the trend of the function, but significantly
increases the number of free parameters to take into account. When using this approach the bins are usually connected
with hyperbolic tangents to preserve continuity. As a function it can be represented as
N −1
wi+1 − wi  z − z 

i
X
w(z) = w1 + 1 + tanh , (3.27)
i=1
2 ξ

where N is the number of bins, wi the amplitude of the bin, zi the position where the bin begins in the z axis
and ξ is a hyperparameter which governs how “smooth” is the transition from one bin to the other (a higher value
represents a higher smoothness). Being one of the most utilized model-independent interpolation methods it has a
wide range of applications such as in reconstructing: the neutrino mass [2090]; the primordial power spectrum and
inflaton potential [2091]; the DE energy density [2092, 2093]; the DE EoS [2092, 2094, 2095]; the interaction kernel in
an IDE model [2010, 2096]; a DE perfect-fluid model [2097]. Another example of a binned analysis (or interpolation)
is the reconstruction of the functions µ(z) and Σ(z), which describe deviations from General Relativity (e.g., see
Refs. [2098–2100]). In Refs. [2101, 2102], joint constraints on these functions and on the DE density parameter
ΩDE (z) were obtained combining CMB, BAO, and SN data. The results of this analysis show hints for deviations
from General Relativity, mainly in the Σ(z) function which encodes deviations from the gravitational lensing effect
expected in the standard model, while also reducing the both H0 tension to ∼ 2σ (when comparing the inferred value
of H0 = 69.44 ± 1.3 km s−1 Mpc−1 with the SH0ES one) and the S8 tension (by inferring a value of S8 = 0.780 ± 0.033
which is in line with DES Y3’s measured value of S8 = 0.769 ± 0.016).
Another choice of interpolation is commonly referred to as “nodal reconstruction” [2103]. Here the interpolations
are done using linear, cubic or higher order splines, to fill in the gaps between a certain number of nodes. The simplest
example is the linear interpolation, that is, given two coordinates (zi , fi ) and (zi+1 , fi+1 ), the function behaves as
follows
fi+1 − fi
f (z) = fi + (z − zi ), z ∈ [zi , zi+1 ] . (3.28)
zi+1 − zi
This method has been used to reconstruct several important cosmological quantities, i.e., the primordial power spec-
trum [2104–2107], the DE energy density and the DE EoS [2092, 2108], the expansion rate of the Universe [2109] and
the neutrino mass [2090]; just to mention a few.
An unusual type of interpolation makes use of GP. The mathematical formalism behind it was already presented
in Section 3.1, and further details of its applications in reconstruction will be discussed in the following subsection.
136

This interpolation works in a similar manner to the nodal reconstruction, but with the obvious distinction that the
connection between the varying nodes is made through GP. This approach can mitigate some issues that the other
two approaches present, such as the choice of the bins’ width and positions [2110], while also presenting the advantage
of being infinitely differentiable. Given its relatively new implementation it has not been used widely. In Ref. [2110]
was first implemented with the equation of state parameter, and in Ref. [1984] was used to reconstruct the interaction
kernel of an IDE model.
3.3.5. Cosmography
The degeneracy among various theories proposed in recent years to explain DE has driven the investigation of
methods that enable the study of cosmic expansion without relying on predefined cosmological models. This is the
case of the well-known cosmographic approach [2077–2080], which depends only on the cosmological principle and
uses series expansions of the luminosity distance around the current time. The power of the cosmographic method
arises from the fact that it involves observables that can be directly compared with data and guarantees independence
from any assumed DE equation of state. Thus, the cosmographic method has been extensively utilized to distinguish
between various theoretical models that appear similar when compared to observations [2111–2114].
However, the cosmographic method faces two main challenges that may limit its effectiveness as an accurate tool
for describing cosmic expansion. The first challenge is the necessity for a substantial amount of data to distinguish
between the cosmological constant and a DDE. The second issue concerns the use of high-redshift data to explore
potential deviations from the standard cosmological model. The latter aspect conflicts with the core principle of
the standard cosmographic technique, which relies on a Taylor expansion series centered around z = 0. As a result,
large error propagations due to convergence issues often compromise the effectiveness of the method. Over time,
various alternatives to the standard cosmographic method have been explored to address these limitations. One
approach involves using auxiliary variables and developing expansion series for cosmological observables based on
re-parametrizations of the redshift variable converging as z ≫ 1 [434, 1389, 2115].
An alternative way to tackle the aforementioned challenges is to utilize rational polynomials that can heal conver-
gence problems typical of the standard Taylor-based cosmography. A prominent example of this technique is offered
by Padè polynomials, which can be calibrated to maximize the convergence radius, thereby enabling a more stable
fitting process at high redshifts [2082–2084]. The benefits of Padè polynomials have recently been exploited in various
theoretical contexts to explore potential deviations from Einstein’s gravity [2085, 2116, 2117]. Furthermore, the cos-
mographic method based on Padè polynomials, as well as the auxiliary y-redhift parametrization, has been adopted
to evaluate in a model-independent way the cosmological tensions in the H0 and σ8 measurements suggested by recent
observations [2118]. Another strategy to extend the convergence radius of the cosmographic series and address the
subjectivity in truncating the expansion, which may still be an issue with the Padé method, involves using Chebyshev
polynomials. It has been shown that these polynomials can significantly reduce the uncertainties in higher-order terms
of the cosmographic series, thus offering a precise description of the Universe’s evolution at late-times [2119, 2120].
The Weighted Function Regression Method, first employed in Ref. [422], addresses the subjectivity involved in
choosing the truncation order of cosmographical expressions. It does so by considering all cosmographical orders
when reconstructing cosmological functions, and it applies robust Bayesian weights to penalize the use of additional
parameters. In practice, the contribution of higher orders can be neglected since their weights are significantly
suppressed. This method has been used to reconstruct the Hubble and deceleration functions, as well as to estimate
the Hubble and deceleration parameters in a model-independent way with low-redshift data, see Ref. [422] and [2121],
respectively.
3.3.6. Non-parametric methods
Non-parametric reconstructions are data-driven methods that do not rely on a specific functional form of the
reconstructed function and hence strive to minimize model assumptions. In other words, the aim is to describe the
data without imposing strong a priori constraints. Among the most common applications of non-parametric methods
we find GP regression, AGBs, the local regression smoothing method and the iterative smoothing method. These
methods vary in their flexibility to reconstruct highly non-linear relations between variables and how the uncertainties
on the reconstruction are obtained. We proceed to describe them and their applications in cosmology.
3.3.6.a. Gaussian Processes Both powerful and versatile tools in ML and statistical analysis [2122, 2123], GP
are a Bayesian technique that generalizes distributions over functions, extending Gaussian distributions into function
space [2124]. This approach allows for the reconstruction of a function f (x) at every point x using an observational
data set {(xi , f (xi ) + σi ) | i = 1, . . . , N }, without needing to assume a predetermined specific functional parameteri-
zation. The reconstructed function and its derivatives are Gaussian random variables with a mean µ(x) and variance
cov [f (x), f (x)] at each data point x. The functions at different points x and x̃ are related by a covariance function
K(x, x̃), commonly referred to as a kernel, which depends on a small set of hyperparameters. Although there is a
wide range of covariance functions available in the literature [2123, 2125], the hyperparameters are generally con-
137

stant, as their values characterize the function’s smoothness rather than modeling such behavior. Most cosmological
applications of GP use a zero mean function and optimize the hyperparameters [2126]. A strict Bayesian approach
is to marginalize over the hyperparameters. As for the mean function, while stationary processes can be described
by a constant mean function, non-stationary ones such as the cosmic distances or growth may not. The authors
in Ref. [2127] showed how marginalizing over a reasonable family of mean functions and over the hyperparameters
yields a robust and unbiased result. An application of this full marginalization approach is to use GP as a forward
model. One can generate untrained GP by sampling the hyperparameters, and calculate the likelihood. This allows
forward modeling of quantities such as the growth rate f σ8 (z) and distances, while combining several datasets [2127–
2133]. In cosmology, GP techniques have been successfully employed to reconstruct the dynamics of cosmological
functions with minimal physical assumptions about the Universe’s geometry or the nature of its main components.
These include the background the expansion rate of the Universe H(z) [419, 422, 423, 429, 2037, 2134–2138] as can
be seen in Fig. 64, the deceleration parameter q(z) [1371, 2139–2141], the distance duality relation [2142–2144], and
the cosmological jerk parameter [2145]. To explore the dynamics of DE, examples include reconstructions of the DE
energy density [2036, 2048, 2131], the scalar field dynamics [2146], the equation of state w(z) [425, 2147, 2148], and
interaction between DE and DM [2149–2152]. These applications have served to study MG models [2136, 2153–2165],
cosmic curvature [2166–2173], and GWs [671, 2174–2176], inflaton speed of sound’s profile [2177], and many other
perspectives [424, 430, 2090, 2178–2187]. In particular, the authors in Ref. [2177] adopt a slightly different approach.
They demonstrated how mildly-informative physical priors can be imposed on a GP reconstruction in a Bayesian way
so that robust constraints on physical parameters can be extracted along with a non-parametric reconstruction. In
summary, all these applications have proven particularly valuable for testing the robustness of ΛCDM and identifying
potential deviations from our current understanding of the Universe.
3.3.6.b. Artificial neural networks AGBs have also seen a steep increase in their application to cosmology (for
further information on this please refer to Sec. 3.2.1, Sec. 3.2.2, Sec. 3.2.3 and Sec. 3.2.4). The two main advantages
of AGBs over other non-parametric techniques are that they can capture (highly) non-linear relations in complex
datasets, and in principle do not require the data to follow a specific statistical distribution. For this reason, the
cosmological community has developed various neural network reconstructions of cosmological functions. Some cos-
mological functions reconstructed with AGBs include the distance modulus µ(z) from the Pantheon and Pantheon+
compilation [1984, 2009, 2012, 2014, 2022, 2024] and the Hubble function H(z) [1991, 2009, 2015, 2188]. In particular,
a novel approach was developed in Ref. [1984] where the authors train a recurrent neural network to learn the mapping
from redshift z to distance modulus µ(z) using the Pantheon SNIa sample. This model, combined with a BNN, is
also able to propagate the uncertainties into the predicted function µ̂(z) and thereby into cosmological constraints.
This approach was also implemented to calibrate GRBs at high redshift [2010]. Other applications include AGBs
reconstructions using LSST simulated data [2189], CNNs to reconstruct the BAO signal [2190], AGBs for growth
data f σ8 [1991, 2188], and velocities of rotation curves [2191]. For visual reference, in Fig. 64 an AGB was used
to reconstruct H(z). Other approaches generate AGB models for small datasets and their errors using feedforward
neural networks with Monte Carlo Dropout and hyperparameter grid optimization [1991] or GAs [2012].
As with other non-parametric reconstruction methods, the goal is to generate an AGB model based on observational
data and then compare the resulting reconstruction with theoretical predictions from parametric methods to allow
for robust conclusions. In this direction, different cosmological models have been analyzed using neural network
reconstructions. Examples include ΛCDM [1991, 2009, 2173], the CPL model [1984, 1991, 2009, 2189], Chaplygin gas
models [1984], and Horndeski gravity [2015], among others.
3.3.6.c. Iterative smoothing method The authors in Ref. [2192] used this non-parametric method to smooth
supernova data using a Gaussian smoothing function. This method aims to reconstruct cosmological quantities such
as the expansion rate, H(z), and the equation of state of DE, w(z), in a model-independent manner. The only
assumptions made by this method are the smoothing scale and a guess background model for the quantity under
study. However, it is necessary to use a bootstrapping method to determine the optimal guess model. In Ref. [2192],
the authors used an iterative method to estimate the guess model. They started with a simple cosmological model,
such as ΛCDM, as the initial guess model and then the results were used as the next step in the iteration. With each
iteration, it was expected the guess model to become more accurate, thus giving a result that is less and less biased
towards the initial guess model used. Indeed, it was noted that using different models for the initial guess does not
affect the final result as long as the process is iterated several times.
The method requires careful consideration of the smoothing scale. A very small smoothing scale gives an accurate
but noisy guess model, therefore after a few iterations, the result will become too noisy to be of any use. It’s better to
use a larger smoothing scale for smoother results. On the other hand, the bias of the final result decreases with each
iteration, since with each iteration one gets closer to the true model. The bias decreases non-linearly with the number
of iterations. In Ref. [2192], points out that after about 10 iterations, for moderate values of the smoothing scale, the
bias is acceptably small. Also, beyond this, the bias still decreases with the number of iterations but the decrease
138

250
2
1
CC data

200
H(z) [Kms 1Mpc 1]

150

100

50

0.25 0.50 0.75 1.00 1.25 1.50 1.75


Redshift z

FIG. 64: Non-parametric reconstructions of the Hubble function H(z) using CCs. Two different approaches
were used: GP regression (left) and AGBs (right). The AGB reconstruction was made following the approach
in Ref. [1991]. The analysis models observational uncertainties using Monte Carlo Dropout, as also employed in
Refs. [1984, 2173, 2189]. The cosmic chronometer data from [410] and theoretical predictions from the ΛCDM
model are included for reference.

is negligible while the process takes more time and results in larger errors on the parameters. In Refs. [2193–2195]
the method was improved to reconstruct the distance modulus in a model-independent way and then employed in
Refs. [2196, 2197], to name just a few examples. In particular, this method was applied to reconstruct the cosmic
expansion history and growth to test curvature, DE, and GR [2195, 2198–2200].
3.3.6.d. Local regression smoothing and simulation extrapolation (Loess+Simex) The LOcally wEighted Scatter-
plot Smoothing (Loess), also known as local polynomial regression, is a non-parametric method for data analysis that
does not require specifying a predefined relationship between dependent and independent variables. It generalizes
standard least-squares methods and is widely used for non-parametric simple regression in various disciplines. Loess
aims to depict the global trend of a dataset by fitting low-degree polynomials to subsets of data around each observa-
tion, giving more weight to points closer to the target observation using a Kernel-based weighting system. Typically,
first or second-order polynomials are used since higher orders do not significantly enhance results and increase com-
putational complexity. The process is applied iteratively to cover the entire data range, resulting in a comprehensive
trend depiction. Important aspects of the method include selecting the number of data points for each fit through
a smoothing parameter called span, s, determining the polynomial degree, and choosing the weight function form.
Additionally, confidence intervals are constructed by calculating the variance of the fitted values, assuming Gaussian
distributed errors. Despite some bias in the estimation, the cross-validation technique provides accurate confidence
intervals around the Loess curve. To account for the observational errors on real data, additional methods are needed.
To this end, the authors in Ref. [2201] proposed to combine the Loess method with the simulation-extrapolation
method, also called Simex, which is a simulation-based method designed to minimize bias resulting from the inclusion
of covariates that are prone to errors. Estimates are derived by introducing additional measurement errors, using a
form of resampling. This resampling helps identify the pattern of measurement error. After estimating the pattern,
final estimates are obtained by extrapolating back to the scenario where there is no measurement error.
Loess and Simex were first used together in cosmology in Ref. [2201] to reconstruct the expansion history H(z).
Later, this technique was also used to reconstruct galaxy rotation curves [2202], the distance modulus [2203], and
other cosmological quantities [2038, 2204, 2205], with also accurate results.

3.4. Bio-inspired algorithms in model selection


Coordinator: Reginald Christian Bernardo
Contributors: Anto Idicherian Lonappan, Antonio da Silva, Arrianne Crystal Velasc, Celia Escamilla-Rivera, David
Valls-Gabaud, Dinko Milakovic, Erika Antonette Enriquez, Filippo Bouche, Isidro Gómez-Vargas, J. Alberto Vázquez,
Jenny G. Sorce, John K. Webb, Jurgen Mifsud, and Renier Mendoza
139

We briefly discuss GAs (Secs. 3.4.1-3.4.2), their applications to cosmology (Sec. 3.4.3) and potential relevance to
the understanding of cosmological tensions (Sec. 3.4.4).
3.4.1. Genetic algorithm
A GA is a biology-inspired optimization strategy that mainly takes elements of natural evolution in order to single
out one solution that is the fittest from a pool of similarly naturally selected individual solutions. GAs are powerful
optimization methods, called meta-heuristic because they do not use derivatives to find the optimum, and they
guarantee to find the best solution under certain conditions, despite the challenges posed by local optimality [2206].
This has been used in a wide variety of scientific problems such as in high energy physics [2207] and GW astronomy
[2208], and is known to be able to find global optimum and resolve tiny differences between seemingly degenerate
solutions to a problem. GA is particularly well-suited to bypass issues related to complex, high-dimensional parameter
spaces and multimodal functions. In cosmology, it was introduced as a means to overcome the bias in selecting a
cosmological model in order to infer the properties of DE [2209]. This was then followed by Ref. [2210] and Ref. [2211],
which has further marketed GA as an alternative tool for cosmological analysis through the estimation of uncertainty.
An excellent recent introduction to GA for cosmological parameter estimation is given in Ref. [2212].
GAs operate with a population of individuals (possible solutions), where each individual is characterized by a
chromosome, which in turn is described by a set of genes. These individuals make up the population that evolves
over successive generations. The key ingredients of GA are described in Table IV, while the evolutionary process is
illustrated in Fig. 65.

TABLE IV: Key ingredients of GAs [2213].

Fitness-function Determines the fitness of individuals in a population. This function can be tailored to the specific
problem, such as minimizing the Euclidean distance or optimizing the likelihood in cosmological
applications. The fitness function is a predefined metric used to rank solutions.
Selection Defines the portion of the population that will advance to the next generation. A common
method is the ‘roulette wheel’ selection, where fitter individuals have higher chances of being
selected for reproduction.
Elitism Ensures that the individuals that passed the selection process can be directly added to the
next generation. Elitism guarantees that the best fitness value will not get worse after every
generation.
Crossover Refers to the process of combining the genetic information of two parents to produce offspring.
This mechanism allows for the exchange of genes between individuals, promoting the inheritance
of favorable traits while introducing genetic diversity into the population.
Mutation Involves the random alteration of an individual’s genes. This step is crucial for introducing
new genetic variations, ensuring that the population does not stagnate and continues to evolve
towards better solutions. This allows GAs to escape local solutions and obtain global solutions.

Initialization

Selection

Elitism Crossover

Mutation

New Generation

FIG. 65: Flowchart illustrating the key steps in GAs [2213].

In essence, GAs sample from a population of individuals ranked by a fitness scale. The fittest individuals are
more likely to survive and reproduce through crossover and mutation, creating offspring with improved traits. The
best or elite individuals are stored to guarantee that the population is improving through the generations. This
iterative process continues until the optimal solution, or the fittest individual, emerges. In cosmology, the gene
can be thought of as each cosmological parameter, such as w0 and wa in CPL cosmology, and the chromosome as
the string of genes, e.g., (H0 , Ωm,0 , w0 , wa ) [2212]. A GA can of course be applied in a variety of ways; notably,
140

in the pioneering works [2209–2211] the GA has been used in the context of grammatical evolution (GE), where
the elements and key ingredients of GAs are utilized to search for solutions within function spaces, alleviating the
arbitrariness of cosmological parametrization. More recently, GAs have been used to eliminate the arbitrary choice
of a kernel function in GP regression [2214] and to optimize neural networks [2012]. It has also found applications in
spectroscopic modeling [2215–2217], and in conjunction with information criteria for model selection [2218].
3.4.2. Machine learning and GA variants

Algorithm 1 GA pseudocode.
Notation: f : fitness function, [b1 , b2 ]: population range, D: population dimension, r: mutation rate, N : population size,
G: max. number of generations, m: individual index, t: generation index, M (t): population at generation t, m∗ : best solution
Input: f , [b1 , b2 ], D, r, N , G
Output: m∗ , f (m∗ )
1: Define the maximum number G of generations and the mutation rate r
2: Set the generation counter t ← 0
3: Generate an initial random population M (0) consisting of N individuals having dimension D and range [b1 , b2 ]
4: while t < G do
5: Compute and save the fitness f (m) of each individual m in the current population M (t)
6: Sort M (t) with respect to the fitness of each m
7: Define selection probabilities p(m) for each m in M (t)
8: Choose a proportion of M (t) based on the selection probabilities p(m)
9: Apply crossover on parent individuals to produce offspring
10: Apply mutation on offspring with a probability based on the mutation rate r
11: Store the new generation M (t + 1) of individuals
12: Select the elite individuals in M (t + 1) to be preserved for the next generations
13: t←t+1
14: end while
15: Return the best individual m∗ of the final generation and its fitness f (m∗ )

In this section, we dig a little further into the details on GAs in order to get a better grip of its intricacies, and
use it as a template to introduce some of its more familiar variants that have also been considered in the field. The
basic steps of GAs are summarized in the pseudocode shown in Algorithm 1, based on the notation in Ref. [2211] and
Ref. [2213].
Algorithm 1 highlights that GAs are fundamentally an optimization strategy, inspired by the principles of natural
selection and genetics. In cosmology, however, particularly through the lens of grammatical evolution, GAs transcend
traditional optimization by serving as an ML method. The resulting combined method utilizes the adaptive search
capabilities of GA and GE to incrementally evolve a grammar that approximates cosmological functions, such as
cosmic expansion and growth rates, from input data. By discovering functional forms that capture the underlying
dynamics of cosmological phenomena, GAs in this context demonstrate their dual role in both optimizing and learning,
revealing intricate patterns that may be governing our Universe. This versatility is notably owed to the pioneering
works of Refs. [2209–2211], which fully fleshed out GA’s flexibility for extracting insights on the elusive DE given
late-time data. This opens up GAs to the same wealth of applications ML has been applied to in cosmology, such as in
the calibration of very high redshift observables [2010], non-parametric cosmological reconstructions [2203], Bayesian
deep learning for DE [1984, 2219, 2220], and in testing the validity of routines such as the cosmographic approach
[2221].
Other than ML, another broad scientific discipline where GAs are prominent is optimization, specifically under
the categories of meta-heuristic, nature-inspired, and evolutionary algorithms. Within this realm, GAs have various
variants, such as memetic algorithms, which are designed to prevent premature convergence by incorporating local
search strategies [2222]. Given their inherently distributed nature, GAs stand to benefit significantly from the advent
of quantum computers and their parallelism. Quantum GAs exploit quantum mechanical phenomena such as superpo-
sition and entanglement to implement quantum evolutionary concepts and operators [2223]: quantum chromosomes,
entangled crossovers, quantum elitism, etc. The quantum nature of these algorithms introduces a non-zero probability
of adding new genetic material to the population, thereby mitigating premature convergence [2224].
In recent years, several quantum GAs have been developed [2224–2227], featuring either a fully quantum architecture
or the integration of quantum operators within a classical GA framework. Implementing these innovative GAs on
actual quantum devices has yielded enhanced performances compared to their classical counterparts, despite the
limitations of today’s noisy quantum hardware [2224, 2226]. The first application of quantum GAs in cosmological
analysis is currently underway, aiming at minimizing the chi-squared function of different cosmological probes: SNIa,
baryon acoustic oscillations, CMB [2228]. This development marks a significant step forward, combining the strengths
141

of quantum computing and GAs to tackle complex problems in cosmology.


One of the more familiar variants of GAs in the astrophysics and cosmology community is particle swarm optimiza-
tion (PSO). PSO is inspired by the social behavior of birds flocking or fish schooling, where each particle (or candidate
solution) adjusts its position in the search space based on its own experience and the experience of neighboring par-
ticles. The algorithm involves iteratively updating the velocities and positions of the particles, guiding them towards
the best solutions found so far. This method has been used in Ref. [2229] to infer cosmological parameters given CMB
data and in Ref. [2230] to calibrate semi-analytic galaxy formation models in a fixed cosmological background. PSO
has been employed even earlier in astrophysics, such as in Ref. [2231] to search for periodic orbits in three-dimensional
galactic potentials. Recently, the method has found application in GW analysis as seen in Refs. [2232].
It is now becoming clearer that GA and PSO approaches are only the tip of the iceberg, or perhaps an opening
to Pandora’s box, of mathematical optimization techniques available for astrophysical and cosmological research.
The recent review by Refs. [2232] discusses several optimization methods, including particle swarm optimization,
in high energy and astrophysics applications. In the following sections, we shall introduce a Bayesian tool inspired
by PSO, the Approximate Bayesian Computation-Sequential Monte Carlo (ABC-SMC) [2233–2235], which has been
applied recently in cosmological model selection and parameter estimation [2236, 2237]. Work on a comparative
study of nature-inspired optimization methods applied to cosmological parameter estimation, including MCMC, GA,
Improved Multi-Operator Differential Evolution [2238], and the Philippine Eagle Optimization Algorithm [2239], is
in progress, showing optimistic results for parameter estimation and cosmological reconstruction [2240].
3.4.3. GA for cosmology
In cosmology, GAs have mostly been used in the context of grammatical evolution [2209–2211]. This was motivated
by the fact that the Universe’s most dominant component at late-times is dark, and a non-parametric reconstruction
via GA-GE seems particularly well suited for such a problem. In grammatical evolution, one deals with a set of
grammar functions, often polynomials are considered as well as trigonometric, logarithm, and exponential functions,
that are exposed to the GA optimization scheme. The output of GA-GE is thus a functional approximation of the
“true” function that has spawned the observed data; this can then be characterized with a fundamental model, such
as ΛCDM or w0 wa CDM CPL, if the data were cosmological in nature. Uncertainty estimation in the context of
GA-GE was also studied in Refs. [2209–2211], suggesting various converging estimates by a Fisher matrix formalism,
bootstrapping a la Monte Carlo, and a path integral approach.
Recently, GA-GE and other ML strategies were used in Ref. [2241] to draw insights from late-time expansion
data. As with any ML approach, however, the output of the reconstruction via GA-GE can generally be quite
dependent on the choice of the hyperparameters, for GA-GE, this would be the grammar functions. This can be
put to good use if there is prior knowledge of the functional behavior of the underlying fundamental model, such
as in Ref. [2242] where trigonometric grammar functions have been utilized to search for oscillating features in the
primordial power spectrum. This turned out to be a promising avenue to address simultaneously the tensions in
the Hubble constant and the matter power spectrum [2243, 2244]. Furthermore, the GA-GE approach has been
applied to SN, baryon acoustic oscillations, CC, RSD, and normalized Hubble rate measurements to symbolically
reconstruct the cosmological functions in Ref. [2245], and was also used to forecast constraints on the ΛCDM model
in Refs. [2246–2248].
Another way to use GAs in cosmology that has gained some traction recently is to directly use it to estimate
the cosmological parameters, e.g., phenomenological DE parameters w0 and wa , in a given model [2212]. In this
way, the uncertainty can similarly be estimated using a Fisher matrix approach or via bootstrapping, resulting in
an approximation to the posterior of the cosmological parameters, that can be used with MCMC to strengthen
the robustness of the analysis. We illustrate this very briefly with a quick GA computation (Fig. 66), to show
the constraints obtained by GAs and MCMC in spatially-curved ΛCDM and flat CPL given cosmic chronometer
measurements [2249] and SN observations [32, 33, 2250]. For the GA parameters, we consider the likelihood L as a
fitness function, a “tournament” type selection with a rate of 30%, an adaptive mutation rate of (80%, 20%), and a
“scattered” crossover type with a probability of 50%. In our notation for adaptive mutation, the first number in the
tuple (a, b) denotes the fraction a of the genes in a chromosome that are going to be mutated for low quality solutions
(ranked accordingly by fitness), and the second number corresponds to the fraction b of chromosomes that are to be
mutated for high quality ones.
In non-flat ΛCDM, it can be seen that the GA and MCMC results agree with each other on the inferred cosmological
parameter space to a high degree, including correlations [2251]. The GA output population can furthermore be used
to chip in to the results; as shown in this case it is also consistent with MCMC. However, in the more complex
CPL model, the analogous constraints reveal a slight hint of disassociation between the results of MCMC and GA,
particularly with the DE parameters. The disagreement is not to be worried about regardless since the deviations
are well within the inferred confidence regions of either methods. This quick exercise nonetheless shows GA in action
in cosmological parameter estimation, and reinforces one of our messages that GAs can act as a supporting tool to
142

MCMC MCMC
GA-Fisher GA-Fisher
GA
GA

0
0.6
5
0.4
m
4 0 6 2
0.2 0.3 0.3 0.4

0
0.3
5
m0

0.1
5 0 5 0
1.2 1.0 0.7 0.5
w0
0.5 0.6 0.7 0.8 0.9

2.5
de0

0.0
wa
2.5
5.0
.5
.0
.5
.0
8
4
0
6
2
0.5
0.6
0.7
0.8
0.9
0.1
0.2
0.3
0.3
0.4
72
73
73
74

.0
.5
.0
.5
.0

5
0
5
0

5
0
5
0

5.0
2.5
0.0
2.5
0.1
0.3
0.4
0.6

1.2
1.0
0.7
0.5
72
72
73
73
74
H0 m0 de0 H0 [km s 1Mpc 1] m w0 wa

(a) Curved ΛCDM (b) CPL (w0 wa CDM)

FIG. 66: Constraints on a spatially-curved ΛCDM and a flat CPL (w0 wa CDM) models with CCs [2249] and SN
[32, 33, 2250]; blue corresponds to MCMC results; black to GA-Fisher (GA Fisher matrix uncertainty estimation
hybrid); orange are the localized samples (trimmed outliers) in the GA final population containing the best solu-
tion.

MCMC for cosmological analysis. This is especially true for models that are more complex than ΛCDM, despite the
different nature of MCMC methods that sample the posterior distribution, while GAs maximize likelihood.
GAs and their variants have been considered in applications that assess improve the output of other methods,
such as in optimizing Neural Networks [2012, 2018] and speeding up Bayesian inference [1932]. Similarly ABC-SMC
and GA have been applied to tackle the issue of kernel selection of GP [2252] when it is applied for cosmological
reconstruction [2214, 2253]; leading toward an ML hybrid that no longer depends on any one arbitrary choice of
a kernel function. Furthermore, ABC-SMC has shown to be quite useful for parametric reconstruction and model
selection, as shown in Ref. [2236], in the end catering its users with one model and its parameters that best suit the
data. Another notable application of GAs is given in Ref. [2254] to test the viability of an emergent universe scenario
using Planck data and BBN.
3.4.4. GA for cosmological tensions
We emphasize that GAs for cosmological parameter estimation are not meant as a replacement for MCMC. GAs are
rather a support tool for the traditional MCMC, offering unique advantages given its ability to navigate complex, non-
linear, and high-dimensional parameter space, for example, to have faster parameter estimations and more information
to define the priors of an MCMC. Subsequently, GAs can become very useful in tackling cosmological tensions and
systematics, especially given the fact that biases might not be traceable by other methods. This has been teased out
in recent results to be discussed.
Ref. [2242] used GAs to search for local features in the primordial power spectrum using Planck data. GAs
were applied as a reconstruction tool by using its main ingredients in a grammar space of functions, such that the
overall implementation looks for the most suitable parametrization of the power spectrum all while constraining the
parameters of each parametrization. It resulted in significant improvements to the fit, compared with prior Planck
likelihoods [2242]. Such local features were shown to be able to reduce simultaneously both the Hubble and the σ8
tensions, giving a possible unified fundamental solution to the cosmic tensions traceable to inflation [2243, 2244].
Notably, GA’s capacity to break degeneracy between otherwise acceptable solutions has proven to be an invaluable
lens for identifying local features in the primordial power spectrum.
On the other hand, Ref. [2255] used GAs to constrain the comoving sound horizon at the baryon drag epoch. They
also checked it against traditional recombination codes or numerical fitting via the Eisenstein-Hu approach. GAs were
again applied in an independent cosmological model fashion by working on grammar functions. Subsequently, GA
143

constraints to the sound horizon at baryon drag prevented usual cosmological parameter biases that affect traditional
methods. Because the Hubble constant from BAO measurements relies on constraints to the sound horizon at baryon
drag epoch, GAs may well be able to help in understanding the cosmic tensions and the role of systematics. The
observable is also directly influenced by the matter density, giving this a route for GA to chip into the σ8 tension.
Ref. [2256] furthermore used GAs to test cosmology at the perturbative level. They used GAs to constrain growth
rate observations, based on a synthetic ΛCDM cosmological model. GAs led easily to the reconstruction of the
underlying cosmology and ruled out other often considered phenomenological and MG models including wCDM,
designer f (R), and the Hu-Sawicki model. This impressive application highlights another aspect of GAs—the ability
to see a signal through noise, even when the noise is as huge as in growth rate observations. [2257] used GAs together
with cosmological background and perturbations data, leading to a support for phantom DE behavior in the dark
sector in order to alleviate the cosmic tensions.
Ref. [2237] used a GA variant ABC to pit together Hubble constant priors that are represent of the Hubble tension.
This has seen the solution consistently evolve toward the direction of the Planck values, independent of the shape of
the priors and the data sets (CCs, SN, and RSD) used in combination. However, this is heavily influenced by the
use of CCs which were considered as a baseline data set in the analysis, since CCs prefer low values of the Hubble
constant. Nonetheless, this serves as another example of when GAs can be used to choose which priors are more
consistent for given data sets. An earlier work has used ABC in the same vein [2236], but in a more extensive way by
using the GA variant to select between ΛCDM and phenomenological parameterizations of DE, sometimes referred
to as XCDM. The results of this work have been quite intriguing, all evolved toward XCDM with a preference for
Hubble constant values consistent with the Planck constraint, regardless of the choice of priors, including Hubble
constant priors, and data sets considered. However, this has similarly relied on CCs as an anchor of the data, and the
resulting low values of the Hubble constant may be traced to this.

CDM-MCMC wCDM-GA-Fisher *
CDM-GA-Fisher
wCDM-MCMC
wCDM-GA-Fisher wCDM-MCMC *
CDM-MCMC *
CDM-GA-Fisher *
wCDM-MCMC * CDM-GA-Fisher *
wCDM-GA-Fisher *
ABC (wCDM/ CDM) CDM-MCMC *
ABC (wCDM/ CDM) *
wCDM-GA-Fisher
8
0.8

wCDM-MCMC
m0/0.3

CDM-GA-Fisher
0.8
8

CDM-MCMC
2
S8 =

0.7

ABC (wCDM/ CDM) *


4
0.6

64

68

72

76

ABC (wCDM/ CDM)


0.6

0.7

0.8

0.8

67.5 70.0 72.5 0.72 0.75 0.78 0.81 0.84


H0 [km s 1Mpc 1] S8 = 8 m0/0.3 H0 [km s 1 Mpc 1] S8 = 8 m0/0.3

FIG. 67: H0 and S8 = σ8 Ωm,0 /0.3 constraints on a spatially-flat ΛCDM and wCDM models using MCMC, GA,
p
and ABC. Data sets used are CCs [2249], standardized distances (Pantheon+/SH0ES [32, 33, 2250]), growth rate
measurements (compiled in Ref. [2258]), and baryon acoustic oscillations (DESI year-1 [698]). Red vertical bands
are from Planck [192], blue from KiDS-450 [696, 849, 857, 899], and gray from Ref. [34]. An asterisk in the super-
script in the label denote that the Pantheon+/SH0ES data was not considered. In the ABC label, the bold faced
model is the preferred one.

Fig. 67 presents a comparison of H0 and S8 = σ8 Ωm,0 /0.3 [6] constraints from MCMC, GA, and ABC-SMC using
p
a combination of late-time data. The results highlight the ability of GAs, discussed extensively in this review, to
address cosmological tensions by corroborating MCMC outcomes and even providing its own distinctive results. All
cases examined show remarkable consistency between GAs and MCMC, with GAs successfully capturing correlations
within the parameter space. Notably, all independent methods considered generally support a low S8 and a high H0 ,
although this conclusion is contingent on the datasets used. Additionally, ABC, which compares ΛCDM and wCDM
144

while constraining their parameters, demonstrated narrower posterior distributions compared to MCMC, particularly
when the algorithm extends over numerous generations before selecting a preferred model [2236]. Nonetheless, ABC
constraints remain consistent due to the method’s capacity to mitigate prior dependence [2236, 2237].
To sum up, GAs offer several significant advantages as they are particularly effective for solving complex optimization
problems where traditional methods falter. GAs are robust and flexible, capable of finding optimal or near-optimal
solutions in large, multi-dimensional search spaces. GAs are also inherently parallel, allowing to explore multiple
solutions simultaneously and increase the likelihood of discovering high quality solutions efficiently. Understandably,
GAs also have some disadvantages, such as computational efficiency, requiring significant processing power and time,
especially for large-scale problems. Its performance also relies heavily on the choice of its hyperparameters such as
population size, mutation rate, and crossover rate. Regardless, this is a certainly one method that the community
may find worth further investing into, for its flexibility and applicability to an incredible range of scenarios, embodied
by the various ways GAs and their variants have been applied in astrophysics and cosmology.
When looking for a needle in a haystack, one may turn to MCMC to identify the most probably region where the
needle could be. Reversely, one may turn to GAs to identify this needle, systematically sifting through the haystack
and evolving towards the solution with precision and persistence. This needle may well be the solution to the cosmic
tensions, if not a huge hint to this cosmological conundrum.

3.5. Inference from cosmological simulations


Coordinator: Lei Zu
Contributors: Alan Heavens, Andrew Liddle, Anto Idicherian Lonappan, Antonio da Silva, Benjamin L’Huillier,
Chi Zhang, Daniela Grandón, David Benisty, Elena Giusarma, Filippo Bouchè, Houzun Chen, Jenny G. Sorce, Jenny
Wagner, Jurgen Mifsud, Luz Ángela García, Marika Asgari, Nikolaos E. Mavromatos, Oleksii Sokoliuk, Ruth Lazkoz,
Vasiliki A. Mitsou, and Wojciech Hellwing

In this section, we explore cosmological simulation and their connection to cosmological tensions. We begin by
introducing the role of cosmological simulations in modeling nonlinear scale physics and the relevant tools. Next, we
examine how cosmological simulations serve as a powerful tool for investigating physics beyond the ΛCDM. Finally,
we discuss key observations and associated tensions at small scales, showing how cosmological simulations bridge
theoretical predictions and observational data.
3.5.1. Cosmological simulation
When studying the large-scale structure evolution of the Universe, the density perturbation δ = δρρm becomes much
m

greater than unity (δ ≫ 1) at small scales like k ≳ 1h/Mpc. In this nonlinear regime, linear perturbation theory fails
to accurately describe the evolution of structures. To overcome this limitation, N -body simulations, a particle based
method, are employed to model the behavior of phase space particles-representing a group of DM particles within
a given cosmological framework. Additionally, some cosmological simulations incorporate hydrodynamics to account
for the effects of baryons. These simulations produce detailed predictions of structure formation, enabling direct
comparisons with observational data. While computationally intensive, cosmological simulations are essential for
understanding the Universe at scales dominated by nonlinear effects. They serve as a critical link between theoretical
cosmological models and observations. By providing insights into nonlinear physics, cosmological simulations play a
pivotal role in addressing tensions between the standard ΛCDM model and observational data.
3.5.2. Nonlinear effects
N -body simulations describe structure formation at scales dominated by the nonlinear effects of gravity. In these
simulations, “particles” represent DM, characterized by properties such as mass, position and velocity. The initial
conditions for these simulations are derived from theoretical models of the early Universe, typically set at high red-
shifts (z ∼ 100) when the Universe was nearly homogeneous. On small scales like k ≳ 10h/Mpc, hydrodynamic effects
become significant. The inclusion of hydrodynamics is necessary in simulations to accurately capture the influence
of baryonic processes. GADGET (GAlaxies with DM and Gas intEracT) is one of the most widely used N -body
and hydrodynamics simulation codes [2259–2261]. It combines N -body methods for gravitational interactions with
Smoothed Particle Hydrodynamics (SPH) for modeling gas dynamics, enabling the study of both collisionless systems
(e.g., DM) and hydrodynamic phenomena. Building on GADGET, other advanced codes have been developed,
such as AREPO [2262], which uses a moving mesh approach instead of SPH, and GIZMO [2263], which replaces
modern SPH with a mesh-free hydrodynamics solver for improved accuracy. SWIFT is also an fully open-source
highly-parallel, versatile, and modular coupled hydrodynamics, gravity, cosmology, and galaxy-formation code [2264].
In addition, Adaptive Mesh Refinement based codes such as RAMSES (Refined Adaptive Mesh with Static Ex-
pansion) [2265] are also extensively used for cosmological and astrophysical simulations. Due to the computational
intensity of high-resolution simulations, several groups have made their results publicly available. Examples include
145

GIZMO
104 linear
Halofit

P (k)(Mpc/h)3
103 HMCode

102

101

100
1
10 100 101
k(h/Mpc)
FIG. 68: The matter power spectrum with the same ΛCDM cosmological parameters for k ∼ 0.1 − 10 h/Mpc at
z = 0. The green solid line shows the linear results. The black solid line represents the N -body simulation results
by using the publicly available code GIZMO while the orange dotted (blue dash-dotted) line are calculated by the
nonlinear analytical code Halofit (HMCode).

the Millennium Simulation [2266], Horizon Run Simulations [2267], Illustris and IllustrisTNG [2268, 2269],
Bolshoi Simulation [2270], EAGLE [2271, 2272], Quijote Simulations [2273]. Researchers can utilize these sim-
ulated results for their own science targets.
Using some N -body simulation results as benchmarks [1018, 2274], several semi-analytic models are developed to
predict the nonlinear matter power spectrum, such as Halofit [885, 886] and HMCode [887, 2275]. These semi-analytic
models significantly reduce computational time while remaining applicable across a wide range of cosmological param-
eters. Halofit employs simulation results to construct fitting formulas, while HMCode, a refined version of the halo
model, introduces modifications that enhance its consistency with simulation outcomes. Additionally, ML techniques
have been increasingly applied as emulator to predict nonlinear structure formation. For example, the Quijote Simu-
lations have been widely employed in ML applications for cosmology, providing large-scale training datasets for deep
learning models aimed at improving parameter inference and nonlinear structure formation predictions [2276, 2277].
Fig. 68 shows the matter power spectrum in the standard ΛCDM model for k ∼ 0.1 − 10h/Mpc at z = 0. The linear
matter power spectrum (green solid line) significantly underestimates the power at nonlinear scales k > 0.1h/Mpc.
Nonlinear results from N -body simulations (black solid line, GIZMO) align closely with analytical models such as
Halofit (orange dotted line) and HMCode (blue dash-dotted line). Within the ΛCDM framework, both Halofit
and HMCode achieve accuracy within 10% compared to the N -body simulation. HMCode also exhibits superior
performance when incorporating extensions beyond ΛCDM, such as baryonic feedback, WDM, massive neutrinos,
DDE, and MG [887, 2275]. However, if the initial matter power spectrum or the late-time dynamical evolution
significantly deviates from the ΛCDM scenario, the reliability of HMCode predictions is doubtful. As shown in
Fig. 69, when addressing new physics beyond ΛCDM, at nonlinear scales, both HMCode and Halofit fail to provide
accurate nonlinear corrections, see Ref. [2278] for details. In such cases, conducting cosmological simulations correctly
becomes essential to study the structure evolution in the specific model.
3.5.3. Cosmological simulations beyond ΛCDM
To explore new physics beyond the ΛCDM framework at nonlinear scales, various groups have developed cosmo-
logical simulations in different cosmological scenarios. One of the most studied extensions is the WDM model. The
free-streaming motion of WDM, due to its thermal velocity, suppresses structure formation on small scales. Numerous
cosmological simulations of WDM [1715, 2279–2281] have investigated the resulting mass distributions and halo prop-
erties. These studies reveal suppressed mass distributions below a scale determined by the mass of WDM particle,
and shallower density slopes systematically compared to CDM.
Another critical area of interest is the total mass of neutrinos, which influences the evolution of the total energy
density and introduces free-streaming effects. Several studies have incorporated massive neutrinos into cosmological
146

1.4
Emulator
N-body simulation
1.2 HMCode
HaloFit
1.0

Pχp (k)/PΛCDM (k)


0.8 z=0
0.6

0.4
n = 4, mχ = 1 MeV
0.2
σ4 = 8 × 10 −17 cm 2
0.0 1
10 100 101
k [h/Mpc]
FIG. 69: The ratio of the nonlinear matter power spectrum at z = 0 in DM-proton scattering scenario [2278]. The
red solid lines denote the emulator developed in Ref. [2278], the black dashed lines are the N -body simulation re-
sults from GIZMO, the blue dashed-dotted lines are the correction from HMCode and the orange dotted lines are
the correction from HaloFit.

simulation codes using additional particle sets or relativistic approaches [2282–2284]. A higher total neutrino mass
results in a stronger suppression of the matter power spectrum at small scales. For instance, the suppression of
the total matter power spectrum at k ∼ 1h/Mpc is approximately 5% for mν = 0.06 eV, and increases to 25% for
mν = 0.3 eV, see Ref. [2283] for details.

In addition to neutrino effects, cosmological simulations have also been extended to study alternative gravity models
and DE effects beyond CDM. Various computational tools have been developed to simulate structure formation under
MG theories, including ecosmog [2285], mg-gadget [2286], me-gadget [2287], and mg-arepo [2288], mg-pcola
[2289], mg-quijote [2290]. ECOSMOG incorporates MG models, including f (R) gravity, chameleon models, and
symmetron theories, using an AMR framework to achieve high-resolution nonlinear simulations. me-gadget allows
for an arbitrary phenomenological (with varying H(a) and GN (a)) and IDM (with varying mDE (a)) models to be
used in the gravity-only simulation setting. mg-pcola is a fast N-body solver extending the COLA method to MG,
balancing speed and accuracy for large-scale structure studies. Recent mg-pcola simulations incorporating massive
neutrinos reveal that their inclusion leads to additional suppression of structure growth. In particular, in scalar-tensor
models, this suppression enhances the damping of small-scale power beyond what is expected from standard neutrino
free-streaming, reinforcing the interplay between neutrinos and MG. mg-quijote is a large suite of MG simulations
tailored for ML applications and parameter inference, covering various MG scenarios and offering a benchmark dataset
for cosmological studies.

Furthermore, if DM has additional non-gravitational interactions—for instance, with baryons/electrons [2278, 2291,
2292], neutrinos [2293–2295], or dark radiation [2296–2298]—this will introduce additional pressure that suppresses
small-scale structure formation. These interactions modify the Boltzmann equation in the early Universe by adding
energy and momentum exchange terms, creating a pressure that counteracts gravitational collapse. This suppression
occurs primarily during the early Universe, when matter densities are high enough to trigger collisions. As the Universe
expands, the interaction rate falls below the Hubble rate, i.e., DM decouples. After that, the interactions become
negligible on galactic scales. Thus the primary deviations between IDM and CDM arise during the early Universe,
influencing the initial conditions for structure formation. A common approach involves calculating the linear evolution
of the matter power spectrum first, including DM interactions, during the early phase when density perturbations
remain small and linear theory is applicable. These results are then used as initial conditions (at z ∼ 100) for
later cosmological simulations, when DM interactions can be neglected. Notably, interactions often imprint unique
oscillation features in the initial power spectrum, similar to those observed in baryons. However, these oscillations
diminish during nonlinear evolution as the Universe evolves. As a result, these distinctive oscillation structures are
expected to be observable primarily at high redshifts [2278, 2298, 2299].
147

3.5.4. Cosmological simulations in nonlinear observation and cosmic tension


Cosmological simulations model particle behavior and predict DM properties at nonlinear scales, providing a crucial
tool for comparing theoretical predictions with observations such as WL, Lyman-α, halo density profile, upcoming
cosmic 21cm, and so on. Various discrepancies between ΛCDM model and observations have emerged at different
scales, including the S8 tension, core-cusp problem, too-big-to-fail problem, the diversity of dwarf galaxy rotation
curves, missing satellites problem, and so on. Cosmological simulations serve as a powerful bridge between theory and
observations, helping to address discrepancies and exploring the physics beyond standard ΛCDM. In this section, we
discuss key nonlinear observations and the related tensions between ΛCDM and the measurements.
3.5.4.a. Weak lensing WL provides a direct method for mapping the late-time large-scale structure of the Universe
by statistically analyzing the shape distortions of numerous background galaxies induced by foreground matter fields.
By calculating the two-point correlation functions of the shapes of galaxy pairs, WL measurements have widely
been used for exploring the matter distributions in our local Universe. WL is sensitive to the scale down to k ≳
1 h/Mpc at low redshift(typically z < 2), where nonlinear gravitational evolution significantly affects the matter power
spectrum [854, 855, 859, 861, 867]. On smaller scales (k ≳ 10 h/Mpc), baryonic processes, such as gas cooling, feedback
from star formation, and AGN, leading significantly systematic uncertainties to WL measurement. To mitigate
uncertainties in baryonic physics, many analyses mask or marginalize over these small-scale regions [861, 867, 985].
Several other approaches have been developed to address this uncertainty. For example, one method that preserves
small-scale measurements is the Bayesian Model Averaging (BMA) framework presented in Ref. [848]. In this approach,
BMA combines the individual posterior distributions of multiple competing baryonic feedback models derived from
hydrodynamical simulations to provide robust cosmological constraints. Furthermore, the nonlinear evolution of
cosmic structures at late-times introduces strong non-Gaussianities in the matter density field, requiring estimators
beyond two-point functions to capture this additional information [828]. These estimators, often referred to as
higher-order or non-Gaussian statistics, are useful for breaking parameter degeneracies [2300, 2301], self-calibrating
systematics [979], and overall improving cosmological constraints compared to two-point functions [836, 841, 843,
846, 2302, 2303]. The theoretical modeling of these estimators relies heavily on cosmological simulations, both
to characterize their dependence on cosmological parameters and to build the likelihoods, where the covariance
matrix is typically estimated from hundreds of simulations at a fixed cosmology [879]. An alternative to adding
estimators of higher-order statistics is to use the full information from the field using Bayesian hierarchical modelling,
e.g., see Ref. [2304–2308], or Bayesian simulation-based inference based on neural compression of the field, e.g., see
Ref. [2309, 2310], which avoid the need to estimate covariance matrices. Both these approaches rely heavily on
simulations. We refer the reader to Sec. 2.1 for a detailed discussion on WL.
In ΛCDM framework, WL surveys have reported an S8 tension (see Sec. 2.1 for details): the S8 parameter measured
from WL survey is lower than the value inferred from Planck CMB data, with a significance of 2-3σ [858, 867, 2311].
Despite some uncertainties in observational measurements, various beyond ΛCDM models have been proposed to
address this issue, such as suppressed small-scale matter power due to interactions between DM and dark radiation
[2298, 2312–2314], DM-baryon interactions [2315], or DM-neutrino scattering [2316, 2317]. Cosmological simulations
play a crucial role in addressing these issues. Upcoming more precise large scale surveys like Vera C. Rubin Observatory
[893], Euclid [437] and China Space Station Telescope (CSST) [2318] are expected to provide crucial insights and help
resolve these tensions.
3.5.4.b. Lyman-α forest Lyman-α forest refers to a series of absorption lines observed in the spectra of distant
RGBs. These lines are caused by the interaction of light from the RGB with intervening clouds of neutral hydrogen
gas along the line of sight. As a result, the Lyman-α forest is widely used as a tracer of neutral hydrogen gas
distribution and, by extension, the underlying matter distribution, assuming a correlation between mass and neutral
gas. The Lyman-α forest probes the matter power spectrum in the mildly nonlinear regime over a broad range of
redshifts, down to scales of 1 − 80 h/Mpc (at z = 2 − 6 in ground-based observation) [2319–2321]. Compared to WL
surveys, the Lyman-α forest is sensitive to smaller scale structures at higher redshifts. Since the Lyman-α forest
mainly traces the IGM in the cosmic web, directly correlated with the neutral hydrogen gas distribution, its signal is
highly influenced by the baryonic physics [2322]. High-precision hydrodynamical simulations that accurately account
for baryonic physics are essential for studying the Lyman-α forest. Using these simulation results, researchers can
investigate physics beyond the ΛCDM, including WDM [1543, 1546, 2321, 2323], fuzzy DM [1544, 1556, 2324] and
neutrino mass [2325, 2326]. Recent simulations suggest that small tensions may exist between ΛCDM cosmological
simulation and Lyman-α observation (see Sec. 2.3.8 for the anomalies in Lyman-α measurements), which can be
explained by mechanisms such as the dark photon heating or DM-neutrino interaction [2327, 2328]. An improved
understanding of baryonic physics is crucial to addressing these issues and refining our interpretation of Lyman-α
observation.
3.5.4.c. Cosmic 21 cm The cosmic 21 cm signal refers to the hyperfine transition of neutral hydrogen atoms,
which emits or absorbs radiation at a wavelength of 21 cm (frequency of 1.42 GHz). This signal is a powerful tool
148

for probing the evolution of the Universe at high redshift, across a wide range of cosmic epochs, from the Dark Ages
(prior to star formation) to the EoR and beyond [2329, 2330]. The 21 cm signal traces the distribution of matter
and can map the evolution of large-scale structures over cosmic time. Although the angular resolution of the 21 cm
signal is limited to tracing ionized bubbles on scales of several Mpc, which corresponds to very large structures, it
remains a powerful tool for exploring the nonlinear scale structure. The small scale structure significantly impacts
the halo collapsing. In general, any model that suppresses or modifies the amplitude of DM fluctuations on small
scales could affect the 21 cm cosmic dawn signal. Several previous works have discussed using the future cosmic 21 cm
data to probe the physics beyond ΛCDM [2331–2336]. One key advantage of the cosmic 21 cm signal is its ability to
trace the mass distribution at very high redshifts, where the original information is preserved due to limited impact
from structure formation [2299, 2337]. By modeling the dynamics of DM and the feedback of baryons, cosmological
simulations play a pivotal role in understanding and interpreting the cosmic 21 cm signal [2299, 2336, 2338]. Upcoming
cosmic 21 cm observations, such as the Low-Frequency Array (LOFAR) [2339], the Hydrogen Epoch of Reionization
Array (HERA) [2340], and the SKAO [2341], are expected to open a new window into the high redshift Universe and
provide novel insights into the nature of DM.
3.5.4.d. Dwarf galaxy scale puzzles At small scales, such as those of dwarf galaxies, significant tensions exist
between predictions from N -body simulations and observational data [2342]. Several key discrepancies are summarized
below:
Core-cusp Problem: Gravity only N -body simulations of DM in a CDM Universe predict that DM halos should
have a “cusp”—a sharp rise in density toward their center, known as Navarro-Frenk-White profile, or NFW [2343].
However, observations of low-mass galaxies of different luminosity and morphology [358] indicate that their DM
halos instead have a core, a central region with nearly constant density [2344, 2345].
Diversity Problem for Rotation Curves: Simulations predict that galaxies with the same maximum circular ve-
locity should have similar halo density profiles. However, observations reveal a significant variation in the central
densities of these galaxies [2346, 2347].
Missing Satellites Problem: Simulations predict a large number of small DM halos, but observations show a much
smaller number of dwarf satellite galaxies around larger galaxies, such as the MW [2348, 2349].
Too-Big-to-Fail Problem: The number of massive DM subhalos predicted by simulations is inconsistent with the
number of observed bright satellite galaxies in systems like the MW [2350, 2351].
These small-scale issues highlight a disconnect between N -body simulations and observations, particularly in our
local group. They suggest that gravity-only CDM N -body simulations are insufficient to fully describe the Universe
on these scales. Several hypotheses have been proposed to resolve these discrepancies, involving both astrophysical
processes and modifications to the standard ΛCDM model. Stellar and supernova feedback can redistribute bary-
onic matter, influencing the gravitational potential and altering the density profile of the DM and the halo mass
function [2352, 2353]. The feedback process can also eject gas from massive subhalos, suppressing star formation to
explain the too big to fail problem [2354–2358]. Another way to solve this small-scale puzzle requires modifications
to ΛCDM. Several models have been proposed such as Self-Interacting DM (SIDM) [2072] and fuzzy DM [2359].
SIDM introduces DM-DM interactions that redistribute energy, forming cores in halos, and leading to more diverse
halo density profiles and a different subhalo mass function. N -body simulations incorporating SIDM have been used
to test these predictions and compare it with the observations. Many SIDM simulations were performed and used
to discuss the DM halo distribution after the first proposal [2360–2365]. Early studies constrained the SIDM cross
section through observations, such as strong lensing [2366]. The results show that the parameter space for the SIDM
model to solve the small scale problem have already been constrained. However, subsequent simulations with higher
resolution and improved algorithms revealed that earlier constraints may have been overestimated [2367–2370]. While
for fuzzy DM, the introduction of ultra-light bosons with kiloparsec-scale wavelength naturally suppresses structure
formation at these scales. Unlike classical particles, fuzzy DM behaves more like a wave, requiring the solution of
quantum-mechanical wave equations in N -body simulations. Studies have shown that fuzzy DM naturally forms
solitonic cores at the centers of halos due to quantum pressure while simultaneously suppressing small-scale structure
formation as a consequence of the uncertainty principle [2371–2373]. This provides a compelling explanation for the
core-cusp problem and the missing satellites problem. For now, resolving small-scale issues remains an ongoing chal-
lenge, as both baryonic feedback processes and alternative DM properties beyond CDM have their own strengths and
limitations. High-resolution cosmological simulations, combined with advanced algorithms to incorporate baryonic
physics and additional DM interactions, provide a crucial tool for understanding these small-scale puzzles. These
simulations become a bridge between the theoretical predictions from proposed models and the observational data.
They help us better understand the existing cosmological tensions.
149

3.6. Profile likelihoods in cosmology


Coordinator: Adrià Gómez-Valent
Contributors: Elisa G. M. Ferreira, Emil Brinch Holm, Giacomo Galloni, Laura Herold, Paolo Campeti, Sophie
Henrot-Versillé, Vivian Poulin, and William Giarè

3.6.1. Motivation
In cosmology, it is very common to encounter high-dimensional parameter spaces. Efficient exploration of these
spaces through the sampling of their corresponding multivariate distributions is commonly achieved using Monte
Carlo techniques. In these analyses, statistical information is contained in the resulting MCMC, from which one can
infer the posterior distribution and other derived Bayesian products. In parameter spaces of dimension greater than
two one cannot visualize the posterior distribution in the original parameter space. One needs to apply the so-called
marginalization technique to project the results into spaces of lower dimension and compute the marginalized posterior
distributions of the individual parameters and their constraints at the desired confidence level. This is also done to
obtain the contour plots in all the two-dimensional planes of interest. Marginalization is extremely efficient and is a
very practical tool that eases the visualization and interpretation of the results in Bayesian analyses. However, when
the original (non-marginalized) PDF has non-Gaussian features, the marginalized posterior distributions obtained do
not exclusively reflect how the parameters are distributed based on their ability to explain the data but also according
to their integrated probability weight. This introduces volume (or marginalization) effects, which can potentially bias
conclusions in a significant way. In other words, if deviations from Gaussianity are large, the marginalized posterior
can hide points in parameter space that fit the data well. Another potential drawback of the marginalization method
is the impact of the prior, which, even if assumed to be flat, is still informative in all cases. For instance, a flat
prior on a variable x or on its logarithm, log x, might induce important changes in the constraints of a model, even
if the boundaries of the priors are perfectly consistent. This is due to volume effects, again. Hence, volume effects
can significantly impact the interpretation of results extracted from MCMCs. Detecting these biases is of utmost
importance. In the context of cosmology, accurately quantifying these biases is essential, e.g., for a correct assessment
of cosmological tensions in the ΛCDM model and beyond, as well as in data compression performed by galaxy and
WL surveys, see Sec. 3.6.4 for details.
A mismatch between the point in parameter space that maximizes the original posterior distribution and the point
constructed with the values of the parameters that maximize the individual one-dimensional posteriors already hints
at the non-negligible impact of marginalization effects. One can investigate the impact of marginalization effects
using the profile likelihood (PL). The PL is a method from frequentist statistics to construct parameter intervals,
and therefore, by construction inherently independent of a prior. It ensures, in particular, that the results (on the
parameter inference) does not depend on the choice of the parameterization of the problem. This part of the work
aims to explain the basics of PLs and is organized as follows. In Sec. 3.6.2 we describe how to build confidence intervals
using PLs. In Sec. 3.6.3 we describe several methods to compute them, both approximately from the MCMCs or, more
precisely, using advanced numerical optimization techniques with the help of concrete codes. Finally, in Sec. 3.6.4 we
give some examples of applications of the PLs in cosmology, providing a rich list of references.
3.6.2. Confidence intervals from Profile Likelihoods
While Bayesian credible intervals are statements about the degree of belief of the value of the true parameter,
frequentist confidence intervals are based on the coverage: frequentist confidence intervals are said to cover the true
value of the parameter with a confidence level α if – after repetition of the experiment – a fraction α of the experiments
contain the true value of the parameter (e.g., see Ref. [1462] for a review, or Ref. [2374] in the context of cosmology).
Confidence intervals with correct coverage can be constructed using the Neyman construction [2375]. This construction
requires knowledge of the full PDF P (µ̂|µtrue ), where µ̂ are the maximum likelihood estimates (MLE) of the model
parameters µ estimated from observations and µtrue the (hypothetical) true values of these parameters. Since this
probability is typically not known a priori, it needs to be estimated by generating mock realizations of the data. This
makes the Neyman construction computationally expensive. Fortunately, a theorem by Wilks [2376] helps in many
practical applications, which states: In the limit of a large data set, the log-likelihood ratio

L(x|µ)
∆χ2 = −2 log R(x, µ) , R(x, µ) = , (3.29)
L(x|µ̂)

follows a χ2 -distribution. Here L is the likelihood function and x is the data vector. In this case, the Neyman
construction can be replaced by a much simpler graphical construction based on the profile likelihood. Wilks theorem
holds trivially if the p.d.f. is Gaussian (for every possible value of the true parameters). We note, however, that Wilks’
theorem holds for the full-dimensional likelihood (for any possible value of the true parameters) and not the profile
likelihood. Although this effect too vanishes in the large data limit, it is often very difficult to assess whether this
150

limit is reached in practice [2377], and the confidence intervals are, therefore, only approximate (see Ref. [2374] for
checks of the asymptotic assumptions in cosmology).
The profile likelihood for a parameter of interest µ (in one dimension) can be obtained by evaluating Eq. (3.29) for
different values of µ. In the presence of other cosmological parameters and nuisance parameters, ν, the profile likelihood
is calculated by setting µ to various values within a range of interest and minimizing χ2 (µ, ν) = −2 log L(µ, ν) with
respect to all parameters ν. The global MLE, or “best-fit”, corresponds to the minimum value χ2min by design. A
graphical frequentist confidence interval for µ can now be constructed using the difference ∆χ2 (µ) = χ2 (µ) − χ2min .
When µ is far from its physical boundary, a confidence interval at confidence level α is obtained by applying a
fixed threshold ∆χ2th to ∆χ2 (µ). This threshold is chosen such that the cumulative distribution function of the χ2
distribution with one degree of freedom equals α. Notable values are ∆χ2th = 1 for 68% CL and ∆χ2th = 3.84 for 95%
CL [2378]. This method is applicable to both parabolic (i.e., for a Gaussian-distributed parameter) and non-parabolic
∆χ2 (µ) because of the invariance of the MLE under reparameterization.
When the parameter estimate is near its physical boundary, the conventional graphical profile likelihood construction
for frequentist confidence intervals can be inadequate. It may result in empty intervals and fail to maintain the
frequentist coverage property if the choice between an upper limit or a two-sided interval is made based on the data
(the so-called “flip-flopping”). An adapted Neyman construction, also known as Feldman-Cousins (FC) prescription
[2379] in the particle-physics literature, addresses these issues. To describe this method, we restrict the discussion to
one-dimensional data x and a single model parameter µ for clarity. For each value of µ (with an unknown true value)
and each observable x, we compute the likelihood ratio R(x, µ), with µ (and its MLE µ̂ ≡ µbest ) restricted to its
physically allowed values. The confidence belt at the desired CL α is constructed by selecting an acceptance interval
[x1 , x2 ] for each µ satisfying
(
R(x1 , µ) = R(x2 , µ) ,
´ x2 (3.30)
x1
P (x|µ) dx = α ,

where P (x|µ) is the p.d.f. of x given µ, and values of x are added to the acceptance interval in order of decreasing
likelihood ratio. The confidence belt is formed by the union of all acceptance intervals [x1 (µ), x2 (µ)]. Intersecting
the confidence belt with a line at x = x0 , where x0 is the value of x minimizing χ2 (i.e., the value observed in an
experiment), yields the confidence interval [µ1 , µ2 ] for µ. The adapted Neyman or FC prescription provides a method
to determine the endpoints of confidence intervals, which smoothly transitions between an upper limit and a two-sided
interval, ensuring exact frequentist coverage even for parameters with non-Gaussian p.d.f.’s (and if Wilks’ theorem
does not hold). This is in contrast to the conservatism (overcoverage) inherent in Bayesian limits in the same context
[2379, 2380]. While overcoverage might not be as problematic as undercoverage, it reduces the ability to reject false
hypotheses. This highlights the value of examining frequentist intervals.
If the p.d.f. of the parameter is Gaussian and the MLE of the parameter near a physical boundary at µ = 0, then
µbest = max(0, x) and the likelihood ratio in simply becomes [2379]
(
exp(−(x − µ)2 /2), for x ≥ 0 ,
R(x, µ) = (3.31)
exp(xµ − µ2 /2), for x < 0 ,

where x and µ are in units of σ, the width of the parabolic fit to ∆χ2 (µ). Then, the confidence interval can be
derived solving Eq. (3.30), with P (x|µ) being a Gaussian with mean µ and unit variance. If, however, the p.d.f. of the
parameter is non-Gaussian and Wilks’ theorem doesn’t hold, a full Neyman construction of the confidence intervals
using simulations for several input values of the parameter of interest is necessary, e.g., see Ref. [2381, 2382] for details.
With the necessary modifications, this entire treatment can be generalized to multiple dimensions. For example,
under the assumption of Gaussianity, the iso-χ2 curves for a two-dimensional case will be ellipses on the parameter
space, whose angle of the major axis depends on the correlation of the two parameters of interest. Furthermore, if no
physical boundary is close to the absolute minimum of χ2 and if we consider the limit of a large data set, it is possible
to adopt a graphical approach to derive the 68% and 95% confidence regions, as we mentioned for the one-dimensional
case. This time, one must search for ∆χ2th = 2.3 for 68% CL and ∆χ2th = 5.99 for 95% CL regions. Similarly to the
one-dimensional case, if Wilks’ theorem doesn’t hold, one should adopt a full Neyman construction to extract the
confidence regions.
3.6.3. Calculation of profile likelihoods
Computing a profile likelihood for a single parameter involves a series of M optimizations, typically with M ∼ O(10),
in the N − 1 dimensional parameter space. Furthermore, these cannot be reused for profiles in other parameters,
unlike an MCMC, which simultaneously produces parameter constraints for all varied parameters. Since the opti-
mizations themselves are moderately difficult, constructing a full profile likelihood analysis of several parameters can
151

be very computationally demanding. In this subsection, we present some of the strategies adopted in the literature
to accommodate the often high computational cost of computing profile likelihoods.
3.6.3.a. Optimisation strategies A converged set of MCMC chains of course contains information about the like-
lihood values at a large number of points in parameter space. Although MCMC algorithms are poor optimisers [1964],
they can be used to estimate profile likelihoods by binning their points along the parameter θ being profiled. The
bin around a fixed value of θ = θ0 can be defined, for example, by a homogeneous central binning, as is done in
Refs. [1935, 1966, 2383–2385], or by taking the n points of the MCMC that have a value of θ closest to the fixed value
θ0 , as done in Ref. [2386], with n a fixed fraction of the total amount of MCMC points. Due to the finite sampling
of the MCMC, the largest likelihood value found in each bin will generally be smaller than its optimized value at
θ0 , which leads to noise in the profile likelihood curve. The great advantage of this approach, however, is that it
has no computational cost if an MCMC is already given, so it can be used as an inexpensive initial test to check for
marginalization and prior effects [1966], see also Ref. [2384]. Furthermore, such a profile can be used as the starting
point for explicit optimizations at each fixed value of the profiled parameter.
When the binned MCMC is not adequately accurate, the profile likelihood can be computed by explicit optimization
in the parameter subspace with the profiled parameter taking on a series of fixed values. Generally speaking, the most
efficient optimization algorithms explore the parameter space for the highest likelihood value by using the gradient of
the likelihood with respect to parameter space. Examples of such gradient-based optimizations algorithms often used
in cosmology include the MIGRAD optimizer of the MINUIT package, which is a variable metric algorithm using the
finite-difference estimates of the first derivative of the likelihood function [2387]. References [2388–2392], amongst
others, used MINUIT to construct profile likelihoods and found it to be computationally efficient.
In some cases, however, the cosmological likelihood functions are noisy, for example, due to approximation switching
in the underlying theory code or insufficient precision settings, leading to inefficiency in the gradient-based algorithms.
For example, Refs. [709, 1959, 2385, 2393] found that gradient-free methods outperformed the gradient-based ones. The
most popular gradient-free method employed in the literature is simulated annealing [2394], which works by running
an MCMC using the modified likelihood L̃ ≡ L1/T with an iteratively decreasing temperature T . The decreasing
temperature enhances the peak structures in the likelihoods and eventually traps the MCMC close to the best fit.
Simulated annealing has been used for profile likelihood construction in Refs. [1950, 1957–1960, 1962, 1967, 2385, 2386],
amongst others. Unfortunately, simulated annealing is very sensitive to its hyperparameters, which include the initial
point of the optimization, the covariance matrix of the proposal distribution of the MCMC, and the rate at which
the temperature decreases. Broadly speaking, three approaches have been suggested to inform the initial point and
covariance matrix:
• Running fixed-parameter MCMCs to obtain information. Refs. [1957–1959] run individual MCMCs for each fixed
value of the profiled parameter and use the best fit and covariance matrices of their chains in the optimisation.
This method produces great initial points and local covariance matrices, although with the disadvantage of the
added computational cost of running the additional MCMCs.
• Using information from an existing MCMC. The CAMEL [2395] and PROSPECT codes [2386] assume the
user to have a converged MCMC in the full parameter space before constructing the profile. Using this, it uses
the binned MCMC profile likelihood estimate discussed in the preceding subsection, and furthermore constructs
individual covariance matrices from the points in each bin. This method usually gives great initial points and
local covariances matrices, but can be sensitive to the quality of the initial MCMC.
• Optimising the profile sequentially and reusing information from previous optimizations. The Procoli
code [2385] also assumes the availability of a converged MCMC at the beginning and uses its maximum a
posteriori point and covariance matrix to find the global best fit. Procoli then carries out optimizations at
equidistant points along the profiled parameter direction sequentially, using the result of the previous iteration
as the initial point of the next. This method yields great initial points, at the sacrifice of the parallellization of
the optimizations.
3.6.3.b. Publicly available profile likelihood codes At the time of writing, there exist several publicly available
tools for computing profile likelihoods in cosmology; here, we present them in order of the recency of their release.
• cobaya fork [2388]: Can be used with cobaya [1911]. Initialises points from an existing MCMC as described
above. Usable as a sampler inside cobaya, making it easy to use if an MCMC has already been run with
cobaya43 .

43 https://github.com/ggalloni/cobaya/tree/profile_sampler
152

• Procoli [2385]: Can be used with MontePython [1909, 1910]. Uses the sequential strategy described above.
Additionally allows for the extraction of individual χ2 values for each experiment used in the data44 .
• Prospect [2386]: Can be used with MontePython and cobaya. Uses an assumed precomputed MCMC as
initialisation, as described in the last section. Introduces an automatic step size tuning scheme, and comes with
a snapshot checkpoint scheme that enables efficient parallellization and inspections of the run45 .
• Camel [2395]: Code written in C++. The minimisation algorithm is based on MINUIT [2387]46 .
All of the above codes (except Camel) work directly with the cosmology-specifying parameter files of MontePython
or cobaya, respectively, that have been used, for example, in a complementary Bayesian analysis. They therefore
also work with the theory codes already available in these, such as CLASS [1940] and, in the case of cobaya,
CAMB [1939]. They also employ the GetDist code [1943] for analysis.
While the brute-force computation of full triangle plots of profile likelihoods, analogous to the most common product
of the Bayesian analysis, is computationally unfeasible, future advances in gradient-based inference and emulators of
likelihood codes are expected to make extensive profile likelihood analyses more broadly available. For example,
already available is the neural network emulator CONNECT47 [2396] has a run-mode that efficiently computes full
triangle plots of profile likelihoods using gradient-based optimization, as explored in reference [2397].
3.6.4. Applications of profile likelihood in cosmology
• EDE & NEDE: Due to the complicated parameter structure of EDE (see Sec. 4.1.1), volume effects can have
a strong impact on the constraints of this model. Using PLs, Refs. [1957, 1958] showed that tight upper limits
on EDE from CMB and LSS in Bayesian analyses, which questioned the ability of EDE to resolve the Hubble
tension, are partially driven by prior volume effects, while in a frequentist analysis EDE presents a viable solution
to the tension. Ref. [1959] use PLs to explore the interplay between EDE and massive neutrinos. In Ref. [1960],
both a Bayesian and PL analysis under an updated Planck CMB pipeline give tight upper limits on the fraction
of EDE. Moreover, Ref. [1962] constrained the NEDE (see Sec. 4.1.2) model and showed that the frequentist
analysis gives greater evidence of the model due to volume effects in the Bayesian posterior associated with the
ΛCDM limit.
• EFTofLSS: Ref. [1950] studied the impacts of priors on the nuisance parameters of the EFTofLSS analysis method
applied to full-shape power spectrum from galaxy surveys. They found that the priors usually employed to
constrain the parameters to their physical regime were informative and had a significant effect on the constraining
power of the analysis method.
• Decaying dark matter: Ref. [1967] studied a model where a fraction of the DM is allowed to decay into dark
radiation. With two new parameters, the abundance of the decaying part of the CDM and its lifetime, series
volume effects are found in the limits of parameter space that recover the ΛCDM model; these include the case
of infinite lifetime and vanishing abundance. In particular, whereas the Bayesian analyses prefer either of these
limits, [1967] shows that the best fit is associated with a narrow log-likelihood peak at a lifetime corresponding
to decays just around recombination.
• Phenomenological transition of dark matter to dark radiation: Ref. [2386] tested the PROSPECT code on a
model where a fraction of the DM transitions to dark radiation, characterized by three parameters: The fraction
of DM that transitions, the central redshift of the transition, and a parameter that controls the shape of the
transition [2398]. The ΛCDM limit is recovered in many limits of parameter space, leading to significant volume
effects and parameterization dependencies that were diagnosed with profile likelihoods.
• Lensing: A complete study of the so-called AL tension in the ΛCDM + AL has been done in Ref. [726].
• Neutrino: One of the prototypical applicationsPof profile likelihood in cosmology is for neutrinos. Since neutrino
mass constraints are close to the boundary mν > 0 and the non-Gaussian posteriors found MCMC data
analysis, profile likelihood was used to obtain the confidence interval on the sum of the neutrino mass. This
has been done in the context of WMAP5, Sloan Digital Sky Survey (SDSS), and HST in Ref. [2383] and for
Planck data in Ref. [2392]. In Ref. [2399] and Ref. [2400], they study the effect of priors in constraining the
number of neutrino species using cosmological data and no evidence for deviations from the standard number
of neutrino using profile likelihood. In Ref. [2399] was also highlighted the link with Alens on the current limit.
Refs. [1281, 2401] explore the preference for “negative” neutrino masses with profile likelihoods.

44 https://github.com/tkarwal/procoli
45 https://github.com/AarhusCosmology/prospect_public
46 http://camel.in2p3.fr/wiki/pmwiki.php
47 https://github.com/AarhusCosmology/connect_public
153

• Number of relativistic species, Neff : Profile likelihood was used to test the possibility that extra relativistic
species are present beyond the standard model prediction. A discrepancy between the posterior and profile
likelihood confidence interval was found in Neff in Refs. [1963, 1964], indicating volume effects when using
Wilkinson Microwave Anisotropy Probe (WMAP7) and HST data. It was applied to Planck data in Ref. [1965].
• Gravitational wave energy density: Constraints on the primordial GW density using CMB data and a study of
the impact on cosmic string models has also been derived in Ref. [2402].
• Tensor-to-scalar ratio: Profile likelihood was used to analyse the tensor-to-scalar ratio, r, with current data. In
Ref. [2391] the most recent Planck and BICEP/Keck CMB data was used in order to gain a clearer perspec-
tive on the discrepancy between Bayesian and frequentist upper limits on r previously found by the SPIDER
collaboration [2382]. A similar analysis was done in Ref. [2388] combining CMB data from Planck PR4 and
BICEP/Keck, and GW data from LIGO-Virgo-KAGRA. With the goal of testing the impact of MG theories
on r, in Ref. [2403] combining data from CMB from Planck DR4 and BICEP/Keck and BAO data from 6dF
Galaxy Survey, SDSS DR7 and eBOSS, the Bayesian and frequentist upper limits on r were obtained.
• Inflation: The advantages of using a profile likelihood, and more in general, of constructing frequentist confidence
intervals with the FC prescription, are particularly evident when trying to constrain inflationary models. This
typically allows for more informative constraints and physics insights, especially on models with numerous
additional parameters, which are typically degenerate and hard to constrain simultaneously, especially with
current data. An example is provided by the latest constraints from Planck and BICEP/Keck on axion-U(1)
inflation [2390] and by forecasted constraints on axion-SU(2) inflation parameters for the future CMB satellite
survey LiteBIRD [2381]. Another example is given by the attempt to constrain the spectral tilt of tensor
primordial perturbations with the latest CMB and GW interferometers data [2388]. Profile likelihoods were
also used to search for features in the primordial power spectrum imprinted in CMB and large-scale structure
formation observables, using data from Planck and WiggleZ [2404].
• Brans-Dicke model with cosmological constant: The Brans-Dicke model with a constant positive vacuum energy
density, the so-called BD-ΛCDM model, has been confronted with observations multiple times in the last years
[2405–2409]. In Ref. [2408] the authors showed that in the absence of the Planck 2018 high-ℓ CMB polarization
and lensing data, but still under a very rich dataset including the Planck 2018 full temperature and low-ℓ
polarization likelihoods, together with the state-of-the-art data on SNIa, BAO, CC and RSD, it is possible to
loosen the H0 tension in BD-ΛCDM, while keeping σ8 ∼ 0.790. In Ref. [1966] the author tested the impact of
volume effects in BD-ΛCDM through the analysis of the profile distribution and showed that they do not play
a major role in this model.
• Coupled quintessence: Coupled quintessence was firstly studied in Refs. [2041, 2410] considering a scalar-
mediated interaction (fifth force) between DM particles, with the scalar field playing the role of DE. Constraints
on this model have been reported e.g., in Refs. [2411–2417]. Volume effects have been recently quantified in
Ref. [1966] for the case of a Peebles-Ratra potential [295, 2418] using the profile distributions built directly from
the MCMC. Their impact on the results are found to be small.
• Neutrino-Dark Matter Interactions: Ref. [2384] studied a model featuring scatter-like interactions between
neutrinos and DM. Allowing the interaction strength and the total neutrino mass to vary, the model can have
up to two parameters more than ΛCDM. A profile likelihood analysis was employed to test possible volume
effects. In this case, no relevant volume effects were found. The profile likelihood supported the results derived
by marginalizing over the parameter space, confirming a mild preference for non-vanishing interactions when
small-scale CMB data are considered [2419, 2420].
• Baryon Acoustic Oscillations: Profile likelihood has been used, sometimes in combination with Bayesian meth-
ods, to quantify BAO errors in determining the fitting scale in surveys. This has been done in the context of
many surveys, like the Sloan Digital Sky Survey (SDSS) BOSS [2421] and the eBOSS [2422], or DES [2423, 2424].
In [2425], they extend this method to add noise and compare different approaches to determine the error in
determining the BAO scale. Profile likelihood was also used to fit the Lyman-α (Lyα) forest 3D correlation
function [2426]. Using mocks, they found that profile likelihood is a good approximation for fitting the BAO
peak from the Lyα forest correlation functions and is in full agreement with the Bayesian analysis (which is not
the case for frequentist maximum likelihood estimators, which assume Gaussianity).
154

4. Fundamental physics
Coordinator: Vivian Poulin

Cosmology stands at a crossroads. The concordance ΛCDM model, which had shown a great level of success in
describing a wide variety of cosmological observations, is being challenged by high precision data as reviewed in Sec. 2.
However, the ΛCDM model has raised profound questions about the nature of its dominant components –DM and
DE– as well as the mechanism at the origin of the primordial fluctuations (usually described by inflation) that remains
largely unknown. Barring the existence of unknown systematic errors, these observational issues thus present a crucial
opportunity to explore new physics beyond the ΛCDM paradigm. This section examines the collective efforts of the
scientific community to interpret the implications of observations that challenge the ΛCDM model, with a particular
focus on what the H0 and S8 tensions may reveal about fundamental physics.
At the core of the ΛCDM value of H0 is the angular size of the sound horizon at recombination θs (z∗ ) =
rs (z∗ )/DA (z∗ ), that CMB data have determined at O ∼ 0.1% precision [192]. In the flat FLRW metric, the an-
gular diameter distance to recombination DA (z∗ ) that carries information about H0 can be computed as
ˆ z∗
dz
DA (z∗ ) = (1 + z∗ )−1 , (4.1)
0 H(z)

where under flat ΛCDM, H(z) = H0 Ωr (1 + z)4 + Ωm,0 (1 + z)3 + ΩΛ . Extracting the angular diameter distance
p
from θs (z∗ ) requires knowledge of the sound horizon
ˆ ∞
cs (z)dz 1
rs (z∗ ) = , with cs (z) = p . (4.2)
z∗ H(z) 3(1 + 4ρb (z)/3ρr (z))

A value of rs (z∗ ) can be computed within a model, given knowledge of the cosmological parameters. Within ΛCDM,
Planck data have allowed us to determine rs (z∗ ) = 144.39 ± 0.3 Mpc (TTTEEE+lowE) from the complex structure
of peaks and troughs [192]. The value of H0 is then adjusted in order to match the corresponding value of θs (z∗ ),
which yields H0 = 67.27 ± 0.6 km s−1 Mpc−1 . Similarly, the value of S8 can be inferred from the linear matter power
spectrum P (k) given a set of cosmological parameters reconstructed from a fit to a given dataset as
r  ˆ ∞ 3 
Ωm,0 k
S8 ≡ × σ8 ≡ P (k)W 2
8 (k)d ln k , (4.3)
0.3 0 2π 3

where W8 (k) is a window function describing a sphere of radius R = 8Mpc/h in Fourier space. Under the ΛCDM
model fit to Planck data, it yields S8 = 0.834 ± 0.016.
The solutions to the H0 discrepancy generally fall into two categories based on the requirement to leave the
precisely measured value of θs (z∗ ) unaffected [708, 709, 712]. First, there are those that reduce the value of the
sound horizon rs (z∗ ) such as to compensate for the shorter angular diameter distance implied by a larger H0 . Those
typically correspond to early-time (pre-recombination) modifications and include for instance include, new relativistic
species, EDE, or modified recombination scenarios. Second, late-time solutions modify the expansion history after
recombination such that the angular diameter distance to recombination remains unaffected. These include e.g.,
phantom DE, IDE, or decaying DM. Both avenues offer advantages and disadvantages: the former scenarios may
more easily adjust the shape of the expansion history that is constrained by BAO and SNIa (beyond the mere value of
H0 ), but will be strongly constrained by CMB observations, while the latter scenarios can more easily leave the CMB
unaffected by exploiting the geometrical degeneracy, but are strongly constrained by late-time observations [709].
Resolving the S8 tension, on the other hand, requires suppressing structure growth, which can be done through new
DM interactions, evolving DE models, or modifications to general relativity. Yet, it remains to be established precisely
what scales and what cosmic era must be affected in order to achieve concordance between all our LSS observations:
is the tension solely appearing at small scales, that are potentially affected by large uncertainties related to non-linear
and baryonic effects, or does it also appear on large-scales at late-times, where the physics is much more linear and
under control? Answering these questions will drive the community effort to find a solution to the tension. Moreover,
there is a clear interplay between the H0 and S8 tensions. First, simply because the definition of the physical scale
k involved in S8 in Eq. (4.3) is trivially influenced by the value of h: as the power spectrum as a strong dependence
on the scale k (falling like k −3 ) tension in S8 may appear due to different values of h in the analysis, fooling us in
comparing different physical scales [2427–2429]. However, even when corrected (e.g., through the use of R = 12 Mpc as
an absolute scale [2427]), the tension can be influenced by the new physics introduced to resolve the Hubble tension.
This stems from the observation that distance indicators, that provide measurements of Ωm,0 and H0 , necessarily
imply a measurement of the physical matter density ωm,0 ≡ Ωm,0 h2 [803, 2430, 2431]. Hence, a larger value of h
155

will automatically imply a larger value of ωm,0 , leading to earlier matter domination and thus a larger amplitude of
fluctuations today.
Finding a solution to either or both tensions is an active area of research. As we illustrate within each section,
understanding and resolving these tensions will not only improve the precision in measuring fundamental cosmological
parameters but also offer a window into the fundamental physics governing the Universe [708]. Whether these tensions
indicate systematic biases in data or genuine hints of new physics, they represent one of the most exciting frontiers
in modern cosmology, potentially unlocking new insights into DM, DE, and the laws of gravity.
4.1. Early-time proposals
4.1.1. Early dark energy and variants
Coordinator: Laura Herold
Contributors: Adrià Gómez-Valent, Alexander Reeves, Alireza Talebian, David Benisty, Elisa G. M. Ferreira,
Gabriela Garcia-Arroyo, Ivonne Zavala, Joan Solà Peracaula, Lu Yin, Luis Anchordoqui, Luz Ángela García, Matteo
Forconi, Nikolaos Mavromatos, Rafaela Gsponer, and Vivian Poulin

4.1.1.a. Introduction and motivation EDE models (for reviews see Refs. [2432–2434]) are a promising class of
models proposed to resolve the Hubble tension with new physics. The central idea behind EDE is to introduce
an additional component to the energy density of the Universe at early-times, which increases the expansion rate
´ ∞ before recombination z . This leads to a decrease of the (comoving) size of the sound horizon: rs (z ) =
∗ ∗
H(z)
z∗ s
c (z)/H(z)dz, where cs is the sound speed in the baryon-photon plasma. To understand why this increases H0 at
present times, note that the angular size of the sound horizon θs = rs /DA is directly and precisely observed by the
CMB, e.g., see Ref. [192]. Hence, to maintain the same size of θs , the (comoving) angular diameter distance reduces,
´ z∗
DA (z ∗ ) = 0 dz/H(z), and leads to an increase of H0 .
All EDE-type solutions to the Hubble tension have in common a DE-like rapid expansion with an equation of state
w ≃ −1 around z ∗ , which peaks at some critical redshift zc and is followed by a phase in which the background
energy density decays or dilutes (see Fig. 70, left). EDE models often take the form of a cosmological scalar, whose
background dynamics follows the homogeneous Klein-Gordon equation: the field is initially frozen in its potential,
such that the background energy density is constant. The fractional contribution of the field to the total energy
density, ρEDE /ρTot , increases over time in a manner similar to that of DE. Eventually, Hubble friction dropping below
a critical value, or a phase transition changing the shape of the potential, releases the scalar, at which point the field
becomes dynamical, and the background energy density dilutes away faster than matter. The contribution of EDE
to the Hubble rate is therefore localized before recombination, where it efficiently acts to reduce the size of the sound
horizon.
EDE-like models have already been studied before the appearance of the Hubble tension, e.g., in the context of
quintessence models [2435–2438] or in the context of recurring DE [2439]. The (axion-like) EDE model has gained
significant attention as a proposed solution to the Hubble tension [2440–2442]. To fully resolve the tension, a fractional
EDE contribution of ∼ 10% around the time of matter-radiation equality is required [2443], reflecting the percentage
difference in H0 estimates between Planck [192] and SH0ES [34]. Although EDE models present a promising class of
solutions when compared to other proposed solutions [709, 712, 2444], the improvement in observational constraints
from the CMB and LSS in the last few years poses challenges to this class of models.
4.1.1.b. The axion-like EDE model The originally proposed EDE model [2440–2442] consists of a scalar field ϕ
in a potential of the form
V (ϕ) = V0 [1 − cos(ϕ/f )]n , (4.4)
where V0 = m f with m and f being referred to as “mass” and “decay constant”, respectively. The index n is
2 2

typically fixed to n = 3 as this provides the best fit to the data [2433, 2443]. The parameters m and f , along with
the initial value of the scalar field in the potential, θi = ϕi /f , need to be fit from the data. This “particle-physics
parameterization” is often traded for the “phenomenological parameterization”, which consists of fEDE , the maximum
fraction of EDE, fEDE ≡ ρEDE (zc )/ρTot (zc ), at the critical redshift, zc , which is determined by the onset of oscillations,
along with the initial value of the scalar field θi .
Predictions for EDE cosmologies have been found using the effective fluid theory approach [2442] and by directly
solving the linearized scalar field evolution equations [2443, 2447]. Extensions of CLASS [1940] using the latter approach
are publicly available4849 . CAMB [1939] includes EDE using the effective fluid approximation50 .

48 https://github.com/PoulinV/AxiCLASS
49 https://github.com/mwt5345/class_ede
50 https://github.com/cmbant/CAMB/tree/master
156

105 total
70.6 ± 1.3
EDE P lanck + BAO + H0
102 matter 72.1 ± 0.8
ρ(z) [Mpc−2]

radiation 68.3+1.0
−1.0
10−1 Λ
P lanck
70.0 ± 1.5
10−4 68.9+0.6
−0.6
P lanck + BAO + S8
70.3 ± 1.3
10−7
73.4+2.6
−3.4
0.10
WMAP + ACT + BAO + SNe
z∗ zc
68.1+0.5
ΩEDE(z)

P lanck NPIPE −0.8


0.05
68.4 ± 0.8

0.00 0 62 64 66 68 70 72 74
10 101 102 103 104 −1
z H0 [km(sMpc) ]

FIG. 70: Left: Energy densities, ρ(z), of different components of the Universe as indicated in the legend. EDE
(green) plays a subdominant rule at all redshifts, z. The fractional energy density ΩEDE (z) (bottom subplot) peaks
at the critical redshift, zc , and decays around recombination, z ∗ . Right: A comparison of constraints on H0 within
the axion-like EDE model under different data sets as indicated in the figure, where solid (dashed) lines represent
Bayesian (frequentist) constraints. The constraints are taken from Refs. [1958, 1960, 2442, 2445, 2446] (see text).

The axion-like EDE model faces some theoretical challenges: the desired degeneracy between fEDE and H0 , and
with that, the ability of EDE to solve the tension sensitively depends on the choice of the parameters zc and θi . Since
there is no a priori reason for these parameters to have specific values, this leads to a “fine-tuning” problem. Moreover,
the shape of the potential with index n = 3 is considered fine-tuned itself since it requires that lower-order terms in
the axion-like potential are zero. Therefore, the axion-like EDE model is often seen as a phenomenological model of
EDE. Nonetheless, there are approaches to derive the axion-like EDE model from fundamental theories such as string
theory [2448–2450] or to motivate n = 3 by higher-order instanton corrections [2451].
4.1.1.c. EDE variants From the axion-like EDE model, we learned valuable lessons about the challenges of EDE
as a solution to the Hubble tension. In order to have the maximal possible leverage to resolve the tension, we need
a specific form of potential, EDE needs to become relevant around matter-radiation equality and it needs to decay
sufficiently fast before recombination. These theoretical challenges, or “fine-tunings”, have inspired many variants of
EDE, which will be presented here.
Several EDE models consist of slow-rolling scalar fields minimally coupled to the metric, with different shapes
of potentials [2435, 2452–2454]. Rock ’n’ Roll EDE assumes a power law potential of the form V (ϕ) ∼ ϕ2n , which
approximates the axion-like potential for small ϕ/f [2455]. Acoustic DE generalizes the EDE model by using a
phenomenological fluid description of EDE parameterized by the sound speed c2s and equation-of-state parameter
wADE [2456].
A class of axion-like EDE models arises in the string-inspired Chern-Simons (CS) gravity [2457], which
characterizes the low energy limit of string theory [2458]. In the simplest of such models [2459–2462], it leads to the
generic form of Running Vacuum Model (RVM) inflation [2463–2466]. In the specific context of string-inspired CS
gravity, it is called Stringy RVM (StRVM) [2461, 2467], in which there is a pre-RVM-inflationary era dominated by a
stiff-fluid of gravitational (Kalb-Ramond) axions that characterize the massless gravitational string multiplet. At the
end of such eras, condensation of primordial chiral GWs leads to a condensation of the gravitational anomaly CS term
that couples the model to string axions that are either of Kalb-Ramond type or arise from compactification [2468].
The condensed CS term leads to linear terms in the axion potential [2469, 2470], responsible for RVM inflation, while
the nonperturbative instanton effects lead to additional periodic modulations. In the context of the StRVM [2459–
2461], the coefficient of the linear axion term is not a constant, but an RVM-type functional of even powers of the
Hubble parameter. This differentiates the model from other linear axion monodromy models that arise from brane
compactification [2471]. In Refs. [2469, 2470] it was shown that the CS condensate leads to a metastable EDE, as
there are imaginary parts in the effective action, pointing to a finite-lifetime of the respective vacua. Notably, different
forms of the StRVM can also help alleviate, in modern eras, the Hubble and growth-of-structure tensions [2472, 2473].
In the framework of α-attractors in inflation [2474–2476], a potential for EDE was proposed in Ref. [2477] that can
reproduce various cases based on its functional form.
Unified models of EDE and late DE have been proposed in Refs. [2478–2480]. Ref. [2478] introduces a model
where a scalar field can explain both EDE and late DE in a joined manner in the context of α-attractors. In Ref. [2479],
157

a unified model of EDE and late DE is introduced, which can also enhance the primordial GW spectrum at PTA
scales. Moreover, the EDE scalar potential can arise from a D-brane moving in the compact dimensions of warped
compactifications in the context of D-brane models of cosmology. Ref. [2480] used a dynamical system approach to
analyze a unified model of EDE and late DE. Scaling EDE is characterized by a constant EDE fraction in both the
radiation- and matter-dominated epochs. This behavior is naturally found in quintessence models with exponential
potentials and dates back to the concept of scalar field DE [296]; see also Ref. [2481, 2482]. The very tight constraints
imposed by the CMB disfavor this type of models as a solution to the tension [2436, 2438, 2483, 2484], even when
allowing for a coupling between EDE and DM [2416]. In the case of tracker solutions, [2485] studied EDE using
quintessence and K-essence in the tracker regimes.
The possibility of EDE field coupling to other particles was also explored. Refs. [2447, 2486, 2487] introduce
a model with a coupling between EDE and DM. The coupling modifies the scalar potential, including a matter
contribution and a modulation of the DM mass. In a similar spirit, the coupled scalar DM to EDE has been studied
in Refs. [2416, 2488, 2489]. The waterfall EDE model was explored in Ref. [2490], in which the EDE phase ends through
spontaneous symmetry breaking. This mechanism has the potential to generate a dark-photon-DM component. In
Ref. [2491], the EDE phase is realized and ended similarly to the warm inflation model. The possibility of EDE
coupling to neutrinos was studied in Refs. [2492, 2493] with the coupling to neutrinos being a natural trigger for the
EDE field at around the matter-radiation equality when the neutrinos become non-relativistic. The possibility of
these models not being able to solve the Hubble tension was raised in Ref. [2494] and later disputed in Ref. [2495].
Emerging DE models consider a non-negligible contribution of the DE density during the radiation domination era.
This effect is introduced through an evolving equation of state that, at early-times, mimics radiation and at late-times
asymptotically reaches ω → −1 [2496–2498]. This phenomenological emergent DE model [2499] shows promising
results when confronted with data [807].
Instead of using a specific EDE model, one can use a model-independent reconstruction of the EDE expansion
history. Ref. [2483] uses a tomographic approach, in which ρEDE is grouped into redshift bins z and constrained
by confronting this model with data. They find that the tightest upper bound on ρEDE occurs around the CMB
decoupling, whereas much weaker constraints are obtained before and after recombination, leaving room for the EDE
model to alleviate the Hubble tension. Ref. [2097] determine ρEDE in scale-factor bins, finding that the maximum
ρEDE is reached at about z ∼ 105 , which is different from the zc derived with the standard approach. Alternatively,
a parametric equation of state can be fit [2500]. Moreover, different microphysics of EDE could have important
effects [2501]. In Ref. [2502], a phenomenological model is proposed that can mimic the EDE phase through multiple
transitions in the vacuum DE density.
4.1.1.d. Constraints on EDE from CMB, BAO and SN In the following sections, we will discuss observational
constraints on the axion-like EDE model from different data sets (see Fig. 70, right, for a whisker plot). When
analyzing EDE with Planck primary and lensing CMB [725], BAO [776, 2503, 2504], SNIa [2023], and the H0
measurement of the SH0ES collaboration [48, 2505], these data sets generally prefer high values of fEDE and high
values of H0 consistent with direct measurements.51 The goodness of fit, i.e., the χ2 , reduces compared to ΛCDM,
albeit with three extra parameters compared to ΛCDM. This is mainly driven by the ability of EDE to accommodate
higher H0 values by reducing rs , which in turn leads to higher values of ns and ωcdm . Ref. [2442] uses the effective
fluid description of the axion-like EDE model, Planck 2015 [725] and SH0ES 2018 [2505] and obtains fEDE ∼ 5% at
zc ∼ 5000 and H0 = 70.6 ± 1.3 km s−1 Mpc−1 (∆χ2 = −14.2, first errorbar in Fig. 70), whereas in Ref. [2443] using
the scalar field description and updating to the SH0ES 2019 value [48] results in fEDE ∼ 10% at zc ∼ 3700 and
H0 = 71.49 ± 1.20 km s−1 Mpc−1 (∆χ2 = −20.3).52 Updating to Planck 2018 data, [1955, 2447] also use the scalar
field description yielding fEDE ∼ 14% at zc ∼ 3800 and H0 = 71.2 ± 1.1 km s−1 Mpc−1 (∆χ2 = −16.2). Finally, those
constraints were recently updated with Planck 2020 NPIPE data [1960] yielding H0 = 71.24 ± 0.77 km s−1 Mpc−1 for
fEDE = 0.107 ± 0.023 at log10 (zc ) = 3.585+0.049
−0.15 (∆χ = −28.0).
2

4.1.1.e. EDE and LSS Confronting the EDE model with LSS data can impose stringent constraints on fEDE .
This can be understood as LSS data typically prefer lower values of the amplitude of clustering σ8 and the related
parameter S8 compared to CMB data (the so-called “S8 discrepancy”). However, fEDE typically correlates positively
with the amount of late-time clustering due to increased ωcdm in models with higher fEDE which compensates for the
EDE-induced boost to the early Integrated Sachs-Wolfe (eISW) effect to maintain a good fit to the CMB [2443, 2506–
2508]. Including LSS data thus weakens the evidence for EDE compared to using CMB data alone by providing an
independent constraint on the clustering amplitude [2445].
Several studies have shown that the combination of BOSS full-shape and BAO data [776, 1520, 2503, 2509–2511]
(based on the effective field theory of LSS, e.g., see Refs. [1173, 1174]) alongside the CMB give tight upper bounds

51 All data combinations and model implementations in this section use the same SNIa sample and BAO data, but differ in H0 and Planck
data.
52 All central intervals in this section are quoted at 68% CL while all upper limits are quoted at 95% CL
158

on fEDE [1169, 2512, 2513] when not including any direct measurements of H0 . WL data have also been explored
in the context of the EDE model. For example, see Ref. [2445] where DES-YI 3x2pt data [1161], as well as S8
priors from KiDS and HSC [854, 857, 858], were used to derive tight upper limits of fEDE < 0.060 and H0 =
−1
68.92+0.57
−0.59 km s Mpc−1 when combined with Planck [192] leading to the conclusion that LSS data can rule out the
EDE model (fifth errorbar in Fig. 70). In contrast, [1955] examine the EDE model in the context of KiDS [858]
and DES [697] data and find that a one-parameter EDE model, which fixes {zc , θi }, does not lead to a significantly
worsened S8 discrepancy compared to ΛCDM and advocate for caution when combining datasets that are statistically
discrepant. Furthermore, [1956] argue that the stringent constraints found when combining CMB data with BOSS
full-shape and BAO likelihoods can be traced to a potential tension in the power spectrum amplitude As between
Planck and BOSS which is present even in ΛCDM. Thus, while it is clear that LSS data can play an important role in
constraining the EDE model, the existence of subtle inconsistencies between CMB and LSS data, present even within
the ΛCDM framework, complicates the interpretation.
It is also worth mentioning the important role that the use of σ8 as a cosmological parameter might play in this
discussion. The σ8 parameter is computed as a convolution of the matter power spectrum and a window function at a
characteristic scale R8 ≡ 8h−1 Mpc. Thus, σ8 provides information about the amplitude of the power spectrum at a
scale that changes in models with different values of H0 . To address this issue, [2427] suggested using the parameter
σ12 , which is computed at a fixed scale R12 = 12 Mpc that does not depend on h. In Ref. [1966] it is shown that despite
the positive correlation between σ8 and fEDE in fitting analyses with CMB+BAO+SNIa, the correlation between σ12
and fEDE is negligible.
Attempts have been made to address both the H0 and S8 tensions simultaneously by expanding the EDE model
parameter space to include components that reduce the amplitude of clustering at late-times. For example, Refs. [1959,
2416] allow for a free neutrino mass in the EDE model finding that current data limits the effectiveness of massive
neutrinos to suppress structure formation at small scales due to the stringent constraint placed on the sum of neutrino
masses from the Planck 2018 data which is not much weakened in the context of EDE. Other works have also
considered a coupling between EDE and DM in order to try to decrease the value of ωcdm and make the growth of
large-scale structures less efficient at late-times [2416, 2447, 2486].
4.1.1.f. Prior effects and frequentist constraints Due to the complicated parameter structure of the EDE model,
the tight upper limits on EDE from CMB and LSS are not only driven by physical effects but also affected by prior
volume (or projection) effects in the MCMC posterior. These prior effects arise due to the nested parameter structure
of the model: while fEDE controls the fraction of EDE, zc and θi are auxiliary parameters describing the model in more
detail. When fEDE approaches zero, the ΛCDM limit is recovered, and zc , θi becomes redundant and unconstrained.
This leads to a larger prior volume at fEDE ≈ 0 than at fEDE > 0 and a non-Gaussian posterior, which in turn
can lead to a preference for the ΛCDM limit in the marginalized posterior. In Refs. [1955, 1956] (and Ref. [1961]
for NEDE) the influence of prior effects was illustrated by fixing zc , θi in the analysis and varying only fEDE along
with the ΛCDM parameters, which leads to larger allowed values of fEDE = 0.0523+0.026 −0.036 when using the same BOSS
full-shape data [719, 1141, 1168] with the effective field theory of LSS [1173, 1174] combined with BAO and CMB
[1429].
Since these effects are related to the prior, which is an inherently Bayesian quantity, Refs. [1957, 1958] constructed
frequentist constraints based on the profile likelihood, which are independent of priors and only depend on the
likelihood (top three dashed errorbars in Fig. 70). Using Planck CMB [1429] and BOSS full-shape power spectrum
data [719, 1141, 1168, 2514] based on the effective field theory of LSS, the profile likelihood yields higher values
of fEDE = 0.087 ± 0.037 and H0 = 70.57 ± 1.36 km s−1 Mpc−1 (∆χ2 = −5.6) than the constraints from Bayesian
posteriors. Even when additionally including 3x2pt data from DES [867], the profile likelihood yields moderately high
values of fEDE ≈ 6 %. This shows that the Bayesian constraints from CMB and LSS are partially driven by prior
effects and indicates that more data is necessary to fully rule out or detect EDE. These results agree with Ref. [1966],
which uses an approximate profile distribution obtained from MCMC samples, which is computationally cheaper than
computing a profile likelihood. In a Bayesian context, [1169] explores the impact of priors on constraints from the
effective field theory of LSS on EDE. Using a Jeffreys prior on nuisance parameters mitigates projection effects from
the effective field theory of LSS, resulting in smaller overall volume projection effects on EDE when combined with
other datasets. Ref. [2515] utilizes a neural network to obtain a data-driven parameterization of EDE.
4.1.1.g. Ground-based CMB experiments and alternative CMB pipelines The Planck experiment offers high-
precision measurements of the large- to intermediate-scale CMB temperature and polarization power spectra. Ground-
based CMB experiments such as ACT and SPT can improve over Planck data, especially at small scales (ℓ ≳ 2000)
and in polarization measurements [731]. Constraining EDE with data from the ACT Data Release 4 (TT, TE,
and EE power spectra) [731] combined with large-scale Planck TT [192, 728, 1429], Planck CMB lensing [693],
−1
and BAO data [776, 1141, 2503] yields fEDE = 0.091+0.020 +1.0
−0.036 and H0 = 70.9−2.0 km s Mpc−1 [709, 2446, 2516]. The
combination of ACT, WMAP [1189], BAO, Pantheon [2023] in a Bayesian analysis leads to fEDE = 0.158+0.015 −0.094 and
159

−1
H0 = 73.43+2.6
−3.4 km s Mpc−1 [2446] (seventh errorbar in Fig. 70). The combination of data sets ACT DR4, SPT-3G
TEEE 2018 [724, 2517], and Planck TT650TEEE [192] gives a similar preference for EDE [2517–2519]. Hence,
replacing Planck data with ACT data leads to a preference of high values of fEDE and H0 without the inclusion
of any H0 prior [2520–2522]. However, the inclusion of SPT temperature and Planck temperature TT at ℓ > 650
strongly decreases this preference, with fEDE < 0.071 in a Bayesian analysis [1954].
This preference for EDE diminishes with recent ACT DR6 data, finding fEDE < 0.12 for ACT combined with Planck,
and even lower for ACT alone [702, 705]. For ACT DR6 alone, the EDE model does not show any improvement in fit
compared to ΛCDM, suggesting that the preference for EDE in the previous ACT DR4 could be due to a statistical
fluctuation in the EE and TE power spectra.
The most commonly used Planck likelihood for multipoles ℓ > 30 is the Plik likelihood [728]. The CAMSPEC [2523]
and Hillipop [700] likelihoods are alternative likelihoods, which differ in data cuts and analysis choices. Ref. [1960] use
the CAMSPEC likelihood with the Planck PR4 NPIPE maps [2524] along with BAO data from BOSS [1141], SDSS [2503]
and 6dFGS [776] and supernova data from Pantheon+ [33] to constrain the EDE model. They find low values of H0 =
−1
68.11+0.47
−0.82 km s Mpc−1 and fEDE < 0.061 in a Bayesian analysis, and similarly H0 = 68.37 ± 0.75 km s−1 Mpc−1 and
fEDE < 0.094 in a frequentist analysis (bottom two errorbars in Fig. 70). Using the Hillipop likelihood, Ref. [2434]
find similarly tight constraints on EDE, as well as Ref. [2525]. Hence, both alternative Planck likelihoods pose a
challenge to the axion-like EDE model’s ability to solve the H0 tension.
4.1.1.h. Complementary EDE observables If EDE existed in the early Universe, it is natural to expect evidence
other than the H0 tension of this deviation from ΛCDM. In this section, we discuss possible such signatures in data
sets different from the commonly used data discussed above. Galaxies and small-scale LSS probes could provide
important insight into EDE models. For example, EDE cosmologies predict more massive clusters and an
increased abundance of galaxy-mass haloes at higher redshifts [1749]. Refs. [1748, 1750, 2526] propose that
EDE cosmologies could accommodate the massive high-redshift JWST galaxies. However, while the fit to JWST data
improves, EDE possibly worsens the fit to Hubble Space Telescope data. Lyman-α forest data from SDSS eBOSS
[1549] and MIKE/HIRAS [2321] give tight upper limits on EDE [1554] due to the preference of lower values of ns
than Planck. However, the Lyman-α forest data shows a discrepancy with Planck data [1555], which warrants further
investigation. Furthermore, an expansion rate differing from ΛCDM would affect the relation between redshift and
age, which can be constrained by the age of ancient objects in the Universe [2527]. Hence, improved age constraints
will be able to constrain EDE cosmologies.
Measurements of the CMB anisotropies along with other CMB observables could hold clues about deviations from
ΛCDM. For example, if EDE would couple to photons in a parity-violating manner, it could explain the tentative
evidence of cosmic birefringence observed in Planck data [688, 692, 2528]. The EDE field with the potential
in Eq. (4.4) predicts a unique shape of the EB power spectra, which is not favoured by Planck data and hence
constrains a possible parity-violating coupling of EDE to photons [690, 2529–2531]. Moreover, spectral distortions
are sensitive to physics in the (pre-) recombination era. Since EDE is dynamical during this epoch, spectral distortion
measurements could be sensitive to distinct EDE signatures [2532].
Theoretical considerations and the requirement of a consistent UV completion can place constraints on EDE models.
For example, Ref. [2533] employed the axion weak gravity conjecture 53 to derive an upper bound on the axion
decay constant. Specifically, for n = 1 in the potential Eq. (4.4), this leads to f < 0.008 MPl , which is not consistent
with the typically assumed parameter values. Potential ways to circumvent or relax this constraint exist [2533].
Moreover, EDE cosmologies typically prefer higher values of ns close to unity, which would point to a scale invariant
Harrison-Zeldovich spectrum [2507, 2519, 2536–2538] and would have important implications for inflation [2539, 2540].
Hence, while EDE models provide a promising class of models to resolve the Hubble tension, theoretical consider-
ations and new data provide challenges to these models.

4.1.2. New early dark energy and variants


Coordinator: Florian Niedermann
Contributors: Martin S. Sloth, and Mathias Garny

NEDE is a theoretical framework to address the Hubble tension where a scalar field ψ undergoes a phase transition
between BBN and recombination on a time scale that is short compared to the Hubble time, such as, e.g., a first-order
phase transition [2541, 2542]. In contrast to other phenomenological approaches to the Hubble tension, NEDE is rooted
in concrete microphysics, which is operative at the eV scale. Current CMB temperature and polarization measurements

53 The axion weak gravity conjecture is part of a broader program known as the “swampland”. This program aims to define the boundary
between effective field theories that are consistent with quantum gravity and those that exist in the “swampland” of incompatible
theories. The precise statement of the axion weak gravity conjecture remains a topic of debate [2534, 2535].
160

are sensitive to the precise mechanism triggering the phase transition, and, therefore, the phenomenology differs
depending on the concrete NEDE realization. The two main examples are Cold and Hot NEDE, where the trigger
of the phase transition is a second scalar field or the temperature of a dark sector, respectively. In the case of
Cold NEDE, the phenomenology shares some features with other EDE-type models, but differs both at background
and perturbation level, particularly after the phase transition. On the other hand, Hot NEDE is more similar to
strongly interacting dark radiation (SIDR) models, and can be viewed as a UV completion of these types of models,
explaining naturally the creation of extra radiation after BBN in a supercooled phase transition. In a nutshell, the
idea is summarized in Fig. 71, where the left panel explains the field-theoretic idea and the right panel depicts the
cosmological model. Initially, the NEDE field ψ is stuck in a false minimum of its potential. It is separated from
the true minimum by a large potential barrier. The (false) vacuum energy associated with ψ constitutes an early DE
component (orange line). However, as the trigger mechanism kicks in, the barrier shrinks, initiating a strong first-
order phase transition. This corresponds to the nucleation of bubbles that separate the true from the false vacuum.
Eventually, all of space is converted to the true vacuum. In the case of a fast phase transition, this can be described
in a fluid model in terms of an abruptly decaying early DE component (red line). It leaves behind a residual vacuum
energy, which can be identified with DE (green line). As a closely related idea, we mention ChainEDE, which relies
on a series of a large number of first-order phase transitions to address the H0 tension [2543].

Cold NEDE

Hot NEDE

trigger

early dark energy


V(ψ)

phase transition
false vacuum

true vacuum

phase tr de
ansition ca
yin
g
flu
id

dark energy
10 -6
10 -4 0.01 1
ψ
a
(a) Microscopic picture (b) Cosmological fluid model

FIG. 71: Schematic representation of the NEDE framework, explaining its microphysics and phenomenology. Left:
The NEDE field decays in a triggered first-order phase transition, giving rise to an energy injection before recombi-
nation. EDE is identified with the field’s vacuum energy before the transition, and DE with the true vacuum after
the transition. Right: On cosmological length scales, a fluid description can be applied. Cold and Hot NEDE energy
densities are represented as the solid and dashed lines, respectively.
4.1.2.a. Cold New Early Dark Energy Cold NEDE introduces an additional, ultralight scalar field ϕ to trigger
the phase transition [1961, 2542]. Its effective potential valid at low energies is
λ 4 β 2 2 1 1 1
V (ψ, ϕ) = ψ + M ψ − αM ψ 3 + m2 ϕ2 + λ̃ϕ2 ψ 2 . . . , (4.5)
4 2 3 2 2
where α, β, λ, and λ̃ are dimensionless EFT parameters. The model introduces a hierarchy of scales, where the trigger
field ϕ is an ultralight field with mass m ∼ 10−27 eV and the tunneling field ψ is much heavier with mass M ∼ eV.
A possible high-energy completion has been argued to be possible within a multi-axion framework, where the small
masses are protected through an approximate shift symmetry [2544]. In terms of its cosmological fluid description,
Cold NEDE introduces four additional parameters: the fluid’s maximal energy fraction, fNEDE , the redshift z∗ of the
phase transition, the equation of state parameter of the post-phase transition fluid, wNEDE , and the relic abundance
of the trigger field today Ωϕ .
The phenomenological model has been implemented in the Boltzmann code TriggerCLASS54 and tested extensively
against cosmological data in the literature. Most work has been done in the limit where Ωϕ ≪ 1, corresponding to a
subdominant trigger field. For example, using Planck 2018 [192], Pantheon [2023], and BAO [776, 1141, 2503] data,

54 https://github.com/NEDE-Cosmo/TriggerCLASS.git
161

0.8
BBN
bound ΔN IR
SIDR
0.6
ΔN NEDE req. for H 0
WZDR

ΔN eff
0.4
Hot NEDE
SU (3)→SU (2)
0.2

ΔN BBN a* (PT) at Td =m ψ 


0.0 - 10
10 10 - 9 10 - 8 10 - 7 10 - 6 10 - 5 10 - 4 10 - 3 10 - 2
a

FIG. 72: Evolution of ∆Neff in Hot NEDE with a supercooled Coleman-Weinberg phase transition (PT) between
the BBN and recombination epochs, leading to a strong increase (∆NBBN → ∆NNEDE ) due to the latent heat be-
ing converted into relativistic SU (2) dark gauge and Higgs bosons during the PT, and a second slight increase due
to the dark Higgs becoming non-relativistic (∆NNEDE → ∆NIR ). This evolution is intimately linked to the under-
lying Coleman-Weinberg model. Also shown are bounds from BBN, and the values required to solve the H0 ten-
sion around recombination, as well as two models featuring extra dark radiation but no phase transition (SIDR and
WZDR [2549]).

the H0 tension was found to be reduced to 2σ when employing the QDMAP tension measure [709] (with a value for H0
taken from Ref. [78]). This result was supported later in Ref. [1962] in a profile likelihood approach, which gets around
prior volume issues, that are otherwise are known to drive the posteriors for fNEDE towards small values [1961] (see
also the discussion in Sec. 4.1.1 for more details). Without including a prior on H0 , it was found that H0 = 69.56+1.16
−1.29
km s−1 Mpc−1 and fNEDE = 0.076+0.040 −0.035 at 68% CL, amounting to a 2σ indication for a non-vanishing fraction of
−1
NEDE. Including the SH0ES prior, H0 = 71.62+0.78 −0.76 km s Mpc−1 and fNEDE = 0.136+0.024−0.026 was obtained at 68%
CL, which amounts to a 5σ evidence for a non-vanishing fraction of NEDE. The model has also been studied in the
presence of additional LSS [1962, 2545] and ground-based CMB data [2446, 2546, 2547], broadly matching the baseline
results cited above, although it should be noted that axiEDE and Cold NEDE respond differently to ACT and SPT
data [2446]. Recently the assumption Ωϕ ≪ 1 was dropped in Ref. [2544]. In that case, the trigger field can make
a sizable contribution to the energy budget today in the form of a fuzzy DM component (for a phenomenologically
similar model see also Ref. [2548]). With their non-vanishing pressure fluctuations, the trigger field’s perturbations
act against gravitational collapse on small scales as required for solving also the S8 tension. As a result, the model
was found to simultaneously reduce both tensions55 below 2σ (with the above choice of datasets S8 = 0.818+0.023−0.017 was
obtained).
4.1.2.b. Hot New Early Dark Energy In Hot NEDE [2550, 2551], the phase transition is triggered by the de-
creasing dark sector temperature Td , and the sound horizon is lowered due to the dark radiation generated from the
latent heat, i.e., the false vacuum energy released into the dark sector during the transition [2552] (for a comparison
to Cold NEDE see Table V). If the phase transition occurs between BBN and recombination, this setup naturally
reconciles SIDR-type solutions to the Hubble tension with BBN bounds on extra radiation [2552] (see Fig. 72). It fur-
thermore differs from SIDR at the perturbation level, and makes complementary predictions. In particular, the model
predicts oscillatory features on small scales in the matter power spectrum alongside a Stochastic Gravitational Wave
Background (SGWB) signal that can be looked for in Lyman-α forest data and pulsar timing arrays, respectively.
Hot NEDE can be realized by the well-known Coleman-Weinberg mechanism of spontaneous symmetry breaking
within a dark sector featuring a dark Higgs field ψ that transforms under a dark gauge symmetry SU (N ). This setup

55 This is another important difference at the phenomenological level between axiEDE (and other EDE type models) and NEDE. While
axiEDE is known to worsen the S8 tension, NEDE has the potential to resolve both the H0 tension and the S8 tension.
162

TABLE V: Comparison of Cold and Hot NEDE, both relying on a fast-triggered phase transition before recombi-
nation to address the Hubble tension.

NEDE model Cold NEDE Hot NEDE


model family EDE (stepped) strongly interacting dark radiation
physics lowering rs false vacuum energy latent heat released as SIDR
trigger ultralight field dark sector temperature
redshift of phase transition 103 ≲ z∗ ≲ 104 105 ≲ z∗ ≲ 109
equation of state parameter wNEDE > 1/3 (∼ 2/3) wNEDE = 1/3
microphysics axion-like particle dark sector Higgs

naturally leads to the required supercooled first-order phase transition, breaking the initial SU (N ) to SU (N − 1), as
well as to self-interactions described by the non-Abelian gauge symmetry. The dynamics is captured by the effective
potential [2552]

ψ2 µ2 ψ2
   
1
V (ψ; Td ) = Bψ 4 ln 2 − − eff ψ 2 1 − 2 + ∆Vthermal (ψ; Td ) , (4.6)
v 2 2 2v

where the dimensionless parameter B ∼ g 4 depends on the gauge coupling g. During the NEDE phase transition,
ψ picks up a super-eV expectation value v. Due to an approximate conformal symmetry (controlled by the soft
breaking parameter µeff ∼ eV), the phase transition exhibits a significant amount of supercooling, where the latent
heat dominates over the dark radiation plasma, giving rise to an EDE component. In a more fundamental picture,
ψ is the modulus of a dark sector Higgs multiplet, breaking an initial SU (N ) gauge symmetry to SU (N − 1). The
corresponding massless gauge bosons alongside the light Higgs degrees of freedom are then populated during the phase
transition, reheating the dark sector and introducing a sizable increase in the effective number of relativistic degrees
of freedom, ∆Neff . Although both Cold and Hot NEDE feature an EDE component, their phenomenology differs
significantly. In particular, the Hot NEDE phase transition occurs before the CMB epoch at redshifts z∗ ≳ 105 (see
the right panel in Fig. 71). As a result, Hot NEDE lowers the sound horizon rs (and thus increases H0 as explained
in Sec. 4.1.1) through the energy injection provided by the strongly interacting dark sector plasma after the phase
transition and, on a phenomenological level, is thus more akin to SIDR models (with mass threshold effects) such as
Ref. [2549]. However, its main phenomenological advantage is that it gets around BBN constraints on ∆Neff , which
otherwise would rule out SIDR models as solutions to the Hubble tension [2552, 2553], see Fig. 72 (for other ideas
to avoid BBN bounds see Refs. [2554–2562]). A first test against cosmological data [2552] showed that the model
−1
reduces the QDMAP tension to 2.8σ. Moreover, it was found that H0 = 96.13+0.62
−1.00 km s Mpc−1 and H0 = 71.17±0.83
km s−1 Mpc−1 without and with including a Gaussian prior on H0 , respectively (datasets as for cold NEDE except for
the prior on H0 taken from Ref. [34]).

4.1.3. Extra relativistic degrees of freedom


Coordinator: Sunny Vagnozzi
Contributors: Adèle Poudou, Alexander Bonilla Rivera, Biswajit Karmakar, Branko Dragovic, Davide Pedrotti,
Diego Rubiera-Garcia, László Árpád Gergely, Leila L. Graef, Luca Visinelli, Luis Anchordoqui, Marcin Postolak,
Margus Saal, Mariana Melo, Matteo Forconi, Özgür Akarsu, Salvatore Capozziello, Sebastian Bahamonde, Shouvik
Roy Choudhury, Simony Santos da Costa, Stefano Gariazzo, Thejs Brinckmann, Utkarsh Kumar, Vivian Poulin,
William Giarè, and Wojciech Hellwing

Introducing extra relativistic species (ERS, also referred to as “dark radiation”) is one of the most well-motivated
and straightforward extensions of the ΛCDM model, but also one of the simplest ways of raising the inferred value of
H0 [2563]. The ESR energy density is most easily parametrized via the effective number of neutrino species Neff [2564],
controlling the relation between the total radiation energy density ρr and the photon energy density ργ [2565, 2566]
"  4 #
7 4 3
ρr = ργ 1 + Neff . (4.7)
8 11

In the presence of only three standard neutrinos undergoing instantaneous decoupling, Neff = 3. However, in the
standard cosmological model, Neff
SM
is slightly larger than 3, owing to the fact that the neutrino decoupling process is
163

not instantaneous. The latest determinations yield Neff SM


≃ 3.044 [2567–2570]. After electron-positron annihilation,
the only relativistic particles in the standard cosmological model are neutrinos and photons. However, several well-
motivated extensions of the Standard Model of particle physics predict the existence of additional particles which
could be relativistic at epochs of cosmological interest, including but not limited to light sterile neutrinos [2571–2575],
Goldstone bosons [2576, 2577], axions, axion-like particles, and ultra-light DM [2441, 2578–2588], light or massless
dark photons [2589–2597], and so on [2598, 2599]): such particles lead to a value of ∆Neff = Neff − Neff SM
̸= 0. It is
worth noting that a non-zero value of ∆Neff does not necessarily imply the existence of new particles. For instance, it
could be associated with a non-standard distribution of standard neutrinos, or the presence of neutrino asymmetries
(i.e., a non-zero chemical potential) [2600, 2601]. In addition, a SGWB, possibly at the origin of the signal detected
in pulsar timing array data [2602–2605], could also provide a significant contribution to Neff [2606–2616]. Moreover,
∆Neff can in principle be negative, for instance, in low-reheating scenarios following inflation, where the reheating
temperature Trh can be as low as a few MeV, leading to incomplete thermalization of standard neutrinos [2617–2620].
A non-zero value of ∆Neff leads to a host of signatures at both the times of BBN and recombination. ERS increases
the physical radiation density and thereby raises the pre-recombination expansion rate H(z) before recombination
(as well as after, although by then the radiation component is sub-dominant). At the BBN epoch, this pushes the
freeze-out of nuclear interactions towards higher temperatures, raising the yield of light elements. At recombination,
the presence of ERS leads to a variety of direct and indirect signatures on the CMB: however, once the redshift
of matter-radiation equality zeq and the acoustic angular scales θs are fixed, the only remaining effects of ∆Neff on
the CMB power spectra consist of an increased (Silk) damping of the higher acoustic peaks and a phase-shift of the
acoustic peaks towards larger scales (e.g., see Refs. [2621–2623]). All these effects indeed lead to some of the tightest
cosmological constraints on ∆Neff : current BBN and CMB inferences both broadly indicate ∆Neff ≲ 0.2 − 0.3,
while next-generation cosmological surveys will be able to improve the sensitivity of ∆Neff by almost an order of
magnitude [680, 681].

ΛCDM SH0ES (R21)


Neff = 3.15
Neff = 3.35
Neff = 3.55
Neff = 3.75
Neff = 3.95

66 68 70 72 74

H0 [km/s/Mpc]
FIG. 73: Posterior distributions for H0 inferred from a combination of Planck 2018, SDSS BAO, and PantheonPlus
SNIa data for different fixed values of Neff . The green shaded region corresponds to the R21 measurement H0 =
73.04 ± 1.04 km s−1 Mpc−1 as inferred by the SH0ES team in Ref. [34].

Increasing Neff while keeping the physical baryon and DM densities ωb and ωc fixed has the effect of increasing
the pre-recombination expansion rate, thus decreasing the comoving sound horizon at recombination rs , which then
requires a higher value of H0 in order to keep θs fixed. This is the reason why ERS are one of the most economical
ways of raising H0 and are generally considered a benchmark model in this sense. These effects lead to a positive
correlation between Neff and H0 [711, 2624], with an increase in the former leading to an increase in the latter, as
164

captured by the following linear relation [2625], as can be seen in Fig. 73


∆H0 = H0 |∆Neff ̸=0 − H0 |ΛCDM ≃ 5.9∆Neff , (4.8)
as calibrated to Planck 2018 CMB, BAO, and Pantheon SNIa measurements in 2020. This relation, first obtained
in Ref. [2625], was illustrated in Ref.[2508] in the H0 -σ8 plane, together with the analogous relation for several
tension-resolving candidate models.
From Eq. (4.8) we see that increasing H0 to a level that solves the Hubble tension requires Neff ≳ 4, corresponding
to the effect of an extra fully thermalized neutrino which, importantly, is completely excluded by BBN and CMB
considerations. The main reason why the CMB excludes high values of Neff is that they disproportionately alter
the damping scale rd , in turn altering the ratio rs /rd , which is tightly constrained especially by high-multipole
polarization measurements to be close to its ΛCDM value (see e.g., the discussion in Ref. [712]). Moreover, the
neutrino drag/phase shift effect [2626, 2627] is also altered to an extent that is incompatible with the data. This
is why the simplest, 7-parameter ΛCDM+Neff model fails to solve the Hubble tension, despite constituting a useful
benchmark [708, 2628, 2629]. Within the simplest 7-parameter ΛCDM+Neff model one infers Neff = 2.89 ± 0.19,
H0 = 66.3 ± 1.4 km s−1 Mpc−1 , and S8 = 0.831 ± 0.013 from Planck 2018 temperature, polarization, and lensing
data, whereas further adding SDSS BAO and Pantheon SNIa data change these numbers to Neff = 3.00 ± 0.17,
H0 = 67.4 ± 1.1 km s−1 Mpc−1 , and S8 = 0.823 ± 0.011 respectively (in what follows, the switch from Pantheon to
PantheonPlus SNIa data is not expected to bring about significant changes).
The cosmological evolution of ERS characterized by Eq. (4.7) is de facto equivalent to that of massless neutrinos,
and for this reason we can refer to the ΛCDM+Neff model as describing free-streaming ERS. It is worth noting that
the above considerations hold for some of the simplest dark radiation models, including light sterile neutrinos [2630],
models with lepton asymmetries/non-zero neutrino chemical potential [2631, 2632], thermal axions [2581], and so
on. Relaxing the assumption that ERS are free-streaming can potentially help address the above shortcomings.
The simplest possibility for such a self-interacting dark radiation is one where the ERS component is strongly self-
coupled and therefore constitutes a perfect fluid [2633], whose Boltzmann hierarchy can be truncated to the first
moment (with vanishing anisotropic stress). Nevertheless, when confronted with cosmological data, such a scenario
fails to solve the Hubble tension [709], as it is unable to reduce the neutrino drag and Silk damping effects to an
extent which is allowed by the data while raising H0 sufficiently. When confronted with Planck 2018 and SDSS
BAO data, one finds H0 = 68.67 ± 0.84 km s−1 Mpc−1 . Other natural extensions envisage the possibility of both
free-streaming and self-interacting ERS, for instance within the “dark sector equilibration” model of Ref. [2634], and
the “recoupling” or “instantaneous decoupling” models (such as in atomic DM [2635–2637] or twin Higgs [2638]) of
Ref. [2313], which, however, can only partially alleviate the tension, leading to figures very similar to the previous
ones. Yet another possibility to undo the unwanted effects of free-streaming ERS and improve the tension-solving
ability of self-interacting dark radiation features scattering between the latter and the DM component: however, this
model can only partially alleviate the Hubble tension and is disfavored from a model-comparison perspective due to
its higher statistical complexity [709]. In this case, one finds H0 = 68.55 ± 0.92 km s−1 Mpc−1 from a fit to Planck
2018 and SDSS BAO data [709].
A class of models that has received considerable interest in this context features non-standard interactions that
do not involve the dark radiation components, but the otherwise standard neutrino species. Non-standard neutrino
interactions, for instance via a four-point contact interaction mediated by a sufficiently massive (m ≳ 1 keV) mediator,
as studied in [1314, 2313, 2639–2651] can suppress or delay neutrino free-streaming, and, with it, the free-streaming-
induced phase shift. for fixed values of the sound horizon rs (which is mostly unaffected by these interactions), the
position of the first acoustic peak would therefore be at higher multipoles. Reversing it to the observed position requires
a higher value of H0 . It has been shown that a combination of self-interacting neutrinos, additional dark radiation
components (Neff ∼ 4), and large neutrino masses (Mν ∼ 1 eV) can offset the unwanted effects of dark radiation
on the damping tail of the CMB, and allow for large values of H0 while not spoiling the fit to CMB temperature
data [1313]. A fit to the latter and BAO data reveals a bi-modal distribution for the interaction strength, with stronger
values of the interaction strength (∼ 109 times that of the weak interaction strength) being the tension-solving ones.
Nevertheless, this solution is strongly challenged from both the data and theoretical perspectives. From the data side,
small-scale polarization data strongly disfavor these models with strong self-interaction strength [2645, 2646]), which
are also partially challenged by BAO data (due to the fact that rs is unchanged) [1314]. In particular, when confronted
−1
against Planck 2018, SDSS BAO, and PantheonPlus SNIa data, one infers H0 = 67.3+2.2 −2.1 km s Mpc−1 and H0 =
+2.2 −1 −1
66.7−2.1 km s Mpc for the moderately interacting and strongly interacting modes respectively [2645, 2646]. Finally,
from the model-building side, a model in which all three neutrinos self-interact equally is excluded by laboratory
constraints [2652], while constructing a phenomenologically viable model has proven challenging [2651]. EFT analyses
of BOSS data reveal support for a strongly interacting neutrino mode [2649], although an analysis combined with
CMB data still disfavors the strongly self-interacting neutrino model where all 3 neutrinos are interacting [1314].
Interestingly, parameter degeneracies allow for a lower value of the scalar spectral index in the interacting neutrino
165

model, which can be leveraged to allow for inflationary models that are otherwise ruled out within ΛCDM, such as
natural inflation and Coleman-Weinberg inflation [2643, 2647, 2651]. It is worth noticing that large values of Neff are
in tension with BBN, unless this extra dark radiation is generated after the BBN epoch. This is the scenario studied
in Ref. [2653], where a fraction of the DM abundance results from the decay of an unstable and initially thermally
decoupled heavy particle, which decays into a DM particle (initially relativistic and therefore contributing to Neff ,
before becoming cold once the Universe has expanded sufficiently) and a Standard Model particle such as neutrinos
or photons. In this case one finds in the best case H0 = 69.08 ± 0.71 km s−1 Mpc−1 and S8 = 0.850 ± 0.008 from a fit
to Planck 2018, SDSS BAO, and Pantheon SNIa data.
Although the simplest self-interacting neutrino models appear to be ruled out, the framework remains a very
interesting benchmark one and gives a good idea of possible ingredients one might want to consider in order to
make non-minimal ERS models viable. Motivated by these considerations, various works have explored models of
eV-scale Majorons, pseudo-Goldstone bosons associated to the spontaneous breaking of a global U (1) lepton number
symmetry. The Majoron enjoys weak couplings to neutrinos, although in the theoretically best motivated coupling
limit, it cannot induce the four-point contact interactions mentioned earlier. The resulting phenomenology is instead
dominated by inverse neutrino decays and Majoron decays. Overall, these effects still damp neutrino free-streaming,
although with a different time dependence, which leads to a better performance of the model when confronted
against cosmological data. It should be noted that Majoron decays also lead to a value of ∆Neff ∼ 0.11 [2654].
Nevertheless, although relatively successful, Majoron models fall short of completely solving the Hubble tension [2655],
+2.21
with H0 = 70.06−2.31 km s−1 Mpc−1 from a fit to Planck 2018, SDSS BAO, and Pantheon SNIa data.
The above are only examples of some of the ERS models considered in the literature, and it is beyond our scope
to give a full list of examples. Other interesting possibilities which have been considered in the literature involve self-
interacting sterile neutrinos [2656–2659], or the so-called stepped fluids [2549, 2660, 2661]. The latter are a particular
class of self-interacting ERS models, consisting of a mix of massless and massive particles: when the temperature of
the dark sector drops below the mass of a massive species, the latter deposits entropy in the lighter ones, leading to
a “step” in the ERS energy density. Concrete Lagrangian realizations of these classes of models have been studied,
for instance, within the context of “Wess-Zumino dark radiation”, and have been found to be relatively promising in
alleviating the Hubble tension, although current data sets do not show a strong preference for the model [2553]. From a
−1
fit to Planck 2018, SDSS BAO, and Pantheon SNIa data, one finds H0 = 69.3+0.9 −1.3 km s Mpc−1 and S8 = 0.829±0.011
for the Wess-Zumino dark radiation model, whereas for a more general stepped dark radiation model these change to
−1
H0 = 68.8+1.3
−1.5 km s Mpc−1 and S8 = 0.823 ± 0.013 respectively [2549].
Overall, despite the fact that there is no consensus ERS model that solves the Hubble tension, many of these
models possess interesting features that could guide the community in the right direction. It would therefore not be
surprising if ERS were to play at least some role in the Hubble tension. These and related aspects are being actively
investigated.

4.2. Late-time proposals


4.2.1. Late dark energy
Coordinator: Rafael C. Nunes
Contributors: Adrià Gómez-Valent, Anil Kumar Yadav, Anjan Ananda Sen, Anne Christine Davis, Arianna Favale,
Benjamin L’Huillier, Branko Dragovic, Brooks Thomas, Carlos G. Boiza, Carsten van de Bruck, Celia Escamilla-
Rivera, Cláudio Gomes, Cristian Moreno, Daniele Oriti, David Benisty, David Tamayo, Davide Pedrotti, Elsa Teixeira,
Emmanuel N. Saridakis, Gaetano Lambiase, Giulia De Somma, Hsu-Wen Chiang, Hussain Gohar, Ilim Cimdiker,
Jaume Haro, Javier Rubio, Joan Solà Peracaula, Juan García-Bellido, Jurgen Mifsud, Kathleen Sammut, Keith R. Di-
enes, Laur Järv, Laura Mersin, Leandros Perivolaropoulos, Leila L. Graef, Lilia Anguelova, Luis A. Escamilla, Luz
Ángela García, Mahdi Najafi, Margus Saal, Mariam Bouhmadi-López, Masoume Reyhani, Mina Ghodsi Yengejeh,
Nikolaos E. Mavromatos, Oem Trivedi, Özgür Akarsu, Paloma Morilla, Paolo Salucci, Purba Mukherjee, Rahul Shah,
Ronaldo C. Batista, Ruth Lazkoz, Salvatore Capozziello, Sanjay Mandal, Sebastian Bahamonde, Simony Santos da
Costa, Suresh Kumar, Utkarsh Kumar, Víctor H. Cárdenas, Vasiliki A. Mitsou, Vincenzo Salzano, Vivian Poulin,
and Wojciech Hellwing

4.2.1.a. Current status of simplified DE model parameterizations circa 2024 The key property of DE is its EoS,
defined as w ≡ Pρxx . Extensions of the ΛCDM model, where w can either be a constant or a dynamical function of cosmic
time, represent the simplest parametric approaches for testing deviations from the ΛCDM framework. Fig. 74 presents
the current state-of-the-art constraints on the EoS, w, at the 68% CL, inferred from various dataset combinations
under the assumption of the wCDM model [2662]. The figure is adapted from Ref. [1293]. As widely discussed in the
literature, models with w < −1, when constrained using CMB data alone, tend to predict higher values of H0 , due
166

CMB
CMB+BAO
CMB+SN
CMB+CC
CMB+BAO+SN
CMB+BAO+SN+CC
CMB+BAO+SN+SH0ES
CMB+FS
BAO+SN
BAO+SN+CC
BAO+CC

−2.0 −1.8 −1.6 −1.4 −1.2 −1.0 −0.8


w

FIG. 74: The whisker plot summarizes the 68% CL intervals for the DE EoS parameter, w, inferred from various
dataset combinations under the assumption of the wCDM model. Results that incorporate CMB data are repre-
sented by blue bars, except for the final consensus dataset combination (CMB+BAO+SN+CC), which is shown in
green. In contrast, results obtained without the CMB data are depicted in red. The green band highlights the 68%
CL interval derived from the final consensus dataset combination, yielding w = −1.013+0.038
−0.043 . The grey vertical
dashed line marks the cosmological constant value, w = −1. Figure taken from Ref. [1293].

to a strong degeneracy between w and H0 [2663]. To achieve a more robust analysis, a joint dataset combination of
−1
CMB+BAO+SN+CC yields H0 = 68.6+1.7 −1.5 km s Mpc−1 , which is in good agreement with the ΛCDM.
Robust and up-to-date analyses of simple extensions to the well-known dynamic w(z)CDM models are presented
in Refs. [1278, 2664]. The results include the most recent BAO measurements obtained by DESI. The preference
for the w(z)CDM model remains robust, regardless of the parameterization used, but none of the parameterizations
are able to resolve the H0 tension. Similar or identical simple parametric models within the w(z)CDM class have
been discussed in the literature. In general, while certain datasets may favor a dynamic parametrization of DE, these
models are not capable of resolving the tension in H0 .
4.2.1.b. Challenges for late dark energy models from BAO and SNIa Data on Deformations of H(z) and the
Ultra-late Physics Transition Approach LDE models aim to resolve the Hubble tension by introducing degrees of
freedom that deform the Planck18 ΛCDM form of E(z) ≡ H(z)/H0 , reconciling the early sound horizon scale with
late-time Cepheid calibrators. However, these models face significant challenges in simultaneously fitting the SH0ES
H0 value, the sound horizon scale at recombination, and distances from SNIa and anisotropic BAO at z ∈ [0.01, 2.5].
Many models also exacerbate the growth tension [3, 768, 1293, 1358, 2454, 2520–2522, 2665–2679].
Recent analyses also propose a rapid transition in the effective gravitational constant, Geff , at ultra-late times
(zt ≲ 0.01) to address the tension [72, 386, 2668, 2670, 2680–2684]. This is based on the mismatch in absolute
magnitude (MB ) between CMB and local Universe observations. The luminosity distance-redshift relation
 
dL (z)
µ(z) = mB (z) − MB = 5 log10 ,
10 pc
suggests SNIa luminosities at z > zt are consistent with constant Geff [1358]. A sudden transition in Geff at zt could
explain the ∆MB ≈ −0.2 mag deviation in local SN needed to reconcile high- and low-redshift H0 values. This model
addresses the H0 tension and suggests reduced growth of density perturbations without affecting ΛCDM background
expansion [386, 1358, 2680, 2681]. Future GW, astrophysical, and cosmological perturbation analyses may test this
hypothesis in the context of physical models [2681, 2682, 2685]. The ultra-late Geff transition provides a compelling
alternative framework warranting further investigation.
On the other hand, when the inverse distance ladder is built with angular (or 2D) BAO, instead of anisotropic
(or 3D) BAO, the phenomenology required to solve the Hubble tension is completely different. Angular BAO data
are claimed to be less subject to model dependencies, since in contrast to anisotropic BAO analyses, no fiducial
167

model is used to convert redshifts and angles into distances of a tracer map. Despite being obtained from the same
parent catalogs, 2D and 3D BAO data are known to be in tension [763–765, 767, 768, 2179], pointing to the existence
of unaccounted-for systematic errors in one or both data sets or an underestimation of their uncertainties. When
angular BAO is employed in the fitting analysis, it is possible to find a late-time solution to both the Hubble and
growth tensions without violating the constancy of MB [768]. The latter requires the effective DE density to become
negative at z ≳ 1.5 − 2 to compensate for the late-time increase of H(z) and not spoil the description of the CMB.
This possibility has been recently realized in the context of a model with a sign-switching cosmological constant
[800, 801, 810, 811, 2686–2688] and even more effectively in the model presented in Ref. [2473].
4.2.1.c. Holographic dark energy and gravity-thermodynamics correspondence models The holographic principle,
rooted in quantum gravity, proposes that a system’s entropy is determined by its surface area rather than its volume.
Holographic dark energy (HDE) utilizes this principle to describe late-time DE as an infrared cutoff for specific HDE
models. Various alternative HDE formulations exist, including the Tsallis and Barrow models, each with distinct
energy density equations and parameters [2689–2707]. Moreover, choices for the infrared cutoff can vary, ranging from
simple options such as the Hubble horizon to more intricate ones like the Granda-Oliveros or Nojiri-Odintsov cutoffs.
Beyond theoretical motivations, HDE models potentially resolve tensions in H0 measurements. Fitting these models
to comprehensive datasets yields values consistent with recent observations, possibly enhanced by incorporating sterile
neutrinos [2708].
On the other hand, modified cosmology based on gravity-thermodynamics correspondence [2709–2712] is another
way to explore the Universe’s evolution. The relationship between black hole quantities and conventional thermo-
dynamics [2713, 2714] has been extended to cosmic horizons [2715], with a range of applications [2709, 2711, 2712].
In particular, these concepts have been used for entropic force cosmological models [2716–2721], in which entropic
force terms account for the Universe’s accelerated expansion. Furthermore, modifications to holographic dark energy
models have been made using different notions of entropy [2722–2724] coming from statistical mechanics and ther-
modynamics.Indeed, the fundamental limit established by the Bekenstein entropy bound [2725] has had important
consequences for understanding fundamental aspects of geometry and possible ways of reducing the Hubble tension
[2726] consistent with data from Refs. [2727, 2728].
The H0 and σ8 tension can be concurrently relieved by the modified cosmology employing Tsallis-Cirto nonextensive
entropy [2722], as demonstrated in Ref. [2729]. The authors demonstrate how a phantom effective DE, which is
recognized as one of the adequate processes that might relieve H0 tension, can be obtained for specific Tsallis-Cirto
parameter choices. Furthermore, given the same parameter choice, they find an enhanced friction term and an
effective Newton’s constant smaller than the conventional one, therefore solving the σ8 tension. In Ref. [2730], the
σ8 tension is examined using Jacobson’s method [2709] in the nonextensive setting. Additionally, Hubble tension
can be relieved in entropic force cosmological models [2717, 2731] because the interaction terms naturally arise from
reversible and irreversible processes across the horizons in the evolution equations. Nevertheless, further work needs
to be done in developing these approaches. Moreover, a wide range of holographic dark energy models have been
utilized to examine these cosmological tensions, for example, Tsallis holographic dark energy [2732]. Addressing
these cosmological challenges will require further research on gravity-thermodynamic relationship to cosmological
applications.
4.2.1.d. Modeling late dark energy: Examining the phantom, tracker, quintom, multifield descriptions and their
clustering properties Phantom-scalar-field DE cosmologies are usually denoted by an EoS that satisfies w < −1
[2054, 2733]. However, the current observations have shown a good agreement with the standard ΛCDM model. Within
these observational analyses, some suggestions invoke the presence of a phantom divide boundary. Furthermore, these
phantom scalar field DE models have introduced scalar fields ϕ to study the dynamics of evolving DE through this
phantom divided boundary, however, within this scheme, we experiment with issues like fine-tuning [2734]. To find
a viable solution, tracker DE scenarios have been analyzed [295, 2735], where ϕ controls the energy density and it
is possible to obtain attractor background solutions. Some parametrizations in this regard [2736–2738] are used to
verify if this model can solve (or relax) the H0 tension. To introduce a phantom scalar field term, we consider a
gravitational action as
ˆ
1 √
S= 2 d4 x −g (R + Lϕ (ϕ, ∂µ ϕ) + Lm ) , (4.9)

where Lϕ is the Lagrangian for a phantom scalar field. Considering a flat FLRW metric, we can vary Eq. (4.9) with
respect to the metric and ϕ to obtain the gravitational field equations and the Klein-Gordon equation. In this scheme,
it is standard to consider a new set of hyperbolic polar coordinates which can ease the numerical solutions [2737].
From the perspective of this ansatz, it was reported that this kind of phantom/tracker model does not address
−1
the Hubble tension when considering a compressed Planck likelihood, given a H0 = 69.1+0.5 −0.6 km s Mpc−1 [2737]
value, which is in a 3σ CL tension with the latest local result [34]. However, Ref. [2739] explored the possibility of
168

reconstructing the constraints using the full Planck [192] likelihood in addition to model-independent CMB baselines,
e.g., ACTPol DR4 [701], the SPT-3G [724] and the WMAP9 [2740] datasets, showing that phantom tracker scalar
field cosmologies can reduce the statistical tension on H0 to below 3σ.
However, quintom DE, featuring two scalar fields with distinct kinetic energies, blends quintessence and phantom
characteristics, and has been shown to be a framework of interest. Model-independent techniques reveal a fluctuating
EoS, crossing the phantom divide line (wDE = −1) multiple times. Quintom models offer a promising alternative to
standard quintessence and phantom models, potentially providing a more accurate fit with fewer parameters. Several
studies explore quintom’s potential in resolving the H0 tension, with analyses coupling the phantom component to
DM and examining various quintessence, phantom, and quintom models. These studies yield various estimates of the
Hubble constant, suggesting the possibility of reducing the tension by up to 2.6 σ [1258, 2056, 2057, 2059, 2081, 2092,
2095, 2741–2751].
Multifield models of cosmic acceleration arise from the coupling of several scalar fields to gravity. The standard
action that describes this system is
ˆ
1 √ 
d4 x −g R − GIJ ({ϕL })∂µ ϕI ∂ µ ϕJ − V ({ϕL }) , (4.10)

S= 2

where gµν is the four-dimensional space-time metric and GIJ is the metric on the manifold parameterized by the scalars
ϕI with I = 1, 2, ..., n . When the field-space metric GIJ has curvature, such multifield models can lead to novel effects,
compared to single-scalar field models, in the context of both cosmological inflation and LDE. Specifically, for LDE
models of this type, see Refs. [2752–2757].
The qualitatively new features of the multifield case arise from solutions of the background equations of motion,
whose field-space trajectories are (strongly) non-geodesic. In the case of two fields, that is, when n = 2, the deviation
from a geodesic is measured by the turning rate of the background trajectory (ϕ10 (t), ϕ20 (t)) in field space, Ω =
−NI Dt T I , where T I and NI are unit vectors, respectively, tangent and normal to the trajectory; also Dt T I ≡
ϕ̇J0 ∇J T I . Strongly non-geodesic trajectories are characterized by large turning rates. Multifield models of LDE,
whose background solutions have such rapid-turning field-space trajectories, can have an equation of state parameter
w very close to −1 and, despite that, can be observationally distinct from a cosmological constant [2754–2757]. This is
because the speed of sound cs of the dark-energy perturbations around the background solution can be (significantly)
reduced compared to the speed of light.
DE perturbations around the exact background solution of Ref. [2756] were studied in Ref. [2757]. It was shown
there that the speed of sound of these perturbations is significantly reduced. Combining that with an equation of
state parameter w ≈ −1 , as well as incorporating matter in the exact background solution, it was further argued
in Ref. [2757] that this model of multifield DE is very promising for alleviating simultaneously the σ8 and Hubble
tensions.
More generally, a low sound speed of DE perturbations is a natural possibility of any k-essence model [1258, 2758,
2759] and also phantom models [1259]. In this case, DE perturbations can grow on small scales and impact the
structure formation, see Ref. [2760] for a review. Recent efforts in understanding the effects of DE perturbations in
structure formation include Refs. [2761–2764], but we still need an accurate determination of their impact on nonlinear
observables.
With more freedom in the perturbative sector, clustering DE models could circumvent the usual trend seen in
late DE models of worsening the growth tension. However, some explorations of this case were conducted only for
non-phantom EoS. Recent analyses of this scenario indicate that DE with low sound speed worsens the S8 tension
[2765] and the very low cs values are not allowed by current data [2766].
4.2.1.e. Running vacuum models Quantum fluctuations in the expanding Universe induce the running of the
vacuum energy density (VED). An example, motivated by dualities in string theory is Ref. [2767] where the spacetime
vacuum is such that its trans-Planckian modes are coupled and track the Hubble expansion rate [2768–2772]. The
running character of the vacuum was originally motivated using renormalization group arguments in curved spacetime
[2773–2775], but more formal calculations have been recently performed in the context of QFT in curved spacetime
using adiabatic regularization and renormalization techniques [2776–2778] as well as effective low-energy string theory
[2459, 2460, 2467]. See Refs. [2779, 2780] for dedicated reviews of the RVM within the QFT approach and [2461] for
the stringy version. The running VED takes the following form
3
ρvac (H) = (c0 + νH 2 + ν̃ Ḣ) + O(H 4 ) . (4.11)
8πG
In contrast to the rigid energy density ρΛ = Λ/8πG associated with the cosmological constant in ΛCDM, ρvac
also receives dynamical contributions from the quantized matter fields. For ν, ν̃ ̸= 0 the VED becomes dynamical.
Eq. (4.11) lies at the core of the class of RVMs. The higher-order terms O(H 4 ) can explain inflation with a graceful
169

exit without the need of ad hoc scalar fields [2463–2465, 2781, 2782]. Here we just focus briefly on the low-energy
terms appearing in Eq. (4.11), which are controlled by the parameters ν and ν̃. Their values depend on the masses
and non-minimal couplings (in the case of the scalar fields) of the various particle species of the theory, and they are
expected to be ≲ O(10−4 − 10−2 ), although they are ultimately determined from observations. In spite of being small,
the running of the vacuum can leave a non-negligible imprint on the post-inflationary stages of the cosmic expansion
both at the background and perturbation levels. The time evolution of ρvac may imply either the non-conservation of
matter and/or the time-evolution of the gravitational coupling G and other fundamental “constants” of nature, and
this can lead to a variety of scenarios explored in the literature [2783–2788]. The RVMs are able to accommodate
the wealth of cosmological data at our disposal and have a positive bearing on the cosmological tensions. The RVMs
were actually the first to point out significant hints of DE (vacuum) dynamics in the Universe using a large amount
of cosmological data, almost one decade before the advent of DESI [2789, 2790], see also Refs. [2789–2795] and
Refs. [2796–2799]. The most updated constraints on these models can be found in the recent works Refs. [2800, 2801],
where the authors analyze the RVMs in light of a complete dataset under the simplified assumption ν̃ = ν/2 and
show that coupling between DM and the running vacuum can significantly losen the growth tension, especially when
the vacuum dynamics is activated at z ∼ 1, and also that a joint running of ρvac and G can strongly mitigate both
the growth and H0 tensions. This is also true in the Brans-Dicke model [2406–2408, 2802].
The stringy version of the RVM [2459–2461, 2467] provides both early- and late-dark-energy models within string
theory [2803, 2804]. This model, a Chern-Simons modification of GR [2457], results in a vacuum energy density with
logarithmic corrections in the Hubble parameter, H 2 ln H. These corrections, arising from quantum graviton effects
in an expanding Universe [2805], help alleviate the H0 and growth-of-structure tensions [2472], and can dominate
logarithmic corrections from matter QFT effects [2776–2778].
4.2.1.f. Emergent dark energy Dynamic DE models are gaining attention over the simple cosmological constant,
Λ, to reconcile discrepancies between CMB and SH0 ES data [2496, 2806]. One such model is Critically Emergent
Dark Energy (CEDE), where DE emerges as a critical phenomenon [2807–2809]. In CEDE, DE density grows from
zero in a phase transition, with an order parameter sensitive to photon thermal baths. Below a critical redshift, zc ,
DE self-interaction dominates, leading to universal order. This model uses Ginzburg-Landau theory to describe phase
transitions and derives DE density dependence below zc as [2809, 2810]

(4.12)
p
κ2 ρDE (z)/3H02 = (1 − Ωm,0 − Ωr,0 ) 1 − z/zc .
−1
Confronting this phantom-like model with the CMB data [728], yields H0 = 70.0+1.2 −2.7 km s Mpc−1 with less χ2
than ΛCDM. Adding SN data [2023] and/or BAO data [775, 776, 2421, 2503], pushes zc toward larger values where
practically ΛCDM is restored; and the H0 posterior shrinks to the tension zone of Planck18. CEDE offers richer
predictions beyond phantom-like DE. Considering the specific scale dependence of DE density fluctuations in this
model [2811], it alleviates both low-ℓ/high-ℓ inconsistencies in the CMB angular power spectrum and the AL anomaly
reported by Planck [192]. Assuming CEDE, 1σ posterior intervals of cosmological parameters are constrained with
ℓ < 800 and ℓ > 800 overlap. Additionally, in this framework, AL is consistent with unity well within the 1σ region.
Another Phenomenologically Emergent Dark Energy (PEDE) model developed to address the Hubble tension is
[2496]

ΩDE (z) = ΩDE,0 [1 − tanh (log10 (1 + z))] . (4.13)

The lack of additional parameters makes the model especially intriguing. However, Ref. [2812] argues that while
the PEDE model fits background data well, it fails to match cluster-scale observations compared to ΛCDM at the
perturbation level. Various authors further analyzed this model statistically with diverse cosmological datasets and
in extended frameworks [1265, 2176, 2181, 2499, 2676, 2813–2819]. However, it offers a worse fit to the combined
datasets in comparison to ΛCDM [709]. As extensions of the PEDE model, include the generalized emergent dark
energy (GEDE) [2499, 2806] and the modified emergent dark energy (MEDE) models [2820]. Both the MEDE &
GEDE frameworks can help in reducing the H0 tension to ≲ 3σ CL.
4.2.1.g. Vacuum metamorphosis The Vacuum Metamorphosis (VM) model, proposed to explain the late-time
accelerated expansion of the Universe, is based on non-perturbative quantum gravitational effects [2821–2823]. It
features a minimally coupled, ultra-light scalar field with mass m ∼ 10−33 eV. The model’s key aspect is a gravitational
phase transition occurring when the Ricci scalar curvature R reaches the order of the scalar field’s squared mass, m2 ,
which is related to the current matter density parameter Ωm,0 . In this framework, the gravitational phase transition
3Ωm,0
occurs at the critical redshift zt = −1 + 4(1−M ) , where M is a free parameter of the theory.
This transition leads to an effective radiation component in the Universe post-transition, resulting in a De-Sitter
phase, distinct from the ΛCDM scenario. In the original VM scenario, there is no cosmological constant at high
170
h i1/4
3
redshifts, and the matter density parameter Ωm,0 is given by 43 (3M (1 − M − Ωk − Ωr,0 )) , where Ωk = −k/H02
and Ωr,0 is the radiation density parameter. Joint analysis with Planck 2018 + BAO + Pantheon yields H0 =
74.21 ± 0.66 km s−1 Mpc−1 [2824], which resolves the H0 tension within 1σ. An extension of the VM model, namely
the VM-VEV model, includes a cosmological constant at high redshifts, represented by the vacuum expectation value
of the massive scalar field. This extension requires additional conditions such as the transition occurring in the past
(zt ≥ 0) and the DE density being non-negative post-transition i.e., ΩDE (z > zt ) ≥ 0, respectively. Joint analysis
with Planck 2018 + BAO + Pantheon leads to H0 = 73.26 ± 0.32 km s−1 Mpc−1 [2824]. While the VM & VM-VEV
models can align H0 with the SH0ES value, they exhibit a poorer fit to low-redshift data, such as BAO and Pantheon,
compared to ΛCDM.
4.2.1.h. Others DE frameworks and their implications for the H0 tension Rather than initially adopting a par-
ticular physical model for DE, given our lack of clear insight into its origin, we can opt for an exploratory approach
utilising a probing function to represent the DE density. The DDE model, generalized to w(z) = n wn x(z), where
P
the parameters wn are fixed by observations and x(z) is a function of redshift that can correspond to both phantom
and non-phantom fields, has been shown to alleviate the σ8 tension [2825] as well as the H0 tension. It is worth
noting that the H0 tension improves further when the neutrino mass ν mν is taken into account [2826]. Another
P
approach was initiated in Refs. [2045, 2046, 2827], where the authors used both linear and quadratic probe functions
for X(z) = ρρDE (z)
DE (0)
. For ΛCDM, X(z) = 1, while X(z) ̸= 1 for any redshift z indicates DE evolution. A quadratic
probe function used in the literature is proposed in Ref. [2047]. This perspective shows that such an approach may
prefer X(z) decreasing with z, even taking negative values for z > 1. Using the most recent BAO observations from
the DESI [698], the same trend was found in Ref. [808]. This trend, including the possibility of a negative X(z),
poses a challenge for DE modeling and may have profound implications in light of recent cosmological tensions and
the foundations of standard cosmology.
A pseudo-Rip DE scenario, leading to asymptotically de Sitter behavior, is proposed in Ref. [2828]. This model
appears to be statistically favored over the consensus ΛCDM model according to some Bayesian discriminators. The
ηλsech(2λ)
equation of state parameter, typically evaluated at the present time, is given by w0 = −1− 3 arctan(sinh(λ)) , where η is a
free parameter of the model. It is concluded that the pseudo-Rip ΛCDM model significantly improves the description
of late-time (z ≤ 2.5) data more decisively than that at higher redshifts.
Late-time DE transitions at redshifts z ≪ 0.1 [2520] can make the predicted value of H0 compatible with SH0ES
measurements. Conversely, in Ref. [2829], a model-independent constraint on late-time models shows strong evidence
against homogeneous new physics over the ΛCDM model. Surprisingly, despite the absence of H0 and MB tensions in
the local Universe, Ref. [2830] argues that late-time data solutions to the H0 tension require a smaller sound horizon
at the recombination era. Additional discussions on rd scales and H0 tension can be found in Refs. [768, 2831]. The
authors in Ref. [2832] argue that new physics at low redshifts cannot resolve the H0 tension.
It is important to emphasize that simple scalar field quintessence models have been widely studied for their ability
to address the coincidence problem through tracking behavior. A possibility is to use generalized axion-like models
with inverse-cosine-like potential [2833–2835], which provides natural entrance and exit of the tracking regime within
a single-field framework. While it is not ruled out, the Hubble tension can be slightly reduced [2836]. Also, a different
approach to tackle the late acceleration of the Universe is by invoking a modified Chaplygin gas [2837–2840], being
the simplest case when its equation of state deviates from that of a cosmological constant by a simple power law of
its energy density. This was analyzed from a perturbation point of view in Ref. [2841].
We also highlight that the cosmological stasis framework [2842–2844] is a phenomenon arising within many BSM
cosmologies in which the abundances Ωi of multiple cosmological energy components with different equation of state
parameters wi remain constant across extended periods despite cosmic expansion. This occurs due to dynamical
feedback within the equations of motion for the corresponding energy densities, with the stasis state serving as a
global dynamical attractor. Such systems evolve toward stasis irrespective of initial conditions and remain in stasis
until some underlyingP mechanism alters the dynamical feedback. Most importantly, the effective equation of state
parameter weff = i Ωi wi for the Universe remains constant during stasis but nevertheless generically differs from the
canonical values associated with, e.g., matter, radiation, or vacuum energy. Stasis at either early or late-times (relative
to recombination) can therefore potentially broaden the scope of possibilities for reconciling cosmological tensions. In
a similar context, we can also mention the effects of the backreaction of super-Hubble cosmological fluctuations on
the late-time accelerated expansion [2845]. Some works suggested that this mechanism could drive a dynamical self-
regulating relaxation of the cosmological constant, as first speculated in Ref. [2846]. This behaviour could potentially
result in an oscillatory effective DE [2847]. A discussion on the tension problem within these scenarios can be found
in Ref. [2847], which also provides further references on these models.
171

4.2.2. Dark energy models exhibiting a rapid density transition from negative to positive values in the late Universe

Coordinator: Özgür Akarsu, Rafael C. Nunes


Contributors: Adrià Gómez-Valent, Alexander Zhuk, Anil Kumar Yadav, Anjan Ananda Sen, Antonio De Felice,
Branko Dragovic, Davide Pedrotti, Emmanuel N. Saridakis, Emre Özülker, Eoin Ó Colgáin, Hanyu Cheng, J. Alberto
Vázquez, Joan Solà Peracaula, Jurgen Mifsud, Laur Järv, Leandros Perivolaropoulos, Leila L. Graef, Luca Visinelli,
Luis Anchordoqui, M.M. Sheikh-Jabbari, Nihan Katırcı, Ruth Lazkoz, and Suresh Kumar

The possible need for DE assuming negative density values at high redshifts was first highlighted by BOSS col-
laboration [788], detecting Λ > 0 for z < 1 but favoring ρDE < 0 for z > 1.6, particularly based on Ly-α BAO
measurements at zeff ≈ 2.34, which exhibited a 2.5σ tension with Planck-ΛCDM [192] and suggested a non-monotonic
H(z) evolution that is difficult to reconcile with strictly ρDE ≥ 0 in GR. Independently, Ref. [444] proposed that this
discrepancy with Planck-ΛCDM, known as the Ly-α BAO anomaly, could be explained by an MG model in which
Λ > 0 is dynamically screened, leading to an effective ρDE passing below zero, accompanied by a singularity in its
EoS parameter wDE , at z ∼ 2.3. This discrepancy was later reduced to ∼ 1.5σ with the completed eBOSS [1526].
However, model-independent/non-parametric reconstructions of ρDE , using the BAO data, consistently favor ρDE
crossing below zero or vanishing for z ≳ 1.5 − 2. Moreover, such reconstructions of wDE —though inherently unable
to capture a sign change in DE density—persistently favor wDE ∼ −1 for z ≲ 1 − 1.5 but trend toward large negative
values below −1 at z ∼ 1.5 − 2—which is expected right after a DE density smoothly transitions from negative
to positive as the Universe expands, therefore potentially indicating a sign change in ρDE . The recent DESI Ly-α
BAO data shows no tension with Planck-ΛCDM [698]. However, while DESI BAO data—when analyzed using the
CPL parametrization—provides more than 3σ evidence for dynamical DE [698], a less noted but significant finding
is that non-parametric reconstructions of DE density also indicate the possibility of vanishing or negative values for
z ≳ 1.5 − 2 [808, 2848], consistent with trends observed in pre-DESI BAO data, e.g., SDSS BAO [2134, 2151, 2848].
It was not immediately recognized that addressing the Ly-α anomaly could naturally be linked to the H0 and S8
tensions, both of which emerged sometime after the Ly-α anomaly was first identified. A brief explanation of this
connection is as follows [800, 810]: If a DE model leaves the pre-recombination Universe ´ ∞ remains unaltered, as in the
standard ΛCDM model, then the comoving sound horizon at last scattering, r∗ = z∗ cs H(z)−1 dz, is expected to
remain effectively unchanged from its ΛCDM counterpart. The Planck CMB spectra provide precise, nearly model-
independent measurements of the angular scale of the sound horizon, θ∗ = r∗ /DM (z∗ ), and the present-day physical
matter density, Ωm,0 h2 , derived from the peak structure and damping tail. Consequently, in DE models featuring
negative DE ´densities at high redshifts, say, for z > z† , both the comoving angular diameter distance to last scattering,
z
DM (z∗ ) = c 0 ∗ H(z)−1 dz, and Ωm,0 h2 must remain consistent with their Planck-ΛCDM inferred values. This requires
that any suppression of H(z) at z > z† , due to negative DE density, must be compensated by an enhancement of H(z)
at z < z† to maintain consistency with the Planck-ΛCDM-inferred DM (z∗ ). As a result, this mechanism naturally
increases H0 (implying a fainter MB ) and decreases Ωm,0 relative to Planck-ΛCDM. A later transition (i.e., smaller
z† ) leads to a prolonged phase of suppression, or similarly, more negative DE density values lead to more suppression;
both effects amplify the enhancement in H0 and reduction in Ωm,0 , provided the transition occurs before the negative
DE density becomes dominant—beyond which expansion would halt and contraction would ensue. This framework
also predicts higher σ8 values because the suppressed H(z) at z > z† reduces cosmic friction, thereby enhancing
structurepformation at high redshifts [2682, 2849]. However, the lower present-day matter density leads to reduced
S8 = σ8 Ωm,0 /0.3 values, potentially resolving the S8 tension. This reduced cosmic friction enhances structure
formation for z > z† [800], potentially explaining the JWST anomaly [1748, 1752, 2850, 2851]—where deep-space
observations at z ≳ 5 indicate stronger structure growth than predicted by Planck-ΛCDM.
In the rapidly growing literature, fitting within the framework outlined above, Refs. [2852] and [2686] are the earliest
studies explicitly connecting the Ly-α anomaly with the H0 tension. Ref. [2852] assumed that the Universe is consistent
with Planck-ΛCDM for z ≳ 4 and reconstructed H(z) using low-redshift data, including Ly-α BAO. They found that
the DE density reaches a minimum within a certain redshift range and becomes negative for z ≳ 2, accompanied by
higher H0 values, and argued that this behavior can be most simply explained by an AdS-like cosmological constant
(Λ < 0) combined with an evolving DE component. Ref. [2686] proposed that inertial mass density, ϱ ≡ ρ + p, maybe
more fundamental than energy density and introduced a DE parametrization with a minimal dynamical deviation
from the usual vacuum energy/cosmological constant (ϱ = 0) in the form ϱ ∝ ρλ , referred to as graduated dark energy
(gDE) (ϱ = const, simple-gDE [2853]). gDE exhibits a wide range of behaviors depending on λ; notably, for large
negative λ, it serves as a phenomenological model for an AdS-to-dS-like transition in DE. It was shown via gDE
that joint observational data, including but not limited to Planck-CMB and Ly-α BAO, suggest λ ∼ −18, indicating
a behavior analogous to a rapid AdS-to-dS transition at z† ∼ 2, consequently alleviating both the H0 tension and
the Ly-α anomaly. With this constraint on λ, gDE resembles a smooth step-like function, yielding ρgDE /ρc0 ∼ −0.7
with wDE ≳ −1 for z† ≳ 2 before rapidly switching sign at z† ∼ 2 and settling at ρgDE /ρc0 ≈ 0.7 with wDE ≲ −1
172

for 0 ≤ z ≲ 2.56 This led to the conjecture that around z† ∼ 2, the Universe underwent a period of rapid mirror
AdS-to-dS transition in vacuum energy—rapid sign-switching cosmological constant, Λs , from negative to positive
while preserving the magnitude—or a similar phenomenon.
The Λs CDM framework [800, 801, 810, 811, 1320, 2856] extends ΛCDM by replacing the usual cosmological constant
(Λ) with a dynamically evolving counterpart (Λs ) that undergoes a mirror AdS-to-dS transition in the late Universe,
while leaving other standard cosmological components—such as CDM, baryons, pre-recombination physics, BBN, and
inflation paradigm—unchanged. This transition can typically be described using sigmoid-like functions, such as the
smooth approximation of the signum function, sgn x ≈ tanh(kx), where k > 1 and x represents either redshift (z) or
scale (a). An example is Λs (z) = Λs0 tanh[ν(z† − z)]/ tanh[νz† ], where ν > 1 controls the sharpness of the transition,
Λs0 > 0 is the present-day value, and z† denotes the transition redshift. For a rapid transition (e.g., ν ≳ 10) occurring
at z† ∼ 2, this function effectively behaves as Λs ≈ Λs0 for z ≲ 2 and Λs ≈ −Λs0 for z ≳ 2. In the limiting case
ν → ∞, the transition becomes instantaneous, Λs (z) → Λs0 sgn[z† − z], defining the abrupt Λs CDM model, which
extends ΛCDM by a single additional parameter, serving as an idealized representation of a rapid mirror AdS-to-dS
transition. Abrupt Λs CDM [800, 801, 810, 811, 1320, 2856] has emerged as one of the most promising and economical
extensions of ΛCDM; introduces only a single additional parameter, z† ∼ 2 (estimated through robust observational
analyses), beyond ΛCDM, resolving major cosmological tensions—including those in H0 , MB , and S8 —as well as the
BAO Ly-α anomaly, while yielding an age of the Universe consistent with estimates from the oldest globular clusters.
Additionally, when allowing variations in mν and Neff , it predicts values consistent with the standard model of particle
physics [1320], suggesting that it may avoid the recently proposed anomaly in which cosmological data (such as DESI
BAO), within the ΛCDM framework, appears to favor mν < 0 [1302, 1319, 2374, 2857–2859] (see Sec. 2.3.5). From
a physical perspective, Λs CDM is identical to ΛCDM for z < z† , featuring a dS-like cosmological constant after the
transition, but introduces a minimal modification by adopting an AdS-like cosmological constant for z > z† , i.e., for
all redshifts prior to the transition. However, from a phenomenological perspective—viz., in terms of the Universe’s
expansion dynamics and observational signatures—the impact of this modification is effectively confined to redshifts
z ≲ z† ∼ 2. Specifically, Λs CDM replicates the H(z) of ΛCDM for z < z† —albeit with systematically larger values—
introduces H(z) deformation around z ∼ z† , and becomes nearly indistinguishable from ΛCDM at higher redshifts
(z ≳ 3). Consequently, from a phenomenological standpoint, Λs CDM is a post-recombination/late-time modification
to ΛCDM. Abrupt Λs CDM, the simplest phenomenological realization of the Λs CDM framework, has been extensively
studied [800, 801, 810, 811, 1320]. However, its abrupt transition at z = z† introduces a discontinuity, leading to
a type II (sudden) singularity [2860] at z = z† , though Ref. [2682] has shown this has negligible impact on cosmic
structure formation and evolution. Nonetheless, this singularity suggests that the abrupt Λs CDM model should be
interpreted as an idealized approximation, effectively serving as a proxy for a rapid yet smooth transition.57
We refer readers to Refs. [425, 444, 768, 788, 801, 1283, 1286, 1287, 1320, 2036, 2092, 2093, 2212, 2236, 2453, 2473,
2496, 2675, 2687, 2807, 2808, 2852–2854, 2856, 2862–2894] for further theoretical and observational studies—including
model agnostic reconstructions (see Sec. 3.3)—that explore DE with negative densities, often consistent with an AdS-
like cosmological constant, at z ≳ 1.5 − 2, and aimed at addressing major cosmological tensions. Phantom DE models,
which typically feature ρDE that decreases with redshift but are conventionally assumed to yield ρDE > 0, are known
to mitigate the H0 tension. Among these, the phantom crossing model, proposed phenomenologically in Ref. [2895]
(DMS20 [812]), stands out. A recent analysis, which considered this model as a particular example of the broader
Omnipotent DE class [812], reaffirmed its success while also revealing that its ability to assume negative densities for
z ≳ 2—mimicking an AdS-like cosmological constant beyond sufficiently high z—is central to its effectiveness. IDE
models [767, 799, 2415, 2896–2906] (see Sec. 4.2.3) offer an alternative approach to resolving the H0 tension; however,
model-independent reconstructions of the IDE kernel [2151] suggest that negative DE densities at z ≳ 2 persist as a
possibility.
While late-Universe rapid AdS-to-dS (or analogous) transitions in DE, as proposed by Λs CDM, were initially
viewed as challenging to reconcile with a robust physical mechanism, the remarkable phenomenological success of
this approach—despite its simplicity—has prompted deeper theoretical inquiries. Even established frameworks, upon
re-examination, have been found to accommodate such transitions within previously overlooked solution spaces,

56 For a minimally interacting DE with ρDE transitioning from negative to positive at z† as the Universe expands (z decreases), physical
consistency demands ρDE < 0 with wDE > −1 just before the transition and ρDE > 0 with wDE < −1 just after, ensuring a smooth
crossing through ρDE = 0 while maintaining the continuity equation. Since pDE remains finite, wDE exhibits a singularity at z = z† ,
with limz→z± w(z) = ±∞. However, this is a safe singularity—ρDE remains finite, and wDE diverges solely because ρDE crosses zero

while pDE remains finite and nonzero. Consequently, given that H 2 ∝ ρm + ρDE , H and other kinematical parameters evolve smoothly;
in scalar field realizations, the sound speed remains luminal, further reinforcing the regularity of the crossing [2854]. By contrast, the
alternative scenario of sign-changing ρDE —where ρDE < 0 with wDE < −1 just before the transition and ρDE > 0 with wDE > −1
just afterward—leads to singularities in both wDE and ρDE , as limz→z± wDE (z) = ∓∞ and limz→z± ρDE (z) = ∓∞, thereby leading
† †
to a breakdown of the spacetime metric. Thus, only the first scenario is physically viable, predicting a phantom-like phase (wDE < −1)
following the transition. Consequently, the tendency of model-agnostic reconstructions of wDE (z) and various EoS parameterizations
(e.g., CPL) to favor large negative values beyond −1 for z ∼ 1.5 − 2 may indicate negative or vanishing DE densities for z ≳ 1.5 − 2.
For further discussion, see Refs. [812, 2682, 2854, 2855]
57 A well-defined formulation necessitates a smooth transition, enabling a detailed study of perturbations but introducing theoretical
challenges. A rapid AdS-to-dS transition can sharply increase Ḣ, inducing transient super-acceleration (Ḣ > 0), which may affect
cosmological observables and, within GR, is often linked to ghost instabilities and WEC violations. These impose constraints on the
transition’s rapidity in GR but may be circumvented in type-II minimally MG theories such as VCDM [2861, 2862]. For further
discussion, see [801, 2854, 2856].
173

prompting researchers to adopt a fresh perspective on familiar theories. For instance, Λs CDM+ [2687, 2883, 2884]
proposes a stringy realization of Λs CDM; it was shown that, despite the AdS swampland conjecture suggesting that
a late-universe AdS-to-dS transition is unlikely due to the arbitrarily large separation between AdS and dS vacua in
moduli space, such a transition can nonetheless be realized through the Casimir forces of fields inhabiting the bulk.58
Building on the same theoretical framework, Λs CDM± [2910] extends Λs CDM+ by allowing the AdS phase to have
arbitrary depths, considering different curvature radii in the AdS and dS phases. It was demonstrated in Ref. [2881]
that, in various formulations of GR, a Λs can arise naturally through an overall signature change of the metric.
Λs VCDM [801, 2856] advances the Λs CDM framework into a theoretically complete physical cosmology, offering a fully
predictive description of the Universe, including the AdS-to-dS transition epoch itself. It was shown that the mirror
AdS-to-dS transition can be effectively realized within a type-II minimally MG framework called VCDM [2861, 2862],
through a specific Lagrangian incorporating an auxiliary scalar field with a smoothly sewed two-segmented linear
potential. Ref. [2887] demonstrated that the teleparallel f (T ) gravity, specifically its exponential infrared form [2911],
which has shown significant promise in addressing the H0 tension [2912, 2913] in its solution space giving phantom-
like effective DE, admits previously overlooked solution spaces, which accommodate an alternative scenario where
the effective DE transitions smoothly from negative to positive at z† ∼ 1.5, while remaining consistent with CMB
power spectra. Building on these insights, f (T )-Λs CDM successfully maps the background dynamics of Λs CDM into
the f (T ) gravity framework [2888], further establishing a theoretical foundation for AdS-to-dS-like transitions in the
late Universe. Ph-Λs CDM [2854], introduced a phantom DE model, which is a specific realization of a general scalar
field with a hyperbolic tangent potential that induces smooth AdS-to-dS, 0-to-dS, and dS-to-dS transitions. Despite
its negative kinetic term, the step-like potential prevents pathologies like unbounded energy growth, Big Rip, and
WEC violations, ensuring smooth evolution of cosmological parameters. Notably, the AdS-to-dS transition in DE
density does not precisely parallel that of the potential, persisting longer due to the kinetic term’s contribution.
Even if different realizations of Λs CDM yield identical background dynamics, they still exhibit differences. GR-based
Λs CDM models [800, 810, 811, 2854], Λs VCDM [801, 2856], and f (T )-Λs CDM differ in their predictions for linear
perturbations, while the string-inspired Λs CDM+ [2687, 2883, 2884, 2910] predicts a modest excess in the total effective
number of neutrino species, with Neff = 3.294. Such distinguishing features are invaluable, as they provide a means
to compare and ultimately discriminate between these alternative realizations using observational data. Other than
the models explicitly realizing/resembling the background dynamics of Λs CDM, there exist various other approaches
that fit within the broader paradigm outlined earlier. These include brane-world models [444, 2865, 2866], multiple
axion models [2914, 2915], energy-momentum log gravity [2871], bimetric gravity [2893], Horndeski gravity [2882],
holographic DE [2890], Granda–Oliveros holographic DE [2891], composite DE (wXCDM) [2473, 2892], inspired by the
ΛXCDM framework [2916]—extends the abrupt Λs CDM model by introducing two free parameters that allow wDE to
take arbitrary constant values before and after the transition59 —, the Omnipotent DE concept [812], with its specific
realization DMS20 [2895], represents a class of DE models characterized by non-monotonic densities and transitions
across wDE = −1, DE pressure parametrizations [2917, 2918], which yield wDE similar DMS20, the Lotka-Volterra
model of two interacting fluids [2880], running Barrow entropy [2879], multiple-transition vacuum DE models [2502],
and scenarios invoking a modification of the gravitational constant between super- and sub-horizon regimes, motivated
by the Hořava–Lifshitz proposal or the Einstein-aether framework [2919].
If future observations continue to support these models, the implications for theoretical physics would be profound,
as an AdS-like cosmological constant is a theoretical sweet spot, favored by the AdS/CFT correspondence [2920]
and string theory frameworks [2921]. Recent work [2922] has demonstrated that a supersymmetric vacuum in string
theory can naturally produce an AdS-like cosmological constant at present-day energy scales, motivating DE scenarios
where the field evolves on a potential with an AdS minimum rather than the standard dS minimum. Such a negative
cosmological constant could shape both the current cosmic acceleration and the Universe’s long-term evolution, offering
a compelling link between fundamental theory and observations. In this direction, Ref. [2872] examined a DE model
consisting of an AdS-like cosmological constant and a DE component with an EoS parameter wϕ , finding no evidence
for an AdS-like cosmological constant and a mild preference for an effective phantom DE component, though ΛCDM
remains favored. Ref. [2923] (see also [2875]) extends wCDM and CPL-CDM by introducing an AdS-like cosmological

58 By combining swampland conjectures with observational data, it was proposed that the cosmological hierarchy problem—i.e., the
smallness of DE density in Planck units—could be understood as an asymptotic field-space limit corresponding to the decompactification
of a micron-sized extra (dark) dimension [2907]. Within this framework, Casimir forces from fields inhabiting the dark dimension can
drive an AdS-to-dS transition [2883], forming the basis of the Λs CDM± and Λs CDM+ . Specifically, a 5D Einstein-de Sitter gravity
action compactified on a circle induces a runaway potential inherited from the 5D cosmological term [2908]. If the 5D cosmological
constant is small, the quantum contribution of the lightest 5D modes—identified with Casimir energy [2909]—becomes significant. A
minimal setup requires a 5D mass spectrum comprising the graviton, three generations of light right-handed neutrinos, and a real scalar
field φ with a potential featuring two local minima [2883]. At z† ∼ 2, φ undergoes quantum tunneling from the false to the true
vacuum [2884], acquiring a larger mass and suppressing its Casimir energy contribution. This modifies the balance of fermionic and
bosonic degrees of freedom, triggering the AdS-to-dS transition. The deep infrared fields of the dark sector contribute to the effective
number of relativistic neutrino-like species.
59 The wXCDM model [2473, 2892], inspired by ΛXCDM [2916], introduces a composite DE scenario with two components: X (for z > z† )
and Y (for z < z† ). Y behaves as running vacuum energy with a quintessence-like EoS (wY ≳ −1), while X, termed ‘phantom matter’
(PM), has a phantom-like EoS (wX ≤ −1) but negative energy density, mimicking effective string action terms at low energies [2461].
The Kalb-Ramond axion and gravitational Chern-Simons term generate a ‘phantom vacuum’ [2467], where PM enhances structure
formation at high z, potentially explaining the JWST anomaly. The de Sitter vacuum is restored via gChS condensates, and data
support wY > −2, wX < −1, aligning with DESI results [698].
174

constant, demonstrating improvements in both data fits and alleviation of the H0 tension, while further studies [1753–
1755] show that this model can also address the JWST anomaly. Indeed, the models discussed in this section are
expected to generally enhance structure formation at high z, making them natural candidates for explaining the JWST
anomaly [1748, 1752, 2850, 2851]. However, their rigorous quantitative analysis of perturbation evolution and cosmic
structuring is necessary. The evolution of cosmic structures and linear perturbations in the (abrupt) Λs CDM model
has been studied in Refs. [1756, 2682, 2849]. Specifically, Ref. [1756] found that for ΩΛ,0 = −0.7 and Ωm,0 = 0.3,
Λs CDM predicts up to an 80% increase in cluster density for turnaround redshifts zmax ≳ 2, suggesting a potential
explanation to the JWST anomaly. Ref. [2682] demonstrated that even in abrupt Λs CDM (the most extreme case),
the transition itself has no significant impact on bound structures, preserving model viability. Additionally, Ref. [2849]
showed that the growth index remains γ ∼ 0.55 as in ΛCDM. However, Planck CMB data predicts Ωm,0 = 0.28 for
Λs CDM and Ωm,0 = 0.32 for ΛCDM, leading to growth rates of f = 0.49 and f = 0.53 at z = 0, respectively. Notably,
Λs CDM predicts a value closer to f = 0.48, recently obtained from LSS data when γ is treated as a free parameter
in ΛCDM [2050], indicating its potential to resolve the structure growth anomaly. DESI data (combination with
other datasets) have further validated DE models incorporating an AdS-like cosmological constant [2885, 2889], while
obviating the need for phantom DE in the DE sector [2889]. The post-reionization HI 21-cm signal is explored as a
probe for an AdS-like cosmological constant in the DE sector [2924]. In summary, a negative cosmological constant
in the DE sector is physically motivated, consistent with current data, and may yield distinctive cosmological and
astrophysical signatures—including the possibility of a bouncing future universe [2925].
Ref. [2926] suggested that a promising approach to resolving cosmological tensions may involve combining early-
and late-time new physics to better fit the data. Ref. [2886] explored a hybrid model integrating abrupt Λs CDM
with a varying electron mass mechanism, identifying it as a promising candidate. However, they found that this
combination does not P improve the tension, as theP two models push Ωm,0 in opposite directions. P Ref. [1320] analyzed
the Λs CDM+Neff + mν and ΛCDM + Neff + mν models, allowing variations in N eff and mν within abrupt
Λs CDM and ΛCDM. They found that Λs CDM+Neff + mν consistently fits the data better while P preserving the
P
success of Λs CDM and predicting standard neutrino properties. In contrast, when ΛCDM + N eff + mν yields high
H0 values, it does so at the cost of large ∆Neff , whereas Λs CDM+Neff + mν achieves similarly high 0 values while
P
H
remaining consistent with the standard Neff . This suggests that the Λs CDM+Neff + mν alleviates the need for
P
pre-recombination new physics, at least concerning neutrino properties.

4.2.3. Interacting dark energy


Coordinator: Carsten van de Bruck
Contributors: Amare Abebe, Dario Bettoni, David Benisty, David Tamayo, Denitsa Staicova, Diego Rubiera-Garcia,
Emmanuel Saridakis, Emre Özülker, Kay Lehnert, Leila L. Graef, Lu Yin, Luis Anchordoqui, Marcel A. van der
Westhuizen, Marco de Cesare, Nikolaos E. Mavromatos, Oem Trivedi, Oleksii Sokoliuk, Purba Mukherjee, Rahul
Shah, Sveva Castello, Vitor da Fonseca, and Yuejia Zhai

4.2.3.a. Introduction In this subsection, we discuss models with an interaction between DE and DM. Such
IDE models have been introduced to address problems of the ΛCDM model, including the H0 and S8 tensions (for
comprehensive reviews, e.g., see Ref. [708, 2927, 2928]). Some IDE models can reduce the H0 tension from 5σ to 3.6σ
[708].
In a general model, the conservation equations for DM and DE take the form
µν µν
∇µ T(DM) = Qν , and ∇µ T(DE) = −Qν , (4.14)
µν µν
ensuring that the total energy-momentum tensor of DM and DE remains conserved, namely, ∇µ (T(DM) + T(DE) ) = 0.
Often, the resulting background equations are written in the form

ρ̇DM + 3HρDM = Q , and ρ̇DE + 3H(ρDE + pDE ) = −Q . (4.15)

In order for IDE models to be viable candidates for addressing the H0 and σ8 tensions, special attention must be
given to the physicality of the parameter space [2929]. In these models, there is not always a mechanism to halt
the energy transfer when either the DM or DE density becomes zero (i.e., Q ̸= 0 in Eq. (4.15) when ρDE/DM = 0),
which can lead to negative energies. The case of Q > 0 corresponds to an energy transfer from DE to DM. It has
been reported that CMB observations seem to favour an energy transfer from DE to DM, WL measurements and
thermodynamical considerations suggest an energy transfer from DM to DE [2040, 2930–2932].
µ µ
One well studied class of models is coupled quintessence, in which Qν = βϕ,ν T(DM)µ and Q = −βT(DM)µ ϕ̇ , where
µ
ϕ is the DE scalar field and T(DM)µ denotes the trace of the DM energy–momentum tensor [2041, 2410, 2933]. There
175

also exists an extension of these models in which DM and DE are disformally coupled [2934–2937]. Such scalar field
theories of IDE may alleviate the Hubble tension but do not fully solve it [2415, 2417, 2938].
Other, more phenomenological approaches propose a coupling of the form Qµ = QuµDM /a or Qµ = QuµDE /a
[2939, 2940] (in these expressions, a is the scale factor and uµDM or uµDE are the four-velocity of DM or DE fluid,
respectively). The interaction Q is usually written as Q = ξHρDE or ξHρDM , where ξ is a dimensionless coupling
constant, determining the size and direction of energy/momentum flow. The case Q = ξHρDE was recently studied in
Ref. [2941] (see also [2897] for a previous study), assuming an equation of state of DE wDE = −0.999, using the CMB
data provided by Planck, WMAP and ACT. Different dataset combinations resulted in consistent results, showing
evidence for a non–zero coupling and a Hubble expansion rate H0 consistent with local measurements. Using Planck
alone, the analysis results in H0 = 71.6 ± 2.1 km s−1 Mpc−1 , ACT alone results in H0 = 72.6+3.4
−2.6 km s
−1
Mpc−1 and
+2.5 −1 −1
the combination of ACT and Planck results in H0 = 71.4−2.8 km s Mpc . In Ref. [2942], the assumption about the
equation of state was relaxed, allowing for a dynamical evolution via the parametrization wDE (a) = w0 + wa (1 − a).
It was found that models featuring a dynamical phantom equation of state (w0 < −1) perform worse than the non-
dynamical case. On the other hand, models featuring a dynamical quintessence equation of state (w(a) > −1 at any
redshift) perform better in attempting to increase the value of H0 compared to the respective non-dynamical case.
However, when considering the joint analysis of CMB, BAO SDSS, and SN data, no significant increase in H0 to solve
the Hubble tension was found, while it is a promising solution considering DESI BAO data [799].
In the following, we present a number of other IDE models.
4.2.3.b. Models
Nonlinear interacting dark energy This class of models is an extension of the models discussed above. A
general form of the nonlinear IDE models is given by Q = 3HξF (ρDE , ρDM ), where F is a nonlinear function of the
energy densities and ξ is a free constant parameter. Generally, when Q > 0, the H0 tension worsens and the S8
alleviates; conversely, when Q < 0, the H0 tension alleviates and the S8 worsens [2929]. The authors of Ref. [2943]
investigate nonlinear IDE as an interaction between DM and a scalar field, in Ref. [2944] a model is discussed that
could be interpreted as a particular case of a running vacuum model Λ(H).
Power laws IDE models with the general form Q = 3HξρpDM ρsDE (ρDM +ρDE )r , where p, s and r integers, were studied
in Ref. [2945],
 and observational constraints for specific cases provided in Ref. [2946]. Observational constraints of
ρDM ρDE β
Q = 3Hξ ρDM + ρDE + ρDM +ρDE are discussed in Ref. [2947]. The model Q = γρα DM ρDE , using CMB, BAO, RSD,
and SNIa data, predicts lower values of f (z)σ8 (z) at z < 1 comparing to ΛCDM, which alleviates the tension of ΛCDM
with various RSD data [2948]. Other model types, such as exponential models Q = 3HξρDE exp (ρDE /ρDM − 1) [2949],
and logarithmic models Q = 3HξρDE log(ρDE /ρDM ) and Q = 3HξρDE log(ρDM /ρDE ) [2950], were studied but showed
no significant promising results.
Models with the forms Q = 3HξρDE sin(ρDE /ρDM − 1) and Q = 3HξρDE [1 + sin(ρDE /ρDM − 1)] have gained
special attention for their potential to alleviate the H0 tension [2951]. Analysis with Planck 2018 gives H0 =
−1 −1
72.67+5.43
−8.26 km s Mpc−1 at 68% CL, and Planck 2018+BAO gives H0 = 69.17+1.52 −1.71 km s Mpc−1 at 68% CL. Hence,
for Planck 2018 and Planck 2018+BAO, the H0 tension with respect to SH0ES (H0 = 73.0 ± 1.0 km s−1 Mpc−1 ) is
reduced down to 0.05σ, 2σ respectively; notably, for Planck 2018 alone case the tension is completely solved. However,
these results should be interpreted with caution due to potential biases and uncertainties. Additionally, a theoretical
justification for the specific form of Q is necessary to ensure that it is not merely an ad hoc model.
Diffusion interactions A diffusive interaction between DE and DM was introduced in Refs. [2952–2954]. The
diffusion of energy density between DE into DM uses a non-conserved stress energy tensor T µν with a source current
µν
j µ , ∇µ T(m) = γ 2 j ν , where γ 2 is the coupling diffusion coefficient of the fluid. The current j µ is a time-like covariant
conserved vector field ∇µ j µ = 0 which describes the conservation of the number of particles in the system. In a
homogeneous expansion, the modified Friedman equations read
γ γ
ρ̇DM + 3HρDM = , ρ̇Λ = − , (4.16)
a3 a3
The contribution of the current goes as ∼ a−3 since the current is covariantly conserved. In this way, there is a
compensation between DE and DM. Ref. [2955] introduces cases with a diffusion constant that could be represented
for a scalar field ϕ or a perfect fluid, leading to late-time forms of DE and a DE density parameter that could be
interpreted as a perturbation of the ΛCDM model.
Another example of diffusion interactions is represented by models where the energy-momentum transfer vector
comes from a potential, Qµ = ∇µ J. This class of models finds a natural embedding within unimodular gravity. In
fact, in unimodular gravity the total energy-momentum of matter fields needs not be conserved in general [2956]. The
energy density of DE is then identified with the potential J up to an arbitrary integration constant, ρDE = J +Λ0 , and
the equation of state is wDE = −1 identically. Models of this kind have been proposed to describe effective diffusion
176

processes due to an underlying discrete spacetime structure [2957, 2958]. In Ref. [2873] it was argued that such
diffusive interactions —with energy transfer from DM to DE— can alleviate the Hubble tension, see Refs. [2959–2961]
for a more detailed analysis. This class of models is also formally equivalent to “interacting vacuum models”, which
have been studied within the context of general relativity [2962, 2963]. Embedding minimally IDE with wDE = −1
within UG has the following advantages: i ) the vacuum energy does not gravitate in UG removing the age-old
problem regarding the puzzling absence of the vacuum energy density in cosmological observations [2964, 2965], ii )
the well-known perturbative large-scale instability studied in Ref. [2939] does not exist in UG [2966], iii ) since the
DE is an effective source in UG rather than a physical one, ρDE can attain negative values opening up a large
phenomenological landscape to address the cosmological tensions. Moreover, it opens up a path for quantum gravity
motivated interaction terms; e.g., see Refs. [2956–2958].
An important aspect often forgotten in the study of such diffusive models is to study the effect of large-scale structure
formation in such a fluid environment. Growth-rate analyses can shed light on whether such a model alleviates the
σ8 tension or not.
Metastable dark energy Models of metastable DE have been analyzed in several works, see for instance
Refs. [708, 2967–2977]. In particular, in Refs. [2967] and [2969], phenomenological models in this context were
investigated, considering the case of a constant DE decay rate, depending only on the intrinsic properties of DE and
the type of decay channel. The following cases of metastable DE decaying in three distinct ways were considered: (I)
exponentially; (II) into DM; (III) into dark radiation. Among these models, model II showed slightly better consis-
tency [2969]. This model is sometimes called metastable IDE [708], since it is described by Eq. (4.15) with Q = ΓρDE ,
where Γ is the DE decay rate.
In Ref. [2969] it was shown that despite the fit for this model against Pantheon + BAO data providing higher
values for H0 , when the CMB distance priors from Planck 2018 are included, the Hubble tension is restored. Later
on, in Ref. [2970], a full analysis was performed using different combinations of data from Planck 2018, BAO, DES,
R19 [48, 192]. A larger value of H0 was then supported, solving the tension with R20 [78] within 1σ. Future data
with higher precision must provide us with a clearer understanding of the performance of these models concerning
the tension. Some discussion on motivations for metastable IDE arising from quantum mechanics can be found in
Refs. [2971, 2972]. We can also find some examples from field theory descriptions, as in Refs. [2968, 2973–2975], for
instance.
Another model of metastable EDE is provided by the so-called Chern-Simons gravity inspired from string theory,
studied in Refs. [2459, 2461, 2469]. It has been shown that, under some conditions, condensation of chiral GWs in
early epochs of the Universe, can lead to a non-trivial condensate of the Chern-Simons gravitational anomaly term.
The latter is characterized by the presence of imaginary parts, which are such that they lead to a metastable EDE,
with a lifetime that can be in the phenomenologically right ballpark of larger than 50-60 e-foldings. The model, which
also leads to RVM type inflation, due to the non-linear H 4 dependence of the (real part of the) pertinent vacuum
energy density, can also help alleviate the Hubble and growth-of-structure tensions in modern eras [2472].
Elastic dark energy interactions In DE-DM elastic interactions models the two species interact via velocity
or momentum transfer. There are various ways of implementing these exchanges [2978–2980] but here, for the sake of
concreteness, we focus on the particular realization presented in Ref. [2980] and further developed in Refs. [2981–2983]
(see Ref. [2984] for a short review), which is based on the following phenomenological implementation of the coupling
µν
∇ν TDM = α (uµDM − uµDE ) , µν
∇ν Tde = −α (uµDM − uµDE ) , (4.17)

where α measures the strength of the interaction. It is clear that the interaction is active only when there is relative
motion between the two species, i.e., at sub-Hubble scales while, at the largest scales, where the two fluids share a
common reference frame, the interaction is absent. This means, in turn, that no modifications to the background
evolution is produced. As a consequence, these models are not designed to address the H0 tension but have an
impact on relaxing the σ8 one. In fact, at the level of linear perturbation, Euler equation receives a new contribution
proportional to the relative velocity between the two species with a time dependent interaction rate Γα = αa4 /ΩDM
in the case of DM and further weighted by the relative abundance of DM to DE R = a3w ΩDM /(1 + w)/ΩDE in the
case of DE. Since the interaction grows with the Universe expansion as a4 , eventually, Γα ∼ H and the interaction
with DE becomes relevant starting to exert a drag on DM which suppresses its clustering. This allows for the onset
of a dark-coupling epoch at late-time as desired and as suggested by data which points towards a suppression in the
clustering at z ≲ 2. By considering Planck 2018 data releases, SNIa and Baryon Acoustic Oscillations data it has been
+0.011+0.022
shown that the interaction can indeed reduce the value of σ8 and does so without affecting Ωm (σ8 = 0.753−0.011−0.021 ,
+0.007+0.014
Ωm = 0.311−0.007−0.013 ) and that the presence of the interaction is strongly favoured [2981, 2982]. The fact that no
correlations between the interacting parameter and background ones is introduced is a particularly welcome feature
since it implies no worsening of the H0 tension. Remarkably, the inclusion of S8 data allows for a constraint at
percent level of the interaction by surveys like JPAS, Euclid and DESI [2981] while SKAO-like experiments should
177

have enough sensitivity to detect the interaction parameter [2985]. Finally, it is worth mentioning that the inclusion
of massive neutrinos does not spoil the detection prospects for this model [2986] and that other implementations and
analyses reach similar conclusions [2983, 2987, 2988].
Fading dark matter If the swampland conjecture for DE is correct, it would lead to the prediction that the
ΛCDM model, where DE is constant, cannot be veracious [2989, 2990]. Quite independently of swampland consider-
ations, there have been proposals for a rolling scalar field as the source ofDE. If this were the case, the swampland
distance conjecture suggests that the rolling field would lead at late-times to an exponentially light tower of states.
Identifying this tower as residing in the dark sector yields a natural coupling of the scalar field to the DM, leading
to a continually reducing DM mass as the scalar field rolls in the recent cosmological epoch. It has been shown that
the way in which the tower of light states evolves over time could help to reduce (though not fully eliminate) the H0
tension [2991]. More unambiguously, the coupling of the scalar field to the DM leads to a reduction of mass, or fading
of DM, which is compensated by a bigger value of DE. The latter becomes more noticeable in the present accelerating
epoch, leading to an increase in H0 . Salam-Sezgin 6-dimensional supergravity [2992] and its string realization of
Cvetič-Gibbons-Pope [2993] can bring to fruition the fading DM model [2994, 2995].
Late-Time interacting constant/dynamical dark energy Late-time dark energy interacting with DM has
been extensively explored in the context of the H0 and σ8 tensions. Numerous studies have been carried out in the
context of the Hubble tension using different forms of the interaction term and Λ/constant EoS late-time dark energy,
and by imposing various theoretical bounds on the parameter space to prevent instabilities [767, 2897–2900, 2902–
2904, 2926, 2928, 2951, 2996–3012]. Similarly in the context of the clustering tension significant studies include
[2702, 2900–2902, 2904, 3000, 3010, 3013–3017]. In Ref. [2942] an interacting setup with interaction proportional to
the DE density (Q ∝ ρde ), was investigated in the light of late-time DDE assuming a CPL parametrization in the
context of the Hubble tension, and found no significant scope of alleviation of the tension with various combinations
of datasets in either the phantom or quintessence regimes. In Ref. [2814], using the same form of interaction, various
other parametrizations of LDE (MEDE, CPL, JBP) were considered. It was reported here that a quintessence EoS
in such an interacting scenario does not help with the Hubble tension, and noticeably makes the σ8 tension worse.
However, a phantom EoS, although did not affect the Hubble tension, showed significant scope in alleviating the σ8
tension (especially for the CPL and JBP parametrizations). The inclusion of RSD data in this scenario helped obtain
tighter constraints on certain parameters, but resulted in a relatively less pronounced relaxation of the clustering
tension, which might indicate ΛCDM bias in RSD data. A phenomenological parametrization, where a scalar field
depends linearly on the number of e-folds, was used to test its interaction with DM [2932]. However, the constrained
interacting dark energy (CIDER) model, with a background identical to the standard ΛCDM model, appears to better
address the σ8 tension [2414, 3018].
Late-time singularities Depending on the choice for the interacting term Q, late-time singularities, for which
infinities in the Hubble factor (or on its derivatives) appear, may arise at finite future times [2112, 3019–3025]. Since
some of them represent possible ends for the Universe, this undermines the consistency of the theory. Furthermore,
they have been shown to occur for both linear and non-linear forms of Q.
4.2.3.c. Modelling and Observational Aspects of IDE Models In this section, we present some other aspects of
IDE models, such as N -body structure formation simulations of certain IDE models and comparing IDE models to
data.
Marginalization approach to IDE A recent estimate of IDE contribution from Pantheon and Pantheon Plus,
CC and transversal BAO [2996] has shown evidences for up to 2σ deviation from ΛCDM. This study has been done
in the framework of the marginalization approach which aims to circumvent the Hubble tension by integrating H0
and rd out of the model, leaving only Ωm , w and ξ as parameters. The results show a strong dependence on the used
SNIa dataset and whether it is Cepheids calibrated or not, with the latter remaining compatible with ΛCDM.
N -body simulations of IDE Several simulations of the structure formation were run up to date under the
IDE ansatz, namely works of Refs. [2287, 3006], where they have considered several cases of IDE and derived such
probes of the LSS as matter power spectrum, halo mass function, concentration-mass relation and halo bias. The
most interesting parameter is the concentration-mass relation, namely c200 that can be related to the X-ray data and
galaxy rotation curves, which via the maximum likelihood method imposes the redshift evolving constraints on the
DE-DM coupling parameter ξ, with the joint posterior distribution giving a best fit of ξ = 0.071 ± 0.034.
Distinguishing IDE from modified gravity A word of caution is required when constraining IDE models
based on large-scale structure observables. Models involving a fifth force acting on DM lead to a breaking of the weak
equivalence principle for this component and thus a deviation in the Euler equation, affecting the growth of cosmic
structure f . However, the resulting observational signatures on current and forecasted measurements of f σ8 present
a strong degeneracy with gravity modifications in the Poisson equation [3026, 3027]. WL, which is sensitive to the
sum of the two gravitational potentials describing the geometry of the Universe, is also generically unable to break
178

this degeneracy. This indicates that standard large-scale structure analyses may not be able to distinguish IDE from
gravity modifications, leading to a potential ambiguity among models proposed to address cosmological tensions.
Luckily, the two scenarios can be disentangled by considering measurements of gravitational redshift [3028, 3029],
an observable accessible in two-point correlations by the coming generation of galaxy surveys [3030, 3031]. This effect
is sensitive to the gravitational potential encoding the distortion of time and appearing in the Euler equation, thus
providing a direct test of the weak equivalence principle and a clear way to discriminate between the two scenarios.
Model-Independent Reconstructions of IDE Kernel In Ref. [2151], the interaction kernel Q(z), which
governs the energy exchange between DM and DE, is reconstructed within a fully model-independent framework,
without imposing any predefined functional form. Employing both binned step-function reconstructions and GP
interpolation, the authors infer the redshift evolution of Q(z) directly from observational data, including BAO, CC,
and the Pantheon+ supernova sample. The reconstructed kernel displays nontrivial features—most notably, multiple
sign changes and mild oscillatory behavior at the 1σ level—suggesting dynamic and potentially reversible energy
transfer between the dark components. As a result of this evolving interaction, the effective DE density is found
to cross zero and become negative at redshifts z ≳ 2, demonstrating that IDE scenarios do not necessarily exclude
negative DE densities at high redshift (see Sec. 4.2.2), despite the conventional assumption that DE remains positive
in IDE models.
Application of the parametrized post-Friedmann framework in IDE Understanding the interaction of
DE necessitates examining both the expansion history of the Universe and the growth of cosmic structures [3032,
3033]. It has been observed that, in the IDE scenario, calculating perturbations often results in divergent curvature
perturbations on superhorizon scales, posing a significant problem for IDE cosmology [3034–3036].
To address this issue, the parametrized post-Friedmann (PPF) framework is an effective theory that treats DE
perturbations based solely on the fundamental properties of DE [3037–3041]. This approach extends the parametrized
PPF framework, originally designed for uncoupled DE, and can mitigate the instability in IDE cosmology [3042].
Consequently, the entire parameter space of IDE models can be explored using observational data [2708, 3007, 3043–
3046].
A generalized framework to study IDE models is also provided by the effective theory of IDE [3047]. This formalism
relies on a model-independent approach at the level of the action and provides a description of linear cosmological
perturbations in scalar-tensor theories of gravity with a homogeneous and isotropic background. This allows for a
direct comparison with cosmological data to constrain a rich non-standard phenomenology [3027, 3048].

4.3. Modified gravity


4.3.1. Modified gravity in light of cosmic tensions
Coordinator: Francesco Bajardi, Micol Benetti, Salvatore Capozziello
Contributors: Adrià Gómez-Valent, Alessandro Vadalà, Ali Övgün, Amare Abebe, Andronikos Paliathanasis, Anil
Kumar Yadav, Araceli Soler Oficial, Christian Pfeifer, Daniel Blixt, David Benisty, Diego Rubiera-Garcia, Duško
Borka, Emmanuel N. Saridakis, Erik Jensko, Francisco S. N. Lobo, Gabriel Farrugia, Gaetano Lambiase, Giuseppe
Sarracino, Hussain Gohar, Ilim Cimdiker, Inês S. Albuquerque, Ismael Ayuso, Joan Solà Peracaula, Konstantinos F.
Dialektopoulos, László Á. Gergely, Marcin Postolak, Marco de Cesare, Maria Caruana, Mariaveronica De Angelis, Ni-
han Katırcı, Nils A. Nilsson, Noemi Frusciante, Pierros Ntelis, Predrag Jovanović, Rebecca Briffa, Rocco D’Agostino,
Saeed Rastgoo, Sanjay Mandal, Sergei D. Odintsov, Tiago B. Gonçalves, Tiziano Schiavone, Vesna Borka Jovanović,
and Vincenzo Salzano

The shortcomings exhibited by GR on different energy scales question its validity as the best theory to describe the
gravitational interaction [3049]. Indeed, at the quantum level it shows problems with its extension to a quantum theory
of gravity [3050, 3051], since it is non-renormalizable [3052], and the prediction of infinite gravitational tidal forces, by
the existence of singularities [3053]. In the large-scale scenario, the observed Universe’s accelerating expansion [27, 34]
cannot be predicted within the context of the standard model of cosmology without introducing the cosmological
constant, whose theoretical value assessed by Quantum Field Theory differs from the value inferred by the Friedman
equations [2964, 3054]. The repulsive force driving the Universe expansion is dubbed “Dark Energy”, whose existence,
to date, is only based on indirect observations [3055, 3056]. In addition, on galactic scales, the dynamics of the farthest
stars orbiting around the center of galaxies (and, generally, incompatibilities on the LSS) represents a further open
issue [3057, 3058] currently addressed to a never-detected form of matter called “Dark Matter” [3059, 3060].
These incompatibilities therefore open up the possibility of new (or rather, extended) scenarios, and here we will
examine possible modifications to gravity.
Among the most important and widely considered approaches to address the aforementioned problems related to
the unknown nature of DE and DM, as well as cosmological tensions, are modifications to GR itself. These can
broadly be split into two main categories, whose features will be considered in detail in the next sections. The
179

Modified Gravity

Extended Models Alternative Models

f(R) Gravity Teleparallel

Higher-Order Symmetric Teleparallel

Horndeski Finsler Gravity

Gauss-Bonnet Non-Local Gravity

Bimetric Extensions of TEGR/STEGR

FIG. 75: A summary of the MG roadmap, outlining the potential extensions of GR.

first category, referred to as extended theories of gravity, includes extensions of the Einstein-Hilbert action by other
curvature terms (see Ref. [3061] for a review on the topic). The second category, referred to as alternative theories
of gravity consists of models modifying the basic principles of GR, such as the Equivalence principle or Lorentz
invariance. The main purpose of modified theories of gravity is to address GR shortcomings on different scales,
providing a different view to describe gravity, potentially able to fix issues related to cosmology [3062] and the quantum
formalism of the gravitational interaction [3050]. Extended and alternative theories of gravity naturally introduce extra
degrees of freedom with respect to GR, which might potentially accommodate cosmic tensions. Several promising MG
models involve modifications to the Einstein-Hilbert gravitational action, such as incorporating higher-order curvature
invariants [3063–3065], establishing connections between geometry and scalar fields [2062, 3066–3068], or introducing
additional geometric features to spacetime beyond curvature, such as torsion (Poincaré gauge gravity and teleparallel
gravity) [3069, 3070] or non-metricity (metric-affine gravity and symmetric teleparallel gravity) [3071, 3072], or non-
linear connections (Finsler gravity) [3073, 3074]. Different classes of MG theories are outlined in Fig. 75.

4.3.1.a. Extended theories of gravity One of the famous examples of GR extensions is the f (R) gravity models,
where the Ricci scalar R in the Einstein-Hilbert action is replaced by a general function f (R). This modification
leads to fourth-order field equations, which can explain cosmic acceleration without invoking DE [3022, 3075–3084].
Various formulations of this theory have been explored, some of which can produce deviations in the Newtonian
potential [3085], potentially addressing phenomena like the galaxy rotation curves without resorting to DM. Notably,
the Starobinsky model [2064], which introduces a quadratic term in the scalar curvature to explain cosmic inflation,
has attracted significant attention within this framework. Extra degrees of freedom in f (R) gravity affect the Hubble
function H(z), providing deviations from the ΛCDM paradigm. In the equivalent scalar-tensor representation in the
Jordan frame of f (R) gravity, in Ref. [3086] it is shown that the scalar field dynamics may lead to a definition of an
effective Hubble constant H0eff (z) that depends on the redshift z. In this regard, the non-minimal coupling between
the scalar field and the metric plays a crucial role. Notably, H0eff (z) may successfully address the Hubble tension since
it could reconcile measurements and estimates of H0 obtained at different redshifts, from local probes at 0 ≲ z ≲ 2 to
CMB at z ∼ 1100. Accordingly, the general scalar field potential is computed in Ref. [3086], as well as the respective
f (R) functional form in the low-redshift regime within f (R) quadratic gravity. Furthermore, Ref. [3087, 3088] shows
that a rescaling of the Universe’s expansion rate can emerge without presupposing a specific dependence on the redshift
(as done in Ref. [3086]). Indeed, it is just the non-minimally coupled scalar field of the scalar-tensor representation
responsible for the Hubble constant scaling. Then, the additional DDE source was implemented in the f (R) gravity
to provide a smooth transition between a quintessence regime and phantom fluids, resulting in an effective Hubble
constant that exhibits a plateau behavior for z ≳ 5 matching the Planck prediction. For equivalent approaches,
180

see also Refs. [3089–3091]. These scenarios with H0eff (z) were motivated by binned analysis of the SNIa Pantheon
sample and BAOs [515, 1663, 3092], pointing out a slowly decreasing trend of the Hubble constant. In Ref. [3093],
by exploiting the f (R) gravity correspondence with a scalar-tensor theory, the authors provide a condition in which
the H0 tension is alleviated. Specifically, this condition is based on the existence of a metastable de Sitter point that
occurs for redshifts near the recombination.
The Einstein-Hilbert action could be extended also by introducing dynamical scalar fields non-minimally coupled
with the geometry [3094]. Scalar-tensor theories are often taken into account to address the dynamics of the early
Universe, in the framework of the inflationary paradigm [3095]. In fact, inflation is usually conceived as generated
by a scalar field ϕ, called the inflaton, which is supposed to be responsible for the accelerating expansion of the
early Universe. According to this picture, inflation should be led by some scalar field driving the cosmic acceleration
between 10−34 and 10−35 s after the Big Bang, generating an isotropic and homogeneous Universe. The theory of
inflation describes also the production of particles after the early-time accelerating expansion (reheating) [3096–3098].
Moreover, to tackle the cosmological constant problem or the evolution of cosmological vacuum energy, new degrees
of freedom for the gravitational field have to be considered.
In this scenario, Horndeski theory represents the most general scalar-tensor theory with second-order field equations
in four dimensions [2061, 3099]. It encompasses a wide range of models and offers a comprehensive framework for
modifying the gravitational interaction by incorporating an extra degree of freedom through a scalar field [3100–
3102]. This results in significant and intriguing consequences for both inflation and DE physics. The detection of the
GW170817 event together with the electromagnetic counterpart [586] provides one of the most stringent constraints on
scalar-tensor theories that propose an unusual speed of GWs. This imposes severe limits on Galileons and extends to
other scalar-tensor theories, including the quartic and quintic sectors of Horndeski’s theory [3103, 3104]. Nevertheless,
it has been noted that since the energy scales observed at LIGO lie very close to the typical cutoff of these types of
theories, the translation of the bound on the speed of GWs to a cosmological setting might be non-trivial [3105].
Scalar-Tensor models have been taken into account also to address the H0 tension problem. For instance, in
Ref. [3106] Planck 2018 data are used to constrain the simplest models of scalar-tensor theories, accommodating
a higher value for H0 and therefore alleviating the tension between Planck/BAO and distance-ladder measurement
from SNIa data. The capability of scalar-tensor and bi-scalar–tensor modified theories to alleviate the tension is also
considered in Refs. [3107, 3108].
Even when limiting the analysis to Horndeski models that respect the GWs’ speed bound, observational constraints
on H0 using Planck data alone have been shown to yield values that are consistent with local determinations at 2σ for
the Galileon Ghost Condensate [3109] and 1σ for the Generalized Cubic Covariant Galileon [2893, 3110]. More recent
studies have also considered Horndeski gravity to specifically address cosmic tensions, showing that particular setups
free of ghost instabilities could lead to interesting late-time features possibly alleviating the H0 tension [2882, 3111].
Cosmic tensions have also been addressed in the context of the time-honored Brans-Dicke (BD) theory [3112, 3113],
which is the prototype of the large family of scalar-tensor theories (see also Ref. [2062]). BD gravity embodies the
first historical attempt at incorporating a dynamical gravitational coupling to GR. It also constitutes the simplest
scalar-tensor theory which can be embedded within the much broader class of Horndeski models [2061] and, in fact,
many of these models reduce to BD at cosmological scales [2405]. Apart from the metric tensor, BD gravity involves
a new (scalar) degree of freedom, φ, which controls the strength (and the possible time evolution) of the gravitational
interaction Gφ = G/φ, with G being the (locally measured) Newton constant transforming the gravitational coupling
constant into a new dynamic variable [3114]. The scalar field φ is non-minimally coupled to gravity via the Ricci scalar
and its dynamics is controlled by a dimensionless parameter ωBD , the BD parameter (or equivalently ϵBD ≡ 1/ωBD ).
Let us denote by BD-ΛCDM the Brans-Dicke counterpart of the standard ΛCDM, which represents the BD model
with a cosmological constant, Λ. The cosmological term Λ is actually not present in the original BD theory, but in
modern times Λ is obviously needed to trigger the late-time observed acceleration of the Universe. The BD-ΛCDM
extension of the original BD model has been recently tested in the light of a large set of cosmological data of various
sorts [2408], such as SNIa+BAO+H(z)+CMB+LSS, where H(z) may include or not the SH0ES prior on H0 . If the
latter is included, one finds [2408] that the BD-ΛCDM model is remarkably successful in reducing the H0 and σ8
tensions to inconspicuous levels of ∼ 1.5σ (when the high-l polarization and lensing data from Planck 2018 are not
considered) [2408]. Notice that The RVM is mimicked by the BD-ΛCDM model, particularly the type-II RVM version
with variable Gφ [2406, 2800–2802]. The BD-ΛCDM turns out to fit the cosmological data significantly better than
the standard ΛCDM model, and this is confirmed by different statistical criteria (see Refs. [2407, 2408, 2800, 2802] for
details). On the other hand, adding a constant potential for the scalar field does not correspond to the usual vacuum
energy in BD theory. An explicit detailed theoretical and observational investigation of this BD extension of ΛCDM
model, studied in Ref. [3115], reveals that the model does not alleviate the tensions with no significant deviations
from ΛCDM.
Values of the effective cosmological gravitational coupling Gφ about ∼ 7 − 9% larger than G are preferred at ∼ 3σ
CL This is possible thanks to the fact that φ remains below 1 throughout the entire cosmic evolution, while preserving
181

the matter and radiation energy densities very close to the typical ΛCDM values. As a result, one finds higher H(z)
values during all the stages of the cosmic history without changing dramatically the abundances of the matter species
in the pre- and post-recombination epochs. The lowering of the sound horizon at the baryon-drag epoch, rd , is
accompanied by an increase of the Hubble parameter. This helps to alleviate the H0 tension in good agreement with
other relevant background datasets, such as e.g., BAO and SNIa. In order not to spoil the correct fit to the CMB
temperature data, the model is prone to yield values of the spectral index ns that are considerably larger (viz. closer
to 1 from below) than the standard ΛCDM model. This causes, however, no problem since the dynamics of φ can
compensate the changes in the matter power spectrum introduced by this fact (cf. Sec. 3 of Ref. [2408]). Remarkably,
the model is able to cut back the σ8 tension at the same time. The relief of this tension can be accomplished by
means of a negative value of ϵBD , or alternatively through a positive value, provided that sufficiently massive (light)
neutrinos are allowed [2408]. In both cases one finds |ϵBD | ≲ O(10−3 ). The BD parameter has a direct impact on the
LSS data through the linear perturbation equations. For instance, for ϵBD < 0 the friction term in the equation of the
matter density contrast grows, and the source term decreases. Both effects contribute in sync to the lowering of σ8 ,
thus providing a physical explanation for a possible solution to the growth tension. It goes without saying that the
compatibility between the larger cosmological value obtained for the gravitational coupling and the value measured
locally, G, should be possible provided one can find an appropriate screening mechanism. Different studies show that
these mechanisms are possible, although other works find that their implementation may not be so straightforward
[3116].
Another way to extend the GR action is to introduce higher-order curvature invariants [3117–3127]. This is the
case of Gauss-Bonnet gravity, where the action includes a specific combination of curvature invariants known as the
Gauss-Bonnet term. In this context, particularly intriguing are theories incorporating quadratic curvature terms,
as they stem from GR viewed as an Effective Field Theory aiming to construct effective models towards quantum
gravity. Notably, certain combinations of higher-order contractions of the Riemann tensor yield a topological surface
term in four dimensions, known as the Gauss-Bonnet topological scalar, G. In four dimensions, this scalar equates to
the Euler density, which, according to the generalized Gauss-Bonnet theorem, provides the Euler characteristic when
integrated over the manifold. The Gauss-Bonnet invariant is often utilized for its topological properties to simplify
dynamics. However, in four dimensions, to render its contribution non-trivial, it is typically coupled to a dynamical
scalar field [3128–3132] or incorporated as a function in the gravitational action [3133–3136]. This approach, where a
function f (G) is added to the scalar curvature, can mimic the behavior of a cosmological constant at late-times and, in
the limit f (G) = 0, GR is restored. Another avenue involves considering higher dimensions where G is non-trivial and
coupling the Gauss-Bonnet invariant to a constant diverging when D = 4. Gauss-Bonnet gravity has been studied to
address the H0 -tension [3137, 3138], finding that can be greatly resolved within 2σ level.
Bimetric gravity [3139] is a consistent theory of non-linearly interacting spin-2 fields, one massless and one massive,
which allows for a broad range of cosmological expansion histories and is compatible with cosmological as well as local
tests of gravity [3140–3142]. In its most general form, the theory features four additional parameters in addition to the
standard ΛCDM ones. There are no free functions, unlike several other MG theories. Studies based on a restricted
class of two-parameter models have shown that the value of H0 inferred from the inverse distance ladder is only
slightly increased compared to ΛCDM [2868]; however, such models do not comply with the constraints required for
a working screening mechanism [3141]. A more recent analysis [2893] based on more general three-parameter models
has shown that, using (model-independent) transverse 2D BAO data in combination with SN and CMB, results in
a value of H0 which is closer to the SH0ES value. On the other hand, using 3D BAO data produces a lower H0
value, suggesting a potential bias in the latter. A study of the S8 tension has not been performed at present, since a
framework for structure formation in bimetric gravity has not yet been developed.
Functors of actions theories (FAT), is part of the extensions of gravity theories, and they predict the existence
of actionions, i.e., the actionic fluctuations and field-particles, which are analogues of energetic fluctuations and
field particles. In light of deviations from DE equations of state, w0 ≃ −1.1, FAT predict the existence of actionic
fluctuations which are of the order of one tenth of the observed volume [3143, 3144].
4.3.1.b. Alternative theories of gravity The second class of modifications revises the basic foundations of GR, such
as the Equivalence Principle, metric compatibility, Lorentz invariance, etc.. One such approach involves considering a
more general geometric framework with an affine connection different from the Levi-Civita one, thereby introducing
both torsion and non-metricity into spacetime. Torsion is associated with the antisymmetric part of the connection,
while non-metricity arises from the non-vanishing covariant derivative of the metric tensor. Within this framework,
the Riemann tensor and its contractions in GR can be expressed in terms of torsion, non-metricity, or a combination
of both. The gravity formulations characterized by curvature, torsion, and non-metricity are dynamically equivalent,
differing only by a boundary term in their respective actions, reason for which they are often termed the “geometric
trinity of gravity” [3145]. This classical equivalence involving boundary terms that do not affect the field equations
(and are thus omitted from the action formulation), is viewed from the perspective of GR as motivated by the
possibility of formulating GR as a gauge theory [3069, 3070, 3146, 3147].
182

In this framework, a consistent gravitational theory can be formulated by promoting torsion over curvature as the
sole governing factor of spacetime, resulting in a theory that precisely mirrors the dynamics of GR. This theory, known
as the “Teleparallel Equivalent of General Relativity” (TEGR) [3148], has garnered significant attention in recent years,
undergoing extensive analysis [3069, 3070, 3147, 3149–3151]. TEGR presents a theoretical framework that interprets
gravity as a consequence of torsion within the fabric of spacetime. In this context, gravitational interactions are
described through a set of tetrad fields, also referred to as “vierbeins”, which form the basis for depicting spacetime
geometry. These tetrad fields define a torsion tensor, serving as the source of gravity in the theory and representing
the antisymmetric component of the Christoffel connection.
An extensively studied model alongside TEGR is the “Symmetric Teleparallel Equivalent of General Relativity”
(STEGR), which describes spacetime through non-metricity. Non-metricity allows for the consideration that spacetime
may not adhere to the metric compatibility condition, a fundamental assumption in GR. While torsion arises directly
from the antisymmetry of the affine connection, non-metricity emerges when the covariant derivative of the metric
tensor is non-zero, denoted as ∇σ gµν ̸= 0. As consisting of field equations entirely equivalent to those of GR, both
TEGR and STEGR fall short of addressing the limitations GR imposes on larger scales. Consequently, similar to
f (R) gravity within the metric formalism, modifications to the Lagrangian density of TEGR can be explored in
various ways [3152], including introducing new torsion [3153, 3154] and non-metricity invariants [3155], coupling these
to additional scalar or axion fields [3156] or introducing an arbitrary function of the torsion scalar, leading to f (T )
gravity [3157, 3158]. The latter has been proposed as a potential solution to late-time cosmological issues, such as
the Universe’s accelerated expansion [3159, 3160]. The characteristics of gravity theories including torsion and non-
metricity, such as f (T ), f (Q) (with T being the torsion scalar and Q the non-metricity scalar) gravity, new general
relativity and many more, are currently under investigation, particularly concerning their applications in cosmology
and astrophysics [2116, 3161–3183]. Specifically, in the context of cosmological tensions, Refs. [2164, 2165, 3184]
explores methods to address the H0 tension within f (T ) models; in Refs. [3185, 3186] it is shown how to alleviate
both the H0 and σ8 tensions simultaneously within torsional gravity from the perspective of effective field theory; in
Ref. [3187] the evolution of scalar perturbations in f (T ) gravity and its effects on the CMB anisotropy is evaluated
to show that f (T ) models do not provide tension on the Hubble constant that prevails in the ΛCDM cosmology; in
Ref. [3188] the authors demonstrate that the f (T ) models prefer a higher value of H0 with respect to the Planck
prediction, in better agreement with local estimates. Similar effects can be achieved in some classes of the beyond-
generalized Proca model that, at the background level, resembles f (T ) theory [3189]. Finally, in Refs. [3190–3192]
solutions to the tensions were proposed in the framework of f (Q) gravity.
In another class of these models, Finsler gravity [3073, 3193–3197] offers an intriguing new way to derive the
gravitational field of many particle systems, when they are modeled as a kinetic gas, instead of as a (perfect) fluid
[3198, 3199]. The dynamics of a kinetic gas is described by a 1-particle distribution function (1PDF) on the 1-particle
phase space of spacetime, which takes the velocity distribution of the gas particles into account. Usually, when the
gravitational field of a kinetic gas is derived, one obtains the energy-momentum tensor of the gas by averaging the
1PDF over the velocity distribution of the particles. By means of this averaging procedure, the information about the
velocity distribution of the gas particles are lost and the gravitational field is derived from an effective fluid energy
momentum-tensor as a source term in the Einstein equations. Finsler spacetime geometry describes gravity in terms
of a curved 1-particle phase space of a curved spacetime (technically the tangent or co-tangent bundle). Here, the
1PDF directly sources the gravitational field of the kinetic gas without losing information through velocity averaging.
The contribution of the velocity (kinetic energy) distribution of the gas to its gravitational field is fully taken into
account in Finslerian gravity, rather than being averaged away [3198, 3199]. The first promising solutions of Finsler
gravity in the cosmological context have been obtained [3200], which show a linearly expanding Universe as a vacuum
background solution. Further studies are ongoing to demonstrate that the contribution of the kinetic energy of a
many-particle system to its gravitational field can be the solution to the Hubble tension and our understanding of
DE.
It is worth pointing out that some alternative theories propose that gravity is mediated by a graviton with a small
non-zero mass mg . Experimental limits on mg are set using models such as the Yukawa potential and dispersion
relations, tested on astrophysical and cosmological scales. In addition to detectors like LIGO/Virgo or LISA, graviton
mass limits can be derived from electromagnetic observations of gravitational systems, such as constraints on S-stars
orbits [3201, 3202], or from Schwarzschild precession of S2 in Yukawa gravity [3203], consistent with LIGO data.
Some theories of gravity relax the assumption of nonminimal coupling between the geometry and matter sources.
This happens, for instance, in f (R, Lm ) [368, 3204], f (R, T ) [3205], f (R, Tµν T µν ) [3206–3209], and f (R, T , Rµν T µν )
[3210, 3211] theories of gravity, where Lm is the matter Lagrangian density, and T the trace of the energy-momentum
tensor Tµν (in its standard definition). Additionally, one can construct similar nonminimal coupled theories in the
torsional framework, such as in f (T, Lm ) [3212, 3213] and in f (T, T ) [3214] theories of gravity, where T is the torsion
scalar.
A characteristic of these theories with nonminimal geometry-matter coupling is that the covariant divergence of the
183

matter energy-momentum tensor can in general be nonzero (though the standard continuity equation can be recovered
in particular models). A new avenue has been recently opened by modifying the introduction of the material source
in the usual EH action, such as matter-type modified theories of gravity with f (Lm , T , Tµν T µν ), since these theories
are equivalent to nonminimal interaction models in GR [3215]. Depending on the form of interaction determined by
f function, some nontrivial dynamics easing cosmological tensions [2151], difficult to achieve via simple interaction
kernels can be achieved, these DE models have effects on early and late-times of the Universe and hence may address
the current tensions within ΛCDM model. The extra two degrees of freedom have the potential to offer more room
to accommodate cosmic tensions. However, this also provides a challenge, given that Einstein-Boltzmann codes are
commonly prepared to deal with at most one extra degree of freedom, it is, at the moment, more challenging to
compute the evolution of perturbations and constrain these theories.
So far, studies have focused on the simplest classes of models. An example of studying dynamics of scalar pertur-
bations, using the quasistatic approximation, is in Ref. [3216] for f (R, T ) = R + f2 (T ) models with f2 (T ) ∝ T 1/2 to
ensure the usual matter continuity equation. This work finds that there is a strong scale k dependence of the matter
perturbations, not supported by observations.
A more recent work Ref. [3217] reports on the possibility of f (R, T ) = R + λT alleviating the σ8 tension, while
increasing the H0 tension. In the Hu–Sawicki model of f (R) gravity the σ8 -tension observations worsen [2825], while
it might be alleviated in viscosity in the DM model (in modified cosmological models, massive neutrinos suppress the
matter power spectrum on the small length scales, which implies that the bounds on neutrino mass get modified too)
[3218–3221].
To conclude, it is worth noticing that most alternative/extended theories of gravity may be subject to screening
mechanisms, such as the chameleon mechanism [3222, 3223], which can be encountered in several scalar-tensor theories
[3224, 3225] as well as in MG models [3226–3230] that can be recast in terms of an additional scalar field (e.g.,
see Ref. [3231]), like f (R) gravity. In this case, the scalar field is coupled to the matter fields so that its mass
is environment-dependent. As long as the matter density is high enough, the scalar field acquires a heavy mass
around the potential minimum, strongly reducing the range of the fifth force it mediates, and making it essentially
unobservable, see however [3232, 3233].
On the contrary, on cosmological scales, far from regions with higher densities, the scalar field has a lighter mass, ef-
fectively mediating the additional gravitational interaction. Another screening mechanism is the symmetron screening
[3234], where the potential of the scalar field V (ϕ) breaks the parity symmetry in low-density regions, which instead
is restored at smaller scales. In this mechanism, the scalar field has a vacuum expectation value (VEV) that depends
on local matter density, becoming larger in low-density regions and smaller in high-density regions. Moreover, the
coupling of the scalar field to matter is proportional to the VEV, so that the scalar couples with gravitational strength
in the low-mass-density regions, while it is decoupled, and then screened, in the high-density ones. In Ref. [3235] it
has been concluded that symmetron screening cannot be used to alleviate the Hubble tension.
Finally, the Vainshtein mechanism [3236] is important for some MG models, such as massive gravity, Dvali–Gabadadze–Porra
(DGP) or Galileon theory. The Vainshtein mechanism manifests itself at non-linear scales, below a certain radius
known as the Vainshtein radius. Nearby massive bodies GR is recovered through a strong kinetic self-coupling,
weakening the interaction with matter and reducing the propagation of the extra degree of freedom. Recently, it was
shown that bimetric gravity can alleviate the H0 tension (refer to Sec. 4.2), and could potentially alleviate also the
S8 tension [2893].60
According to the latest observations, one of the most promising extended gravity models is the generalized cubic
−1
covariant Galileon model. Using only CMB data, provides H0 = 72+8 −5 km s Mpc−1 , σ8 = 0.88+0.07
−0.05 at a 95%
confidence level [3110] and S8 = 0.84 ± 0.10. At the same time, the tensions are restored by combining the CMB data
with external datasets.
In the framework of alternative theories, particularly encouraging is the ∆4 model, which is a specific non-local
gravity model providing non-local quantum corrections to standard GR. This model predicts the values H0 =
70.3 ± 0.9 km s−1 Mpc−1 and σ8 = 0.82 ± 0.02 [3239], using the CMB, BAO and SN datasets, which is higher than
Planck predictions, in better agreement with the local estimates. The related S8 value is S8 = 0.81 ± 0.02.
All in all, although there is not a universally accepted MG model that resolves the cosmological tensions, many
of these models offer promising aspects that could help guide future research. Most of the models considered in
this section provides a higher value for H0 , though implications on σ8 should be further explored. As cosmological
data continue to improve, these theories will be subject to increasingly stringent tests, potentially leading to a
deeper understanding of the fundamental nature of gravity and the Universe. Therefore, it is quite possible that
extended/alternative theories could contribute, at least in part, to addressing the H0 and σ8 tensions.

60 However, it must be noted that drawing accurate predictions for the formation of LSS in bimetric gravity is challenging, so at present
the S8 tension can only be addressed under some simplifying assumptions. Regarding the H0 tension the alleviation comes with the
drawback that more parameters are introduced and bimetric gravity is preferred by AIC [3237] but ΛCDM is preferred by BIC [3238]
with the latter giving a stronger penalty for introducing more parameters to the model in consideration.
184

4.3.2. Early modifications to gravity


Coordinator: Emmanuel N. Saridakis, Konstantinos F. Dialektopoulos
Contributors: Andrzej Borowiec, Alfio Maurizio Bonanno, Ali Övgün, Andreas Lymperis, Andreas Papatriantafyl-
lou, Anil Kumar Yadav, Antonio Racioppi, Athanasios Bakopoulos, Athanasios Chatzistavrakidis, Branko Dragovic,
Carlos G. Boiza, Charalampos Tzerefos, Chiara De Leo, Christian Pfeifer, Cláudio Gomes, Damianos Iosifidis, Daniele
Oriti, David Benisty, Davide Pedrotti, Despoina Totolou, Diego Rubiera-Garcia, Elias Vagenas, Elisa Fazzari, Elvis
Baraković, Flavio Bombacigno, Fotios K. Anagnostopoulos, Francisco S. N. Lobo, Gaetano Lambiase, Genly Leon,
Gerasimos Kouniatalis, Giannis Papagiannopoulos, Giovanni Otalora, Giulia De Somma, Giuseppe Gaetano Luciano,
Gonzalo J. Olmo, José Pedro Mimoso, Jurgen Mifsud, Kathleen Sammut, László Árpád Gergely, Laur Järv, Manolis
Plionis, Manuel Gonzalez-Espinoza, Manuel Hohmann, Margus Saal, Maria Petronikolou, Miguel A. S. Pinto, Niko-
laos E. Mavromatos, Nikolaos Petropoulos, Oem Trivedi, Özgür Akarsu, Petros Asimakis, Saeed Rastgoo, Salvatore
Capozziello, Saurya Das, Sebastian Bahamonde, Simony Santos da Costa, Spyros Basilakos, Stylianos A. Tsilioukas,
Supriya Pan, Thanasis Karakasis, Tiago B. Gonçalves, Tomi Koivisto, Vasilios Zarikas, Vedad Pasic, Xin Ren, Yuejia
Zhai, and Yu-Min Hu

MG is a quite general framework that can offer alleviation to the tensions, since apart from late-time solutions
(which were analyzed in the previous subsection) one can apply it at early-times too [3240]. In order to solve the
Hubble tension, it is possible to introduce modifications to the standard cosmological model. The general idea is that
new assumptions at pre- and post-recombination are needed. For this reason, changing some of these assumptions,
i.e., changing the cosmological model before or after the recombination epoch, maybe the solution to obtain a different
value of H0 . One possible solution is to consider a MG scenario before the recombination epoch. Since the BAO data
constrain Hrd , the final goal is to obtain a lower value of the sound horizon at the drag epoch rd and thus a higher
value of H0 . Some constraints from models adhering´ ∞ to this approach are shown in Table VI.
The expression of the sound horizon is rd = zd dzcs (z)/H(z), where at sufficiently high redshift the sound speed

cs (z) can be considered almost constant and equal to c/ 3. To decrease the value of rd we need that, in a certain
redshift interval z ∈ (z1 , z2 ), the expansion of the Universe is higher with respect to the ΛCDM case in the same
value. An example of this approach can be found in Ref. [3241], where it is shown that the Hubble tension can
be reduced by assuming a MG model acting at early-times with a scalar field non-minimally coupled to the Ricci
scalar. Furthermore, in Ref. [3242], they explore the possibility of obtaining a lower value of the sound horizon, and
consequently a higher value of H0 , by using a MG model that allows for a phenomenological shift in the effective Planck
mass on cosmological scales before recombination, through a non-minimal coupling of a scalar field with gravity.
In Refs. [2459, 2461, 2467] it was proposed that the dynamics of the early Universe is described by a Chern-
Simons (CS) gravitational model, characterized by gravitational anomalies, which arises in the low-energy limit of
string theory [2458, 2468], after appropriate compactification to (3+1) spacetime dimensions. The model involves
the graviton and Kalb-Ramond (KR), or gravitational, axion field (assuming a constant dilaton). The latter couples
to the anomalous terms. Chiral primordial GW can condense under some circumstances, leading to a non-trivial
condensate of the CS terms [2469, 3243, 3244], and hence a linear axion potential, resembling the axion monodromy
situation in string-theory models [2471]. Such linear-axion potentials lead to inflation [2469, 2804], which however
are of the so-called RVM type [2464]. This means that the corresponding vacuum energy density contains dominant
terms of order H 4 (where H is the Hubble parameter, which is almost constant during inflation).
It is such non-linear terms that drive the RVM inflation. In Ref. [2469], it was shown that the CS condensates
are metastable, in the sense that they contain imaginary parts [2470], which can determine the duration of the RVM
inflation. The model contains a pre-RVM-inflationary era characterized by stiff-KR-axion-matter dominance. It is
during such an epoch that chiral GWs are formed, by either non-spherically-symmetric mergers of primordial (rotating)
black holes, or collapses of domain walls, the latter arising, for instance, from dynamical breaking of supergravity [3245,
3246], which can characterize the early epochs of such string-inspired models after the Big Bang [2461]).
The detailed transition from such an era to the RVM inflation is discussed in Ref. [2469], using a dynamical system
approach. The model can also lead to post-inflationary late-eras DE modifications, involving vacuum energy terms
of the form H 2 ln(H), which are held responsible for the simultaneous alleviation of H0 and growth-of-structure
tensions [2472].
The simplest generalization of Einstein’s theory are the so-called f (R) theories, which can be further generalized
by including a non-minimal coupling between the matter Lagrangian and another generic function of the curvature
[368]
ˆ

S = d4 x −g (f1 (R) + f2 (R)L) .

This model gives rise to an extra force term in the geodesics for a perfect fluid. In fact, stemming from the Liouville
185

theorem in phase space, the flux density of particles is shown to be covariantly conserved, and the Boltzmann H-
theorem is still preserved as the entropy vector flux is a non-decreasing function of the spacetime coordinates likewise
in general relativity. Despite the distribution function being formally equivalent to Einstein theory for a homogeneous
and isotropic universe, some quantities differ, namely as the effects of the non-minimal coupling appear on the radiation
density evolution and on the matter Lagrangian choice, which is no longer degenerate, for instance Ref. [3247]. In
these theories, the baryon asymmetry is generated through an effective coupling between the Ricci scalar and the
net baryon current, which dynamically breaks CPT invariance, thus matching the observed asymmetry for very small
deviations from general relativity. Moreover, the resulting temperatures are compatible with the subsequent formation
of the primordial abundances of light elements [3248]. Hence, among others, this scenario allows for a good agreement
with SN distance data and the BAO data for different exponents in a power-law type non-minimal coupling function,
thus alleviating the Hubble tension [372].
A consistent gauge theory of gravity and spacetime based on the Lorentz symmetry was found rather recently
[3249–3251]. Addressing the problem of time, the theory predicts that the space that emerges via spontaneous
breaking of the Lorentz symmetry could be massive [3249, 3252, 3253]. This would be the minimal explanation of the
missing mass in the Universe, and a slight extension of the scenario might alleviate the H0 tension through a cosmic
spin as has been discussed also in the context of Poincaré models of gravity [3253–3256]. An ultraviolet completion
mentioned in Ref. [3257] leads to modifications of early cosmology, whereas the relevance of a different description
of nonperturbative effects to the cosmological tensions was speculated in Ref. [3258]. Ref. [3259] rewrites the higher
curvature theories using a gauge theory of gravity.
As a subcase of metric-affine and gauge theories of gravity, teleparallel gravity has also been well-studied in the
early Universe. The Teleparallel Equivalent of General Relativity (TEGR) describes gravitational interactions as the
torsion of the connection of spacetime. Several teleparallel inflationary scenarios have been proposed in the literature,
in order to change the early-Universe evolution and eventually alleviate the H0 tension, and here we briefly summarize
some of them. Born-Infeld inspired inflation originates from the work of Born and Infeld, who proposed a finite self-
energy for point-like particles to avoid divergences in physics. This approach has influenced various gravity theories,
including Eddington-inspired Born-Infeld gravity, which addresses singularities like black holes and the big bang. In
teleparallel cosmology, Born-Infeld inspired models use an f (T ) gravity Lagrangian, featuring a scaling parameter λBI
that induces inflation without an inflaton field [3159, 3260, 3261]. This model aligns with the CMB and transitions
to standard ΛCDM at later times, providing a comprehensive framework for early Universe dynamics [3262–3266].
In Loop Quantum Cosmology within the FLRW framework, the matter bounce scenario is described by a modified
Friedmann equation, resulting in a non-singular bounce at t = 0. This model can be reformulated in terms of
f (T ) gravity [3267–3275]. The Lagrangian can be adjusted to fit observations by varying the critical density ρcr .
Scalar fields, affecting spectral indices and power spectra, are crucial for matching CMB observations and ensuring
reheating. Perturbations reveal additional degrees of freedom, impacting the scalar and tensor power spectra, with
the tensor-scalar ratio potentially exceeding observed bounds, dependent on scalar field dynamics.
Higgs inflation explores the coupling of the Higgs field with gravitational theories like GR, TEGR, and STEGR.
The Higgs potential V (ϕ) = λ4 (ϕ2 − v 2 )2 can be coupled with the torsion scalar T to investigate inflationary scenarios.
In this context, a particular action is used, incorporating kinetic and potential functions [3276]. Some studies reveal
issues, such as a high tensor-to-scalar ratio or failing to achieve inflationary models. Despite these challenges, alter-
native scalar field couplings and potential functions may yield more practical inflationary models, suggesting further
exploration is necessary for natural inflation in teleparallel gravity.
When the canonical quantization techniques of Loop Quantum Gravity are applied to cosmological scenarios, one
finds that the big bang singularity is generically resolved by a big bounce [3277–3282]. This happens even when
different quantization prescriptions are used, which leads to various nonsingular models of loop quantum cosmologies
[3283–3287] (known as LQC, mLQC-I, and mLQC-II [3285]) that differ in their concrete pre-bounce and post-bounce
dynamics. In Ref. [3288], it has been found that the effective dynamics of those three quantum-corrected models can
be accurately approximated by a 3-parameter family of metric-affine f (R) theories of gravity [3289], where two of
the parameters can be set by constraints at the bounce and at low curvatures (GR limit), while the third one can
be determined (numerically) by fitting the background evolution. Interestingly, the non-perturbative dynamics of all
three models is dominated by a logarithmic correction [3290]. Additionally, the best fit value of the free parameter can
be well approximated by elementary combinations of the bounce density and the Barbero-Immirzi parameter. On the
other hand, it is possible to recover regular big-bounce scenarios also in the presence of a dynamical Barbero-Immirzi
field, when a purely torsional Nieh-Yan term on a FLRW background [3291, 3292] and its non-metrical generalization
in Bianchi-I cosmologies [3293] are taken into account. This idea naturally leads to projective-invariant metric-affine
formulations of Chern-Simons MG, for which big bounce and de Sitter solutions arise [3294]. A peculiar aspect
of these models is that they predict the coupling of metric tensor modes to torsion tensor components, leading to
torsional birefringence and the possibility of distinguishing the usual metric version of Chern-Simons gravity from its
metric-affine counterpart via the quasinormal mode emission of compact objects [3295] (see also Ref. [3296]).
186

Axions and Axion-Like Particles (ALPs) provide well-motivated candidates to address important problems in cos-
mology, maintaining at the same time a significant discovery potential in current and future experiments. Among
their uses, they can serve as candidates for DE within the quintessence scenario [3297, 3298]. ALPs arise naturally in
string-inspired models [2468], often not single but as a plethora of light pseudoscalar fields [3299]. Their characteristic
sinusoidal potential generates models of EDE that can potentially resolve the Hubble tension [2442, 2443, 3300].
Remarkably, although these models are loosely inspired by string theory, recent attempts highlight the potential
of embedding them into stable type IIB string compactifications [2449, 2450], which would provide a fundamental
starting point. The appearance of multiple ALPs in stringy scenarios allows for alignment mechanisms that combine
them and can lead to models where at different stages of the cosmological evolution different ALPs act as inflaton,
QCD axion, and quintessence [3301, 3302]. Pseudoscalar fields were also considered in the context of teleparallel
and symmetric teleparallel gravity, where they couple to the CP-violating quadratic terms that can be formed with
the torsion tensor or/and the non-metricity tensor [3156, 3303, 3304]. Such additional fields could contribute to the
cosmological dynamics in the early Universe [3156, 3305] and can also lead to GW velocity birefringence [3306, 3307],
which can be tested using multi-messenger approaches [3308].
DM creation during or after inflation can be associated with irreversible thermodynamic processes [3309–3311].
Refs. [3312, 3313] proposed an alternative cosmology based on the irreversible thermodynamics of open systems,
in which the explanation for the production of macroscopic matter and entropy in the early Universe relies on a
reinterpretation of the matter energy-momentum tensor that includes an irreversible creation term. However, this
contrasts with the covariant conservation of the energy-momentum tensor in GR.
In the last decade, several MG theories that contain nonminimal couplings between geometry and matter have been
proposed, such as f (R, Lm ) [3204], f (R, T ) [3205], f (R, Tµν T µν ) [3206–3209] and f (R, T, Rµν T µν ) [3210] theories of
gravity, where R and Rµν are the Ricci scalar and tensor, Lm the matter Lagrangian density, and T the trace of the
energy-momentum tensor Tµν . A property of all these theories is that the matter energy-momentum tensor is not
conserved. This feature allowed Harko [3314] to physically interpret such non-conservation as an irreversible energy
flow from the gravitational field to the matter sector that could result in particle creation, by using the thermodynamics
of open systems approach. The effects and implications of the irreversible matter creation processes on the late
cosmological evolution have been studied on some of these modified theories of gravity [3314–3317] (see Ref. [3318]
for a review). It was shown in Ref. [3215] that if such theories include the Ricci scalar R in the Einstein–Hilbert form,
i.e., assuming minimal geometry-matter coupling—they are effectively equivalent to nonminimal interaction models
within the framework of GR. These MG theories could be further explored in the context of the early Universe DM
creation that can eventually lead to alleviation of the H0 tension.
ΛCDM paradigm is very efficient in describing our Universe; however, it is broadly acknowledged that the model
ought to break down and exhibit limitations in at least the two extreme phases of the Universe, namely its very
early and very late phases. In the very early Universe, Planck scale or quantum gravity effects are expected to set
in, resulting in potentially significant conclusions, some of which may be measurable. While proposals for quantum
gravity abound, and one can in principle take any theory and examine its implications for the above stages of the
Universe, here we will primarily be interested in a virtual model-independent prediction of theories of quantum gravity,
namely the Generalized Uncertainty Principle (GUP). The GUP predicts significant changes to several thermodynamic
quantities, while it leaves some others intact [3319–3322].
A generic and often-used version of GUP is
   
pi pj 2 2
(4.18)

[xi , pj ] = iℏ 1 − α pδij + + α p δij + 3pi pj
p
ℏ
1 − 2α⟨p⟩ + 4α2 ⟨p⟩2 , (4.19)

∆x∆p ≥
2

where the dimensional parameter α can be traded off with a dimensional one α0 , via α = α0 ℓP l /ℏ , where ℓP l ≃ 10−35 m
is the standard Planck length and 1 ≤ α0 ≤ 1015 , the upper bound dictated by the fact that no Planck scale effects
have been observed at the LHC, that has a characteristic length scale of 10−20 m ≃ 1015 ℓP l (corresponding to 10 TeV
of energy), which the “new length scale” α0 ℓP l should not exceed.
It follows that quantum gravity effects can in principle take effect right from the electroweak to the Planck scale.
Since our focus is in the very early Universe, we will be concerned near the Planck scale. As explained in Ref. [3319]
the standard Heisenberg Uncertainty Principle and the GUP can be translated into a corresponding time-energy
uncertainty relation. This, coupled with the first Friedmann equation, can in turn be used to estimate the Planck
energy, mass, density, time, temperature, and effective number density. Since the GUP terms can differ significantly
from the standard Heisenberg one near the Planck scale, it is not surprising that many of the quantities alluded to
above undergo significant changes in the early Universe. Perhaps the most significant among them is the Planck
energy density, whose ratio to the observed DE density increases by at least a factor of 4 to about 10119 .
187

Additionally, the same uncertainty principle can in fact affect the Friedmann equations and cosmological predictions
in the radiation dominated era, as shown in Refs. [3320, 3321]. There, an interaction term coupling baryon current and
space-time was used to satisfy the first two Sakharov conditions, which along with the modified Friedmann equations
break thermal equilibrium and thus satisfy the third and last Sakharov condition. The results can provide a rich
phenomenology.
In the GUP approach, where a modification to the canonical algebra of the system leads to minimal uncertainties in
the configuration and/or momenta, the evolution of the FLRW Universe with canonical variables (c, p) in the presence
of scalar matter field ϕ with its momentum pϕ , can be describedby the action of the GUP modified Wheeler-deWitt
equation on the wave function of the Universe Ψ, ∂ϕ2 + Θ b GUP Ψ = 0, where Θ b GUP is a self-adjoint operator. In
one approach [3323], which is equivalent to a cut-off in the length, [q, p] = i 1 + βp2 , using suitable wave-packets

for the deep Planckian regime, it is found that the wave-packets do not fall in the classical singularity and the GUP
Universe exhibit a stationary behavior in approaching the Planckian region. This implies that no bounce is present
in this model. In another approach using GUP-modified Friedmann equations [3324], it is found that the DE density
scales as α̃H 4 , where α̃ is a GUP parameter associated to the minimum length, and H is the Hubble parameter. This
means that this approach to GUP cannot explain the acceleration of the present Universe, since the energy density
decreases very quickly. It is worth mentioning that GUP comes in various forms and a recent proposal based on a
different modification of algebra (minimal momenta) [3325] together with an improved prescription may be a better
approach. The reason is that 1) the momenta of gravitational field are related to the metric or triads so it is more
natural to have minimal values in those quantities, and 2) without an improved scheme, just like LQC, one might
obtain incorrect classical or asymptotic limits and hence the theory might not also be correct in the Planckian regime.
The Asymptotic Safety approach to Quantum Gravity offers a potential solution to the problem of quantizing the
gravitational field. This approach is grounded in the existence of an ultraviolet non-Gaussian fixed point for the
√ √
couplings of the gR and g operators in the Euclidean theory [3326]. A direct consequence of this fixed point is
the dynamic suppression of Newton constant G at very high energies, which has significant implications for early
cosmological models. In Ref. [3327], a comprehensive cosmological evolution from the inflationary phase to a late
de Sitter phase is discussed. Additionally, [3328] proposes the idea that the Universe originates from a state of zero
entropy and evolves towards an accelerated expansion phase.
Following Ref. [3329], one can also expect various modifications to the Effective Action at the inflationary scale. In
R2
particular, it has been shown in Ref. [3330] that an f (R) Lagrangian of the form L = R + αR3/2 + 6m 2 − Λ represents

a modification of the well-known Starobinsky quadratic Lagrangian. This modification arises due to the presence
of higher curvature relevant operators generated by the renormalized flow near the non-Gaussian fixed point. The
predicted scalar-to-tensor ratio r is significantly higher than in the standard Starobinsky model, making it a potential
target for testing in future CMB missions like LiteBIRD.
The cosmological evolution of the early Universe is tested at temperature T ≲ 1 MeV through the prediction of
BBN. For temperatures T ≫ 1 MeV, it is not excluded that new physics occurred, offering the possibility that
non-standard cosmologies might have played a relevant role in the interplay between particle physics and cosmology.
The Ice Cube experiment, for example, has revealed a high-energy astrophysical neutrino flux, with energies ∼ O(1)
PeV [3331, 3332]. Among the various scenarios to explain this phenomenon, it has been proposed the minimal model
of DM decay [3333]. This interaction is also able to generate the correct abundance of DM in the Universe [3333]. In
the standard cosmological model, the rate of DM decay, needed to obtain the correct DM relic abundance, differs by
several orders of magnitude as compared with that needed to explain the Ice Cube data, making the four-dimensional
operator unsuitable. However, such a discrepancy can be reconciled in modified cosmologies, so that the Ice Cube
neutrino rate and DM relic density can be consistently explained [3334–3336].
In Fig. 76 we summarize the various early modified gravity theories that have been proposed to alleviate the
tensions, while some corresponding predictions for H0 and σ8 are summarized in Table VI.

TABLE VI: Early modifications to gravity: predictions.

Model H0 σ8
Early MG and the sound horizon [2433] > 70 < 0.8
String-inspired Chern-Simons modifications [2472] 71.27+0.76 +0.017
−0.73 0.816−0.015
Non-minimally coupled models [372] +0.52
70.59−0.53 -
Axions and Axion-like particles [3308] 70.6 ± 1.3 -
188

FIG. 76: Possible early modifications to gravity.

4.4. Matter sector solutions


4.4.1. Cold dark matter
Coordinator: Luca Visinelli
Contributors: Andrzej Borowiec, Anil Kumar Yadav, Branko Dragovic, Davide Pedrotti, Goran S. Djordjevic,
Ioannis D. Gialamas, Ippocratis Saltas, Ivan De Martino, Janusz Gluza, Jenny G. Sorce, Jenny Wagner, Krishna
Naidoo, Marcin Postolak, Martti Raidal, Mehmet Demirci, Paolo Salucci, Pran Nath, Reggie C. Pantig, Riccardo
Della Monica, Rishav Roshan, Sebastian Trojanowski, Torsten Bringmann, Venus Keus, and Wojciech Hellwing

4.4.1.a. Motivation and evidences The circular velocity of stars in a spiral galaxy can be inferred from the
measurements of the Doppler shift of atomic lines at different distances R from the galactic center, leading to a
velocity distribution v = v(R) known as the rotation curve. The distribution of the luminous matter in virialized
objects does not match that of the gravitational matter, demanding an additional non-luminous component.
Historically, this “missing mass problem” in the late Universe led to the proposition that there is a yet undetected
form of matter [3337–3341]. Subsequently, CMB observations revealed a missing mass in the power spectrum already
at redshift z = 1100, which could also be attributed to the same form of matter we call DM today. Evidence for the
presence of DM in the Universe span from galactic (rotation curves of spiral galaxies) to extra-galactic (clusters of
galaxies, weak gravitational lensing observations) to cosmological scales (LSS, acoustic peaks in the CMB, BBN, non-
linear growth of inhomogeneities. See Ref. [3342] for a historic reconstruction and Refs. [358, 3343–3347] for reviews.
N-body simulations with DM only reproduce well the LSS [2343, 3057, 3348] and are advancing to incorporate the
essential gastrophysics effects [2268, 3349].
The role of DM in addressing cosmic tensions is of profound interest from both theoretical and observational
perspectives. If discovered, a DM particle would be a relic from a period before BBN, for which we currently have
no direct data. The properties of DM—such as its clustering power spectrum, free-streaming velocity, and particle
characteristics—would enable us to better model the pre-BBN epoch and differentiate between various “early Universe”
proposals aimed at resolving these tensions. Furthermore, some theories that address the H0 tension explicitly involve
189

1030
1025
1020
1015
1010
105
100
10-5
10-10
10-15
10-20
10-25
10-30
- 40 - 30 - 20 - 10 0 10 20 30 40 50

FIG. 77: An estimate of the scattering cross section for some DM candidates, as a function of the DM mass. The
plot is inspired from Refs. [3350–3352]. As a comparison, the SM neutrino is also included.

interactions between DM and other components, including DE. Pinning down the properties of DM is crucial for
understanding its interactions with other particles, including those in exotic DE models.
4.4.1.b. Theory The evidence supporting the existence of non-baryonic DM is compelling across all observed
astrophysical scales. While alternative theories are still under consideration, the notion of CDM has emerged as
the dominant framework. DM might exist as macroscopic objects or as a particle, either fundamental or composite,
proposed in extensions beyond the SM framework. A non-exhaustive summary of the properties (scattering cross
section and target mass) of some candidates is provided in Fig. 77, inspired by previous work in Refs. [3350–3352].
Macroscopic DM One of the earliest proposition on the nature of DM is that it could appear in the form of
known astrophysical objects of mass M such as planets or stellar black holes [3353, 3354]. This conjecture has been
challenged by gravitational microlensing surveys, which collectively exclude these objects as the main DM component
over the mass range 10−11 ≲ M/M⊙ ≲ 10 [3355–3358].
Along with known astrophysical objects, massive DM candidates consist of primordial black holes which are pro-
duced in the early Universe by a plethora of mechanisms at different mass scales. Stable primordial black holes possess
a mass M ≳ 5 × 1014 g so that their Page time [3359] exceeds the age of the Universe, making them suitable DM
candidates. A more stringent bound M ≳ 1017 g is obtained from the non-observation of Hawking emission in the
extragalactic γ-ray flux and in X-ray detectors [3360–3363]. Lensing constraints also apply to primordial black holes,
leaving an open window of opportunity where primordial black holes are the DM in the mass range M ∈ [1017 -1022 ] g
under the assumption of a monochromatic mass function.
Particle DM An intriguing possibility is that DM results from a completely new sector of physics not com-
prised within the SM. This includes candidates such as sterile neutrinos, axions, supersymmetric particles (gravitinos,
neutralinos, axinos), weakly/feebly/strongly interacting massive particles, Q-balls, or ultralight scalars. Generally
speaking, DM particles should be i) non-relativistic at recombination, ii) neutral or feebly charged, iii) stable over
cosmological timescales, iv) feebly interacting with the SM. While the upper limit for the mass m of a particle can-
190

didate is the Planck mass, the lower limit for a fermionic species is provided by the Gunn-Tremaine condition [3364],
which reads m ≳ 100 eV from consideration on dwarf spheroidal galaxies [3365]. For a bosonic species, the lower mass
limit is often considered as the Hubble rate at matter-radiation equality, m ≳ 10−27 eV, from the consideration that
the ultralight field should oscillate by the time at which the matter component becomes relevant over radiation. Some
particle models that have been explored include the followings. Explaining the missing mass across cosmic times and
scales by a single extension of the known cosmological model has motivated the search for further corroborations of
its existence and led particle physicists to investigate potential particle candidates.
Flavor discrete symmetry models The SM fails to explain neutrino masses, mixings, and DM with the WIMP
paradigm comprehensively. Standing at this juncture, it is certainly a tempting challenge to find a common origin of
these two seemingly uncorrelated sectors and much effort goes beyond the SM of particle physics to explore scenarios
that can accommodate a candidate of DM and explain non-zero neutrino masses and mixing [3366]. Typically,
models based on non-Abelian discrete flavor or modular symmetries are considered to explain observed neutrino
mixing, and breaking of the same symmetry may lead to the residual symmetry, which at the same time stabilizes the
DM [3366, 3367]. For example, in models based on the A4 discrete group, light neutrino masses can be associated with
tree-level type-I seesaw mechanism and the one-loop contributions accommodating viable DM candidates responsible
for observed relic abundance of DM [3368]. In such models, constraints coming from the lepton flavor violating
processes are also important, restricting DM masses and leaving ranges of DM parameters that can be probed in
collider studies.
Supersymmetry Since B and L-violating operators in the SM have dimension d ≥ 5, the renormalizable La-
grangian possesses an accidental global B-L symmetry [3369–3372], which is imposed in supersymmetric extensions of
the SM to avoid experimental consequences such as a large proton decay rate. Such a symmetry leads to an R-parity
which is conserved in scattering and decay processes, leading to supersymmetric particles to be produced in pairs.
The same R-parity leads to the stability of the lightest supersymmetric particle, which could be a DM candidate if it
is neutral. Supersymmetry, first postulated to address the hierarchy problem, has thus long provided us with a long
list of DM candidates [3373]. These include neutralinos [3374–3380], sneutrinos [3381, 3382], gravitinos [3383–3393],
and axinos [3394, 3395], massive spin-2 particles [3396–3400] resulting from modifications of gravity, and Kaluza-Klein
excitations [3401–3403] found in theories incorporating extra dimensions. However, some of the simplest versions of
the theory are under stress due to the non-observation of any supersymmetric particles in collider experiments. The
lightest neutralino, among others, is an example of a much broader category of particle DM candidates dubbed weakly
interacting massive particles (WIMPs), produced thermally in the early Universe.
Weakly interacting massive particles (WIMPs) WIMPs, some of the most studied DM candidates, are
hypothesized to have been produced in the early Universe as they were initially in thermal equilibrium with the
primordial bath. As the Universe expanded and cooled, the production of WIMPs ceased and they “froze out” of
thermal equilibrium. The freeze-out temperature is defined as the point when the Hubble expansion rate exceeds the
WIMP annihilation rate. The leftover density of WIMPs after freeze-out, i.e., their relic density, depends on their
annihilation cross-section. The annihilation cross-section needed for WIMPs to achieve the observed DM density
today is naturally of the same order of magnitude as the weak nuclear force cross-section. This coincidence is known
as the “WIMP miracle” because it suggests that WIMPs could naturally account for DM without needing highly
tuned parameters. Since the WIMP miracle provides a compelling argument for them as the leading DM candidate,
many direct and indirect detection experiments are built to probe the typical WIMP mass range of ∼ O(GeV) to
∼ O(100 GeV) [3343, 3344]. Even though direct and indirect experiments continue to constrain the parameter space
for the vanilla WIMPs scenario - as for most DM models - exciting new directions in the WIMP paradigm, such as
CP-violating DM [3404], multi-component DM [3405], pseudo-Goldstone DM [3406], have been proposed which not
only provide viable DM candidates but also address other shortcomings of the SM [3407]. Even if the WIMP miracle
is not realized, loopholes exist such as coannihilation [3374, 3408, 3409], annihilation to slightly heavier states [3408],
p-wave annihilation, resonances [3408, 3410], and Sommerfeld enhancement [3411, 3412].
Sub-GeV thermal DM The WIMP paradigm can be generalized to lower DM masses and coupling constants,
and various sub-GeV DM candidates have been proposed that can also be produced thermally in the early Uni-
verse [3413–3415]. In this case, their mass is typically bounded from below to m > O(MeV) by astrophysical con-
straints and BBN, while even stronger indirect detection bounds are present for light DM candidates annihilating
to SM species via s-wave channels, see the recent review in Ref. [3416]. In other models, however, dedicated di-
rect detection techniques are needed to probe them [3417], and multiple accelerator-based searches are ongoing and
planned [3418].
Sterile neutrinos A minimal extension of the SM with the addition of a Weyl fermion N as a sterile neutrino
could provide viable DM candidates. This particle couples to the SM through a Yukawa term such as ∼ yN LH, where
L is a left-handed lepton doublet and H is the SM Higgs doublet, which could give rise to the small neutrino masses
191

through the seesaw mechanism [3419–3422] and possible lead to baryogenesis via the leptogenesis mechanism [3423].
Sterile neutrinos of mass m < me might decay into three active neutrinos, thus constraining the sterile mixing angle
θ for a given m [3424]. The abundance of sterile neutrino DM is fixed through a freeze-in mechanism in which active
neutrinos oscillate into N before decoupling [3425]. Possible caveats include a finite lepton asymmetry [3426] or the
presence of self-interactions [3427–3429].
The QCD axion The QCD axion [3430, 3431] is a light pseudo-scalar coupled with the SM gluons, introduced
to solve the strong-CP problem through the Peccei-Quinn mechanism [3432]. The interaction with the gluon leads to
a mixing of the axion with the neutral pions and to a small mass ma ∼ 6 µeV(1012 GeV/fa ), where fa is the energy
scale at which the infrared description breaks. Because of its interaction with the gluon that induces a coupling
with the charged pions, this particle inevitably couples with the SM photon. Models that generalize the Peccei-
Quinn mechanism leading to a very light axion and a prediction for the axion-photon coupling gaγγ have been long
studied and constitute the benchmark for laboratory axion searches, such as the KSVZ [3433, 3434] and the DFSZ
axion [3435, 3436]. A nonrelativistic population of cosmic QCD axions is produced through the vacuum misalignment
mechanism [3437–3439], leading to a number density in axions with a large occupation number at the time Tosc ∼ GeV
at which coherent oscillations in the field begin. This population of axions can possibly address the DM abundance
for given values of the mass ma and of the initial value of the axion field θi fa [3440–3442]. The uncertainties in the
assessment of the axion parameters can be divided into cosmological [3443, 3444], particle content [3445, 3446], and
theoretical [3447], see Refs. [2579, 3448] for reviews. A robust experimental search is currently ongoing worldwide to
try and detect this particle [2584, 3449].
Ultralight bosons Ultralight bosons or “fuzzy” DM (FDM) with masses in the range m ∼ 10−23 -10−18 eV would
possess a de Broglie wavelength of astronomical size [2359, 3450, 3451]. An ultralight scalar, naturally generated in
string theory, is generally considered to be real or pseudo-scalar field minimally coupled to the metric, and acquires a
small mass from various UV mechanisms [3299, 3452]. When the Universe cools below a certain critical temperature,
the field starts to roll down and oscillates about the minimum of its potential [2578, 3453–3455]. N-body cosmo-
logical simulations have shown the formation of a DM core within each virialised halo, which is supported against
gravitational collapse by the internal quantum pressure due to the Heisenberg uncertainty principle on the de Broglie
scale, surrounded by an interference pattern representing the oscillation in the density and velocity fields of the DM
particles [2371, 3456–3460]. A lower bound on the boson mass is derived by considering the quantum pressure intro-
duced at cosmological scales, which suppresses the power spectrum of linear inhomogeneities at large wavelengths and
leads to a lower bound m ≳ 10−21 eV [1544, 1545, 1556, 3461]. Stronger constraints on the mass and the abundance
of ultralight scalars are placed from a variety of observations which involve detection strategies at astronomical or
cosmological scales [2324, 3346, 3462–3468], which could nevertheless depend on the details of cosmology and the
astrophysics involved. An issue with FDM candidates lies in that such particles cannot produce the very large regions
of about constant density observed at the center of large galaxies [1735, 3469].
4.4.1.c. Experimental efforts
Laboratory searches Given the vast landscape of particle DM candidates, several techniques have been imple-
mented for their searches. Here, we highlight the detection of WIMPs and axions in underground detectors.
If DM is in the form of WIMPs with a small coupling to SM nucleons, it is possible to detect a small event rate
R based on DM-nucleon elastic scattering, totaling R ∼ NT nDM σA v. Here, NT is the number of nucleon targets in
the experiment, nDM is the number density of DM in the Solar system, v ∼ 220 km/s is the relative velocity, and
σA is the scattering cross section off a nucleon of atomic number A. For spin-independent (SI) interactions, the DM
scattering cross section is often quoted in terms of σSI ≈ σA /A4 [3343, 3470]. For spin-dependent (SD) scatterings,
the enhancement with the target mass is more modest and reads σA ≈ σSD A2 (J + 1)/J, where J is the spin of
the nucleus. The most sensitive experiments include liquid xenon-based detectors: XENONnT [3471], LUX-ZEPLIN
(LZ) [3472], and PandaX-4T [3473], which all set a bound on the SI DM-nucleon cross section σSI ≲ 10−47 cm2 for
the DM mass m ∼ 40 GeV. The argon-based detector DarkSide-50 [3474] is the most sensitive in targeting the SI
cross section for WIMP masses below about 3 GeV, with proposed experiments including DarkSide-20k [3475] and
ArDM [3476]. Note, that the results depend on several assumptions on: i) the distribution of DM in the solar system
and around the Earth, ii) the universal coupling between the WIMP and the nucleons, iii) the mass range of the
mediator, iv) the knowledge of the nuclear form factors, v) the effective theory used to model the interaction, vi)
the absence of self-interactions, vii) the scattering process being elastic. All of these caveats have been thoroughly
investigated in the literature in relation to realistic DM models.
The laboratory search for the QCD axion and other light bosons proceeds by means of resonant cavities in a
“haloscope” [3477, 3478], light shining through wall experiments [3479–3482], dielectric haloscopes [3483], and DM
radio searches [3484]. In a resonant cavity, a galactic axion converts into a microwave photon through the interac-
tion with the virtual photon of an external magnetic field [3485–3491]. DM radio searches is employed in DMRa-
dio/ABRACADABRA [3492] and, for light axions not related to the QCD theory, with SuperMAG [3493]. Scalar
192

fields produced in the Sun are searched through helioscopes: CAST [3494] and IAXO [3495].
Indirect detection Indirect detection of DM involves observing the potential byproducts of DM interactions
rather than revealing a new particle. This method relies on DM particles annihilating or decaying into SM particles,
which are then detected in astronomical and terrestrial instruments. The byproducts of these interactions typically
include (1) gamma rays [3496, 3497], high-energy photons that can travel vast distances and are easier to trace back to
their source; (2) Neutrinos [3498–3500], which are nearly massless and interact very weakly with matter, making them
hard to detect but potentially useful for indirect detection; and (3) Charged cosmic rays [3501–3503] such as electrons,
positrons, protons, and antiprotons. An excess of these particles in cosmic rays could indicate DM interactions [3504].
Different detectors and observatories are used to identify the potential signals from DM interactions. Instruments
like the Fermi Gamma-ray Space Telescope observe the sky for gamma-ray emissions, focusing on high DM density
regions like the Galactic Center or dwarf spheroidal and dwarf spirals [3505–3509]. Facilities such as IceCube in
Antarctica detect neutrinos by observing the Cherenkov radiation produced when neutrinos interact with ice [3510,
3511]. Instruments like the Alpha Magnetic Spectrometer (AMS-02) on the International Space Station measure the
flux of cosmic rays, looking for anomalies that could indicate DM annihilation or decay [3512].
One key challenge for indirect detection is to distinguish a potential DM signals from astrophysical backgrounds.
For instance, gamma rays can be produced by various astrophysical sources like pulsars, supernova remnants, or AGN,
while neutrinos are produced in the Sun and in the atmosphere. Cosmic rays have numerous sources, including the
Sun and SN, making it difficult to attribute any observed excess to DM. For this, fingerprints are searched through
cross-correlation studies using data from different wavelengths and types of particles to identify coincident excesses
that could point at DM, and spectrum analysis where the energy spectrum of detected particles is analyzed to identify
features characteristic of DM annihilation or decay.
Currently, there is no definitive detection of DM through indirect methods, but some intriguing signals and excesses
warrant further investigation. For instance, Fermi has observed a gamma-ray excess in the Galactic Center which could
be due to DM annihilation [3513]. Also, the AMS-02 has detected an excess of positrons in cosmic rays, which might
indicate DM [3512], although astrophysical sources like pulsars are a competing explanation. Future instruments and
missions, such as the Cherenkov Telescope Array (CTA) [3514–3518] and the next-generation neutrino observatories
will provide more sensitive measurements and potentially offer clearer insights into the nature of DM.
Inferring the distribution and amount of DM by its gravitational interactions is a complementary route to the
detection of byproducts. Gravitational lensing is a very suitable probe because it only relies on light deflection,
with no further need for the modeling of luminous matter effects. Since the first detection of a strongly lensed
RGB [3519, 3520], strong gravitational lensing has successfully constrained masses of galaxies and galaxy clusters to
about 10% [3521, 3522]. While the initial lens modeling of the light-deflecting mass density profiles overestimated the
DM inferred from strong lensing, increasing complexity of the lens models has reduced the amount of DM necessary
to produce the observed, highly magnified and distorted images of background galaxies [3523]. Yet, open questions
still remain whether the amount of substructures observed in galaxy clusters agree with ΛCDM [3524–3528].

4.4.2. Warm dark matter


Coordinator: Supriya Pan
Contributors: Biswajit Karmakar, Branko Dragovic, Cláudio Gomes, Cora Uhlemann, Davide Pedrotti, Emmanuel
N. Saridakis, Ioannis D. Gialamas, Janusz Gluza, Massimiliano Romanello, Mehmet Demirci, Pran Nath, Reggie C.
Pantig, Riccardo Della Monica, Torsten Bringmann, Venus Keus, Weiqiang Yang, and Wojciech Hellwing

4.4.2.a. Constraints on WDM from observation of high-redshift galaxies and reionization The introduction of
WDM particles, with mass of the order of a few keV, can alleviate some of the challenges of the ΛCDM model at the
kpc scales, namely the Missing Satellites problem or the Too-Big-to-Fail problem [2342]. Indeed, during the structure
formation process, the higher thermal velocity of non-collisional WDM particles determines a flow from overdense to
underdense regions, canceling the cosmic perturbations with a physical size smaller than the so-called free streaming
length, λFS . This leads to a suppression in the matter power spectrum, P (k), which becomes more important with
decreasing WDM particle mass, mχ , and has macroscopic consequences on the number density of low-mass haloes.
In addition to baryonic effects related to the star formation efficiency in DM haloes, small-scale modifications in
the halo mass function can also alter the high-z number density of faint galaxies, resulting in a turn-over in the rest
frame galaxy ultraviolet LF (UV LF) [3529, 3530] and in a delay in the galaxy formation [3531, 3532]. From the
comparison between the abundance of faint galaxies and the number density of DM haloes predicted in the context
of WDM cosmologies, it is possible to derive a lower limit on mχ , independent of the baryonic physics involved in the
galaxy formation process [3533].
In turn, the progressive steepening of the UV LF faint-end slope at high redshift confirms the fundamental role
of faint galaxies during the EoR, when the intergalactic hydrogen passed from a neutral to an ionized state for the
193

effect of energetic photons emitted by the primeval sources of light, at z ≳ 6. Indeed, the UV LF is directly linked
to the star formation rate (SFR) [3534] and so to the existence of a hot stellar population that acts as a source of
ionizing photons responsible for the reheating of the IGM [3535–3537]. In particular, their number density is given
by Ṅion = fesc ξion ρUV , where the UV luminosity density, ρUV , derived from the integral of the UV LF, is multiplied
by two quantities, namely the ionizing photon production efficiency, ξion , and the escape fraction, fesc . The former
describes how efficiently it is possible to get ionizing photons from a UV continuum radiation field and depends on
different astrophysical quantities, such as the initial mass function (IMF), the metallicity and the age of the stellar
population. The latter is defined as the fraction of ionizing photons that escape from the emitting galaxies and
contribute to the phase transition of the IGM. Due to the lack of information about the geometry of the ISM in
high-z galaxies, fesc is not well constrained and it is difficult to model properly. Moreover, as a multiplicative factor
of the total ionizing photons budget, it degenerates with the WDM particles mass, thus summarizing most of our
uncertainties about the reconstruction of the reionization history in both ΛCDM and ΛWDM scenarios [3538, 3539].
Despite these astrophysical uncertainties, constraints on mχ can be obtained through observations of the LF [3540–
3542] and the SFR density [1747] up to high redshift, and from the comparison of the expected electron scattering
optical depth, τes , positive with an ionized IGM, with empirical data from Planck [3543].
4.4.2.b. Candidates Traditionally, WDM is defined as a fermion with 2 degrees of freedom that decouples rela-
tivistically and has a temperature to match the observed DM relic density with zero chemical potential. This allows
to express observational limits on the free-streaming length as constraints on the mass of the DM particle. Lyman-α
observations, for example, firmly exclude mWDM > 1.9 keV [3544] (though sometimes more aggressive bounds have
been derived, reaching up to mWDM > 5.3 keV [1546]).
From the theoretical point of view, an excellent WDM candidate consists of a fermion that is a singlet under the
standard model gauge group, commonly referred to as sterile (or right-handed) neutrino [3545, 3546]. The resulting
mixing with active neutrinos, through the neutrino portal, is constrained to mixing angles too small for sterile neutrinos
to fully thermalize in the early Universe. Instead, DM production proceeds via oscillations, combined with neutral
and charged current interactions of the active neutrinos with the SM heat bath [3425]. Parameter regions for which
this mechanism produces the correct DM abundance in its simplest form are excluded [3547]. These bounds can be
evaded by enhancing the production through oscillations by introducing a large primordial lepton asymmetry [3426],
new self-interactions of the SM [3428, 3548] or sterile neutrinos [3429, 3549]. Further alternative scenarios that remain
viable include the decay of a new scalar state combined with subsequent entropy injection [3550–3552] and thermal
production via extended gauge sectors [3553–3556]. Common to these scenarios is that the WDM mass bounds in
general become model-dependent, and therefore have to be re-expressed by using model-independent bounds on the
free-streaming length (for the case of self-interacting sterile neutrinos, notably, the bound on the sound-horizon due
to late kinetic decoupling can be more stringent [3429]).
Fuzzy dark matter (FDM): Fuzzy DM (FDM), also known as wave DM, quantum wave dark matter, or ultra-
light axion DM, is an interesting area of research that bridges concepts from quantum mechanics, astrophysics, and
cosmology, which offers potential solutions to some longstanding puzzles in our understanding of the Universe. It is a
theoretical model proposing that DM consists of ultra-light particles with boson masses on the order of mb = 10−22 eV
[2371, 3453].
Unlike traditional particle DM models, fuzzy DM behaves more like a wave due to its extremely low mass. This
wave-like nature leads to unique quantum mechanical effects on large scales [3557, 3558]. This wave-like behavior
of FDM results in a quantum pressure that opposes gravitational collapse on small scales. This can suppress the
formation of small-scale structures, potentially addressing the “missing satellites problem” observed in the MW and
other galaxies. In terms of halo structure, the density of DM halos is predicted to have a central core with a soliton-
like structure. This differs from the sharp cusps predicted by CDM models and can help resolve some discrepancies
between observations and simulations of galactic cores. While ultra-light, FDM behaves like a CDM on cosmological
scales, providing a good fit for large-scale structure formation and CMB observations. Some works show hints that
ultralight axions can improve consistency between CMB and galaxy clustering data, potentially reducing the S8
tension [3559]. Tensions between the former and Lyman-alpha constraints could be alleviated with a component of
ultralight or WDM [1555]. Regarding astrophysical tests, precise measurements of the distribution and movement
of stars in galaxies, the structure of DM halos, and gravitational lensing effects can be done. The low mass end
for ultralight DM has been ruled out by the dynamics of ultra-faint dwarf galaxies based on stellar kinematics in
virialised halos [3466] and the Lyman-alpha forest [1544] using hydrodynamical simulations, where uncertainties in
the modelling remain. Recently, constraints on the soliton mass using the Event Horizon Telescope (EHT) results for
the black hole shadows of M87* and Sgr. A* was found [3560], confirming the theoretical model of FDM. It was also
tested and constrained using the motion of the S2 star around Sgr. A* [3468]. How the FDM soliton evolves while
being accreted on a black hole helps determine the stages of accretion flow [3561]. Not only for black holes, the effect
of solitonic quantum wave DM was also explored through wormhole solutions [3562]. Indeed, these studies explored
how the effects of FDM can manifest through astrophysical objects, acting as an alternative method for Earth-based
194

laboratory detection.
Based on the existing literature, it is clear that the WDM models are potentially reach since some indications
for alleviating the S8 tension has been noted. However, the models in this category are not fully explored widely
unlike other cosmological models. We anticipate that WDM remains to be a potentially rich area that needs to be
investigated widely aiming to understand how the tensions on H0 and S8 can be relieved.

4.4.3. Interacting and decaying dark matter


Coordinator: Elsa M. Teixeira
Contributors: Amare Abebe, Amin Aboubrahim, Andrzej Borowiec, Anil Kumar Yadav, Branko Dragovic, Brooks
Thomas, Davide Pedrotti, Emmanuel N. Saridakis, Emre Özülker, Gaspard Poulot, Ioannis D. Gialamas, Ivonne
Zavala Carrasco, Jose A. R. Cembranos, Keith R. Dienes, Luis Anchordoqui, Marcin Postolak, Nikolaos E. Mavro-
matos, Paolo Salucci, Pran Nath, Rafael C. Nunes, Reggie C. Pantig, Riccardo Della Monica, Rishav Roshan, Simony
Santos da Costa, Tomi Koivisto, Torsten Bringmann, Víctor H. Cárdenas, Vasiliki A. Mitsou, Venus Keus, Vivian
Poulin, William Giarè, and Wojciech Hellwing

4.4.3.a. Interacting dark matter Models of IDM replace the CDM component in ΛCDM by posing either a self-
interaction between DM particles or interactions between DM and the other species in the Universe. Consequently,
the mass of the DM particles bears some dependence on such interaction, defined in an FLRW background as

ρ̇IDM + 3HρIDM = Qi , (4.20)

where Qi embodies the interaction between DM and the i-th species with energy density ρi (whose conservation
equation will include a symmetric interacting term to ensure energy conservation). One key feature in favour of
several models of IDM is that they can exhibit suppression of the matter power spectrum in relation to its ΛCDM
counterpart. This effect is relevant in terms of attempts to address the cosmological tensions, particularly the S8
tension, even upon including constraints from the CMB data. In the following, we present various types of IDM
models and discuss their relevance for the resolution of the cosmic tensions.
Dark matter interactions with baryonic matter Theoretical models often consider DM-baryonic matter
(DM-BM) scattering to modify the evolution of density perturbations. For instance, velocity-dependent cross-sections
in these interactions can impact the growth rate of cosmic structures, affecting the matter power spectrum and, hence,
the derived value for the S8 parameter [2315, 3563, 3564]. These interactions may also modify the reionisation history
through changes in the ionisation fraction, optical depth, and CMB data-derived cosmological parameters [3565].
Recent simulations suggest even weak DM-BM interactions can leave observable effects on large-scale structures and
the Lyman-α forest [3566]. Screened DM models propose local environment-dependent interaction strengths [3231]
while models like matter bounce scenarios propose alternatives to the standard inflationary paradigm [3567, 3568].
These DM-BM interaction models must balance improving cosmological fits without violating constraints from other
astrophysical observations, such as galaxy clustering and WL surveys, thereby representing a promising but challenging
research avenue [3569]. Further observational data from upcoming missions and theoretical refinement are needed to
assess their viability as a potential solution to the cosmological discrepancies [3570].
Recent astrophysical evidence shows discrepancies between observed mass distributions in galaxies and predictions
from the collisionless ΛCDM model [3343]. Notably, a constant density region is often found in galaxies’ innermost
areas, contrary to the expected cuspy density profile [358, 2343, 2345, 2352, 3571, 3572]. Additionally, DM halo
parameters correlate with the structural parameters of the luminous component, such as the half-light radius R50
and stellar mass M⋆ [3573–3577]. This entanglement between the dark and luminous components suggests some
interaction between DM and BM [3578], occurring only where the product of DM and BM densities exceeds a threshold
[3578, 3579]. Outside these regions, DM halos retain their original NFW profiles. The DM-BM interaction, observed
over several Gyrs, does not affect the formation of galaxy dark halos but gradually modifies the halo density over time,
erasing the initial cusp. This process may be observable in the recent mass distribution of high-redshift galaxies [3578].
While the macroscopic effects of this interaction are clearer, the microscopic processes remain uncertain. Numerous
candidates for IDM have been proposed (e.g., see Refs. [3580–3583]), and investigations continue. Observations suggest
these interactions could occur in secluded regions of galaxies, such as stars, molecular clouds, neutron stars, black
holes, and central supermassive black holes, aligning with various studies (e.g., see Refs. [3584, 3585]).
Dark matter interactions with photons In Ref. [2661], an extension of the ΛCDM model is proposed where
DM couples with photons (γ), resulting in a nonconservation of particle numbers. This model suggests that DM
particles gradually decay into photons over time without significant scattering processes, slightly deviating from
the standard cosmic evolution. It offers a potential solution to the cosmic tensions (see Fig. 4). Based on the
theoretical study of couplings between DM and relativistic relics in Ref. [3586], a lower limit on the DM mass has
195

been derived [3587] for DM-γ interactions: mχ > 8.7 keV at 95% CL (assuming it saturates 100% of the observed
relic density). Alternatively, one could derive a bound on the abundance of such DM candidates depending on the
mass [3556]. Observational constraints on DM-γ scattering cross-section were derived in Ref. [3588], with σDM−γ ≤
2.25 × 10−6 σTh , (mDM /GeV) at 95% CL. A multi-IDM framework, including interactions with photons, was explored
in Ref. [3589], simultaneously addressing the H0 and S8 tensions. The prospects for testing DM-γ interactions through
CMB spectral distortions have been discussed in Ref. [3590]. The black hole shadow, a dark region observed against
a bright background, is closely related to the photonsphere, the region where photons orbit the black hole. By
performing backward tracing techniques, one can study how the shadow deviates from the classical prediction due
to DM effects. Studies like Refs. [3591, 3592] have shown that DM-γ interactions in the photonsphere can alter the
shadow’s shape, illustrating how astrophysical objects can be used as independent DM probes. Numerous studies
have since explored different DM models and their effects on black hole properties [3468, 3560, 3593–3622], concluding
that significant DM concentrations are necessary for observable effects on a black hole’s shadow [3623].
Dark matter interaction with dark radiation The possibility that the multiple cosmic discrepancies are
tied to the existence of a dark sector filled with a thermal bath of dark relativistic species, also known as dark
radiation (DR), interacting with DM has been widely explored (e.g., see Ref. [2549, 2553, 2660, 2661, 3624–3631]). In
the standard model, the effective number of neutrinos accounting for the three neutrino species, Neff = 3. Still, the
presence of DR can modify this value, reducing the sound horizon and adjusting the inverse distance ladder calibration
of the BAO and SNIa. This is discussed extensively in Sec. 4.1.3.
The DR-DM interaction has two primary effects: it introduces a drag term in the DM Euler equation, reducing
the growth of DM perturbations and potentially explaining the low S8 values measured by WL surveys; and it
influences DR perturbation dynamics, reducing their free-streaming, which can result in a larger contribution to Neff
and align with H0 measurements from the SH0 ES experiment. Various models embody these ideas, such as the “non-
Abelian DM model” [3624, 3625, 3627], “partially acoustic DM" [3626], “cannibal DM” [3628], stepped-dark radiation
[2549, 2553, 2660, 3629–3631], and neutrino-DM interaction [2317]. Interestingly, these models may also address other
CDM issues, such as the cusp/core problem, the so-called “too big to fail" (TBTF) issue, and the missing satellites
problem (see Ref. [2342] for a review). Indeed, Ref. [3632] is the first particle-physics proposal to address all three
problems simultaneously. The model in Ref. [3633] integrates this into DR as sterile neutrinos, introducing late-time
effects that influence H0 /σ8 . Ref. [3634] offers a comprehensive Lagrangian-level classification of DM-DR interaction
models, showing how CDM can mimic a WDM cut-off in the matter power spectrum. Alternatively, model-independent
analyses can be conducted through the “effective field theory of structure formation” (ETHOS) [2296, 2297], which
provides a mapping from Lagrangian parameters of DM-DR models to phenomenological parameters of the linear
power spectrum, with numerical simulations extending this to the full non-linear power spectrum.
Dark matter interaction with dark energy Interactions within the dark sector may help resolve various
tensions simultaneously by leveraging on the unknown nature of DE and DM (see Sec. 4.2.3 for a more thorough
account of IDE models). For instance, in Refs. [3635–3637], the authors show that by applying fluid approximation
methods for rapidly oscillating scalar field DM coupled to scalar field DE through a Lagrangian approach, DM can
effectively be treated as a fluid on cosmological scales under the appropriate Hubble-averaged behaviour. These
models are expected to help alleviate cosmic tensions, namely on S8 , due to their contributions in suppressing the
matter power spectrum on small scales, and on H0 through the modified background dynamics.
A recent paper proposed a cosmological model of DM and DE based on a particle physics Lagrangian [3638] to
address cosmic tensions. The model involves two interacting ultralight scalar fields, one for DM and one for DE, with a
coupling between them and was tested against various cosmological data sets, including Planck [192, 728, 1429], BAO
[776, 796, 1141, 2503, 3639, 3640], and Pantheon and SH0ES data [33, 34]. It demonstrated an ability to alleviate the
H0 tension, reducing the discrepancy to about ∼ 2σ, and also resolved the S8 tension.
4.4.3.b. Decaying dark matter Another approach posits instead that a fraction of the DM sector is undergoing a
decay process into another species, falling under the broad heading of decaying DM (DDM) models. This scenario is
motivated by theoretical considerations, can explain specific experimental results, and addresses small-scale issues in
the CDM paradigm. Initially developed in the 1980s, the DDM model has gained renewed interest in light of cosmic
tensions. Accordingly, such models conventionally contain two extra parameters: the lifetime of DDM τ plus the
initial fraction of DDM to total DM fDDM . The decay is expressed in FLRW as

ρ̇DDM + 3HρDDM = −Γi ρDDM , (4.21)

where Γi is the decay width of DDM (related to its lifetime as Γ = τ −1 ) into the species ρi . There are several
variations of the DDM hypothesis. However, these scenarios alone are unlikely to resolve the H0 tension, which would
require, for example, assuming an earlier decay of DM.
196

Dark matter decays into dark decay products Various scenarios involving the decay of an unstable compo-
nent of multicomponent DM into dark radiation have been proposed to address the H0 and S8 tensions [2559, 3641–
3648]. For short-lived particles (Γ ≳ 106 Gyr−1 ), DM decays into dark radiation at early-times (t ≪ tLS ∼ 1.17×1013 s,
where tLS denotes the time of last scattering), increasing the expansion rate and reducing the sound horizon size rs ,
thus increasing H0 [2559, 3641, 3642]. Since the angular size of the sound horizon at the last scattering surface
θLS ≡ rLS /DM (tLS ) is a CMB observable that must be kept fixed, a reduction of rLS simultaneously decreases the
comoving angular diameter distance from a present-day observer to the last scattering surface DM (tLS ), and in-
creases H0 . For long-lived particles, DM is depleted into radiation after tLS , shifting matter-DE equality to earlier
times and allowing a late-time increase in H0 [3643, 3644]. Moreover, two-body decays that transfer energy from
DM to DR, reducing the late Universe matter content, help accommodate local S8 measurements [3645–3647]. For
Γ ≳ H0 ∼ 0.7 Gyr−1 , most unstable DM particles have decayed by redshift z = 3, impacting observations like those
from IceCube if sterile neutrinos play the role of DR [3649, 3650]. If Γ ≲ H0 , only a fraction of DM particles have
decayed. Recent CMB data severely constrain the fraction of unstable DM in these scenarios [3556, 3651–3656]. Short-
lived particles are constrained by CMB polarisation [3655, 3656], and low redshift DM depletion reduces CMB lensing
power, conflicting with Planck data [3651–3653]. Including BAO measurements further tightens constraints on the
fraction of long-lived particles [3654, 3655]. Current bounds make a decaying DM solution to the H0 tension unlikely,
though a combination of scenarios with multiple decaying DM particles (for example decays into a combination CDM
and WDM [3657]) at different epochs may effectively ease this tension.
More promising is the role that DDM can play regarding the S8 tension, especially in models where the (dark)
decay products are massive particles. These products get a velocity kick ϵ ≡ v/c from the decay, suppressing structure
growth on scales below their free-streaming length while behaving like regular CDM at the background level, thus
avoiding the strong constraints from BAO data. These scenarios have been tested against various cosmological data
[3646, 3647], including the full shape of BOSS power spectrum [3658] and KiDS WL measurements [3659], and shown
to reduce the S8 tension to ∼ 1.5σ for lifetimes Γ−1 ≃ 120 Gyr and velocity kick ϵ ≃ 1.2% [3658]. In addition,
depending on the lifetime and kinetic energy of the decay, DDM can address or mitigate various tensions related
to small-scale structures and galaxy formation [3660–3664] such as the missing satellites problem [2348, 2349], by
estimating the power spectrum cut-off scale.
It is also worthwhile to move beyond the specific idea of decaying particle DM and consider a more generic pa-
rameterization to describe the potential conversion of a DM component to DR, with a transition rate that may not
necessarily follow exponential decay [2314, 2398]. This approach includes traditional decaying DM but also encom-
passes scenarios such as merging primordial black holes (PBHs) or late-time DM self-annihilation in the presence of
long-range forces. In these generalised scenarios, the S8 and H0 tensions can be mitigated in somewhat different ways,
though a complete resolution of these tensions remains unlikely.
Dynamical dark matter DDM models the decay of a DM ensemble across epochs, balancing the collective
lifetimes against cosmological abundances among various DM components [3665]. One realization of such a DDM
ensemble that has been extensively studied consists of a tower of dark Kaluza-Klein states associated with a large
extra dimension; such a tower then allows for the possibility of “intra-ensemble” decays from heavier to lighter dark
states [3666, 3667]. Such decays can modify the DM velocity distribution over time [3667] and potentially address
the S8 tension [3668, 3669]. This idea has recently found new relevance within the Swampland program [3670], which
suggests an extra mesoscopic dimension (“dark dimension”) in the micron range [2907], leading to a tower of weakly
interacting light DM particles which are KK excitations of the graviton [3671]. While DDM scenarios of this sort may
potentially address the S8 tension, they are also subject to constraints on the expansion rate at late-times from, e.g.,
SNIa [3672].
Dark sectors with varying equations of state In the case of soft DM and soft DE [3673, 3674] one considers
that effectively the dark sectors have a different equation of state at large scales, namely at scales entering the
Friedmann equations, and a different one at intermediate scales, namely at scales entering the perturbation equations,
features that are typical in soft materials [3675]. In this case, the perturbative-level properties are slightly changed,
and one can easily alleviate the S8 tension. It has also been shown that such scenarios can be compatible with
JWST observations, offering a potential explanation for the unexpectedly rapid emergence of massive galaxies at high
redshifts [3676].
Some applications of decaying dark matter DM decaying during or after recombination can alter the CMB
power spectrum by injecting energy that reionises the IGM, constraining the DM lifetime to τDM ≳ 1025 s for
photon or e+ e− decay products [192, 3556, 3677, 3678]. Additionally, null detections of diffuse X/γ-rays impose
strict constraints on the DM lifetime. In DM-dominated galaxies or clusters, e+ e− decay products can produce radio
waves through electromagnetic interactions, observable by radio telescopes like the SKAO [3679], providing a better
probe for decaying DM compared to current gamma-ray observations. A recent study [3680] found that DM decay
width ΓDM ≳ 10−30 s−1 is detectable with SKAO. These constraints were used in Refs. [3681, 3682] to understand
197

the scale of Quantum Gravity, assuming that it breaks all approximate global symmetries, including DM stability.
Moreover, null detection of X-ray signals has ruled out minimal DM models like νMSM [3683] for DM produced from
active-sterile oscillation and has severely constrained the DM mass above the MeV scale for feebly interacting massive
particles produced from gauge boson decays [3684].

4.5. Other solutions


4.5.1. Estimates based on a possible local void
Coordinator: Indranil Banik
Contributors: Alireza Talebian, Bahman Khanpour, Ebrahim Yusofi, Harry Desmond, Paolo Salucci, Richard
Stiskalek, Sahar Mohammadi, Sandeep Haridasu, Sergij Mazurenko, Sveva Castello, and Vasileios Kalaitzidis

Cosmic voids may help to alleviate the Hubble tension. A void effectively adds negative mass to a localized region
of an otherwise homogeneous universe. The repulsive gravitational effect of this negative mass causes matter to flow
away from the void, further deepening it and enhancing its repulsive effect. Recent studies in a void-dominated
late universe [3685–3688] indicate that cosmic voids achieve maximum entropy and thermal equilibrium [3689, 3690].
This hierarchical evolution may lead to the formation of large spherical cosmic voids through the merging of smaller
sub-voids via the void-in-void process [1782, 1791, 3691]. Studies suggest that clusters and voids can exert positive
pressure individually [3687, 3690]. Their coexistence in the Universe’s web-like structure results in opposing pressures:
overdensities generate positive pressure that pulls them closer together, while voids create negative expansion pressure
that separates galaxies. The outflow from a local void could inflate observed redshifts and thus estimates of H0 , though
this effect is not sufficient to solve the Hubble tension in ΛCDM [3692, 3693].
Near-infrared observations covering 90% of the sky suggest that we live inside the Keenan-Barger-Cowie (KBC)
void [3694]. The results are deep enough to cover most of the galaxy LF, making them representative of the underlying
matter distribution. The KBC void or Local Hole is evident throughout the electromagnetic spectrum, including in
X-rays [3695, 3696], optical [3697, 3698], infrared [3694, 3699–3706], and radio wavelengths [1402, 3707]. Comparison
with the Millennium XXL simulation [3708] suggests that the KBC void is 6σ discrepant with ΛCDM [3709]. Since
the underlying observations are galaxy number counts in redshift space, the results are unaffected by the assumed
H0 .
Various proposals have been made arguing that gravitationally driven outflows from the KBC void could solve the
Hubble tension [3710–3714]. In a homogeneously expanding universe, we would expect that cz ′ = ȧ = H0 , where
z ′ ≡ dz/dr is the rate at which z rises with r. However, if we are located near the centre of a large void, the outflow
velocity would rise from zero at the void centre to some maximum further out, before gradually decaying away. The
initial rise with r would inflate cz ′ . To estimate how much, we note that the KBC void has an underdensity in
redshift space of δobs = 46 ± 6% (see figure 11 of Ref. [3694]). The actual underdensity is about half as much because
of RSD [3715], the reduction in distance to any fixed redshift due to outflow from a local void inflating cz ′ . The
reduced volume reduces the number counts, making the underdensity appear deeper in redshift space than it actually
is. Accounting for this and assuming the required reduction in comoving density from the nearly homogeneous early
Universe must be due to an increase in comoving volume, [3709] argued in their equation 5 that

cz ′ −1/6
= (1 − δ obs ) . (4.22)
H0
Thus, we would generally expect the locally estimated H0 (actually cz ′ ) to exceed H0 estimated from the CMB or
other high-redshift probes by 11 ± 2%. This is only a rough estimate – the actual result depends on how the void has
evolved over cosmic history. This must differ from the growth of voids in ΛCDM given that density fluctuations on
the KBC void’s comoving scale of 300 Mpc are very well observed at the time of recombination and match ΛCDM
expectations [700], implying that density perturbations on these scales grow faster than in ΛCDM. It is suggested
that this is caused by a modification to gravity on length scales ≳ 100 Mpc, which would not affect the early Universe
as its smaller age reduced the cosmic horizon, the distance that light and gravity could have travelled [3709].
A further argument for this scenario is the observed bulk flow curve out to z = 0.083 or about 350 Mpc [1421] based
on the CF4 galaxy catalogue [1419]. The bulk flow measures the average vector velocity of tracers within a spherical
region centred on the Sun, albeit using only line of sight velocities. Ref. [1421] argue that the observed bulk flow
on the largest probed scales is roughly quadruple the ΛCDM expectation and thus in > 5σ tension with it (see also
Sec. 2.3.14). The bulk flow curve in a full-sky survey is independent of the assumed H0 , which sets the monopole of the
velocity field, while the bulk flow is a function of the orthogonal dipole. There are additional subtleties when the sky
coverage is not complete, but since CF4 covers most of the sky, a slight adjustment to the estimator is sufficient to deal
with the small unobserved regions of the sky, which mainly lie at low Galactic latitudes (see figure 2 of Ref. [1419]).
However, the sky coverage becomes less complete at z ≳ 0.05 as CF4 mostly relies on SDSS. There can be spurious
198

effects even at lower z because CF4 is reliant on combining surveys, each with their own zero-point and limited sky
coverage.
The local void scenario was explored in great detail by Ref. [3709], who constructed semi-analytic models of a
small initial underdensity at z = 9 and evolved it to today. Those authors used Milgromian MOND ([364–366]) to
enhance the gravity from any given (under)density distribution, allowing the exploration of self-consistent models
with enhanced structure growth on the relevant scales. A good joint fit was obtained to the density profile of the
KBC void and the observed magnitude of the Hubble tension. An important conclusion was that our location in the
void needs to be fine-tuned at the 2% level, which is not a strong argument against the model [3709]. The velocity
field in their model was recently calculated in more detail [3716]. The predicted bulk flow curve was found to agree
well with the observations of Ref. [1421], which are themselves in good agreement with the measurements reported by
Ref. [1422]. The rising bulk flow curve implies we are located fairly close to the void centre, minimizing any impact on
the CMB quadrupole due to gravitational lensing [3717, 3718]. Ongoing work aims to perform this peculiar velocity
inference at the field level rather than relying on the bulk flow summary statistic.
The H0 tension should disappear at high z once the density returns to the cosmic mean and the void’s impact
gradually decays. Indeed, the ages of the oldest Galactic stars and CCs (Sec. 2.1.13) suggest a low H0 consistent with
the CMB value and well below the local cz ′ [438, 441, 3719, 3720]. In fact, the Hubble tension is largely a mismatch
between the local cz ′ and H0 estimated in other ways, typically from higher z data [3721]. This can be done assuming
ΛCDM, leading to the concept of H0 (z), the value of H0 inferred from data in a narrow redshift range centred on z.
Although ΛCDM predicts a flat curve, there is evidence for a declining trend [515, 1285, 3086, 3087, 3092, 3722, 3723].
The most recent analyses carefully minimize covariance between H0 estimates from different redshift bins, finding 6σ
evidence for a declining trend [3724, 3725]. Similar techniques can be applied to construct simulated H0 (z) curves
that map the decay of the void’s impact on z [3726]. Their figure 3 shows good agreement with the observed H0 (z)
[3724, 3725].
BAO observables also deviate from ΛCDM in the manner predicted by the local void scenario [3727]. These
discrepancies (especially in DESI data; [698]) have alternatively been interpreted in terms of the DE density varying
with time [1278, 1279, 3728]. Regardless of the interpretation, the Hubble tension seemingly becomes weaker at high
redshift, and not merely due to larger uncertainties [2675]. A recent study examined a variable DE fluid with a
quadratic EoS in a late-time context [3688, 3690, 3729]. Such an EoS may emerge from cluster/void mergers and
could represent a redshift-dependent EoS parameter for cosmic acceleration [3686–3688, 3690].
The main issue facing the void model at the moment is that the Hubble tension does not clearly decay at high z
in analysis of SN [3730], though there are some hints of such a decay [515, 3092, 3724]. While the BAO ruler has
a fixed comoving size at z ≲ 1000, SN are not standard candles but merely standardizable [3731]. The intrinsic SN
luminosity depends on colour and stretch, or how rapidly the light curve decays [33, 218]. There are not enough
SN in host galaxies with Cepheid distances to reliably constrain these dependences, which must be obtained jointly
with the other cosmological parameters in the Hubble flow sample. This is usually done assuming some parametrized
form for the relation between luminosity distance and redshift derived from an isotropic and homogeneous Friedmann
cosmological model [3732, 3733], though with an additional free parameter in the DE equation of state w [3734–3736].
SN calibrated through the distance ladder, in combination with BAO data and the Planck distance priors [192],
suggest compatibility with a homogeneous cosmology and indicate that a potential local void would not significantly
impact the measurement of H0 from SN [3737–3739] – but this would need to be confirmed by less model-dependent
approaches as a cross-check [3740, 3741]. A particular concern is that the typical stretch and colour of SN exhibit
strong trends with z [3742–3744]. Due to the need to calibrate how both affect the intrinsic SN luminosity, these
trends may mask the decaying away of the excess redshift induced by a local void [3726]. One might think that
neglecting an actually present local void would cause the optimal homogeneous Friedmann model to provide a poor
fit to the data, but the SN magnitude uncertainties are artificially inflated such that the reduced χ2 = 1, completely
hiding any such issue [3745]. The results also change slightly if we assume that it is the colour rather than magnitude
that has an extra unknown source of noise, for instance, due to dust [3746]. There is ongoing work on revising the
SN fitting procedure to allow a local void to inflate redshifts in the nearby Universe, jointly considering cosmological
and standardization parameters.
While the KBC void could potentially solve the Hubble tension, it is not evident in the reconstructed local density
field [2305] and its existence has not been validated kinematically [3730, 3747]. A recent analysis of galaxies using
the Radial Tully-Fisher (RTF) relation [355] provides a very promising way to investigate the possibility [356]. They
demonstrated comprehensively that a local void capable of alleviating the Hubble tension is highly unlikely to reside
at z ≲ 0.01 (as suggested by Ref. [3704]) because it would have been evident in the analysis of the local cosmic
expansion through the RTF distance indicator. This agrees with previous studies over the same redshift range using
standard SN distance indicators and datasets [3730, 3747] – though a plausible local void would inflate redshifts much
further out [3726]. Remarkably, the RTF indicator could easily become the best distance indicator for z ≲ 0.15, where
it could be applied to a galaxy sample several times larger than SN datasets. In each galaxy, the RTF method would
199

provide multiple distance measurements, each as accurate as those obtained from SN.
The KBC void is unique among solutions to the Hubble tension in that it was proposed long before the Hubble
tension became known. Its large size and depth and the anomalous bulk flow curve all suggest that structure formation
on scales of several hundred Mpc is more efficient than expected in ΛCDM, as also suggested by the high redshift,
mass, and collision velocity of the El Gordo interacting galaxy clusters [3748, 3749]. If so, the Hubble tension might
be resolved without adjusting the background cosmology, preserving its excellent fit to high-redshift datasets and
the ages of the oldest stars [441, 456, 3719]. In the future, important tests will be provided by techniques which
extend beyond the void radius. For instance, peculiar velocity reconstructions using galaxy scaling relations already
constrain cz ′ out to z = 0.1 [389, 390, 394, 397]. Extending them to z ≳ 0.3 would test the predicted decaying of the
void’s effect at larger z [3726]. BAO measurements suffer the opposite problem: they are most accurate at z ≳ 0.5,
well beyond the void’s influence [698]. It is crucial to extend BAO measurements to z ≲ 0.2, where a local void
should have an appreciable impact [3727]. More detailed analyses of existing datasets would also help, for instance
obtaining bulk flow measurements using SN to check results obtained so far using galaxies and extend them further
out [1383, 1385, 1386, 2829, 3750, 3751]. The inferred bulk flows can serve as a consistency check on measurements
using the far more numerous galaxies.

4.5.2. Primordial magnetic fields


Coordinator: Karsten Jedamzik
Contributors: Alireza Talebian, Anto Idicherian Lonappan, Cláudio Gomes, Gaetano Lambiase, Iryna Vavilova,
Levon Pogosian, and Tom Abel

The possible existence of PMFs in the early Universe has been a research topic for decades (for further reviews see
Refs. [3752–3754]). The underlying question is if magnetic fields observed in galaxies, clusters of galaxies, filaments,
and in particular voids of the extragalactic medium [3755–3761] could be the result of a magnetogenesis process in the
very early Universe which magnetized the Universe already well before recombination. The alternative would be that
all/or most of the observed magnetic fields are due to astrophysical processes such as dynamos and outflows. The
impact of a putative PMFs on the CMB has therefore been studied in great detail [686, 3762–3808]. Most proposed
effects result in upper limits around comoving ∼ 1 nG, which is however a field strength about two orders of magnitude
stronger than that required for cluster magnetic fields ∼ 1µG to be explained entirely by a PMFs [3809]. It has been
realized that the anisotropies in the CMB are particularly sensitive to one effect, the creation of small-scale baryon
inhomogeneities by PMFs [3810, 3811].
A PMF of comoving field strength ∼ 0.05 nG on comoving length scales ∼ kpc produces slightly non-linear baryon
inhomogeneities on such scales during recombination. This is possible since kpc scales are well below the photon
mean free path such that magnetic pressure is counter-acted only by the small baryon pressure and not the photon
pressure. This process is often referred to as “baryon clumping”.
Note that any PMF is characterized by a continuous spectrum, containing initially magnetic power on a large range
of scales, either with a scale-invariant spectrum due to inflationary magnetogenesis or a very blue Batchelor spectrum
due to magnetogenesis during cosmic phase transitions. In the latter scenario the bulk of the PMF is dissipated
as the Universe expands, and a comoving ∼ 0.05 nG field left shortly before recombination results in a ∼ 0.01 nG
field at the present epoch, approximately the value to explain cluster magnetic fields. The process of PMF-induced
baryon clumping has been studied numerically and allowed to set stringent upper limits on the final total present
PMF strength of ∼ 0.01 nG for a Batchelor spectrum [3812]. This upper limit may be subject to factor two changes
when the newer Planck CMB data and/or a more realistic treatment of the PMF effects on the CMB are employed.
Recombination in a clumpy baryon medium proceeds faster than in a homogeneous medium due to the non-
linearity in the recombination rate (i.e., αe ⟨ne np ⟩ > αe ⟨ne ⟩⟨np ⟩ where αe is the recombination coefficient and ne , np
are electron- and proton- density, respectively, with ne ≈ np ). Recombination therefore occurs earlier when PMFs are
present, such that the sound horizon r⋆ is reduced [3813]. A reduction of the sound horizon compared to ΛCDM is
the ingredient utilized by essentially all early-time solutions for the Hubble tension (see Sec. 4.1.1 for an explanation).
The effect of baryon clumping on the CMB has been analyzed by several groups [3813–3816]. Here the clumping
was treated in toy three-zone models, averaging the ionization fraction from three different and independent regions
with the average density the same as in an homogeneous universe. Due to the absence of detailed knowledge of the
PMFs generated baryon PDF function, i.e., the probability to find a baryon at a certain density, and it’s evolution,
non-evolutionary models with an ad hoc PDF were chosen. In Ref. [3813] it was found that when Planck data in
combination with three late-time H0 determinations was confronted to such clumping models, clumping was preferred
at the ∼ 3σ CL with inferred H0 ≈ 69.8 − 71 km s−1 Mpc−1 depending on the chosen baryon PDF. Though the fit
to the CMB data was slightly worse than in ΛCDM it was still reasonably good (a PTE value of 0.17 compared to
0.20 for ΛCDM without PMFs). On the other hand, when other data is included, such as BAO and SN, slightly
lower H0 ≈ 69.7 − 70.6 km s−1 Mpc−1 were inferred. It is now known that the combination of Planck CMB data and
200

BAO data essentially disallows H0 as high as the SH0ES value H0 ≈ 73 km s−1 Mpc−1 in all early-type solutions to
the Hubble tension if the physical matter density parameter Ωm h2 has a value comparable to the one in the best-fit
ΛCDM model [803]. When the model was confronted only to CMB data (Planck in combination with SPT-3G and/or
ACT), but no late-time H0 measurements were used, clumping was no further detected but only limited from above
[3814–3816]. It was argued that clumping is heavily constrained by the reduced Silk damping in conflict with the
ACT data [3814], see below however.

z = 1000
74
PMF Planck+DESI+H0
2

72

Projected Over Density

H0 [km/s/Mpc]
y (24 kpc)

70

68

0.6

66
0.005 0.01 0.015
x (24 kpc) bpmf [nG]

FIG. 78: Left: Results of an MHD simulation [3817] starting at redshift z = 4500 with an initial comoving mag-
netic field of 0.526 nG with Batchelor spectrum and initially vanishing density perturbations and peculiar flows.
The figure shows the (projected) density fluctuations generated at z = 1000 by the stochastic magnetic field. Right:
Preliminary results of a MCMC comparing the theoretical prediction of a ΛCDM model with a PMF with Batche-
lor spectrum to the combination of Planck CMB data [192], DESI-1Y BAO data [698], and the SH0ES [34] Hubble
constant determination. It is seen that a final magnetic field of B ≈ 0.01 nG is preferred leading to Hubble con-
stants of H0 ≈ 70 km s−1 Mpc−1 , significantly larger than those predicted in ΛCDM without a PMF. This field
strength is close to that needed to explain observed cluster magnetic fields entirely by a PMF.

These promising trends tempted a recent much more sophisticated study of recombination in the presence of PMFs
[3817]. Instead of using three-zone models, MHD simulations of PMFs before, during, and after recombination were
performed automatically leading to the correct baryon PDF and its evolution. The simulations were performed down
to redshift z = 10, providing a direct connection between the magnitude of clumping during recombination and the
PMF total strength at the onset of structure formation. A chemistry solver coupled to the MHD simulations were used
to self-consistently compute the ionization fraction in each volume element. The average perturbation of the ionization
fraction compared to a homogeneous no-PMF model was computed and can be used to compute the anisotropies in
the CMB in a magnetized universe. The study revealed the possible importance of two further effects, namely the
mixing of Lyman-alpha photons between different regions, and enhanced Lyman-alpha photon loss due to peculiar
motions of the baryon fluid. Whereas the second effect was found to be subdominant for weak magnetic fields (final
field ∼ 0.01 nG for Batchelor spectrum) the almost complete mixing of Lyman-alpha photons from regions at different
densities leads to a further reduction of the ionization fraction during recombination. It also results in an enhancement
of Silk damping such that magnetized models often have virtually identical Silk damping than non-magnetized ones.
The left panel of Fig. 78 shows the projected baryon overdensity, i.e., (ρ − ⟨ρ⟩)/⟨ρ⟩) of a magnetized universe at
z = 1000, with density clumps entirely created by the existing stochastic PMF with Batchelor spectrum. There
is currently no other than PMF explanation for the putative existence of baryon isocurvature density fluctuations
before recombination as other scenarios (pre-existing isocurvature fluctuations) are ruled out by BBN or are not
volume-filling enough (isocurvature fluctuations due to cosmic strings). The right panel of Fig. 78 shows the result if
a magnetized ΛCDM model is compared to Planck, DESI, and SH0ES data. It is found that a universe without PMFs
is essentially ruled out by this data combination. Moreover, the predicted Hubble constant is around 70 km s−1 Mpc−1 .
Though this value does not match the SH0ES value it is in accordance with other observational determinations of H0
[174, 252]. It is intriguing to realize that the final present day magnetic field B0 ≈ 0.01 nG is essentially that required
to explain the origin of cluster magnetic fields. It furthermore is also an explanation of the putative observations of an
extragalactic field by γ-ray observations. Further observational advances in small-scale CMB anisotropy experiments
201

(ACT, SPT-3G, S4) and and γ-ray observations of a compact-size halo of powerful blazars (CTA with sensitivity less
0.1 TeV and sufficient angular resolution [3818]) have the potential to establish that the early Universe had already
been magnetized.

4.5.3. Feasibility of inflationary models to ameliorate the cosmic tensions


Coordinator: Jaume de Haro, Luis A. Anchordoqui
Contributors: Ali Övgün, Alireza Talebian, Andrew R. Liddle, Andrzej Borowiec, Anil Kumar Yadav, Anto
Idicherian Lonappan, Antonio Racioppi, Benjamin L’Huillier, Carsten Van De Bruck, Cláudio Gomes, Cristian
Moreno-Pulido, Dario Bettoni, David Benitsy, Deng Wang, Denitsa Staicova, Diego Rubiera-Garcia, Elias C. Vage-
nas, Emanuela Dimastrogiovanni, Emmanuel N. Saridakis, Giulia Gubitosi, Giuseppe Fanizza, Goran S. Djordjević,
Ido Ben-Dayan, Ignatios Antoniadis, Ioannis D. Gialamas, Javier Rubio, Joan Solà Peracaula, Juan Garcia-Bellido,
Laura Mersini-Houghton, Leila L. Graef, M.M. Sheikh-Jabbari, Margus Saal, Marina Cortês, Micol Benetti, Milan
Milošević, Nikolaos E. Mavromatos, Nils A. Nilsson, Octavian Postavaru, Oem Trivedi, Özgür Akarsu, Petar Suman,
Rishav Roshan, Salvatore Samuele Sirletti, Saurya Das, Simony Santos Da Costa, and Venus Keus

4.5.3.a. Modeling the inflationary expansion and its generalities in connection to the cosmic tensions Inflation
was first suggested to solve the flatness and horizon problems [3819]. The simplest inflationary model, known as
“vanilla” or single-field inflation, describes the expansion of the Universe using a single scalar field ϕ with a suitable
potential V (ϕ). The dynamics of this field are governed by the conservation and Friedmann equations. The most
popular approximation for the scalar field is the so-called slow-roll approximation, with the potential energy
dominating the energy density of the Universe, i.e., ϕ̇2 ≪ V (ϕ) and ϕ̈ ≪ 3H ϕ̇, where H is the Hubble rate. This leads
to an approximately exponential expansion, with the scale factor evolving as a(t) ∼ eHt . It is possible to define the
slow roll parameters: ϵ = MPl 2
[V ′ (ϕ)/V (ϕ)]2 /2 and η = MPl 2
[V ′′ (ϕ)/V (ϕ)]. Using them, one can relate inflationary
observables, such as the scalar spectral index ns ≈ 1 − 6ϵ + 2η and the tensor-to-scalar ratio r ≈ 16ϵ, to the shape of
the inflaton potential V (ϕ) [3820]. A point worth noting at this juncture is that an approach to addressing the H0
tension involves studying the slow-rolling dynamics of a self-interacting scalar field, which can lead to an emerging
H as a function of redshift; however, it should also be noted that while the scalar field could be responsible for a
quintessence slow-rolling, it has not been associated thus far to inflationary models [3090].
Vanilla inflation predicts a nearly scale-invariant spectrum of scalar perturbations with a slight red tilt, quantified
by the scalar spectral index ns ≈ 0.96 − 0.98. However, the slow-roll approximation is only one way to go, there
are other examples such as fast-roll approximation [3821], constant-roll [3822, 3823], rapid-turn inflation [3824, 3825],
scalar field inflation in the presence of a non-minimal coupling between matter and curvature [3826] symmergent
inflation [3827, 3828], gauge-field inflation [1442, 3829–3853], s-dual inflation [3854, 3855], nonlinear electrodynamics
inflation [3856–3864], and symmergent gravity theory [3827, 3828].
Many inflationary models predict primordial non-Gaussianities, which play a crucial role in understanding the early
Universe’s evolution and testing the inflationary paradigm. Non-Gaussianities are quantified by the non-linearity
parameter fNL , which measures the amplitude of non-Gaussian fluctuations in the primordial curvature perturbations.
These fluctuations require analysis of higher-order correlation functions, such as the bispectrum, beyond the two-point
correlation function used for Gaussian fluctuations. Previous CMB experiments, including WMAP and early releases
of the Planck data, aimed to detect these non-Gaussianities. The most recent constraints from the Planck 2018 results
equilateral orthogonal
provide fNL values for the standard templates: fNL local
= −0.9 ± 5.1, fNL = −26 ± 47, and fNL = −38 ± 24,
what indicates no statistically significant detection of primordial non-Gaussianities yet. However, it is important
to keep in mind that multi-field or exotic single field inflationary models can lead to strongly scale-dependent non-
Gaussianity [3865], which could evade large-scale constraints while still significantly affecting small scales. Such scale-
dependent primordial non-Gaussianities could actually alleviate the S8 tension in the non-linear regime of structure
formation, without affecting the linear regime [3866].
Future CMB surveys, such as those planned with the Simons Observatory, CMB-S4, and the Euclid mission, promise
significantly improved constraints on fNL . The Simons Observatory aims to achieve sensitivity levels that could reduce
the uncertainties on fNL to around ∆fNL ∼ 1. Similarly, CMB-S4, with its planned deep and wide CMB observa-
tions, is expected to further tighten these constraints and provide unprecedented precision in measuring primordial
non-Gaussianities. The Euclid mission, primarily designed for large-scale structure surveys, will complement CMB
observations by probing the distribution of galaxies and the large-scale structure of the Universe, thereby offering
additional avenues to constrain fNL . These upcoming observations are expected to enhance our ability to test a
broader range of inflationary models and deepen our understanding of models that could alleviate the S8 tension.
4.5.3.b. Classes of inflationary models with ramifications for the cosmic tensions There are numerous extensions
or alternatives to vanilla inflation, justified in various ways. In this subsection we summarize the generalities of
202

inflationary models that have a connection to the cosmic tensions, and call attention to the role these specific models
have in ameliorating the tensions.
Higgs inflation In the simplest Higgs inflation model [3867], the Standard Model Higgs field is coupled non-
minimally to gravity through a large dimensionless parameter, leading effectively to an almost flat Einstein-frame
potential able to support inflation for a sufficient number of e-folds (for a review, see Ref. [3868]). One of the
appealing aspects of this scenario is its minimalistic approach, as it does not require the introduction of new fields or
couplings beyond the Standard Model content. In particular, the reheating stage can be computed in detail, leading,
in terms of the number of last e-folds N⋆ , to precise predictions for the spectral tilt ns = 1 − 2/N⋆ ≃ 0.966 and
the tensor-to-scalar ratio r = 12/N⋆2 ≃ 0.0033 [3869–3871]. Additionally, the scenario offers a compelling connection
with accelerator experiments, providing a detailed description of the power spectrum across a vast range of scales
[3872, 3873]. The MCMC analysis of current cosmological data confirms the model’s observational viability and
demonstrates that for the interval N⋆ ∈ [54.5, 60], it can break down the H0 − σ8 correlation [3874]. Specifically,
considering an instantaneous transition to radiation-domination, the H0 tension is reduced to approximately 3σ [76],
constraining H0 = 67.94 ± 0.45 km s−1 Mpc−1 , while the value of σ8 = 0.793 ± 0.003 is in complete agreement with
the KiDS-1000 results [3875]. The constrained S8 = 0.8019 ± 0.0097 is in 2σ with Planck, DES and KiDS-1000 data.
Several extensions of the minimal Higgs inflation scenario have been considered in the literature [3876–3883]. The
so-called Higgs-Dilaton model addresses the current accelerated expansion of the Universe by introducing a unimodular
constraint that manifests itself as the strength of a runaway Einstein-frame potential for a dilaton field [3876]. A
significant feature of this scenario is the existence of specific consistency relations between the spectral tilt of primordial
density perturbations, the tensor-to-scalar ratio, and the DE equation of state [3877].Predictions of these relations are
in tension with direct measurements of H0 using low-redshift SN, but they align well with current CMB observations
and large-scale structure data, offering distinctive signatures that could be tested by future galaxy surveys [3884–3886].
Starobinsky inflation Starobinsky cosmology is renowned as an excellent large-field slow-roll inflationary
model [2064], aligning very well with CMB observations [3887, 3888]. The key feature of the Starobinsky potential,
essential for achieving slow-roll inflation, is its asymptotic behavior, where Vs (ϕ) → const. as ϕ → ∞. Despite its
celebrated success, the model does not seem to align well with proposals that could resolve the cosmic tensions. For
example, it was noted in Ref. [3889] that Starobinsky inflation can hardly coexist with an EDE fraction ∼ 0.06, which
is only able to reduce the H0 tension down to about 3σ.
Brane inflation A theory that stems from the string theory, involving the interaction between branes in higher-
dimensional space. Inflation occurs due to the motion of branes within a compactified extra-dimensional space, such
as in the Klebanov-Strassler throat. The separation between branes acts as the inflaton, and their eventual collision
ends inflation. This scenario can lead to observable levels of non-Gaussianity in the CMB power spectrum and the
distribution of large-scale structures in the Universe, as well as potentially detectable GWs. The dynamics of the
inflaton in brane inflation are governed by the Dirac–Born–Infeld (DBI) action, which is a type of non-canonical
action [3890–3894]. Matrix inflation (a.k.a. M-flation) [3895] may also be viewed as a multi-brane version of the
brane-inflation scenarios.
Tachyon cosmological models of inflation, based on the dynamics of a D3-brane in the bulk of the second Randall-
Sundrum model, have been extensively considered [3896]. Particle creation and reheating in a braneworld inflationary
scenario with a tachyon field were extended to include matter in the bulk. It has been demonstrated how the interaction
of the tachyon with the U(1) gauge field drives the cosmological creation of massless particles and where estimates
have been found for the resulting reheating at the end of inflation [3897]. A holographic braneworld scenario with
a D3-brane located at the holographic boundary of an asymptotic AdS5 bulk [3898] has shown significantly better
agreement with Planck data [3887].
The cosmic tensions have been addressed in several models based on brane inflation. For example, in Ref. [2866] it
was shown that phantom braneworld prefers a higher value of H0 providing a much better fit to the local measurements.
In addition, in Ref. [3899] it was pointed out that the simplest D-term inflation can be consistent with the cosmic
string bound provided by observations of gravitational. Consistency with observations requires a spectral index ns = 1
in accordance with the pre-recombination proposals to alleviate the H0 tension (e.g., EDE).
Quintessential inflation After the discovery of the current cosmic acceleration [48, 2964, 3900, 3901], several
theoretical mechanisms were developed to explain it. Quintessence models broadly encompass scenarios with one or
more scalar fields with appropriate potential/dynamics that can drive the late-time cosmic acceleration, allowing for
much richer possibilities beyond that simple case, in which one gets a dynamical equation of state parameter. An
important feature of quintessential models is the need to examine ghosts and other non-physical features of their
perturbations [1259]. For recent applications in cosmology, see Refs. [3902–3904]. One of the simplest ways to do it is
the so-called quintessential inflation [3905–3911], where the inflaton field is the only responsible for both accelerated
phases. Several authors developed and improved the original Peebles–Vilenkin model [3912–3915], also embedding
it in more fundamental frameworks [3916–3919]. In particular, α-attractors quintessential-inflation parameters can
203

be constrained by stage IV surveys [3920] and they have been tested using Planck and low-redshift data [3921] and
scale invariant models [3886]. Of particular interest here, the quintessential inflation, coming from the Lorentzian
distribution introduced in Refs. [3922, 3923], agrees with the recent observations and is in agreement with the SH0ES
H0 measurement [3924]. Indeed by confronting the modelto cosmological observations, one obtains H0 = 72.25 ±
0.74 km s−1 Mpc−1 at 68% CL [3924, 3925].
K-essence, multifield inflation, and their generalizations K-essence extends the standard model by includ-
ing a non-canonical kinetic term in the Lagrangian, P (ϕ, X), where X = 12 ∂µ ϕ∂ µ ϕ [2060, 2758, 3926–3928]. K-essence
models can help address cosmic tensions through their unique equation of state properties and field dynamics affecting
both early and late Universe evolution. In Ref. [3929] researchers demonstrate how k-essence can modify the expansion
history in ways that specifically target both the Hubble and structure growth tensions. Similarly, Ref. [3930] examines
how an EDE model triggered by the radiation-matter transition can be realized in k-essence. K-essence’s theoretical
flexibility makes it particularly valuable for tension resolution, as shown in Refs. [3931, 3932], where k-essence is shown
to maintain consistency with both observational constraints and theoretical requirements. The multifield dynamics
has been further studied in Ref. [3933]. The generalized k-essence can further contribute to models affecting the sound
speed of perturbations and the equation of state, which can modify cosmological distance ladder and thus influence
tension-related parameters [2766, 3934, 3935].
Uniform five-dimensional inflation Compact extra dimensions can obtain large size by higher dimensional
inflation, relating the weakness of the actual gravitational force to the size of the observable Universe [2908, 3668].
A solution to the horizon problem implies that the fundamental scale of gravity must be smaller than 1013 GeV
which can be realized in a brane-world framework for any number of extra dimensions. However, the requirement of
(approximate) flat power spectrum of primordial density fluctuations consistent with present observations makes this
simple proposal possible only for one extra dimension at around the micron scale [2908, 3936]. Consistency with 2018
Planck CMB data is supported by numerical simulations [3937–3939]. After the end of five-dimensional inflation, the
radion modulus can be stabilized at a vacuum with positive energy of the order of the present DE scale. An attractive
mechanism for radion stabilization, is based on the contribution to the Casimir energy of fields propagating in the
bulk: the graviton, a real scalar field, and three right-handed neutrinos [2883]. The scalar field has a potential holding
two local minima with very small differences in vacuum energy and bigger curvature (mass) of the lower one, and
thus when the false vacuum tunnels to its true vacuum state, the field becomes more massive and its contribution
to the Casimir energy becomes exponentially suppressed [2884]. The tunneling process then changes the difference
between the total number of fermionic and bosonic degrees of freedom contributing to the quantum corrections of
the vacuum energy, yielding a sign-switching cosmological constant, Λs , see Sec. 4.2.2. Despite the fact that when
inflation comes to an end radion stabilization necessitates a dark sector with ∆Neff ∼ 0.25 (i.e., saturating Planck’s
upper limit [192]), in the style of Ref. [811], uniform five-dimensional inflation could help simultaneously resolve the
H0 , S8 , and MB tensions [2687, 2910].
Running vacuum models The RVM approach (see Refs. [2465, 2779, 2780] and references therein), is a frame-
work for an effective description of cosmology, and it has implications for the time evolution of the fundamental
constants [2783, 2788]. The RVM framework [2461] is based upon the idea that the cosmological vacuum energy
densities and pressure are functions of even powers of the Hubble parameter H(t) (and, under circumstances, also
of logarithmic non-polynomial corrections of H(t)). In the cosmic vacuum, which characterises inflation, a de Sitter
equation of state p(H(t)) = −ρ(H(t)), with
Λ(H(t)) 3 H4 
ρ(H(t)) = 2
= 2 c0 + ν H 2 + α 2 + . . . , (4.23)
κ κ HI
is satisfied, although, during each post-inflationary era, the total energy density and pressure is supplemented by
the corresponding contributions of matter and/or radiation, as recent quantum-field-theory calculations in the RVM
context have shown [2776–2778, 3940]. We remark that the coefficients ν and α are considered as constants during
each epoch of the Universe’s evolution; HI is the scale of inflation, which, according to data in Ref. [3887], is currently
set to HI ≲ 10−5 MPl , in order of magnitude. The important feature of the RVM framework is that inflation does
not require any external inflaton fields, but it is mainly due to the non-linear gravitational dynamics due to the H 4
and higher-order terms in the expression of ρ(H(t)) [2463, 2464], which dominate the early Universe eras. Note that
the RVM can provide a smooth evolution of the Universe, and can also account for its thermodynamic history and
properties [2466, 3941].
This generic framework finds a concrete realisation in the context of Quantum Field Theory (QFT) in curved
spacetime when the vacuum energy density is renormalized using the adiabatic renormalization method, in which a
subtraction is performed at an off-shell scale M subsequently identified with the Hubble rate H(t) at each cosmic
epoch. This eliminates the quartic mass contributions ∼ m4 (hence the need for fine tuning) and induces a dynamics of
the vacuum energy of the above form [2776–2778, 3940]. An alternative realisation, termed “Stringy RVM” (StRVM),
204

is found when RVM is embedded in string theory cosmologies, under the assumption [2459, 2461, 2467] that the
early Universe consists of fields appearing in the bosonic massless ground-state multiplet of superstrings. The model
assumes that the dilaton is stabilised to a constant value, through minimisation of an appropriate potential, possibly
induced by string loops [2460]. The pertinent string-inspired, low-energy gravitational theory is a Chern-Simons
(CS) gravity [2457, 3942]. The CS nature of gravity arises from the linear coupling of the string-model independent
axion [2458, 2468] to the CS gravitational anomaly, which is due to the Green–Schwarz counterterms [3943] that
modify the antisymmetric-tensor field strength as a consequence of the requirement of cancellation of gauge and
gravitational anomalies in the extra dimensional string space. In the StRVM, the gravitational anomalies in the
primordial Universe are not assumed to be cancelled, although such a cancellation occurs at the exit from RVM
inflation, after the generation of chiral fermionic matter [2459, 2461, 2467], due to the decay of the RVM vacuum.
The RVM inflation in the StRVM is induced by condensates of the gravitational CS term due to condensation of
primordial GW, which are formed during a pre-inflationary epoch [2461, 2467]. An estimate for the CS condensate
can be provided within a second quantisation of weak GW formalism [2469, 3243, 3244]. It is found to be proportional
to H 4 (t), which during inflation varies very slowly with the cosmic time. As shown in Ref. [2469], the formation of
the CS condensate provides an approximately linear KR-axion monodromy potential, leading to inflation in the sense
that the latter corresponds to a saddle point of the evolution of an appropriate dynamical system of variables. This
result is general and refers to all scenarios with linear axion potentials, such as those obtained, for instance, from
appropriate compactifications of string theories [2471].
Finally, let us notice the intriguing potential implication of this framework on the alleviation of cosmological
tensions [2472, 2473]. Indeed, in the context of StRVM, quantum graviton corrections induce in general corrections
to the effective vacuum energy density of the non-polynomial form H 2n ln(H), n ∈ Z+ . Such corrections have been
computed explicitly in the context of dynamically broken supergravity models [2805], which can also be cast in an
RVM format [3944] and can describe a pre-RVM inflationary epoch of StRVM. In cases such corrections survive in
modern epochs, they can provide a resolution of the Hubble and, under some circumstances, also of the growth-of-
structure tension: H0 = 71.27+0.76 +0.017
−0.73 and σ8 = 0.816−0.015 [2472]. See also the “phantom matter” proposal of Ref. [2473],
inspired by the StRVM approach and providing a potential resolution of both tensions. Subdominant logarithmic
corrections on H can also arise within the conventional RVM, by integrating out matter fields in QFT of fermions
and bosons in expanding spacetimes [2776–2778, 3940].
In a related context to the RVM, we can mention the works from Ref. [3945] and [3946] on the back-reaction effect
of super-Hubble cosmological fluctuations. Ref. [3945] examines a scenario in which a large bare cosmological constant
drives early accelerated expansion and shows that the backreaction effect of infrared fluctuations can reduce it. Such
kind of models are expected to have an impact on the H0 tension problem, as discussed in Ref. [2847].

4.6. The cosmological principle


Coordinator: Eoin Ó Colgáin
Contributors: Alexander Bonilla Rivera, András Kovács, Anil Kumar Yadav, Antonio da Silva, Asta Heinesen,
Daniele Oriti, Dario Bettoni, David Benisty, David L. Wiltshire, Diego Rubiera-Garcia, Dinko Milakovic, Duško
Borka, Ebrahim Yusofi, Elias C. Vagenas, Giuseppe Fanizza, Goran S. Djordjević, Hassan Abdalla, Ido Ben-Dayan,
Iryna Vavilova, Jenny G. Sorce, Jenny Wagner, Jessica Santiago, John K. Webb, José Pedro Mimoso, Juan Garcia-
Bellido, Laura Mersini-Houghton, Leandros Perivolaropoulos, M.M. Sheikh-Jabbari, Maciej Bilicki, Marie-Noëlle Cé
lérier, Marina Cortês, Milan Milošević, Nihan Katırcı, Nils A. Nilsson, Özgür Akarsu, Predrag Jovanović, Saurya Das,
Tajron Jurić, Víctor H. Cárdenas, Valerio Marra, Vesna Borka Jovanović, and Wojciech Hellwing

In recent years the status of the cosmological principle (CP), a fundamental assumption in modern cosmology, has
become less clear cut. However, what stymies debate is the absence of a consensus definition of the CP. The CP is
commonly presented as the idea that the Universe is isotropic (the same in every direction for a comoving observer
who is at rest with respect to the cosmic fluid of the expanding Universe) and homogeneous (the same at every point
on a hypersurface corresponding to a given cosmic time) at suitably large scales. Homogeneity implies that there
is no special location, such as a preferred observing position or center, while isotropy states that there is no special
direction, such as an axis. These two statements do not automatically imply one another, and the possible departures
from each of them are currently among the most prominent research topics in cosmology.
Unfortunately, the terminology is loose and it is always possible to find a definition that is perfectly untestable.
First, what is meant by large scales is rarely discussed. Secondly, given that we have only one vantage point in
the Universe, testing homogeneity is difficult. In particular, depending on how one defines the latter, one could end
up with a claimed “large scale” varying from ∼ 70h−1 Mpc [3947–3949] up to ∼ 260h−1 Mpc [3950]. Thirdly, even
when scientists argue that there is an anomaly with the CP, great care is required to separate ΛCDM predictions
from generic predictions of isotropic/homogeneous Universes. For example, ∼ 260h−1 Mpc [3950] is a claimed upper
bound for the homogeneity scale in a ΛCDM universe. Moreover, homogeneity is also challenged by large structures
205

[1600, 3951–3954], but one inevitably must compare to the predictions of a specific FLRW model, i.e., ΛCDM, and
often, one either finds comparable structures in simulations or one can question the statistics [1604, 1605, 3955–3957].
Finally, a point very rarely discussed, and which may pass by as a minimal assumption, is the very existence of
a “cosmic fluid”. Besides how harmless those words might sound, what this actually assumes is the existence of a
unique reference frame which is at rest with respect to the average distribution of the CMB, baryonic matter, DM
and DE simultaneously in different time periods of the history of the Universe. Looking closely at this definition,
one then opens Pandora’s box, in which to start, we have a lack of proper and robust definition of averages in GR,
followed by questions regarding averaging scales and even if the averaging scales should be the same or not for different
fluids.61 Admittedly, these are excellent theoretical questions, but if one approaches cosmology at a granular scale,
constructing an observationally testable model is impossible. In cosmology one should not lose track of the grossly
simplified assumptions - justifiable on the relatively poor quality of astronomical data - being made.

(a) HST Key Project (b) Galaxy cluster scaling

FIG. 79: Left: Angular variations of H0 on the sky in galactic coordinates in the aftermath of the HST Key
Project. Reproduced from Figure 1 (a) of Ref. [3958]. Right: Angular variations of H0 on the sky in galactic co-
ordinates from galaxy cluster scaling relations. Reproduced from Figure 7 of Ref. [3959].
The Universe is a gravitating system. Gravity is GR, at least in a minimal setting. This necessitates a choice for a
spacetime metric gµν to insert on the left hand side of Einstein’s equation. Here the modus operandi is to start from
the most symmetric possibility. This leads one to a maximally symmetric spacetime in 4D, namely Minkowski, anti-de
Sitter, and de Sitter metrics. From these three, the latter is the only physically relevant possibility and appears in
the asymptotic future of any universe with late-time accelerated expansion. The pursuit of symmetry also leads one
to the perfect cosmological principle (PCP), according to which the Universe is homogeneous and isotropic in space
and time. This means that the Universe in the large is also unchanging with time. The PCP was the basis of the
steady-state cosmology, an alternative to the Big Bang theory [3960, 3961], but this required a physically contrived
scenario whereby the density of matter remained unchanged due to a continuous creation of matter.
4.6.1. FLRW spacetime
In pursuit of a more physical model, one can always pursue classes of less restrictive spacetimes. The concession
one makes is reducing maximal symmetry in the 4D spacetime to maximal symmetry in a 3D space. This leads one
to the FLRW metric, a cornerstone of modern cosmology. The FLRW spacetime metric gµν is given by Ref. [3732]
dr2
 
2 µ ν 2 2
ds = gµν dx dx = −c dt + a(t) 2 2 2 2 2
+ r (dθ + sin θ dϕ ) , (4.24)
1 − kr2
where a(t) is the scale factor, k represents the spatial curvature, and (t, r, θ, ϕ) are the comoving coordinates. Combin-
ing FLRW with GR, the Einstein equation reduces to the Friedmann equation, 62 which has an important consequence
that the Hubble constant H0 = H(z = 0) arises as an integration constant [3723]. Note, Eq. (4.24) simply specifies the
spacetime and not the model. However, once one fully specifies the model, e.g., ΛCDM, the onus is on the astronomy
and cosmology communities to demonstrate that H0 is a constant in cosmic time or redshift. If H0 is not constant in
a given model, one runs into a “Hubble tension” problem, e.g., see Refs. [34, 192].
This then brings us to the simplest way to test the metric in Eq. (4.24). H0 can be determined model independently
in the local Universe z ≲ 0.2 using Cepheids, TRGB, or other possible primary distance indicators in order to calibrate

61 Applying the CP for a void-dominated cosmology [3685–3688], involving superclusters and cosmic voids, necessitates that cosmic voids
tend towards achieving maximum symmetry. The result of such an evolution would be a vast spherical cosmic void, which is formed
by the merging of smaller sub-voids (void-in-void process) [1782, 1791, 3691]. The density of the largest cosmic voids will be closer and
closer to that of the Universe, making the CP more applicable to the present void-dominated Universe.
62 The Friedmann equation can be derived by moving to the accelerated frame using the scale factor [3962, 3963].
206

SNIa [34, 144]. Note, working locally, one can safely decouple the cosmological model [220]. In short, one should
make sure that there are no statistically significant angular variations of H0 between different directions on the sky,
because as discussed, H0 must be a constant. Already, galaxy cluster scaling relations point to ∼ 9% changes in H0
in the local Universe [3959, 3964, 3965]. Evidently, if the precision on H0 is ∼ 5 − 10%, this is no problem, but it
may become problematic if one pursues ∼ 1% precision. Since nobody builds the distance ladder on galaxy clusters,
it is imperative to recover this result in SNIa. Preliminary results exist, but given the relatively small size of SNIa
samples, which become worse when restricted to lower redshifts, angular H0 variations at low redshifts are consistent
with statistical fluctuations (< 2σ) [1339, 1369] (see also Refs. [1375, 1380, 1385, 3958, 3966]).63 Given the better
statistics in cosmicflows (CF) data, a data set comprising SNIa alongside other distance indicators, the observation
of an anisotropic H0 has recently been pushed to 3.9σ [3970]. It should be stressed that there is no guarantee that
one can get to a > 3σ variation in H0 from SNIa alone, but the test can and should be done. Note, a statistically
significant angular variation in H0 would make both the Hubble constant and the Hubble tension ill-defined. Since,
S8 is a parameter within the ΛCDM model, and the model builds upon FLRW, a breakdown of FLRW would also
make S8 tension ill-defined 64 .
In addition, locally, there are also claims from the CF program that there are anisotropic bulk flows [1338], which
may no longer converge to ΛCDM expectations [1421] (see also Refs. [1420, 1422]), thereby echoing a kinematic SZ
anomaly, one disputed by the Planck collaboration [3972], from a decade previously [1424, 1426, 1427, 3973]. An
upgrade of CF data to include WALLABY [391, 3974, 3975], FAST [393] and DESI data [394] recently found that
the homogeneity scale is not yet recovered at 300h−1 Mpc [3976], thereby strictly violating the 260h−1 bound [3950].
In tandem, the inferred bulk flow was larger than ΛCDM predictions, but remained consistent with the model. Note,
in Ref. [1421] an isotropic expansion or constant H0 is assumed, but the bulk flow becomes anomalously large. CF
data makes use of a host of distance indicators, including SNIa, so one can in principle recover the same result from
SNIa alone.65 Here it is worth noting that SNIa redshifts are typically corrected for peculiar velocities of the host
galaxy, so it is imperative to make sure that anomalously large bulk flows are not simply swept under the rug to
reinforce an isotropic expansion (see criticism of SNIa peculiar velocity corrections in Refs. [3977, 3978]). Another
cautionary note on making peculiar velocities corrections regards the non-linearities which arise on local scales. This
is an extremely delicate subject, especially given that the boundary scales where linear and non-linear regimes are
applicable, together with the regime where perturbation theory is no longer valid, are still under investigation. In
order to avoid this point, one must go to high redshift sources.
The other way to test the CP model independently is through aberration and Doppler effects with a large sample of
sources. Although the Ellis-Baldwin test [1357] was outlined last century, the test only became feasible this century
[1393, 1399–1402, 1405–1409, 1414, 3979, 3980]. While there may be outliers [1409, 1412, 1413, 1458, 3981, 3982],
there is a growing consensus that radio galaxy and RGB data sets do not inhabit the same rest frame as the CMB
[1393, 1400–1402, 1406–1408, 1414, 3979, 3980]. Nevertheless, these are challenging studies, both technically and
sociologically, thus it is always conceivable that observational [1397, 1458, 3982, 3983] and or theoretical systematics
[1355, 3984] (however see Ref. [3985]) are at play. The earliest studies have typically employed frequentist statistics,
but later papers have provided a Bayesian interpretation [1395, 1414, 1458]. Taken at face value, the disagreement
between CMB and radio galaxies/RGBs on the cosmic dipole points to a breakdown of the CP – and, therefore, of
FLRW. The problem is interpreting this result. At what scales or redshifts does this happen? How badly is FLRW
broken? None of this would be obvious without corroborating FLRW anomalies in the local Universe.
So where are we now on FLRW? Both in the local Universe z ≲ 0.2 and large samples of radio galaxies/RGBs with
assumed median redshifts of z ∼ 1 we see hints of a violation of FLRW. Given the rich structures in the local Universe
[1338, 3986, 3987], variations in H0 , or alternatively bulk flows, are expected. See Fig. 79. The important question
is whether these effects deviate from the expectations of the ΛCDM model. On the other hand, a mismatch in the
cosmic dipole, could easily be explained by a relatively local effect and not an FLRW violation at large scales. Thus, it
is of interest to study the Hubble diagram out to z ∼ 1 in order to check if one recovers an FLRW universe. One could
adopt the ΛCDM model, e.g., see Refs. [1380, 1398, 3966, 3988], and check that variations in H0 become less significant
as effective redshifts of SNIa samples, etc, increase, but one can also tackle this problem model independently through
cosmography [2076] up to z = 1, the radius of convergence of the expansion [2115]. See Refs. [1377, 3989] for
preliminary studies in this direction. Note, cosmography introduces additional parameters, especially if one wants
to describe models beyond FLRW [1353], so one will require large data sets to conduct these studies. Nevertheless,
the motivation should be clear. Current hints of FLRW violation at low and high redshift need to be connected

63 One can also fit a dipole to SNIa [1386, 3967–3969]. One may find [3969] that SNIa and CMB are not in the same frame, but that the
difference can be interpreted as a bulk peculiar velocity vbulk consistent with ΛCDM expectations. Note, these analyses are typically
biased towards the lowest redshift SNIa, so one needs to study the redshift dependence of vbulk in shells, where overlap is expected with
the cosmicflows program [1421].
64 As an aside, since an anisotropy makes a contribution to the energy budget of the Universe scaling as a−6 , which is relevant in the
early Universe, it is plausible a new anisotropic cosmological model can be found that alleviates the tension with larger local H0
determinations, e.g., see Refs. [1359, 3971].
65 Indeed, restricting the CF data to SNIa, one confirms the large bulk flow, but large errors make this consistent with ΛCDM expectations.
We thank Rick Watkins for private communication.
207

in a common framework. The big question that needs to be addressed is how good is FLRW as an approximation
to the physical Universe? While few scientists outside the field of cosmology, especially relativists, would challenge
the assertion that FLRW is an approximate symmetry, since there is no argument that it is fundamental, this may
be cold comfort for precision cosmology. ∼ 9% variations in H0 [3959, 3964, 3965], suggest that FLRW is a good
approximation to ≲ 10%. This is easy to square with the perceived isotropy of the CMB, when one factors in residual
asymmetries in CMB data or CMB anomalies [1325]. A subset of the latter has persisted through three independent
satellite experiments, namely COBE [676], WMAP [2740] and Planck [192], allowing one to make the case that the
Universe is unlikely to be statistically isotropic based on CMB data alone [1327]. Unsurprisingly, H0 can vary by
up to ∼ 10% along the axis of the hemispherical power asymmetry [1328, 1329], a recognized CMB anomaly, and
attempts to recover the CMB dipole from aberration and Doppler boosts are contaminated at lower multipoles (larger
scales) by this anomaly (see Fig. 3 of Ref. [3990]).66 In addition, it is evident that WMAP/Planck CMB data prefers
a phenomenological Bianchi model over ΛCDM [3993–3996],67 which is only possible if there is a residual anisotropy
in the data. Nevertheless, it is not clear if beyond FLRW models, especially the models in Sec. 4.6.2, can rival FLRW
cosmology. This is not a statement about fits to data, this is the statement that at a purely technical level one would
need a complete understanding of cosmological perturbation theory in these models. Given the anomalies in this
white paper, it is hard not to contemplate that precision cosmology, defined through models where fitting parameters
are constrained at the ∼ 1% error, may be at a crossroads. See Ref. [1323] for a recent overview of observations and
the key claim that we have exceeded the precision where the approximate nature of the FLRW assumption is evident.
4.6.2. Moving beyond FLRW
In light of these FLRW anomalies, there is a growing interest in exploring more general metrics that relax the
assumptions of homogeneity and isotropy. In this respect, if the CP is relaxed so that the homogeneity holds but not
isotropy, then the corresponding exact solutions of Einstein’s field equations are given by the class of cosmological
models called Bianchi universes, named after the Italian mathematician Luigi Bianchi who first classified these spaces
[1359, 3997–3999]. Homogeneity in space implies that Einstein’s field equations reduce from partial to ordinary
differential equations in time, making Bianchi models exact solutions. A complete list of all such solutions for all
Bianchi cosmological models from type I to type IX and for perfect fluid is given by Ref. [4000]. Bianchi universes also
contain, as a subclass, the standard isotropic FLRW universes. Although Bianchi models were previously considered
to be inconsistent with observations, recent studies highlighted above challenge the isotropy assumption and revive
interest in these models, which therefore remain widely-studied [4001, 4002]. Ref. [1359, 3999] present an observational
analysis of Bianchi type-I, -V, and -XI spacetime extensions of the ΛCDM model. This section discusses some of the
alternative metrics that have been proposed to address these issues.
4.6.2.a. Lemaître-Tolman-Bondi (LTB) metric The LTB metric [4003–4005] describes a spherically symmetric
but radially inhomogeneous universe. It is given by
R′ (r, t)2
ds2 = −c2 dt2 + dr2 + R(r, t)2 (dθ2 + sin2 θ dϕ2 ) , (4.25)
1 + 2E(r)
where R(r, t) is the radius of the spherical shell at time t and radial coordinate r, and E(r) is an arbitrary function
related to the energy of the shells. The LTB metric allows for variations in the density and expansion rate along
different radial directions, providing a more flexible framework to model inhomogeneities in the Universe [4003–4005].
The series of N -body simulations discussed in Ref. [4006] explores the development of large-scale structures against
a LTB background that incorporates a cosmological constant, specifically within the ΛLTB model framework.
Spherical inhomogeneities in LTB models can emerge either due to matter voids or due to Hubble scale spherical
inhomogeneities of DE. In the latter case, such inhomogeneities can be supported by topological considerations,
particularly in the context of recently formed global monopoles. These topological defects arise when the vacuum
manifold of the quintessence scalar field has non-trivial π2 topological properties [4007, 4008]. This scenario corre-
sponds to the “Topological Quintessence” class of models, which may support deviations from the FLRW metric on
cosmological scales.
In the context of topological quintessence, a global monopole formed during a recent phase transition with a core
size comparable to the present Hubble scale could induce the observed accelerated expansion of the Universe. The
monopole’s scalar field is trapped near a local maximum of its potential in a cosmologically large region of space. This
setup leads to an inhomogeneous but isotropic DE distribution, where the core of the monopole exhibits an effective
cosmological constant-like behavior, while the outer regions revert to a matter-dominated Einstein-de Sitter universe.

66 For this reason, it is customary to remove large scales, e.g. see Refs.[3991, 3992].
67 These phenomenological Bianchi models cannot be seen as deformations of the ΛCDM model since the cosmological parameters adopt
different values.
208

An off-center observer within such a topological quintessence framework would naturally observe cosmic dipoles
due to the asymmetric positioning relative to the monopole core. This asymmetry can manifest as dipoles in various
cosmological observations, including the CMB and large-scale velocity flows [1356, 4009, 4010].
Recent numerical simulations have explored the dynamics of such global monopoles minimally coupled to gravity
in an expanding universe with initially homogeneous matter. These studies show that when the energy density of the
monopole core starts dominating the background density, the spacetime in the core begins to accelerate its expansion
in accordance with a ΛCDM model with an effective inhomogeneous spherical DE density parameter ΩΛ (r) [1435].
Away from the core, the Universe appears as an Einstein-de Sitter Universe, while near the core, ΩΛ (r) reaches a
maximum.
These findings suggest that topological quintessence models could provide viable explanations for certain cosmo-
logical observations and anomalies. The key is that the presence of a large-scale topological defect could naturally
introduce the observed anisotropies without requiring exotic modifications of gravity or the introduction of multiple
new fields.
4.6.2.b. Bianchi metrics Bianchi models [3997, 4011–4016] generalize the FLRW metric by allowing for anisotropies
while maintaining homogeneity. They are classified into nine different types based on their symmetry properties. For
example, the Bianchi I metric is given by

ds2 = −c2 dt2 + a(t)2 dx2 + b(t)2 dy 2 + c(t)2 dz 2 , (4.26)

where a(t), b(t), and c(t) are scale factors along the x, y, and z axes, respectively. These models can describe universes
that have directional dependencies in their expansion rates, which might help explain certain observed anisotropies
[1323, 4017–4019]. See Ref. [1359, 3999] for observational analysis of Bianchi type-I extension of the ΛCDM model.
4.6.3. Mc-Vittie spacetime
The McVittie metric is the exact solution of Einstein’s field equations describing a black hole or a massive object
immersed in an expanding cosmological spacetime [4020, 4021]. The metric can be written by

dr2
 
2
ds2 = −(1 − ΦN ) dt2 + a (t) + r2 dΩ2 , (4.27)
1 − ΦN

where ΦN = 2GM/(rc2 ). In the low-energy limit (ΦN ≪ 1 and rH ≪ 1) the equation of motion reads [4022–4024]:
r̈/r = −GM/r3 + ä/a . For the de Sitter–Schwarzschild metric the ä/a changes to Λc2 /3. For z ≈ 0.67 the ä/a
acts as an attractive force for early cosmic times while for the de Sitter–Schwarzschild metric only the constant term
Λc2 /3 remains. Galaxy groups in the local Universe are bounds on the interplay between the expansion force and the
Newtonian attraction [4025]. This gives another setup to determine the Hubble constant [4026–4032].
4.6.3.a. Inhomogeneous general relativity solutions Beyond the specific cases of LTB and Bianchi metrics, there
are more general inhomogeneous solutions in GR that do not assume any specific symmetry. These solutions can be
constructed using various techniques, such as analytic or perturbative methods or numerical relativity. One notable
example of an exact solution of the GR field equations, obtained with a dust gravitational source, is the Szekeres
model, which generalizes the matter (+ cosmological constant) dominated FLRW solution by dropping any symmetry
in the metric

ds2 = −c2 dt2 + e2α dx2 + e2β (dy 2 + dz 2 ) , (4.28)

where α and β are functions of both spatial and temporal coordinates. These models allow for full inhomogeneity and
anisotropy, providing a more general framework to describe the Universe [4033]. The Szekeres models can be divided
into three classes: quasi-spherical, quasi-plane and quasi-hyperbolic. The quasi-spherical Szekeres class of solutions,
which includes the LTB model as a subclass, can therefore be viewed as a generalisation of this LTB model where
the spheres of constant {t, x} coordinates are non-concentric and exhibit each a given mass dipole whose direction
is rotated from one sphere to the other. However, the surfaces x = const within a space t = const might be quasi-
spherical in regions of the space and quasi-hyperbolic elsewhere, with a zero curvature at the boundaries. Every class
possesses the FLRW solutions as a limit for particular forms of the functions defining its metric which can be obtained
at some large scale for any of the classes, quasi-spherical, quasi-plane and quasi-hyperbolic. Hence, the possibility to
use the Szekeres model for representing the inhomogeneous late Universe with a transition to homogeneity (FLRW)
at some larger scale. These models have been used to study various cosmological phenomena, including the formation
of large-scale structures and the impact of inhomogeneities on the CMB. Now, in the era of precision cosmology, their
application to the analysis of the largest cosmological data sets will have to be performed [4034].
209

4.6.3.b. Fully model independent cosmographic approach Another possibility to test the CP in a fully covariant,
model independent way without imposing any a priori metric, is using the general cosmographic approach developed by
Kristian and Sachs [4035] (see also Refs. [1353, 1356, 4036]). Its only assumption is that the galaxies can be described
in terms of a pressureless dust fluid following flow lines that can be described in terms of a unique congruence – an
assumption made by all of the other theoretical approaches as well. The idea then lies on the fact that the matter
flow, and its kinematical parameters – the expansion Θ, the shear σab and the vorticity ωab – are a direct probe
to the expansion of the Universe. One then can Taylor expand the generalized luminosity distance in redshift and
obtain the generalized Hubble and deceleration parameters, which are now naturally directional dependent quantities.
For an observer boosted with respect to the matter frame (one must always assume this is the case and directly
measure the boost velocity), the generalized Hubble parameter has been proven to contain a monopole, a dipole
induced only by the boost velocity, and a quadrupole component [1356]. This approach also provides a fully covariant
non-perturbative generalization of the perturbative result obtained by Ref. [4037] in which they develop a way to
disentangle the intrinsic dipole from a kinematically originated one [1356], a very important need in order to test the
CP.
4.6.3.c. Timescape cosmology A new study of the Pantheon+ catalogue has now claimed that an inhomogeneous
alternative [4038–4041] to the models discussed above fits better than ΛCDM with strong to very strong Bayesian
evidence (ln B ∼ 3–5) [3741]. Positive evidence (ln B ∼ 1–2) remains even when restricted to SNIa with z > zmin =
0.06. In contrast to models based on a single metric, timescape combines small scale FLRW geometries—regionally
valid on scales of ∼ 3–30 Mpc—via a Buchert average [4042–4044]. A closure condition is needed to supplement the
Buchert equations, without which its physical interpretation is open to debate [1595, 4045]. Timescape extends the
Strong Equivalence Principle to a Cosmological Equivalence Principle to apply on small scales over which average
isotropic motion in empty space is operationally indistinguishable from average isotropic expansion in nonempty space
[4046], leading to the quasilocal Hubble expansion condition as a closure relation for Buchert averages. The predicted
variance in local Hubble expansion can be calibrated [4040] and tested observationally, leading to a potential resolution
of the Hubble tension [3740], and a natural framework for resolving dipole anomalies in terms of small scale non-
kinematic differential expansion [4047]. This appears to be consistent with new analysis of void statistics in numerical
relativity simulations using the full Einstein equations [1597].
The timescape and ΛCDM expansion histories differ by ∼ 1–3% over small redshift ranges, but can be distinguished
with a long enough lever arm. Independent projections made for the Euclid mission in 2014 [4048] show that with
Euclid BAOs plus 1000 independent SNIa distances, the FLRW expansion history and the timescape alternative can
be definitively tested via the Clarkson–Bassett-Lu (CBL) test [4049]. The new Pantheon+ results for SNIa [3741] have
been independently confirmed independently with the DES survey [4050], where it was found that for events with
z > zmin = 0.033 timespace is preferred over ΛCDM with ln B = 1.7. Based on simple geometric scaling arguments,
the DES analysis [4050] finds that BAOs strongly favour ΛCDM over timescape. However, in timescape the BAO
must be extracted directly from raw galaxy clustering data [4051], and independently recalibrated from the CMB
anisotropy spectrum and the sound speed in the primordial plasma. The nonbaryonic to baryonic matter densities
ratio for timescape still has large uncertainties [4040]. Thus reanalysis of the matter matter model in the early
Universe in conjunction with constraints from a variety of forthcoming datasets is an urgent goal for implementing
the CBL test to definitively decide between FLRW and timescape by 2030.
4.6.3.d. Constraints from observations Detailed cosmological observations have imposed constraints on some of
the alternative models described above. For instance, the Generalized LTB model with inhomogeneous isotropic
DE has been studied to understand its consistency with observations such as the Union2 SNIa data and the CMB
multipoles [1430]. More recently, Ref. [3739] confronted ΛLTB models to a host of data sets finding that the models
could not resolve Hubble tension. It has been shown that for such models to be consistent with observations, the size
of the inhomogeneity must be large, typically on the order of a few Gpc. Additionally, the observer must be located
relatively close to the center of the inhomogeneity to avoid large dipole anisotropies.
The exploration of these alternative metrics is crucial for addressing the current tensions in cosmology. By consid-
ering more general spacetimes, we can test the robustness of the CP and potentially uncover new physics that could
reconcile discrepancies between observations and the standard model. As high-precision data continue to pour in, the
development and testing of these models will be an essential part of the future of cosmology.
4.6.4. What needs to happen going forward
Even though a more precise definition of the CP took a little while to be developed [4052] following the pioneering
works of Vesto Slipher [4053], Henrietta Swan Leavitt [4054], Georges Lemaître [4055] and Edwin Hubble [4056], it
was formulated in a time when, more than a principle, it was a need for cosmology to go forward. Its historical
importance and scientific contribution to the advancement of cosmology is undeniable. But, given the amount of
present and future coming data, continuing to apply it without a proper investigation of its validity and limitations
is no longer a matter of science, but faith.
210

Here we develop a brief list of the necessary requirements in order to properly test the CP:
• Robust data with large sky coverage, in order to test for possible directional dependent effects;
• Model independent techniques for cleaning up possible foreground contamination which does not assume (or
imposes) the CP;
• Model independent analysis and parameter estimation;
• The practice of clearly and forthrightly stating all of the (explicit and implicit) assumptions made throughout
the analysis;
• A robust model independent theoretical method connected to observations, in a way that allows for the inter-
pretation of the data.
Note that angular variations of H0 on the sky and the cosmic dipole can both be studied independently. The former
will benefit from large SNIa samples with excellent sky coverage from ZTF, Vera Rubin observatory, and Romans
space telescope. The SKAO will provide large radio galaxy samples allowing a definitive conclusion on the cosmic
dipole anomaly [1403, 4057].

4.7. Quantum gravity phenomenology


Coordinator: Giulia Gubitosi
Contributors: Christian Pfeifer, Elias C. Vagenas, Gaetano Lambiase, Manuel Hohmann, Nikolaos E. Mavromatos,
Saurya Das, and Vasiliki A. Mitsou

One might reasonably conjecture that at least some of the MG models discussed in the previous sections of this review
might emerge in an appropriate limit from quantum gravity. And in fact, for some of the models the connection is more
apparent, as is the case, for example, of Hořava-Lifshitz gravity [4058], that introduces anisotropic scaling between
space and time at high energies. This leads to a power-counting renormalizable theory of gravity that deviates from
GR at short distances. This model has been proposed by Hořava in Ref. [4058], where an effective Quantum Gravity
approach not requiring the Lorentz invariance at fundamental ultra-violet scales has been formulated. This invariance,
however, emerges at large distances. It mainly aims to solve the high-energy issues suffered by GR through a spacetime
foliation capable of reproducing the causal structure out of the quantum regime. Basic foundations and applications
of this approach can be found e.g., in Refs. [3567, 4059–4069]. This model is potentially capable of addressing the
H0 tension, as demonstrated in Ref. [4070]. Specifically, the authors obtain a positive result on the cosmic tensions
between the Hubble constant H0 and the cosmic shear S8 due to a shift of H0 towards a higher value. Moreover, in
Ref. [4071] the authors show that up to 36% of the Hubble tension can be explained by Lorentz-violating effects in
a Hořava–Lifshitz scenario. This, in fact, is a common feature of theories involving gravitational Lorentz violation,
where local G no longer equals cosmological G [4072].
However, a full-fledged fundamental theory of quantum gravity is still elusive. This motivates the adoption of a
bottom-up approach, that is complementary to the top-down approach attempting to formulate fundamental theories
and then working out their predictions in specific limits. In this bottom-up approach, possible features of the quantum
interaction and dynamics of gravity as well as particles and fields propagating on a quantum spacetime are described
at an effective level via phenomenological models. This field of research, called quantum gravity phenomenology,
successfully leads to observable predictions whose confirmation or constraints serve as guidelines for building a fun-
damental theory of quantum gravity, as it is explained in detail in the review in Ref. [3050] and white papers in
Refs. [3051, 4073].
In the following, we will outline how phenomenological models of quantum gravity can affect the understanding of
cosmological tensions and how cosmological observations may lead to the discovery of quantum aspects of gravity.
4.7.1. QG modified gravitational dynamics
4.7.1.a. Hubble tension and generalized uncertainty principle The GUP was originally motivated by consider-
ations coming from string theory [4074] and from the analysis of the relation between gravitational and quantum
theories in black hole physics [4075]. Subsequent work found relations to noncommutative geometry [4076, 4077]. Re-
cently, in Ref. [4078] it was shown that the deformation parameter entering the GUP can be related to the coefficients
of the standard model extension (SME) in the gravity sector [4079] and in this case stringent bounds on the GUP
parameter can be inferred from SME parameters.
The idea of GUP was suggested as a possible explanation of the Hubble tension in Refs. [4080, 4081]. GUP
introduces Planck-scale corrections to the standard relations between canonical variables, due to the interplay between
the quantum theory and general relativity. Therefore, it is expected that such GUP corrections will be relevant during
211

the very early/Planck epochs of cosmology, and leave their fingerprints on the quantum fluctuations. As these GUP-
modified quantum fluctuations propagate in a cosmological spacetime, they affect primordial fluctuations during
cosmic inflation. Since these fluctuations are encoded in the CMB anisotropies [4082, 4083], one expects to read the
GUP corrections in the CMB power spectrum. In particular, one can select the cosmological spacetime to be the
isotropic and homogeneous FLRW Universe with Hamiltonian, in natural units, of the form [4084, 4085]

p2a
HFLRW (pa , a) = N + 6N ka − N ρa3 + κΠ , (4.29)
24a
where a and pa are, respectively, the scale factor which plays the role of the generalized coordinate operator and
the generalized canonical momentum conjugate to the scale factor. Note that N is the lapse function and Π is its
conjugate momentum, while κ is its corresponding coefficient. Then, one introduces the GUP-modified canonical
momentum [4086–4089]
 
Pa = pa 1 + λ1 pa + λ2 p2a + O(p3a ) , (4.30)

and substitutes it in the Hamiltonian HFLRW (pa , a), in order to obtain the GUP-modified Hamiltonian (keeping terms
up to 2nd order in momentum)

1 p2a (1 + 2λ1 pa + 2λ2 p2a + λ21 p2a )


GUP
HFLRW (pa , a) = + 6ka − ρa3 + κΠ . (4.31)
24 a
By combining the GUP-modified Hamiltonian equations, one finds the GUP-modified Hubble function to be [4090]
h i1/2
HGUP = H 1 + 48λ1 a2 H + 864λ2 + 576λ21 a4 H 2 (4.32)

,

where H is the standard Hubble function.


As already mentioned, CMB will include signatures of the GUP corrections, and thus, the fingerprints of quantum
gravity. Therefore, on the one hand, the GUP-modified Hubble parameter, i.e., HGUP , can represent the one obtained
by the Planck collaboration which utilises the CMB data [192], HCM B . On the other hand, the unmodified Hubble
parameter H, can be assumed to be the one obtained from the HST which utilizes the SNIa data [48]. In this case,
we dub it HSN . Based on this, the above expression for the GUP-modified Hubble parameter becomes
h i1/2
HCM B = HSN 1 + 48λ1 a2 HSN + 864λ2 + 576λ21 a4 HSN
2
(4.33)

.

It is evident that the GUP can, in principle, provide at least a partial explanation of the Hubble tension problem.
Detailed analyses are currently ongoing.
4.7.2. Quantum gravity effects on the physics of particles and fields
4.7.2.a. Propagation of particles and fields on quantum spacetime One intensively studied aspect of the interac-
tion between quantum spacetime and particles and fields is a modification of their propagation properties. Usually,
this is studied in terms of a modified dispersion relation (MDR) of the particles, which encodes a modified light cone
and mass shell structure. One distinguishes two cases: the Lorentz invariance violating (LIV) case [4091], where just
an MDR is considered, and the deformed relativity (DSR) case [4092, 4093], where the MDR is supplemented by a
modified energy-momentum conservation law and deformed (most often non-linear) Lorentz transformations between
observers which leave the MDR invariant and transform the deformed energy-momentum conservation in a covariant
way.
In the context of cosmology, the most interesting consequence of an MDR is a different time of arrival of photons
of different energy, when they are emitted simultaneously at the same spacetime event at redshift z [4094], such as
GRBs [4095, 4096] or AGNs [4097]. Such a time delay can be thought of analogously as to what happens when
electromagnetic radiation propagates through an optical non-trivial medium and photons of different energy are
affected differently by the medium.
In general, such an MDR for photons can be parametrized in the form [4098]

E 2 = p2 (1 + f (E, p, EQG , z)) , (4.34)

where f is a function that parametrizes the Planck scale modification in terms of the energy E of the photon at
emission, its comoving momentum p = P/a(t), the quantum gravity energy scale EQG (which might or might not be
212

the Planck scale) and possibly the redshift. For two different photons, emitted with an energy E1 and E2 , to leading
non-vanishing order in EQG , this leads to a time delay of the form

1 E2n − E1n
∆tQG = t2 − t1 = n κ(z) , (4.35)
H0 EQG

where κ(z) is the redshift distance function to the source at redshift z. It depends on the choice of the MDR model,
i.e., in the choice of the function f and on the choice of the cosmological model and its parameters. Some examples
are
• non-critical string-inspired models with redshift dependent EQG [4099–4101] or the Jacob-Piran model [4102]
´z ′ n
with constant EQG , with κ(z) = 0 E n(1+z )
(z ′ )H(z ′ ) dz .

QG

• For n = 1 in Eq. (4.35), DSR´ models like κ-Poincare in the bicrossproduct basis [4103], where the redshift
z
distance function reads κ(z) = 0 dz ′ (z′ +1)H(z
1
′ ) or in more general DSR models in FLRW spacetime, where the

redshift distance function is described by the three-parameter (η1 , η2 , η3 ) model [4104]


´z ´ z′ dz′′
   2   4 

H(z ′ )I(z ′ ) H(z ′ )I(z ′ )
 
κ(z) = 0 dz ′ (z +1)
H(z ′ ) η 1 + η2 1 − 1 − z ′ +1 + η3 1 − 1 − z ′ +1 , I(z ′ ) = 0 H(z ′′ ) .

Taking into account these time delay effects in the analysis of the Hubble tension might lead to an alleviation of
the present tensions [4105–4107]. More detailed analyses remain a future prospect.
4.7.2.b. Interactions of particles and fields on quantum spacetime Another relevant feature related to quantum
gravity effects on particles concerns modifications of relativistic interactions, that might induce changes to the pre-
dictions of the particle content of the Universe.
In LIV models, the combination of MDR with the expected standard conservation of energy and momentum gives
rise to strong modifications to interaction thresholds [4091], with effects for example on the opacity of the Universe
to high-energy particles [4108, 4109].
Going beyond LIV, an important ingredient for self-consistent DSR models is a modified energy momentum con-
servation. In a process in which two particles with 4-momenta p and q collide, the center of mass energy is not given
by their simple sum of their momenta, since this would not be invariant under modified Lorentz transformations, but
given by a modified addition law, to first order of the form [4093]

1
(p ⊕ q)µ = pµ + qµ + f (p, q)µ , (4.36)
EQG

where the precise expression of f (p, q)µ depends on the model under consideration. In general, it is a non-linear
function of p and q.
This modified energy-momentum conservation law leads as well to modifications in relativistic threshold reactions,
however they would be much weaker than in the LIV case, typically only relevant for particles of Planck-scale energy.
While these effects have no direct link to cosmological tensions, their concomitant presence with the propagation effects
discussed above might help constrain the specific form of MDR modification that is compatible with observations.

4.8. Varying fundamental constants and their role in the Hubble tension
Coordinator: Jens Chluba
Contributors: Catarina Marques, Dan Grin, Gabriel Lynch, Leo Vacher, Nils Schöneberg, Ruchika Kaushik, and
Vitor da Fonseca

Fundamental physical constants need not be constant, neither spatially nor temporally. This seemingly simple state-
ment has profound implications for a wide range of physical processes and interactions, and can be probed through
a number of observations, see Ref. [4110–4112] for a broad review. Studies of fundamental constants (FCs) and
their possible temporal and spatial variations are thus of utmost importance, and could provide a glimpse at physics
beyond the standard model, possibly shedding light on the presence of additional scalar fields and their couplings to
the standard sector, e.g., see Ref. [4113, 4114].
In the cosmological context, the fine-structure constant, αEM , and electron rest mass, me , are the most interesting,
although variations of Newton’s constant have also been considered [4115–4117] subject to some recent theoretical and
observational constraints [4118, 4119]. The former can be directly probed with measurements of the CMB temperature
and polarization anisotropies, e.g., see Ref. [4120–4127] through their effect on the cosmological recombination history
213

αEM me
TT + WP TT + WP
Planck 2013
+ lensing + lensing

Planck 2013
TT + WP TT + WP
+ lensing + + BAO DR7
BAO DR7

TT + WP
TTTEEE + lowTEB + Riess 2011
+ lensing
Planck 2015

TTTEEE + lowTEB
+ lensing
TTTEEE + lowTEB

Planck 2015
+ lensing
+ BAO DR12
TTTEEE + lowTEB
+ lensing
TTTEEE + lowTEB + BAO DR12
+ lensing
+ BAO DR12 TTTEEE + lowTEB
+ Riess 2019 + lensing
+ BAO DR12
+ Riess 2019
TTTEEE + lowl
+ lowE + lensing TTTEEE + lowl
Planck 2018

+ lowE + lensing

Planck 2018
Planck 2018

Planck 2018
Riess 2019

Riess 2019
TTTEEE + lowl
+ lowE + lensing TTTEEE + lowl
+ BAO DR12 + lowE + lensing
+ BAO DR12
TTTEEE + lowl
+ lowE + lensing TTTEEE + lowl + lowE
+ BAO DR12 + lensing + BAO DR12
+ Riess 2019 + Riess 2019

0.98 0.99 1.00 1.01 1.02 65.0 67.5 70.0 72.5 75.0 0.90 0.95 1.00 1.05 1.10 65.0 67.5 70.0 72.5 75.0
αEM /αEM,0 H0 me /me,0 H0

FIG. 80: Constraints on the fundamental constants using various combinations of Planck data together with their
H0 values and errors. Left: results from the fine structure constant αEM . Right: similar results but for the effec-
tive electron mass me . Here, we have redacted the constraint for H0 from CMB data only because the error bars
are so large. For the me MCMC analysis, we have widened the prior on the Hubble constant such that H0 >
20 km s−1 Mpc−1 . Figure is from Ref. [4130], which illustrated that varying me can alleviate the Hubble tension.

and photon scattering rate. A detailed description of individual physical effects on the CMB power spectra is given in
Ref. [4128], with calculations of the cosmological recombination history carried out using CosmoRec [4129]. In short,
increasing αEM and/or me leads to earlier recombination. This is primarily driven by the changes to the atomic
energy levels, which scale as E ∝ αEM2
me , thereby enforcing a higher temperature for recombination to occur. Beyond
this leading dependence, several subtle effects are encountered leading to differences in how αEM and me variations
affect the CMB signals. Crucially, the effect of αEM and me on the Thomson scattering rate, σT ∝ αEM 2
/m2e , has to
be carefully taken into account to yield consistent constraints [4128, 4130]. For additional discussion of the effects on
recombination see Ref. [4131].
Analysing Planck 2013 data, the values of αEM and me around recombination were proven to coincide with those
obtained in the lab to within ≃ 0.4% for αEM and ≃ 1% − 6% for me [4132]. These limits are ≃ 2 − 3 orders of
magnitude weaker than constraints obtained from other “local” measurements [4133–4136] and those from BBN [4137];
however, the CMB is sensitive to very different phases in the history of the Universe, centered around the time of last
scattering some 380, 000 years after the Big Bang, thereby complementing these measurements. In addition, CMB
measurements can be used to probe spatial variations of the FCs at cosmological distances [4138, 4139], opening yet
another avenue for exploration.
With the Planck 2013 results in mind, no significant surprises were expected from the analysis of improved CMB
data of the Planck 2015 and 2018 releases. It, however, turned out that when considering models with varying me ,
the geometric degeneracy becomes significant and can accommodate shifts in the value of the Hubble parameter when
multiple probes are combined [4130]. The same geometric freedom is not encountered when varying αEM due to
the modified dependence of the visibility function on this parameter. Indeed, when allowing me to depart from the
standard (local) value, a non-standard value of ∆me /me = 1.0191 ± 0.0059 (≃ 3.2σ significance) can be traded for
a reduction of the Hubble tension, suggesting that new physics may be at work (see Fig. 80 for illustration). The
addition of extra degrees of freedom that influence the post-recombination Universe such as Ωk or w0 /wa to variations
of the electron mass allows to further ease the Hubble tension with a ∼ 3.6σ preference, robust to SN data and BBN
constraints [709, 1523, 2444].
This finding has spurred an increased interest in studying VFCs in this context, with scenarios that allow for
varying me , e.g., see Ref. [4130, 4140–4143] for additional discussion, ranking high in model comparisons [709]. Since
VFCs can be caused by the presence of scalar fields, e.g., see Ref. [4113, 4144, 4145], a natural question is whether
the same scalar field could also be causing effects relating to EDE, possibly indicating a ‘two sides of the same coin’
interplay. Constraints using the synergy of cosmological and local data have been put on a variety of well-motivated
scalar fields models allowing for a physical changes in αEM [4146–4150]. However, local data, such as atomic clocks,
are putting very strong constraints on such models such that they are mostly unable to produce significant variations
214

of αEM during the recombination era [4151], while such constraints on varying me models are less restrictive [1523].
In addition, VFCs could play a role in solving the Hubble tension even in light of the recent DESI measurements
[1523, 2886, 4152–4154], although a general mechanism causing early recombination could simply be the main cause
of the tensions [4154].
Given the state of affairs, it will be extremely important to ask how different cosmological probes can be combined
to shed light on the physical origin of the Hubble tension. One important avenue forward is to directly constrain the
electron recombination history, given its crucial role in the formation of the CMB anisotropies [738, 4155–4160]. This
has recently been achieved using a non-perturbative and model-independent approach [4152], providing a clear target
for theoretical exploration. Indeed, the cosmological data prefers early recombination over the standard recombination
history obtained using CosmoRec, very much like what is caused by varying me or models with early structure formation
[4152]. This data-driven result may therefore indicate new physics in the redshift one thousand Universe, but at this
point cannot distinguish the physical cause of this finding.
One way of directly probing the recombination history is through measurements of the cosmological recombination
radiation (CRR) [2532, 4139, 4161–4163]. This tiny spectral distortion signal is created by photons emitted in the
hydrogen and helium recombination eras, and thus directly depends on the time and duration of the recombination
process. It was shown in Refs. [2532, 4139] that various extensions to ΛCDM can in principle be distinguished with
future CMB spectrometer measurements as envisioned for the ESA Voyage 2050 program [4164]. Should the Hubble
tension persist, then this will provide the ultimate test for various theoretical models. In addition, one can expect
the observational uncertainties in the cosmological recombination history to hamper our ability to answer questions
about extensions to ΛCDM. A measurement of the CRR is therefore highly motivated even beyond questions about
the Hubble tension, and provide a direct probe of one of the main pillars in our interpretation of CMB data.

4.9. Local New physics solutions to the Hubble and growth tensions
Coordinator: Leandros Perivolaropoulos
Contributors: Bhuvnesh Jain, Harry Desmond, Indranil Banik, Jeremy Sakstein, Nick Samaras, and Ruchika
Kaushik

A recent comprehensive analysis provides new insights into the Hubble tension, suggesting that the core of the
problem may lie in the distance ladder measurements rather than in conflicts between early and late Universe ob-
servations [3721]. This study compiled and analyzed two distinct groups of H0 measurements: those based on
the distance ladder approach, and those derived from one-step methods independent of both the distance ladder
and the sound horizon scale. The analysis revealed a significant discrepancy between these two groups. Distance
ladder-based measurements yielded a best-fit H0 = 72.8 ± 0.5 km s−1 Mpc−1 , while one-step measurements resulted
in H0 = 69.0 ± 0.48 km s−1 Mpc−1 . Notably, when two outlier measurements were removed from the one-step sam-
ple, the best-fit value reduced to H0 = 68.3 ± 0.5 km s−1 Mpc−1 , which is fully consistent with sound horizon-based
measurements like those from Planck CMB observations. A Kolmogorov-Smirnov test yielded a p-value of 0.0001,
indicating that the two samples are not drawn from the same underlying distribution.
Since the distance ladder is the only method for measuring H0 that is based on local physics (via calibrators of
SNIa), these findings lend support to the hypothesis of local physics solutions to the Hubble tension. They suggest
that the discrepancy may not be between early and late-time measurements, but rather between distance ladder
measurements and all other H0 determinations. This points to either a systematic effect influencing all distance
ladder measurements or new physics differentially impacting rungs of the ladder.
Such an intriguing class of potential solutions to the Hubble tension involves new physics acting in the local part of
the distance ladder (first and second rungs). This idea was first developed by Ref. [4165], who identified that screened
fifth forces (for reviews see Refs. [4166–4170]) may have a differential impact on different rungs of the ladder such
that their neglect could bias the inference of H0 . Ref. [4165] developed a range of phenomenological screening models,
in which the degree of screening (and hence effective value of Newton’s constant, Geff ) is set by various gravitational
properties of the galaxies used to calibrate SNIa and their environments. This included a novel screening mechanism
governed by the local DM density arising from the viable DE candidate of baryon–DM interactions [3568, 4171].
Ref. [4165] determined ranges for the screening proxies such that the anchor galaxies of the Cepheid PLR (the MW,
LMC, and N4258) are screened while some of the galaxies used to infer the SNIa absolute magnitude are not. Since
Cepheid pulsation periods are reduced and luminosities increased by an unscreened fifth force, the standard analysis
neglecting the fifth force would then underestimate the distance to the SNIa-calibrator galaxies, which at fixed redshift
would overestimate H0 .
A detailed modification to the SH0ES analysis pipeline allowing for this effect revealed that the Hubble tension could
be eliminated with a fifth-force strength (Geff − G)/G ≈ 0.1 in some of the screening models. However, an important
constraint derives from the consistency of Cepheid and TRGB distances to galaxies where both can be measured. The
fifth force affects TRGB distances oppositely to Cepheid distances [4165], so the fact that these distances agree under
215

GR implies that they will disagree under a strong fifth force. This requires (Geff − G)/G ≲ 0.05 in the most effective
models, preventing this theory from reducing the H0 tension below the 2σ level. Other constraints were studied, but
found not to yield such a strong bound. Nevertheless, in conjunction with some other effect (of which a myriad are
presented in this white paper), the tension could be resolved.
Ref. [4165] only addressed the Cepheid-calibrated distance ladder. The principal alternative is TRGB, which yields
a slightly smaller H0 value of 70−72 km s−1 Mpc−1 under the assumption of GR. The model was extended to the TRGB
calibration in Ref. [4172], where it was shown to be fully effective at solving the Hubble tension in that case. This
requires the LMC (the anchor galaxy of the TRGB absolute magnitude) to be less screened than the SNIa calibrators.
As the LMC is the least massive of the SH0ES anchors – and TRGB stars tend to be found in higher-mass galaxies
than Cepheids, which are young stars found in star-forming disks – significant regions of the MG parameter space can
solve both the TRGB tension and reduce the Cepheid tension to ∼ 2.5σ. Note that the SNIa themselves are assumed
screened in all these analyses, as would typically be expected in viable MG models due to their dense environments –
discussion of possible effects on SNIa may be found in Appendix B of Ref. [4165]. Variants of the model are studied
in Ref. [3235, 4173].
An alternative to screening is the more recent idea that new physical phenomena or transitions in the form of
physical laws occur locally, specifically at distances ≲ 40 Mpc or redshifts z ≲ 0.01, affecting the connection between
the distance ladder’s second rung (SNIa calibration by Cepheids or TRGB) and third rung (Hubble flow SNIa).
These Local Physics Transitions (LPTs) are assumed to involve abrupt changes of fundamental constants like the
gravitational constant G or the fine structure constant. They can affect the calibrators and SNIa used in the first and
second rungs of the distance ladder, leading to a different behavior compared to that in the Hubble flow (at larger
scales).
The simplest class of this paradigm involves an abrupt transition in the value of G while the other fundamental
constants remain fixed. Refs. [2668] and [2681] propose that a rapid transition in G at a redshift zt ≈ 0.01 could
resolve the Hubble tension. This G-step model (GSM) could imply that SNIa have lower luminosity L< at local scales
(second rung of distance ladder) than the luminosity L> of distant SNIa in the Hubble flow (third rung). The ratio
that would be required to solve the Hubble tension is L> /L< = 1.15, leading to a higher inferred value of H0 due to
the degeneracy between the SNIa L> and H0 in the Hubble flow.
The LPT paradigm was studied in Ref. [2681] through a reanalysis of the SH0ES data. By allowing for a transition
in the absolute magnitude of SNIa at a distance of about 50 Mpc, they find that the best-fit value of H0 drops
significantly from H0 = 73.04 ± 1.04 km s−1 Mpc−1 to H0 = 67.32 ± 4.64 km s−1 Mpc−1 , in full consistency with the
Planck value. This model also shows a substantial improvement in the fit to the data when an additional constraint
from the inverse distance ladder is included. Another study further tested and confirmed that the Cepheid and SNIa
datasets do not disfavour such a transition [4174].
Assuming L − A ∝ Gγ (where A, γ are constants and L is the SNIa absolute luminosity), an exponent γ < 0 (γ > 0)
would mean that a lower (higher) value of G is required at z ≳ 0.01 to solve the Hubble tension, leading also to a
potential resolution (exacerbation) of the growth rate tension [2668].
Recent constraints on G variation between the present time and recombination suggest that |∆G/G| < 0.05 at 2σ,
where ∆G ≡ G> − G< [4117]. These studies assume ΛCDM expansion and that all other fundamental constants
remain unchanged; relaxing these could weaken the constraints. Using the strong constraints of Ref. [4117], requiring
L> /L< = 1.15 with A = 0 requires γ ∈ / [−4.5, 2.8].
Early studies assumed that L ∝ MCh (where MCh is the Chandrasekhar mass), which implies γ = −3/2 [4175–4177].
However, this does not account for the standardization procedure required for turning SNIa into standard candles;
building a semi-analytic model to account for this, more recent studies use a simplified SNIa standardization procedure
(no full use of light curve stretch and no use of color) and approximate semi-analytical arguments (no hydrodynamical
simulation) to suggest that γ ≃ 1.46 with A = 0 [4165, 4178, 4179]. Both approaches make significant simplifications
to the underlying physics; a fully reliable analysis should use full 3D hydrodynamical simulations of SNIa with a
range of properties and obtain their light curves, properly treat MG effects on white dwarf structure, and adopt a
standardization protocol identical to that employed empirically, also producing uncertainties on γ and A.
It is thus possible that γ ∈/ [−4.5, 2.8] could be viable in the GSM. γ < 0 is however also strongly constrained: a
sharp rise in G up to 150 Myr ago (z ≲ 0.01) would have dramatic unobserved effects on neutron stars by causing them
to contract, thereby releasing vast amounts of energy [4180]. This is however also based on significant simplifying
assumptions that may not hold in practice. In particular, roughly constant neutron star and binary formation rate
over time is assumed, all galaxies are assumed similar to the MW, higher black hole formation at early-times due to
more massive stars and low metallicity is ignored, etc.
Further strong challenges to the GSM are presented in Ref. [4119]. Their main criticisms relate to the expansion of
the Earth’s orbit around the Sun coupled with lower G leading to a 10% increase in the number of days per year that
is not seen in the geochronometric and cyclostratigraphic records, a higher helioseismic age of the Sun due to higher
G over the vast majority of its history creating a mismatch with the ages of the oldest meteorite samples from the
216

early Solar System, and a sharp drop in Solar insolation on the Earth in the geologically recent past due to a drop in
G, leading to a large temperature drop.
Even if γ ∈/ [−4.5, 2.8] were to be conclusively ruled out or SNIa modeling found to require γ < 0 which is then
ruled out by neutron stars, or the tests in Ref. [4119] were verified, the possibility of a simultaneous transition of G
with other fundamental constants [4181–4183] like the fine structure constant would remain, keeping the general LPT
paradigm alive. Transitions of the type envisaged by the paradigm could be produced by theories such as Dilaton and
Kaluza-Klein theories [4144], TeVeS-like theories [4184], or varying α theories [4185]. Note that the above constraints
do not apply to the screening models discussed at the start of this section because screening naturally hides the fifth
force locally and there is no universal change to fundamental constants like G.
In conclusion, the proposed local physics modifications discussed here, including both the screened fifth force model
and transitions in the effective gravitational constant, appear to constitute a promising approach to the resolution
of the Hubble tension. These results underscore the importance of further investigation of new physics in the local
Universe that could affect the calibration of the cosmic distance ladder, and of constraints on such physics from other
sources.
217

5. Discussion and future opportunities


The problem of cosmic tensions poses several challenges and a number of opportunities in more deeply assessing pos-
sible systematics in observational surveys, developing new data analysis tool kits to refine our treatment of these data
products, and developing more robust physical models that can build on the successes of the concordance model and
pose new questions and give new solutions to the next decade of cosmology. To meet these challenges, a community-
wide effort will be needed involving advances in each separate field, but also strongly and more interconnected
relationships between these disciplines.
The coming decade will bring a number of pivotal surveys and observatories that aim to confront the central
questions posed by the last decades of concordance physics. This may bring additional unforeseen discoveries in the
cosmic history of the Universe, as well as new tools from the increasingly dominant spectrum of ML and statistical
physics toolkits, and possibly a new paradigm in our perspective of fundamental physics as an explanation of the
driver of cosmic evolution. In Sec. 5.1 the most significant observatories and survey prospects are discussed in the
context of the coming decade of observational cosmology. In Sec. 5.2 several key areas of development are identified
which will be crucial to meeting the growing observational, statistical, and fundamental physics challenges of the
upcoming decade of cosmology. Finally, we close in Sec. 5.3 with a summary of the central themes of this work and
an outlook perspective on the impact of cosmic tensions over the next few years.
5.1. Future survey prospects
Cosmology has undergone astounding advancements in the last few decades, due in part to the rapid progress of
robust surveys, including unprecedented instrument precision and impressive new approaches to statistical analyses.
These advancements have enabled a wider spectrum of tests of theoretical models, as well as unforeseen discoveries
such as the accelerating expansion of the Universe in the late 1990s and the unexpectedly high number of galaxies
imaged in the Hubble Deep Field. The next decade is set to extend this discovery potential with a swath of exciting
missions nearing completion.
The dynamics of future surveys have pivoted to extreme precision surveys with vastly expanded mission objectives.
These include measurements reaching much closer to the early Universe, such as the potential detection of the first
luminous objects in the Universe through the 21-cm hydrogen line, the possible detection of signatures of inflation
in the CMB, and significant reductions in the uncertainty of current CMB measurements. There are also a host of
surveys that will track the spatial distribution of galaxies in the Universe to unprecedented levels in terms of the
number of galaxies and the precision of these measurements. These surveys will facilitate even deeper investigations
into the large-scale structure of the cosmos, providing more information on the nature and evolution of DE and DM.
There is a vast array of planned observational surveys and measurements for the coming decade, which will add to
the already considerable number of active survey collaborations. Below, we describe some of the major planned and
ongoing missions and survey analyses.
5.1.1. Cosmic microwave background
A key focus of the next generation of cosmological surveys is the study of the CMB. Building on decades of ground-
breaking discoveries, future CMB experiments are poised to address some of the most pressing challenges in modern
cosmology, including the persistent tensions in the Hubble constant and the amplitude of matter fluctuations (S8 ).
Upcoming CMB missions aim to achieve unprecedented precision in measurements of the CMB’s temperature and
polarization anisotropies, offering unique insights into the physics of the early Universe and its evolution.
The H0 tension stems from a significant discrepancy between the value of the Hubble constant inferred from early
Universe observations, such as the CMB, and local measurements using the cosmic distance ladder. Future CMB
experiments will refine measurements of the sound horizon at the baryon drag epoch—a critical calibration scale for
early-time H0 estimates. These high-precision data will test proposed solutions to the H0 tension, including EDE
models, extra relativistic species, or modifications to pre-recombination physics, by either validating or ruling out
scenarios that alter the expansion history before recombination.
The S8 tension, which reflects a discrepancy in the amplitude of matter clustering, is another key focus of upcoming
CMB experiments. By producing high-resolution maps of the CMB lensing power spectrum, they will directly probe
the distribution of DM and the growth of large-scale structures. These measurements will be critical for cross-checking
WL and galaxy clustering observations, helping to clarify whether the S8 tension arises from unknown systematics in
late-time surveys or signals a breakdown in the standard cosmological model.
Additionally, CMB experiments will probe primordial B-mode polarization, providing a unique window into the
inflationary era and the energy scale of the early Universe. These observations could also inform models that connect
the physics of inflation to current cosmological tensions, such as scenarios involving new scalar fields or MG.
Future CMB missions will also contribute to our understanding of DE and neutrino physics. Improved measurements
of secondary anisotropies, such as the Sunyaev–Zel’dovich effect, will enable detailed studies of galaxy clusters and
baryonic feedback processes, while precise constraints on the sum of neutrino masses and extra relativistic species will
218

help refine models of the early Universe. By combining these capabilities with cross-correlations to other cosmological
probes, such as BAO and WL surveys, CMB experiments will play a pivotal role in addressing cosmological tensions
and advancing our understanding of the fundamental properties of the Universe.
Below, we highlight some of the major planned and ongoing CMB-focused missions and their scientific goals:

• South Pole Telescope – 3rd Generation (SPT-3G): The SPT-3G [4186] is the latest upgrade to the SPT,
focusing on high-resolution studies of the CMB from the exceptional observational site at the South Pole. This
third-generation camera, featuring over 16,000 detectors, has significantly improved sensitivity, enabling deeper
and more precise measurements of CMB temperature and polarization anisotropies. What sets SPT-3G apart is
its ability to perform high-resolution surveys over small sky patches, optimized for studies of galaxy clusters via
the Sunyaev–Zel’dovich effect, as well as gravitational lensing of the CMB. These measurements are critical for
understanding the distribution of DM and large-scale structure formation. Additionally, SPT-3G contributes to
constraints on the sum of neutrino masses and potential new physics beyond the standard cosmological model.
The compact field-of-view and deep integration capabilities make SPT-3G uniquely suited for detecting subtle
signals, such as CMB lensing, and studying small-scale anisotropies.

• Atacama Cosmology Telescope (ACT): Located in the Atacama Desert of Chile, ACT [4187] is a ground-
based observatory dedicated to high-resolution studies of the CMB. The telescope has been instrumental in
mapping CMB temperature and polarization anisotropies and probing large-scale structure. The ACT Data
Release 4 (DR4) provided high-sensitivity maps of the CMB, including measurements of temperature, E-mode
polarization, and cross-correlations, covering over 17,000 square degrees of the sky. DR4 data significantly
refined estimates of cosmological parameters, including constraints on H0 , σ8 , and the sum of neutrino masses.
It also included improved measurements of the lensing power spectrum, advancing our understanding of the
distribution of matter in the Universe. The forthcoming ACT Data Release 6 (DR6) is expected to feature
even more precise maps with reduced noise levels and expanded sky coverage. DR6 aims to provide improved
constraints on primordial B-mode polarization, enhancing our ability to test inflationary physics. Additionally,
it will enable more detailed studies of secondary anisotropies, such as the Sunyaev–Zel’dovich effect, and cross-
correlations with galaxy surveys.

• PolarBear and Simons Array: PolarBear, along with its successor, the Simons Array [4188], is focused on
high-resolution measurements of the CMB’s polarization anisotropies. Located in the Atacama Desert, Chile,
these experiments aim to detect B-mode polarization caused by gravitational lensing and primordial GWs. The
Simons Array, an upgraded version of PolarBear, consists of multiple telescopes with advanced detectors for
improved sensitivity. These measurements will refine constraints on the tensor-to-scalar ratio r, test inflationary
models, and map the lensing power spectrum, shedding light on the distribution of DM and the evolution of
large-scale structure.

• SPIDER: SPIDER [2382] is a balloon-borne experiment designed to detect the large-scale polarization of the
CMB, focusing on B-modes linked to primordial GWs. By operating above the atmosphere, SPIDER achieves
reduced contamination from ground-based noise and atmospheric effects. Its payload consists of multiple tele-
scopes equipped with cryogenic polarimeters, optimized for observing large angular scales. SPIDER has already
completed successful flights, with future missions planned to improve sensitivity and expand sky coverage.

• Ali Cosmic Polarization Telescope (AliCPT): AliCPT [4189] is a ground-based CMB experiment located
at the high-altitude Ali Observatory in Tibet. AliCPT focuses on measuring the polarization of the CMB
at large angular scales, particularly the B-modes associated with primordial GWs. Its high-altitude location
minimizes atmospheric contamination, allowing for precise measurements.

• Q&U Bolometric Interferometer for Cosmology (QUBIC): QUBIC [4190] is an international collabo-
ration designed to measure the B-mode polarization of the CMB using bolometric interferometry. It uniquely
combines the sensitivity of bolometric detectors with the spatial filtering capabilities of interferometry, allowing
for precise measurements of polarization anisotropies. Located at high altitude in Argentina, QUBIC targets
primordial B-modes associated with inflation, as well as lensing-induced B-modes, providing critical constraints
on the tensor-to-scalar ratio r. The first module began operations in 2022.

• The Simons Observatory (SO): The Simons Observatory [680] is a next-generation ground-based experiment
located in the Atacama Desert, Chile. It will consist of three small-aperture telescopes (SATs) and one large-
aperture telescope (LAT), all equipped with advanced cryogenic detector arrays. The primary goal of the
observatory is to measure the CMB temperature and polarization anisotropies with high precision across a wide
range of angular scales. SO will target the primordial B-mode polarization of the CMB, providing critical tests of
219

inflationary physics and insights into the early Universe’s energy scale. It will also probe secondary anisotropies,
such as the Sunyaev–Zel’dovich effect, to map the growth of large-scale structure. The observatory will place
constraints on the sum of neutrino masses and search for new physics beyond the standard model. Additionally,
SO will refine measurements of cosmological parameters like H0 and S8 , addressing key tensions in the ΛCDM
model. The Simons Observatory is expected to begin operations in the mid-2020s.
• Cosmology Large Angular Scale Surveyor (CLASS): The CLASS experiment [4191] is designed to study
the polarization of the CMB on large angular scales. Located in the Atacama Desert, Chile, CLASS employs a
series of telescopes equipped with cryogenic detectors to measure the faint polarization signals from the early
Universe. The experiment focuses on detecting the reionization and recombination bumps in the CMB polar-
ization power spectrum, aiming to constrain primordial GWs and the optical depth of reionization. CLASS’s
innovative strategy includes observing large sky patches with rapid rotation to minimize systematic errors caused
by atmospheric and instrumental noise.
• GroundBIRD: GroundBIRD [4192] is a ground-based experiment designed to measure the polarization of the
CMB on large angular scales. Located at the Teide Observatory in Spain, it uses fast rotation and superconduct-
ing detectors to reduce atmospheric noise and systematic errors. The experiment focuses on detecting B-mode
polarization signals, particularly those associated with primordial GWs. GroundBIRD’s innovative observing
strategy, which involves continuous scanning with a rotating cryostat, enhances its sensitivity to large-scale
polarization while minimizing contamination.
• BICEP/Keck Array and BICEP Array: The BICEP/Keck Array and its successor, the BICEP Array
[4193], are focused on detecting B-mode polarization of the CMB with high sensitivity. Operating from the
South Pole, these experiments target large-scale polarization signals to constrain the tensor-to-scalar ratio r
and probe the inflationary epoch. The BICEP Array incorporates advanced detector technology and expanded
frequency coverage to improve sensitivity and reduce systematic uncertainties.
• The Lite satellite for the study of B-mode polarization and Inflation from cosmic background
Radiation Detection (LiteBIRD): The LiteBIRD probe [2381, 4194–4197] aims to measure the unique
imprints of B-mode polarization in CMB photons, which are related to primordial GWs and inflation. It will
succeed the Planck mission in conducting full-sky surveys. Expected to launch in late 2029, LiteBIRD is a
Japanese initiative by JAXA involving collaborations with agencies in North America and Europe. Positioned
at the Lagrange point L2 in the Sun-Earth system, it will conduct a three-year survey. The relic CMB radiation
features E-mode polarization, linked to scalar perturbations, and B-mode polarization, associated with tensor
perturbations. While the Cosmic Background Imager provided the first detailed E-mode polarization map,
the B-mode signal remains undetected. LiteBIRD will take cosmic-variance-limited measurements of E-mode
polarization to study large-scale correlations, shedding light on initial conditions for cosmological perturbations
and providing insights into physics beyond the standard model. Regarding B-mode polarization, LiteBIRD
will aim for a tensor-to-scalar ratio limit of r < 0.001. A positive detection would have a profound impact on
fundamental physics, offering insights into inflationary physics, parity violation, and primordial cosmological
magnetism.
• CMB—Stage 4 (CMB-S4): The CMB-S4 observatory [681] will operate telescopes at both the South Pole
and the Atacama Desert in Chile, enabling deep microwave observations over small and large sky fields. It
will employ advanced superconducting detector array technologies to reduce galactic foreground contamination.
Expected to start operations in the late 2020s, though currently on hold, CMB-S4 aims to measure primordial
GWs associated with early rapid expansion of density fluctuations, characterized by the B-mode polarization of
the CMB. This would provide critical information about cosmic inflation and fundamental physics, targeting
a tensor-to-scalar ratio of r < 0.002, an order of magnitude improvement over current limits. CMB-S4 will
constrain the sum of neutrino masses, critical for understanding the sterile neutrino theory and the inverted
neutrino mass hierarchy. Beyond neutrino physics, it will impose stringent limits on possible light particles
beyond the standard model, addressing extra effective degrees of freedom. On DM, CMB-S4 will probe CMB
anisotropies for potential signals of WIMP annihilation, placing constraints on their masses, while exploring non-
thermal DM effects on lensing power spectra. Another key goal is probing DE, with precision measurements
of ΛCDM cosmological parameters such as H0 and S8 , enabling comparisons with other early-time probes.
Additionally, CMB-S4 will generate higher-resolution maps of matter distribution in the Universe by measuring
distortions in CMB photons caused by gravitational lensing from the surface of last scattering. This will
significantly enhance future galaxy surveys.
• Cosmic Microwave Background - High Definition (CMB-HD): CMB-HD [4198, 4199] is a proposed
next-generation ground-based observatory that aims to revolutionize cosmology with its unparalleled resolution
220

and sensitivity. Unlike other CMB experiments, CMB-HD focuses on small-scale anisotropies, achieving an
angular resolution of 0.5 arcminutes and surveying over 50% of the sky. This capability allows it to probe
previously inaccessible signals and extend our understanding of the Universe to finer detail. CMB-HD’s primary
science goals include precise measurements of the Sunyaev–Zel’dovich effects to map galaxy clusters, their gas
content, and baryonic physics. It will also produce detailed maps of the small-scale CMB lensing power spectrum,
offering new insights into the distribution of DM and the growth of large-scale structure. By detecting primordial
B-mode polarization, CMB-HD will constrain the tensor-to-scalar ratio r to test inflationary physics at energy
scales far beyond current limits. Additionally, CMB-HD will place stringent constraints on the sum of neutrino
masses, light relics, and possible deviations from the ΛCDM model.
• Probe of Inflation and Cosmic Origins (PICO): PICO [4200] is a proposed next-generation satellite mission
designed to provide a comprehensive, all-sky measurement of the CMB. Operating from space, PICO will avoid
the challenges of atmospheric contamination and cross-calibration issues that affect ground-based experiments.
Its observations will span a broad frequency range from 20 GHz to 800 GHz, enabling highly accurate foreground
removal and calibration consistency across the entire sky. A key advantage of PICO’s space-based platform is its
ability to precisely measure the optical depth to reionization, τ . This measurement, critical for understanding the
early history of star formation and the growth of cosmic structures, is less accessible to ground-based telescopes
due to the need for absolute calibration. By combining this capability with its unparalleled sensitivity to B-
mode polarization, PICO will constrain the tensor-to-scalar ratio r to levels below 0.001, providing definitive
tests of inflationary models. In addition to probing inflation, PICO will deliver high-resolution maps of the
E-mode polarization and CMB lensing power spectrum, offering insights into the distribution of DM and large-
scale structure. It will also place stringent constraints on the sum of neutrino masses, detect potential signals
from light relics, and test extensions to the ΛCDM model. If approved, PICO will complement ground-based
experiments by addressing systematic challenges and providing the calibration accuracy and sensitivity only
possible from space.
• Polarized Radiation Imaging and Spectroscopy Mission (PRISM): PRISM [4201] is a proposed satellite
mission designed to provide extremely sensitive, all-sky measurements of the CMB’s temperature and polariza-
tion anisotropies. With its advanced instrumentation and broad frequency coverage, PRISM aims to enhance
our understanding of both early and late-time cosmology. A key goal of the mission is to detect primordial B-
mode polarization, which would provide direct evidence of inflation and constrain the energy scale of the early
Universe. PRISM’s ability to refine measurements of the CMB power spectrum and polarization will enable
tighter constraints on the sound horizon at the baryon drag epoch, directly addressing the H0 tension by testing
EDE models and other modifications to pre-recombination physics. Additionally, PRISM’s high sensitivity to
secondary anisotropies, such as CMB lensing, will improve measurements of the lensing power spectrum, aiding
investigations into the S8 tension and the growth of large-scale structure. If approved, PRISM would comple-
ment ground-based and balloon-borne experiments by avoiding atmospheric contamination, offering a clean and
highly detailed dataset for cosmological studies.

5.1.2. Baryon acoustic oscillations


The study of BAO has become a cornerstone of modern cosmology, providing a powerful method for measuring
the expansion history of the Universe. BAO are the relic imprints of sound waves that propagated in the early
Universe, leaving a characteristic scale in the large-scale distribution of galaxies and matter. This standard ruler
has proven invaluable for calibrating cosmic distances and constraining key cosmological parameters, including the
Hubble constant and the DE equation of state.
As the precision of cosmological measurements continues to improve, future BAO experiments are poised to deliver
new insights into the late-time Universe. By mapping the three-dimensional distribution of galaxies, RGBs, and other
tracers across an extensive redshift range, these experiments will probe the dynamics of cosmic expansion and the
growth of large-scale structure. This enhanced precision will enable a deeper exploration of the H0 and S8 tensions,
testing whether they arise from unaccounted systematics or new physics beyond the ΛCDM model.
In addition to refining our understanding of DE and its influence on cosmic acceleration, upcoming BAO surveys
will provide complementary constraints on modifications to gravity and potential extensions to standard cosmology.
Synergies with other probes, such as WL and CMB lensing, will further bolster the ability to test fundamental physics
and the consistency of the cosmological model.
Below, we describe some of the major planned and ongoing BAO-focused missions and their scientific objectives:

• Dark Energy Spectroscopic Instrument (DESI): DESI [813] is a state-of-the-art spectroscopic survey
currently operating on the Mayall 4-meter telescope at Kitt Peak National Observatory. Its primary goal is to
create the most detailed three-dimensional map of the Universe by observing tens of millions of galaxies and
221

RGBs across a wide redshift range. DESI’s high-precision BAO measurements will provide critical constraints
on the cosmic distance scale and the expansion history of the Universe, directly addressing the H0 tension.
Additionally, RSD analyses from DESI will help probe the growth rate of large-scale structures, contributing to
our understanding of the S8 tension and testing potential modifications to gravity.

• Canadian Hydrogen Intensity Mapping Experiment (CHIME): CHIME [4202] is a revolutionary radio
interferometer located at the Dominion Radio Astrophysical Observatory in Canada. Designed to operate in the
400–800 MHz band, it maps the large-scale structure of the Universe using the 21 cm hydrogen line, covering a
redshift range of 0.8 < z < 2.5. By using intensity mapping techniques, CHIME provides precise measurements
of the BAO scale, complementing optical surveys like DESI and HIRAX. The experiment’s wide field-of-view
and innovative cylindrical reflector design enable it to conduct highly efficient surveys, making CHIME a key
player in refining the cosmic distance scale. .

• BAO from Integrated Neutral Gas Observations (BINGO): BINGO [4203] is a 21 cm intensity mapping
experiment specifically designed to measure BAO in the redshift range 0.13 < z < 0.48. Located in Brazil, it
utilizes a compact array of radio antennas to map the large-scale distribution of neutral hydrogen. BINGO
is optimized to minimize instrumental noise and systematics, providing precise constraints on the expansion
history of the Universe and DE. The project is currently under construction and expected to begin operations
in the mid-2020s.

• Euclid: The Euclid mission [4204], led by ESA, is designed to create a detailed three-dimensional map of the
Universe, enabling precise measurements of the BAO scale across a wide redshift range (0.7 < z < 2.0). Its
near-infrared spectroscopic survey will provide accurate redshifts for tens of millions of galaxies, establishing
a robust standard ruler for cosmological distances. By focusing on high-redshift galaxy clustering, Euclid will
refine the expansion history and DE equation of state. Its spectroscopic data will also complement optical BAO
studies, bridging gaps in redshift coverage and enhancing multi-probe analyses with other cosmological surveys.

• Prime Focus Spectrograph (PFS): The PFS survey [4205, 4206] is a spectroscopic survey operating on the
8-meter Subaru Telescope. It is designed to observe 2,400 objects simultaneously within a 1.2 deg2 field of view,
covering wavelengths from the near-ultraviolet to the near-infrared. With its wide field and spectral coverage,
PFS focuses on three primary science programs: cosmology, galaxy evolution, and galactic archaeology. For
cosmology, PFS will observe approximately 4 million emission-line galaxies (ELGs) over 1,200 deg2 , covering
redshifts from 0.8 to 2.4. Unlike other surveys, it will uniquely map ELGs at 2.0 < z < 2.4, complementing
DESI. PFS will provide high-precision measurements of BAO and the Alcock–Paczyński (AP) effect, enabling
constraints on the Hubble expansion history and testing potential evolution in dark energy out to z = 2.4. By
combining these results with lower-redshift BAO constraints, PFS will determine the dark energy density to
approximately 7% accuracy per redshift bin. Additionally, RSD measurements will reconstruct the growth rate
of cosmic structure, f σ8 (z), with 6% accuracy
P up to z = 2.4, allowing for a precise determination of the sum of
neutrino masses, with an uncertainty of σ( mν ) = 0.02.

• Roman Space Telescope (formerly WFIRST): The Roman Space Telescope [892], led by NASA, is a highly
ambitious mission designed to address key questions about DE, exoplanets, and the structure of the Universe.
Through its High Latitude Survey, Roman will map the large-scale distribution of galaxies and perform WL
and BAO analyses with unprecedented precision. This mission will provide robust constraints on the expansion
history, DE dynamics, and the growth of cosmic structures.

• Rubin Observatory’s Legacy Survey of Space and Time (LSST): The Rubin Observatory [4207], in its
final construction phase in northern Chile, will conduct the LSST, a 10-year survey of 18, 000 deg2 of the sky
across six wavelength bands. Public data release is expected approximately two years after first light in 2025.
LSST will address a broad range of fundamental questions, including: (1) Is DE dynamical, as characterized by
the w0 -wa parametrization? (2) Can the expansion history and large-scale structure help distinguish between
exotic energy densities and MG? (3) What are the properties of DM, as revealed by microlensing searches?
(4) Do DM halos exist without hosting galaxies? (5) Are matter fluctuations in the late Universe consistent
with CMB-derived constraints? LSST is expected to profoundly impact our understanding of DE, DM, and
fundamental physics.

• SKAO (Square Kilometer Array Observatory): SKAO [4208] is a next-generation radio telescope array
under development, with sites in South Africa and Australia. Its unprecedented sensitivity and angular resolution
will enable precise 21-cm intensity mapping, allowing for detailed BAO measurements across a wide redshift
range. These data will be instrumental in tracing the evolution of cosmic expansion and structure growth,
222

providing crucial information for resolving the H0 and S8 tensions and testing the consistency of the ΛCDM
model. (see Sec. 5.1.5)
• The Hydrogen Intensity and Real-time Analysis eXperiment (HIRAX): HIRAX [1839] is a radio
interferometer array under construction in South Africa, optimized for intensity mapping of the 21-cm hydrogen
line. HIRAX will measure BAO across a redshift range of 0.8 < z < 2.5, directly probing the expansion history
during the period of cosmic acceleration driven by DE. These measurements will complement optical BAO
surveys, refining constraints on DE models and addressing potential new physics.
• Dark Energy Spectroscopic Instrument - Phase II (DESI-II): DESI-II [4209] is a proposed extension of
the highly successful DESI survey, aiming to build on its existing infrastructure to further refine cosmological
measurements. By expanding its redshift coverage and increasing the volume of observed galaxies and RGBs,
DESI-II will provide even more precise measurements of the BAO and RSD. The extended survey will tar-
get fainter galaxies and higher redshift objects, allowing for a deeper exploration of the late-time Universe’s
expansion history and the growth of cosmic structures.
• The 4-metre Multi-Object Spectroscopic Telescope (4MOST): 4MOST [4210], based at the ESO’s
VISTA telescope in Chile, is designed to conduct massive spectroscopic surveys of galaxies and RGBs over large
sky areas. Its multi-object spectrograph will enable precise BAO measurements and RSD analyses, improving
constraints on the cosmic expansion rate and structure growth. 4MOST will work synergistically with imaging
surveys like LSST to provide redshift information critical for cosmological studies, including cross-correlation
analyses to probe the S8 tension.
• Spectro-Photometer for the History of the Universe, Epoch of Reionization, and Ices Explorer
(SPHEREx): SPHEREx [4211] is a NASA mission designed to perform an all-sky spectral survey. By mapping
galaxies across a wide range of redshifts, SPHEREx will measure BAO and RSD signals, enabling precise
constraints on the expansion history and the growth of cosmic structure. Its unique spectral coverage will
complement optical surveys and provide new insights into the physics.
• CO Mapping Array Project (COMAP): COMAP [4212] is a pioneering experiment focused on intensity
mapping of carbon monoxide (CO) lines at high redshift. COMAP will trace the large-scale structure of the
Universe during the epoch of galaxy formation, providing a complementary approach to BAO measurements.
These data will help refine constraints on the expansion history and test models of DE and MG.
• Packed Ultra-Wideband Mapping Array (PUMA): PUMA [4213] is a proposed next-generation radio
interferometer optimized for detecting BAO and measuring RSD over a wide redshift range of 2 < z < 6.
Using 21 cm intensity mapping, PUMA will map the large-scale distribution of neutral hydrogen, providing
unprecedented insights into the Universe’s expansion history and the growth of cosmic structure. Its innovative
design, featuring a densely packed array of antennas, will enable high sensitivity and wide bandwidth, making
it a critical experiment for studying DE, testing modifications to gravity, and resolving tensions in cosmological
parameters. Expected to begin operations in the 2030s, PUMA will complement optical BAO surveys and push
the boundaries of cosmological research.

5.1.3. Weak lensing experiments


WL is a cornerstone tool for investigating cosmological tensions, particularly the S8 discrepancy, which reflects a
persistent difference between the amplitude of matter fluctuations inferred from early- and late-Universe observations.
WL measures the subtle distortions in galaxy shapes caused by the gravitational lensing effect of intervening mass
distributions, providing a direct probe of the growth of cosmic structures. pBy combining the clustering amplitude
σ8 with the matter density parameter Ωm,0 , the derived parameter S8 = σ8 Ωm,0 /0.3 becomes a critical test of the
ΛCDM model.
In addition to its crucial role in addressing the S8 tension, WL data also contribute to resolving the H0 tension.
Synergistic analyses that combine WL with other cosmological probes, such as CMB lensing and BAO measurements,
enable a multi-probe approach to jointly constrain both early- and late-Universe parameters. This integrated strategy
improves the robustness of results and provides deeper insights into the underlying physics driving these tensions.
Below, we highlight planned and ongoing WL experiments and their contributions to addressing the S8 tension:

• Kilo-Degree Survey (KiDS): KiDS [859] provided high-precision WL and photometric redshift measure-
ments across 1, 350 deg2 , with a strong focus on controlling systematics such as shear calibration, photometric
redshifts, and intrinsic galaxy alignments. Using high-quality imaging from the Very Large Telescope (VLT) and
tomographic redshift binning, KiDS delivered some of the most precise S8 constraints, highlighting a persistent
223

tension with ΛCDM predictions. Its integration with external datasets, such as galaxy clustering and CMB
lensing maps, enabled multi-probe analyses and established a strong foundation for addressing the S8 tension
in future surveys.
• Dark Energy Survey (DES): DES [4214] observed 5, 000 deg2 of the southern sky, combining WL and galaxy
clustering to provide robust S8 constraints. With precise photometric redshift calibration and careful control
of systematics, DES advanced studies of large-scale structure growth. Its cross-correlations with CMB lensing
maps and tomographic analyses tested ΛCDM predictions and explored extensions like evolving DE. DES’s
extensive data and methodological innovations set a high standard for future WL surveys.
• Hyper Suprime-Cam (HSC): HSC [853], conducted on the Subaru Telescope, delivered WL data over
1, 400 deg2 with exceptional resolution and depth, enabling precise studies of cosmic shear. Its high-resolution
imaging allowed detailed analyses of smaller-scale structures and provided key insights into S8 , revealing per-
sistent tensions with ΛCDM predictions. The survey’s innovative techniques for photometric redshift estima-
tion and systematic error control ensured high accuracy in its results. HSC also contributed significantly to
cross-correlation studies with galaxy clustering and CMB lensing, further refining constraints on the growth of
structure and testing extensions to the standard cosmological model.
• Super-pressure Balloon-borne Imaging Telescope (SuperBIT): SuperBIT [4215] is a stratospheric,
balloon-borne telescope designed for high-resolution, wide-field imaging, enabling precise WL measurements
of galaxy clusters. By operating above most of Earth’s atmosphere, it minimizes atmospheric distortions, de-
livering exceptional data quality. SuperBIT’s observations are critical for mapping the distribution of DM and
studying large-scale structure formation, providing complementary insights to ground- and space-based surveys.
Its unique capabilities contribute to addressing the S8 tension by offering an independent probe of structure
growth.
• Euclid: Euclid’s high-resolution optical and near-infrared imaging capabilities [4204] are designed to map the
weak gravitational lensing of billions of galaxies over 15, 000 deg2 of the sky. Its ability to probe redshifts up
to z ∼ 2.5 makes it uniquely suited for tomographic studies of structure growth, providing tight constraints
on the S8 parameter. By achieving unparalleled depth and resolution in WL, Euclid will play a crucial role
in testing modifications to ΛCDM, such as evolving DE models and MG theories. Its combination of imaging
and spectroscopic data will enable synergy with Rubin LSST and CMB lensing maps, further enhancing our
understanding of cosmic structure formation.
• Rubin Observatory’s Legacy Survey of Space and Time (LSST): The Rubin Observatory [4207] will
survey 18, 000 deg2 of the sky over a 10-year period, providing deep, multi-band imaging across six optical filters.
Rubin’s LSST high precision in WL and galaxy clustering measurements will allow detailed tomographic studies
of structure growth and yield tighter constraints on S8 . Its ability to detect millions of faint galaxies at high
redshifts will improve our understanding of the evolution of cosmic structures and test potential extensions
to ΛCDM. Rubin’s LSST rich dataset will also facilitate cross-correlations with CMB and other WL surveys,
strengthening the multi-probe approach to resolving cosmological tensions.
• Roman Space Telescope (formerly WFIRST): The Roman Space Telescope’s High Latitude Survey [892]
will deliver high-resolution WL data over 2, 000 deg2 , focusing on the distribution of DM and the growth of cosmic
structures. Its combination of near-infrared imaging and spectroscopy enables precise photometric redshift
estimation, essential for tomographic studies of WL. Roman’s unparalleled sensitivity at high redshifts (z ∼ 2)
will refine measurements of the S8 parameter, providing stringent tests of ΛCDM and its alternatives, such as
evolving DE and MG models. Its synergy with Euclid and Rubin’s LSST will enhance cross-calibration efforts
and improve constraints on cosmic structure formation across a wide range of scales and epochs.
• Square Kilometer Array Observatory (SKAO): SKAO [4208], utilizing 21-cm intensity mapping and
WL, will provide unique and independent constraints on S8 by probing the distribution of DM and the growth
of structures. Its radio-based approach will extend WL studies to higher redshifts and larger scales, offering
a complementary perspective to optical surveys. By addressing potential systematics in traditional probes,
SKAO will play a crucial role in testing deviations from ΛCDM and enhancing multi-probe strategies to resolve
cosmological tensions.
• Einstein Telescope (ET) and Cosmic Explorer (CE): These proposed third-generation GW observato-
ries [593, 4216–4218], expected to begin operations in the mid to late 2030s, will achieve unprecedented sensitivity
in detecting GWs. In addition to their primary focus on GW astrophysics, they will enable measurements of
WL effects on GW signals. This innovative approach provides a novel and independent method to probe the S8
224

parameter, offering insights into the growth of cosmic structures and testing extensions to the standard cosmo-
logical model. Their unique capabilities will complement traditional WL surveys, further enriching multi-probe
cosmological studies. (see Sec. 5.1.4)

5.1.4. Gravitational waves as probes of cosmological tensions


GWs offer a revolutionary perspective in addressing cosmological tensions, functioning as “standard sirens” that
enable the measurement of cosmic distances independently of electromagnetic calibrations. By directly determining
the luminosity distance to GW events, particularly those accompanied by electromagnetic counterparts (e.g., binary
neutron star mergers), this method bypasses systematics associated with traditional distance ladder techniques. For
a recent review on the capabilities of GW observatories see [653].
In addition to these “bright sirens,” GWs from events lacking identifiable electromagnetic counterparts, termed
“dark sirens,” can also contribute to cosmology. By correlating the GW signal with galaxy catalogs to infer host
redshifts, dark sirens expand the scope of GW cosmology, providing complementary constraints on the Hubble constant
and other cosmological parameters. Furthermore, advancements in spectroscopic observations of host galaxies have
introduced “spectroscopic sirens,” which refine redshift measurements associated with GW events. These high-precision
techniques enhance the reliability of constraints on the expansion history and reduce uncertainties linked to host galaxy
identification.
Beyond H0 , GWs contribute to understanding large-scale structure formation through WL of GWs by intervening
matter. These lensing effects provide unique insights into the amplitude of matter fluctuations, addressing the
S8 tension. Additionally, third-generation observatories such as the ET and the Cosmic Explorer will extend GW
observations to higher redshifts (z ∼ 10), probing the early Universe’s expansion history and potential deviations from
the ΛCDM model. The integration of GW observations with traditional cosmological probes, such as BAO, CMB,
and WL, opens new pathways for multi-messenger cosmology. Together, these complementary approaches promise to
resolve key tensions and deepen our understanding of the fundamental physics of the Universe.
Below, we highlight ongoing and planned GW experiments and their contributions to addressing cosmological
tensions.

• LIGO-Virgo-KAGRA Network: The collaboration between the LIGO, Virgo, and KAGRA interferome-
ters [589] has already achieved a remarkable number of significant detections, including the merger of a neutron
star with an unknown compact object. This success is attributed to substantial sensitivity improvements, stem-
ming from upgrades to individual detectors and the synergistic effect of combined observations. The fourth
observation run (O4) is scheduled to conclude in the summer of 2025, after which further upgrades will be im-
plemented. These upgrades will include advancements such as reduced thermal noise, new test mass mirrors, and
the installation of a larger beamsplitter in the LIGO detectors. These enhancements will significantly increase
the network’s overall sensitivity, enabling a larger detection volume and improved precision in measurement. As
the number of detections grows, the uncertainty in key cosmological parameters like H0 will decrease. Specif-
ically, for neutron star-neutron
√ star (NS-NS) mergers, the uncertainty in H0 estimates is expected to improve
by approximately 15%/ N , where N is the number of detections. This progress underscores the pivotal role of
the LIGO-Virgo-KAGRA network in refining our understanding of the Universe.
• Laser Interferometer Space Antenna (LISA): LISA [596] will be the first space-based GW detector,
consisting of three spacecraft in a triangular configuration, separated by millions of kilometers, and following a
heliocentric orbit. Scheduled for launch in the mid-2030s, LISA will primarily focus on astrophysical phenomena,
including high-redshift mergers, extreme mass ratio inspirals, galactic binaries, and planetary objects. However,
its contributions to cosmology will be equally groundbreaking. One of LISA’s key cosmological objectives is
to dramatically expand the catalog of dark sirens. These GW events, devoid of electromagnetic counterparts,
serve as alternative standard candles, independent of the traditional cosmological distance ladder. By leveraging
statistical methods with galaxy catalogs or identifying electromagnetic counterparts where possible, LISA will
test the distance-redshift relation on cosmological scales, providing new constraints on DE models. With sensi-
tivity extending to redshifts as high as z ∼ 10, LISA will open a window to the Universe’s distant past, offering
unprecedented opportunities to explore the nature of DE at scales inaccessible to current methods. Further-
more, LISA’s observations will enable precise and independent measurements of H0 , contributing to resolving
the persistent cosmological tensions.
• Einstein Telescope (ET): ET [4216–4218] is a proposed next-generation GW observatory designed to achieve
unprecedented sensitivity through advancements in quantum noise suppression, cryogenics, and interferometry.
With a triangular configuration and a planned underground location to minimize seismic noise, the ET represents
a significant leap forward from current ground-based detectors. A pathfinder prototype [4219], established at
Maastricht University in 2021, has already demonstrated promising results in key enabling technologies, paving
225

the way for the full-scale project. Like LISA, the ET will test the distance-redshift relation on cosmological
scales and provide independent measurements of H0 , free from the assumptions of the traditional distance ladder
[653]. This will primarily be achieved through the detection of binary compact object coalescences, which serve
as robust standard sirens. Beyond H0 , the ET will be uniquely equipped to detect SGWBs of cosmological
origin, offering a direct window into the early Universe. These observations could shed light on fundamental
processes such as inflation, the formation of primordial black holes, phase transitions in the early Universe,
and potential topological defects. The ET’s ability to probe these phenomena would significantly deepen our
understanding of the Universe’s origins and evolution, making it a cornerstone of future GW astronomy and
cosmology.
• Cosmic Explorer (CE): The proposed CE observatory [593, 594] will build on the design principles of LIGO,
but with significantly enhanced capabilities. Its arms, each spanning 40 km, will dramatically improve sensitivity,
particularly in the low-frequency range. This enhanced sensitivity will enable the detection of black hole-black
hole mergers at unprecedented distances and increase the detection rate to as many as 105 events per year.
With its ability to observe GWs in the 5–4000 Hz frequency band, CE will produce an unparalleled map of
the GW sky, reaching back to high redshifts of z ∼ 20 [4220, 4221]. By probing this deep into the Universe’s
history, CE will offer complementary insights into binary mergers, tracing their origins to the first stars and
providing critical information on the star formation rate and galaxy evolution over cosmic time. In addition to
astrophysical discoveries, CE will play a pivotal role in cosmology. By observing a vast catalog of dark sirens,
CE will enable precise inferences of H0 and provide a valuable independent check on the expansion history of
the Universe. This makes CE an essential component of next-generation GW astronomy and its intersection
with cosmology.
• TianQin: TianQin [4222, 4223] is a proposed Chinese space-based GW observatory designed to detect GWs in
the millihertz frequency range. It will consist of three spacecraft in a geocentric orbit, forming an equilateral
triangle with arm lengths of approximately 105 km. TianQin’s primary scientific objectives include the detection
of signals from supermassive black hole mergers, extreme mass ratio inspirals (EMRIs), and SGWBs. One of
TianQin’s unique features is its orbit near the Earth, which facilitates precise laser interferometry while mini-
mizing challenges associated with deep-space communication. Similar to LISA, TianQin will employ advanced
laser metrology and drag-free control to achieve exceptional sensitivity to low-frequency GWs. In the context of
cosmology, TianQin will contribute significantly to resolving cosmological tensions. By expanding the catalog of
both bright and dark sirens, TianQin will enable independent measurements of H0 and test the distance-redshift
relation on cosmological scales. Additionally, its observations will probe the early Universe, offering insights
into the SGWB and potential new physics beyond the standard model. TianQin is expected to launch in the
2030s, complementing other space-based observatories and advancing the era of precision GW cosmology.
• Taiji: Taiji [4224] is a proposed Chinese space-based GW observatory designed to detect GWs in the millihertz
frequency band. Unlike TianQin, which operates in a geocentric orbit, Taiji will be positioned at the Sun-Earth
Lagrange point L2 , providing a quieter observational environment and allowing for longer baselines. Its design
includes three spacecraft forming an equilateral triangle with arm lengths of 3 × 106 km, optimized for detecting
low-frequency GWs. Taiji’s scientific goals include probing the evolution of the Universe at redshifts as high
as z ∼ 20 and detecting GWs from sources such as massive black hole binaries, intermediate-mass black hole
mergers, and the SGWB. Its high sensitivity will also enable the study of rare cosmic phenomena inaccessible to
other observatories, offering an unparalleled view of early-Universe processes. In cosmology, Taiji will contribute
to addressing key tensions, including the H0 discrepancy, by expanding the catalog of dark sirens and providing
precise measurements of the distance-redshift relation. Additionally, Taiji will enhance our understanding of DE
and inflation by exploring signals from primordial GWs, phase transitions, and primordial black holes. Scheduled
for launch in the 2030s, Taiji complements the capabilities of both TianQin and LISA, focusing on high-redshift
phenomena and deepening our understanding of the early Universe while advancing China’s leadership in GW
astronomy.
• DECi-hertz Interferometer Gravitational Wave Observatory (DECIGO): DECIGO [4225] is a pro-
posed Japanese space-based GW observatory designed to fill the frequency gap between ground-based detectors
like LIGO and Virgo and low-frequency observatories like LISA. With a target frequency range centered around
decihertz (0.1–10 Hz), DECIGO will enable the study of intermediate-mass black hole binaries, early Universe
GW backgrounds, and other phenomena inaccessible to existing detectors. The DECIGO mission will consist
of three spacecraft forming a triangular configuration with arm lengths of 1, 000 km, operating in a heliocen-
tric orbit. The observatory will use highly sensitive laser interferometry and drag-free control technologies to
achieve unprecedented precision in GW detection. A prototype mission, B-DECIGO, is planned to demonstrate
key technologies in Earth orbit before the full-scale DECIGO mission launches. In cosmology, DECIGO will
226

provide unique insights by detecting GWs from primordial sources such as phase transitions, cosmic strings, and
inflation. Its sensitivity to the SGWB will allow for constraints on the physics of the early Universe, offering
a direct window into energy scales far beyond those probed by CMB. DECIGO will also expand the catalog
of standard sirens, enabling precise measurements of H0 and the distance-redshift relation, which are crucial
for resolving cosmological tensions. DECIGO’s ability to study GWs from z ∼ 1000 to the present will bridge
the gap between early-Universe physics and late-time structure formation, providing complementary data to
missions like LISA, Taiji, and TianQin. Expected to launch in the mid-2030s, DECIGO represents a major
advancement in GW astronomy, with profound implications for cosmology and fundamental physics.
• Pulsar Timing Arrays (PTAs): Pulsar Timing Arrays are an innovative approach to detecting low-frequency
GWs in the nanohertz frequency band. Unlike ground- or space-based interferometers, PTAs utilize the pre-
cise timing of millisecond pulsars as natural clocks to measure distortions in spacetime caused by passing
GWs. International collaborations, such as the North American Nanohertz Observatory for Gravitational
Waves (NANOGrav) [4226], the European Pulsar Timing Array (EPTA), the Parkes Pulsar Timing Array
(PPTA) [4227], and the International Pulsar Timing Array (IPTA) [4228], form a global network to maximize
sensitivity and coverage. PTAs are particularly sensitive to GWs produced by supermassive black hole binaries
(SMBHBs), which emit at nanohertz frequencies during their inspiral phase. Additionally, PTAs can probe the
SGWB arising from the superposition of signals from numerous SMBHBs or from cosmological sources, such
as cosmic strings, phase transitions, or inflation. In cosmology, PTAs provide unique insights by constraining
the evolution of structure and the formation of massive galaxies. These measurements complement other GW
observatories by covering a distinct frequency range, extending the spectrum of observable GWs. PTAs can
also indirectly contribute to addressing the H0 tension by improving our understanding of galaxy mergers and
structure growth. Recent breakthroughs, such as the NANOGrav 15-year data release, have hinted at the first
detection of a SGWB. These results, if confirmed, would open a new window into the Universe’s evolution and
provide evidence for processes occurring at very high energy scales, far beyond the reach of current particle
accelerators. PTAs represent an essential component of the global GW detection strategy, providing a unique
and complementary perspective to ground- and space-based observatories. With the advent of next-generation
radio telescopes like SKAO, the sensitivity of PTAs is expected to dramatically improve, enabling more precise
measurements and extending the range of detectable sources.

5.1.5. 21 cm Cosmology
The 21 cm line, arising from the hyperfine transition of neutral hydrogen (HI), provides a powerful and versatile tool
for probing the Universe’s structure and evolution across a wide range of redshifts. This signal serves as a unique
tracer of the IGM and large-scale structure, enabling the study of key epochs such as the cosmic dawn, the EoR,
and the post-reionization era. Unlike traditional probes, 21 cm cosmology captures three-dimensional information,
offering tomographic insights into the distribution of matter and its interaction with radiation fields over cosmic time.
The potential of the 21 cm line extends beyond mapping large-scale structure. It offers a novel avenue to address
key cosmological tensions, including constraints on H0 and the amplitude of matter fluctuations S8 . Moreover, the
sensitivity of 21 cm surveys to high-redshift phenomena makes them an invaluable complement to other cosmological
probes such as CMB, BAO, and WL.
In this part, we outline the current and future 21 cm experiments, their capabilities, and their expected contributions
to cosmology. Special emphasis is placed on their role in addressing fundamental questions about the early Universe,
the formation of cosmic structures, and the underlying physics driving cosmic tensions:

• Low-Frequency Array (LOFAR): LOFAR [2339] is a cutting-edge radio interferometer designed to observe
the Universe at low radio frequencies, operating primarily in the range of 10 MHz to 240 MHz. With its core
located in the Netherlands and stations distributed across Europe, LOFAR provides high-resolution imaging
and wide-field observations. A key focus of LOFAR is probing the 21 cm hydrogen signal from the EoR, offering
insights into the period when the first stars and galaxies ionized the IGM. By mapping the structure of neutral
hydrogen during this era, LOFAR aims to uncover the processes governing cosmic reionization and the emergence
of the first luminous sources. Additionally, LOFAR’s sensitivity to low-frequency signals makes it a valuable tool
for studying other phenomena, including cosmic magnetism, DM, and the large-scale structure of the Universe.
The array’s modular and scalable design allows it to complement ongoing and future surveys.
• Murchison Widefield Array (MWA): MWA [4229] is a cutting-edge low-frequency radio interferometer
located in Western Australia, designed to observe the Universe at frequencies between 80 MHz and 300 MHz.
With its compact configuration of 4,096 dipole antennas spread across 1.5 kilometers, the MWA excels at
capturing wide-field, low-frequency observations. A primary objective of the MWA is to detect the 21 cm
hydrogen signal from the EoR, enabling a detailed study of the processes that ionized the early Universe
227

and the formation of the first stars and galaxies. Additionally, the MWA probes large-scale cosmic structures,
providing insights into the distribution of matter and the role of DE in cosmic evolution. The array is a precursor
instrument to the SKAO and serves as a testbed for developing advanced techniques in calibration, foreground
removal, and signal extraction. Its location at the radio-quiet Murchison Radio-astronomy Observatory ensures
minimal interference, enhancing its sensitivity to faint cosmological signals. Beyond cosmology, the MWA is
also used to study solar physics, ionospheric science, and transient phenomena, making it a versatile instrument
for advancing astrophysical research.

• Hydrogen Epoch of Reionization Array (HERA): HERA [2340] is a dedicated radio interferometer de-
signed specifically to study the 21 cm hydrogen signal from the EoR. Located in the radio-quiet Karoo region of
South Africa, HERA consists of 350 closely packed parabolic dishes, each 14 m in diameter, optimized for high
sensitivity to faint cosmological signals at redshifts corresponding to the EoR (6 < z < 12). HERA’s primary
goal is to provide a detailed characterization of the IGM during the EoR, tracing the formation and evolution
of the first luminous sources—stars, galaxies, and black holes—and their impact on the ionization state of the
Universe. By mapping fluctuations in the 21 cm emission, HERA will constrain the timing and progression of
reionization, offering insights into the astrophysical processes driving it. The array’s design prioritizes simplicity
and precision calibration, which are critical for mitigating systematics and isolating the faint EoR signal from
overwhelming foreground contamination. As a successor to PAPER, HERA incorporates lessons learned from
earlier experiments, significantly improving sensitivity and resolution. HERA is expected to complement other
21 cm experiments like LOFAR, MWA, and SKAO by focusing specifically on the EoR. Its measurements will
also provide valuable data for testing models of cosmic structure formation, the properties of the first galaxies,
and potential new physics beyond the standard cosmological model.

• The Radio Experiment for the Analysis of Cosmic Hydrogen (REACH): REACH [4230] is an inno-
vative experiment designed to detect the globally averaged 21 cm signal from neutral hydrogen, targeting the
cosmic dawn and the EoR. What sets REACH apart is its novel calibration techniques and an advanced antenna
design aimed at mitigating the impact of instrumental systematics and foreground contamination, long-standing
challenges in 21 cm cosmology. By achieving unprecedented precision in isolating the faint cosmological sig-
nal, REACH opens a new window into the thermal and ionization history of the early Universe. Beyond its
methodological advancements, REACH has the potential to address broader cosmological tensions. Its ability
to constrain the timing and physics of reionization provides indirect insights into the nature of DM and DE.
Additionally, deviations in the 21 cm signal from standard theoretical predictions could indicate new physics,
such as interactions between baryons and DM or non-standard early heating mechanisms. These contributions
place REACH at the forefront of innovative probes for resolving foundational challenges in cosmology.

• Canadian Hydrogen Intensity Mapping Experiment (CHIME): Located at the Dominion Radio Astro-
physical Observatory in Canada, CHIME [4202] operates in the 400–800 MHz band, corresponding to a redshift
range of 0.8 to 2.5. Its primary goal is to map neutral hydrogen intensity across this range to measure the
expansion history of the Universe. By producing the largest three-dimensional map of hydrogen intensity in this
critical epoch, CHIME provides unprecedented insights into the large-scale structure of the Universe and probes
cosmic evolution during the dark ages through the 21 cm transition line. In addition to tracing the evolution
of structure, CHIME offers statistical constraints on BAO using the Hydrogen Intensity Mapping technique.
These measurements enable synergistic analyses with other cosmological surveys, enhancing the precision of
constraints on fundamental parameters like H0 and the amplitude of matter fluctuations S8 . Since beginning
operations in 2018, CHIME has already yielded significant data, with planned upgrades set to further improve
the precision and resolution of intensity maps.

• Giant Metrewave Radio Telescope (GMRT): Located near Pune, India, the GMRT [4231] is an array
of 30 fully steerable parabolic dishes, each 45 m in diameter, operating in the frequency range of 50 MHz to
1.5 GHz. It is one of the most sensitive radio telescopes for low-frequency observations and plays a significant
role in 21 cm cosmology. GMRT has been used extensively to study the 21 cm hydrogen signal from the Cosmic
Dawn and the EoR. Its ability to probe neutral hydrogen at different redshifts provides critical insights into
the thermal and ionization history of the IGM and the formation of the first luminous sources. In addition to
EoR studies, GMRT contributes to understanding large-scale structures, galaxy formation, and the properties
of DE and DM. As an established facility, GMRT continues to produce high-impact results and serves as a vital
complement to ongoing and upcoming experiments such as LOFAR, HERA, and SKAO.

• Five-hundred-meter Aperture Spherical Telescope (FAST): FAST [4232], located in Guizhou, China, is
the world’s largest single-dish radio telescope, with an aperture of 500 m. Operating in the frequency range of
228

70 MHz to 3 GHz, FAST is a highly versatile instrument capable of a broad range of astrophysical studies, includ-
ing 21 cm observations. Its massive collecting area provides unparalleled sensitivity, making it an invaluable tool
for studying neutral hydrogen in the Universe. FAST contributes to 21 cm cosmology by probing the large-scale
distribution of hydrogen gas, investigating the EoR, and tracing the evolution of cosmic structures during the
Cosmic Dawn. Its ability to perform both single-dish and interferometric measurements allows it to complement
other instruments in the global effort to constrain models of cosmic evolution. Beyond 21 cm studies, FAST is
used for pulsar searches, extragalactic surveys, and studies of ISM dynamics. With its extraordinary sensitivity
and high precision, FAST is poised to make new contributions to our understanding of the early Universe and
fundamental physics.
• Hydrogen Intensity and Real-time Analysis eXperiment (HIRAX): Building on the technology of
CHIME, HIRAX [1839] is located near the SKAO site in South Africa and is designed to achieve similar
scientific objectives in the southern sky. Operating in the 400–800 MHz frequency band, HIRAX aims to map
neutral hydrogen intensity over the redshift range of 0.8 to 2.5, providing critical insights into the evolution of
the Universe during this epoch. The array consists of 1, 000 closely packed 6 m radio dishes, which will create
a high-resolution hydrogen intensity map of the southern sky using the 21 cm emission line. Initially conceived
as a demonstration of SKAO technology, HIRAX has evolved into an ambitious project, poised to deliver a
detailed three-dimensional map of large-scale structure in this redshift range. A primary science goal of HIRAX
is to measure BAO signature in the matter distribution, enabling precise constraints on the expansion history
of the Universe. HIRAX’s southern sky coverage complements CHIME’s northern sky observations, offering an
opportunity for joint analyses to refine measurements of cosmological parameters, including H0 and S8 .
• Square Kilometre Array Observatory (SKAO): The SKAO [4208] is expected to begin operations for
Phase 1 in the late 2020s, with sites located in Australia (covering low-frequency observations) and South Africa
(mid-frequency observations). Together, these will form the world’s largest radio telescope system, with a total
collecting area of approximately 1 km2 , enabling unprecedented sensitivity and resolution. Operating across
a frequency range of 50 MHz to 14 GHz, the SKAO will span a physical radius of roughly 3, 000 km, hosting
hundreds of radio dishes and over 100,000 antennas. By producing the most comprehensive 21 cm hydrogen
intensity maps to date, the SKAO will probe the large-scale structure of the Universe, reaching back to the
first galaxies and stars. This ambitious project will provide new insights into the EoR, the emergence of the
first luminous sources, and the role of DE in shaping cosmic evolution. Additionally, it will complement other
observational probes, offering synergies with WL, BAO, and CMB experiments to refine our understanding of
cosmological tensions.

5.1.6. Type Ia supernovae and distance ladder


SNIa play a pivotal role in modern cosmology as highly reliable standard candles for measuring cosmic distances.
By leveraging the uniform intrinsic luminosity of these SNIa, calibrated through the cosmic distance ladder, precise
determinations of H0 are possible. The cosmic distance ladder comprises several hierarchical steps: (1) geometric
distance measurements using Cepheid variables, TRGB stars, masers, etc.; (2) calibration of SNIa luminosities based
on these measurements; and (3) application of SNIa to measure H0 in the Hubble flow, where cosmic expansion
dominates. Recent advancements in distance ladder measurements have refined the determination of H0 , notably
through Cepheid-calibrated SNIa in the SH0ES project, which reports a value significantly higher than early Universe
estimates from the CMB. This discrepancy, known as the Hubble tension, remains one of the most pressing challenges
in cosmology. Complementary approaches, including TRGB, Mira variables, and SBF methods, offer alternative
calibrations and independent cross-checks, helping to ensure robustness and reduce systematic biases. However, all
these methods consistently yield higher H0 values than those derived from early Universe probes, highlighting the
tension.
Future surveys and advanced instrumentation promise to transform SNIa measurements and distance ladder stud-
ies. JWST’s unprecedented resolution has already begun improving Cepheid observations in crowded fields, while
upcoming missions aim to reduce systematic uncertainties further. New facilities will also increase the number and
diversity of SNIa observations, extending measurements to higher redshifts and improving statistical power. Below, we
highlight ongoing and planned SN and distance ladder experiments and their contributions to addressing cosmological
tensions.

• Dark Energy Survey (DES): DES [4214], conducted from 2013 to 2019, used the 4-meter Blanco Telescope
at Cerro Tololo Inter-American Observatory (CTIO) to observe approximately 5, 000 deg2 of the southern sky.
DES contributed significantly to supernova cosmology by discovering and monitoring thousands of SNIa across
a broad redshift range (z ∼ 0.01 to 1.2). Its high-quality photometry enabled precise constraints on H0 and DE
properties. DES complements other distance ladder experiments by providing a statistically robust dataset of
229

SNIa at intermediate redshifts, improving constraints on the cosmic expansion history and addressing potential
systematics in supernova cosmology.
• The Gaia mission: Gaia [143], operating since 2013, has revolutionized the cosmic distance ladder by providing
precise parallaxes for millions of stars, including Cepheid variables and TRGB stars. These highly accurate
measurements are critical for calibrating the absolute luminosities of SNIa and improving the determination
of the Hubble constant. Gaia’s unparalleled astrometric precision significantly reduces uncertainties in the
local distance scale and mitigates systematic biases in SNIa calibrations. By anchoring the first rung of the
distance ladder, Gaia plays a key role in addressing the Hubble tension and refining our understanding of cosmic
expansion.
• Zwicky Transient Facility (ZTF): The ZTF [4233], operating since 2018 at the Palomar Observatory, utilizes
a 48-inch Schmidt Telescope with a wide field of view (47 deg2 ) for high-cadence transient surveys. ZTF has
been instrumental in discovering and monitoring nearby SNIa, providing critical data for calibrating the distance
ladder. By targeting local SNIa, ZTF supports precise measurements of H0 and cross-validates calibrations
derived from other standard candles, such as Cepheids and TRGB stars. Its extensive dataset complements
space-based missions like Gaia and JWST by increasing the sample size of well-observed SNIa.
• Nearby Supernova Factory (SNfactory): The SNfactory [4234], an international collaboration, focuses on
obtaining detailed spectrophotometry of low-redshift SNIa. Using the SuperNova Integral Field Spectrograph
(SNIFS) on the University of Hawaii’s 2.2-meter telescope, SNfactory has been essential in characterizing the
diversity of SNIa and improving standardization techniques. SNfactory’s precise spectrophotometric data reduce
systematic uncertainties in SNIa luminosities, refining the calibration of the distance ladder and improving
constraints on H0 .
• Asteroid Terrestrial-impact Last Alert System (ATLAS): Although primarily designed for asteroid de-
tection, ATLAS [4235], operating on multiple telescopes in Hawaii, South Africa, and Chile, has made substantial
contributions to supernova cosmology. Its wide field of view and high-cadence observations are well-suited for
identifying and monitoring nearby SNIa. ATLAS’s discoveries provide valuable data for calibrating the distance
ladder, especially at low redshifts, complementing ground-based and space-based missions in improving the
accuracy of H0 measurements and addressing cosmological tensions.
• James Webb Space Telescope (JWST): JWST [291], launched in 2021, provides unparalleled near-infrared
imaging and spectroscopy, enabling precise observations of Cepheid variables and TRGB stars in crowded fields.
These measurements significantly enhance the calibration of SNIa and improve the accuracy of H0 . JWST’s
ability to observe faint objects at high resolution allows it to extend the cosmic distance ladder to greater
distances and reduce systematic uncertainties. By complementing ground- and space-based surveys, JWST
plays a crucial role in addressing the Hubble tension and probing potential new physics beyond ΛCDM.
• Rubin Observatory’s Legacy Survey of Space and Time (LSST): The Rubin Observatory [4207], cur-
rently under construction in Chile, will conduct the LSST, a ten-year survey covering approximately 18, 000 deg2
of the southern sky. With its 8.4-meter aperture and a 9.6-square-degree field of view, Rubin’s LSST will discover
and monitor hundreds of thousands of SNIa across a broad redshift range (z ∼ 0.01 to 1.2). Rubin’s LSST will
revolutionize supernova cosmology by significantly increasing the sample size and improving statistical precision
in measuring H0 . Its high-cadence, multi-band photometry will mitigate systematics such as dust extinction
and host-galaxy effects, and its synergy with spectroscopic follow-up surveys will ensure robust classifications
and redshift measurements.
• Roman Space Telescope (formerly WFIRST): The Roman Space Telescope [892], scheduled for launch
in the late 2020s, will perform a High Latitude Survey to observe thousands of SNIa across a broad redshift
range (z ∼ 0.1 to 2.0). With its wide field of view and high-resolution imaging in the near-infrared, Roman
will extend the cosmic distance ladder, improving the precision and accuracy of the Hubble constant. Ro-
man’s deep multi-band photometry will reduce systematics like dust extinction and enhance the calibrations of
standard candles. By complementing ground-based surveys such as Rubin’s LSST and incorporating follow-up
spectroscopy, Roman will play a critical role in resolving the Hubble tension.
• Extremely Large Telescope (ELT): The ELT [4236], currently under construction in Chile, will leverage its
39-meter aperture and advanced adaptive optics to achieve unprecedented precision in SNIa observations. By
providing high-resolution imaging and spectroscopy, the ELT will play a vital role in calibrating the distance
ladder through Cepheids, TRGB stars, and direct measurements of SNIa in the local Universe. The ELT’s
ability to observe SNIa at higher redshifts will complement space-based surveys like Rubin’s LSST and Roman,
reducing systematic uncertainties and improving the determination of the Hubble constant.
230

• Future Extremely Large Telescopes (GMT and TMT): Beyond the ELT, upcoming ground-based obser-
vatories such as the Giant Magellan Telescope (GMT) [4237] and the Thirty Meter Telescope (TMT) [4238] will
significantly enhance SNIa and distance ladder measurements. With their large apertures of 24.5 meters (GMT)
and 30 meters (TMT), these telescopes will provide unparalleled resolution and sensitivity, enabling precise ob-
servations of Cepheids, TRGB stars, and SNIa in both the local and high-redshift Universe. These telescopes will
extend the reach of the distance ladder by detecting fainter Cepheids and SNIa at greater distances, improving
calibration precision and reducing systematic uncertainties. Their advanced instrumentation will complement
space-based missions like JWST and Roman by providing high-resolution follow-up spectroscopy and imaging,
further refining H0 and contributing to resolving the Hubble tension.

5.1.7. Time delay cosmography


The strong lensing of RGBs and SN offers a unique opportunity to measure the expansion rate of the Universe across
different redshifts. This approach leverages the variable nature of these objects, enabling the use of the time-delay
method—an independent high-redshift cosmological probe that does not rely on the distance ladder. The technique
involves observing multiple images of a strongly lensed source, such as a RGB or SN, produced by an intermediary
lensing galaxy. The mass profile of this galaxy, combined with the respective distances between the observer, lens,
and source, determines the offsets of the lensed images and is crucial for obtaining precise estimates of cosmological
parameters.
This method relies on the time delays between light rays traveling along different paths, caused by both geometric
differences in the light geodesics and the distribution of matter within the lensing galaxy. By accurately measuring
these delays, the time-delay method provides an estimate of the Hubble constant, independent of large-scale structure
parameters.
Strong lensing is inherently challenging due to the specific alignment required between the observer, lens, and
source. Moreover, the rarity of suitable RGBs or SN reduces the probability of detection. Despite these challenges,
over 300 strongly lensed RGBs have been identified to date, typically in the redshift range ∼ 1 − 3, with lensing
galaxies found at redshifts ∼ 0.2 − 0.8. The lensed images are usually separated by a few arcseconds, well within
the resolution capabilities of modern ground-based telescopes. Time delays ranging from a few days to a year are
sufficient for accurate measurements, making the time-delay method a powerful tool for cosmology.
Below, we highlight ongoing and planned experiments leveraging the time-delay method and their contributions to
measuring H0 and addressing cosmological tensions.

• Very Large Telescope Interferometer (VLTI): The VLTI [4239], operated by the European Southern Ob-
servatory (ESO), provides ultra-high angular resolution imaging through its interferometric array of telescopes.
With its ability to resolve fine details in strongly lensed RGBs, VLTI plays a critical role in time-delay cosmog-
raphy by enabling precise measurements of the lensing geometry and time delays. Its high spatial resolution is
particularly useful for modeling lensing galaxies and reducing systematic uncertainties in the determination of
H0 .

• Atacama Large Millimeter/submillimeter Array (ALMA): ALMA [4240], located in Chile, offers un-
paralleled sensitivity and resolution for observing lensed submillimeter galaxies and RGBs. By measuring time
delays in radio and submillimeter wavelengths, ALMA provides a complementary approach to optical time-delay
cosmography, reducing systematic biases and refining constraints on the Hubble constant.

• Keck Observatory: The Keck Observatory [4241], located on Maunakea, Hawaii, features two 10-meter tele-
scopes equipped with advanced adaptive optics systems. These capabilities allow Keck to obtain high-resolution
imaging and spectroscopy of strongly lensed RGBs and SN, essential for precise time-delay measurements. Keck’s
sensitivity and angular resolution enable detailed modeling of lensing galaxies and accurate determination of
light path geometries, contributing to independent constraints on H0 . Its role as a follow-up instrument for
time-delay cosmography complements surveys like Rubin’s LSST and Euclid, providing critical data for refining
cosmological models.

• Subaru Telescope: The Subaru Telescope [4205], a 8.2-meter optical-infrared telescope located on Maunakea,
Hawaii, plays a key role in observing strong lensing systems for time-delay cosmography. Equipped with instru-
ments such as the HSC, Subaru excels at wide-field imaging and detailed follow-up of lensed systems. Subaru’s
capabilities allow it to identify and resolve lensed RGB systems and SN, particularly in crowded fields, enabling
precise time-delay measurements. Its contributions to modeling lensing mass distributions and improving light
path geometries are vital for reducing systematic uncertainties in the determination of H0 . Subaru serves as a
valuable complement to space-based missions like Roman and ground-based facilities like Keck.
231

• Euclid: The Euclid mission [4204], covering a 14, 000 deg2 area of the sky, is expected to observe approximately
170, 000 strong lensing events during its lifetime. These events will provide valuable insights into the properties
of DM halos in dwarf galaxies, the mass distribution within galaxies, and galaxy cluster structures in the
redshift range 0 < z < 2. Strong lensing events are inherently rare and require robust detection methods.
To address this challenge, Euclid employs CNNs, which are trained on extensive simulated datasets to extract
image features and identify strong lensing properties. These advanced pipelines allow for precise identification
of lensing phenomena, including sharp Einstein rings, which are expected to be detectable down to an angular
resolution of 0.5 arcseconds. Additionally, Euclid is anticipated to detect up to ∼ 2300 strongly lensed RGB
sources, enabling precise measurements of time delays. The time-delay cosmography enabled by Euclid’s high-
resolution imaging will place stringent constraints on H0 , making it one of the most powerful tools for addressing
cosmological tensions.
• Rubin Observatory’s Legacy Survey of Space and Time (LSST): The Rubin Observatory’s [4207]
unparalleled depth and sky coverage will enable the discovery of up to ten million RGBs over its operational
lifetime, spanning a redshift range of z ∼ 0 − 7. This extensive dataset will have profound implications for
RGB-related physics, including strong lensing and time-delay cosmography. Rubin’s LSST is expected to detect
approximately ∼ 2600 time-delayed lensing systems, assuming optimistic projections. These systems will play
a pivotal role in refining measurements of H0 and probing the dynamical properties of DE at high redshifts.
With its deep and wide survey capabilities, Rubin’s LSST will significantly enhance the statistical precision of
time-delay cosmography, providing critical insights into the expansion history of the Universe.
• Roman Space Telescope (formerly WFIRST): The Roman Space Telescope [892], set to launch in the late
2020s, will make significant contributions to time-delay cosmography through its deep, high-resolution near-
infrared imaging and wide field of view. Roman is expected to detect thousands of strongly lensed RGBs and
SN, enabling precise measurements of time delays across a broad redshift range. With its exceptional sensitivity,
Roman will identify and characterize lensing systems with high precision, including detailed modeling of lensing
galaxies. These observations will reduce uncertainties in the Hubble constant H0 and provide crucial constraints
on the properties of DE. Roman’s synergy with other surveys, such as Rubin’s LSST, will further enhance the
statistical power and accuracy of time-delay cosmography.
• Square Kilometre Array Observatory (SKAO): The SKAO [4208], with its radio arrays in South Africa and
Australia, will enable high-resolution observations of strongly lensed radio sources. By detecting and monitoring
time-delayed radio-lensed systems, SKAO will provide independent measurements of H0 . Its sensitivity to faint
radio signals and ability to resolve sub-arcsecond structures make it a powerful tool for time-delay cosmography.
• Extremely Large Telescope (ELT): The ELT [4236], currently under construction in Chile, will leverage its
39-meter aperture to achieve unparalleled angular resolution and sensitivity. This will allow for detailed imaging
and spectroscopy of strongly lensed RGBs and supernovae, facilitating precise time-delay measurements. The
ELT’s advanced instrumentation will help model lensing galaxies with greater accuracy, reducing uncertainties
in the determination of the Hubble constant.
• Thirty Meter Telescope (TMT): The TMT [4238], planned for construction in Hawaii or the Canary Islands,
will provide high-resolution imaging and spectroscopic capabilities, enabling detailed observations of strongly
lensed systems. By resolving lensed images with extreme precision, the TMT will contribute to time-delay
cosmography by improving measurements of H0 and probing the dynamics of DE.
• Giant Magellan Telescope (GMT): The GMT [4237], under construction in Chile, will utilize its segmented
24.5-meter primary mirror to achieve exceptional imaging and spectroscopic capabilities. Its high resolution and
sensitivity will support time-delay cosmography by enabling precise modeling of lensed systems and accurate
measurements of H0 , particularly at intermediate redshifts.

5.1.8. Fast Radio Bursts (FRBs)


FRBs are extremely bright, short-duration radio transients lasting only a few milliseconds, with the potential to serve
as powerful tools for probing cosmology. Their extragalactic origin is supported by their large dispersion measures,
which are observed to evolve with redshift. A significant number of FRBs have been localized to host galaxies, allowing
their redshifts to be determined and their origins, such as possible magnetar emissions, to be studied.
Localized FRBs provide valuable insights into several cosmological phenomena, including the baryon fraction in
the IGM, strong lensing events, the equivalence principle, the expansion history of the Universe, and the cosmic
reionization history. The interaction of FRBs with the IGM makes them excellent probes of baryonic matter in
these regions, as the interaction of radio signals with gas distributions offers a new perspective on the “missing
232

baryon problem”. Regarding the Hubble constant H0 , FRBs alone face degeneracies that limit their ability to directly
constrain H0 . However, combining FRB data with complementary datasets such as CMB, BBN, or GW measurements
breaks these degeneracies and enables precise cosmological constraints.
Below, we highlight the list of ongoing and planned experiments leveraging FRBs to address cosmological tensions
and improve our understanding of the Universe’s evolution.

• Canadian Hydrogen Intensity Mapping Experiment (CHIME): CHIME [4202], situated in British
Columbia, Canada, is a pioneering radio telescope designed to monitor transient radio signals, including FRBs.
Its fixed-cylinder design and wide field of view allow CHIME to observe the sky continuously, detecting thousands
of FRBs annually across a range of redshifts. This high detection rate enables the identification of rare FRBs
suitable for cosmological studies, such as those with high dispersion measures or unique signatures that indicate
strong lensing events. CHIME’s advanced time-domain sensitivity is optimized for detecting short-duration
radio transients, making it a valuable tool for measuring the arrival times of FRBs with high precision. These
capabilities allow CHIME to constrain propagation effects, such as delays caused by intervening structures, which
can be cross-referenced with GW studies to extract cosmological information. Its unparalleled FRB detection
volume provides a complementary dataset for investigating the evolution of the cosmic expansion when paired
with optical and GW experiments.

• Australian Square Kilometre Array Pathfinder (ASKAP): ASKAP [4242], located in Western Australia,
is a cutting-edge radio telescope optimized for detecting and localizing FRBs. Its unique combination of wide
field-of-view and interferometric imaging capabilities allows ASKAP to efficiently detect and pinpoint FRBs
to sub-arcsecond accuracy. This precise localization is critical for identifying host galaxies and determining
redshifts of FRB sources. ASKAP’s advanced beamforming technology enables the detection of high-redshift
FRBs, providing insights into the IGM and the evolution of the cosmic expansion. Its ability to associate FRBs
with host galaxies allows for independent constraints on H0 when combined with other cosmological probes.

• Upgrade of the Molonglo Observatory Synthesis Telescope (UTMOST): UTMOST [4243], located in
Australia, is a reconfigured radio telescope optimized for detecting FRBs. Its high-cadence observing strategy
makes it ideal for the real-time discovery and timing of FRBs, particularly at low redshifts. Although less sensi-
tive than newer facilities, UTMOST’s rapid response capability enables timely follow-up studies and statistical
surveys of FRBs. By characterizing nearby FRBs, UTMOST complements high-sensitivity instruments such
as ASKAP and FAST, contributing to our understanding of the local FRB population and its cosmological
implications.

• MeerKAT: MeerKAT [4244], situated in South Africa, is a highly sensitive radio telescope designed for high-
resolution imaging and precision timing, making it an invaluable tool for studying FRBs. Its wide frequency
coverage and long baselines enable MeerKAT to detect faint and distant FRBs, particularly those at high
redshifts, offering insights into the early Universe. MeerKAT excels in pinpointing FRB host galaxies and
measuring dispersion measures with high precision. This allows for a detailed study of the large-scale structure
of the Universe and the evolution of baryonic matter over cosmic time. MeerKAT also plays a complementary role
to SKAO by serving as a pathfinder and refining techniques for future large-scale FRB surveys. Its contributions
to the Hubble constant determination and reionization studies position MeerKAT as a crucial instrument for
resolving cosmological tensions.

• Five-hundred-meter Aperture Spherical Telescope (FAST): FAST [4232], located in Guizhou, China,
is the world’s largest single-dish radio telescope, offering unparalleled sensitivity for detecting faint and distant
FRBs. Its ability to detect FRBs with extremely high signal-to-noise ratios enables precise measurements of
dispersion measures and localization for follow-up studies. FAST contributes significantly to cosmology by
detecting high-redshift FRBs and analyzing their interaction with the IGM. Its sensitivity allows it to probe the
cosmic expansion and test reionization models, complementing large-scale surveys such as SKAO and CHIME.
FAST’s unique capabilities make it a critical instrument for advancing FRB-based cosmological studies.

• Deep Synoptic Array (DSA-110): The DSA-110 [4245], located in California, is a dedicated radio array
designed exclusively for detecting and localizing FRBs. With 110 antennas operating in the 1.28 GHz band,
the array achieves sub-arcsecond localization accuracy, enabling precise identification of FRB host galaxies.
DSA-110’s focus on real-time detection and localization provides critical insights into the redshift distribution
of FRBs and their cosmological applications. By linking FRBs to their host environments, DSA-110 contributes
to breaking degeneracies in cosmological parameter estimation, particularly when combined with other datasets
such as GW observations and optical surveys.
233

• Giant Metrewave Radio Telescope (GMRT): The GMRT [4231], located in Pune, India, has been instru-
mental in studying FRBs at low radio frequencies. Its unique sensitivity and wide frequency coverage make it
a valuable tool for characterizing the spectral properties of FRBs and their interaction with the IGM. GMRT
has been used for follow-up observations of FRBs detected by other instruments, particularly at frequencies
below 800 MHz. These observations complement higher-frequency detections by telescopes like CHIME and
ASKAP, providing new insights into FRB emission mechanisms and their propagation through cosmic struc-
tures. While not a dedicated FRB facility, GMRT’s versatility and precision continue to support advancements
in FRB cosmology.
• Very Large Array Low-band Ionosphere and Transient Experiment (VLITE): VLITE [4246], operat-
ing on the Very Large Array (VLA), focuses on low-frequency transient detection, including FRBs. Its ability to
monitor the sky continuously allows it to identify faint FRBs and investigate their spectral properties at lower
frequencies.
• Square Kilometre Array Observatory (SKAO): The SKAO [4208], under construction in South Africa and
Australia, is poised to become the largest and most sensitive radio telescope array in the world. Its exceptional
capabilities will enable precise localization of FRBs at sub-arcsecond resolution and direct measurement of their
redshifts, significantly advancing FRB cosmology. By detecting and studying FRBs across a broad redshift range,
SKAO will provide insights into the evolution of the IGM, reionization history, and cosmic expansion. Its ability
to map dispersion measures over cosmic time will complement other datasets, such as GW observations and
optical surveys, to break degeneracies in cosmological parameters, including H0 . While its primary focus is on
large-scale structure studies through 21 cm intensity mapping, SKAO’s FRB-specific contributions, including
redshift localization and host galaxy characterization, make it a new instrument for addressing cosmological
tensions.
• Next Generation Deep Synoptic Array (DSA-2000): DSA-2000 [4247], a planned expansion of the
DSA-110, will feature 2,000 antennas, dramatically increasing sensitivity and detection rates for FRBs. This
next-generation array will provide precise localization for tens of thousands of FRBs across a wide redshift
range. With its enhanced capabilities, DSA-2000 will enable detailed studies of the cosmic expansion, baryon
distribution, and reionization history.
• Low-Frequency Array (LOFAR2.0): LOFAR2.0 [1144], an upgraded version of the Low-Frequency Array,
will provide enhanced sensitivity and temporal resolution for detecting FRBs at low frequencies. Its broad
frequency coverage and improved capabilities will allow it to probe previously inaccessible redshift ranges for
FRBs, making it a unique tool for studying the early Universe. LOFAR2.0’s ability to detect low-frequency
FRBs will contribute to understanding their emission mechanisms and their interaction with the IGM.
• APERTure Tile In Focus (Apertif ): Apertif [4248], an upgrade to the Westerbork Synthesis Radio Tele-
scope (WSRT) in the Netherlands, provides wide-field imaging for detecting FRBs. Its combination of sensitivity
and survey speed enables it to identify transient events, contributing to population studies of FRBs and com-
plementing other instruments in the field.
• Burst Transient Telescope (BURSTT): BURSTT is a planned experiment designed to search for FRBs
with enhanced detection rates and localization capabilities. Its wide field of view will make it particularly
effective in identifying high-redshift FRBs and improving constraints on their cosmological applications.
• Canadian Hydrogen Observatory and Radio-transient Detector (CHORD): CHORD [4249], a planned
successor to CHIME, will combine higher sensitivity with improved localization capabilities for FRBs. CHORD
is expected to detect tens of thousands of FRBs across a broad redshift range, providing critical insights into
the evolution of cosmic structures and addressing cosmological tensions.
• Pathfinder for Real-time Extragalactic Cosmic Intensity Survey Experiment (PRECISE): PRE-
CISE is a dedicated pathfinder for FRB studies, focusing on real-time detection and precise localization of
FRBs. It will enhance the study of their host galaxies and the IGM, making it a valuable addition to the next
generation of FRB experiments.

5.1.9. Line-intensity mapping


The large-scale structure of the Universe can be surveyed through the radiation intensity emitted by gas clouds, in
a manner similar to galaxy surveys, but without resolving individual sources. This approach significantly accelerates
the survey process. While many surveys focus exclusively on the 21 cm line associated with the hyperfine transition
of neutral hydrogen, line-intensity mapping (LIM) extends this technique to include additional spectral lines, such as
234

carbon monoxide (CO) and ionized carbon (CII), which provide complementary tracers of large-scale structure and
galaxy evolution. The intensity distribution maps derived from LIM can be translated into matter fluctuation power
spectra and enable studies across a broader range of cosmic epochs.
Unlike traditional 21 cm experiments, which focus specifically on the neutral hydrogen distribution, LIM captures
the integrated emission of multiple spectral lines, making it a more versatile approach. This broader capability allows
LIM to probe different phases of the ISM and the EoR, as well as to reduce parameter degeneracies by spanning a
wider range of redshifts and tracers.
This innovative technique employs angular resolution observations to trace the spatial distribution of matter, even
at high redshifts where traditional galaxy surveys face challenges. If foreground contamination can be effectively
mitigated, line-intensity mapping has the potential to (1) provide wide redshift coverage, averaging out cosmological
parameter degeneracies across different redshifts, (2) yield large-scale information that could reveal signatures of
inflation and gravitational effects, and (3) access matter density modes that are less affected by non-linear physics.
By complementing 21 cm experiments, LIM plays a critical role in addressing cosmological tensions, such as the H0
and S8 discrepancies. Its ability to combine data from multiple spectral lines and explore different epochs of cosmic
evolution allows it to test new models of DE, modifications to gravity, and other potential extensions to the standard
cosmological model.
The further development of this technique will require advancements in millimeter-wave detectors and the refinement
of data analysis methods to mitigate astrophysical and modeling systematics. Below, we identify the key ongoing and
planned experiments leveraging line-intensity mapping to study cosmic large-scale structures and the EoR using this
method.

• The Hobby-Eberly Telescope Dark Energy Experiment (HETDEX): HETDEX [4250] is a ground-
breaking survey leveraging line-intensity mapping of Lyman-α emitters to study the large-scale structure of the
Universe. By mapping the intensity distribution of Lyman-α emission over the redshift range 1.9 < z < 3.5,
HETDEX provides insights into the matter power spectrum and the evolution of cosmic structures during a
pivotal epoch in the Universe’s history. HETDEX’s line-intensity mapping measurements play a crucial role
in addressing cosmological tensions by offering independent constraints on large-scale structure and its growth,
as well as the expansion history of the Universe. Additionally, HETDEX’s use of intensity mapping reduces
reliance on resolving individual sources, enabling efficient surveys of faint, distant cosmic structures.
• CO Mapping Array Project (COMAP): COMAP [4212] is a pioneering experiment designed to map
the large-scale structure of the Universe through intensity mapping of the CO rotational transition lines. By
targeting the z ∼ 2.4 − 3.4 redshift range, COMAP traces the distribution of molecular gas in galaxies during
the peak of star formation activity, offering insights into galaxy evolution and the assembly of cosmic structures.
COMAP plays a crucial role in addressing cosmological tensions by providing independent constraints on the
large-scale matter distribution and the expansion history.
• CarbON CII line in post-rEionization and ReionizaTiOn epoch (CONCERTO): CONCERTO [4251]
is a state-of-the-art experiment designed to study the large-scale structure of the Universe using line-intensity
mapping of the [CII] emission line. Operating in the redshift range 4.5 < z < 8.5, CONCERTO aims to trace
the distribution of ionized carbon during the EoR and the early stages of galaxy formation. By capturing the
integrated [CII] signal from unresolved sources, it provides a unique probe of the ISM and star formation in high-
redshift galaxies. CONCERTO’s sensitivity to high-redshift structures complements other intensity mapping
experiments and traditional surveys, enabling a more detailed understanding of reionization-era physics and
its connection to late-time cosmological parameters. Furthermore, by probing large-scale modes, CONCERTO
helps refine models of DE, inflation, and modifications to gravity, providing new insights into the underlying
causes of cosmological tensions.
• Spectro-Photometer for the History of the Universe, Epoch of Reionization, and Ices Explorer
(SPHEREx): This proposed all-sky survey aims to provide spatial distribution data using various tracers to
explore different cosmic eras [4211]. The survey will utilize the Hα, Hβ, and [OIII] lines to probe 0.1 < z < 5,
0.5 < z < 2, and 0.5 < z < 3, respectively, mapping low-redshift gas clouds. The higher redshift range, spanning
5.2 < z < 8, will be investigated using the Lyman-α line, providing a direct probe of the EoR. By spanning a
wide range of redshifts and utilizing multiple spectral lines, SPHEREx can help address cosmological tensions,
such as the H0 and S8 discrepancies, by providing independent measurements of large-scale structure growth
and the expansion history of the Universe.
• Tomographic Ionized-Carbon Mapping Experiment (TIME): TIME [4252] is designed to probe the
EoR by mapping the redshifted 157.7 µm emission line of singly ionized carbon ([CII]) over the redshift range
5 ≲ z ≲ 9. In addition, it observes carbon monoxide (CO) rotational transitions from galaxies at intermediate
235

redshifts 0.5 ≲ z ≲ 2. These measurements aim to trace the early stages of galaxy formation and star formation
history, providing a complementary view of large-scale structure during key epochs in cosmic evolution.
• The Cosmic Dawn Intensity Mapper (CDIM): CDIM [4253] is a proposed space-based mission designed
to explore the large-scale structure of the Universe through line-intensity mapping of multiple spectral lines.
By targeting key tracers such as hydrogen Lyman-α, Hα, Hβ, and [OIII], CDIM will probe a wide redshift
range, spanning from the EoR (6 < z < 10) to the era of peak star formation (z ∼ 2). These measurements
will provide a comprehensive view of the distribution of ionized and molecular gas, as well as star formation
processes. CDIM’s ability to map the matter distribution across a wide range of redshifts and tracers reduces
parameter degeneracies and systematic uncertainties. By bridging the gap between early- and late-time probes,
CDIM provides complementary data to galaxy surveys and CMB experiments. Additionally, its sensitivity to
faint, diffuse emission makes it a powerful tool for studying large-scale modes that are less accessible with
traditional surveys.
• Fred Young Submillimeter Telescope (CCAT-prime): CCAT-prime [4254], equipped with the advanced
Prime-Cam instrument, is designed to perform line-intensity mapping of emission lines such as [CII] at redshifts
relevant to the EoR and cosmic dawn. By tracing the [CII] line, along with other potential tracers, CCAT-prime
aims to explore the star formation history, the ISM, and the role of galaxies in reionizing the early Universe. Its
high-altitude site and state-of-the-art instrumentation provide exceptional sensitivity for mapping large-scale
structure during key phases of cosmic evolution.
5.1.10. Potential new probes
As our understanding of the Universe deepens, new observational techniques and theoretical frameworks are being
developed to challenge existing paradigms and explore unresolved tensions in cosmology. These potential new probes
leverage advanced technologies and novel methodologies to address persistent discrepancies in key cosmological pa-
rameters, such as the Hubble constant H0 , the amplitude of matter fluctuations S8 , etc. By pushing the boundaries
of precision measurements and theoretical modeling, these probes aim to uncover new physics or systematic effects
that may underlie these tensions.
This section explores these innovative probes, highlighting their potential contributions to resolving the current
challenges in cosmology and paving the way for new discoveries about the Universe.
• Visible and Infrared Survey Telescope for Astronomy (VISTA): VISTA [4255] is a ground-based facility
dedicated to wide-field surveys in the visible and near-infrared wavelengths. Equipped with an advanced infrared
camera, VISTA is uniquely suited to mapping the large-scale structure of the Universe by detecting galaxies and
RGBs at high redshifts. Its unprecedented sensitivity and wide-field capabilities allow it to probe the cosmic web,
trace the distribution of matter, and explore the evolution of structures over cosmic time. By providing precise
measurements of galaxy clustering and WL signals, VISTA contributes to addressing key cosmological tensions,
such as those related to the amplitude of matter fluctuations S8 and the Hubble constant H0 . Furthermore, its
ability to survey large areas of the sky with high resolution offers unique opportunities to identify deviations
from standard cosmological models, potentially unveiling signatures of new physics.
• Transient High-Energy Sky and Early Universe Surveyor (THESEUS): THESEUS [654] is a proposed
space-based mission designed to explore the high-energy transient sky and study the early Universe. By detecting
GRBs from the first generation of stars (Pop III stars) and other energetic phenomena, THESEUS provides a
unique probe into the era of cosmic reionization and early star formation. THESEUS combines wide-field
monitoring and precise localization of GRBs with follow-up spectroscopy, enabling measurements of the cosmic
star formation rate and the chemical enrichment history of the Universe. Its ability to observe the high-energy
sky at unprecedented sensitivity offers a complementary approach to resolving cosmological tensions, such as
refining constraints on the matter density parameter (Ωm,0 ) and testing models of DE. By bridging the gap
between high-energy astrophysics and cosmology, THESEUS represents a transformative tool for uncovering
new physics and deepening our understanding of the early Universe.
• Quasars as Geometric Rulers: RGBs, the luminous cores of AGN, offer a novel approach to cosmological
measurements by serving as geometric rulers. Their variability over time, combined with precise measurements of
time delays in gravitationally lensed RGBs, enables the determination of absolute distances in the Universe. This
method bypasses many of the traditional assumptions in standard candles or rulers, providing an independent
means of measuring H0 and testing the expansion history of the Universe. In addition to their role in time-delay
cosmography, RGBs also probe the large-scale structure of the Universe through clustering and absorption-line
studies. This dual utility allows RGBs to contribute to resolving tensions in cosmological parameters, such as
H0 and S8 . Their utility as geometric rulers and tracers of structure represents a powerful avenue for exploring
deviations from the standard cosmological model.
236

• Redshift Drift with ANDES (ArmazoNes High Dispersion Echelle Spectrograph): ANDES, a
planned instrument for the ELT, offers a unique opportunity to measure the redshift drift—a direct, model-
independent observation of the Universe’s expansion history. By observing spectral lines of distant RGBs and
tracking their subtle changes over decades, ANDES aims to detect the tiny redshift drift predicted by the stan-
dard cosmological model. This groundbreaking method provides a direct probe of the dynamics of the Universe,
bypassing the need for distance ladder calibrations or assumptions about standard candles or rulers. By directly
measuring the acceleration or deceleration of cosmic expansion, ANDES will place new constraints on DE, mod-
ifications to gravity, and the fundamental properties of the Universe. Its unparalleled precision makes it a new
tool for addressing unresolved tensions in cosmological parameters.
• CMB Spectral Distortions: Deviations from a perfect blackbody spectrum of the CMB encode information
about energy injections in the early Universe. These distortions arise due to processes such as Silk damping of
primordial perturbations, decaying or annihilating particles, and early structure formation. The two primary
types of distortions, µ-distortions (sensitive to energy injections at z ≳ 105 ) and y-distortions (from later energy
injections), provide a complementary approach to testing the standard ΛCDM model. Measurements of spectral
distortions could reveal signatures of new physics, including primordial non-Gaussianity, modifications to the
recombination history, or interactions between DM and radiation. Future space missions such as PIXIE and
PRISM aim to achieve the required sensitivity to detect these distortions, opening new avenues for cosmological
exploration.
• Lunar Astronomy: Lunar astronomy [4256], leveraging the unique environment of the Moon, presents a
groundbreaking opportunity to probe the cosmos in ways that are challenging or impossible from Earth. The
Moon’s lack of atmosphere eliminates atmospheric distortions, enabling precise observations across a wide range
of wavelengths, including radio, infrared, and X-rays. Furthermore, the Moon’s far side provides an unparalleled
radio-quiet environment, ideal for low-frequency observations of the early Universe, including the cosmic dawn
and the EoR. Lunar observatories have the potential to address cosmological tensions by providing high-precision
measurements of large-scale structure, the CMB, and the 21 cm signal from neutral hydrogen. Additionally,
their stable environment offers the possibility of ultra-long-term observations, enabling direct measurements
of cosmological parameters such as the Hubble constant and constraints on the nature of DE. By extending
observational capabilities beyond Earth-based and space-based platforms, lunar astronomy could transform our
understanding of the Universe.

5.2. Recommendations
The open question of cosmic tensions presents a unique opportunity to reassess how observational measurements
in cosmology are reconciled with physical models through diverse data analysis approaches. This may signal a turning
point in our understanding of how to interpret measurements, particularly in the presence of potential systematics at
a fundamental level. Such a systematic effect would need to manifest across all phenomena in local distance ladder
measurements or be ubiquitous in early-time measurements. Another possibility is the emergence of new, independent
measurements that do not rely on either the distance ladder or early Universe constraints. In particular, advances in
GW astronomy, where standard sirens provide a novel approach to constructing distance scales, hold great promise.
While this method remains in its early stages and has yet to reach the precision required to compete with traditional
approaches, its continued development—particularly with upcoming detectors such as LISA and third-generation
ground-based observatories—could play a crucial role in resolving cosmic tensions in the future.
A key tool in analyzing observational data and potential systematics, alongside the exploration of fundamental
physics theories, is the application of data analysis methods. Traditionally, data analysis has relied on Bayesian
statistical techniques, both for interpreting observational data and for comparing physical models. However, there
is growing interest in developing frequentist approaches more broadly, given their reduced dependence on prior in-
formation. At the same time, there is strong motivation to advance statistical methods that leverage the broad
range of ML techniques. These include supervised learning approaches such as GP, as well as unsupervised methods
like neural networks. The development of toolkits based on these statistical frameworks holds significant potential,
both in enhancing the extraction of information from observational data and in refining model-building strategies.
In particular, such tools could provide valuable insights into fundamental physics by systematically addressing the
open issues surrounding cosmic tensions. It will be essential to integrate these methodologies into standard analysis
pipelines for upcoming large-scale surveys.
A plethora of physically motivated fundamental physics theories have been studied in recent years, with particular
attention given to their connection to resolving the open problem of cosmic tensions and associated questions within
the standard model of cosmology. These efforts have focused on further developing geometric gravity, exploring
potential physics beyond the standard model of particle physics, and investigating emergent phenomenology and
237

possible quantum gravity effects. On the other hand, many practical theories in these approaches were originally
motivated by theoretical or semi-physical considerations. The further integration of the growing body of observational
evidence into the development of new classes of physical theories will drive the next generation of cosmological models,
aligning them with the most pressing questions from observational surveys. Additionally, the use and advancement of
model-building toolkits will contribute to the creation of physically motivated theoretical models that address cosmic
tensions while preserving the successes of the current standard model of cosmology.
The breadth of observational surveys has been reviewed in Sec. 2, where their potential systematics are discussed
in detail. In Sec. 3, recent developments in traditional data analysis approaches are examined alongside discussions of
emerging statistical tools that are expected to become competitive in the coming years, including those based on ML
methods. The development of new physically motivated fundamental theories is explored in Sec. 4, with a focus on
their potential to resolve cosmological tensions while preserving the successful elements of the current standard model
of cosmology. The future of the field will depend not only on the advancement of each of these areas individually
but also on the ability of the community to integrate these diverse approaches into a coherent framework to address
the growing challenges posed by cosmic tensions. This effort will require close collaboration between observational
cosmologists, data analysts, and fundamental physicists. In the coming decade, it will be essential to:
• Leverage next-generation observational surveys to improve the precision of key cosmological parameters and
systematically test for hidden systematics.
• Incorporate advanced statistical techniques, including ML-based inference methods and profile likelihood ap-
proaches, into standard cosmological pipelines.
• Continue developing theoretical models that remain consistent with observational constraints while addressing
the root causes of cosmological tensions.
• Encourage interdisciplinary collaboration between observational, computational, and theoretical communities
to ensure that different approaches are synthesized into a cohesive framework.
Resolving the cosmic tensions represents a defining challenge for the cosmology community. Success in this endeavor
will require a concerted effort across multiple fronts, balancing innovative new techniques with a rigorous assessment
of existing methodologies. By systematically integrating new data, analysis techniques, and theoretical advance-
ments, the next decade of cosmological research has the potential to transform our understanding of the Universe’s
fundamental properties.

Challenges in observational cosmology:


• Cepheid variables – The next generation of ground- and space-based optical surveys will provide deep time-
series observations of distant galaxies, expanding the volume over which the universality of the Period-Luminosity
relation can be tested. These observations will also improve our understanding of the impact of metallicity
in different environments. As new Cepheids are discovered and observational precision increases, reducing
uncertainties in Cepheid measurements will require better constraints on dust reddening, zero-point corrections
in Cepheid parallaxes, and new geometric anchors in the distance ladder.
• Masers – Systematics in megamaser-hosting galaxies, such as non-circular orbits and non-gravitational forces,
do not significantly impact constraints on cosmological parameters. However, the primary challenge for maser-
based techniques lies in discovering new high-quality megamaser-hosting galaxies suitable for precise distance
measurements.
• Tip of the Red Giant Branch (TRGB) – This method relies on a statistical approach to standardizing
TRGB stars, making the calibration process a critical aspect of the analysis. The main challenges in applying the
TRGB method include (1) potential imbalances between TRGB and calibrating galaxy brightness intensities,
(2) the need for accurate calibration of the target field, and (3) effective outlier detection methods for poorly
sampled LFs.
• Mira variables – Oxygen-rich Mira stars can serve as anchors in the first rung of the distance ladder for
calibrating absolute magnitudes. However, their luminosity, color, large angular size, and surface variations
introduce significant uncertainties in parallax measurements. On the second rung, the number of known Mira
stars remains low, presenting an opportunity for future observations to expand their use in distance calibration.
• Type Ia supernovae – SNIa primarily appear in the second and third rungs of the three-rung distance ladder.
Their systematics are among the most extensively studied and best mitigated. However, the effects of (1)
calibration systematics, (2) redshift-dependent biases in the third rung, and (3) stellar physics and sample
238

selection may each contribute to potential changes in the estimates of certain cosmological parameters. Another
challenge in this sector is the potential expansion of the distance ladder framework to include higher or lower
rungs.
• J-regions of the asymptotic giant branch (JAGB) – This is a relatively new but promising method
for extending the distance ladder. The primary challenges for establishing this approach as a robust distance
indicator include a better understanding of the effects of metallicity, atmospheric influences on photometry, and
variations in pulsation across different stellar populations.
• Type II supernovae – SNII offer an independent path to circumvent SNIa systematics while introducing
different systematic considerations. This method has the potential to achieve greater precision with the addition
of further calibrators and a better understanding of how spectroscopic data can be leveraged to improve the
standardization procedure.
• Surface brightness fluctuations (SBF) methods – SBF-based methods have shown promise in constraining
distances to very distant galaxies, relying primarily on imaging data. The next generation of optical telescopes
will provide opportunities to refine these techniques further. Given the statistical nature of the approach, a
deeper understanding of stellar population studies and peculiar evolutionary stages will help reduce system-
atic uncertainties. Addressing these factors will enable SBF to continue improving as a robust method for
constraining cosmological parameters.
• Cosmic chronometers – This method has the advantage of being theoretically economical in terms of its
underlying assumptions while providing a direct numerical measurement of the Hubble parameter in redshift
space. Beyond the need to increase the number of measurements using this approach, some systematic challenges
include (1) uncertainties in stellar metallicity estimates, (2) errors in star formation history modeling and
underlying assumptions, and (3) contamination by young or star-forming outliers.
• HII galaxies – These galaxies are characterized by bursts of massive star formation, and their strong emission
in specific spectral lines makes them valuable probes for high-redshift cosmology. However, for this method
to become more robust, several systematics need to be addressed, including (1) improved physical modeling of
emission line width profiles, (2) better understanding of stellar age effects, (3) corrections for extinction-induced
deformations in emission lines, and (4) refined metallicity distribution calibrations.
• Baryonic Tully-Fisher relation methods – The Tully-Fisher and BTFR can be applied to cosmology to
constrain the Hubble constant, given an underlying cosmological model. While this method offers good control
over systematics, the relatively small number of observed galactic objects currently limits its competitiveness
compared to other techniques. This limitation will be addressed with upcoming large-scale surveys. On the
modeling side, further development may help reduce model dependence, for instance, by incorporating ML
techniques into the analysis.
• Strong lensing and time-delay cosmography – Strong lensing provides a distance-ladder-independent
method for measuring H0 through time-delay cosmography. However, several systematics currently limit its
precision, including uncertainties in lens mass modeling, the impact of line-of-sight structures, and degeneracies
in the inferred time delays. Additionally, assumptions about mass profile parametrizations can introduce biases
in cosmological constraints. Improving this technique will require higher-resolution imaging, better character-
izations of lens environments, and larger statistical samples of well-measured lensing systems to reduce both
statistical and systematic uncertainties.
• Gravitational wave astronomy – Free from electromagnetic radiation, GW cosmology provides uniquely
independent methods to constrain cosmological parameters, separate from all other approaches. Standard
sirens offer the additional advantage of serving as absolute distance indicators, making them independent of a
cosmological model in determining the Hubble constant. However, an electromagnetic counterpart is required to
identify the host galaxy, which is then used to infer certain galactic properties; alternatively, statistical methods
utilizing galaxy catalogs can be employed. On the other hand, dark sirens suffer from potential systematics
related to assumptions about galactic properties.
• Cosmic microwave background (CMB) radiation – Like other observatories, CMB experiments are subject
to instrumentation systematics and internal tensions, which can be more pronounced given the precision of the
measurements and the vast volume of data collected. Ensuring consistency between high- and low-multipole
analyses in cosmological parameter estimation remains a critical challenge for future CMB surveys. The next
generation of CMB experiments will aim to minimize instrumental and astrophysical systematics, improving
239

constraints on fundamental physics. By reducing these contamination effects, it may even become possible to
robustly measure B-mode polarization, which remains a crucial target for detecting primordial GWs and probing
inflationary physics.
• Baryon acoustic oscillations (BAO) – This method enables robust cosmological analyses based on millions of
extragalactic objects. However, its statistical nature introduces a number of systematics that must be carefully
accounted for. One key challenge is that the source density field is displaced due to peculiar motions, particularly
in low-redshift regions, which is corrected using density field reconstruction techniques. As more galaxies are
observed with higher precision, these reconstruction methods must continue to improve. Another important
aspect of BAO surveys is the reliance on an underlying fiducial cosmology to transform between coordinate
systems. While the influence of this model is relatively small, developing techniques to further reduce this
dependence remains a priority. Additionally, BAO measurements are taken relative to the sound horizon,
introducing another model-dependent feature that should be further refined to minimize its impact.
• Galaxy clustering – The growth of large-scale structures can be inferred from the statistical clustering of
galaxies in redshift space, which depends on estimates of correlation functions and power spectra. The statistical
nature of these measurements introduces several systematics, including those originating from measurement
techniques (such as atmospheric and extragalactic extinction corrections), galaxy survey systematics (such as
calibration errors and foreground contamination), and uncertainties in nonlinear modeling for redshift-space
galaxy clustering.
• Large-scale structure and higher-order statistics – Beyond two-point statistics such as the galaxy corre-
lation function and power spectrum, higher-order statistics and cosmic web topology provide additional tools
to probe cosmology. Measurements of cosmic voids, filament structures, and the bispectrum offer sensitivity
to non-Gaussianities and potential modifications to gravity. The main challenges in this sector include model-
ing nonlinear structure formation, correcting for observational systematics, and integrating these statistics into
existing cosmological frameworks.
• Weak lensing – Weak gravitational lensing surveys rely on statistically significant samples of galaxies, in-
troducing several challenges that must be addressed. These include uncertainties in individual galaxy redshift
estimates due to the inherent limitations of photometric redshift determinations, systematic biases in shear and
shape ellipticity measurements, and difficulties in accurately modeling nonlinearities in the matter power spec-
trum. Improving these aspects will be essential for maximizing the precision of WL constraints on cosmological
parameters.
• Galaxy cluster counts – Galaxy clusters are the largest gravitationally bound structures in the Universe and
are highly sensitive to cosmic evolution and the underlying cosmological parameters. Aside from the challenge
of distinguishing how different cosmological models affect cluster evolution, a major issue in this method is
the mass calibration problem. This refers to the relationship between the total halo mass and observable
cluster properties, which is crucial for using cluster counts as a cosmological probe. The primary method for
determining halo masses is WL, which itself introduces additional systematic uncertainties. As survey data
become more precise, the mass calibration problem will increasingly become a limiting factor in the effectiveness
of this approach.
• Quasar and GRB distance ladder – Both RGBs and GRBs have shown strong potential as probes of the
high-redshift Universe, providing constraints across multiple cosmological parameters. However, their precision
remains limited compared to other methods, partly due to systematics related to the physical nature of the
sources, which remain an active area of discussion in the literature. Resolving these open questions would
significantly enhance the statistical power of cosmological parameter constraints when these data are used in
combination with other survey observations.
• Fast radio bursts (FRBs) – The relatively recent discovery of FRBs provides a powerful new probe of cosmo-
logical models. These events have shown promising potential to independently constrain the Hubble constant,
as well as other cosmological parameters, particularly offering insights into the matter density parameter within
ΛCDM. However, systematics related to the modeling of DM contributions and the IGM remain crucial open
questions for this technique. Refinements in these areas will be necessary to improve its precision.
• Radio background excess – The discrepancy between the expected emission from known astrophysical and
cosmological sources and the observed brightness in the electromagnetic spectrum presents an anomaly in the
radio segment of the spectrum. Resolving this requires a careful balance in accounting for infrared radiation
production in early galaxies, X-ray contributions affecting inverse Compton scattering of CMB photons, the
240

depth of the 21 cm signal, and constraints from the cross-correlated angular power spectrum with optical sources.
The next generation of observatories has the potential to refine and test possible explanations for this excess
signal.

• Bulk flow measurements – The average peculiar velocity of a statistically significant sample of galaxies
should diminish at sufficiently high redshifts. However, measurements of bulk flows face several challenges,
including large uncertainties in peculiar velocity estimates and biases in the spatial distribution of observed
galaxies. Recent studies have suggested that bulk flow magnitudes increase with distance, a result that appears
counterintuitive. Addressing this challenge will require the continued development of bulk flow measurement
techniques, as well as improved observational data quality at higher redshifts.

• Cosmic dipole and anisotropies – The observed CMB dipole is expected to be consistent with the dipole
measured in other astrophysical tracers. However, discrepancies in dipole measurements using different observa-
tional datasets suggest the possibility of systematic effects or new physics. Additionally, potential anisotropies
in cosmological parameters such as H0 and S8 challenge the assumption of statistical isotropy. Addressing
these issues requires improved methodologies and cross-correlations between independent datasets to test for
underlying systematics or signatures of new physics.

• Cross-correlations between cosmological probes – Combining different observational techniques provides


a means to reduce systematics and improve cosmological parameter constraints. Key cross-correlations include
(1) WL and galaxy clustering to calibrate galaxy bias and structure formation models, (2) CMB lensing and
large-scale structure to refine DM distribution models, and (3) standard sirens and BAO measurements to
independently test the cosmic expansion history. As observational precision improves, cross-correlations will
become increasingly important in resolving cosmological tensions.

• 21 cm intensity mapping – The neutral hydrogen 21 cm signal provides a powerful method to trace the
large-scale structure of the Universe across a wide range of redshifts. This technique has the potential to probe
the cosmic dawn, reionization, and the post-reionization epochs, offering insights into the distribution of matter
over cosmic time. However, several challenges must be addressed, including foreground contamination from
astrophysical sources, calibration uncertainties, and the need for precise theoretical modeling of the signal’s
evolution.

• Line-intensity mapping (LIM) – LIM provides a promising method to trace large-scale structure by mea-
suring integrated emission from spectral lines such as CO and Lyman-α rather than relying on resolved galaxy
surveys. This method has the potential to map cosmic structure at high redshifts, but faces challenges such as
foreground contamination and the need for precise theoretical modeling.

• Sunyaev-Zel’dovich (SZ) and X-ray surveys – Measurements of galaxy clusters using the SZ effect and
X-ray emission offer valuable constraints on cosmic structure formation and expansion. However, systematic
uncertainties related to gas physics, temperature profiles, and mass calibration remain key challenges.

• Multi-messenger cosmology – Observations across multiple cosmic messengers, including electromagnetic


signals, GWs, and high-energy particles, provide new opportunities for constraining cosmological parameters and
testing fundamental physics. By combining information from different observational channels, multi-messenger
approaches can help break degeneracies inherent in individual methods. However, challenges remain, including
the low detection rates of some sources, difficulties in precise localization, and uncertainties in the modeling of
extreme astrophysical environments.

• Neutrino Cosmology – Neutrinos remain among the least understood fundamental particles, and their ex-
tremely weak interactions make direct measurements of their properties exceptionally difficult. A key open
question is whether these particles have a nonzero mass. Moreover, terrestrial neutrino oscillation experiments
provide constraints that appear to contrast with indirect cosmological measurements, potentially hinting at
physics beyond the Standard Model in the neutrino sector. Reconciling these discrepancies will require further
advancements in both laboratory-based and cosmological neutrino studies.

Challenges in the development of numerical and data analysis tools:


• Model constraints in conjunction with Bayesian statistics – Bayesian statistical methods have been
widely used to constrain cosmological model parameters against observational data, and they now feature a broad
suite of tools designed to meet these challenges. However, as survey data volumes increase and cosmological
241

models become more complex, the computational demands of traditional Bayesian approaches are becoming
increasingly problematic. This motivates the development of more efficient algorithms and novel implementations
of Bayesian inference techniques to maintain accuracy while improving computational feasibility.
• Frequentist approaches – Profile likelihoods provide a standard statistical method for analyzing cosmological
models with high-dimensional parameter spaces, particularly when the likelihood surfaces exhibit non-Gaussian
features. This approach involves running multiple traditional MCMC analyses for fixed parameters to construct
a profile likelihood. Several community tools already implement this method, some incorporating ML techniques
to enhance performance. However, due to the computationally intensive nature of this approach, further work on
accelerating the sampling process will be crucial for making it more practical for a wider range of applications.
• Hybrid Bayesian-Frequentist approaches – While Bayesian inference dominates cosmological data analysis,
frequentist methods such as profile likelihoods and bootstrapping offer complementary insights, particularly
in scenarios where priors strongly influence results. Hybrid approaches that combine these frameworks can
provide more robust parameter constraints, reduce sensitivity to prior assumptions, and improve statistical
interpretability in high-dimensional cosmological models.
• Fast approximate inference techniques – Traditional Bayesian inference methods such as MCMC can be
computationally prohibitive, especially for high-dimensional cosmological models. Alternative approaches such
as Hamiltonian Monte Carlo, Variational Inference, and Neural Density Estimation offer promising routes to
accelerate parameter estimation while maintaining accuracy. Developing and benchmarking these techniques
for cosmological applications is an open challenge.
• Cosmology simulator inference techniques – The vast amount of information contained in the large-
scale structure of the Universe presents a significant challenge for traditional simulation and analysis methods.
Extracting cosmological parameters from these simulations is particularly difficult, as non-Gaussian and higher-
order effects may become increasingly relevant for improving accuracy. Recent advances, including neural
network architectures and decision tree learning approaches, have shown promise in addressing these challenges.
However, the availability of community tools and a robust comparison of different inference techniques remain
open issues that need further exploration.
• Computational challenges in large-scale simulations – Cosmological simulations are essential for under-
standing structure formation and model testing. However, the increasing size of observational datasets and
complexity of simulations present major computational challenges. Key issues include (1) optimizing numerical
algorithms for high-performance computing architectures, (2) improving resolution and scalability in N-body
and hydrodynamical simulations, and (3) reducing computational costs while maintaining accuracy in statistical
inference from simulations.
• Reconstruction techniques – The use of observational survey data to reconstruct physical models is still in
its early stages; however, powerful techniques have already been developed and made publicly available. The
development of more mature and statistically robust reconstruction methods, including those leveraging ML
approaches, will be an important aspect of progress in this direction. A major challenge for these techniques
lies in effectively utilizing the wide variety of available datasets, which may include background, perturbative,
and other physical aspects of cosmological models.
• Genetic algorithm model selection – GAs provide a mechanism to explore an otherwise infinite model
space within a finite time by applying selection criteria. These methods primarily rely on Bayesian constraint
analyses, but practical toolkits for their implementation are not yet publicly available in the literature. The
development of such toolkits is essential for advancing this technique into a more comprehensive classification
framework. Additionally, combining GAs with other, faster constraint methods may enhance their suitability
for practical model selection analyses.
• Machine learning for theoretical model selection – The increasing complexity of cosmological models
necessitates efficient methods for exploring large parameter spaces. ML techniques, such as deep learning
classifiers and GP regression, are being developed to identify viable theoretical models by learning patterns in
high-dimensional likelihood landscapes. However, challenges remain in ensuring interpretability, robustness, and
generalization across different datasets.
• Calibration and validation of machine learning in cosmology – ML algorithms are increasingly used for
inference, parameter estimation, and anomaly detection in cosmological datasets. However, ensuring that these
models generalize well across different observational conditions, remain unbiased, and do not overfit to specific
242

datasets is a key challenge. Benchmarking ML models against traditional methods and developing interpretable
AI approaches will be crucial for their broader adoption in cosmology.

• Dark siren cosmology – The number of standard sirens without an associated electromagnetic counterpart
continues to grow. These sources can also be used to constrain cosmological parameters, albeit through a statis-
tical reconstruction of their population properties using galaxy catalogs. However, the simultaneous modeling of
cosmology and mass distributions introduces computational challenges that must be addressed to extract robust
constraints.

• Spectroscopical sirens – The use of spectroscopic data from host galaxies associated with GW events offers a
promising avenue for improving the precision of standard siren measurements. By obtaining redshift information
through spectroscopic observations, it is possible to reduce uncertainties in distance measurements and break
degeneracies in cosmological parameter estimation. However, challenges remain in accurately identifying the
host galaxy among multiple candidates, mitigating observational systematics in spectroscopic measurements,
and optimizing survey strategies to maximize the detection of viable spectroscopical sirens.

• Systematics mitigation in statistical inference – The increasing precision of cosmological surveys requires
robust methods to identify and mitigate systematics in data analysis. Key challenges include (1) the impact
of prior choices on Bayesian inference, (2) biases introduced by incomplete or inhomogeneous datasets, and (3)
uncertainties in the modeling of astrophysical and instrumental effects. The development of statistical techniques
that can cross-validate results across different datasets while minimizing these biases remains an important open
problem.

• Big Bang Nucleosynthesis numerical estimates – The precision of BBN predictions for primordial element
abundances has significantly improved in recent years. Advanced numerical codes have revealed tensions in
some species ratios that were not apparent in previous lower-order calculations. Future numerical developments
incorporating these higher-order effects will be crucial for reassessing these tensions and refining BBN constraints
on early Universe physics.

• Cross-correlations and multi-probe statistical methods – Combining multiple cosmological probes is


essential for breaking degeneracies and improving parameter constraints. However, this requires sophisticated
statistical frameworks to account for systematics, survey-specific biases, and correlated uncertainties across
datasets. Developing robust cross-correlation pipelines that integrate WL, large-scale structure, and CMB
measurements is a key challenge for upcoming surveys.

• Data-driven approaches for systematics mitigation – Systematic uncertainties remain a major challenge in
cosmology, particularly in photometric redshift estimation, instrumental biases, and survey calibration. ML, GP,
and hybrid statistical methods can provide real-time corrections for systematics, but require careful validation
to ensure robustness across different datasets.

• High-volume and high-dimensional data processing – The increasing size of cosmological datasets from
next-generation surveys requires the development of scalable numerical methods and high-performance comput-
ing techniques. Challenges include managing high-dimensional parameter spaces, optimizing data compression
without loss of information, and implementing real-time data processing pipelines for massive observational
datasets.

Challenges in fundamental physics:


• Early-time physics – The early Universe provides a compelling regime for introducing exotic physics that
remains shielded by recombination. Mechanisms such as early scalar fields, additional neutrino species, or mod-
ified expansion histories primarily affect the sound horizon at recombination, which in turn impacts inferences
of the Hubble constant. These early modifications also present opportunities for testing inflationary scenarios.
However, a significant challenge remains in explaining the evolution of large-scale structure, ensuring that any
proposed early-time physics can simultaneously produce the correct growth rate to remain consistent with all
cosmological observations.

• Inflation – Phenomenological models successfully provide a spectrum of scenarios that can alleviate some
tensions even under the assumption of Gaussian fields. However, it remains crucial to construct physically
realistic explicit models that respect fundamental physical principles such as locality, causality, and unitarity,
while also approximating key features of these phenomenological models. As future galaxy surveys improve
243

precision on possible non-Gaussian features associated with primordial conditions, it will become increasingly
important to develop accurate numerical simulations and theoretical descriptions of the nonlinear aspects of the
evolution of these fields.
• Big Bang Nucleosynthesis (BBN) tensions – BBN predictions generally exhibit strong agreement with
ΛCDM in the production rates of primordial species. However, discrepancies remain, particularly for lithium
isotopes, and increasingly for helium and deuterium, when compared with their relative abundances inferred
from observations of the oldest stars. These discrepancies may stem from stellar astrophysics rather than
cosmological considerations, but they could also signal challenges for early Universe dynamics in non-standard
cosmological models. Resolving these tensions requires improved modeling of stellar evolution, nuclear reaction
rates, and alternative BBN scenarios.
• Primordial magnetic fields – The possible existence of non-negligible PMFs could have influenced CMB
anisotropies and small-scale inhomogeneities in the large-scale structure of the Universe. Various realizations
of these fields have shown promise in addressing aspects of cosmic tensions, particularly in modifying early-
Universe physics. However, more precise small-scale measurements of CMB anisotropies are needed to robustly
test these scenarios and establish whether PMFs play a significant cosmological role.
• Quantum decoherence in the early Universe – The classical nature of primordial perturbations is assumed
in standard cosmology, yet their quantum mechanical origin requires a consistent explanation of decoherence
processes. Understanding how quantum fluctuations transitioned to classical density perturbations and testing
for potential residual quantum signatures remains an unresolved issue in fundamental physics.
• Parity violations in cosmology – Parity-violating interactions, such as those arising in extensions to general
relativity or axion-like particles coupled to gravity, could leave unique imprints on the CMB polarization or large-
scale structure. The detection of a nonzero cosmic birefringence signal, or deviations in E/B-mode power spectra,
could offer evidence for such interactions. However, current observational constraints remain inconclusive, and
refining the observational tests for parity violation remains an open challenge.
• Nonflat spatial curvature – The cosmological curvature parameter is a fundamental aspect of the FLRW
geometric background and is predicted to be zero in most inflationary models. However, any possible nonzero
value remains highly degenerate with other cosmological parameters in standard analyses. Methods to break
this degeneracy, such as BAO measurements or other observational techniques, could provide important insights
into potential departures from the standard model and their role in addressing cosmological tensions.
• Testing the Cosmological Principle – The standard model of cosmology assumes large-scale homogeneity and
isotropy. However, several observations, including cosmic dipoles, bulk flows, and peculiar velocity anomalies,
suggest that these assumptions may need to be tested more rigorously. Developing more precise observational
tests and statistical methodologies to assess the validity of the cosmological principle remains an important
challenge in the field.
• Nonvanishing cosmic dipoles – The cosmic dipole represents the largest expected deviation from isotropy
if the cosmological principle is broken at any level. Potential dipole signals have been investigated in various
datasets, including the CMB, RGB distributions, fine-structure constant measurements, SNIa, radio galax-
ies, GRBs, and bulk flow studies. Recent observations have hinted at slight nonvanishing dipoles in some of
these probes. To assess the validity and implications of such findings, it is necessary to develop more robust
cosmological backgrounds that can accommodate or rule out these potential departures from isotropy.
• CMB anisotropy anomalies – Several features in the CMB anisotropic power spectrum exhibit mild tensions
with the predictions of the ΛCDM model. Early Universe modifications that explain or incorporate these
features more naturally may offer new insights into fundamental physics beyond standard cosmology. These
anomalies include: (1) a stronger-than-expected ISW effect in certain analyses, (2) non-Gaussian features in
the CMB, such as the Cold Spot, (3) a slight hemispherical asymmetry in the CMB power spectrum, (4) an
unexpected lack of large-scale CMB temperature correlations, (5) a higher-than-predicted lensing parameter,
and (6) a moderate inconsistency between parameter values derived from high- and low-multipole regions of the
CMB power spectrum. Understanding whether these features stem from new physics or residual systematics
remains an important challenge for upcoming surveys.
• Anomalous Integrated Sachs-Wolfe (ISW) effect – The ISW effect arises from the interaction of CMB
photons with evolving gravitational potentials associated with large-scale structure. Observations suggest a
discrepancy between the predicted ISW signal in ΛCDM and its measured imprint in cosmic superstructures
244

and supervoids, both in amplitude and sign. This may indicate a more complex growth history than expected
in the concordance model. Understanding whether this excess signal can be reconciled within ΛCDM, or if it
points to new physics, remains an open question.
• The nature of dark matter – The connection between the fundamental properties of DM and cosmic evolution
has significant implications both for the early Universe, through thermal and non-thermal relic physics, and for
the late Universe, influencing the formation of large-scale structure. Modeling how the thermal history and
other fundamental characteristics of DM may lead to deviations from standard cosmology via solutions of the
Boltzmann equation will be crucial in assessing whether DM plays a role in alleviating cosmological tensions.
• Axion-like particles and ultralight fields – The existence of ultralight scalar fields, such as axion-like
particles, could impact the evolution of the Universe through novel interactions with DM, DE, or radiation.
These fields have been proposed to address cosmic tensions by altering the expansion rate or structure growth.
Developing observational tests for their presence and impact remains a crucial task.
• Complex dark sector interactions – Many extensions to ΛCDM propose that DM and DE may not be
independent, but instead part of a larger dark sector with self-interactions or couplings to other fields. These
models introduce new forces in the dark sector that could impact structure formation, cosmic expansion, and
small-scale clustering. However, constraining these interactions remains difficult, as current probes (e.g., CMB,
BAO, WL) primarily test gravitational effects rather than direct couplings.
• Modified gravity – Extensions beyond ΛCDM may involve modifications to gravity, ranging from mild to
radical departures from General Relativity. The literature is rich with competing formulations of MG theories,
but many are disfavored at smaller scales and exhibit strong degeneracies both among themselves and with scalar-
tensor theories. Additionally, MG theories often introduce challenges such as unwanted fifth-force effects. On
the other hand, these theories provide an avenue for constructing alternative cosmological models independent
of ΛCDM, which could yield departures from concordance cosmology in both early- and late-time evolution.
• Late-time physics – Measurements in the local Universe place strict constraints on the cosmic expansion
profile, leaving little room for significant deviations from ΛCDM. However, recent results from large-scale surveys
suggest possible hints of dynamical properties in DE at late-times. While numerous competing theories focus
on late-time modifications, many of their observational properties remain highly degenerate. To address this
issue, the development of more distinct and testable theoretical models, as well as more precise predictions from
specific scenarios, would be beneficial.
• Exotic dark energy models – The majority of DE models assume a scalar-field-driven component, but
alternative formulations involving vector or tensor fields have been proposed. These include vector-tensor
theories, emergent gravity models, and Lorentz-violating DE frameworks. Understanding the observational
signatures and constraints on these exotic models is an open challenge.
• Variation of fundamental constants – The possible variation of fundamental constants could arise from
various exotic physics scenarios, including non-Standard Model particle interactions, MG, or coupled scalar
fields, among others. Temporal or spatial variations in fundamental constants would have profound consequences
for fundamental physics, potentially impacting atomic structure, nuclear reactions, and cosmological evolution.
Observational efforts have primarily focused on detecting variations in the fine-structure constant and the
electron mass, as these have shown promise in providing an alternative explanation for the Hubble tension.
However, current CMB observations lack the precision to distinguish between models that allow for varying
fundamental constants. Addressing this challenge will require next-generation observatories with significantly
improved sensitivity to such effects.
• Higher-dimensional cosmological models – Theories incorporating extra spatial dimensions, such as brane-
world scenarios and string-inspired cosmologies, propose modifications to gravity that could influence cosmic
expansion, the dark sector, or structure formation. Developing testable predictions for these theories and
determining their viability in explaining cosmic tensions remains a key theoretical challenge.
• Cosmological quantum gravity – It is expected that classical gravity breaks down at high energy scales, but
the impact of UV-complete quantum gravity theories on cosmological observables remains unclear. In particular,
it is uncertain whether such effects could resolve the Hubble tension or other cosmic anomalies. Possible quantum
gravitational effects, such as the GUP, Lorentz invariance violations, or nonlinear modifications, could provide a
new perspective on cosmic tensions as manifestations of non-classical physics. However, fully developed quantum
gravity models that remain consistent with small-scale physics while simultaneously addressing these challenges
are still in early stages of development.
245

• Testing the equivalence principle on cosmological scales – The equivalence principle, a cornerstone
of general relativity, has been tested at small scales with high precision. However, cosmological tests remain
limited. Anomalies in the behavior of standard sirens, WL, or cosmic dipoles could indicate potential violations.
Future large-scale surveys will play a crucial role in refining these tests.

• Relic neutrinos and direct detection – The cosmic neutrino background (CνB) is a fundamental predic-
tion of Big Bang cosmology, yet it has never been directly detected. Cosmological observations place strong
constraints on its properties, but upcoming laboratory experiments and indirect probes may provide the first
evidence for relic neutrinos. If detected, these neutrinos could offer crucial insights into early-Universe physics,
including possible deviations from the Standard Model.

By confronting these challenges, the cosmology community will be poised to revolutionize our understanding of
the Universe, ensuring that future cosmological models remain consistent with increasingly precise observations while
harnessing the full power of advanced statistical and computational methodologies.

5.3. Conclusions
The issue of cosmic tensions has emerged over the past decade as a significant challenge to the ΛCDM concordance
model, which also faces underlying consistency issues and additional anomalies. While this is not unexpected for
a phenomenological framework, the mounting observational discrepancies suggest that a new concordance model
may ultimately be required to resolve these tensions. Despite the long-standing success of ΛCDM, the breadth
and persistence of cosmic tensions represent the most serious and potentially insurmountable test for the standard
cosmological paradigm.
Among these challenges, the continued statistical significance of the Hubble constant tension has solidified its status
as a defining question in cosmology. This tension not only tests the robustness of the concordance model but also
probes the ability of the cosmology community to reconcile local (late-time) measurements with global (early-time)
constraints. The latter, being model-dependent, introduces additional complexities, while the former is subject to
astrophysical systematics that may exert only a limited influence on the magnitude of the discrepancy. However, the
widespread and consistent detection of the Hubble tension across independent measurement techniques casts doubt
on the plausibility of a single cross-survey systematic effect that could fully account for the discrepancy. In Sec. 2.1,
we review the diverse range of methods used to infer H0 . Many of these techniques rely on the distance ladder to
measure the Hubble flow, including refined SNIa analyses, Cepheid variable calibrations, TRGB and JAGB methods,
SNII measurements, Mira variables, and the SBF technique. Additionally, CCs provide a direct means of measuring
the Hubble parameter to comparable redshifts. Collectively, these approaches yield a consistently high value of H0 ,
a result reinforced by distance-ladder-independent methods such as maser-driven constraints and estimates based
on the BTFR. Furthermore, higher-redshift techniques—including HII galaxy distance indicators, time-delay strong
lensing, FRBs, RGBs and GRBs Hubble diagrams—further substantiate the discrepancy. Another promising avenue
for constraining H0 lies in GW-based measurements using standard sirens, which provide an independent approach free
from electromagnetic systematics. This method continues to improve and has already delivered encouraging results,
potentially adding new insights into the nature and expression of the Hubble tension. On the other hand, every
CMB-based survey—including Planck, ACT, and SPT—consistently yields a significantly lower value of the Hubble
constant, assuming a ΛCDM cosmological framework. This lower range of H0 values is further reinforced through the
combination of baryon acoustic oscillation data and constraints from BBN. The core of the Hubble tension thus lies
in the stark contrast between low-redshift measurements, which suggest a higher H0 , and early-Universe estimates
derived from the CMB and similarly high-redshift probes, which favor a lower value.
The Hubble tension serves as a critical test of ΛCDM—or any alternative cosmological model—spanning from
the primordial Universe to the present day. A closely related parameter that traces the expansion and structure
formation history of the Universe is S8 , which encapsulates both the total matter density and the amplitude of
matter fluctuations on scales of 8 h−1 Mpc. This parameter is widely regarded as an effective measure of cosmic
structure growth and is examined in Sec. 2.2. Similar to the Hubble tension, a statistical discrepancy emerges in
the observed values of S8 when comparing CMB-based primary anisotropy measurements with local probes such as
WL, galaxy clustering, and galaxy cluster abundance studies. The S8 parameter is also closely connected to the f σ8
parameter, which is constrained through RSD, as well as other structure growth parameters, including the growth
index, growth parameter, and growth rate. CMB-based measurements yield a relatively high and stable estimate of
S8 across multiple experiments, including different data subsets from the Planck mission’s legacy release. In contrast,
independent analyses from WL surveys, galaxy clustering, RSD, and galaxy count-based approaches predominantly
indicate a lower S8 value at a statistically significant level. When considered alongside the Hubble tension, the
potential S8 tension presents an even greater challenge to ΛCDM and alternative cosmological models, further testing
their predictive power and ability to fully reconcile current observational data.
246

Additionally, several anomalies and challenges arising from the confrontation between ΛCDM and observational
data have become increasingly apparent. These issues, described in Sec. 3, include the Alens anomaly, which suggests
an excess lensing amplitude in the Planck data and a potential deviation from ΛCDM predictions. This is further
compounded by a slight preference for a closed universe in spatial curvature parameter constraints within the ΛCDM
framework. Other CMB anisotropy anomalies include the low quadrupole moment and its alignment with the octupole,
hemispherical asymmetry, and the Cold Spot anomaly. Beyond the CMB, additional unresolved questions concern
the exact nature of cosmological neutrino physics, potential deviations from the cosmological principle due to large-
scale bulk flows, and anomalies in the primordial abundances of certain species in the early Universe. More broadly,
discrepancies have emerged in Lyman-α measurements, the ISW effect, radio synchrotron background observations,
and the possible existence of large cosmic voids.
The vast amount of observational data related to cosmic expansion, large-scale structure formation, and the evo-
lution of fundamental parameters presents a formidable challenge for the data analysis and theoretical modeling
community. Traditional implementations of MCMC methods are increasingly struggling to distinguish between com-
peting cosmological scenarios due to the complexity of their parameter dependencies and the high dimensionality of
their parameter spaces. In Sec. 3, novel approaches to MCMC methods are discussed alongside alternative data anal-
ysis techniques. Frequentist parameter inference through profile likelihoods offers an intriguing pathway to mitigating
prior volume effects that can dominate posterior constraints in Bayesian approaches. However, this method is even
more computationally demanding than standard MCMC, making it impractical for certain classes of models beyond
ΛCDM. Similarly, extracting cosmological information from large-scale simulations has become computationally in-
tractable using conventional methods, necessitating the development of novel tools. ML techniques, particularly deep
generative models, have shown significant promise in constraining cosmological parameters with improved efficiency.
More broadly, ML inference frameworks have begun replacing traditional MCMC techniques, leveraging various neural
network architectures to refine parameter estimation. In parallel, model selection methodologies have incorporated
GAs as a systematic approach for exploring complex model spaces against observational constraints. These rapidly
evolving techniques are proving essential in addressing the challenges posed by next-generation observatories, which
are producing unprecedented volumes of data and introducing a vast array of potential new physics scenarios.
The increasingly robust expression of cosmic tensions—spanning the Hubble constant, large-scale structure growth,
and various other anomalies—necessitates a reevaluation of potential new physics within both current and future
observational surveys. The cosmology community has proposed a wide range of possible directions for addressing
these challenges, which are reviewed in Sec. 4. Early-Universe modifications have the advantage of introducing
exotic dynamics prior to recombination, beyond the reach of direct electromagnetic observations. However, while
altering the sound horizon can help reconcile certain tensions, these modifications may introduce inconsistencies
with the early growth of large-scale structure. Addressing this issue requires exploring different realizations of new
physics in this sector alongside novel data analysis techniques, such as frequentist parameter inference. On the
other hand, late-time modifications to cosmology have also shown promise in addressing cosmic tensions through
diverse mechanisms, including additional fields, modifications to gravitational physics, and other extensions to the
standard paradigm. However, the significantly more detailed and structured nature of late-Universe observations
places stringent constraints on the evolution of such models, restricting the amplitude of any new physics in this
regime. An alternative to modifying gravity in the field equations is to reconsider the behavior of the matter sector,
such as through interacting or decaying DM scenarios, or more intricate physics governing DM properties. Beyond
standard modifications to cosmic evolution, alternative explanations have been explored, including the role of large
cosmic voids, the influence of PMFs, and variations in the dynamics of cosmic inflation. The severity of these
tensions has also revived interest in non-standard cosmological geometries that challenge aspects of the cosmological
principle, though such proposals remain in early stages of development. Similarly, while quantum gravity theories
have not yet demonstrated a direct resolution to cosmic tensions, the potential variation of fundamental constants
has been shown to meaningfully impact both the inferred Hubble constant and the growth of large-scale structures,
with broader implications across multiple areas of fundamental physics. Ultimately, reconciling early- and late-time
cosmological observations, alongside non-standard phenomenology, may necessitate a combination of modifications to
the concordance model. However, it remains crucial to break the strong degeneracy among different classes of modified
cosmologies and ensure that proposed models not only fit existing data but also yield new, testable predictions while
preserving the well-established successes of the concordance framework across both astrophysical and cosmological
scales.
The coming years will see an explosion of high-precision observational data from upcoming surveys, as reviewed
in Sec. 5.1. New CMB experiments, such as LiteBIRD and CMB-S4, have the potential to detect primordial GWs
and refine measurements of CMB anisotropies, which may provide key insights into fundamental physics. In parallel,
large-scale structure surveys—including those conducted by Euclid, the Roman Space Telescope, and the Rubin
Observatory—will probe the evolution of cosmic structure and the nature of DE in unprecedented detail. WL studies,
RSD, and clustering analyses will provide complementary information on cosmic structure formation, while 21 cm
247

intensity mapping from SKAO, FAST, and CHIME will open a new observational window into the cosmic dawn.
Additionally, radio astronomy efforts, such as FRB surveys with UTMOST and MeerKAT, will offer independent
constraints on cosmic expansion and structure evolution. Beyond electromagnetic probes, the next generation of
GW observatories—including the LIGO-Virgo-KAGRA network, the ET, Cosmic Explorer, and the space-based
LISA mission—will provide direct insight into gravitational radiation, potentially offering novel constraints on early-
Universe physics and cosmic expansion. Additionally, advances in standard siren measurements may provide an
independent means of resolving the Hubble tension. Meanwhile, improvements in the cosmic distance ladder, enabled
by JWST, ELT, and other facilities, will refine local measurements of H0 with increasing precision. As the volume of
observational data continues to grow exponentially, computational advancements will play an increasingly critical role
in extracting cosmological information. The limitations of traditional MCMC methods necessitate the development
of alternative approaches, such as frequentist inference through profile likelihoods, deep generative models, and other
ML techniques. These innovations have already shown promise in improving parameter estimation and accelerating
statistical inference, while model selection algorithms, including GAs, offer systematic strategies for exploring vast
cosmological parameter spaces. The integration of artificial intelligence and high-performance computing will be
essential in processing and analyzing the unprecedented data streams expected from next-generation observatories.
The comprehensive review presented in this White Paper, along with the collaborative efforts that shaped it, reflects
a growing consensus within the cosmology community on the key directions that must be pursued in the coming years.
As detailed in Sec. 5.2, these efforts span observational challenges, advancements in data analysis techniques, and
the theoretical development of new physics models. Crucially, resolving cosmic tensions will require a synergistic
approach, combining observational precision, computational innovation, and theoretical creativity. The next decade
of cosmology will be pivotal in shaping our understanding of the Universe, with the potential to refine the current
paradigm or uncover new physics that reshapes our fundamental model of cosmic evolution.
248

6. Conventions
The White Paper spans research communities in observational, data analysis, and fundamental physics areas which
observe vastly different notational traditions. A big effort was made to homogenize this notation across the breadth
of communities in the separate sections of this project. These notation conventions are defined in Table VII. Any
deviations from this notation convection are noted in the specific subsections in which they occur.

Definition Meaning
ℏ = c = kB = 1 Natural units
−2
κ2 ≡ 8πG = MPl Gravitational constant
ln := loge , log := log10 Logarithm conventions
(− + + +) Metric signature
gµν Metric tensor
Gµν ≡ Rµν − 21 gµν R Einstein tensor
h Λ i Cosmological constant
dr 2
ds2 = −dt2 + a2 (t) 1−kr 2
dθ2 + sin2 θdϕ2 Friedmann-Lemaître-Robertson-Walker (FLRW) spacetime metric

2 + r

a(t) Scale factor


a0 = 1 Scale factor today (set to unity)
t Cosmic (proper) time
´ t dt′
τ (t) = 0 a(t ′) Conformal time
˙ ≡ dt d
Cosmic time derivative

≡ dτ d
Conformal time derivative
T µν = √2−g δg δLm
µν
Energy-momentum tensor of the Lagrangian density L
z = −1 + a1 Cosmological redshift
H(z) = ȧa Hubble parameter
H0 Hubble constant
h ≡ H0 /100 km s−1 Mpc−1 Dimensionless reduced Hubble constant
ρm , ρb , ρr Energy density of total matter, baryonic matter, and radiation
ρDM , ρDE Energy density of dark matter and dark energy
Ωm Present-day matter density parameter
Ωr = 2.469 × 10−5 h−2 (1 + 0.2271Neff ) Present-day radiation density parameter
ΩDM , ΩDE Present-day density parameters of dark matter and dark energy
ΩCDM Present-day density parameters of cold dark matter
2
Ωm (z) = κ3Hρm 2 Matter density parameter
2
Ωr (z) = κ3Hρ2r Relativistic content density parameter
κ2 ρDE
ΩDE (z) = 3H 2 Dark energy density parameter
w ≡ pρ Equation of state (EoS) parameter
´τ c s Sound speed
rs ≡ 0 cs (τ ′ )dτ ′ Sound horizon
rd ≡ rs (τd ) Sound horizon at drag epoch
σp 8 Amplitude of mass fluctuations on scales of 8 h−1 Mpc
S8 = σ8 Ωm /0.3 Weighted amplitude of matter fluctuations

TABLE VII: List of notation conventions used in the White Paper (unless otherwise stated).
249

7. List of Acronyms
ACT Atacama Cosmology Telescope
AGN Active Galactic Nucleus
AGB Asymptotic Giant Branch
AGB Artificial Neural Network
BAO Baryon Acoustic Oscillations
BTFR Baryonic Tully-Fisher Relation
BNN Bayesian Neural Network
BBN Big Bang Nucleosynthesis
BOSS Baryon Oscillation Spectroscopic Survey
CC Cosmic chronometers
CDM Cold Dark Matter
CMB Cosmic Microwave Background Radiation
CNN Convolutional Neural Network
CPL Chevallier-Polarski-Linder
DDE Dynamical Dark Energy
DE Dark Energy
DM Dark Matter
eBOSS Extended Baryon Oscillation Spectroscopic Survey
EDE Early Dark Energy
EFTofLSS Effective Field Theory of Large Scale Structure
ELT Extremely Large Telescope
EoR Epoch of Reionization
EoS Equation of State
ESA European Space Agency
ET Einstein Telescope
DES Dark Energy Survey
DESI The Dark Energy Spectroscopic Instrument
FLRW Friedmann-Lemaître-Robertson-Walker
FRB Fast Radio Burst
CMB-S4 CMB—Stage 4
GA Genetic Algorithm
GP Gaussian processes
GR General relativity
GRB Gamma-ray burst
GUP Generalized Uncertainty Principle
GW Gravitational Wave
HST Hubble Space Telescope
HSC Hyper Suprime-Cam
IDE Interacting Dark Energy
IDM Interacting Dark Matter
IGM Intergalactic Medium
ISM Interstellar Medium
ISW Integrated Sachs–Wolfe
250

LDE Late Dark Energy


LF Luminosity Function
LISA Laser Interferometer Space Antenna
LMC Large Magellanic Clouds
JAGB J-region Asymptotic Giant Branch
JWST James Webb Space Telescope
KiDS Kilo-Degree Survey
LIGO Laser Interferometer Gravitational Wave Observatory
LSS Large Scale Structure
LSST Legacy Survey of Space and Time
MG Modified Gravity
MCMC Markov chain Monte Carlo
ML Machine Learning
MW Milky Way galaxy
MOND Modified Newtonian dynamics
NEDE New Early Dark Energy
PMF Primordial Magnetic Field
PDF Probability Density function
QSO QSO
RGB Red Giant Branch
RSD Redshift-Space Distortions
RVM Running Vacuum Model
SBF Surface Brightness Fluctuations
SGWB Stochastic Gravitational Wave Background
SKAO Square Kilometer Array Observatory
SMC Small Magellanic Clouds
SN Supernovae
SNIa Type Ia supernovae
SNII Type II supernovae
SPT South Pole Telescope
SPT-3G South Pole Telescope - 3rd Generation
SZ Sunyaev–Zeldovich
TRGB Tip of the Red Giant Branch
WDM Warm Dark Matter
WL Weak Lensing
WMAP Wilkinson Microwave Anisotropy Probe
ZTF Zwicky Transient Facility
251

Acknowledgements
This paper is based upon work from COST Action CA21136 Addressing observational tensions in cosmology with sys-
tematics and fundamental physics (CosmoVerse) supported by COST (European Cooperation in Science and Technology).
EDV is supported by a Royal Society Dorothy Hodgkin Research Fellowship. JLS would also like to acknowledge funding
from “Xjenza Malta” as part of the “Technology Development Programme” DTP-2024-014 (CosmicLearning) Project and the
“XM-TÜBİTAK R&I Programme” BridgingCosmology project. AGV is funded by “la Caixa” Foundation (ID 100010434)
and the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agree-
ment No 847648, with fellowship code LCF/BQ/PI23/11970027. AP acknowledges support from the Polish National Science
Centre through the grant 2023/50/A/ST9/00579. AP is supported by NSF Grant No. 2308193. CvdB is supported by the
Lancaster–Sheffield Consortium for Fundamental Physics under STFC grant: ST/X000621/1. CU was supported by UKRI
STFC ST/W001020/1 and European Union ERC StG, LSS_BeyondAverage, 101075919. DE acknowledges support from the
Swiss National Science Foundation (SNSF) under grant agreement 200021_212576. EMT is supported by funding from the
European Research Council (ERC) under the European Union’s HORIZON-ERC-2022 (grant agreement no. 101076865). ENS
acknowledges the contribution of the LISA CosWG. This project has received funding from the European Research Council
under the European Union’s Horizon 2020 research and innovation programme (grant agreement 853291). FB acknowledges
the support of the Royal Society through the University Research Fellowship. The work of FN is supported by VR Starting
Grant 2022-03160 of the Swedish Research Council. GB is supported by the Spanish grants CIPROM/2021/054 (Generali-
tat Valenciana) and PID2023-151418NB-I00 funded by MCIU/AEI/10.13039/501100011033. IS acknowledges NASA grants
N4-ADAP24-0021 and 24-ADAP24-0074, and this research was supported in part by grant NSF PHY-2309135 to the Kavli
Institute for Theoretical Physics (KITP). JT is supported by a Ramón y Cajal contract by the Spanish Ministry for Science,
Innovation and Universities with Ref. RYC2023-045660-I. JM would also like to acknowledge funding from “Xjenza Malta” as
part of the “Technology Development Programme” DTP-2024-014 (CosmicLearning) Project. KS acknowledges support from
the Australian Government through the Australian Research Council Centre of Excellence for Gravitational Wave Discovery
(OzGrav), through project number CE230100016. K.F.D. was supported by the PNRR-III-C9-2022–I9 call, with project num-
ber 760016/27.01.2023, and funding from “Xjenza Malta” as part of the “XM-TÜBİTAK R&I Programme” BridgingCosmology
project. LZ is supported by the NAWA Ulam fellowship (No. BPN/ULM/2023/1/00107/U/00001) and the National Science
Centre, Poland (research grant No. 2021/42/E/ST2/00031). LG acknowledges financial support from AGAUR, CSIC, MCIN
and AEI 10.13039/501100011033 under projects PID2023-151307NB-I00, PIE 20215AT016, CEX2020-001058-M, ILINK23001,
COOPB2304, and 2021-SGR-01270. LV acknowledges support by the National Natural Science Foundation of China (NSFC)
through the grant No. 12350610240 “Astrophysical Axion Laboratories”, and also thanks INFN through the “QGSKY” In-
iziativa Specifica project. The work of LAA is supported by the US National Science Foundation Grant PHY-2412679. MA
acknowledges the UK Science and Technology Facilities Council (STFC) under grant number ST/Y002652/1 and the Royal
Society under grant numbers RGSR2222268 and ICAR1231094. MG acknowledges support from the European Union (ERC,
RELiCS, project number 101116027) and the PRIN (Progetti di ricerca di Rilevante Interesse Nazionale) number 2022WJ9J33.
MF is funded by the PRIN (Progetti di ricerca di Rilevante Interesse Nazionale) number 2022WJ9J33. MC and RH acknowl-
edge support from the project “INAF-EDGE” (Large Grant 12-2022, P.I. L. Hunt), from the ASI-INAF agreement “Scientific
Activity for the Euclid Mission” (n.2024-10-HH.0; WP8420) and from the INAF “Astrofisica Fondamentale” GO-grant 2024
(PI M. Cantiello). MM acknowledges support from MIUR, PRIN 2022 (grant 2022NY2ZRS 001) and from the grant ASI
n. 2024-10-HH.0 “Attività scientifiche per la missione Euclid – fase E”. RCN thanks the financial support from the CNPq
under the project No. 304306/2022-3 and FAPERGS under the project No. 23/2551-0000848-3. RCB is supported by an
appointment to the JRG Program at the APCTP through the Science and Technology Promotion Fund and Lottery Fund of
the Korean Government, and was also supported by the Korean Local Governments in Gyeongsangbuk-do Province and Pohang
City. RC acknowledges support from the CONAHCYT research grant CF2022-320152. This project has received funding from
the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant
Agreement No. 947660). RIA is funded by the SNSF Eccellenza Professorial Fellowship PCEFP2_194638. SK acknowledges
funding by the National Center for Science, Poland, grant no. 2023/49/B/ST9/02777. SL is supported by the National
Science Foundation Graduate Research Fellowship Program under grant No. DGE2139757. SV acknowledges support from
the Istituto Nazionale di Fisica Nucleare (INFN) through the Commissione Scientifica Nazionale 4 (CSN4) Iniziativa Specifica
“Quantum Fields in Gravity, Cosmology and Black Holes” (FLAG), and from the University of Trento and the Provincia
Autonoma di Trento (PAT, Autonomous Province of Trento) through the UniTrento Internal Call for Research 2023 grant
“Searching for Dark Energy off the beaten track” (DARKTRACK, grant agreement no. E63C22000500003). SP acknowledges
the financial support from the Department of Science and Technology (DST), Govt. of India under the Scheme “Fund for
Improvement of S&T Infrastructure (FIST)” (File No. SR/FST/MS-I/2019/41). TT acknowledges support from the National
Science Foundation, the National Areonautics and Space Administration, and the Gordon and Betty Moore Foundation. VP is
supported by funding from the European Research Council (ERC) under the European Union’s HORIZON-ERC-2022 (grant
agreement no. 101076865) and from the European Union’s Horizon 2020 research and innovation program under the Marie
Skłodowska-Curie grant agreement no. 860881-HIDDeN. AD acknowledges the support of the European Union’s Horizon 2021
research and innovation programme under the Marie Sklodowska-Curie grant agreement No. 101068013 (QGRANT). AK has
252

been supported by a Lendület excellence grant by the Hungarian Academy of Sciences (MTA). This project has received fund-
ing from the European Union’s Horizon Europe research and innovation programme under the Marie Skłodowska-Curie grant
agreement number 101130774. Funding for this project was also available in part through the Hungarian Ministry of Innovation
and Technology NRDI Office grant OTKA NN147550. ARL was supported by FCT through the Investigador FCT Contract
CEECIND/02854/2017 and the research project PTDC/FIS-AST/0054/2021. AP is grateful for the support of Vicerrectoría
de Investigación y Desarrollo Tecnológico (Vridt) at Universidad Católica del Norte through Núcleo de Investigación Geometría
Diferencial y Aplicaciones, Resolución Vridt No - 096/2022 and Resolución Vridt No - 098/2022. AP was supported by the
Proyecto Fondecyt Regular 2024, Folio 1240514, Etapa 2024. AB acknowledges support from the National Science Center,
project no. UMO-2022/45/B/ST2/01067. ABR is supported by the Fundação Carlos Chagas Filho de Amparo à Pesquisa
do Estado do Rio de Janeiro (FAPERJ), Grant No E-26/200.149/2025 é 200.150/2025 (304809) AJS received support from
NASA through STScI grants HST-GO-16773 and JWST-GO-2974. AM acknowledges the financial support by Conacyt-Mexico
through the Post- doc Project I1200/311/2023. AB acknowledges the fellowship of the Brazilian research agency CNPq. AH
is funded by the Carlsberg foundation. BSS acknowledges National Science Foundation awards AST-2307147, PHY-2207638,
PHY-2308886 and PHY-2309064. The research activities of BT is supported in part by the U.S. National Science Foundation
under Grant PHY-2014104. CGB is supported by the Spanish Grant PID2023-149016NB-I00 (MINECO/AEI/FEDER, UE)
and the Basque government Grant No. IT1628-22 (Spain). CM is supported by an FCT fellowship, grant number 2023.03984.
CZ is supported by the China Scholarship Council for 1 year study at SISSA. CP acknowledges the financial support by the
excellence cluster QuantumFrontiers of the German Research Foundation (Deutsche Forschungsgemeinschaft, DFG) under
Germany’s Excellence Strategy – EXC-2123 QuantumFrontiers – 390837967 and was funded by the Deutsche Forschungsge-
meinschaft (DFG, German Research Foundation) - Project Number 420243324. D.B. acknowledges support from projects
PID2021-122938NB-I00 funded by the Spanish “Ministerio de Ciencia e Innovación” and FEDER “A way of making Europe”,
PID2022-139841NB-I00 funded by the Spanish “Ministerio de Ciencia e Innovación” and SA097P24 funded by Junta de Castilla
y León. DS acknowledges support from Bulgarian National Science Fund grant number KP-06-N58/5. DRG acknowledges
support from grant PID2022-138607NBI00, funded by MCIN/AEI/10.13039/501100011033. DB acknowledges funding from
the Ministry of Science, Technological Development and Innovations of the Republic of Serbia, Project contract No. 451-03-
136/2025-03/200017. The work of EG was supported in part by grant NSF PHY-2309135 to the Kavli Institute for Theoretical
Physics (KITP). EF is supported by “Theoretical Astroparticle Physics” (TAsP), iniziativa specifica INFN and by the research
grant number 2022E2J4RK “PANTHEON: Perspectives in Astroparticle and Neutrino THEory with Old and New messen-
gers”under the program PRIN 2022 funded by the Italian Ministero dell’Università e della Ricerca (MUR). EDMF is supported
by World Premier International Research Center Initiative (WPI Initiative), MEXT, Japan. FA thanks CNPq and Fundação
Carlos Chagas Filho de Amparo à Pesquisa do Estado do Rio de Janeiro (FAPERJ), Processo SEI 260003/014913/2023 for
financial support. The work of FB was supported by the postdoctoral grant CIAPOS/2021/169. FP acknowledges par-
tial support from the INFN grant InDark and from the Italian Ministry of University and Research (mur), PRIN 2022
‘EXSKALIBUR – Euclid-Cross-SKA: Likelihood Inference Building for Universe’s Research’, Grant No. 20222BBYB9, CUP
C53D2300131 0006, and from the European Union – Next Generation EU. FP also acknowledges support from the FCT project
“BEYLA – BEYond LAmbda” with ref. number PTDC/FIS-AST/0054/2021. FSNL acknowledges support from the Fundação
para a Ciência e a Tecnologia (FCT) Scientific Employment Stimulus contract with reference CEECINST/00032/2018, and
funding through the research grants UIDB/04434/2020, UIDP/04434/2020 and PTDC/FIS-AST/0054/2021. GL was funded
by Agencia Nacional de Investigación y Desarrollo (ANID) through Proyecto Fondecyt Regular 2024, Folio 1240514, Etapa
2024. He also thanks Vicerrectoría de Investigación y Desarrollo Tecnológico (VRIDT) at Universidad Católica del Norte for
support through Núcleo de Investigación Geometría Diferencial y Aplicaciones (Resolución VRIDT N°096/2022). GG acknowl-
edges the financial support from the COSMOS network (www.cosmosnet.it) through the ASI (Italian Space Agency) Grants
2016-24-H.0 and 2016-24-H.1- 2018. GDS acknowledges support from INAF-ASTROFIT fellowship and Istituto Nazionale di
Fisica Nucleare (INFN), Naples Section, for specific initiatives QGSKY and Moonlight2, as well as GAIA DPAC funds from
INAF and ASI (PI: M.Lattanzi). GJO acknowledges financial support from the Spanish Grants PID2020-116567GB-C21,
PID2023-149560NB-C21 funded by MCIN/AEI/10.13039/501100011033, by CEX2023-001292-S funded by MCIU/AEI, and
by i-COOP23096 funded by CSIC. GSDj acknowledges the support by the Ministry of Science, Technological Development
and Innovation of the Republic of Serbia under contract 451-03-137/2025-03/ 200124 and support by the ICTP - SEENET-
MTP NT03 Project TECOM-GRASP. GCH acknowledges support through the ESA research fellowship programm. HAF was
partially supported by NSF grant AST-1907404. The work of IDG was supported by the Estonian Research Council grants
MOB3JD1202, RVTT3, RVTT7, and by the CoE program TK202 “Fundamental Universe”. The work of IAM is supported by
the Basque Government Grant IT1628-22, by Grant PID2021-123226NB-I00 (funded by MCIN/AEI/10.13039/501100011033
and by “ERDF A way of making Europe”). IDM acknowledges support from the grant PID2021-122938NB-I00 funded by
MCIN/AEI/10.13039/501100011033 and from the grant SA097P24 funded by Junta de Castilla y León and by “ERDF A way
of making Europe”. JA acknowledges support from the Diputación General de Aragón-Fondo Social Europeo (DGA-FSE)
Grant No. 2020-E21-17R of the Aragon Government. JG and BK have been supported in part by the Polish National Science
Center (NCN) under grant 2020/37/B/ST2/02371. JR is supported by a Ramón y Cajal contract of the Spanish Ministry of
Science and Innovation with Ref. RYC2020-028870-I. This research was further supported by the project PID2022-139841NB-
253

I00 of MICIU/AEI/10.13039/501100011033 and FEDER, UE. JS is supported by the Taiwan National Science & Technology
Council. JARC is supported by the project PID2022-139841NB-I00 funded by MICIU/AEI/10.13039/501100011033 and by
ERDF/EU. JJ acknowledge partial support from STScI under grants HST-GO-16262 and JWST-GO-03055. JGB acknowledges
support from the Spanish Research Project PID2021-123012NB-C43 [MICINN-FEDER], and the Centro de Excelencia Severo
Ochoa Program, Spain CEX2020-001007-S at IFT. KL is a recipient of the John and Pat Hume Scholarship and acknowledges
support from the Friedrich Naumann Foundation for Freedom and from the Swiss Study Foundation. The research activities
of KRD are supported in part by the U.S. National Science Foundation through its employee IR/D program as well as by the
U.S. Department of Energy under Grant DE-FG02-13ER41976 / DE-SC0009913. LLG acknowledges support from Conselho
Nacional de Desenvolvimento Cientıfico e Tecnologico (CNPq), Grant No. 307636/2023-2 and from the Fundacao Carlos
Chagas Filho de Amparo a Pesquisa do Estado do Rio de Janeiro (FAPERJ) Grant No. E-26/204.598/2024. LY acknowledges
support from YST Program of APCTP and Natural Science Foundation of Shanghai 24ZR1424600. LC acknowledges support
from FCT - Fundação para a Ciência e a Tecnologia through the projects with DOI identifiers 10.54499/2023.11681.PEX and
10.54499/2024.00249.CERN. MB is supported by the Polish National Science Center through grants no. 2020/38/E/ST9/00395
and 2020/39/B/ST9/03494. MGE acknowledges the financial support of FONDECYT de Postdoctorado, N° 3230801. MdC
acknowledges support from INFN iniziativa specifica GeoSymQFT. The work of MR is supported by the European Union – Next
Generation EU and by the Italian Ministry of University and Research (MUR) via the PRIN 2022 project n. 20228WHTYC.
ME was supported by the Estonian Ministry of Education and Research (grant TK202, “Foundations of the Universe”), by
Estonian Research Council grant PRG2172, and by the European Union’s Horizon Europe research and innovation programme
(EXCOSM, grant No. 101159513). The work of MBL, CGB & PM is supported by the Spanish Grant PID2023-149016NB-
I00 (MINECO/AEI/FEDER, UE). This work is also supported by the Basque government Grant No. IT1628-22 (Spain).
MC was supported by FCT through the Investigador FCT Contract No. CEECIND/02581/2018 and the research project
PTDC/FIS-AST/0054/2021. MM acknowledges funding by the Agenzia Spaziale Italiana (asi) under agreement no. 2018-23-
HH.0 and support from INFN/Euclid Sezione di Roma. MASP acknowledges support from the FCT through the Fellowship
UI/BD/154479/2022, through the Research Grants UIDB/04434/2020 and UIDP/04434/2020, and through the project with
reference PTDC/FIS-AST/0054/2021 (“BEYond LAmbda”). MASP also acknowledges support from the MICINU through
the project with reference PID2023-149560NB-C21. MAS received support from the CAPES scholarship. MK was funded
through SASPRO2 project AGE of Gravity: Alternative Geometries of Gravity, which has received funding from the European
Union’s Horizon 2020 research and innovation programme under the Marie Skłodowska-Curie grant agreement No. 945478.
MM acknowledges the support by the Ministry of Science, Technological Development and Innovation of the Republic of Serbia
under contract 451-03-137/2025-03/ 200124. MI acknowledges that this material is based upon work supported in part by the
Department of Energy, Office of Science, under Award Number DE-SC0022184 and also in part by the U.S. National Science
Foundation under grant AST2327245. NF acknowledge support from the Research grant TAsP (Theoretical Astroparticle
Physics) funded by INFN. NF is supported by the European Union – Next Generation EU and by the Italian Ministry of
University and Research (MUR) via the PRIN 2022 project n. 20228WHTYC. The Work of NEM. is supported in part by the
UK Science and Technology Facilities research Council (STFC) under the research grant ST/X000753/1. NAN was financed
by IBS under the project code IBS-R018-D3, and acknowledges support from PSL/Observatoire de Paris. NS is supported
by the Charles University Grant Agency (GAUK) - 94224. PS is supported by Science and Technology Facilities Council
(STFC) training grant ST/X508287/1. PJ acknowledges funding from the Ministry of Science, Technological Development and
Innovations of the Republic of Serbia, Project contract No. 451-03-136/2025-03/200002. RG acknowledges the support from
the SNF 200020_175751 and 200020_207379 “Cosmology with 3D Maps of the Universe” research grant. RH work has been
supported by the Spanish grants FPU19/03348 of MU, MCIN/AEI/10.13039/501100011033 grants PID2020-113644GB-I00,
PID2023-148162NB-C22, by the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-
Curie grants HORIZON-MSCA-2021-SE-01/101086085-ASYMMETRY and H2020-MSCA-ITN-2019/860881-HIDDeN and by
the Generalitat Valenciana grants PROMETEO/2019/083 and CIPROM/2022/69. RH acknowledges funding from the Italian
National Institute of Astrophysics (INAF) through large grant PRIN 12-2022 “INAF-EDGE” (PI L. Hunt). RR acknowl-
edges financial support from the STFC Consolidated Grant ST/X000583/1. RBN acknowledges funding by the Royal Society
through the University Research Fellowship Renewal URF R 221005. RK is supported by Project SA097P24 funded by
Junta de Castilla y Leon. SB is supported by “Agencia Nacional de Investigación y Desarrollo” (ANID), Grant “Becas Chile
postdoctorado al extranjero” No. 74220006. ST is supported by the National Science Centre, Poland (research grant No.
2021/42/E/ST2/00031). SSC acknowledges support from the Istituto Nazionale di Fisica Nucleare (INFN) through the Com-
missione Scientifica Nazionale 4 (CSN4) Iniziativa Specifica “Quantum Fields in Gravity, Cosmology and Black Holes” (FLAG)
and from the Fondazione Cassa di Risparmio di Trento e Rovereto (CARITRO Foundation) through a Caritro Fellowship
(project “Inflation and dark sector physics in light of next-generation cosmological surveys”). SJL is supported by grant PIP
11220200100729CO CONICET and grant 20020170100129BA UBACYT. SP is supported by the international Gemini Obser-
vatory, a program of NSF NOIRLab, which is managed by the Association of Universities for Research in Astronomy (AURA)
under a cooperative agreement with the U.S. National Science Foundation, on behalf of the Gemini partnership of Argentina,
Brazil, Canada, Chile, the Republic of Korea, and the United States of America. SP acknowledges the financial support of
the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) Fellowships 300936/2023-0 and 301628/2024-6.
254

TP acknowledges the financial support from the Slovenian Research Agency (grants I0-0033, P1-0031, J1-8136, J1-2460 and
Z1-1853). TB was supported by ICSC – Centro Nazionale di Ricerca in High Performance Computing, Big Data and Quantum
Computing, funded by European Union – NextGenerationEU. TS is supported by the Della Riccia foundation grant 2025 and
the Galileo Galilei Institute Boost Fellowship 2024. TK was supported by the Estonian Research Council grants CoE TK202
“Foundations of the Universe” and PRG2608 “Space - Time - Matter”. VM thanks CNPq (Brazil) and FAPES (Brazil) for partial
financial support. VAM is supported by Generalitat Valenciana via the Excellence Grant CIPROM/2021/073, by the Spanish
MICIN/AEI/10.13039/501100011033 and the EU/FEDER via grant PID2021-122134NB-C21, and by the Spanish MCIU/AEI
via the Severo Ochoa project CEX2023-001292-S. VP acknowledges the support by the Ministry of Education and Science
of the Federation of Bosnia and Herzegovina, under Project Number 05-35-2467-1/23. VK acknowledges financial support
from Research Ireland Grant 21/PATH-S/9475 (MOREHIGGS) under the SFI-IRC Pathway Programme. Report number:
DIAS-STP-25-05. VM acknowledges support from ANID FONDECYT Regular grant number 1231418, Millennium Science
Initiative AIM23-0001, and Centro de Astrofísica de Valparaíso CIDI N21. VBJ acknowledges funding from the Ministry of Sci-
ence, Technological Development and Innovations of the Republic of Serbia, Project contract No. 451-03-136/2025-03/200017.
WY has been supported by the National Natural Science Foundation of China under Grant No. 12175096, and Liaoning
Revitalization Talents Program under Grant no. XLYC1907098. WY is supported via the research projects ‘COLAB’ funded
by the National Science Center, Poland, under agreement number UMO-2020/39/B/ST9/03494. “YH is funded by NSFC
under the grant No.12347137 and the China Postdoctoral Science Foundation under Grant No. 2024M753076”.

Afternote: The opinions and conclusions expressed herein are those of the authors, and do not represent any funding
agencies.
255

[1] CosmoVerse Network, “CA21136 - Addressing observational tensions in cosmology with systematics and fundamental
physics.” https://www.cost.eu/actions/CA21136.
[2] CosmoVerse Network, “CA21136 - Addressing observational tensions in cosmology with systematics and fundamental
physics.” https://cosmoversetensions.eu/.
[3] E. Abdalla et al., “Cosmology intertwined: A review of the particle physics, astrophysics, and cosmology associated
with the cosmological tensions and anomalies,” JHEAp 34 (2022) 49–211, arXiv:2203.06142 [astro-ph.CO].
[4] E. Di Valentino et al., “Snowmass2021 - Letter of interest cosmology intertwined I: Perspectives for the next decade,”
Astropart. Phys. 131 (2021) 102606, arXiv:2008.11283 [astro-ph.CO].
[5] E. Di Valentino et al., “Snowmass2021 - Letter of interest cosmology intertwined II: The hubble constant tension,”
Astropart. Phys. 131 (2021) 102605, arXiv:2008.11284 [astro-ph.CO].
[6] E. Di Valentino et al., “Cosmology Intertwined III: f σ8 and S8 ,” Astropart. Phys. 131 (2021) 102604,
arXiv:2008.11285 [astro-ph.CO].
[7] E. Di Valentino et al., “Snowmass2021 - Letter of interest cosmology intertwined IV: The age of the universe and its
curvature,” Astropart. Phys. 131 (2021) 102607, arXiv:2008.11286 [astro-ph.CO].
[8] CosmoVerse Network Seminars, “CA21136 - Addressing observational tensions in cosmology with systematics and
fundamental physics.” https://cosmoversetensions.eu/for-scientists/cosmoverse-seminars/.
[9] CosmoVerse Network Conferences, “CA21136 - Addressing observational tensions in cosmology with systematics and
fundamental physics.” https://cosmoversetensions.eu/category/event/conferences/.
[10] CosmoVerse Network School, “CA21136 - Addressing observational tensions in cosmology with systematics and
fundamental physics.” https://cosmoversetensions.eu/event/cosmoverseschoolcorfu/.
[11] CosmoVerse Network Training Series, “CA21136 - Addressing observational tensions in cosmology with systematics
and fundamental physics.” https://cosmoversetensions.eu/for-scientists/training-series/.
[12] CosmoVerse Network Training Series, “CA21136 - Addressing observational tensions in cosmology with systematics
and fundamental physics.” https://cosmoversetensions.eu/for-scientists/journal-club/.
[13] G. Rácz, L. Dobos, R. Beck, I. Szapudi, and I. Csabai, “Concordance cosmology without dark energy,” Mon. Not. Roy.
Astron. Soc. 469 (2017) no. 1, L1–L5, arXiv:1607.08797 [astro-ph.CO].
[14] R. Beck, I. Csabai, G. Rácz, and I. Szapudi, “The integrated Sachs–Wolfe effect in the AvERA cosmology,” Mon. Not.
Roy. Astron. Soc. 479 (2018) no. 3, 3582–3591, arXiv:1801.08566 [astro-ph.CO].
[15] W. L. Freedman et al., “The Carnegie-Chicago Hubble Program. VIII. An Independent Determination of the Hubble
Constant Based on the Tip of the Red Giant Branch,” Astrophys. J. 882 (2019) 34, arXiv:1907.05922 [astro-ph.CO].
[16] G. S. Anand et al., “Tip of the Red Giant Branch Distances with JWST: An Absolute Calibration in NGC 4258 and
First Applications to Type Ia Supernova Hosts,” Astrophys. J. 966 (2024) no. 1, 89, arXiv:2401.04776
[astro-ph.CO].
[17] A. J. Lee, W. L. Freedman, I. S. Jang, B. F. Madore, and K. A. Owens, “First JWST Observations of JAGB Stars in
the SN Ia Host Galaxies: NGC 7250, NGC 4536, NGC 3972,” Astrophys. J. 961 (2024) no. 1, 132, arXiv:2312.02282
[astro-ph.GA].
[18] S. Li, A. G. Riess, S. Casertano, G. S. Anand, D. M. Scolnic, W. Yuan, L. Breuval, and C. D. Huang, “Reconnaissance
with JWST of the J-region Asymptotic Giant Branch in Distance Ladder Galaxies: From Irregular Luminosity
Functions to Approximation of the Hubble Constant,” Astrophys. J. 966 (2024) no. 1, 20, arXiv:2401.04777
[astro-ph.CO].
[19] C. D. Huang, A. G. Riess, W. Yuan, L. M. Macri, N. L. Zakamska, S. Casertano, P. A. Whitelock, S. L. Hoffmann,
A. V. Filippenko, and D. Scolnic, “Hubble Space Telescope Observations of Mira Variables in the Type Ia Supernova
Host NGC 1559: An Alternative Candle to Measure the Hubble Constant,” Astrophys. J. 889 (2020) 5,
arXiv:1908.10883 [astro-ph.CO].
[20] C. D. Huang et al., “The Mira Distance to M101 and a 4% Measurement of H 0 ,” Astrophys. J. 963 (2024) no. 2, 83,
arXiv:2312.08423 [astro-ph.CO].
[21] J. P. Blakeslee, J. B. Jensen, C.-P. Ma, P. A. Milne, and J. E. Greene, “The Hubble Constant from Infrared Surface
Brightness Fluctuation Distances,” Astrophys. J. 911 (2021) no. 1, 65, arXiv:2101.02221 [astro-ph.CO].
[22] G. S. Anand, R. B. Tully, Y. Cohen, D. I. Makarov, L. N. Makarova, J. B. Jensen, J. P. Blakeslee, M. Cantiello,
E. Kourkchi, and G. Raimondo, “The Population II Extragalactic Distance Scale: A TRGB Distance to the Fornax
Cluster with JWST,” Astrophys. J. 973 (2024) no. 2, 83, arXiv:2405.03743 [astro-ph.CO].
[23] T. de Jaeger, L. Galbany, A. G. Riess, B. E. Stahl, B. J. Shappee, A. V. Filippenko, and W. Zheng, “A 5 per cent
measurement of the Hubble–Lemaître constant from Type II supernovae,” Mon. Not. Roy. Astron. Soc. 514 (2022)
no. 3, 4620–4628, arXiv:2203.08974 [astro-ph.CO].
[24] G. Csörnyei, C. Vogl, S. Taubenberger, A. Flörs, S. Blondin, M. G. Cudmani, A. Holas, S. Kressierer, B. Leibundgut,
and W. Hillebrandt, “Consistency of Type IIP supernova sibling distances,” Astron. Astrophys. 672 (2023) A129,
arXiv:2302.03112 [astro-ph.SR].
[25] E. Kourkchi, R. B. Tully, H. M. Courtois, A. Dupuy, and D. Guinet, “Cosmicflows-4: the baryonic Tully–Fisher
relation providing ∼10 000 distances,” Mon. Not. Roy. Astron. Soc. 511 (2022) no. 4, 6160–6178, arXiv:2201.13023
[astro-ph.GA].
[26] A. G. Riess et al., “JWST Validates HST Distance Measurements: Selection of Supernova Subsample Explains
256

Differences in JWST Estimates of Local H 0 ,” Astrophys. J. 977 (2024) no. 1, 120, arXiv:2408.11770 [astro-ph.CO].
[27] HST Collaboration, W. L. Freedman et al., “Final results from the Hubble Space Telescope key project to measure the
Hubble constant,” Astrophys. J. 553 (2001) 47–72, arXiv:astro-ph/0012376.
[28] Gaia Collaboration, “Gaia Data Release 3. Summary of the content and survey properties,” A&A 674 (2023) A1,
arXiv:2208.00211 [astro-ph.GA].
[29] G. Pietrzyński et al., “A distance to the Large Magellanic Cloud that is precise to one per cent,” Nature 567 (2019)
no. 7747, 200–203.
[30] D. Graczyk, G. Pietrzyński, I. B. Thompson, W. Gieren, B. Zgirski, S. Villanova, M. Górski, P. Wielgórski,
P. Karczmarek, W. Narloch, B. Pilecki, M. Taormina, R. Smolec, K. Suchomska, A. Gallenne, N. Nardetto, J. Storm,
R.-P. Kudritzki, M. Kałuszyński, and W. Pych, “A Distance Determination to the Small Magellanic Cloud with an
Accuracy of Better than Two Percent Based on Late-type Eclipsing Binary Stars,” Astrophys. J. 904 (2020) no. 1, 13,
arXiv:2010.08754 [astro-ph.GA].
[31] M. J. Reid, D. W. Pesce, and A. G. Riess, “An Improved Distance to NGC 4258 and its Implications for the Hubble
Constant,” Astrophys. J. Lett. 886 (2019) no. 2, L27, arXiv:1908.05625 [astro-ph.GA].
[32] D. Scolnic et al., “The Pantheon+ Analysis: The Full Data Set and Light-curve Release,” Astrophys. J. 938 (2022)
no. 2, 113, arXiv:2112.03863 [astro-ph.CO].
[33] D. Brout et al., “The Pantheon+ Analysis: Cosmological Constraints,” Astrophys. J. 938 (2022) no. 2, 110,
arXiv:2202.04077 [astro-ph.CO].
[34] A. G. Riess et al., “A Comprehensive Measurement of the Local Value of the Hubble Constant with 1 km/s/Mpc
Uncertainty from the Hubble Space Telescope and the SH0ES Team,” Astrophys. J. Lett. 934 (2022) no. 1, L7,
arXiv:2112.04510 [astro-ph.CO].
[35] L. Breuval, A. G. Riess, S. Casertano, W. Yuan, L. M. Macri, M. Romaniello, Y. S. Murakami, D. Scolnic, G. S.
Anand, and I. Soszyński, “Small Magellanic Cloud Cepheids Observed with the Hubble Space Telescope Provide a New
Anchor for the SH0ES Distance Ladder,” Astrophys. J. 973 (2024) no. 1, 30, arXiv:2404.08038 [astro-ph.CO].
[36] L. Galbany et al., “An updated measurement of the Hubble constant from near-infrared observations of Type Ia
supernovae,” Astron. Astrophys. 679 (2023) A95, arXiv:2209.02546 [astro-ph.CO].
[37] V. Scowcroft, W. L. Freedman, B. F. Madore, A. Monson, S. E. Persson, J. Rich, M. Seibert, and J. R. Rigby, “The
Carnegie Hubble Program: The Distance and Structure of the SMC as Revealed by Mid-infrared Observations of
Cepheids,” Astrophys. J. 816 (2016) no. 2, 49, arXiv:1502.06995 [astro-ph.GA].
[38] I. Soszyński, A. Udalski, M. K. Szymański, D. Skowron, G. Pietrzyński, R. Poleski, P. Pietrukowicz, J. Skowron,
P. Mróz, S. Kozłowski, Ł. Wyrzykowski, K. Ulaczyk, and M. Pawlak, “The OGLE Collection of Variable Stars. Classical
Cepheids in the Magellanic System,” Acta Astron. 65 (2015) no. 4, 297–312, arXiv:1601.01318 [astro-ph.SR].
[39] V. Ripepi, M.-R. L. Cioni, M. I. Moretti, M. Marconi, K. Bekki, G. Clementini, R. de Grijs, J. Emerson, M. A. T.
Groenewegen, V. D. Ivanov, R. Molinaro, T. Muraveva, J. M. Oliveira, A. E. Piatti, S. Subramanian, and J. T. van
Loon, “The VMC survey - XXV. The 3D structure of the Small Magellanic Cloud from Classical Cepheids,” MNRAS
472 (2017) no. 1, 808–827, arXiv:1707.04500 [astro-ph.GA].
[40] S. A. Zhevakin, “Physical basis of the pulsation theory of variable stars,” Ann. Rev. Astron. Astrophys. 1 (1963)
367–400.
[41] J. P. Cox, Theory of Stellar Pulsation. (PSA-2), Volume 2, vol. 2. Princeton University Press, 1980.
[42] H. S. Leavitt and E. C. Pickering, “Periods of 25 Variable Stars in the Small Magellanic Cloud,” Harvard Obs. Circ.
173 (1912) 1–3.
[43] H. Shapley, “Studies based on the colors and magnitudes in stellar clusters. IX. Three notes on cepheid variation.,”
Astrophys. J. 49 (1919) 24–41.
[44] S. L. Hoffmann et al., “Optical Identification of Cepheids in 19 Host Galaxies of Type Ia Supernovae and NGC 4258
with the Hubble Space Telescope,” Astrophys. J. 830 (2016) no. 1, 10, arXiv:1607.08658 [astro-ph.SR].
[45] L. Inno, N. Matsunaga, M. Romaniello, G. Bono, A. Monson, I. Ferraro, G. Iannicola, E. Persson, R. Buonanno,
W. Freedman, W. Gieren, M. A. T. Groenewegen, Y. Ita, C. D. Laney, B. Lemasle, B. F. Madore, T. Nagayama,
Y. Nakada, M. Nonino, G. Pietrzyński, F. Primas, V. Scowcroft, I. Soszyński, T. Tanabé, and A. Udalski, “New NIR
light-curve templates for classical Cepheids,” A&A 576 (2015) A30, arXiv:1410.5460 [astro-ph.SR].
[46] L. Breuval, A. G. Riess, L. M. Macri, S. Li, W. Yuan, S. Casertano, T. Konchady, B. Trahin, M. J. Durbin, and B. F.
Williams, “A 1.3% Distance to M33 from Hubble Space Telescope Cepheid Photometry,” Astrophys. J. 951 (2023)
no. 2, 118, arXiv:2304.00037 [astro-ph.CO].
[47] R. I. Anderson, “On Cepheid distances in the H0 measurement,” arXiv e-prints (2024) arXiv:2403.02801,
arXiv:2403.02801 [astro-ph.SR].
[48] A. G. Riess, S. Casertano, W. Yuan, L. M. Macri, and D. Scolnic, “Large Magellanic Cloud Cepheid Standards Provide
a 1% Foundation for the Determination of the Hubble Constant and Stronger Evidence for Physics beyond ΛCDM,”
Astrophys. J. 876 (2019) no. 1, 85, arXiv:1903.07603 [astro-ph.CO].
[49] A. G. Riess, G. Narayan, and A. Calamida, “Calibration of the WFC3-IR Count-rate Nonlinearity, Sub-percent
Accuracy for a Factor of a Million in Flux.” Instrument science report wfc3 2019-1, 13 pages, 2019.
[50] A. G. Riess, G. S. Anand, W. Yuan, S. Casertano, A. Dolphin, L. M. Macri, L. Breuval, D. Scolnic, M. Perrin, and
R. I. Anderson, “Crowded No More: The Accuracy of the Hubble Constant Tested with High-resolution Observations
of Cepheids by JWST,” Astrophys. J. Lett. 956 (2023) no. 1, L18, arXiv:2307.15806 [astro-ph.CO].
[51] A. G. Riess, G. S. Anand, W. Yuan, S. Casertano, A. Dolphin, L. M. Macri, L. Breuval, D. Scolnic, M. Perrin, and
I. R. Anderson, “JWST Observations Reject Unrecognized Crowding of Cepheid Photometry as an Explanation for the
257

Hubble Tension at 8σ Confidence,” Astrophys. J. Lett. 962 (2024) no. 1, L17, arXiv:2401.04773 [astro-ph.CO].
[52] R. C. Kennicutt, Jr. et al., “The HST Key Project on the Extragalactic Distance Scale. 13. The Metallicity dependence
of the Cepheid distance scale,” Astrophys. J. 498 (1998) 181, arXiv:astro-ph/9712055.
[53] S. Sakai, L. Ferrarese, R. Kennicutt, and A. Saha, “The Effect of metallicity on Cepheid - based distances,” Astrophys.
J. 608 (2004) 42–61, arXiv:astro-ph/0402499.
[54] L. M. Macri, K. Z. Stanek, D. Bersier, L. Greenhill, and M. Reid, “A new Cepheid distance to the maser-host galaxy
NGC 4258 and its implications for the Hubble Constant,” Astrophys. J. 652 (2006) 1133–1149,
arXiv:astro-ph/0608211.
[55] V. Ripepi, G. Catanzaro, R. Molinaro, M. Gatto, G. De Somma, M. Marconi, M. Romaniello, S. Leccia, I. Musella,
E. Trentin, G. Clementini, V. Testa, F. Cusano, and J. Storm, “Cepheid metallicity in the Leavitt law (C-metall)
survey - I. HARPS-N@TNG spectroscopy of 47 classical Cepheids and 1 BL Her variables,” MNRAS 508 (2021) no. 3,
4047–4071, arXiv:2108.11391 [astro-ph.GA].
[56] A. Bhardwaj et al., “High-resolution Spectroscopic Metallicities of Milky Way Cepheid Standards and Their Impact on
the Leavitt Law and the Hubble Constant,” Astrophys. J. Lett. 955 (2023) no. 1, L13, arXiv:2309.03263
[astro-ph.SR].
[57] E. Trentin, V. Ripepi, R. Molinaro, G. Catanzaro, J. Storm, G. De Somma, M. Marconi, A. Bhardwaj, M. Gatto,
V. Testa, I. Musella, G. Clementini, and S. Leccia, “Cepheid Metallicity in the Leavitt Law (C- MetaLL) survey. IV.
The metallicity dependence of Cepheid period-luminosity relations,” A&A 681 (2024) A65, arXiv:2310.03603
[astro-ph.SR].
[58] L. Breuval, A. G. Riess, P. Kervella, R. I. Anderson, and M. Romaniello, “An Improved Calibration of the Wavelength
Dependence of Metallicity on the Cepheid Leavitt Law,” Astrophys. J. 939 (2022) no. 2, 89, arXiv:2205.06280
[astro-ph.GA].
[59] R. I. Anderson, H. Saio, S. Ekström, C. Georgy, and G. Meynet, “On the Effect of Rotation on Populations of Classical
Cepheids II. Pulsation Analysis for Metallicities 0.014, 0.006, and 0.002,” Astron. Astrophys. 591 (2016) A8,
arXiv:1604.05691 [astro-ph.SR].
[60] G. De Somma, M. Marconi, R. Molinaro, V. Ripepi, S. Leccia, and I. Musella, “An Updated Metal-dependent
Theoretical Scenario for Classical Cepheids,” ApJS 262 (2022) no. 1, 25, arXiv:2206.11154 [astro-ph.SR].
[61] B. F. Madore, “The period-luminosity relation. IV - Intrinsic relations and reddenings for the Large Magellanic Cloud
Cepheids,” Astrophys. J. 253 (1982) 575–579.
[62] M. Romaniello et al., “The iron and oxygen content of LMC Classical Cepheids and its implications for the
extragalactic distance scale and Hubble constant - Equivalent width analysis with Kurucz stellar atmosphere models,”
Astron. Astrophys. 658 (2022) A29, arXiv:2110.08860 [astro-ph.CO]. [Erratum: Astron.Astrophys. 662, C1 (2022)].
[63] D. Zaritsky, R. C. Kennicutt, and J. P. Huchra, “H II regions and the abundance properties of spiral galaxies,”
Astrophys. J. 420 (1994) 87.
[64] A. G. Riess, W.-D. Li, P. B. Stetson, A. V. Filippenko, S. Jha, R. P. Kirshner, P. M. Challis, P. M. Garnavich, and
R. Chornock, “Cepheid calibrations from the Hubble Space Telescope of the luminosity of two recent Type Ia
supernovae and a re-determination of the Hubble constant,” Astrophys. J. 627 (2005) 579–607,
arXiv:astro-ph/0503159.
[65] A. G. Riess et al., “Cepheid Calibrations of Modern Type Ia Supernovae:Implications for the Hubble Constant,”
Astrophys. J. Suppl. 183 (2009) 109–141, arXiv:0905.0697 [astro-ph.CO].
[66] L. Galbany et al., “Nearby supernova host galaxies from the CALIFA Survey: II. SN environmental metallicity,”
Astron. Astrophys. 591 (2016) A48, arXiv:1603.07808 [astro-ph.GA].
[67] P. Kervella, A. Gallenne, N. Remage Evans, L. Szabados, F. Arenou, A. Mérand, Y. Proto, P. Karczmarek,
N. Nardetto, W. Gieren, and G. Pietrzynski, “Multiplicity of Galactic Cepheids and RR Lyrae stars from Gaia DR2. I.
Binarity from proper motion anomaly,” A&A 623 (2019) A116, arXiv:1903.03632 [astro-ph.SR].
[68] P. Karczmarek, G. Hajdu, G. Pietrzyński, W. Gieren, W. Narloch, R. Smolec, G. Wiktorowicz, and K. Belczynski,
“Synthetic Population of Binary Cepheids. II. The Effect of Companion Light on the Extragalactic Distance Scale,”
Astrophys. J. 950 (2023) no. 2, 182, arXiv:2303.15664 [astro-ph.GA].
[69] R. I. Anderson and A. G. Riess, “On Cepheid Distance Scale Bias Due to Stellar Companions and Cluster
Populations,” Astrophys. J. 861 (2018) no. 1, 36, arXiv:1712.01065 [astro-ph.SR].
[70] B. J. Mochejska, L. M. Macri, D. D. Sasselov, and K. Z. Stanek, “The direct project: influence of blending on the
cepheid distance scale. I. cepheids in m31,” Astron. J. 120 (2000) 810, arXiv:astro-ph/9908293.
[71] B. Follin and L. Knox, “Insensitivity of the distance ladder Hubble constant determination to Cepheid calibration
modelling choices,” Mon. Not. Roy. Astron. Soc. 477 (2018) no. 4, 4534–4542, arXiv:1707.01175 [astro-ph.CO].
[72] L. Perivolaropoulos and F. Skara, “Hubble tension or a transition of the Cepheid SnIa calibrator parameters?,” Phys.
Rev. D 104 (2021) no. 12, 123511, arXiv:2109.04406 [astro-ph.CO].
[73] C. Hahn, T. K. Starkenburg, D. Anglés-Alcázar, E. Choi, R. Davé, C. Dickey, K. G. Iyer, A. H. Maller, R. S.
Somerville, J. L. Tinker, and L. Y. A. Yung, “IQ Collaboratory. III. The Empirical Dust Attenuation
Framework-Taking Hydrodynamical Simulations with a Grain of Dust,” Astrophys. J. 926 (2022) no. 2, 122,
arXiv:2106.09741 [astro-ph.GA].
[74] G. F. Benedict, B. E. McArthur, M. W. Feast, T. G. Barnes, T. E. Harrison, R. J. Patterson, J. W. Menzies, J. L.
Bean, and W. L. Freedman, “Hubble Space Telescope Fine Guidance Sensor Parallaxes of Galactic Cepheid Variable
Stars: Period-Luminosity Relations,” Astron. J. 133 (2007) 1810–1827, arXiv:astro-ph/0612465. [Erratum: Astron.J.
133, 2980 (2007)].
258

[75] S. Casertano et al., “Parallax of Galactic Cepheids from Spatially Scanning the Wide Field Camera 3 on the Hubble
Space Telescope: The Case of SS Canis Majoris,” Astrophys. J. 825 (2016) no. 1, 11, arXiv:1512.09371
[astro-ph.SR].
[76] A. G. Riess et al., “New Parallaxes of Galactic Cepheids from Spatially Scanning the Hubble Space Telescope:
Implications for the Hubble Constant,” Astrophys. J. 855 (2018) no. 2, 136, arXiv:1801.01120 [astro-ph.SR].
[77] L. Lindegren, U. Bastian, M. Biermann, A. Bombrun, A. de Torres, E. Gerlach, R. Geyer, J. Hernández, T. Hilger,
D. Hobbs, S. A. Klioner, U. Lammers, P. J. McMillan, M. Ramos-Lerate, H. Steidelmüller, C. A. Stephenson, and
F. van Leeuwen, “Gaia Early Data Release 3. Parallax bias versus magnitude, colour, and position,” A&A 649 (2021)
A4, arXiv:2012.01742 [astro-ph.IM].
[78] A. G. Riess, S. Casertano, W. Yuan, J. B. Bowers, L. Macri, J. C. Zinn, and D. Scolnic, “Cosmic Distances Calibrated
to 1% Precision with Gaia EDR3 Parallaxes and Hubble Space Telescope Photometry of 75 Milky Way Cepheids
Confirm Tension with ΛCDM,” Astrophys. J. Lett. 908 (2021) no. 1, L6, arXiv:2012.08534 [astro-ph.CO].
[79] S. Li, S. Casertano, and A. G. Riess, “A Maximum Likelihood Calibration of the Tip of the Red Giant Branch
Luminosity from High Latitude Field Giants Using Gaia Early Data Release 3 Parallaxes,” Astrophys. J. 939 (2022)
no. 2, 96, arXiv:2202.11110 [astro-ph.GA].
[80] R. I. Anderson, L. Eyer, and N. Mowlavi, “Cepheids in Open Clusters: An 8-D All-sky Census,” Mon. Not. Roy.
Astron. Soc. 434 (2013) 2238, arXiv:1212.5119 [astro-ph.GA].
[81] L. Breuval et al., “The Milky Way Cepheid Leavitt law based on Gaia DR2 parallaxes of companion stars and host
open cluster populations,” Astron. Astrophys. 643 (2020) A115, arXiv:2006.08763 [astro-ph.SR].
[82] A. G. Riess, L. Breuval, W. Yuan, S. Casertano, L. M. Macri, J. B. Bowers, D. Scolnic, T. Cantat-Gaudin, R. I.
Anderson, and M. C. Reyes, “Cluster Cepheids with High Precision Gaia Parallaxes, Low Zero-point Uncertainties, and
Hubble Space Telescope Photometry,” Astrophys. J. 938 (2022) no. 1, 36, arXiv:2208.01045 [astro-ph.CO].
[83] M. C. Reyes and R. I. Anderson, “A 0.9% calibration of the Galactic Cepheid luminosity scale based on Gaia DR3 data
of open clusters and Cepheids,” Astron. Astrophys. 672 (2023) A85, arXiv:2208.09403 [astro-ph.GA].
[84] W. Yuan, L. M. Macri, A. G. Riess, T. G. Brink, S. Casertano, A. V. Filippenko, S. L. Hoffmann, C. D. Huang, and
D. Scolnic, “Absolute Calibration of Cepheid Period–Luminosity Relations in NGC 4258,” Astrophys. J. 940 (2022)
no. 1, 64, arXiv:2203.06681 [astro-ph.GA].
[85] S. Li, A. G. Riess, M. P. Busch, S. Casertano, L. M. Macri, and W. Yuan, “A Sub-2% Distance to M31 from
Photometrically Homogeneous Near-infrared Cepheid Period–Luminosity Relations Measured with the Hubble Space
Telescope,” Astrophys. J. 920 (2021) no. 2, 84, arXiv:2107.08029 [astro-ph.CO].
[86] A. L. Argon, L. J. Greenhill, J. M. Moran, M. J. Reid, K. M. Menten, and M. Inoue, “The IC133 water vapor maser in
the galaxy M33: A Geometric distance,” Astrophys. J. 615 (2004) 702–719, arXiv:astro-ph/0407486.
[87] A. Z. Bonanos et al., “The First DIRECT Distance Determination to a Detached Eclipsing Binary in M33,” Astrophys.
J. 652 (2006) 313–322, arXiv:astro-ph/0606279.
[88] M. Taormina, R.-P. Kudritzki, J. Puls, B. Pilecki, E. Sextl, G. Pietrzyński, M. A. Urbaneja, and W. Gieren, “Toward
Early-type Eclipsing Binaries as Extragalactic Milestones. II. NLTE Spectral Analysis and Stellar Parameters of the
Detached O-type System OGLE-LMC-ECL-06782 in the LMC,” Astrophys. J. 890 (2020) no. 2, 137,
arXiv:2001.04762 [astro-ph.SR].
[89] A. Salsi, N. Nardetto, D. Mourard, D. Graczyk, M. Taormina, O. Creevey, V. Hocdé, F. Morand, K. Perraut,
G. Pietrzynski, and G. H. Schaefer, “Progress on the calibration of surface brightness-color relations for early- and
late-type stars,” A&A 652 (2021) A26, arXiv:2106.01073 [astro-ph.SR].
[90] C.-C. Ngeow, A. Bhardwaj, J.-Y. Henderson, M. J. Graham, R. R. Laher, M. S. Medford, J. Purdum, and
B. Rusholme, “Zwicky Transient Facility and Globular Clusters: The Period-Luminosity and Period-Wesenheit
Relations for Type II Cepheids,” Astrophys. J. 164 (2022) no. 4, 154, arXiv:2208.03404 [astro-ph.SR].
[91] W. Narloch, G. Hajdu, G. Pietrzyński, W. Gieren, P. Wielgórski, B. Zgirski, P. Karczmarek, M. Górski, and
D. Graczyk, “Period-Luminosity Relations for Galactic Classical Cepheids in the Sloan Bands,” Astrophys. J. 953
(2023) no. 1, 14, arXiv:2306.06326 [astro-ph.GA].
[92] F. Bresolin, W. Gieren, R.-P. Kudritzki, G. Pietrzynski, M. A. Urbaneja, and G. Carraro, “Extragalactic chemical
abundances: do HII regions and young stars tell the same story? The case of the spiral galaxy NGC 300,” Astrophys.
J. 700 (2009) 309–330, arXiv:0905.2791 [astro-ph.CO].
[93] F. Bresolin, R.-P. Kudritzki, and M. A. Urbaneja, “The Metallicity and Distance of NGC 2403 from Blue Supergiants,”
Astrophys. J. 940 (2022) no. 1, 32, arXiv:2209.13135 [astro-ph.GA].
[94] C.-C. Ngeow, S. M. Kanbur, S. Nikolaev, J. Buonaccorsi, K. H. Cook, and D. L. Welch, “Further empirical evidence for
the non-linearity of the period-luminosity relations as seen in the Large Magellanic Cloud Cepheids,” Mon. Not. Roy.
Astron. Soc. 363 (2005) 831–846, arXiv:astro-ph/0507601.
[95] A. Sandage, G. A. Tammann, and B. Reindl, “New period-luminosity and period-color relations of classical Cepheids:
III. Cepheids in SMC,” Astron. Astrophys. 493 (2009) 471–479, arXiv:0810.1780 [astro-ph].
[96] M. Kodric et al., “Properties of M31. II: A Cepheid disk sample derived from the first year of PS1 PAndromeda data,”
AJ 145 (2013) 106, arXiv:1301.6170 [astro-ph.CO].
[97] A. Bhardwaj, S. M. Kanbur, L. M. Macri, H. P. Singh, C.-C. Ngeow, and E. E. O. Ishida, “Large Magellanic Cloud
Near-Infrared Synoptic Survey - III. A statistical study of non-linearity in the Leavitt Laws,” MNRAS 457 (2016)
no. 2, 1644–1665, arXiv:1601.00953 [astro-ph.GA].
[98] D. Kushnir and A. Sharon, “A Cepheid systematics-free test of H0 to ≲ 2.5% accuracy using SH0ES photometry,”
arXiv:2404.16102 [astro-ph.CO].
259

[99] C. Kuo, J. A. Braatz, M. J. Reid, F. K. Y. Lo, J. J. Condon, C. M. V. Impellizzeri, and C. Henkel, “The Megamaser
Cosmology Project. V. An Angular Diameter Distance to NGC 6264 at 140 Mpc,” Astrophys. J. 767 (2013) 155,
arXiv:1207.7273 [astro-ph.CO].
[100] C. Y. Kuo, J. A. Braatz, K. Y. Lo, M. J. Reid, S. H. Suyu, D. W. Pesce, J. J. Condon, C. Henkel, and C. M. V.
Impellizzeri, “The Megamaser Cosmology Project. VI. Observations of NGC 6323,” Astrophys. J. 800 (2015) no. 1, 26,
arXiv:1411.5106 [astro-ph.GA].
[101] B. Cooke and M. Elitzur, “Water masers in late-type stars,” Astrophys. J. 295 (1985) 175–182.
[102] K. Y. Lo, “Mega-Masers and Galaxies,” Annu. Rev. Astron. Astrophys. 43 (2005) no. 1, 625–676.
[103] A. L. Argon, L. J. Greenhill, M. J. Reid, J. M. Moran, and E. M. L. Humphreys, “Toward a New Geometric Distance
to the Active Galaxy NGC4258. 1. VLBI Monitoring of Water Maser Emission,” Astrophys. J. 659 (2007) 1040–1062,
arXiv:astro-ph/0701396.
[104] J. Kormendy and L. C. Ho, “Coevolution (Or Not) of Supermassive Black Holes and Host Galaxies,” Ann. Rev.
Astron. Astrophys. 51 (2013) 511–653, arXiv:1304.7762 [astro-ph.CO].
[105] J. R. Herrnstein, J. M. Moran, L. J. Greenhill, P. J. Diamond, M. Inoue, N. Nakai, M. Miyoshi, C. Henkel, and
A. Riess, “A Geometric distance to the galaxy NGC 4258 from orbital motions in a nuclear gas disk,” Nature 400
(1999) 539–541, arXiv:astro-ph/9907013.
[106] C. Y. Kuo, J. A. Braatz, J. J. Condon, C. M. V. Impellizzeri, K. Y. Lo, I. Zaw, M. Schenker, C. Henkel, M. J. Reid, and
J. E. Greene, “The Megamaser Cosmology Project. III. Accurate Masses of Seven Supermassive Black Holes in Active
Galaxies with Circumnuclear Megamaser Disks,” Astrophys. J. 727 (2011) 20, arXiv:1008.2146 [astro-ph.CO].
[107] F. Gao, J. A. Braatz, M. J. Reid, J. J. Condon, J. E. Greene, C. Henkel, C. M. V. Impellizzeri, K. Y. Lo, C. Y. Kuo,
D. W. Pesce, J. Wagner, and W. Zhao, “The Megamaser Cosmology Project. IX. Black Hole Masses for Three Maser
Galaxies,” Astrophys. J. 834 (2017) no. 1, 52, arXiv:1610.06802 [astro-ph.GA].
[108] M. J. Reid, J. A. Braatz, J. J. Condon, K. Y. Lo, C. Y. Kuo, C. M. V. Impellizzeri, and C. Henkel, “The Megamaser
Cosmology Project: IV. A Direct Measurement of the Hubble Constant from UGC 3789,” Astrophys. J. 767 (2013)
154, arXiv:1207.7292 [astro-ph.CO].
[109] F. Gao, J. A. Braatz, M. J. Reid, K. Y. Lo, J. J. Condon, C. Henkel, C. Y. Kuo, C. M. V. Impellizzeri, D. W. Pesce,
and W. Zhao, “The Megamaser Cosmology Project VIII. A Geometric Distance to NGC 5765b,” Astrophys. J. 817
(2016) no. 2, 128, arXiv:1511.08311 [astro-ph.GA].
[110] D. W. Pesce et al., “The Megamaser Cosmology Project. XIII. Combined Hubble constant constraints,” Astrophys. J.
Lett. 891 (2020) no. 1, L1, arXiv:2001.09213 [astro-ph.CO].
[111] M. J. Reid, J. A. Braatz, J. J. Condon, L. J. Greenhill, C. Henkel, and K. Y. Lo, “The Megamaser Cosmology Project:
I. VLBI observations of UGC 3789,” Astrophys. J. 695 (2009) 287–291, arXiv:0811.4345 [astro-ph].
[112] J. A. Braatz, M. J. Reid, E. M. L. Humphreys, C. Henkel, J. J. Condon, and K. Y. Lo, “The Megamaser Cosmology
Project. II. The Angular-Diameter Distance to UGC 3789,” Astrophys. J. 718 (2010) 657–665, arXiv:1005.1955
[astro-ph.CO].
[113] C. Y. Kuo, A. Constantin, J. A. Braatz, H. H. Chung, C. A. Witherspoon, D. Pesce, C. M. V. Impellizzeri, F. Gao,
L. Hao, J. H. Woo, and I. Zaw, “Enhancing the H2 O Megamaser Detection Rate Using Optical and Mid-infrared
Photometry,” Astrophys. J. 860 (2018) no. 2, 169, arXiv:1712.04204 [astro-ph.GA].
[114] C. Y. Kuo, J. Y. Hsiang, H. H. Chung, A. Constantin, Y. Y. Chang, E. d. Cunha, D. Pesce, W. T. Chien, B. Y. Chen,
J. A. Braatz, I. Zaw, S. Matsushita, and J. C. Lin, “A More Efficient Search for H2 O Megamaser Galaxies: The Power
of X-Ray and Mid-infrared Photometry,” Astrophys. J. 892 (2020) no. 1, 18, arXiv:1911.10721 [astro-ph.GA].
[115] D. W. Pesce, J. A. Braatz, J. J. Condon, F. Gao, C. Henkel, E. Litzinger, K. Y. Lo, and M. J. Reid, “The Megamaser
Cosmology Project. VII. Investigating disk physics using spectral monitoring observations,” Astrophys. J. 810 (2015)
no. 1, 65, arXiv:1507.07904 [astro-ph.GA].
[116] A. E. Bragg, L. J. Greenhill, J. M. Moran, and C. Henkel, “Accelerations of water masers in ngc4258,” Astrophys. J.
535 (2000) 73, arXiv:astro-ph/0001543.
[117] E. M. L. Humphreys, M. J. Reid, L. J. Greenhill, J. M. Moran, and A. L. Argon, “Toward a New Distance to the
Active Galaxy NGC 4258: II. Centripetal Accelerations and Investigation of Spiral Structure,” Astrophys. J. 672
(2008) 800–816, arXiv:0709.0925 [astro-ph].
[118] J. Braatz, D. Pesce, J. Condon, and M. Reid, “H2O Megamaser Cosmology with the ngVLA,” Bulletin of the AAS 51
(2019) no. 3, 446.
[119] M. G. Lee, W. L. Freedman, and B. F. Madore, “The Tip of the Red Giant Branch as a Distance Indicator for
Resolved Galaxies,” Astrophys. J. 417 (1993) 553.
[120] J. B. Jensen, J. P. Blakeslee, M. Cantiello, M. Cowles, G. S. Anand, R. B. Tully, E. Kourkchi, and G. Raimondo, “The
TRGB-SBF Project. III. Refining the HST Surface Brightness Fluctuation Distance Scale Calibration with JWST,”
arXiv:2502.15935 [astro-ph.CO].
[121] G. S. Anand, R. B. Tully, L. Rizzi, A. G. Riess, and W. Yuan, “Comparing Tip of the Red Giant Branch Distance
Scales: An Independent Reduction of the Carnegie-Chicago Hubble Program and the Value of the Hubble Constant,”
Astrophys. J. 932 (2022) no. 1, 15, arXiv:2108.00007 [astro-ph.CO].
[122] D. Scolnic, A. G. Riess, J. Wu, S. Li, G. S. Anand, R. Beaton, S. Casertano, R. I. Anderson, S. Dhawan, and X. Ke,
“CATS: The Hubble Constant from Standardized TRGB and Type Ia Supernova Measurements,” Astrophys. J. Lett.
954 (2023) no. 1, L31, arXiv:2304.06693 [astro-ph.CO].
[123] S. Li and R. L. Beaton, “The Tip of the Red Giant Branch Distance Ladder and the Hubble Constant,”
arXiv:2403.17048 [astro-ph.CO].
260

[124] J. B. Jensen et al., “Infrared Surface Brightness Fluctuation Distances for MASSIVE and Type Ia Supernova Host
Galaxies,” Astrophys. J. Supp. 255 (2021) no. 2, 21, arXiv:2105.08299 [astro-ph.CO].
[125] J. Wu, D. Scolnic, A. G. Riess, G. S. Anand, R. Beaton, S. Casertano, X. Ke, and S. Li, “Comparative Analysis of
TRGBs (CATs) from Unsupervised, Multi-halo-field Measurements: Contrast is Key,” Astrophys. J. 954 (2023) no. 1,
87, arXiv:2211.06354 [astro-ph.CO].
[126] R. I. Anderson, N. W. Koblischke, and L. Eyer, “Small-amplitude Red Giants Elucidate the Nature of the Tip of the
Red Giant Branch as a Standard Candle,” Astrophys. J. Lett. 963 (2024) no. 2, L43, arXiv:2303.04790
[astro-ph.CO].
[127] D. Hatt et al., “The Carnegie-Chicago Hubble Program. II. The Distance to IC 1613: The Tip of the Red Giant Branch
and RR Lyrae Period–luminosity Relations,” Astrophys. J. 845 (2017) no. 2, 146, arXiv:1703.06468 [astro-ph.CO].
[128] S. Nikolaev and M. D. Weinberg, “Stellar populations in the large magellanic cloud from 2mass,” Astrophys. J. 542
(2000) 804–818, arXiv:astro-ph/0003012.
[129] B. F. Madore and W. L. Freedman, “Astrophysical Distance Scale: The AGB J-band Method. I. Calibration and a
First Application,” Astrophys. J. 899 (2020) no. 1, 66, arXiv:2005.10792 [astro-ph.GA].
[130] R. I. Anderson, “The span of space,” Nature Physics 20 (2024) no. 11, 1841–1841.
[131] J. Tang, A. Bressan, P. Rosenfield, A. Slemer, P. Marigo, L. Girardi, and L. Bianchi, “New PARSEC evolutionary
tracks of massive stars at low metallicity: testing canonical stellar evolution in nearby star-forming dwarf galaxies,”
MNRAS 445 (2014) no. 4, 4287–4305, arXiv:1410.1745 [astro-ph.SR].
[132] Y. Chen, A. Bressan, L. Girardi, P. Marigo, X. Kong, and A. Lanza, “PARSEC evolutionary tracks of massive stars up
to 350 M⊙ at metallicities 0.0001 ≤ Z ≤ 0.04,” MNRAS 452 (2015) no. 1, 1068–1080, arXiv:1506.01681
[astro-ph.SR].
[133] M. Salaris and S. Cassisi, “The tip of the red giant branch as a distance indicator: results from evolutionary models,”
Mon. Not. Roy. Astron. Soc. 289 (1997) 406, arXiv:astro-ph/9703186.
[134] M. Salaris, S. Cassisi, and A. Weiss, “Red giant branch stars: the theoretical framework,” Publ. Astron. Soc. Pac. 114
(2002) 375, arXiv:astro-ph/0201387.
[135] M. Salaris and S. Cassisi, Evolution of Stars and Stellar Populations. Wiley, 2005.
[136] S. Cassisi and M. Salaris, Old Stellar Populations: How to Study the Fossil Record of Galaxy Formation. Wiley-VCH
Verlag GmbH, 2013.
[137] A. Serenelli, A. Weiss, S. Cassisi, M. Salaris, and A. Pietrinferni, “The brightness of the red giant branch tip.
Theoretical framework, a set of reference models, and predicted observables,” A&A 606 (2017) A33,
arXiv:1706.09910 [astro-ph.SR].
[138] L. Rizzi, R. B. Tully, D. Makarov, L. Makarova, A. E. Dolphin, S. Sakai, and E. J. Shaya, “Tip of the Red Giant
Branch Distances. 2. Zero-Point Calibration,” Astrophys. J. 661 (2007) 815–829, arXiv:astro-ph/0701518.
[139] I. S. Jang and M. G. Lee, “The Tip of the Red Giant Branch Distances to Type Ia Supernova Host Galaxies. IV. Color
Dependence and Zero-point Calibration,” Astrophys. J. 835 (2017) no. 1, 28, arXiv:1611.05040 [astro-ph.GA].
[140] E. Farag, F. X. Timmes, M. T. Chidester, S. Anandagoda, and D. H. Hartmann, “Stellar Neutrino Emission across the
Mass–Metallicity Plane,” Astrophys. J. Suppl. 270 (2024) no. 1, 5, arXiv:2310.13142 [astro-ph.SR].
[141] S. Cassisi, “Modelling of Red Giant Stars: The state-of-the-art,” in European Physical Journal Web of Conferences,
vol. 160 of European Physical Journal Web of Conferences, p. 04002. 2017.
[142] I. D. Saltas and E. Tognelli, “New calibrated models for the tip of the red giant branch luminosity and a thorough
analysis of theoretical uncertainties,” Mon. Not. Roy. Astron. Soc. 514 (2022) no. 2, 3058–3073, arXiv:2203.02499
[astro-ph.SR].
[143] Gaia Collaboration, T. Prusti et al., “The Gaia Mission,” Astron. Astrophys. 595 (2016) no. Gaia Data Release 1, A1,
arXiv:1609.04153 [astro-ph.IM].
[144] W. L. Freedman, “Measurements of the Hubble Constant: Tensions in Perspective,” Astrophys. J. 919 (2021) no. 1, 16,
arXiv:2106.15656 [astro-ph.CO].
[145] B. F. Madore, W. L. Freedman, and K. Owens, “Astrophysical Distance Scale VII: A Self-Consistent,
Multi-Wavelength Calibration of the Slopes and Relative Zero Points for the Run of Luminosity with Color of Stars
Defining the Tip of the Red Giant Branch,” arXiv:2311.05048 [astro-ph.GA].
[146] G. Csörnyei, R. I. Anderson, C. Vogl, S. Taubenberger, S. Blondin, B. Leibundgut, and W. Hillebrandt, “Reeling in the
Whirlpool galaxy: Distance to M 51 clarified through Cepheids and the type IIP supernova 2005cs,” Astron.
Astrophys. 678 (2023) A44, arXiv:2305.13943 [astro-ph.GA].
[147] N. W. Koblischke and R. I. Anderson, “Calibrating and Standardizing the Tip of the Red Giant Branch in the Small
Magellanic Cloud Using Small-amplitude Red Giants,” Astrophys. J. 974 (2024) no. 2, 181, arXiv:2406.19375
[astro-ph.SR].
[148] S. Li, S. Casertano, and A. G. Riess, “A Gaia Data Release 3 View on the Tip of the Red Giant Branch Luminosity,”
Astrophys. J. 950 (2023) no. 2, 83, arXiv:2304.06695 [astro-ph.GA].
[149] J. Maíz Apellániz, M. Pantaleoni González, and R. H. Barbá, “Validation of the accuracy and precision of Gaia EDR3
parallaxes with globular clusters,” A&A 649 (2021) A13, arXiv:2101.10206 [astro-ph.IM].
[150] Y. Huang, H. Yuan, T. C. Beers, and H. Zhang, “The Parallax Zero-point of Gaia Early Data Release 3 from
LAMOST Primary Red Clump Stars,” ApJL 910 (2021) no. 1, L5, arXiv:2101.09691 [astro-ph.GA].
[151] F. Ren, X. Chen, H. Zhang, R. de Grijs, L. Deng, and Y. Huang, “Gaia EDR3 Parallax Zero-point Offset Based on W
Ursae Majoris-type Eclipsing Binaries,” ApJL 911 (2021) no. 2, L20, arXiv:2103.16096 [astro-ph.SR].
[152] S. Khan, R. I. Anderson, A. Miglio, B. Mosser, and Y. P. Elsworth, “Investigating Gaia EDR3 parallax systematics
261

using asteroseismology of cool giant stars observed by Kepler, K2, and TESS. II. Deciphering Gaia parallax
systematics using red clump stars,” A&A 680 (2023) A105, arXiv:2310.03654 [astro-ph.SR].
[153] J. Soltis, S. Casertano, and A. G. Riess, “The Parallax of ω Centauri Measured from Gaia EDR3 and a Direct,
Geometric Calibration of the Tip of the Red Giant Branch and the Hubble Constant,” Astrophys. J. Lett. 908 (2021)
no. 1, L5, arXiv:2012.09196 [astro-ph.GA].
[154] S. Jang, A. P. Milone, E. P. Lagioia, M. Tailo, M. Carlos, E. Dondoglio, M. Martorano, A. Mohandasan, A. F. Marino,
G. Cordoni, and Y. W. Lee, “Integrated Photometry of Multiple Stellar Populations in Globular Clusters,” Astrophys.
J. 920 (2021) no. 2, 129, arXiv:2107.14246 [astro-ph.GA].
[155] D. M. Skowron, J. Skowron, A. Udalski, M. K. Szymański, I. Soszyński, Ł. Wyrzykowski, K. Ulaczyk, R. Poleski,
S. Kozłowski, P. Pietrukowicz, P. Mróz, K. Rybicki, P. Iwanek, M. Wrona, and M. Gromadzki, “OGLE-ing the
Magellanic System: Optical Reddening Maps of the Large and Small Magellanic Clouds from Red Clump Stars,” ApJS
252 (2021) no. 2, 23, arXiv:2006.02448 [astro-ph.SR].
[156] D. J. Schlegel, D. P. Finkbeiner, and M. Davis, “Maps of dust IR emission for use in estimation of reddening and
CMBR foregrounds,” Astrophys. J. 500 (1998) 525, arXiv:astro-ph/9710327.
[157] E. F. Schlafly and D. P. Finkbeiner, “Measuring Reddening with SDSS Stellar Spectra and Recalibrating SFD,”
Astrophys. J. 737 (2011) 103, arXiv:1012.4804 [astro-ph.GA].
[158] R. I. Anderson, “Relativistic corrections for measuring Hubble’s constant to 1% using stellar standard candles,”
Astron. Astrophys. 658 (2022) A148, arXiv:2108.09067 [astro-ph.CO].
[159] J. E. G. Peek, B. Ménard, and L. Corrales, “Dust in the Circumgalactic Medium of Low-redshift Galaxies,” Astrophys.
J. 813 (2015) no. 1, 7, arXiv:1411.3333 [astro-ph.GA].
[160] B. Mendez, M. Davis, J. Moustakas, J. Newman, B. F. Madore, and W. L. Freedman, “Deviations from the local
Hubble flow. 1. The Tip of the red giant branch as a distance indicator,” Astron. J. 124 (2002) 213,
arXiv:astro-ph/0204192.
[161] D. Makarov, L. Makarova, L. Rizzi, R. B. Tully, A. E. Dolphin, S. Sakai, and E. J. Shaya, “Tip of the red giant branch
distances. 1. optimization of a maximum likelihood algorithm,” Astron. J. 132 (2006) 2729–2742,
arXiv:astro-ph/0603073.
[162] B. F. Madore, V. Mager, and W. L. Freedman, “Sharpening the Tip of the Red Giant Branch,” Astrophys. J. 690
(2009) 389–393, arXiv:0809.2598 [astro-ph].
[163] R. L. Beaton, G. Bono, V. F. Braga, M. Dall’Ora, G. Fiorentino, I. S. Jang, C. E. Martínez-Vázquez, N. Matsunaga,
M. Monelli, J. R. Neeley, and M. Salaris, “Old-Aged Primary Distance Indicators,” Space Sci. Rev. 214 (2018) no. 8,
113, arXiv:1808.09191 [astro-ph.GA].
[164] Euclid Collaboration, Y. Mellier et al., “Euclid. I. Overview of the Euclid mission,” arXiv:2405.13491
[astro-ph.CO].
[165] M. J. B. Newman, K. B. W. McQuinn, E. D. Skillman, M. L. Boyer, R. E. Cohen, A. E. Dolphin, and O. G. Telford,
“An Empirical Calibration of the Tip of the Red Giant Branch Distance Method in the Near Infrared. I. Hubble Space
Telescope WFC3/IR F110W and F160W Filters,” Astrophys. J. 966 (2024) no. 2, 175, arXiv:2403.03086
[astro-ph.GA].
[166] M. J. B. Newman, K. B. W. McQuinn, E. D. Skillman, M. L. Boyer, R. E. Cohen, A. E. Dolphin, and O. G. Telford,
“An Empirical Calibration of the Tip of the Red Giant Branch Distance Method in the Near Infrared. II. JWST
NIRCam Wide Filters,” Astrophys. J. 975 (2024) no. 2, 195, arXiv:2406.03532 [astro-ph.CO].
[167] E. Valenti, F. R. Ferraro, and L. Origlia, “The Red giant branch in the near - infrared color - magnitude diagrams. 1:
The Calibration of photometric indices,” Mon. Not. Roy. Astron. Soc. 351 (2004) 1204, arXiv:astro-ph/0403563.
[168] J. J. Dalcanton et al., “Resolved Near-Infrared Stellar Populations in Nearby Galaxies,” Astrophys. J. Suppl. 198
(2012) 6, arXiv:1109.6893 [astro-ph.CO].
[169] P.-F. Wu, R. B. Tully, L. Rizzi, A. E. Dolphin, B. A. Jacobs, and I. D. Karachentsev, “Infrared Tip of the Red Giant
Branch and Distances to the Maffei/IC 342 Group,” Astron. J. 148 (2014) 7, arXiv:1404.2987 [astro-ph.GA].
[170] B. F. Madore et al., “The Near-Infrared Tip of the Red Giant Branch. I. A Calibration in the Isolated Dwarf Galaxy
IC 1613,” Astrophys. J. 858 (2018) no. 1, 11, arXiv:1803.01278 [astro-ph.GA].
[171] T. J. Hoyt, W. L. Freedman, B. F. Madore, M. Seibert, R. L. Beaton, D. Hatt, I. S. Jang, M. G. Lee, A. J. Monson,
and J. A. Rich, “The Near-Infrared Tip of the Red Giant Branch. II. An Absolute Calibration in the Large Magellanic
Cloud,” Astrophys. J. 858 (2018) no. 1, 12, arXiv:1803.01277 [astro-ph.GA].
[172] M. J. Durbin, R. L. Beaton, J. J. Dalcanton, B. F. Williams, and M. L. Boyer, “MCR-TRGB: A
Multiwavelength-covariant, Robust Tip of the Red Giant Branch Measurement Method,” Astrophys. J. 898 (2020)
no. 1, 57, arXiv:2006.08559 [astro-ph.GA].
[173] K. B. W. McQuinn, M. Boyer, E. D. Skillman, and A. E. Dolphin, “Using the Tip of the Red Giant Branch As a
Distance Indicator in the Near Infrared,” Astrophys. J. 880 (2019) no. 1, 63, arXiv:1904.01571 [astro-ph.GA].
[174] W. L. Freedman, B. F. Madore, I. S. Jang, T. J. Hoyt, A. J. Lee, and K. A. Owens, “Status Report on the
Chicago-Carnegie Hubble Program (CCHP): Three Independent Astrophysical Determinations of the Hubble Constant
Using the James Webb Space Telescope,” arXiv:2408.06153 [astro-ph.CO].
[175] Gaia Collaboration, “Gaia Data Release 3. The Galaxy in your preferred colours: Synthetic photometry from Gaia
low-resolution spectra,” A&A 674 (2023) A33, arXiv:2206.06215 [astro-ph.SR].
[176] D. R. Weisz et al., “The JWST Resolved Stellar Populations Early Release Science Program. II. Survey Overview,”
ApJS 268 (2023) no. 1, 15, arXiv:2301.04659 [astro-ph.GA].
[177] A. J. Lee, L. Rousseau-Nepton, W. L. Freedman, B. F. Madore, M.-R. L. Cioni, T. J. Hoyt, I. S. Jang, A. Javadi, and
262

K. A. Owens, “The Astrophysical Distance Scale. V. A 2% Distance to the Local Group Spiral M33 via the JAGB
Method, Tip of the Red Giant Branch, and Leavitt Law,” Astrophys. J. 933 (2022) no. 2, 201, arXiv:2205.11323
[astro-ph.GA].
[178] J. Tonry and D. P. Schneider, “A New Technique for Measuring Extragalactic Distances,” AJ 96 (1988) 807.
[179] J. L. Tonry, E. A. Ajhar, and G. A. Luppino, “Observations of Surface-Brightness Fluctuations in Virgo,” AJ 100
(1990) 1416.
[180] J. P. Blakeslee, A. Jordan, S. Mei, P. Cote, L. Ferrarese, L. Infante, E. W. Peng, J. L. Tonry, and M. J. West, “The
ACS Fornax Cluster Survey. V. Measurement and Recalibration of Surface Brightness Fluctuations and a Precise Value
of the Fornax–Virgo Relative Distance,” Astrophys. J. 694 (2009) 556–572, arXiv:0901.1138 [astro-ph.CO].
[181] J. B. Jensen, J. P. Blakeslee, Z. Gibson, H.-c. Lee, M. Cantiello, G. Raimondo, N. Boyer, and H. Cho, “Measuring
Infrared Surface Brightness Fluctuation Distances with HST WFC3: Calibration and Advice,” Astrophys. J. 808
(2015) no. 1, 91, arXiv:1505.00400 [astro-ph.GA].
[182] M. Moresco et al., “Unveiling the Universe with emerging cosmological probes,” Living Rev. Rel. 25 (2022) no. 1, 6,
arXiv:2201.07241 [astro-ph.CO].
[183] M. Cantiello and J. P. Blakeslee, “Surface Brightness Fluctuations,” arXiv:2307.03116 [astro-ph.CO].
[184] J. L. Tonry, A. Dressler, J. P. Blakeslee, E. A. Ajhar, A. B. Fletcher, G. A. Luppino, M. R. Metzger, and C. B. Moore,
“The sbf survey of galaxy distances. 4. Sbf magnitudes, colors, and distances,” Astrophys. J. 546 (2001) 681–693,
arXiv:astro-ph/0011223.
[185] G. Raimondo, “Joint Analysis of near-infrared properties and surface brightness fluctuations of LMC star clusters,”
Astrophys. J. 700 (2009) 1247–1261, arXiv:0907.1408 [astro-ph.GA].
[186] M. Cantiello et al., “The Next Generation Virgo Cluster Survey (NGVS). III. A Catalog of Surface Brightness
Fluctuation Distances and the Three-dimensional Distribution of Galaxies in the Virgo Cluster,” Astrophys. J. 966
(2024) no. 1, 145, arXiv:2403.16235 [astro-ph.GA].
[187] L. Ferrarese, P. Côté, J.-C. Cuillandre, S. D. J. Gwyn, E. W. Peng, L. A. MacArthur, P.-A. Duc, A. Boselli, S. Mei,
T. Erben, A. W. McConnachie, P. R. Durrell, J. C. Mihos, A. Jordán, A. Lançon, T. H. Puzia, E. Emsellem, M. L.
Balogh, J. P. Blakeslee, L. van Waerbeke, R. Gavazzi, B. Vollmer, J. J. Kavelaars, D. Woods, N. M. Ball, S. Boissier,
S. Courteau, E. Ferriere, G. Gavazzi, H. Hildebrandt, P. Hudelot, M. Huertas-Company, C. Liu, D. McLaughlin,
Y. Mellier, M. Milkeraitis, D. Schade, C. Balkowski, F. Bournaud, R. G. Carlberg, S. C. Chapman, H. Hoekstra,
C. Peng, M. Sawicki, L. Simard, J. E. Taylor, R. B. Tully, W. van Driel, C. D. Wilson, T. Burdullis, B. Mahoney, and
N. Manset, “The Next Generation Virgo Cluster Survey (NGVS). I. Introduction to the Survey,” ApJS 200 (2012)
no. 1, 4.
[188] L. V. Sales, A. Wetzel, and A. Fattahi, “Baryonic solutions and challenges for cosmological models of dwarf galaxies,”
Nature Astron. 6 (2022) no. 8, 897–910, arXiv:2206.05295 [astro-ph.GA].
[189] J. P. Blakeslee and M. Cantiello, “Independent Analysis of the Distance to NGC 1052-DF2,” Research Notes of the
American Astronomical Society 2 (2018) no. 3, 146, arXiv:1808.02176 [astro-ph.GA].
[190] S. G. Carlsten, R. L. Beaton, J. P. Greco, and J. E. Greene, “Using Surface Brightness Fluctuations to Study Nearby
Satellite Galaxy Systems: Calibration and Methodology,” Astrophys. J. 879 (2019) no. 1, 13, arXiv:1901.07575
[astro-ph.GA].
[191] Y. J. Kim and M. G. Lee, “Calibration of Surface Brightness Fluctuations for Dwarf Galaxies in the Hyper
Suprime-Cam gi Filter System,” Astrophys. J. 923 (2021) no. 2, 152, arXiv:2110.02522 [astro-ph.GA].
[192] Planck Collaboration, N. Aghanim et al., “Planck 2018 results. VI. Cosmological parameters,” Astron. Astrophys. 641
(2020) A6, arXiv:1807.06209 [astro-ph.CO]. [Erratum: Astron.Astrophys. 652, C4 (2021)].
[193] N. Khetan et al., “A new measurement of the Hubble constant using Type Ia supernovae calibrated with surface
brightness fluctuations,” Astron. Astrophys. 647 (2021) A72, arXiv:2008.07754 [astro-ph.CO].
[194] P. Garnavich, C. M. Wood, P. Milne, J. B. Jensen, J. P. Blakeslee, P. J. Brown, D. Scolnic, B. Rose, and D. Brout,
“Connecting Infrared Surface Brightness Fluctuation Distances to Type Ia Supernova Hosts: Testing the Top Rung of
the Distance Ladder,” Astrophys. J. 953 (2023) no. 1, 35, arXiv:2204.12060 [astro-ph.CO].
[195] C. Chung, S.-J. Yoon, H. Cho, S.-Y. Lee, and Y.-W. Lee, “Yonsei Evolutionary Population Synthesis (YEPS) Model.
III. Surface Brightness Fluctuation of Normal and Helium-enhanced Simple Stellar Populations,” ApJS 250 (2020)
no. 2, 33, arXiv:2009.00625 [astro-ph.GA].
[196] I. S. Glass and T. L. Evans, “A period-luminosity relation for Mira variables in the Large Magellanic Cloud,” Nature
(London) 291 (1981) no. 5813, 303–304.
[197] M. W. Feast, I. S. Glass, P. A. Whitelock, and R. M. Catchpole, “A period-luminosity-colour relation for Mira
variables.,” MNRAS 241 (1989) 375–392.
[198] P. A. Whitelock, “Asymptotic Giant Branch Variables as Extragalactic Distance Indicators,” IAU Symp. 289 (2013)
209, arXiv:1210.7307 [astro-ph.CO].
[199] P. A. Whitelock, “Asymptotic Giant Branch Variables in Nearby Galaxies,” in Why Galaxies Care About AGB Stars:
A Continuing Challenge through Cosmic Time, F. Kerschbaum, M. Groenewegen, and H. Olofsson, eds., vol. 343 of
IAU Symposium, pp. 275–282. 2019. arXiv:1809.10077 [astro-ph.GA].
[200] C. D. Huang, “The Mira Distance Ladder,” arXiv:2401.09581 [astro-ph.CO].
[201] P. Iwanek, R. Poleski, S. Kozłowski, I. Soszyński, P. Pietrukowicz, M. Ban, J. Skowron, P. Mróz, M. Wrona,
A. Udalski, M. K. Szymański, D. M. Skowron, K. Ulaczyk, M. Gromadzki, K. Rybicki, and M. Ratajczak, “A
Three-dimensional Map of the Milky Way Using 66,000 Mira Variable Stars,” ApJS 264 (2023) no. 1, 20,
arXiv:2212.00035 [astro-ph.GA].
263

[202] M. L. Boyer, K. B. W. McQuinn, M. A. T. Groenewegen, A. A. Zijlstra, P. A. Whitelock, J. T. van Loon,


G. Sonneborn, G. C. Sloan, E. D. Skillman, M. Meixner, I. McDonald, O. C. Jones, A. Javadi, R. D. Gehrz,
N. Britavskiy, and A. Z. Bonanos, “An Infrared Census of DUST in Nearby Galaxies with Spitzer (DUSTiNGS). IV.
Discovery of High-redshift AGB Analogs,” Astrophys. J. 851 (2017) no. 2, 152, arXiv:1711.02129 [astro-ph.SR].
[203] S. R. Goldman, M. L. Boyer, K. B. W. McQuinn, P. A. Whitelock, I. McDonald, J. T. van Loon, E. D. Skillman, R. D.
Gehrz, A. Javadi, G. C. Sloan, O. C. Jones, M. A. T. Groenewegen, and J. W. Menzies, “An Infrared Census of DUST
in Nearby Galaxies with Spitzer (DUSTiNGS). V. The Period-Luminosity Relation for Dusty Metal-poor AGB Stars,”
Astrophys. J. 877 (2019) no. 1, 49, arXiv:1902.07362 [astro-ph.SR].
[204] M. Trabucchi, P. R. Wood, N. Mowlavi, G. Pastorelli, P. Marigo, L. Girardi, and T. Lebzelter, “Modelling long-period
variables - II. Fundamental mode pulsation in the non-linear regime,” MNRAS 500 (2021) no. 2, 1575–1591,
arXiv:2010.13654 [astro-ph.SR].
[205] W. Yuan, L. M. Macri, S. He, J. Z. Huang, S. M. Kanbur, and C.-C. Ngeow, “Large Magellanic Cloud Near-infrared
Synoptic Survey. V. Period-Luminosity Relations of Miras,” AJ 154 (2017) no. 4, 149, arXiv:1708.04742
[astro-ph.SR].
[206] W. Yuan, L. M. Macri, A. Javadi, Z. Lin, and J. Z. Huang, “Near-infrared Mira Period-Luminosity Relations in M33,”
AJ 156 (2018) no. 3, 112, arXiv:1807.03544 [astro-ph.SR].
[207] C. D. Huang et al., “A Near-infrared Period–Luminosity Relation for Miras in NGC 4258, an Anchor for a New
Distance Ladder,” Astrophys. J. 857 (2018) no. 1, 67, arXiv:1801.02711 [astro-ph.CO].
[208] A. Chiavassa, B. Freytag, and M. Schultheis, “Heading Gaia to measure atmospheric dynamics in AGB stars,” A&A
617 (2018) L1, arXiv:1808.02548 [astro-ph.SR].
[209] A. Chiavassa, K. Kravchenko, F. Millour, G. Schaefer, M. Schultheis, B. Freytag, O. Creevey, V. Hocdé, F. Morand,
R. Ligi, S. Kraus, J. D. Monnier, D. Mourard, N. Nardetto, N. Anugu, J. B. Le Bouquin, C. L. Davies, J. Ennis,
T. Gardner, A. Labdon, C. Lanthermann, B. R. Setterholm, and T. ten Brummelaar, “Optical interferometry and Gaia
measurement uncertainties reveal the physics of asymptotic giant branch stars,” A&A 640 (2020) A23,
arXiv:2006.07318 [astro-ph.SR].
[210] M. Andriantsaralaza, S. Ramstedt, W. H. T. Vlemmings, and E. De Beck, “Distance estimates for AGB stars from
parallax measurements,” A&A 667 (2022) A74, arXiv:2209.03906 [astro-ph.SR].
[211] A. Ahmad, B. Freytag, and S. Höfner, “Properties of self-excited pulsations in 3D simulations of AGB stars and red
supergiants,” A&A 669 (2023) A49, arXiv:2211.07682 [astro-ph.SR].
[212] T. Konchady, L. M. Macri, X. Yan, and J. Z. Huang, “The M33 synoptic stellar survey. III. Miras and LPVs in
griJHKS ,” MNRAS 531 (2024) no. 1, 110–132, arXiv:2405.00503 [astro-ph.GA].
[213] W. Yuan, Period-Luminosity Relations of Cepheid and Mira Variables and Their Application to the Extragalactic
Distance Scale. PhD thesis, Texas A&M University, 2017.
https://oaktrust.library.tamu.edu/items/26f6e9a7-5e4c-47ea-ac3e-1645a5b6ab5a/full.
[214] SNLS Collaboration, J. Guy et al., “SALT2: Using distant supernovae to improve the use of Type Ia supernovae as
distance indicators,” Astron. Astrophys. 466 (2007) 11–21, arXiv:astro-ph/0701828.
[215] W. D. Kenworthy et al., “SALT3: An Improved Type Ia Supernova Model for Measuring Cosmic Distances,”
Astrophys. J. 923 (2021) no. 2, 265, arXiv:2104.07795 [astro-ph.CO].
[216] C. R. Burns et al., “The Carnegie Supernova Project: Light Curve Fitting with SNooPy,” Astron. J. 141 (2011) 19,
arXiv:1010.4040 [astro-ph.CO].
[217] K. S. Mandel, G. Narayan, and R. P. Kirshner, “Type Ia Supernova Light Curve Inference: Hierarchical Models in the
Optical and Near Infrared,” Astrophys. J. 731 (2011) 120, arXiv:1011.5910 [astro-ph.CO].
[218] R. Tripp, “A two-parameter luminosity correction for Type IA supernovae,” A&A 331 (1998) 815–820.
[219] SNLS Collaboration, A. Conley et al., “Supernova Constraints and Systematic Uncertainties from the First 3 Years of
the Supernova Legacy Survey,” Astrophys. J. Suppl. 192 (2011) 1, arXiv:1104.1443 [astro-ph.CO].
[220] S. Dhawan, D. Brout, D. Scolnic, A. Goobar, A. G. Riess, and V. Miranda, “Cosmological Model Insensitivity of Local
H0 from the Cepheid Distance Ladder,” Astrophys. J. 894 (2020) no. 1, 54, arXiv:2001.09260 [astro-ph.CO].
[221] Y. S. Murakami, A. G. Riess, B. E. Stahl, W. D. Kenworthy, D.-M. A. Pluck, A. Macoretta, D. Brout, D. O. Jones,
D. M. Scolnic, and A. V. Filippenko, “Leveraging SN Ia spectroscopic similarity to improve the measurement of H 0 ,”
JCAP 11 (2023) 046, arXiv:2306.00070 [astro-ph.CO].
[222] E. R. Peterson et al., “The Pantheon+ Analysis: Evaluating Peculiar Velocity Corrections in Cosmological Analyses
with Nearby Type Ia Supernovae,” Astrophys. J. 938 (2022) no. 2, 112, arXiv:2110.03487 [astro-ph.CO].
[223] S. R. Brownsberger, D. Brout, D. Scolnic, C. W. Stubbs, and A. G. Riess, “Dependence of Cosmological Constraints on
Gray Photometric Zero-point Uncertainties of Supernova Surveys,” Astrophys. J. 944 (2023) no. 2, 188,
arXiv:2110.03486 [astro-ph.CO].
[224] D. O. Jones et al., “Should Type Ia Supernova Distances be Corrected for their Local Environments?,” Astrophys. J.
867 (2018) no. 2, 108, arXiv:1805.05911 [astro-ph.CO].
[225] CSP Collaboration, C. R. Burns et al., “The Carnegie Supernova Project: Absolute Calibration and the Hubble
Constant,” Astrophys. J. 869 (2018) no. 1, 56, arXiv:1809.06381 [astro-ph.CO].
[226] S. Dhawan et al., “A Uniform Type Ia Supernova Distance Ladder with the Zwicky Transient Facility: Absolute
Calibration Based on the Tip of the Red Giant Branch Method,” Astrophys. J. 934 (2022) no. 2, 185,
arXiv:2203.04241 [astro-ph.CO].
[227] S. Dhawan, S. W. Jha, and B. Leibundgut, “Measuring the Hubble constant with Type Ia supernovae as near-infrared
standard candles,” Astron. Astrophys. 609 (2018) A72, arXiv:1707.00715 [astro-ph.CO].
264

[228] D. O. Jones et al., “Cosmological Results from the RAISIN Survey: Using Type Ia Supernovae in the Near Infrared as
a Novel Path to Measure the Dark Energy Equation of State,” Astrophys. J. 933 (2022) no. 2, 172, arXiv:2201.07801
[astro-ph.CO].
[229] S. Dhawan, S. Thorp, K. S. Mandel, S. M. Ward, G. Narayan, S. W. Jha, and T. Chant, “A BayeSN distance ladder:
H0 from a consistent modelling of Type Ia supernovae from the optical to the near-infrared,” Mon. Not. Roy. Astron.
Soc. 524 (2023) no. 1, 235–244, arXiv:2211.07657 [astro-ph.CO].
[230] A. Carr, T. M. Davis, D. Scolnic, D. Scolnic, K. Said, D. Brout, E. R. Peterson, and R. Kessler, “The Pantheon+
analysis: Improving the redshifts and peculiar velocities of Type Ia supernovae used in cosmological analyses,” Publ.
Astron. Soc. Austral. 39 (2022) e046, arXiv:2112.01471 [astro-ph.CO].
[231] C. L. Steinhardt, A. Sneppen, and B. Sen, “VizieR Online Data Catalog: Revised redshifts of the Pantheon supernovae
Ia (Steinhardt+, 2020).” Vizier on-line data catalog: J/apj/902/14. originally published in: 2020apj...902...14s, 2022.
[232] B. Popovic, D. Brout, R. Kessler, and D. Scolnic, “The Pantheon+ Analysis: Forward Modeling the Dust and Intrinsic
Color Distributions of Type Ia Supernovae, and Quantifying Their Impact on Cosmological Inferences,” Astrophys. J.
945 (2023) no. 1, 84, arXiv:2112.04456 [astro-ph.CO].
[233] M. Rigault et al., “Confirmation of a Star Formation Bias in Type Ia Supernova Distances and its Effect on
Measurement of the Hubble Constant,” Astrophys. J. 802 (2015) no. 1, 20, arXiv:1412.6501 [astro-ph.CO].
[234] A. G. Riess, L. Macri, S. Casertano, H. Lampeitl, H. C. Ferguson, A. V. Filippenko, S. W. Jha, W. Li, and
R. Chornock, “A 3% Solution: Determination of the Hubble Constant with the Hubble Space Telescope and Wide Field
Camera 3,” Astrophys. J. 730 (2011) 119, arXiv:1103.2976 [astro-ph.CO]. [Erratum: Astrophys.J. 732, 129 (2011)].
[235] R. Wojtak and J. Hjorth, “Consistent extinction model for type Ia supernovae in Cepheid-based calibration galaxies
and its impact on H0,” Mon. Not. Roy. Astron. Soc. 533 (2024) no. 2, 2319–2334, arXiv:2403.10388 [astro-ph.CO].
[236] DES Collaboration, E. Macaulay et al., “First Cosmological Results using Type Ia Supernovae from the Dark Energy
Survey: Measurement of the Hubble Constant,” Mon. Not. Roy. Astron. Soc. 486 (2019) no. 2, 2184–2196,
arXiv:1811.02376 [astro-ph.CO].
[237] S. M. Feeney, H. V. Peiris, A. R. Williamson, S. M. Nissanke, D. J. Mortlock, J. Alsing, and D. Scolnic, “Prospects for
resolving the Hubble constant tension with standard sirens,” Phys. Rev. Lett. 122 (2019) no. 6, 061105,
arXiv:1802.03404 [astro-ph.CO].
[238] DES Collaboration, R. Camilleri et al., “The Dark Energy Survey Supernova Program: An updated measurement of
the Hubble constant using the Inverse Distance Ladder,” Mon. Not. Roy. Astron. Soc. 537 (2025) no. 2, 1818–1825,
arXiv:2406.05049 [astro-ph.CO].
[239] P. Ruiz-Lapuente and J. I. González Hernández, ““SNe Ia Twins for Life”: Toward a Precise Determination of H 0 ,”
Astrophys. J. 977 (2024) no. 2, 180, arXiv:2312.10334 [astro-ph.CO].
[240] H. B. Richer, “Observations of a complete sample of C stars in the Large Magellanic Cloud.,” Astrophys. J. 243 (1981)
744.
[241] H. B. Richer, D. R. Crabtree, and C. J. Pritchet, “Luminous late-type stars in NGC 205.,” Astrophys. J. 287 (1984)
138–147.
[242] H. B. Richer, C. J. Pritchet, and D. R. Crabtree, “Luminous late-type stars in NGC 300.,” Astrophys. J. 298 (1985)
240–248.
[243] C. J. Pritchet, H. B. Richer, D. Schade, D. Crabtree, and H. K. C. Yee, “The Late-Type Stellar Content of NGC 55,”
Astrophys. J. 323 (1987) 79.
[244] K. H. Cook, M. Aaronson, and J. Norris, “Carbon and M Stars in Nearby Galaxies: A Preliminary Survey Using a
Photometric Technique,” Astrophys. J. 305 (1986) 634.
[245] P. Battinelli and S. Demers, “The standard candle aspect of carbon stars,” A&A 442 (2005) no. 1, 159–163.
[246] P. Ripoche, J. Heyl, J. Parada, and H. Richer, “Carbon stars as standard candles: I. The luminosity function of carbon
stars in the Magellanic Clouds,” MNRAS 495 (2020) no. 3, 2858–2866, arXiv:2005.05539 [astro-ph.SR].
[247] W. L. Freedman and B. F. Madore, “Astrophysical Distance Scale II. Application of the JAGB Method: A Nearby
Galaxy Sample,” Astrophys. J. 899 (2020) no. 1, 67, arXiv:2005.10793 [astro-ph.GA].
[248] B. Zgirski, G. Pietrzyński, W. Gieren, M. Górski, P. Wielgórski, P. Karczmarek, F. Bresolin, P. Kervella, R.-P.
Kudritzki, J. Storm, D. Graczyk, G. Hajdu, W. Narloch, B. Pilecki, K. Suchomska, and M. Taormina, “The Araucaria
Project. Distances to Nine Galaxies Based on a Statistical Analysis of their Carbon Stars (JAGB Method),” Astrophys.
J. 916 (2021) no. 1, 19.
[249] A. J. Lee, “Carbon Stars as Standard Candles: An Empirical Test for the Reddening, Metallicity, and Age Sensitivity
of the J-region Asymptotic Giant Branch (JAGB) Method,” Astrophys. J. 956 (2023) no. 1, 15, arXiv:2305.02453
[astro-ph.GA].
[250] L. M. Macri, C.-C. Ngeow, S. M. Kanbur, S. Mahzooni, and M. T. Smitka, “Large Magellanic Cloud Near-Infrared
Synoptic Survey. I. Cepheid variables and the calibration of the Leavitt Law,” Astron. J. 149 (2015) 117,
arXiv:1412.1511 [astro-ph.SR].
[251] J. Parada, J. Heyl, H. Richer, P. Ripoche, and L. Rousseau-Nepton, “Carbon stars as standard candles – II. The
median J magnitude as a distance indicator,” Mon. Not. Roy. Astron. Soc. 501 (2021) no. 1, 933–947,
arXiv:2011.11681 [astro-ph.GA].
[252] A. J. Lee, W. L. Freedman, B. F. Madore, I. S. Jang, K. A. Owens, and T. J. Hoyt, “The Chicago-Carnegie Hubble
Program: The JWST J-region Asymptotic Giant Branch (JAGB) Extragalactic Distance Scale,” arXiv:2408.03474
[astro-ph.GA].
[253] J. Parada, J. Heyl, H. Richer, P. Ripoche, and L. Rousseau-Nepton, “Carbon stars as standard candles – III.
265

Un-binned maximum likelihood fitting and comparison with TRGB estimations,” Mon. Not. Roy. Astron. Soc. 522
(2023) no. 1, 195–210, arXiv:2303.16934 [astro-ph.GA].
[254] S. Li, A. G. Riess, D. Scolnic, S. Casertano, and G. S. Anand, “JAGB 2.0: Improved Constraints on the J-region
Asymptotic Giant Branch-based Hubble Constant from an Expanded Sample of JWST Observations,”
arXiv:2502.05259 [astro-ph.CO].
[255] S. Li, A. G. Riess, D. Scolnic, G. S. Anand, J. Wu, S. Casertano, W. Yuan, R. Beaton, and R. I. Anderson,
“Standardized Luminosity of the Tip of the Red Giant Branch Utilizing Multiple Fields in NGC 4258 and the CATs
Algorithm,” Astrophys. J. 956 (2023) no. 1, 32, arXiv:2306.10103 [astro-ph.GA].
[256] S. J. Smartt, “Progenitors of core-collapse supernovae,” Ann. Rev. Astron. Astrophys. 47 (2009) 63–106,
arXiv:0908.0700 [astro-ph.SR].
[257] C. D. Kilpatrick et al., “SN 2023ixf in Messier 101: A Variable Red Supergiant as the Progenitor Candidate to a Type
II Supernova,” Astrophys. J. Lett. 952 (2023) no. 1, L23, arXiv:2306.04722 [astro-ph.SR].
[258] R. P. Kirshner and J. Kwan, “Distances to extragalactic supernovae,” Astrophys. J. 193 (1974) 27.
[259] E. A. Baron, P. E. Nugent, D. Branch, and P. H. Hauschildt, “Type IIP supernovae as cosmological probes: A SEAM
distance to SN 1999em,” Astrophys. J. Lett. 616 (2004) L91–L94, arXiv:astro-ph/0410153.
[260] L. Dessart and D. J. Hillier, “Distance determinations using Type II supernovae and the expanding photosphere
method,” Astron. Astrophys. 439 (2005) 671, arXiv:astro-ph/0505465.
[261] M. Hamuy and P. A. Pinto, “Type II supernovae as standardized candles,” Astrophys. J. Lett. 566 (2002) L63–L65,
arXiv:astro-ph/0201279.
[262] E. E. E. Gall et al., “An updated Type II supernova Hubble diagram,” Astron. Astrophys. 611 (2018) A25,
arXiv:1705.10806 [astro-ph.CO].
[263] R. V. Wagoner, “Effects of scattering on continuous radiation from supernovae and determination of their distances,”
ApJL 250 (1981) L65–L69.
[264] R. G. Eastman, B. P. Schmidt, and R. Kirshner, “The Atmospheres of Type II Supernovae and the Expanding
Photosphere Method,” Astrophys. J. 466 (1996) 911.
[265] L. Dessart and D. J. Hillier, “Quantitative spectroscopy of photospheric-phase Type II supernovae,” Astron. Astrophys.
437 (2005) 667, arXiv:astro-ph/0504028.
[266] C. Vogl, S. A. Sim, U. M. Noebauer, W. E. Kerzendorf, and W. Hillebrandt, “Spectral modeling of type II supernovae.
I. Dilution factors,” Astron. Astrophys. 621 (2019) A29, arXiv:1811.02543 [astro-ph.HE].
[267] M. I. Jones, “Distance determination to 12 Type II Supernovae using the Expanding Photosphere Method,” Rev. Mex.
Astron. Astrof. Ser. Conf. 35 (2009) 310, arXiv:0810.5538 [astro-ph].
[268] E. E. E. Gall, R. Kotak, B. Leibundgut, S. Taubenberger, W. Hillebrandt, and M. Kromer, “Applying the expanding
photosphere and standardized candle methods to Type II-Plateau supernovae at cosmologically significant redshifts:
the distance to SN 2013eq,” Astron. Astrophys. 592 (2016) A129, arXiv:1603.04730 [astro-ph.CO].
[269] G. Dhungana et al., “Cosmological Distance Measurement of Twelve Nearby Supernovae IIP with ROTSE-IIIb,”
Astrophys. J. 962 (2024) no. 1, 60, arXiv:2308.00916 [astro-ph.HE].
[270] L. Dessart et al., “Using Quantitative Spectroscopic Analysis to Determine the Properties and Distances of Type
II-Plateau Supernovae: SNe 2005cs and 2006bp,” Astrophys. J. 675 (2008) 644, arXiv:0711.1815 [astro-ph].
[271] C. Vogl, W. E. Kerzendorf, S. A. Sim, U. M. Noebauer, S. Lietzau, and W. Hillebrandt, “Spectral modeling of type II
supernovae II. A machine learning approach to quantitative spectroscopic analysis,” Astron. Astrophys. 633 (2020)
A88, arXiv:1911.04444 [astro-ph.HE].
[272] C. Vogl et al., “No rungs attached: A distance-ladder free determination of the Hubble constant through type II
supernova spectral modelling,” arXiv:2411.04968 [astro-ph.CO].
[273] T. de Jaeger, B. E. Stahl, W. Zheng, A. V. Filippenko, A. G. Riess, and L. Galbany, “A measurement of the Hubble
constant from Type II supernovae,” Mon. Not. Roy. Astron. Soc. 496 (2020) no. 3, 3402–3411, arXiv:2006.03412
[astro-ph.CO].
[274] M. A. Hamuy, Type II supernovae as distance indicators. PhD thesis, University of Arizona, 2001.
[275] M. Hamuy, “Observed and physical properties of core-collapse supernovae,” Astrophys. J. 582 (2003) 905–914,
arXiv:astro-ph/0209174.
[276] SNLS Collaboration, P. Nugent et al., “Towards a Cosmological Hubble Diagram for Type II-P Supernovae,”
Astrophys. J. 645 (2006) 841–850, arXiv:astro-ph/0603535.
[277] T. de Jaeger and L. Galbany, “The pursuit of the Hubble Constant using Type II Supernovae,” arXiv:2305.17243
[astro-ph.CO].
[278] S. Bose and B. Kumar, “Distance Determination To Eight Galaxies Using Expanding Photosphere Method,”
Astrophys. J. 782 (2014) 98, arXiv:1401.5115 [astro-ph.CO].
[279] L. Searle and W. L. W. Sargent, “Inferences from the Composition of Two Dwarf Blue Galaxies,” Astrophys. J. 173
(1972) 25.
[280] J. Bergeron, “Characteristics of the blue stars in the dwarf galaxies I Zw 18 and II Zw 40.,” Astrophys. J. 211 (1977)
62–67.
[281] R. Terlevich and J. Melnick, “The dynamics and chemical composition of giant extragalactic H II regions,” MNRAS
195 (1981) 839–851.
[282] D. Kunth and G. Ostlin, “The most metal-poor galaxies,” Astron. Astrophys. Rev. 10 (2000) 1–79,
arXiv:astro-ph/9911094.
[283] R. Chávez, R. Terlevich, E. Terlevich, F. Bresolin, J. Melnick, M. Plionis, and S. Basilakos, “The L–σ relation for
266

massive bursts of star formation,” Mon. Not. Roy. Astron. Soc. 442 (2014) no. 4, 3565–3597, arXiv:1405.4010
[astro-ph.GA].
[284] J. Melnick, R. Terlevich, and M. Moles, “Giant H II regions as distance indicators- II. Application to H II galaxies and
the value of the Hubble constant.,” MNRAS 235 (1988) 297–313.
[285] V. Bordalo and E. Telles, “The L-sigma Relation of Local HII Galaxies,” Astrophys. J. 735 (2011) 52,
arXiv:1104.4719 [astro-ph.CO].
[286] M. Plionis, R. Terlevich, S. Basilakos, F. Bresolin, E. Terlevich, J. Melnick, and R. Chavez, “A Strategy to Measure the
Dark Energy Equation of State using the HII galaxy Hubble Relation & X-ray AGN Clustering: Preliminary Results,”
Mon. Not. Roy. Astron. Soc. 416 (2011) 2981, arXiv:1106.4558 [astro-ph.CO].
[287] R. Chávez, M. Plionis, S. Basilakos, R. Terlevich, E. Terlevich, J. Melnick, F. Bresolin, and A. L. González-Morán,
“Constraining the dark energy equation of state with HII galaxies,” Mon. Not. Roy. Astron. Soc. 462 (2016) no. 3,
2431–2439, arXiv:1607.06458 [astro-ph.CO].
[288] A. L. González-Morán, R. Chávez, R. Terlevich, E. Terlevich, F. Bresolin, D. Fernández-Arenas, M. Plionis,
S. Basilakos, J. Melnick, and E. Telles, “Independent cosmological constraints from high-z H ii galaxies,” Mon. Not.
Roy. Astron. Soc. 487 (2019) no. 4, 4669–4694, arXiv:1906.02195 [astro-ph.GA].
[289] R. Chávez, R. Terlevich, E. Terlevich, A. L. González-Morán, D. Fernández-Arenas, F. Bresolin, M. Plionis,
S. Basilakos, R. Amorín, and M. Llerena, “Mapping the Hubble flow from z~0 to z~7.5 with H II Galaxies,” MNRAS
538 (2025) no. 2, 1264–1271, arXiv:2404.16261 [astro-ph.CO].
[290] B. Dorner, G. Giardino, P. Ferruit, C. Alves de Oliveira, S. M. Birkmann, T. Böker, G. De Marchi, X. Gnata,
J. Köhler, M. Sirianni, and P. Jakobsen, “A model-based approach to the spatial and spectral calibration of NIRSpec
onboard JWST,” A&A 592 (2016) A113, arXiv:1606.05640 [astro-ph.IM].
[291] J. P. Gardner et al., “The James Webb Space Telescope,” Space Sci. Rev. 123 (2006) 485, arXiv:astro-ph/0606175.
[292] F. Feroz and M. P. Hobson, “Multimodal nested sampling: an efficient and robust alternative to MCMC methods for
astronomical data analysis,” Mon. Not. Roy. Astron. Soc. 384 (2008) 449, arXiv:0704.3704 [astro-ph].
[293] F. Feroz, M. P. Hobson, and M. Bridges, “MultiNest: an efficient and robust Bayesian inference tool for cosmology and
particle physics,” Mon. Not. Roy. Astron. Soc. 398 (2009) 1601–1614, arXiv:0809.3437 [astro-ph].
[294] F. Feroz, M. P. Hobson, E. Cameron, and A. N. Pettitt, “Importance Nested Sampling and the MultiNest Algorithm,”
Open J. Astrophys. 2 (2019) no. 1, 10, arXiv:1306.2144 [astro-ph.IM].
[295] B. Ratra and P. J. E. Peebles, “Cosmological Consequences of a Rolling Homogeneous Scalar Field,” Phys. Rev. D 37
(1988) 3406.
[296] C. Wetterich, “Cosmology and the Fate of Dilatation Symmetry,” Nucl. Phys. B 302 (1988) 668–696,
arXiv:1711.03844 [hep-th].
[297] M. Chevallier and D. Polarski, “Accelerating universes with scaling dark matter,” Int. J. Mod. Phys. D 10 (2001)
213–224, arXiv:gr-qc/0009008.
[298] E. V. Linder, “Exploring the expansion history of the universe,” Phys. Rev. Lett. 90 (2003) 091301,
arXiv:astro-ph/0208512.
[299] P. J. E. Peebles and B. Ratra, “The Cosmological Constant and Dark Energy,” Rev. Mod. Phys. 75 (2003) 559–606,
arXiv:astro-ph/0207347.
[300] D. Fernández Arenas, E. Terlevich, R. Terlevich, J. Melnick, R. Chávez, F. Bresolin, E. Telles, M. Plionis, and
S. Basilakos, “An independent determination of the local Hubble constant,” Mon. Not. Roy. Astron. Soc. 474 (2018)
no. 1, 1250–1276, arXiv:1710.05951 [astro-ph.CO].
[301] M. Llerena, R. Amorín, L. Pentericci, A. Calabrò, A. E. Shapley, K. Boutsia, E. Pérez-Montero, J. M. Vílchez, and
K. Nakajima, “Ionized gas kinematics and chemical abundances of low-mass star-forming galaxies at z ∼ 3,” A&A 676
(2023) A53, arXiv:2303.01536 [astro-ph.GA].
[302] A. de Graaff et al., “Ionised gas kinematics and dynamical masses of z ≳ 6 galaxies from JADES/NIRSpec
high-resolution spectroscopy,” A&A 684 (2024) A87, arXiv:2308.09742 [astro-ph.GA].
[303] S. S. McGaugh, J. M. Schombert, G. D. Bothun, and W. J. G. de Blok, “The Baryonic Tully-Fisher relation,”
Astrophys. J. Lett. 533 (2000) L99–L102, arXiv:astro-ph/0003001.
[304] E. F. Bell and R. S. de Jong, “Stellar mass-to-light ratios and the Tully-Fisher relation,” Astrophys. J. 550 (2001)
212–229, arXiv:astro-ph/0011493.
[305] S. Gurovich, S. S. McGaugh, K. C. Freeman, H. Jerjen, L. Staveley-Smith, and W. J. G. De Blok, “The Baryonic Tully
Fisher relation,” Publ. Astron. Soc. Austral. 21 (2004) no. 4, 412, arXiv:astro-ph/0411521.
[306] S. S. McGaugh, “The Baryonic Tully-Fisher relation of galaxies with extended rotation curves and the stellar mass of
rotating galaxies,” Astrophys. J. 632 (2005) 859–871, arXiv:astro-ph/0506750.
[307] D. Pfenniger and Y. Revaz, “The Baryonic Tully-Fisher relation revisited,” Astron. Astrophys. 431 (2005) 511,
arXiv:astro-ph/0409621.
[308] A. Begum, J. N. Chengalur, I. D. Karachentsev, and M. E. Sharina, “Baryonic Tully-Fisher Relation for Extremely
Low Mass Galaxies,” Mon. Not. Roy. Astron. Soc. 386 (2008) 138, arXiv:0801.3606 [astro-ph].
[309] C. Trachternach, W. J. G. de Blok, S. S. McGaugh, J. M. van der Hulst, and R. J. Dettmar, “The baryonic
Tully-Fisher relation and its implication for dark matter halos,” Astron. Astrophys. 505 (2009) 577, arXiv:0907.5533
[astro-ph.CO].
[310] D. V. Stark, S. S. McGaugh, and R. A. Swaters, “A First Attempt to Calibrate the Baryonic Tully-Fisher Relation
with Gas Dominated Galaxies,” Astron. J. 138 (2009) 392, arXiv:0905.4528 [astro-ph.CO].
[311] S. Gurovich, K. Freeman, H. Jerjen, L. Staveley-Smith, and I. Puerari, “The slope of the Baryonic Tully-Fisher
267

relation,” Astron. J. 140 (2010) 663, arXiv:1004.4365 [astro-ph.CO].


[312] D. Zaritsky et al., “The Baryonic Tully-Fisher Relationship for S4 G Galaxies and the ”Condensed” Baryon Fraction of
Galaxies,” Astron. J. 147 (2014) 134, arXiv:1402.6315 [astro-ph.GA].
[313] R. B. Tully and J. R. Fisher, “A New method of determining distances to galaxies,” Astron. Astrophys. 54 (1977)
661–673.
[314] M. J. Pierce and R. B. Tully, “Distances to the Virgo and Ursa Major Clusters and a Determination of H 0,”
Astrophys. J. 330 (1988) 579.
[315] P. Teerikorpi, “The inverse Tully-Fisher relation,” Astrophysical Letters and Communications 31 (1995) 263.
[316] R. Giovanelli, M. Haynes, T. Herter, N. Vogt, L. d. Costa, W. Freudling, J. Salzer, and G. Wegner, “The i-band
tully-fisher relation for cluster galaxies: data presentation,” Astron. J. 113 (1997) 22–52, arXiv:astro-ph/9610117.
[317] R. Giovanelli, M. Haynes, T. Herter, N. Vogt, L. d. Costa, W. Freudling, J. Salzer, and G. Wegner, “The i-band
tully-fisher relation for cluster galaxies: a template relation, its scatter and bias corrections,” Astron. J. 113 (1997)
53–79, arXiv:astro-ph/9610118.
[318] S. Courteau and H.-W. Rix, “Maximal disks and the tully-fisher relation,” ASP Conf. Ser. 136 (1998) 196,
arXiv:astro-ph/9707290.
[319] T. R. Brent and M. J. Pierce, “Distances to galaxies from the correlation between luminosities and linewidths. 3.
Cluster template and global measurements of h0,” Astrophys. J. 533 (2000) 744–780, arXiv:astro-ph/9911052.
[320] I. D. Karachentsev, S. N. Mitronova, V. E. Karachentseva, Y. N. Kudrya, and T. H. Jarrett, “The 2MASS Tully-Fisher
relation for flat edge - on galaxies,” Astron. Astrophys. 396 (2002) 431–438, arXiv:astro-ph/0209189.
[321] A. G. Bedregal, A. Aragon-Salamanca, and M. R. Merrifield, “The Tully-Fisher relation for S0 galaxies,” Mon. Not.
Roy. Astron. Soc. 373 (2006) 1125–1140, arXiv:astro-ph/0609076.
[322] E. Noordermeer and M. A. W. Verheijen, “The high mass end of the Tully-Fisher relation,” Mon. Not. Roy. Astron.
Soc. 381 (2007) 1463, arXiv:0708.2822 [astro-ph].
[323] C. M. Springob, K. L. Masters, M. P. Haynes, R. Giovanelli, and C. Marinoni, “SFI++ II: A New I-band Tully-Fisher
Catalog, Derivation of Peculiar Velocities and Dataset Properties,” Astrophys. J. Suppl. 172 (2007) 599–614,
arXiv:0705.0647 [astro-ph]. [Erratum: Astrophys.J.Suppl. 182, 474–475 (2009)].
[324] M. J. Williams, M. Bureau, and M. Cappellari, “The Tully-Fisher relations of early-type spiral and S0 galaxies,” Mon.
Not. Roy. Astron. Soc. 409 (2010) 1330, arXiv:1007.4072 [astro-ph.GA].
[325] P. Mocz, A. Green, M. Malacari, and K. Glazebrook, “The Tully-Fisher Relation for 25,000 SDSS Galaxies as Function
of Environment,” Mon. Not. Roy. Astron. Soc. 425 (2012) 296, arXiv:1206.1662 [astro-ph.CO].
[326] R. B. Tully and H. M. Courtois, “Cosmicflows-2: I-band Luminosity - HI Linewidth Calibration,” Astrophys. J. 749
(2012) 78, arXiv:1202.3191 [astro-ph.CO].
[327] T. D. Rawle, J. R. Lucey, R. J. Smith, and J. T. C. G. Head, “S0 galaxies in the Coma cluster: Environmental
dependence of the S0 offset from the Tully-Fisher relation,” Mon. Not. Roy. Astron. Soc. 433 (2013) 2667,
arXiv:1305.6929 [astro-ph.CO].
[328] J. G. Sorce, H. M. Courtois, R. B. Tully, M. Seibert, V. Scowcroft, W. L. Freedman, B. F. Madore, S. E. Persson,
A. Monson, and J. Rigby, “Calibration of the Mid-Infrared Tully-Fisher Relation,” Astrophys. J. 765 (2013) 94,
arXiv:1301.4833 [astro-ph.CO].
[329] J. G. Sorce, R. B. Tully, H. M. Courtois, T. H. Jarrett, J. D. Neill, and E. J. Shaya, “From Spitzer Galaxy photometry
to Tully–Fisher distances,” Mon. Not. Roy. Astron. Soc. 444 (2014) no. 1, 527–541, arXiv:1408.0729 [astro-ph.GA].
[330] S. Torres-Flores, C. Mendes de Oliveira, H. Plana, P. Amram, and B. Epinat, “The Tully-Fisher relations for Hickson
Compact Group galaxies,” Mon. Not. Roy. Astron. Soc. 432 (2013) 3085, arXiv:1304.4493 [astro-ph.CO].
[331] K. Said, R. C. Kraan-Korteweg, and T. H. Jarrett, “On how to extend the NIR Tully–Fisher relation to be truly
all-sky,” Mon. Not. Roy. Astron. Soc. 447 (2015) no. 2, 1618–1629, arXiv:1411.7361 [astro-ph.GA].
[332] K. Said, R. C. Kraan-Korteweg, T. H. Jarrett, L. Staveley-Smith, and W. L. Williams, “NIR Tully–Fisher in the Zone
of Avoidance – III. Deep NIR catalogue of the HIZOA galaxies,” Mon. Not. Roy. Astron. Soc. 462 (2016) no. 3,
3386–3400, arXiv:1607.08596 [astro-ph.GA].
[333] E. Kourkchi, R. B. Tully, G. S. Anand, H. M. Courtois, A. Dupuy, J. D. Neill, L. Rizzi, and M. Seibert,
“Cosmicflows-4: The Calibration of Optical and Infrared Tully–Fisher Relations,” Astrophys. J. 896 (2020) no. 1, 3,
arXiv:2004.14499 [astro-ph.GA].
[334] R. Bell, K. Said, T. Davis, and T. H. Jarrett, “Calibration of the Tully–Fisher relation in the WISE W1 (3.4 µm) and
W2 (4.6 µm) bands,” Mon. Not. Roy. Astron. Soc. 519 (2022) no. 1, 102–120, arXiv:2205.13136 [astro-ph.GA].
[335] P. Boubel, M. Colless, K. Said, and L. Staveley-Smith, “Large-scale motions and growth rate from forward-modelling
Tully–Fisher peculiar velocities,” Mon. Not. Roy. Astron. Soc. 531 (2024) no. 1, 84–109, arXiv:2301.12648
[astro-ph.CO].
[336] D. J. Eisenstein and A. Loeb, “An Analytical model for the triaxial collapse of cosmological perturbations,” Astrophys.
J. 439 (1995) 520, arXiv:astro-ph/9405012.
[337] S. S. McGaugh and W. J. G. de Blok, “Testing the dark matter hypothesis with low surface brightness galaxies and
other evidence,” Astrophys. J. 499 (1998) 41, arXiv:astro-ph/9801123.
[338] S. S. McGaugh and W. J. G. de Blok, “Testing the hypothesis of modified dynamics with low surface brightness
galaxies and other evidence,” Astrophys. J. 499 (1998) 66–81, arXiv:astro-ph/9801102.
[339] H. J. Mo, S. Mao, and S. D. M. White, “The Formation of galactic disks,” Mon. Not. Roy. Astron. Soc. 295 (1998)
319, arXiv:astro-ph/9707093.
[340] M. Steinmetz and J. Navarro, “The cosmological origin of the tully-fisher relation,” Astrophys. J. 513 (1999) 555–560,
268

arXiv:astro-ph/9808076.
[341] F. C. van den Bosch, “Semi-analytical models for the formation of disk galaxies: I. constraints from the tully-fisher
relation,” Astrophys. J. 530 (2000) 177, arXiv:astro-ph/9909298.
[342] L. Mayer and B. Moore, “The baryonic mass - velocity relation: Clues to feedback processes during structure formation
and the cosmic baryon inventory,” Mon. Not. Roy. Astron. Soc. 354 (2004) 477, arXiv:astro-ph/0309500.
[343] O. Y. Gnedin, D. H. Weinberg, J. Pizagno, F. Prada, and H.-W. Rix, “Dark Matter Halos of Disk Galaxies:
Constraints from the Tully-Fisher Relation,” Astrophys. J. 671 (2007) 1115–1134, arXiv:astro-ph/0607394.
[344] F. Governato, B. Willman, L. Mayer, A. Brooks, G. Stinson, O. Valenzuela, J. Wadsley, and T. R. Quinn, “Forming
disk galaxies in lambda-cdm simulations,” Mon. Not. Roy. Astron. Soc. 374 (2007) 1479–1494,
arXiv:astro-ph/0602351.
[345] V. Avila-Reese, J. Zavala, C. Firmani, and H. M. Hernandez-Toledo, “On the baryonic, stellar, and luminous scaling
relations of disk galaxies,” Astron. J. 136 (2008) 1340–1360, arXiv:0807.0636 [astro-ph].
[346] A. A. Dutton, “The baryonic Tully-Fisher relation and galactic outflows,” Mon. Not. Roy. Astron. Soc. 424 (2012)
3123, arXiv:1206.1855 [astro-ph.CO].
[347] S. McGaugh, “The Baryonic Tully-Fisher Relation of Gas Rich Galaxies as a Test of LCDM and MOND,” Astron. J.
143 (2012) 40, arXiv:1107.2934 [astro-ph.CO].
[348] M. Aumer, S. White, T. Naab, and C. Scannapieco, “Towards a more realistic population of bright spiral galaxies in
cosmological simulations,” Mon. Not. Roy. Astron. Soc. 434 (2013) 3142, arXiv:1304.1559 [astro-ph.GA].
[349] H. Desmond and R. H. Wechsler, “The Tully–Fisher and mass–size relations from halo abundance matching,” Mon.
Not. Roy. Astron. Soc. 454 (2015) no. 1, 322–343, arXiv:1506.00169 [astro-ph.GA].
[350] F. Lelli et al., “Gas dynamics in tidal dwarf galaxies: Disc formation at z = 0,” Astron. Astrophys. 584 (2015) A113,
arXiv:1509.05404 [astro-ph.GA].
[351] P. Salucci, C. S. Frenk, and M. Persic, “A physical distance indicator for spiral galaxies and the determination of
H(0),” Mon. Not. Roy. Astron. Soc. 262 (1993) 392.
[352] A. Lapi, P. Salucci, and L. Danese, “Precision Scaling Relations for Disk Galaxies in the Local Universe,” Astrophys. J.
859 (2018) no. 1, 2, arXiv:1804.06086 [astro-ph.GA].
[353] K. Said, “Tully-Fisher relation,” arXiv:2310.16053 [astro-ph.CO].
[354] S. S. McGaugh, “A Novel Test of the Modified Newtonian Dynamics with Gas Rich Galaxies,” Phys. Rev. Lett. 106
(2011) 121303, arXiv:1102.3913 [astro-ph.CO]. [Erratum: Phys.Rev.Lett. 107, 229901 (2011)].
[355] I. A. Yegorova and P. Salucci, “The Radial Tully-Fisher relation for spiral galaxies. 1.,” Mon. Not. Roy. Astron. Soc.
377 (2007) 507–515, arXiv:astro-ph/0612434.
[356] B. S. Haridasu, P. Salucci, and G. Sharma, “Radial Tully–Fisher relation and the local variance of Hubble parameter,”
Mon. Not. Roy. Astron. Soc. 532 (2024) no. 2, 2234–2247, arXiv:2403.06859 [astro-ph.CO].
[357] J.-P. Fontaine, P. Salucci, and E. Karukes, “The radial tully-fisher relations in dwarf spiral galaxies,” in Proceedings of
the 53rd Rencontres de Moriond on Cosmology 2018, pp. 363 – 364. 2020. https://www.scopus.com/inward/record.
uri?eid=2-s2.0-85089624578&partnerID=40&md5=e364fc57db93bf5425ab83b4aa5fadbd.
[358] P. Salucci, “The distribution of dark matter in galaxies,” Astron. Astrophys. Rev. 27 (2019) no. 1, 2,
arXiv:1811.08843 [astro-ph.GA].
[359] F. Lelli, S. S. McGaugh, and J. M. Schombert, “The Small Scatter of the Baryonic Tully–fisher Relation,” Astrophys.
J. Lett. 816 (2016) no. 1, L14, arXiv:1512.04543 [astro-ph.GA].
[360] V. Borka Jovanović, S. Capozziello, P. Jovanović, and D. Borka, “Recovering the fundamental plane of galaxies by
f (R) gravity,” Phys. Dark Univ. 14 (2016) 73–83, arXiv:1610.03336 [astro-ph.GA].
[361] S. Capozziello, P. Jovanović, V. B. Jovanović, and D. Borka, “Addressing the missing matter problem in galaxies
through a new fundamental gravitational radius,” JCAP 06 (2017) 044, arXiv:1702.03430 [gr-qc].
[362] S. Capozziello, D. Borka, V. Borka Jovanović, and P. Jovanović, “Galactic Structures from Gravitational Radii,”
Galaxies 6 (2018) no. 1, 22.
[363] S. Capozziello, V. B. Jovanović, D. Borka, and P. Jovanović, “Constraining theories of gravity by fundamental plane of
elliptical galaxies,” Phys.Dark Univ. 29 (2020) 100573, arXiv:2004.11557 [gr-qc].
[364] M. Milgrom, “A Modification of the Newtonian dynamics as a possible alternative to the hidden mass hypothesis,”
Astrophys. J. 270 (1983) 365–370.
[365] B. Famaey and S. McGaugh, “Modified Newtonian Dynamics (MOND): Observational Phenomenology and Relativistic
Extensions,” Living Rev. Rel. 15 (2012) 10, arXiv:1112.3960 [astro-ph.CO].
[366] I. Banik and H. Zhao, “From Galactic Bars to the Hubble Tension: Weighing Up the Astrophysical Evidence for
Milgromian Gravity,” Symmetry 14 (2022) no. 7, 1331, arXiv:2110.06936 [astro-ph.CO].
[367] B. Famaey and A. Durakovic, “Modified Newtonian Dynamics (MOND),” arXiv:2501.17006 [astro-ph.GA].
[368] O. Bertolami, C. G. Boehmer, T. Harko, and F. S. N. Lobo, “Extra force in f(R) modified theories of gravity,” Phys.
Rev. D 75 (2007) 104016, arXiv:0704.1733 [gr-qc].
[369] O. Bertolami and C. Gomes, “The Layzer-Irvine equation in theories with non-minimal coupling between matter and
curvature,” JCAP 09 (2014) 010, arXiv:1406.5990 [astro-ph.CO].
[370] C. Gomes, “Jeans instability in non-minimal matter-curvature coupling gravity,” Eur. Phys. J. C 80 (2020) no. 7, 633,
arXiv:2008.10026 [gr-qc].
[371] C. Gomes and K. Ourabah, “Quantum kinetic theory of Jeans instability in non-minimal matter-curvature coupling
gravity,” Eur. Phys. J. C 83 (2023) no. 1, 40, arXiv:2204.07871 [gr-qc].
[372] M. Barroso Varela and O. Bertolami, “Hubble tension in a nonminimally coupled curvature-matter gravity model,”
269

JCAP 06 (2024) 025, arXiv:2403.11683 [gr-qc].


[373] M. H. P. M. van Putten, “Evidence for Galaxy Dynamics Tracing Background Cosmology Below the de Sitter Scale of
Acceleration,” Astrophys. J. 848 (2017) no. 1, 28, arXiv:1709.05944 [astro-ph.GA].
[374] M. H. P. M. van Putten, “Self-similar galaxy dynamics below the de Sitter scale of acceleration,” MNRAS 481 (2018)
no. 1, L26–L29, arXiv:1804.06212 [astro-ph.GA].
[375] G.-M. Lee and M. H. P. M. van Putten, “Prospects for high-resolution probes of galaxy dynamics tracing background
cosmology in MaNGA,” New Astron. 117 (2025) 102360, arXiv:2501.11882 [astro-ph.CO].
[376] M. H. P. M. van Putten, “Galaxy dynamics tracing quantum cosmology beyond <mml:math altimg=“si13.svg”
display=“inline” id=“d1e23”><mml:mi>Λ</mml:mi></mml:math>CDM below the de Sitter scale of acceleration,”
Chinese Journal of Physics 91 (2024) 377–381, arXiv:2408.06399 [gr-qc].
[377] M. H. P. M. van Putten, “The Fast and Furious in JWST high- z galaxies,” Physics of the Dark Universe 43 (2024)
101417, arXiv:2312.16692 [astro-ph.CO].
[378] F. Lelli, S. S. McGaugh, J. M. Schombert, and M. S. Pawlowski, “The Relation between Stellar and Dynamical Surface
Densities in the Central Regions of Disk Galaxies,” Astrophys. J. Lett. 827 (2016) no. 1, L19, arXiv:1607.02145
[astro-ph.GA].
[379] F. Lelli, S. S. McGaugh, and J. M. Schombert, “SPARC: Mass Models for 175 Disk Galaxies with Spitzer Photometry
and Accurate Rotation Curves,” Astron. J. 152 (2016) 157, arXiv:1606.09251 [astro-ph.GA].
[380] J. Schombert, S. McGaugh, and F. Lelli, “Using the Baryonic Tully–Fisher Relation to Measure H o,” Astron. J. 160
(2020) no. 2, 71, arXiv:2006.08615 [astro-ph.CO].
[381] M. Puech, F. Hammer, H. Flores, R. Delgado-Serrano, M. Rodrigues, and Y. Yang, “The baryonic content and
Tully-Fisher relation at z~0.6,” Astron. Astrophys. 510 (2010) A68, arXiv:0903.3961 [astro-ph.CO].
[382] S. H. Miller, K. Bundy, M. Sullivan, R. S. Ellis, and T. Treu, “The Assembly History of Disk Galaxies: I - The
Tully-Fisher Relation to z~1.3 from Deep Exposures with DEIMOS,” Astrophys. J. 741 (2011) 115, arXiv:1102.3911
[astro-ph.CO].
[383] E. M. Di Teodoro, F. Fraternali, and S. H. Miller, “Flat rotation curves and low velocity dispersions in KMOS
star-forming galaxies at z ~1,” A&A 594 (2016) A77, arXiv:1602.04942 [astro-ph.GA].
[384] I. Shivaei et al., “The MOSDEF Survey: Dynamical and Baryonic Masses and Kinematic Structures of Star-Forming
Galaxies at 1.4 ≤ z ≤ 2.6,” Astrophys. J. 819 (2016) 80, arXiv:1511.03272 [astro-ph.GA].
[385] C. M. S. Straatman et al., “ZFIRE: The Evolution of the Stellar Mass Tully-Fisher Relation to Redshift 2.0 < Z < 2.5
with MOSFIRE,” Astrophys. J. 839 (2017) 57, arXiv:1703.00016 [astro-ph.GA].
[386] G. Alestas, I. Antoniou, and L. Perivolaropoulos, “Hints for a Gravitational Transition in Tully–Fisher Data,” Universe
7 (2021) no. 10, 366, arXiv:2104.14481 [astro-ph.CO].
[387] C. Stone, S. Courteau, N. Arora, M. Frosst, and T. Jarrett, “PROBES-I: A Compendium of Deep Rotation Curves and
Matched multiband Photometry,” Astrophys. J. Suppl. 262 (2022) 33, arXiv:2209.09912 [astro-ph.GA].
[388] G. Sharma, V. Upadhyaya, P. Salucci, and S. Desai, “Tully-Fisher Relation of Late-type Galaxies at 0.6 ≤ z ≤ 2.5,”
Astron. Astrophys. 689 (2024) A318, arXiv:2406.08934 [astro-ph.GA].
[389] P. Boubel, M. Colless, K. Said, and L. Staveley-Smith, “An improved Tully–Fisher estimate of H0,” Mon. Not. Roy.
Astron. Soc. 533 (2024) no. 2, 1550–1559, arXiv:2408.03660 [astro-ph.CO].
[390] D. Scolnic, P. Boubel, J. Byrne, A. G. Riess, and G. S. Anand, “Calibrating the Tully-Fisher Relation to Measure the
Hubble Constant,” arXiv:2412.08449 [astro-ph.CO].
[391] B. S. Koribalski et al., “WALLABY – an SKA Pathfinder H i survey,” Astrophys. Space Sci. 365 (2020) no. 7, 118,
arXiv:2002.07311 [astro-ph.GA].
[392] C. Saulder et al., “Target selection for the DESI Peculiar Velocity Survey,” Mon. Not. Roy. Astron. Soc. 525 (2023)
no. 1, 1106–1125, arXiv:2302.13760 [astro-ph.CO].
[393] C.-P. Zhang, M. Zhu, P. Jiang, C. Cheng, J. Wang, J. Wang, J.-L. Xu, X.-L. Liu, N.-P. Yu, L. Qian, H. Yu, M. Ai,
Y. Jing, C. Xu, Z. Liu, X. Guan, C. Sun, Q. Yang, M. Huang, Q. Hao, and FAST Collaboration, “The FAST all sky H
I survey (FASHI): The first release of catalog,” Science China Physics, Mechanics, and Astronomy 67 (2024) no. 1,
219511, arXiv:2312.06097 [astro-ph.GA].
[394] K. Said et al., “DESI Peculiar Velocity Survey – Fundamental Plane,” arXiv:2408.13842 [astro-ph.CO].
[395] S. Djorgovski and M. Davis, “Fundamental properties of elliptical galaxies,” Astrophys. J. 313 (1987) 59.
[396] A. Dressler, S. M. Faber, D. Burstein, R. L. Davies, D. Lynden-Bell, R. J. Terlevich, and G. Wegner, “Spectroscopy
and photometry of elliptical galaxies: A Large scale streaming motion in the local universe,” Astrophys. J. Lett. 313
(1987) L37–L42.
[397] D. Scolnic et al., “The Hubble Tension in Our Own Backyard: DESI and the Nearness of the Coma Cluster,”
Astrophys. J. Lett. 979 (2025) no. 1, L9, arXiv:2409.14546 [astro-ph.CO].
[398] A. Rest, H. J. Weiland, C. W. Stubbs, S. J. Smartt, K. W. Smith, B. Stalder, A. N. Heinze, L. Denneau, and J. L.
Tonry, “ATLAS: A High-cadence All-sky Survey System,” Publ. Astron. Soc. Pac. 130 (2018) no. 988, 064505.
[399] Young Supernova Experiment Collaboration, D. O. Jones et al., “The Young Supernova Experiment: Survey
Goals, Overview, and Operations,” Astrophys. J. 908 (2021) no. 2, 143, arXiv:2010.09724 [astro-ph.HE].
[400] D. Carter et al., “The HST/ACS Coma Cluster Survey: I - Survey Objectives and Design,” Astrophys. J. Suppl. 176
(2008) 424, arXiv:0801.3745 [astro-ph].
[401] J. B. Jensen, J. L. Tonry, and G. A. Luppino, “The infrared surface brightness fluctuation distances to the hydra and
coma clusters,” Astrophys. J. 510 (1999) 71, arXiv:astro-ph/9807326.
[402] B. Thomsen, W. A. Baum, M. Hammergren, and G. Worthey, “The distance to the coma cluster from surface
270

brightness fluctuations,” Astrophys. J. Lett. 483 (1997) L37–L40, arXiv:astro-ph/9704121.


[403] M. D. Gregg, “The coma - leo I distance ratio and the hubble constant,” New Astron. 1 (1997) 363,
arXiv:astro-ph/9703031.
[404] J. Hjorth and N. R. Tanvir, “Calibration of the fundamental plane zero-point in the Leo-I group and an estimate of the
Hubble constant,” Astrophys. J. 482 (1997) 68–74, arXiv:astro-ph/9701025.
[405] J. J. Kavelaars, W. E. Harris, D. A. Hanes, J. E. Hesser, and C. J. Pritchet, “The globular cluster systems in the coma
ellipticals. I: the luminosity function in ngc 4874, and implications for hubble’s constant,” Astrophys. J. 533 (2000)
125, arXiv:astro-ph/9911206.
[406] R. de Grijs and G. Bono, “Clustering of Local Group distances: publication bias or correlated measurements? VII. A
distance framework out to 100 Mpc,” Astrophys. J. Suppl. 248 (2020) no. 1, 6, arXiv:2004.00114 [astro-ph.GA].
[407] WST Collaboration, V. Mainieri et al., “The Wide-field Spectroscopic Telescope (WST) Science White Paper,”
arXiv:2403.05398 [astro-ph.IM].
[408] R. Jimenez and A. Loeb, “Constraining cosmological parameters based on relative galaxy ages,” Astrophys. J. 573
(2002) 37–42, arXiv:astro-ph/0106145.
[409] J. Simon, L. Verde, and R. Jimenez, “Constraints on the redshift dependence of the dark energy potential,” Phys. Rev.
D 71 (2005) 123001, arXiv:astro-ph/0412269.
[410] D. Stern, R. Jimenez, L. Verde, M. Kamionkowski, and S. A. Stanford, “Cosmic Chronometers: Constraining the
Equation of State of Dark Energy. I: H(z) Measurements,” JCAP 02 (2010) 008, arXiv:0907.3149 [astro-ph.CO].
[411] M. Moresco, “Raising the bar: new constraints on the Hubble parameter with cosmic chronometers at z ∼ 2,” Mon.
Not. Roy. Astron. Soc. 450 (2015) no. 1, L16–L20, arXiv:1503.01116 [astro-ph.CO].
[412] M. Moresco, L. Pozzetti, A. Cimatti, R. Jimenez, C. Maraston, L. Verde, D. Thomas, A. Citro, R. Tojeiro, and
D. Wilkinson, “A 6% measurement of the Hubble parameter at z ∼ 0.45: direct evidence of the epoch of cosmic
re-acceleration,” JCAP 05 (2016) 014, arXiv:1601.01701 [astro-ph.CO].
[413] A. L. Ratsimbazafy, S. I. Loubser, S. M. Crawford, C. M. Cress, B. A. Bassett, R. C. Nichol, and P. Väisänen,
“Age-dating Luminous Red Galaxies observed with the Southern African Large Telescope,” Mon. Not. Roy. Astron.
Soc. 467 (2017) no. 3, 3239–3254, arXiv:1702.00418 [astro-ph.CO].
[414] N. Borghi, M. Moresco, A. Cimatti, A. Huchet, S. Quai, and L. Pozzetti, “Toward a Better Understanding of Cosmic
Chronometers: Stellar Population Properties of Passive Galaxies at Intermediate Redshift,” Astrophys. J. 927 (2022)
no. 2, 164, arXiv:2106.14894 [astro-ph.GA].
[415] K. Jiao, N. Borghi, M. Moresco, and T.-J. Zhang, “New Observational H(z) Data from Full-spectrum Fitting of Cosmic
Chronometers in the LEGA-C Survey,” Astrophys. J. Suppl. 265 (2023) no. 2, 48, arXiv:2205.05701 [astro-ph.CO].
[416] E. Tomasetti, M. Moresco, N. Borghi, K. Jiao, A. Cimatti, L. Pozzetti, A. C. Carnall, R. J. McLure, and L. Pentericci,
“A new measurement of the expansion history of the Universe at z = 1.26 with cosmic chronometers in VANDELS,”
Astron. Astrophys. 679 (2023) A96, arXiv:2305.16387 [astro-ph.CO].
[417] R. Jimenez, M. Moresco, L. Verde, and B. D. Wandelt, “Cosmic chronometers with photometry: a new path to H(z),”
JCAP 11 (2023) 047, arXiv:2306.11425 [astro-ph.CO].
[418] M. Moresco, “Addressing the Hubble tension with cosmic chronometers,” arXiv:2307.09501 [astro-ph.CO].
[419] V. C. Busti, C. Clarkson, and M. Seikel, “Evidence for a Lower Value for H0 from Cosmic Chronometers Data?,” Mon.
Not. Roy. Astron. Soc. 441 (2014) 11, arXiv:1402.5429 [astro-ph.CO].
[420] S. Cao, X. Zheng, M. Biesiada, J. Qi, Y. Chen, and Z.-H. Zhu, “Ultra-compact structure in intermediate-luminosity
radio quasars: building a sample of standard cosmological rulers and improving the dark energy constraints up to z ~
3,” Astron. Astrophys. 606 (2017) A15, arXiv:1708.08635 [astro-ph.CO].
[421] H. Yu, B. Ratra, and F.-Y. Wang, “Hubble Parameter and Baryon Acoustic Oscillation Measurement Constraints on
the Hubble Constant, the Deviation from the Spatially Flat ΛCDM Model, the Deceleration–Acceleration Transition
Redshift, and Spatial Curvature,” Astrophys. J. 856 (2018) no. 1, 3, arXiv:1711.03437 [astro-ph.CO].
[422] A. Gómez-Valent and L. Amendola, “H0 from cosmic chronometers and Type Ia supernovae, with Gaussian Processes
and the novel Weighted Polynomial Regression method,” JCAP 04 (2018) 051, arXiv:1802.01505 [astro-ph.CO].
[423] B. S. Haridasu, V. V. Luković, M. Moresco, and N. Vittorio, “An improved model-independent assessment of the
late-time cosmic expansion,” JCAP 10 (2018) 015, arXiv:1805.03595 [astro-ph.CO].
[424] A. Gómez-Valent, “Measuring the sound horizon and absolute magnitude of SNIa by maximizing the consistency
between low-redshift data sets,” Phys. Rev. D 105 (2022) no. 4, 043528, arXiv:2111.15450 [astro-ph.CO].
[425] A. Bonilla, S. Kumar, and R. C. Nunes, “Measurements of H0 and reconstruction of the dark energy properties from a
model-independent joint analysis,” Eur. Phys. J. C 81 (2021) no. 2, 127, arXiv:2011.07140 [astro-ph.CO].
[426] E. Ó Colgáin and M. M. Sheikh-Jabbari, “Elucidating cosmological model dependence with H0 ,” Eur. Phys. J. C 81
(2021) no. 10, 892, arXiv:2101.08565 [astro-ph.CO].
[427] T. Liu, S. Cao, S. Zhang, C. Zheng, and W. Guo, “Revisiting the Hubble Constant, Spatial Curvature, and
Cosmography with Strongly Lensed Quasar and Hubble Parameter Observations,” Astrophys. J. 939 (2022) no. 1, 37,
arXiv:2204.07365 [astro-ph.CO].
[428] Y. Yang, X. Lu, L. Qian, and S. Cao, “Potentialities of Hubble parameter and expansion rate function data to alleviate
Hubble tension,” Mon. Not. Roy. Astron. Soc. 519 (2023) no. 4, 4938–4950, arXiv:2204.01020 [astro-ph.CO].
[429] A. Favale, A. Gómez-Valent, and M. Migliaccio, “Cosmic chronometers to calibrate the ladders and measure the
curvature of the Universe. A model-independent study,” Mon. Not. Roy. Astron. Soc. 523 (2023) no. 3, 3406–3422,
arXiv:2301.09591 [astro-ph.CO].
[430] A. Favale, M. G. Dainotti, A. Gómez-Valent, and M. Migliaccio, “Towards a new model-independent calibration of
271

Gamma-Ray Bursts,” JHEAp 44 (2024) 323–339, arXiv:2402.13115 [astro-ph.CO].


[431] Y. Yang, T. Liu, J. Huang, X. Cheng, M. Biesiada, and S.-m. Wu, “Simultaneous measurements on cosmic curvature
and opacity using latest HII regions and H(z) observations,” Eur. Phys. J. C 84 (2024) no. 1, 3, arXiv:2401.03413
[astro-ph.CO].
[432] N. Rani, D. Jain, S. Mahajan, A. Mukherjee, and M. Biesiada, “Revisiting dark energy models using differential ages of
galaxies,” JCAP 03 (2017) 005, arXiv:1612.07492 [astro-ph.CO].
[433] Z. Li, J. E. Gonzalez, H. Yu, Z.-H. Zhu, and J. S. Alcaniz, “Constructing a cosmological model-independent Hubble
diagram of type Ia supernovae with cosmic chronometers,” Phys. Rev. D 93 (2016) no. 4, 043014, arXiv:1504.03269
[astro-ph.CO].
[434] S. Capozziello, R. D’Agostino, and O. Luongo, “High-redshift cosmography: auxiliary variables versus Padé
polynomials,” Mon. Not. Roy. Astron. Soc. 494 (2020) no. 2, 2576–2590, arXiv:2003.09341 [astro-ph.CO].
[435] M. Moresco, L. Verde, L. Pozzetti, R. Jimenez, and A. Cimatti, “New constraints on cosmological parameters and
neutrino properties using the expansion rate of the Universe to z~1.75,” JCAP 07 (2012) 053, arXiv:1201.6658
[astro-ph.CO].
[436] M. Moresco, R. Jimenez, L. Verde, A. Cimatti, L. Pozzetti, C. Maraston, and D. Thomas, “Constraining the time
evolution of dark energy, curvature and neutrino properties with cosmic chronometers,” JCAP 12 (2016) 039,
arXiv:1604.00183 [astro-ph.CO].
[437] EUCLID Collaboration, R. Laureijs et al., “Euclid Definition Study Report,” arXiv:1110.3193 [astro-ph.CO].
[438] F. Cogato, M. Moresco, L. Amati, and A. Cimatti, “An analytical late–Universe approach to the weaving of modern
cosmology,” Mon. Not. Roy. Astron. Soc. 527 (2023) no. 3, 4874–4888, arXiv:2309.01375 [astro-ph.CO].
[439] A. Heavens, R. Jimenez, and L. Verde, “Standard rulers, candles, and clocks from the low-redshift Universe,” Phys.
Rev. Lett. 113 (2014) no. 24, 241302, arXiv:1409.6217 [astro-ph.CO].
[440] L. Verde, J. L. Bernal, A. F. Heavens, and R. Jimenez, “The length of the low-redshift standard ruler,” Mon. Not. Roy.
Astron. Soc. 467 (2017) no. 1, 731–736, arXiv:1607.05297 [astro-ph.CO].
[441] W. Guo, Q. Wang, S. Cao, M. Biesiada, T. Liu, Y. Lian, X. Jiang, C. Mu, and D. Cheng, “Newest Measurements of
Hubble Constant from DESI 2024 Baryon Acoustic Oscillation Observations,” Astrophys. J. Lett. 978 (2025) no. 2,
L33, arXiv:2412.13045 [astro-ph.CO].
[442] L. Amati, R. D’Agostino, O. Luongo, M. Muccino, and M. Tantalo, “Addressing the circularity problem in the
Ep − Eiso correlation of gamma-ray bursts,” Mon. Not. Roy. Astron. Soc. 486 (2019) no. 1, L46–L51,
arXiv:1811.08934 [astro-ph.HE].
[443] J.-J. Wei and F. Melia, “Model-independent Distance Calibration and Curvature Measurement Using Quasars and
Cosmic Chronometers,” Astrophys. J. 888 (2020) no. 2, 99, arXiv:1912.00668 [astro-ph.CO].
[444] V. Sahni, A. Shafieloo, and A. A. Starobinsky, “Model independent evidence for dark energy evolution from Baryon
Acoustic Oscillations,” Astrophys. J. Lett. 793 (2014) no. 2, L40, arXiv:1406.2209 [astro-ph.CO].
[445] X. Zheng, X. Ding, M. Biesiada, S. Cao, and Z. Zhu, “What are Omh2 (z1 , z2 ) and Om(z1 , z2 ) diagnostics telling us in
light of H(z) data?,” Astrophys. J. 825 (2016) no. 1, 17, arXiv:1604.07910 [astro-ph.CO].
[446] X. Zheng, M. Biesiada, X. Ding, S. Cao, S. Zhang, and Z.-H. Zhu, “Statistical analysis with cosmic-expansion-rate
measurements and two-point diagnostics,” Eur. Phys. J. C 78 (2018) no. 3, 274, arXiv:1803.09106 [astro-ph.CO].
[447] T. Treu and P. J. Marshall, “Time Delay Cosmography,” Astron. Astrophys. Rev. 24 (2016) no. 1, 11,
arXiv:1605.05333 [astro-ph.CO].
[448] S. H. Suyu, T.-C. Chang, F. Courbin, and T. Okumura, “Cosmological distance indicators,” Space Sci. Rev. 214 (2018)
no. 5, 91, arXiv:1801.07262 [astro-ph.CO].
[449] T. Treu, S. H. Suyu, and P. J. Marshall, “Strong lensing time-delay cosmography in the 2020s,” Astron. Astrophys.
Rev. 30 (2022) no. 1, 8, arXiv:2210.15794 [astro-ph.CO].
[450] S. H. Suyu, A. Goobar, T. Collett, A. More, and G. Vernardos, “Strong Gravitational Lensing and Microlensing of
Supernovae,” Space Sci. Rev. 220 (2024) no. 1, 13, arXiv:2301.07729 [astro-ph.CO].
[451] S. Birrer, M. Millon, D. Sluse, A. J. Shajib, F. Courbin, S. Erickson, L. V. E. Koopmans, S. H. Suyu, and T. Treu,
“Time-Delay Cosmography: Measuring the Hubble Constant and Other Cosmological Parameters with Strong
Gravitational Lensing,” Space Sci. Rev. 220 (2024) no. 5, 48, arXiv:2210.10833 [astro-ph.CO].
[452] S. Refsdal, “On the Possibility of Determining Hubble’s Parameter and the Masses of Galaxies from the Gravitational
Lens Effect,” Mon. Not. Roy. Astron. Soc. 128 (1964) no. 4, 307–310.
[453] XXX Collaboration, R. M. Quimby, M. Oguri, A. More, S. More, T. J. Moriya, M. C. Werner, M. Tanaka,
G. Folatelli, M. C. Bersten, and K. Nomoto, “Detection of the Gravitational Lens Magnifying a Type Ia Supernova,”
Science 344 (2014) no. 6, 396–399, arXiv:1404.6014 [astro-ph.CO].
[454] P. L. Kelly et al., “Multiple Images of a Highly Magnified Supernova Formed by an Early-Type Cluster Galaxy Lens,”
Science 347 (2015) 1123, arXiv:1411.6009 [astro-ph.CO].
[455] P. L. Kelly et al., “Constraints on the Hubble constant from supernova Refsdal’s reappearance,” Science 380 (2023)
no. 6649, abh1322, arXiv:2305.06367 [astro-ph.CO].
[456] C. Grillo, L. Pagano, P. Rosati, and S. H. Suyu, “Cosmography with supernova Refsdal through time-delay cluster
lensing: Independent measurements of the Hubble constant and geometry of the Universe,” Astron. Astrophys. 684
(2024) L23, arXiv:2401.10980 [astro-ph.CO].
[457] I. Jee, E. Komatsu, S. H. Suyu, and D. Huterer, “Time-delay Cosmography: Increased Leverage with Angular
Diameter Distances,” JCAP 04 (2016) 031, arXiv:1509.03310 [astro-ph.CO].
[458] E. V. Linder, “Lensing Time Delays and Cosmological Complementarity,” Phys. Rev. D 84 (2011) 123529,
272

arXiv:1109.2592 [astro-ph.CO].
[459] S. H. Suyu et al., “Cosmology from gravitational lens time delays and Planck data,” Astrophys. J. Lett. 788 (2014)
L35, arXiv:1306.4732 [astro-ph.CO].
[460] T. Collett, F. Montanari, and S. Rasanen, “Model-Independent Determination of H0 and ΩK0 from Strong Lensing
and Type Ia Supernovae,” Phys. Rev. Lett. 123 (2019) no. 23, 231101, arXiv:1905.09781 [astro-ph.CO].
[461] A. J. Shajib, T. Treu, and A. Agnello, “Improving time-delay cosmography with spatially resolved kinematics,” Mon.
Not. Roy. Astron. Soc. 473 (2018) no. 1, 210–226, arXiv:1709.01517 [astro-ph.CO].
[462] DES Collaboration, A. J. Shajib et al., “STRIDES: a 3.9 per cent measurement of the Hubble constant from the
strong lens system DES J0408−5354,” Mon. Not. Roy. Astron. Soc. 494 (2020) no. 4, 6072–6102, arXiv:1910.06306
[astro-ph.CO].
[463] M. Millon et al., “TDCOSMO. I. An exploration of systematic uncertainties in the inference of H0 from time-delay
cosmography,” Astron. Astrophys. 639 (2020) A101, arXiv:1912.08027 [astro-ph.CO].
[464] E. E. Falco, M. V. Gorenstein, and I. I. Shapiro, “On model-dependent bounds on H 0 from gravitational images :
application to Q 0957+561 A, B.,” ApJL 289 (1985) L1–L4.
[465] J. Wagner, “Generalised model-independent characterisation of strong gravitational lenses IV: formalism-intrinsic
degeneracies,” Astron. Astrophys. 620 (2018) A86, arXiv:1809.03505 [astro-ph.CO].
[466] S. Birrer et al., “TDCOSMO - IV. Hierarchical time-delay cosmography – joint inference of the Hubble constant and
galaxy density profiles,” Astron. Astrophys. 643 (2020) A165, arXiv:2007.02941 [astro-ph.CO].
[467] A. Goobar et al., “iPTF16geu: A multiply imaged, gravitationally lensed type Ia supernova,” Science 356 (2017)
291–295, arXiv:1611.00014 [astro-ph.CO].
[468] A. Goobar et al., “Uncovering a population of gravitational lens galaxies with magnified standard candle SN Zwicky,”
Nature Astron. 7 (2023) 1098–1107, arXiv:2211.00656 [astro-ph.CO]. [Erratum: Nature Astron. 7, 1137–1137
(2023)].
[469] S. A. Rodney, G. B. Brammer, J. D. R. Pierel, J. Richard, S. Toft, K. F. O’Connor, M. Akhshik, and K. E. Whitaker,
“A gravitationally lensed supernova with an observable two-decade time delay,” Nature Astron. 5 (2021) no. 11,
1118–1125, arXiv:2106.08935 [astro-ph.CO].
[470] W. Chen, P. L. Kelly, M. Oguri, T. J. Broadhurst, J. M. Diego, N. Emami, A. V. Filippenko, T. L. Treu, and
A. Zitrin, “Shock cooling of a red-supergiant supernova at redshift 3 in lensed images,” Nature (London) 611 (2022)
no. 7935, 256–259, arXiv:2306.12985 [astro-ph.GA].
[471] P. Kelly, A. Zitrin, M. Oguri, J. Diego, H. Williams, T. Broadhurst, W. Chen, A. Koekemoer, J. Pierel, L. Strolger, and
T. Treu, “Strongly Lensed SN in MACS 2129 Galaxy-Cluster Field,” Transient Name Server AstroNote 169 (2022) 1.
[472] B. L. Frye et al., “The JWST Discovery of the Triply Imaged Type Ia “Supernova H0pe” and Observations of the
Galaxy Cluster PLCK G165.7+67.0,” Astrophys. J. 961 (2024) no. 2, 171, arXiv:2309.07326 [astro-ph.GA].
[473] J. D. R. Pierel et al., “Lensed Type Ia Supernova “Encore” at z = 2: The First Instance of Two Multiply Imaged
Supernovae in the Same Host Galaxy,” Astrophys. J. Lett. 967 (2024) no. 2, L37, arXiv:2404.02139 [astro-ph.CO].
[474] J. D. R. Pierel et al., “JWST Photometric Time-delay and Magnification Measurements for the Triply Imaged Type Ia
“SN H0pe” at z = 1.78,” Astrophys. J. 967 (2024) no. 1, 50, arXiv:2403.18954 [astro-ph.CO].
[475] J. D. R. Pierel, S. Rodney, G. Vernardos, M. Oguri, R. Kessler, and T. Anguita, “Projected Cosmological Constraints
from Strongly Lensed Supernovae with the Roman Space Telescope,” Astrophys. J. 908 (2021) no. 2, 190,
arXiv:2010.12399 [astro-ph.CO].
[476] S. Birrer and T. Treu, “TDCOSMO - V. Strategies for precise and accurate measurements of the Hubble constant with
strong lensing,” Astron. Astrophys. 649 (2021) A61, arXiv:2008.06157 [astro-ph.CO].
[477] M. Oguri and P. J. Marshall, “Gravitationally lensed quasars and supernovae in future wide-field optical imaging
surveys,” Mon. Not. Roy. Astron. Soc. 405 (2010) 2579–2593, arXiv:1001.2037 [astro-ph.CO].
[478] T. Petrushevska, “Strongly Lensed Supernovae in Well-Studied Galaxy Clusters with the Vera C. Rubin Observatory,”
Symmetry 12 (2020) no. 12, 1966, arXiv:2011.14122 [astro-ph.CO].
[479] LSST Dark Enrgy Science Collaboration, N. Arendse, S. Dhawan, A. S. Carracedo, H. V. Peiris, A. Goobar,
R. Wojtak, C. Alves, R. Biswas, S. Huber, and S. Birrer, “Detecting strongly lensed type Ia supernovae with LSST,”
Mon. Not. Roy. Astron. Soc. 531 (2024) no. 3, 3509–3523, arXiv:2312.04621 [astro-ph.CO].
[480] T. E. Collett, “The population of galaxy-galaxy strong lenses in forthcoming optical imaging surveys,” Astrophys. J.
811 (2015) no. 1, 20, arXiv:1507.02657 [astro-ph.CO].
[481] X.-L. Meng, T. Treu, A. Agnello, M. W. Auger, K. Liao, and P. J. Marshall, “Precision cosmology with time delay
lenses: high resolution imaging requirements,” JCAP 09 (2015) 059, arXiv:1506.07640 [astro-ph.CO].
[482] K. Liao et al., “Strong Lens Time Delay Challenge: II. Results of TDC1,” Astrophys. J. 800 (2015) no. 1, 11,
arXiv:1409.1254 [astro-ph.IM].
[483] P. Jovanovic, A. F. Zakharov, L. C. Popovic, and T. Petrovic, “Microlensing of the X-ray, UV and optical emission
regions of quasars: Simulations of the time-scales and amplitude variations of microlensing events,” Mon. Not. Roy.
Astron. Soc. 386 (2008) 397, arXiv:0801.4473 [astro-ph].
[484] M. Millon et al., “TDCOSMO - II. Six new time delays in lensed quasars from high-cadence monitoring at the MPIA
2.2 m telescope,” Astron. Astrophys. 642 (2020) A193, arXiv:2006.10066 [astro-ph.CO].
[485] D. A. Goldstein, P. E. Nugent, D. N. Kasen, and T. E. Collett, “Precise Time Delays from Strongly Gravitationally
Lensed Type Ia Supernovae with Chromatically Microlensed Images,” Astrophys. J. 855 (2018) no. 1, 22,
arXiv:1708.00003 [astro-ph.CO].
[486] J. Bayer, S. Huber, C. Vogl, S. H. Suyu, S. Taubenberger, D. Sluse, J. H. H. Chan, and W. E. Kerzendorf,
273

“HOLISMOKES - V. Microlensing of type II supernovae and time-delay inference through spectroscopic phase
retrieval,” Astron. Astrophys. 653 (2021) A29, arXiv:2101.06229 [astro-ph.CO].
[487] A. Etherington et al., “Automated galaxy–galaxy strong lens modelling: No lens left behind,” Mon. Not. Roy. Astron.
Soc. 517 (2022) no. 3, 3275–3302, arXiv:2202.09201 [astro-ph.CO].
[488] DES Collaboration, T. Schmidt et al., “STRIDES: automated uniform models for 30 quadruply imaged quasars,”
Mon. Not. Roy. Astron. Soc. 518 (2023) no. 1, 1260–1300, arXiv:2206.04696 [astro-ph.CO].
[489] R. Legin, Y. Hezaveh, L. Perreault-Levasseur, and B. Wandelt, “A Framework for Obtaining Accurate Posteriors of
Strong Gravitational Lensing Parameters with Flexible Priors and Implicit Likelihoods Using Density Estimation,”
Astrophys. J. 943 (2023) no. 1, 4, arXiv:2212.00044 [astro-ph.IM].
[490] A. J. Shajib, T. Treu, S. Birrer, and A. Sonnenfeld, “Dark matter haloes of massive elliptical galaxies at z ∼ 0.2 are
well described by the Navarro-Frenk-White profile,” MNRAS 503 (2021) no. 2, 2380–2405, arXiv:2008.11724
[astro-ph.GA].
[491] A. Adam, L. Perreault-Levasseur, Y. Hezaveh, and M. Welling, “Pixelated Reconstruction of Foreground Density and
Background Surface Brightness in Gravitational Lensing Systems Using Recurrent Inference Machines,” Astrophys. J.
951 (2023) no. 1, 6, arXiv:2301.04168 [astro-ph.IM].
[492] LSST Dark Energy Science Collaboration, J. W. Park, S. Wagner-Carena, S. Birrer, P. J. Marshall, J. Y.-Y. Lin,
and A. Roodman, “Large-Scale Gravitational Lens Modeling with Bayesian Neural Networks for Accurate and Precise
Inference of the Hubble Constant,” Astrophys. J. 910 (2021) no. 1, 39, arXiv:2012.00042 [astro-ph.IM].
[493] S. Schuldt, R. Cañameras, Y. Shu, S. H. Suyu, S. Taubenberger, T. Meinhardt, and L. Leal-Taixé, “HOLISMOKES –
IX. Neural network inference of strong-lens parameters and uncertainties from ground-based images,” Astron.
Astrophys. 671 (2023) A147, arXiv:2206.11279 [astro-ph.IM].
[494] S. Ertl, S. Schuldt, S. H. Suyu, T. Schmidt, T. Treu, S. Birrer, A. J. Shajib, and D. Sluse, “TDCOSMO X. Automated
Modeling of 9 Strongly Lensed Quasars and Comparison Between Lens Modeling Software,” Astron. Astrophys. 672
(2023) A2, arXiv:2209.03094 [astro-ph.CO].
[495] T. Treu and L. V. E. Koopmans, “The internal structure of the lens pg1115+080: breaking degeneracies in the value of
the hubble constant,” Mon. Not. Roy. Astron. Soc. 337 (2002) L6, arXiv:astro-ph/0210002.
[496] A. Yıldırım, S. H. Suyu, and A. Halkola, “Time-delay cosmographic forecasts with strong lensing and JWST stellar
kinematics,” Mon. Not. Roy. Astron. Soc. 493 (2020) no. 4, 4783–4807, arXiv:1904.07237 [astro-ph.CO].
[497] TDCOSMO Collaboration, A. J. Shajib et al., “TDCOSMO. XII. Improved Hubble constant measurement from
lensing time delays using spatially resolved stellar kinematics of the lens galaxy,” Astron. Astrophys. 673 (2023) A9,
arXiv:2301.02656 [astro-ph.CO].
[498] S. Birrer, S. Dhawan, and A. J. Shajib, “The Hubble Constant from Strongly Lensed Supernovae with Standardizable
Magnifications,” Astrophys. J. 924 (2022) no. 1, 2, arXiv:2107.12385 [astro-ph.CO].
[499] N. Khadka, S. Birrer, A. Leauthaud, and H. Nix, “Breaking the mass-sheet degeneracy in strong lensing mass
modelling with weak lensing observations,” Mon. Not. Roy. Astron. Soc. 533 (2024) no. 1, 795–806, arXiv:2404.01513
[astro-ph.CO].
[500] A. Yıldırım, S. H. Suyu, G. C. F. Chen, and E. Komatsu, “TDCOSMO - XIII. Cosmological distance measurements in
light of the mass-sheet degeneracy: Forecasts from strong lensing and integral field unit stellar kinematics,” Astron.
Astrophys. 675 (2023) A21, arXiv:2109.14615 [astro-ph.CO].
[501] A. Sonnenfeld, S.-S. Li, G. Despali, R. Gavazzi, A. J. Shajib, and E. N. Taylor, “Strong lensing selection effects,” A&A
678 (2023) A4, arXiv:2301.13230 [astro-ph.GA].
[502] E. Lusso et al., “Quasars as standard candles III. Validation of a new sample for cosmological studies,” Astron.
Astrophys. 642 (2020) A150, arXiv:2008.08586 [astro-ph.GA].
[503] F. Wang, J. Yang, X. Fan, J. F. Hennawi, A. J. Barth, E. Banados, F. Bian, K. Boutsia, T. Connor, F. B. Davies,
R. Decarli, A.-C. Eilers, E. P. Farina, R. Green, L. Jiang, J.-T. Li, C. Mazzucchelli, R. Nanni, J.-T. Schindler,
B. Venemans, F. Walter, X.-B. Wu, and M. Yue, “A Luminous Quasar at Redshift 7.642,” ApJL 907 (2021) no. 1, L1,
arXiv:2101.03179 [astro-ph.GA].
[504] D. Watson, K. D. Denney, M. Vestergaard, and T. M. Davis, “A new cosmological distance measure using AGN,”
Astrophys. J. Lett. 740 (2011) L49, arXiv:1109.4632 [astro-ph.CO].
[505] M. Haas, R. Chini, M. Ramolla, F. Pozo Nuñez, C. Westhues, R. Watermann, V. Hoffmeister, and M. Murphy,
“Photometric AGN reverberation mapping - an efficient tool for BLR sizes, black hole masses, and host-subtracted
AGN luminosities,” A&A 535 (2011) A73, arXiv:1109.1848 [astro-ph.CO].
[506] F. La Franca, S. Bianchi, G. Ponti, E. Branchini, and G. Matt, “A New Cosmological Distance Measure Using Active
Galactic Nucleus X-Ray Variability,” Astrophys. J. Lett. 787 (2014) L12, arXiv:1404.2607 [astro-ph.CO].
[507] M. L. Martínez-Aldama, A. del Olmo, P. Marziani, J. W. Sulentic, C. A. Negrete, D. Dultzin, M. D’Onofrio, and
J. Perea, “Extreme quasars at high redshift,” A&A 618 (2018) A179, arXiv:1807.11006 [astro-ph.GA].
[508] P. Marziani, D. Dultzin, J. W. Sulentic, A. Del Olmo, C. A. Negrete, M. L. Martínez-Aldama, M. D’Onofrio, E. Bon,
N. Bon, and G. M. Stirpe, “A main sequence for quasars,” Frontiers in Astronomy and Space Sciences 5 (2018) 6,
arXiv:1802.05575 [astro-ph.GA].
[509] S. Panda, P. Marziani, and B. Czerny, “The Quasar Main Sequence explained by the combination of Eddington ratio,
metallicity and orientation,” Astrophys. J. 882 (2019) no. 2, 79, arXiv:1905.01729 [astro-ph.HE].
[510] S. Panda and P. Marziani, “High Eddington quasars as discovery tools: current state and challenges,” Frontiers in
Astronomy and Space Sciences 10 (2023) , arXiv:2210.15041 [astro-ph.GA].
https://www.frontiersin.org/journals/astronomy-and-space-sciences/articles/10.3389/fspas.2023.1130103.
274

[511] R. Prince, K. Hryniewicz, S. Panda, B. Czerny, and A. Pollo, “Viewing Angle Observations and Effects of Evolution
with Redshift, Black Hole Mass, and Eddington Ratio in Quasar-based Cosmology,” Astrophys. J. 925 (2022) no. 2,
215, arXiv:2106.03877 [astro-ph.GA].
[512] M. G. Dainotti, G. Bardiacchi, A. L. Lenart, S. Capozziello, E. O. Colgain, R. Solomon, D. Stojkovic, and M. M.
Sheikh-Jabbari, “Quasar Standardization: Overcoming Selection Biases and Redshift Evolution,” Astrophys. J. 931
(2022) no. 2, 106, arXiv:2203.12914 [astro-ph.HE].
[513] S. Cao, M. Zajaček, S. Panda, M. L. Martínez-Aldama, B. Czerny, and B. Ratra, “Standardizing
reverberation-measured C iv time-lag quasars, and using them with standardized Mg ii quasars to constrain
cosmological parameters,” Mon. Not. Roy. Astron. Soc. 516 (2022) no. 2, 1721–1740, arXiv:2205.15552
[astro-ph.CO].
[514] S. Cao, M. Zajaček, B. Czerny, S. Panda, and B. Ratra, “Effects of heterogeneous data sets and time-lag measurement
techniques on cosmological parameter constraints from Mg ii and C iv reverberation-mapped quasar data,” Mon. Not.
Roy. Astron. Soc. 528 (2024) no. 4, 6444–6469, arXiv:2309.16516 [astro-ph.CO].
[515] M. G. Dainotti, B. De Simone, T. Schiavone, G. Montani, E. Rinaldi, G. Lambiase, M. Bogdan, and S. Ugale, “On the
Evolution of the Hubble Constant with the SNe Ia Pantheon Sample and Baryon Acoustic Oscillations: A Feasibility
Study for GRB-Cosmology in 2030,” Galaxies 10 (2022) no. 1, 24, arXiv:2201.09848 [astro-ph.CO].
[516] G. Bargiacchi, M. G. Dainotti, S. Nagataki, and S. Capozziello, “Gamma-ray bursts, quasars, baryonic acoustic
oscillations, and supernovae Ia: new statistical insights and cosmological constraints,” Mon. Not. Roy. Astron. Soc.
521 (2023) no. 3, 3909–3924, arXiv:2303.07076 [astro-ph.CO].
[517] M. G. Dainotti, G. Bargiacchi, M. Bogdan, A. L. Lenart, K. Iwasaki, S. Capozziello, B. Zhang, and N. Fraija,
“Reducing the Uncertainty on the Hubble Constant up to 35% with an Improved Statistical Analysis: Different Best-fit
Likelihoods for Type Ia Supernovae, Baryon Acoustic Oscillations, Quasars, and Gamma-Ray Bursts,” Astrophys. J.
951 (2023) no. 1, 63, arXiv:2305.10030 [astro-ph.CO].
[518] A. L. Lenart, G. Bargiacchi, M. G. Dainotti, S. Nagataki, and S. Capozziello, “A Bias-free Cosmological Analysis with
Quasars Alleviating H 0 Tension,” Astrophys. J. Suppl. 264 (2023) no. 2, 46, arXiv:2211.10785 [astro-ph.CO].
[519] T. Yang, A. Banerjee, and E. O. Colgáin, “Cosmography and flat ΛCDM tensions at high redshift,” Phys. Rev. D 102
(2020) no. 12, 123532, arXiv:1911.01681 [astro-ph.CO].
[520] N. Khadka and B. Ratra, “Quasar X-ray and UV flux, baryon acoustic oscillation, and Hubble parameter measurement
constraints on cosmological model parameters,” Mon. Not. Roy. Astron. Soc. 492 (2020) no. 3, 4456–4468,
arXiv:1909.01400 [astro-ph.CO].
[521] G. Bargiacchi, M. Benetti, S. Capozziello, E. Lusso, G. Risaliti, and M. Signorini, “Quasar cosmology: dark energy
evolution and spatial curvature,” Mon. Not. Roy. Astron. Soc. 515 (2022) no. 2, 1795–1806, arXiv:2111.02420
[astro-ph.CO].
[522] E. O. Colgáin, M. M. Sheikh-Jabbari, R. Solomon, G. Bargiacchi, S. Capozziello, M. G. Dainotti, and D. Stojkovic,
“Revealing intrinsic flat ΛCDM biases with standardizable candles,” Phys. Rev. D 106 (2022) no. 4, L041301,
arXiv:2203.10558 [astro-ph.CO].
[523] M. Zajaček, B. Czerny, N. Khadka, M. L. Martínez-Aldama, R. Prince, S. Panda, and B. Ratra, “Effect of Extinction
on Quasar Luminosity Distances Determined from UV and X-Ray Flux Measurements,” Astrophys. J. 961 (2024)
no. 2, 229, arXiv:2305.08179 [astro-ph.GA].
[524] A. Sacchi et al., “Quasars as high-redshift standard candles,” Astron. Astrophys. 663 (2022) L7, arXiv:2206.13528
[astro-ph.CO].
[525] H. Shabani, A. De, T.-H. Loo, and E. N. Saridakis, “Cosmology of f(Q) gravity in non-flat Universe,” Eur. Phys. J. C
84 (2024) no. 3, 285, arXiv:2306.13324 [gr-qc].
[526] M. Sabiee, M. Malekjani, and D. Mohammad Zadeh Jassur, “f(T) cosmology against the cosmographic method: A new
study using mock and observational data,” Mon. Not. Roy. Astron. Soc. 516 (2022) no. 2, 2597–2613,
arXiv:2212.04113 [astro-ph.CO].
[527] E. Lusso, E. Piedipalumbo, G. Risaliti, M. Paolillo, S. Bisogni, E. Nardini, and L. Amati, “Tension with the flat
ΛCDM model from a high-redshift Hubble diagram of supernovae, quasars, and gamma-ray bursts,” Astron. Astrophys.
628 (2019) L4, arXiv:1907.07692 [astro-ph.CO].
[528] M. G. Dainotti, A. L. Lenart, M. G. Yengejeh, S. Chakraborty, N. Fraija, E. Di Valentino, and G. Montani, “A new
binning method to choose a standard set of Quasars,” Phys. Dark Univ. 44 (2024) 101428, arXiv:2401.12847
[astro-ph.HE].
[529] G. Bargiacchi, G. Risaliti, M. Benetti, S. Capozziello, E. Lusso, A. Saccardi, and M. Signorini, “Cosmography by
orthogonalized logarithmic polynomials,” Astron. Astrophys. 649 (2021) A65, arXiv:2101.08278 [astro-ph.CO].
[530] G. Bargiacchi, M. G. Dainotti, and S. Capozziello, “Tensions with the flat ΛCDM model from high-redshift
cosmography,” Mon. Not. Roy. Astron. Soc. 525 (2023) no. 2, 3104–3116, arXiv:2307.15359 [astro-ph.CO].
[531] R. A. Preston, D. D. Morabito, J. G. Williams, J. Faulkner, D. L. Jauncey, and G. Nicolson, “A VLBI survey at 2.29
GHz.,” AJ 90 (1985) 1599–1603.
[532] J.-M. Wang, Y.-Y. Songsheng, Y.-R. Li, P. Du, and Z.-X. Zhang, “A parallax distance to 3C 273 through
spectroastrometry and reverberation mapping,” Nature Astron. 4 (2020) 517–525, arXiv:1906.08417 [astro-ph.CO].
[533] R. Sandoval-Orozco, C. Escamilla-Rivera, R. Briffa, and J. Levi Said, “f(T) cosmology in the regime of quasar
observations,” Phys. Dark Univ. 43 (2024) 101407, arXiv:2309.03675 [astro-ph.CO].
[534] A. Sandage, “The Change of Redshift and Apparent Luminosity of Galaxies due to the Deceleration of Selected
Expanding Universes.,” Astrophys. J. 136 (1962) 319.
275

[535] G. Calderone, K. Boutsia, S. Cristiani, A. Grazian, R. Amorin, V. D’Odorico, G. Cupani, F. Fontanot, and M. Salvato,
“Finding the brightest cosmic beacons in the Southern Hemisphere,” Astrophys. J. 887 (2019) no. 2, 268,
arXiv:1909.06391 [astro-ph.IM].
[536] J. Liske et al., “Cosmic dynamics in the era of Extremely Large Telescopes,” Mon. Not. Roy. Astron. Soc. 386 (2008)
1192–1218, arXiv:0802.1532 [astro-ph].
[537] C. M. J. Marques, C. J. A. P. Martins, and C. S. Alves, “Fundamental cosmology from ANDES precision
spectroscopy,” Mon. Not. Roy. Astron. Soc. 522 (2023) no. 4, 5973–5979, arXiv:2305.01446 [astro-ph.IM].
[538] ANDES Collaboration, C. J. A. P. Martins et al., “Cosmology and fundamental physics with the ELT-ANDES
spectrograph,” Exper. Astron. 57 (2024) no. 1, 5, arXiv:2311.16274 [astro-ph.CO].
[539] M. G. Dainotti, G. Bargiacchi, A. L. Lenart, S. Nagataki, and S. Capozziello, “Quasars: Standard Candles up to z =
7.5 with the Precision of Supernovae Ia,” Astrophys. J. 950 (2023) no. 1, 45, arXiv:2305.19668 [astro-ph.CO].
[540] M. G. Dainotti, G. Bargiacchi, A. L. Lenart, and S. Capozziello, “The Scavenger Hunt for Quasar Samples to Be Used
as Cosmological Tools,” Galaxies 12 (2024) no. 1, 4, arXiv:2401.11998 [astro-ph.CO].
[541] L. Amati et al., “Intrinsic spectra and energetics of BeppoSAX gamma-ray bursts with known redshifts,” Astron.
Astrophys. 390 (2002) 81, arXiv:astro-ph/0205230.
[542] L. Amati, “The E(p,i) - E(iso) correlation in grbs: updated observational status, re-analysis and main implications,”
Mon. Not. Roy. Astron. Soc. 372 (2006) 233–245, arXiv:astro-ph/0601553.
[543] L. Amati, F. Frontera, and C. Guidorzi, “Spectrum-energy correlations in Gamma-Ray Bursts confront extremely
energetic Fermi GRBs,” Astron. Astrophys. 508 (2009) 173–180, arXiv:0907.0384 [astro-ph.HE].
[544] F. F. Dirirsa, S. Razzaque, F. Piron, M. Arimoto, M. Axelsson, D. Kocevski, F. Longo, M. Ohno, and S. Zhu, “Spectral
analysis of Fermi-LAT gamma-ray bursts with known redshift and their potential use as cosmological standard
candles,” Astrophys. J. 887 (2019) 13, arXiv:1910.07009 [astro-ph.HE].
[545] D. Yonetoku, T. Murakami, T. Nakamura, R. Yamazaki, A. K. Inoue, and K. Ioka, “Gamma-ray burst formation rates
inferred from the spectral peak energy-peak luminosity relation,” Astrophys. J. 609 (2004) 935,
arXiv:astro-ph/0309217.
[546] T. Aldowma and S. Razzaque, “Deep Neural Networks for estimation of gamma-ray burst redshifts,” Mon. Not. Roy.
Astron. Soc. 529 (2024) no. 3, 2676–2685, arXiv:2401.11005 [astro-ph.HE].
[547] M. G. Dainotti, V. F. Cardone, and S. Capozziello, “A time - luminosity correlation for Gamma Ray Bursts in the X -
rays,” Mon. Not. Roy. Astron. Soc. 391 (2008) 79, arXiv:0809.1389 [astro-ph].
[548] B. D. Metzger, D. Giannios, T. A. Thompson, N. Bucciantini, and E. Quataert, “The Proto-Magnetar Model for
Gamma-Ray Bursts,” Mon. Not. Roy. Astron. Soc. 413 (2011) 2031, arXiv:1012.0001 [astro-ph.HE].
[549] A. Rowlinson, P. T. O’Brien, B. D. Metzger, N. R. Tanvir, and A. J. Levan, “Signatures of magnetar central engines in
short GRB lightcurves,” Mon. Not. Roy. Astron. Soc. 430 (2013) 1061, arXiv:1301.0629 [astro-ph.HE].
[550] A. Rowlinson, B. P. Gompertz, M. Dainotti, P. T. O’Brien, R. A. M. J. Wijers, and A. J. van der Horst, “Constraining
properties of GRB magnetar central engines using the observed plateau luminosity and duration correlation,” Mon.
Not. Roy. Astron. Soc. 443 (2014) no. 2, 1779–1787, arXiv:1407.1053 [astro-ph.HE].
[551] G. Stratta, M. G. Dainotti, S. Dall’Osso, X. Hernandez, and G. De Cesare, “On the magnetar origin of the GRBs
presenting X-ray afterglow plateaus,” Astrophys. J. 869 (2018) no. 2, 155, arXiv:1804.08652 [astro-ph.HE].
[552] Z. L. Uhm and A. M. Beloborodov, “On the Mechanism of Gamma-Ray Burst Afterglows,” Astrophys. J. Lett. 665
(2007) L93, arXiv:astro-ph/0701205.
[553] Z. L. Uhm, B. Zhang, R. Hascoet, F. Daigne, R. Mochkovitch, and I. H. Park, “Dynamics and Afterglow Light Curves
of GRB Blast Waves with a Long-lived Reverse Shock,” Astrophys. J. 761 (2012) 147, arXiv:1208.2347
[astro-ph.HE].
[554] R. Hascoet, F. Daigne, and R. Mochkovitch, “The prompt–early afterglow connection in gamma-ray bursts:
implications for the early afterglow physics,” Mon. Not. Roy. Astron. Soc. 442 (2014) no. 1, 20–27, arXiv:1401.0751
[astro-ph.HE].
[555] M. G. Dainotti, R. Willingale, S. Capozziello, V. F. Cardone, and M. Ostrowski, “Discovery of a tight correlation for
gamma ray burst afterglows with ‘canonical’ light curves,” Astrophys. J. Lett. 722 (2010) L215, arXiv:1009.1663
[astro-ph.HE].
[556] M. G. Dainotti, M. Ostrowski, and R. Willingale, “Toward a standard Gamma Ray Burst: tight correlations between
the prompt and the afterglow plateau phase emission,” Mon. Not. Roy. Astron. Soc. 418 (2011) 2202,
arXiv:1103.1138 [astro-ph.HE].
[557] M. G. Dainotti, V. Petrosian, J. Singal, and M. Ostrowski, “Determination of the Intrinsic Luminosity Time
Correlation in the X-Ray Afterglows of Gamma-Ray Bursts,” Astrophys. J. 774 (2013) 157, arXiv:1307.7297
[astro-ph.HE].
[558] M. G. Dainotti, V. Petrosian, R. Willingale, P. O’Brien, M. Ostrowski, and S. Nagataki, “Luminosity–time and
luminosity–luminosity correlations for GRB prompt and afterglow plateau emissions,” Mon. Not. Roy. Astron. Soc.
451 (2015) no. 4, 3898–3908, arXiv:1506.00702 [astro-ph.HE].
[559] M. G. Dainotti, S. Postnikov, X. Hernandez, and M. Ostrowski, “A fundamental plane for long gamma-ray bursts with
X-ray plateaus,” Astrophys. J. Lett. 825 (2016) no. 2, L20, arXiv:1604.06840 [astro-ph.HE].
[560] M. G. Dainotti, X. Hernandez, S. Postnikov, S. Nagataki, P. Obrien, R. Willingale, and S. Striegel, “A Study of the
Gamma-Ray Burst Fundamental Plane,” Astrophys. J. 848 (2017) no. 2, 88, arXiv:1704.04908 [astro-ph.HE].
[561] M. G. Dainotti, S. Nagataki, K. Maeda, S. Postnikov, and E. Pian, “A study of gamma ray bursts with afterglow
plateau phases associated with supernovae,” Astron. Astrophys. 600 (2017) A98, arXiv:1612.02917 [astro-ph.HE].
276

[562] M. G. Dainotti, A. Lenart, G. Sarracino, S. Nagataki, S. Capozziello, and N. Fraija, “The X-ray fundamental plane of
the Platinum Sample, the Kilonovae and the SNe Ib/c associated with GRBs,” Astrophys. J. 904 (2020) no. 2, 97,
arXiv:2010.02092 [astro-ph.HE].
[563] V. F. Cardone, S. Capozziello, and M. G. Dainotti, “An updated Gamma Ray Bursts Hubble diagram,” Mon. Not.
Roy. Astron. Soc. 400 (2009) no. 2, 775–790, arXiv:0901.3194 [astro-ph.CO].
[564] M. G. Dainotti, V. F. Cardone, E. Piedipalumbo, and S. Capozziello, “Slope evolution of GRB correlations and
cosmology,” Mon. Not. Roy. Astron. Soc. 436 (2013) 82, arXiv:1308.1918 [astro-ph.HE].
[565] M. G. Dainotti, V. Nielson, G. Sarracino, E. Rinaldi, S. Nagataki, S. Capozziello, O. Y. Gnedin, and G. Bargiacchi,
“Optical and X-ray GRB Fundamental Planes as cosmological distance indicators,” Mon. Not. Roy. Astron. Soc. 514
(2022) no. 2, 1828–1856, arXiv:2203.15538 [astro-ph.CO].
[566] M. G. Dainotti, A. L. Lenart, A. Chraya, G. Sarracino, S. Nagataki, N. Fraija, S. Capozziello, and M. Bogdan, “The
Gamma-ray Bursts fundamental plane correlation as a cosmological tool,” Mon. Not. Roy. Astron. Soc. 518 (2023)
no. 2, 2201–2240, arXiv:2209.08675 [astro-ph.HE].
[567] M. G. Dainotti et al., “Inferring the Redshift of More than 150 GRBs with a Machine-learning Ensemble Model,”
Astrophys. J. Suppl. 271 (2024) no. 1, 22, arXiv:2401.03589 [astro-ph.CO].
[568] M. G. Dainotti, A. Narendra, A. Pollo, V. Petrosian, M. Bogdan, K. Iwasaki, J. X. Prochaska, E. Rinaldi, and
D. Zhou, “Gamma-Ray Bursts as Distance Indicators by a Statistical Learning Approach,” Astrophys. J. Lett. 967
(2024) no. 2, L30, arXiv:2402.04551 [astro-ph.HE].
[569] M. G. Dainotti, R. Sharma, A. Narendra, D. Levine, E. Rinaldi, A. Pollo, and G. Bhatta, “A Stochastic Approach To
Reconstruct Gamma Ray Burst Lightcurves,” Astrophys. J. 267 (2023) no. 2, 42, arXiv:2305.12126 [astro-ph.HE].
[570] N. Liang, Z. Li, X. Xie, and P. Wu, “Calibrating Gamma-Ray Bursts by Using a Gaussian Process with Type Ia
Supernovae,” Astrophys. J. 941 (2022) no. 1, 84, arXiv:2211.02473 [astro-ph.CO].
[571] O. Luongo and M. Muccino, “Kinematic constraints beyond z ≃ 0 using calibrated GRB correlations,” Astron.
Astrophys. 641 (2020) A174, arXiv:2010.05218 [astro-ph.CO].
[572] Y. Mu, B. Chang, and L. Xu, “Cosmography via Gaussian process with gamma ray bursts,” JCAP 09 (2023) 041,
arXiv:2302.02559 [astro-ph.CO].
[573] N. Rea, M. Gullon, J. A. Pons, R. Perna, M. G. Dainotti, J. A. Miralles, and D. F. Torres, “Constraining the
GRB-magnetar model by means of the Galactic pulsar population,” Astrophys. J. 813 (2015) no. 2, 92,
arXiv:1510.01430 [astro-ph.HE].
[574] A. C. Alfano, S. Capozziello, O. Luongo, and M. Muccino, “Cosmological transition epoch from gamma-ray burst
correlations,” JHEAp 42 (2024) 178–196, arXiv:2402.18967 [astro-ph.CO].
[575] Z. Cano, “Gamma-ray burst supernovae as standardizable candles,” Astrophys. J. 794 (2014) 121, arXiv:1407.2589
[astro-ph.HE].
[576] M. G. Dainotti, B. De Simone, M. I. Khadir, K. Kawaguchi, T. J. Moriya, T. Takiwaki, N. Tominaga, and
A. Gangopadhyay, “The Quest for New Correlations in the Realm of the Gamma-Ray Burst—Supernova Connection,”
Astrophys. J. 938 (2022) no. 1, 41, arXiv:2208.10958 [astro-ph.HE].
[577] D. Staicova, “Impact of cosmology on Lorentz Invariance Violation constraints from GRB time-delays,” Class. Quant.
Grav. 40 (2023) no. 19, 195012, arXiv:2305.06504 [gr-qc].
[578] D. Staicova, “Probing for Lorentz Invariance Violation in Pantheon Plus Dominated Cosmology,” Universe 10 (2024)
no. 2, 75, arXiv:2401.06068 [gr-qc].
[579] Fermi-LAT Collaboration, S. Abdollahi et al., “A gamma-ray determination of the Universe’s star formation history,”
Science 362 (2018) no. 6418, 1031–1034, arXiv:1812.01031 [astro-ph.HE].
[580] MAGIC Collaboration, V. A. Acciari et al., “Measurement of the extragalactic background light using MAGIC and
Fermi-LAT gamma-ray observations of blazars up to z = 1,” Mon. Not. Roy. Astron. Soc. 486 (2019) no. 3, 4233–4251,
arXiv:1904.00134 [astro-ph.HE].
[581] A. Domínguez, R. Wojtak, J. Finke, M. Ajello, K. Helgason, F. Prada, A. Desai, V. Paliya, L. Marcotulli, and
D. Hartmann, “A new measurement of the Hubble constant and matter content of the Universe using extragalactic
background light γ-ray attenuation,” Astrophys. J. 885 (2019) no. 2, 137, arXiv:1903.12097 [astro-ph.CO].
[582] A. Domínguez et al., “A new derivation of the Hubble constant from γ-ray attenuation using improved optical depths
for the Fermi and CTA era,” Mon. Not. Roy. Astron. Soc. 527 (2023) no. 3, 4632–4642, arXiv:2306.09878
[astro-ph.HE].
[583] LIGO Scientific Collaboration, J. Aasi et al., “Advanced LIGO,” Class. Quant. Grav. 32 (2015) 074001,
arXiv:1411.4547 [gr-qc].
[584] VIRGO Collaboration, F. Acernese et al., “Advanced Virgo: a second-generation interferometric gravitational wave
detector,” Class. Quant. Grav. 32 (2015) no. 2, 024001, arXiv:1408.3978 [gr-qc].
[585] KAGRA Collaboration, T. Akutsu et al., “Overview of KAGRA: Calibration, detector characterization, physical
environmental monitors, and the geophysics interferometer,” PTEP 2021 (2021) no. 5, 05A102, arXiv:2009.09305
[gr-qc].
[586] LIGO Scientific, Virgo Collaboration, B. P. Abbott et al., “GW170817: Observation of Gravitational Waves from a
Binary Neutron Star Inspiral,” Phys. Rev. Lett. 119 (2017) no. 16, 161101, arXiv:1710.05832 [gr-qc].
[587] LIGO Scientific, KAGRA, VIRGO Collaboration, R. Abbott et al., “Observation of Gravitational Waves from Two
Neutron Star–Black Hole Coalescences,” Astrophys. J. Lett. 915 (2021) no. 1, L5, arXiv:2106.15163 [astro-ph.HE].
[588] LIGO Scientific, Virgo Collaboration, B. P. Abbott et al., “Observation of Gravitational Waves from a Binary Black
Hole Merger,” Phys. Rev. Lett. 116 (2016) no. 6, 061102, arXiv:1602.03837 [gr-qc].
277

[589] KAGRA, LIGO Scientific, Virgo Collaboration, B. P. Abbott et al., “Prospects for observing and localizing
gravitational-wave transients with Advanced LIGO, Advanced Virgo and KAGRA,” Living Rev. Rel. 19 (2016) 1,
arXiv:1304.0670 [gr-qc].
[590] P. Fritschel, K. Kuns, J. Driggers, A. Effler, B. Lantz, D. Ottaway, S. Ballmer, K. Dooley, R. Adhikari, M. Evans,
B. Farr, G. Gonzalez, P. Schmidt, and S. Raja, “Report from the LSC post-o5 study group,” Tech. Rep. T2200287,
LIGO, 2022. https://dcc.ligo.org/LIGO-T2200287/public.
[591] VIRGO Collaboration, F. Acernese et al., “Advanced Virgo Plus: Future Perspectives,” J. Phys. Conf. Ser. 2429
(2023) no. 1, 012040.
[592] D. Reitze et al., “Cosmic Explorer: The U.S. Contribution to Gravitational-Wave Astronomy beyond LIGO,” Bull.
Am. Astron. Soc. 51 (2019) no. 7, 035, arXiv:1907.04833 [astro-ph.IM].
[593] M. Evans et al., “A Horizon Study for Cosmic Explorer: Science, Observatories, and Community,” arXiv:2109.09882
[astro-ph.IM].
[594] M. Evans et al., “Cosmic Explorer: A Submission to the NSF MPSAC ngGW Subcommittee,” arXiv:2306.13745
[astro-ph.IM].
[595] ET Steering Committee, “ET design report update 2020,” Tech. Rep. ET-0007A-20, Einstein Telescope, 2020.
https://apps.et-gw.eu/tds/?content=3&r=17245.
[596] LISA Collaboration, P. Amaro-Seoane et al., “Laser Interferometer Space Antenna,” arXiv:1702.00786
[astro-ph.IM].
[597] B. F. Schutz, “Determining the Hubble Constant from Gravitational Wave Observations,” Nature 323 (1986) 310–311.
[598] D. E. Holz and S. A. Hughes, “Using gravitational-wave standard sirens,” Astrophys. J. 629 (2005) 15–22,
arXiv:astro-ph/0504616.
[599] LIGO Scientific, Virgo, Fermi GBM, INTEGRAL, IceCube, AstroSat Cadmium Zinc Telluride Imager
Team, IPN, Insight-Hxmt, ANTARES, Swift, AGILE Team, 1M2H Team, Dark Energy Camera
GW-EM, DES, DLT40, GRAWITA, Fermi-LAT, ATCA, ASKAP, Las Cumbres Observatory Group,
OzGrav, DWF (Deeper Wider Faster Program), AST3, CAASTRO, VINROUGE, MASTER, J-GEM,
GROWTH, JAGWAR, CaltechNRAO, TTU-NRAO, NuSTAR, Pan-STARRS, MAXI Team, TZAC
Consortium, KU, Nordic Optical Telescope, ePESSTO, GROND, Texas Tech University, SALT Group,
TOROS, BOOTES, MWA, CALET, IKI-GW Follow-up, H.E.S.S., LOFAR, LWA, HAWC, Pierre
Auger, ALMA, Euro VLBI Team, Pi of Sky, Chandra Team at McGill University, DFN, ATLAS
Telescopes, High Time Resolution Universe Survey, RIMAS, RATIR, SKA South Africa/MeerKAT
Collaboration, B. P. Abbott et al., “Multi-messenger Observations of a Binary Neutron Star Merger,” Astrophys. J.
Lett. 848 (2017) no. 2, L12, arXiv:1710.05833 [astro-ph.HE].
[600] LIGO Scientific, Virgo, 1M2H, Dark Energy Camera GW-E, DES, DLT40, Las Cumbres Observatory,
VINROUGE, MASTER Collaboration, B. P. Abbott et al., “A gravitational-wave standard siren measurement of
the Hubble constant,” Nature 551 (2017) no. 7678, 85–88, arXiv:1710.05835 [astro-ph.CO].
[601] C. Nicolaou, O. Lahav, P. Lemos, W. Hartley, and J. Braden, “The Impact of Peculiar Velocities on the Estimation of
the Hubble Constant from Gravitational Wave Standard Sirens,” Mon. Not. Roy. Astron. Soc. 495 (2020) no. 1, 90–97,
arXiv:1909.09609 [astro-ph.CO].
[602] S. Mukherjee, G. Lavaux, F. R. Bouchet, J. Jasche, B. D. Wandelt, S. M. Nissanke, F. Leclercq, and K. Hotokezaka,
“Velocity correction for Hubble constant measurements from standard sirens,” Astron. Astrophys. 646 (2021) A65,
arXiv:1909.08627 [astro-ph.CO].
[603] C. Howlett and T. M. Davis, “Standard siren speeds: improving velocities in gravitational-wave measurements of H0 ,”
Mon. Not. Roy. Astron. Soc. 492 (2020) no. 3, 3803–3815, arXiv:1909.00587 [astro-ph.CO].
[604] C. Guidorzi et al., “Improved Constraints on H0 from a Combined Analysis of Gravitational-wave and Electromagnetic
Emission from GW170817,” Astrophys. J. Lett. 851 (2017) no. 2, L36, arXiv:1710.06426 [astro-ph.CO].
[605] K. Hotokezaka, E. Nakar, O. Gottlieb, S. Nissanke, K. Masuda, G. Hallinan, K. P. Mooley, and A. T. Deller, “A
Hubble constant measurement from superluminal motion of the jet in GW170817,” Nature Astron. 3 (2019) no. 10,
940–944, arXiv:1806.10596 [astro-ph.CO].
[606] S. Dhawan, M. Bulla, A. Goobar, A. S. Carracedo, and C. N. Setzer, “Constraining the observer angle of the kilonova
AT2017gfo associated with GW170817: Implications for the Hubble constant,” Astrophys. J. 888 (2020) no. 2, 67,
arXiv:1909.13810 [astro-ph.HE].
[607] A. Palmese, R. Kaur, A. Hajela, R. Margutti, A. McDowell, and A. MacFadyen, “Standard siren measurement of the
Hubble constant using GW170817 and the latest observations of the electromagnetic counterpart afterglow,” Phys.
Rev. D 109 (2024) no. 6, 063508, arXiv:2305.19914 [astro-ph.CO].
[608] A. Palmese, O. Graur, J. T. Annis, S. BenZvi, E. Di Valentino, J. Garcia-Bellido, S. G. A. Gontcho, R. Keeley,
A. Kim, O. Lahav, S. Nissanke, K. Paterson, M. Sako, A. Shafieloo, and Y.-D. Tsai, “Gravitational wave cosmology
and astrophysics with large spectroscopic galaxy surveys,” Bulletin of the AAS 51 (2019) no. 3, 310,
arXiv:1903.04730 [astro-ph.CO].
[609] H.-Y. Chen, “Systematic Uncertainty of Standard Sirens from the Viewing Angle of Binary Neutron Star Inspirals,”
Phys. Rev. Lett. 125 (2020) no. 20, 201301, arXiv:2006.02779 [astro-ph.HE].
[610] H.-Y. Chen, C. Talbot, and E. A. Chase, “Mitigating the Counterpart Selection Effect for Standard Sirens,” Phys. Rev.
Lett. 132 (2024) no. 19, 191003, arXiv:2307.10402 [astro-ph.CO].
[611] G. Gianfagna, L. Piro, F. Pannarale, H. Van Eerten, F. Ricci, and G. Ryan, “Potential biases and prospects for the
Hubble constant estimation via electromagnetic and gravitational-wave joint analyses,” Mon. Not. Roy. Astron. Soc.
278

528 (2024) no. 2, 2600–2613, arXiv:2309.17073 [astro-ph.HE].


[612] J. Heinzel, M. W. Coughlin, T. Dietrich, M. Bulla, S. Antier, N. Christensen, D. A. Coulter, R. J. Foley, L. Issa, and
N. Khetan, “Comparing inclination dependent analyses of kilonova transients,” Mon. Not. Roy. Astron. Soc. 502
(2021) no. 2, 3057–3065, arXiv:2010.10746 [astro-ph.HE].
[613] M. Müller, S. Mukherjee, and G. Ryan, “Be Careful in Multimessenger Inference of the Hubble Constant: A Path
Forward for Robust Inference,” Astrophys. J. Lett. 977 (2024) no. 2, L45, arXiv:2406.11965 [astro-ph.CO].
[614] S. Vitale and H.-Y. Chen, “Measuring the Hubble constant with neutron star black hole mergers,” Phys. Rev. Lett. 121
(2018) no. 2, 021303, arXiv:1804.07337 [astro-ph.CO].
[615] M. J. Graham et al., “Candidate Electromagnetic Counterpart to the Binary Black Hole Merger Gravitational Wave
Event S190521g,” Phys. Rev. Lett. 124 (2020) no. 25, 251102, arXiv:2006.14122 [astro-ph.HE].
[616] M. J. Graham et al., “A Light in the Dark: Searching for Electromagnetic Counterparts to Black Hole–Black Hole
Mergers in LIGO/Virgo O3 with the Zwicky Transient Facility,” Astrophys. J. 942 (2023) no. 2, 99, arXiv:2209.13004
[astro-ph.HE].
[617] T. Cabrera et al., “Searching for electromagnetic emission in an AGN from the gravitational wave binary black hole
merger candidate S230922g,” Phys. Rev. D 110 (2024) no. 12, 123029, arXiv:2407.10698 [astro-ph.HE].
[618] G. Ashton, K. Ackley, I. M. n. Hernandez, and B. Piotrzkowski, “Current observations are insufficient to confidently
associate the binary black hole merger GW190521 with AGN J124942.3 + 344929,” Class. Quant. Grav. 38 (2021)
no. 23, 235004, arXiv:2009.12346 [astro-ph.HE].
[619] A. Palmese, M. Fishbach, C. J. Burke, J. T. Annis, and X. Liu, “Do LIGO/Virgo Black Hole Mergers Produce AGN
Flares? The Case of GW190521 and Prospects for Reaching a Confident Association,” Astrophys. J. Lett. 914 (2021)
no. 2, L34, arXiv:2103.16069 [astro-ph.HE].
[620] S. Mukherjee, A. Ghosh, M. J. Graham, C. Karathanasis, M. M. Kasliwal, I. Magaña Hernandez, S. M. Nissanke,
A. Silvestri, and B. D. Wandelt, “First measurement of the Hubble parameter from bright binary black hole
GW190521,” arXiv:2009.14199 [astro-ph.CO].
[621] V. Gayathri, J. Healy, J. Lange, B. O’Brien, M. Szczepanczyk, I. Bartos, M. Campanelli, S. Klimenko, C. O. Lousto,
and R. O’Shaughnessy, “Measuring the Hubble Constant with GW190521 as an Eccentric black hole Merger and Its
Potential Electromagnetic Counterpart,” Astrophys. J. Lett. 908 (2021) no. 2, L34, arXiv:2009.14247 [astro-ph.HE].
[622] H.-Y. Chen, C.-J. Haster, S. Vitale, W. M. Farr, and M. Isi, “A standard siren cosmological measurement from the
potential GW190521 electromagnetic counterpart ZTF19abanrhr,” Mon. Not. Roy. Astron. Soc. 513 (2022) no. 2,
2152–2157, arXiv:2009.14057 [astro-ph.CO].
[623] C. R. Bom and A. Palmese, “Standard siren cosmology with gravitational waves from binary black hole mergers in
active galactic nuclei,” Phys. Rev. D 110 (2024) no. 8, 083005, arXiv:2307.01330 [astro-ph.CO].
[624] L. M. B. Alves, A. G. Sullivan, Y. Yang, G. V., Z. Marka, S. Marka, and I. Bartos, “Determining the Hubble constant
with AGN-assisted black hole mergers,” Mon. Not. Roy. Astron. Soc. 531 (2024) no. 3, 3679–3683, arXiv:2009.13739
[astro-ph.HE].
[625] J. R. Gair et al., “The Hitchhiker’s Guide to the Galaxy Catalog Approach for Dark Siren Gravitational-wave
Cosmology,” Astron. J. 166 (2023) no. 1, 22, arXiv:2212.08694 [gr-qc].
[626] R. Gray et al., “Cosmological inference using gravitational wave standard sirens: A mock data analysis,” Phys. Rev. D
101 (2020) no. 12, 122001, arXiv:1908.06050 [gr-qc].
[627] W. M. Farr, M. Fishbach, J. Ye, and D. Holz, “A Future Percent-Level Measurement of the Hubble Expansion at
Redshift 0.8 With Advanced LIGO,” Astrophys. J. Lett. 883 (2019) no. 2, L42, arXiv:1908.09084 [astro-ph.CO].
[628] J. M. Ezquiaga and D. E. Holz, “Spectral Sirens: Cosmology from the Full Mass Distribution of Compact Binaries,”
Phys. Rev. Lett. 129 (2022) no. 6, 061102, arXiv:2202.08240 [astro-ph.CO].
[629] S. Mastrogiovanni, D. Laghi, R. Gray, G. C. Santoro, A. Ghosh, C. Karathanasis, K. Leyde, D. A. Steer, S. Perries,
and G. Pierra, “Joint population and cosmological properties inference with gravitational waves standard sirens and
galaxy surveys,” Phys. Rev. D 108 (2023) no. 4, 042002, arXiv:2305.10488 [astro-ph.CO].
[630] R. Gray et al., “Joint cosmological and gravitational-wave population inference using dark sirens and galaxy
catalogues,” JCAP 12 (2023) 023, arXiv:2308.02281 [astro-ph.CO].
[631] N. Borghi, M. Mancarella, M. Moresco, M. Tagliazucchi, F. Iacovelli, A. Cimatti, and M. Maggiore, “Cosmology and
Astrophysics with Standard Sirens and Galaxy Catalogs in View of Future Gravitational Wave Observations,”
Astrophys. J. 964 (2024) no. 2, 191, arXiv:2312.05302 [astro-ph.CO].
[632] DES, LIGO Scientific, Virgo Collaboration, M. Soares-Santos et al., “First Measurement of the Hubble Constant
from a Dark Standard Siren using the Dark Energy Survey Galaxies and the LIGO/Virgo Binary–Black-hole Merger
GW170814,” Astrophys. J. Lett. 876 (2019) no. 1, L7, arXiv:1901.01540 [astro-ph.CO].
[633] LIGO Scientific, Virgo, VIRGO Collaboration, B. P. Abbott et al., “A Gravitational-wave Measurement of the
Hubble Constant Following the Second Observing Run of Advanced LIGO and Virgo,” Astrophys. J. 909 (2021) no. 2,
218, arXiv:1908.06060 [astro-ph.CO].
[634] LIGO Scientific, Virgo Collaboration, R. Abbott et al., “GW190814: Gravitational Waves from the Coalescence of a
23 Solar Mass Black Hole with a 2.6 Solar Mass Compact Object,” Astrophys. J. Lett. 896 (2020) no. 2, L44,
arXiv:2006.12611 [astro-ph.HE].
[635] LIGO Scientific, Virgo, KAGRA Collaboration, R. Abbott et al., “Constraints on the Cosmic Expansion History
from GWTC–3,” Astrophys. J. 949 (2023) no. 2, 76, arXiv:2111.03604 [astro-ph.CO].
[636] A. Palmese, C. R. Bom, S. Mucesh, and W. G. Hartley, “A Standard Siren Measurement of the Hubble Constant Using
Gravitational-wave Events from the First Three LIGO/Virgo Observing Runs and the DESI Legacy Survey,”
279

Astrophys. J. 943 (2023) no. 1, 56, arXiv:2111.06445 [astro-ph.CO].


[637] DESI Collaboration, W. Ballard et al., “A Dark Siren Measurement of the Hubble Constant with the LIGO/Virgo
Gravitational Wave Event GW190412 and DESI Galaxies,” Res. Notes AAS 7 (2023) no. 11, 250, arXiv:2311.13062
[astro-ph.CO].
[638] V. Alfradique et al., “A dark siren measurement of the Hubble constant using gravitational wave events from the first
three LIGO/Virgo observing runs and DELVE,” Mon. Not. Roy. Astron. Soc. 528 (2024) no. 2, 3249–3259,
arXiv:2310.13695 [astro-ph.CO].
[639] C. R. Bom, V. Alfradique, A. Palmese, G. Teixeira, L. Santana-Silva, A. Santos, and P. Darc, “A dark standard siren
measurement of the Hubble constant following LIGO/Virgo/KAGRA O4a and previous runs,” MNRAS 535 (2024)
no. 1, 961–975, arXiv:2404.16092 [astro-ph.CO].
[640] A. Finke, S. Foffa, F. Iacovelli, M. Maggiore, and M. Mancarella, “Cosmology with LIGO/Virgo dark sirens: Hubble
parameter and modified gravitational wave propagation,” JCAP 08 (2021) 026, arXiv:2101.12660 [astro-ph.CO].
[641] S. Borhanian, A. Dhani, A. Gupta, K. G. Arun, and B. S. Sathyaprakash, “Dark Sirens to Resolve the
Hubble–Lemaître Tension,” Astrophys. J. Lett. 905 (2020) no. 2, L28, arXiv:2007.02883 [astro-ph.CO].
[642] C. C. Diaz and S. Mukherjee, “Mapping the cosmic expansion history from LIGO-Virgo-KAGRA in synergy with DESI
and SPHEREx,” Mon. Not. Roy. Astron. Soc. 511 (2022) no. 2, 2782–2795, arXiv:2107.12787 [astro-ph.CO].
[643] J. R. Bond, W. D. Arnett, and B. J. Carr, “The Evolution and fate of Very Massive Objects,” Astrophys. J. 280 (1984)
825–847.
[644] M. Zevin, S. S. Bavera, C. P. L. Berry, V. Kalogera, T. Fragos, P. Marchant, C. L. Rodriguez, F. Antonini, D. E. Holz,
and C. Pankow, “One Channel to Rule Them All? Constraining the Origins of Binary Black Holes Using Multiple
Formation Pathways,” Astrophys. J. 910 (2021) no. 2, 152, arXiv:2011.10057 [astro-ph.HE].
[645] KAGRA, VIRGO, LIGO Scientific Collaboration, R. Abbott et al., “Population of Merging Compact Binaries
Inferred Using Gravitational Waves through GWTC-3,” Phys. Rev. X 13 (2023) no. 1, 011048, arXiv:2111.03634
[astro-ph.HE].
[646] I. Magaña Hernandez and A. Palmese, “A new bump in the night: evidence of a new feature in the binary black hole
mass distribution at 70 M⊙ from gravitational-wave observations,” arXiv:2407.02460 [astro-ph.HE].
[647] M. Mancarella, E. Genoud-Prachex, and M. Maggiore, “Cosmology and modified gravitational wave propagation from
binary black hole population models,” Phys. Rev. D 105 (2022) no. 6, 064030, arXiv:2112.05728 [gr-qc].
[648] K. Leyde, S. Mastrogiovanni, D. A. Steer, E. Chassande-Mottin, and C. Karathanasis, “Current and future constraints
on cosmology and modified gravitational wave friction from binary black holes,” JCAP 09 (2022) 012,
arXiv:2202.00025 [gr-qc].
[649] DES Collaboration, A. Palmese et al., “A statistical standard siren measurement of the Hubble constant from the
LIGO/Virgo gravitational wave compact object merger GW190814 and Dark Energy Survey galaxies,” Astrophys. J.
Lett. 900 (2020) no. 2, L33, arXiv:2006.14961 [astro-ph.CO].
[650] C. Turski, M. Bilicki, G. Dálya, R. Gray, and A. Ghosh, “Impact of modelling galaxy redshift uncertainties on the
gravitational-wave dark standard siren measurement of the Hubble constant,” Mon. Not. Roy. Astron. Soc. 526 (2023)
no. 4, 6224–6233, arXiv:2302.12037 [gr-qc].
[651] G. Perna, S. Mastrogiovanni, and A. Ricciardone, “Investigating the impact of galaxies’ compact binary hosting
probability for gravitational-wave cosmology,” arXiv:2405.07904 [astro-ph.CO].
[652] A. G. Hanselman, A. Vijaykumar, M. Fishbach, and D. E. Holz, “Gravitational-wave Dark Siren Cosmology
Systematics from Galaxy Weighting,” Astrophys. J. 979 (2025) no. 1, 9, arXiv:2405.14818 [astro-ph.CO].
[653] H.-Y. Chen, J. M. Ezquiaga, and I. Gupta, “Cosmography with next-generation gravitational wave detectors,” Class.
Quant. Grav. 41 (2024) no. 12, 125004, arXiv:2402.03120 [gr-qc].
[654] THESEUS Collaboration, L. Amati et al., “The THESEUS space mission: science goals, requirements and mission
concept,” Exper. Astron. 52 (2021) no. 3, 183–218, arXiv:2104.09531 [astro-ph.IM].
[655] M. Califano, I. de Martino, D. Vernieri, and S. Capozziello, “Constraining ΛCDM cosmological parameters with
Einstein Telescope mock data,” Mon. Not. Roy. Astron. Soc. 518 (2023) 3372–3385, arXiv:2205.11221
[astro-ph.CO].
[656] M. Califano, I. de Martino, D. Vernieri, and S. Capozziello, “Exploiting the Einstein Telescope to solve the Hubble
tension,” Phys. Rev. D 107 (2023) no. 12, 123519, arXiv:2208.13999 [astro-ph.CO].
[657] A. Palmese and A. G. Kim, “Probing gravity and growth of structure with gravitational waves and galaxies’ peculiar
velocity,” Phys. Rev. D 103 (2021) no. 10, 103507, arXiv:2005.04325 [astro-ph.CO].
[658] LISA Collaboration, M. Colpi et al., “LISA Definition Study Report,” arXiv:2402.07571 [astro-ph.CO].
[659] LISA Cosmology Working Group Collaboration, P. Auclair et al., “Cosmology with the Laser Interferometer
Space Antenna,” Living Rev. Rel. 26 (2023) no. 1, 5, arXiv:2204.05434 [astro-ph.CO].
[660] A. Klein et al., “Science with the space-based interferometer eLISA: Supermassive black hole binaries,” Phys. Rev. D
93 (2016) no. 2, 024003, arXiv:1511.05581 [gr-qc].
[661] A. Mangiagli, C. Caprini, M. Volonteri, S. Marsat, S. Vergani, N. Tamanini, and H. Inchauspé, “Massive black hole
binaries in LISA: Multimessenger prospects and electromagnetic counterparts,” Phys. Rev. D 106 (2022) no. 10,
103017, arXiv:2207.10678 [astro-ph.HE].
[662] S. Babak, J. Gair, A. Sesana, E. Barausse, C. F. Sopuerta, C. P. L. Berry, E. Berti, P. Amaro-Seoane, A. Petiteau, and
A. Klein, “Science with the space-based interferometer LISA. V: Extreme mass-ratio inspirals,” Phys. Rev. D 95 (2017)
no. 10, 103012, arXiv:1703.09722 [gr-qc].
[663] N. Tamanini, C. Caprini, E. Barausse, A. Sesana, A. Klein, and A. Petiteau, “Science with the space-based
280

interferometer eLISA. III: Probing the expansion of the Universe using gravitational wave standard sirens,” JCAP 04
(2016) 002, arXiv:1601.07112 [astro-ph.CO].
[664] LISA Cosmology Working Group Collaboration, E. Belgacem et al., “Testing modified gravity at cosmological
distances with LISA standard sirens,” JCAP 07 (2019) 024, arXiv:1906.01593 [astro-ph.CO].
[665] A. Mangiagli, C. Caprini, S. Marsat, L. Speri, R. R. Caldwell, and N. Tamanini, “Massive black hole binaries in LISA:
constraining cosmological parameters at high redshifts,” arXiv:2312.04632 [astro-ph.CO].
[666] C. L. MacLeod and C. J. Hogan, “Precision of Hubble constant derived using black hole binary absolute distances and
statistical redshift information,” Phys. Rev. D 77 (2008) 043512, arXiv:0712.0618 [astro-ph].
[667] D. Laghi, N. Tamanini, W. Del Pozzo, A. Sesana, J. Gair, S. Babak, and D. Izquierdo-Villalba, “Gravitational-wave
cosmology with extreme mass-ratio inspirals,” Mon. Not. Roy. Astron. Soc. 508 (2021) no. 3, 4512–4531,
arXiv:2102.01708 [astro-ph.CO].
[668] M. Toscani, O. Burke, C. Liu, N. B. Zamel, N. Tamanini, and F. Pozzoli, “Strongly lensed extreme mass-ratio
inspirals,” Phys. Rev. D 109 (2024) no. 6, 063505, arXiv:2307.06722 [astro-ph.CO].
[669] N. Tamanini, “Late time cosmology with LISA: probing the cosmic expansion with massive black hole binary mergers
as standard sirens,” J. Phys. Conf. Ser. 840 (2017) no. 1, 012029, arXiv:1612.02634 [astro-ph.CO].
[670] C. Caprini and N. Tamanini, “Constraining early and interacting dark energy with gravitational wave standard sirens:
the potential of the eLISA mission,” JCAP 10 (2016) 006, arXiv:1607.08755 [astro-ph.CO].
[671] R.-G. Cai, N. Tamanini, and T. Yang, “Reconstructing the dark sector interaction with LISA,” JCAP 05 (2017) 031,
arXiv:1703.07323 [astro-ph.CO].
[672] M. Corman, A. Ghosh, C. Escamilla-Rivera, M. A. Hendry, S. Marsat, and N. Tamanini, “Constraining cosmological
extra dimensions with gravitational wave standard sirens: From theory to current and future multimessenger
observations,” Phys. Rev. D 105 (2022) no. 6, 064061, arXiv:2109.08748 [gr-qc].
[673] L. Speri, N. Tamanini, R. R. Caldwell, J. R. Gair, and B. Wang, “Testing the Quasar Hubble Diagram with LISA
Standard Sirens,” Phys. Rev. D 103 (2021) no. 8, 083526, arXiv:2010.09049 [astro-ph.CO].
[674] C. Liu, D. Laghi, and N. Tamanini, “Probing modified gravitational-wave propagation with extreme mass-ratio
inspirals,” Phys. Rev. D 109 (2024) no. 6, 063521, arXiv:2310.12813 [astro-ph.CO].
[675] D. J. Fixsen, “The Temperature of the Cosmic Microwave Background,” Astrophys. J. 707 (2009) 916–920,
arXiv:0911.1955 [astro-ph.CO].
[676] COBE Collaboration, G. F. Smoot et al., “Structure in the COBE differential microwave radiometer first year maps,”
Astrophys. J. Lett. 396 (1992) L1–L5.
[677] J. C. Mather et al., “A Preliminary measurement of the Cosmic Microwave Background spectrum by the Cosmic
Background Explorer (COBE) satellite,” Astrophys. J. Lett. 354 (1990) L37–L40.
[678] WMAP Collaboration, C. L. Bennett et al., “First year Wilkinson Microwave Anisotropy Probe (WMAP)
observations: Preliminary maps and basic results,” Astrophys. J. Suppl. 148 (2003) 1–27, arXiv:astro-ph/0302207.
[679] LiteBIRD Collaboration, E. Allys et al., “Probing Cosmic Inflation with the LiteBIRD Cosmic Microwave
Background Polarization Survey,” PTEP 2023 (2023) no. 4, 042F01, arXiv:2202.02773 [astro-ph.IM].
[680] Simons Observatory Collaboration, P. Ade et al., “The Simons Observatory: Science goals and forecasts,” JCAP 02
(2019) 056, arXiv:1808.07445 [astro-ph.CO].
[681] CMB-S4 Collaboration, K. N. Abazajian et al., “CMB-S4 Science Book, First Edition,” arXiv:1610.02743
[astro-ph.CO].
[682] M. Kamionkowski and E. D. Kovetz, “The Quest for B Modes from Inflationary Gravitational Waves,” Ann. Rev.
Astron. Astrophys. 54 (2016) 227–269, arXiv:1510.06042 [astro-ph.CO].
[683] M. C. Guzzetti, N. Bartolo, M. Liguori, and S. Matarrese, “Gravitational waves from inflation,” Riv. Nuovo Cim. 39
(2016) no. 9, 399–495, arXiv:1605.01615 [astro-ph.CO].
[684] M. Gerbino, A. Gruppuso, P. Natoli, M. Shiraishi, and A. Melchiorri, “Testing chirality of primordial gravitational
waves with Planck and future CMB data: no hope from angular power spectra,” JCAP 07 (2016) 044,
arXiv:1605.09357 [astro-ph.CO].
[685] BICEP2, Keck Arrary Collaboration, P. A. R. Ade et al., “BICEP2 / Keck Array IX: New bounds on anisotropies
of CMB polarization rotation and implications for axionlike particles and primordial magnetic fields,” Phys. Rev. D 96
(2017) no. 10, 102003, arXiv:1705.02523 [astro-ph.CO].
[686] L. Pogosian and A. Zucca, “Searching for Primordial Magnetic Fields with CMB B-modes,” Class. Quant. Grav. 35
(2018) no. 12, 124004, arXiv:1801.08936 [astro-ph.CO].
[687] N. Bartolo, G. Orlando, and M. Shiraishi, “Measuring chiral gravitational waves in Chern-Simons gravity with CMB
bispectra,” JCAP 01 (2019) 050, arXiv:1809.11170 [astro-ph.CO].
[688] Y. Minami and E. Komatsu, “New Extraction of the Cosmic Birefringence from the Planck 2018 Polarization Data,”
Phys. Rev. Lett. 125 (2020) no. 22, 221301, arXiv:2011.11254 [astro-ph.CO].
[689] T. Namikawa et al., “Atacama Cosmology Telescope: Constraints on cosmic birefringence,” Phys. Rev. D 101 (2020)
no. 8, 083527, arXiv:2001.10465 [astro-ph.CO].
[690] G. Choi, W. Lin, L. Visinelli, and T. T. Yanagida, “Cosmic birefringence and electroweak axion dark energy,” Phys.
Rev. D 104 (2021) no. 10, L101302, arXiv:2106.12602 [hep-ph].
[691] A. Greco, N. Bartolo, and A. Gruppuso, “Cosmic birefrigence: cross-spectra and cross-bispectra with CMB
anisotropies,” JCAP 03 (2022) no. 03, 050, arXiv:2202.04584 [astro-ph.CO].
[692] E. Komatsu, “New physics from the polarized light of the cosmic microwave background,” Nature Rev. Phys. 4 (2022)
no. 7, 452–469, arXiv:2202.13919 [astro-ph.CO].
281

[693] Planck Collaboration, N. Aghanim et al., “Planck 2018 results. VIII. Gravitational lensing,” Astron. Astrophys. 641
(2020) A8, arXiv:1807.06210 [astro-ph.CO].
[694] ACT Collaboration, M. S. Madhavacheril et al., “The Atacama Cosmology Telescope: DR6 Gravitational Lensing
Map and Cosmological Parameters,” Astrophys. J. 962 (2024) no. 2, 113, arXiv:2304.05203 [astro-ph.CO].
[695] SPT-3G Collaboration, L. Balkenhol et al., “Constraints on ΛCDM extensions from the SPT-3G 2018 EE and TE
power spectra,” Phys. Rev. D 104 (2021) no. 8, 083509, arXiv:2103.13618 [astro-ph.CO].
[696] S. Joudaki et al., “KiDS-450: Testing extensions to the standard cosmological model,” Mon. Not. Roy. Astron. Soc.
471 (2017) no. 2, 1259–1279, arXiv:1610.04606 [astro-ph.CO].
[697] M. Asgari et al., “KiDS+VIKING-450 and DES-Y1 combined: Mitigating baryon feedback uncertainty with
COSEBIs,” Astron. Astrophys. 634 (2020) A127, arXiv:1910.05336 [astro-ph.CO].
[698] DESI Collaboration, A. G. Adame et al., “DESI 2024 VI: cosmological constraints from the measurements of baryon
acoustic oscillations,” JCAP 02 (2025) 021, arXiv:2404.03002 [astro-ph.CO].
[699] E. Rosenberg, S. Gratton, and G. Efstathiou, “CMB power spectra and cosmological parameters from Planck PR4
with CamSpec,” Mon. Not. Roy. Astron. Soc. 517 (2022) no. 3, 4620–4636, arXiv:2205.10869 [astro-ph.CO].
[700] M. Tristram et al., “Cosmological parameters derived from the final Planck data release (PR4),” Astron. Astrophys.
682 (2024) A37, arXiv:2309.10034 [astro-ph.CO].
[701] ACT Collaboration, S. Aiola et al., “The Atacama Cosmology Telescope: DR4 Maps and Cosmological Parameters,”
JCAP 12 (2020) 047, arXiv:2007.07288 [astro-ph.CO].
[702] T. Louis et al., “The Atacama Cosmology Telescope: DR6 Power Spectra, Likelihoods and ΛCDM Parameters,”
arXiv:2503.14452 [astro-ph.CO].
[703] SPT-3G Collaboration, L. Balkenhol et al., “Measurement of the CMB temperature power spectrum and constraints
on cosmology from the SPT-3G 2018 TT, TE, and EE dataset,” Phys. Rev. D 108 (2023) no. 2, 023510,
arXiv:2212.05642 [astro-ph.CO].
[704] N. Schöneberg, J. Lesgourgues, and D. C. Hooper, “The BAO+BBN take on the Hubble tension,” JCAP 10 (2019)
029, arXiv:1907.11594 [astro-ph.CO].
[705] E. Calabrese et al., “The Atacama Cosmology Telescope: DR6 Constraints on Extended Cosmological Models,”
arXiv:2503.14454 [astro-ph.CO].
[706] SPT-3G Collaboration, F. Ge et al., “Cosmology From CMB Lensing and Delensed EE Power Spectra Using
2019-2020 SPT-3G Polarization Data,” arXiv:2411.06000 [astro-ph.CO].
[707] SPT-3G Collaboration, K. Prabhu et al., “Testing the ΛCDM Cosmological Model with Forthcoming Measurements
of the Cosmic Microwave Background with SPT-3G,” Astrophys. J. 973 (2024) no. 1, 4, arXiv:2403.17925
[astro-ph.CO].
[708] E. Di Valentino, O. Mena, S. Pan, L. Visinelli, W. Yang, A. Melchiorri, D. F. Mota, A. G. Riess, and J. Silk, “In the
realm of the Hubble tension—a review of solutions,” Class. Quant. Grav. 38 (2021) no. 15, 153001, arXiv:2103.01183
[astro-ph.CO].
[709] N. Schöneberg, G. Franco Abellán, A. Pérez Sánchez, S. J. Witte, V. Poulin, and J. Lesgourgues, “The H0 Olympics:
A fair ranking of proposed models,” Phys. Rept. 984 (2022) 1–55, arXiv:2107.10291 [astro-ph.CO].
[710] J. M. Berryman et al., “Neutrino self-interactions: A white paper,” Phys. Dark Univ. 42 (2023) 101267,
arXiv:2203.01955 [hep-ph].
[711] J. L. Bernal, L. Verde, and A. G. Riess, “The trouble with H0 ,” JCAP 10 (2016) 019, arXiv:1607.05617
[astro-ph.CO].
[712] L. Knox and M. Millea, “Hubble constant hunter’s guide,” Phys. Rev. D 101 (2020) no. 4, 043533, arXiv:1908.03663
[astro-ph.CO].
[713] G. Efstathiou and J. R. Bond, “Cosmic confusion: Degeneracies among cosmological parameters derived from
measurements of microwave background anisotropies,” Mon. Not. Roy. Astron. Soc. 304 (1999) 75–97,
arXiv:astro-ph/9807103.
[714] Planck Collaboration, N. Aghanim et al., “Planck intermediate results. LI. Features in the cosmic microwave
background temperature power spectrum and shifts in cosmological parameters,” Astron. Astrophys. 607 (2017) A95,
arXiv:1608.02487 [astro-ph.CO].
[715] D. J. Eisenstein, H.-j. Seo, and M. J. White, “On the Robustness of the Acoustic Scale in the Low-Redshift Clustering
of Matter,” Astrophys. J. 664 (2007) 660–674, arXiv:astro-ph/0604361.
[716] DESI Collaboration, M. A. Karim et al., “DESI DR2 Results II: Measurements of Baryon Acoustic Oscillations and
Cosmological Constraints,” arXiv:2503.14738 [astro-ph.CO].
[717] O. H. E. Philcox, G. S. Farren, B. D. Sherwin, E. J. Baxter, and D. J. Brout, “Determining the Hubble constant
without the sound horizon: A 3.6% constraint on H0 from galaxy surveys, CMB lensing, and supernovae,” Phys. Rev.
D 106 (2022) no. 6, 063530, arXiv:2204.02984 [astro-ph.CO].
[718] R. Zhao et al., “A multitracer analysis for the eBOSS galaxy sample based on the effective field theory of large-scale
structure,” Mon. Not. Roy. Astron. Soc. 532 (2024) no. 1, 783–804, arXiv:2308.06206 [astro-ph.CO].
[719] G. D’Amico, J. Gleyzes, N. Kokron, K. Markovic, L. Senatore, P. Zhang, F. Beutler, and H. Gil-Marín, “The
Cosmological Analysis of the SDSS/BOSS data from the Effective Field Theory of Large-Scale Structure,” JCAP 05
(2020) 005, arXiv:1909.05271 [astro-ph.CO].
[720] SimBIG Collaboration, C. Hahn, M. Eickenberg, S. Ho, J. Hou, P. Lemos, E. Massara, C. Modi,
A. Moradinezhad Dizgah, L. Parker, and B. R.-S. Blancard, “Cosmological constraints from the nonlinear galaxy
bispectrum,” Phys. Rev. D 109 (2024) no. 8, 083534, arXiv:2310.15243 [astro-ph.CO].
282

[721] G. D’Amico, Y. Donath, M. Lewandowski, L. Senatore, and P. Zhang, “The BOSS bispectrum analysis at one loop
from the Effective Field Theory of Large-Scale Structure,” JCAP 05 (2024) 059, arXiv:2206.08327 [astro-ph.CO].
[722] J. de Cruz Perez, C.-G. Park, and B. Ratra, “Updated observational constraints on spatially flat and nonflat ΛCDM
and XCDM cosmological models,” Phys. Rev. D 110 (2024) no. 2, 023506, arXiv:2404.19194 [astro-ph.CO].
[723] E. Calabrese, A. Slosar, A. Melchiorri, G. F. Smoot, and O. Zahn, “Cosmic Microwave Weak lensing data as a test for
the dark universe,” Phys. Rev. D 77 (2008) 123531, arXiv:0803.2309 [astro-ph].
[724] SPT-3G Collaboration, D. Dutcher et al., “Measurements of the E-mode polarization and temperature-E-mode
correlation of the CMB from SPT-3G 2018 data,” Phys. Rev. D 104 (2021) no. 2, 022003, arXiv:2101.01684
[astro-ph.CO].
[725] Planck Collaboration, N. Aghanim et al., “Planck 2015 results. XI. CMB power spectra, likelihoods, and robustness of
parameters,” Astron. Astrophys. 594 (2016) A11, arXiv:1507.02704 [astro-ph.CO].
[726] F. Couchot, S. Henrot-Versillé, O. Perdereau, S. Plaszczynski, B. Rouillé d’Orfeuil, M. Spinelli, and M. Tristram,
“Relieving tensions related to the lensing of the cosmic microwave background temperature power spectra,” Astron.
Astrophys. 597 (2017) A126, arXiv:1510.07600 [astro-ph.CO].
[727] G. E. Addison, Y. Huang, D. J. Watts, C. L. Bennett, M. Halpern, G. Hinshaw, and J. L. Weiland, “Quantifying
discordance in the 2015 Planck CMB spectrum,” Astrophys. J. 818 (2016) no. 2, 132, arXiv:1511.00055
[astro-ph.CO].
[728] Planck Collaboration, N. Aghanim et al., “Planck 2018 results. V. CMB power spectra and likelihoods,” Astron.
Astrophys. 641 (2020) A5, arXiv:1907.12875 [astro-ph.CO].
[729] Planck Collaboration, Y. Akrami et al., “Planck 2018 results. II. Low Frequency Instrument data processing,” Astron.
Astrophys. 641 (2020) A2, arXiv:1807.06206 [astro-ph.CO].
[730] Planck Collaboration, N. Aghanim et al., “Planck 2018 results. III. High Frequency Instrument data processing and
frequency maps,” Astron. Astrophys. 641 (2020) A3, arXiv:1807.06207 [astro-ph.CO].
[731] ACT Collaboration, S. K. Choi et al., “The Atacama Cosmology Telescope: a measurement of the Cosmic Microwave
Background power spectra at 98 and 150 GHz,” JCAP 12 (2020) 045, arXiv:2007.07289 [astro-ph.CO].
[732] S. Giardiello, A. J. Duivenvoorden, E. Calabrese, G. Galloni, M. Hasselfield, J. C. Hill, A. La Posta, T. Louis,
M. Madhavacheril, and L. Pagano, “Modeling beam chromaticity for high-resolution CMB analyses,” Phys. Rev. D 111
(2025) no. 4, 043502, arXiv:2411.10124 [astro-ph.CO].
[733] S. Giardiello et al., “The Simons Observatory: impact of bandpass, polarization angle and calibration uncertainties on
small-scale power spectrum analysis,” JCAP 09 (2024) 008, arXiv:2403.05242 [astro-ph.CO].
[734] E. Di Valentino and S. Bridle, “Exploring the Tension between Current Cosmic Microwave Background and Cosmic
Shear Data,” Symmetry 10 (2018) no. 11, 585.
[735] Planck Collaboration, P. A. R. Ade et al., “Planck 2015 results. XXIV. Cosmology from Sunyaev-Zeldovich cluster
counts,” Astron. Astrophys. 594 (2016) A24, arXiv:1502.01597 [astro-ph.CO].
[736] Planck Collaboration, P. A. R. Ade et al., “Planck 2015 results. XXVII. The Second Planck Catalogue of
Sunyaev-Zeldovich Sources,” Astron. Astrophys. 594 (2016) A27, arXiv:1502.01598 [astro-ph.CO].
[737] SPT Collaboration, T. de Haan et al., “Cosmological Constraints from Galaxy Clusters in the 2500 square-degree
SPT-SZ Survey,” Astrophys. J. 832 (2016) no. 1, 95, arXiv:1603.06522 [astro-ph.CO].
[738] P. J. E. Peebles and J. T. Yu, “Primeval adiabatic perturbation in an expanding universe,” Astrophys. J. 162 (1970)
815–836.
[739] R. A. Sunyaev and Y. B. Zeldovich, “Small-scale fluctuations of relic radiation,” Astrophys. Space Sci. 7 (1970) no. 1,
3–19.
[740] D. J. Eisenstein and W. Hu, “Baryonic features in the matter transfer function,” Astrophys. J. 496 (1998) 605,
arXiv:astro-ph/9709112.
[741] B. A. Bassett and R. Hlozek, “Baryon Acoustic Oscillations,” arXiv:0910.5224 [astro-ph.CO].
[742] D. H. Weinberg, M. J. Mortonson, D. J. Eisenstein, C. Hirata, A. G. Riess, and E. Rozo, “Observational Probes of
Cosmic Acceleration,” Phys. Rept. 530 (2013) 87–255, arXiv:1201.2434 [astro-ph.CO].
[743] SDSS Collaboration, D. J. Eisenstein et al., “Detection of the Baryon Acoustic Peak in the Large-Scale Correlation
Function of SDSS Luminous Red Galaxies,” Astrophys. J. 633 (2005) 560–574, arXiv:astro-ph/0501171.
[744] DESI Collaboration, A. G. Adame et al., “DESI 2024 IV: Baryon Acoustic Oscillations from the Lyman alpha forest,”
JCAP 01 (2025) 124, arXiv:2404.03001 [astro-ph.CO].
[745] 2dFGRS Collaboration, S. Cole et al., “The 2dF Galaxy Redshift Survey: Power-spectrum analysis of the final
dataset and cosmological implications,” Mon. Not. Roy. Astron. Soc. 362 (2005) 505–534, arXiv:astro-ph/0501174.
[746] SDSS Collaboration, W. J. Percival et al., “Baryon Acoustic Oscillations in the Sloan Digital Sky Survey Data Release
7 Galaxy Sample,” Mon. Not. Roy. Astron. Soc. 401 (2010) 2148–2168, arXiv:0907.1660 [astro-ph.CO].
[747] BOSS Collaboration, F. Beutler et al., “The clustering of galaxies in the completed SDSS-III Baryon Oscillation
Spectroscopic Survey: baryon acoustic oscillations in the Fourier space,” Mon. Not. Roy. Astron. Soc. 464 (2017)
no. 3, 3409–3430, arXiv:1607.03149 [astro-ph.CO].
[748] BOSS Collaboration, L. Anderson et al., “The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic
Survey: baryon acoustic oscillations in the Data Releases 10 and 11 Galaxy samples,” Mon. Not. Roy. Astron. Soc. 441
(2014) no. 1, 24–62, arXiv:1312.4877 [astro-ph.CO].
[749] K. T. Mehta, H.-J. Seo, J. Eckel, D. J. Eisenstein, M. Metchnik, P. Pinto, and X. Xu, “Galaxy Bias and its Effects on
the Baryon Acoustic Oscillations Measurements,” Astrophys. J. 734 (2011) 94, arXiv:1104.1178 [astro-ph.CO].
[750] N. Padmanabhan, X. Xu, D. J. Eisenstein, R. Scalzo, A. J. Cuesta, K. T. Mehta, and E. Kazin, “A 2 per cent distance
283

to z=0.35 by reconstructing baryon acoustic oscillations - I. Methods and application to the Sloan Digital Sky Survey,”
Mon. Not. Roy. Astron. Soc. 427 (2012) no. 3, 2132–2145, arXiv:1202.0090 [astro-ph.CO].
[751] D. J. Eisenstein, H.-j. Seo, E. Sirko, and D. Spergel, “Improving Cosmological Distance Measurements by
Reconstruction of the Baryon Acoustic Peak,” Astrophys. J. 664 (2007) 675–679, arXiv:astro-ph/0604362.
[752] J. L. Bernal, T. L. Smith, K. K. Boddy, and M. Kamionkowski, “Robustness of baryon acoustic oscillation constraints
for early-Universe modifications of ΛCDM cosmology,” Phys. Rev. D 102 (2020) no. 12, 123515, arXiv:2004.07263
[astro-ph.CO].
[753] S. Sanz-Wuhl, H. Gil-Marín, A. J. Cuesta, and L. Verde, “BAO cosmology in non-spatially flat background geometry
from BOSS+eBOSS and lessons for future surveys,” JCAP 05 (2024) 116, arXiv:2402.03427 [astro-ph.CO].
[754] J. Pan, D. Huterer, F. Andrade-Oliveira, and C. Avestruz, “Compressed baryon acoustic oscillation analysis is robust
to modified-gravity models,” JCAP 06 (2024) 051, arXiv:2312.05177 [astro-ph.CO].
[755] BOSS Collaboration, M. Vargas-Magaña et al., “The clustering of galaxies in the completed SDSS-III Baryon
Oscillation Spectroscopic Survey: theoretical systematics and Baryon Acoustic Oscillations in the galaxy correlation
function,” Mon. Not. Roy. Astron. Soc. 477 (2018) no. 1, 1153–1188, arXiv:1610.03506 [astro-ph.CO].
[756] P. Carter, F. Beutler, W. J. Percival, J. DeRose, R. H. Wechsler, and C. Zhao, “The impact of the fiducial cosmology
assumption on BAO distance scale measurements,” Mon. Not. Roy. Astron. Soc. 494 (2020) no. 2, 2076–2089,
arXiv:1906.03035 [astro-ph.CO].
[757] S.-F. Chen et al., “Baryon acoustic oscillation theory and modelling systematics for the DESI 2024 results,” Mon. Not.
Roy. Astron. Soc. 534 (2024) no. 1, 544–574, arXiv:2402.14070 [astro-ph.CO].
[758] D. Baumann, D. Green, and M. Zaldarriaga, “Phases of New Physics in the BAO Spectrum,” JCAP 11 (2017) 007,
arXiv:1703.00894 [astro-ph.CO].
[759] D. Baumann, F. Beutler, R. Flauger, D. Green, A. Slosar, M. Vargas-Magaña, B. Wallisch, and C. Yèche, “First
constraint on the neutrino-induced phase shift in the spectrum of baryon acoustic oscillations,” Nature Phys. 15 (2019)
465–469, arXiv:1803.10741 [astro-ph.CO].
[760] E. Sanchez, A. Carnero, J. Garcia-Bellido, E. Gaztanaga, F. de Simoni, M. Crocce, A. Cabre, P. Fosalba, and
D. Alonso, “Tracing The Sound Horizon Scale With Photometric Redshift Surveys,” Mon. Not. Roy. Astron. Soc. 411
(2011) 277–288, arXiv:1006.3226 [astro-ph.CO].
[761] R. Menote and V. Marra, “Baryon acoustic oscillations in thin redshift shells from BOSS DR12 and eBOSS DR16
galaxies,” Mon. Not. Roy. Astron. Soc. 513 (2022) no. 2, 1600–1608, arXiv:2112.10000 [astro-ph.CO].
[762] G. C. Carvalho, A. Bernui, M. Benetti, J. C. Carvalho, E. de Carvalho, and J. S. Alcaniz, “The transverse baryonic
acoustic scale from the SDSS DR11 galaxies,” Astropart. Phys. 119 (2020) 102432, arXiv:1709.00271 [astro-ph.CO].
[763] D. Camarena and V. Marra, “A new method to build the (inverse) distance ladder,” Mon. Not. Roy. Astron. Soc. 495
(2020) no. 3, 2630–2644, arXiv:1910.14125 [astro-ph.CO].
[764] R. C. Nunes, S. K. Yadav, J. F. Jesus, and A. Bernui, “Cosmological parameter analyses using transversal BAO data,”
Mon. Not. Roy. Astron. Soc. 497 (2020) no. 2, 2133–2141, arXiv:2002.09293 [astro-ph.CO].
[765] R. C. Nunes and A. Bernui, “BAO signatures in the 2-point angular correlations and the Hubble tension,” Eur. Phys.
J. C 80 (2020) no. 11, 1025, arXiv:2008.03259 [astro-ph.CO].
[766] R. Arjona and S. Nesseris, “Novel null tests for the spatial curvature and homogeneity of the Universe and their
machine learning reconstructions,” Phys. Rev. D 103 (2021) no. 10, 103539, arXiv:2103.06789 [astro-ph.CO].
[767] A. Bernui, E. Di Valentino, W. Giarè, S. Kumar, and R. C. Nunes, “Exploring the H0 tension and the evidence for
dark sector interactions from 2D BAO measurements,” Phys. Rev. D 107 (2023) no. 10, 103531, arXiv:2301.06097
[astro-ph.CO].
[768] A. Gómez-Valent, A. Favale, M. Migliaccio, and A. A. Sen, “Late-time phenomenology required to solve the H0 tension
in view of the cosmic ladders and the anisotropic and angular BAO datasets,” Phys. Rev. D 109 (2024) no. 2, 023525,
arXiv:2309.07795 [astro-ph.CO].
[769] M. Benetti, L. L. Graef, and J. S. Alcaniz, “The H0 and σ8 tensions and the scale invariant spectrum,” JCAP 07
(2018) 066, arXiv:1712.00677 [astro-ph.CO].
[770] DES Collaboration, J. Mena-Fernández et al., “Dark Energy Survey: Galaxy sample for the baryonic acoustic
oscillation measurement from the final dataset,” Phys. Rev. D 110 (2024) no. 6, 063514, arXiv:2402.10697
[astro-ph.CO].
[771] DES Collaboration, T. M. C. Abbott et al., “Dark Energy Survey: A 2.1% measurement of the angular baryonic
acoustic oscillation scale at redshift zeff=0.85 from the final dataset,” Phys. Rev. D 110 (2024) no. 6, 063515,
arXiv:2402.10696 [astro-ph.CO].
[772] P. McDonald and D. Eisenstein, “Dark energy and curvature from a future baryonic acoustic oscillation survey using
the Lyman-alpha forest,” Phys. Rev. D 76 (2007) 063009, arXiv:astro-ph/0607122.
[773] 2dFGRS Collaboration, W. J. Percival et al., “The 2dF Galaxy Redshift Survey: The Power spectrum and the matter
content of the Universe,” Mon. Not. Roy. Astron. Soc. 327 (2001) 1297, arXiv:astro-ph/0105252.
[774] BOSS Collaboration, A. Slosar et al., “Measurement of Baryon Acoustic Oscillations in the Lyman-alpha Forest
Fluctuations in BOSS Data Release 9,” JCAP 04 (2013) 026, arXiv:1301.3459 [astro-ph.CO].
[775] C. Blake et al., “The WiggleZ Dark Energy Survey: mapping the distance-redshift relation with baryon acoustic
oscillations,” Mon. Not. Roy. Astron. Soc. 418 (2011) 1707–1724, arXiv:1108.2635 [astro-ph.CO].
[776] F. Beutler, C. Blake, M. Colless, D. H. Jones, L. Staveley-Smith, L. Campbell, Q. Parker, W. Saunders, and
F. Watson, “The 6dF Galaxy Survey: Baryon Acoustic Oscillations and the Local Hubble Constant,” Mon. Not. Roy.
Astron. Soc. 416 (2011) 3017–3032, arXiv:1106.3366 [astro-ph.CO].
284

[777] BOSS Collaboration, L. Anderson et al., “The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic
Survey: measuring DA and H at z = 0.57 from the baryon acoustic peak in the Data Release 9 spectroscopic Galaxy
sample,” Mon. Not. Roy. Astron. Soc. 439 (2014) no. 1, 83–101, arXiv:1303.4666 [astro-ph.CO].
[778] BOSS Collaboration, A. Font-Ribera et al., “Quasar-Lyman α Forest Cross-Correlation from BOSS DR11 : Baryon
Acoustic Oscillations,” JCAP 05 (2014) 027, arXiv:1311.1767 [astro-ph.CO].
[779] BOSS Collaboration, A. J. Ross et al., “The clustering of galaxies in the completed SDSS-III Baryon Oscillation
Spectroscopic Survey: Observational systematics and baryon acoustic oscillations in the correlation function,” Mon.
Not. Roy. Astron. Soc. 464 (2017) no. 1, 1168–1191, arXiv:1607.03145 [astro-ph.CO].
[780] eBOSS Collaboration, R. Neveux et al., “The completed SDSS-IV extended Baryon Oscillation Spectroscopic Survey:
BAO and RSD measurements from the anisotropic power spectrum of the quasar sample between redshift 0.8 and 2.2,”
Mon. Not. Roy. Astron. Soc. 499 (2020) no. 1, 210–229, arXiv:2007.08999 [astro-ph.CO].
[781] eBOSS Collaboration, A. de Mattia et al., “The Completed SDSS-IV extended Baryon Oscillation Spectroscopic
Survey: measurement of the BAO and growth rate of structure of the emission line galaxy sample from the anisotropic
power spectrum between redshift 0.6 and 1.1,” Mon. Not. Roy. Astron. Soc. 501 (2021) no. 4, 5616–5645,
arXiv:2007.09008 [astro-ph.CO].
[782] DESI Collaboration, A. G. Adame et al., “DESI 2024 III: Baryon Acoustic Oscillations from Galaxies and Quasars,”
arXiv:2404.03000 [astro-ph.CO].
[783] D. Bianchi, A. Burden, W. J. Percival, D. Brooks, R. N. Cahn, J. E. Forero-Romero, M. Levi, A. J. Ross, and
G. Tarle, “Unbiased clustering estimates with the DESI fibre assignment,” Mon. Not. Roy. Astron. Soc. 481 (2018)
no. 2, 2338–2348, arXiv:1805.00951 [astro-ph.CO].
[784] G. C. Carvalho, A. Bernui, M. Benetti, J. C. Carvalho, and J. S. Alcaniz, “Baryon Acoustic Oscillations from the SDSS
DR10 galaxies angular correlation function,” Phys. Rev. D 93 (2016) no. 2, 023530, arXiv:1507.08972 [astro-ph.CO].
[785] J. S. Alcaniz, G. C. Carvalho, A. Bernui, J. C. Carvalho, and M. Benetti, “Measuring baryon acoustic oscillations with
angular two-point correlation function,” Fundam. Theor. Phys. 187 (2017) 11–19, arXiv:1611.08458 [astro-ph.CO].
[786] E. de Carvalho, A. Bernui, G. C. Carvalho, C. P. Novaes, and H. S. Xavier, “Angular Baryon Acoustic Oscillation
measure at z = 2.225 from the SDSS quasar survey,” JCAP 04 (2018) 064, arXiv:1709.00113 [astro-ph.CO].
[787] E. de Carvalho, A. Bernui, F. Avila, C. P. Novaes, and J. P. Nogueira-Cavalcante, “BAO angular scale at zeff = 0.11
with the SDSS blue galaxies,” Astron. Astrophys. 649 (2021) A20, arXiv:2103.14121 [astro-ph.CO].
[788] BOSS Collaboration, E. Aubourg et al., “Cosmological implications of baryon acoustic oscillation measurements,”
Phys. Rev. D 92 (2015) no. 12, 123516, arXiv:1411.1074 [astro-ph.CO].
[789] G. S. Farren, O. H. E. Philcox, and B. D. Sherwin, “Determining the Hubble constant without the sound horizon:
Perspectives with future galaxy surveys,” Phys. Rev. D 105 (2022) no. 6, 063503, arXiv:2112.10749 [astro-ph.CO].
[790] O. H. E. Philcox, B. D. Sherwin, G. S. Farren, and E. J. Baxter, “Determining the Hubble Constant without the Sound
Horizon: Measurements from Galaxy Surveys,” Phys. Rev. D 103 (2021) no. 2, 023538, arXiv:2008.08084
[astro-ph.CO].
[791] E. J. Baxter and B. D. Sherwin, “Determining the Hubble Constant without the Sound Horizon Scale: Measurements
from CMB Lensing,” Mon. Not. Roy. Astron. Soc. 501 (2021) no. 2, 1823–1835, arXiv:2007.04007 [astro-ph.CO].
[792] S. Brieden, H. Gil-Marín, and L. Verde, “A tale of two (or more) h’s,” JCAP 04 (2023) 023, arXiv:2212.04522
[astro-ph.CO].
[793] A. Krolewski, W. J. Percival, and A. Woodfinden, “New Method to Determine the Hubble Parameter from
Cosmological Energy-Density Measurements,” Phys. Rev. Lett. 134 (2025) no. 10, 101002, arXiv:2403.19227
[astro-ph.CO].
[794] N. Arendse et al., “Cosmic dissonance: are new physics or systematics behind a short sound horizon?,” Astron.
Astrophys. 639 (2020) A57, arXiv:1909.07986 [astro-ph.CO].
[795] J.-Q. Jiang and Y.-S. Piao, “Can the sound horizon-free measurement of H0 constrain early new physics?,”
arXiv:2501.16883 [astro-ph.CO].
[796] eBOSS Collaboration, S. Alam et al., “Completed SDSS-IV extended Baryon Oscillation Spectroscopic Survey:
Cosmological implications from two decades of spectroscopic surveys at the Apache Point Observatory,” Phys. Rev. D
103 (2021) no. 8, 083533, arXiv:2007.08991 [astro-ph.CO].
[797] N. Schöneberg, “The 2024 BBN baryon abundance update,” JCAP 06 (2024) 006, arXiv:2401.15054 [astro-ph.CO].
[798] W. J. Percival, S. Cole, D. J. Eisenstein, R. C. Nichol, J. A. Peacock, A. C. Pope, and A. S. Szalay, “Measuring the
Baryon Acoustic Oscillation scale using the SDSS and 2dFGRS,” Mon. Not. Roy. Astron. Soc. 381 (2007) 1053–1066,
arXiv:0705.3323 [astro-ph].
[799] W. Giarè, M. A. Sabogal, R. C. Nunes, and E. Di Valentino, “Interacting Dark Energy after DESI Baryon Acoustic
Oscillation Measurements,” Phys. Rev. Lett. 133 (2024) no. 25, 251003, arXiv:2404.15232 [astro-ph.CO].
[800] O. Akarsu, S. Kumar, E. Özülker, J. A. Vazquez, and A. Yadav, “Relaxing cosmological tensions with a sign switching
cosmological constant: Improved results with Planck, BAO, and Pantheon data,” Phys. Rev. D 108 (2023) no. 2,
023513, arXiv:2211.05742 [astro-ph.CO].
[801] O. Akarsu, A. De Felice, E. Di Valentino, S. Kumar, R. C. Nunes, E. Özülker, J. A. Vazquez, and A. Yadav,
“Cosmological constraints on ΛsCDM scenario in a type II minimally modified gravity,” Phys. Rev. D 110 (2024)
no. 10, 103527, arXiv:2406.07526 [astro-ph.CO].
[802] T. L. Smith, V. Poulin, and T. Simon, “Assessing the robustness of sound horizon-free determinations of the Hubble
constant,” Phys. Rev. D 108 (2023) no. 10, 103525, arXiv:2208.12992 [astro-ph.CO].
[803] K. Jedamzik, L. Pogosian, and G.-B. Zhao, “Why reducing the cosmic sound horizon alone can not fully resolve the
285

Hubble tension,” Commun. in Phys. 4 (2021) 123, arXiv:2010.04158 [astro-ph.CO].


[804] L. Pogosian, G.-B. Zhao, and K. Jedamzik, “Recombination-independent determination of the sound horizon and the
Hubble constant from BAO,” Astrophys. J. Lett. 904 (2020) no. 2, L17, arXiv:2009.08455 [astro-ph.CO].
[805] D. Benisty and D. Staicova, “Testing late-time cosmic acceleration with uncorrelated baryon acoustic oscillation
dataset,” Astron. Astrophys. 647 (2021) A38, arXiv:2009.10701 [astro-ph.CO].
[806] D. Staicova, “DE Models with Combined H0 · rd from BAO and CMB Dataset and Friends,” Universe 8 (2022) no. 12,
631, arXiv:2211.08139 [astro-ph.CO].
[807] D. Staicova and D. Benisty, “Constraining the dark energy models using baryon acoustic oscillations: An approach
independent of H0 · rd,” Astron. Astrophys. 668 (2022) A135, arXiv:2107.14129 [astro-ph.CO].
[808] DESI Collaboration, R. Calderon et al., “DESI 2024: reconstructing dark energy using crossing statistics with DESI
DR1 BAO data,” JCAP 10 (2024) 048, arXiv:2405.04216 [astro-ph.CO].
[809] F. Sinigaglia, F.-S. Kitaura, K. Nagamine, and Y. Oku, “The Negative Baryon Acoustic Oscillation Shift in the Lyα
Forest from Cosmological Simulations,” Astrophys. J. Lett. 971 (2024) no. 1, L22, arXiv:2407.03918 [astro-ph.CO].
[810] O. Akarsu, S. Kumar, E. Özülker, and J. A. Vazquez, “Relaxing cosmological tensions with a sign switching
cosmological constant,” Phys. Rev. D 104 (2021) no. 12, 123512, arXiv:2108.09239 [astro-ph.CO].
[811] O. Akarsu, E. Di Valentino, S. Kumar, R. C. Nunes, J. A. Vazquez, and A. Yadav, “Λs CDM model: A promising
scenario for alleviation of cosmological tensions,” arXiv:2307.10899 [astro-ph.CO].
[812] S. A. Adil, O. Akarsu, E. Di Valentino, R. C. Nunes, E. Özülker, A. A. Sen, and E. Specogna, “Omnipotent dark
energy: A phenomenological answer to the Hubble tension,” Phys. Rev. D 109 (2024) no. 2, 023527, arXiv:2306.08046
[astro-ph.CO].
[813] DESI Collaboration, A. Aghamousa et al., “The DESI Experiment Part I: Science,Targeting, and Survey Design,”
arXiv:1611.00036 [astro-ph.IM].
[814] D. J. Schlegel et al., “The MegaMapper: A Stage-5 Spectroscopic Instrument Concept for the Study of Inflation and
Dark Energy,” arXiv:2209.04322 [astro-ph.IM].
[815] D. J. Schlegel et al., “Astro2020 APC White Paper: The MegaMapper: a z > 2 Spectroscopic Instrument for the Study
of Inflation and Dark Energy,” Bull. Am. Astron. Soc. 51 (2019) no. 7, 229, arXiv:1907.11171 [astro-ph.IM].
[816] R. Ellis et al., “SpecTel: A 10-12 meter class Spectroscopic Survey Telescope,” arXiv:1907.06797 [astro-ph.IM].
[817] DESI Collaboration, A. G. Adame et al., “Validation of the Scientific Program for the Dark Energy Spectroscopic
Instrument,” Astron. J. 167 (2024) no. 2, 62, arXiv:2306.06307 [astro-ph.CO].
[818] B. Jain and U. Seljak, “Cosmological model predictions for weak lensing: Linear and nonlinear regimes,” Astrophys. J.
484 (1997) 560, arXiv:astro-ph/9611077.
[819] N. Kaiser, “Weak gravitational lensing of distant galaxies,” Astrophys. J. 388 (1992) 272.
[820] N. Kaiser, “Weak lensing by galaxy clusters,” arXiv:astro-ph/9912569.
[821] J. Miralda-Escude, “The Correlation Function of Galaxy Ellipticities Produced by Gravitational Lensing,” Astrophys.
J. 380 (1991) 1.
[822] R. D. Blandford, A. B. Saust, T. G. Brainerd, and J. V. Villumsen, “The distortion of distant galaxy images by
large-scale structure,” Mon. Not. Roy. Astron. Soc. 251 (1991) no. 4, 600–627.
[823] L. Van Waerbeke, F. Bernardeau, and Y. Mellier, “Efficiency of weak lensing surveys to probe cosmological models,”
Astron. Astrophys. 342 (1999) 15–33, arXiv:astro-ph/9807007.
[824] M. Bartelmann and M. Maturi, “Weak gravitational lensing,” in Ground-based and Airborne Telescopes VII, vol. 12,
p. 32440. 2017. arXiv:1612.06535 [astro-ph.CO].
[825] M. Kilbinger, “Cosmology with cosmic shear observations: a review,” Rept. Prog. Phys. 78 (2015) 086901,
arXiv:1411.0115 [astro-ph.CO].
[826] J. Prat and D. Bacon, “Weak Gravitational Lensing,” arXiv:2501.07938 [astro-ph.CO].
[827] P. Schneider, L. van Waerbeke, M. Kilbinger, and Y. Mellier, “Analysis of two-point statistics of cosmic shear: I.
estimators and covariances,” Astron. Astrophys. 396 (2002) 1–20, arXiv:astro-ph/0206182.
[828] Euclid Collaboration, V. Ajani et al., “Euclid Preparation. XXVIII. Forecasts for ten different higher-order weak
lensing statistics,” Astron. Astrophys. 675 (2023) A120, arXiv:2301.12890 [astro-ph.CO].
[829] L. Porth, S. Heydenreich, P. Burger, L. Linke, and P. Schneider, “A road map to cosmological parameter analysis with
third-order shear statistics - III. Efficient estimation of third-order shear correlation functions and an application to
the KiDS-1000 data,” Astron. Astrophys. 689 (2024) A227, arXiv:2309.08601 [astro-ph.CO].
[830] M. Takada and B. Jain, “The Three - point correlation function in cosmology,” Mon. Not. Roy. Astron. Soc. 340
(2003) 580–608, arXiv:astro-ph/0209167.
[831] S. Heydenreich, L. Linke, P. Burger, and P. Schneider, “A roadmap to cosmological parameter analysis with third-order
shear statistics - I. Modelling and validation,” Astron. Astrophys. 672 (2023) A44, arXiv:2208.11686 [astro-ph.CO].
[832] A. Halder, O. Friedrich, S. Seitz, and T. N. Varga, “The integrated three-point correlation function of cosmic shear,”
Mon. Not. Roy. Astron. Soc. 506 (2021) no. 2, 2780–2803, arXiv:2102.10177 [astro-ph.CO].
[833] W. R. Coulton, J. Liu, M. S. Madhavacheril, V. Böhm, and D. N. Spergel, “Constraining Neutrino Mass with the
Tomographic Weak Lensing Bispectrum,” JCAP 05 (2019) 043, arXiv:1810.02374 [astro-ph.CO].
[834] I. Kayo, M. Takada, and B. Jain, “Information content of weak lensing power spectrum and bispectrum: including the
non-Gaussian error covariance matrix,” Mon. Not. Roy. Astron. Soc. 429 (2013) 344–371, arXiv:1207.6322
[astro-ph.CO].
[835] L. Castiblanco, C. Uhlemann, J. Harnois-Déraps, and A. Barthelemy, “Unleashing cosmic shear information with the
tomographic weak lensing PDF,” arXiv:2405.09651 [astro-ph.CO].
286

[836] L. Thiele, G. A. Marques, J. Liu, and M. Shirasaki, “Cosmological constraints from the Subaru Hyper Suprime-Cam
year 1 shear catalogue lensing convergence probability distribution function,” Phys. Rev. D 108 (2023) no. 12, 123526,
arXiv:2304.05928 [astro-ph.CO].
[837] B. Giblin, Y.-C. Cai, and J. Harnois-Déraps, “Enhancing cosmic shear with the multiscale lensing probability density
function,” Mon. Not. Roy. Astron. Soc. 520 (2023) no. 2, 1721–1737, arXiv:2211.05708 [astro-ph.CO].
[838] J. Liu and M. S. Madhavacheril, “Constraining neutrino mass with the tomographic weak lensing one-point probability
distribution function and power spectrum,” Phys. Rev. D 99 (2019) no. 8, 083508, arXiv:1809.10747 [astro-ph.CO].
[839] DES Collaboration, M. Gatti et al., “Dark Energy Survey Year 3 results: Cosmology with moments of weak lensing
mass maps,” Phys. Rev. D 106 (2022) no. 8, 083509, arXiv:2110.10141 [astro-ph.CO].
[840] J. Harnois-Deraps et al., “KiDS-1000 and DES-Y1 combined: cosmology from peak count statistics,” Mon. Not. Roy.
Astron. Soc. 534 (2024) no. 4, 3305–3330, arXiv:2405.10312 [astro-ph.CO].
[841] G. A. Marques, J. Liu, M. Shirasaki, L. Thiele, D. Grandón, K. M. Huffenberger, S. Cheng, J. Harnois-Déraps,
K. Osato, and W. R. Coulton, “Cosmology from weak lensing peaks and minima with Subaru Hyper Suprime-Cam
Survey first-year data,” Mon. Not. Roy. Astron. Soc. 528 (2024) no. 3, 4513–4527, arXiv:2308.10866 [astro-ph.CO].
[842] N. Martinet, J. G. Bartlett, A. Kiessling, and B. Sartoris, “Constraining cosmology with shear peak statistics:
tomographic analysis,” Astron. Astrophys. 581 (2015) A101, arXiv:1506.02192 [astro-ph.CO].
[843] J. Liu, A. Petri, Z. Haiman, L. Hui, J. M. Kratochvil, and M. May, “Cosmology constraints from the weak lensing peak
counts and the power spectrum in CFHTLenS data,” Phys. Rev. D 91 (2015) no. 6, 063507, arXiv:1412.0757
[astro-ph.CO].
[844] DES Collaboration, D. Gruen et al., “Density Split Statistics: Cosmological Constraints from Counts and Lensing in
Cells in DES Y1 and SDSS Data,” Phys. Rev. D 98 (2018) no. 2, 023507, arXiv:1710.05045 [astro-ph.CO].
[845] S. Cheng, G. A. Marques, D. Grandón, L. Thiele, M. Shirasaki, B. Ménard, and J. Liu, “Cosmological constraints from
weak lensing scattering transform using HSC Y1 data,” JCAP 01 (2025) 006, arXiv:2404.16085 [astro-ph.CO].
[846] D. Grandón, G. A. Marques, L. Thiele, S. Cheng, M. Shirasaki, and J. Liu, “Impact of baryonic feedback on HSC-Y1
weak lensing non-Gaussian statistics,” Phys. Rev. D 110 (2024) no. 10, 103539, arXiv:2403.03807 [astro-ph.CO].
[847] J. Armijo, G. A. Marques, C. P. Novaes, L. Thiele, J. A. Cowell, D. Grandón, M. Shirasaki, and J. Liu, “Cosmological
constraints using Minkowski functionals from the first year data of the Hyper Suprime-Cam,” Mon. Not. Roy. Astron.
Soc. 537 (2025) no. 4, 3553–3560, arXiv:2410.00401 [astro-ph.CO].
[848] D. Grandón and E. Sellentin, “Stage IV baryonic feedback correction for non-Gaussianity inference,” Mon. Not. Roy.
Astron. Soc. 536 (2025) 2064–2071, arXiv:2407.20448 [astro-ph.CO].
[849] K. Kuijken et al., “Gravitational Lensing Analysis of the Kilo Degree Survey,” Mon. Not. Roy. Astron. Soc. 454 (2015)
no. 4, 3500–3532, arXiv:1507.00738 [astro-ph.CO].
[850] K. Kuijken et al., “The fourth data release of the Kilo-Degree Survey: ugri imaging and nine-band optical-IR
photometry over 1000 square degrees,” Astron. Astrophys. 625 (2019) A2, arXiv:1902.11265 [astro-ph.GA].
[851] DES Collaboration, M. Gatti et al., “Dark energy survey year 3 results: weak lensing shape catalogue,” Mon. Not.
Roy. Astron. Soc. 504 (2021) no. 3, 4312–4336, arXiv:2011.03408 [astro-ph.CO].
[852] DES, NOAO Data Lab Collaboration, T. M. C. Abbott et al., “The Dark Energy Survey Data Release 1,”
Astrophys. J. Suppl. 239 (2018) no. 2, 18, arXiv:1801.03181 [astro-ph.IM].
[853] H. Aihara et al., “The Hyper Suprime-Cam SSP Survey: Overview and Survey Design,” Publ. Astron. Soc. Jap. 70
(2018) S4, arXiv:1704.05858 [astro-ph.IM].
[854] HSC Collaboration, C. Hikage et al., “Cosmology from cosmic shear power spectra with Subaru Hyper Suprime-Cam
first-year data,” Publ. Astron. Soc. Jap. 71 (2019) no. 2, 43, arXiv:1809.09148 [astro-ph.CO].
[855] T. Hamana et al., “Cosmological constraints from cosmic shear two-point correlation functions with HSC survey
first-year data,” Publ. Astron. Soc. Jap. 72 (2020) no. 1, 16, arXiv:1906.06041 [astro-ph.CO]. [Erratum:
Publ.Astron.Soc.Jap. 74, 488-491 (2022)].
[856] D. Anbajagane et al., “The DECADE cosmic shear project IV: cosmological constraints from 107 million galaxies
across 5,400 deg2 of the sky,” arXiv:2502.17677 [astro-ph.CO].
[857] H. Hildebrandt et al., “KiDS-450: Cosmological parameter constraints from tomographic weak gravitational lensing,”
Mon. Not. Roy. Astron. Soc. 465 (2017) 1454, arXiv:1606.05338 [astro-ph.CO].
[858] H. Hildebrandt et al., “KiDS+VIKING-450: Cosmic shear tomography with optical and infrared data,” Astron.
Astrophys. 633 (2020) A69, arXiv:1812.06076 [astro-ph.CO].
[859] KiDS Collaboration, M. Asgari et al., “KiDS-1000 Cosmology: Cosmic shear constraints and comparison between two
point statistics,” Astron. Astrophys. 645 (2021) A104, arXiv:2007.15633 [astro-ph.CO].
[860] DES Collaboration, L. F. Secco et al., “Dark Energy Survey Year 3 results: Cosmology from cosmic shear and
robustness to modeling uncertainty,” Phys. Rev. D 105 (2022) no. 2, 023515, arXiv:2105.13544 [astro-ph.CO].
[861] DES Collaboration, A. Amon et al., “Dark Energy Survey Year 3 results: Cosmology from cosmic shear and
robustness to data calibration,” Phys. Rev. D 105 (2022) no. 2, 023514, arXiv:2105.13543 [astro-ph.CO].
[862] DES Collaboration, M. A. Troxel et al., “Dark Energy Survey Year 1 results: Cosmological constraints from cosmic
shear,” Phys. Rev. D 98 (2018) no. 4, 043528, arXiv:1708.01538 [astro-ph.CO].
[863] R. Dalal et al., “Hyper Suprime-Cam Year 3 results: Cosmology from cosmic shear power spectra,” Phys. Rev. D 108
(2023) no. 12, 123519, arXiv:2304.00701 [astro-ph.CO].
[864] T. Erben, H. Hildebrandt, L. Miller, L. van Waerbeke, C. Heymans, H. Hoekstra, T. D. Kitching, Y. Mellier,
J. Benjamin, C. Blake, C. Bonnett, O. Cordes, J. Coupon, L. Fu, R. Gavazzi, B. Gillis, E. Grocutt, S. D. J. Gwyn,
K. Holhjem, M. J. Hudson, M. Kilbinger, K. Kuijken, M. Milkeraitis, B. T. P. Rowe, T. Schrabback, E. Semboloni,
287

P. Simon, M. Smit, O. Toader, S. Vafaei, E. van Uitert, and M. Velander, “CFHTLenS: the Canada-France-Hawaii
Telescope Lensing Survey - imaging data and catalogue products,” MNRAS 433 (2013) no. 3, 2545–2563,
arXiv:1210.8156 [astro-ph.CO].
[865] Planck Collaboration, P. A. R. Ade et al., “Planck 2013 results. XVI. Cosmological parameters,” Astron. Astrophys.
571 (2014) A16, arXiv:1303.5076 [astro-ph.CO].
[866] Kilo-Degree Survey, DES Collaboration, T. M. C. Abbott et al., “DES Y3 + KiDS-1000: Consistent cosmology
combining cosmic shear surveys,” Open J. Astrophys. 6 (2023) 2305.17173, arXiv:2305.17173 [astro-ph.CO].
[867] DES Collaboration, T. M. C. Abbott et al., “Dark Energy Survey Year 3 results: Cosmological constraints from
galaxy clustering and weak lensing,” Phys. Rev. D 105 (2022) no. 2, 023520, arXiv:2105.13549 [astro-ph.CO].
[868] DES Collaboration, C. Doux et al., “Dark energy survey year 3 results: cosmological constraints from the analysis of
cosmic shear in harmonic space,” Mon. Not. Roy. Astron. Soc. 515 (2022) no. 2, 1942–1972, arXiv:2203.07128
[astro-ph.CO].
[869] X. Li et al., “Hyper Suprime-Cam Year 3 results: Cosmology from cosmic shear two-point correlation functions,” Phys.
Rev. D 108 (2023) no. 12, 123518, arXiv:2304.00702 [astro-ph.CO].
[870] S. Sugiyama et al., “Hyper Suprime-Cam Year 3 results: Cosmology from galaxy clustering and weak lensing with HSC
and SDSS using the minimal bias model,” Phys. Rev. D 108 (2023) no. 12, 123521, arXiv:2304.00705 [astro-ph.CO].
[871] H. Miyatake et al., “Hyper Suprime-Cam Year 3 results: Cosmology from galaxy clustering and weak lensing with HSC
and SDSS using the emulator based halo model,” Phys. Rev. D 108 (2023) no. 12, 123517, arXiv:2304.00704
[astro-ph.CO].
[872] C. Heymans et al., “KiDS-1000 Cosmology: Multi-probe weak gravitational lensing and spectroscopic galaxy clustering
constraints,” Astron. Astrophys. 646 (2021) A140, arXiv:2007.15632 [astro-ph.CO].
[873] KiDS, Euclid Collaboration, A. Loureiro et al., “KiDS and Euclid: Cosmological implications of a pseudo angular
power spectrum analysis of KiDS-1000 cosmic shear tomography,” Astron. Astrophys. 665 (2022) A56,
arXiv:2110.06947 [astro-ph.CO].
[874] SPT Collaboration, F. Bianchini et al., “Constraints on Cosmological Parameters from the 500 deg2 SPTpol Lensing
Power Spectrum,” Astrophys. J. 888 (2020) 119, arXiv:1910.07157 [astro-ph.CO].
[875] DES Collaboration, M. Gatti et al., “Dark Energy Survey Year 3 results: Simulation-based cosmological inference with
wavelet harmonics, scattering transforms, and moments of weak lensing mass maps. Validation on simulations,” Phys.
Rev. D 109 (2024) no. 6, 063534, arXiv:2310.17557 [astro-ph.CO].
[876] M. von Wietersheim-Kramsta, K. Lin, N. Tessore, B. Joachimi, A. Loureiro, R. Reischke, and A. H. Wright,
“KiDS-SBI: Simulation-based inference analysis of KiDS-1000 cosmic shear,” Astron. Astrophys. 694 (2025) A223,
arXiv:2404.15402 [astro-ph.CO].
[877] C. P. Novaes, L. Thiele, J. Armijo, S. Cheng, J. A. Cowell, G. A. Marques, E. G. M. Ferreira, M. Shirasaki, K. Osato,
and J. Liu, “Cosmology from HSC Y1 Weak Lensing with Combined Higher-Order Statistics and Simulation-based
Inference,” arXiv:2409.01301 [astro-ph.CO].
[878] R. Reischke et al., “KiDS-Legacy: Covariance validation and the unified OneCovariance framework for projected
large-scale structure observables,” arXiv:2410.06962 [astro-ph.CO].
[879] E. Sellentin and A. F. Heavens, “Parameter inference with estimated covariance matrices,” Mon. Not. Roy. Astron.
Soc. 456 (2016) no. 1, L132–L136, arXiv:1511.05969 [astro-ph.CO].
[880] E. Sellentin, C. Heymans, and J. Harnois-Déraps, “The skewed weak lensing likelihood: why biases arise, despite data
and theory being sound,” Mon. Not. Roy. Astron. Soc. 477 (2018) no. 4, 4879–4895, arXiv:1712.04923
[astro-ph.CO].
[881] N. E. Chisari, M. L. A. Richardson, J. Devriendt, Y. Dubois, A. Schneider, A. L. Brun, M. C., R. S. Beckmann,
S. Peirani, A. Slyz, and C. Pichon, “The impact of baryons on the matter power spectrum from the Horizon-AGN
cosmological hydrodynamical simulation,” Mon. Not. Roy. Astron. Soc. 480 (2018) no. 3, 3962–3977,
arXiv:1801.08559 [astro-ph.CO].
[882] Euclid Collaboration, M. Knabenhans et al., “Euclid preparation: IX. EuclidEmulator2 – power spectrum emulation
with massive neutrinos and self-consistent dark energy perturbations,” Mon. Not. Roy. Astron. Soc. 505 (2021) no. 2,
2840–2869, arXiv:2010.11288 [astro-ph.CO].
[883] G. Aricò, R. E. Angulo, S. Contreras, L. Ondaro-Mallea, M. Pellejero-Ibañez, and M. Zennaro, “The BACCO
simulation project: a baryonification emulator with neural networks,” Mon. Not. Roy. Astron. Soc. 506 (2021) no. 3,
4070–4082, arXiv:2011.15018 [astro-ph.CO].
[884] M. Asgari, A. J. Mead, and C. Heymans, “The halo model for cosmology: a pedagogical review,” arXiv:2303.08752
[astro-ph.CO].
[885] VIRGO Consortium Collaboration, R. E. Smith, J. A. Peacock, A. Jenkins, S. D. M. White, C. S. Frenk, F. R.
Pearce, P. A. Thomas, G. Efstathiou, and H. M. P. Couchmann, “Stable clustering, the halo model and nonlinear
cosmological power spectra,” Mon. Not. Roy. Astron. Soc. 341 (2003) 1311, arXiv:astro-ph/0207664.
[886] R. Takahashi, M. Sato, T. Nishimichi, A. Taruya, and M. Oguri, “Revising the Halofit Model for the Nonlinear Matter
Power Spectrum,” Astrophys. J. 761 (2012) 152, arXiv:1208.2701 [astro-ph.CO].
[887] A. Mead, S. Brieden, T. Tröster, and C. Heymans, “hmcode-2020: improved modelling of non-linear cosmological
power spectra with baryonic feedback,” Mon. Not. Roy. Astron. Soc. 502 (2021) no. 1, 1401–1422, arXiv:2009.01858
[astro-ph.CO].
[888] A. Mead, C. Heymans, L. Lombriser, J. Peacock, O. Steele, and H. Winther, “Accurate halo-model matter power
spectra with dark energy, massive neutrinos and modified gravitational forces,” Mon. Not. Roy. Astron. Soc. 459
288

(2016) no. 2, 1468–1488, arXiv:1602.02154 [astro-ph.CO].


[889] F. Bernardeau, T. Nishimichi, and A. Taruya, “Cosmic shear full nulling: sorting out dynamics, geometry and
systematics,” Mon. Not. Roy. Astron. Soc. 445 (2014) no. 2, 1526–1537, arXiv:1312.0430 [astro-ph.CO].
[890] P. L. Taylor, F. Bernardeau, and T. D. Kitching, “k-cut Cosmic Shear: Tunable Power Spectrum Sensitivity to Test
Gravity,” Phys. Rev. D 98 (2018) no. 8, 083514, arXiv:1809.03515 [astro-ph.CO].
[891] A. Barthelemy, S. Codis, C. Uhlemann, F. Bernardeau, and R. Gavazzi, “A nulling strategy for modelling lensing
convergence in cones with large deviation theory,” Mon. Not. Roy. Astron. Soc. 492 (2020) no. 3, 3420–3439,
arXiv:1909.02615 [astro-ph.CO].
[892] D. Spergel et al., “Wide-Field InfrarRed Survey Telescope-Astrophysics Focused Telescope Assets WFIRST-AFTA
2015 Report,” arXiv:1503.03757 [astro-ph.IM].
[893] LSST Collaboration, v. Ivezić et al., “LSST: from Science Drivers to Reference Design and Anticipated Data
Products,” Astrophys. J. 873 (2019) no. 2, 111, arXiv:0805.2366 [astro-ph].
[894] R. Massey et al., “Origins of weak lensing systematics, and requirements on future instrumentation (or knowledge of
instrumentation),” Mon. Not. Roy. Astron. Soc. 429 (2013) 661, arXiv:1210.7690 [astro-ph.CO].
[895] N. Kaiser, G. Squires, and T. J. Broadhurst, “A Method for weak lensing observations,” Astrophys. J. 449 (1995)
460–475, arXiv:astro-ph/9411005.
[896] L. Miller, T. D. Kitching, C. Heymans, A. F. Heavens, and L. Van Waerbeke, “Bayesian Galaxy Shape Measurement
for Weak Lensing Surveys. 1. Methodology and a Fast Fitting Algorithm,” Mon. Not. Roy. Astron. Soc. 382 (2007)
315, arXiv:0708.2340 [astro-ph].
[897] G. M. Bernstein and R. Armstrong, “Bayesian Lensing Shear Measurement,” Mon. Not. Roy. Astron. Soc. 438 (2014)
no. 2, 1880–1893, arXiv:1304.1843 [astro-ph.CO].
[898] R. Mandelbaum, “Weak lensing for precision cosmology,” Ann. Rev. Astron. Astrophys. 56 (2018) 393–433,
arXiv:1710.03235 [astro-ph.CO].
[899] I. Fenech Conti, R. Herbonnet, H. Hoekstra, J. Merten, L. Miller, and M. Viola, “Calibration of weak-lensing shear in
the Kilo-Degree Survey,” Mon. Not. Roy. Astron. Soc. 467 (2017) no. 2, 1627–1651, arXiv:1606.05337
[astro-ph.CO].
[900] E. Huff and R. Mandelbaum, “Metacalibration: Direct Self-Calibration of Biases in Shear Measurement,”
arXiv:1702.02600 [astro-ph.CO].
[901] DES Collaboration, N. MacCrann et al., “Dark Energy Survey Y3 results: blending shear and redshift biases in image
simulations,” Mon. Not. Roy. Astron. Soc. 509 (2021) no. 3, 3371–3394, arXiv:2012.08567 [astro-ph.CO].
[902] X. Li et al., “The three-year shear catalog of the Subaru Hyper Suprime-Cam SSP Survey,” Publ. Astron. Soc. Jap. 74
(2022) no. 2, 421–459–459, arXiv:2107.00136 [astro-ph.CO].
[903] S.-S. Li et al., “KiDS-Legacy calibration: unifying shear and redshift calibration with the SKiLLS multi-band image
simulations,” Astron. Astrophys. 670 (2023) A100, arXiv:2210.07163 [astro-ph.CO].
[904] H. Hoekstra, A. Kannawadi, and T. D. Kitching, “Accounting for object detection bias in weak gravitational lensing
studies,” Astron. Astrophys. 646 (2021) A124, arXiv:2010.04178 [astro-ph.CO].
[905] N. Kaiser, “A new shear estimator for weak lensing observations,” Astrophys. J. 537 (2000) 555,
arXiv:astro-ph/9904003.
[906] G. M. Bernstein and M. Jarvis, “Shapes and shears, stars and smears: optimal measurements for weak lensing,”
Astron. J. 123 (2002) 583–618, arXiv:astro-ph/0107431.
[907] C. M. Hirata and U. Seljak, “Shear calibration biases in weak lensing surveys,” Mon. Not. Roy. Astron. Soc. 343
(2003) 459–480, arXiv:astro-ph/0301054.
[908] J. Hartlap, S. Hilbert, P. Schneider, and H. Hildebrandt, “A bias in cosmic shear from galaxy selection: results from
ray-tracing simulations,” Astron. Astrophys. 528 (2011) A51, arXiv:1010.0010 [astro-ph.CO].
[909] E. S. Sheldon, M. R. Becker, N. MacCrann, and M. Jarvis, “Mitigating Shear-dependent Object Detection Biases with
Metacalibration,” Astrophys. J. 902 (2020) no. 2, 138, arXiv:1911.02505 [astro-ph.CO].
[910] C. Heymans et al., “The Shear TEsting Programme. 1. Weak lensing analysis of simulated ground-based observations,”
Mon. Not. Roy. Astron. Soc. 368 (2006) 1323–1339, arXiv:astro-ph/0506112.
[911] R. Massey et al., “The Shear TEsting Programme 2: Factors affecting high precision weak lensing analyses,” Mon. Not.
Roy. Astron. Soc. 376 (2007) 13–38, arXiv:astro-ph/0608643.
[912] S. Bridle et al., “Results of the GREAT08 Challenge: An image analysis competition for cosmological lensing,” Mon.
Not. Roy. Astron. Soc. 405 (2010) 2044, arXiv:0908.0945 [astro-ph.CO].
[913] T. D. Kitching et al., “Image Analysis for Cosmology: Results from the GREAT10 Star Challenge,” Astrophys. J.
Supp. 205 (2013) 12, arXiv:1210.1979 [astro-ph.IM].
[914] R. Mandelbaum et al., “The Third Gravitational Lensing Accuracy Testing (GREAT3) Challenge Handbook,”
Astrophys. J. Suppl. 212 (2014) 5, arXiv:1308.4982 [astro-ph.CO].
[915] S.-S. Li et al., “KiDS-1000: Cosmology with improved cosmic shear measurements,” Astron. Astrophys. 679 (2023)
A133, arXiv:2306.11124 [astro-ph.CO].
[916] J. A. Newman and D. Gruen, “Photometric Redshifts for Next-Generation Surveys,” Ann. Rev. Astron. Astrophys. 60
(2022) 363–414, arXiv:2206.13633 [astro-ph.CO].
[917] M. Salvato, O. Ilbert, and B. Hoyle, “The many flavours of photometric redshifts,” Nature Astronomy 3 (2019)
212–222, arXiv:1805.12574 [astro-ph.GA].
[918] R. Tagliaferri, G. Longo, S. Andreon, S. Capozziello, C. Donalek, and G. Giordano, Neural Networks for Photometric
Redshifts Evaluation, pp. 226–234. Springer Berlin Heidelberg, Berlin, Heidelberg, 2003.
289

[919] A. A. Collister and O. Lahav, “ANNz: Estimating Photometric Redshifts Using Artificial Neural Networks,”
Publications of the Astronomical Society of the Pacific 116 (2004) 345–351, arXiv:astro-ph/0311058 [astro-ph].
[920] D. W. Gerdes, A. J. Sypniewski, T. A. McKay, J. Hao, M. R. Weis, R. H. Wechsler, and M. T. Busha, “ArborZ:
Photometric Redshifts Using Boosted Decision Trees,” Astrophys. J. 715 (2010) 823–832, arXiv:0908.4085
[astro-ph.CO].
[921] M. Carrasco Kind and R. J. Brunner, “TPZ: photometric redshift PDFs and ancillary information by using prediction
trees and random forests,” MNRAS 432 (2013) no. 2, 1483–1501, arXiv:1303.7269 [astro-ph.CO].
[922] C. Bonnett, “Using neural networks to estimate redshift distributions. An application to CFHTLenS,” MNRAS 449
(2015) 1043–1056, arXiv:1312.1287 [astro-ph.CO].
[923] M. M. Rau, S. Seitz, F. Brimioulle, E. Frank, O. Friedrich, D. Gruen, and B. Hoyle, “Accurate photometric redshift
probability density estimation - method comparison and application,” MNRAS 452 (2015) no. 4, 3710–3725,
arXiv:1503.08215 [astro-ph.CO].
[924] B. Hoyle, “Measuring photometric redshifts using galaxy images and Deep Neural Networks,” Astronomy and
Computing 16 (2016) 34–40, arXiv:1504.07255 [astro-ph.IM].
[925] S. Arnouts, S. Cristiani, L. Moscardini, S. Matarrese, F. Lucchin, A. Fontana, and E. Giallongo, “Measuring and
modelling the redshift evolution of clustering: the Hubble Deep Field North,” MNRAS 310 (1999) 540–556,
arXiv:astro-ph/9902290 [astro-ph].
[926] N. Benitez, “Bayesian photometric redshift estimation,” Astrophys. J. 536 (2000) 571–583, arXiv:astro-ph/9811189.
[927] O. Ilbert et al., “Accurate photometric redshifts for the cfht legacy survey calibrated using the vimos vlt deep survey,”
Astron. Astrophys. 457 (2006) 841–856, arXiv:astro-ph/0603217.
[928] R. Feldmann, C. M. Carollo, C. Porciani, S. J. Lilly, P. Capak, Y. Taniguchi, O. Le Fèvre, A. Renzini, N. Scoville,
M. Ajiki, H. Aussel, T. Contini, H. McCracken, B. Mobasher, T. Murayama, D. Sanders, S. Sasaki, C. Scarlata,
M. Scodeggio, Y. Shioya, J. Silverman, M. Takahashi, D. Thompson, and G. Zamorani, “The Zurich Extragalactic
Bayesian Redshift Analyzer and its first application: COSMOS,” MNRAS 372 (2006) 565–577,
arXiv:astro-ph/0609044 [astro-ph].
[929] N. Greisel, S. Seitz, N. Drory, R. Bender, R. P. Saglia, and J. Snigula, “Photometric redshifts and model spectral
energy distributions of galaxies from the SDSS-III BOSS DR10 data,” MNRAS 451 (2015) 1848–1867,
arXiv:1505.01157.
[930] B. Leistedt, D. J. Mortlock, and H. V. Peiris, “Hierarchical Bayesian inference of galaxy redshift distributions from
photometric surveys,” MNRAS 460 (2016) 4258–4267, arXiv:1602.05960.
[931] A. I. Malz and D. W. Hogg, “How to Obtain the Redshift Distribution from Probabilistic Redshift Estimates,”
Astrophys. J. 928 (2022) no. 2, 127, arXiv:2007.12178 [astro-ph.CO].
[932] C. E. Cunha, D. Huterer, H. Lin, M. T. Busha, and R. H. Wechsler, “Spectroscopic failures in photometric redshift
calibration: cosmological biases and survey requirements,” MNRAS 444 (2014) 129–146, arXiv:1207.3347
[astro-ph.CO].
[933] J. A. Newman et al., “Spectroscopic needs for imaging dark energy experiments,” Astropart. Phys. 63 (2015) 81–100,
arXiv:1309.5384 [astro-ph.CO]. [Erratum: Astropart.Phys. 65, 112–113 (2015)].
[934] D. C. Masters, D. K. Stern, J. G. Cohen, P. L. Capak, J. D. Rhodes, F. J. Castander, and S. Paltani, “The Complete
Calibration of the Color-Redshift Relation (C3R2) Survey: Survey Overview and Data Release 1,” Astrophys. J. 841
(2017) no. 2, 111, arXiv:1704.06665 [astro-ph.CO].
[935] D. C. Masters, D. K. Stern, J. G. Cohen, P. L. Capak, S. A. Stanford, N. Hernitschek, A. Galametz, I. Davidzon, J. D.
Rhodes, D. Sanders, B. Mobasher, F. Castander, K. Pruett, and S. Fotopoulou, “The Complete Calibration of the
Color-Redshift Relation (C3R2) Survey: Analysis and Data Release 2,” Astrophys. J. 877 (2019) no. 2, 81,
arXiv:1904.06394 [astro-ph.GA].
[936] W. G. Hartley et al., “The impact of spectroscopic incompleteness in direct calibration of redshift distributions for
weak lensing surveys,” MNRAS 496 (2020) no. 4, 4769–4786, arXiv:2003.10454 [astro-ph.GA].
[937] Z. Ma, W. Hu, and D. Huterer, “Effects of Photometric Redshift Uncertainties on Weak-Lensing Tomography,”
Astrophys. J. 636 (2006) no. 1, 21–29, arXiv:astro-ph/0506614 [astro-ph].
[938] D. Masters, P. Capak, D. Stern, O. Ilbert, M. Salvato, S. Schmidt, G. Longo, J. Rhodes, S. Paltani, B. Mobasher,
H. Hoekstra, H. Hildebrandt, J. Coupon, C. Steinhardt, J. Speagle, A. Faisst, A. Kalinich, M. Brodwin, M. Brescia,
and S. Cavuoti, “Mapping the Galaxy Color-Redshift Relation: Optimal Photometric Redshift Calibration Strategies
for Cosmology Surveys,” Astrophys. J. 813 (2015) 53, arXiv:1509.03318 [astro-ph.CO].
[939] M. J. Jee, J. A. Tyson, M. D. Schneider, D. Wittman, S. Schmidt, and S. Hilbert, “Cosmic shear results from the deep
lens survey - I: Joint constraints on omega_m and sigma_8 with a two-dimensional analysis,” Astrophys. J. 765
(2013) 74, arXiv:1210.2732 [astro-ph.CO].
[940] J. Benjamin, L. Van Waerbeke, C. Heymans, M. Kilbinger, T. Erben, H. Hildebrandt, H. Hoekstra, T. D. Kitching,
Y. Mellier, L. Miller, B. Rowe, T. Schrabback, F. Simpson, J. Coupon, L. Fu, J. Harnois-Déraps, M. J. Hudson,
K. Kuijken, E. Semboloni, S. Vafaei, and M. Velander, “CFHTLenS tomographic weak lensing: quantifying accurate
redshift distributions,” MNRAS 431 (2013) 1547–1564, arXiv:1212.3327 [astro-ph.CO].
[941] N. Benítez, “BPZ: Bayesian Photometric Redshift Code.” Astrophysics source code library, record ascl:1108.011, 2011.
[942] M. Lima, C. E. Cunha, H. Oyaizu, J. Frieman, H. Lin, and E. S. Sheldon, “Estimating the redshift distribution of
photometric galaxy samples,” MNRAS 390 (2008) no. 1, 118–130, arXiv:0801.3822 [astro-ph].
[943] DES Collaboration, B. Hoyle et al., “Dark Energy Survey Year 1 Results: Redshift distributions of the weak lensing
source galaxies,” Mon. Not. Roy. Astron. Soc. 478 (2018) no. 1, 592–610, arXiv:1708.01532 [astro-ph.CO].
290

[944] A. H. Wright, H. Hildebrandt, J. L. van den Busch, and C. Heymans, “Photometric redshift calibration with
self-organising maps,” Astron. Astrophys. 637 (2020) A100, arXiv:1909.09632 [astro-ph.CO].
[945] H. Hildebrandt, J. L. van den Busch, A. H. Wright, C. Blake, B. Joachimi, K. Kuijken, T. Tröster, M. Asgari,
M. Bilicki, J. T. A. de Jong, A. Dvornik, T. Erben, F. Getman, B. Giblin, C. Heymans, A. Kannawadi, C. A. Lin, and
H. Y. Shan, “KiDS-1000 catalogue: Redshift distributions and their calibration,” A&A 647 (2021) A124,
arXiv:2007.15635 [astro-ph.CO].
[946] A. H. Wright et al., “The fifth data release of the Kilo Degree Survey: Multi-epoch optical/NIR imaging covering wide
and legacy-calibration fields,” A&A 686 (2024) A170.
[947] J. Myles et al., “Dark Energy Survey Year 3 results: redshift calibration of the weak lensing source galaxies,” MNRAS
505 (2021) no. 3, 4249–4277, arXiv:2012.08566 [astro-ph.CO].
[948] R. Cawthon et al., “Dark Energy Survey Year 3 results: calibration of lens sample redshift distributions using
clustering redshifts with BOSS/eBOSS,” MNRAS 513 (2022) no. 4, 5517–5539, arXiv:2012.12826 [astro-ph.CO].
[949] M. Gatti et al., “Dark Energy Survey Year 3 Results: clustering redshifts - calibration of the weak lensing source
redshift distributions with redMaGiC and BOSS/eBOSS,” MNRAS 510 (2022) no. 1, 1223–1247, arXiv:2012.08569
[astro-ph.CO].
[950] M. M. Rau, R. Dalal, T. Zhang, X. Li, A. J. Nishizawa, S. More, R. Mandelbaum, H. Miyatake, M. A. Strauss, and
M. Takada, “Weak lensing tomographic redshift distribution inference for the Hyper Suprime-Cam Subaru Strategic
Program three-year shape catalogue,” Mon. Not. Roy. Astron. Soc. 524 (2023) no. 4, 5109–5131, arXiv:2211.16516
[astro-ph.CO].
[951] M. Schneider, L. Knox, H. Zhan, and A. Connolly, “Using Galaxy Two-Point Correlation Functions to Determine the
Redshift Distributions of Galaxies Binned by Photometric Redshift,” Astrophys. J. 651 (2006) no. 1, 14–23,
arXiv:astro-ph/0606098 [astro-ph].
[952] J. A. Newman, “Calibrating Redshift Distributions beyond Spectroscopic Limits with Cross-Correlations,” Astrophys.
J. 684 (2008) 88–101, arXiv:0805.1409 [astro-ph].
[953] J. L. v. d. Busch et al., “KiDS-1000: Cosmic shear with enhanced redshift calibration,” Astron. Astrophys. 664 (2022)
A170, arXiv:2204.02396 [astro-ph.CO].
[954] LSST Dark Energy Science Collaboration, I. Moskowitz, E. Gawiser, J. F. Crenshaw, B. H. Andrews, A. I. Malz,
and S. Schmidt, “Improving Photometric Redshift Estimates with Training Sample Augmentation,” Astrophys. J. Lett.
967 (2024) no. 1, L6, arXiv:2402.15551 [astro-ph.IM].
[955] T. Zhang, M. M. Rau, R. Mandelbaum, X. Li, and B. Moews, “Photometric redshift uncertainties in weak
gravitational lensing shear analysis: models and marginalization,” Mon. Not. Roy. Astron. Soc. 518 (2022) no. 1,
709–723, arXiv:2206.10169 [astro-ph.CO].
[956] M. P. van Daalen, J. Schaye, C. M. Booth, and C. D. Vecchia, “The effects of galaxy formation on the matter power
spectrum: A challenge for precision cosmology,” Mon. Not. Roy. Astron. Soc. 415 (2011) 3649–3665, arXiv:1104.1174
[astro-ph.CO].
[957] J. Sunseri, Z. Li, and J. Liu, “Effects of baryonic feedback on the cosmic web,” Phys. Rev. D 107 (2023) no. 2, 023514,
arXiv:2212.05927 [astro-ph.CO].
[958] A. Schneider, R. Teyssier, J. Stadel, N. E. Chisari, A. M. C. Le Brun, A. Amara, and A. Refregier, “Quantifying
baryon effects on the matter power spectrum and the weak lensing shear correlation,” JCAP 2019 (2019) no. 3, 020,
arXiv:1810.08629 [astro-ph.CO].
[959] G. Aricò, R. E. Angulo, C. Hernández-Monteagudo, S. Contreras, M. Zennaro, M. Pellejero-Ibañez, and
Y. Rosas-Guevara, “Modelling the large-scale mass density field of the universe as a function of cosmology and
baryonic physics,” MNRAS 495 (2020) no. 4, 4800–4819, arXiv:1911.08471 [astro-ph.CO].
[960] A. Schneider, N. Stoira, A. Refregier, A. J. Weiss, M. Knabenhans, J. Stadel, and R. Teyssier, “Baryonic effects for
weak lensing. Part I. Power spectrum and covariance matrix,” JCAP 2020 (2020) no. 4, 019, arXiv:1910.11357
[astro-ph.CO].
[961] J. M. Sullivan, U. Seljak, and S. Singh, “An analytic hybrid halo + perturbation theory model for small-scale
correlators: baryons, halos, and galaxies,” JCAP 2021 (2021) no. 11, 026, arXiv:2104.10676 [astro-ph.CO].
[962] A. M. C. Le Brun, I. G. McCarthy, J. Schaye, and T. J. Ponman, “Towards a realistic population of simulated galaxy
groups and clusters,” MNRAS 441 (2014) no. 2, 1270–1290, arXiv:1312.5462 [astro-ph.CO].
[963] I. G. McCarthy, S. Bird, J. Schaye, J. Harnois-Deraps, A. S. Font, and L. van Waerbeke, “The BAHAMAS project:
the CMB-large-scale structure tension and the roles of massive neutrinos and galaxy formation,” MNRAS 476 (2018)
no. 3, 2999–3030, arXiv:1712.02411 [astro-ph.CO].
[964] S. Kaviraj, C. Laigle, T. Kimm, J. E. G. Devriendt, Y. Dubois, C. Pichon, A. Slyz, E. Chisari, and S. Peirani, “The
Horizon-AGN simulation: evolution of galaxy properties over cosmic time,” MNRAS 467 (2017) no. 4, 4739–4752,
arXiv:1605.09379 [astro-ph.GA].
[965] M. P. van Daalen, I. G. McCarthy, and J. Schaye, “Exploring the effects of galaxy formation on matter clustering
through a library of simulation power spectra,” Mon. Not. Roy. Astron. Soc. 491 (2020) no. 2, 2424–2446,
arXiv:1906.00968 [astro-ph.CO].
[966] J. Schaye, R. Kugel, M. Schaller, J. C. Helly, J. Braspenning, W. Elbers, I. G. McCarthy, M. P. van Daalen,
B. Vandenbroucke, C. S. Frenk, J. Kwan, J. Salcido, Y. M. Bahé, J. Borrow, E. Chaikin, O. Hahn, F. Huško,
A. Jenkins, C. G. Lacey, and F. S. J. Nobels, “The FLAMINGO project: cosmological hydrodynamical simulations for
large-scale structure and galaxy cluster surveys,” MNRAS 526 (2023) no. 4, 4978–5020, arXiv:2306.04024
[astro-ph.CO].
291

[967] N. E. Chisari et al., “Modelling baryonic feedback for survey cosmology,” Open J. Astrophys. 2 (2019) no. 1, 4,
arXiv:1905.06082 [astro-ph.CO].
[968] A. Schneider and R. Teyssier, “A new method to quantify the effects of baryons on the matter power spectrum,” JCAP
2015 (2015) no. 12, 049–049, arXiv:1510.06034 [astro-ph.CO].
[969] K. R. Moran, K. Heitmann, E. Lawrence, S. Habib, D. Bingham, A. Upadhye, J. Kwan, D. Higdon, and R. Payne,
“The Mira–Titan Universe – IV. High-precision power spectrum emulation,” Mon. Not. Roy. Astron. Soc. 520 (2023)
no. 3, 3443–3458, arXiv:2207.12345 [astro-ph.CO].
[970] LSST Dark Energy Science Collaboration, N. E. Chisari et al., “Core Cosmology Library: Precision Cosmological
Predictions for LSST,” Astrophys. J. Suppl. 242 (2019) no. 1, 2, arXiv:1812.05995 [astro-ph.CO].
[971] Šarčević, Nikolina, “Modelling and Mitigating the Systematics in Weak Lensing Measurements.”
https://doi.org/10.5281/zenodo.14602273, 2024.
[972] G. Aricò, R. E. Angulo, M. Zennaro, S. Contreras, A. Chen, and C. Hernández-Monteagudo, “DES Y3 cosmic shear
down to small scales: Constraints on cosmology and baryons,” Astron. Astrophys. 678 (2023) A109, arXiv:2303.05537
[astro-ph.CO].
[973] C. Fedeli, “The clustering of baryonic matter. I: a halo-model approach,” JCAP 2014 (2014) no. 4, 028,
arXiv:1401.2997 [astro-ph.CO].
[974] H.-J. Huang, T. Eifler, R. Mandelbaum, and S. Dodelson, “Modelling baryonic physics in future weak lensing surveys,”
Mon. Not. Roy. Astron. Soc. 488 (2019) no. 2, 1652–1678, arXiv:1809.01146 [astro-ph.CO].
[975] J. C. Broxterman et al., “The FLAMINGO project: baryonic impact on weak gravitational lensing convergence peak
counts,” Mon. Not. Roy. Astron. Soc. 529 (2024) no. 3, 2309–2326, arXiv:2312.08450 [astro-ph.CO].
[976] S. Foreman, W. Coulton, F. Villaescusa-Navarro, and A. Barreira, “Baryonic effects on the matter bispectrum,” Mon.
Not. Roy. Astron. Soc. 498 (2020) no. 2, 2887–2911, arXiv:1910.03597 [astro-ph.CO].
[977] N. Martinet, T. Castro, J. Harnois-Déraps, E. Jullo, C. Giocoli, and K. Dolag, “Impact of baryons in cosmic shear
analyses with tomographic aperture mass statistics,” Astron. Astrophys. 648 (2021) A115, arXiv:2012.09614
[astro-ph.CO].
[978] T. Lu and Z. Haiman, “The impact of baryons on cosmological inference from weak lensing statistics,” Mon. Not. Roy.
Astron. Soc. 506 (2021) no. 3, 3406–3417, arXiv:2104.04165 [astro-ph.CO].
[979] E. Semboloni, H. Hoekstra, and J. Schaye, “Effect of baryonic feedback on two- and three-point shear statistics:
prospects for detection and improved modelling,” Mon. Not. Roy. Astron. Soc. 434 (2013) 148, arXiv:1210.7303
[astro-ph.CO].
[980] W. Elbers et al., “The FLAMINGO project: the coupling between baryonic feedback and cosmology in light of the S8
tension,” Mon. Not. Roy. Astron. Soc. 537 (2025) no. 2, 2160–2178, arXiv:2403.12967 [astro-ph.CO].
[981] DES Collaboration, L. Bigwood et al., “Weak lensing combined with the kinetic Sunyaev–Zel’dovich effect: a study of
baryonic feedback,” Mon. Not. Roy. Astron. Soc. 534 (2024) no. 1, 655–682, arXiv:2404.06098 [astro-ph.CO].
[982] I. S. Khrykin et al., “FLIMFLAM DR1: The First Constraints on the Cosmic Baryon Distribution from Eight Fast
Radio Burst Sight Lines,” Astrophys. J. 973 (2024) no. 2, 151, arXiv:2402.00505 [astro-ph.GA].
[983] A. Amon et al., “Consistent lensing and clustering in a low-S8 Universe with BOSS, DES Year 3, HSC Year 1, and
KiDS-1000,” MNRAS 518 (2023) no. 1, 477–503, arXiv:2202.07440 [astro-ph.CO].
[984] K.-G. Lee, M. Ata, I. S. Khrykin, Y. Huang, J. X. Prochaska, J. Cooke, J. Zhang, and A. Batten, “Constraining the
Cosmic Baryon Distribution with Fast Radio Burst Foreground Mapping,” Astrophys. J. 928 (2022) no. 1, 9,
arXiv:2109.00386 [astro-ph.CO].
[985] C. García-García, M. Zennaro, G. Aricò, D. Alonso, and R. E. Angulo, “Cosmic shear with small scales: DES-Y3,
KiDS-1000 and HSC-DR1,” JCAP 08 (2024) 024, arXiv:2403.13794 [astro-ph.CO].
[986] R. Terasawa et al., “Exploring the baryonic effect signature in the Hyper Suprime-Cam Year 3 cosmic shear two-point
correlations on small scales: The S8 tension remains present,” Phys. Rev. D 111 (2025) no. 6, 063509,
arXiv:2403.20323 [astro-ph.CO].
[987] A. Amon and G. Efstathiou, “A non-linear solution to the S8 tension?,” Mon. Not. Roy. Astron. Soc. 516 (2022) no. 4,
5355–5366, arXiv:2206.11794 [astro-ph.CO].
[988] T. Nishimichi et al., “Dark Quest. I. Fast and Accurate Emulation of Halo Clustering Statistics and Its Application to
Galaxy Clustering,” Astrophys. J. 884 (2019) 29, arXiv:1811.09504 [astro-ph.CO].
[989] S. Amodeo et al., “Atacama Cosmology Telescope: Modeling the gas thermodynamics in BOSS CMASS galaxies from
kinematic and thermal Sunyaev-Zel’dovich measurements,” Phys. Rev. D 103 (2021) no. 6, 063514,
arXiv:Amodeo:2020mmu [astro-ph.CO]. [Erratum: Phys.Rev.D 107, 063514 (2023)].
[990] C. Lamman, E. Tsaprazi, J. Shi, N. N. Šarčević, S. Pyne, E. Legnani, and T. Ferreira, “The IA Guide: A Breakdown of
Intrinsic Alignment Formalisms,” arXiv:2309.08605 [astro-ph.CO].
[991] M. A. Troxel and M. Ishak, “The Intrinsic Alignment of Galaxies and its Impact on Weak Gravitational Lensing in an
Era of Precision Cosmology,” Phys. Rept. 558 (2014) 1–59, arXiv:1407.6990 [astro-ph.CO].
[992] J. Prat, J. Zuntz, C. Chang, T. Tröster, E. Pedersen, C. García-García, E. Phillips-Longley, J. Sanchez, D. Alonso,
X. Fang, E. Gawiser, K. Heitmann, M. Ishak, M. Jarvis, E. Kovacs, P. Larsen, Y. Y. Mao, L. Medina Varela,
M. Paterno, S. D. Vitenti, Z. Zhang, and LSST Dark Energy Science Collaboration, “The catalog-to-cosmology
framework for weak lensing and galaxy clustering for LSST,” The Open Journal of Astrophysics 6 (2023) 13,
arXiv:2212.09345 [astro-ph.CO].
[993] DES Collaboration, C. Sánchez et al., “Dark Energy Survey Year 3 results: Exploiting small-scale information with
lensing shear ratios,” Phys. Rev. D 105 (2022) no. 8, 083529, arXiv:2105.13542 [astro-ph.CO].
292

[994] D. N. Cross and C. Sánchez, “Inverse galaxy-galaxy lensing: Magnification, intrinsic alignments, and cosmology,” Phys.
Rev. D 110 (2024) no. 12, 123534, arXiv:2410.00714 [astro-ph.CO].
[995] S. Bridle and L. King, “Dark energy constraints from cosmic shear power spectra: impact of intrinsic alignments on
photometric redshift requirements,” New J. Phys. 9 (2007) 444, arXiv:0705.0166 [astro-ph].
[996] J. Blazek, N. MacCrann, M. A. Troxel, and X. Fang, “Beyond linear galaxy alignments,” Phys. Rev. D 100 (2019)
no. 10, 103506, arXiv:1708.09247 [astro-ph.CO].
[997] S. Samuroff, R. Mandelbaum, and J. Blazek, “Advances in constraining intrinsic alignment models with hydrodynamic
simulations,” Mon. Not. Roy. Astron. Soc. 508 (2021) no. 1, 637–664, arXiv:2009.10735 [astro-ph.CO].
[998] DES Collaboration, S. Samuroff et al., “The Dark Energy Survey Year 3 and eBOSS: constraining galaxy intrinsic
alignments across luminosity and colour space,” Mon. Not. Roy. Astron. Soc. 524 (2023) no. 2, 2195–2223,
arXiv:2212.11319 [astro-ph.CO].
[999] DES Collaboration, K. Hoffmann et al., “Modeling intrinsic galaxy alignment in the MICE simulation,” Phys. Rev. D
106 (2022) no. 12, 123510, arXiv:2206.14219 [astro-ph.CO].
[1000] M. C. Fortuna, H. Hoekstra, B. Joachimi, H. Johnston, N. E. Chisari, C. Georgiou, and C. Mahony, “The halo model
as a versatile tool to predict intrinsic alignments,” Mon. Not. Roy. Astron. Soc. 501 (2021) no. 2, 2983–3002,
arXiv:2003.02700 [astro-ph.CO].
[1001] T. Bakx, T. Kurita, N. E. Chisari, Z. Vlah, and F. Schmidt, “Effective field theory of intrinsic alignments at one loop
order: a comparison to dark matter simulations,” JCAP 10 (2023) 005, arXiv:2303.15565 [astro-ph.CO].
[1002] C. D. Leonard, M. M. Rau, and R. Mandelbaum, “Photometric redshifts and intrinsic alignments: Degeneracies and
biases in the 3×2pt analysis,” Phys. Rev. D 109 (2024) no. 8, 083528, arXiv:2401.06060 [astro-ph.CO].
[1003] LSST Dark Energy Science Collaboration, Šarčević, Nikolina and Leonard, C. Danielle and Rau, Markus M., “Joint
Modelling of Astrophysical Systematics for Cosmology with LSST Cosmic Shear,” arXiv:2406.03352 [astro-ph.CO].
[1004] A. Campos, S. Samuroff, and R. Mandelbaum, “An empirical approach to model selection: weak lensing and intrinsic
alignments,” Mon. Not. Roy. Astron. Soc. 525 (2023) no. 2, 1885–1901, arXiv:2211.02800 [astro-ph.CO].
[1005] P. M. Merkel and B. M. Schäfer, “A theoretical estimate of intrinsic ellipticity bispectra induced by angular momenta
alignments,” Mon. Not. Roy. Astron. Soc. 445 (2014) no. 3, 2918–2929, arXiv:1409.5197 [astro-ph.CO].
[1006] P. A. Burger et al., “KiDS-1000 cosmology: Combined second- and third-order shear statistics,” Astron. Astrophys.
683 (2024) A103, arXiv:2309.08602 [astro-ph.CO].
[1007] A. Barthelemy, A. Halder, Z. Gong, and C. Uhlemann, “Making the leap. Part I. Modelling the reconstructed lensing
convergence PDF from cosmic shear with survey masks and systematics,” JCAP 03 (2024) 060, arXiv:2307.09468
[astro-ph.CO].
[1008] J. Harnois-Déraps, N. Martinet, and R. Reischke, “Cosmic shear beyond 2-point statistics: Accounting for galaxy
intrinsic alignment with projected tidal fields,” Mon. Not. Roy. Astron. Soc. 509 (2022) 3868–3868, arXiv:2107.08041
[astro-ph.CO].
[1009] DES Collaboration, M. Gatti et al., “Detection of the significant impact of source clustering on higher order statistics
with DES Year 3 weak gravitational lensing data,” Mon. Not. Roy. Astron. Soc. 527 (2024) no. 1, L115–L121,
arXiv:2307.13860 [astro-ph.CO].
[1010] KiDS-1000, Euclid Collaboration, L. Linke et al., “Euclid and KiDS-1000: Quantifying the impact of source-lens
clustering on cosmic shear analyses,” Astron. Astrophys. 693 (2025) A210, arXiv:2407.09810 [astro-ph.CO].
[1011] E. Krause, T. Eifler, and J. Blazek, “The impact of intrinsic alignment on current and future cosmic shear surveys,”
Mon. Not. Roy. Astron. Soc. 456 (2016) no. 1, 207–222, arXiv:1506.08730 [astro-ph.CO].
[1012] DES Collaboration, S. Samuroff et al., “Dark Energy Survey Year 1 Results: Constraints on Intrinsic Alignments and
their Colour Dependence from Galaxy Clustering and Weak Lensing,” Mon. Not. Roy. Astron. Soc. 489 (2019) no. 4,
5453–5482, arXiv:1811.06989 [astro-ph.CO].
[1013] DES Collaboration, J. McCullough et al., “Dark Energy Survey Year 3: Blue Shear,” arXiv:2410.22272
[astro-ph.CO].
[1014] DES Collaboration, J. Muir et al., “Blinding multiprobe cosmological experiments,” Mon. Not. Roy. Astron. Soc. 494
(2020) no. 3, 4454–4470, arXiv:1911.05929 [astro-ph.CO].
[1015] S. W. Allen, A. E. Evrard, and A. B. Mantz, “Cosmological Parameters from Observations of Galaxy Clusters,” Ann.
Rev. Astron. Astrophys. 49 (2011) 409–470, arXiv:1103.4829 [astro-ph.CO].
[1016] W. H. Press and P. Schechter, “Formation of galaxies and clusters of galaxies by selfsimilar gravitational
condensation,” Astrophys. J. 187 (1974) 425–438.
[1017] R. K. Sheth and G. Tormen, “Large scale bias and the peak background split,” Mon. Not. Roy. Astron. Soc. 308
(1999) 119, arXiv:astro-ph/9901122.
[1018] A. Jenkins, C. S. Frenk, S. D. M. White, J. M. Colberg, S. Cole, A. E. Evrard, H. M. P. Couchman, and N. Yoshida,
“The Mass function of dark matter halos,” Mon. Not. Roy. Astron. Soc. 321 (2001) 372, arXiv:astro-ph/0005260.
[1019] J. L. Tinker, A. V. Kravtsov, A. Klypin, K. Abazajian, M. S. Warren, G. Yepes, S. Gottlober, and D. E. Holz, “Toward
a halo mass function for precision cosmology: The Limits of universality,” Astrophys. J. 688 (2008) 709–728,
arXiv:0803.2706 [astro-ph].
[1020] W. A. Watson, I. T. Iliev, A. D’Aloisio, A. Knebe, P. R. Shapiro, and G. Yepes, “The halo mass function through the
cosmic ages,” Mon. Not. Roy. Astron. Soc. 433 (2013) 1230, arXiv:1212.0095 [astro-ph.CO].
[1021] G. Despali, C. Giocoli, R. E. Angulo, G. Tormen, R. K. Sheth, G. Baso, and L. Moscardini, “The universality of the
virial halo mass function and models for non-universality of other halo definitions,” Mon. Not. Roy. Astron. Soc. 456
(2016) no. 3, 2486–2504, arXiv:1507.05627 [astro-ph.CO].
293

[1022] N. Kaiser, “Evolution and clustering of rich clusters,” Mon. Not. Roy. Astron. Soc. 222 (1986) 323–345.
[1023] N. Truong et al., “Cosmological hydrodynamical simulations of galaxy clusters: X-ray scaling relations and their
evolution,” Mon. Not. Roy. Astron. Soc. 474 (2018) no. 3, 4089–4111, arXiv:1607.00019 [astro-ph.CO].
[1024] A.-R. Pop et al., “Sunyaev-Zel’dovich effect and X-ray scaling relations of galaxies, groups and clusters in the
IllustrisTNG simulations,” arXiv:2205.11528 [astro-ph.GA].
[1025] A. Pellissier, O. Hahn, and C. Ferrari, “Rhapsody-C simulations – Anisotropic thermal conduction, black hole physics,
and the robustness of massive galaxy cluster scaling relations,” Mon. Not. Roy. Astron. Soc. 522 (2023) no. 1, 721–749,
arXiv:2301.02684 [astro-ph.CO].
[1026] J. Braspenning et al., “The FLAMINGO project: galaxy clusters in comparison to X-ray observations,” Mon. Not.
Roy. Astron. Soc. 533 (2024) no. 3, 2656–2676, arXiv:2312.08277 [astro-ph.GA].
[1027] A. B. Mantz et al., “Weighing the giants – IV. Cosmology and neutrino mass,” Mon. Not. Roy. Astron. Soc. 446
(2015) 2205–2225, arXiv:1407.4516 [astro-ph.CO].
[1028] SPT Collaboration, S. Bocquet et al., “Mass Calibration and Cosmological Analysis of the SPT-SZ Galaxy Cluster
Sample Using Velocity Dispersion σv and X-ray YX Measurements,” Astrophys. J. 799 (2015) no. 2, 214,
arXiv:1407.2942 [astro-ph.CO].
[1029] SPT Collaboration, S. Bocquet et al., “Cluster Cosmology Constraints from the 2500 deg2 SPT-SZ Survey: Inclusion
of Weak Gravitational Lensing Data from Magellan and the Hubble Space Telescope,” Astrophys. J. 878 (2019) no. 1,
55, arXiv:1812.01679 [astro-ph.CO].
[1030] I.-N. Chiu, M. Klein, J. Mohr, and S. Bocquet, “Cosmological constraints from galaxy clusters and groups in the
eROSITA final equatorial depth survey,” Mon. Not. Roy. Astron. Soc. 522 (2023) no. 2, 1601–1642, arXiv:2207.12429
[astro-ph.CO].
[1031] DES, SPT Collaboration, S. Bocquet et al., “SPT clusters with DES and HST weak lensing. I. Cluster lensing and
Bayesian population modeling of multiwavelength cluster datasets,” Phys. Rev. D 110 (2024) no. 8, 083509,
arXiv:2310.12213 [astro-ph.CO].
[1032] SPT, DES Collaboration, S. Bocquet et al., “SPT clusters with DES and HST weak lensing. II. Cosmological
constraints from the abundance of massive halos,” Phys. Rev. D 110 (2024) no. 8, 083510, arXiv:2401.02075
[astro-ph.CO].
[1033] V. Ghirardini et al., “The SRG/eROSITA all-sky survey: Cosmology constraints from cluster abundances in the
western Galactic hemisphere,” Astron. Astrophys. 689 (2024) A298, arXiv:2402.08458 [astro-ph.CO].
[1034] XXL Collaboration, F. Pacaud et al., “The XXL Survey: XXV. Cosmological analysis of the C1 cluster number
counts,” Astron. Astrophys. 620 (2018) A10, arXiv:1810.01624 [astro-ph.CO].
[1035] DES Collaboration, M. Costanzi et al., “Modelling projection effects in optically selected cluster catalogues,” Mon.
Not. Roy. Astron. Soc. 482 (2019) no. 1, 490–505, arXiv:1807.07072 [astro-ph.CO].
[1036] DES Collaboration, T. M. C. Abbott et al., “Dark Energy Survey Year 1 Results: Cosmological constraints from
cluster abundances and weak lensing,” Phys. Rev. D 102 (2020) no. 2, 023509, arXiv:2002.11124 [astro-ph.CO].
[1037] DES Collaboration, C. To et al., “Dark Energy Survey Year 1 Results: Cosmological Constraints from Cluster
Abundances, Weak Lensing, and Galaxy Correlations,” Phys. Rev. Lett. 126 (2021) 141301, arXiv:2010.01138
[astro-ph.CO].
[1038] C. Garrel et al., “The XXL survey - XLVI. Forward cosmological analysis of the C1 cluster sample,” Astron. Astrophys.
663 (2022) A3, arXiv:2109.13171 [astro-ph.CO].
[1039] G. F. Lesci et al., “AMICO galaxy clusters in KiDS-DR3: Cosmological constraints from counts and stacked weak
lensing,” Astron. Astrophys. 659 (2022) A88, arXiv:2012.12273 [astro-ph.CO].
[1040] Y. Park, T. Sunayama, M. Takada, Y. Kobayashi, H. Miyatake, S. More, T. Nishimichi, and S. Sugiyama, “Cluster
cosmology with anisotropic boosts: validation of a novel forward modelling analysis and application on SDSS
redMaPPer clusters,” Mon. Not. Roy. Astron. Soc. 518 (2022) no. 4, 5171–5189, arXiv:2112.09059 [astro-ph.CO].
[1041] T. Sunayama et al., “Optical cluster cosmology with SDSS redMaPPer clusters and HSC-Y3 lensing measurements,”
Phys. Rev. D 110 (2024) no. 8, 083511, arXiv:2309.13025 [astro-ph.CO].
[1042] W. Voges et al., “The ROSAT all - sky survey bright source catalogue,” Astron. Astrophys. 349 (1999) 389,
arXiv:astro-ph/9909315.
[1043] S. Borgani, P. Rosati, P. Tozzi, and C. Norman, “Cosmological constraints from the rosat deep cluster survey,”
Astrophys. J. 517 (1999) 40, arXiv:astro-ph/9901017.
[1044] R. Piffaretti, M. Arnaud, G. W. Pratt, E. Pointecouteau, and J. B. Melin, “The MCXC: a Meta-Catalogue of X-ray
detected Clusters of galaxies,” Astron. Astrophys. 534 (2011) A109, arXiv:1007.1916 [astro-ph.CO].
[1045] S. Borgani, P. Rosati, P. Tozzi, S. A. Stanford, P. E. Eisenhardt, C. Lidman, B. Holden, R. Della Ceca, C. Norman,
and G. Squires, “Measuring omega_m with the rosat deep cluster survey,” Astrophys. J. 561 (2001) 13–21,
arXiv:astro-ph/0106428.
[1046] H. Bohringer et al., “The Representative XMM-Newton Cluster Structure Survey (REXCESS) of an X-ray Luminosity
Selected Galaxy Cluster Sample,” Astron. Astrophys. 469 (2007) 363–377, arXiv:astro-ph/0703553.
[1047] T. H. Reiprich and H. Boehringer, “The Mass function of an X-ray flux-limited sample of galaxy clusters,” Astrophys.
J. 567 (2002) 716–740, arXiv:astro-ph/0111285.
[1048] G. Schellenberger and T. H. Reiprich, “HICOSMO – cosmology with a complete sample of galaxy clusters – I. Data
analysis, sample selection and luminosity–mass scaling relation,” Mon. Not. Roy. Astron. Soc. 469 (2017) no. 3,
3738–3761, arXiv:1705.05842 [astro-ph.CO].
[1049] A. Vikhlinin et al., “Chandra Cluster Cosmology Project III: Cosmological Parameter Constraints,” Astrophys. J. 692
294

(2009) 1060–1074, arXiv:0812.2720 [astro-ph].


[1050] M. Pierre et al., “The XXL Survey - I. Scientific motivations − XMM-Newton observing plan − Follow-up observations
and simulation programme,” Astron. Astrophys. 592 (2016) A1, arXiv:1512.04317 [astro-ph.CO].
[1051] R. A. Sunyaev and Y. B. Zeldovich, “The Observations of relic radiation as a test of the nature of X-Ray radiation
from the clusters of galaxies,” Comments Astrophys. Space Phys. 4 (1972) 173–178.
[1052] T. Mroczkowski et al., “Astrophysics with the Spatially and Spectrally Resolved Sunyaev-Zeldovich Effects: A
Millimetre/Submillimetre Probe of the Warm and Hot Universe,” Space Sci. Rev. 215 (2019) no. 1, 17,
arXiv:1811.02310 [astro-ph.CO].
[1053] A. V. Kravtsov, A. Vikhlinin, and D. Nagai, “A new robust low-scatter x-ray mass indicator for clusters of galaxies,”
Astrophys. J. 650 (2006) 128–136, arXiv:astro-ph/0603205.
[1054] D. Nagai, A. V. Kravtsov, and A. Vikhlinin, “Effects of Galaxy Formation on Thermodynamics of the Intracluster
Medium,” Astrophys. J. 668 (2007) 1–14, arXiv:astro-ph/0703661.
[1055] M. Arnaud, E. Pointecouteau, and G. W. Pratt, “Calibration of the galaxy cluster M_500-Y_X relation with
XMM-Newton,” Astron. Astrophys. 474 (2007) L37, arXiv:0709.1561 [astro-ph].
[1056] D. P. Marrone et al., “LoCuSS: The Sunyaev-Zel’dovich Effect and Weak Lensing Mass Scaling Relation,” Astrophys.
J. 754 (2012) 119, arXiv:1107.5115 [astro-ph.CO].
[1057] Planck Collaboration, P. A. R. Ade et al., “Planck Early Results XI: Calibration of the local galaxy cluster
Sunyaev-Zeldovich scaling relations,” Astron. Astrophys. 536 (2011) A11, arXiv:1101.2026 [astro-ph.CO].
[1058] Planck Collaboration, P. A. R. Ade et al., “Planck 2013 results. XXIX. The Planck catalogue of Sunyaev-Zeldovich
sources,” Astron. Astrophys. 571 (2014) A29, arXiv:1303.5089 [astro-ph.CO].
[1059] R. Williamson et al., “An SZ-selected sample of the most massive galaxy clusters in the 2500-square-degree South Pole
Telescope survey,” Astrophys. J. 738 (2011) 139, arXiv:1101.1290 [astro-ph.CO].
[1060] SPT Collaboration, L. E. Bleem et al., “Galaxy Clusters Discovered via the Sunyaev-Zel’dovich Effect in the
2500-square-degree SPT-SZ survey,” Astrophys. J. Suppl. 216 (2015) no. 2, 27, arXiv:1409.0850 [astro-ph.CO].
[1061] SPT, DES Collaboration, L. E. Bleem et al., “Galaxy Clusters Discovered via the Thermal Sunyaev-Zel’dovich Effect
in the 500-square-degree SPTpol Survey,” Open J. Astrophys. 7 (2024) astro.2311.07512, arXiv:2311.07512
[astro-ph.CO].
[1062] M. Hasselfield et al., “The Atacama Cosmology Telescope: Sunyaev-Zel’dovich selected galaxyclusters at 148 GHz from
three seasons of data,” JCAP 07 (2013) 008, arXiv:1301.0816 [astro-ph.CO].
[1063] ACT, DES Collaboration, M. Hilton et al., “The Atacama Cosmology Telescope: A Catalog of >4000
Sunyaev–Zel’dovich Galaxy Clusters,” Astrophys. J. Suppl. 253 (2021) no. 1, 3, arXiv:2009.11043 [astro-ph.CO].
[1064] G. O. Abell, H. G. Corwin, Jr., and R. P. Olowin, “A Catalog of rich clusters of galaxies,” Astrophys. J. Suppl. 70
(1989) 1.
[1065] SDSS Collaboration, B. Koester et al., “A MaxBCG Catalog of 13,823 Galaxy Clusters from the Sloan Digital Sky
Survey,” Astrophys. J. 660 (2007) 239–255, arXiv:astro-ph/0701265.
[1066] SDSS Collaboration, E. S. Rykoff et al., “REDMAPPER I: Algorithm and SDSS DR8 Catalog,” Astrophys. J. 785
(2014) 104, arXiv:1303.3562 [astro-ph.CO].
[1067] M. Oguri, “A cluster finding algorithm based on the multiband identification of red sequence galaxies,” Mon. Not. Roy.
Astron. Soc. 444 (2014) no. 1, 147–161, arXiv:1407.4693 [astro-ph.CO].
[1068] C. P. Haines et al., “LoCuSS: The slow quenching of star formation in cluster galaxies and the need for pre-processing,”
Astrophys. J. 806 (2015) no. 1, 101, arXiv:1504.05604 [astro-ph.GA].
[1069] P. A. A. Lopes, A. L. B. Ribeiro, and S. B. Rembold, “NoSOCS in SDSS – IV. The role of environment beyond the
extent of galaxy clusters,” Mon. Not. Roy. Astron. Soc. 437 (2014) no. 3, 2430–2447, arXiv:1310.6309
[astro-ph.CO].
[1070] F. Sarron, N. Martinet, F. Durret, and C. Adami, “Evolution of the cluster optical galaxy luminosity function in the
CFHTLS : breaking the degeneracy between mass and redshift,” Astron. Astrophys. 613 (2018) A67,
arXiv:1712.09481 [astro-ph.GA].
[1071] Euclid Collaboration, R. Adam et al., “Euclid preparation. III. Galaxy cluster detection in the wide photometric
survey, performance and algorithm selection,” Astron. Astrophys. 627 (2019) no. 627, A23, arXiv:1906.04707
[astro-ph.CO].
[1072] DES Collaboration, M. Costanzi et al., “Methods for cluster cosmology and application to the SDSS in preparation for
DES Year 1 release,” Mon. Not. Roy. Astron. Soc. 488 (2019) no. 4, 4779–4800, arXiv:1810.09456 [astro-ph.CO].
[1073] Z. L. Wen and J. L. Han, “Clusters of galaxies up to z = 1.5 identified from photometric data of the Dark Energy
Survey and unWISE,” Mon. Not. Roy. Astron. Soc. 513 (2022) no. 3, 3946–3959, arXiv:2204.11215 [astro-ph.CO].
[1074] M. Maturi, F. Bellagamba, M. Radovich, M. Roncarelli, M. Sereno, L. Moscardini, S. Bardelli, and E. Puddu, “AMICO
galaxy clusters in KiDS-DR3: sample properties and selection function,” Mon. Not. Roy. Astron. Soc. 485 (2019) 498,
arXiv:1810.02811 [astro-ph.CO].
[1075] M. Oguri et al., “An optically-selected cluster catalog at redshift 0.1< z< 1.1 from the Hyper Suprime-Cam Subaru
Strategic Program S16A data,” Publ. Astron. Soc. Jap. 70 (2018) S20, arXiv:1701.00818 [astro-ph.CO].
[1076] Z. L. Wen and J. L. Han, “A Catalog of 1.58 Million Clusters of Galaxies Identified from the DESI Legacy Imaging
Surveys,” Astrophys. J. Suppl. 272 (2024) no. 2, 39, arXiv:2404.02002 [astro-ph.CO].
[1077] M. Simet, T. McClintock, R. Mandelbaum, E. Rozo, E. Rykoff, E. Sheldon, and R. H. Wechsler, “Weak lensing
measurement of the mass–richness relation of SDSS redMaPPer clusters,” Mon. Not. Roy. Astron. Soc. 466 (2017)
no. 3, 3103–3118, arXiv:1603.06953 [astro-ph.CO].
295

[1078] DES Collaboration, T. McClintock et al., “Dark Energy Survey Year 1 Results: Weak Lensing Mass Calibration of
redMaPPer Galaxy Clusters,” Mon. Not. Roy. Astron. Soc. 482 (2019) no. 1, 1352–1378, arXiv:1805.00039
[astro-ph.CO].
[1079] C. Ge, M. Sun, E. Rozo, N. Sehgal, A. Vikhlinin, W. Forman, C. Jones, and D. Nagai, “X-ray scaling relations from a
complete sample of the richest maxBCG clusters,” Mon. Not. Roy. Astron. Soc. 484 (2019) no. 2, 1946–1971,
arXiv:1803.05007 [astro-ph.GA].
[1080] DES Collaboration, V. Wetzell et al., “Velocity dispersions of clusters in the Dark Energy Survey Y3 redMaPPer
catalogue,” Mon. Not. Roy. Astron. Soc. 514 (2022) no. 4, 4696–4717, arXiv:2107.07631 [astro-ph.CO].
[1081] S. Miyazaki, T. Hamana, R. S. Ellis, N. Kashikawa, R. J. Massey, J. Taylor, and A. Refregier, “A Subaru Weak
Lensing Survey I: Cluster Candidates and Spectroscopic Verification,” Astrophys. J. 669 (2007) 714, arXiv:0707.2249
[astro-ph].
[1082] S. Miyazaki et al., “A large sample of shear selected clusters from the Hyper Suprime-Cam Subaru Strategic Program
S16A wide field mass maps,” Publ. Astron. Soc. Jap. 70 (2018) S27, arXiv:1802.10290 [astro-ph.CO].
[1083] K.-F. Chen et al., “Weak-Lensing Shear-Selected Galaxy Clusters from the Hyper Suprime-Cam Subaru Strategic
Program: I. Cluster Catalog, Selection Function and Mass–Observable Relation,” arXiv:2406.11966 [astro-ph.CO].
[1084] G. Leroy, S. Pires, G. W. Pratt, and C. Giocoli, “Fast multi-scale galaxy cluster detection with weak lensing: Towards
a mass-selected sample,” Astron. Astrophys. 678 (2023) A125, arXiv:2304.01812 [astro-ph.CO].
[1085] M. E. Ramos-Ceja et al., “The eROSITA Final Equatorial-Depth Survey (eFEDS) - A complete census of X-ray
properties of Subaru Hyper Suprime-Cam weak lensing shear-selected clusters in the eFEDS footprint,” Astron.
Astrophys. 661 (2022) A14, arXiv:2109.07836 [astro-ph.CO].
[1086] I.-N. Chiu et al., “Weak-lensing Shear-selected Galaxy Clusters from the Hyper Suprime-Cam Subaru Strategic
Program: II. Cosmological Constraints from the Cluster Abundance,” arXiv:2406.11970 [astro-ph.CO].
[1087] K.-F. Chen, M. Oguri, Y.-T. Lin, and S. Miyazaki, “Mass Bias of Weak Lensing Shear-selected Galaxy Cluster
Samples,” Astrophys. J. 891 (2020) no. 2, 139, arXiv:1911.11480 [astro-ph.GA].
[1088] K. Umetsu, “Cluster–galaxy weak lensing,” Astron. Astrophys. Rev. 28 (2020) no. 1, 7, arXiv:2007.00506
[astro-ph.CO].
[1089] DES Collaboration, H.-Y. Wu et al., “Optical selection bias and projection effects in stacked galaxy cluster weak
lensing,” Mon. Not. Roy. Astron. Soc. 515 (2022) no. 3, 4471–4486, arXiv:2203.05416 [astro-ph.CO].
[1090] DES, eROSITA-DE Collaboration, S. Grandis et al., “The SRG/eROSITA All-Sky Survey - Dark Energy Survey
year 3 weak gravitational lensing by eRASS1 selected galaxy clusters,” Astron. Astrophys. 687 (2024) A178,
arXiv:2402.08455 [astro-ph.CO].
[1091] H. Hoekstra, “How well can we determine cluster mass profiles from weak lensing?,” Mon. Not. Roy. Astron. Soc. 339
(2003) 1155, arXiv:astro-ph/0208351.
[1092] D. Gruen, S. Seitz, M. R. Becker, O. Friedrich, and A. Mana, “Cosmic variance of the galaxy cluster weak lensing
signal,” Mon. Not. Roy. Astron. Soc. 449 (2015) no. 4, 4264–4276, arXiv:1501.01632 [astro-ph.CO].
[1093] D. E. Applegate, A. von der Linden, P. L. Kelly, M. T. Allen, S. W. Allen, P. R. Burchat, D. L. Burke, H. Ebeling,
A. Mantz, and R. G. Morris, “Weighing the Giants – III. Methods and measurements of accurate galaxy cluster
weak-lensing masses,” Mon. Not. Roy. Astron. Soc. 439 (2014) no. 1, 48–72, arXiv:1208.0605 [astro-ph.CO].
[1094] SPT Collaboration, J. P. Dietrich et al., “Sunyaev–Zel’dovich effect and X-ray scaling relations from weak lensing
mass calibration of 32 South Pole Telescope selected galaxy clusters,” Mon. Not. Roy. Astron. Soc. 483 (2019) no. 3,
2871–2906, arXiv:1711.05344 [astro-ph.CO].
[1095] S. Grandis, S. Bocquet, J. J. Mohr, M. Klein, and K. Dolag, “Calibration of bias and scatter involved in cluster mass
measurements using optical weak gravitational lensing,” Mon. Not. Roy. Astron. Soc. 507 (2021) no. 4, 5671–5689,
arXiv:2103.16212 [astro-ph.CO].
[1096] M. R. Becker and A. V. Kravtsov, “On the Accuracy of Weak Lensing Cluster Mass Reconstructions,” Astrophys. J.
740 (2011) 25, arXiv:1011.1681 [astro-ph.CO].
[1097] Y. M. Bahe, I. G. McCarthy, and L. J. King, “Mock weak lensing analysis of simulated galaxy clusters: bias and scatter
in mass and concentration,” Mon. Not. Roy. Astron. Soc. 421 (2012) 1073–1088, arXiv:1106.2046 [astro-ph.CO].
[1098] Euclid Collaboration, C. Giocoli et al., “Euclid preparation - XXXII. Evaluating the weak-lensing cluster mass biases
using the Three Hundred Project hydrodynamical simulations,” Astron. Astrophys. 681 (2024) A67,
arXiv:2302.00687 [astro-ph.CO].
[1099] I.-N. Chiu et al., “The eROSITA Final Equatorial-Depth Survey (eFEDS) - X-ray observable-to-mass-and-redshift
relations of galaxy clusters and groups with weak-lensing mass calibration from the Hyper Suprime-Cam Subaru
Strategic Program survey,” Astron. Astrophys. 661 (2022) A11, arXiv:2107.05652 [astro-ph.CO].
[1100] F. Kleinebreil et al., “The SRG/eROSITA All-Sky Survey: Weak-Lensing of eRASS1 Galaxy Clusters in KiDS-1000
and Consistency Checks with DES Y3 & HSC-Y3,” arXiv:2402.08456 [astro-ph.CO].
[1101] F. Bellagamba et al., “AMICO galaxy clusters in KiDS-DR3: weak-lensing mass calibration,” Mon. Not. Roy. Astron.
Soc. 484 (2019) no. 2, 1598–1615, arXiv:1810.02827 [astro-ph.CO].
[1102] W. Cui, S. Borgani, K. Dolag, G. Murante, and L. Tornatore, “The effects of baryons on the halo mass function,” Mon.
Not. Roy. Astron. Soc. 423 (2012) 2279, arXiv:1111.3066 [astro-ph.CO].
[1103] S. Bocquet, A. Saro, K. Dolag, and J. J. Mohr, “Halo mass function: Baryon impact, fitting formulae and implications
for cluster cosmology,” Mon. Not. Roy. Astron. Soc. 456 (2016) no. 3, 2361–2373, arXiv:1502.07357 [astro-ph.CO].
[1104] T. Castro, S. Borgani, K. Dolag, V. Marra, M. Quartin, A. Saro, and E. Sefusatti, “On the impact of baryons on the
halo mass function, bias, and cluster cosmology,” Mon. Not. Roy. Astron. Soc. 500 (2020) no. 2, 2316–2335,
296

arXiv:2009.01775 [astro-ph.CO].
[1105] DES, SPT Collaboration, M. Costanzi et al., “Cosmological constraints from DES Y1 cluster abundances and SPT
multiwavelength data,” Phys. Rev. D 103 (2021) no. 4, 043522, arXiv:2010.13800 [astro-ph.CO].
[1106] SPT Collaboration, A. B. Mantz et al., “Cosmological constraints from gas mass fractions of massive, relaxed galaxy
clusters,” Mon. Not. Roy. Astron. Soc. 510 (2021) no. 1, 131–145, arXiv:2111.09343 [astro-ph.CO].
[1107] F. Schmidt, “Cosmological Simulations of Normal-Branch Braneworld Gravity,” Phys. Rev. D 80 (2009) 123003,
arXiv:0910.0235 [astro-ph.CO].
[1108] M. Kopp, S. A. Appleby, I. Achitouv, and J. Weller, “Spherical collapse and halo mass function in f (R) theories,”
Phys. Rev. D 88 (2013) no. 8, 084015, arXiv:1306.3233 [astro-ph.CO].
[1109] S. Hagstotz, M. Costanzi, M. Baldi, and J. Weller, “Joint halo-mass function for modified gravity and massive
neutrinos – I. Simulations and cosmological forecasts,” Mon. Not. Roy. Astron. Soc. 486 (2019) no. 3, 3927–3941,
arXiv:1806.07400 [astro-ph.CO].
[1110] S. Gupta, W. A. Hellwing, M. Bilicki, and J. E. García-Farieta, “Universality of the halo mass function in modified
gravity cosmologies,” Phys. Rev. D 105 (2022) no. 4, 043538, arXiv:2112.03699 [astro-ph.CO].
[1111] M. Sereno, A. Veropalumbo, F. Marulli, G. Covone, L. Moscardini, and A. Cimatti, “New constraints on σ8 from a
joint analysis of stacked gravitational lensing and clustering of galaxy clusters,” Mon. Not. Roy. Astron. Soc. 449
(2015) no. 4, 4147–4161, arXiv:1410.5438 [astro-ph.CO].
[1112] F. Marulli, A. Veropalumbo, J. E. García-Farieta, M. Moresco, L. Moscardini, and A. Cimatti, “C3 Cluster Clustering
Cosmology I. New Constraints on the Cosmic Growth Rate at z ∼ 0.3 from Redshift-space Clustering Anisotropies,”
Astrophys. J. 920 (2021) no. 1, 13, arXiv:2010.11206 [astro-ph.CO].
[1113] G. F. Lesci et al., “AMICO galaxy clusters in KiDS-DR3: Constraints on cosmological parameters and on the
normalisation of the mass-richness relation from clustering,” Astron. Astrophys. 665 (2022) A100, arXiv:2203.07398
[astro-ph.CO].
[1114] M. Romanello et al., “AMICO galaxy clusters in KiDS-DR3: Cosmological constraints from the angular power
spectrum and correlation function,” Astron. Astrophys. 682 (2024) A72, arXiv:2310.12224 [astro-ph.CO].
[1115] M. C. A. Cerbolini, B. Sartoris, J.-Q. Xia, A. Biviano, S. Borgani, and M. Viel, “Constraining neutrino properties with
a Euclid-like galaxy cluster survey,” JCAP 06 (2013) 020, arXiv:1303.4550 [astro-ph.CO].
[1116] B. Sartoris et al., “Next Generation Cosmology: Constraints from the Euclid Galaxy Cluster Survey,” Mon. Not. Roy.
Astron. Soc. 459 (2016) no. 2, 1764–1780, arXiv:1505.02165 [astro-ph.CO].
[1117] L. Moscardini, S. Matarrese, and H. J. Mo, “Constraining cosmological parameters with the clustering properties of
galaxy clusters in optical and x-ray bands,” Mon. Not. Roy. Astron. Soc. 327 (2001) 422, arXiv:astro-ph/0009006.
[1118] R. K. Sheth, H. J. Mo, and G. Tormen, “Ellipsoidal collapse and an improved model for the number and spatial
distribution of dark matter haloes,” Mon. Not. Roy. Astron. Soc. 323 (2001) 1, arXiv:astro-ph/9907024.
[1119] E. Branchini, S. Camera, A. Cuoco, N. Fornengo, M. Regis, M. Viel, and J.-Q. Xia, “Cross-correlating the γ-ray sky
with Catalogs of Galaxy Clusters,” Astrophys. J. Suppl. 228 (2017) no. 1, 8, arXiv:1612.05788 [astro-ph.CO].
[1120] K. Paech, N. Hamaus, B. Hoyle, M. Costanzi, T. Giannantonio, S. Hagstotz, G. Sauerwein, and J. Weller,
“Cross-correlation of galaxies and galaxy clusters in the Sloan Digital Sky Survey and the importance of non-Poissonian
shot noise,” Mon. Not. Roy. Astron. Soc. 470 (2017) no. 3, 2566–2577, arXiv:1612.02018 [astro-ph.CO].
[1121] J. L. Tinker, B. E. Robertson, A. V. Kravtsov, A. Klypin, M. S. Warren, G. Yepes, and S. Gottlober, “The Large Scale
Bias of Dark Matter Halos: Numerical Calibration and Model Tests,” Astrophys. J. 724 (2010) 878–886,
arXiv:1001.3162 [astro-ph.CO].
[1122] N. Okabe, Y.-Y. Zhang, A. Finoguenov, M. Takada, G. P. Smith, K. Umetsu, and T. Futamase, “LoCuSS: Calibrating
Mass-Observables Scaling Relations for Cluster Cosmology with Subaru Weak Lensing Observations,” Astrophys. J.
721 (2010) 875–885, arXiv:1007.3816 [astro-ph.CO].
[1123] S. Giodini, L. Lovisari, E. Pointecouteau, S. Ettori, T. H. Reiprich, and H. Hoekstra, “Scaling relations for galaxy
clusters: properties and evolution,” Space Sci. Rev. 177 (2013) 247–282, arXiv:1305.3286 [astro-ph.CO].
[1124] A. Veropalumbo, F. Marulli, L. Moscardini, M. Moresco, and A. Cimatti, “An improved measurement of baryon
acoustic oscillations from the correlation function of galaxy clusters at z ∼ 0.3,” Mon. Not. Roy. Astron. Soc. 442
(2014) no. 4, 3275–3283, arXiv:1311.5895 [astro-ph.CO].
[1125] F. Marulli, A. Veropalumbo, L. Moscardini, A. Cimatti, and K. Dolag, “Redshift-space distortions of galaxies, clusters
and AGN: testing how the accuracy of growth rate measurements depends on scales and sample selections,” Astron.
Astrophys. 599 (2017) A106, arXiv:1505.01170 [astro-ph.CO].
[1126] M. Moresco, A. Veropalumbo, F. Marulli, L. Moscardini, and A. Cimatti, “C3: Cluster Clustering Cosmology. ii. First
Detection of the Baryon Acoustic Oscillations Peak in the Three-point Correlation Function of Galaxy Clusters,”
Astrophys. J. 919 (2021) no. 2, 144, arXiv:2011.04665 [astro-ph.CO].
[1127] C. Alcock and B. Paczynski, “An evolution free test for non-zero cosmological constant,” Nature 281 (1979) 358–359.
[1128] F. Marulli, D. Bianchi, E. Branchini, L. Guzzo, L. Moscardini, and R. E. Angulo, “Cosmology with clustering
anisotropies: disentangling dynamic and geometric distortions in galaxy redshift surveys,” Mon. Not. Roy. Astron. Soc.
426 (2012) 2566, arXiv:1203.1002 [astro-ph.CO].
[1129] BOSS Collaboration, F. Beutler et al., “The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic
Survey: Testing gravity with redshift-space distortions using the power spectrum multipoles,” Mon. Not. Roy. Astron.
Soc. 443 (2014) no. 2, 1065–1089, arXiv:1312.4611 [astro-ph.CO].
[1130] F. Marulli et al., “The XXL Survey: XVI. The clustering of X-ray selected galaxy clusters at z~0.3,” Astron.
Astrophys. 620 (2018) A1, arXiv:1807.04760 [astro-ph.CO].
297

[1131] A. Veropalumbo, F. Marulli, L. Moscardini, M. Moresco, and A. Cimatti, “Measuring the distance–redshift relation
with the baryon acoustic oscillations of galaxy clusters,” Mon. Not. Roy. Astron. Soc. 458 (2016) no. 2, 1909–1920,
arXiv:1510.08852 [astro-ph.CO].
[1132] G. F. Lesci, A. Veropalumbo, M. Sereno, F. Marulli, L. Moscardini, and C. Giocoli, “Mass bias and cosmological
constraints from Planck cluster clustering,” Astron. Astrophys. 674 (2023) A80, arXiv:2302.14074 [astro-ph.CO].
[1133] V. Lindholm, A. Finoguenov, J. Comparat, C. C. Kirkpatrick, E. Rykoff, N. Clerc, C. Collins, S. Damsted, J. I.
Chitham, and N. Padilla, “Clustering of CODEX clusters,” Astron. Astrophys. 646 (2021) A8, arXiv:2012.00090
[astro-ph.CO].
[1134] P. J. E. Peebles, “Statistical Analysis of Catalogs of Extragalactic Objects. I. Theory,” Astrophys. J. 185 (1973)
413–440.
[1135] eBOSS Collaboration, A. Tamone et al., “The Completed SDSS-IV extended Baryon Oscillation Spectroscopic
Survey: Growth rate of structure measurement from anisotropic clustering analysis in configuration space between
redshift 0.6 and 1.1 for the Emission Line Galaxy sample,” Mon. Not. Roy. Astron. Soc. 499 (2020) no. 4, 5527–5546,
arXiv:2007.09009 [astro-ph.CO].
[1136] Z. Li, Y. P. Jing, P. Zhang, and D. Cheng, “Measurement of Redshift-Space Power Spectrum for BOSS galaxies and
the Growth Rate at redshift 0.57,” Astrophys. J. 833 (2016) no. 2, 287, arXiv:1609.03697 [astro-ph.CO].
[1137] S. Alam, S. Ho, M. Vargas-Magaña, and D. P. Schneider, “Testing general relativity with growth rate measurement
from Sloan Digital Sky Survey – III. Baryon Oscillations Spectroscopic Survey galaxies,” Mon. Not. Roy. Astron. Soc.
453 (2015) no. 2, 1754–1767, arXiv:1504.02100 [astro-ph.CO].
[1138] ACT Collaboration, G. S. Farren et al., “The Atacama Cosmology Telescope: Cosmology from Cross-correlations of
unWISE Galaxies and ACT DR6 CMB Lensing,” Astrophys. J. 966 (2024) no. 2, 157, arXiv:2309.05659
[astro-ph.CO].
[1139] G. Piccirilli, G. Fabbian, D. Alonso, K. Storey-Fisher, J. Carron, A. Lewis, and C. García-García, “Growth history and
quasar bias evolution at z < 3 from Quaia,” JCAP 06 (2024) 012, arXiv:2402.05761 [astro-ph.CO].
[1140] D. Alonso, G. Fabbian, K. Storey-Fisher, A.-C. Eilers, C. García-García, D. W. Hogg, and H.-W. Rix, “Constraining
cosmology with the Gaia-unWISE Quasar Catalog and CMB lensing: structure growth,” JCAP 11 (2023) 043,
arXiv:2306.17748 [astro-ph.CO].
[1141] BOSS Collaboration, S. Alam et al., “The clustering of galaxies in the completed SDSS-III Baryon Oscillation
Spectroscopic Survey: cosmological analysis of the DR12 galaxy sample,” Mon. Not. Roy. Astron. Soc. 470 (2017)
no. 3, 2617–2652, arXiv:1607.03155 [astro-ph.CO].
[1142] DESI Collaboration, A. Dey et al., “Overview of the DESI Legacy Imaging Surveys,” Astron. J. 157 (2019) no. 5, 168,
arXiv:1804.08657 [astro-ph.IM].
[1143] C. S. Saraf, P. Bielewicz, and M. Chodorowski, “Effect of redshift bin mismatch on the cross correlation between the
DESI Legacy Imaging Survey and the Planck CMB lensing potential,” Astron. Astrophys. 690 (2024) A338,
arXiv:2406.02857 [astro-ph.CO].
[1144] S. J. Nakoneczny et al., “Cosmology from LOFAR Two-metre Sky Survey Data Release 2: Cross-correlation with the
cosmic microwave background,” Astron. Astrophys. 681 (2024) A105, arXiv:2310.07642 [astro-ph.CO]. [Erratum:
Astron.Astrophys. 686, C2 (2024)].
[1145] DES, SPT Collaboration, C. Chang et al., “Joint analysis of Dark Energy Survey Year 3 data and CMB lensing from
SPT and Planck. II. Cross-correlation measurements and cosmological constraints,” Phys. Rev. D 107 (2023) no. 2,
023530, arXiv:2203.12440 [astro-ph.CO].
[1146] M. White et al., “Cosmological constraints from the tomographic cross-correlation of DESI Luminous Red Galaxies
and Planck CMB lensing,” JCAP 02 (2022) no. 02, 007, arXiv:2111.09898 [astro-ph.CO].
[1147] Z. Sun, J. Yao, F. Dong, X. Yang, L. Zhang, and P. Zhang, “Cross-correlation of Planck cosmic microwave background
lensing with DESI galaxy groups,” Mon. Not. Roy. Astron. Soc. 511 (2022) no. 3, 3548–3560, arXiv:2109.07387
[astro-ph.CO].
[1148] A. Krolewski, S. Ferraro, and M. White, “Cosmological constraints from unWISE and Planck CMB lensing
tomography,” JCAP 12 (2021) no. 12, 028, arXiv:2105.03421 [astro-ph.CO].
[1149] Q. Hang, S. Alam, J. A. Peacock, and Y.-C. Cai, “Galaxy clustering in the DESI Legacy Survey and its imprint on the
CMB,” Mon. Not. Roy. Astron. Soc. 501 (2021) no. 1, 1481–1498, arXiv:2010.00466 [astro-ph.CO].
[1150] J. A. Peacock and M. Bilicki, “Wide-area tomography of CMB lensing and the growth of cosmological density
fluctuations,” Mon. Not. Roy. Astron. Soc. 481 (2018) no. 1, 1133–1148, arXiv:1805.11525 [astro-ph.CO].
[1151] DES Collaboration, T. Giannantonio et al., “CMB lensing tomography with the DES Science Verification galaxies,”
Mon. Not. Roy. Astron. Soc. 456 (2016) no. 3, 3213–3244, arXiv:1507.05551 [astro-ph.CO].
[1152] N. Sailer et al., “Cosmological constraints from the cross-correlation of DESI Luminous Red Galaxies with CMB
lensing from Planck PR4 and ACT DR6,” arXiv:2407.04607 [astro-ph.CO].
[1153] C. S. Saraf and P. Bielewicz, “Tomographic cross correlations between galaxy surveys and the CMB gravitational
lensing potential - Effect of the redshift bin mismatch,” Astron. Astrophys. 687 (2024) A150, arXiv:2311.15261
[astro-ph.CO].
[1154] T. Simon, P. Zhang, and V. Poulin, “Cosmological inference from the EFTofLSS: the eBOSS QSO full-shape analysis,”
JCAP 07 (2023) 041, arXiv:2210.14931 [astro-ph.CO].
[1155] Y. Kobayashi, T. Nishimichi, M. Takada, and H. Miyatake, “Full-shape cosmology analysis of the SDSS-III BOSS
galaxy power spectrum using an emulator-based halo model: A 5% determination of σ8,” Phys. Rev. D 105 (2022)
no. 8, 083517, arXiv:2110.06969 [astro-ph.CO].
298

[1156] S.-F. Chen, Z. Vlah, and M. White, “A new analysis of galaxy 2-point functions in the BOSS survey, including
full-shape information and post-reconstruction BAO,” JCAP 02 (2022) no. 02, 008, arXiv:2110.05530 [astro-ph.CO].
[1157] M. M. Ivanov, O. H. E. Philcox, G. Cabass, T. Nishimichi, M. Simonović, and M. Zaldarriaga, “Cosmology with the
galaxy bispectrum multipoles: Optimal estimation and application to BOSS data,” Phys. Rev. D 107 (2023) no. 8,
083515, arXiv:2302.04414 [astro-ph.CO].
[1158] O. H. E. Philcox and M. M. Ivanov, “BOSS DR12 full-shape cosmology: ΛCDM constraints from the large-scale galaxy
power spectrum and bispectrum monopole,” Phys. Rev. D 105 (2022) no. 4, 043517, arXiv:2112.04515
[astro-ph.CO].
[1159] S. Joudaki et al., “KiDS-450 + 2dFLenS: Cosmological parameter constraints from weak gravitational lensing
tomography and overlapping redshift-space galaxy clustering,” Mon. Not. Roy. Astron. Soc. 474 (2018) no. 4,
4894–4924, arXiv:1707.06627 [astro-ph.CO].
[1160] E. van Uitert et al., “KiDS+GAMA: cosmology constraints from a joint analysis of cosmic shear, galaxy–galaxy lensing,
and angular clustering,” Mon. Not. Roy. Astron. Soc. 476 (2018) no. 4, 4662–4689, arXiv:1706.05004 [astro-ph.CO].
[1161] DES Collaboration, T. M. C. Abbott et al., “Dark Energy Survey year 1 results: Cosmological constraints from galaxy
clustering and weak lensing,” Phys. Rev. D 98 (2018) no. 4, 043526, arXiv:1708.01530 [astro-ph.CO].
[1162] BOSS Collaboration, A. J. Ross et al., “The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic
Survey: Analysis of potential systematics,” Mon. Not. Roy. Astron. Soc. 424 (2012) 564, arXiv:1203.6499
[astro-ph.CO].
[1163] C. Hahn, R. Scoccimarro, M. R. Blanton, J. L. Tinker, and S. A. Rodríguez-Torres, “The effect of fibre collisions on
the galaxy power spectrum multipoles,” Mon. Not. Roy. Astron. Soc. 467 (2017) no. 2, 1940–1956, arXiv:1609.01714
[astro-ph.CO].
[1164] A. de Mattia and V. Ruhlmann-Kleider, “Integral constraints in spectroscopic surveys,” JCAP 08 (2019) 036,
arXiv:1904.08851 [astro-ph.CO].
[1165] C. S. Saraf, P. Bielewicz, and M. Chodorowski, “Cross-correlation between Planck CMB lensing potential and galaxy
catalogues from HELP,” Mon. Not. Roy. Astron. Soc. 515 (2022) no. 2, 1993–2007, arXiv:2106.02551 [astro-ph.CO].
[1166] A. R. Pullen, S. Alam, S. He, and S. Ho, “Constraining Gravity at the Largest Scales through CMB Lensing and
Galaxy Velocities,” Mon. Not. Roy. Astron. Soc. 460 (2016) no. 4, 4098–4108, arXiv:1511.04457 [astro-ph.CO].
[1167] S. Joudaki et al., “CFHTLenS revisited: assessing concordance with Planck including astrophysical systematics,” Mon.
Not. Roy. Astron. Soc. 465 (2017) no. 2, 2033–2052, arXiv:1601.05786 [astro-ph.CO].
[1168] M. M. Ivanov, M. Simonović, and M. Zaldarriaga, “Cosmological Parameters from the BOSS Galaxy Power Spectrum,”
JCAP 05 (2020) 042, arXiv:1909.05277 [astro-ph.CO].
[1169] R. Gsponer, R. Zhao, J. Donald-McCann, D. Bacon, K. Koyama, R. Crittenden, T. Simon, and E.-M. Mueller,
“Cosmological constraints on early dark energy from the full shape analysis of eBOSS DR16,” Mon. Not. Roy. Astron.
Soc. 530 (2024) no. 3, 3075–3099, arXiv:2312.01977 [astro-ph.CO].
[1170] S.-F. Chen, Z. Vlah, E. Castorina, and M. White, “Redshift-Space Distortions in Lagrangian Perturbation Theory,”
JCAP 03 (2021) 100, arXiv:2012.04636 [astro-ph.CO].
[1171] DESI Collaboration, S. Ramirez-Solano et al., “Full Modeling and parameter compression methods in configuration
space for DESI 2024 and beyond,” JCAP 01 (2025) 129, arXiv:2404.07268 [astro-ph.CO].
[1172] M. Maus et al., “A comparison of effective field theory models of redshift space galaxy power spectra for DESI 2024
and future surveys,” JCAP 01 (2025) 134, arXiv:2404.07272 [astro-ph.CO].
[1173] D. Baumann, A. Nicolis, L. Senatore, and M. Zaldarriaga, “Cosmological Non-Linearities as an Effective Fluid,” JCAP
07 (2012) 051, arXiv:1004.2488 [astro-ph.CO].
[1174] J. J. M. Carrasco, M. P. Hertzberg, and L. Senatore, “The Effective Field Theory of Cosmological Large Scale
Structures,” JHEP 09 (2012) 082, arXiv:1206.2926 [astro-ph.CO].
[1175] R. A. Porto, L. Senatore, and M. Zaldarriaga, “The Lagrangian-space Effective Field Theory of Large Scale
Structures,” JCAP 05 (2014) 022, arXiv:1311.2168 [astro-ph.CO].
[1176] M. Lewandowski, L. Senatore, F. Prada, C. Zhao, and C.-H. Chuang, “EFT of large scale structures in redshift space,”
Phys. Rev. D 97 (2018) no. 6, 063526, arXiv:1512.06831 [astro-ph.CO].
[1177] A. Lewis and A. Challinor, “Weak gravitational lensing of the CMB,” Phys. Rept. 429 (2006) 1–65,
arXiv:astro-ph/0601594.
[1178] J. Carron, M. Mirmelstein, and A. Lewis, “CMB lensing from Planck PR4 maps,” JCAP 09 (2022) 039,
arXiv:2206.07773 [astro-ph.CO].
[1179] ACT Collaboration, F. J. Qu et al., “The Atacama Cosmology Telescope: A Measurement of the DR6 CMB Lensing
Power Spectrum and Its Implications for Structure Growth,” Astrophys. J. 962 (2024) no. 2, 112, arXiv:2304.05202
[astro-ph.CO].
[1180] SPT Collaboration, Z. Pan et al., “Measurement of gravitational lensing of the cosmic microwave background using
SPT-3G 2018 data,” Phys. Rev. D 108 (2023) no. 12, 122005, arXiv:2308.11608 [astro-ph.CO].
[1181] E. Di Valentino, A. Melchiorri, and J. Silk, “Planck evidence for a closed Universe and a possible crisis for cosmology,”
Nature Astron. 4 (2019) no. 2, 196–203, arXiv:1911.02087 [astro-ph.CO].
[1182] W. Handley, “Curvature tension: evidence for a closed universe,” Phys. Rev. D 103 (2021) no. 4, L041301,
arXiv:1908.09139 [astro-ph.CO].
[1183] G. Efstathiou and S. Gratton, “The evidence for a spatially flat Universe,” Mon. Not. Roy. Astron. Soc. 496 (2020)
no. 1, L91–L95, arXiv:2002.06892 [astro-ph.CO].
[1184] A. Glanville, C. Howlett, and T. M. Davis, “Full-shape galaxy power spectra and the curvature tension,” Mon. Not.
299

Roy. Astron. Soc. 517 (2022) no. 2, 3087–3100, arXiv:2205.05892 [astro-ph.CO].


[1185] S. Vagnozzi, A. Loeb, and M. Moresco, “Eppur è piatto? The Cosmic Chronometers Take on Spatial Curvature and
Cosmic Concordance,” Astrophys. J. 908 (2021) no. 1, 84, arXiv:2011.11645 [astro-ph.CO].
[1186] S. Vagnozzi, E. Di Valentino, S. Gariazzo, A. Melchiorri, O. Mena, and J. Silk, “The galaxy power spectrum take on
spatial curvature and cosmic concordance,” Phys. Dark Univ. 33 (2021) 100851, arXiv:2010.02230 [astro-ph.CO].
[1187] C.-G. Park and B. Ratra, “Using the tilted flat-ΛCDM and the untilted non-flat ΛCDM inflation models to measure
cosmological parameters from a compilation of observational data,” Astrophys. J. 882 (2019) 158, arXiv:1801.00213
[astro-ph.CO].
[1188] C.-G. Park and B. Ratra, “Measuring the Hubble constant and spatial curvature from supernova apparent magnitude,
baryon acoustic oscillation, and Hubble parameter data,” Astrophys. Space Sci. 364 (2019) no. 8, 134,
arXiv:1809.03598 [astro-ph.CO].
[1189] WMAP Collaboration, C. L. Bennett et al., “Nine-Year Wilkinson Microwave Anisotropy Probe (WMAP)
Observations: Final Maps and Results,” Astrophys. J. Suppl. 208 (2013) 20, arXiv:1212.5225 [astro-ph.CO].
[1190] Planck Collaboration, R. Adam et al., “Planck 2015 results. I. Overview of products and scientific results,” Astron.
Astrophys. 594 (2016) A1, arXiv:1502.01582 [astro-ph.CO].
[1191] WMAP Collaboration, G. Hinshaw et al., “First year Wilkinson Microwave Anisotropy Probe (WMAP) observations:
The Angular power spectrum,” Astrophys. J. Suppl. 148 (2003) 135, arXiv:astro-ph/0302217.
[1192] C. R. Contaldi, M. Peloso, L. Kofman, and A. D. Linde, “Suppressing the lower multipoles in the CMB anisotropies,”
JCAP 07 (2003) 002, arXiv:astro-ph/0303636.
[1193] J. M. Cline, P. Crotty, and J. Lesgourgues, “Does the small CMB quadrupole moment suggest new physics?,” JCAP 09
(2003) 010, arXiv:astro-ph/0304558.
[1194] D. J. Schwarz, G. D. Starkman, D. Huterer, and C. J. Copi, “Is the low-l microwave background cosmic?,” Phys. Rev.
Lett. 93 (2004) 221301, arXiv:astro-ph/0403353.
[1195] C. J. Copi, D. Huterer, D. J. Schwarz, and G. D. Starkman, “On the large-angle anomalies of the microwave sky,”
Mon. Not. Roy. Astron. Soc. 367 (2006) 79–102, arXiv:astro-ph/0508047.
[1196] A. de Oliveira-Costa and M. Tegmark, “Cmb multipole measurements in the presence of foregrounds,” Phys. Rev. D 74
(2006) 023005, arXiv:astro-ph/0603369.
[1197] A. Slosar and U. Seljak, “Assessing the effects of foregrounds and sky removal in WMAP,” Phys. Rev. D 70 (2004)
083002, arXiv:astro-ph/0404567.
[1198] A. Hajian, “Analysis of the apparent lack of power in the cosmic microwave background anisotropy at large angular
scales,” arXiv:astro-ph/0702723.
[1199] Planck Collaboration, P. A. R. Ade et al., “Planck 2013 results. XXIII. Isotropy and statistics of the CMB,” Astron.
Astrophys. 571 (2014) A23, arXiv:1303.5083 [astro-ph.CO].
[1200] H. K. Eriksen, F. K. Hansen, A. J. Banday, K. M. Gorski, and P. B. Lilje, “Asymmetries in the Cosmic Microwave
Background anisotropy field,” Astrophys. J. 605 (2004) 14–20, arXiv:astro-ph/0307507. [Erratum: Astrophys.J. 609,
1198 (2004)].
[1201] Planck Collaboration, P. A. R. Ade et al., “Planck 2015 results. XVI. Isotropy and statistics of the CMB,” Astron.
Astrophys. 594 (2016) A16, arXiv:1506.07135 [astro-ph.CO].
[1202] A. L. Erickcek, M. Kamionkowski, and S. M. Carroll, “A Hemispherical Power Asymmetry from Inflation,” Phys. Rev.
D 78 (2008) 123520, arXiv:0806.0377 [astro-ph].
[1203] F. Schmidt and L. Hui, “Cosmic Microwave Background Power Asymmetry from Non-Gaussian Modulation,” Phys.
Rev. Lett. 110 (2013) 011301, arXiv:1210.2965 [astro-ph.CO]. [Erratum: Phys.Rev.Lett. 110, 059902 (2013)].
[1204] F. K. Hansen, T. Trombetti, N. Bartolo, U. Natale, M. Liguori, A. J. Banday, and K. M. Gorski, “Isotropic
non-Gaussian gNL-like toy models that reproduce cosmic microwave background anomalies,” Astron. Astrophys. 626
(2019) A13, arXiv:1806.08531 [astro-ph.CO].
[1205] C. Gordon and W. Hu, “A Low CMB quadrupole from dark energy isocurvature perturbations,” Phys. Rev. D 70
(2004) 083003, arXiv:astro-ph/0406496.
[1206] K. Land and J. Magueijo, “The Axis of evil,” Phys. Rev. Lett. 95 (2005) 071301, arXiv:astro-ph/0502237.
[1207] J. Kim and P. Naselsky, “Lack of angular correlation and odd-parity preference in CMB data,” Astrophys. J. 739
(2011) 79, arXiv:1011.0377 [astro-ph.CO].
[1208] A. Lue, L.-M. Wang, and M. Kamionkowski, “Cosmological signature of new parity violating interactions,” Phys. Rev.
Lett. 83 (1999) 1506–1509, arXiv:astro-ph/9812088.
[1209] S. H. S. Alexander, “Is cosmic parity violation responsible for the anomalies in the WMAP data?,” Phys. Lett. B 660
(2008) 444–448, arXiv:hep-th/0601034.
[1210] P. Vielva, E. Martinez-Gonzalez, R. B. Barreiro, J. L. Sanz, and L. Cayon, “Detection of non-Gaussianity in the
WMAP 1 - year data using spherical wavelets,” Astrophys. J. 609 (2004) 22–34, arXiv:astro-ph/0310273.
[1211] M. Cruz, L. Cayon, E. Martinez-Gonzalez, P. Vielva, and J. Jin, “The non-gaussian cold spot in the 3-year wmap
data,” Astrophys. J. 655 (2007) 11–20, arXiv:astro-ph/0603859.
[1212] S. Nadathur, M. Lavinto, S. Hotchkiss, and S. Räsänen, “Can a supervoid explain the Cold Spot?,” Phys. Rev. D 90
(2014) no. 10, 103510, arXiv:1408.4720 [astro-ph.CO].
[1213] K. T. Inoue and J. Silk, “Local voids as the origin of large-angle cosmic microwave background anomalies I,”
Astrophys. J. 648 (2006) 23–30, arXiv:astro-ph/0602478.
[1214] R. Holman, L. Mersini-Houghton, and T. Takahashi, “Cosmological avatars of the landscape I: Bracketing the SUSY
breaking scale,” Phys. Rev. D 77 (2008) 063510, arXiv:hep-th/0611223.
300

[1215] M. Cruz, M. Tucci, E. Martinez-Gonzalez, and P. Vielva, “The non-gaussian cold spot in wmap: significance,
morphology and foreground contribution,” Mon. Not. Roy. Astron. Soc. 369 (2006) 57–67, arXiv:astro-ph/0601427.
[1216] L. Rudnick, S. Brown, and L. R. Williams, “Extragalactic Radio Sources and the WMAP Cold Spot,” Astrophys. J.
671 (2007) 40–44, arXiv:0704.0908 [astro-ph].
[1217] D. G. Lambas, F. K. Hansen, F. Toscano, H. E. Luparello, and E. F. Boero, “The CMB Cold Spot as predicted by
foregrounds around nearby galaxies,” Astron. Astrophys. 681 (2024) A2, arXiv:2310.13755 [astro-ph.CO].
[1218] I. Szapudi et al., “Detection of a Supervoid Aligned with the Cold Spot of the Cosmic Microwave Background,” Mon.
Not. Roy. Astron. Soc. 450 (2015) no. 1, 288–294, arXiv:1405.1566 [astro-ph.CO].
[1219] R. Mackenzie, T. Shanks, M. N. Bremer, Y.-C. Cai, M. L. P. Gunawardhana, A. Kovács, P. Norberg, and I. Szapudi,
“Evidence against a supervoid causing the CMB Cold Spot,” Mon. Not. Roy. Astron. Soc. 470 (2017) no. 2,
2328–2338, arXiv:1704.03814 [astro-ph.CO].
[1220] H. M. Courtois, R. B. Tully, Y. Hoffman, D. Pomarede, R. Graziani, and A. Dupuy, “Cosmicflows-3: Cold Spot
Repeller?,” Astrophys. J. Lett. 847 (2017) no. 1, L6, arXiv:1708.07547 [astro-ph.CO].
[1221] DES Collaboration, A. Kovács et al., “The DES view of the Eridanus supervoid and the CMB cold spot,” Mon. Not.
Roy. Astron. Soc. 510 (2022) no. 1, 216–229, arXiv:2112.07699 [astro-ph.CO].
[1222] H. E. Luparello, E. F. Boero, M. Lares, A. G. Sánchez, and D. G. Lambas, “The cosmic shallows – I. Interaction of
CMB photons in extended galaxy haloes,” Mon. Not. Roy. Astron. Soc. 518 (2022) no. 4, 5643–5652,
arXiv:2206.14217 [astro-ph.CO].
[1223] F. K. Hansen, E. F. Boero, H. E. Luparello, and D. G. Lambas, “A possible common explanation for several cosmic
microwave background (CMB) anomalies: A strong impact of nearby galaxies on observed large-scale CMB
fluctuations,” Astron. Astrophys. 675 (2023) L7, arXiv:2305.00268 [astro-ph.CO].
[1224] J. P. Huchra et al., “The 2MASS Redshift Survey - Description and Data Release,” Astrophys. J. Suppl. 199 (2012) 26,
arXiv:1108.0669 [astro-ph.CO].
[1225] M. Cruz, E. Martínez-González, C. Gimeno-Amo, B. J. Kavanagh, and M. Tucci, “Unexplained correlation between the
Cosmic Microwave Background temperature and the local matter density distribution,” arXiv:2407.17599
[astro-ph.CO].
[1226] F. K. Hansen, D. G. Lambas, H. E. Luparello, F. Toscano, and L. A. Pereyra, “A 5.7σ detection confirming the
existence of a possibly dark matter related CMB foreground in nearby cosmic filaments,” arXiv:2411.15307
[astro-ph.CO].
[1227] F. Toscano, F. K. Hansen, D. G. Lambas, H. Luparello, P. Fosalba, and E. Gaztañaga, “Are CMB derived cosmological
parameters affected by foregrounds associated to nearby galaxies?,” arXiv:2410.24026 [astro-ph.CO].
[1228] C. J. Copi, D. Huterer, D. J. Schwarz, and G. D. Starkman, “No large-angle correlations on the non-Galactic
microwave sky,” Mon. Not. Roy. Astron. Soc. 399 (2009) 295–303, arXiv:0808.3767 [astro-ph].
[1229] D. Sarkar, D. Huterer, C. J. Copi, G. D. Starkman, and D. J. Schwarz, “Missing Power vs low-l Alignments in the
Cosmic Microwave Background: No Correlation in the Standard Cosmological Model,” Astropart. Phys. 34 (2011)
591–594, arXiv:1004.3784 [astro-ph.CO].
[1230] J. P. Luminet, J. Weeks, A. Riazuelo, R. Lehoucq, and J. P. Uzan, “Dodecahedral space topology as an explanation for
weak wide - angle temperature correlations in the cosmic microwave background,” Nature 425 (2003) 593,
arXiv:astro-ph/0310253.
[1231] R. Aurich, H. S. Janzer, S. Lustig, and F. Steiner, “Do we Live in a Small Universe?,” Class. Quant. Grav. 25 (2008)
125006, arXiv:0708.1420 [astro-ph].
[1232] C. Copi, D. Huterer, D. Schwarz, and G. Starkman, “The Uncorrelated Universe: Statistical Anisotropy and the
Vanishing Angular Correlation Function in WMAP Years 1-3,” Phys. Rev. D 75 (2007) 023507,
arXiv:astro-ph/0605135.
[1233] C. J. Hogan, “Gravitational Waves from Light Cosmic Strings: Backgrounds and Bursts with Large Loops,” Phys. Rev.
D 74 (2006) 043526, arXiv:astro-ph/0605567.
[1234] L. P. Chimento and M. I. Forte, “Anisotropic k-essence cosmologies,” Phys. Rev. D 73 (2006) 063502,
arXiv:astro-ph/0510726.
[1235] S. Tsujikawa, M. Sami, and R. Maartens, “Observational constraints on braneworld inflation: The Effect of a
Gauss-Bonnet term,” Phys. Rev. D 70 (2004) 063525, arXiv:astro-ph/0406078.
[1236] M.-a. Watanabe, S. Kanno, and J. Soda, “Inflationary Universe with Anisotropic Hair,” Phys. Rev. Lett. 102 (2009)
191302, arXiv:0902.2833 [hep-th].
[1237] J. Delabrouille et al., “The pre-launch Planck Sky Model: a model of sky emission at submillimetre to centimetre
wavelengths,” Astron. Astrophys. 553 (2013) A96, arXiv:1207.3675 [astro-ph.CO].
[1238] L. Mersini-Houghton, “Cosmological Implications of the String Theory Landscape,” AIP Conf. Proc. 878 (2006) no. 1,
315–322, arXiv:hep-ph/0609157.
[1239] L. Mersini-Houghton, “Can we predict Lambda for the non-SUSY sector of the landscape,” Class. Quant. Grav. 22
(2005) 3481–3490, arXiv:hep-th/0504026.
[1240] A. Kobakhidze and L. Mersini-Houghton, “Birth of the universe from the landscape of string theory,” Eur. Phys. J. C
49 (2007) 869–873, arXiv:hep-th/0410213.
[1241] R. Holman and L. Mersini-Houghton, “Why the universe started from a low entropy state,” Phys. Rev. D 74 (2006)
123510, arXiv:hep-th/0511102.
[1242] R. Holman, L. Mersini-Houghton, and T. Takahashi, “Cosmological Avatars of the Landscape. II. CMB and LSS
Signatures,” Phys. Rev. D 77 (2008) 063511, arXiv:hep-th/0612142.
301

[1243] E. Di Valentino and L. Mersini-Houghton, “Testing Predictions of the Quantum Landscape Multiverse 1: The
Starobinsky Inflationary Potential,” JCAP 03 (2017) 002, arXiv:1612.09588 [astro-ph.CO].
[1244] E. Di Valentino and L. Mersini-Houghton, “Testing Predictions of the Quantum Landscape Multiverse 2: The
Exponential Inflationary Potential,” JCAP 03 (2017) 020, arXiv:1612.08334 [astro-ph.CO].
[1245] E. Di Valentino and L. Mersini-Houghton, “Testing Predictions of the Quantum Landscape Multiverse 3: The Hilltop
Inflationary Potential,” Symmetry 11 (2019) no. 4, 520, arXiv:1807.10833 [astro-ph.CO].
[1246] L. Mersini-Houghton, “Predictions of the Quantum Landscape Multiverse,” Class. Quant. Grav. 34 (2017) no. 4,
047001, arXiv:1612.07129 [hep-th].
[1247] DESI Collaboration, K. Lodha et al., “DESI 2024: Constraints on physics-focused aspects of dark energy using DESI
DR1 BAO data,” Phys. Rev. D 111 (2025) no. 2, 023532, arXiv:2405.13588 [astro-ph.CO].
[1248] DESI Collaboration, M. A. Karim et al., “DESI DR2 Results I: Baryon Acoustic Oscillations from the Lyman Alpha
Forest,” arXiv:2503.14739 [astro-ph.CO].
[1249] DESI Collaboration, U. Andrade et al., “Validation of the DESI DR2 Measurements of Baryon Acoustic Oscillations
from Galaxies and Quasars,” arXiv:2503.14742 [astro-ph.CO].
[1250] A. Brodzeller et al., “Construction of the Damped Lyα Absorber Catalog for DESI DR2 Lyα BAO,”
arXiv:2503.14740 [astro-ph.CO].
[1251] DESI Collaboration, K. Lodha et al., “Extended Dark Energy analysis using DESI DR2 BAO measurements,”
arXiv:2503.14743 [astro-ph.CO].
[1252] D. Rubin et al., “Union Through UNITY: Cosmology with 2,000 SNe Using a Unified Bayesian Framework,”
arXiv:2311.12098 [astro-ph.CO].
[1253] DES Collaboration, T. M. C. Abbott et al., “The Dark Energy Survey: Cosmology Results with ∼1500 New
High-redshift Type Ia Supernovae Using the Full 5 yr Data Set,” Astrophys. J. Lett. 973 (2024) no. 1, L14,
arXiv:2401.02929 [astro-ph.CO].
[1254] DES Collaboration, B. O. Sánchez et al., “The Dark Energy Survey Supernova Program: Light Curves and 5 Yr Data
Release,” Astrophys. J. 975 (2024) no. 1, 5, arXiv:2406.05046 [astro-ph.CO].
[1255] DES Collaboration, M. Vincenzi et al., “The Dark Energy Survey Supernova Program: Cosmological Analysis and
Systematic Uncertainties,” Astrophys. J. 975 (2024) no. 1, 86, arXiv:2401.02945 [astro-ph.CO].
[1256] DESI Collaboration, K. Lodha et al., “Extended Dark Energy analysis using DESI DR2 BAO measurements,”
arXiv:2503.14743 [astro-ph.CO].
[1257] S. M. Carroll, M. Hoffman, and M. Trodden, “Can the dark energy equation-of-state parameter w be less than −1?,”
Phys. Rev. D 68 (2003) 023509, arXiv:astro-ph/0301273.
[1258] W. Hu, “Crossing the phantom divide: Dark energy internal degrees of freedom,” Phys. Rev. D 71 (2005) 047301,
arXiv:astro-ph/0410680.
[1259] P. Creminelli, G. D’Amico, J. Norena, and F. Vernizzi, “The Effective Theory of Quintessence: the w<-1 Side
Unveiled,” JCAP 02 (2009) 018, arXiv:0811.0827 [astro-ph].
[1260] A. Chudaykin and M. Kunz, “Modified gravity interpretation of the evolving dark energy in light of DESI data,” Phys.
Rev. D 110 (2024) no. 12, 123524, arXiv:2407.02558 [astro-ph.CO].
[1261] G. Ye, M. Martinelli, B. Hu, and A. Silvestri, “Non-minimally coupled gravity as a physically viable fit to DESI 2024
BAO,” arXiv:2407.15832 [astro-ph.CO].
[1262] M. Ishak et al., “Modified Gravity Constraints from the Full Shape Modeling of Clustering Measurements from DESI
2024,” arXiv:2411.12026 [astro-ph.CO].
[1263] A. Chudaykin, M. Kunz, and J. Carron, “Modified gravity constraints with Planck ISW-lensing bispectrum,”
arXiv:2503.09893 [astro-ph.CO].
[1264] W. J. Wolf, P. G. Ferreira, and C. García-García, “Matching current observational constraints with nonminimally
coupled dark energy,” Phys. Rev. D 111 (2025) no. 4, L041303, arXiv:2409.17019 [astro-ph.CO].
[1265] A. Hernández-Almada, M. L. Mendoza-Martínez, M. A. García-Aspeitia, and V. Motta, “Phenomenological emergent
dark energy in the light of DESI Data Release 1,” Phys. Dark Univ. 46 (2024) 101668, arXiv:2407.09430
[astro-ph.CO].
[1266] S. Pourojaghi, M. Malekjani, and Z. Davari, “Cosmological constraints on dark energy parametrizations after DESI
2024: Persistent deviation from standard ΛCDM cosmology,” arXiv:2407.09767 [astro-ph.CO].
[1267] O. F. Ramadan, J. Sakstein, and D. Rubin, “DESI constraints on exponential quintessence,” Phys. Rev. D 110 (2024)
no. 4, L041303, arXiv:2405.18747 [astro-ph.CO].
[1268] Y. Carloni, O. Luongo, and M. Muccino, “Does dark energy really revive using DESI 2024 data?,” Phys. Rev. D 111
(2025) no. 2, 023512, arXiv:2404.12068 [astro-ph.CO].
[1269] K. V. Berghaus, J. A. Kable, and V. Miranda, “Quantifying scalar field dynamics with DESI 2024 Y1 BAO
measurements,” Phys. Rev. D 110 (2024) no. 10, 103524, arXiv:2404.14341 [astro-ph.CO].
[1270] F. J. Qu, K. M. Surrao, B. Bolliet, J. C. Hill, B. D. Sherwin, and H. T. Jense, “Accelerated inference on accelerated
cosmic expansion: New constraints on axion-like early dark energy with DESI BAO and ACT DR6 CMB lensing,”
arXiv:2404.16805 [astro-ph.CO].
[1271] A. Notari, M. Redi, and A. Tesi, “Consistent theories for the DESI dark energy fit,” JCAP 11 (2024) 025,
arXiv:2406.08459 [astro-ph.CO].
[1272] P. Adolf, M. Hirsch, S. Krieg, H. Päs, and M. Tabet, “Fitting the DESI BAO data with dark energy driven by the
Cohen-Kaplan-Nelson bound,” JCAP 08 (2024) 048, arXiv:2406.09964 [astro-ph.CO].
[1273] J.-Q. Jiang, D. Pedrotti, S. S. da Costa, and S. Vagnozzi, “Nonparametric late-time expansion history reconstruction
302

and implications for the Hubble tension in light of recent DESI and type Ia supernovae data,” Phys. Rev. D 110 (2024)
no. 12, 123519, arXiv:2408.02365 [astro-ph.CO].
[1274] B. R. Dinda and R. Maartens, “Model-agnostic assessment of dark energy after DESI DR1 BAO,” JCAP 01 (2025)
120, arXiv:2407.17252 [astro-ph.CO].
[1275] W. J. Wolf, C. García-García, and P. G. Ferreira, “Robustness of Dark Energy Phenomenology Across Different
Parameterizations,” arXiv:2502.04929 [astro-ph.CO].
[1276] A. Sousa-Neto, C. Bengaly, J. E. González, and J. Alcaniz, “No evidence for dynamical dark energy from DESI and SN
data: a symbolic regression analysis,” arXiv:2502.10506 [astro-ph.CO].
[1277] R. de Putter and E. V. Linder, “Calibrating Dark Energy,” JCAP 10 (2008) 042, arXiv:0808.0189 [astro-ph].
[1278] W. Giarè, M. Najafi, S. Pan, E. Di Valentino, and J. T. Firouzjaee, “Robust preference for Dynamical Dark Energy in
DESI BAO and SN measurements,” JCAP 10 (2024) 035, arXiv:2407.16689 [astro-ph.CO].
[1279] Z. Wang, S. Lin, Z. Ding, and B. Hu, “The role of LRG1 and LRG2’s monopole in inferring the DESI 2024 BAO
cosmology,” Mon. Not. Roy. Astron. Soc. 534 (2024) no. 4, 3869–3875, arXiv:2405.02168 [astro-ph.CO].
[1280] E. O. Colgáin, M. G. Dainotti, S. Capozziello, S. Pourojaghi, M. M. Sheikh-Jabbari, and D. Stojkovic, “Does DESI
2024 Confirm ΛCDM?,” arXiv:2404.08633 [astro-ph.CO].
[1281] D. Naredo-Tuero, M. Escudero, E. Fernández-Martínez, X. Marcano, and V. Poulin, “Critical look at the cosmological
neutrino mass bound,” Phys. Rev. D 110 (2024) no. 12, 123537, arXiv:2407.13831 [astro-ph.CO].
[1282] D. Sapone and S. Nesseris, “Outliers in DESI BAO: robustness and cosmological implications,” arXiv:2412.01740
[astro-ph.CO].
[1283] W. Giarè, T. Mahassen, E. Di Valentino, and S. Pan, “An overview of what current data can (and cannot yet) say
about evolving dark energy,” arXiv:2502.10264 [astro-ph.CO].
[1284] DES Collaboration, T. M. C. Abbott et al., “Dark Energy Survey: implications for cosmological expansion models
from the final DES Baryon Acoustic Oscillation and Supernova data,” arXiv:2503.06712 [astro-ph.CO].
[1285] E. O. Colgáin, M. M. Sheikh-Jabbari, R. Solomon, M. G. Dainotti, and D. Stojkovic, “Putting flat ΛCDM in the
(Redshift) bin,” Phys. Dark Univ. 44 (2024) 101464, arXiv:2206.11447 [astro-ph.CO].
[1286] M. Malekjani, R. M. Conville, E. O. Colgáin, S. Pourojaghi, and M. M. Sheikh-Jabbari, “On redshift evolution and
negative dark energy density in Pantheon + Supernovae,” Eur. Phys. J. C 84 (2024) no. 3, 317, arXiv:2301.12725
[astro-ph.CO].
[1287] E. O. Colgáin, S. Pourojaghi, and M. M. Sheikh-Jabbari, “Implications of DES 5YR SNe Dataset for ΛCDM,” Eur.
Phys. J. C 85 (2025) no. 3, 286, arXiv:2406.06389 [astro-ph.CO].
[1288] E. O. Colgáin and M. M. Sheikh-Jabbari, “DESI and SNe: Dynamical Dark Energy, Ωm Tension or Systematics?,”
arXiv:2412.12905 [astro-ph.CO].
[1289] A. Notari, M. Redi, and A. Tesi, “BAO vs. SN evidence for evolving dark energy,” arXiv:2411.11685 [astro-ph.CO].
[1290] G. Efstathiou, “Evolving Dark Energy or Supernovae Systematics?,” arXiv:2408.07175 [astro-ph.CO].
[1291] DES Collaboration, M. Vincenzi et al., “Comparing the DES-SN5YR and Pantheon+ SN cosmology analyses:
Investigation based on ”Evolving Dark Energy or Supernovae systematics?”,” arXiv:2501.06664 [astro-ph.CO].
[1292] W. Giarè, “Dynamical Dark Energy Beyond Planck? Constraints from multiple CMB probes, DESI BAO and Type-Ia
Supernovae,” arXiv:2409.17074 [astro-ph.CO].
[1293] L. A. Escamilla, W. Giarè, E. Di Valentino, R. C. Nunes, and S. Vagnozzi, “The state of the dark energy equation of
state circa 2023,” JCAP 05 (2024) 091, arXiv:2307.14802 [astro-ph.CO].
[1294] W. Giarè, E. Di Valentino, and A. Melchiorri, “Measuring the reionization optical depth without large-scale CMB
polarization,” Phys. Rev. D 109 (2024) no. 10, 103519, arXiv:2312.06482 [astro-ph.CO].
[1295] I. Ben-Dayan, U. Kumar, M. Shimon, and A. Verma, “Impact of low ell’s on large scale structure anomalies,” JCAP 02
(2025) 069, arXiv:2409.15457 [astro-ph.CO].
[1296] Z.-Y. Peng and Y.-S. Piao, “Dark energy and lensing anomaly in Planck CMB data,” arXiv:2502.04641
[astro-ph.CO].
[1297] C.-G. Park and B. Ratra, “Is excess smoothing of Planck CMB ansiotropy data partially responsible for evidence for
dark energy dynamics in other w(z)CDM parametrizations?,” arXiv:2501.03480 [astro-ph.CO].
[1298] C.-G. Park, J. de Cruz Perez, and B. Ratra, “Is the w0 wa CDM cosmological parameterization evidence for dark energy
dynamics partially caused by the excess smoothing of Planck CMB anisotropy data?,” arXiv:2410.13627
[astro-ph.CO].
[1299] E. Di Valentino, S. Gariazzo, and O. Mena, “Most constraining cosmological neutrino mass bounds,” Phys. Rev. D 104
(2021) no. 8, 083504, arXiv:2106.15267 [astro-ph.CO].
[1300] J.-Q. Jiang, W. Giarè, S. Gariazzo, M. G. Dainotti, E. Di Valentino, O. Mena, D. Pedrotti, S. S. da Costa, and
S. Vagnozzi, “Neutrino cosmology after DESI: tightest mass upper limits, preference for the normal ordering, and
tension with terrestrial observations,” JCAP 01 (2025) 153, arXiv:2407.18047 [astro-ph.CO].
[1301] E. Di Valentino, S. Gariazzo, W. Giarè, and O. Mena, “Impact of the damping tail on neutrino mass constraints,”
Phys. Rev. D 108 (2023) no. 8, 083509, arXiv:2305.12989 [astro-ph.CO].
[1302] D. Wang, O. Mena, E. Di Valentino, and S. Gariazzo, “Updating neutrino mass constraints with background
measurements,” Phys. Rev. D 110 (2024) no. 10, 103536, arXiv:2405.03368 [astro-ph.CO].
[1303] J. Hamann and J. Hasenkamp, “A new life for sterile neutrinos: resolving inconsistencies using hot dark matter,”
JCAP 10 (2013) 044, arXiv:1308.3255 [astro-ph.CO].
[1304] P. F. de Salas, D. V. Forero, S. Gariazzo, P. Martínez-Miravé, O. Mena, C. A. Ternes, M. Tórtola, and J. W. F. Valle,
“2020 global reassessment of the neutrino oscillation picture,” JHEP 02 (2021) 071, arXiv:2006.11237 [hep-ph].
303

[1305] I. Esteban, M. C. Gonzalez-Garcia, M. Maltoni, T. Schwetz, and A. Zhou, “The fate of hints: updated global analysis
of three-flavor neutrino oscillations,” JHEP 09 (2020) 178, arXiv:2007.14792 [hep-ph].
[1306] S. Gariazzo, O. Mena, and T. Schwetz, “Quantifying the tension between cosmological and terrestrial constraints on
neutrino masses,” Phys. Dark Univ. 40 (2023) 101226, arXiv:2302.14159 [hep-ph].
[1307] E. di Valentino, S. Gariazzo, and O. Mena, “Model marginalized constraints on neutrino properties from cosmology,”
Phys. Rev. D 106 (2022) no. 4, 043540, arXiv:2207.05167 [astro-ph.CO].
[1308] C. S. Lorenz, L. Funcke, E. Calabrese, and S. Hannestad, “Time-varying neutrino mass from a supercooled phase
transition: current cosmological constraints and impact on the Ωm -σ8 plane,” Phys. Rev. D 99 (2019) no. 2, 023501,
arXiv:1811.01991 [astro-ph.CO].
[1309] M. Escudero, J. Lopez-Pavon, N. Rius, and S. Sandner, “Relaxing Cosmological Neutrino Mass Bounds with Unstable
Neutrinos,” JHEP 12 (2020) 119, arXiv:2007.04994 [hep-ph].
[1310] Z. Chacko, A. Dev, P. Du, V. Poulin, and Y. Tsai, “Determining the Neutrino Lifetime from Cosmology,” Phys. Rev. D
103 (2021) no. 4, 043519, arXiv:2002.08401 [astro-ph.CO].
[1311] Z. Chacko, A. Dev, P. Du, V. Poulin, and Y. Tsai, “Cosmological Limits on the Neutrino Mass and Lifetime,” JHEP
04 (2020) 020, arXiv:1909.05275 [hep-ph].
[1312] G. Franco Abellán, Z. Chacko, A. Dev, P. Du, V. Poulin, and Y. Tsai, “Improved cosmological constraints on the
neutrino mass and lifetime,” JHEP 08 (2022) 076, arXiv:2112.13862 [hep-ph].
[1313] C. D. Kreisch, F.-Y. Cyr-Racine, and O. Doré, “Neutrino puzzle: Anomalies, interactions, and cosmological tensions,”
Phys. Rev. D 101 (2020) no. 12, 123505, arXiv:1902.00534 [astro-ph.CO].
[1314] D. Camarena and F.-Y. Cyr-Racine, “Strong constraints on a simple self-interacting neutrino cosmology,” Phys. Rev. D
111 (2025) no. 2, 023504, arXiv:2403.05496 [astro-ph.CO].
[1315] I. Esteban and J. Salvado, “Long Range Interactions in Cosmology: Implications for Neutrinos,” JCAP 05 (2021) 036,
arXiv:2101.05804 [hep-ph].
[1316] I. M. Oldengott, G. Barenboim, S. Kahlen, J. Salvado, and D. J. Schwarz, “How to relax the cosmological neutrino
mass bound,” JCAP 04 (2019) 049, arXiv:1901.04352 [astro-ph.CO].
[1317] G. Dvali and L. Funcke, “Small neutrino masses from gravitational θ-term,” Phys. Rev. D 93 (2016) no. 11, 113002,
arXiv:1602.03191 [hep-ph].
[1318] G. Barenboim, H. Sanchis, W. H. Kinney, and D. Rios, “Bound on thermal y distortion of the cosmic neutrino
background,” Phys. Rev. D 110 (2024) no. 12, 123535, arXiv:2407.18102 [astro-ph.CO].
[1319] N. Craig, D. Green, J. Meyers, and S. Rajendran, “No νs is Good News,” JHEP 09 (2024) 097, arXiv:2405.00836
[astro-ph.CO].
[1320] A. Yadav, S. Kumar, C. Kibris, and O. Akarsu, “Λs CDM cosmology: alleviating major cosmological tensions by
predicting standard neutrino properties,” JCAP 01 (2025) 042, arXiv:2406.18496 [astro-ph.CO].
[1321] Katrin Collaboration, M. Aker et al., “Direct neutrino-mass measurement based on 259 days of KATRIN data,”
arXiv:2406.13516 [nucl-ex].
[1322] G. Drexlin, V. Hannen, S. Mertens, and C. Weinheimer, “Current direct neutrino mass experiments,” Adv. High Energy
Phys. 2013 (2013) 293986, arXiv:1307.0101 [physics.ins-det].
[1323] P. K. Aluri et al., “Is the observable Universe consistent with the cosmological principle?,” Class. Quant. Grav. 40
(2023) no. 9, 094001, arXiv:2207.05765 [astro-ph.CO].
[1324] G. Ellis, R. Maartens, and M. MacCallum, Relativistic Cosmology. Cambridge University Press, 2012.
https://books.google.it/books?id=IgkhAwAAQBAJ.
[1325] D. J. Schwarz, C. J. Copi, D. Huterer, and G. D. Starkman, “CMB Anomalies after Planck,” Class. Quant. Grav. 33
(2016) no. 18, 184001, arXiv:1510.07929 [astro-ph.CO].
[1326] P. Fleury, C. Clarkson, and R. Maartens, “How does the cosmic large-scale structure bias the Hubble diagram?,” JCAP
03 (2017) 062, arXiv:1612.03726 [astro-ph.CO].
[1327] J. Jones, C. J. Copi, G. D. Starkman, and Y. Akrami, “The Universe is not statistically isotropic,” arXiv:2310.12859
[astro-ph.CO].
[1328] P. Fosalba and E. Gaztanaga, “Explaining Cosmological Anisotropy: Evidence for Causal Horizons from CMB data,”
Mon. Not. Roy. Astron. Soc. 504 (2021) no. 4, 5840–5862, arXiv:2011.00910 [astro-ph.CO].
[1329] S. Yeung and M.-C. Chu, “Directional variations of cosmological parameters from the Planck CMB data,” Phys. Rev.
D 105 (2022) no. 8, 083508, arXiv:2201.03799 [astro-ph.CO].
[1330] Euclid Collaboration, A. Kashlinsky et al., “Euclid preparation - XLVI. The near-infrared background dipole
experiment with Euclid,” Astron. Astrophys. 689 (2024) A294, arXiv:2401.17945 [astro-ph.CO].
[1331] M. Plionis, “Large-scale optical dipole anisotropy,” MNRAS 234 (1988) 401–416.
[1332] M. Plionis and R. Valdarnini, “Evidence for large-scale structure on scales ≡300 h−1 Mpc.,” MNRAS 249 (1991) 46.
[1333] R. Scaramella, G. Vettolani, and G. Zamorani, “The Distribution of Clusters of Galaxies within 300 MPC H -1 and the
Crossover to an Isotropic and Homogeneous Universe,” ApJL 376 (1991) L1.
[1334] E. Branchini and M. Plionis, “Reconstructing positions and peculiar velocities of galaxy clusters within 20000 km/sec.
I: the cluster 3-D dipole,” Astrophys. J. 460 (1996) 569, arXiv:astro-ph/9501028.
[1335] M. Plionis and E. Kolokotronis, “The x-ray cluster dipole,” Astrophys. J. 500 (1998) 1, arXiv:astro-ph/9707147.
[1336] M. Rowan-Robinson et al., “The IRAS PSCz dipole,” Mon. Not. Roy. Astron. Soc. 314 (2000) 375,
arXiv:astro-ph/9912223.
[1337] D. D. Kocevksi, H. Ebeling, B. Tully, and C. R. Mullis, “The Dipole Anisotropy of the First All-Sky X-ray Cluster
Sample,” in Nearby Large-Scale Structures and the Zone of Avoidance, A. P. Fairall and P. A. Woudt, eds., vol. 329 of
304

Astronomical Society of the Pacific Conference Series, p. 89. 2005.


[1338] Y. Hoffman, D. Pomarede, R. Brent Tully, and H. Courtois, “The Dipole Repeller,” arXiv:1702.02483 [astro-ph.CO].
[1339] M. Lopes, A. Bernui, C. Franco, and F. Avila, “Bulk Flow Motion Detection in the Local Universe with Pantheon+
Type Ia Supernovae,” Astrophys. J. 967 (2024) no. 1, 47, arXiv:2405.11077 [astro-ph.CO].
[1340] J. A. R. Cembranos, A. L. Maroto, and H. Villarrubia-Rojo, “Non-comoving Cosmology,” JCAP 06 (2019) 041,
arXiv:1903.11009 [astro-ph.CO].
[1341] C. G. Tsagas, M. I. Kadiltzoglou, and K. Asvesta, “The deceleration parameter in “tilted” Friedmann universes:
Newtonian vs relativistic treatment,” Astrophys. Space Sci. 366 (2021) no. 9, 90, arXiv:2105.09267 [gr-qc].
[1342] C. G. Tsagas, “The deceleration parameter in ‘tilted’ universes: generalising the Friedmann background,” Eur. Phys. J.
C 82 (2022) no. 6, 521, arXiv:2112.04313 [gr-qc].
[1343] K. Asvesta, L. Kazantzidis, L. Perivolaropoulos, and C. G. Tsagas, “Observational constraints on the deceleration
parameter in a tilted universe,” Mon. Not. Roy. Astron. Soc. 513 (2022) no. 2, 2394–2406, arXiv:2202.00962
[astro-ph.CO].
[1344] J. Santiago and C. G. Tsagas, “Timelike vs null deceleration parameter in tilted Friedmann universes,”
arXiv:2203.01126 [gr-qc].
[1345] A. R. King and G. F. R. Ellis, “Tilted homogeneous cosmological models,” Commun. Math. Phys. 31 (1973) 209–242.
[1346] C. G. Tsagas, A. Challinor, and R. Maartens, “Relativistic cosmology and large-scale structure,” Phys. Rept. 465
(2008) 61–147, arXiv:0705.4397 [astro-ph].
[1347] C. Krishnan, R. Mondol, and M. M. Sheikh-Jabbari, “Dipole cosmology: the Copernican paradigm beyond FLRW,”
JCAP 07 (2023) 020, arXiv:2209.14918 [astro-ph.CO].
[1348] C. Krishnan, R. Mondol, and M. M. Sheikh-Jabbari, “A tilt instability in the cosmological principle,” Eur. Phys. J. C
83 (2023) no. 9, 874, arXiv:2211.08093 [astro-ph.CO].
[1349] E. Ebrahimian, C. Krishnan, R. Mondol, and M. M. Sheikh-Jabbari, “Towards a realistic dipole cosmology: the dipole
ΛCDM model,” Class. Quant. Grav. 41 (2024) no. 14, 145007, arXiv:2305.16177 [astro-ph.CO].
[1350] A. Allahyari, E. Ebrahimian, R. Mondol, and M. M. Sheikh-Jabbari, “Big Bang in dipole cosmology,” Eur. Phys. J. C
85 (2025) no. 2, 119, arXiv:2307.15791 [astro-ph.CO].
[1351] C. G. Tsagas and M. I. Kadiltzoglou, “Deceleration parameter in tilted Friedmann universes,” Phys. Rev. D 92 (2015)
no. 4, 043515, arXiv:1507.04266 [gr-qc].
[1352] A. Heinesen and T. Buchert, “Solving the curvature and Hubble parameter inconsistencies through structure
formation-induced curvature,” Class. Quant. Grav. 37 (2020) no. 16, 164001, arXiv:2002.10831 [gr-qc]. [Erratum:
Class.Quant.Grav. 37, 229601 (2020)].
[1353] A. Heinesen, “Multipole decomposition of the general luminosity distance ’Hubble law’ – a new framework for
observational cosmology,” JCAP 05 (2021) 008, arXiv:2010.06534 [astro-ph.CO].
[1354] A. Heinesen and H. J. Macpherson, “A prediction for anisotropies in the nearby Hubble flow,” JCAP 03 (2022) no. 03,
057, arXiv:2111.14423 [astro-ph.CO].
[1355] C. Guandalin, J. Piat, C. Clarkson, and R. Maartens, “Theoretical Systematics in Testing the Cosmological Principle
with the Kinematic Quasar Dipole,” Astrophys. J. 953 (2023) no. 2, 144, arXiv:2212.04925 [astro-ph.CO].
[1356] R. Maartens, J. Santiago, C. Clarkson, B. Kalbouneh, and C. Marinoni, “Covariant cosmography: the
observer-dependence of the Hubble parameter,” JCAP 09 (2024) 070, arXiv:2312.09875 [astro-ph.CO].
[1357] G. F. R. Ellis and J. E. Baldwin, “On the expected anisotropy of radio source counts,” MNRAS 206 (1984) 377–381.
[1358] L. Perivolaropoulos and F. Skara, “Challenges for ΛCDM: An update,” New Astron. Rev. 95 (2022) 101659,
arXiv:2105.05208 [astro-ph.CO].
[1359] O. Akarsu, E. Di Valentino, S. Kumar, M. Ozyigit, and S. Sharma, “Testing spatial curvature and anisotropic
expansion on top of the ΛCDM model,” Phys. Dark Univ. 39 (2023) 101162, arXiv:2112.07807 [astro-ph.CO].
[1360] A. Mariano and L. Perivolaropoulos, “CMB Maximum temperature asymmetry Axis: Alignment with other cosmic
asymmetries,” Phys. Rev. D 87 (2013) no. 4, 043511, arXiv:1211.5915 [astro-ph.CO].
[1361] J. K. Webb, J. A. King, M. T. Murphy, V. V. Flambaum, R. F. Carswell, and M. B. Bainbridge, “Indications of a
spatial variation of the fine structure constant,” Phys. Rev. Lett. 107 (2011) 191101, arXiv:1008.3907 [astro-ph.CO].
[1362] M. R. Wilczynska et al., “Four direct measurements of the fine-structure constant 13 billion years ago,” Sci. Adv. 6
(2020) no. 17, eaay9672, arXiv:2003.07627 [astro-ph.CO].
[1363] R. Mohayaee, M. Rameez, and S. Sarkar, “Do supernovae indicate an accelerating universe?,” Eur. Phys. J. ST 230
(2021) no. 9, 2067–2076, arXiv:2106.03119 [astro-ph.CO].
[1364] J. Colin, R. Mohayaee, M. Rameez, and S. Sarkar, “Evidence for anisotropy of cosmic acceleration,” Astron. Astrophys.
631 (2019) L13, arXiv:1808.04597 [astro-ph.CO].
[1365] R. Mohayaee, M. Rameez, and S. Sarkar, “Cosmological Inference from within the Peculiar Local Universe,” Universe
10 (2024) no. 5, 209, arXiv:2003.10420 [astro-ph.CO].
[1366] A. K. Singal, “Peculiar motion of Solar system from the Hubble diagram of supernovae Ia and its implications for
cosmology,” Mon. Not. Roy. Astron. Soc. 515 (2022) no. 4, 5969–5980, arXiv:2106.11968 [astro-ph.CO].
[1367] N. Horstmann, Y. Pietschke, and D. J. Schwarz, “Inference of the cosmic rest-frame from supernovae Ia,” Astron.
Astrophys. 668 (2022) A34, arXiv:2111.03055 [astro-ph.CO].
[1368] J. A. Cowell, S. Dhawan, and H. J. Macpherson, “Potential signature of a quadrupolar hubble expansion in
Pantheon+supernovae,” Mon. Not. Roy. Astron. Soc. 526 (2023) no. 1, 1482–1494, arXiv:2212.13569 [astro-ph.CO].
[1369] R. Mc Conville and E. O. Colgáin, “Anisotropic distance ladder in Pantheon+supernovae,” Phys. Rev. D 108 (2023)
no. 12, 123533, arXiv:2304.02718 [astro-ph.CO].
305

[1370] U. Andrade, C. A. P. Bengaly, J. S. Alcaniz, and B. Santos, “Isotropy of low redshift type Ia Supernovae: A Bayesian
analysis,” Phys. Rev. D 97 (2018) no. 8, 083518, arXiv:1711.10536 [astro-ph.CO].
[1371] C. A. P. Bengaly, “Evidence for cosmic acceleration with next-generation surveys: A model-independent approach,”
Mon. Not. Roy. Astron. Soc. 499 (2020) no. 1, L6–L10, arXiv:1912.05528 [astro-ph.CO].
[1372] W. Rahman, R. Trotta, S. S. Boruah, M. J. Hudson, and D. A. van Dyk, “New constraints on anisotropic expansion
from supernovae Type Ia,” Mon. Not. Roy. Astron. Soc. 514 (2022) no. 1, 139–163, arXiv:2108.12497 [astro-ph.CO].
[1373] A. Salehi, H. Farajollahi, M. Motahari, P. Pashamokhtari, M. Yarahmadi, and S. Fathi, “Are Type Ia supernova
powerful tool to detect anisotropic expansion of the Universe?,” Eur. Phys. J. C 80 (2020) no. 8, 753.
[1374] J. P. Hu, Y. Y. Wang, and F. Y. Wang, “Testing cosmic anisotropy with Pantheon sample and quasars at high
redshifts,” Astron. Astrophys. 643 (2020) A93, arXiv:2008.12439 [astro-ph.CO].
[1375] S. Dhawan, A. Borderies, H. J. Macpherson, and A. Heinesen, “The quadrupole in the local Hubble parameter: first
constraints using Type Ia supernova data and forecasts for future surveys,” Mon. Not. Roy. Astron. Soc. 519 (2023)
no. 4, 4841–4855, arXiv:2205.12692 [astro-ph.CO].
[1376] D. Sapone, S. Nesseris, and C. A. P. Bengaly, “Is there any measurable redshift dependence on the SN Ia absolute
magnitude?,” Phys. Dark Univ. 32 (2021) 100814, arXiv:2006.05461 [astro-ph.CO].
[1377] C. A. P. Bengaly, J. S. Alcaniz, and C. Pigozzo, “Testing the isotropy of cosmic acceleration with the Pantheon+ and
SH0ES datasets: A cosmographic analysis,” Phys. Rev. D 109 (2024) no. 12, 123533, arXiv:2402.17741
[astro-ph.CO].
[1378] I. Antoniou and L. Perivolaropoulos, “Searching for a Cosmological Preferred Axis: Union2 Data Analysis and
Comparison with Other Probes,” JCAP 12 (2010) 012, arXiv:1007.4347 [astro-ph.CO].
[1379] A. Mariano and L. Perivolaropoulos, “Is there correlation between Fine Structure and Dark Energy Cosmic Dipoles?,”
Phys. Rev. D 86 (2012) 083517, arXiv:1206.4055 [astro-ph.CO].
[1380] C. Krishnan, R. Mohayaee, E. O. Colgáin, M. M. Sheikh-Jabbari, and L. Yin, “Hints of FLRW breakdown from
supernovae,” Phys. Rev. D 105 (2022) no. 6, 063514, arXiv:2106.02532 [astro-ph.CO].
[1381] B. Bahr-Kalus, D. J. Schwarz, M. Seikel, and A. Wiegand, “Constraints on anisotropic cosmic expansion from
supernovae,” Astron. Astrophys. 553 (2013) A56, arXiv:1212.3691 [astro-ph.CO].
[1382] D. Zhao, Y. Zhou, and Z. Chang, “Anisotropy of the Universe via the Pantheon supernovae sample revisited,” Mon.
Not. Roy. Astron. Soc. 486 (2019) no. 4, 5679–5689, arXiv:1903.12401 [astro-ph.CO].
[1383] B. Kalbouneh, C. Marinoni, and J. Bel, “Multipole expansion of the local expansion rate,” Phys. Rev. D 107 (2023)
no. 2, 023507, arXiv:2210.11333 [astro-ph.CO].
[1384] L. Tang, H.-N. Lin, L. Liu, and X. Li, “Consistency of Pantheon+ supernovae with a large-scale isotropic universe*,”
Chin. Phys. C 47 (2023) no. 12, 125101, arXiv:2309.11320 [astro-ph.CO].
[1385] J. P. Hu, Y. Y. Wang, J. Hu, and F. Y. Wang, “Testing the cosmological principle with the Pantheon+ sample and the
region-fitting method,” Astron. Astrophys. 681 (2024) A88, arXiv:2310.11727 [astro-ph.CO].
[1386] F. Sorrenti, R. Durrer, and M. Kunz, “The dipole of the Pantheon+SH0ES data,” JCAP 11 (2023) 054,
arXiv:2212.10328 [astro-ph.CO].
[1387] L. Perivolaropoulos, “Isotropy properties of the absolute luminosity magnitudes of SnIa in the Pantheon+ and SH0ES
samples,” Phys. Rev. D 108 (2023) no. 6, 063509, arXiv:2305.12819 [astro-ph.CO].
[1388] G. Risaliti and E. Lusso, “A Hubble Diagram for Quasars,” Astrophys. J. 815 (2015) 33, arXiv:1505.07118
[astro-ph.CO].
[1389] G. Risaliti and E. Lusso, “Cosmological constraints from the Hubble diagram of quasars at high redshifts,” Nature
Astron. 3 (2019) no. 3, 272–277, arXiv:1811.02590 [astro-ph.CO].
[1390] N. Khadka and B. Ratra, “Do quasar X-ray and UV flux measurements provide a useful test of cosmological models?,”
Mon. Not. Roy. Astron. Soc. 510 (2022) no. 2, 2753–2772, arXiv:2107.07600 [astro-ph.CO].
[1391] N. Khadka, M. Zajaček, R. Prince, S. Panda, B. Czerny, M. L. Martínez-Aldama, V. K. Jaiswal, and B. Ratra,
“Quasar UV/X-ray relation luminosity distances are shorter than reverberation-measured radius–luminosity relation
luminosity distances,” Mon. Not. Roy. Astron. Soc. 522 (2023) no. 1, 1247–1264, arXiv:2212.10483 [astro-ph.CO].
[1392] E. O. Colgáin, M. M. Sheikh-Jabbari, and L. Yin, “Do high redshift QSOs and GRBs corroborate JWST?,”
arXiv:2405.19953 [astro-ph.CO].
[1393] N. J. Secrest, S. von Hausegger, M. Rameez, R. Mohayaee, S. Sarkar, and J. Colin, “A Test of the Cosmological
Principle with Quasars,” Astrophys. J. Lett. 908 (2021) no. 2, L51, arXiv:2009.14826 [astro-ph.CO].
[1394] A. K. Singal, “Peculiar motion of the solar system derived from a dipole anisotropy in the redshift distribution of
distant quasars,” Mon. Not. Roy. Astron. Soc. 488 (2019) no. 1, L104–L108, arXiv:1405.4796 [astro-ph.CO].
[1395] L. Dam, G. F. Lewis, and B. J. Brewer, “Testing the cosmological principle with CatWISE quasars: a bayesian analysis
of the number-count dipole,” Mon. Not. Roy. Astron. Soc. 525 (2023) no. 1, 231–245, arXiv:2212.07733
[astro-ph.CO].
[1396] R. Kothari, M. Panwar, G. Singh, P. Tiwari, and P. Jain, “A study of dipolar signal in distant Quasars with various
observables,” Eur. Phys. J. C 84 (2024) no. 1, 75, arXiv:2208.14397 [astro-ph.CO].
[1397] A. Abghari, E. F. Bunn, L. T. Hergt, B. Li, D. Scott, R. M. Sullivan, and D. Wei, “Reassessment of the dipole in the
distribution of quasars on the sky,” JCAP 11 (2024) 067, arXiv:2405.09762 [astro-ph.CO].
[1398] O. Luongo, M. Muccino, E. O. Colgáin, M. M. Sheikh-Jabbari, and L. Yin, “Larger H0 values in the CMB dipole
direction,” Phys. Rev. D 105 (2022) no. 10, 103510, arXiv:2108.13228 [astro-ph.CO].
[1399] C. Blake and J. Wall, “Detection of the velocity dipole in the radio galaxies of the nrao vla sky survey,” Nature 416
(2002) 150–152, arXiv:astro-ph/0203385.
306

[1400] J. Colin, R. Mohayaee, M. Rameez, and S. Sarkar, “High redshift radio galaxies and divergence from the CMB dipole,”
Mon. Not. Roy. Astron. Soc. 471 (2017) no. 1, 1045–1055, arXiv:1703.09376 [astro-ph.CO].
[1401] A. K. Singal, “Large peculiar motion of the solar system from the dipole anisotropy in sky brightness due to distant
radio sources,” Astrophys. J. Lett. 742 (2011) L23, arXiv:1110.6260 [astro-ph.CO].
[1402] M. Rubart and D. J. Schwarz, “Cosmic radio dipole from NVSS and WENSS,” Astron. Astrophys. 555 (2013) A117,
arXiv:1301.5559 [astro-ph.CO].
[1403] C. A. P. Bengaly, T. M. Siewert, D. J. Schwarz, and R. Maartens, “Testing the standard model of cosmology with the
SKA: the cosmic radio dipole,” Mon. Not. Roy. Astron. Soc. 486 (2019) no. 1, 1350–1357, arXiv:1810.04960
[astro-ph.CO].
[1404] A. K. Singal, “Large disparity in cosmic reference frames determined from the sky distributions of radio sources and
the microwave background radiation,” Phys. Rev. D 100 (2019) no. 6, 063501, arXiv:1904.11362 [physics.gen-ph].
[1405] T. M. Siewert, M. Schmidt-Rubart, and D. J. Schwarz, “Cosmic radio dipole: Estimators and frequency dependence,”
Astron. Astrophys. 653 (2021) A9, arXiv:2010.08366 [astro-ph.CO].
[1406] N. J. Secrest, S. von Hausegger, M. Rameez, R. Mohayaee, and S. Sarkar, “A Challenge to the Standard Cosmological
Model,” Astrophys. J. Lett. 937 (2022) no. 2, L31, arXiv:2206.05624 [astro-ph.CO].
[1407] J. D. Wagenveld, H.-R. Klöckner, and D. J. Schwarz, “The cosmic radio dipole: Bayesian estimators on new and old
radio surveys,” Astron. Astrophys. 675 (2023) A72, arXiv:2305.15335 [astro-ph.CO].
[1408] A. K. Singal, “Resolution of the incongruency of dipole asymmetries within various large radio surveys – implications
for the Cosmological Principle,” Mon. Not. Roy. Astron. Soc. 528 (2024) no. 4, 5679–5691, arXiv:2312.12785
[astro-ph.CO].
[1409] C. Gibelyou and D. Huterer, “Dipoles in the Sky,” Mon. Not. Roy. Astron. Soc. 427 (2012) 1994–2021,
arXiv:1205.6476 [astro-ph.CO].
[1410] C. A. P. Bengaly, R. Maartens, N. Randriamiarinarivo, and A. Baloyi, “Testing the Cosmological Principle in the radio
sky,” JCAP 09 (2019) 025, arXiv:1905.12378 [astro-ph.CO].
[1411] C. Murray, “The effects of lensing by local structures on the dipole of radio source counts,” Mon. Not. Roy. Astron.
Soc. 510 (2022) no. 2, 3098–3101, arXiv:2112.06689 [astro-ph.CO].
[1412] Y.-T. Cheng, T.-C. Chang, and A. Lidz, “Is the Radio Source Dipole from NVSS Consistent with the Cosmic
Microwave Background and ΛCDM?,” Astrophys. J. 965 (2024) no. 1, 32, arXiv:2309.02490 [astro-ph.CO].
[1413] P. da Silveira Ferreira and V. Marra, “Tomographic redshift dipole: testing the cosmological principle,” JCAP 09
(2024) 077, arXiv:2403.14580 [astro-ph.CO].
[1414] O. T. Oayda, V. Mittal, G. F. Lewis, and T. Murphy, “A Bayesian approach to the cosmic dipole in radio galaxy
surveys: joint analysis of NVSS & RACS,” Mon. Not. Roy. Astron. Soc. 531 (2024) no. 4, 4545–4559,
arXiv:2406.01871 [astro-ph.CO].
[1415] B. E. Schaefer, “The Hubble Diagram to Redshift >6 from 69 Gamma-Ray Bursts,” Astrophys. J. 660 (2007) 16–46,
arXiv:astro-ph/0612285.
[1416] S. Basilakos and L. Perivolaropoulos, “Testing GRBs as Standard Candles,” Mon. Not. Roy. Astron. Soc. 391 (2008)
411–419, arXiv:0805.0875 [astro-ph].
[1417] N. Liang, W. K. Xiao, Y. Liu, and S. N. Zhang, “A Cosmology Independent Calibration of Gamma-Ray Burst
Luminosity Relations and the Hubble Diagram,” Astrophys. J. 685 (2008) 354, arXiv:0802.4262 [astro-ph].
[1418] S. Cao, N. Khadka, and B. Ratra, “Standardizing Dainotti-correlated gamma-ray bursts, and using them with
standardized Amati-correlated gamma-ray bursts to constrain cosmological model parameters,” Mon. Not. Roy.
Astron. Soc. 510 (2022) no. 2, 2928–2947, arXiv:2110.14840 [astro-ph.CO].
[1419] R. B. Tully et al., “Cosmicflows-4,” Astrophys. J. 944 (2023) no. 1, 94, arXiv:2209.11238 [astro-ph.CO].
[1420] Y. Hoffman, A. Valade, N. I. Libeskind, J. G. Sorce, R. B. Tully, S. Pfeifer, S. Gottlöber, and D. Pomarède, “The
large-scale velocity field from the Cosmicflows-4 data,” MNRAS 527 (2024) no. 2, 3788–3805, arXiv:2311.01340
[astro-ph.CO].
[1421] R. Watkins, T. Allen, C. J. Bradford, A. Ramon, A. Walker, H. A. Feldman, R. Cionitti, Y. Al-Shorman, E. Kourkchi,
and R. B. Tully, “Analysing the large-scale bulk flow using cosmicflows4: increasing tension with the standard
cosmological model,” Mon. Not. Roy. Astron. Soc. 524 (2023) no. 2, 1885–1892, arXiv:2302.02028 [astro-ph.CO].
[1422] A. M. Whitford, C. Howlett, and T. M. Davis, “Evaluating bulk flow estimators for CosmicFlows–4 measurements,”
Mon. Not. Roy. Astron. Soc. 526 (2023) no. 2, 3051–3071, arXiv:2306.11269 [astro-ph.CO].
[1423] S. Peery, R. Watkins, and H. A. Feldman, “Easily Interpretable Bulk Flows: Continuing Tension with the Standard
Cosmological Model,” Mon. Not. Roy. Astron. Soc. 481 (2018) no. 1, 1368–1375, arXiv:1808.07772 [astro-ph.CO].
[1424] A. Kashlinsky, F. Atrio-Barandela, D. Kocevski, and H. Ebeling, “A measurement of large-scale peculiar velocities of
clusters of galaxies: results and cosmological implications,” Astrophys. J. Lett. 686 (2009) L49–L52, arXiv:0809.3734
[astro-ph].
[1425] F. Atrio-Barandela, A. Kashlinsky, H. Ebeling, D. J. Fixsen, and D. Kocevski, “Probing the Dark Flow Signal in
Wmap 9 -year and Planck Cosmic Microwave Background Maps,” Astrophys. J. 810 (2015) no. 2, 143,
arXiv:1411.4180 [astro-ph.CO].
[1426] A. Kashlinsky, F. Atrio-Barandela, H. Ebeling, A. Edge, and D. Kocevski, “A new measurement of the bulk flow of
X-ray luminous clusters of galaxies,” Astrophys. J. Lett. 712 (2010) L81–L85, arXiv:0910.4958 [astro-ph.CO].
[1427] A. Kashlinsky, F. Atrio-Barandela, and H. Ebeling, “Measuring the dark flow with public X-ray cluster data,”
Astrophys. J. 732 (2011) 1, arXiv:1012.3214 [astro-ph.CO].
[1428] F. Atrio-Barandela, “On the Statistical Significance of the Bulk Flow Measured by the PLANCK Satellite,” Astron.
307

Astrophys. 557 (2013) A116, arXiv:1303.6614 [astro-ph.CO].


[1429] Planck Collaboration, N. Aghanim et al., “Planck 2018 results. I. Overview and the cosmological legacy of Planck,”
Astron. Astrophys. 641 (2020) A1, arXiv:1807.06205 [astro-ph.CO].
[1430] J. Grande and L. Perivolaropoulos, “Generalized LTB model with Inhomogeneous Isotropic Dark Energy:
Observational Constraints,” Phys. Rev. D 84 (2011) 023514, arXiv:1103.4143 [astro-ph.CO].
[1431] H. Alnes, M. Amarzguioui, and O. Gron, “Can a dust dominated Universe have accelerated expansion?,” JCAP 01
(2007) 007, arXiv:astro-ph/0506449.
[1432] L. Ackerman, S. M. Carroll, and M. B. Wise, “Imprints of a Primordial Preferred Direction on the Microwave
Background,” Phys. Rev. D 75 (2007) 083502, arXiv:astro-ph/0701357. [Erratum: Phys.Rev.D 80, 069901 (2009)].
[1433] R. Aurich, S. Lustig, and F. Steiner, “CMB anisotropy of the Poincare dodecahedron,” Class. Quant. Grav. 22 (2005)
2061–2083, arXiv:astro-ph/0412569.
[1434] COMPACT Collaboration, Y. Akrami et al., “Promise of Future Searches for Cosmic Topology,” Phys. Rev. Lett.
132 (2024) no. 17, 171501, arXiv:2210.11426 [astro-ph.CO].
[1435] J. C. Bueno Sanchez and L. Perivolaropoulos, “Topological Quintessence,” Phys. Rev. D 84 (2011) 123516,
arXiv:1110.2587 [astro-ph.CO].
[1436] L. Perivolaropoulos, “Topological quintessence: Generalizing Lambda CDM with inhomogeneous dark energy,” Rom. J.
Phys. 57 (2012) 950–968.
[1437] T. Koivisto and D. F. Mota, “Dark energy anisotropic stress and large scale structure formation,” Phys. Rev. D 73
(2006) 083502, arXiv:astro-ph/0512135.
[1438] R. Battye and A. Moss, “Anisotropic dark energy and CMB anomalies,” Phys. Rev. D 80 (2009) 023531,
arXiv:0905.3403 [astro-ph.CO].
[1439] J. D. Barrow, “Cosmological limits on slightly skew stresses,” Phys. Rev. D 55 (1997) 7451–7460,
arXiv:gr-qc/9701038.
[1440] J. D. Barrow, P. G. Ferreira, and J. Silk, “Constraints on a primordial magnetic field,” Phys. Rev. Lett. 78 (1997)
3610–3613, arXiv:astro-ph/9701063.
[1441] L. Campanelli, “A Model of Universe Anisotropization,” Phys. Rev. D 80 (2009) 063006, arXiv:0907.3703
[astro-ph.CO].
[1442] A. Maleknejad, M. M. Sheikh-Jabbari, and J. Soda, “Gauge Fields and Inflation,” Phys. Rept. 528 (2013) 161–261,
arXiv:1212.2921 [hep-th].
[1443] K. Dimopoulos, M. Karciauskas, and J. M. Wagstaff, “Vector Curvaton with varying Kinetic Function,” Phys. Rev. D
81 (2010) 023522, arXiv:0907.1838 [hep-ph].
[1444] M. Karciauskas, K. Dimopoulos, and D. H. Lyth, “Anisotropic non-Gaussianity from vector field perturbations,” Phys.
Rev. D 80 (2009) 023509, arXiv:0812.0264 [astro-ph]. [Erratum: Phys.Rev.D 85, 069905 (2012)].
[1445] C. G. Tsagas, “Peculiar motions, accelerated expansion and the cosmological axis,” Phys. Rev. D 84 (2011) 063503,
arXiv:1107.4045 [astro-ph.CO].
[1446] T. Anton and T. Clifton, “Modelling the emergence of cosmic anisotropy from non-linear structures,” Class. Quant.
Grav. 40 (2023) no. 14, 145004, arXiv:2302.05715 [gr-qc].
[1447] E. Pastén and V. H. Cárdenas, “Testing ΛCDM cosmology in a binned universe: Anomalies in the deceleration
parameter,” Phys. Dark Univ. 40 (2023) 101224, arXiv:2301.10740 [astro-ph.CO].
[1448] C. G. Tsagas, “The peculiar Jeans length,” Eur. Phys. J. C 81 (2021) no. 8, 753, arXiv:2103.15884 [gr-qc].
[1449] O. Roldan, A. Notari, and M. Quartin, “Interpreting the CMB aberration and Doppler measurements: boost or
intrinsic dipole?,” JCAP 06 (2016) 026, arXiv:1603.02664 [astro-ph.CO].
[1450] S. Yasini and E. Pierpaoli, “Beyond the Boost: Measuring the Intrinsic Dipole of the Cosmic Microwave Background
Using the Spectral Distortions of the Monopole and Quadrupole,” Phys. Rev. Lett. 119 (2017) no. 22, 221102,
arXiv:1610.00015 [astro-ph.CO].
[1451] P. d. S. Ferreira and M. Quartin, “First Constraints on the Intrinsic CMB Dipole and Our Velocity with Doppler and
Aberration,” Phys. Rev. Lett. 127 (2021) no. 10, 101301, arXiv:2011.08385 [astro-ph.CO].
[1452] M. I. Khan and R. Saha, “Detection of Dipole Modulation in CMB Temperature Anisotropy Maps from WMAP and
Planck using Artificial Intelligence,” Astrophys. J. 947 (2023) no. 2, 47, arXiv:2212.04438 [astro-ph.CO].
[1453] C. E. Kester, A. Bernui, and W. S. Hipólito-Ricaldi, “Probing the statistical isotropy of the universe with Planck data
of the cosmic microwave background,” Astron. Astrophys. 683 (2024) A176, arXiv:2310.02928 [astro-ph.CO].
[1454] J. A. King, J. K. Webb, M. T. Murphy, V. V. Flambaum, R. F. Carswell, M. B. Bainbridge, M. R. Wilczynska, and
F. E. Koch, “Spatial variation in the fine-structure constant – new results from VLT/UVES,” Mon. Not. Roy. Astron.
Soc. 422 (2012) 3370–3413, arXiv:1202.4758 [astro-ph.CO].
[1455] D.-C. Qiang, H.-K. Deng, and H. Wei, “Cosmic Anisotropy and Fast Radio Bursts,” Class. Quant. Grav. 37 (2020)
no. 18, 185022, arXiv:1902.03580 [astro-ph.CO].
[1456] H.-N. Lin and Y. Sang, “Probing the anisotropic distribution of baryon matter in the Universe using fast radio bursts
*,” Chin. Phys. C 45 (2021) no. 12, 125101, arXiv:2111.12934 [astro-ph.CO].
[1457] O. T. Oayda and G. F. Lewis, “Testing the cosmological principle: on the time dilation of distant sources,” Mon. Not.
Roy. Astron. Soc. 523 (2023) no. 1, 667–675, arXiv:2305.06771 [astro-ph.CO].
[1458] V. Mittal, O. T. Oayda, and G. F. Lewis, “The Cosmic Dipole in the Quaia Sample of Quasars: A Bayesian Analysis,”
Mon. Not. Roy. Astron. Soc. 527 (2024) no. 3, 8497–8510, arXiv:2311.14938 [astro-ph.CO]. [Erratum:
Mon.Not.Roy.Astron.Soc. 530, 4763–4764 (2024)].
[1459] F. Iocco, G. Mangano, G. Miele, O. Pisanti, and P. D. Serpico, “Primordial Nucleosynthesis: from precision cosmology
308

to fundamental physics,” Phys. Rept. 472 (2009) 1–76, arXiv:0809.0631 [astro-ph].


[1460] R. H. Cyburt, B. D. Fields, K. A. Olive, and T.-H. Yeh, “Big Bang Nucleosynthesis: 2015,” Rev. Mod. Phys. 88 (2016)
015004, arXiv:1505.01076 [astro-ph.CO].
[1461] I. Tanihata, H. Toki, and T. Kajino, eds., Big Bang Nucleosynthesis, pp. 1–21. Springer Nature Singapore, 2023.
arXiv:2301.12299 [astro-ph.CO].
[1462] Particle Data Group Collaboration, R. L. Workman et al., “Review of Particle Physics,” PTEP 2022 (2022) 083C01.
[1463] T.-H. Yeh, K. A. Olive, and B. D. Fields, “The impact of new d(p, γ)3 rates on Big Bang Nucleosynthesis,” JCAP 03
(2021) 046, arXiv:2011.13874 [astro-ph.CO].
[1464] A. Coc, S. Goriely, Y. Xu, M. Saimpert, and E. Vangioni, “Standard Big-Bang Nucleosynthesis up to CNO with an
improved extended nuclear network,” Astrophys. J. 744 (2012) 158, arXiv:1107.1117 [astro-ph.CO].
[1465] R. Consiglio, P. F. de Salas, G. Mangano, G. Miele, S. Pastor, and O. Pisanti, “PArthENoPE reloaded,” Comput.
Phys. Commun. 233 (2018) 237–242, arXiv:1712.04378 [astro-ph.CO].
[1466] C. Pitrou, A. Coc, J.-P. Uzan, and E. Vangioni, “A new tension in the cosmological model from primordial
deuterium?,” Mon. Not. Roy. Astron. Soc. 502 (2021) no. 2, 2474–2481, arXiv:2011.11320 [astro-ph.CO].
[1467] V. Singh, D. Bhowmick, and D. N. Basu, “Re-examining the Lithium abundance problem in Big-Bang
nucleosynthesis,” Astropart. Phys. 162 (2024) 102995, arXiv:2304.08032 [nucl-th].
[1468] S. Gariazzo, P. F. de Salas, O. Pisanti, and R. Consiglio, “PArthENoPE revolutions,” Comput. Phys. Commun. 271
(2022) 108205, arXiv:2103.05027 [astro-ph.IM].
[1469] L. Sbordone et al., “The metal-poor end of the Spite plateau. 1: Stellar parameters, metallicities and lithium
abundances,” Astron. Astrophys. 522 (2010) A26, arXiv:1003.4510 [astro-ph.GA].
[1470] J. C. Howk, N. Lehner, B. D. Fields, and G. J. Mathews, “The detection of interstellar lithium in a low-metallicity
galaxy,” Nature 489 (2012) 121, arXiv:1207.3081 [astro-ph.CO].
[1471] L. Izzo, P. Molaro, G. Cescutti, E. Aydi, P. Selvelli, E. Harvey, A. Agnello, P. Bonifacio, M. Della Valle, E. Guido, and
M. Hernanz, “Detection of 7 Be II in the Small Magellanic Cloud,” MNRAS 510 (2022) no. 4, 5302–5314,
arXiv:2112.11859 [astro-ph.SR].
[1472] B. D. Fields, “The primordial lithium problem,” Ann. Rev. Nucl. Part. Sci. 61 (2011) 47–68, arXiv:1203.3551
[astro-ph.CO].
[1473] F. Spite and M. Spite, “Abundance of lithium in unevolved halo stars and old disk stars: Interpretation and
consequences,” Astron. Astrophys. 115 (1982) 357–366.
[1474] R. Rebolo, P. Molaro, and J. E. Beckman, “Lithium abundances in metal-deficient dwarfs.,” A&A 192 (1988) 192–205.
[1475] F. Matteucci, M. Molero, D. S. Aguado, and D. Romano, “The evolution of Lithium: implications of a universal Spite
plateau,” MNRAS 505 (2021) no. 1, 200–206, arXiv:2104.11504 [astro-ph.GA].
[1476] T. Makki, M. E. Eid, and G. Mathews, “New Insight Concerning Primordial Lithium Production,” arXiv:2402.17871
[astro-ph.CO].
[1477] X. Fu, A. Bressan, P. Molaro, and P. Marigo, “Lithium evolution in metal-poor stars: from pre-main sequence to the
Spite plateau,” MNRAS 452 (2015) no. 3, 3256–3265, arXiv:1506.05993 [astro-ph.SR].
[1478] B. D. Fields and K. A. Olive, “Implications of the non-observation of 6 Li in halo stars for the primordial 7 Li problem,”
JCAP 10 (2022) 078, arXiv:2204.03167 [astro-ph.GA].
[1479] V. V. Smith, D. L. Lambert, and P. E. Nissen, “The 6Li/7Li ratio in the metal-poor halo dwarfs HD 19445 and HD
84937,” Astrophys. J. 408 (1993) 262–276.
[1480] L. M. Hobbs and J. A. Thorburn, “Lithium Isotope Ratios in Six Halo Stars,” ApJL 428 (1994) L25.
[1481] L. M. Hobbs and J. A. Thorburn, “Lithium Isotope Ratios in Halo Stars. II.,” Astrophys. J. 491 (1997) no. 2, 772–788.
[1482] V. V. Smith, D. L. Lambert, and P. E. Nissen, “Isotopic Lithium Abundances in Nine Halo Stars,” Astrophys. J. 506
(1998) no. 1, 405–423.
[1483] L. M. Hobbs, J. A. Thorburn, and L. M. Rebull, “Lithium Isotope Ratios in Halo Stars. III.,” Astrophys. J. 523 (1999)
no. 2, 797–804.
[1484] R. Cayrel, M. Spite, F. Spite, E. Vangioni-Flam, M. Casse, J. A. R. Cayrel, and J. Audouze, “New high s/n
observations of the li-6/li-7 blend in hd 84937 and two other metal-poor stars,” Astron. Astrophys. 343 (1999) 923,
arXiv:astro-ph/9901205.
[1485] M. Asplund, D. L. Lambert, P. E. Nissen, F. Primas, and V. V. Smith, “Lithium isotopic abundances in metal-poor
halo stars,” Astrophys. J. 644 (2006) 229–259, arXiv:astro-ph/0510636.
[1486] M. Steffen, R. Cayrel, P. Bonifacio, H. G. Ludwig, and E. Caffau, “6Li in metal-poor halo stars: real or spurious?,”
IAU Symp. 265 (2010) 23, arXiv:0910.5917 [astro-ph.SR].
[1487] A. E. G. Perez, W. Aoki, S. Inoue, S. G. Ryan, T. K. Suzuki, and M. Chiba, “6Li/7Li estimates for metal-poor stars,”
Astron. Astrophys. 504 (2009) 213, arXiv:0909.5163 [astro-ph.SR].
[1488] K. Lind, J. Melendez, M. Asplund, R. Collet, and Z. Magic, “The lithium isotopic ratio in very metal-poor stars,”
Astron. Astrophys. 554 (2013) A96, arXiv:1305.6564 [astro-ph.SR].
[1489] E. X. Wang, T. Nordlander, M. Asplund, K. Lind, Y. Zhou, and H. Reggiani, “Non-detection of 6Li in Spite plateau
stars with ESPRESSO,” Mon. Not. Roy. Astron. Soc. 509 (2021) no. 1, 1521–1535, arXiv:2110.03822.
[1490] A. Korn, “The ups and downs of inferred cosmological lithium,” EPJ Web Conf. 297 (2024) 01007, arXiv:2406.09974
[astro-ph.SR].
[1491] GALAH Collaboration, X. Gao et al., “The GALAH Survey: A new constraint on cosmological lithium and Galactic
lithium evolution from warm dwarf stars,” Mon. Not. Roy. Astron. Soc. 497 (2020) no. 1, L30–L34, arXiv:2006.05173
[astro-ph.SR].
309

[1492] K. A. Olive, P. Petitjean, E. Vangioni, and J. Silk, “Higher D or Li: Probes of Physics beyond the Standard Model,”
Mon. Not. Roy. Astron. Soc. 426 (2012) 1427, arXiv:1203.5701 [astro-ph.CO].
[1493] A. Coc, M. Pospelov, J.-P. Uzan, and E. Vangioni, “Modified big bang nucleosynthesis with nonstandard neutron
sources,” Phys. Rev. D 90 (2014) no. 8, 085018, arXiv:1405.1718 [hep-ph].
[1494] K. Jedamzik, “Neutralinos and Big Bang nucleosynthesis,” Phys. Rev. D 70 (2004) 083510, arXiv:astro-ph/0405583.
[1495] K. Jedamzik, “Did something decay, evaporate, or annihilate during Big Bang nucleosynthesis?,” Phys. Rev. D 70
(2004) 063524, arXiv:astro-ph/0402344.
[1496] A. Goudelis, M. Pospelov, and J. Pradler, “Light Particle Solution to the Cosmic Lithium Problem,” Phys. Rev. Lett.
116 (2016) no. 21, 211303, arXiv:1510.08858 [hep-ph].
[1497] S. Q. Hou, J. J. He, A. Parikh, D. Kahl, C. A. Bertulani, T. Kajino, G. J. Mathews, and G. Zhao, “Non-extensive
Statistics to the Cosmological Lithium Problem,” Astrophys. J. 834 (2017) no. 2, 165, arXiv:1701.04149
[astro-ph.CO].
[1498] R. Nakamura, M.-a. Hasahimoto, R. Ichimasa, and K. Arai, “Big-Bang nucleosynthesis: Constraints on nuclear
reaction rates, neutrino degeneracy, inhomogeneous and Brans–Dicke models,” Int. J. Mod. Phys. E 26 (2017) no. 08,
1741003, arXiv:1710.08153 [astro-ph.CO].
[1499] D. G. Yamazaki, M. Kusakabe, T. Kajino, G. J. Mathews, and M.-K. Cheoun, “The new hybrid BBN model with the
photon cooling, X particle, and the primordial magnetic field,” Int. J. Mod. Phys. E 26 (2017) no. 08, 1741006.
[1500] T. R. Makki, M. F. El Eid, and G. J. Mathews, “Impact of Neutrino Properties and Dark Matter on the Primordial
Lithium Production,” Int. J. Mod. Phys. E 28 (2019) no. 08, 1950065, arXiv:1901.03726 [astro-ph.CO].
[1501] T. R. Makki, M. F. El Eid, and G. J. Mathews, “A critical analysis of the Big Bang Nucleosynthesis,” Mod. Phys. Lett.
A 34 (2019) no. 24, 1950194.
[1502] A.-K. Burns, V. Keus, M. Sher, and T. M. P. Tait, “Constraints on variation of the weak scale from big bang
nucleosynthesis,” Phys. Rev. D 109 (2024) no. 12, 123506, arXiv:2402.08626 [hep-ph].
[1503] G. J. Mathews, A. Kedia, N. Sasankan, M. Kusakabe, Y. Luo, T. Kajino, D. Yamazaki, T. Makki, and M. E. Eid,
“Cosmological Solutions to the Lithium Problem,” JPS Conf. Proc. 31 (2020) 011033, arXiv:1909.01245
[astro-ph.CO].
[1504] T. M. Bania, R. T. Rood, and D. S. Balser, “The cosmological density of baryons from observations of 3He+ in the
Milky Way,” Nature 415 (2002) 54–57.
[1505] E. Aver, D. A. Berg, K. A. Olive, R. W. Pogge, J. J. Salzer, and E. D. Skillman, “Improving helium abundance
determinations with Leo P as a case study,” JCAP 03 (2021) 027, arXiv:2010.04180 [astro-ph.CO].
[1506] M. Valerdi, A. Peimbert, M. Peimbert, and A. Sixtos, “Determination of the Primordial Helium Abundance Based on
NGC 346, an H ii Region of the Small Magellanic Cloud,” Astrophys. J. 876 (2019) no. 2, 98, arXiv:1904.01594
[astro-ph.GA].
[1507] V. Fernández, E. Terlevich, A. I. Díaz, and R. Terlevich, “A Bayesian direct method implementation to fit emission line
spectra: Application to the primordial He abundance determination,” Mon. Not. Roy. Astron. Soc. 487 (2019) no. 3,
3221–3238, arXiv:1905.09215 [astro-ph.GA].
[1508] O. A. Kurichin, P. A. Kislitsyn, V. V. Klimenko, S. A. Balashev, and A. V. Ivanchik, “A new determination of the
primordial helium abundance using the analyses of H II region spectra from SDSS,” Mon. Not. Roy. Astron. Soc. 502
(2021) no. 2, 3045–3056, arXiv:2101.09127 [astro-ph.CO].
[1509] T. Hsyu, R. J. Cooke, J. X. Prochaska, and M. Bolte, “The PHLEK Survey: A New Determination of the Primordial
Helium Abundance,” Astrophys. J. 896 (2020) no. 1, 77, arXiv:2005.12290 [astro-ph.GA].
[1510] M. Valerdi, A. Peimbert, and M. Peimbert, “Chemical abundances in seven metal-poor H II regions and a
determination of the primordial helium abundance,” MNRAS 505 (2021) no. 3, 3624–3634, arXiv:2105.12260
[astro-ph.GA].
[1511] A. Matsumoto et al., “EMPRESS. VIII. A New Determination of Primordial He Abundance with Extremely
Metal-poor Galaxies: A Suggestion of the Lepton Asymmetry and Implications for the Hubble Tension,” Astrophys. J.
941 (2022) no. 2, 167, arXiv:2203.09617 [astro-ph.CO].
[1512] T. Takahashi and S. Yamashita, “Big bang nucleosynthesis and early dark energy in light of the EMPRESS Yp results
and the H0 tension,” Phys. Rev. D 107 (2023) no. 10, 103520, arXiv:2211.04087 [astro-ph.CO].
[1513] M. Escudero, A. Ibarra, and V. Maura, “Primordial lepton asymmetries in the precision cosmology era: Current status
and future sensitivities from BBN and the CMB,” Phys. Rev. D 107 (2023) no. 3, 035024, arXiv:2208.03201
[hep-ph].
[1514] Y. I. Izotov, T. X. Thuan, and N. G. Guseva, “A new determination of the primordial He abundance using the He i
λ10830 Å emission line: cosmological implications,” Mon. Not. Roy. Astron. Soc. 445 (2014) no. 1, 778–793,
arXiv:1408.6953 [astro-ph.CO].
[1515] C. Pitrou, A. Coc, J.-P. Uzan, and E. Vangioni, “Resolving conclusions about the early Universe requires accurate
nuclear measurements,” Nature Rev. Phys. 3 (2021) no. 4, 231–232, arXiv:2104.11148 [astro-ph.CO].
[1516] O. Pisanti, G. Mangano, G. Miele, and P. Mazzella, “Primordial Deuterium after LUNA: concordances and error
budget,” JCAP 04 (2021) 020, arXiv:2011.11537 [astro-ph.CO].
[1517] E. Di Valentino, C. Gustavino, J. Lesgourgues, G. Mangano, A. Melchiorri, G. Miele, and O. Pisanti, “Probing nuclear
rates with Planck and BICEP2,” Phys. Rev. D 90 (2014) no. 2, 023543, arXiv:1404.7848 [astro-ph.CO].
[1518] G. E. Addison, G. Hinshaw, and M. Halpern, “Cosmological constraints from baryon acoustic oscillations and clustering
of large-scale structure,” Mon. Not. Roy. Astron. Soc. 436 (2013) 1674–1683, arXiv:1304.6984 [astro-ph.CO].
[1519] G. E. Addison, D. J. Watts, C. L. Bennett, M. Halpern, G. Hinshaw, and J. L. Weiland, “Elucidating ΛCDM: Impact
310

of Baryon Acoustic Oscillation Measurements on the Hubble Constant Discrepancy,” Astrophys. J. 853 (2018) no. 2,
119, arXiv:1707.06547 [astro-ph.CO].
[1520] eBOSS Collaboration, M. Blomqvist et al., “Baryon acoustic oscillations from the cross-correlation of Lyα absorption
and quasars in eBOSS DR14,” Astron. Astrophys. 629 (2019) A86, arXiv:1904.03430 [astro-ph.CO].
[1521] A. Cuceu, J. Farr, P. Lemos, and A. Font-Ribera, “Baryon Acoustic Oscillations and the Hubble Constant: Past,
Present and Future,” JCAP 10 (2019) 044, arXiv:1906.11628 [astro-ph.CO].
[1522] N. Schöneberg, L. Verde, H. Gil-Marín, and S. Brieden, “BAO+BBN revisited — growing the Hubble tension with a
0.7 km/s/Mpc constraint,” JCAP 11 (2022) 039, arXiv:2209.14330 [astro-ph.CO].
[1523] N. Schöneberg and L. Vacher, “The mass effect — variations of the electron mass and their impact on cosmology,”
JCAP 03 (2025) 004, arXiv:2407.16845 [astro-ph.CO].
[1524] B. D. Fields, K. A. Olive, T.-H. Yeh, and C. Young, “Big-Bang Nucleosynthesis after Planck,” JCAP 03 (2020) 010,
arXiv:1912.01132 [astro-ph.CO]. [Erratum: JCAP 11, E02 (2020)].
[1525] M. McQuinn, “The Evolution of the Intergalactic Medium,” Ann. Rev. Astron. Astrophys. 54 (2016) 313–362,
arXiv:1512.00086 [astro-ph.CO].
[1526] eBOSS Collaboration, H. du Mas des Bourboux et al., “The Completed SDSS-IV Extended Baryon Oscillation
Spectroscopic Survey: Baryon Acoustic Oscillations with Lyα Forests,” Astrophys. J. 901 (2020) no. 2, 153,
arXiv:2007.08995 [astro-ph.CO].
[1527] A. Cuceu, A. Font-Ribera, B. Joachimi, and S. Nadathur, “Cosmology beyond BAO from the 3D distribution of the
Lyman-α forest,” Mon. Not. Roy. Astron. Soc. 506 (2021) no. 4, 5439–5450, arXiv:2103.14075 [astro-ph.CO].
[1528] R. de Belsunce, O. H. E. Philcox, V. Irsic, P. McDonald, J. Guy, and N. Palanque-Delabrouille, “The 3D Lyman-α
forest power spectrum from eBOSS DR16,” Mon. Not. Roy. Astron. Soc. 533 (2024) no. 3, 3756–3770,
arXiv:2403.08241 [astro-ph.CO].
[1529] DESI Collaboration, C. Ravoux et al., “The Dark Energy Spectroscopic Instrument: one-dimensional power spectrum
from first Ly α forest samples with Fast Fourier Transform,” Mon. Not. Roy. Astron. Soc. 526 (2023) no. 4, 5118–5140,
arXiv:2306.06311 [astro-ph.CO].
[1530] N. G. Karaçaylı et al., “Optimal 1D Lyα Forest Power Spectrum Estimation – III. DESI early data,” Mon. Not. Roy.
Astron. Soc. 528 (2024) no. 3, 3941–3963, arXiv:2306.06316 [astro-ph.CO].
[1531] V. K. Narayanan, D. N. Spergel, R. Dave, and C.-P. Ma, “Constraints on the mass of warm dark matter particles and
the shape of the linear power spectrum from the Lyα forest,” Astrophys. J. Lett. 543 (2000) L103–L106,
arXiv:astro-ph/0005095.
[1532] U. Seljak, A. Slosar, and P. McDonald, “Cosmological parameters from combining the Lyman-alpha forest with CMB,
galaxy clustering and SN constraints,” JCAP 10 (2006) 014, arXiv:astro-ph/0604335.
[1533] M. Viel, J. Schaye, and C. M. Booth, “The impact of feedback from galaxy formation on the Lyman-alpha transmitted
flux,” Mon. Not. Roy. Astron. Soc. 429 (2013) 1734, arXiv:1207.6567 [astro-ph.CO].
[1534] V. Iršič et al., “The Lyman α forest power spectrum from the XQ-100 Legacy Survey,” Mon. Not. Roy. Astron. Soc.
466 (2017) no. 4, 4332–4345, arXiv:1702.01761 [astro-ph.CO].
[1535] E. Boera, G. D. Becker, J. S. Bolton, and F. Nasir, “Revealing Reionization with the Thermal History of the
Intergalactic Medium: New Constraints from the Lyα Flux Power Spectrum,” Astrophys. J. 872 (2019) no. 1, 101,
arXiv:1809.06980 [astro-ph.CO].
[1536] N. G. Karaçaylı et al., “Optimal 1D Ly α forest power spectrum estimation – II. KODIAQ, SQUAD, and XQ-100,”
Mon. Not. Roy. Astron. Soc. 509 (2022) no. 2, 2842–2855, arXiv:2108.10870 [astro-ph.CO].
[1537] B. Villasenor, B. Robertson, P. Madau, and E. Schneider, “New constraints on warm dark matter from the Lyman-α
forest power spectrum,” Phys. Rev. D 108 (2023) no. 2, 023502, arXiv:2209.14220 [astro-ph.CO].
[1538] E. Puchwein et al., “The Sherwood–Relics simulations: overview and impact of patchy reionization and pressure
smoothing on the intergalactic medium,” Mon. Not. Roy. Astron. Soc. 519 (2023) no. 4, 6162–6183, arXiv:2207.13098
[astro-ph.CO].
[1539] C. C. Doughty, J. F. Hennawi, F. B. Davies, Z. Lukić, and J. Oñorbe, “Convergence of small scale Ly α structure at
high-z under different reionization scenarios,” Mon. Not. Roy. Astron. Soc. 525 (2023) no. 3, 3790–3805,
arXiv:2305.16200 [astro-ph.CO].
[1540] S. Bird, M. Fernandez, M.-F. Ho, M. Qezlou, R. Monadi, Y. Ni, N. Chen, R. Croft, and T. Di Matteo, “PRIYA: a new
suite of Lyman-α forest simulations for cosmology,” JCAP 10 (2023) 037, arXiv:2306.05471 [astro-ph.CO].
[1541] SDSS Collaboration, P. McDonald et al., “The Linear theory power spectrum from the Lyman-alpha forest in the
Sloan Digital Sky Survey,” Astrophys. J. 635 (2005) 761–783, arXiv:astro-ph/0407377.
[1542] C. Pedersen, A. Font-Ribera, and N. Y. Gnedin, “Compressing the Cosmological Information in One-dimensional
Correlations of the Lyα Forest,” Astrophys. J. 944 (2023) no. 2, 223, arXiv:2209.09895 [astro-ph.CO].
[1543] M. Viel, J. Lesgourgues, M. G. Haehnelt, S. Matarrese, and A. Riotto, “Constraining warm dark matter candidates
including sterile neutrinos and light gravitinos with WMAP and the Lyman-alpha forest,” Phys. Rev. D 71 (2005)
063534, arXiv:astro-ph/0501562.
[1544] V. Iršič, M. Viel, M. G. Haehnelt, J. S. Bolton, and G. D. Becker, “First constraints on fuzzy dark matter from
Lyman-α forest data and hydrodynamical simulations,” Phys. Rev. Lett. 119 (2017) no. 3, 031302, arXiv:1703.04683
[astro-ph.CO].
[1545] E. Armengaud, N. Palanque-Delabrouille, C. Yèche, D. J. E. Marsh, and J. Baur, “Constraining the mass of light
bosonic dark matter using SDSS Lyman-α forest,” Mon. Not. Roy. Astron. Soc. 471 (2017) no. 4, 4606–4614,
arXiv:1703.09126 [astro-ph.CO].
311

[1546] N. Palanque-Delabrouille, C. Yèche, N. Schöneberg, J. Lesgourgues, M. Walther, S. Chabanier, and E. Armengaud,


“Hints, neutrino bounds and WDM constraints from SDSS DR14 Lyman-α and Planck full-survey data,” JCAP 04
(2020) 038, arXiv:1911.09073 [astro-ph.CO].
[1547] K. K. Rogers and H. V. Peiris, “General framework for cosmological dark matter bounds using N -body simulations,”
Phys. Rev. D 103 (2021) no. 4, 043526, arXiv:2007.13751 [astro-ph.CO].
[1548] V. Iršič et al., “Unveiling dark matter free streaming at the smallest scales with the high redshift Lyman-alpha forest,”
Phys. Rev. D 109 (2024) no. 4, 043511, arXiv:2309.04533 [astro-ph.CO].
[1549] eBOSS Collaboration, S. Chabanier et al., “The one-dimensional power spectrum from the SDSS DR14 Lyα forests,”
JCAP 07 (2019) 017, arXiv:1812.03554 [astro-ph.CO].
[1550] M. A. Fernandez, S. Bird, and M.-F. Ho, “Cosmological constraints from the eBOSS Lyman-α forest using the PRIYA
simulations,” JCAP 07 (2024) 029, arXiv:2309.03943 [astro-ph.CO].
[1551] V. Iršič et al., “New Constraints on the free-streaming of warm dark matter from intermediate and small scale
Lyman-α forest data,” Phys. Rev. D 96 (2017) no. 2, 023522, arXiv:1702.01764 [astro-ph.CO].
[1552] C. Yèche, N. Palanque-Delabrouille, J. Baur, and H. du Mas des Bourboux, “Constraints on neutrino masses from
Lyman-alpha forest power spectrum with BOSS and XQ-100,” JCAP 06 (2017) 047, arXiv:1702.03314
[astro-ph.CO].
[1553] M. Esposito, V. Iršič, M. Costanzi, S. Borgani, A. Saro, and M. Viel, “Weighing cosmic structures with clusters of
galaxies and the intergalactic medium,” Mon. Not. Roy. Astron. Soc. 515 (2022) no. 1, 857–870, arXiv:2202.00974
[astro-ph.CO].
[1554] S. Goldstein, J. C. Hill, V. Iršič, and B. D. Sherwin, “Canonical Hubble-Tension-Resolving Early Dark Energy
Cosmologies Are Inconsistent with the Lyman-α Forest,” Phys. Rev. Lett. 131 (2023) no. 20, 201001,
arXiv:2303.00746 [astro-ph.CO].
[1555] K. K. Rogers and V. Poulin, “5σ tension between Planck cosmic microwave background and eBOSS Lyman-alpha
forest and constraints on physics beyond ΛCDM,” Phys. Rev. Res. 7 (2025) no. 1, L012018, arXiv:2311.16377
[astro-ph.CO].
[1556] T. Kobayashi, R. Murgia, A. De Simone, V. Iršič, and M. Viel, “Lyman-α constraints on ultralight scalar dark matter:
Implications for the early and late universe,” Phys. Rev. D 96 (2017) no. 12, 123514, arXiv:1708.00015
[astro-ph.CO].
[1557] J. Baur, N. Palanque-Delabrouille, C. Yeche, A. Boyarsky, O. Ruchayskiy, E. Armengaud, and J. Lesgourgues,
“Constraints from Ly-α forests on non-thermal dark matter including resonantly-produced sterile neutrinos,” JCAP 12
(2017) 013, arXiv:1706.03118 [astro-ph.CO].
[1558] P. Gaikwad et al., “Probing the thermal state of the intergalactic medium at z > 5 with the transmission spikes in
high-resolution Ly α forest spectra,” Mon. Not. Roy. Astron. Soc. 494 (2020) no. 4, 5091–5109, arXiv:2001.10018
[astro-ph.CO].
[1559] P. Montero-Camacho, C. M. Hirata, P. Martini, and K. Honscheid, “Impact of inhomogeneous reionization on the
Lyman-α forest,” Mon. Not. Roy. Astron. Soc. 487 (2019) no. 1, 1047–1056, arXiv:1902.02892 [astro-ph.CO].
[1560] M. Molaro, V. Iršič, J. S. Bolton, M. Lieu, L. C. Keating, E. Puchwein, M. G. Haehnelt, and M. Viel, “Possible
evidence for a large-scale enhancement in the Lyman-α forest power spectrum at redshift z ≥ 4,” Mon. Not. Roy.
Astron. Soc. 521 (2023) no. 1, 1489–1501, arXiv:2303.05167 [astro-ph.CO].
[1561] J. Einasto, Dark Matter and Cosmic Web Story. World Scientific Publishing Co., 2014.
[1562] M. Einasto, J. Einasto, P. Tenjes, S. Korhonen, R. Kipper, E. Tempel, L. J. Liivamägi, and P. Heinämäki, “Galaxy
groups and clusters and their brightest galaxies within the cosmic web,” Astron. Astrophys. 681 (2024) A91,
arXiv:2311.01868 [astro-ph.CO].
[1563] S. Sankhyayan, J. Bagchi, E. Tempel, S. More, M. Einasto, P. Dabhade, S. Raychaudhury, R. Athreya, and
P. Heinämäki, “Identification of Superclusters and Their Properties in the Sloan Digital Sky Survey Using the WHL
Cluster Catalog,” Astrophys. J. 958 (2023) no. 1, 62, arXiv:2309.06251 [astro-ph.CO].
[1564] J. Einasto, I. Suhhonenko, L. J. Liivamägi, and M. Einasto, “Evolution of superclusters in the cosmic web,” Astron.
Astrophys. 623 (2019) A97, arXiv:1901.09378 [astro-ph.CO].
[1565] M. Einasto, J. Einasto, E. Tago, G. B. Dalton, and H. Andernach, “The structure of the universe traced by rich
clusters of galaxies.,” MNRAS 269 (1994) 301–322.
[1566] H. Lietzen, E. Tempel, L. J. Liivamägi, A. Montero-Dorta, M. Einasto, A. Streblyanska, C. Maraston, J. A. Rubiño
Martín, and E. Saar, “Discovery of a massive supercluster system at z ∼ 0.47,” Astron. Astrophys. 588 (2016) L4,
arXiv:1602.08498 [astro-ph.CO].
[1567] M. Einasto, R. Kipper, P. Tenjes, H. Lietzen, E. Tempel, L. J. Liivamägi, J. Einasto, A. Tamm, P. Heinämäki, and
P. Nurmi, “The Corona Borealis supercluster: connectivity, collapse, and evolution,” Astron. Astrophys. 649 (2021)
A51, arXiv:2103.02326 [astro-ph.CO].
[1568] N. Aghanim, T. Tuominen, V. Bonjean, C. Gouin, T. Bonnaire, and M. Einasto, “Dissecting a miniature universe: A
multi-wavelength view of galaxy quenching in the Shapley supercluster,” Astron. Astrophys. 689 (2024) A332,
arXiv:2402.18455 [astro-ph.CO].
[1569] L. J. Liivamagi, E. Tempel, and E. Saar, “SDSS DR7 superclusters. The catalogues,” Astron. Astrophys. 539 (2012)
A80, arXiv:1012.1989 [astro-ph.CO].
[1570] R. B. Tully, H. Courtois, Y. Hoffman, and D. Pomarède, “The Laniakea supercluster of galaxies,” Nature 513 (2014)
no. 7516, 71, arXiv:1409.0880 [astro-ph.CO].
[1571] A. Dupuy and H. M. Courtois, “Dynamic cosmography of the local Universe: Laniakea and five more watershed
312

superclusters,” Astron. Astrophys. 678 (2023) A176, arXiv:2305.02339 [astro-ph.CO].


[1572] M. Einasto et al., “The Sloan Great Wall. Morphology and galaxy content,” Astrophys. J. 736 (2011) 51,
arXiv:1105.1632 [astro-ph.CO].
[1573] J. M. Zúñiga, C. A. Caretta, and H. Andernach, “Nucleation regions in the Large-Scale Structure I: A catalogue of
cores in nearby rich superclusters,” Publ. Astron. Soc. Austral. 41 (2024) e078, arXiv:2405.13280 [astro-ph.CO].
[1574] S. Lim and J. Lee, “An Analytic Formula for the Supercluster Mass Function,” Astrophys. J. 783 (2014) 39,
arXiv:1201.1382 [astro-ph.CO].
[1575] S. Nadathur, S. Hotchkiss, and S. Sarkar, “The integrated Sachs-Wolfe imprints of cosmic superstructures: a problem
for ΛCDM,” JCAP 06 (2012) 042, arXiv:1109.4126 [astro-ph.CO].
[1576] B. R. Granett, M. C. Neyrinck, and I. Szapudi, “An Imprint of Super-Structures on the Microwave Background due to
the Integrated Sachs-Wolfe Effect,” Astrophys. J. Lett. 683 (2008) L99–L102, arXiv:0805.3695 [astro-ph].
[1577] I. D. Gialamas, G. Hütsi, K. Kannike, A. Racioppi, M. Raidal, M. Vasar, and H. Veermäe, “Interpreting DESI 2024
BAO: Late-time dynamical dark energy or a local effect?,” Phys. Rev. D 111 (2025) no. 4, 043540, arXiv:2406.07533
[astro-ph.CO].
[1578] M. Reyhani, M. Najafi, J. T. Firouzjaee, and E. Di Valentino, “Structure formation in various dynamical dark energy
scenarios,” Phys. Dark Univ. 44 (2024) 101477, arXiv:2403.15202 [astro-ph.CO].
[1579] R. K. Sachs and A. M. Wolfe, “Perturbations of a cosmological model and angular variations of the microwave
background,” Astrophys. J. 147 (1967) 73–90.
[1580] M. J. Rees and D. W. Sciama, “Large scale Density Inhomogeneities in the Universe,” Nature 217 (1968) 511–516.
[1581] S. Flender, S. Hotchkiss, and S. Nadathur, “The stacked ISW signal of rare superstructures in Λ CDM,” JCAP 02
(2013) 013, arXiv:1212.0776 [astro-ph.CO].
[1582] C. Hernandez-Monteagudo and R. E. Smith, “On the signature of nearby superclusters and voids in the Integrated
Sachs-Wolfe effect,” Mon. Not. Roy. Astron. Soc. 435 (2013) 1094, arXiv:1212.1174 [astro-ph.CO].
[1583] S. Aiola, A. Kosowsky, and B. Wang, “Gaussian Approximation of Peak Values in the Integrated Sachs-Wolfe Effect,”
Phys. Rev. D 91 (2015) 043510, arXiv:1410.6138 [astro-ph.CO].
[1584] A. M. Sołtan, “ISW in ΛCDM or something else?,” Mon. Not. Roy. Astron. Soc. 488 (2019) no. 2, 2732–2742,
arXiv:1812.09348 [astro-ph.CO].
[1585] Planck Collaboration, P. A. R. Ade et al., “Planck 2015 results. XXI. The integrated Sachs-Wolfe effect,” Astron.
Astrophys. 594 (2016) A21, arXiv:1502.01595 [astro-ph.CO].
[1586] Y.-C. Cai, M. Neyrinck, Q. Mao, J. A. Peacock, I. Szapudi, and A. A. Berlind, “The lensing and temperature imprints
of voids on the Cosmic Microwave Background,” Mon. Not. Roy. Astron. Soc. 466 (2017) no. 3, 3364–3375,
arXiv:1609.00301 [astro-ph.CO].
[1587] A. Kovács, “The part and the whole: voids, supervoids, and their ISW imprint,” Mon. Not. Roy. Astron. Soc. 475
(2018) no. 2, 1777–1790, arXiv:1701.08583 [astro-ph.CO].
[1588] S. Nadathur and R. Crittenden, “A detection of the integrated Sachs-Wolfe imprint of cosmic superstructures using a
matched-filter approach,” Astrophys. J. Lett. 830 (2016) no. 1, L19, arXiv:1608.08638 [astro-ph.CO].
[1589] DES Collaboration, A. Kovács et al., “Imprint of DES superstructures on the cosmic microwave background,” Mon.
Not. Roy. Astron. Soc. 465 (2017) no. 4, 4166–4179, arXiv:1610.00637 [astro-ph.CO].
[1590] DES Collaboration, A. Kovács et al., “More out of less: an excess integrated Sachs-Wolfe signal from supervoids
mapped out by the Dark Energy Survey,” Mon. Not. Roy. Astron. Soc. 484 (2019) 5267–5277, arXiv:1811.07812
[astro-ph.CO].
[1591] M. Cruz, E. Martinez-Gonzalez, P. Vielva, and L. Cayon, “Detection of a non-gaussian spot in wmap,” Mon. Not. Roy.
Astron. Soc. 356 (2005) 29–40, arXiv:astro-ph/0405341.
[1592] F. Finelli, J. García-Bellido, A. Kovács, F. Paci, and I. Szapudi, “Supervoids in the WISE–2MASS catalogue
imprinting cold spots in the cosmic microwave background,” Mon. Not. Roy. Astron. Soc. 455 (2016) no. 2, 1246–1256,
arXiv:1405.1555 [astro-ph.CO].
[1593] A. Pataki, P. Raffai, I. Csabai, G. Rácz, and I. Szapudi, “Constraints on AvERA Cosmologies from Cosmic
Chronometers and Type Ia Supernovae,” arXiv:2503.21369 [astro-ph.CO].
[1594] S. Räsänen, “Backreaction: directions of progress,” Class. Quant. Grav. 28 (2011) 164008, arXiv:1102.0408
[astro-ph.CO].
[1595] T. Buchert et al., “Is there proof that backreaction of inhomogeneities is irrelevant in cosmology?,” Class. Quant. Grav.
32 (2015) 215021, arXiv:1505.07800 [gr-qc].
[1596] D. L. Wiltshire, “Solution to the cosmological constant problem,” arXiv:2404.02129 [gr-qc].
[1597] M. J. Williams, H. J. Macpherson, D. L. Wiltshire, and C. Stevens, “First investigation of void statistics in numerical
relativity simulations,” Mon. Not. Roy. Astron. Soc. 536 (2025) 2645–2660, arXiv:2403.15134 [astro-ph.CO].
[1598] A. Kovács, R. Beck, A. Smith, G. Rácz, I. Csabai, and I. Szapudi, “Evidence for a high-z ISW signal from supervoids
in the distribution of eBOSS quasars,” Mon. Not. Roy. Astron. Soc. 513 (2022) no. 1, 15–26, arXiv:2107.13038
[astro-ph.CO].
[1599] eBOSS Collaboration, A. D. Myers et al., “The SDSS-IV extended Baryon Oscillation Spectroscopic Survey: Quasar
Target Selection,” Astrophys. J. Suppl. 221 (2015) no. 2, 27, arXiv:1508.04472 [astro-ph.CO].
[1600] J. R. Gott, III, M. Juric, D. Schlegel, F. Hoyle, M. Vogeley, M. Tegmark, N. A. Bahcall, and J. Brinkmann, “A map of
the universe,” Astrophys. J. 624 (2005) 463, arXiv:astro-ph/0310571.
[1601] M. Einasto, H. Lietzen, M. Gramann, E. Tempel, E. Saar, L. J. Liivamägi, P. Heinämäki, P. Nurmi, and J. Einasto,
“Sloan Great Wall as a complex of superclusters with collapsing cores,” Astron. Astrophys. 595 (2016) A70,
313

arXiv:1608.04988 [astro-ph.CO].
[1602] M. Einasto, P. Tenjes, M. Gramann, H. Lietzen, R. Kipper, L. J. Liivamägi, E. Tempel, S. Sankhyayan, and J. Einasto,
“The evolution of high-density cores of the BOSS Great Wall superclusters,” Astron. Astrophys. 666 (2022) A52,
arXiv:2204.08918 [astro-ph.CO].
[1603] J. Einasto et al., “Luminous superclusters: remnants from inflation,” Astron. Astrophys. 459 (2006) L1,
arXiv:astro-ph/0605393.
[1604] R. K. Sheth and A. Diaferio, “How unusual are the Shapley Supercluster and the Sloan Great Wall?,” Mon. Not. Roy.
Astron. Soc. 417 (2011) 2938–2949, arXiv:1105.3378 [astro-ph.CO].
[1605] C. Park, Y.-Y. Choi, J. Kim, J. R. Gott, III, S. S. Kim, and K.-S. Kim, “The Challenge of the Largest Structures in
the Universe to Cosmology,” Astrophys. J. Lett. 759 (2012) L7, arXiv:1209.5659 [astro-ph.CO].
[1606] J. Einasto and R. H. Miller, “Neighboring Superclusters and Their Environs,” in Early Evolution of the Universe and
its Present Structure, G. O. Abell and G. Chincarini, eds., vol. 104 of IAU Symposium, p. 405. 1983.
[1607] P. A. Shaver, “Radio Surveys and Large Scale Structure,” Australian Journal of Physics 44 (1991) no. 6, 759.
[1608] H. Boehringer, G. Chon, and J. Truemper, “The Cosmic Large-Scale Structure in X-rays (CLASSIX) Cluster Survey -
II. Unveiling a pancake structure with a 100 Mpc radius in the local Universe,” Astron. Astrophys. 651 (2021) A15,
arXiv:2105.13999 [astro-ph.CO].
[1609] P. J. E. Peebles, “Flat patterns in cosmic structure,” Mon. Not. Roy. Astron. Soc. 526 (2023) no. 3, 4490–4501,
arXiv:2308.04245 [astro-ph.CO].
[1610] M. Einasto, E. Tago, J. Jaaniste, J. Einasto, and H. Andernach, “The supercluster-void network I. the supercluster
catalogue and large scale distribution,” Astron. Astrophys. Suppl. Ser. 123 (1997) 119, arXiv:astro-ph/9610088.
[1611] K. Dolag, J. G. Sorce, S. Pilipenko, E. Hernández-Martínez, M. Valentini, S. Gottlöber, N. Aghanim, and
I. Khabibullin, “Simulating the LOcal Web (SLOW) - I. Anomalies in the local density field,” Astron. Astrophys. 677
(2023) A169, arXiv:2302.10960 [astro-ph.CO].
[1612] P. J. E. Peebles, “Improving Physical Cosmology: An Empiricist’s Assessment,” arXiv e-prints (2021)
arXiv:2106.02672, arXiv:2106.02672 [astro-ph.CO].
[1613] O. Toomet, H. Andernach, J. Einasto, M. Einasto, E. Kasak, A. A. Starobinsky, and E. Tago, “The supercluster-void
network v. alternative evidence for its regularity,” Astron. Astrophys. 393 (2002) 1, arXiv:astro-ph/9907238.
[1614] J. Einasto, M. Einasto, S. Gottlober, V. Mller, V. Saar, A. A. Starobinsky, E. Tago, D. Tucker, H. Andernach, and
P. Frisch, “A 120 MPC Periodicity in the Three-Dimensional Distribution of Galaxy Superclusters,” Nature 385 (1997)
139–141, arXiv:astro-ph/9701018.
[1615] J. Einasto, M. Einasto, P. Frisch, S. Gottlober, V. Mller, V. Saar, A. A. Starobinsky, E. Tago, D. Tucker, and
H. Andernach, “The supercluster-void network. 2. An oscillating cluster correlation function,” Mon. Not. Roy. Astron.
Soc. 289 (1997) 801–812, arXiv:astro-ph/9704127.
[1616] E. Tago, J. Einasto, M. Einasto, V. Muller, and H. Andernach, “Optical and x-ray clusters as tracers of the
supercluster-void network. II the spatial correlation function,” Astron. J. 123 (2002) 37, arXiv:astro-ph/0012537.
[1617] U. Lindner, J. Einasto, M. Einasto, W. Freudling, K. Fricke, and E. Tago, “The structure of supervoids - I: void
hierarchy in the northern local supervoid,” Astron. Astrophys. 301 (1995) 329, arXiv:astro-ph/9503044.
[1618] A. Kovács and J. García-Bellido, “Cosmic troublemakers: the Cold Spot, the Eridanus Supervoid, and the Great
Walls,” Mon. Not. Roy. Astron. Soc. 462 (2016) no. 2, 1882–1893, arXiv:1511.09008 [astro-ph.CO].
[1619] M. Einasto, E. Tago, H. Lietzen, C. Park, P. Heinamaki, E. Saar, H. Song, L. J. Liivamagi, and J. Einasto, “Tracing a
high redshift cosmic web with quasar systems,” Astron. Astrophys. 568 (2014) A46, arXiv:1406.5578 [astro-ph.CO].
[1620] M. Einasto, P. Heinämäki, L. J. Liivamägi, V. J. Martinez, L. Hurtado-Gil, P. Arnalte-Mur, P. Nurmi, J. Einasto, and
E. Saar, “Shell-like structures in our cosmic neighbourhood,” Astron. Astrophys. 587 (2016) A116, arXiv:1506.05295
[astro-ph.CO].
[1621] M. Einasto, L. J. Liivamagi, E. Tago, E. Saar, E. Tempel, J. Einasto, V. J. Martinez, and P. Heinamaki, “SDSS DR7
superclusters. Morphology,” Astron. Astrophys. 532 (2011) A5, arXiv:1105.2124 [astro-ph.CO].
[1622] R. B. Tully, C. Howlett, and D. Pomarede, “Ho’oleilana: An Individual Baryon Acoustic Oscillation?,” Astrophys. J.
954 (2023) no. 2, 169, arXiv:2309.00677 [astro-ph.CO].
[1623] P. Arnalte-Mur, A. Labatie, N. Clerc, V. J. Martinez, J. L. Starck, M. Lachieze-Rey, E. Saar, and S. Paredes, “Wavelet
analysis of baryon acoustic structures in the galaxy distribution,” Astron. Astrophys. 542 (2012) A34,
arXiv:1101.1911 [astro-ph.CO].
[1624] M. Einasto, M. Gramann, E. Saar, L. J. Liivamagi, E. Tempel, J. Nevalainen, P. Heinamaki, C. Park, and J. Einasto,
“Unusual A2142 supercluster with a collapsing core: distribution of light and mass,” Astron. Astrophys. 580 (2015)
A69, arXiv:1505.07233 [astro-ph.CO].
[1625] P. Heinämäki, P. Teerikorpi, M. Douspis, P. Nurmi, M. Einasto, M. Gramann, J. Nevalainen, and E. Saar,
“Quasi-spherical superclusters,” Astron. Astrophys. 668 (2022) A37, arXiv:2210.13294 [astro-ph.CO].
[1626] A. Vikhlinin, M. Markevitch, S. S. Murray, C. Jones, W. Forman, and L. Van Speybroeck, “Chandra temperature
profiles for a sample of nearby relaxed galaxy clusters,” Astrophys. J. 628 (2005) 655–672, arXiv:astro-ph/0412306.
[1627] E. Tempel, A. Tamm, M. Gramann, T. Tuvikene, L. J. Liivamägi, I. Suhhonenko, R. Kipper, M. Einasto, and E. Saar,
“Flux- and volume-limited groups/clusters for the SDSS galaxies: catalogues and mass estimation,” Astron. Astrophys.
566 (2014) A1, arXiv:1402.1350 [astro-ph.CO].
[1628] M. J. Rieke, D. Kelly, and S. Horner, “Overview of James Webb Space Telescope and NIRCam’s Role,” in Cryogenic
Optical Systems and Instruments XI, J. B. Heaney and L. G. Burriesci, eds., vol. 5904 of Society of Photo-Optical
Instrumentation Engineers (SPIE) Conference Series, pp. 1–8. 2005.
314

[1629] P. Santini, A. Fontana, M. Castellano, N. Leethochawalit, M. Trenti, T. Treu, D. Belfiori, S. Birrer, A. Bonchi,
E. Merlin, C. Mason, T. Morishita, M. Nonino, D. Paris, G. Polenta, P. Rosati, L. Yang, K. Boyett, M. Bradac,
A. Calabrò, A. Dressler, K. Glazebrook, D. Marchesini, S. Mascia, T. Nanayakkara, L. Pentericci, G. Roberts-Borsani,
C. Scarlata, B. Vulcani, and X. Wang, “Early Results from GLASS-JWST. XI. Stellar Masses and Mass-to-light Ratio
of z > 7 Galaxies,” ApJL 942 (2023) no. 2, L27, arXiv:2207.11379 [astro-ph.GA].
[1630] M. Castellano, A. Fontana, T. Treu, P. Santini, E. Merlin, N. Leethochawalit, M. Trenti, E. Vanzella, U. Mestric,
A. Bonchi, D. Belfiori, M. Nonino, D. Paris, G. Polenta, G. Roberts-Borsani, K. Boyett, M. Bradač, A. Calabrò,
K. Glazebrook, C. Grillo, S. Mascia, C. Mason, A. Mercurio, T. Morishita, T. Nanayakkara, L. Pentericci, P. Rosati,
B. Vulcani, X. Wang, and L. Yang, “Early Results from GLASS-JWST. III. Galaxy Candidates at z 9-15,” ApJL 938
(2022) no. 2, L15, arXiv:2207.09436 [astro-ph.GA].
[1631] S. L. Finkelstein, M. B. Bagley, H. C. Ferguson, S. M. Wilkins, J. S. Kartaltepe, et al., “CEERS Key Paper. I. An
Early Look into the First 500 Myr of Galaxy Formation with JWST,” Astrophys. J. Lett. 946 (2023) no. 1, L13,
arXiv:2211.05792 [astro-ph.GA].
[1632] R. P. Naidu, P. A. Oesch, P. van Dokkum, E. J. Nelson, K. A. Suess, G. Brammer, K. E. Whitaker, G. Illingworth,
R. Bouwens, S. Tacchella, J. Matthee, N. Allen, R. Bezanson, C. Conroy, I. Labbe, J. Leja, E. Leonova, D. Magee,
S. H. Price, D. J. Setton, V. Strait, M. Stefanon, S. Toft, J. R. Weaver, and A. Weibel, “Two Remarkably Luminous
Galaxy Candidates at z ≈ 10-12 Revealed by JWST,” The Astrophysical Journal Letters 940 (2022) no. 1, L14,
arXiv:2207.09434 [astro-ph.GA].
[1633] T. Treu et al., “The GLASS-JWST Early Release Science Program. I. Survey Design and Release Plans,” Astrophys. J.
935 (2022) no. 2, 110, arXiv:2206.07978 [astro-ph.GA].
[1634] Y. Harikane, M. Ouchi, M. Oguri, Y. Ono, K. Nakajima, Y. Isobe, H. Umeda, K. Mawatari, and Y. Zhang, “A
Comprehensive Study of Galaxies at z ∼ 9–16 Found in the Early JWST Data: Ultraviolet Luminosity Functions and
Cosmic Star Formation History at the Pre-reionization Epoch,” Astrophys. J. Suppl. 265 (2023) no. 1, ,
arXiv:2208.01612 [astro-ph.GA].
[1635] P. G. Pérez-González et al., “CEERS Key Paper. IV. A Triality in the Nature of HST-dark Galaxies,” ApJL 946
(2023) no. 1, L16, arXiv:2211.00045 [astro-ph.GA].
[1636] C. Papovich et al., “CEERS Key Paper. V. Galaxies at 4 < z < 9 Are Bluer than They Appear–Characterizing Galaxy
Stellar Populations from Rest-frame ∼1 µm Imaging,” The Astrophysical Journal Letters 949 (2023) no. 2, L18,
arXiv:2301.00027 [astro-ph.GA].
[1637] R. J. Bouwens, M. Stefanon, G. Brammer, P. A. Oesch, T. Herard-Demanche, G. D. Illingworth, J. Matthee, R. P.
Naidu, P. G. van Dokkum, and I. F. van Leeuwen, “Evolution of the UV LF from z 15 to z 8 using new JWST
NIRCam medium-band observations over the HUDF/XDF,” Mon. Not. Roy. Astron. Soc. 523 (2023) no. 1, 1036–1055,
arXiv:2211.02607 [astro-ph.GA].
[1638] N. J. Adams, C. J. Conselice, L. Ferreira, D. Austin, J. A. A. Trussler, I. Juodžbalis, S. M. Wilkins, J. Caruana,
P. Dayal, A. Verma, and A. P. Vijayan, “Discovery and properties of ultra-high redshift galaxies (9 < z < 12) in the
JWST ERO SMACS 0723 Field,” Mon. Not. Roy. Astron. Soc. 518 (2023) no. 3, 4755–4766, arXiv:2207.11217
[astro-ph.GA].
[1639] R. Bouwens, G. Illingworth, P. Oesch, M. Stefanon, R. Naidu, I. van Leeuwen, and D. Magee, “UV luminosity density
results at z > 8 from the first JWST/NIRCam fields: limitations of early data sets and the need for spectroscopy,”
Mon. Not. Roy. Astron. Soc. 523 (2023) no. 1, 1009–1035, arXiv:2212.06683 [astro-ph.CO].
[1640] S. L. Finkelstein et al., “The Complete CEERS Early Universe Galaxy Sample: A Surprisingly Slow Evolution of the
Space Density of Bright Galaxies at z ∼ 8.5–14.5,” The Astrophysical Journal Letters 969 (2024) no. 1, L2,
arXiv:2311.04279 [astro-ph.GA].
[1641] D. J. McLeod, C. T. Donnan, R. J. McLure, J. S. Dunlop, D. Magee, R. Begley, A. C. Carnall, F. Cullen, R. S. Ellis,
M. L. Hamadouche, and T. M. Stanton, “The galaxy UV luminosity function at z ≃ 11 from a suite of public JWST
ERS, ERO, and Cycle-1 programs,” Mon. Not. Roy. Astron. Soc. 527 (2024) no. 3, 5004–5022, arXiv:2304.14469
[astro-ph.GA].
[1642] L. Y. A. Yung, R. S. Somerville, S. L. Finkelstein, S. M. Wilkins, and J. P. Gardner, “Are the ultra-high-redshift
galaxies at z > 10 surprising in the context of standard galaxy formation models?,” Mon. Not. Roy. Astron. Soc. 527
(2024) no. 3, 5929–5948, arXiv:2304.04348 [astro-ph.GA].
[1643] Y.-Y. Wang, L. Lei, S.-P. Tang, G.-W. Yuan, and Y.-Z. Fan, “Digging into the Ultraviolet Luminosity Functions of
Galaxies at High Redshifts: Galaxies Evolution, Reionization, and Cosmological Parameters,” The Astrophysical
Journal 975 (2024) no. 2, 285, arXiv:2405.09350 [astro-ph.CO].
[1644] N. Sabti, J. B. Muñoz, and M. Kamionkowski, “Insights from HST into Ultramassive Galaxies and Early-Universe
Cosmology,” Phys. Rev. Lett. 132 (2024) no. 6, 061002, arXiv:2305.07049 [astro-ph.CO].
[1645] C. T. Donnan, D. J. McLeod, J. S. Dunlop, R. J. McLure, A. C. Carnall, R. Begley, F. Cullen, M. L. Hamadouche,
R. A. A. Bowler, D. Magee, H. J. McCracken, B. Milvang-Jensen, A. Moneti, and T. Targett, “The evolution of the
galaxy UV luminosity function at redshifts z ≃ 8 − 15 from deep JWST and ground-based near-infrared imaging,”
Mon. Not. Roy. Astron. Soc. 518 (2023) no. 4, 6011–6040, arXiv:2207.12356 [astro-ph.GA].
[1646] L. J. Furtak, M. Shuntov, H. Atek, A. Zitrin, J. Richard, M. D. Lehnert, and J. Chevallard, “Constraining the physical
properties of the first lensed z 9 - 16 galaxy candidates with JWST,” Mon. Not. Roy. Astron. Soc. 519 (2023) no. 2,
3064–3075, arXiv:2208.05473 [astro-ph.GA].
[1647] J. A. Zavala et al., “Dusty Starbursts Masquerading as Ultra-high Redshift Galaxies in JWST CEERS Observations,”
The Astrophysical Journal Letters 943 (2023) no. 2, L9, arXiv:2208.01816 [astro-ph.GA].
315

[1648] H. Yan, Z. Ma, C. Ling, C. Cheng, and J.-s. Huang, “First Batch of z ≈ 11–20 Candidate Objects Revealed by the
James Webb Space Telescope Early Release Observations on SMACS 0723-73,” Astrophys. J. Lett. 942 (2023) no. 1,
L9, arXiv:2207.11558 [astro-ph.GA].
[1649] I. Labbe et al., “A population of red candidate massive galaxies ~600 Myr after the Big Bang,” Nature 616 (2023)
no. 7956, 266–269, arXiv:2207.12446 [astro-ph.GA].
[1650] M. Xiao et al., “Accelerated formation of ultra-massive galaxies in the first billion years,” Nature (London) 635 (2024)
no. 8038, 311–315, arXiv:2309.02492 [astro-ph.GA].
[1651] C. M. Casey, H. B. Akins, M. Shuntov, O. Ilbert, L. Paquereau, M. Franco, C. C. Hayward, S. L. Finkelstein,
M. Boylan-Kolchin, B. E. Robertson, N. Allen, M. Brinch, O. R. Cooper, X. Ding, N. E. Drakos, A. L. Faisst,
S. Fujimoto, S. Gillman, S. Harish, M. Hirschmann, S. Jin, J. S. Kartaltepe, A. M. Koekemoer, V. Kokorev, D. Liu,
A. S. Long, G. Magdis, C. Maraston, C. L. Martin, H. J. McCracken, J. McKinney, B. Mobasher, J. Rhodes, R. M.
Rich, D. B. Sanders, J. D. Silverman, S. Toft, A. P. Vijayan, J. R. Weaver, S. M. Wilkins, L. Yang, and J. A. Zavala,
“COSMOS-Web: Intrinsically Luminous z ≳ 10 Galaxy Candidates Test Early Stellar Mass Assembly,” The
Astrophysical Journal 965 (2024) no. 1, 98, arXiv:2308.10932 [astro-ph.GA].
[1652] A. Weibel, P. A. Oesch, L. Barrufet, R. Gottumukkala, R. S. Ellis, P. Santini, J. R. Weaver, N. Allen, R. Bouwens,
R. A. A. Bowler, G. Brammer, A. C. Carnall, F. Cullen, P. Dayal, M. Dickinson, C. T. Donnan, J. S. Dunlop,
M. Giavalisco, N. A. Grogin, G. D. Illingworth, A. M. Koekemoer, I. Labbe, D. Marchesini, D. J. McLeod, R. J.
McLure, R. P. Naidu, P. G. Pérez-González, M. Shuntov, M. Stefanon, S. Toft, and M. Xiao, “Galaxy build-up in the
first 1.5 Gyr of cosmic history: insights from the stellar mass function at z 4-9 from JWST NIRCam observations,”
Mon. Not. Roy. Astron. Soc. 533 (2024) no. 2, 1808–1838, arXiv:2403.08872 [astro-ph.GA].
[1653] K. Chworowsky, S. L. Finkelstein, M. Boylan-Kolchin, E. J. McGrath, K. G. Iyer, C. Papovich, M. Dickinson, A. J.
Taylor, L. Y. A. Yung, P. Arrabal Haro, M. B. Bagley, B. E. Backhaus, R. Bhatawdekar, Y. Cheng, N. J. Cleri, J. W.
Cole, M. C. Cooper, L. Costantin, A. Dekel, M. Franco, S. Fujimoto, C. C. Hayward, B. W. Holwerda,
M. Huertas-Company, M. Hirschmann, T. A. Hutchison, A. M. Koekemoer, R. L. Larson, Z. Li, A. S. Long, R. A.
Lucas, N. Pirzkal, G. Rodighiero, R. S. Somerville, B. N. Vanderhoof, A. de la Vega, S. M. Wilkins, G. Yang, and J. A.
Zavala, “Evidence for a Shallow Evolution in the Volume Densities of Massive Galaxies at z = 4–8 from CEERS,” The
Astronomical Journal 168 (2024) no. 3, 113, arXiv:2311.14804 [astro-ph.GA].
[1654] C. C. Lovell, I. Harrison, Y. Harikane, S. Tacchella, and S. M. Wilkins, “Extreme value statistics of the halo and stellar
mass distributions at high redshift: are JWST results in tension with ΛCDM?,” Mon. Not. Roy. Astron. Soc. 518
(2022) no. 2, 2511–2520, arXiv:2208.10479 [astro-ph.GA].
[1655] M. Boylan-Kolchin, “Stress testing ΛCDM with high-redshift galaxy candidates,” Nature Astron. 7 (2023) no. 6,
731–735, arXiv:2208.01611 [astro-ph.CO].
[1656] M. Forconi, Ruchika, A. Melchiorri, O. Mena, and N. Menci, “Do the early galaxies observed by JWST disagree with
Planck’s CMB polarization measurements?,” JCAP 10 (2023) 012, arXiv:2306.07781 [astro-ph.CO].
[1657] J. Lu, L. Wang, X. Chen, D. Rubin, S. Perlmutter, D. Baade, J. Mould, J. Vinko, E. Regős, and A. M. Koekemoer,
“Constraints on Cosmological Parameters with a Sample of Type Ia Supernovae from JWST,” Astrophys. J. 941
(2022) no. 1, 71, arXiv:2210.00746 [astro-ph.CO].
[1658] J. D. R. Pierel et al., “Testing for Intrinsic Type Ia Supernova Luminosity Evolution at z > 2 with JWST,” Astrophys.
J. Lett. 981 (2025) no. 1, L9, arXiv:2411.11953 [astro-ph.CO].
[1659] J. Vinko and E. Regos, “SN 2023adsy – a normal Type Ia Supernova at z=2.9, discovered by JWST,”
arXiv:2411.10427 [astro-ph.HE].
[1660] D. A. Coulter et al., “Discovery of a likely Type II SN at z=3.6 with JWST,” arXiv:2501.05513 [astro-ph.HE].
[1661] T. J. Moriya et al., “Properties of high-redshift Type II supernovae discovered by the JADES transient survey,”
arXiv:2501.08969 [astro-ph.HE].
[1662] T. J. Moriya, Y. Harikane, and A. K. Inoue, “Constraint on the event rate of general relativistic instability supernovae
from the early JWST deep field data,” Mon. Not. Roy. Astron. Soc. 526 (2023) no. 2, 2400–2402, arXiv:2309.12049
[astro-ph.HE].
[1663] M. G. Dainotti et al., “A New Master Supernovae Ia sample and the investigation of the H0 tension,”
arXiv:2501.11772 [astro-ph.CO].
[1664] P. M. M. Alonso, C. Escamilla-Rivera, and R. Sandoval-Orozco, “Constraining dark energy cosmologies with spatial
curvature using Supernovae JWST forecasting,” JCAP 04 (2024) 084, arXiv:2309.12292 [astro-ph.CO].
[1665] C. C. Williams, S. Alberts, Z. Ji, K. N. Hainline, J. Lyu, G. Rieke, R. Endsley, K. A. Suess, F. Sun, B. D. Johnson,
M. Florian, I. Shivaei, W. Rujopakarn, W. M. Baker, R. Bhatawdekar, K. Boyett, A. J. Bunker, A. J. Cameron,
S. Carniani, S. Charlot, E. Curtis-Lake, C. DeCoursey, A. de Graaff, E. Egami, D. J. Eisenstein, J. L. Gibson,
R. Hausen, J. M. Helton, R. Maiolino, M. V. Maseda, E. J. Nelson, P. G. Pérez-González, M. J. Rieke, B. E.
Robertson, A. Saxena, S. Tacchella, C. N. A. Willmer, and C. J. Willott, “The Galaxies Missed by Hubble and ALMA:
The Contribution of Extremely Red Galaxies to the Cosmic Census at 3 < z < 8,” The Astrophysical Journal 968
(2024) no. 1, 34, arXiv:2311.07483 [astro-ph.GA].
[1666] C. L. Steinhardt, V. Kokorev, V. Rusakov, E. Garcia, and A. Sneppen, “Templates for Fitting Photometry of
Ultra-high-redshift Galaxies,” The Astrophysical Journal Letters 951 (2023) no. 2, L40, arXiv:2208.07879
[astro-ph.GA].
[1667] R. Endsley, D. P. Stark, L. Whitler, M. W. Topping, Z. Chen, A. Plat, J. Chisholm, and S. Charlot, “A
JWST/NIRCam study of key contributors to reionization: the star-forming and ionizing properties of UV-faint z 7-8
galaxies,” Mon. Not. Roy. Astron. Soc. 524 (2023) no. 2, 2312–2330, arXiv:2208.14999 [astro-ph.GA].
316

[1668] P. Arrabal Haro, M. Dickinson, S. L. Finkelstein, S. Fujimoto, V. Fernández, J. S. Kartaltepe, I. Jung, J. W. Cole,
D. Burgarella, K. Chworowsky, T. A. Hutchison, A. M. Morales, C. Papovich, R. C. Simons, R. O. Amorín, B. E.
Backhaus, M. B. Bagley, L. Bisigello, A. Calabrò, M. Castellano, N. J. Cleri, R. Davé, A. Dekel, H. C. Ferguson,
A. Fontana, E. Gawiser, M. Giavalisco, S. Harish, N. P. Hathi, M. Hirschmann, B. W. Holwerda,
M. Huertas-Company, A. M. Koekemoer, R. L. Larson, R. A. Lucas, B. Mobasher, P. G. Pérez-González, N. Pirzkal,
C. Rose, P. Santini, J. R. Trump, A. de la Vega, X. Wang, B. J. Weiner, S. M. Wilkins, G. Yang, L. Y. A. Yung, and
J. A. Zavala, “Spectroscopic Confirmation of CEERS NIRCam-selected Galaxies at ≃ 8 − 10,” The Astrophysical
Journal Letters 951 (2023) no. 1, L22, arXiv:2304.05378 [astro-ph.GA].
[1669] B. Wang, S. Fujimoto, I. Labbé, L. J. Furtak, T. B. Miller, D. J. Setton, A. Zitrin, H. Atek, R. Bezanson, G. Brammer,
J. Leja, P. A. Oesch, S. H. Price, I. Chemerynska, S. E. Cutler, P. Dayal, P. van Dokkum, A. D. Goulding, J. E.
Greene, Y. Fudamoto, G. Khullar, V. Kokorev, D. Marchesini, R. Pan, J. R. Weaver, K. E. Whitaker, and C. C.
Williams, “UNCOVER: Illuminating the Early Universe-JWST/NIRSpec Confirmation of z > 12 Galaxies,” The
Astrophysical Journal Letters 957 (2023) no. 2, L34, arXiv:2308.03745 [astro-ph.GA].
[1670] B. E. Robertson et al., “Identification and properties of intense star-forming galaxies at redshifts z > 10,” Nature
Astron. 7 (2023) no. 5, 611–621, arXiv:2212.04480 [astro-ph.GA].
[1671] E. Curtis-Lake, S. Carniani, A. Cameron, S. Charlot, P. Jakobsen, R. Maiolino, A. Bunker, J. Witstok, R. Smit,
J. Chevallard, C. Willott, P. Ferruit, S. Arribas, N. Bonaventura, M. Curti, F. D’Eugenio, M. Franx, G. Giardino, T. J.
Looser, N. Lützgendorf, M. V. Maseda, T. Rawle, H.-W. Rix, B. Rodríguez del Pino, H. Übler, M. Sirianni,
A. Dressler, E. Egami, D. J. Eisenstein, R. Endsley, K. Hainline, R. Hausen, B. D. Johnson, M. Rieke, B. Robertson,
I. Shivaei, D. P. Stark, S. Tacchella, C. C. Williams, C. N. A. Willmer, R. Bhatawdekar, R. Bowler, K. Boyett,
Z. Chen, A. de Graaff, J. M. Helton, R. E. Hviding, G. C. Jones, N. Kumari, J. Lyu, E. Nelson, M. Perna, L. Sandles,
A. Saxena, K. A. Suess, F. Sun, M. W. Topping, I. E. B. Wallace, and L. Whitler, “Spectroscopic confirmation of four
metal-poor galaxies at z = 10.3-13.2,” Nature Astronomy 7 (2023) 622–632, arXiv:2212.04568 [astro-ph.GA].
[1672] S. Fujimoto et al., “CEERS Spectroscopic Confirmation of NIRCam-selected z ≳ 8 Galaxy Candidates with
JWST/NIRSpec: Initial Characterization of Their Properties,” Astrophys. J. Lett. 949 (2023) no. 2, L25,
arXiv:2301.09482 [astro-ph.GA].
[1673] S. Carniani, K. Hainline, F. D’Eugenio, D. J. Eisenstein, P. Jakobsen, J. Witstok, B. D. Johnson, J. Chevallard,
R. Maiolino, J. M. Helton, C. Willott, B. Robertson, S. Alberts, S. Arribas, W. M. Baker, R. Bhatawdekar, K. Boyett,
A. J. Bunker, A. J. Cameron, P. A. Cargile, S. Charlot, M. Curti, E. Curtis-Lake, E. Egami, G. Giardino, K. Isaak,
Z. Ji, G. C. Jones, N. Kumari, M. V. Maseda, E. Parlanti, P. G. Pérez-González, T. Rawle, G. Rieke, M. Rieke, B. R.
Del Pino, A. Saxena, J. Scholtz, R. Smit, F. Sun, S. Tacchella, H. Übler, G. Venturi, C. C. Williams, and C. N. A.
Willmer, “Spectroscopic confirmation of two luminous galaxies at a redshift of 14,” Nature (London) 633 (2024)
no. 8029, 318–322, arXiv:2405.18485 [astro-ph.GA].
[1674] M. Castellano, L. Napolitano, A. Fontana, G. Roberts-Borsani, T. Treu, E. Vanzella, J. A. Zavala, P. Arrabal Haro,
A. Calabrò, M. Llerena, S. Mascia, E. Merlin, D. Paris, L. Pentericci, P. Santini, T. J. L. C. Bakx, P. Bergamini,
G. Cupani, M. Dickinson, A. V. Filippenko, K. Glazebrook, C. Grillo, P. L. Kelly, M. A. Malkan, C. A. Mason,
T. Morishita, T. Nanayakkara, P. Rosati, E. Sani, X. Wang, and I. Yoon, “JWST NIRSpec Spectroscopy of the
Remarkable Bright Galaxy GHZ2/GLASS-z12 at Redshift 12.34,” The Astrophysical Journal 972 (2024) no. 2, 143,
arXiv:2403.10238 [astro-ph.GA].
[1675] C. Giménez-Arteaga et al., “Spatially Resolved Properties of Galaxies at 5 < z < 9 in the SMACS 0723 JWST ERO
Field,” Astrophys. J. 948 (2023) no. 2, 126, arXiv:2212.08670 [astro-ph.GA].
[1676] A. Ferrara, “Super-early JWST galaxies, outflows, and Lyα visibility in the Epoch of Reionization,” Astron. Astrophys.
684 (2024) A207, arXiv:2310.12197 [astro-ph.GA].
[1677] C. A. Mason, M. Trenti, and T. Treu, “The brightest galaxies at cosmic dawn,” Mon. Not. Roy. Astron. Soc. 521
(2023) no. 1, 497–503, arXiv:2207.14808 [astro-ph.GA].
[1678] A. Kravtsov and V. Belokurov, “Stochastic star formation and the abundance of z > 10 UV-bright galaxies,”
arXiv:2405.04578 [astro-ph.GA].
[1679] S. Hegde, M. M. Wyatt, and S. R. Furlanetto, “A hidden population of active galactic nuclei can explain the
overabundance of luminous z > 10 objects observed by JWST,” JCAP 08 (2024) 025, arXiv:2405.01629
[astro-ph.GA].
[1680] V. Gelli, C. Mason, and C. C. Hayward, “The Impact of Mass-dependent Stochasticity at Cosmic Dawn,” The
Astrophysical Journal 975 (2024) no. 2, 192, arXiv:2405.13108 [astro-ph.GA].
[1681] G. Desprez et al., “ΛCDM not dead yet: massive high-z Balmer break galaxies are less common than previously
reported,” Mon. Not. Roy. Astron. Soc. 530 (2024) no. 3, 2935–2952, arXiv:2310.03063 [astro-ph.GA].
[1682] A. Dekel, K. C. Sarkar, Y. Birnboim, N. Mandelker, and Z. Li, “Efficient formation of massive galaxies at cosmic dawn
by feedback-free starbursts,” Mon. Not. Roy. Astron. Soc. 523 (2023) no. 3, 3201–3218, arXiv:2303.04827
[astro-ph.GA].
[1683] A. Ferrara, A. Pallottini, and P. Dayal, “On the stunning abundance of super-early, luminous galaxies revealed by
JWST,” Mon. Not. Roy. Astron. Soc. 522 (2023) no. 3, 3986–3991, arXiv:2208.00720 [astro-ph.GA].
[1684] B. Wang, J. Leja, H. Atek, I. Labbé, Y. Li, R. Bezanson, G. Brammer, S. E. Cutler, P. Dayal, L. J. Furtak, J. E.
Greene, V. Kokorev, R. Pan, S. H. Price, K. A. Suess, J. R. Weaver, K. E. Whitaker, and C. C. Williams, “Quantifying
the Effects of Known Unknowns on Inferred High-redshift Galaxy Properties: Burstiness, IMF, and Nebular Physics,”
The Astrophysical Journal 963 (2024) no. 1, 74, arXiv:2310.06781 [astro-ph.GA].
[1685] D. Narayanan, S. Lower, P. Torrey, G. Brammer, W. Cui, R. Davé, K. G. Iyer, Q. Li, C. C. Lovell, L. V. Sales, D. P.
317

Stark, F. Marinacci, and M. Vogelsberger, “Outshining by Recent Star Formation Prevents the Accurate Measurement
of High-z Galaxy Stellar Masses,” The Astrophysical Journal 961 (2024) no. 1, 73, arXiv:2306.10118 [astro-ph.GA].
[1686] R. K. Cochrane, H. Katz, R. Begley, C. C. Hayward, and P. N. Best, “High-z Stellar Masses Can Be Recovered
Robustly with JWST Photometry,” The Astrophysical Journall 978 (2025) no. 2, L42, arXiv:2412.02622
[astro-ph.GA].
[1687] M. Haslbauer, P. Kroupa, A. H. Zonoozi, and H. Haghi, “Has JWST Already Falsified Dark-matter-driven Galaxy
Formation?,” Astrophys. J. Lett. 939 (2022) no. 2, L31, arXiv:2210.14915 [astro-ph.GA].
[1688] E. M. Ventura, Y. Qin, S. Balu, and J. S. B. Wyithe, “Semi-analytic modelling of Pop. III star formation and
metallicity evolution – I. Impact on the UV luminosity functions at z = 9–16,” Mon. Not. Roy. Astron. Soc. 529
(2024) no. 1, 628–646, arXiv:2401.07396 [astro-ph.GA].
[1689] A. Trinca, R. Schneider, R. Valiante, L. Graziani, A. Ferrotti, K. Omukai, and S. Chon, “Exploring the nature of
UV-bright z ≳ 10 galaxies detected by JWST: star formation, black hole accretion, or a non-universal IMF?,” Mon.
Not. Roy. Astron. Soc. 529 (2024) no. 4, 3563–3581, arXiv:2305.04944 [astro-ph.GA].
[1690] E. E. Salpeter, “The Luminosity function and stellar evolution,” Astrophys. J. 121 (1955) 161–167.
[1691] E. R. Cueto, A. Hutter, P. Dayal, S. Gottlöber, K. E. Heintz, C. Mason, M. Trebitsch, and G. Yepes, “ASTRAEUS -
IX. Impact of an evolving stellar initial mass function on early galaxies and reionisation,” Astron. Astrophys. 686
(2024) A138, arXiv:2312.12109 [astro-ph.GA].
[1692] S. Tacchella, S. L. Finkelstein, M. Bagley, M. Dickinson, H. C. Ferguson, M. Giavalisco, L. Graziani, N. A. Grogin,
N. Hathi, T. A. Hutchison, I. Jung, A. M. Koekemoer, R. L. Larson, C. Papovich, N. Pirzkal, S. Rojas-Ruiz, M. Song,
R. Schneider, R. S. Somerville, S. M. Wilkins, and L. Y. A. Yung, “On the Stellar Populations of Galaxies at z = 9-11:
The Growth of Metals and Stellar Mass at Early Times,” The Astrophysical Journal 927 (2022) no. 2, 170,
arXiv:2111.05351 [astro-ph.GA].
[1693] L. Whitler, D. P. Stark, R. Endsley, J. Leja, S. Charlot, and J. Chevallard, “Star formation histories of UV-luminous
galaxies at z ≃ 6.8: implications for stellar mass assembly at early cosmic times,” Mon. Not. Roy. Astron. Soc. 519
(2023) no. 4, 5859–5881, arXiv:2206.05315 [astro-ph.GA].
[1694] F. Iocco and L. Visinelli, “Compatibility of JWST results with exotic halos,” Phys. Dark Univ. 44 (2024) 101496,
arXiv:2403.13068 [astro-ph.CO].
[1695] A. Pallottini and A. Ferrara, “Stochastic star formation in early galaxies: Implications for the James Webb Space
Telescope,” Astron. Astrophys. 677 (2023) L4, arXiv:2307.03219 [astro-ph.GA].
[1696] D. Ceverino, Y. Nakazato, N. Yoshida, R. S. Klessen, and S. C. O. Glover, “Redshift-dependent galaxy formation
efficiency at z = 5 - 13 in the FirstLight Simulations,” Astron. Astrophys. 689 (2024) A244, arXiv:2404.02537
[astro-ph.GA].
[1697] C. Turner, S. Tacchella, F. D’Eugenio, S. Carniani, M. Curti, K. Glazebrook, B. D. Johnson, S. Lim, T. Looser,
R. Maiolino, T. Nanayakkara, and J. Wan, “Age-dating early quiescent galaxies: high star formation efficiency, but
consistent with direct, higher-redshift observations,” Mon. Not. Roy. Astron. Soc. 537 (2025) no. 2, 1826–1848,
arXiv:2410.05377 [astro-ph.GA].
[1698] T. Harvey, C. J. Conselice, N. J. Adams, D. Austin, I. Juodžbalis, J. Trussler, Q. Li, K. Ormerod, L. Ferreira, C. C.
Lovell, Q. Duan, L. Westcott, H. Harris, R. Bhatawdekar, D. Coe, S. H. Cohen, J. Caruana, C. Cheng, S. P. Driver,
B. Frye, L. J. Furtak, N. A. Grogin, N. P. Hathi, B. W. Holwerda, R. A. Jansen, A. M. Koekemoer, M. A. Marshall,
M. Nonino, A. P. Vijayan, S. M. Wilkins, R. Windhorst, C. N. A. Willmer, H. Yan, and A. Zitrin, “EPOCHS. IV. SED
Modeling Assumptions and Their Impact on the Stellar Mass Function at 6.5 ≤ z ≤ 13.5 Using PEARLS and Public
JWST Observations,” The Astrophysical Journal 978 (2025) no. 1, 89, arXiv:2403.03908 [astro-ph.GA].
[1699] Y. Chen, H. J. Mo, and K. Wang, “Massive dark matter haloes at high redshift: implications for observations in the
JWST era,” Mon. Not. Roy. Astron. Soc. 526 (2023) no. 2, 2542–2559, arXiv:2304.13890 [astro-ph.GA].
[1700] J. Wang, Z. Huang, L. Huang, and J. Liu, “Quantifying the Tension between Cosmological Models and JWST Red
Candidate Massive Galaxies,” Res. Astron. Astrophys. 24 (2024) no. 4, 045001, arXiv:2311.02866 [astro-ph.CO].
[1701] Y. Qin, S. Balu, and J. S. B. Wyithe, “Implications of z ≳ 12 JWST galaxies for galaxy formation at high redshift,”
Mon. Not. Roy. Astron. Soc. 526 (2023) no. 1, 1324–1342, arXiv:2305.17959 [astro-ph.GA].
[1702] H. Atek, M. Shuntov, L. J. Furtak, J. Richard, J.-P. Kneib, G. Mahler, A. Zitrin, H. J. McCracken, S. Charlot,
J. Chevallard, and I. Chemerynska, “Revealing galaxy candidates out to z 16 with JWST observations of the lensing
cluster SMACS0723,” Mon. Not. Roy. Astron. Soc. 519 (2023) no. 1, 1201–1220, arXiv:2207.12338 [astro-ph.GA].
[1703] R. Navarro-Carrera, P. Rinaldi, K. I. Caputi, E. Iani, V. Kokorev, and S. E. van Mierlo, “Constraints on the Faint End
of the Galaxy Stellar Mass Function at z ≃ 4–8 from Deep JWST Data,” Astrophys. J. 961 (2024) no. 2, 207,
arXiv:2305.16141 [astro-ph.GA].
[1704] M. Maggiore and A. Riotto, “The Halo mass function from excursion set theory. II. The diffusing barrier,” Astrophys.
J. 717 (2010) 515–525, arXiv:0903.1250 [astro-ph.CO].
[1705] I. E. Achitouv and P. S. Corasaniti, “Primordial Bispectrum and Trispectrum Contributions to the Non-Gaussian
Excursion Set Halo Mass Function with Diffusive Drifting Barrier,” Phys. Rev. D 86 (2012) 083011, arXiv:1207.4796
[astro-ph.CO].
[1706] D. Reed, R. Bower, C. Frenk, A. Jenkins, and T. Theuns, “The halo mass function from the dark ages through the
present day,” Mon. Not. Roy. Astron. Soc. 374 (2007) 2–15, arXiv:astro-ph/0607150.
[1707] S. Basilakos, M. Plionis, and J. A. S. Lima, “Confronting Dark Energy Models using Galaxy Cluster Number Counts,”
Phys. Rev. D 82 (2010) 083517, arXiv:1006.3418 [astro-ph.CO].
[1708] S. Bhattacharya, K. Heitmann, M. White, Z. Lukic, C. Wagner, and S. Habib, “Mass Function Predictions Beyond
318

LCDM,” Astrophys. J. 732 (2011) 122, arXiv:1005.2239 [astro-ph.CO].


[1709] S. Tacchella, S. Bose, C. Conroy, D. J. Eisenstein, and B. D. Johnson, “A Redshift-independent Efficiency Model: Star
Formation and Stellar Masses in Dark Matter Halos at z ≳ 4,” Astrophys. J. 868 (2018) no. 2, 92, arXiv:1806.03299
[astro-ph.GA].
[1710] S. W. Allen, R. W. Schmidt, H. Ebeling, A. C. Fabian, and L. van Speybroeck, “Constraints on dark energy from
Chandra observations of the largest relaxed galaxy clusters,” Mon. Not. Roy. Astron. Soc. 353 (2004) 457,
arXiv:astro-ph/0405340.
[1711] A. Vikhlinin, A. Kravtsov, W. Forman, C. Jones, M. Markevitch, S. S. Murray, and L. Van Speybroeck, “Chandra
sample of nearby relaxed galaxy clusters: Mass, gas fraction, and mass-temperature relation,” Astrophys. J. 640 (2006)
691–709, arXiv:astro-ph/0507092.
[1712] A. V. Kravtsov, D. Nagai, and A. A. Vikhlinin, “Effects of cooling and star formation on the baryon fractions in
clusters,” Astrophys. J. 625 (2005) 588–598, arXiv:astro-ph/0501227.
[1713] S. Borgani and A. Kravtsov, “Cosmological simulations of galaxy clusters,” Adv. Sci. Lett. 4 (2011) 204,
arXiv:0906.4370 [astro-ph.CO].
[1714] A. B. Mantz, S. W. Allen, R. G. Morris, D. A. Rapetti, D. E. Applegate, P. L. Kelly, A. von der Linden, and R. W.
Schmidt, “Cosmology and astrophysics from relaxed galaxy clusters – II. Cosmological constraints,” Mon. Not. Roy.
Astron. Soc. 440 (2014) no. 3, 2077–2098, arXiv:1402.6212 [astro-ph.CO].
[1715] U. Maio and M. Viel, “The First Billion Years of a Warm Dark Matter Universe,” Mon. Not. Roy. Astron. Soc. 446
(2015) 2760–2775, arXiv:1409.6718 [astro-ph.CO].
[1716] K. Panchal and S. Desai, “Comparison of ΛCDM and Rh = ct with updated galaxy cluster fgas measurements using
Bayesian inference,” JHEAp 43 (2024) 15–19, arXiv:2401.11138 [astro-ph.CO].
[1717] H.-L. Huang, J.-Q. Jiang, and Y.-S. Piao, “High-redshift JWST massive galaxies and the initial clustering of
supermassive primordial black holes,” Phys. Rev. D 110 (2024) no. 10, 103540, arXiv:2407.15781 [astro-ph.CO].
[1718] G.-W. Yuan, L. Lei, Y.-Z. Wang, B. Wang, Y.-Y. Wang, C. Chen, Z.-Q. Shen, Y.-F. Cai, and Y.-Z. Fan, “Rapidly
growing primordial black holes as seeds of the massive high-redshift JWST Galaxies,” Sci. China Phys. Mech. Astron.
67 (2024) no. 10, 109512, arXiv:2303.09391 [astro-ph.CO].
[1719] B. Carr, S. Clesse, J. Garcia-Bellido, M. Hawkins, and F. Kuhnel, “Observational evidence for primordial black holes:
A positivist perspective,” Phys. Rept. 1054 (2024) 1–68, arXiv:2306.03903 [astro-ph.CO].
[1720] A. D. Dolgov, “Tension between HST/JWST and ΛCDM Cosmology, PBH, and Antimatter in the Galaxy,” in 14th
Frascati workshop on Multifrequency Behaviour of High Energy Cosmic Sources. 2023. arXiv:2310.00671
[astro-ph.CO].
[1721] A. Trinca, R. Schneider, R. Maiolino, R. Valiante, L. Graziani, and M. Volonteri, “Seeking the growth of the first black
hole seeds with JWST,” Mon. Not. Roy. Astron. Soc. 519 (2023) no. 3, 4753–4764, arXiv:2211.01389 [astro-ph.GA].
[1722] F. Pacucci, B. Nguyen, S. Carniani, R. Maiolino, and X. Fan, “JWST CEERS and JADES Active Galaxies at z = 4–7
Violate the Local M ⊙ –M ⊙ Relation at >3σ: Implications for Low-mass Black Holes and Seeding Models,” Astrophys.
J. Lett. 957 (2023) no. 1, L3, arXiv:2308.12331 [astro-ph.GA].
[1723] R. Maiolino et al., “A small and vigorous black hole in the early Universe,” Nature 627 (2024) no. 8002, 59–63,
arXiv:2305.12492 [astro-ph.GA]. [Erratum: Nature 630, E2 (2024)].
[1724] H. Mo, Y. Chen, and H. Wang, “A two-phase model of galaxy formation: I. The growth of galaxies and supermassive
black holes,” Mon. Not. Roy. Astron. Soc. 532 (2024) no. 4, 3808–3838, arXiv:2311.05030 [astro-ph.GA].
[1725] CEERS Team Collaboration, R. L. Larson et al., “A CEERS Discovery of an Accreting Supermassive Black Hole 570
Myr after the Big Bang: Identifying a Progenitor of Massive z > 6 Quasars,” Astrophys. J. Lett. 953 (2023) no. 2, L29,
arXiv:2303.08918 [astro-ph.GA].
[1726] A. Bogdan et al., “Evidence for heavy-seed origin of early supermassive black holes from a z ≈ 10 X-ray quasar,” Nature
Astron. 8 (2024) no. 1, 126–133, arXiv:2305.15458 [astro-ph.GA].
[1727] R. Schneider, R. Valiante, A. Trinca, L. Graziani, M. Volonteri, and R. Maiolino, “Are we surprised to find SMBHs
with JWST at z ≥ 9?,” Mon. Not. Roy. Astron. Soc. 526 (2023) no. 3, 3250–3261, arXiv:2305.12504 [astro-ph.GA].
[1728] H.-L. Huang, Y.-T. Wang, and Y.-S. Piao, “Supermassive primordial black holes for the GHZ9 and UHZ1 observed by
the JWST,” arXiv:2410.05891 [astro-ph.GA].
[1729] A. Trinca, R. Valiante, R. Schneider, I. Juodžbalis, R. Maiolino, L. Graziani, A. Lupi, P. Natarajan, M. Volonteri, and
T. Zana, “Episodic super-Eddington accretion as a clue to Overmassive Black Holes in the early Universe,”
arXiv:2412.14248 [astro-ph.GA].
[1730] S. Kumar and N. Weiner, “Early Galaxies from Rare Inflationary Processes and JWST Observations,”
arXiv:2502.08701 [astro-ph.CO].
[1731] P. Parashari and R. Laha, “Primordial power spectrum in light of JWST observations of high redshift galaxies,” Mon.
Not. Roy. Astron. Soc. 526 (2023) no. 1, L63–L69, arXiv:2305.00999 [astro-ph.CO].
[1732] S. Hirano and N. Yoshida, “Early Structure Formation from Primordial Density Fluctuations with a Blue, Tilted
Power Spectrum: High-redshift Galaxies,” Astrophys. J. 963 (2024) no. 1, 2, arXiv:2306.11993 [astro-ph.GA].
[1733] H. Padmanabhan and A. Loeb, “Alleviating the Need for Exponential Evolution of JWST Galaxies in 1010 M ⊙ Haloes
at z > 10 by a Modified ΛCDM Power Spectrum,” Astrophys. J. Lett. 953 (2023) no. 1, L4, arXiv:2306.04684
[astro-ph.CO].
[1734] M. Biagetti, G. Franciolini, and A. Riotto, “High-redshift JWST Observations and Primordial Non-Gaussianity,”
Astrophys. J. 944 (2023) no. 2, 113, arXiv:2210.04812 [astro-ph.CO].
[1735] M. De Laurentis and P. Salucci, “The Accurate Mass Distribution of M87, the Giant Galaxy with Imaged Shadow of
319

Its Supermassive Black Hole, as a Portal to New Physics,” Astrophys. J. 929 (2022) no. 1, 17, arXiv:2206.01997
[astro-ph.CO].
[1736] S.-Y. Guo, M. Khlopov, X. Liu, L. Wu, Y. Wu, and B. Zhu, “Footprints of axion-like particle in pulsar timing array
data and James Webb Space Telescope observations,” Sci. China Phys. Mech. Astron. 67 (2024) no. 11, 111011,
arXiv:2306.17022 [hep-ph].
[1737] G. Hütsi, M. Raidal, J. Urrutia, V. Vaskonen, and H. Veermäe, “Did JWST observe imprints of axion miniclusters or
primordial black holes?,” Phys. Rev. D 107 (2023) no. 4, 043502, arXiv:2211.02651 [astro-ph.CO].
[1738] S. Bird, C.-F. Chang, Y. Cui, and D. Yang, “Enhanced early galaxy formation in JWST from axion dark matter?,”
Phys. Lett. B 858 (2024) 139062, arXiv:2307.10302 [hep-ph].
[1739] Y. Gong, B. Yue, Y. Cao, and X. Chen, “Fuzzy Dark Matter as a Solution to Reconcile the Stellar Mass Density of
High-z Massive Galaxies and Reionization History,” Astrophys. J. 947 (2023) no. 1, 28, arXiv:2209.13757
[astro-ph.CO].
[1740] A. Zhitnitsky, “Structure formation paradigm and axion quark nugget dark matter model,” Phys. Dark Univ. 40
(2023) 101217, arXiv:2302.00010 [hep-ph].
[1741] Q. Zhang, S. Li, X.-H. Tan, and J.-Q. Xia, “Constraints on Primordial Magnetic Fields from High Redshift Stellar
Mass Density,” Astrophys. J. 972 (2024) no. 1, 117, arXiv:2408.03584 [astro-ph.CO].
[1742] H. Jiao, R. Brandenberger, and A. Refregier, “Early structure formation from cosmic string loops in light of early
JWST observations,” Phys. Rev. D 108 (2023) no. 4, 043510, arXiv:2304.06429 [astro-ph.CO].
[1743] H. Jiao, R. Brandenberger, and A. Refregier, “N-body simulation of early structure formation from cosmic string
loops,” Phys. Rev. D 109 (2024) no. 12, 123524, arXiv:2402.06235 [astro-ph.CO].
[1744] S. M. Koehler, H. Jiao, and R. Kannan, “Investigating cosmic strings using large-volume hydrodynamical simulations
in the context of JWST’s massive UV-bright galaxies,” arXiv:2412.00182 [astro-ph.CO].
[1745] P. Dayal and S. K. Giri, “Warm dark matter constraints from the JWST,” Mon. Not. Roy. Astron. Soc. 528 (2024)
no. 2, 2784–2789, arXiv:2303.14239 [astro-ph.CO].
[1746] U. Maio and M. Viel, “JWST high-redshift galaxy constraints on warm and cold dark matter models,” Astron.
Astrophys. 672 (2023) A71, arXiv:2211.03620 [astro-ph.CO].
[1747] G. Gandolfi, A. Lapi, T. Ronconi, and L. Danese, “Astroparticle Constraints from the Cosmic Star Formation Rate
Density at High Redshift: Current Status and Forecasts for JWST,” Universe 8 (2022) no. 11, 589, arXiv:2211.02840
[astro-ph.CO].
[1748] M. Forconi, W. Giarè, O. Mena, Ruchika, E. Di Valentino, A. Melchiorri, and R. C. Nunes, “A double take on early
and interacting dark energy from JWST,” JCAP 05 (2024) 097, arXiv:2312.11074 [astro-ph.CO].
[1749] A. Klypin, V. Poulin, F. Prada, J. Primack, M. Kamionkowski, V. Avila-Reese, A. Rodriguez-Puebla, P. Behroozi,
D. Hellinger, and T. L. Smith, “Clustering and Halo Abundances in Early Dark Energy Cosmological Models,” Mon.
Not. Roy. Astron. Soc. 504 (2021) no. 1, 769–781, arXiv:2006.14910 [astro-ph.CO].
[1750] X. Shen, M. Vogelsberger, M. Boylan-Kolchin, S. Tacchella, and R. P. Naidu, “Early galaxies and early dark energy: a
unified solution to the hubble tension and puzzles of massive bright galaxies revealed by JWST,” Mon. Not. Roy.
Astron. Soc. 533 (2024) no. 4, 3923–3936, arXiv:2406.15548 [astro-ph.GA].
[1751] J.-Q. Jiang, W. Liu, H. Zhan, and B. Hu, “Explanation of high redshift luminous galaxies from JWST by an early dark
energy model,” Phys. Rev. D 111 (2025) no. 2, 023519, arXiv:2409.19941 [astro-ph.CO].
[1752] N. Menci, M. Castellano, P. Santini, E. Merlin, A. Fontana, and F. Shankar, “High-redshift Galaxies from Early JWST
Observations: Constraints on Dark Energy Models,” Astrophys. J. Lett. 938 (2022) no. 1, L5, arXiv:2208.11471
[astro-ph.CO].
[1753] N. Menci, A. A. Sen, and M. Castellano, “The Excess of JWST Bright Galaxies: A Possible Origin in the Ground
State of Dynamical Dark Energy in the Light of DESI 2024 Data,” Astrophys. J. 976 (2024) no. 2, 227,
arXiv:2410.22940 [astro-ph.CO].
[1754] N. Menci, S. A. Adil, U. Mukhopadhyay, A. A. Sen, and S. Vagnozzi, “Negative cosmological constant in the dark
energy sector: tests from JWST photometric and spectroscopic observations of high-redshift galaxies,” JCAP 07
(2024) 072, arXiv:2401.12659 [astro-ph.CO].
[1755] S. A. Adil, U. Mukhopadhyay, A. A. Sen, and S. Vagnozzi, “Dark energy in light of the early JWST observations: case
for a negative cosmological constant?,” JCAP 10 (2023) 072, arXiv:2307.12763 [astro-ph.CO].
[1756] E. A. Paraskevas and L. Perivolaropoulos, “The density of virialized clusters as a probe of dark energy,” Mon. Not.
Roy. Astron. Soc. 531 (2024) no. 1, 1021–1033, arXiv:2308.07046 [astro-ph.CO].
[1757] P. Wang, B.-Y. Su, L. Zu, Y. Yang, and L. Feng, “Exploring the dark energy equation of state with JWST,” Eur. Phys.
J. Plus 139 (2024) no. 8, 711, arXiv:2307.11374 [astro-ph.CO].
[1758] M. Jõeveer, J. Einasto, and E. Tago, “Spatial distribution of galaxies and of clusters of galaxies in the southern galactic
hemisphere,” MNRAS 185 (1978) 357–370.
[1759] S. A. Gregory and L. A. Thompson, “The Coma/A1367 supercluster and its environs.,” Astrophys. J. 222 (1978)
784–799.
[1760] R. P. Kirshner, A. Oemler, Jr., P. L. Schechter, and S. A. Shectman, “A million cubic megaparsec void in Bootes,”
Astrophys. J. Lett. 248 (1981) L57–60.
[1761] V. de Lapparent, M. J. Geller, and J. P. Huchra, “A Slice of the universe,” Astrophys. J. Lett. 302 (1986) L1–L5.
[1762] Y. B. Zeldovich, J. Einasto, and S. F. Shandarin, “Giant Voids in the Universe,” Nature 300 (1982) 407–413.
[1763] P. J. E. Peebles, “The void phenomenon,” Astrophys. J. 557 (2001) 495–504, arXiv:astro-ph/0101127.
[1764] N. Padilla, D. Paz, M. Lares, L. Ceccarelli, D. G. Lambas, Y.-C. Cai, and B. Li, “Void Dynamics,” IAU Symp. 308
320

(2014) 530–537, arXiv:1410.8186 [astro-ph.CO].


[1765] A. Pisani et al., “Cosmic voids: a novel probe to shed light on our Universe,” arXiv:1903.05161 [astro-ph.CO].
[1766] N. Hamaus, B. D. Wandelt, P. M. Sutter, G. Lavaux, and M. S. Warren, “Cosmology with Void-Galaxy Correlations,”
Phys. Rev. Lett. 112 (2014) no. 4, 041304, arXiv:1307.2571 [astro-ph.CO].
[1767] K. C. Chan, N. Hamaus, and V. Desjacques, “Large-Scale Clustering of Cosmic Voids,” Phys. Rev. D 90 (2014) no. 10,
103521, arXiv:1409.3849 [astro-ph.CO].
[1768] G. Pollina, N. Hamaus, K. Dolag, J. Weller, M. Baldi, and L. Moscardini, “On the linearity of tracer bias around
voids,” Mon. Not. Roy. Astron. Soc. 469 (2017) no. 1, 787–799, arXiv:1610.06176 [astro-ph.CO].
[1769] S. Contarini, T. Ronconi, F. Marulli, L. Moscardini, A. Veropalumbo, and M. Baldi, “Cosmological exploitation of the
size function of cosmic voids identified in the distribution of biased tracers,” Mon. Not. Roy. Astron. Soc. 488 (2019)
no. 3, 3526–3540, arXiv:1904.01022 [astro-ph.CO].
[1770] J. Einasto, L. J. Liivamägi, I. Suhhonenko, and M. Einasto, “The biasing phenomenon,” Astron. Astrophys. 630 (2019)
A62, arXiv:1906.03617 [astro-ph.CO].
[1771] J. M. Colberg et al., “The Aspen–Amsterdam Void Finder Comparison Project,” Mon. Not. Roy. Astron. Soc. 387
(2008) 933, arXiv:0803.0918 [astro-ph].
[1772] M. C. Neyrinck, “ZOBOV: a parameter-free void-finding algorithm,” Mon. Not. Roy. Astron. Soc. 386 (2008) 2101,
arXiv:0712.3049 [astro-ph].
[1773] P. M. Sutter, G. Lavaux, N. Hamaus, A. Pisani, B. D. Wandelt, M. S. Warren, F. Villaescusa-Navarro, P. Zivick,
Q. Mao, and B. B. Thompson, “VIDE: The Void IDentification and Examination toolkit,” Astron. Comput. 9 (2015)
1–9, arXiv:1406.1191 [astro-ph.CO].
[1774] A. Elyiv, F. Marulli, G. Pollina, M. Baldi, E. Branchini, A. Cimatti, and L. Moscardini, “Cosmic voids detection
without density measurements,” Mon. Not. Roy. Astron. Soc. 448 (2015) no. 1, 642–653, arXiv:1410.4559
[astro-ph.CO].
[1775] D. J. Paz, C. M. Correa, S. R. Gualpa, A. N. Ruiz, C. S. Bederián, R. D. Graña, and N. D. Padilla, “Guess the cheese
flavour by the size of its holes: a cosmological test using the abundance of popcorn voids,” Mon. Not. Roy. Astron.
Soc. 522 (2023) no. 2, 2553–2569, arXiv:2212.06849 [astro-ph.CO].
[1776] A. N. Ruiz, D. J. Paz, M. Lares, H. E. Luparello, L. Ceccarelli, and D. Garcia Lambas, “Clues on void evolution – III.
Structure and dynamics in void shells,” Mon. Not. Roy. Astron. Soc. 448 (2015) no. 2, 1471–1482, arXiv:1501.02120
[astro-ph.CO].
[1777] D. G. Lambas, M. Lares, L. Ceccarelli, A. N. Ruiz, D. J. Paz, V. E. Maldonado, and H. E. Luparello, “The sparkling
Universe: the coherent motions of cosmic voids,” Mon. Not. Roy. Astron. Soc. 455 (2016) no. 1, L99–L103,
arXiv:1510.00712 [astro-ph.CO].
[1778] N. Schuster, N. Hamaus, K. Dolag, and J. Weller, “Why cosmic voids matter: mitigation of baryonic physics,” JCAP
08 (2024) 065, arXiv:2312.11241 [astro-ph.CO].
[1779] D. Paz, M. Lares, L. Ceccarelli, N. Padilla, and D. G. Lambas, “Clues on void evolution II: Measuring density and
velocity profiles on SDSS galaxy redshift space distortions,” Mon. Not. Roy. Astron. Soc. 436 (2013) 3480,
arXiv:1306.5799 [astro-ph.CO].
[1780] N. Schuster, N. Hamaus, K. Dolag, and J. Weller, “Why cosmic voids matter: nonlinear structure & linear dynamics,”
JCAP 05 (2023) 031, arXiv:2210.02457 [astro-ph.CO].
[1781] I. Szapudi et al., “The Cold Spot in the Cosmic Microwave Background: the Shadow of a Supervoid,” in 49th
Rencontres de Moriond on Cosmology, pp. 33–41. 2014. arXiv:1406.3622 [astro-ph.CO].
[1782] R. K. Sheth and R. van de Weygaert, “A Hierarchy of voids: Much ado about nothing,” Mon. Not. Roy. Astron. Soc.
350 (2004) 517, arXiv:astro-ph/0311260.
[1783] E. Jennings, Y. Li, and W. Hu, “The abundance of voids and the excursion set formalism,” Mon. Not. Roy. Astron.
Soc. 434 (2013) 2167, arXiv:1304.6087 [astro-ph.CO].
[1784] G. Verza, C. Carbone, A. Pisani, C. Porciani, and S. Matarrese, “The universal multiplicity function: counting haloes
and voids,” JCAP 10 (2024) 079, arXiv:2401.14451 [astro-ph.CO].
[1785] G. Lavaux and B. D. Wandelt, “Precision cosmography with stacked voids,” Astrophys. J. 754 (2012) 109,
arXiv:1110.0345 [astro-ph.CO].
[1786] N. Hamaus, A. Pisani, P. M. Sutter, G. Lavaux, S. Escoffier, B. D. Wandelt, and J. Weller, “Constraints on Cosmology
and Gravity from the Dynamics of Voids,” Phys. Rev. Lett. 117 (2016) no. 9, 091302, arXiv:1602.01784
[astro-ph.CO].
[1787] Y.-C. Cai, A. Taylor, J. A. Peacock, and N. Padilla, “Redshift-space distortions around voids,” Mon. Not. Roy. Astron.
Soc. 462 (2016) no. 3, 2465–2477, arXiv:1603.05184 [astro-ph.CO].
[1788] C. M. Correa, D. J. Paz, N. D. Padilla, A. N. Ruiz, R. E. Angulo, and A. G. Sánchez, “Non-fiducial cosmological test
from geometrical and dynamical distortions around voids,” Mon. Not. Roy. Astron. Soc. 485 (2019) no. 4, 5761–5772,
arXiv:1811.12251 [astro-ph.CO].
[1789] C. M. Correa, D. J. Paz, A. G. Sánchez, A. N. Ruiz, N. D. Padilla, and R. E. Angulo, “Redshift-space effects in voids
and their impact on cosmological tests. Part I: the void size function,” Mon. Not. Roy. Astron. Soc. 500 (2020) no. 1,
911–925, arXiv:2007.12064 [astro-ph.CO].
[1790] C. M. Correa, D. J. Paz, N. D. Padilla, A. G. Sánchez, A. N. Ruiz, and R. E. Angulo, “Redshift-space effects in voids
and their impact on cosmological tests – II. The void-galaxy cross-correlation function,” Mon. Not. Roy. Astron. Soc.
509 (2021) no. 2, 1871–1884, arXiv:2107.01314 [astro-ph.CO].
[1791] S. Contarini, A. Pisani, N. Hamaus, F. Marulli, L. Moscardini, and M. Baldi, “The perspective of voids on rising
321

cosmology tensions,” Astron. Astrophys. 682 (2024) A20, arXiv:2212.07438 [astro-ph.CO].


[1792] N. Hamaus, A. Pisani, J.-A. Choi, G. Lavaux, B. D. Wandelt, and J. Weller, “Precision cosmology with voids in the
final BOSS data,” JCAP 12 (2020) 023, arXiv:2007.07895 [astro-ph.CO].
[1793] Euclid Collaboration, N. Hamaus et al., “Euclid: Forecasts from redshift-space distortions and the Alcock–Paczynski
test with cosmic voids,” Astron. Astrophys. 658 (2022) A20, arXiv:2108.10347 [astro-ph.CO].
[1794] Euclid Collaboration, S. Contarini et al., “Euclid: Cosmological forecasts from the void size function,” Astron.
Astrophys. 667 (2022) A162, arXiv:2205.11525 [astro-ph.CO].
[1795] Euclid Collaboration, M. Bonici et al., “Euclid: Forecasts from the void-lensing cross-correlation,” Astron. Astrophys.
670 (2023) A47, arXiv:2206.14211 [astro-ph.CO].
[1796] S. Raghunathan, S. Nadathur, B. D. Sherwin, and N. Whitehorn, “The Gravitational Lensing Signatures of BOSS
Voids in the Cosmic Microwave Background,” Astrophys. J. 890 (2020) no. 2, 168, arXiv:1911.08475 [astro-ph.CO].
[1797] DES Collaboration, U. Demirbozan et al., “The gravitational lensing imprints of DES Y3 superstructures on the
CMB: a matched filtering approach,” Mon. Not. Roy. Astron. Soc. 534 (2024) no. 3, 2328–2343, arXiv:2404.18278
[astro-ph.CO].
[1798] G. Camacho-Ciurana, P. Lee, N. Arsenov, A. Kovács, I. Szapudi, and I. Csabai, “The cosmic microwave background
lensing imprint of cosmic voids detected in the WISE-Pan-STARRS luminous red galaxy catalog,” Astron. Astrophys.
689 (2024) A171, arXiv:2312.08483 [astro-ph.CO].
[1799] S. Sartori et al., “The imprint of cosmic voids from the DESI Legacy Survey DR9 LRGs in the Planck 2018 lensing
map through spectroscopically calibrated mocks,” arXiv:2412.02761 [astro-ph.CO].
[1800] D. R. Lorimer, M. Bailes, M. A. McLaughlin, D. J. Narkevic, and F. Crawford, “A bright millisecond radio burst of
extragalactic origin,” Science 318 (2007) 777, arXiv:0709.4301 [astro-ph].
[1801] S. P. Tendulkar et al., “The Host Galaxy and Redshift of the Repeating Fast Radio Burst FRB 121102,” Astrophys. J.
Lett. 834 (2017) no. 2, L7, arXiv:1701.01100 [astro-ph.HE].
[1802] A. Walters, A. Weltman, B. M. Gaensler, Y.-Z. Ma, and A. Witzemann, “Future Cosmological Constraints from Fast
Radio Bursts,” Astrophys. J. 856 (2018) no. 1, 65, arXiv:1711.11277 [astro-ph.CO].
[1803] W. Deng and B. Zhang, “Cosmological Implications of Fast Radio Burst/Gamma-Ray Burst Associations,” Astrophys.
J. Lett. 783 (2014) L35, arXiv:1401.0059 [astro-ph.HE].
[1804] A. Walters, Y.-Z. Ma, J. Sievers, and A. Weltman, “Probing Diffuse Gas with Fast Radio Bursts,” Phys. Rev. D 100
(2019) no. 10, 103519, arXiv:1909.02821 [astro-ph.CO].
[1805] J. P. Macquart et al., “A census of baryons in the Universe from localized fast radio bursts,” Nature 581 (2020)
no. 7809, 391–395, arXiv:2005.13161 [astro-ph.CO].
[1806] E. Platts, A. Weltman, A. Walters, S. P. Tendulkar, J. E. B. Gordin, and S. Kandhai, “A Living Theory Catalogue for
Fast Radio Bursts,” Phys. Rept. 821 (2019) 1–27, arXiv:1810.05836 [astro-ph.HE].
[1807] B. Zhang, “The Physical Mechanisms of Fast Radio Bursts,” Nature 587 (2020) 45–53, arXiv:2011.03500
[astro-ph.HE].
[1808] B. Zhang, “The physics of fast radio bursts,” Rev. Mod. Phys. 95 (2023) no. 3, 035005, arXiv:2212.03972
[astro-ph.HE].
[1809] S. Kalita and A. Weltman, “Continuous gravitational wave detection to understand the generation mechanism of fast
radio bursts,” Mon. Not. Roy. Astron. Soc. 520 (2023) no. 3, 3742–3748, arXiv:2211.00940 [astro-ph.HE].
[1810] H.E.S.S. Collaboration, J. O. Chibueze et al., “A MeerKAT, e-MERLIN, H.E.S.S., and Swift search for persistent and
transient emission associated with three localized FRBs,” Mon. Not. Roy. Astron. Soc. 515 (2022) no. 1, 1365–1379,
arXiv:2201.00069 [astro-ph.HE].
[1811] B. Zhou, X. Li, T. Wang, Y.-Z. Fan, and D.-M. Wei, “Fast radio bursts as a cosmic probe?,” Phys. Rev. D 89 (2014)
no. 10, 107303, arXiv:1401.2927 [astro-ph.CO].
[1812] H. Gao, Z. Li, and B. Zhang, “Fast Radio Burst/Gamma-Ray Burst Cosmography,” Astrophys. J. 788 (2014) 189,
arXiv:1402.2498 [astro-ph.CO].
[1813] Z. Pleunis et al., “LOFAR Detection of 110–188 MHz Emission and Frequency-dependent Activity from FRB
20180916B,” Astrophys. J. Lett. 911 (2021) no. 1, L3, arXiv:2012.08372 [astro-ph.HE].
[1814] V. Gajjar et al., “Highest Frequency Detection of FRB 121102 at 4–8 GHz Using the Breakthrough Listen Digital
Backend at the Green Bank Telescope,” Astrophys. J. 863 (2018) no. 1, 2, arXiv:1804.04101 [astro-ph.HE].
[1815] L. Bonetti, J. Ellis, N. E. Mavromatos, A. S. Sakharov, E. K. G. Sarkisyan-Grinbaum, and A. D. A. M. Spallicci,
“Photon Mass Limits from Fast Radio Bursts,” Phys. Lett. B 757 (2016) 548–552, arXiv:1602.09135 [astro-ph.HE].
[1816] H. Wang, X. Miao, and L. Shao, “Bounding the photon mass with cosmological propagation of fast radio bursts,” Phys.
Lett. B 820 (2021) 136596, arXiv:2103.15299 [astro-ph.HE].
[1817] H.-N. Lin, L. Tang, and R. Zou, “Revised constraints on the photon mass from well-localized fast radio bursts,” Mon.
Not. Roy. Astron. Soc. 520 (2023) no. 1, 1324–1331, arXiv:2301.12103 [gr-qc].
[1818] R. Reischke and S. Hagstotz, “Consistent constraints on the equivalence principle from localized fast radio bursts,”
Mon. Not. Roy. Astron. Soc. 523 (2023) no. 4, 6264–6271, arXiv:2302.10072 [astro-ph.CO].
[1819] S. Kalita, “Constraining fundamental constants with fast radio bursts: unveiling the role of energy scale,” Mon. Not.
Roy. Astron. Soc. 533 (2024) no. 1, L57–L63, arXiv:2407.01736 [gr-qc].
[1820] T. Lemos, R. Gonçalves, J. Carvalho, and J. Alcaniz, “A search for the fine-structure constant evolution from fast
radio bursts and type Ia supernovae data,” JCAP 01 (2025) 059, arXiv:2406.11691 [astro-ph.CO].
[1821] Z.-X. Li, H. Gao, X.-H. Ding, G.-J. Wang, and B. Zhang, “Strongly lensed repeating fast radio bursts as precision
probes of the universe,” Nature Commun. 9 (2018) no. 1, 3833, arXiv:1708.06357 [astro-ph.CO].
322

[1822] S. Hagstotz, R. Reischke, and R. Lilow, “A new measurement of the Hubble constant using fast radio bursts,” Mon.
Not. Roy. Astron. Soc. 511 (2022) no. 1, 662–667, arXiv:2104.04538 [astro-ph.CO].
[1823] Q. Wu, G.-Q. Zhang, and F.-Y. Wang, “An 8 per cent determination of the Hubble constant from localized fast radio
bursts,” Mon. Not. Roy. Astron. Soc. 515 (2022) no. 1, L1–L5, arXiv:2108.00581 [astro-ph.CO]. [Erratum:
Mon.Not.Roy.Astron.Soc. 531, L8 (2024)].
[1824] C. W. James et al., “A measurement of Hubble’s Constant using Fast Radio Bursts,” Mon. Not. Roy. Astron. Soc. 516
(2022) no. 4, 4862–4881, arXiv:2208.00819 [astro-ph.CO].
[1825] J. Baptista, J. X. Prochaska, A. G. Mannings, C. W. James, R. M. Shannon, S. D. Ryder, A. T. Deller, D. R. Scott,
M. Glowacki, and N. Tejos, “Measuring the Variance of the Macquart Relation in Redshift–Extragalactic Dispersion
Measure Modeling,” Astrophys. J. 965 (2024) no. 1, 57, arXiv:2305.07022 [astro-ph.CO].
[1826] Y. Liu, H. Yu, and P. Wu, “Cosmological-model-independent Determination of Hubble Constant from Fast Radio
Bursts and Hubble Parameter Measurements,” Astrophys. J. Lett. 946 (2023) no. 2, L49, arXiv:2210.05202
[astro-ph.CO].
[1827] J.-J. Wei and F. Melia, “Investigating Cosmological Models and the Hubble Tension Using Localized Fast Radio
Bursts,” Astrophys. J. 955 (2023) no. 2, 101, arXiv:2308.05918 [astro-ph.CO].
[1828] Z.-W. Zhao, J.-G. Zhang, Y. Li, J.-F. Zhang, and X. Zhang, “FRB dark sirens: Measuring the Hubble constant with
unlocalized fast radio bursts,” arXiv:2212.13433 [astro-ph.CO].
[1829] J. Gao, Z. Zhou, M. Du, R. Zou, J. Hu, and L. Xu, “A Measurement of Hubble Constant Using Cosmographic
Approach from Fast Radio Bursts and SNe Ia,” arXiv:2307.08285 [astro-ph.CO].
[1830] J. A. S. Fortunato, D. J. Bacon, W. S. Hipólito-Ricaldi, and D. Wands, “Fast Radio Bursts and Artificial Neural
Networks: a cosmological-model-independent estimation of the Hubble constant,” JCAP 01 (2025) 018,
arXiv:2407.03532 [astro-ph.CO].
[1831] S. Kalita, S. Bhatporia, and A. Weltman, “Fast Radio Bursts as probes of the late-time universe: a new insight on the
Hubble tension,” arXiv (2024) , arXiv:2410.01974 [astro-ph.CO].
[1832] E. F. Piratova-Moreno, L. A. García, C. A. Benavides-Gallego, and C. Cabrera, “Fast Radio Bursts as cosmological
proxies: estimating the Hubble constant,” arXiv:2502.08509 [astro-ph.CO].
[1833] J. B. Muñoz, E. D. Kovetz, L. Dai, and M. Kamionkowski, “Lensing of Fast Radio Bursts as a Probe of Compact Dark
Matter,” Phys. Rev. Lett. 117 (2016) no. 9, 091301, arXiv:1605.00008 [astro-ph.CO].
[1834] M. W. Sammons, J.-P. Macquart, R. D. Ekers, R. M. Shannon, H. Cho, J. X. Prochaska, A. T. Deller, and C. K. Day,
“First Constraints on Compact Dark Matter from Fast Radio Burst Microstructure,” Astrophys. J. 900 (2020) no. 2,
122, arXiv:2002.12533 [astro-ph.CO].
[1835] R. Laha, “Lensing of fast radio bursts: Future constraints on primordial black hole density with an extended mass
function and a new probe of exotic compact fermion and boson stars,” Phys. Rev. D 102 (2020) no. 2, 023016,
arXiv:1812.11810 [astro-ph.CO].
[1836] K. Liao, S. B. Zhang, Z. Li, and H. Gao, “Constraints on compact dark matter with fast radio burst observations,”
Astrophys. J. 896 (2020) no. 1, L11, arXiv:2003.13349 [astro-ph.CO].
[1837] C. Leung et al., “Constraining primordial black holes using fast radio burst gravitational-lens interferometry with
CHIME/FRB,” Phys. Rev. D 106 (2022) no. 4, 043017, arXiv:2204.06001 [astro-ph.HE].
[1838] S. Kalita, S. Bhatporia, and A. Weltman, “Gravitational lensing in modified gravity: a case study for Fast Radio
Bursts,” JCAP 11 (2023) 059, arXiv:2308.16604 [gr-qc].
[1839] D. Crichton et al., “Hydrogen Intensity and Real-Time Analysis Experiment: 256-element array status and overview,”
J. Astron. Telesc. Instrum. Syst. 8 (2022) 011019, arXiv:2109.13755 [astro-ph.IM].
[1840] L. Connor and V. Ravi, “Stellar prospects for FRB gravitational lensing,” Mon. Not. Roy. Astron. Soc. 521 (2023)
no. 3, 4024–4038, arXiv:2206.14310 [astro-ph.CO].
[1841] S. C. C. Ho, T. Hashimoto, T. Goto, Y.-W. Lin, S. J. Kim, Y. Uno, and T. Y. Y. Hsiao, “Future Constraints on Dark
Matter with Gravitationally Lensed Fast Radio Bursts Detected by BURSTT,” Astrophys. J. 950 (2023) no. 1, 53,
arXiv:2304.04990 [astro-ph.HE].
[1842] A. Weltman et al., “Fundamental physics with the Square Kilometre Array,” Publ. Astron. Soc. Austral. 37 (2020)
e002, arXiv:1810.02680 [astro-ph.CO].
[1843] M. G. Hauser and E. Dwek, “The cosmic infrared background: measurements and implications,” Ann. Rev. Astron.
Astrophys. 39 (2001) 249–307, arXiv:astro-ph/0105539.
[1844] R. C. Gilmore, P. Madau, J. R. Primack, R. S. Somerville, and F. Haardt, “GeV Gamma-Ray Attenuation and the
High-Redshift UV Background,” Mon. Not. Roy. Astron. Soc. 399 (2009) 1694, arXiv:0905.1144 [astro-ph.CO].
[1845] N. Cappelluti et al., “The Chandra COSMOS legacy survey: Energy Spectrum of the Cosmic X-ray Background and
constraints on undetected populations,” Astrophys. J. 837 (2017) no. 1, 19, arXiv:1702.01660 [astro-ph.HE].
[1846] Fermi-LAT Collaboration, M. Ackermann et al., “Resolving the Extragalactic γ-Ray Background above 50 GeV with
the Fermi Large Area Telescope,” Phys. Rev. Lett. 116 (2016) no. 15, 151105, arXiv:1511.00693 [astro-ph.CO].
[1847] J. Singal et al., “The Second Radio Synchrotron Background Workshop: Conference Summary and Report,” Publ.
Astron. Soc. Pac. 135 (2023) no. 1045, 036001, arXiv:2211.16547 [astro-ph.CO].
[1848] D. J. Fixsen et al., “ARCADE 2 Measurement of the Extra-Galactic Sky Temperature at 3-90 GHz,” Astrophys. J. 734
(2011) 5, arXiv:0901.0555 [astro-ph.CO].
[1849] J. Singal et al., “The ARCADE 2 Instrument,” Astrophys. J. 730 (2011) 138, arXiv:0901.0546 [astro-ph.IM].
[1850] J. Dowell and G. B. Taylor, “The Radio Background Below 100 MHz,” Astrophys. J. Lett. 858 (2018) no. 1, L9,
arXiv:1804.08581 [astro-ph.CO].
323

[1851] G. Jóhannesson and T. A. Porter, “Signatures of Recent Cosmic-Ray Acceleration in the High-latitude Gamma-Ray
Sky,” Astrophys. J. 917 (2021) no. 1, 30, arXiv:2104.13708 [astro-ph.HE].
[1852] C. G. T. Haslam, C. J. Salter, H. Stoffel, and W. E. Wilson, “A 408 MHz all-sky continuum survey. II. The atlas of
contour maps,” Astron. Astrophys. Suppl. Ser. 47 (1982) 1–142.
[1853] K. Maeda, H. Alvarez, J. Aparici, J. May, and P. Reich, “A 45-MHz continuum survey of the northern hemisphere,”
Astron. Astrophys. Suppl. Ser. 140 (1999) 145–154.
[1854] R. S. Roger, C. H. Costain, T. L. Landecker, and C. M. Swerdlyk, “The radio emission from the galaxy at 22 mhz,”
Astron. Astrophys. Suppl. Ser. 137 (1999) 7, arXiv:astro-ph/9902213.
[1855] J. Singal et al., “The Radio Synchrotron Background: Conference Summary and Report,” Publ. Astron. Soc. Pac. 130
(2018) no. 985, 036001, arXiv:1711.09979 [astro-ph.HE].
[1856] J. J. Condon, W. D. Cotton, E. B. Fomalont, K. I. Kellermann, N. Miller, R. A. Perley, D. Scott, T. Vernstrom, and
J. V. Wall, “Resolving the Radio Source Background: Deeper Understanding Through Confusion,” Astrophys. J. 758
(2012) 23, arXiv:1207.2439 [astro-ph.CO].
[1857] M. J. Hardcastle, T. W. Shimwell, C. Tasse, P. N. Best, A. Drabent, M. J. Jarvis, I. Prandoni, H. J. A. Röttgering,
J. Sabater, and D. J. Schwarz, “The contribution of discrete sources to the sky temperature at 144 MHz,” Astron.
Astrophys. 648 (2021) A10, arXiv:2011.08294 [astro-ph.CO].
[1858] C. L. Hale, I. H. Whittam, M. J. Jarvis, P. N. Best, N. L. Thomas, I. Heywood, M. Prescott, N. Adams, J. Afonso,
F. An, R. A. A. Bowler, J. D. Collier, R. H. W. Cook, R. Davé, B. S. Frank, M. Glowacki, P. W. Hatfield, S. Kolwa,
C. C. Lovell, N. Maddox, L. Marchetti, L. K. Morabito, E. Murphy, I. Prandoni, Z. Randriamanakoto, and A. R.
Taylor, “MIGHTEE: deep 1.4 GHz source counts and the sky temperature contribution of star-forming galaxies and
active galactic nuclei,” MNRAS 520 (2023) no. 2, 2668–2691, arXiv:2211.05741 [astro-ph.GA].
[1859] A. Kogut et al., “ARCADE 2 Observations of Galactic Radio Emission,” Astrophys. J. 734 (2011) 4, arXiv:0901.0562
[astro-ph.GA].
[1860] J. Singal, A. Kogut, E. Jones, and H. Dunlap, “Axial Ratio of Edge-on Spiral Galaxies as a Test for Bright Radio
Halos,” Astrophys. J. Lett. 799 (2015) no. 1, L10, arXiv:1501.00499 [astro-ph.GA].
[1861] M. G. H. Krause and M. J. Hardcastle, “Can the Local Bubble explain the radio background?,” Mon. Not. Roy.
Astron. Soc. 502 (2021) no. 2, 2807–2814, arXiv:2101.05255 [astro-ph.HE].
[1862] A. R. Offringa, J. Singal, S. Heston, S. Horiuchi, and D. M. Lucero, “Measurement of the anisotropy power spectrum of
the radio synchrotron background,” Mon. Not. Roy. Astron. Soc. 509 (2021) no. 1, 114–121, arXiv:2110.00499
[astro-ph.CO].
[1863] F. J. Cowie, A. R. Offringa, B. K. Gehlot, J. Singal, S. Heston, S. Horiuchi, and D. M. Lucero, “Diffuse sources,
clustering, and the excess anisotropy of the radio synchrotron background,” Mon. Not. Roy. Astron. Soc. 523 (2023)
no. 4, 5034–5046, arXiv:2306.00829 [astro-ph.CO].
[1864] P. P. Ponente, Y. Ascasibar, and J. M. Diego, “The contribution of star-forming galaxies to the cosmic radio
background,” Mon. Not. Roy. Astron. Soc. 418 (2011) 691, arXiv:1104.3012 [astro-ph.CO].
[1865] J. Singal, L. Stawarz, A. Lawrence, and V. Petrosian, “Sources of the Radio Background Considered,” Mon. Not. Roy.
Astron. Soc. 409 (2010) 1172, arXiv:0909.1997 [astro-ph.CO].
[1866] E. Todarello, M. Regis, F. Bianchini, J. Singal, E. Branchini, F. J. Cowie, S. Heston, S. Horiuchi, D. Lucero, and
A. Offringa, “Constraints on the origin of the radio synchrotron background via angular correlations,” Mon. Not. Roy.
Astron. Soc. 530 (2024) no. 3, 2994–3004, arXiv:2311.17641 [astro-ph.CO].
[1867] A. Fialkov and R. Barkana, “Signature of Excess Radio Background in the 21-cm Global Signal and Power Spectrum,”
Mon. Not. Roy. Astron. Soc. 486 (2019) no. 2, 1763–1773, arXiv:1902.02438 [astro-ph.CO].
[1868] P. L. Biermann, B. B. Nath, L. I. Caramete, B. C. Harms, T. Stanev, and J. Becker Tjus, “Cosmic backgrounds due to
the formation of the first generation of supermassive black holes,” Mon. Not. Roy. Astron. Soc. 441 (2014) no. 2,
1147–1156, arXiv:1403.3804 [astro-ph.CO].
[1869] K. Fang and T. Linden, “Cluster Mergers and the Origin of the ARCADE-2 Excess,” JCAP 10 (2016) 004,
arXiv:1506.05807 [astro-ph.HE].
[1870] N. Fornengo, R. Lineros, M. Regis, and M. Taoso, “Possibility of a Dark Matter Interpretation for the Excess in
Isotropic Radio Emission Reported by ARCADE,” Phys. Rev. Lett. 107 (2011) 271302, arXiv:1108.0569 [hep-ph].
[1871] D. Hooper, A. V. Belikov, T. E. Jeltema, T. Linden, S. Profumo, and T. R. Slatyer, “The Isotropic Radio Background
and Annihilating Dark Matter,” Phys. Rev. D 86 (2012) 103003, arXiv:1203.3547 [astro-ph.CO].
[1872] K. Fang and T. Linden, “Anisotropy of the extragalactic radio background from dark matter annihilation,” Phys. Rev.
D 91 (2015) no. 8, 083501, arXiv:1412.7545 [astro-ph.HE].
[1873] E. C. F. S. Fortes, O. D. Miranda, F. W. Stecker, and C. A. Wuensche, “Some Implications of the Leptonic
Annihilation of Dark Matter: Possible Galactic Radio Emission Signatures and the Excess Radio Flux of Extragalactic
Origin,” JCAP 11 (2019) 047, arXiv:1907.13184 [hep-ph].
[1874] Y. Yang, G. Yang, X. Huang, X. Chen, T. Lu, and H. Zong, “Contribution of ultracompact dark matter minihalos to
the isotropic radio background,” Phys. Rev. D 87 (2013) no. 8, 083519, arXiv:1206.3750 [astro-ph.HE].
[1875] D. Spolyar, P. Bodenheimer, K. Freese, and P. Gondolo, “Dark Stars: a new look at the First Stars in the Universe,”
Astrophys. J. 705 (2009) 1031–1042, arXiv:0903.3070 [astro-ph.CO].
[1876] T. Rindler-Daller, K. Freese, R. H. D. Townsend, and L. Visinelli, “Stability and pulsation of the first dark stars,”
Mon. Not. Roy. Astron. Soc. 503 (2021) no. 3, 3677–3691, arXiv:2011.00231 [astro-ph.CO].
[1877] K. Lawson and A. R. Zhitnitsky, “Isotropic Radio Background from Quark Nugget Dark Matter,” Phys. Lett. B 724
(2013) 17–21, arXiv:1210.2400 [astro-ph.CO].
324

[1878] N. Cappelluti, G. Hasinger, and P. Natarajan, “Exploring the High-redshift PBH-ΛCDM Universe: Early Black Hole
Seeding, the First Stars and Cosmic Radiation Backgrounds,” Astrophys. J. 926 (2022) no. 2, 205, arXiv:2109.08701
[astro-ph.CO].
[1879] S. Mittal and G. Kulkarni, “Background of radio photons from primordial black holes,” Mon. Not. Roy. Astron. Soc.
510 (2022) no. 4, 4992–4997, arXiv:2110.11975 [astro-ph.CO].
[1880] S. K. Acharya, J. Dhandha, and J. Chluba, “Can accreting primordial black holes explain the excess radio
background?,” Mon. Not. Roy. Astron. Soc. 517 (2022) no. 2, 2454–2461, arXiv:2208.03816 [astro-ph.CO].
[1881] M. Pospelov, J. Pradler, J. T. Ruderman, and A. Urbano, “Room for New Physics in the Rayleigh-Jeans Tail of the
Cosmic Microwave Background,” Phys. Rev. Lett. 121 (2018) no. 3, 031103, arXiv:1803.07048 [hep-ph].
[1882] A. Caputo, H. Liu, S. Mishra-Sharma, M. Pospelov, and J. T. Ruderman, “Radio excess from stimulated dark matter
decay,” Phys. Rev. D 107 (2023) no. 12, 123033, arXiv:2206.07713 [hep-ph].
[1883] S. K. Acharya and J. Chluba, “A closer look at dark photon explanations of the excess radio background,” Mon. Not.
Roy. Astron. Soc. 521 (2023) no. 3, 3939–3950, arXiv:2209.09063 [astro-ph.CO].
[1884] S. K. Acharya, B. Cyr, and J. Chluba, “Constraining broad photon spectrum injections from exotic and astrophysical
sources,” Mon. Not. Roy. Astron. Soc. 527 (2023) 2024, arXiv:2309.00975 [astro-ph.CO].
[1885] B. Cyr, J. Chluba, and S. K. Acharya, “Cosmic string solution to the radio synchrotron background,” Phys. Rev. D
109 (2024) no. 12, L121301, arXiv:2308.03512 [astro-ph.CO].
[1886] P. S. B. Dev, P. Di Bari, I. Martínez-Soler, and R. Roshan, “Relic neutrino decay solution to the excess radio
background,” JCAP 04 (2024) 046, arXiv:2312.03082 [hep-ph].
[1887] S. D. Bale, N. Bassett, J. O. Burns, J. Dorigo Jones, K. Goetz, C. Hellum-Bye, S. Hermann, J. Hibbard,
M. Maksimovic, R. McLean, R. Monsalve, P. O’Connor, A. Parsons, M. Pulupa, R. Pund, D. Rapetti, K. M.
Rotermund, B. Saliwanchik, A. Slosar, D. Sundkvist, and A. Suzuki, “LuSEE ’Night’: The Lunar Surface
Electromagnetics Experiment,” arXiv e-prints (2023) arXiv:2301.10345, arXiv:2301.10345 [astro-ph.IM].
[1888] E. Lee, J. Chluba, and G. P. Holder, “Refined modelling of the radio SZ signal: kinematic terms, relativistic
temperature corrections, and anisotropies in the radio background,” Mon. Not. Roy. Astron. Soc. 512 (2022) no. 4,
5153–5164, arXiv:2112.10666 [astro-ph.CO].
[1889] G. Holder and J. Chluba, “The radio SZ effect as a probe of the cosmological radio background,” arXiv:2110.08373
[astro-ph.CO].
[1890] R. Watkins, H. A. Feldman, and M. J. Hudson, “Consistently Large Cosmic Flows on Scales of 100 Mpc/h: a Challenge
for the Standard LCDM Cosmology,” Mon. Not. Roy. Astron. Soc. 392 (2009) 743–756, arXiv:0809.4041 [astro-ph].
[1891] H. A. Feldman, R. Watkins, and M. J. Hudson, “Cosmic Flows on 100 Mpc/h Scales: Standardized Minimum Variance
Bulk Flow, Shear and Octupole Moments,” Mon. Not. Roy. Astron. Soc. 407 (2010) 2328–2338, arXiv:0911.5516
[astro-ph.CO].
[1892] J. C. Bird, K. Z. Stanek, and J. L. Prieto, “Using Ultra Long Period Cepheids to Extend the Cosmic Distance Ladder
to 100 Mpc and Beyond,” Astrophys. J. 695 (2009) 874–882, arXiv:0807.4933 [astro-ph].
[1893] G. Fiorentino, R. C. Ramos, G. Clementini, M. Marconi, I. Musella, A. Aloisi, F. Annibali, A. Saha, M. Tosi, and R. P.
van der Marel, “Multi-Epoch HST Observations of IZw18: Characterization of Variable Stars at Ultra-Low
Metallicities,” Astrophys. J. 711 (2010) 808–817, arXiv:1001.4044 [astro-ph.SR].
[1894] M. Marconi, I. Musella, G. Fiorentino, G. Clementini, A. Aloisi, F. Annibali, R. C. Ramos, A. Saha, M. Tosi, and R. P.
van der Marel, “Pulsation Models for Ultra-Low (Z=0.0004) Metallicity Classical Cepheids,” Astrophys. J. 713 (2010)
615–625, arXiv:1002.4752 [astro-ph.SR].
[1895] I. Musella, M. Marconi, R. Molinaro, G. Fiorentino, V. Ripepi, G. De Somma, and M. I. Moretti, “New insights into
the use of Ultra Long Period Cepheids as cosmological standard candles,” MNRAS 501 (2021) no. 1, 866–874,
arXiv:2011.10533 [astro-ph.SR].
[1896] I. Musella, “Ultra Long Period Cepheids: Observation and Theory,” Universe 8 (2022) no. 6, 335.
[1897] C.-C. Ngeow, C.-H. Lee, M. T.-C. Yang, C.-S. Lin, H.-Y. Hsiao, Y.-C. Cheng, Z.-Y. Lin, I. L. Lin, S. M. Kanbur, and
W.-H. Ip, “VI-Band Follow-Up Observations of Ultra-Long-Period Cepheid Candidates in M31,” AJ 149 (2015) no. 2,
66, arXiv:1501.02456 [astro-ph.SR].
[1898] I. Musella, S. Leccia, R. Molinaro, M. Marconi, F. Cusano, M. Di Criscienzo, G. Fiorentino, V. Braga, V. Ripepi,
G. De Somma, M. Gatto, E. Luongo, and T. Sicignano, “Ultra-long-period Cepheids as Standard Candles from Gaia to
Rubin-LSST,” ApJS 275 (2024) no. 2, 26, arXiv:2410.12017 [astro-ph.GA].
[1899] DESI Collaboration, A. G. Adame et al., “The Early Data Release of the Dark Energy Spectroscopic Instrument,”
Astron. J. 168 (2024) no. 2, 58, arXiv:2306.06308 [astro-ph.CO].
[1900] DESI Collaboration, M. Abdul Karim et al., “Data Release 1 of the Dark Energy Spectroscopic Instrument,”
arXiv:2503.14745 [astro-ph.CO].
[1901] DESI Collaboration, W. Elbers et al., “Constraints on Neutrino Physics from DESI DR2 BAO and DR1 Full Shape,”
arXiv:2503.14744 [astro-ph.CO].
[1902] DES Collaboration, B. Flaugher et al., “The Dark Energy Camera,” Astron. J. 150 (2015) 150, arXiv:1504.02900
[astro-ph.IM].
[1903] Linea Science Server, DES Collaboration, T. M. C. Abbott et al., “The Dark Energy Survey Data Release 2,”
Astrophys. J. Supp. 255 (2021) no. 2, 20, arXiv:2101.05765 [astro-ph.IM].
[1904] S. Miyazaki et al., “Hyper Suprime-Cam: System design and verification of image quality,” Publications of the
Astronomical Society of Japan 70 (2018) S1.
[1905] H. Aihara et al., “First Data Release of the Hyper Suprime-Cam Subaru Strategic Program,” Publ. Astron. Soc. Jap.
325

70 (2018) S8, arXiv:1702.08449 [astro-ph.IM].


[1906] H. Aihara et al., “Third data release of the Hyper Suprime-Cam Subaru Strategic Program,” Publ. Astron. Soc. Jap.
74 (2022) no. 2, 247–272–272, arXiv:2108.13045 [astro-ph.IM].
[1907] K. Moriwaki, T. Nishimichi, and N. Yoshida, “Machine learning for observational cosmology,” Rept. Prog. Phys. 86
(2023) no. 7, 076901, arXiv:2303.15794 [astro-ph.IM].
[1908] A. Lewis, “Efficient sampling of fast and slow cosmological parameters,” Phys. Rev. D 87 (2013) no. 10, 103529,
arXiv:1304.4473 [astro-ph.CO].
[1909] B. Audren, J. Lesgourgues, K. Benabed, and S. Prunet, “Conservative Constraints on Early Cosmology: an illustration
of the Monte Python cosmological parameter inference code,” JCAP 02 (2013) 001, arXiv:1210.7183 [astro-ph.CO].
[1910] T. Brinckmann and J. Lesgourgues, “MontePython 3: boosted MCMC sampler and other features,” Phys. Dark Univ.
24 (2019) 100260, arXiv:1804.07261 [astro-ph.CO].
[1911] J. Torrado and A. Lewis, “Cobaya: Code for Bayesian Analysis of hierarchical physical models,” JCAP 05 (2021) 057,
arXiv:2005.05290 [astro-ph.IM].
[1912] D. Foreman-Mackey, D. W. Hogg, D. Lang, and J. Goodman, “emcee: The MCMC Hammer,” Publ. Astron. Soc. Pac.
125 (2013) 306–312, arXiv:1202.3665 [astro-ph.IM].
[1913] M. Karamanis, F. Beutler, J. A. Peacock, D. Nabergoj, and U. Seljak, “Accelerating astronomical and cosmological
inference with preconditioned Monte Carlo,” Mon. Not. Roy. Astron. Soc. 516 (2022) no. 2, 1644–1653,
arXiv:2207.05652 [astro-ph.IM].
[1914] G. Papamakarios, E. Nalisnick, D. J. Rezende, S. Mohamed, and B. Lakshminarayanan, “Normalizing flows for
probabilistic modeling and inference,” Journal of Machine Learning Research 22 (2021) no. 57, 1–64,
arXiv:1912.02762 [stat.ML]. http://jmlr.org/papers/v22/19-1028.html.
[1915] R. M. Neal, “Slice sampling,” The Annals of Statistics 31 (2003) no. 3, 705–767.
[1916] M. Karamanis, F. Beutler, and J. A. Peacock, “zeus: a python implementation of ensemble slice sampling for efficient
Bayesian parameter inference,” Mon. Not. Roy. Astron. Soc. 508 (2021) no. 3, 3589–3603, arXiv:2105.03468
[astro-ph.IM].
[1917] M. D. Hoffman and A. Gelman, “The no-u-turn sampler: Adaptively setting path lengths in hamiltonian monte carlo,”
Journal of Machine Learning Research 15 (2014) no. 47, 1593–1623, arXiv:1111.4246 [stat.CO].
http://jmlr.org/papers/v15/hoffman14a.html.
[1918] J. Bradbury, R. Frostig, P. Hawkins, M. J. Johnson, C. Leary, D. Maclaurin, G. Necula, A. Paszke, J. VanderPlas,
S. Wanderman-Milne, and Q. Zhang, “JAX: composable transformations of Python+NumPy programs.”
http://github.com/jax-ml/jax, 2018.
[1919] J.-E. Campagne, F. Lanusse, J. Zuntz, A. Boucaud, S. Casas, M. Karamanis, D. Kirkby, D. Lanzieri, Y. Li, and
A. Peel, “JAX-COSMO: An End-to-End Differentiable and GPU Accelerated Cosmology Library,” Open J. Astrophys.
6 (2023) 1–15, arXiv:2302.05163 [astro-ph.CO].
[1920] O. Hahn, F. List, and N. Porqueres, “DISCO-DJ I: a differentiable Einstein-Boltzmann solver for cosmology,” JCAP
06 (2024) 063, arXiv:2311.03291 [astro-ph.CO].
[1921] L. Balkenhol, C. Trendafilova, K. Benabed, and S. Galli, “candl: cosmic microwave background analysis with a
differentiable likelihood,” Astron. Astrophys. 686 (2024) A10, arXiv:2401.13433 [astro-ph.CO].
[1922] D. Piras and A. Spurio Mancini, “CosmoPower-JAX: high-dimensional Bayesian inference with differentiable
cosmological emulators,” arXiv:2305.06347 [astro-ph.CO].
[1923] J. Ruiz-Zapatero, D. Alonso, C. García-García, A. Nicola, A. Mootoovaloo, J. M. Sullivan, M. Bonici, and P. G.
Ferreira, “LimberJack.jl: auto-differentiable methods for angular power spectra analyses,” arXiv:2310.08306
[astro-ph.CO].
[1924] E. Bingham, J. P. Chen, M. Jankowiak, F. Obermeyer, N. Pradhan, T. Karaletsos, R. Singh, P. Szerlip, P. Horsfall,
and N. D. Goodman, “Pyro: Deep Universal Probabilistic Programming,” arXiv e-prints (2018) arXiv:1810.09538,
arXiv:1810.09538 [cs.LG].
[1925] D. Piras, A. Polanska, A. Spurio Mancini, M. A. Price, and J. D. McEwen, “The future of cosmological
likelihood-based inference: accelerated high-dimensional parameter estimation and model comparison,” Open J.
Astrophys. 7 (2024) , arXiv:2405.12965 [astro-ph.CO].
[1926] A. Mootoovaloo, J. Ruiz-Zapatero, C. García-García, and D. Alonso, “Assessment of gradient-based samplers in
standard cosmological likelihoods,” Mon. Not. Roy. Astron. Soc. 534 (2024) no. 3, 1668–1681, arXiv:2406.04725
[astro-ph.IM].
[1927] J. Skilling, “Nested sampling for general Bayesian computation,” Bayesian Analysis 1 (2006) no. 4, 833 – 859.
https://doi.org/10.1214/06-BA127.
[1928] J. Buchner, “Ultranest - a robust, general purpose bayesian inference engine,” Journal of Open Source Software 6
(2021) no. 60, 3001. https://doi.org/10.21105/joss.03001.
[1929] W. J. Handley, M. P. Hobson, and A. N. Lasenby, “PolyChord: nested sampling for cosmology,” Mon. Not. Roy.
Astron. Soc. 450 (2015) no. 1, L61–L65, arXiv:1502.01856 [astro-ph.CO].
[1930] J. S. Speagle, “dynesty: a dynamic nested sampling package for estimating Bayesian posteriors and evidences,” Mon.
Not. Roy. Astron. Soc. 493 (2020) no. 3, 3132–3158, arXiv:1904.02180 [astro-ph.IM].
[1931] J. U. Lange, “nautilus: boosting Bayesian importance nested sampling with deep learning,” Mon. Not. Roy. Astron.
Soc. 525 (2023) no. 2, 3181–3194, arXiv:2306.16923 [astro-ph.IM].
[1932] I. Gómez-Vargas and J. A. Vázquez, “Deep learning and genetic algorithms for cosmological Bayesian inference
speed-up,” Phys. Rev. D 110 (2024) no. 8, 083518, arXiv:2405.03293 [astro-ph.IM].
326

[1933] M. J. Williams, J. Veitch, and C. Messenger, “Importance nested sampling with normalising flows,” Mach. Learn. Sci.
Tech. 4 (2023) no. 3, 035011, arXiv:2302.08526 [astro-ph.IM].
[1934] J. Buchner, “Nested sampling methods,” Statistics Surveys (2021) , arXiv:2101.09675 [stat.CO].
[1935] E. O. Colgáin, S. Pourojaghi, M. M. Sheikh-Jabbari, and D. Sherwin, “A comparison of Bayesian and frequentist
confidence intervals in the presence of a late Universe degeneracy,” Eur. Phys. J. C 85 (2025) no. 2, 124,
arXiv:2307.16349 [astro-ph.CO].
[1936] J. Albert, C. Balazs, A. Fowlie, W. Handley, N. Hunt-Smith, R. R. de Austri, and M. White, “A comparison of
Bayesian sampling algorithms for high-dimensional particle physics and cosmology applications,” arXiv:2409.18464
[hep-ph].
[1937] D. Staicova, “Modern Bayesian Sampling Methods for Cosmological Inference: A Comparative Study,” Universe 11
(2025) 68, arXiv:2501.06022 [astro-ph.CO].
[1938] S. Raghvendra, P. Shirzadian, and K. Zhang, “A New Robust Partial p-Wasserstein-Based Metric for Comparing
Distributions,” arXiv e-prints (2024) arXiv:2405.03664, arXiv:2405.03664 [cs.LG].
[1939] A. Lewis, A. Challinor, and A. Lasenby, “Efficient computation of CMB anisotropies in closed FRW models,”
Astrophys. J. 538 (2000) 473–476, arXiv:astro-ph/9911177.
[1940] D. Blas, J. Lesgourgues, and T. Tram, “The Cosmic Linear Anisotropy Solving System (CLASS) II: Approximation
schemes,” JCAP 07 (2011) 034, arXiv:1104.2933 [astro-ph.CO].
[1941] B. Moser, C. S. Lorenz, U. Schmitt, A. Refregier, J. Fluri, R. Sgier, F. Tarsitano, and L. Heisenberg, “Symbolic
implementation of extensions of the PyCosmo Boltzmann solver,” Astron. Comput. 40 (2022) 100603,
arXiv:2112.08395 [astro-ph.CO].
[1942] J. Zuntz, M. Paterno, E. Jennings, D. Rudd, A. Manzotti, S. Dodelson, S. Bridle, S. Sehrish, and J. Kowalkowski,
“CosmoSIS: modular cosmological parameter estimation,” Astron. Comput. 12 (2015) 45–59, arXiv:1409.3409
[astro-ph.CO].
[1943] A. Lewis, “GetDist: a Python package for analysing Monte Carlo samples,” arXiv:1910.13970 [astro-ph.IM].
[1944] L. Senatore and M. Zaldarriaga, “The IR-resummed Effective Field Theory of Large Scale Structures,” JCAP 02
(2015) 013, arXiv:1404.5954 [astro-ph.CO].
[1945] L. Senatore, “Bias in the Effective Field Theory of Large Scale Structures,” JCAP 11 (2015) 007, arXiv:1406.7843
[astro-ph.CO].
[1946] L. Senatore and M. Zaldarriaga, “Redshift Space Distortions in the Effective Field Theory of Large Scale Structures,”
arXiv:1409.1225 [astro-ph.CO].
[1947] P. Carrilho, C. Moretti, and A. Pourtsidou, “Cosmology with the EFTofLSS and BOSS: dark energy constraints and a
note on priors,” JCAP 01 (2023) 028, arXiv:2207.14784 [astro-ph.CO].
[1948] T. Simon, P. Zhang, V. Poulin, and T. L. Smith, “Consistency of effective field theory analyses of the BOSS power
spectrum,” Phys. Rev. D 107 (2023) no. 12, 123530, arXiv:2208.05929 [astro-ph.CO].
[1949] M. Maus, S.-F. Chen, and M. White, “A comparison of template vs. direct model fitting for redshift-space distortions
in BOSS,” JCAP 06 (2023) 005, arXiv:2302.07430 [astro-ph.CO].
[1950] E. B. Holm, L. Herold, T. Simon, E. G. M. Ferreira, S. Hannestad, V. Poulin, and T. Tram, “Bayesian and frequentist
investigation of prior effects in EFT of LSS analyses of full-shape BOSS and eBOSS data,” Phys. Rev. D 108 (2023)
no. 12, 123514, arXiv:2309.04468 [astro-ph.CO].
[1951] J. Donald-McCann, R. Gsponer, R. Zhao, K. Koyama, and F. Beutler, “Analysis of unified galaxy power spectrum
multipole measurements,” Mon. Not. Roy. Astron. Soc. 526 (2023) no. 3, 3461–3481, arXiv:2307.07475
[astro-ph.CO].
[1952] M. M. Ivanov, A. Obuljen, C. Cuesta-Lazaro, and M. W. Toomey, “Full-shape analysis with simulation-based priors:
cosmological parameters and the structure growth anomaly,” arXiv:2409.10609 [astro-ph.CO].
[1953] A. Chudaykin, M. M. Ivanov, and T. Nishimichi, “On priors and scale cuts in EFT-based full-shape analyses,”
arXiv:2410.16358 [astro-ph.CO].
[1954] T. L. Smith and V. Poulin, “Current small-scale CMB constraints to axionlike early dark energy,” Phys. Rev. D 109
(2024) no. 10, 103506, arXiv:2309.03265 [astro-ph.CO].
[1955] R. Murgia, G. F. Abellán, and V. Poulin, “Early dark energy resolution to the Hubble tension in light of weak lensing
surveys and lensing anomalies,” Phys. Rev. D 103 (2021) no. 6, 063502, arXiv:2009.10733 [astro-ph.CO].
[1956] T. L. Smith, V. Poulin, J. L. Bernal, K. K. Boddy, M. Kamionkowski, and R. Murgia, “Early dark energy is not
excluded by current large-scale structure data,” Phys. Rev. D 103 (2021) no. 12, 123542, arXiv:2009.10740
[astro-ph.CO].
[1957] L. Herold, E. G. M. Ferreira, and E. Komatsu, “New Constraint on Early Dark Energy from Planck and BOSS Data
Using the Profile Likelihood,” Astrophys. J. Lett. 929 (2022) no. 1, L16, arXiv:2112.12140 [astro-ph.CO].
[1958] L. Herold and E. G. M. Ferreira, “Resolving the Hubble tension with early dark energy,” Phys. Rev. D 108 (2023)
no. 4, 043513, arXiv:2210.16296 [astro-ph.CO].
[1959] A. Reeves, L. Herold, S. Vagnozzi, B. D. Sherwin, and E. G. M. Ferreira, “Restoring cosmological concordance with
early dark energy and massive neutrinos?,” Mon. Not. Roy. Astron. Soc. 520 (2023) no. 3, 3688–3695,
arXiv:2207.01501 [astro-ph.CO].
[1960] G. Efstathiou, E. Rosenberg, and V. Poulin, “Improved Planck Constraints on Axionlike Early Dark Energy as a
Resolution of the Hubble Tension,” Phys. Rev. Lett. 132 (2024) no. 22, 221002, arXiv:2311.00524 [astro-ph.CO].
[1961] F. Niedermann and M. S. Sloth, “Resolving the Hubble tension with new early dark energy,” Phys. Rev. D 102 (2020)
no. 6, 063527, arXiv:2006.06686 [astro-ph.CO].
327

[1962] J. S. Cruz, S. Hannestad, E. B. Holm, F. Niedermann, M. S. Sloth, and T. Tram, “Profiling cold new early dark
energy,” Phys. Rev. D 108 (2023) no. 2, 023518, arXiv:2302.07934 [astro-ph.CO].
[1963] J. Hamann, S. Hannestad, G. G. Raffelt, and Y. Y. Y. Wong, “Observational bounds on the cosmic radiation density,”
JCAP 08 (2007) 021, arXiv:0705.0440 [astro-ph].
[1964] J. Hamann, “Evidence for extra radiation? Profile likelihood versus Bayesian posterior,” JCAP 03 (2012) 021,
arXiv:1110.4271 [astro-ph.CO].
[1965] S. Henrot-Versillé, F. Couchot, X. Garrido, H. Imada, T. Louis, M. Tristram, and S. Vanneste, “Comparison of results
on Neff from various Planck likelihoods,” Astron. Astrophys. 623 (2019) A9, arXiv:1807.05003 [astro-ph.CO].
[1966] A. Gómez-Valent, “Fast test to assess the impact of marginalization in Monte Carlo analyses and its application to
cosmology,” Phys. Rev. D 106 (2022) no. 6, 063506, arXiv:2203.16285 [astro-ph.CO].
[1967] E. B. Holm, L. Herold, S. Hannestad, A. Nygaard, and T. Tram, “Decaying dark matter with profile likelihoods,” Phys.
Rev. D 107 (2023) no. 2, L021303, arXiv:2211.01935 [astro-ph.CO].
[1968] A. R. Liddle, “Information criteria for astrophysical model selection,” Mon. Not. Roy. Astron. Soc. 377 (2007)
L74–L78, arXiv:astro-ph/0701113.
[1969] H. Jeffreys, The Theory of Probability. Oxford Classic Texts in the Physical Sciences. Oxford University Press, 1939.
[1970] D. J. Spiegelhalter, N. G. Best, B. P. Carlin, and A. van der Linde, “Bayesian measures of model complexity and fit,”
J. Roy. Statist. Soc. B 64 (2002) no. 4, 583–639.
[1971] R. Trotta, “Bayes in the sky: Bayesian inference and model selection in cosmology,” Contemp. Phys. 49 (2008) 71–104,
arXiv:0803.4089 [astro-ph].
[1972] P. Marshall, N. Rajguru, and A. Slosar, “Bayesian evidence as a tool for comparing datasets,” Physical Review D 73
(2006) no. 6, 067302, arXiv:astro-ph/0412535 [astro-ph].
[1973] W. Handley and P. Lemos, “Quantifying tensions in cosmological parameters: Interpreting the DES evidence ratio,”
Phys. Rev. D 100 (2019) no. 4, 043504, arXiv:1902.04029 [astro-ph.CO].
[1974] M. Raveri and W. Hu, “Concordance and discordance in cosmology,” Phys. Rev. D 99 (2019) 043506.
https://link.aps.org/doi/10.1103/PhysRevD.99.043506.
[1975] M. Raveri, G. Zacharegkas, and W. Hu, “Quantifying concordance of correlated cosmological data sets,” Phys. Rev. D
101 (2020) 103527. https://link.aps.org/doi/10.1103/PhysRevD.101.103527.
[1976] M. Raveri and C. Doux, “Non-gaussian estimates of tensions in cosmological parameters,” Phys. Rev. D 104 (2021)
043504. https://link.aps.org/doi/10.1103/PhysRevD.104.043504.
[1977] M. Leizerovich, S. J. Landau, and C. G. Scóccola, “Tensions in cosmology: A discussion of statistical tools to
determine inconsistencies,” Physics Letters B 855 (2024) 138844, arXiv:2312.08542 [astro-ph.CO].
[1978] P. Lemos, M. Raveri, A. Campos, et al., and DES Collaboration, “Assessing tension metrics with dark energy survey
and Planck data,” MNRAS 505 (2021) no. 4, 6179–6194, arXiv:2012.09554 [astro-ph.CO].
[1979] M. Ho, M. M. Rau, M. Ntampaka, A. Farahi, H. Trac, and B. Poczos, “A Robust and Efficient Deep Learning Method
for Dynamical Mass Measurements of Galaxy Clusters,” Astrophys. J. 887 (2019) 25, arXiv:1902.05950
[astro-ph.CO].
[1980] A. Peel, F. Lalande, J.-L. Starck, V. Pettorino, J. Merten, C. Giocoli, M. Meneghetti, and M. Baldi, “Distinguishing
standard and modified gravity cosmologies with machine learning,” Phys. Rev. D 100 (2019) no. 2, 023508,
arXiv:1810.11030 [astro-ph.CO].
[1981] J. a. Caldeira, W. L. K. Wu, B. Nord, C. Avestruz, S. Trivedi, and K. T. Story, “DeepCMB: Lensing Reconstruction of
the Cosmic Microwave Background with Deep Neural Networks,” Astron. Comput. 28 (2019) 100307,
arXiv:1810.01483 [astro-ph.CO].
[1982] S. He, Y. Li, Y. Feng, S. Ho, S. Ravanbakhsh, W. Chen, and B. Póczos, “Learning to Predict the Cosmological
Structure Formation,” Proc. Nat. Acad. Sci. 116 (2019) no. 28, 13825–13832, arXiv:1811.06533 [astro-ph.CO].
[1983] S. Ravanbakhsh, F. Lanusse, R. Mandelbaum, J. Schneider, and B. Poczos, “Enabling Dark Energy Science with Deep
Generative Models of Galaxy Images,” arXiv:1609.05796 [astro-ph.IM].
[1984] C. Escamilla-Rivera, M. A. C. Quintero, and S. Capozziello, “A deep learning approach to cosmological dark energy
models,” JCAP 03 (2020) 008, arXiv:1910.02788 [astro-ph.CO].
[1985] ANTARES Collaboration, G. Narayan et al., “Machine Learning-based Brokers for Real-time Classification of the
LSST Alert Stream,” Astrophys. J. Suppl. 236 (2018) no. 1, 9, arXiv:1801.07323 [astro-ph.IM].
[1986] F. Lanusse, Q. Ma, N. Li, T. E. Collett, C.-L. Li, S. Ravanbakhsh, R. Mandelbaum, and B. Poczos, “CMU DeepLens:
deep learning for automatic image-based galaxy–galaxy strong lens finding,” Mon. Not. Roy. Astron. Soc. 473 (2018)
no. 3, 3895–3906, arXiv:1703.02642 [astro-ph.IM].
[1987] G. Carleo, I. Cirac, K. Cranmer, L. Daudet, M. Schuld, N. Tishby, L. Vogt-Maranto, and L. Zdeborová, “Machine
learning and the physical sciences,” Rev. Mod. Phys. 91 (2019) no. 4, 045002, arXiv:1903.10563 [physics.comp-ph].
[1988] P. Mehta, M. Bukov, C.-H. Wang, A. G. R. Day, C. Richardson, C. K. Fisher, and D. J. Schwab, “A high-bias,
low-variance introduction to Machine Learning for physicists,” Phys. Rept. 810 (2019) 1–124, arXiv:1803.08823
[physics.comp-ph].
[1989] C. Aggarwal, Neural Networks and Deep Learning: A Textbook. Springer International Publishing, 2018.
https://books.google.com.mt/books?id=achqDwAAQBAJ.
[1990] Y.-C. Wang, Y.-B. Xie, T.-J. Zhang, H.-C. Huang, T. Zhang, and K. Liu, “Likelihood-free Cosmological Constraints
with Artificial Neural Networks: An Application on Hubble Parameters and SNe Ia,” Astrophys. J. Supp. 254 (2021)
no. 2, 43, arXiv:2005.10628 [astro-ph.CO].
[1991] I. Gómez-Vargas, R. M. Esquivel, R. García-Salcedo, and J. A. Vázquez, “Neural network reconstructions for the
328

Hubble parameter, growth rate and distance modulus,” Eur. Phys. J. C 83 (2023) no. 4, 304, arXiv:2104.00595
[astro-ph.CO].
[1992] G.-J. Wang, C. Cheng, Y.-Z. Ma, J.-Q. Xia, A. Abebe, and A. Beesham, “CoLFI: Cosmological Likelihood-free
Inference with Neural Density Estimators,” Astrophys. J. Suppl. 268 (2023) no. 1, 7, arXiv:2306.11102
[astro-ph.CO].
[1993] A. Manrique-Yus and E. Sellentin, “Euclid-era cosmology for everyone: neural net assisted MCMC sampling for the
joint 3 × 2 likelihood,” Mon. Not. Roy. Astron. Soc. 491 (2020) no. 2, 2655–2663, arXiv:1907.05881 [astro-ph.CO].
[1994] A. Spurio Mancini, D. Piras, J. Alsing, B. Joachimi, and M. P. Hobson, “CosmoPower: emulating cosmological power
spectra for accelerated Bayesian inference from next-generation surveys,” Mon. Not. Roy. Astron. Soc. 511 (2022)
no. 2, 1771–1788, arXiv:2106.03846 [astro-ph.CO].
[1995] J. Albers, C. Fidler, J. Lesgourgues, N. Schöneberg, and J. Torrado, “CosmicNet. Part I. Physics-driven implementation
of neural networks within Einstein-Boltzmann Solvers,” JCAP 09 (2019) 028, arXiv:1907.05764 [astro-ph.CO].
[1996] S. Günther, J. Lesgourgues, G. Samaras, N. Schöneberg, F. Stadtmann, C. Fidler, and J. Torrado, “CosmicNet II:
emulating extended cosmologies with efficient and accurate neural networks,” JCAP 11 (2022) 035, arXiv:2207.05707
[astro-ph.CO].
[1997] T. Auld, M. Bridges, M. P. Hobson, and S. F. Gull, “Fast cosmological parameter estimation using neural networks,”
Mon. Not. Roy. Astron. Soc. 376 (2007) L11–L15, arXiv:astro-ph/0608174.
[1998] D. Grandón and E. Sellentin, “Bayesian error propagation for neural-net based parameter inference,” The Open
Journal of Astrophysics (2022) , arXiv:2205.11587 [astro-ph.IM].
[1999] G.-J. Wang, S.-Y. Li, and J.-Q. Xia, “ECoPANN: A Framework for Estimating Cosmological Parameters using
Artificial Neural Networks,” Astrophys. J. Suppl. 249 (2020) no. 2, 25, arXiv:2005.07089 [astro-ph.CO].
[2000] J. Asorey, M. Crocce, E. Gaztanaga, and A. Lewis, “Recovering 3D clustering information with angular correlations,”
Mon. Not. Roy. Astron. Soc. 427 (2012) 1891, arXiv:1207.6487 [astro-ph.CO].
[2001] C. G. Sabiu, K. Kadota, J. Asorey, and I. Park, “Probing ultra-light axion dark matter from 21 cm tomography using
Convolutional Neural Networks,” JCAP 01 (2022) no. 01, 020, arXiv:2108.07972 [astro-ph.CO].
[2002] M. Andrés-Carcasona, M. Martinez, and L. M. Mir, “Fast Bayesian gravitational wave parameter estimation using
convolutional neural networks,” Mon. Not. Roy. Astron. Soc. 527 (2023) no. 2, 2887–2894, arXiv:2309.04303 [gr-qc].
[2003] E. Jones, T. Do, B. Boscoe, J. Singal, Y. Wan, and Z. Nguyen, “Improving Photometric Redshift Estimation for
Cosmology with LSST Using Bayesian Neural Networks,” Astrophys. J. 964 (2024) no. 2, 130, arXiv:2306.13179
[astro-ph.CO].
[2004] H. Aihara et al., “Second Data Release of the Hyper Suprime-Cam Subaru Strategic Program,” Publ. Astron. Soc. Jap.
71 (2019) no. 6, 114, arXiv:1905.12221 [astro-ph.IM].
[2005] M. Mancarella, J. Kennedy, B. Bose, and L. Lombriser, “Seeking new physics in cosmology with Bayesian neural
networks: Dark energy and modified gravity,” Phys. Rev. D 105 (2022) no. 2, 023531, arXiv:2012.03992
[astro-ph.CO].
[2006] L. Thummel, B. Bose, A. Pourtsidou, and L. Lombriser, “Classifying modified gravity and dark energy theories with
Bayesian neural networks: massive neutrinos, baryonic feedback, and the theoretical error,” Mon. Not. Roy. Astron.
Soc. 535 (2024) no. 4, 3141–3161, arXiv:2403.16949 [astro-ph.CO].
[2007] G. R. Dvali, G. Gabadadze, and M. Porrati, “4-D gravity on a brane in 5-D Minkowski space,” Phys. Lett. B 485
(2000) 208–214, arXiv:hep-th/0005016.
[2008] P. Mukherjee, J. Levi Said, and J. Mifsud, “Neural network reconstruction of H’(z) and its application in teleparallel
gravity,” JCAP 12 (2022) 029, arXiv:2209.01113 [astro-ph.CO].
[2009] G.-J. Wang, X.-J. Ma, S.-Y. Li, and J.-Q. Xia, “Reconstructing Functions and Estimating Parameters with Artificial
Neural Networks: A Test with a Hubble Parameter and SNe Ia,” Astrophys. J. Suppl. 246 (2020) no. 1, 13,
arXiv:1910.03636 [astro-ph.CO].
[2010] C. Escamilla-Rivera, M. Carvajal, C. Zamora, and M. Hendry, “Neural networks and standard cosmography with
newly calibrated high redshift GRB observations,” JCAP 04 (2022) no. 04, 016, arXiv:2109.00636 [astro-ph.CO].
[2011] J. d. D. R. Olvera, I. Gómez-Vargas, and J. A. Vázquez, “Observational Cosmology with Artificial Neural Networks,”
Universe 8 (2022) no. 2, 120, arXiv:2112.12645 [astro-ph.CO].
[2012] I. Gómez-Vargas, J. B. Andrade, and J. A. Vázquez, “Neural networks optimized by genetic algorithms in cosmology,”
Phys. Rev. D 107 (2023) no. 4, 043509, arXiv:2209.02685 [astro-ph.IM].
[2013] L. Giambagli, D. Fanelli, G. Risaliti, and M. Signorini, “Nonparametric analysis of the Hubble diagram with neural
networks,” Astron. Astrophys. 678 (2023) A13, arXiv:2302.12582 [astro-ph.CO].
[2014] K. F. Dialektopoulos, P. Mukherjee, J. Levi Said, and J. Mifsud, “Neural network reconstruction of cosmology using
the Pantheon compilation,” Eur. Phys. J. C 83 (2023) no. 10, 956, arXiv:2305.15499 [gr-qc].
[2015] K. F. Dialektopoulos, P. Mukherjee, J. Levi Said, and J. Mifsud, “Neural network reconstruction of scalar-tensor
cosmology,” Phys. Dark Univ. 43 (2024) 101383, arXiv:2305.15500 [gr-qc].
[2016] P. Mukherjee, K. F. Dialektopoulos, J. Levi Said, and J. Mifsud, “A possible late-time transition of M B inferred via
neural networks,” JCAP 09 (2024) 060, arXiv:2402.10502 [astro-ph.CO].
[2017] B. Zhang, H. Wang, X. Nong, G. Wang, P. Wu, and N. Liang, “Model-independent gamma-ray bursts constraints on
cosmological models using machine learning,” Astrophys. Space Sci. 370 (2025) no. 1, 10, arXiv:2312.09440
[astro-ph.CO].
[2018] J.-C. Zhang, Y. Hu, K. Jiao, H.-F. Wang, Y.-B. Xie, B. Yu, L.-L. Zhao, and T.-J. Zhang, “A Nonparametric
Reconstruction of the Hubble Parameter H(z) Based on Radial Basis Function Neural Networks,” Astrophys. J. Suppl.
329

270 (2024) no. 2, 23, arXiv:2311.13938 [astro-ph.CO].


[2019] H. Xie, X. Nong, H. Wang, B. Zhang, Z. Li, and N. Liang, “Constraints on cosmological models with gamma-ray bursts
in cosmology-independent way,” Int. J. Mod. Phys. D 34 (2025) no. 02, 2450073, arXiv:2307.16467 [astro-ph.CO].
[2020] L. Tang, X. Li, H.-N. Lin, and L. Liu, “Model-independently calibrating the luminosity correlations of gamma-ray
bursts using deep learning,” Astrophys. J. 907 (2021) no. 2, 121, arXiv:2011.14040 [astro-ph.CO].
[2021] L. Liu, L.-J. Hu, L. Tang, and Y. Wu, “Constraining the Spatial Curvature of the Local Universe with Deep Learning,”
Res. Astron. Astrophys. 23 (2023) no. 12, 125012, arXiv:2309.11334 [astro-ph.CO].
[2022] R. Shah, S. Saha, P. Mukherjee, U. Garain, and S. Pal, “LADDER: Revisiting the Cosmic Distance Ladder with Deep
Learning Approaches and Exploring Its Applications,” Astrophys. J. Suppl. 273 (2024) no. 2, 27, arXiv:2401.17029
[astro-ph.CO].
[2023] Pan-STARRS1 Collaboration, D. M. Scolnic et al., “The Complete Light-curve Sample of Spectroscopically
Confirmed SNe Ia from Pan-STARRS1 and Cosmological Constraints from the Combined Pantheon Sample,”
Astrophys. J. 859 (2018) no. 2, 101, arXiv:1710.00845 [astro-ph.CO].
[2024] R. Shah, P. Mukherjee, S. Saha, U. Garain, and S. Pal, “Deep Learning Based Recalibration of SDSS and DESI BAO
Alleviates Hubble and Clustering Tensions,” arXiv:2412.14750 [astro-ph.CO].
[2025] G. Papamakarios, E. Nalisnick, D. Jimenez Rezende, S. Mohamed, and B. Lakshminarayanan, “Normalizing Flows for
Probabilistic Modeling and Inference,” arXiv e-prints (2019) arXiv:1912.02762, arXiv:1912.02762 [stat.ML].
[2026] J. Sohl-Dickstein, E. A. Weiss, N. Maheswaranathan, and S. Ganguli, “Deep Unsupervised Learning using
Nonequilibrium Thermodynamics,” arXiv e-prints (2015) arXiv:1503.03585, arXiv:1503.03585 [cs.LG].
[2027] J. Song, C. Meng, and S. Ermon, “Denoising Diffusion Implicit Models,” arXiv e-prints (2020) arXiv:2010.02502,
arXiv:2010.02502 [cs.LG].
[2028] H. T. J. Bevins, W. J. Handley, P. Lemos, P. H. Sims, E. d. L. Acedo, A. Fialkov, and J. Alsing, “Marginal
post-processing of Bayesian inference products with normalizing flows and kernel density estimators,” Mon. Not. Roy.
Astron. Soc. 526 (2023) no. 3, 4613–4626, arXiv:2205.12841 [astro-ph.IM].
[2029] R. Friedman and S. Hassan, “HIGlow: Conditional Normalizing Flows for High-Fidelity HI Map Modeling,”
arXiv:2211.12724 [astro-ph.CO].
[2030] A. Mootoovaloo, C. García-García, D. Alonso, and J. Ruiz-Zapatero, “emuflow: normalizing flows for joint
cosmological analysis,” Mon. Not. Roy. Astron. Soc. 536 (2024) no. 1, 190–202, arXiv:2409.01407 [astro-ph.CO].
[2031] M. Raveri and C. Doux, “Non-Gaussian estimates of tensions in cosmological parameters,” Phys. Rev. D 104 (2021)
no. 4, 043504, arXiv:2105.03324 [astro-ph.CO].
[2032] B. Dai and U. Seljak, “Translation and rotation equivariant normalizing flow (TRENF) for optimal cosmological
analysis,” Mon. Not. Roy. Astron. Soc. 516 (2022) no. 2, 2363–2373, arXiv:2202.05282 [astro-ph.CO].
[2033] B. Dai and U. Seljak, “Multiscale Flow for robust and optimal cosmological analysis,” Proc. Nat. Acad. Sci. 121 (2024)
no. 9, e2309624121, arXiv:2306.04689 [astro-ph.CO].
[2034] N. Mudur, C. Cuesta-Lazaro, and D. P. Finkbeiner, “Cosmological Field Emulation and Parameter Inference with
Diffusion Models,” in 37th Conference on Neural Information Processing Systems. 2023. arXiv:2312.07534
[astro-ph.CO].
[2035] N. Mudur, C. Cuesta-Lazaro, and D. P. Finkbeiner, “Diffusion-HMC: Parameter Inference with Diffusion-model-driven
Hamiltonian Monte Carlo,” Astrophys. J. 978 (2025) no. 1, 64, arXiv:2405.05255 [astro-ph.CO].
[2036] R. C. Bernardo, D. Grandón, J. Said Levi, and V. H. Cárdenas, “Parametric and nonparametric methods hint dark
energy evolution,” Phys. Dark Univ. 36 (2022) 101017, arXiv:2111.08289 [astro-ph.CO].
[2037] M. Wang, X. Fu, B. Xu, Y. Yang, and Z. Chen, “Testing the FLRW metric with Hubble and transverse BAO
measurements,” Phys. Rev. D 108 (2023) no. 10, 103506, arXiv:2305.01268 [gr-qc].
[2038] N. Rani, D. Jain, S. Mahajan, A. Mukherjee, and N. Pires, “Transition Redshift: New constraints from parametric and
nonparametric methods,” JCAP 12 (2015) 045, arXiv:1503.08543 [gr-qc].
[2039] H. K. Jassal, J. S. Bagla, and T. Padmanabhan, “WMAP constraints on low redshift evolution of dark energy,” Mon.
Not. Roy. Astron. Soc. 356 (2005) L11–L16, arXiv:astro-ph/0404378.
[2040] B. Wang, E. Abdalla, F. Atrio-Barandela, and D. Pavon, “Dark Matter and Dark Energy Interactions: Theoretical
Challenges, Cosmological Implications and Observational Signatures,” Rept. Prog. Phys. 79 (2016) no. 9, 096901,
arXiv:1603.08299 [astro-ph.CO].
[2041] L. Amendola, “Coupled quintessence,” Phys. Rev. D 62 (2000) 043511, arXiv:astro-ph/9908023.
[2042] V. H. Cárdenas, D. Grandón, and S. Lepe, “Dark energy and Dark matter interaction in light of the second law of
thermodynamics,” Eur. Phys. J. C 79 (2019) no. 4, 357, arXiv:1812.03540 [astro-ph.CO].
[2043] D. Grandón and V. H. Cárdenas, “Exploring evidence of interaction between dark energy and dark matter,” General
Relativity and Gravitation 51 (2019) no. 3, 42, arXiv:1804.03296 [astro-ph.CO].
[2044] C.-Y. Sun and R.-H. Yue, “New Interaction between Dark Energy and Dark Matter Changes Sign during Cosmological
Evolution,” Phys. Rev. D 85 (2012) 043010, arXiv:1009.1214 [gr-qc].
[2045] Y. Wang and P. M. Garnavich, “Measuring time dependence of dark energy density from type Ia supernova data,”
Astrophys. J. 552 (2001) 445, arXiv:astro-ph/0101040.
[2046] Y. Wang and G. Lovelace, “Unbiased estimate of dark energy density from type IA supernova data,” Astrophys. J.
Lett. 562 (2001) L115–L120, arXiv:astro-ph/0109233.
[2047] V. H. Cardenas, “Exploring hints for dark energy density evolution in light of recent data,” Phys. Lett. B 750 (2015)
128–134, arXiv:1405.5116 [astro-ph.CO].
[2048] D. Grandon and V. H. Cardenas, “Studies on dark energy evolution,” Class. Quant. Grav. 38 (2021) no. 14, 145008,
330

arXiv:2107.04876 [astro-ph.CO].
[2049] E. V. Linder, “Cosmic growth history and expansion history,” Phys. Rev. D 72 (2005) 043529,
arXiv:astro-ph/0507263.
[2050] N.-M. Nguyen, D. Huterer, and Y. Wen, “Evidence for Suppression of Structure Growth in the Concordance
Cosmological Model,” Phys. Rev. Lett. 131 (2023) no. 11, 111001, arXiv:2302.01331 [astro-ph.CO].
[2051] R. Caldwell, A. Cooray, and A. Melchiorri, “Constraints on a New Post-General Relativity Cosmological Parameter,”
Phys. Rev. D 76 (2007) 023507, arXiv:astro-ph/0703375.
[2052] L. Amendola, S. Fogli, A. Guarnizo, M. Kunz, and A. Vollmer, “Model-independent constraints on the cosmological
anisotropic stress,” Phys. Rev. D 89 (2014) no. 6, 063538, arXiv:1311.4765 [astro-ph.CO].
[2053] M. Ranjbar, S. Akhshabi, and M. Shadmehri, “Gravitational slip parameter and gravitational waves in
Einstein–Cartan theory,” Eur. Phys. J. C 84 (2024) no. 3, 316, arXiv:2401.02129 [gr-qc].
[2054] R. R. Caldwell, “A Phantom menace?,” Phys. Lett. B 545 (2002) 23–29, arXiv:astro-ph/9908168.
[2055] E. Elizalde, S. Nojiri, and S. D. Odintsov, “Late-time cosmology in (phantom) scalar-tensor theory: Dark energy and
the cosmic speed-up,” Phys. Rev. D 70 (2004) 043539, arXiv:hep-th/0405034.
[2056] Y.-F. Cai, E. N. Saridakis, M. R. Setare, and J.-Q. Xia, “Quintom Cosmology: Theoretical implications and
observations,” Phys. Rept. 493 (2010) 1–60, arXiv:0909.2776 [hep-th].
[2057] Z.-K. Guo, Y.-S. Piao, X.-M. Zhang, and Y.-Z. Zhang, “Cosmological evolution of a quintom model of dark energy,”
Phys. Lett. B 608 (2005) 177–182, arXiv:astro-ph/0410654.
[2058] J. Tot, B. Yildirim, A. Coley, and G. Leon, “The dynamics of scalar-field Quintom cosmological models,” Phys. Dark
Univ. 39 (2023) 101155, arXiv:2204.06538 [gr-qc].
[2059] J. A. Vázquez, D. Tamayo, G. Garcia-Arroyo, I. Gómez-Vargas, I. Quiros, and A. A. Sen, “Coupled multiscalar field
dark energy,” Phys. Rev. D 109 (2024) no. 2, 023511, arXiv:2305.11396 [astro-ph.CO].
[2060] C. Armendariz-Picon, V. F. Mukhanov, and P. J. Steinhardt, “Essentials of k essence,” Phys. Rev. D 63 (2001) 103510,
arXiv:astro-ph/0006373.
[2061] G. W. Horndeski, “Second-order scalar-tensor field equations in a four-dimensional space,” Int. J. Theor. Phys. 10
(1974) 363–384.
[2062] T. Clifton, P. G. Ferreira, A. Padilla, and C. Skordis, “Modified Gravity and Cosmology,” Phys. Rept. 513 (2012)
1–189, arXiv:1106.2476 [astro-ph.CO].
[2063] J. Gleyzes, D. Langlois, F. Piazza, and F. Vernizzi, “Healthy theories beyond Horndeski,” Phys. Rev. Lett. 114 (2015)
no. 21, 211101, arXiv:1404.6495 [hep-th].
[2064] A. A. Starobinsky, “A New Type of Isotropic Cosmological Models Without Singularity,” Phys. Lett. B 91 (1980)
99–102.
[2065] M. Leizerovich, L. Kraiselburd, S. J. Landau, and C. G. Scóccola, “Testing f(R) gravity models with quasar x-ray and
UV fluxes,” Phys. Rev. D 105 (2022) no. 10, 103526, arXiv:2112.01492 [astro-ph.CO].
[2066] Y. Pan, Y. He, J. Qi, J. Li, S. Cao, T. Liu, and J. Wang, “Testing f(R) gravity with the simulated data of gravitational
waves from the Einstein Telescope,” Astrophys. J. 911 (2021) no. 2, 135, arXiv:2103.05212 [astro-ph.CO].
[2067] P. Bode, J. P. Ostriker, and N. Turok, “Halo formation in warm dark matter models,” Astrophys. J. 556 (2001)
93–107, arXiv:astro-ph/0010389.
[2068] B. Liu, H. Shan, and J. Zhang, “New Galaxy UV Luminosity Constraints on Warm Dark Matter from JWST,”
Astrophys. J. 968 (2024) no. 2, 79, arXiv:2404.13596 [astro-ph.CO].
[2069] S. Lin et al., “Can We Constrain Warm Dark Matter Masses with Individual Galaxies?,” Astrophys. J. 970 (2024)
no. 2, 170, arXiv:2401.17940 [astro-ph.CO].
[2070] K. A. Oman, C. S. Frenk, R. A. Crain, M. R. Lovell, and J. Pfeffer, “A warm dark matter cosmogony may yield more
low-mass galaxy detections in 21-cm surveys than a cold dark matter one,” Mon. Not. Roy. Astron. Soc. 533 (2024)
no. 1, 67–78, arXiv:2401.11878 [astro-ph.CO].
[2071] J. C. Rose, P. Torrey, F. Villaescusa-Navarro, M. Vogelsberger, S. O’Neil, M. V. Medvedev, R. Low, R. Adhikari, and
D. Angles-Alcazar, “Inferring warm dark matter masses with deep learning,” Mon. Not. Roy. Astron. Soc. 527 (2023)
no. 1, 739–755, arXiv:2304.14432 [astro-ph.CO].
[2072] D. N. Spergel and P. J. Steinhardt, “Observational evidence for selfinteracting cold dark matter,” Phys. Rev. Lett. 84
(2000) 3760–3763, arXiv:astro-ph/9909386.
[2073] DES Collaboration, D. Cross et al., “Examining the self-interaction of dark matter through central cluster galaxy
offsets,” Mon. Not. Roy. Astron. Soc. 529 (2024) no. 1, 52–58, arXiv:2304.10128 [astro-ph.CO].
[2074] D. Yang, E. O. Nadler, and H.-B. Yu, “Testing the parametric model for self-interacting dark matter using matched
halos in cosmological simulations,” Phys. Dark Univ. 47 (2025) 101807, arXiv:2406.10753 [astro-ph.CO].
[2075] T. Bringmann, F. Kahlhoefer, K. Schmidt-Hoberg, and P. Walia, “Strong constraints on self-interacting dark matter
with light mediators,” Phys. Rev. Lett. 118 (2017) no. 14, 141802, arXiv:1612.00845 [hep-ph].
[2076] M. Visser, “Jerk and the cosmological equation of state,” Class. Quant. Grav. 21 (2004) 2603–2616,
arXiv:gr-qc/0309109.
[2077] M. Visser, “Cosmography: Cosmology without the Einstein equations,” Gen. Rel. Grav. 37 (2005) 1541–1548,
arXiv:gr-qc/0411131.
[2078] C. Cattoen and M. Visser, “Cosmographic Hubble fits to the supernova data,” Phys. Rev. D 78 (2008) 063501,
arXiv:0809.0537 [gr-qc].
[2079] P. K. S. Dunsby and O. Luongo, “On the theory and applications of modern cosmography,” Int. J. Geom. Meth. Mod.
Phys. 13 (2016) no. 03, 1630002, arXiv:1511.06532 [gr-qc].
331

[2080] S. Capozziello, R. D’Agostino, and O. Luongo, “Extended Gravity Cosmography,” Int. J. Mod. Phys. D 28 (2019)
no. 10, 1930016, arXiv:1904.01427 [gr-qc].
[2081] D. Tamayo and J. A. Vazquez, “Fourier-series expansion of the dark-energy equation of state,” Mon. Not. Roy. Astron.
Soc. 487 (2019) no. 1, 729–736, arXiv:1901.08679 [astro-ph.CO].
[2082] C. Gruber and O. Luongo, “Cosmographic analysis of the equation of state of the universe through Padé
approximations,” Phys. Rev. D 89 (2014) no. 10, 103506, arXiv:1309.3215 [gr-qc].
[2083] A. Aviles, A. Bravetti, S. Capozziello, and O. Luongo, “Precision cosmology with Padé rational approximations:
Theoretical predictions versus observational limits,” Phys. Rev. D 90 (2014) no. 4, 043531, arXiv:1405.6935 [gr-qc].
[2084] H. Wei, X.-P. Yan, and Y.-N. Zhou, “Cosmological Applications of Padé Approximant,” JCAP 01 (2014) 045,
arXiv:1312.1117 [astro-ph.CO].
[2085] S. Capozziello and R. D’Agostino, “Reconstructing the distortion function of non-local cosmology: A
model-independent approach,” Phys. Dark Univ. 42 (2023) 101346, arXiv:2310.03136 [gr-qc].
[2086] S. Capozziello, R. D’Agostino, and O. Luongo, “Rational approximations of f (R) cosmography through Padé
polynomials,” JCAP 05 (2018) 008, arXiv:1709.08407 [gr-qc].
[2087] A. Hojjati, L. Pogosian, and G.-B. Zhao, “Detecting Features in the Dark Energy Equation of State: A Wavelet
Approach,” JCAP 04 (2010) 007, arXiv:0912.4843 [astro-ph.CO].
[2088] J. Wagner and S. Meyer, “Generalised model-independent characterisation of strong gravitational lenses V:
reconstructing the lensing distance ratio by supernovae for a general Friedmann universe,” Mon. Not. Roy. Astron.
Soc. 490 (2019) no. 2, 1913–1927, arXiv:1812.04002 [astro-ph.CO].
[2089] C. Mignone and M. Bartelmann, “Model-independent determination of the cosmic expansion rate. I. Application to
type-Ia supernovae,” A&A 481 (2008) no. 2, 295–303, arXiv:0711.0370 [astro-ph].
[2090] C. S. Lorenz, L. Funcke, M. Löffler, and E. Calabrese, “Reconstruction of the neutrino mass as a function of redshift,”
Phys. Rev. D 104 (2021) no. 12, 123518, arXiv:2102.13618 [astro-ph.CO].
[2091] D. K. Hazra, A. Shafieloo, and G. F. Smoot, “Reconstruction of broad features in the primordial spectrum and inflaton
potential from Planck,” JCAP 12 (2013) 035, arXiv:1310.3038 [astro-ph.CO].
[2092] L. A. Escamilla and J. A. Vazquez, “Model selection applied to reconstructions of the Dark Energy,” Eur. Phys. J. C
83 (2023) no. 3, 251, arXiv:2111.10457 [astro-ph.CO].
[2093] Y. Wang, L. Pogosian, G.-B. Zhao, and A. Zucca, “Evolution of dark energy reconstructed from the latest
observations,” Astrophys. J. Lett. 869 (2018) L8, arXiv:1807.03772 [astro-ph.CO].
[2094] G.-B. Zhao, R. G. Crittenden, L. Pogosian, and X. Zhang, “Examining the evidence for dynamical dark energy,” Phys.
Rev. Lett. 109 (2012) 171301, arXiv:1207.3804 [astro-ph.CO].
[2095] G.-B. Zhao et al., “Dynamical dark energy in light of the latest observations,” Nature Astron. 1 (2017) no. 9, 627–632,
arXiv:1701.08165 [astro-ph.CO].
[2096] R.-G. Cai and Q. Su, “On the dark sector interactions,” Physical Review D—Particles, Fields, Gravitation, and
Cosmology 81 (2010) no. 10, 103514.
[2097] A. Moss, E. Copeland, S. Bamford, and T. Clarke, “A model-independent reconstruction of dark energy to very high
redshift,” arXiv:2109.14848 [astro-ph.CO].
[2098] L. Amendola, M. Kunz, and D. Sapone, “Measuring the dark side (with weak lensing),” JCAP 04 (2008) 013,
arXiv:0704.2421 [astro-ph].
[2099] E. Bertschinger and P. Zukin, “Distinguishing Modified Gravity from Dark Energy,” Phys. Rev. D 78 (2008) 024015,
arXiv:0801.2431 [astro-ph].
[2100] L. Pogosian, A. Silvestri, K. Koyama, and G.-B. Zhao, “How to optimally parametrize deviations from general
relativity in the evolution of cosmological perturbations,” Phys. Rev. D 81 (2010) no. 10, 104023, arXiv:1002.2382
[astro-ph.CO].
[2101] M. Raveri, L. Pogosian, M. Martinelli, K. Koyama, A. Silvestri, and G.-B. Zhao, “Principal reconstructed modes of
dark energy and gravity,” JCAP 02 (2023) 061, arXiv:2107.12990 [astro-ph.CO].
[2102] L. Pogosian, M. Raveri, K. Koyama, M. Martinelli, A. Silvestri, G.-B. Zhao, J. Li, S. Peirone, and A. Zucca, “Imprints
of cosmological tensions in reconstructed gravity,” Nature Astron. 6 (2022) no. 12, 1484–1490, arXiv:2107.12992
[astro-ph.CO].
[2103] S. Hee, J. A. Vázquez, W. J. Handley, M. P. Hobson, and A. N. Lasenby, “Constraining the dark energy equation of
state using Bayes theorem and the Kullback–Leibler divergence,” Mon. Not. Roy. Astron. Soc. 466 (2017) no. 1,
369–377, arXiv:1607.00270 [astro-ph.CO].
[2104] Z.-K. Guo, D. J. Schwarz, and Y.-Z. Zhang, “Reconstruction of the primordial power spectrum from CMB data,”
JCAP 08 (2011) 031, arXiv:1105.5916 [astro-ph.CO].
[2105] A. Ravenni, L. Verde, and A. J. Cuesta, “Red, Straight, no bends: primordial power spectrum reconstruction from
CMB and large-scale structure,” JCAP 08 (2016) 028, arXiv:1605.06637 [astro-ph.CO].
[2106] G. Aslanyan, L. C. Price, K. N. Abazajian, and R. Easther, “The Knotted Sky I: Planck constraints on the primordial
power spectrum,” JCAP 08 (2014) 052, arXiv:1403.5849 [astro-ph.CO].
[2107] W. J. Handley, A. N. Lasenby, H. V. Peiris, and M. P. Hobson, “Bayesian inflationary reconstructions from Planck
2018 data,” Phys. Rev. D 100 (2019) no. 10, 103511, arXiv:1908.00906 [astro-ph.CO].
[2108] J. Alberto Vazquez, M. Bridges, M. P. Hobson, and A. N. Lasenby, “Reconstruction of the Dark Energy equation of
state,” JCAP 09 (2012) 020, arXiv:1205.0847 [astro-ph.CO].
[2109] I. Tutusaus, B. Lamine, and A. Blanchard, “Model-independent cosmic acceleration and redshift-dependent intrinsic
luminosity in type-Ia supernovae,” Astron. Astrophys. 625 (2019) A15, arXiv:1803.06197 [astro-ph.CO].
332

[2110] F. Gerardi, M. Martinelli, and A. Silvestri, “Reconstruction of the Dark Energy equation of state from latest data: the
impact of theoretical priors,” JCAP 07 (2019) 042, arXiv:1902.09423 [astro-ph.CO].
[2111] A. Aviles, C. Gruber, O. Luongo, and H. Quevedo, “Cosmography and constraints on the equation of state of the
Universe in various parametrizations,” Phys. Rev. D 86 (2012) 123516, arXiv:1204.2007 [astro-ph.CO].
[2112] K. Bamba, S. Capozziello, S. Nojiri, and S. D. Odintsov, “Dark energy cosmology: the equivalent description via
different theoretical models and cosmography tests,” Astrophys. Space Sci. 342 (2012) 155–228, arXiv:1205.3421
[gr-qc].
[2113] S. Capozziello, R. D’Agostino, and O. Luongo, “Model-independent reconstruction of f (T ) teleparallel cosmology,”
Gen. Rel. Grav. 49 (2017) no. 11, 141, arXiv:1706.02962 [gr-qc].
[2114] S. Capozziello, R. D’Agostino, and O. Luongo, “Thermodynamic parametrization of dark energy,” Phys. Dark Univ.
36 (2022) 101045, arXiv:2202.03300 [astro-ph.CO].
[2115] C. Cattoen and M. Visser, “The Hubble series: Convergence properties and redshift variables,” Class. Quant. Grav. 24
(2007) 5985–5998, arXiv:0710.1887 [gr-qc].
[2116] S. Capozziello and R. D’Agostino, “Model-independent reconstruction of f(Q) non-metric gravity,” Phys. Lett. B 832
(2022) 137229, arXiv:2204.01015 [gr-qc].
[2117] S. Capozziello and R. D’Agostino, “A Cosmographic outlook on Dark Energy and Modified Gravity,” Frascati Phys.
Ser. 74 (2022) 193–208, arXiv:2211.17194 [astro-ph.CO].
[2118] R. D’Agostino and R. C. Nunes, “Cosmographic view on the H0 and σ8 tensions,” Phys. Rev. D 108 (2023) no. 2,
023523, arXiv:2307.13464 [astro-ph.CO].
[2119] S. Capozziello, R. D’Agostino, and O. Luongo, “Cosmographic analysis with Chebyshev polynomials,” Mon. Not. Roy.
Astron. Soc. 476 (2018) no. 3, 3924–3938, arXiv:1712.04380 [astro-ph.CO].
[2120] S. Capozziello, R. D’Agostino, and O. Luongo, “Cosmographic Reconstruction to Discriminate Between Modified
Gravity and Dark Energy,” Acta Phys. Polon. Supp. 13 (2020) 271.
[2121] A. Gómez-Valent, “Quantifying the evidence for the current speed-up of the Universe with low and
intermediate-redshift data. A more model-independent approach,” JCAP 05 (2019) 026, arXiv:1810.02278
[astro-ph.CO].
[2122] C. E. Rasmussen, C. K. Williams, et al., Gaussian processes for machine learning, vol. 1. Springer, 2006.
[2123] C. K. Williams, Gaussian processes for machine learning. MIT press, 2005.
[2124] M. Seikel, C. Clarkson, and M. Smith, “Reconstruction of dark energy and expansion dynamics using Gaussian
processes,” JCAP 06 (2012) 036, arXiv:1204.2832 [astro-ph.CO].
[2125] W. Sun, K. Jiao, and T.-J. Zhang, “Influence of the Bounds of the Hyperparameters on the Reconstruction of the
Hubble Constant with the Gaussian Process,” Astrophys. J. 915 (2021) no. 2, 123, arXiv:2105.12618 [astro-ph.CO].
[2126] S. Dhawan, J. Alsing, and S. Vagnozzi, “Non-parametric spatial curvature inference using late-Universe cosmological
probes,” Mon. Not. Roy. Astron. Soc. 506 (2021) no. 1, L1–L5, arXiv:2104.02485 [astro-ph.CO].
[2127] S.-g. Hwang, B. L’Huillier, R. E. Keeley, M. J. Jee, and A. Shafieloo, “How to use GP: effects of the mean function and
hyperparameter selection on Gaussian process regression,” JCAP 02 (2023) 014, arXiv:2206.15081 [astro-ph.CO].
[2128] F. Avila, A. Bernui, A. Bonilla, and R. C. Nunes, “Inferring S8 (z) and γ(z) with cosmic growth rate measurements
using machine learning,” Eur. Phys. J. C 82 (2022) no. 7, 594, arXiv:2201.07829 [astro-ph.CO].
[2129] S. Joudaki, M. Kaplinghat, R. Keeley, and D. Kirkby, “Model independent inference of the expansion history and
implications for the growth of structure,” Phys. Rev. D 97 (2018) no. 12, 123501, arXiv:1710.04236 [astro-ph.CO].
[2130] R. E. Keeley, A. Shafieloo, B. L’Huillier, and E. V. Linder, “Debiasing Cosmic Gravitational Wave Sirens,” Mon. Not.
Roy. Astron. Soc. 491 (2020) no. 3, 3983–3989, arXiv:1905.10216 [astro-ph.CO].
[2131] R. Calderón, B. L’Huillier, D. Polarski, A. Shafieloo, and A. A. Starobinsky, “Joint reconstructions of growth and
expansion histories from stage-IV surveys with minimal assumptions: Dark energy beyond Λ,” Phys. Rev. D 106
(2022) no. 8, 083513, arXiv:2206.13820 [astro-ph.CO].
[2132] R. Calderón, B. L’Huillier, D. Polarski, A. Shafieloo, and A. A. Starobinsky, “Joint reconstructions of growth and
expansion histories from stage-IV surveys with minimal assumptions. II. Modified gravity and massive neutrinos,”
Phys. Rev. D 108 (2023) no. 2, 023504, arXiv:2301.00640 [astro-ph.CO].
[2133] M. Reyes and C. Escamilla-Rivera, “On the Degeneracy between fσ 8 Tension and Its Gaussian Process Forecasting,”
Universe 8 (2022) no. 8, 394, arXiv:2203.03574 [astro-ph.CO].
[2134] M. A. Sabogal, O. Akarsu, A. Bonilla, E. Di Valentino, and R. C. Nunes, “Exploring new physics in the late Universe’s
expansion through non-parametric inference,” Eur. Phys. J. C 84 (2024) no. 7, 703, arXiv:2407.04223
[astro-ph.CO].
[2135] M.-J. Zhang and H. Li, “Gaussian processes reconstruction of dark energy from observational data,” Eur. Phys. J. C
78 (2018) no. 6, 460, arXiv:1806.02981 [astro-ph.CO].
[2136] R. Briffa, S. Capozziello, J. Levi Said, J. Mifsud, and E. N. Saridakis, “Constraining teleparallel gravity through
Gaussian processes,” Class. Quant. Grav. 38 (2020) no. 5, 055007, arXiv:2009.14582 [gr-qc].
[2137] N. Banerjee, P. Mukherjee, and D. Pavón, “Spatial curvature and thermodynamics,” Mon. Not. Roy. Astron. Soc. 521
(2023) no. 4, 5473–5482, arXiv:2301.09823 [astro-ph.CO].
[2138] N. Banerjee, P. Mukherjee, and D. Pavón, “Checking the second law at cosmic scales,” JCAP 11 (2023) 092,
arXiv:2309.12298 [astro-ph.CO].
[2139] P. Mukherjee and N. Banerjee, “Revisiting a non-parametric reconstruction of the deceleration parameter from
combined background and the growth rate data,” Phys. Dark Univ. 36 (2022) 100998, arXiv:2007.15941
[astro-ph.CO].
333

[2140] A. Shafieloo, A. G. Kim, and E. V. Linder, “Gaussian process cosmography,” Phys. Rev. D 85 (2012) no. 12, 123530,
arXiv:1204.2272 [astro-ph.CO].
[2141] M.-J. Zhang and J.-Q. Xia, “Test of the cosmic evolution using Gaussian processes,” JCAP 12 (2016) 005,
arXiv:1606.04398 [astro-ph.CO].
[2142] K. Liao, A. Shafieloo, R. E. Keeley, and E. V. Linder, “A model-independent determination of the Hubble constant
from lensed quasars and supernovae using Gaussian process regression,” Astrophys. J. Lett. 886 (2019) no. 1, L23,
arXiv:1908.04967 [astro-ph.CO].
[2143] P. Mukherjee and A. Mukherjee, “Assessment of the cosmic distance duality relation using Gaussian process,” Mon.
Not. Roy. Astron. Soc. 504 (2021) no. 3, 3938–3946, arXiv:2104.06066 [astro-ph.CO].
[2144] F. Renzi, N. B. Hogg, and W. Giarè, “The resilience of the Etherington–Hubble relation,” Mon. Not. Roy. Astron. Soc.
513 (2022) no. 3, 4004–4014, arXiv:2112.05701 [astro-ph.CO].
[2145] P. Mukherjee and N. Banerjee, “Non-parametric reconstruction of the cosmological jerk parameter,” Eur. Phys. J. C
81 (2021) no. 1, 36, arXiv:2007.10124 [astro-ph.CO].
[2146] J. F. Jesus, R. Valentim, A. A. Escobal, S. H. Pereira, and D. Benndorf, “Gaussian processes reconstruction of the dark
energy potential,” JCAP 11 (2022) 037, arXiv:2112.09722 [astro-ph.CO].
[2147] D. Wang and X.-H. Meng, “Improved constraints on the dark energy equation of state using Gaussian processes,” Phys.
Rev. D 95 (2017) no. 2, 023508, arXiv:1708.07750 [astro-ph.CO].
[2148] T. Holsclaw, U. Alam, B. Sanso, H. Lee, K. Heitmann, S. Habib, and D. Higdon, “Nonparametric Dark Energy
Reconstruction from Supernova Data,” Phys. Rev. Lett. 105 (2010) 241302, arXiv:1011.3079 [astro-ph.CO].
[2149] M. Aljaf, D. Gregoris, and M. Khurshudyan, “Constraints on interacting dark energy models through cosmic
chronometers and Gaussian process,” Eur. Phys. J. C 81 (2021) no. 6, 544, arXiv:2005.01891 [astro-ph.CO].
[2150] A. Bonilla, S. Kumar, R. C. Nunes, and S. Pan, “Reconstruction of the dark sectors’ interaction: A model-independent
inference and forecast from GW standard sirens,” Mon. Not. Roy. Astron. Soc. 512 (2022) no. 3, 4231–4238,
arXiv:2102.06149 [astro-ph.CO].
[2151] L. A. Escamilla, O. Akarsu, E. Di Valentino, and J. A. Vazquez, “Model-independent reconstruction of the interacting
dark energy kernel: Binned and Gaussian process,” JCAP 11 (2023) 051, arXiv:2305.16290 [astro-ph.CO].
[2152] P. Mukherjee and N. Banerjee, “Nonparametric reconstruction of interaction in the cosmic dark sector,” Phys. Rev. D
103 (2021) no. 12, 123530, arXiv:2105.09995 [astro-ph.CO].
[2153] A. M. Pinho, S. Casas, and L. Amendola, “Model-independent reconstruction of the linear anisotropic stress η,” JCAP
11 (2018) 027, arXiv:1805.00027 [astro-ph.CO].
[2154] Y. Mu, E.-K. Li, and L. Xu, “Model-independent reconstruction of growth index via Gaussian process,” Classical and
Quantum Gravity 40 (2023) no. 22, 225003.
[2155] J. A. S. Fortunato, P. H. R. S. Moraes, J. G. d. L. Júnior, and E. Brito, “Search for the f(R, T) gravity functional form
via gaussian processes,” Eur. Phys. J. C 84 (2024) no. 2, 198, arXiv:2305.01325 [gr-qc].
[2156] A. K. Singha, A. Sardar, and U. Debnath, “f(Q) Reconstruction: In the light of various modified gravity models,”
Physics of the Dark Universe 41 (2023) 101240.
[2157] F. Oliveira, F. Avila, A. Bernui, A. Bonilla, and R. C. Nunes, “Reconstructing the growth index γ with Gaussian
processes,” Eur. Phys. J. C 84 (2024) no. 6, 636, arXiv:2311.14216 [astro-ph.CO].
[2158] G. N. Gadbail, S. Mandal, and P. K. Sahoo, “Gaussian Process Approach for Model-independent Reconstruction of
f(Q) Gravity with Direct Hubble Measurements,” Astrophys. J. 972 (2024) no. 2, 174, arXiv:2404.13095 [gr-qc].
[2159] Y. Mu, E.-K. Li, and L. Xu, “Data-driven and almost model-independent reconstruction of modified gravity,” JCAP
06 (2023) 022, arXiv:2302.09777 [astro-ph.CO].
[2160] J. Sultana, M. K. Yennapureddy, F. Melia, and D. Kazanas, “Constraining f(R) models with cosmic chronometers and
the H ii galaxy Hubble diagram,” Mon. Not. Roy. Astron. Soc. 514 (2022) no. 4, 5827–5839, arXiv:2206.10761
[astro-ph.CO].
[2161] E. Elizalde and M. Khurshudyan, “Swampland criteria for f(R) gravity derived with a Gaussian process,” European
Physical Journal C 82 (2022) no. 9, 811.
[2162] D. Wang, “Pantheon+ constraints on dark energy and modified gravity: An evidence of dynamical dark energy,” Phys.
Rev. D 106 (2022) no. 6, 063515, arXiv:2207.07164 [astro-ph.CO].
[2163] A. Sardar and U. Debnath, “Reconstruction of extended f(P) cubic gravity from other modified gravity models,”
Physics of the Dark Universe 35 (2022) 100926.
[2164] X. Ren, S.-F. Yan, Y. Zhao, Y.-F. Cai, and E. N. Saridakis, “Gaussian processes and effective field theory of f (T )
gravity under the H0 tension,” Astrophys. J. 932 (2022) 2, arXiv:2203.01926 [astro-ph.CO].
[2165] Y.-F. Cai, M. Khurshudyan, and E. N. Saridakis, “Model-independent reconstruction of f (T ) gravity from Gaussian
Processes,” Astrophys. J. 888 (2020) 62, arXiv:1907.10813 [astro-ph.CO].
[2166] Y. Liu, S. Cao, T. Liu, X. Li, S. Geng, Y. Lian, and W. Guo, “Model-independent constraints on cosmic curvature:
implication from updated Hubble diagram of high-redshift standard candles,” Astrophys. J. 901 (2020) no. 2, 129,
arXiv:2008.08378 [astro-ph.CO].
[2167] P. Mukherjee and N. Banerjee, “Constraining the curvature density parameter in cosmology,” Phys. Rev. D 105 (2022)
no. 6, 063516, arXiv:2202.07886 [astro-ph.CO].
[2168] Y. Pan, J. Diao, J.-Z. Qi, J. Li, S. Cao, and Q.-Q. Jiang, “Testing the spatial geometry of the Universe with TianQin:
Prospect of using supermassive black hole binaries,” Astron. Astrophys. 683 (2024) A91, arXiv:2310.14723
[astro-ph.CO].
[2169] J.-Z. Qi, P. Meng, J.-F. Zhang, and X. Zhang, “Model-independent measurement of cosmic curvature with the latest
334

H(z) and SNe Ia data: A comprehensive investigation,” Phys. Rev. D 108 (2023) no. 6, 063522, arXiv:2302.08889
[astro-ph.CO].
[2170] Y.-J. Wang, J.-Z. Qi, B. Wang, J.-F. Zhang, J.-L. Cui, and X. Zhang, “Cosmological model-independent measurement
of cosmic curvature using distance sum rule with the help of gravitational waves,” Mon. Not. Roy. Astron. Soc. 516
(2022) no. 4, 5187–5195, arXiv:2201.12553 [astro-ph.CO].
[2171] P.-J. Wu, J.-Z. Qi, and X. Zhang, “Null test for cosmic curvature using Gaussian process*,” Chin. Phys. C 47 (2023)
no. 5, 055106, arXiv:2209.08502 [astro-ph.CO].
[2172] Y. Yang and Y. Gong, “Measurement on the cosmic curvature using the Gaussian process method,” Mon. Not. Roy.
Astron. Soc. 504 (2021) no. 2, 3092–3097, arXiv:2007.05714 [astro-ph.CO].
[2173] G.-J. Wang, X.-J. Ma, and J.-Q. Xia, “Machine learning the cosmic curvature in a model-independent way,” Mon. Not.
Roy. Astron. Soc. 501 (2021) no. 4, 5714–5722, arXiv:2004.13913 [astro-ph.CO].
[2174] E. Belgacem, S. Foffa, M. Maggiore, and T. Yang, “Gaussian processes reconstruction of modified gravitational wave
propagation,” Phys. Rev. D 101 (2020) no. 6, 063505, arXiv:1911.11497 [astro-ph.CO].
[2175] X. Zheng, S. Cao, Y. Liu, M. Biesiada, T. Liu, S. Geng, Y. Lian, and W. Guo, “Model-independent constraints on
cosmic curvature: implication from the future space gravitational-wave antenna DECIGO,” Eur. Phys. J. C 81 (2021)
no. 1, 14, arXiv:2012.14607 [astro-ph.CO].
[2176] R. Shah, A. Bhaumik, P. Mukherjee, and S. Pal, “A thorough investigation of the prospects of eLISA in addressing the
Hubble tension: Fisher forecast, MCMC and Machine Learning,” JCAP 06 (2023) 038, arXiv:2301.12708
[astro-ph.CO].
[2177] G. Cañas Herrera, J. Torrado, and A. Achúcarro, “Bayesian reconstruction of the inflaton’s speed of sound using CMB
data,” Phys. Rev. D 103 (2021) 123531, arXiv:2012.04640 [astro-ph.CO].
[2178] B. R. Dinda and N. Banerjee, “Model independent bounds on type Ia supernova absolute peak magnitude,” Phys. Rev.
D 107 (2023) no. 6, 063513, arXiv:2208.14740 [astro-ph.CO].
[2179] A. Favale, A. Gómez-Valent, and M. Migliaccio, “Quantification of 2D vs 3D BAO tension using SNIa as a redshift
interpolator and test of the Etherington relation,” Phys. Lett. B 858 (2024) 139027, arXiv:2405.12142
[astro-ph.CO].
[2180] G. Rodrigues and C. Bengaly, “A model-independent test of speed of light variability with cosmological observations,”
JCAP 07 (2022) no. 07, 029, arXiv:2112.01963 [astro-ph.CO].
[2181] P. Mukherjee, R. Shah, A. Bhaumik, and S. Pal, “Reconstructing the Hubble Parameter with Future
Gravitational-wave Missions Using Machine Learning,” Astrophys. J. 960 (2024) no. 1, 61, arXiv:2303.05169
[astro-ph.CO].
[2182] J.-L. Li, Y.-P. Yang, S.-X. Yi, J.-P. Hu, F.-Y. Wang, and Y.-K. Qu, “Constraints on the Cosmological Parameters with
Three-Parameter Correlation of Gamma-Ray Bursts,” Astrophys. J. 953 (2023) no. 1, 58, arXiv:2306.12840
[astro-ph.HE].
[2183] P. Mukherjee and A. A. Sen, “Model-independent cosmological inference post DESI DR1 BAO measurements,” Phys.
Rev. D 110 (2024) no. 12, 123502, arXiv:2405.19178 [astro-ph.CO].
[2184] M. Khurshudyan and E. Elizalde, “Constraints on Prospective Deviations from the Cold Dark Matter Model Using a
Gaussian Process,” Galaxies 12 (2024) no. 4, 31, arXiv:2402.08630 [gr-qc].
[2185] R. von Marttens, J. E. Gonzalez, J. Alcaniz, V. Marra, and L. Casarini, “Model-independent reconstruction of dark
sector interactions,” Phys. Rev. D 104 (2021) no. 4, 043515, arXiv:2011.10846 [astro-ph.CO].
[2186] P. Mukherjee, Non-parametric Reconstruction Of Some Cosmological Parameters. PhD thesis, IISER, Kolkata, 2022.
arXiv:2207.07857 [astro-ph.CO].
[2187] P. Mukherjee, G. Rodrigues, and C. Bengaly, “Examining the validity of the minimal varying speed of light model
through cosmological observations: Relaxing the null curvature constraint,” Phys. Dark Univ. 43 (2024) 101380,
arXiv:2302.00867 [astro-ph.CO].
[2188] K. Dialektopoulos, J. L. Said, J. Mifsud, J. Sultana, and K. Z. Adami, “Neural network reconstruction of late-time
cosmology and null tests,” JCAP 02 (2022) no. 02, 023, arXiv:2111.11462 [astro-ph.CO].
[2189] A. Mitra, I. Gómez-Vargas, and V. Zarikas, “Dark energy reconstruction analysis with artificial neural networks:
Application on simulated Supernova Ia data from Rubin Observatory,” Phys. Dark Univ. 46 (2024) 101706,
arXiv:2402.18124 [astro-ph.CO].
[2190] T.-X. Mao, J. Wang, B. Li, Y.-C. Cai, B. Falck, M. Neyrinck, and A. Szalay, “Baryon acoustic oscillations
reconstruction using convolutional neural networks,” Mon. Not. Roy. Astron. Soc. 501 (2021) no. 1, 1499–1510,
arXiv:2002.10218 [astro-ph.CO].
[2191] G. Garcia-Arroyo, I. Gómez-Vargas, and J. A. Vázquez, “Reconstructing rotation curves with artificial neural
networks,” arXiv:2404.05833 [astro-ph.GA].
[2192] A. Shafieloo, U. Alam, V. Sahni, and A. A. Starobinsky, “Smoothing Supernova Data to Reconstruct the Expansion
History of the Universe and its Age,” Mon. Not. Roy. Astron. Soc. 366 (2006) 1081–1095, arXiv:astro-ph/0505329.
[2193] A. Shafieloo, “Model Independent Reconstruction of the Expansion History of the Universe and the Properties of Dark
Energy,” Mon. Not. Roy. Astron. Soc. 380 (2007) 1573–1580, arXiv:astro-ph/0703034.
[2194] A. Shafieloo and C. Clarkson, “Model independent tests of the standard cosmological model,” Phys. Rev. D 81 (2010)
083537, arXiv:0911.4858 [astro-ph.CO].
[2195] A. Shafieloo, B. L’Huillier, and A. A. Starobinsky, “Falsifying ΛCDM: Model-independent tests of the concordance
model with eBOSS DR14Q and Pantheon,” Phys. Rev. D 98 (2018) no. 8, 083526, arXiv:1804.04320 [astro-ph.CO].
[2196] A. Shafieloo, “Crossing Statistic: Reconstructing the Expansion History of the Universe,” JCAP 08 (2012) 002,
335

arXiv:1204.1109 [astro-ph.CO].
[2197] H. Koo, A. Shafieloo, R. E. Keeley, and B. L’Huillier, “Model selection and parameter estimation using the iterative
smoothing method,” JCAP 03 (2021) 034, arXiv:2009.12045 [astro-ph.CO].
[2198] B. L’Huillier and A. Shafieloo, “Model-independent test of the FLRW metric, the flatness of the Universe, and
non-local measurement of H0 rd ,” JCAP 01 (2017) 015, arXiv:1606.06832 [astro-ph.CO].
[2199] B. L’Huillier, A. Shafieloo, and H. Kim, “Model-independent cosmological constraints from growth and expansion,”
Mon. Not. Roy. Astron. Soc. 476 (2018) no. 3, 3263–3268, arXiv:1712.04865 [astro-ph.CO].
[2200] B. L’Huillier, A. Mitra, A. Shafieloo, R. E. Keeley, and H. Koo, “Litmus tests of the flat ΛCDM model and
model-independent measurement of H0 rd with LSST and DESI,” arXiv:2407.07847 [astro-ph.CO].
[2201] A. Montiel, R. Lazkoz, I. Sendra, C. Escamilla-Rivera, and V. Salzano, “Nonparametric reconstruction of the cosmic
expansion with local regression smoothing and simulation extrapolation,” Phys. Rev. D 89 (2014) no. 4, 043007,
arXiv:1401.4188 [astro-ph.CO].
[2202] L. M. Fernández-Hernández, A. Montiel, and M. A. Rodríguez-Meza, “Galaxy rotation curves using a non-parametric
regression method: core, cuspy and fuzzy scalar field dark matter models,” MNRAS 488 (2019) no. 4, 5127–5144,
arXiv:1809.06875 [astro-ph.GA].
[2203] C. Escamilla-Rivera, J. Levi Said, and J. Mifsud, “Performance of non-parametric reconstruction techniques in the
late-time universe,” JCAP 10 (2021) 016, arXiv:2105.14332 [astro-ph.CO].
[2204] A. Rana, D. Jain, S. Mahajan, and A. Mukherjee, “Revisiting the distance duality relation using a non-parametric
regression method,” JCAP 07 (2016) 026, arXiv:1511.09223 [astro-ph.CO].
[2205] C. Escamilla-Rivera and J. C. Fabris, “Nonparametric reconstruction of the Om diagnostic to test ΛCDM,” Galaxies 4
(2016) no. 4, 76, arXiv:1511.07066 [astro-ph.CO].
[2206] G. Rudolph, “Convergence analysis of canonical genetic algorithms,” IEEE transactions on neural networks 5 (1994)
no. 1, 96–101.
[2207] Y. Akrami, P. Scott, J. Edsjo, J. Conrad, and L. Bergstrom, “A Profile Likelihood Analysis of the Constrained MSSM
with Genetic Algorithms,” JHEP 04 (2010) 057, arXiv:0910.3950 [hep-ph].
[2208] J. Crowder, N. J. Cornish, and L. Reddinger, “Darwin meets Einstein: LISA data analysis using genetic algorithms,”
Phys. Rev. D 73 (2006) 063011, arXiv:gr-qc/0601036.
[2209] C. Bogdanos and S. Nesseris, “Genetic Algorithms and Supernovae Type Ia Analysis,” JCAP 05 (2009) 006,
arXiv:0903.2805 [astro-ph.CO].
[2210] S. Nesseris and A. Shafieloo, “A model independent null test on the cosmological constant,” Mon. Not. Roy. Astron.
Soc. 408 (2010) 1879–1885, arXiv:1004.0960 [astro-ph.CO].
[2211] S. Nesseris and J. Garcia-Bellido, “A new perspective on Dark Energy modeling via Genetic Algorithms,” JCAP 11
(2012) 033, arXiv:1205.0364 [astro-ph.CO].
[2212] R. Medel-Esquivel, I. Gómez-Vargas, A. A. M. Sánchez, R. García-Salcedo, and J. Alberto Vázquez, “Cosmological
Parameter Estimation with Genetic Algorithms,” Universe 10 (2024) no. 1, 11, arXiv:2311.05699 [astro-ph.CO].
[2213] A. F. Gad, “Pygad: An intuitive genetic algorithm python library,” Multimedia Tools and Applications (2023) 1–14.
[2214] R. C. Bernardo and J. Levi Said, “Towards a model-independent reconstruction approach for late-time Hubble data,”
JCAP 08 (2021) 027, arXiv:2106.08688 [astro-ph.CO].
[2215] M. B. Bainbridge and J. K. Webb, “Artificial intelligence applied to the automatic analysis of absorption spectra.
Objective measurement of the fine structure constant,” MNRAS 468 (2017) no. 2, 1639–1670, arXiv:1606.07393
[astro-ph.IM].
[2216] M. B. Bainbridge and J. K. Webb, “Evaluating the New Automatic Method for the Analysis of Absorption Spectra
Using Synthetic Spectra,” Universe 3 (2017) no. 2, 34, arXiv:1704.08710 [astro-ph.IM].
[2217] C.-C. Lee, J. K. Webb, R. F. Carswell, and D. Milaković, “Artificial intelligence and quasar absorption system
modelling; application to fundamental constants at high redshift,” Mon. Not. Roy. Astron. Soc. 504 (2021) no. 2,
1787–1800, arXiv:2008.02583 [astro-ph.CO].
[2218] J. K. Webb, C.-C. Lee, R. F. Carswell, and D. Milaković, “Getting the model right: an information criterion for
spectroscopy,” MNRAS 501 (2021) no. 2, 2268–2278, arXiv:2009.08336 [astro-ph.IM].
[2219] C. Escamilla-Rivera, “Bayesian deep learning for dark energy,” arXiv:2005.06412 [astro-ph.CO].
[2220] C. Escamilla-Rivera, “Deep learning for cosmology,” PoS AISIS2019 (2020) 021.
[2221] C. Z. Munõz and C. Escamilla-Rivera, “Inverse Cosmography: testing the effectiveness of cosmographic polynomials
using machine learning,” JCAP 12 (2020) 007, arXiv:2005.02807 [gr-qc].
[2222] M. Gendreau and J.-Y. Potvin, eds., A Modern Introduction to Memetic Algorithms, pp. 141–183. Springer US,
Boston, MA, 2010. https://doi.org/10.1007/978-1-4419-1665-5_6.
[2223] O. H. M. Ross, “A Review of Quantum-Inspired Metaheuristics: Going From Classical Computers to Real Quantum
Computers,” IEEE Access 8 (2020) 814–838.
[2224] G. Acampora, R. Schiattarella, and A. Vitiello, “Using quantum amplitude amplification in genetic algorithms,” Expert
Systems With Applications 209 (2022) 118203.
[2225] G. Acampora and A. Vitiello, “Implementing evolutionary optimization on actual quantum processors,” Info. Sci. 575
(2021) 542–562.
[2226] R. Ibarrondo, G. Gatti, and M. Sanz, “Quantum vs classical genetic algorithms: A numerical comparison shows faster
convergence,” in 2022 IEEE Symposium Series on Computational Intelligence. 2022. arXiv:2207.09251 [quant-ph].
[2227] R. Ibarrondo, G. Gatti, and M. Sanz, “Quantum Genetic Algorithm with Individuals in Multiple Registers,” IEEE
Trans. Evol. Comput. 28 (2024) 788–797, arXiv:2203.15039 [quant-ph].
336

[2228] G. Sarracino et al., “A Quantum Genetic Algorithm for Cosmological Functions,” In preparation (2024) .
[2229] J. Prasad and T. Souradeep, “Cosmological parameter estimation using Particle Swarm Optimization (PSO),” Phys.
Rev. D 85 (2012) no. 12, 123008, arXiv:1108.5600 [astro-ph.CO]. [Erratum: Phys.Rev.D 90, 109903 (2014)].
[2230] A. N. Ruiz et al., “Calibration of semi-analytic models of galaxy formation using Particle Swarm Optimization,”
Astrophys. J. 801 (2015) no. 2, 139, arXiv:1310.7034 [astro-ph.CO].
[2231] C. Skokos, K. E. Parsopoulos, P. A. Patsis, and M. N. Vrahatis, “Particle swarm optimization: An Efficient method for
tracing periodic orbits in 3-D Galactic potentials,” Mon. Not. Roy. Astron. Soc. 359 (2005) 251–260,
arXiv:astro-ph/0502164.
[2232] DarkMachines High Dimensional Sampling Group Collaboration, C. Balázs et al., “A comparison of
optimisation algorithms for high-dimensional particle and astrophysics applications,” JHEP 05 (2021) 108,
arXiv:2101.04525 [hep-ph].
[2233] T. Toni, D. Welch, N. Strelkowa, A. Ipsen, and M. P. H. Stumpf, “Approximate bayesian computation scheme for
parameter inference and model selection in dynamical systems,” Journal of The Royal Society Interface 6 (2010)
no. 31, 187–202, arXiv:0901.1925. https://royalsocietypublishing.org/doi/abs/10.1098/rsif.2008.0172.
[2234] T. Toni and M. P. H. Stumpf, “Simulation-based model selection for dynamical systems in systems and population
biology,” Bioinformatics 26 (2009) no. 1, 104–110, arXiv:0911.1705.
https://doi.org/10.1093/bioinformatics/btp619.
[2235] T. Toni and M. P. H. Stumpf, “Tutorial on ABC rejection and ABC SMC for parameter estimation and model
selection,” arXiv e-prints (2009) arXiv:0910.4472, arXiv:0910.4472 [stat.CO].
[2236] R. C. Bernardo, D. Grandón, J. Levi Said, and V. H. Cárdenas, “Dark energy by natural evolution: Constraining dark
energy using Approximate Bayesian Computation,” Phys. Dark Univ. 40 (2023) 101213, arXiv:2211.05482
[astro-ph.CO].
[2237] R. C. Bernardo and Y.-R. Lee, “Hubble constant by natural selection: Evolution chips in the Hubble tension,” Astron.
Comput. 44 (2023) 100740, arXiv:2212.02203 [astro-ph.CO].
[2238] K. M. Sallam, S. M. Elsayed, R. K. Chakrabortty, and M. J. Ryan, “Improved multi-operator differential evolution
algorithm for solving unconstrained problems,” in 2020 IEEE Congress on Evolutionary Computation (CEC), pp. 1–8.
2020.
[2239] E. A. T. Enriquez, R. G. Mendoza, and A. C. T. Velasco, “Philippine Eagle Optimization Algorithm,” arXiv e-prints
(2021) arXiv:2112.10318, arXiv:2112.10318 [math.OC].
[2240] R. C. Bernardo, R. Mendoza, E. A. Enriquez, A. C. Velasco, and R. Reyes, “Metaheuristic cosmological parameter
estimation: The Philippine Eagle hunts down cosmic expansion,” In preparation (2025) .
[2241] K. Y. Kim and H. W. Lee, “Investigating the suitability of data-driven methods for extracting physical parameters in
cosmological models,” Astron. Comput. 45 (2023) 100762.
[2242] K. Lodha, L. Pinol, S. Nesseris, A. Shafieloo, W. Sohn, and M. Fasiello, “Searching for local features in primordial
power spectrum using genetic algorithms,” Mon. Not. Roy. Astron. Soc. 530 (2024) no. 2, 1424–1435,
arXiv:2308.04940 [astro-ph.CO].
[2243] A. Antony, F. Finelli, D. K. Hazra, and A. Shafieloo, “Discordances in Cosmology and the Violation of Slow-Roll
Inflationary Dynamics,” Phys. Rev. Lett. 130 (2023) no. 11, 111001, arXiv:2202.14028 [astro-ph.CO].
[2244] D. K. Hazra, A. Antony, and A. Shafieloo, “One spectrum to cure them all: signature from early Universe solves major
anomalies and tensions in cosmology,” JCAP 08 (2022) no. 08, 063, arXiv:2201.12000 [astro-ph.CO].
[2245] G. Alestas, L. Kazantzidis, and S. Nesseris, “Machine learning constraints on deviations from general relativity from
the large scale structure of the Universe,” Phys. Rev. D 106 (2022) no. 10, 103519, arXiv:2209.12799 [astro-ph.CO].
[2246] EUCLID Collaboration, M. Martinelli et al., “Euclid: Forecast constraints on the cosmic distance duality relation
with complementary external probes,” Astron. Astrophys. 644 (2020) A80, arXiv:2007.16153 [astro-ph.CO].
[2247] Euclid Collaboration, M. Martinelli et al., “Euclid: Constraining dark energy coupled to electromagnetism using
astrophysical and laboratory data,” Astron. Astrophys. 654 (2021) A148, arXiv:2105.09746 [astro-ph.CO].
[2248] Euclid Collaboration, S. Nesseris et al., “Euclid: Forecast constraints on consistency tests of the ΛCDM model,”
Astron. Astrophys. 660 (2022) A67, arXiv:2110.11421 [astro-ph.CO].
[2249] M. Moresco, R. Jimenez, L. Verde, A. Cimatti, and L. Pozzetti, “Setting the Stage for Cosmic Chronometers. II.
Impact of Stellar Population Synthesis Models Systematics and Full Covariance Matrix,” Astrophys. J. 898 (2020)
no. 1, 82, arXiv:2003.07362 [astro-ph.GA].
[2250] D. Brout et al., “The Pantheon+ Analysis: SuperCal-fragilistic Cross Calibration, Retrained SALT2 Light-curve Model,
and Calibration Systematic Uncertainty,” Astrophys. J. 938 (2022) no. 2, 111, arXiv:2112.03864 [astro-ph.CO].
[2251] R. C. Bernardo and Y. Chen, “Demystifying genetic algorithm for cosmological parameter estimation and such,” In
preparation (2024) .
[2252] A. B. Abdessalem, N. Dervilis, D. J. Wagg, and K. Worden, “Automatic kernel selection for gaussian processes
regression with approximate bayesian computation and sequential monte carlo,” Frontiers in Built Environment 3
(2017) 52.
[2253] H. Zhang, Y.-C. Wang, T.-J. Zhang, and T.-t. Zhang, “Kernel Selection for Gaussian Process in Cosmology: With
Approximate Bayesian Computation Rejection and Nested Sampling,” Astrophys. J. Suppl. 266 (2023) no. 2, 27,
arXiv:2304.03911 [astro-ph.CO].
[2254] V. K. Oikonomou and G. Kafanelis, “Primordial cosmology of an emergent-like universe from modified gravity:
Reconstruction and phenomenology optimization with a genetic algorithm,” International Journal of Modern Physics
D 33 (2024) no. 1, 2350114, arXiv:2312.16324 [gr-qc].
337

[2255] A. Aizpuru, R. Arjona, and S. Nesseris, “Machine learning improved fits of the sound horizon at the baryon drag
epoch,” Phys. Rev. D 104 (2021) no. 4, 043521, arXiv:2106.00428 [astro-ph.CO].
[2256] R. Arjona, A. Melchiorri, and S. Nesseris, “Testing the ΛCDM paradigm with growth rate data and machine learning,”
JCAP 05 (2022) no. 05, 047, arXiv:2107.04343 [astro-ph.CO].
[2257] M. R. Gangopadhyay, M. Sami, and M. K. Sharma, “Phantom dark energy as a natural selection of evolutionary
processes a^ la genetic algorithm and cosmological tensions,” Phys. Rev. D 108 (2023) no. 10, 103526,
arXiv:2303.07301 [astro-ph.CO].
[2258] L. Kazantzidis and L. Perivolaropoulos, “Evolution of the f σ8 tension with the Planck15/ΛCDM determination and
implications for modified gravity theories,” Phys. Rev. D 97 (2018) no. 10, 103503, arXiv:1803.01337 [astro-ph.CO].
[2259] V. Springel, N. Yoshida, and S. D. M. White, “GADGET: A Code for collisionless and gasdynamical cosmological
simulations,” New Astron. 6 (2001) 79, arXiv:astro-ph/0003162.
[2260] V. Springel, “The Cosmological simulation code GADGET-2,” Mon. Not. Roy. Astron. Soc. 364 (2005) 1105–1134,
arXiv:astro-ph/0505010.
[2261] V. Springel, R. Pakmor, O. Zier, and M. Reinecke, “Simulating cosmic structure formation with the gadget-4 code,”
Mon. Not. Roy. Astron. Soc. 506 (2021) no. 2, 2871–2949, arXiv:2010.03567 [astro-ph.IM].
[2262] V. Springel, “E pur si muove: Galiliean-invariant cosmological hydrodynamical simulations on a moving mesh,” Mon.
Not. Roy. Astron. Soc. 401 (2010) 791, arXiv:0901.4107 [astro-ph.CO].
[2263] P. F. Hopkins, “A new class of accurate, mesh-free hydrodynamic simulation methods,” Mon. Not. Roy. Astron. Soc.
450 (2015) no. 1, 53–110, arXiv:1409.7395 [astro-ph.CO].
[2264] SWIFT Collaboration, M. Schaller et al., “SWIFT: A modern highly-parallel gravity and smoothed particle
hydrodynamics solver for astrophysical and cosmological applications,” arXiv:2305.13380 [astro-ph.IM].
[2265] R. Teyssier, “Cosmological hydrodynamics with adaptive mesh refinement: a new high resolution code called
RAMSES,” Astron. Astrophys. 385 (2002) 337–364, arXiv:astro-ph/0111367.
[2266] V. Springel et al., “Simulating the joint evolution of quasars, galaxies and their large-scale distribution,” Nature 435
(2005) 629–636, arXiv:astro-ph/0504097.
[2267] J. Kim, C. Park, R. Gott, III, and J. Dubinski, “The Horizon Run N-body Simulation: Baryon Acoustic Oscillations
and Topology of Large Scale Structure of the Universe,” Astrophys. J. 701 (2009) 1547–1559, arXiv:0812.1392
[astro-ph].
[2268] M. Vogelsberger, S. Genel, V. Springel, P. Torrey, D. Sijacki, D. Xu, G. F. Snyder, D. Nelson, and L. Hernquist,
“Introducing the Illustris Project: Simulating the coevolution of dark and visible matter in the Universe,” Mon. Not.
Roy. Astron. Soc. 444 (2014) no. 2, 1518–1547, arXiv:1405.2921 [astro-ph.CO].
[2269] V. Springel et al., “First results from the IllustrisTNG simulations: matter and galaxy clustering,” Mon. Not. Roy.
Astron. Soc. 475 (2018) no. 1, 676–698, arXiv:1707.03397 [astro-ph.GA].
[2270] A. Klypin, S. Trujillo-Gomez, and J. Primack, “Halos and galaxies in the standard cosmological model: results from
the Bolshoi simulation,” Astrophys. J. 740 (2011) 102, arXiv:1002.3660 [astro-ph.CO].
[2271] J. Schaye et al., “The EAGLE project: Simulating the evolution and assembly of galaxies and their environments,”
Mon. Not. Roy. Astron. Soc. 446 (2015) 521–554, arXiv:1407.7040 [astro-ph.GA].
[2272] R. A. Crain et al., “The EAGLE simulations of galaxy formation: calibration of subgrid physics and model variations,”
Mon. Not. Roy. Astron. Soc. 450 (2015) no. 2, 1937–1961, arXiv:1501.01311 [astro-ph.GA].
[2273] F. Villaescusa-Navarro et al., “The Quijote simulations,” Astrophys. J. Suppl. 250 (2020) no. 1, 2, arXiv:1909.05273
[astro-ph.CO].
[2274] G. Kauffmann, J. M. Colberg, A. Diaferio, and S. D. M. White, “Clustering of galaxies in a hierarchical universe: 1.
Methods and results at z=0,” Mon. Not. Roy. Astron. Soc. 303 (1999) 188–206, arXiv:astro-ph/9805283.
[2275] A. Mead, J. Peacock, C. Heymans, S. Joudaki, and A. Heavens, “An accurate halo model for fitting non-linear
cosmological power spectra and baryonic feedback models,” Mon. Not. Roy. Astron. Soc. 454 (2015) no. 2, 1958–1975,
arXiv:1505.07833 [astro-ph.CO].
[2276] H. J. Hortua, “Constraining cosmological parameters from N-body simulations with Bayesian Neural Networks,” in
35th Conference on Neural Information Processing Systems. 2021. arXiv:2112.11865 [astro-ph.CO].
[2277] A. Lazanu, “Extracting cosmological parameters from N-body simulations using machine learning techniques,” JCAP
09 (2021) 039, arXiv:2106.11061 [astro-ph.CO].
[2278] C. Zhang, L. Zu, H.-Z. Chen, Y.-L. S. Tsai, and Y.-Z. Fan, “Weak lensing constraints on dark matter-baryon
interactions with N-body simulations and machine learning,” JCAP 08 (2024) 003, arXiv:2402.18880 [astro-ph.CO].
[2279] M. R. Lovell, C. S. Frenk, V. R. Eke, A. Jenkins, L. Gao, and T. Theuns, “The properties of warm dark matter
haloes,” Mon. Not. Roy. Astron. Soc. 439 (2014) 300–317, arXiv:1308.1399 [astro-ph.CO].
[2280] S. Bose, W. A. Hellwing, C. S. Frenk, A. Jenkins, M. R. Lovell, J. C. Helly, and B. Li, “The COpernicus COmplexio:
Statistical Properties of Warm Dark Matter Haloes,” Mon. Not. Roy. Astron. Soc. 455 (2016) no. 1, 318–333,
arXiv:1507.01998 [astro-ph.CO].
[2281] Y. Shtanov and V. I. Zhdanov, “Discreteness effects in N-body simulations of warm dark matter,” Phys. Rev. D 109
(2024) no. 6, 063031, arXiv:2307.07778 [astro-ph.CO].
[2282] F. Villaescusa-Navarro, F. Marulli, M. Viel, E. Branchini, E. Castorina, E. Sefusatti, and S. Saito, “Cosmology with
massive neutrinos I: towards a realistic modeling of the relation between matter, haloes and galaxies,” JCAP 03 (2014)
011, arXiv:1311.0866 [astro-ph.CO].
[2283] J. Adamek, R. Durrer, and M. Kunz, “Relativistic N-body simulations with massive neutrinos,” JCAP 11 (2017) 004,
arXiv:1707.06938 [astro-ph.CO].
338

[2284] J. Liu, S. Bird, J. M. Z. Matilla, J. C. Hill, Z. Haiman, M. S. Madhavacheril, A. Petri, and D. N. Spergel,
“MassiveNuS: Cosmological Massive Neutrino Simulations,” JCAP 03 (2018) 049, arXiv:1711.10524 [astro-ph.CO].
[2285] B. Li, G.-B. Zhao, R. Teyssier, and K. Koyama, “ECOSMOG: An Efficient Code for Simulating Modified Gravity,”
JCAP 01 (2012) 051, arXiv:1110.1379 [astro-ph.CO].
[2286] E. Puchwein, M. Baldi, and V. Springel, “Modified Gravity-GADGET: A new code for cosmological hydrodynamical
simulations of modified gravity models,” Mon. Not. Roy. Astron. Soc. 436 (2013) 348, arXiv:1305.2418
[astro-ph.CO].
[2287] J. Zhang, R. An, S. Liao, W. Luo, Z. Li, and B. Wang, “Fully self-consistent cosmological simulation pipeline for
interacting dark energy models,” Phys. Rev. D 98 (2018) no. 10, 103530, arXiv:1811.01519 [astro-ph.CO].
[2288] C. Arnold, M. Leo, and B. Li, “Realistic simulations of galaxy formation in f (R) modified gravity,” Nature Astron. 3
(2019) no. 10, 945–954, arXiv:1907.02977 [astro-ph.CO].
[2289] B. S. Wright, H. A. Winther, and K. Koyama, “COLA with massive neutrinos,” JCAP 10 (2017) 054,
arXiv:1705.08165 [astro-ph.CO].
[2290] Baldi, Marco and Villaescusa-Navarro, Francisco, “Quijote Simulations: Modified Gravity.”
https://quijote-simulations.readthedocs.io/en/latest/mg.html, 2025. Accessed: 2025-01-30.
[2291] X.-l. Chen, S. Hannestad, and R. J. Scherrer, “Cosmic microwave background and large scale structure limits on the
interaction between dark matter and baryons,” Phys. Rev. D 65 (2002) 123515, arXiv:astro-ph/0202496.
[2292] D. V. Nguyen, D. Sarnaaik, K. K. Boddy, E. O. Nadler, and V. Gluscevic, “Observational constraints on dark matter
scattering with electrons,” Phys. Rev. D 104 (2021) no. 10, 103521, arXiv:2107.12380 [astro-ph.CO].
[2293] P. Serra, F. Zalamea, A. Cooray, G. Mangano, and A. Melchiorri, “Constraints on neutrino – dark matter interactions
from cosmic microwave background and large scale structure data,” Phys. Rev. D 81 (2010) 043507, arXiv:0911.4411
[astro-ph.CO].
[2294] G. Mangano, A. Melchiorri, P. Serra, A. Cooray, and M. Kamionkowski, “Cosmological bounds on dark
matter-neutrino interactions,” Phys. Rev. D 74 (2006) 043517, arXiv:astro-ph/0606190.
[2295] R. J. Wilkinson, C. Boehm, and J. Lesgourgues, “Constraining Dark Matter-Neutrino Interactions using the CMB and
Large-Scale Structure,” JCAP 05 (2014) 011, arXiv:1401.7597 [astro-ph.CO].
[2296] F.-Y. Cyr-Racine, K. Sigurdson, J. Zavala, T. Bringmann, M. Vogelsberger, and C. Pfrommer, “ETHOS—an effective
theory of structure formation: From dark particle physics to the matter distribution of the Universe,” Phys. Rev. D 93
(2016) no. 12, 123527, arXiv:1512.05344 [astro-ph.CO].
[2297] M. Vogelsberger, J. Zavala, F.-Y. Cyr-Racine, C. Pfrommer, T. Bringmann, and K. Sigurdson, “ETHOS – an effective
theory of structure formation: dark matter physics as a possible explanation of the small-scale CDM problems,” Mon.
Not. Roy. Astron. Soc. 460 (2016) no. 2, 1399–1416, arXiv:1512.05349 [astro-ph.CO].
[2298] L. Zu, C. Zhang, H.-Z. Chen, W. Wang, Y.-L. S. Tsai, Y. Tsai, W. Luo, and Y.-Z. Fan, “Exploring mirror twin Higgs
cosmology with present and future weak lensing surveys,” JCAP 08 (2023) 023, arXiv:2304.06308 [astro-ph.CO].
[2299] S. Bohr, J. Zavala, F.-Y. Cyr-Racine, M. Vogelsberger, T. Bringmann, and C. Pfrommer, “ETHOS – an effective
parametrization and classification for structure formation: the non-linear regime at z ≳ 5,” Mon. Not. Roy. Astron.
Soc. 498 (2020) no. 3, 3403–3419, arXiv:2006.01842 [astro-ph.CO].
[2300] V. Ajani, A. Peel, V. Pettorino, J.-L. Starck, Z. Li, and J. Liu, “Constraining neutrino masses with weak-lensing
multiscale peak counts,” Phys. Rev. D 102 (2020) no. 10, 103531, arXiv:2001.10993 [astro-ph.CO].
[2301] A. Peel, V. Pettorino, C. Giocoli, J.-L. Starck, and M. Baldi, “Breaking degeneracies in modified gravity with higher
(than 2nd) order weak-lensing statistics,” A& A 619 (2018) A38, arXiv:1805.05146 [astro-ph.CO].
[2302] V. Ajani, J.-L. Starck, and V. Pettorino, “Starlet ℓ1 -norm for weak lensing cosmology,” A& A 645 (2021) L11,
arXiv:2101.01542 [astro-ph.CO].
[2303] J. Harnois-Déraps, N. Martinet, T. Castro, K. Dolag, B. Giblin, C. Heymans, H. Hildebrandt, and Q. Xia, “Cosmic
shear cosmology beyond two-point statistics: a combined peak count and correlation function analysis of DES-Y1,”
MNRAS 506 (2021) no. 2, 1623–1650, arXiv:2012.02777 [astro-ph.CO].
[2304] J. Jasche and B. D. Wandelt, “Bayesian physical reconstruction of initial conditions from large scale structure surveys,”
Mon. Not. Roy. Astron. Soc. 432 (2013) 894, arXiv:1203.3639 [astro-ph.CO].
[2305] J. Jasche and G. Lavaux, “Physical Bayesian modelling of the non-linear matter distribution: new insights into the
Nearby Universe,” Astron. Astrophys. 625 (2019) A64, arXiv:1806.11117 [astro-ph.CO].
[2306] N. Porqueres, A. Heavens, D. Mortlock, and G. Lavaux, “Bayesian forward modelling of cosmic shear data,” Mon. Not.
Roy. Astron. Soc. 502 (2021) no. 2, 3035–3044, arXiv:2011.07722 [astro-ph.CO].
[2307] N. Porqueres, A. Heavens, D. Mortlock, and G. Lavaux, “Lifting weak lensing degeneracies with a field-based
likelihood,” Mon. Not. Roy. Astron. Soc. 509 (2021) no. 3, 3194–3202, arXiv:2108.04825 [astro-ph.CO].
[2308] N. Porqueres, A. Heavens, D. Mortlock, G. Lavaux, and T. L. Makinen, “Field-level inference of cosmic shear with
intrinsic alignments and baryons,” arXiv:2304.04785 [astro-ph.CO].
[2309] T. L. Makinen, C. Sui, B. D. Wandelt, N. Porqueres, and A. Heavens, “Hybrid Summary Statistics,”
arXiv:2410.07548 [stat.ML].
[2310] DES Collaboration, N. Jeffrey et al., “Dark Energy Survey Year 3 results: likelihood-free, simulation-based wCDM
inference with neural compression of weak-lensing map statistics,” Mon. Not. Roy. Astron. Soc. 536 (2025) no. 2,
1303–1322, arXiv:2403.02314 [astro-ph.CO].
[2311] N. MacCrann, J. Zuntz, S. Bridle, B. Jain, and M. R. Becker, “Cosmic Discordance: Are Planck CMB and CFHTLenS
weak lensing measurements out of tune?,” Mon. Not. Roy. Astron. Soc. 451 (2015) no. 3, 2877–2888, arXiv:1408.4742
[astro-ph.CO].
339

[2312] H. Rubira, A. Mazoun, and M. Garny, “Full-shape BOSS constraints on dark matter interacting with dark radiation
and lifting the S8 tension,” JCAP 01 (2023) 034, arXiv:2209.03974 [astro-ph.CO].
[2313] T. Brinckmann, J. H. Chang, P. Du, and M. LoVerde, “Confronting interacting dark radiation scenarios with
cosmological data,” Phys. Rev. D 107 (2023) no. 12, 123517, arXiv:2212.13264 [astro-ph.CO].
[2314] DES Collaboration, A. Chen et al., “Constraints on dark matter to dark radiation conversion in the late universe with
DES-Y1 and external data,” Phys. Rev. D 103 (2021) no. 12, 123528, arXiv:2011.04606 [astro-ph.CO].
[2315] A. He, M. M. Ivanov, R. An, and V. Gluscevic, “S8 Tension in the Context of Dark Matter–Baryon Scattering,”
Astrophys. J. Lett. 954 (2023) no. 1, L8, arXiv:2301.08260 [astro-ph.CO].
[2316] E. Di Valentino, C. Bøehm, E. Hivon, and F. R. Bouchet, “Reducing the H0 and σ8 tensions with Dark
Matter-neutrino interactions,” Phys. Rev. D 97 (2018) no. 4, 043513, arXiv:1710.02559 [astro-ph.CO].
[2317] L. Zu, W. Giarè, C. Zhang, E. Di Valentino, Y.-L. S. Tsai, and S. Trojanowski, “Can νDM interactions solve the S8
discrepancy?,” arXiv:2501.13785 [astro-ph.CO].
[2318] H. Zhan, “The wide-field multiband imaging and slitless spectroscopy survey to be carried out by the survey space
telescope of china manned space program,” Chinese Science Bulletin 66 (2021) 1290–1298.
[2319] R. A. C. Croft, D. H. Weinberg, M. Bolte, S. Burles, L. Hernquist, N. Katz, D. Kirkman, and D. Tytler, “Towards a
precise measurement of matter clustering: Lyman alpha forest data at redshifts 2-4,” Astrophys. J. 581 (2002) 20–52,
arXiv:astro-ph/0012324.
[2320] M. Viel, “The Lyman-α Forest as a Probe of the Coldness of Dark Matter,” in 19th Conference on High Energy
Physics, pp. 255–260. 2008.
[2321] M. Viel, G. D. Becker, J. S. Bolton, and M. G. Haehnelt, “Warm dark matter as a solution to the small scale crisis:
New constraints from high redshift Lyman-α forest data,” Phys. Rev. D 88 (2013) 043502, arXiv:1306.2314
[astro-ph.CO].
[2322] M. Viel, M. G. Haehnelt, and V. Springel, “Inferring the dark matter power spectrum from the Lyman-alpha forest in
high-resolution QSO absorption spectra,” Mon. Not. Roy. Astron. Soc. 354 (2004) 684, arXiv:astro-ph/0404600.
[2323] A. Boyarsky, J. Lesgourgues, O. Ruchayskiy, and M. Viel, “Lyman-alpha constraints on warm and on warm-plus-cold
dark matter models,” JCAP 05 (2009) 012, arXiv:0812.0010 [astro-ph].
[2324] K. K. Rogers and H. V. Peiris, “Strong Bound on Canonical Ultralight Axion Dark Matter from the Lyman-Alpha
Forest,” Phys. Rev. Lett. 126 (2021) no. 7, 071302, arXiv:2007.12705 [astro-ph.CO].
[2325] M. Garny, T. Konstandin, L. Sagunski, and M. Viel, “Neutrino mass bounds from confronting an effective model with
BOSS Lyman-α data,” JCAP 03 (2021) 049, arXiv:2011.03050 [astro-ph.CO].
[2326] N. Palanque-Delabrouille et al., “Neutrino masses and cosmology with Lyman-alpha forest power spectrum,” JCAP 11
(2015) 011, arXiv:1506.05976 [astro-ph.CO].
[2327] J. S. Bolton, A. Caputo, H. Liu, and M. Viel, “Comparison of Low-Redshift Lyman-α Forest Observations to
Hydrodynamical Simulations with Dark Photon Dark Matter,” Phys. Rev. Lett. 129 (2022) no. 21, 211102,
arXiv:2206.13520 [hep-ph].
[2328] D. C. Hooper and M. Lucca, “Hints of dark matter-neutrino interactions in Lyman-α data,” Phys. Rev. D 105 (2022)
no. 10, 103504, arXiv:2110.04024 [astro-ph.CO].
[2329] S. Furlanetto, S. P. Oh, and F. Briggs, “Cosmology at Low Frequencies: The 21 cm Transition and the High-Redshift
Universe,” Phys. Rept. 433 (2006) 181–301, arXiv:astro-ph/0608032.
[2330] J. R. Pritchard and A. Loeb, “21-cm cosmology,” Rept. Prog. Phys. 75 (2012) 086901, arXiv:1109.6012
[astro-ph.CO].
[2331] A. Schneider, “Constraining noncold dark matter models with the global 21-cm signal,” Phys. Rev. D 98 (2018) no. 6,
063021, arXiv:1805.00021 [astro-ph.CO].
[2332] L. Lopez-Honorez, O. Mena, and P. Villanueva-Domingo, “Dark matter microphysics and 21 cm observations,” Phys.
Rev. D 99 (2019) no. 2, 023522, arXiv:1811.02716 [astro-ph.CO].
[2333] M. Escudero, L. Lopez-Honorez, O. Mena, S. Palomares-Ruiz, and P. Villanueva-Domingo, “A fresh look into the
interacting dark matter scenario,” JCAP 06 (2018) 007, arXiv:1803.08427 [astro-ph.CO].
[2334] J. B. Muñoz, C. Dvorkin, and F.-Y. Cyr-Racine, “Probing the Small-Scale Matter Power Spectrum with Large-Scale
21-cm Data,” Phys. Rev. D 101 (2020) no. 6, 063526, arXiv:1911.11144 [astro-ph.CO].
[2335] S. Yoshiura, K. Takahashi, and T. Takahashi, “Probing Small Scale Primordial Power Spectrum with 21cm Line Global
Signal,” Phys. Rev. D 101 (2020) no. 8, 083520, arXiv:1911.07442 [astro-ph.CO].
[2336] J. B. Muñoz, S. Bohr, F.-Y. Cyr-Racine, J. Zavala, and M. Vogelsberger, “ETHOS - an effective theory of structure
formation: Impact of dark acoustic oscillations on cosmic dawn,” Phys. Rev. D 103 (2021) no. 4, 043512,
arXiv:2011.05333 [astro-ph.CO].
[2337] M. R. Lovell, J. Zavala, M. Vogelsberger, X. Shen, F.-Y. Cyr-Racine, C. Pfrommer, K. Sigurdson, M. Boylan-Kolchin,
and A. Pillepich, “ETHOS – an effective theory of structure formation: predictions for the high-redshift Universe –
abundance of galaxies and reionization,” Mon. Not. Roy. Astron. Soc. 477 (2018) no. 3, 2886–2899, arXiv:1711.10497
[astro-ph.CO].
[2338] S. Hassan, R. Davé, K. Finlator, and M. G. Santos, “Simulating the 21 cm signal from reionization including non-linear
ionizations and inhomogeneous recombinations,” Mon. Not. Roy. Astron. Soc. 457 (2016) no. 2, 1550–1567,
arXiv:1510.04280 [astro-ph.CO].
[2339] LOFAR Collaboration, M. P. van Haarlem et al., “LOFAR: The LOw-Frequency ARray,” Astron. Astrophys. 556
(2013) A2, arXiv:1305.3550 [astro-ph.IM].
[2340] D. R. DeBoer et al., “Hydrogen Epoch of Reionization Array (HERA),” Publ. Astron. Soc. Pac. 129 (2017) no. 974,
340

045001, arXiv:1606.07473 [astro-ph.IM].


[2341] L. V. E. Koopmans et al., “The Cosmic Dawn and Epoch of Reionization with the Square Kilometre Array,” PoS
AASKA14 (2015) 001, arXiv:1505.07568 [astro-ph.CO].
[2342] J. S. Bullock and M. Boylan-Kolchin, “Small-Scale Challenges to the ΛCDM Paradigm,” Ann. Rev. Astron. Astrophys.
55 (2017) 343–387, arXiv:1707.04256 [astro-ph.CO].
[2343] J. F. Navarro, C. S. Frenk, and S. D. M. White, “The Structure of cold dark matter halos,” Astrophys. J. 462 (1996)
563–575, arXiv:astro-ph/9508025.
[2344] W. J. G. de Blok, “The Core-Cusp Problem,” Adv. Astron. 2010 (2010) 789293, arXiv:0910.3538 [astro-ph.CO].
[2345] G. Gentile, P. Salucci, U. Klein, D. Vergani, and P. Kalberla, “The Cored distribution of dark matter in spiral
galaxies,” Mon. Not. Roy. Astron. Soc. 351 (2004) 903, arXiv:astro-ph/0403154.
[2346] J. S. Bullock, T. S. Kolatt, Y. Sigad, R. S. Somerville, A. V. Kravtsov, A. A. Klypin, J. R. Primack, and A. Dekel,
“Profiles of dark haloes. Evolution, scatter, and environment,” Mon. Not. Roy. Astron. Soc. 321 (2001) 559–575,
arXiv:astro-ph/9908159.
[2347] K. A. Oman et al., “The unexpected diversity of dwarf galaxy rotation curves,” Mon. Not. Roy. Astron. Soc. 452
(2015) no. 4, 3650–3665, arXiv:1504.01437 [astro-ph.GA].
[2348] A. A. Klypin, A. V. Kravtsov, O. Valenzuela, and F. Prada, “Where are the missing Galactic satellites?,” Astrophys. J.
522 (1999) 82–92, arXiv:astro-ph/9901240.
[2349] B. Moore, S. Ghigna, F. Governato, G. Lake, T. R. Quinn, J. Stadel, and P. Tozzi, “Dark matter substructure within
galactic halos,” Astrophys. J. Lett. 524 (1999) L19–L22, arXiv:astro-ph/9907411.
[2350] M. Boylan-Kolchin, J. S. Bullock, and M. Kaplinghat, “Too big to fail? The puzzling darkness of massive Milky Way
subhaloes,” Mon. Not. Roy. Astron. Soc. 415 (2011) L40, arXiv:1103.0007 [astro-ph.CO].
[2351] M. Boylan-Kolchin, J. S. Bullock, and M. Kaplinghat, “The Milky Way’s bright satellites as an apparent failure of
LCDM,” Mon. Not. Roy. Astron. Soc. 422 (2012) 1203–1218, arXiv:1111.2048 [astro-ph.CO].
[2352] J. F. Navarro, V. R. Eke, and C. S. Frenk, “The cores of dwarf galaxy halos,” Mon. Not. Roy. Astron. Soc. 283 (1996)
L72–L78, arXiv:astro-ph/9610187.
[2353] S.-H. Oh, C. Brook, F. Governato, E. Brinks, L. Mayer, W. J. G. de Blok, A. Brooks, and F. Walter, “The central
slope of dark matter cores in dwarf galaxies: Simulations vs. THINGS,” Astron. J. 142 (2011) 24, arXiv:1011.2777
[astro-ph.CO].
[2354] A. Zolotov, A. M. Brooks, B. Willman, F. Governato, A. Pontzen, C. Christensen, A. Dekel, T. Quinn, S. Shen, and
J. Wadsley, “Baryons Matter: Why Luminous Satellite Galaxies Have Reduced Central Masses,” Astrophys. J. 761
(2012) 71, arXiv:1207.0007 [astro-ph.CO].
[2355] K. S. Arraki, A. Klypin, S. More, and S. Trujillo-Gomez, “Effects of baryon removal on the structure of dwarf
spheroidal galaxies,” Mon. Not. Roy. Astron. Soc. 438 (2014) no. 2, 1466–1482, arXiv:1212.6651 [astro-ph.CO].
[2356] A. M. Brooks and A. Zolotov, “Why Baryons Matter: The Kinematics of Dwarf Spheroidal Satellites,” Astrophys. J.
786 (2014) 87, arXiv:1207.2468 [astro-ph.CO].
[2357] C. B. Brook and A. Di Cintio, “Expanded haloes, abundance matching and too-big-to-fail in the Local Group,” Mon.
Not. Roy. Astron. Soc. 450 (2015) no. 4, 3920–3934, arXiv:1410.3825 [astro-ph.GA].
[2358] A. A. Dutton, A. V. Macciò, J. Frings, L. Wang, G. S. Stinson, C. Penzo, and X. Kang, “NIHAO V: too big does not
fail – reconciling the conflict between ACDM predictions and the circular velocities of nearby field galaxies,” Mon. Not.
Roy. Astron. Soc. 457 (2016) no. 1, L74–L78, arXiv:1512.00453 [astro-ph.GA].
[2359] W. Hu, R. Barkana, and A. Gruzinov, “Cold and fuzzy dark matter,” Phys. Rev. Lett. 85 (2000) 1158–1161,
arXiv:astro-ph/0003365.
[2360] B. Moore, S. Gelato, A. Jenkins, F. R. Pearce, and V. Quilis, “Collisional versus collisionless dark matter,” Astrophys.
J. Lett. 535 (2000) L21–L24, arXiv:astro-ph/0002308.
[2361] N. Yoshida, V. Springel, S. D. M. White, and G. Tormen, “Collisional dark matter and the structure of dark halos,”
Astrophys. J. Lett. 535 (2000) L103, arXiv:astro-ph/0002362.
[2362] A. Burkert, “The Structure and evolution of weakly selfinteracting cold dark matter halos,” Astrophys. J. Lett. 534
(2000) L143–L146, arXiv:astro-ph/0002409.
[2363] C. S. Kochanek and M. J. White, “A Quantitative study of interacting dark matter in halos,” Astrophys. J. 543 (2000)
514, arXiv:astro-ph/0003483.
[2364] N. Yoshida, V. Springel, S. D. M. White, and G. Tormen, “Weakly self-interacting dark matter and the structure of
dark halos,” Astrophys. J. Lett. 544 (2000) L87–L90, arXiv:astro-ph/0006134.
[2365] R. Dave, D. N. Spergel, P. J. Steinhardt, and B. D. Wandelt, “Halo properties in cosmological simulations of
selfinteracting cold dark matter,” Astrophys. J. 547 (2001) 574–589, arXiv:astro-ph/0006218.
[2366] J. Miralda-Escude, “A test of the collisional dark matter hypothesis from cluster lensing,” Astrophys. J. 564 (2002) 60,
arXiv:astro-ph/0002050.
[2367] M. Rocha, A. H. G. Peter, J. S. Bullock, M. Kaplinghat, S. Garrison-Kimmel, J. Onorbe, and L. A. Moustakas,
“Cosmological Simulations with Self-Interacting Dark Matter I: Constant Density Cores and Substructure,” Mon. Not.
Roy. Astron. Soc. 430 (2013) 81–104, arXiv:1208.3025 [astro-ph.CO].
[2368] A. H. G. Peter, M. Rocha, J. S. Bullock, and M. Kaplinghat, “Cosmological Simulations with Self-Interacting Dark
Matter II: Halo Shapes vs. Observations,” Mon. Not. Roy. Astron. Soc. 430 (2013) 105, arXiv:1208.3026
[astro-ph.CO].
[2369] M. Vogelsberger, J. Zavala, and A. Loeb, “Subhaloes in Self-Interacting Galactic Dark Matter Haloes,” Mon. Not. Roy.
Astron. Soc. 423 (2012) 3740, arXiv:1201.5892 [astro-ph.CO].
341

[2370] O. D. Elbert, J. S. Bullock, S. Garrison-Kimmel, M. Rocha, J. Oñorbe, and A. H. G. Peter, “Core formation in dwarf
haloes with self-interacting dark matter: no fine-tuning necessary,” Mon. Not. Roy. Astron. Soc. 453 (2015) no. 1,
29–37, arXiv:1412.1477 [astro-ph.GA].
[2371] H.-Y. Schive, T. Chiueh, and T. Broadhurst, “Cosmic Structure as the Quantum Interference of a Coherent Dark
Wave,” Nature Phys. 10 (2014) 496–499, arXiv:1406.6586 [astro-ph.GA].
[2372] P. Mocz, M. Vogelsberger, V. H. Robles, J. Zavala, M. Boylan-Kolchin, A. Fialkov, and L. Hernquist, “Galaxy
formation with BECDM – I. Turbulence and relaxation of idealized haloes,” Mon. Not. Roy. Astron. Soc. 471 (2017)
no. 4, 4559–4570, arXiv:1705.05845 [astro-ph.CO].
[2373] J. Veltmaat, J. C. Niemeyer, and B. Schwabe, “Formation and structure of ultralight bosonic dark matter halos,” Phys.
Rev. D 98 (2018) no. 4, 043509, arXiv:1804.09647 [astro-ph.CO].
[2374] L. Herold, E. G. M. Ferreira, and L. Heinrich, “Profile Likelihoods in Cosmology: When, Why and How illustrated
with ΛCDM, Massive Neutrinos and Dark Energy,” arXiv:2408.07700 [astro-ph.CO].
[2375] J. Neyman, “Outline of a Theory of Statistical Estimation Based on the Classical Theory of Probability,” Phil. Trans.
Roy. Soc. Lond. A 236 (1937) no. 767, 333–380.
[2376] S. S. Wilks, “The Large-Sample Distribution of the Likelihood Ratio for Testing Composite Hypotheses,” Annals Math.
Statist. 9 (1938) no. 1, 60–62.
[2377] Y. Pawitan, In All Likelihood: Statistical Modelling and Inference Using Likelihood. Oxford science publications. OUP
Oxford, 2001. https://books.google.com/books?id=M-3pSCVxV5oC.
[2378] R. Trotta, “Bayesian Methods in Cosmology,” arXiv e-prints (2017) arXiv:1701.01467, arXiv:1701.01467
[astro-ph.CO].
[2379] G. J. Feldman and R. D. Cousins, “A Unified approach to the classical statistical analysis of small signals,” Phys. Rev.
D 57 (1998) 3873–3889, arXiv:physics/9711021.
[2380] R. D. Cousins, “Why isn’t every physicist a Bayesian?,” Am. J. Phys. 63 (1995) 398.
[2381] LiteBIRD Collaboration, P. Campeti et al., “LiteBIRD science goals and forecasts. A case study of the origin of
primordial gravitational waves using large-scale CMB polarization,” JCAP 06 (2024) 008, arXiv:2312.00717
[astro-ph.CO].
[2382] SPIDER Collaboration, P. A. R. Ade et al., “A Constraint on Primordial B-modes from the First Flight of the Spider
Balloon-borne Telescope,” Astrophys. J. 927 (2022) no. 2, 174, arXiv:2103.13334 [astro-ph.CO].
[2383] B. A. Reid, L. Verde, R. Jimenez, and O. Mena, “Robust Neutrino Constraints by Combining Low Redshift
Observations with the CMB,” JCAP 01 (2010) 003, arXiv:0910.0008 [astro-ph.CO].
[2384] W. Giarè, A. Gómez-Valent, E. Di Valentino, and C. van de Bruck, “Hints of neutrino dark matter scattering in the
CMB? Constraints from the marginalized and profile distributions,” Phys. Rev. D 109 (2024) no. 6, 063516,
arXiv:2311.09116 [astro-ph.CO].
[2385] T. Karwal, Y. Patel, A. Bartlett, V. Poulin, T. L. Smith, and D. N. Pfeffer, “Procoli: Profiles of cosmological
likelihoods,” arXiv:2401.14225 [astro-ph.CO].
[2386] E. B. Holm, A. Nygaard, J. Dakin, S. Hannestad, and T. Tram, “PROSPECT: a profile likelihood code for frequentist
cosmological parameter inference,” Mon. Not. Roy. Astron. Soc. 535 (2024) no. 4, 3686–3699, arXiv:2312.02972
[astro-ph.CO].
[2387] F. James and M. Roos, “Minuit - a system for function minimization and analysis of the parameter errors and
correlations,” Computer Physics Communications 10 (1975) no. 6, 343–367.
https://www.sciencedirect.com/science/article/pii/0010465575900399.
[2388] G. Galloni, S. Henrot-Versillé, and M. Tristram, “Robust constraints on tensor perturbations from cosmological data:
A comparative analysis from Bayesian and frequentist perspectives,” Phys. Rev. D 110 (2024) no. 6, 063511,
arXiv:2405.04455 [astro-ph.CO].
[2389] C. Moretti, M. Tsedrik, P. Carrilho, and A. Pourtsidou, “Modified gravity and massive neutrinos: constraints from the
full shape analysis of BOSS galaxies and forecasts for Stage IV surveys,” JCAP 12 (2023) 025, arXiv:2306.09275
[astro-ph.CO].
[2390] P. Campeti, O. Özsoy, I. Obata, and M. Shiraishi, “New constraints on axion-gauge field dynamics during inflation
from Planck and BICEP/Keck data sets,” JCAP 07 (2022) no. 07, 039, arXiv:2203.03401 [astro-ph.CO].
[2391] P. Campeti and E. Komatsu, “New Constraint on the Tensor-to-scalar Ratio from the Planck and BICEP/Keck Array
Data Using the Profile Likelihood,” Astrophys. J. 941 (2022) no. 2, 110, arXiv:2205.05617 [astro-ph.CO].
[2392] Planck Collaboration, P. A. R. Ade et al., “Planck intermediate results. XVI. Profile likelihoods for cosmological
parameters,” Astron. Astrophys. 566 (2014) A54, arXiv:1311.1657 [astro-ph.CO].
[2393] S. Hannestad, “Stochastic optimization methods for extracting cosmological parameters from cosmic microwave
background radiation power spectra,” Phys. Rev. D 61 (2000) 023002, arXiv:astro-ph/9911330.
[2394] S. Kirkpatrick, C. D. Gelatt, and M. P. Vecchi, “Optimization by Simulated Annealing,” Science 220 (1983) 671–680.
[2395] S. Henrot-Versillé, O. Perdereau, S. Plaszczynski, B. R. d’Orfeuil, M. Spinelli, and M. Tristram, “Agnostic cosmology
in the CAMEL framework,” arXiv:1607.02964 [astro-ph.CO].
[2396] A. Nygaard, E. B. Holm, S. Hannestad, and T. Tram, “CONNECT: a neural network based framework for emulating
cosmological observables and cosmological parameter inference,” JCAP 05 (2023) 025, arXiv:2205.15726
[astro-ph.IM].
[2397] A. Nygaard, E. B. Holm, S. Hannestad, and T. Tram, “Fast and effortless computation of profile likelihoods using
CONNECT,” JCAP 11 (2023) 064, arXiv:2308.06379 [astro-ph.CO].
[2398] T. Bringmann, F. Kahlhoefer, K. Schmidt-Hoberg, and P. Walia, “Converting nonrelativistic dark matter to radiation,”
342

Phys. Rev. D 98 (2018) no. 2, 023543, arXiv:1803.03644 [astro-ph.CO].


[2399] F. Couchot, S. Henrot-Versillé, O. Perdereau, S. Plaszczynski, B. Rouillé d’Orfeuil, M. Spinelli, and M. Tristram,
“Cosmological constraints on the neutrino mass including systematic uncertainties,” Astron. Astrophys. 606 (2017)
A104, arXiv:1703.10829 [astro-ph.CO].
[2400] A. X. Gonzalez-Morales, R. Poltis, B. D. Sherwin, and L. Verde, “Are priors responsible for cosmology favoring
additional neutrino species?,” arXiv:1106.5052 [astro-ph.CO].
[2401] L. Herold and M. Kamionkowski, “Revisiting the impact of neutrino mass hierarchies on neutrino mass constraints in
light of recent DESI data,” arXiv:2412.03546 [astro-ph.CO].
[2402] S. Henrot-Versille et al., “Improved constraint on the primordial gravitational-wave density using recent cosmological
data and its impact on cosmic string models,” Class. Quant. Grav. 32 (2015) no. 4, 045003, arXiv:1408.5299
[astro-ph.CO].
[2403] A. a. J. S. Capistrano, R. C. Nunes, and L. A. Cabral, “Lower tensor-to-scalar ratio as possible signature of modified
gravity,” Phys. Rev. D 109 (2024) no. 12, 123517, arXiv:2403.13860 [gr-qc].
[2404] B. Hu and J. Torrado, “Searching for primordial localized features with CMB and LSS spectra,” Phys. Rev. D 91
(2015) no. 6, 064039, arXiv:1410.4804 [astro-ph.CO].
[2405] A. Avilez and C. Skordis, “Cosmological constraints on Brans-Dicke theory,” Phys. Rev. Lett. 113 (2014) no. 1, 011101,
arXiv:1303.4330 [astro-ph.CO].
[2406] J. de Cruz Pérez and J. Solà Peracaula, “Brans–Dicke cosmology mimicking running vacuum,” Mod. Phys. Lett. A 33
(2018) no. 38, 1850228, arXiv:1809.03329 [gr-qc].
[2407] J. Solà Peracaula, A. Gomez-Valent, J. de Cruz Pérez, and C. Moreno-Pulido, “Brans–Dicke Gravity with a
Cosmological Constant Smoothes Out ΛCDM Tensions,” Astrophys. J. Lett. 886 (2019) no. 1, L6, arXiv:1909.02554
[astro-ph.CO].
[2408] J. Solà Peracaula, A. Gómez-Valent, J. de Cruz Pérez, and C. Moreno-Pulido, “Brans–Dicke cosmology with a Λ-term:
a possible solution to ΛCDM tensions,” Class. Quant. Grav. 37 (2020) no. 24, 245003, arXiv:2006.04273
[astro-ph.CO].
[2409] S. Joudaki, P. G. Ferreira, N. A. Lima, and H. A. Winther, “Testing gravity on cosmic scales: A case study of
Jordan-Brans-Dicke theory,” Phys. Rev. D 105 (2022) no. 4, 043522, arXiv:2010.15278 [astro-ph.CO].
[2410] C. Wetterich, “The Cosmon model for an asymptotically vanishing time dependent cosmological ’constant’,” Astron.
Astrophys. 301 (1995) 321–328, arXiv:hep-th/9408025.
[2411] V. Pettorino, L. Amendola, C. Baccigalupi, and C. Quercellini, “Constraints on coupled dark energy using CMB data
from WMAP and SPT,” Phys. Rev. D 86 (2012) 103507, arXiv:1207.3293 [astro-ph.CO].
[2412] V. Pettorino, “Testing modified gravity with Planck: the case of coupled dark energy,” Phys. Rev. D 88 (2013) 063519,
arXiv:1305.7457 [astro-ph.CO].
[2413] Planck Collaboration, P. A. R. Ade et al., “Planck 2015 results. XIV. Dark energy and modified gravity,” Astron.
Astrophys. 594 (2016) A14, arXiv:1502.01590 [astro-ph.CO].
[2414] B. J. Barros, L. Amendola, T. Barreiro, and N. J. Nunes, “Coupled quintessence with a ΛCDM background: removing
the σ8 tension,” JCAP 01 (2019) 007, arXiv:1802.09216 [astro-ph.CO].
[2415] A. Gómez-Valent, V. Pettorino, and L. Amendola, “Update on coupled dark energy and the H0 tension,” Phys. Rev. D
101 (2020) no. 12, 123513, arXiv:2004.00610 [astro-ph.CO].
[2416] A. Gómez-Valent, Z. Zheng, L. Amendola, C. Wetterich, and V. Pettorino, “Coupled and uncoupled early dark energy,
massive neutrinos, and the cosmological tensions,” Phys. Rev. D 106 (2022) no. 10, 103522, arXiv:2207.14487
[astro-ph.CO].
[2417] L. W. K. Goh, A. Gómez-Valent, V. Pettorino, and M. Kilbinger, “Constraining constant and tomographic coupled
dark energy with low-redshift and high-redshift probes,” Phys. Rev. D 107 (2023) no. 8, 083503, arXiv:2211.13588
[astro-ph.CO].
[2418] P. J. E. Peebles and B. Ratra, “Cosmology with a Time Variable Cosmological Constant,” Astrophys. J. Lett. 325
(1988) L17.
[2419] P. Brax, C. van de Bruck, E. Di Valentino, W. Giarè, and S. Trojanowski, “New insights on ν–DM interactions,” Mon.
Not. Roy. Astron. Soc. 527 (2023) no. 1, L122–L126, arXiv:2303.16895 [astro-ph.CO].
[2420] P. Brax, C. van de Bruck, E. Di Valentino, W. Giarè, and S. Trojanowski, “Extended analysis of neutrino-dark matter
interactions with small-scale CMB experiments,” Phys. Dark Univ. 42 (2023) 101321, arXiv:2305.01383
[astro-ph.CO].
[2421] BOSS Collaboration, L. Anderson et al., “The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic
Survey: Baryon Acoustic Oscillations in the Data Release 9 Spectroscopic Galaxy Sample,” Mon. Not. Roy. Astron.
Soc. 427 (2013) no. 4, 3435–3467, arXiv:1203.6594 [astro-ph.CO].
[2422] eBOSS Collaboration, M. Ata et al., “The clustering of the SDSS-IV extended Baryon Oscillation Spectroscopic
Survey DR14 quasar sample: first measurement of baryon acoustic oscillations between redshift 0.8 and 2.2,” Mon.
Not. Roy. Astron. Soc. 473 (2018) no. 4, 4773–4794, arXiv:1705.06373 [astro-ph.CO].
[2423] DES Collaboration, T. M. C. Abbott et al., “Dark Energy Survey Year 1 Results: Measurement of the Baryon
Acoustic Oscillation scale in the distribution of galaxies to redshift 1,” Mon. Not. Roy. Astron. Soc. 483 (2019) no. 4,
4866–4883, arXiv:1712.06209 [astro-ph.CO].
[2424] DES Collaboration, K. C. Chan et al., “BAO from Angular Clustering: Optimization and Mitigation of Theoretical
Systematics,” Mon. Not. Roy. Astron. Soc. 480 (2018) no. 3, 3031–3051, arXiv:1801.04390 [astro-ph.CO].
[2425] R. Ruggeri and C. Blake, “How accurately can we measure the baryon acoustic oscillation feature?,” Mon. Not. Roy.
343

Astron. Soc. 498 (2020) no. 3, 3744–3757, arXiv:1909.13011 [astro-ph.CO].


[2426] A. Cuceu, A. Font-Ribera, and B. Joachimi, “Bayesian methods for fitting Baryon Acoustic Oscillations in the
Lyman-α forest,” JCAP 07 (2020) 035, arXiv:2004.02761 [astro-ph.CO].
[2427] A. G. Sanchez, “Arguments against using h−1 Mpc units in observational cosmology,” Phys. Rev. D 102 (2020) no. 12,
123511, arXiv:2002.07829 [astro-ph.CO].
[2428] L. F. Secco, T. Karwal, W. Hu, and E. Krause, “Role of the Hubble scale in the weak lensing versus CMB tension,”
Phys. Rev. D 107 (2023) no. 8, 083532, arXiv:2209.12997 [astro-ph.CO].
[2429] M. Forconi, A. Favale, and A. Gómez-Valent, “Illustrating the consequences of a misuse of σ8 in cosmology,”
arXiv:2501.11571 [astro-ph.CO].
[2430] V. Poulin, T. L. Smith, R. Calderón, and T. Simon, “On the implications of the ‘cosmic calibration tension’ beyond H0
and the synergy between early- and late-time new physics,” arXiv:2407.18292 [astro-ph.CO].
[2431] D. Pedrotti, J.-Q. Jiang, L. A. Escamilla, S. S. da Costa, and S. Vagnozzi, “Multidimensionality of the Hubble tension:
The roles of Ωm and ωc,” Phys. Rev. D 111 (2025) no. 2, 023506, arXiv:2408.04530 [astro-ph.CO].
[2432] M. Kamionkowski and A. G. Riess, “The Hubble Tension and Early Dark Energy,” Ann. Rev. Nucl. Part. Sci. 73
(2023) 153–180, arXiv:2211.04492 [astro-ph.CO].
[2433] V. Poulin, T. L. Smith, and T. Karwal, “The Ups and Downs of Early Dark Energy solutions to the Hubble tension: A
review of models, hints and constraints circa 2023,” Phys. Dark Univ. 42 (2023) 101348, arXiv:2302.09032
[astro-ph.CO].
[2434] E. McDonough, J. C. Hill, M. M. Ivanov, A. La Posta, and M. W. Toomey, “Observational constraints on early dark
energy,” Int. J. Mod. Phys. D 33 (2024) no. 11, 2430003, arXiv:2310.19899 [astro-ph.CO].
[2435] A. Albrecht and C. Skordis, “Phenomenology of a realistic accelerating universe using only Planck scale physics,” Phys.
Rev. Lett. 84 (2000) 2076–2079, arXiv:astro-ph/9908085.
[2436] M. Doran, M. J. Lilley, J. Schwindt, and C. Wetterich, “Quintessence and the separation of CMB peaks,” Astrophys. J.
559 (2001) 501–506, arXiv:astro-ph/0012139.
[2437] C. Wetterich, “Phenomenological parameterization of quintessence,” Phys. Lett. B 594 (2004) 17–22,
arXiv:astro-ph/0403289.
[2438] M. Doran and G. Robbers, “Early dark energy cosmologies,” JCAP 06 (2006) 026, arXiv:astro-ph/0601544.
[2439] M. Kamionkowski, J. Pradler, and D. G. E. Walker, “Dark energy from the string axiverse,” Phys. Rev. Lett. 113
(2014) no. 25, 251302, arXiv:1409.0549 [hep-ph].
[2440] T. Karwal and M. Kamionkowski, “Dark energy at early times, the Hubble parameter, and the string axiverse,” Phys.
Rev. D 94 (2016) no. 10, 103523, arXiv:1608.01309 [astro-ph.CO].
[2441] V. Poulin, T. L. Smith, D. Grin, T. Karwal, and M. Kamionkowski, “Cosmological implications of ultralight axionlike
fields,” Phys. Rev. D 98 (2018) no. 8, 083525, arXiv:1806.10608 [astro-ph.CO].
[2442] V. Poulin, T. L. Smith, T. Karwal, and M. Kamionkowski, “Early Dark Energy Can Resolve The Hubble Tension,”
Phys. Rev. Lett. 122 (2019) no. 22, 221301, arXiv:1811.04083 [astro-ph.CO].
[2443] T. L. Smith, V. Poulin, and M. A. Amin, “Oscillating scalar fields and the Hubble tension: a resolution with novel
signatures,” Phys. Rev. D 101 (2020) no. 6, 063523, arXiv:1908.06995 [astro-ph.CO].
[2444] A. R. Khalife, M. B. Zanjani, S. Galli, S. Günther, J. Lesgourgues, and K. Benabed, “Review of Hubble tension
solutions with new SH0ES and SPT-3G data,” JCAP 04 (2024) 059, arXiv:2312.09814 [astro-ph.CO].
[2445] J. C. Hill, E. McDonough, M. W. Toomey, and S. Alexander, “Early dark energy does not restore cosmological
concordance,” Phys. Rev. D 102 (2020) no. 4, 043507, arXiv:2003.07355 [astro-ph.CO].
[2446] V. Poulin, T. L. Smith, and A. Bartlett, “Dark energy at early times and ACT data: A larger Hubble constant without
late-time priors,” Phys. Rev. D 104 (2021) no. 12, 123550, arXiv:2109.06229 [astro-ph.CO].
[2447] E. McDonough, M.-X. Lin, J. C. Hill, W. Hu, and S. Zhou, “Early dark sector, the Hubble tension, and the
swampland,” Phys. Rev. D 106 (2022) no. 4, 043525, arXiv:2112.09128 [astro-ph.CO].
[2448] S. Alexander and E. McDonough, “Axion-Dilaton Destabilization and the Hubble Tension,” Phys. Lett. B 797 (2019)
134830, arXiv:1904.08912 [astro-ph.CO].
[2449] M. Cicoli, M. Licheri, R. Mahanta, E. McDonough, F. G. Pedro, and M. Scalisi, “Early Dark Energy in Type IIB
String Theory,” JHEP 06 (2023) 052, arXiv:2303.03414 [hep-th].
[2450] E. McDonough and M. Scalisi, “Towards Early Dark Energy in string theory,” JHEP 10 (2023) 118,
arXiv:2209.00011 [hep-th].
[2451] R. Kappl, H. P. Nilles, and M. W. Winkler, “Modulated Natural Inflation,” Phys. Lett. B 753 (2016) 653–659,
arXiv:1511.05560 [hep-th].
[2452] L. Yin, “Reducing the H0 tension with exponential acoustic dark energy,” Eur. Phys. J. C 82 (2022) no. 1, 78,
arXiv:2012.13917 [astro-ph.CO].
[2453] G. Ye and Y.-S. Piao, “Is the Hubble tension a hint of AdS phase around recombination?,” Phys. Rev. D 101 (2020)
no. 8, 083507, arXiv:2001.02451 [astro-ph.CO].
[2454] A. Adil, A. Albrecht, and L. Knox, “Quintessential cosmological tensions,” Phys. Rev. D 107 (2023) no. 6, 063521,
arXiv:2207.10235 [astro-ph.CO].
[2455] P. Agrawal, F.-Y. Cyr-Racine, D. Pinner, and L. Randall, “Rock ‘n’ roll solutions to the Hubble tension,” Phys. Dark
Univ. 42 (2023) 101347, arXiv:1904.01016 [astro-ph.CO].
[2456] M.-X. Lin, G. Benevento, W. Hu, and M. Raveri, “Acoustic Dark Energy: Potential Conversion of the Hubble
Tension,” Phys. Rev. D 100 (2019) no. 6, 063542, arXiv:1905.12618 [astro-ph.CO].
[2457] R. Jackiw and S. Y. Pi, “Chern-Simons modification of general relativity,” Phys. Rev. D 68 (2003) 104012,
344

arXiv:gr-qc/0308071.
[2458] M. J. Duncan, N. Kaloper, and K. A. Olive, “Axion hair and dynamical torsion from anomalies,” Nucl. Phys. B 387
(1992) 215–235.
[2459] S. Basilakos, N. E. Mavromatos, and J. Solà Peracaula, “Gravitational and Chiral Anomalies in the Running Vacuum
Universe and Matter-Antimatter Asymmetry,” Phys. Rev. D 101 (2020) no. 4, 045001, arXiv:1907.04890 [hep-ph].
[2460] S. Basilakos, N. E. Mavromatos, and J. Solà Peracaula, “Quantum Anomalies in String-Inspired Running Vacuum
Universe: Inflation and Axion Dark Matter,” Phys. Lett. B 803 (2020) 135342, arXiv:2001.03465 [gr-qc].
[2461] N. E. Mavromatos and J. Solà Peracaula, “Stringy-running-vacuum-model inflation: from primordial gravitational
waves and stiff axion matter to dynamical dark energy,” Eur. Phys. J. ST 230 (2021) no. 9, 2077–2110,
arXiv:2012.07971 [hep-ph].
[2462] E. Guendelman, R. Herrera, and D. Benisty, “Unifying inflation with early and late dark energy with multiple fields:
Spontaneously broken scale invariant two measures theory,” Phys. Rev. D 105 (2022) no. 12, 124035,
arXiv:2201.06470 [gr-qc].
[2463] E. L. D. Perico, J. A. S. Lima, S. Basilakos, and J. Sola, “Complete Cosmic History with a dynamical Λ = Λ(H) term,”
Phys. Rev. D 88 (2013) 063531, arXiv:1306.0591 [astro-ph.CO].
[2464] J. A. S. Lima, S. Basilakos, and J. Sola, “Expansion History with Decaying Vacuum: A Complete Cosmological
Scenario,” Mon. Not. Roy. Astron. Soc. 431 (2013) 923–929, arXiv:1209.2802 [gr-qc].
[2465] J. Solà and A. Gómez-Valent, “The Λ̄CDM cosmology: From inflation to dark energy through running Λ,” Int. J. Mod.
Phys. D 24 (2015) 1541003, arXiv:1501.03832 [gr-qc].
[2466] J. Solà Peracaula and H. Yu, “Particle and entropy production in the Running Vacuum Universe,” Gen. Rel. Grav. 52
(2020) no. 2, 17, arXiv:1910.01638 [gr-qc].
[2467] N. E. Mavromatos and J. Solà Peracaula, “Inflationary physics and trans-Planckian conjecture in the stringy running
vacuum model: from the phantom vacuum to the true vacuum,” Eur. Phys. J. Plus 136 (2021) no. 11, 1152,
arXiv:2105.02659 [hep-th].
[2468] P. Svrcek and E. Witten, “Axions In String Theory,” JHEP 06 (2006) 051, arXiv:hep-th/0605206.
[2469] P. Dorlis, N. E. Mavromatos, and S.-N. Vlachos, “Condensate-induced inflation from primordial gravitational waves in
string-inspired Chern-Simons gravity,” Phys. Rev. D 110 (2024) no. 6, 063512, arXiv:2403.09005 [gr-qc].
[2470] N. E. Mavromatos, P. Dorlis, and S.-N. Vlachos, “Torsion-induced axions in string theory, quantum gravity and the
cosmological tensions,” in Workshop on the Standard Model and Beyond. 2024. arXiv:2404.18741 [gr-qc].
[2471] L. McAllister, E. Silverstein, and A. Westphal, “Gravity Waves and Linear Inflation from Axion Monodromy,” Phys.
Rev. D 82 (2010) 046003, arXiv:0808.0706 [hep-th].
[2472] A. Gómez-Valent, N. E. Mavromatos, and J. Solà Peracaula, “Stringy running vacuum model and current tensions in
cosmology,” Class. Quant. Grav. 41 (2024) no. 1, 015026, arXiv:2305.15774 [gr-qc].
[2473] A. Gomez-Valent and J. Solà Peracaula, “Phantom Matter: A Challenging Solution to the Cosmological Tensions,”
Astrophys. J. 975 (2024) no. 1, 64, arXiv:2404.18845 [astro-ph.CO].
[2474] R. Kallosh and A. Linde, “Universality Class in Conformal Inflation,” JCAP 07 (2013) 002, arXiv:1306.5220
[hep-th].
[2475] R. Kallosh, A. Linde, and D. Roest, “Superconformal Inflationary α-Attractors,” JHEP 11 (2013) 198,
arXiv:1311.0472 [hep-th].
[2476] M. Galante, R. Kallosh, A. Linde, and D. Roest, “Unity of Cosmological Inflation Attractors,” Phys. Rev. Lett. 114
(2015) no. 14, 141302, arXiv:1412.3797 [hep-th].
[2477] M. Braglia, W. T. Emond, F. Finelli, A. E. Gumrukcuoglu, and K. Koyama, “Unified framework for early dark energy
from α-attractors,” Phys. Rev. D 102 (2020) no. 8, 083513, arXiv:2005.14053 [astro-ph.CO].
[2478] L. Brissenden, K. Dimopoulos, and S. Sánchez López, “Non-oscillating early dark energy and quintessence from
α-attractors,” Astropart. Phys. 157 (2024) 102925, arXiv:2301.03572 [astro-ph.CO].
[2479] D. Chowdhury, G. Tasinato, and I. Zavala, “Dark energy, D-branes and pulsar timing arrays,” JCAP 11 (2023) 090,
arXiv:2307.01188 [hep-th].
[2480] O. F. Ramadan, T. Karwal, and J. Sakstein, “Attractive proposal for resolving the Hubble tension: Dynamical
attractors that unify early and late dark energy,” Phys. Rev. D 109 (2024) no. 6, 063525, arXiv:2309.08082
[astro-ph.CO].
[2481] E. J. Copeland, A. R. Liddle, and D. Wands, “Exponential potentials and cosmological scaling solutions,” Phys. Rev. D
57 (1998) 4686–4690, arXiv:gr-qc/9711068.
[2482] T. Barreiro, E. J. Copeland, and N. J. Nunes, “Quintessence arising from exponential potentials,” Phys. Rev. D 61
(2000) 127301, arXiv:astro-ph/9910214.
[2483] A. Gómez-Valent, Z. Zheng, L. Amendola, V. Pettorino, and C. Wetterich, “Early dark energy in the pre- and
postrecombination epochs,” Phys. Rev. D 104 (2021) no. 8, 083536, arXiv:2107.11065 [astro-ph.CO].
[2484] V. Pettorino, L. Amendola, and C. Wetterich, “How early is early dark energy?,” Phys. Rev. D 87 (2013) 083009,
arXiv:1301.5279 [astro-ph.CO].
[2485] E. J. Copeland, A. Moss, S. Sevillano Muñoz, and J. M. M. White, “Scaling solutions as Early Dark Energy resolutions
to the Hubble tension,” JCAP 05 (2024) 078, arXiv:2309.15295 [astro-ph.CO].
[2486] T. Karwal, M. Raveri, B. Jain, J. Khoury, and M. Trodden, “Chameleon early dark energy and the Hubble tension,”
Phys. Rev. D 105 (2022) no. 6, 063535, arXiv:2106.13290 [astro-ph.CO].
[2487] M.-X. Lin, E. McDonough, J. C. Hill, and W. Hu, “Dark matter trigger for early dark energy coincidence,” Phys. Rev.
D 107 (2023) no. 10, 103523, arXiv:2212.08098 [astro-ph.CO].
345

[2488] G. Liu, Z. Zhou, Y. Mu, and L. Xu, “Alleviating cosmological tensions with a coupled scalar fields model,” Phys. Rev.
D 108 (2023) no. 8, 083523, arXiv:2307.07228 [astro-ph.CO].
[2489] G. Garcia-Arroyo, L. A. Ureña López, and J. A. Vázquez, “Interacting scalar fields: Dark matter and early dark
energy,” Phys. Rev. D 110 (2024) no. 2, 023529, arXiv:2402.08815 [astro-ph.CO].
[2490] A. Talebian, “Early dark energy and dark photon dark matter from waterfall symmetry breaking,” Phys. Rev. D 109
(2024) no. 12, 123526, arXiv:2312.08254 [astro-ph.CO].
[2491] K. V. Berghaus and T. Karwal, “Thermal Friction as a Solution to the Hubble Tension,” Phys. Rev. D 101 (2020)
no. 8, 083537, arXiv:1911.06281 [astro-ph.CO].
[2492] J. Sakstein and M. Trodden, “Early Dark Energy from Massive Neutrinos as a Natural Resolution of the Hubble
Tension,” Phys. Rev. Lett. 124 (2020) no. 16, 161301, arXiv:1911.11760 [astro-ph.CO].
[2493] M. Carrillo González, Q. Liang, J. Sakstein, and M. Trodden, “Neutrino-Assisted Early Dark Energy: Theory and
Cosmology,” JCAP 04 (2021) 063, arXiv:2011.09895 [astro-ph.CO].
[2494] D. H. F. de Souza and R. Rosenfeld, “Can neutrino-assisted early dark energy models ameliorate the H0 tension in a
natural way?,” Phys. Rev. D 108 (2023) no. 8, 083512, arXiv:2302.04644 [astro-ph.CO].
[2495] M. Carrillo González, Q. Liang, J. Sakstein, and M. Trodden, “Neutrino-Assisted Early Dark Energy is a Natural
Resolution of the Hubble Tension,” arXiv:2302.09091 [astro-ph.CO].
[2496] X. Li and A. Shafieloo, “A Simple Phenomenological Emergent Dark Energy Model can Resolve the Hubble Tension,”
Astrophys. J. Lett. 883 (2019) no. 1, L3, arXiv:1906.08275 [astro-ph.CO].
[2497] L. A. García, L. Castañeda, and J. M. Tejeiro, “A novel early Dark Energy model,” New Astron. 84 (2021) 101503,
arXiv:2009.07357 [astro-ph.CO].
[2498] H. B. Benaoum, L. A. García, and L. Castañeda, “Early dark energy induced by non-linear electrodynamics,”
arXiv:2307.05917 [gr-qc].
[2499] X. Li and A. Shafieloo, “Evidence for Emergent Dark Energy,” Astrophys. J. 902 (2020) no. 1, 58, arXiv:2001.05103
[astro-ph.CO].
[2500] S. Nojiri, S. D. Odintsov, D. Saez-Chillon Gomez, and G. S. Sharov, “Modeling and testing the equation of state for
(Early) dark energy,” Phys. Dark Univ. 32 (2021) 100837, arXiv:2103.05304 [gr-qc].
[2501] V. I. Sabla and R. R. Caldwell, “Microphysics of early dark energy,” Phys. Rev. D 106 (2022) no. 6, 063526,
arXiv:2202.08291 [astro-ph.CO].
[2502] H. Moshafi, H. Firouzjahi, and A. Talebian, “Multiple Transitions in Vacuum Dark Energy and H 0 Tension,”
Astrophys. J. 940 (2022) no. 2, 121, arXiv:2208.05583 [astro-ph.CO].
[2503] A. J. Ross, L. Samushia, C. Howlett, W. J. Percival, A. Burden, and M. Manera, “The clustering of the SDSS DR7
main Galaxy sample – I. A 4 per cent distance measure at z = 0.15,” Mon. Not. Roy. Astron. Soc. 449 (2015) no. 1,
835–847, arXiv:1409.3242 [astro-ph.CO].
[2504] E.-M. Mueller, W. Percival, E. Linder, S. Alam, G.-B. Zhao, A. G. Sánchez, F. Beutler, and J. Brinkmann, “The
clustering of galaxies in the completed SDSS-III Baryon Oscillation Spectroscopic Survey: constraining modified
gravity,” Mon. Not. Roy. Astron. Soc. 475 (2018) no. 2, 2122–2131, arXiv:1612.00812 [astro-ph.CO].
[2505] A. G. Riess et al., “Milky Way Cepheid Standards for Measuring Cosmic Distances and Application to Gaia DR2:
Implications for the Hubble Constant,” Astrophys. J. 861 (2018) no. 2, 126, arXiv:1804.10655 [astro-ph.CO].
[2506] S. Vagnozzi, “Consistency tests of ΛCDM from the early integrated Sachs-Wolfe effect: Implications for early-time new
physics and the Hubble tension,” Phys. Rev. D 104 (2021) no. 6, 063524, arXiv:2105.10425 [astro-ph.CO].
[2507] G. Ye, B. Hu, and Y.-S. Piao, “Implication of the Hubble tension for the primordial Universe in light of recent
cosmological data,” Phys. Rev. D 104 (2021) no. 6, 063510, arXiv:2103.09729 [astro-ph.CO].
[2508] I. d. O. C. Pedreira, M. Benetti, E. G. M. Ferreira, L. L. Graef, and L. Herold, “Visual tool for assessing
tension-resolving models in the H0-σ8 plane,” Phys. Rev. D 109 (2024) no. 10, 103525, arXiv:2311.04977
[astro-ph.CO].
[2509] BOSS Collaboration, H. Gil-Marín et al., “The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic
Survey: RSD measurement from the LOS-dependent power spectrum of DR12 BOSS galaxies,” Mon. Not. Roy.
Astron. Soc. 460 (2016) no. 4, 4188–4209, arXiv:1509.06386 [astro-ph.CO].
[2510] BOSS Collaboration, H. Gil-Marín et al., “The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic
Survey: BAO measurement from the LOS-dependent power spectrum of DR12 BOSS galaxies,” Mon. Not. Roy.
Astron. Soc. 460 (2016) no. 4, 4210–4219, arXiv:1509.06373 [astro-ph.CO].
[2511] eBOSS Collaboration, V. de Sainte Agathe et al., “Baryon acoustic oscillations at z = 2.34 from the correlations of
Lyα absorption in eBOSS DR14,” Astron. Astrophys. 629 (2019) A85, arXiv:1904.03400 [astro-ph.CO].
[2512] M. M. Ivanov, E. McDonough, J. C. Hill, M. Simonović, M. W. Toomey, S. Alexander, and M. Zaldarriaga,
“Constraining Early Dark Energy with Large-Scale Structure,” Phys. Rev. D 102 (2020) no. 10, 103502,
arXiv:2006.11235 [astro-ph.CO].
[2513] G. D’Amico, L. Senatore, P. Zhang, and H. Zheng, “The Hubble Tension in Light of the Full-Shape Analysis of
Large-Scale Structure Data,” JCAP 05 (2021) 072, arXiv:2006.12420 [astro-ph.CO].
[2514] F. Beutler and P. McDonald, “Unified galaxy power spectrum measurements from 6dFGS, BOSS, and eBOSS,” JCAP
11 (2021) 031, arXiv:2106.06324 [astro-ph.CO].
[2515] D. Piras, L. Herold, L. Lucie-Smith, and E. Komatsu, “ΛCDM and early dark energy in latent space: a data-driven
parametrization of the CMB temperature power spectrum,” arXiv:2502.09810 [astro-ph.CO].
[2516] J. C. Hill et al., “Atacama Cosmology Telescope: Constraints on prerecombination early dark energy,” Phys. Rev. D
105 (2022) no. 12, 123536, arXiv:2109.04451 [astro-ph.CO].
346

[2517] A. La Posta, T. Louis, X. Garrido, and J. C. Hill, “Constraints on prerecombination early dark energy from SPT-3G
public data,” Phys. Rev. D 105 (2022) no. 8, 083519, arXiv:2112.10754 [astro-ph.CO].
[2518] T. L. Smith, M. Lucca, V. Poulin, G. F. Abellan, L. Balkenhol, K. Benabed, S. Galli, and R. Murgia, “Hints of early
dark energy in Planck, SPT, and ACT data: New physics or systematics?,” Phys. Rev. D 106 (2022) no. 4, 043526,
arXiv:2202.09379 [astro-ph.CO].
[2519] J.-Q. Jiang and Y.-S. Piao, “Toward early dark energy and ns=1 with Planck, ACT, and SPT observations,” Phys.
Rev. D 105 (2022) no. 10, 103514, arXiv:2202.13379 [astro-ph.CO].
[2520] G. Benevento, W. Hu, and M. Raveri, “Can Late Dark Energy Transitions Raise the Hubble constant?,” Phys. Rev. D
101 (2020) no. 10, 103517, arXiv:2002.11707 [astro-ph.CO].
[2521] D. Camarena and V. Marra, “On the use of the local prior on the absolute magnitude of Type Ia supernovae in
cosmological inference,” Mon. Not. Roy. Astron. Soc. 504 (2021) 5164–5171, arXiv:2101.08641 [astro-ph.CO].
[2522] G. Efstathiou, “To H0 or not to H0?,” Mon. Not. Roy. Astron. Soc. 505 (2021) no. 3, 3866–3872, arXiv:2103.08723
[astro-ph.CO].
[2523] G. Efstathiou and S. Gratton, “A Detailed Description of the CamSpec Likelihood Pipeline and a Reanalysis of the
Planck High Frequency Maps,” arXiv:1910.00483 [astro-ph.CO].
[2524] Planck Collaboration, Y. Akrami et al., “P lanck intermediate results. LVII. Joint Planck LFI and HFI data
processing,” Astron. Astrophys. 643 (2020) A42, arXiv:2007.04997 [astro-ph.CO].
[2525] J.-Q. Jiang, G. Ye, and Y.-S. Piao, “Impact of the Hubble tension on the r- ns contour,” Phys. Lett. B 851 (2024)
138588, arXiv:2303.12345 [astro-ph.CO].
[2526] W. Liu, H. Zhan, Y. Gong, and X. Wang, “Can early dark energy be probed by the high-redshift galaxy abundance?,”
Mon. Not. Roy. Astron. Soc. 533 (2024) no. 1, 860–871, arXiv:2402.14339 [astro-ph.CO].
[2527] M. Boylan-Kolchin and D. R. Weisz, “Uncertain times: the redshift–time relation from cosmology and stars,” Mon.
Not. Roy. Astron. Soc. 505 (2021) no. 2, 2764–2783, arXiv:2103.15825 [astro-ph.CO].
[2528] P. Diego-Palazuelos et al., “Cosmic Birefringence from Planck Public Release 4,” in 56th Rencontres de Moriond on
Cosmology. 2022. arXiv:2203.04830 [astro-ph.CO].
[2529] K. Murai, F. Naokawa, T. Namikawa, and E. Komatsu, “Isotropic cosmic birefringence from early dark energy,” Phys.
Rev. D 107 (2023) no. 4, L041302, arXiv:2209.07804 [astro-ph.CO].
[2530] J. R. Eskilt, L. Herold, E. Komatsu, K. Murai, T. Namikawa, and F. Naokawa, “Constraints on Early Dark Energy
from Isotropic Cosmic Birefringence,” Phys. Rev. Lett. 131 (2023) no. 12, 121001, arXiv:2303.15369 [astro-ph.CO].
[2531] L. Yin, J. Kochappan, T. Ghosh, and B.-H. Lee, “Is cosmic birefringence model-dependent?,” JCAP 10 (2023) 007,
arXiv:2305.07937 [astro-ph.CO].
[2532] L. Hart and J. Chluba, “Using the cosmological recombination radiation to probe early dark energy and fundamental
constant variations,” Mon. Not. Roy. Astron. Soc. 519 (2023) no. 3, 3664–3680, arXiv:2209.12290 [astro-ph.CO].
[2533] T. Rudelius, “Constraints on early dark energy from the axion weak gravity conjecture,” JCAP 01 (2023) 014,
arXiv:2203.05575 [hep-th].
[2534] A. Hebecker, P. Mangat, S. Theisen, and L. T. Witkowski, “Can Gravitational Instantons Really Constrain Axion
Inflation?,” JHEP 02 (2017) 097, arXiv:1607.06814 [hep-th].
[2535] B. Heidenreich, M. Reece, and T. Rudelius, “Sharpening the Weak Gravity Conjecture with Dimensional Reduction,”
JHEP 02 (2016) 140, arXiv:1509.06374 [hep-th].
[2536] J.-Q. Jiang, G. Ye, and Y.-S. Piao, “Return of Harrison–Zeldovich spectrum in light of recent cosmological tensions,”
Mon. Not. Roy. Astron. Soc. 527 (2023) no. 1, L54–L59, arXiv:2210.06125 [astro-ph.CO].
[2537] Z.-Y. Peng and Y.-S. Piao, “Testing the ns-H0 scaling relation with Planck-independent CMB data,” Phys. Rev. D 109
(2024) no. 2, 023519, arXiv:2308.01012 [astro-ph.CO].
[2538] H. Wang, G. Ye, J.-Q. Jiang, and Y.-S. Piao, “Towards primordial gravitational waves and ns = 1 in light of
BICEP/Keck, DESI BAO and Hubble tension,” arXiv:2409.17879 [astro-ph.CO].
[2539] R. Kallosh and A. Linde, “Hybrid cosmological attractors,” Phys. Rev. D 106 (2022) no. 2, 023522, arXiv:2204.02425
[hep-th].
[2540] G. Ye, J.-Q. Jiang, and Y.-S. Piao, “Toward inflation with ns=1 in light of the Hubble tension and implications for
primordial gravitational waves,” Phys. Rev. D 106 (2022) no. 10, 103528, arXiv:2205.02478 [astro-ph.CO].
[2541] F. Niedermann and M. S. Sloth, “New Early Dark Energy as a solution to the H0 and S8 tensions,” arXiv:2307.03481
[hep-ph].
[2542] F. Niedermann and M. S. Sloth, “New early dark energy,” Phys. Rev. D 103 (2021) no. 4, L041303, arXiv:1910.10739
[astro-ph.CO].
[2543] K. Freese and M. W. Winkler, “Chain early dark energy: A Proposal for solving the Hubble tension and explaining
today’s dark energy,” Phys. Rev. D 104 (2021) no. 8, 083533, arXiv:2102.13655 [astro-ph.CO].
[2544] J. S. Cruz, F. Niedermann, and M. S. Sloth, “Cold New Early Dark Energy pulls the trigger on the H 0 and S 8
tensions: a simultaneous solution to both tensions without new ingredients,” JCAP 11 (2023) 033, arXiv:2305.08895
[astro-ph.CO].
[2545] F. Niedermann and M. S. Sloth, “New Early Dark Energy is compatible with current LSS data,” Phys. Rev. D 103
(2021) no. 10, 103537, arXiv:2009.00006 [astro-ph.CO].
[2546] B. S. Haridasu, H. Khoraminezhad, and M. Viel, “Scrutinizing Early Dark Energy models through CMB lensing,”
arXiv:2212.09136 [astro-ph.CO].
[2547] J. S. Cruz, F. Niedermann, and M. S. Sloth, “A grounded perspective on new early dark energy using ACT, SPT, and
BICEP/Keck,” JCAP 02 (2023) 041, arXiv:2209.02708 [astro-ph.CO].
347

[2548] I. J. Allali, M. P. Hertzberg, and F. Rompineve, “Dark sector to restore cosmological concordance,” Phys. Rev. D 104
(2021) no. 8, L081303, arXiv:2104.12798 [astro-ph.CO].
[2549] D. Aloni, A. Berlin, M. Joseph, M. Schmaltz, and N. Weiner, “A Step in understanding the Hubble tension,” Phys.
Rev. D 105 (2022) no. 12, 123516, arXiv:2111.00014 [astro-ph.CO].
[2550] F. Niedermann and M. S. Sloth, “Hot new early dark energy,” Phys. Rev. D 105 (2022) no. 6, 063509,
arXiv:2112.00770 [hep-ph].
[2551] F. Niedermann and M. S. Sloth, “Hot new early dark energy: Towards a unified dark sector of neutrinos, dark energy
and dark matter,” Phys. Lett. B 835 (2022) 137555, arXiv:2112.00759 [hep-ph].
[2552] M. Garny, F. Niedermann, H. Rubira, and M. S. Sloth, “Hot new early dark energy bridging cosmic gaps: Supercooled
phase transition reconciles stepped dark radiation solutions to the Hubble tension with BBN,” Phys. Rev. D 110
(2024) no. 2, 023531, arXiv:2404.07256 [astro-ph.CO].
[2553] N. Schöneberg and G. Franco Abellán, “A step in the right direction? Analyzing the Wess Zumino Dark Radiation
solution to the Hubble tension,” JCAP 12 (2022) 001, arXiv:2206.11276 [astro-ph.CO].
[2554] A. Berlin and N. Blinov, “Thermal Dark Matter Below an MeV,” Phys. Rev. Lett. 120 (2018) no. 2, 021801,
arXiv:1706.07046 [hep-ph].
[2555] M. Berbig, S. Jana, and A. Trautner, “The Hubble tension and a renormalizable model of gauged neutrino
self-interactions,” Phys. Rev. D 102 (2020) no. 11, 115008, arXiv:2004.13039 [hep-ph].
[2556] M. Escudero, T. Schwetz, and J. Terol-Calvo, “A seesaw model for large neutrino masses in concordance with
cosmology,” JHEP 02 (2023) 142, arXiv:2211.01729 [hep-ph]. [Addendum: JHEP 06, 119 (2024)].
[2557] D. Aloni, M. Joseph, M. Schmaltz, and N. Weiner, “Dark Radiation from Neutrino Mixing after Big Bang
Nucleosynthesis,” Phys. Rev. Lett. 131 (2023) no. 22, 221001, arXiv:2301.10792 [astro-ph.CO].
[2558] W. Fischler and J. Meyers, “Dark Radiation Emerging After Big Bang Nucleosynthesis?,” Phys. Rev. D 83 (2011)
063520, arXiv:1011.3501 [astro-ph.CO].
[2559] D. Hooper, F. S. Queiroz, and N. Y. Gnedin, “Non-Thermal Dark Matter Mimicking An Additional Neutrino Species
In The Early Universe,” Phys. Rev. D 85 (2012) 063513, arXiv:1111.6599 [astro-ph.CO].
[2560] O. E. Bjaelde, S. Das, and A. Moss, “Origin of Delta N_eff as a Result of an Interaction between Dark Radiation and
Dark Matter,” JCAP 10 (2012) 017, arXiv:1205.0553 [astro-ph.CO].
[2561] K. Choi, K.-Y. Choi, and C. S. Shin, “Dark radiation and small-scale structure problems with decaying particles,”
Phys. Rev. D 86 (2012) 083529, arXiv:1208.2496 [hep-ph].
[2562] A. Nygaard, E. B. Holm, T. Tram, and S. Hannestad, “Decaying Dark Matter and the Hubble Tension,”
arXiv:2307.00418 [astro-ph.CO].
[2563] S. Gariazzo and O. Mena, “On the dark radiation role in the Hubble constant tension,” arXiv:2306.15067
[astro-ph.CO].
[2564] G. Steigman, D. N. Schramm, and J. E. Gunn, “Cosmological Limits to the Number of Massive Leptons,” Phys. Lett.
B 66 (1977) 202–204.
[2565] M. Archidiacono, E. Giusarma, S. Hannestad, and O. Mena, “Cosmic dark radiation and neutrinos,” Adv. High Energy
Phys. 2013 (2013) 191047, arXiv:1307.0637 [astro-ph.CO].
[2566] M. Archidiacono and S. Gariazzo, “Two Sides of the Same Coin: Sterile Neutrinos and Dark Radiation, Status and
Perspectives,” Universe 8 (2022) no. 3, 175, arXiv:2201.10319 [hep-ph].
[2567] K. Akita and M. Yamaguchi, “A precision calculation of relic neutrino decoupling,” JCAP 08 (2020) 012,
arXiv:2005.07047 [hep-ph].
[2568] J. Froustey, C. Pitrou, and M. C. Volpe, “Neutrino decoupling including flavour oscillations and primordial
nucleosynthesis,” JCAP 12 (2020) 015, arXiv:2008.01074 [hep-ph].
[2569] J. J. Bennett, G. Buldgen, P. F. De Salas, M. Drewes, S. Gariazzo, S. Pastor, and Y. Y. Y. Wong, “Towards a precision
calculation of Neff in the Standard Model II: Neutrino decoupling in the presence of flavour oscillations and
finite-temperature QED,” JCAP 04 (2021) 073, arXiv:2012.02726 [hep-ph].
[2570] M. Drewes, Y. Georis, M. Klasen, L. P. Wiggering, and Y. Y. Y. Wong, “Towards a precision calculation of N ef f in
the Standard Model. Part III. Improved estimate of NLO contributions to the collision integral,” JCAP 06 (2024) 032,
arXiv:2402.18481 [hep-ph].
[2571] L. García, J. Tejeiro, and L. Castañeda, “Primordial nucleosynthesis in the presence of sterile neutrinos,” Proc. Int.
Sch. Phys. Fermi 178 (2011) 309–316.
[2572] L. A. Anchordoqui and H. Goldberg, “Neutrino cosmology after WMAP 7-Year data and LHC first Z’ bounds,” Phys.
Rev. Lett. 108 (2012) 081805, arXiv:1111.7264 [hep-ph].
[2573] L. A. Anchordoqui, H. Goldberg, and G. Steigman, “Right-Handed Neutrinos as the Dark Radiation: Status and
Forecasts for the LHC,” Phys. Lett. B 718 (2013) 1162–1165, arXiv:1211.0186 [hep-ph].
[2574] T. D. Jacques, L. M. Krauss, and C. Lunardini, “Additional Light Sterile Neutrinos and Cosmology,” Phys. Rev. D 87
(2013) no. 8, 083515, arXiv:1301.3119 [astro-ph.CO]. [Erratum: Phys.Rev.D 88, 109901 (2013)].
[2575] S. Roy Choudhury and S. Choubey, “Constraining light sterile neutrino mass with the BICEP2/Keck Array 2014
B-mode polarization data,” Eur. Phys. J. C 79 (2019) no. 7, 557, arXiv:1807.10294 [astro-ph.CO].
[2576] S. Weinberg, “Goldstone Bosons as Fractional Cosmic Neutrinos,” Phys. Rev. Lett. 110 (2013) no. 24, 241301,
arXiv:1305.1971 [astro-ph.CO].
[2577] W. Lin, L. Visinelli, D. Xu, and T. T. Yanagida, “Neutrino astronomy as a probe of physics beyond the Standard
Model: Decay of sub-MeV B-L gauge boson dark matter,” Phys. Rev. D 106 (2022) no. 7, 075011, arXiv:2202.04496
[hep-ph].
348

[2578] P. Arias, D. Cadamuro, M. Goodsell, J. Jaeckel, J. Redondo, and A. Ringwald, “WISPy Cold Dark Matter,” JCAP 06
(2012) 013, arXiv:1201.5902 [hep-ph].
[2579] D. J. E. Marsh, “Axion Cosmology,” Phys. Rept. 643 (2016) 1–79, arXiv:1510.07633 [astro-ph.CO].
[2580] D. Baumann, D. Green, and B. Wallisch, “New Target for Cosmic Axion Searches,” Phys. Rev. Lett. 117 (2016) no. 17,
171301, arXiv:1604.08614 [astro-ph.CO].
[2581] F. D’Eramo, R. Z. Ferreira, A. Notari, and J. L. Bernal, “Hot Axions and the H0 tension,” JCAP 11 (2018) 014,
arXiv:1808.07430 [hep-ph].
[2582] E. G. M. Ferreira, “Ultra-light dark matter,” Astron. Astrophys. Rev. 29 (2021) no. 1, 7, arXiv:2005.03254
[astro-ph.CO].
[2583] Y. K. Semertzidis and S. Youn, “Axion dark matter: How to see it?,” Sci. Adv. 8 (2022) no. 8, abm9928,
arXiv:2104.14831 [hep-ph].
[2584] F. Chadha-Day, J. Ellis, and D. J. E. Marsh, “Axion dark matter: What is it and why now?,” Sci. Adv. 8 (2022) no. 8,
abj3618, arXiv:2105.01406 [hep-ph].
[2585] D. Green, Y. Guo, and B. Wallisch, “Cosmological implications of axion-matter couplings,” JCAP 02 (2022) no. 02,
019, arXiv:2109.12088 [astro-ph.CO].
[2586] C. A. J. O’Hare, “Cosmology of axion dark matter,” PoS COSMICWISPers (2024) 040, arXiv:2403.17697
[hep-ph].
[2587] L. Caloni, M. Gerbino, M. Lattanzi, and L. Visinelli, “Novel cosmological bounds on thermally-produced axion-like
particles,” JCAP 09 (2022) 021, arXiv:2205.01637 [astro-ph.CO].
[2588] F. D’Eramo, E. Di Valentino, W. Giarè, F. Hajkarim, A. Melchiorri, O. Mena, F. Renzi, and S. Yun, “Cosmological
bound on the QCD axion mass, redux,” JCAP 09 (2022) 022, arXiv:2205.07849 [astro-ph.CO].
[2589] J. M. Cline, Z. Liu, and W. Xue, “Millicharged Atomic Dark Matter,” Phys. Rev. D 85 (2012) 101302,
arXiv:1201.4858 [hep-ph].
[2590] J. Fan, A. Katz, L. Randall, and M. Reece, “Double-Disk Dark Matter,” Phys. Dark Univ. 2 (2013) 139–156,
arXiv:1303.1521 [astro-ph.CO].
[2591] H. Vogel and J. Redondo, “Dark Radiation constraints on minicharged particles in models with a hidden photon,”
JCAP 02 (2014) 029, arXiv:1311.2600 [hep-ph].
[2592] K. Petraki, L. Pearce, and A. Kusenko, “Self-interacting asymmetric dark matter coupled to a light massive dark
photon,” JCAP 07 (2014) 039, arXiv:1403.1077 [hep-ph].
[2593] R. Foot and S. Vagnozzi, “Dissipative hidden sector dark matter,” Phys. Rev. D 91 (2015) 023512, arXiv:1409.7174
[hep-ph].
[2594] R. Foot and S. Vagnozzi, “Diurnal modulation signal from dissipative hidden sector dark matter,” Phys. Lett. B 748
(2015) 61–66, arXiv:1412.0762 [hep-ph].
[2595] R. Foot and S. Vagnozzi, “Solving the small-scale structure puzzles with dissipative dark matter,” JCAP 07 (2016)
013, arXiv:1602.02467 [astro-ph.CO].
[2596] V. V. Flambaum and I. B. Samsonov, “Ultralight dark photon as a model for early universe dark matter,” Phys. Rev.
D 100 (2019) no. 6, 063541, arXiv:1908.09432 [astro-ph.CO].
[2597] L. A. Anchordoqui and S. E. Perez Bergliaffa, “Hot thermal universe endowed with massive dark vector fields and the
Hubble tension,” Phys. Rev. D 100 (2019) no. 12, 123525, arXiv:1910.05860 [astro-ph.CO].
[2598] G. Steigman, “Equivalent Neutrinos, Light WIMPs, and the Chimera of Dark Radiation,” Phys. Rev. D 87 (2013)
no. 10, 103517, arXiv:1303.0049 [astro-ph.CO].
[2599] C. Brust, D. E. Kaplan, and M. T. Walters, “New Light Species and the CMB,” JHEP 12 (2013) 058,
arXiv:1303.5379 [hep-ph].
[2600] R. C. Nunes and A. Bonilla, “Probing the properties of relic neutrinos using the cosmic microwave background, the
Hubble Space Telescope and galaxy clusters,” Mon. Not. Roy. Astron. Soc. 473 (2018) no. 4, 4404–4409,
arXiv:1710.10264 [astro-ph.CO].
[2601] A. Bonilla, R. C. Nunes, and E. M. C. Abreu, “Forecast on lepton asymmetry from future CMB experiments,” Mon.
Not. Roy. Astron. Soc. 485 (2019) no. 2, 2486–2491, arXiv:1810.06356 [astro-ph.CO].
[2602] NANOGrav Collaboration, Z. Arzoumanian et al., “The NANOGrav 12.5 yr Data Set: Search for an Isotropic
Stochastic Gravitational-wave Background,” Astrophys. J. Lett. 905 (2020) no. 2, L34, arXiv:2009.04496
[astro-ph.HE].
[2603] B. Goncharov et al., “On the Evidence for a Common-spectrum Process in the Search for the Nanohertz
Gravitational-wave Background with the Parkes Pulsar Timing Array,” Astrophys. J. Lett. 917 (2021) no. 2, L19,
arXiv:2107.12112 [astro-ph.HE].
[2604] EPTA Collaboration, S. Chen et al., “Common-red-signal analysis with 24-yr high-precision timing of the European
Pulsar Timing Array: inferences in the stochastic gravitational-wave background search,” Mon. Not. Roy. Astron. Soc.
508 (2021) no. 4, 4970–4993, arXiv:2110.13184 [astro-ph.HE].
[2605] J. Antoniadis et al., “The International Pulsar Timing Array second data release: Search for an isotropic gravitational
wave background,” Mon. Not. Roy. Astron. Soc. 510 (2022) no. 4, 4873–4887, arXiv:2201.03980 [astro-ph.HE].
[2606] B. Allen and J. D. Romano, “Detecting a stochastic background of gravitational radiation: Signal processing strategies
and sensitivities,” Phys. Rev. D 59 (1999) 102001, arXiv:gr-qc/9710117.
[2607] T. L. Smith, E. Pierpaoli, and M. Kamionkowski, “A new cosmic microwave background constraint to primordial
gravitational waves,” Phys. Rev. Lett. 97 (2006) 021301, arXiv:astro-ph/0603144.
[2608] L. A. Boyle and A. Buonanno, “Relating gravitational wave constraints from primordial nucleosynthesis, pulsar timing,
349

laser interferometers, and the CMB: Implications for the early Universe,” Phys. Rev. D 78 (2008) 043531,
arXiv:0708.2279 [astro-ph].
[2609] S. Kuroyanagi, T. Takahashi, and S. Yokoyama, “Blue-tilted Tensor Spectrum and Thermal History of the Universe,”
JCAP 02 (2015) 003, arXiv:1407.4785 [astro-ph.CO].
[2610] G. Cabass, L. Pagano, L. Salvati, M. Gerbino, E. Giusarma, and A. Melchiorri, “Updated Constraints and Forecasts on
Primordial Tensor Modes,” Phys. Rev. D 93 (2016) no. 6, 063508, arXiv:1511.05146 [astro-ph.CO].
[2611] I. Ben-Dayan, B. Keating, D. Leon, and I. Wolfson, “Constraints on scalar and tensor spectra from Nef f ,” JCAP 06
(2019) 007, arXiv:1903.11843 [astro-ph.CO].
[2612] M. Aich, Y.-Z. Ma, W.-M. Dai, and J.-Q. Xia, “How much primordial tensor mode is allowed?,” Phys. Rev. D 101
(2020) no. 6, 063536, arXiv:1912.00995 [astro-ph.CO].
[2613] W. Giarè, M. Forconi, E. Di Valentino, and A. Melchiorri, “Towards a reliable calculation of relic radiation from
primordial gravitational waves,” Mon. Not. Roy. Astron. Soc. 520 (2023) 2, arXiv:2210.14159 [astro-ph.CO].
[2614] NANOGrav Collaboration, A. Afzal et al., “The NANOGrav 15 yr Data Set: Search for Signals from New Physics,”
Astrophys. J. Lett. 951 (2023) no. 1, L11, arXiv:2306.16219 [astro-ph.HE]. [Erratum: Astrophys.J.Lett. 971, L27
(2024), Erratum: Astrophys.J. 971, L27 (2024)].
[2615] M. Benetti, L. L. Graef, and S. Vagnozzi, “Primordial gravitational waves from NANOGrav: A broken power-law
approach,” Phys. Rev. D 105 (2022) no. 4, 043520, arXiv:2111.04758 [astro-ph.CO].
[2616] S. Vagnozzi, “Inflationary interpretation of the stochastic gravitational wave background signal detected by pulsar
timing array experiments,” JHEAp 39 (2023) 81–98, arXiv:2306.16912 [astro-ph.CO].
[2617] M. Kawasaki, K. Kohri, and N. Sugiyama, “Cosmological constraints on late time entropy production,” Phys. Rev.
Lett. 82 (1999) 4168, arXiv:astro-ph/9811437.
[2618] G. Gelmini, S. Palomares-Ruiz, and S. Pascoli, “Low reheating temperature and the visible sterile neutrino,” Phys.
Rev. Lett. 93 (2004) 081302, arXiv:astro-ph/0403323.
[2619] P. F. de Salas, M. Lattanzi, G. Mangano, G. Miele, S. Pastor, and O. Pisanti, “Bounds on very low reheating scenarios
after Planck,” Phys. Rev. D 92 (2015) no. 12, 123534, arXiv:1511.00672 [astro-ph.CO].
[2620] M. Gerbino, K. Freese, S. Vagnozzi, M. Lattanzi, O. Mena, E. Giusarma, and S. Ho, “Impact of neutrino properties on
the estimation of inflationary parameters from current and future observations,” Phys. Rev. D 95 (2017) no. 4, 043512,
arXiv:1610.08830 [astro-ph.CO].
[2621] Z. Hou, R. Keisler, L. Knox, M. Millea, and C. Reichardt, “How Massless Neutrinos Affect the Cosmic Microwave
Background Damping Tail,” Phys. Rev. D 87 (2013) 083008, arXiv:1104.2333 [astro-ph.CO].
[2622] D. Baumann, “Primordial Cosmology,” PoS TASI2017 (2018) 009, arXiv:1807.03098 [hep-th].
[2623] S. Vagnozzi, “Cosmological searches for the neutrino mass scale and mass ordering,” arXiv:1907.08010
[astro-ph.CO].
[2624] E. Di Valentino, A. Melchiorri, and J. Silk, “Reconciling Planck with the local value of H0 in extended parameter
space,” Phys. Lett. B 761 (2016) 242–246, arXiv:1606.00634 [astro-ph.CO].
[2625] S. Vagnozzi, “New physics in light of the H0 tension: An alternative view,” Phys. Rev. D 102 (2020) no. 2, 023518,
arXiv:1907.07569 [astro-ph.CO].
[2626] S. Bashinsky and U. Seljak, “Neutrino perturbations in CMB anisotropy and matter clustering,” Phys. Rev. D 69
(2004) 083002, arXiv:astro-ph/0310198.
[2627] B. Follin, L. Knox, M. Millea, and Z. Pan, “First Detection of the Acoustic Oscillation Phase Shift Expected from the
Cosmic Neutrino Background,” Phys. Rev. Lett. 115 (2015) no. 9, 091301, arXiv:1503.07863 [astro-ph.CO].
[2628] E. Di Valentino, A. Melchiorri, and J. Silk, “Cosmological constraints in extended parameter space from the Planck
2018 Legacy release,” JCAP 01 (2020) 013, arXiv:1908.01391 [astro-ph.CO].
[2629] O. Seto and Y. Toda, “Comparing early dark energy and extra radiation solutions to the Hubble tension with BBN,”
Phys. Rev. D 103 (2021) no. 12, 123501, arXiv:2101.03740 [astro-ph.CO].
[2630] S. Hagstotz, P. F. de Salas, S. Gariazzo, M. Gerbino, M. Lattanzi, S. Vagnozzi, K. Freese, and S. Pastor, “Bounds on
light sterile neutrino mass and mixing from cosmology and laboratory searches,” Phys. Rev. D 104 (2021) no. 12,
123524, arXiv:2003.02289 [astro-ph.CO].
[2631] G. Barenboim, W. H. Kinney, and W.-I. Park, “Flavor versus mass eigenstates in neutrino asymmetries: implications
for cosmology,” Eur. Phys. J. C 77 (2017) no. 9, 590, arXiv:1609.03200 [astro-ph.CO].
[2632] O. Seto and Y. Toda, “Hubble tension in lepton asymmetric cosmology with an extra radiation,” Phys. Rev. D 104
(2021) no. 6, 063019, arXiv:2104.04381 [astro-ph.CO].
[2633] D. Baumann, D. Green, J. Meyers, and B. Wallisch, “Phases of New Physics in the CMB,” JCAP 01 (2016) 007,
arXiv:1508.06342 [astro-ph.CO].
[2634] N. Blinov and G. Marques-Tavares, “Interacting radiation after Planck and its implications for the Hubble Tension,”
JCAP 09 (2020) 029, arXiv:2003.08387 [astro-ph.CO].
[2635] D. E. Kaplan, G. Z. Krnjaic, K. R. Rehermann, and C. M. Wells, “Atomic Dark Matter,” JCAP 05 (2010) 021,
arXiv:0909.0753 [hep-ph].
[2636] F.-Y. Cyr-Racine and K. Sigurdson, “Cosmology of atomic dark matter,” Phys. Rev. D 87 (2013) no. 10, 103515,
arXiv:1209.5752 [astro-ph.CO].
[2637] S. Bansal, J. Barron, D. Curtin, and Y. Tsai, “Precision cosmological constraints on atomic dark matter,” JHEP 10
(2023) 095, arXiv:2212.02487 [hep-ph].
[2638] Z. Chacko, D. Curtin, M. Geller, and Y. Tsai, “Cosmological Signatures of a Mirror Twin Higgs,” JHEP 09 (2018) 163,
arXiv:1803.03263 [hep-ph].
350

[2639] F.-Y. Cyr-Racine and K. Sigurdson, “Limits on Neutrino-Neutrino Scattering in the Early Universe,” Phys. Rev. D 90
(2014) no. 12, 123533, arXiv:1306.1536 [astro-ph.CO].
[2640] M. Archidiacono and S. Hannestad, “Updated constraints on non-standard neutrino interactions from Planck,” JCAP
07 (2014) 046, arXiv:1311.3873 [astro-ph.CO].
[2641] L. Lancaster, F.-Y. Cyr-Racine, L. Knox, and Z. Pan, “A tale of two modes: Neutrino free-streaming in the early
universe,” JCAP 07 (2017) 033, arXiv:1704.06657 [astro-ph.CO].
[2642] I. M. Oldengott, T. Tram, C. Rampf, and Y. Y. Y. Wong, “Interacting neutrinos in cosmology: exact description and
constraints,” JCAP 11 (2017) 027, arXiv:1706.02123 [astro-ph.CO].
[2643] G. Barenboim, P. B. Denton, and I. M. Oldengott, “Constraints on inflation with an extended neutrino sector,” Phys.
Rev. D 99 (2019) no. 8, 083515, arXiv:1903.02036 [astro-ph.CO].
[2644] A. Das and S. Ghosh, “Flavor-specific interaction favors strong neutrino self-coupling in the early universe,” JCAP 07
(2021) 038, arXiv:2011.12315 [astro-ph.CO].
[2645] S. Roy Choudhury, S. Hannestad, and T. Tram, “Updated constraints on massive neutrino self-interactions from
cosmology in light of the H0 tension,” JCAP 03 (2021) 084, arXiv:2012.07519 [astro-ph.CO].
[2646] T. Brinckmann, J. H. Chang, and M. LoVerde, “Self-interacting neutrinos, the Hubble parameter tension, and the
cosmic microwave background,” Phys. Rev. D 104 (2021) no. 6, 063523, arXiv:2012.11830 [astro-ph.CO].
[2647] S. Roy Choudhury, S. Hannestad, and T. Tram, “Massive neutrino self-interactions and inflation,” JCAP 10 (2022)
018, arXiv:2207.07142 [astro-ph.CO].
[2648] A. Das and S. Ghosh, “The magnificent ACT of flavor-specific neutrino self-interaction,” JCAP 09 (2023) 042,
arXiv:2303.08843 [astro-ph.CO].
[2649] D. Camarena, F.-Y. Cyr-Racine, and J. Houghteling, “Confronting self-interacting neutrinos with the full shape of the
galaxy power spectrum,” Phys. Rev. D 108 (2023) no. 10, 103535, arXiv:2309.03941 [astro-ph.CO].
[2650] A. He, R. An, M. M. Ivanov, and V. Gluscevic, “Self-interacting neutrinos in light of large-scale structure data,” Phys.
Rev. D 109 (2024) no. 10, 103527, arXiv:2309.03956 [astro-ph.CO].
[2651] N. Bostan and S. Roy Choudhury, “First constraints on non-minimally coupled Natural and Coleman-Weinberg
inflation and massive neutrino self-interactions with Planck+BICEP/Keck,” JCAP 07 (2024) 032, arXiv:2310.01491
[astro-ph.CO].
[2652] N. Blinov, K. J. Kelly, G. Z. Krnjaic, and S. D. McDermott, “Constraining the Self-Interacting Neutrino Interpretation
of the Hubble Tension,” Phys. Rev. Lett. 123 (2019) no. 19, 191102, arXiv:1905.02727 [astro-ph.CO].
[2653] S. S. da Costa, D. R. da Silva, A. S. de Jesus, N. Pinto-Neto, and F. S. Queiroz, “The H 0 trouble: confronting
non-thermal dark matter and phantom cosmology with the CMB, BAO, and Type Ia supernovae data,” JCAP 04
(2024) 035, arXiv:2311.07420 [astro-ph.CO].
[2654] M. Escudero and S. J. Witte, “A CMB search for the neutrino mass mechanism and its relation to the Hubble tension,”
Eur. Phys. J. C 80 (2020) no. 4, 294, arXiv:1909.04044 [astro-ph.CO].
[2655] G. Barenboim and U. Nierste, “Modified majoron model for cosmological anomalies,” Phys. Rev. D 104 (2021) no. 2,
023013, arXiv:2005.13280 [hep-ph].
[2656] M. Archidiacono, S. Hannestad, R. S. Hansen, and T. Tram, “Cosmology with self-interacting sterile neutrinos and
dark matter - A pseudoscalar model,” Phys. Rev. D 91 (2015) no. 6, 065021, arXiv:1404.5915 [astro-ph.CO].
[2657] F. Forastieri, M. Lattanzi, G. Mangano, A. Mirizzi, P. Natoli, and N. Saviano, “Cosmic microwave background
constraints on secret interactions among sterile neutrinos,” JCAP 07 (2017) 038, arXiv:1704.00626 [astro-ph.CO].
[2658] M. Archidiacono, S. Gariazzo, C. Giunti, S. Hannestad, and T. Tram, “Sterile neutrino self-interactions: H0 tension
and short-baseline anomalies,” JCAP 12 (2020) 029, arXiv:2006.12885 [astro-ph.CO].
[2659] M. A. Corona, R. Murgia, M. Cadeddu, M. Archidiacono, S. Gariazzo, C. Giunti, and S. Hannestad, “Pseudoscalar
sterile neutrino self-interactions in light of Planck, SPT and ACT data,” JCAP 06 (2022) no. 06, 010,
arXiv:2112.00037 [astro-ph.CO].
[2660] M. Joseph, D. Aloni, M. Schmaltz, E. N. Sivarajan, and N. Weiner, “A Step in understanding the S8 tension,” Phys.
Rev. D 108 (2023) no. 2, 023520, arXiv:2207.03500 [astro-ph.CO].
[2661] H. Bagherian, M. Joseph, M. Schmaltz, and E. N. Sivarajan, “Confronting interacting radiation models for the Hubble
tension with Lyman-α data,” Phys. Rev. D 111 (2025) no. 4, 043513, arXiv:2405.17554 [astro-ph.CO].
[2662] W. Yang, E. Di Valentino, S. Pan, Y. Wu, and J. Lu, “Dynamical dark energy after Planck CMB final release and H0
tension,” Mon. Not. Roy. Astron. Soc. 501 (2021) no. 4, 5845–5858, arXiv:2101.02168 [astro-ph.CO].
[2663] F. Dong, C. Park, S. E. Hong, J. Kim, H. Seong Hwang, H. Park, and S. Appleby, “Tomographic Alcock–Paczyński
Test with Redshift-space Correlation Function: Evidence for the Dark Energy Equation-of-state Parameter w > −1,”
Astrophys. J. 953 (2023) no. 1, 98, arXiv:2305.00206 [astro-ph.CO].
[2664] M. Najafi, S. Pan, E. Di Valentino, and J. T. Firouzjaee, “Dynamical dark energy confronted with multiple CMB
missions,” Phys. Dark Univ. 45 (2024) 101539, arXiv:2407.14939 [astro-ph.CO].
[2665] S. Vagnozzi, L. Visinelli, O. Mena, and D. F. Mota, “Do we have any hope of detecting scattering between dark energy
and baryons through cosmology?,” Mon. Not. Roy. Astron. Soc. 493 (2020) no. 1, 1139–1152, arXiv:1911.12374
[gr-qc].
[2666] G. Alestas, L. Kazantzidis, and L. Perivolaropoulos, “w − M phantom transition at zt <0.1 as a resolution of the
Hubble tension,” Phys. Rev. D 103 (2021) no. 8, 083517, arXiv:2012.13932 [astro-ph.CO].
[2667] W. Yang, S. Pan, E. Di Valentino, E. N. Saridakis, and S. Chakraborty, “Observational constraints on one-parameter
dynamical dark-energy parametrizations and the H0 tension,” Phys. Rev. D 99 (2019) no. 4, 043543,
arXiv:1810.05141 [astro-ph.CO].
351

[2668] V. Marra and L. Perivolaropoulos, “Rapid transition of Geff at zt≃0.01 as a possible solution of the Hubble and growth
tensions,” Phys. Rev. D 104 (2021) no. 2, L021303, arXiv:2102.06012 [astro-ph.CO].
[2669] G. Alestas, D. Camarena, E. Di Valentino, L. Kazantzidis, V. Marra, S. Nesseris, and L. Perivolaropoulos,
“Late-transition versus smooth H(z)-deformation models for the resolution of the Hubble crisis,” Phys. Rev. D 105
(2022) no. 6, 063538, arXiv:2110.04336 [astro-ph.CO].
[2670] G. Alestas and L. Perivolaropoulos, “Late-time approaches to the Hubble tension deforming H(z), worsen the growth
tension,” Mon. Not. Roy. Astron. Soc. 504 (2021) no. 3, 3956–3962, arXiv:2103.04045 [astro-ph.CO].
[2671] L. Heisenberg, H. Villarrubia-Rojo, and J. Zosso, “Can late-time extensions solve the H0 and σ8 tensions?,” Phys. Rev.
D 106 (2022) no. 4, 043503, arXiv:2202.01202 [astro-ph.CO].
[2672] E. Frion, D. Camarena, L. Giani, T. Miranda, D. Bertacca, V. Marra, and O. F. Piattella, “Bayesian analysis of a
Unified Dark Matter model with transition: can it alleviate the H0 tension?,” Open J. Astrophys. 7 (2024) ,
arXiv:2307.06320 [astro-ph.CO].
[2673] S. Kumar, R. C. Nunes, S. Pan, and P. Yadav, “New late-time constraints on f(R) gravity,” Phys. Dark Univ. 42
(2023) 101281, arXiv:2301.07897 [astro-ph.CO].
[2674] S. Capozziello, G. Sarracino, and G. De Somma, “A Critical Discussion on the H0 Tension †,” Universe 10 (2024)
no. 3, 140, arXiv:2403.12796 [gr-qc].
[2675] D. Bousis and L. Perivolaropoulos, “Hubble tension tomography: BAO vs SN Ia distance tension,” Phys. Rev. D 110
(2024) no. 10, 103546, arXiv:2405.07039 [astro-ph.CO].
[2676] G. Alestas, L. Kazantzidis, and L. Perivolaropoulos, “H0 tension, phantom dark energy, and cosmological parameter
degeneracies,” Phys. Rev. D 101 (2020) no. 12, 123516, arXiv:2004.08363 [astro-ph.CO].
[2677] R. Briffa, C. Escamilla-Rivera, J. Said Levi, J. Mifsud, and N. L. Pullicino, “Impact of H0 priors on f (T ) late time
cosmology,” Eur. Phys. J. Plus 137 (2022) no. 5, 532, arXiv:2108.03853 [astro-ph.CO].
[2678] J. Levi Said, J. Mifsud, J. Sultana, and K. Z. Adami, “Reconstructing teleparallel gravity with cosmic structure growth
and expansion rate data,” JCAP 06 (2021) 015, arXiv:2103.05021 [astro-ph.CO].
[2679] A. Theodoropoulos and L. Perivolaropoulos, “The Hubble Tension, the M Crisis of Late Time H(z) Deformation
Models and the Reconstruction of Quintessence Lagrangians,” Universe 7 (2021) no. 8, 300, arXiv:2109.06256
[astro-ph.CO].
[2680] G. Alestas, L. Perivolaropoulos, and K. Tanidis, “Constraining a late time transition of Gef f using low-z galaxy survey
data,” Phys. Rev. D 106 (2022) no. 2, 023526, arXiv:2201.05846 [astro-ph.CO].
[2681] L. Perivolaropoulos and F. Skara, “Gravitational transitions via the explicitly broken symmetron screening
mechanism,” Phys. Rev. D 106 (2022) no. 4, 043528, arXiv:2203.10374 [astro-ph.CO].
[2682] E. A. Paraskevas, A. Cam, L. Perivolaropoulos, and O. Akarsu, “Transition dynamics in the ΛsCDM model:
Implications for bound cosmic structures,” Phys. Rev. D 109 (2024) no. 10, 103522, arXiv:2402.05908
[astro-ph.CO].
[2683] E. A. Paraskevas and L. Perivolaropoulos, “Effects of a Late Gravitational Transition on Gravitational Waves and
Anticipated Constraints,” Universe 9 (2023) no. 7, 317, arXiv:2307.00298 [astro-ph.CO].
[2684] L. Perivolaropoulos and F. Skara, “A Reanalysis of the Latest SH0ES Data for H0 : Effects of New Degrees of Freedom
on the Hubble Tension,” Universe 8 (2022) no. 10, 502, arXiv:2208.11169 [astro-ph.CO].
[2685] L. Perivolaropoulos, “Is the Hubble Crisis Connected with the Extinction of Dinosaurs?,” Universe 8 (2022) no. 5, 263,
arXiv:2201.08997 [astro-ph.EP].
[2686] O. Akarsu, J. D. Barrow, L. A. Escamilla, and J. A. Vazquez, “Graduated dark energy: Observational hints of a
spontaneous sign switch in the cosmological constant,” Phys. Rev. D 101 (2020) no. 6, 063528, arXiv:1912.08751
[astro-ph.CO].
[2687] L. A. Anchordoqui, I. Antoniadis, D. Lust, N. T. Noble, and J. F. Soriano, “From infinite to infinitesimal: Using the
Universe as a dataset to probe Casimir corrections to the vacuum energy from fields inhabiting the dark dimension,”
Phys. Dark Univ. 46 (2024) 101715, arXiv:2404.17334 [astro-ph.CO].
[2688] A. De Felice, S. Kumar, S. Mukohyama, and R. C. Nunes, “Observational bounds on extended minimal theories of
massive gravity: new limits on the graviton mass,” JCAP 04 (2024) 013, arXiv:2311.10530 [astro-ph.CO].
[2689] G. ’t Hooft, “Dimensional reduction in quantum gravity,” Conf. Proc. C 930308 (1993) 284–296,
arXiv:gr-qc/9310026.
[2690] L. Susskind, “The World as a hologram,” J. Math. Phys. 36 (1995) 6377–6396, arXiv:hep-th/9409089.
[2691] R. Bousso, “A Covariant entropy conjecture,” JHEP 07 (1999) 004, arXiv:hep-th/9905177.
[2692] A. G. Cohen, D. B. Kaplan, and A. E. Nelson, “Effective field theory, black holes, and the cosmological constant,”
Phys. Rev. Lett. 82 (1999) 4971–4974, arXiv:hep-th/9803132.
[2693] S. Wang, Y. Wang, and M. Li, “Holographic Dark Energy,” Phys. Rept. 696 (2017) 1–57, arXiv:1612.00345
[astro-ph.CO].
[2694] S. Nojiri, S. D. Odintsov, and T. Paul, “Different Faces of Generalized Holographic Dark Energy,” Symmetry 13 (2021)
no. 6, 928, arXiv:2105.08438 [gr-qc].
[2695] S. Nojiri, S. D. Odintsov, V. K. Oikonomou, and T. Paul, “Unifying Holographic Inflation with Holographic Dark
Energy: a Covariant Approach,” Phys. Rev. D 102 (2020) no. 2, 023540, arXiv:2007.06829 [gr-qc].
[2696] O. Trivedi, M. Khlopov, and A. V. Timoshkin, “Tsallis Holographic Dark Energy with Power Law Ansatz Approach,”
Symmetry 16 (2024) no. 4, 446, arXiv:2402.05784 [gr-qc].
[2697] M. Tavayef, A. Sheykhi, K. Bamba, and H. Moradpour, “Tsallis Holographic Dark Energy,” Phys. Lett. B 781 (2018)
195–200, arXiv:1804.02983 [gr-qc].
352

[2698] E. N. Saridakis, “Barrow holographic dark energy,” Phys. Rev. D 102 (2020) no. 12, 123525, arXiv:2005.04115
[gr-qc].
[2699] H. Moradpour, S. A. Moosavi, I. P. Lobo, J. P. Morais Graça, A. Jawad, and I. G. Salako, “Thermodynamic approach
to holographic dark energy and the Rényi entropy,” Eur. Phys. J. C 78 (2018) no. 10, 829, arXiv:1803.02195
[physics.gen-ph].
[2700] N. Drepanou, A. Lymperis, E. N. Saridakis, and K. Yesmakhanova, “Kaniadakis holographic dark energy and
cosmology,” Eur. Phys. J. C 82 (2022) no. 5, 449, arXiv:2109.09181 [gr-qc].
[2701] L. N. Granda and A. Oliveros, “Infrared cut-off proposal for the Holographic density,” Phys. Lett. B 669 (2008)
275–277, arXiv:0810.3149 [gr-qc].
[2702] P. Mukherjee, A. Mukherjee, H. K. Jassal, A. Dasgupta, and N. Banerjee, “Holographic dark energy: constraints on
the interaction from diverse observational data sets,” Eur. Phys. J. Plus 134 (2019) no. 4, 147, arXiv:1710.02417
[astro-ph.CO].
[2703] P. Adhikary, S. Das, S. Basilakos, and E. N. Saridakis, “Barrow holographic dark energy in a nonflat universe,” Phys.
Rev. D 104 (2021) no. 12, 123519, arXiv:2104.13118 [gr-qc].
[2704] H. Moradpour, A. H. Ziaie, and M. Kord Zangeneh, “Generalized entropies and corresponding holographic dark energy
models,” Eur. Phys. J. C 80 (2020) no. 8, 732, arXiv:2005.06271 [gr-qc].
[2705] A. Sayahian Jahromi, S. A. Moosavi, H. Moradpour, J. P. Morais Graça, I. P. Lobo, I. G. Salako, and A. Jawad,
“Generalized entropy formalism and a new holographic dark energy model,” Phys. Lett. B 780 (2018) 21–24,
arXiv:1802.07722 [gr-qc].
[2706] O. Trivedi, A. Bidlan, and P. Moniz, “Fractional holographic dark energy,” Phys. Lett. B 858 (2024) 139074,
arXiv:2407.16685 [gr-qc].
[2707] H. Moradpour, S. Jalalzadeh, and U. K. Sharma, “On the thermodynamics of reconciling quantum and gravity,” Eur.
Phys. J. Plus 139 (2024) no. 2, 170, arXiv:2304.06494 [gr-qc].
[2708] R.-Y. Guo, J.-F. Zhang, and X. Zhang, “Can the H0 tension be resolved in extensions to ΛCDM cosmology?,” JCAP
02 (2019) 054, arXiv:1809.02340 [astro-ph.CO].
[2709] T. Jacobson, “Thermodynamics of space-time: The Einstein equation of state,” Phys. Rev. Lett. 75 (1995) 1260–1263,
arXiv:gr-qc/9504004.
[2710] E. P. Verlinde, “On the Origin of Gravity and the Laws of Newton,” JHEP 04 (2011) 029, arXiv:1001.0785 [hep-th].
[2711] T. Padmanabhan, “Gravitational entropy of static space-times and microscopic density of states,” Class. Quant. Grav.
21 (2004) 4485–4494, arXiv:gr-qc/0308070.
[2712] R.-G. Cai and S. P. Kim, “First law of thermodynamics and Friedmann equations of Friedmann-Robertson-Walker
universe,” JHEP 02 (2005) 050, arXiv:hep-th/0501055.
[2713] J. D. Bekenstein, “Black holes and entropy,” Phys. Rev. D 7 (1973) 2333–2346.
[2714] S. W. Hawking, “Black hole explosions,” Nature 248 (1974) 30–31.
[2715] G. W. Gibbons and S. W. Hawking, “Cosmological Event Horizons, Thermodynamics, and Particle Creation,” Phys.
Rev. D 15 (1977) 2738–2751.
[2716] D. A. Easson, P. H. Frampton, and G. F. Smoot, “Entropic Accelerating Universe,” Phys. Lett. B 696 (2011) 273–277,
arXiv:1002.4278 [hep-th].
[2717] H. Gohar and V. Salzano, “Generalized mass-to-horizon relation: A new global approach to entropic cosmologies and
its connection to ΛCDM,” Phys. Rev. D 109 (2024) no. 8, 084075, arXiv:2307.06239 [gr-qc].
[2718] N. Komatsu and S. Kimura, “Entropic cosmology for a generalized black-hole entropy,” Phys. Rev. D 88 (2013)
083534, arXiv:1307.5949 [astro-ph.CO].
[2719] R. C. Nunes, E. M. Barboza, Jr., E. M. C. Abreu, and J. A. Neto, “Probing the cosmological viability of non-gaussian
statistics,” JCAP 08 (2016) 051, arXiv:1509.05059 [gr-qc].
[2720] H. Moradpour, R. C. Nunes, E. M. C. Abreu, and J. A. Neto, “A note on the relations between thermodynamics, energy
definitions and Friedmann equations,” Mod. Phys. Lett. A 32 (2017) no. 13, 1750078, arXiv:1603.01465 [gr-qc].
[2721] E. M. C. Abreu, J. A. Neto, E. M. Barboza, and R. C. Nunes, “Tsallis and Kaniadakis statistics from the viewpoint of
entropic gravity formalism,” Int. J. Mod. Phys. A 32 (2017) no. 05, 1750028, arXiv:1701.06898 [gr-qc].
[2722] C. Tsallis and L. J. L. Cirto, “Black hole thermodynamical entropy,” Eur. Phys. J. C 73 (2013) 2487,
arXiv:1202.2154 [cond-mat.stat-mech].
[2723] J. D. Barrow, “The Area of a Rough Black Hole,” Phys. Lett. B 808 (2020) 135643, arXiv:2004.09444 [gr-qc].
[2724] I. Cimidiker, M. P. Dabrowski, and H. Gohar, “Generalized uncertainty principle impact on nonextensive black hole
thermodynamics,” Class. Quant. Grav. 40 (2023) no. 14, 145001, arXiv:2301.00609 [gr-qc].
[2725] J. D. Bekenstein, “Universal upper bound on the entropy-to-energy ratio for bounded systems,” Phys. Rev. D 23
(1981) no. 2, 287–298.
[2726] M. H. P. M. van Putten, “On the Hubble expansion in a Big Bang quantum cosmology,” JHEAp 45 (2025) 194–199,
arXiv:2403.10865 [astro-ph.CO].
[2727] D. Camarena and V. Marra, “Local determination of the Hubble constant and the deceleration parameter,” Phys. Rev.
Res. 2 (2020) no. 1, 013028, arXiv:1906.11814 [astro-ph.CO].
[2728] B. De Simone, M. H. P. M. van Putten, M. G. Dainotti, and G. Lambiase, “A doublet of cosmological models to
challenge the H0 tension in the Pantheon Supernovae Ia catalog,” JHEAp 45 (2025) 290–298, arXiv:2411.05744
[astro-ph.CO].
[2729] S. Basilakos, A. Lymperis, M. Petronikolou, and E. N. Saridakis, “Alleviating both H0 and σ8 tensions in Tsallis
cosmology,” Eur. Phys. J. C 84 (2024) no. 3, 297, arXiv:2308.01200 [gr-qc].
353

[2730] M. Asghari and A. Sheykhi, “Observational constraints on Tsallis modified gravity,” Mon. Not. Roy. Astron. Soc. 508
(2021) no. 2, 2855–2861, arXiv:2106.15551 [gr-qc].
[2731] H. Gohar and V. Salzano, “Cosmological Constraints on Entropic Cosmology with Matter Creation,” Eur. Phys. J. C
81 (2021) no. 4, 338, arXiv:2008.09635 [gr-qc].
[2732] W. J. C. da Silva and R. Silva, “Cosmological Perturbations in the Tsallis Holographic Dark Energy Scenarios,” Eur.
Phys. J. Plus 136 (2021) no. 5, 543, arXiv:2011.09520 [astro-ph.CO].
[2733] A. Melchiorri, L. Mersini-Houghton, C. J. Odman, and M. Trodden, “The State of the dark energy equation of state,”
Phys. Rev. D 68 (2003) 043509, arXiv:astro-ph/0211522.
[2734] A. Melchiorri and L. Mersini-Houghton, “Does the low CMB quadrupole provide a new cosmic coincidence problem?,”
arXiv:hep-ph/0403222.
[2735] P. J. Steinhardt, L.-M. Wang, and I. Zlatev, “Cosmological tracking solutions,” Phys. Rev. D 59 (1999) 123504,
arXiv:astro-ph/9812313.
[2736] N. Roy, A. X. Gonzalez-Morales, and L. A. Urena-Lopez, “New general parametrization of quintessence fields and its
observational constraints,” Phys. Rev. D 98 (2018) no. 6, 063530, arXiv:1803.09204 [gr-qc].
[2737] F. X. Linares Cedeño, N. Roy, and L. A. Ureña López, “Tracker phantom field and a cosmological constant: Dynamics
of a composite dark energy model,” Phys. Rev. D 104 (2021) no. 12, 123502, arXiv:2105.07103 [astro-ph.CO].
[2738] N. Roy, S. Goswami, and S. Das, “Quintessence or phantom: Study of scalar field dark energy models through a general
parametrization of the Hubble parameter,” Phys. Dark Univ. 36 (2022) 101037, arXiv:2201.09306 [astro-ph.CO].
[2739] J. A. Nájera and C. Escamilla-Rivera, “Phantom Scalar Field Cosmologies Constrained by Early Cosmic
Measurements,” Universe 10 (2024) no. 6, 232, arXiv:2403.16562 [gr-qc].
[2740] WMAP Collaboration, G. Hinshaw et al., “Nine-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations:
Cosmological Parameter Results,” Astrophys. J. Suppl. 208 (2013) 19, arXiv:1212.5226 [astro-ph.CO].
[2741] B. Feng, X.-L. Wang, and X.-M. Zhang, “Dark energy constraints from the cosmic age and supernova,” Phys. Lett. B
607 (2005) 35–41, arXiv:astro-ph/0404224.
[2742] T. Qiu, “Theoretical Aspects of Quintom Models,” Mod. Phys. Lett. A 25 (2010) 909–921, arXiv:1002.3971 [hep-th].
[2743] M.-J. Zhang and H. Li, “Observational constraint on the dark energy scalar field,” Chin. Phys. C 45 (2021) no. 4,
045103, arXiv:1809.08936 [astro-ph.CO].
[2744] G. Leon, A. Paliathanasis, and J. L. Morales-Martínez, “The past and future dynamics of quintom dark energy
models,” Eur. Phys. J. C 78 (2018) no. 9, 753, arXiv:1808.05634 [gr-qc].
[2745] S. Panpanich, P. Burikham, S. Ponglertsakul, and L. Tannukij, “Resolving Hubble Tension with Quintom Dark Energy
Model,” Chin. Phys. C 45 (2021) no. 1, 015108, arXiv:1908.03324 [gr-qc].
[2746] D. Wang, W. Zhang, and X.-H. Meng, “Searching for the evidence of dynamical dark energy,” Eur. Phys. J. C 79
(2019) no. 3, 211, arXiv:1903.08913 [astro-ph.CO].
[2747] L. A. Escamilla, S. Pan, E. Di Valentino, A. Paliathanasis, J. A. Vázquez, and W. Yang, “Testing an oscillatory
behavior of dark energy,” Phys. Rev. D 111 (2025) no. 2, 023531, arXiv:2404.00181 [astro-ph.CO].
[2748] J. A. Vázquez, D. Tamayo, A. A. Sen, and I. Quiros, “Bayesian model selection on scalar ϵ-field dark energy,” Phys.
Rev. D 103 (2021) no. 4, 043506, arXiv:2009.01904 [gr-qc].
[2749] J.-Q. Xia, H. Li, and X. Zhang, “Dark Energy Constraints after Planck,” Phys. Rev. D 88 (2013) 063501,
arXiv:1308.0188 [astro-ph.CO].
[2750] L. Fu, L. Chen, M. Yang, J. Wang, and M.-J. Zhang, “A Better Reconciliation of Hubble Tension in the Dark Energy
Scalar Field,” Res. Astron. Astrophys. 23 (2023) no. 3, 035004.
[2751] N. Roy and L. A. Ureña López, “Tracker behaviour of quintom dark energy and the Hubble tension,”
arXiv:2312.04003 [astro-ph.CO].
[2752] M. Cicoli, G. Dibitetto, and F. G. Pedro, “New accelerating solutions in late-time cosmology,” Phys. Rev. D 101
(2020) no. 10, 103524, arXiv:2002.02695 [gr-qc].
[2753] M. Cicoli, G. Dibitetto, and F. G. Pedro, “Out of the Swampland with Multifield Quintessence?,” JHEP 10 (2020) 035,
arXiv:2007.11011 [hep-th].
[2754] Y. Akrami, M. Sasaki, A. R. Solomon, and V. Vardanyan, “Multi-field dark energy: Cosmic acceleration on a steep
potential,” Phys. Lett. B 819 (2021) 136427, arXiv:2008.13660 [astro-ph.CO].
[2755] J. R. Eskilt, Y. Akrami, A. R. Solomon, and V. Vardanyan, “Cosmological dynamics of multifield dark energy,” Phys.
Rev. D 106 (2022) no. 2, 023512, arXiv:2201.08841 [astro-ph.CO].
[2756] L. Anguelova, J. Dumancic, R. Gass, and L. C. R. Wijewardhana, “Dark energy from inspiraling in field space,” JCAP
03 (2022) no. 03, 018, arXiv:2111.12136 [hep-th].
[2757] L. Anguelova, J. Dumancic, R. Gass, and L. C. R. Wijewardhana, “Dynamics of inspiraling dark energy,” Eur. Phys. J.
C 84 (2024) no. 4, 365, arXiv:2311.07839 [hep-th].
[2758] C. Armendariz-Picon, T. Damour, and V. F. Mukhanov, “k - inflation,” Phys. Lett. B 458 (1999) 209–218,
arXiv:hep-th/9904075.
[2759] J. Garriga and V. F. Mukhanov, “Perturbations in k-inflation,” Phys. Lett. B 458 (1999) 219–225,
arXiv:hep-th/9904176.
[2760] R. C. Batista, “A Short Review on Clustering Dark Energy,” Universe 8 (2021) no. 1, 22, arXiv:2204.12341
[astro-ph.CO].
[2761] J. Dakin, S. Hannestad, T. Tram, M. Knabenhans, and J. Stadel, “Dark energy perturbations in N -body simulations,”
JCAP 08 (2019) 013, arXiv:1904.05210 [astro-ph.CO].
[2762] F. Hassani, J. Adamek, M. Kunz, and F. Vernizzi, “k-evolution: a relativistic N-body code for clustering dark energy,”
354

JCAP 12 (2019) 011, arXiv:1910.01104 [astro-ph.CO].


[2763] L. Blot, P. S. Corasaniti, and F. Schmidt, “Non-linear Eulerian hydrodynamics of dark energy: Riemann problem and
finite volume schemes,” JCAP 05 (2023) 001, arXiv:2210.04800 [astro-ph.CO].
[2764] R. C. Batista, H. P. de Oliveira, and L. R. W. Abramo, “Spherical collapse of non-top-hat profiles in the presence of
dark energy with arbitrary sound speed,” JCAP 02 (2023) 037, arXiv:2210.14769 [astro-ph.CO].
[2765] I. Ben-Dayan and U. Kumar, “Theoretical priors and the dark energy equation of state,” Eur. Phys. J. C 84 (2024)
no. 2, 167, arXiv:2310.03092 [astro-ph.CO].
[2766] B. R. Dinda and N. Banerjee, “Constraints on the speed of sound in the k-essence model of dark energy,” Eur. Phys. J.
C 84 (2024) no. 2, 177, arXiv:2309.10538 [astro-ph.CO].
[2767] M. Bastero-Gil, P. H. Frampton, and L. Mersini-Houghton, “Modified dispersion relations from closed strings in
toroidal cosmology,” Phys. Rev. D 65 (2002) 106002, arXiv:hep-th/0110167.
[2768] L. Mersini-Houghton, M. Bastero-Gil, and P. Kanti, “Relic dark energy from transPlanckian regime,” Phys. Rev. D 64
(2001) 043508, arXiv:hep-ph/0101210.
[2769] L. Mersini-Houghton and M. Bastero-Gil, “Dark energy may probe string theory,” in 3rd International Sakharov
Conference on Physics. 2002. arXiv:hep-th/0212153.
[2770] M. Bastero-Gil and L. Mersini-Houghton, “SN1A data and the CMB of Modified Curvature at short and long
distances,” Phys. Rev. D 65 (2002) 023502, arXiv:astro-ph/0107256.
[2771] M. Bastero-Gil and L. Mersini-Houghton, “Equation of state of the transPlanckian dark energy and the coincidence
problem,” Phys. Rev. D 67 (2003) 103519, arXiv:hep-th/0205271.
[2772] M. Bastero-Gil, K. Freese, and L. Mersini-Houghton, “What can WMAP tell us about the very early universe? New
physics as an explanation of suppressed large scale power and running spectral index,” Phys. Rev. D 68 (2003) 123514,
arXiv:hep-ph/0306289.
[2773] I. L. Shapiro and J. Sola, “Scaling behavior of the cosmological constant: Interface between quantum field theory and
cosmology,” JHEP 02 (2002) 006, arXiv:hep-th/0012227.
[2774] J. Sola, “Dark energy: A Quantum fossil from the inflationary Universe?,” J. Phys. A 41 (2008) 164066,
arXiv:0710.4151 [hep-th].
[2775] I. L. Shapiro and J. Sola, “On the possible running of the cosmological ’constant’,” Phys. Lett. B 682 (2009) 105–113,
arXiv:0910.4925 [hep-th].
[2776] C. Moreno-Pulido and J. Sola, “Running vacuum in quantum field theory in curved spacetime: renormalizing ρvac
without ∼ m4 terms,” Eur. Phys. J. C 80 (2020) no. 8, 692, arXiv:2005.03164 [gr-qc].
[2777] C. Moreno-Pulido and J. Sola Peracaula, “Renormalizing the vacuum energy in cosmological spacetime: implications
for the cosmological constant problem,” Eur. Phys. J. C 82 (2022) no. 6, 551, arXiv:2201.05827 [gr-qc].
[2778] C. Moreno-Pulido, J. Sola Peracaula, and S. Cheraghchi, “Running vacuum in QFT in FLRW spacetime: the dynamics
of ρvac (H) from the quantized matter fields,” Eur. Phys. J. C 83 (2023) no. 7, 637, arXiv:2301.05205 [gr-qc].
[2779] J. Sola, “Cosmological constant and vacuum energy: old and new ideas,” J. Phys. Conf. Ser. 453 (2013) 012015,
arXiv:1306.1527 [gr-qc].
[2780] J. Sola Peracaula, “The cosmological constant problem and running vacuum in the expanding universe,” Phil. Trans.
Roy. Soc. Lond. A 380 (2022) 20210182, arXiv:2203.13757 [gr-qc].
[2781] S. Basilakos, J. A. S. Lima, and J. Sola, “From inflation to dark energy through a dynamical Lambda: an attempt at
alleviating fundamental cosmic puzzles,” Int. J. Mod. Phys. D 22 (2013) 1342008, arXiv:1307.6251 [astro-ph.CO].
[2782] S. Basilakos, N. E. Mavromatos, and J. Solà Peracaula, “Scalar Field Theory Description of the Running Vacuum
Model: the Vacuumon,” JCAP 12 (2019) 025, arXiv:1901.06638 [gr-qc].
[2783] H. Fritzsch and J. Sola, “Matter Non-conservation in the Universe and Dynamical Dark Energy,” Class. Quant. Grav.
29 (2012) 215002, arXiv:1202.5097 [hep-ph].
[2784] A. Gomez-Valent and J. Sola, “Vacuum models with a linear and a quadratic term in H: structure formation and
number counts analysis,” Mon. Not. Roy. Astron. Soc. 448 (2015) 2810–2821, arXiv:1412.3785 [astro-ph.CO].
[2785] A. Gomez-Valent, E. Karimkhani, and J. Sola, “Background history and cosmic perturbations for a general system of
self-conserved dynamical dark energy and matter,” JCAP 12 (2015) 048, arXiv:1509.03298 [gr-qc].
[2786] H. Fritzsch and J. Sola, “Fundamental constants and cosmic vacuum: the micro and macro connection,” Mod. Phys.
Lett. A 30 (2015) no. 22, 1540034, arXiv:1502.01411 [gr-qc].
[2787] H. Fritzsch, J. Solà, and R. C. Nunes, “Running vacuum in the Universe and the time variation of the fundamental
constants of Nature,” Eur. Phys. J. C 77 (2017) no. 3, 193, arXiv:1605.06104 [hep-ph].
[2788] J. Solà Peracaula, “The dynamics of vacuum, gravity and matter: Implications on the fundamental constants,” Int. J.
Mod. Phys. A 39 (2024) no. 09n10, 2441016, arXiv:2308.13349 [gr-qc].
[2789] J. Sola, A. Gomez-Valent, and J. de Cruz Pérez, “Hints of dynamical vacuum energy in the expanding Universe,”
Astrophys. J. Lett. 811 (2015) L14, arXiv:1506.05793 [gr-qc].
[2790] J. Solà, A. Gómez-Valent, and J. de Cruz Pérez, “First evidence of running cosmic vacuum: challenging the
concordance model,” Astrophys. J. 836 (2017) no. 1, 43, arXiv:1602.02103 [astro-ph.CO].
[2791] J. Solà Peracaula, J. de Cruz Pérez, and A. Gómez-Valent, “Dynamical dark energy vs. Λ = const in light of
observations,” EPL 121 (2018) no. 3, 39001, arXiv:1606.00450 [gr-qc].
[2792] J. Solà Peracaula, J. de Cruz Pérez, and A. Gomez-Valent, “Possible signals of vacuum dynamics in the Universe,”
Mon. Not. Roy. Astron. Soc. 478 (2018) no. 4, 4357–4373, arXiv:1703.08218 [astro-ph.CO].
[2793] J. Solà, A. Gómez-Valent, and J. de Cruz Pérez, “The H0 tension in light of vacuum dynamics in the Universe,” Phys.
Lett. B 774 (2017) 317–324, arXiv:1705.06723 [astro-ph.CO].
355

[2794] A. Gomez-Valent and J. Sola, “Relaxing the σ8 -tension through running vacuum in the Universe,” EPL 120 (2017)
no. 3, 39001, arXiv:1711.00692 [astro-ph.CO].
[2795] A. Gómez-Valent and J. Solà Peracaula, “Density perturbations for running vacuum: a successful approach to
structure formation and to the σ8 -tension,” Mon. Not. Roy. Astron. Soc. 478 (2018) no. 1, 126–145, arXiv:1801.08501
[astro-ph.CO].
[2796] A. Gómez-Valent, J. Solà, and S. Basilakos, “Dynamical vacuum energy in the expanding Universe confronted with
observations: a dedicated study,” JCAP 01 (2015) 004, arXiv:1409.7048 [astro-ph.CO].
[2797] C.-Q. Geng, C.-C. Lee, and L. Yin, “Constraints on running vacuum model with H(z) and f σ8 ,” JCAP 08 (2017) 032,
arXiv:1704.02136 [astro-ph.CO].
[2798] P. Tsiapi and S. Basilakos, “Testing dynamical vacuum models with CMB power spectrum from Planck,” Mon. Not.
Roy. Astron. Soc. 485 (2019) no. 2, 2505–2510, arXiv:1810.12902 [astro-ph.CO].
[2799] P. Asimakis, S. Basilakos, N. E. Mavromatos, and E. N. Saridakis, “Big bang nucleosynthesis constraints on
higher-order modified gravities,” Phys. Rev. D 105 (2022) no. 8, 084010, arXiv:2112.10863 [gr-qc].
[2800] J. Solà Peracaula, A. Gómez-Valent, J. de Cruz Perez, and C. Moreno-Pulido, “Running vacuum against the H0 and
σ8 tensions,” EPL 134 (2021) no. 1, 19001, arXiv:2102.12758 [astro-ph.CO].
[2801] J. Sola Peracaula, A. Gomez-Valent, J. de Cruz Perez, and C. Moreno-Pulido, “Running Vacuum in the Universe:
Phenomenological Status in Light of the Latest Observations, and Its Impact on the σ 8 and H0 Tensions,” Universe 9
(2023) no. 6, 262, arXiv:2304.11157 [astro-ph.CO].
[2802] J. de Cruz Perez and J. Sola Peracaula, “Running vacuum in Brans & Dicke theory: A possible cure for the σ8 and H0
tensions,” Phys. Dark Univ. 43 (2024) 101406, arXiv:2302.04807 [astro-ph.CO].
[2803] N. E. Mavromatos, “Geometrical origins of the universe dark sector: string-inspired torsion and anomalies as seeds for
inflation and dark matter,” Phil. Trans. A. Math. Phys. Eng. Sci. 380 (2022) no. 2222, 20210188, arXiv:2108.02152
[gr-qc].
[2804] N. E. Mavromatos, “Lorentz Symmetry Violation in String-Inspired Effective Modified Gravity Theories,” Lect. Notes
Phys. 1017 (2023) 3–48, arXiv:2205.07044 [hep-th].
[2805] J. Alexandre, N. Houston, and N. E. Mavromatos, “Starobinsky-type Inflation in Dynamical Supergravity Breaking
Scenarios,” Phys. Rev. D 89 (2014) no. 2, 027703, arXiv:1312.5197 [gr-qc].
[2806] W. Yang, E. Di Valentino, S. Pan, A. Shafieloo, and X. Li, “Generalized emergent dark energy model and the Hubble
constant tension,” Phys. Rev. D 104 (2021) no. 6, 063521, arXiv:2103.03815 [astro-ph.CO].
[2807] A. Banihashemi, N. Khosravi, and A. H. Shirazi, “Phase transition in the dark sector as a proposal to lessen
cosmological tensions,” Phys. Rev. D 101 (2020) no. 12, 123521, arXiv:1808.02472 [astro-ph.CO].
[2808] A. Banihashemi, N. Khosravi, and A. H. Shirazi, “Ginzburg-Landau Theory of Dark Energy: A Framework to Study
Both Temporal and Spatial Cosmological Tensions Simultaneously,” Phys. Rev. D 99 (2019) no. 8, 083509,
arXiv:1810.11007 [astro-ph.CO].
[2809] A. Banihashemi, N. Khosravi, and A. Shafieloo, “Dark energy as a critical phenomenon: a hint from Hubble tension,”
JCAP 06 (2021) 003, arXiv:2012.01407 [astro-ph.CO].
[2810] V. L. Ginzburg and L. D. Landau, “On the Theory of superconductivity,” Zh. Eksp. Teor. Fiz. 20 (1950) 1064–1082.
[2811] A. Banihashemi and N. Khosravi, “Fluctuations in the Ginzburg–Landau Theory of Dark Energy: Internal
(In)consistencies in the Planck Data Set,” Astrophys. J. 931 (2022) no. 2, 148, arXiv:2201.04119 [astro-ph.CO].
[2812] M. Rezaei, T. Naderi, M. Malekjani, and A. Mehrabi, “A Bayesian comparison between ΛCDM and phenomenologically
emergent dark energy models,” Eur. Phys. J. C 80 (2020) no. 5, 374, arXiv:2004.08168 [astro-ph.CO].
[2813] S. Pan, W. Yang, E. Di Valentino, A. Shafieloo, and S. Chakraborty, “Reconciling H0 tension in a six parameter
space?,” JCAP 06 (2020) no. 06, 062, arXiv:1907.12551 [astro-ph.CO].
[2814] R. Shah, P. Mukherjee, and S. Pal, “Reconciling S8: insights from interacting dark sectors,” Mon. Not. Roy. Astron.
Soc. 536 (2024) no. 3, 2404–2420, arXiv:2404.06396 [astro-ph.CO].
[2815] A. Hernández-Almada, G. Leon, J. Magaña, M. A. García-Aspeitia, and V. Motta, “Generalized Emergent Dark
Energy: observational Hubble data constraints and stability analysis,” Mon. Not. Roy. Astron. Soc. 497 (2020) no. 2,
1590–1602, arXiv:2002.12881 [astro-ph.CO].
[2816] W. Yang, E. Di Valentino, S. Pan, and O. Mena, “Emergent Dark Energy, neutrinos and cosmological tensions,” Phys.
Dark Univ. 31 (2021) 100762, arXiv:2007.02927 [astro-ph.CO].
[2817] Z. Liu and H. Miao, “Update constraints on neutrino mass and mass hierarchy in light of dark energy models,” Int. J.
Mod. Phys. D 29 (2020) no. 13, 2050088, arXiv:2002.05563 [astro-ph.CO].
[2818] E. Di Valentino, S. Gariazzo, C. Giunti, O. Mena, S. Pan, and W. Yang, “Minimal dark energy: Key to sterile neutrino
and Hubble constant tensions?,” Phys. Rev. D 105 (2022) no. 10, 103511, arXiv:2110.03990 [astro-ph.CO].
[2819] M. A. García-Aspeitia, G. Fernandez-Anaya, A. Hernández-Almada, G. Leon, and J. Magaña, “Cosmology under the
fractional calculus approach,” Mon. Not. Roy. Astron. Soc. 517 (2022) no. 4, 4813–4826, arXiv:2207.00878 [gr-qc].
[2820] H. B. Benaoum, W. Yang, S. Pan, and E. Di Valentino, “Modified emergent dark energy and its astronomical
constraints,” Int. J. Mod. Phys. D 31 (2022) no. 03, 2250015, arXiv:2008.09098 [gr-qc].
[2821] L. Parker and A. Raval, “New quantum aspects of a vacuum dominated universe,” Phys. Rev. D 62 (2000) 083503,
arXiv:gr-qc/0003103. [Erratum: Phys.Rev.D 67, 029903 (2003)].
[2822] L. Parker and D. A. T. Vanzella, “Acceleration of the universe, vacuum metamorphosis, and the large time asymptotic
form of the heat kernel,” Phys. Rev. D 69 (2004) 104009, arXiv:gr-qc/0312108.
[2823] R. R. Caldwell, W. Komp, L. Parker, and D. A. T. Vanzella, “A Sudden gravitational transition,” Phys. Rev. D 73
(2006) 023513, arXiv:astro-ph/0507622.
356

[2824] E. Di Valentino, E. V. Linder, and A. Melchiorri, “H0 ex machina: Vacuum metamorphosis and beyond H0 ,” Phys.
Dark Univ. 30 (2020) 100733, arXiv:2006.16291 [astro-ph.CO].
[2825] G. Lambiase, S. Mohanty, A. Narang, and P. Parashari, “Testing dark energy models in the light of σ8 tension,” Eur.
Phys. J. C 79 (2019) no. 2, 141, arXiv:1804.07154 [astro-ph.CO].
[2826] R. A. Battye and A. Moss, “Evidence for Massive Neutrinos from Cosmic Microwave Background and Lensing
Observations,” Phys. Rev. Lett. 112 (2014) no. 5, 051303, arXiv:1308.5870 [astro-ph.CO].
[2827] Y. Wang, K. Freese, P. Gondolo, and M. Lewis, “Future Type IA supernova data as tests of dark energy from modified
Friedmann equations,” Astrophys. J. 594 (2003) 25–32, arXiv:astro-ph/0302064.
[2828] R. Lazkoz, V. Salzano, L. Fernandez-Jambrina, and M. Bouhmadi-López, “Ripped ΛCDM: An observational contender
to the consensus cosmological model,” Phys. Dark Univ. 45 (2024) 101511, arXiv:2311.10526 [astro-ph.CO].
[2829] L. Huang, S.-J. Wang, and W.-W. Yu, “No-go guide for the Hubble tension: Late-time or local-scale new physics,” Sci.
China Phys. Mech. Astron. 68 (2025) no. 2, 220413, arXiv:2401.14170 [astro-ph.CO].
[2830] N. Kitazawa, “Late-time data require smaller sound horizon at recombination,” arXiv:2310.10017 [astro-ph.CO].
[2831] Ruchika, “2D BAO vs 3D BAO: Hints for new physics?,” arXiv:2406.05453 [astro-ph.CO].
[2832] R. E. Keeley and A. Shafieloo, “Ruling Out New Physics at Low Redshift as a Solution to the H0 Tension,” Phys. Rev.
Lett. 131 (2023) no. 11, 111002, arXiv:2206.08440 [astro-ph.CO].
[2833] M. W. Hossain and A. Maqsood, “Comparison between axionlike and power law potentials in a cosmological
background,” Phys. Rev. D 109 (2024) no. 10, 103512, arXiv:2311.17825 [astro-ph.CO].
[2834] C. G. Boiza and M. Bouhmadi-López, “Speeding up the Universe with a generalised axion-like potential,”
arXiv:2409.18184 [astro-ph.CO].
[2835] C. G. Boiza and M. Bouhmadi-López, “Cosmological perturbations in a generalised axion-like dark energy model,”
Phys. Dark Univ. 48 (2025) 101845, arXiv:2410.22467 [astro-ph.CO].
[2836] H.-W. Chiang, C. G. Boiza, and M. Bouhmadi-López, “Observational constraints on generalised axion-like potentials
for the late Universe,” arXiv:2503.04898 [astro-ph.CO].
[2837] H. Benaoum, “Accelerated Universe from Modified Chaplygin Gas and Tachyonic Fluid,” Universe 8 (2022) no. 7, 340,
arXiv:hep-th/0205140.
[2838] H. Stefancic, “Expansion around the vacuum equation of state - Sudden future singularities and asymptotic behavior,”
Phys. Rev. D 71 (2005) 084024, arXiv:astro-ph/0411630.
[2839] M. Bouhmadi-López, M. Brilenkov, R. Brilenkov, J. a. Morais, and A. Zhuk, “Scalar perturbations in the late
Universe: viability of the Chaplygin gas models,” JCAP 12 (2015) 037, arXiv:1509.06963 [gr-qc].
[2840] W. Yang, S. Pan, S. Vagnozzi, E. Di Valentino, D. F. Mota, and S. Capozziello, “Dawn of the dark: unified dark
sectors and the EDGES Cosmic Dawn 21-cm signal,” JCAP 11 (2019) 044, arXiv:1907.05344 [astro-ph.CO].
[2841] I. Albarran, M. Bouhmadi-López, and J. a. Morais, “Cosmological perturbations in an effective and genuinely phantom
dark energy Universe,” Phys. Dark Univ. 16 (2017) 94–108, arXiv:1611.00392 [astro-ph.CO].
[2842] K. R. Dienes, L. Heurtier, F. Huang, D. Kim, T. M. P. Tait, and B. Thomas, “Stasis in an expanding universe: A
recipe for stable mixed-component cosmological eras,” Phys. Rev. D 105 (2022) no. 2, 023530, arXiv:2111.04753
[astro-ph.CO].
[2843] K. R. Dienes, L. Heurtier, F. Huang, T. M. P. Tait, and B. Thomas, “Stasis, Stasis, Triple Stasis,” Phys. Rev. D 109
(2024) no. 8, 083508, arXiv:2309.10345 [astro-ph.CO].
[2844] K. R. Dienes, L. Heurtier, F. Huang, T. M. P. Tait, and B. Thomas, “Cosmological stasis from dynamical scalars:
Tracking solutions and the possibility of a stasis-induced inflation,” Phys. Rev. D 110 (2024) no. 12, 123514,
arXiv:2406.06830 [astro-ph.CO].
[2845] L. R. W. Abramo, R. H. Brandenberger, and V. F. Mukhanov, “The Energy - momentum tensor for cosmological
perturbations,” Phys. Rev. D 56 (1997) 3248–3257, arXiv:gr-qc/9704037.
[2846] R. H. Brandenberger, “Back reaction of cosmological perturbations,” in 3rd International Conference on Particle
Physics and the Early Universe, pp. 198–206. 2000. arXiv:hep-th/0004016.
[2847] M. A. C. Alvarez, L. Graef, and R. Brandenberger, “Back-Reaction of Super-Hubble Fluctuations, Late Time Tracking
and Recent Observational Results,” arXiv:2502.17395 [astro-ph.CO].
[2848] L. A. Escamilla, E. Özülker, O. Akarsu, E. Di Valentino, and J. A. Vázquez, “Do we need wavelets in the late
Universe?,” arXiv:2408.12516 [astro-ph.CO].
[2849] O. Akarsu, A. Çam, E. A. Paraskevas, and L. Perivolaropoulos, “Linear matter density perturbations in the Λs CDM
model: Examining growth dynamics and addressing the S8 tension,” arXiv:2502.20384 [astro-ph.CO].
[2850] M. Biagetti, G. Franciolini, and A. Riotto, “Enhancing massive galaxy formation at high redshift in non-standard
cosmologies,” Astrophys. J. 944 (2023) no. 2, 113, arXiv:2210.04812 [astro-ph.CO].
[2851] D. Wang and Y. Liu, “JWST high redshift galaxy observations have a strong tension with Planck CMB
measurements,” arXiv:2301.00347 [astro-ph.CO].
[2852] K. Dutta, Ruchika, A. Roy, A. A. Sen, and M. M. Sheikh-Jabbari, “Beyond ΛCDM with low and high redshift data:
implications for dark energy,” Gen. Rel. Grav. 52 (2020) no. 2, 15, arXiv:1808.06623 [astro-ph.CO].
[2853] G. Acquaviva, O. Akarsu, N. Katirci, and J. A. Vazquez, “Simple-graduated dark energy and spatial curvature,” Phys.
Rev. D 104 (2021) no. 2, 023505, arXiv:2104.02623 [astro-ph.CO].
[2854] O. Akarsu, L. Perivolaropoulos, A. Tsikoundoura, A. E. Yükselci, and A. Zhuk, “Dynamical dark energy with
AdS-to-dS and dS-to-dS transitions: Implications for the H0 tension,” arXiv:2502.14667 [astro-ph.CO].
[2855] E. Ozulker, “Is the dark energy equation of state parameter singular?,” Phys. Rev. D 106 (2022) no. 6, 063509,
arXiv:2203.04167 [astro-ph.CO].
357

[2856] O. Akarsu, A. De Felice, E. Di Valentino, S. Kumar, R. C. Nunes, E. Ozulker, J. A. Vazquez, and A. Yadav, “Λs CDM
cosmology from a type-II minimally modified gravity,” arXiv:2402.07716 [astro-ph.CO].
[2857] D. Green and J. Meyers, “The Cosmological Preference for Negative Neutrino Mass,” arXiv:2407.07878
[astro-ph.CO].
[2858] W. Elbers, C. S. Frenk, A. Jenkins, B. Li, and S. Pascoli, “Negative neutrino masses as a mirage of dark energy,” Phys.
Rev. D 111 (2025) no. 6, 063534, arXiv:2407.10965 [astro-ph.CO].
[2859] S.-F. Ge and L. Tan, “Capability of Cosmic Gravitational Focusing on Identifying the Neutrino Mass Ordering,”
arXiv:2409.11115 [hep-ph].
[2860] J. D. Barrow, “Sudden future singularities,” Class. Quant. Grav. 21 (2004) L79–L82, arXiv:gr-qc/0403084.
[2861] A. De Felice, A. Doll, and S. Mukohyama, “A theory of type-II minimally modified gravity,” JCAP 09 (2020) 034,
arXiv:2004.12549 [gr-qc].
[2862] A. De Felice, S. Mukohyama, and M. C. Pookkillath, “Addressing H0 tension by means of VCDM,” Phys. Lett. B 816
(2021) 136201, arXiv:2009.08718 [astro-ph.CO]. [Erratum: Phys.Lett.B 818, 136364 (2021)].
[2863] J. A. Vazquez, S. Hee, M. P. Hobson, A. N. Lasenby, M. Ibison, and M. Bridges, “Observational constraints on
conformal time symmetry, missing matter and double dark energy,” JCAP 07 (2018) 062, arXiv:1208.2542
[astro-ph.CO].
[2864] BOSS Collaboration, T. Delubac et al., “Baryon acoustic oscillations in the Lyα forest of BOSS DR11 quasars,”
Astron. Astrophys. 574 (2015) A59, arXiv:1404.1801 [astro-ph.CO].
[2865] V. Sahni and Y. Shtanov, “Brane world models of dark energy,” JCAP 11 (2003) 014, arXiv:astro-ph/0202346.
[2866] S. Bag, V. Sahni, A. Shafieloo, and Y. Shtanov, “Phantom Braneworld and the Hubble Tension,” Astrophys. J. 923
(2021) no. 2, 212, arXiv:2107.03271 [astro-ph.CO].
[2867] E. Di Valentino, E. V. Linder, and A. Melchiorri, “Vacuum phase transition solves the H0 tension,” Phys. Rev. D 97
(2018) no. 4, 043528, arXiv:1710.02153 [astro-ph.CO].
[2868] E. Mörtsell and S. Dhawan, “Does the Hubble constant tension call for new physics?,” JCAP 09 (2018) 025,
arXiv:1801.07260 [astro-ph.CO].
[2869] V. Poulin, K. K. Boddy, S. Bird, and M. Kamionkowski, “Implications of an extended dark energy cosmology with
massive neutrinos for cosmological tensions,” Phys. Rev. D 97 (2018) no. 12, 123504, arXiv:1803.02474
[astro-ph.CO].
[2870] S. Capozziello, Ruchika, and A. A. Sen, “Model independent constraints on dark energy evolution from low-redshift
observations,” Mon. Not. Roy. Astron. Soc. 484 (2019) 4484, arXiv:1806.03943 [astro-ph.CO].
[2871] O. Akarsu, J. D. Barrow, C. V. R. Board, N. M. Uzun, and J. A. Vazquez, “Screening Λ in a new modified gravity
model,” Eur. Phys. J. C 79 (2019) no. 10, 846, arXiv:1903.11519 [gr-qc].
[2872] L. Visinelli, S. Vagnozzi, and U. Danielsson, “Revisiting a negative cosmological constant from low-redshift data,”
Symmetry 11 (2019) no. 8, 1035, arXiv:1907.07953 [astro-ph.CO].
[2873] A. Perez, D. Sudarsky, and E. Wilson-Ewing, “Resolving the H0 tension with diffusion,” Gen. Rel. Grav. 53 (2021)
no. 1, 7, arXiv:2001.07536 [astro-ph.CO].
[2874] O. Akarsu, N. Katırcı, S. Kumar, R. C. Nunes, B. Öztürk, and S. Sharma, “Rastall gravity extension of the standard
ΛCDM model: theoretical features and observational constraints,” Eur. Phys. J. C 80 (2020) no. 11, 1050,
arXiv:2004.04074 [astro-ph.CO].
[2875] R. Calderón, R. Gannouji, B. L’Huillier, and D. Polarski, “Negative cosmological constant in the dark sector?,” Phys.
Rev. D 103 (2021) no. 2, 023526, arXiv:2008.10237 [astro-ph.CO].
[2876] G. Ye and Y.-S. Piao, “T0 censorship of early dark energy and AdS vacua,” Phys. Rev. D 102 (2020) no. 8, 083523,
arXiv:2008.10832 [astro-ph.CO].
[2877] A. Paliathanasis and G. Leon, “Dynamics of a two scalar field cosmological model with phantom terms,” Class. Quant.
Grav. 38 (2021) no. 7, 075013, arXiv:2009.12874 [gr-qc].
[2878] O. Akarsu, E. O. Colgain, E. Özulker, S. Thakur, and L. Yin, “Inevitable manifestation of wiggles in the expansion of
the late Universe,” Phys. Rev. D 107 (2023) no. 12, 123526, arXiv:2207.10609 [astro-ph.CO].
[2879] S. Di Gennaro and Y. C. Ong, “Sign Switching Dark Energy from a Running Barrow Entropy,” Universe 8 (2022)
no. 10, 541, arXiv:2205.09311 [gr-qc].
[2880] Y. C. Ong, “An Effective Sign Switching Dark Energy: Lotka–Volterra Model of Two Interacting Fluids,” Universe 9
(2023) no. 10, 437, arXiv:2212.04429 [gr-qc].
[2881] B. Alexandre, S. Gielen, and J. a. Magueijo, “Overall signature of the metric and the cosmological constant,” JCAP 02
(2024) 036, arXiv:2306.11502 [hep-th].
[2882] Y. Tiwari, B. Ghosh, and R. K. Jain, “Towards a possible solution to the Hubble tension with Horndeski gravity,” Eur.
Phys. J. C 84 (2024) no. 3, 220, arXiv:2301.09382 [astro-ph.CO].
[2883] L. A. Anchordoqui, I. Antoniadis, and D. Lust, “Anti-de Sitter → de Sitter transition driven by Casimir forces and
mitigating tensions in cosmological parameters,” Phys. Lett. B 855 (2024) 138775, arXiv:2312.12352 [hep-th].
[2884] L. A. Anchordoqui, I. Antoniadis, D. Bielli, A. Chatrabhuti, and H. Isono, “Thin-wall vacuum decay in the presence of
a compact dimension meets the H0 and S8 tensions,” arXiv:2410.18649 [hep-th].
[2885] H. Wang, Z.-Y. Peng, and Y.-S. Piao, “Can recent DESI BAO measurements accommodate a negative cosmological
constant?,” Phys. Rev. D 111 (2025) no. 6, L061306, arXiv:2406.03395 [astro-ph.CO].
[2886] Y. Toda, W. Giarè, E. Özülker, E. Di Valentino, and S. Vagnozzi, “Combining pre- and post-recombination new
physics to address cosmological tensions: Case study with varying electron mass and sign-switching cosmological
constant,” Phys. Dark Univ. 46 (2024) 101676, arXiv:2407.01173 [astro-ph.CO].
358

[2887] O. Akarsu, B. Bulduk, A. De Felice, N. Katırcı, and N. M. Uzun, “Unexplored regions in teleparallel f (T ) gravity:
Sign-changing dark energy density,” arXiv:2410.23068 [gr-qc].
[2888] M. S. Souza, A. M. Barcelos, R. C. Nunes, O. Akarsu, and S. Kumar, “Mapping the Λs CDM Scenario to f(T) Modified
Gravity: Effects on Structure Growth Rate,” Universe 11 (2025) no. 1, 2, arXiv:2501.18031 [astro-ph.CO].
[2889] P. Mukherjee, D. Kumar, and A. A. Sen, “Quintessential Implications of the presence of AdS in the Dark Energy
sector,” arXiv:2501.18335 [astro-ph.CO].
[2890] U. K. Tyagi, S. Haridasu, and S. Basak, “Holographic and gravity-thermodynamic approaches in entropic cosmology:
Bayesian assessment using late-time data,” Phys. Rev. D 110 (2024) no. 6, 063503, arXiv:2406.07446 [astro-ph.CO].
[2891] M. T. Manoharan, “Insights on Granda–Oliveros holographic dark energy: possibility of negative dark energy at
z ≳ 2,” Eur. Phys. J. C 84 (2024) no. 5, 552.
[2892] A. Gómez-Valent and J. Solà Peracaula, “Composite Dark Energy and the Cosmological Tensions,” arXiv:2412.15124
[astro-ph.CO]. arXiv:2412.15124.
[2893] S. Dwivedi and M. Högås, “2D BAO vs. 3D BAO: Solving the Hubble Tension with Bimetric Cosmology,” Universe 10
(2024) no. 11, 406, arXiv:2407.04322 [astro-ph.CO].
[2894] R. E. Keeley, K. N. Abazajian, M. Kaplinghat, and A. Shafieloo, “The Preference for Evolving Dark Energy from
Cosmological Distance Measurements and Possible Signatures in the Growth Rate of Perturbations,”
arXiv:2502.12667 [astro-ph.CO].
[2895] E. Di Valentino, A. Mukherjee, and A. A. Sen, “Dark Energy with Phantom Crossing and the H0 Tension,” Entropy 23
(2021) no. 4, 404, arXiv:2005.12587 [astro-ph.CO].
[2896] S. Kumar and R. C. Nunes, “Echo of interactions in the dark sector,” Phys. Rev. D 96 (2017) no. 10, 103511,
arXiv:1702.02143 [astro-ph.CO].
[2897] E. Di Valentino, A. Melchiorri, and O. Mena, “Can interacting dark energy solve the H0 tension?,” Phys. Rev. D 96
(2017) no. 4, 043503, arXiv:1704.08342 [astro-ph.CO].
[2898] W. Yang, A. Mukherjee, E. Di Valentino, and S. Pan, “Interacting dark energy with time varying equation of state and
the H0 tension,” Phys. Rev. D 98 (2018) no. 12, 123527, arXiv:1809.06883 [astro-ph.CO].
[2899] S. Pan, W. Yang, E. Di Valentino, E. N. Saridakis, and S. Chakraborty, “Interacting scenarios with dynamical dark
energy: Observational constraints and alleviation of the H0 tension,” Phys. Rev. D 100 (2019) no. 10, 103520,
arXiv:1907.07540 [astro-ph.CO].
[2900] S. Kumar, R. C. Nunes, and S. K. Yadav, “Dark sector interaction: a remedy of the tensions between CMB and LSS
data,” Eur. Phys. J. C 79 (2019) no. 7, 576, arXiv:1903.04865 [astro-ph.CO].
[2901] E. Di Valentino, A. Melchiorri, O. Mena, and S. Vagnozzi, “Nonminimal dark sector physics and cosmological
tensions,” Phys. Rev. D 101 (2020) no. 6, 063502, arXiv:1910.09853 [astro-ph.CO].
[2902] E. Di Valentino, A. Melchiorri, O. Mena, and S. Vagnozzi, “Interacting dark energy in the early 2020s: A promising
solution to the H0 and cosmic shear tensions,” Phys. Dark Univ. 30 (2020) 100666, arXiv:1908.04281 [astro-ph.CO].
[2903] M. Lucca and D. C. Hooper, “Shedding light on dark matter-dark energy interactions,” Phys. Rev. D 102 (2020)
no. 12, 123502, arXiv:2002.06127 [astro-ph.CO].
[2904] S. Kumar, “Remedy of some cosmological tensions via effective phantom-like behavior of interacting vacuum energy,”
Phys. Dark Univ. 33 (2021) 100862, arXiv:2102.12902 [astro-ph.CO].
[2905] R. C. Nunes, S. Vagnozzi, S. Kumar, E. Di Valentino, and O. Mena, “New tests of dark sector interactions from the
full-shape galaxy power spectrum,” Phys. Rev. D 105 (2022) no. 12, 123506, arXiv:2203.08093 [astro-ph.CO].
[2906] M. A. Sabogal, E. Silva, R. C. Nunes, S. Kumar, and E. Di Valentino, “Sign Switching in Dark Sector Coupling
Interactions as a Candidate for Resolving Cosmological Tensions,” arXiv:2501.10323 [astro-ph.CO].
[2907] M. Montero, C. Vafa, and I. Valenzuela, “The dark dimension and the Swampland,” JHEP 02 (2023) 022,
arXiv:2205.12293 [hep-th].
[2908] L. A. Anchordoqui and I. Antoniadis, “Large extra dimensions from higher-dimensional inflation,” Phys. Rev. D 109
(2024) no. 10, 103508, arXiv:2310.20282 [hep-ph].
[2909] N. Arkani-Hamed, S. Dubovsky, A. Nicolis, and G. Villadoro, “Quantum Horizons of the Standard Model Landscape,”
JHEP 06 (2007) 078, arXiv:hep-th/0703067.
[2910] J. F. Soriano, S. Wohlberg, and L. A. Anchordoqui, “New insights on a sign-switching Λ,” arXiv:2502.19239
[astro-ph.CO].
[2911] A. Awad, W. El Hanafy, G. G. L. Nashed, and E. N. Saridakis, “Phase Portraits of general f(T) Cosmology,” JCAP 02
(2018) 052, arXiv:1710.10194 [gr-qc].
[2912] M. Hashim, W. El Hanafy, A. Golovnev, and A. A. El-Zant, “Toward a concordance teleparallel cosmology. Part I.
Background dynamics,” JCAP 07 (2021) 052, arXiv:2010.14964 [astro-ph.CO].
[2913] M. Hashim, A. A. El-Zant, W. El Hanafy, and A. Golovnev, “Toward a concordance teleparallel cosmology. Part II.
Linear perturbation,” JCAP 07 (2021) 053, arXiv:2104.08311 [astro-ph.CO].
[2914] M. Cicoli, S. De Alwis, A. Maharana, F. Muia, and F. Quevedo, “De Sitter vs Quintessence in String Theory,” Fortsch.
Phys. 67 (2019) no. 1-2, 1800079, arXiv:1808.08967 [hep-th].
[2915] Ruchika, S. A. Adil, K. Dutta, A. Mukherjee, and A. A. Sen, “Observational constraints on axion(s) dark energy with
a cosmological constant,” Phys. Dark Univ. 40 (2023) 101199, arXiv:2005.08813 [astro-ph.CO].
[2916] J. Grande, J. Solà, and H. Stefancic, “LXCDM: A Cosmon model solution to the cosmological coincidence problem?,”
JCAP 08 (2006) 011, arXiv:gr-qc/0604057.
[2917] A. A. Sen, “Deviation From LambdaCDM: Pressure Parametrization,” Phys. Rev. D 77 (2008) 043508,
arXiv:0708.1072 [astro-ph].
359

[2918] S. Kumar, A. Nautiyal, and A. A. Sen, “Deviation from ΛCDM with cosmic strings networks,” Eur. Phys. J. C 73
(2013) no. 9, 2562, arXiv:1207.4024 [astro-ph.CO].
[2919] R. Y. Wen, L. T. Hergt, N. Afshordi, and D. Scott, “A cosmic glitch in gravity,” JCAP 03 (2024) 045,
arXiv:2311.03028 [astro-ph.CO].
[2920] J. M. Maldacena, “The Large N limit of superconformal field theories and supergravity,” Adv. Theor. Math. Phys. 2
(1998) 231–252, arXiv:hep-th/9711200.
[2921] R. Bousso and J. Polchinski, “Quantization of four form fluxes and dynamical neutralization of the cosmological
constant,” JHEP 06 (2000) 006, arXiv:hep-th/0004134.
[2922] M. Demirtas, M. Kim, L. McAllister, J. Moritz, and A. Rios-Tascon, “Exponentially Small Cosmological Constant in
String Theory,” Phys. Rev. Lett. 128 (2022) no. 1, 011602, arXiv:2107.09065 [hep-th].
[2923] A. A. Sen, S. A. Adil, and S. Sen, “Do cosmological observations allow a negative Λ?,” Mon. Not. Roy. Astron. Soc.
518 (2022) no. 1, 1098–1105, arXiv:2112.10641 [astro-ph.CO].
[2924] C. B. V. Dash, T. G. Sarkar, and A. A. Sen, “Post-reionization H i 21-cm signal: a probe of negative cosmological
constant,” MMNRAS 527 (2023) no. 4, 11694–11706, arXiv:2309.01623 [astro-ph.CO].
[2925] C. Andrei, A. Ijjas, and P. J. Steinhardt, “Rapidly descending dark energy and the end of cosmic expansion,” Proc.
Nat. Acad. Sci. 119 (2022) no. 15, e2200539119, arXiv:2201.07704 [astro-ph.CO].
[2926] S. Vagnozzi, “Seven Hints That Early-Time New Physics Alone Is Not Sufficient to Solve the Hubble Tension,”
Universe 9 (2023) no. 9, 393, arXiv:2308.16628 [astro-ph.CO].
[2927] Y. L. Bolotin, A. Kostenko, O. A. Lemets, and D. A. Yerokhin, “Cosmological Evolution With Interaction Between
Dark Energy And Dark Matter,” Int. J. Mod. Phys. D 24 (2014) no. 03, 1530007, arXiv:1310.0085 [astro-ph.CO].
[2928] B. Wang, E. Abdalla, F. Atrio-Barandela, and D. Pavón, “Further understanding the interaction between dark energy
and dark matter: current status and future directions,” Rept. Prog. Phys. 87 (2024) no. 3, 036901, arXiv:2402.00819
[astro-ph.CO].
[2929] M. A. van der Westhuizen and A. Abebe, “Interacting dark energy: clarifying the cosmological implications and
viability conditions,” JCAP 01 (2024) 048, arXiv:2302.11949 [gr-qc].
[2930] J. S. Alcaniz and J. A. S. Lima, “Interpreting cosmological vacuum decay,” Phys. Rev. D 72 (2005) 063516,
arXiv:astro-ph/0507372.
[2931] D. Pavon and B. Wang, “Le Chatelier-Braun principle in cosmological physics,” Gen. Rel. Grav. 41 (2009) 1–5,
arXiv:0712.0565 [gr-qc].
[2932] V. da Fonseca, T. Barreiro, and N. J. Nunes, “A simple parametrisation for coupled dark energy,” Phys. Dark Univ. 35
(2022) 100940, arXiv:2104.14889 [astro-ph.CO].
[2933] M. Carrillo González and M. Trodden, “Field Theories and Fluids for an Interacting Dark Sector,” Phys. Rev. D 97
(2018) no. 4, 043508, arXiv:1705.04737 [astro-ph.CO]. [Erratum: Phys.Rev.D 101, 089901 (2020)].
[2934] M. Zumalacarregui, T. S. Koivisto, D. F. Mota, and P. Ruiz-Lapuente, “Disformal Scalar Fields and the Dark Sector of
the Universe,” JCAP 05 (2010) 038, arXiv:1004.2684 [astro-ph.CO].
[2935] C. van de Bruck and J. Morrice, “Disformal couplings and the dark sector of the universe,” JCAP 04 (2015) 036,
arXiv:1501.03073 [gr-qc].
[2936] E. M. Teixeira, A. Nunes, and N. J. Nunes, “Disformally Coupled Quintessence,” Phys. Rev. D 101 (2020) no. 8,
083506, arXiv:1912.13348 [gr-qc].
[2937] C. van de Bruck and E. M. Teixeira, “Dark D-Brane Cosmology: from background evolution to cosmological
perturbations,” Phys. Rev. D 102 (2020) no. 10, 103503, arXiv:2007.15414 [gr-qc].
[2938] C. Van De Bruck and J. Mifsud, “Searching for dark matter - dark energy interactions: going beyond the conformal
case,” Phys. Rev. D 97 (2018) no. 2, 023506, arXiv:1709.04882 [astro-ph.CO].
[2939] J. Valiviita, E. Majerotto, and R. Maartens, “Instability in interacting dark energy and dark matter fluids,” JCAP 07
(2008) 020, arXiv:0804.0232 [astro-ph].
[2940] M. B. Gavela, D. Hernandez, L. Lopez Honorez, O. Mena, and S. Rigolin, “Dark coupling,” JCAP 07 (2009) 034,
arXiv:0901.1611 [astro-ph.CO]. [Erratum: JCAP 05, E01 (2010)].
[2941] Y. Zhai, W. Giarè, C. van de Bruck, E. Di Valentino, O. Mena, and R. C. Nunes, “A consistent view of interacting
dark energy from multiple CMB probes,” JCAP 07 (2023) 032, arXiv:2303.08201 [astro-ph.CO].
[2942] W. Giarè, Y. Zhai, S. Pan, E. Di Valentino, R. C. Nunes, and C. van de Bruck, “Tightening the reins on nonminimal
dark sector physics: Interacting dark energy with dynamical and nondynamical equation of state,” Phys. Rev. D 110
(2024) no. 6, 063527, arXiv:2404.02110 [astro-ph.CO].
[2943] A. Paliathanasis, S. Pan, and W. Yang, “Dynamics of nonlinear interacting dark energy models,” Int. J. Mod. Phys. D
28 (2019) no. 12, 1950161, arXiv:1903.02370 [gr-qc].
[2944] J. De-Santiago, I. E. Sánchez G., and D. Tamayo, “Non-linear coupling in the dark sector as a running vacuum model,”
Gen. Rel. Grav. 50 (2018) no. 8, 101, arXiv:1612.02836 [astro-ph.CO].
[2945] F. Arevalo, A. P. R. Bacalhau, and W. Zimdahl, “Cosmological dynamics with non-linear interactions,” Class. Quant.
Grav. 29 (2012) 235001, arXiv:1112.5095 [astro-ph.CO].
[2946] E. Ebrahimi, H. Golchin, A. Mehrabi, and S. M. S. Movahed, “Consistency of nonlinear interacting ghost dark energy
with recent observations,” Int. J. Mod. Phys. D 26 (2017) no. 11, 1750124, arXiv:1611.06551 [astro-ph.CO].
[2947] M. Khurshudyan and A. Khurshudyan, “On cosmology of interacting varying polytropic dark fluids,” Mod. Phys. Lett.
A 34 (2019) no. 17, 1950133, arXiv:1707.04116 [gr-qc].
[2948] G. Cheng, Y.-Z. Ma, F. Wu, J. Zhang, and X. Chen, “Testing interacting dark matter and dark energy model with
cosmological data,” Phys. Rev. D 102 (2020) no. 4, 043517, arXiv:1911.04520 [astro-ph.CO].
360

[2949] W. Yang, S. Pan, and A. Paliathanasis, “Cosmological constraints on an exponential interaction in the dark sector,”
Mon. Not. Roy. Astron. Soc. 482 (2019) no. 1, 1007–1016, arXiv:1804.08558 [gr-qc].
[2950] M. Khurshudyan and A. Khurshudyan, “Some interacting dark energy models,” Symmetry 10 (2018) no. 11, 577,
arXiv:1708.02293 [gr-qc].
[2951] S. Pan, W. Yang, and A. Paliathanasis, “Non-linear interacting cosmological models after Planck 2018 legacy release
and the H0 tension,” Mon. Not. Roy. Astron. Soc. 493 (2020) no. 3, 3114–3131, arXiv:2002.03408 [astro-ph.CO].
[2952] Z. Haba, A. Stachowski, and M. Szydłowski, “Dynamics of the diffusive DM-DE interaction – Dynamical system
approach,” JCAP 07 (2016) 024, arXiv:1603.07620 [gr-qc].
[2953] G. Koutsoumbas, K. Ntrekis, E. Papantonopoulos, and E. N. Saridakis, “Unification of Dark Matter - Dark Energy in
Generalized Galileon Theories,” JCAP 02 (2018) 003, arXiv:1704.08640 [gr-qc].
[2954] S. Calogero and H. Velten, “Cosmology with matter diffusion,” JCAP 11 (2013) 025, arXiv:1308.3393 [astro-ph.CO].
[2955] E. F. Piratova Moreno and L. A. García, “LATE ACCELERATED EXPANSION OF THE UNIVERSE IN
DIFFUSIVE SCENARIOS,” Rev. Mex. Astron. Astrofis. 59 (2023) no. 2, 389–399.
[2956] T. Josset, A. Perez, and D. Sudarsky, “Dark Energy from Violation of Energy Conservation,” Phys. Rev. Lett. 118
(2017) no. 2, 021102, arXiv:1604.04183 [gr-qc].
[2957] A. Perez and D. Sudarsky, “Dark energy from quantum gravity discreteness,” Phys. Rev. Lett. 122 (2019) no. 22,
221302, arXiv:1711.05183 [gr-qc].
[2958] A. Perez and D. Sudarsky, “Black holes, Planckian granularity, and the changing cosmological ‘constant’,” Gen. Rel.
Grav. 53 (2021) no. 4, 40, arXiv:1911.06059 [gr-qc].
[2959] C. Corral, N. Cruz, and E. González, “Diffusion in unimodular gravity: Analytical solutions, late-time acceleration,
and cosmological constraints,” Phys. Rev. D 102 (2020) no. 2, 023508, arXiv:2005.06052 [gr-qc].
[2960] F. X. Linares Cedeño and U. Nucamendi, “Revisiting cosmological diffusion models in Unimodular Gravity and the H0
tension,” Phys. Dark Univ. 32 (2021) 100807, arXiv:2009.10268 [astro-ph.CO].
[2961] S. J. Landau, M. Benetti, A. Perez, and D. Sudarsky, “Cosmological constraints on unimodular gravity models with
diffusion,” Phys. Rev. D 108 (2023) no. 4, 043524, arXiv:2211.07424 [astro-ph.CO].
[2962] D. Wands, J. De-Santiago, and Y. Wang, “Inhomogeneous vacuum energy,” Class. Quant. Grav. 29 (2012) 145017,
arXiv:1203.6776 [astro-ph.CO].
[2963] M. Sebastianutti, N. B. Hogg, and M. Bruni, “The interacting vacuum and tensions: A comparison of theoretical
models,” Phys. Dark Univ. 46 (2024) 101546, arXiv:2312.14123 [astro-ph.CO].
[2964] S. Weinberg, “The Cosmological Constant Problem,” Rev. Mod. Phys. 61 (1989) 1–23.
[2965] G. F. R. Ellis, H. van Elst, J. Murugan, and J.-P. Uzan, “On the Trace-Free Einstein Equations as a Viable Alternative
to General Relativity,” Class. Quant. Grav. 28 (2011) 225007, arXiv:1008.1196 [gr-qc].
[2966] M. de Cesare and E. Wilson-Ewing, “Interacting dark sector from the trace-free Einstein equations: Cosmological
perturbations with no instability,” Phys. Rev. D 106 (2022) no. 2, 023527, arXiv:2112.12701 [gr-qc].
[2967] A. Shafieloo, D. K. Hazra, V. Sahni, and A. A. Starobinsky, “Metastable Dark Energy with Radioactive-like Decay,”
Mon. Not. Roy. Astron. Soc. 473 (2018) no. 2, 2760–2770, arXiv:1610.05192 [astro-ph.CO].
[2968] J. S. T. de Souza, G. S. Vicente, and L. L. Graef, “Constraints on Metastable Dark Energy Decaying into Dark
Matter,” Universe 10 (2024) no. 9, 371, arXiv:2403.04970 [astro-ph.CO].
[2969] X. Li, A. Shafieloo, V. Sahni, and A. A. Starobinsky, “Revisiting Metastable Dark Energy and Tensions in the
Estimation of Cosmological Parameters,” Astrophys. J. 887 (2019) 153, arXiv:1904.03790 [astro-ph.CO].
[2970] W. Yang, E. Di Valentino, S. Pan, S. Basilakos, and A. Paliathanasis, “Metastable dark energy models in light of
P lanck 2018 data: Alleviating the H0 tension,” Phys. Rev. D 102 (2020) no. 6, 063503, arXiv:2001.04307
[astro-ph.CO].
[2971] K. Urbanowski, “Cosmological “constant” in a universe born in the metastable false vacuum state,” Eur. Phys. J. C 82
(2022) no. 3, 242, arXiv:2110.11957 [gr-qc].
[2972] K. Urbanowski, “A universe born in a metastable false vacuum state needs not die,” Eur. Phys. J. C 83 (2023) no. 1,
55, arXiv:2207.10965 [gr-qc].
[2973] E. Abdalla, L. L. Graef, and B. Wang, “A Model for Dark Energy decay,” Phys. Lett. B 726 (2013) 786–790,
arXiv:1202.0499 [gr-qc].
[2974] R. G. Landim and E. Abdalla, “Metastable dark energy,” Phys. Lett. B 764 (2017) 271–276, arXiv:1611.00428
[hep-ph].
[2975] R. G. Landim, R. J. F. Marcondes, F. F. Bernardi, and E. Abdalla, “Interacting Dark Energy in the Dark SU (2)R
Model,” Braz. J. Phys. 48 (2018) no. 4, 364–369, arXiv:1711.07282 [astro-ph.CO].
[2976] D. Stojkovic, G. D. Starkman, and R. Matsuo, “Dark energy, the colored anti-de Sitter vacuum, and LHC
phenomenology,” Phys. Rev. D 77 (2008) 063006, arXiv:hep-ph/0703246.
[2977] E. Greenwood, E. Halstead, R. Poltis, and D. Stojkovic, “Dark energy, the electroweak vacua and collider
phenomenology,” Phys. Rev. D 79 (2009) 103003, arXiv:0810.5343 [hep-ph].
[2978] F. Simpson, “Scattering of dark matter and dark energy,” Phys. Rev. D 82 (2010) 083505, arXiv:1007.1034
[astro-ph.CO].
[2979] A. Pourtsidou, C. Skordis, and E. J. Copeland, “Models of dark matter coupled to dark energy,” Phys. Rev. D 88
(2013) no. 8, 083505, arXiv:1307.0458 [astro-ph.CO].
[2980] M. Asghari, J. Beltrán Jiménez, S. Khosravi, and D. F. Mota, “On structure formation from a small-scales-interacting
dark sector,” JCAP 04 (2019) 042, arXiv:1902.05532 [astro-ph.CO].
[2981] D. Figueruelo et al., “J-PAS: Forecasts for dark matter - dark energy elastic couplings,” JCAP 07 (2021) 022,
361

arXiv:2103.01571 [astro-ph.CO].
[2982] J. Beltrán Jiménez, D. Bettoni, D. Figueruelo, F. A. Teppa Pannia, and S. Tsujikawa, “Probing elastic interactions in
the dark sector and the role of S8,” Phys. Rev. D 104 (2021) no. 10, 103503, arXiv:2106.11222 [astro-ph.CO].
[2983] V. Poulin, J. L. Bernal, E. D. Kovetz, and M. Kamionkowski, “Sigma-8 tension is a drag,” Phys. Rev. D 107 (2023)
no. 12, 123538, arXiv:2209.06217 [astro-ph.CO].
[2984] J. B. Jiménez, D. Bettoni, D. Figueruelo, and F. A. Teppa Pannia, “On evidence for elastic interactions in the dark
sector,” Phys. Dark Univ. 47 (2025) 101761, arXiv:2410.18645 [astro-ph.CO].
[2985] J. Beltrán Jiménez, E. Di Dio, and D. Figueruelo, “A smoking gun from the power spectrum dipole for elastic
interactions in the dark sector,” JCAP 11 (2023) 088, arXiv:2212.08617 [astro-ph.CO].
[2986] J. Beltrán Jiménez, D. Figueruelo, and F. A. Teppa Pannia, “Nondegeneracy of massive neutrinos and elastic
interactions in the dark sector,” Phys. Rev. D 110 (2024) no. 2, 023527, arXiv:2403.03216 [astro-ph.CO].
[2987] M. Baldi and F. Simpson, “Structure formation simulations with momentum exchange: alleviating tensions between
high-redshift and low-redshift cosmological probes,” Mon. Not. Roy. Astron. Soc. 465 (2017) no. 1, 653–666,
arXiv:1605.05623 [astro-ph.CO].
[2988] A. Pourtsidou and T. Tram, “Reconciling CMB and structure growth measurements with dark energy interactions,”
Phys. Rev. D 94 (2016) no. 4, 043518, arXiv:1604.04222 [astro-ph.CO].
[2989] P. Agrawal, G. Obied, P. J. Steinhardt, and C. Vafa, “On the Cosmological Implications of the String Swampland,”
Phys. Lett. B 784 (2018) 271–276, arXiv:1806.09718 [hep-th].
[2990] W. H. Kinney, S. Vagnozzi, and L. Visinelli, “The zoo plot meets the swampland: mutual (in)consistency of single-field
inflation, string conjectures, and cosmological data,” Class. Quant. Grav. 36 (2019) no. 11, 117001, arXiv:1808.06424
[astro-ph.CO].
[2991] P. Agrawal, G. Obied, and C. Vafa, “H0 tension, swampland conjectures, and the epoch of fading dark matter,” Phys.
Rev. D 103 (2021) no. 4, 043523, arXiv:1906.08261 [astro-ph.CO].
[2992] A. Salam and E. Sezgin, “Chiral Compactification on Minkowski x S**2 of N=2 Einstein-Maxwell Supergravity in
Six-Dimensions,” Phys. Lett. B 147 (1984) 47.
[2993] M. Cvetic, G. W. Gibbons, and C. N. Pope, “A String and M theory origin for the Salam-Sezgin model,” Nucl. Phys. B
677 (2004) 164–180, arXiv:hep-th/0308026.
[2994] L. A. Anchordoqui, I. Antoniadis, D. Lüst, J. F. Soriano, and T. R. Taylor, “H0 tension and the String Swampland,”
Phys. Rev. D 101 (2020) 083532, arXiv:1912.00242 [hep-th].
[2995] L. A. Anchordoqui, I. Antoniadis, D. Lüst, and J. F. Soriano, “Dark energy, Ricci-nonflat spaces, and the Swampland,”
Phys. Lett. B 816 (2021) 136199, arXiv:2005.10075 [hep-th].
[2996] D. Benisty, S. Pan, D. Staicova, E. Di Valentino, and R. C. Nunes, “Late-time constraints on interacting dark energy:
Analysis independent of H0, rd, and MB,” Astron. Astrophys. 688 (2024) A156, arXiv:2403.00056 [astro-ph.CO].
[2997] G. A. Hoerning, R. G. Landim, L. O. Ponte, R. P. Rolim, F. B. Abdalla, and E. Abdalla, “Constraints on interacting
dark energy revisited: implications for the Hubble tension,” arXiv:2308.05807 [astro-ph.CO].
[2998] S. Pan and W. Yang, “On the interacting dark energy scenarios − the case for Hubble constant tension,”
arXiv:2310.07260 [astro-ph.CO].
[2999] W. Yang, S. Pan, E. Di Valentino, R. C. Nunes, S. Vagnozzi, and D. F. Mota, “Tale of stable interacting dark energy,
observational signatures, and the H0 tension,” JCAP 09 (2018) 019, arXiv:1805.08252 [astro-ph.CO].
[3000] A. Bhattacharyya, U. Alam, K. L. Pandey, S. Das, and S. Pal, “Are H0 and σ8 tensions generic to present cosmological
data?,” Astrophys. J. 876 (2019) no. 2, 143, arXiv:1805.04716 [astro-ph.CO].
[3001] E. Di Valentino, R. Z. Ferreira, L. Visinelli, and U. Danielsson, “Late time transitions in the quintessence field and the
H0 tension,” Phys. Dark Univ. 26 (2019) 100385, arXiv:1906.11255 [astro-ph.CO].
[3002] Y.-H. Yao and X.-H. Meng, “Can interacting dark energy with dynamical coupling resolve the Hubble tension,”
arXiv:2207.05955 [astro-ph.CO].
[3003] S. Gariazzo, E. Di Valentino, O. Mena, and R. C. Nunes, “Late-time interacting cosmologies and the Hubble constant
tension,” Phys. Rev. D 106 (2022) no. 2, 023530, arXiv:2111.03152 [astro-ph.CO].
[3004] R.-Y. Guo, L. Feng, T.-Y. Yao, and X.-Y. Chen, “Exploration of interacting dynamical dark energy model with
interaction term including the equation-of-state parameter: alleviation of the H0 tension,” JCAP 12 (2021) no. 12, 036,
arXiv:2110.02536 [gr-qc].
[3005] R. C. Nunes and E. Di Valentino, “Dark sector interaction and the supernova absolute magnitude tension,” Phys. Rev.
D 104 (2021) no. 6, 063529, arXiv:2107.09151 [astro-ph.CO].
[3006] Y. Zhao, Y. Liu, S. Liao, J. Zhang, X. Liu, and W. Du, “Constraining interacting dark energy models with the halo
concentration–mass relation,” Mon. Not. Roy. Astron. Soc. 523 (2023) no. 4, 5962–5971, arXiv:2212.02050
[astro-ph.CO].
[3007] L.-Y. Gao, Z.-W. Zhao, S.-S. Xue, and X. Zhang, “Relieving the H 0 tension with a new interacting dark energy
model,” JCAP 07 (2021) 005, arXiv:2101.10714 [astro-ph.CO].
[3008] H. Amirhashchi, A. K. Yadav, N. Ahmad, and V. Yadav, “Interacting dark sectors in anisotropic universe:
Observational constraints and H0 tension,” Phys. Dark Univ. 36 (2022) 101043, arXiv:2001.03775 [astro-ph.CO].
[3009] S. Pan, W. Yang, C. Singha, and E. N. Saridakis, “Observational constraints on sign-changeable interaction models
and alleviation of the H0 tension,” Phys. Rev. D 100 (2019) no. 8, 083539, arXiv:1903.10969 [astro-ph.CO].
[3010] L.-Y. Gao, S.-S. Xue, and X. Zhang, “Dark energy and matter interacting scenario to relieve H 0 and S 8 tensions*,”
Chin. Phys. C 48 (2024) no. 5, 051001, arXiv:2212.13146 [astro-ph.CO].
[3011] W. Yang, S. Pan, E. Di Valentino, O. Mena, and A. Melchiorri, “2021-H0 odyssey: closed, phantom and interacting
362

dark energy cosmologies,” JCAP 10 (2021) 008, arXiv:2101.03129 [astro-ph.CO].


[3012] W. Yang, S. Pan, R. C. Nunes, and D. F. Mota, “Dark calling Dark: Interaction in the dark sector in presence of
neutrino properties after Planck CMB final release,” JCAP 04 (2020) 008, arXiv:1910.08821 [astro-ph.CO].
[3013] M. Lucca, “Dark energy–dark matter interactions as a solution to the S8 tension,” Phys. Dark Univ. 34 (2021) 100899,
arXiv:2105.09249 [astro-ph.CO].
[3014] R. An, C. Feng, and B. Wang, “Relieving the Tension between Weak Lensing and Cosmic Microwave Background with
Interacting Dark Matter and Dark Energy Models,” JCAP 02 (2018) 038, arXiv:1711.06799 [astro-ph.CO].
[3015] S. Sinha and N. Banerjee, “Density perturbation in an interacting holographic dark energy model,” Eur. Phys. J. Plus
135 (2020) no. 10, 779, arXiv:1911.06520 [gr-qc].
[3016] S. Sinha, “Differentiating dark interactions with perturbation,” Phys. Rev. D 103 (2021) no. 12, 123547,
arXiv:2101.08959 [astro-ph.CO].
[3017] S. Sinha, M. Banerjee, and S. Das, “Perturbation in an interacting dark Universe,” Phys. Dark Univ. 42 (2023) 101273,
arXiv:2204.05174 [gr-qc].
[3018] B. J. Barros, D. Castelão, V. da Fonseca, T. Barreiro, N. J. Nunes, and I. Tereno, “Is there evidence for CIDER in the
Universe?,” JCAP 01 (2023) 013, arXiv:2209.04468 [astro-ph.CO].
[3019] O. Trivedi, “Recent Advances in Cosmological Singularities,” Symmetry 16 (2024) no. 3, 298, arXiv:2309.08954
[gr-qc].
[3020] J. de Haro, S. Nojiri, S. D. Odintsov, V. K. Oikonomou, and S. Pan, “Finite-time cosmological singularities and the
possible fate of the Universe,” Phys. Rept. 1034 (2023) 1–114, arXiv:2309.07465 [gr-qc].
[3021] S. Nojiri, S. D. Odintsov, and S. Tsujikawa, “Properties of singularities in (phantom) dark energy universe,” Phys. Rev.
D 71 (2005) 063004, arXiv:hep-th/0501025.
[3022] S. Nojiri, S. D. Odintsov, and V. K. Oikonomou, “Modified Gravity Theories on a Nutshell: Inflation, Bounce and
Late-time Evolution,” Phys. Rept. 692 (2017) 1–104, arXiv:1705.11098 [gr-qc].
[3023] S. Nojiri and S. D. Odintsov, “Inhomogeneous equation of state of the universe: Phantom era, future singularity and
crossing the phantom barrier,” Phys. Rev. D 72 (2005) 023003, arXiv:hep-th/0505215.
[3024] K. Bamba, S. Nojiri, and S. D. Odintsov, “The Universe future in modified gravity theories: Approaching the
finite-time future singularity,” JCAP 10 (2008) 045, arXiv:0807.2575 [hep-th].
[3025] S. Nojiri and S. D. Odintsov, “Quantum escape of sudden future singularity,” Phys. Lett. B 595 (2004) 1–8,
arXiv:hep-th/0405078.
[3026] S. Castello, N. Grimm, and C. Bonvin, “Rescuing constraints on modified gravity using gravitational redshift in
large-scale structure,” Phys. Rev. D 106 (2022) no. 8, 083511, arXiv:2204.11507 [astro-ph.CO].
[3027] S. Castello, M. Mancarella, N. Grimm, D. Sobral-Blanco, I. Tutusaus, and C. Bonvin, “Gravitational redshift
constraints on the effective theory of interacting dark energy,” JCAP 05 (2024) 003, arXiv:2311.14425
[astro-ph.CO].
[3028] C. Bonvin and L. E. Pogosian, “Can cosmology distinguish a dark force from a modification of gravity?,” Nature
Astron. 7 (2023) no. 9, 1023–1024.
[3029] S. Castello, Z. Wang, L. Dam, C. Bonvin, and L. Pogosian, “Disentangling modified gravity from a dark force with
gravitational redshift,” Phys. Rev. D 110 (2024) no. 10, 103523, arXiv:2404.09379 [astro-ph.CO].
[3030] D. Sobral-Blanco and C. Bonvin, “Measuring anisotropic stress with relativistic effects,” Phys. Rev. D 104 (2021)
no. 6, 063516, arXiv:2102.05086 [astro-ph.CO].
[3031] D. Sobral-Blanco and C. Bonvin, “Measuring the distortion of time with relativistic effects in large-scale structure,”
Mon. Not. Roy. Astron. Soc. 519 (2022) no. 1, L39–L44, arXiv:2205.02567 [astro-ph.CO].
[3032] Y.-H. Li, J.-F. Zhang, and X. Zhang, “Testing models of vacuum energy interacting with cold dark matter,” Phys. Rev.
D 93 (2016) no. 2, 023002, arXiv:1506.06349 [astro-ph.CO].
[3033] L. Feng, H.-L. Li, J.-F. Zhang, and X. Zhang, “Exploring neutrino mass and mass hierarchy in interacting dark energy
models,” Sci. China Phys. Mech. Astron. 63 (2020) no. 2, 220401, arXiv:1903.08848 [astro-ph.CO].
[3034] Y.-H. Li and X. Zhang, “Large-scale stable interacting dark energy model: Cosmological perturbations and
observational constraints,” Phys. Rev. D 89 (2014) no. 8, 083009, arXiv:1312.6328 [astro-ph.CO].
[3035] L. Feng and X. Zhang, “Revisit of the interacting holographic dark energy model after Planck 2015,” JCAP 08 (2016)
072, arXiv:1607.05567 [astro-ph.CO].
[3036] M.-M. Zhao, D.-Z. He, J.-F. Zhang, and X. Zhang, “Search for sterile neutrinos in holographic dark energy cosmology:
Reconciling Planck observation with the local measurement of the Hubble constant,” Phys. Rev. D 96 (2017) no. 4,
043520, arXiv:1703.08456 [astro-ph.CO].
[3037] Y.-H. Li, J.-F. Zhang, and X. Zhang, “Exploring the full parameter space for an interacting dark energy model with
recent observations including redshift-space distortions: Application of the parametrized post-Friedmann approach,”
Phys. Rev. D 90 (2014) no. 12, 123007, arXiv:1409.7205 [astro-ph.CO].
[3038] Y.-H. Li, J.-F. Zhang, and X. Zhang, “Parametrized Post-Friedmann Framework for Interacting Dark Energy,” Phys.
Rev. D 90 (2014) no. 6, 063005, arXiv:1404.5220 [astro-ph.CO].
[3039] R.-Y. Guo, Y.-H. Li, J.-F. Zhang, and X. Zhang, “Weighing neutrinos in the scenario of vacuum energy interacting
with cold dark matter: application of the parameterized post-Friedmann approach,” JCAP 05 (2017) 040,
arXiv:1702.04189 [astro-ph.CO].
[3040] X. Zhang, “Probing the interaction between dark energy and dark matter with the parametrized post-Friedmann
approach,” Sci. China Phys. Mech. Astron. 60 (2017) no. 5, 050431, arXiv:1702.04564 [astro-ph.CO].
[3041] L. Feng, Y.-H. Li, F. Yu, J.-F. Zhang, and X. Zhang, “Exploring interacting holographic dark energy in a perturbed
363

universe with parameterized post-Friedmann approach,” Eur. Phys. J. C 78 (2018) no. 10, 865, arXiv:1807.03022
[astro-ph.CO].
[3042] Y.-H. Li and X. Zhang, “IDECAMB: an implementation of interacting dark energy cosmology in CAMB,” JCAP 09
(2023) 046, arXiv:2306.01593 [astro-ph.CO].
[3043] J.-J. Guo, J.-F. Zhang, Y.-H. Li, D.-Z. He, and X. Zhang, “Probing the sign-changeable interaction between dark
energy and dark matter with current observations,” Sci. China Phys. Mech. Astron. 61 (2018) no. 3, 030011,
arXiv:1710.03068 [astro-ph.CO].
[3044] H.-L. Li, J.-F. Zhang, L. Feng, and X. Zhang, “Reexploration of interacting holographic dark energy model: Cases of
interaction term excluding the Hubble parameter,” Eur. Phys. J. C 77 (2017) no. 12, 907, arXiv:1711.06159
[astro-ph.CO].
[3045] L. Feng, D.-Z. He, H.-L. Li, J.-F. Zhang, and X. Zhang, “Constraints on active and sterile neutrinos in an interacting
dark energy cosmology,” Sci. China Phys. Mech. Astron. 63 (2020) no. 9, 290404, arXiv:1910.03872 [astro-ph.CO].
[3046] M. Zhang, B. Wang, P.-J. Wu, J.-Z. Qi, Y. Xu, J.-F. Zhang, and X. Zhang, “Prospects for Constraining Interacting
Dark Energy Models with 21 cm Intensity Mapping Experiments,” Astrophys. J. 918 (2021) no. 2, 56,
arXiv:2102.03979 [astro-ph.CO].
[3047] J. Gleyzes, D. Langlois, M. Mancarella, and F. Vernizzi, “Effective Theory of Interacting Dark Energy,” JCAP 08
(2015) 054, arXiv:1504.05481 [astro-ph.CO].
[3048] J. Gleyzes, D. Langlois, M. Mancarella, and F. Vernizzi, “Effective Theory of Dark Energy at Redshift Survey Scales,”
JCAP 02 (2016) 056, arXiv:1509.02191 [astro-ph.CO].
[3049] C. M. Will, “The Confrontation between General Relativity and Experiment,” Living Rev. Rel. 17 (2014) 4,
arXiv:1403.7377 [gr-qc].
[3050] A. Addazi et al., “Quantum gravity phenomenology at the dawn of the multi-messenger era—A review,” Prog. Part.
Nucl. Phys. 125 (2022) 103948, arXiv:2111.05659 [hep-ph].
[3051] R. Alves Batista et al., “White paper and roadmap for quantum gravity phenomenology in the multi-messenger era,”
Class. Quant. Grav. 42 (2025) no. 3, 032001, arXiv:2312.00409 [gr-qc].
[3052] M. H. Goroff and A. Sagnotti, “The Ultraviolet Behavior of Einstein Gravity,” Nucl. Phys. B 266 (1986) 709–736.
[3053] L. Barack et al., “Black holes, gravitational waves and fundamental physics: a roadmap,” Class. Quant. Grav. 36
(2019) no. 14, 143001, arXiv:1806.05195 [gr-qc].
[3054] S. M. Carroll, “The Cosmological constant,” Living Rev. Rel. 4 (2001) 1, arXiv:astro-ph/0004075.
[3055] Supernova Search Team Collaboration, A. G. Riess et al., “Observational evidence from supernovae for an
accelerating universe and a cosmological constant,” Astron. J. 116 (1998) 1009–1038, arXiv:astro-ph/9805201.
[3056] Supernova Cosmology Project Collaboration, S. Perlmutter et al., “Measurements of Ω and Λ from 42 High
Redshift Supernovae,” Astrophys. J. 517 (1999) 565–586, arXiv:astro-ph/9812133.
[3057] M. Davis, G. Efstathiou, C. S. Frenk, and S. D. M. White, “The Evolution of Large Scale Structure in a Universe
Dominated by Cold Dark Matter,” Astrophys. J. 292 (1985) 371–394.
[3058] G. Bertone and D. Hooper, “History of dark matter,” Rev. Mod. Phys. 90 (2018) no. 4, 045002, arXiv:1605.04909
[astro-ph.CO].
[3059] E. Aprile and T. Doke, “Liquid Xenon Detectors for Particle Physics and Astrophysics,” Rev. Mod. Phys. 82 (2010)
2053–2097, arXiv:0910.4956 [physics.ins-det].
[3060] M. Misiaszek and N. Rossi, “Direct Detection of Dark Matter: A Critical Review,” Symmetry 16 (2024) no. 2, 201,
arXiv:2310.20472 [hep-ph].
[3061] F. Bajardi and S. Capozziello, Noether Symmetries in Theories of Gravity. Cambridge Monographs on Mathematical
Physics. Cambridge University Press, 2022.
[3062] CANTATA Collaboration, E. N. Saridakis, R. Lazkoz, V. Salzano, P. Vargas Moniz, S. Capozziello,
J. Beltrán Jiménez, M. De Laurentis, and G. J. Olmo, eds., Modified Gravity and Cosmology. An Update by the
CANTATA Network. Springer, 2021. arXiv:2105.12582 [gr-qc].
[3063] K. S. Stelle, “Renormalization of Higher Derivative Quantum Gravity,” Phys. Rev. D 16 (1977) 953–969.
[3064] F. Bajardi, S. Capozziello, and D. Vernieri, “Non-local curvature and Gauss–Bonnet cosmologies by Noether
symmetries,” Eur. Phys. J. Plus 135 (2020) no. 12, 942, arXiv:2011.01317 [gr-qc].
[3065] F. Bajardi, D. Vernieri, and S. Capozziello, “Exact solutions in higher-dimensional Lovelock and AdS5 Chern-Simons
gravity,” JCAP 11 (2021) no. 11, 057, arXiv:2106.07396 [gr-qc].
[3066] J. J. Halliwell, “Scalar Fields in Cosmology with an Exponential Potential,” Phys. Lett. B 185 (1987) 341.
[3067] J.-P. Uzan, “Cosmological scaling solutions of nonminimally coupled scalar fields,” Phys. Rev. D 59 (1999) 123510,
arXiv:gr-qc/9903004.
[3068] Z. Urban, F. Bajardi, and S. Capozziello, “The Noether–Bessel-Hagen symmetry approach for dynamical systems,” Int.
J. Geom. Meth. Mod. Phys. 17 (2020) no. 14, 2050215, arXiv:2003.13756 [gr-qc].
[3069] M. Krssak, R. J. van den Hoogen, J. G. Pereira, C. G. Böhmer, and A. A. Coley, “Teleparallel theories of gravity:
illuminating a fully invariant approach,” Class. Quant. Grav. 36 (2019) no. 18, 183001, arXiv:1810.12932 [gr-qc].
[3070] S. Bahamonde, K. F. Dialektopoulos, C. Escamilla-Rivera, G. Farrugia, V. Gakis, M. Hendry, M. Hohmann,
J. Levi Said, J. Mifsud, and E. Di Valentino, “Teleparallel gravity: from theory to cosmology,” Rept. Prog. Phys. 86
(2023) no. 2, 026901, arXiv:2106.13793 [gr-qc].
[3071] A. Jiménez Cano, Metric-affine Gauge theories of gravity. Foundations and new insights. PhD thesis, Granada U.,
Theor. Phys. Astrophys., 2021. arXiv:2201.12847 [gr-qc].
[3072] I. Ayuso, R. Lazkoz, and V. Salzano, “Observational constraints on cosmological solutions of f (Q) theories,” Phys. Rev.
364

D 103 (2021) no. 6, 063505, arXiv:2012.00046 [astro-ph.CO].


[3073] M. Hohmann, C. Pfeifer, and N. Voicu, “Mathematical foundations for field theories on Finsler spacetimes,” J. Math.
Phys. 63 (2022) no. 3, 032503, arXiv:2106.14965 [math-ph].
[3074] C. Pfeifer, “Finsler spacetime geometry in Physics,” Int. J. Geom. Meth. Mod. Phys. 16 (2019) no. supp02, 1941004,
arXiv:1903.10185 [gr-qc].
[3075] J. D. Barrow and S. Cotsakis, “Inflation and the Conformal Structure of Higher Order Gravity Theories,” Phys. Lett. B
214 (1988) 515–518.
[3076] T. P. Sotiriou and V. Faraoni, “f(R) Theories Of Gravity,” Rev. Mod. Phys. 82 (2010) 451–497, arXiv:0805.1726
[gr-qc].
[3077] A. De Felice and S. Tsujikawa, “f(R) theories,” Living Rev. Rel. 13 (2010) 3, arXiv:1002.4928 [gr-qc].
[3078] S. Capozziello and M. De Laurentis, “Extended Theories of Gravity,” Phys. Rept. 509 (2011) 167–321,
arXiv:1108.6266 [gr-qc].
[3079] A. A. Starobinsky, “Disappearing cosmological constant in f(R) gravity,” JETP Lett. 86 (2007) 157–163,
arXiv:0706.2041 [astro-ph].
[3080] A. Paliathanasis, M. Tsamparlis, and S. Basilakos, “Constraints and analytical solutions of f (R) theories of gravity
using Noether symmetries,” Phys. Rev. D 84 (2011) 123514, arXiv:1111.4547 [astro-ph.CO].
[3081] S. Nojiri and S. D. Odintsov, “Modified f(R) gravity consistent with realistic cosmology: From matter dominated epoch
to dark energy universe,” Phys. Rev. D 74 (2006) 086005, arXiv:hep-th/0608008.
[3082] G. Papagiannopoulos, S. Basilakos, J. D. Barrow, and A. Paliathanasis, “New integrable models and analytical
solutions in f (R) cosmology with an ideal gas,” Phys. Rev. D 97 (2018) no. 2, 024026, arXiv:1801.01274 [gr-qc].
[3083] F. Bajardi, R. D’Agostino, M. Benetti, V. De Falco, and S. Capozziello, “Early and late time cosmology: the f(R)
gravity perspective,” Eur. Phys. J. Plus 137 (2022) no. 11, 1239, arXiv:2211.06268 [gr-qc].
[3084] S. Nojiri and S. D. Odintsov, “Unified cosmic history in modified gravity: from F(R) theory to Lorentz non-invariant
models,” Phys. Rept. 505 (2011) 59–144, arXiv:1011.0544 [gr-qc].
[3085] S. Capozziello et al., “Constraining Theories of Gravity by GINGER experiment,” Eur. Phys. J. Plus 136 (2021) no. 4,
394, arXiv:2103.15135 [gr-qc]. [Erratum: Eur.Phys.J.Plus 136, 563 (2021)].
[3086] T. Schiavone, G. Montani, and F. Bombacigno, “f(R) gravity in the Jordan frame as a paradigm for the Hubble
tension,” Mon. Not. Roy. Astron. Soc. 522 (2023) no. 1, L72–L77, arXiv:2211.16737 [gr-qc].
[3087] G. Montani, M. De Angelis, F. Bombacigno, and N. Carlevaro, “Metric f(R) gravity with dynamical dark energy as a
scenario for the Hubble tension,” Mon. Not. Roy. Astron. Soc. 527 (2023) no. 1, L156–L161, arXiv:2306.11101
[gr-qc].
[3088] G. Montani, N. Carlevaro, and M. De Angelis, “Modified Gravity in the Presence of Matter Creation: Scenario for the
Late Universe,” Entropy 26 (2024) no. 8, 662, arXiv:2407.12409 [gr-qc].
[3089] G. Montani, N. Carlevaro, and M. G. Dainotti, “Running Hubble constant: Evolutionary Dark Energy,” Phys. Dark
Univ. 48 (2025) 101847, arXiv:2411.07060 [gr-qc].
[3090] G. Montani, N. Carlevaro, and M. G. Dainotti, “Slow-rolling scalar dynamics as solution for the Hubble tension,” Phys.
Dark Univ. 44 (2024) 101486, arXiv:2311.04822 [gr-qc].
[3091] G. Montani, N. Carlevaro, L. A. Escamilla, and E. Di Valentino, “Kinetic model for dark energy—dark matter
interaction: Scenario for the hubble tension,” Phys. Dark Univ. 48 (2025) 101848, arXiv:2404.15977 [gr-qc].
[3092] M. G. Dainotti, B. De Simone, T. Schiavone, G. Montani, E. Rinaldi, and G. Lambiase, “On the Hubble constant
tension in the SNe Ia Pantheon sample,” Astrophys. J. 912 (2021) no. 2, 150, arXiv:2103.02117 [astro-ph.CO].
[3093] S. Nojiri, S. D. Odintsov, and V. K. Oikonomou, “Integral F(R) gravity and saddle point condition as a remedy for the
H0-tension,” Nucl. Phys. B 980 (2022) 115850, arXiv:2205.11681 [gr-qc].
[3094] I. Ayuso and J. A. R. Cembranos, “Nonminimal scalar-tensor theories,” Phys. Rev. D 101 (2020) no. 4, 044007,
arXiv:1411.1653 [gr-qc].
[3095] J. D. Barrow, “Slow roll inflation in scalar - tensor theories,” Phys. Rev. D 51 (1995) 2729–2732.
[3096] Y. Shtanov, J. H. Traschen, and R. H. Brandenberger, “Universe reheating after inflation,” Phys. Rev. D 51 (1995)
5438–5455, arXiv:hep-ph/9407247.
[3097] R. Allahverdi, R. Brandenberger, F.-Y. Cyr-Racine, and A. Mazumdar, “Reheating in Inflationary Cosmology: Theory
and Applications,” Ann. Rev. Nucl. Part. Sci. 60 (2010) 27–51, arXiv:1001.2600 [hep-th].
[3098] P. B. Greene, “Inflationary reheating and fermions,” AIP Conf. Proc. 478 (1999) no. 1, 72–74, arXiv:hep-ph/9905256.
[3099] T. Kobayashi, “Horndeski theory and beyond: a review,” Rept. Prog. Phys. 82 (2019) no. 8, 086901, arXiv:1901.07183
[gr-qc].
[3100] C. Deffayet, X. Gao, D. A. Steer, and G. Zahariade, “From k-essence to generalised Galileons,” Phys. Rev. D 84 (2011)
064039, arXiv:1103.3260 [hep-th].
[3101] R. Kase and S. Tsujikawa, “Dark energy in Horndeski theories after GW170817: A review,” Int. J. Mod. Phys. D 28
(2019) no. 05, 1942005, arXiv:1809.08735 [gr-qc].
[3102] R. D’Agostino and R. C. Nunes, “Probing observational bounds on scalar-tensor theories from standard sirens,” Phys.
Rev. D 100 (2019) no. 4, 044041, arXiv:1907.05516 [gr-qc].
[3103] J. M. Ezquiaga and M. Zumalacárregui, “Dark Energy After GW170817: Dead Ends and the Road Ahead,” Phys. Rev.
Lett. 119 (2017) no. 25, 251304, arXiv:1710.05901 [astro-ph.CO].
[3104] A. Bonilla, R. D’Agostino, R. C. Nunes, and J. C. N. de Araujo, “Forecasts on the speed of gravitational waves at high
z,” JCAP 03 (2020) 015, arXiv:1910.05631 [gr-qc].
[3105] C. de Rham and S. Melville, “Gravitational Rainbows: LIGO and Dark Energy at its Cutoff,” Phys. Rev. Lett. 121
365

(2018) no. 22, 221101, arXiv:1806.09417 [hep-th].


[3106] M. Ballardini, M. Braglia, F. Finelli, D. Paoletti, A. A. Starobinsky, and C. Umiltà, “Scalar-tensor theories of gravity,
neutrino physics, and the H0 tension,” JCAP 10 (2020) 044, arXiv:2004.14349 [astro-ph.CO].
[3107] M. Petronikolou, S. Basilakos, and E. N. Saridakis, “Alleviating H0 tension in Horndeski gravity,” Phys. Rev. D 106
(2022) no. 12, 124051, arXiv:2110.01338 [gr-qc].
[3108] M. Petronikolou and E. N. Saridakis, “Alleviating the H0 Tension in Scalar–Tensor and Bi-Scalar–Tensor Theories,”
Universe 9 (2023) no. 9, 397, arXiv:2308.16044 [gr-qc].
[3109] S. Peirone, G. Benevento, N. Frusciante, and S. Tsujikawa, “Cosmological data favor Galileon ghost condensate over
ΛCDM,” Phys. Rev. D 100 (2019) no. 6, 063540, arXiv:1905.05166 [astro-ph.CO].
[3110] N. Frusciante, S. Peirone, L. Atayde, and A. De Felice, “Phenomenology of the generalized cubic covariant Galileon
model and cosmological bounds,” Phys. Rev. D 101 (2020) no. 6, 064001, arXiv:1912.07586 [astro-ph.CO].
[3111] S. Banerjee, M. Petronikolou, and E. N. Saridakis, “Alleviating the H0 tension with new gravitational scalar tensor
theories,” Phys. Rev. D 108 (2023) no. 2, 024012, arXiv:2209.02426 [gr-qc].
[3112] C. Brans and R. H. Dicke, “Mach’s principle and a relativistic theory of gravitation,” Phys. Rev. 124 (1961) 925–935.
[3113] R. H. Dicke, “Mach’s principle and invariance under transformation of units,” Phys. Rev. 125 (1962) 2163–2167.
[3114] I. Ayuso, J. P. Mimoso, and N. J. Nunes, “What if Newton’s Gravitational Constant Was Negative?,” Galaxies 7
(2019) no. 1, 38, arXiv:1903.07604 [gr-qc].
[3115] O. Akarsu, N. Katırcı, N. Özdemir, and J. A. Vázquez, “Anisotropic massive Brans-Dicke gravity extension of the
standard ΛCDM model,” Eur. Phys. J. C 80 (2020) no. 1, 32, arXiv:1903.06679 [gr-qc].
[3116] A. Gómez-Valent and P. Hassan Puttasiddappa, “Difficulties in reconciling non-negligible differences between the local
and cosmological values of the gravitational coupling in extended Brans-Dicke theories,” JCAP 09 (2021) 040,
arXiv:2105.14819 [astro-ph.CO].
[3117] J. D. Barrow and S. Cotsakis, “Chaotic Behavior in Higher Order Gravity Theories,” Phys. Lett. B 232 (1989) 172–176.
[3118] S. Cotsakis and G. Flessas, “Stability of FRW cosmology in higher order gravity,” Phys. Rev. D 48 (1993) 3577–3584.
[3119] S. Cotsakis, J. Demaret, Y. D. Rop, and L. Querella, “Mixmaster universe in fourth-order gravity theories,” Phys. Rev.
D 48 (1993) no. 10, 4595.
[3120] S. Carloni, J. a. L. Rosa, and J. P. S. Lemos, “Cosmology of f (R, □R) gravity,” Phys. Rev. D 99 (2019) no. 10, 104001,
arXiv:1808.07316 [gr-qc].
[3121] L. Amendola, A. Battaglia Mayer, S. Capozziello, F. Occhionero, S. Gottlober, V. Muller, and H. J. Schmidt,
“Generalized sixth order gravity and inflation,” Class. Quant. Grav. 10 (1993) L43–L47.
[3122] R. R. Cuzinatto, C. A. M. de Melo, L. G. Medeiros, and P. J. Pompeia, “f (R, ∇µ1 R, ..., ∇µ1 ...∇µn R) theories of
gravity in Einstein frame: a higher order modified Starobinsky inflation model in the Palatini approach,” Phys. Rev. D
99 (2019) no. 8, 084053, arXiv:1806.08850 [gr-qc].
[3123] S. Gottlober, H. J. Schmidt, and A. A. Starobinsky, “Sixth Order Gravity and Conformal Transformations,” Class.
Quant. Grav. 7 (1990) 893.
[3124] A. Paliathanasis, “f (R, □R)-gravity and equivalency with the modified GUP Scalar field models,” Eur. Phys. J. C 84
(2024) no. 4, 422, arXiv:2404.04519 [gr-qc].
[3125] D. Wands, “Extended gravity theories and the Einstein-Hilbert action,” Class. Quant. Grav. 11 (1994) 269–280,
arXiv:gr-qc/9307034.
[3126] A. L. Berkin and K.-i. Maeda, “Effects of R**3 and R box R terms on R**2 inflation,” Phys. Lett. B 245 (1990)
348–354.
[3127] F. Bajardi and R. D’Agostino, “Corrections to general relativity with higher-order invariants and cosmological
applications,” Int. J. Geom. Meth. Mod. Phys. 21 (2024) no. 10, 2440006.
[3128] B. M. N. Carter and I. P. Neupane, “Towards inflation and dark energy cosmologies from modified Gauss-Bonnet
theory,” JCAP 06 (2006) 004, arXiv:hep-th/0512262.
[3129] F. Bajardi and S. Capozziello, “Equivalence of nonminimally coupled cosmologies by Noether symmetries,” Int. J.
Mod. Phys. D 29 (2020) no. 14, 2030015, arXiv:2010.07914 [gr-qc].
[3130] A. D. Millano, G. Leon, and A. Paliathanasis, “Phase-Space Analysis of an Einstein–Gauss–Bonnet Scalar Field
Cosmology,” Mathematics 11 (2023) no. 6, 1408, arXiv:2302.09371 [gr-qc].
[3131] A. D. Millano, G. Leon, and A. Paliathanasis, “Global dynamics in Einstein-Gauss-Bonnet scalar field cosmology with
matter,” Phys. Rev. D 108 (2023) no. 2, 023519, arXiv:2304.08659 [gr-qc].
[3132] S. Santos Da Costa, F. V. Roig, J. S. Alcaniz, S. Capozziello, M. De Laurentis, and M. Benetti, “Dynamical analysis
on f (R, G) cosmology,” Class. Quant. Grav. 35 (2018) no. 7, 075013, arXiv:1802.02572 [gr-qc].
[3133] B. Li, J. D. Barrow, and D. F. Mota, “The Cosmology of Modified Gauss-Bonnet Gravity,” Phys. Rev. D 76 (2007)
044027, arXiv:0705.3795 [gr-qc].
[3134] G. Cognola, E. Elizalde, S. Nojiri, S. D. Odintsov, and S. Zerbini, “Dark energy in modified Gauss-Bonnet gravity:
Late-time acceleration and the hierarchy problem,” Phys. Rev. D 73 (2006) 084007, arXiv:hep-th/0601008.
[3135] S. Nojiri and S. D. Odintsov, “Modified Gauss-Bonnet theory as gravitational alternative for dark energy,” Phys. Lett.
B 631 (2005) 1–6, arXiv:hep-th/0508049.
[3136] S. Nojiri, S. D. Odintsov, and M. Sasaki, “Gauss-Bonnet dark energy,” Phys. Rev. D 71 (2005) 123509,
arXiv:hep-th/0504052.
[3137] D. Wang and D. Mota, “4D Gauss–Bonnet gravity: Cosmological constraints, H0 tension and large scale structure,”
Phys. Dark Univ. 32 (2021) 100813, arXiv:2103.12358 [astro-ph.CO].
[3138] M. Benetti, S. Santos da Costa, S. Capozziello, J. S. Alcaniz, and M. De Laurentis, “Observational constraints on
366

Gauss–Bonnet cosmology,” Int. J. Mod. Phys. D 27 (2018) no. 08, 1850084, arXiv:1803.00895 [gr-qc].
[3139] S. F. Hassan and R. A. Rosen, “Bimetric Gravity from Ghost-free Massive Gravity,” JHEP 02 (2012) 126,
arXiv:1109.3515 [hep-th].
[3140] M. Högås and E. Mörtsell, “Constraints on bimetric gravity. Part I. Analytical constraints,” JCAP 05 (2021) 001,
arXiv:2101.08794 [gr-qc].
[3141] M. Högås and E. Mörtsell, “Constraints on bimetric gravity. Part II. Observational constraints,” JCAP 05 (2021) 002,
arXiv:2101.08795 [gr-qc].
[3142] M. Högås, Was Einstein Wrong? : Theoretical and observational constraints on massive gravity. PhD thesis,
Stockholm University, Faculty of Science, Department of Physics., 2022.
[3143] P. Ntelis and A. Morris, “Functors of Actions,” Found. Phys. 53 (2023) no. 1, 29, arXiv:2010.06707
[physics.gen-ph].
[3144] P. Ntelis, “New avenues and observational constraints on functors of actions theories,” PoS EPS-HEP2023 (2024) 104.
[3145] J. Beltrán Jiménez, L. Heisenberg, and T. S. Koivisto, “The Geometrical Trinity of Gravity,” Universe 5 (2019) no. 7,
173, arXiv:1903.06830 [hep-th].
[3146] S. Capozziello, V. De Falco, and C. Ferrara, “Comparing equivalent gravities: common features and differences,” Eur.
Phys. J. C 82 (2022) no. 10, 865, arXiv:2208.03011 [gr-qc].
[3147] M. Hohmann, L. Järv, M. Krššák, and C. Pfeifer, “Teleparallel theories of gravity as analogue of nonlinear
electrodynamics,” Phys. Rev. D 97 (2018) no. 10, 104042, arXiv:1711.09930 [gr-qc].
[3148] J. W. Maluf, “The teleparallel equivalent of general relativity,” Annalen Phys. 525 (2013) 339–357, arXiv:1303.3897
[gr-qc].
[3149] C. Xu, E. N. Saridakis, and G. Leon, “Phase-Space analysis of Teleparallel Dark Energy,” JCAP 07 (2012) 005,
arXiv:1202.3781 [gr-qc].
[3150] Y. N. Obukhov and J. G. Pereira, “Metric affine approach to teleparallel gravity,” Phys. Rev. D 67 (2003) 044016,
arXiv:gr-qc/0212080.
[3151] C.-Q. Geng, C.-C. Lee, and E. N. Saridakis, “Observational Constraints on Teleparallel Dark Energy,” JCAP 01 (2012)
002, arXiv:1110.0913 [astro-ph.CO].
[3152] F. Bajardi and S. Capozziello, “Noether symmetries and quantum cosmology in extended teleparallel gravity,” Int. J.
Geom. Meth. Mod. Phys. 18 (2021) no. supp01, 2140002, arXiv:2101.00432 [gr-qc].
[3153] K. Hayashi and T. Shirafuji, “New general relativity.,” Phys. Rev. D 19 (1979) 3524–3553. [Addendum: Phys.Rev.D 24,
3312–3314 (1982)].
[3154] S. Bahamonde, C. G. Böhmer, and M. Krššák, “New classes of modified teleparallel gravity models,” Phys. Lett. B 775
(2017) 37–43, arXiv:1706.04920 [gr-qc].
[3155] L. Järv, M. Rünkla, M. Saal, and O. Vilson, “Nonmetricity formulation of general relativity and its scalar-tensor
extension,” Phys. Rev. D 97 (2018) no. 12, 124025, arXiv:1802.00492 [gr-qc].
[3156] M. Hohmann and C. Pfeifer, “Teleparallel axions and cosmology,” Eur. Phys. J. C 81 (2021) no. 4, 376,
arXiv:2012.14423 [gr-qc].
[3157] Y.-F. Cai, S. Capozziello, M. De Laurentis, and E. N. Saridakis, “f(T) teleparallel gravity and cosmology,” Rept. Prog.
Phys. 79 (2016) no. 10, 106901, arXiv:1511.07586 [gr-qc].
[3158] B. Li, T. P. Sotiriou, and J. D. Barrow, “f (T ) gravity and local Lorentz invariance,” Phys. Rev. D 83 (2011) 064035,
arXiv:1010.1041 [gr-qc].
[3159] R. Ferraro and F. Fiorini, “Modified teleparallel gravity: Inflation without inflaton,” Phys. Rev. D 75 (2007) 084031,
arXiv:gr-qc/0610067.
[3160] E. V. Linder, “Einstein’s Other Gravity and the Acceleration of the Universe,” Phys. Rev. D 81 (2010) 127301,
arXiv:1005.3039 [astro-ph.CO]. [Erratum: Phys.Rev.D 82, 109902 (2010)].
[3161] S.-H. Chen, J. B. Dent, S. Dutta, and E. N. Saridakis, “Cosmological perturbations in f(T) gravity,” Phys. Rev. D 83
(2011) 023508, arXiv:1008.1250 [astro-ph.CO].
[3162] K. Bamba, C.-Q. Geng, C.-C. Lee, and L.-W. Luo, “Equation of state for dark energy in f (T ) gravity,” JCAP 01
(2011) 021, arXiv:1011.0508 [astro-ph.CO].
[3163] G. Kofinas and E. N. Saridakis, “Teleparallel equivalent of Gauss-Bonnet gravity and its modifications,” Phys. Rev. D
90 (2014) 084044, arXiv:1404.2249 [gr-qc].
[3164] A. Paliathanasis, J. D. Barrow, and P. G. L. Leach, “Cosmological Solutions of f (T ) Gravity,” Phys. Rev. D 94 (2016)
no. 2, 023525, arXiv:1606.00659 [gr-qc].
[3165] A. Finch and J. L. Said, “Galactic Rotation Dynamics in f(T) gravity,” Eur. Phys. J. C 78 (2018) no. 7, 560,
arXiv:1806.09677 [astro-ph.GA].
[3166] G. Farrugia, J. Levi Said, V. Gakis, and E. N. Saridakis, “Gravitational Waves in Modified Teleparallel Theories,”
Phys. Rev. D 97 (2018) no. 12, 124064, arXiv:1804.07365 [gr-qc].
[3167] I. Soudi, G. Farrugia, V. Gakis, J. Levi Said, and E. N. Saridakis, “Polarization of gravitational waves in symmetric
teleparallel theories of gravity and their modifications,” Phys. Rev. D 100 (2019) no. 4, 044008, arXiv:1810.08220
[gr-qc].
[3168] G. Farrugia, J. Levi Said, and A. Finch, “Gravitoelectromagnetism, Solar System Tests, and Weak-Field Solutions in
f (T, B) Gravity with Observational Constraints,” Universe 6 (2020) no. 2, 34, arXiv:2002.08183 [gr-qc].
[3169] N. Dimakis, A. Paliathanasis, and T. Christodoulakis, “Exploring quantum cosmology within the framework of
teleparallel f(T) gravity,” Phys. Rev. D 109 (2024) no. 2, 024031, arXiv:2308.08759 [gr-qc].
[3170] I. S. Albuquerque and N. Frusciante, “A designer approach to f(Q) gravity and cosmological implications,” Phys. Dark
367

Univ. 35 (2022) 100980, arXiv:2202.04637 [astro-ph.CO].


[3171] F. K. Anagnostopoulos, V. Gakis, E. N. Saridakis, and S. Basilakos, “New models and big bang nucleosynthesis
constraints in f(Q) gravity,” Eur. Phys. J. C 83 (2023) no. 1, 58, arXiv:2205.11445 [gr-qc].
[3172] S. Capozziello and M. Shokri, “Slow-roll inflation in f(Q) non-metric gravity,” Phys. Dark Univ. 37 (2022) 101113,
arXiv:2209.06670 [gr-qc].
[3173] F. Bajardi, D. Vernieri, and S. Capozziello, “Bouncing Cosmology in f(Q) Symmetric Teleparallel Gravity,” Eur. Phys.
J. Plus 135 (2020) no. 11, 912, arXiv:2011.01248 [gr-qc].
[3174] A. Banerjee, A. Pradhan, T. Tangphati, and F. Rahaman, “Wormhole geometries in f (Q) gravity and the energy
conditions,” Eur. Phys. J. C 81 (2021) no. 11, 1031, arXiv:2109.15105 [gr-qc].
[3175] R. D’Agostino and R. C. Nunes, “Forecasting constraints on deviations from general relativity in f(Q) gravity with
standard sirens,” Phys. Rev. D 106 (2022) no. 12, 124053, arXiv:2210.11935 [gr-qc].
[3176] M. Hohmann and C. Pfeifer, “Gravitational wave birefringence in spatially curved teleparallel cosmology,” Phys. Lett.
B 834 (2022) 137437, arXiv:2203.01856 [gr-qc].
[3177] W. Khyllep, A. Paliathanasis, and J. Dutta, “Cosmological solutions and growth index of matter perturbations in f (Q)
gravity,” Phys. Rev. D 103 (2021) no. 10, 103521, arXiv:2103.08372 [gr-qc].
[3178] S. Capozziello, M. Caruana, G. Farrugia, J. Levi Said, and J. Sultana, “Cosmic growth in f(T) teleparallel gravity,”
Gen. Rel. Grav. 56 (2024) no. 2, 27, arXiv:2308.15995 [gr-qc].
[3179] A. Paliathanasis, “Dynamical analysis of fQ-cosmology,” Phys. Dark Univ. 41 (2023) 101255, arXiv:2304.04219
[gr-qc].
[3180] A. Paliathanasis, “f(T) cosmology with nonzero curvature,” Mod. Phys. Lett. A 36 (2021) no. 38, 2150261,
arXiv:2107.00620 [gr-qc].
[3181] Y. Yang, X. Ren, Q. Wang, Z. Lu, D. Zhang, Y.-F. Cai, and E. N. Saridakis, “Quintom cosmology and modified
gravity after DESI 2024,” Sci. Bull. 69 (2024) 2698–2704, arXiv:2404.19437 [astro-ph.CO].
[3182] A. Paliathanasis, “Attractors in fQ,B-gravity,” Phys. Dark Univ. 45 (2024) 101519.
[3183] C. Wu, X. Ren, Y. Yang, Y.-M. Hu, and E. N. Saridakis, “Background-dependent and classical correspondences
between f (Q) and f (T ) gravity,” arXiv:2412.01104 [gr-qc].
[3184] M. Aljaf, E. Elizalde, M. Khurshudyan, K. Myrzakulov, and A. Zhadyranova, “Solving the H0 tension in f(T) gravity
through Bayesian machine learning,” Eur. Phys. J. C 82 (2022) no. 12, 1130, arXiv:2205.06252 [astro-ph.CO].
[3185] S.-F. Yan, P. Zhang, J.-W. Chen, X.-Z. Zhang, Y.-F. Cai, and E. N. Saridakis, “Interpreting cosmological tensions from
the effective field theory of torsional gravity,” Phys. Rev. D 101 (2020) no. 12, 121301, arXiv:1909.06388
[astro-ph.CO].
[3186] X. Ren, T. H. T. Wong, Y.-F. Cai, and E. N. Saridakis, “Data-driven Reconstruction of the Late-time Cosmic
Acceleration with f(T) Gravity,” Phys. Dark Univ. 32 (2021) 100812, arXiv:2103.01260 [astro-ph.CO].
[3187] R. C. Nunes, “Structure formation in f (T ) gravity and a solution for H0 tension,” JCAP 05 (2018) 052,
arXiv:1802.02281 [gr-qc].
[3188] R. D’Agostino and R. C. Nunes, “Measurements of H0 in modified gravity theories: The role of lensed quasars in the
late-time Universe,” Phys. Rev. D 101 (2020) no. 10, 103505, arXiv:2002.06381 [astro-ph.CO].
[3189] C. de Rham, S. Garcia-Saenz, L. Heisenberg, and V. Pozsgay, “Cosmology of Extended Proca-Nuevo,” JCAP 03 (2022)
053, arXiv:2110.14327 [hep-th].
[3190] B. J. Barros, T. Barreiro, T. Koivisto, and N. J. Nunes, “Testing F (Q) gravity with redshift space distortions,” Phys.
Dark Univ. 30 (2020) 100616, arXiv:2004.07867 [gr-qc].
[3191] Q. Wang, X. Ren, Y.-F. Cai, W. Luo, and E. N. Saridakis, “Observational Test of f(Q) Gravity with Weak
Gravitational Lensing,” Astrophys. J. 974 (2024) no. 1, 7, arXiv:2406.00242 [astro-ph.CO].
[3192] Z. Sakr and L. Schey, “Investigating the Hubble tension and σ 8 discrepancy in f(Q) cosmology,” JCAP 10 (2024) 052,
arXiv:2405.03627 [astro-ph.CO].
[3193] C. Pfeifer and M. N. R. Wohlfarth, “Finsler geometric extension of Einstein gravity,” Phys. Rev. D 85 (2012) 064009,
arXiv:1112.5641 [gr-qc].
[3194] S. Basilakos and P. Stavrinos, “Cosmological equivalence between the Finsler-Randers space-time and the DGP gravity
model,” Phys. Rev. D 87 (2013) no. 4, 043506, arXiv:1301.4327 [gr-qc].
[3195] S. Basilakos, A. P. Kouretsis, E. N. Saridakis, and P. Stavrinos, “Resembling dark energy and modified gravity with
Finsler-Randers cosmology,” Phys. Rev. D 88 (2013) 123510, arXiv:1311.5915 [gr-qc].
[3196] G. Papagiannopoulos, S. Basilakos, A. Paliathanasis, S. Savvidou, and P. C. Stavrinos, “Finsler–Randers cosmology:
dynamical analysis and growth of matter perturbations,” Class. Quant. Grav. 34 (2017) no. 22, 225008,
arXiv:1709.03748 [gr-qc].
[3197] G. Papagiannopoulos, S. Basilakos, A. Paliathanasis, S. Pan, and P. Stavrinos, “Dynamics in varying vacuum
Finsler–Randers cosmology,” Eur. Phys. J. C 80 (2020) no. 9, 816, arXiv:2005.06231 [gr-qc].
[3198] M. Hohmann, C. Pfeifer, and N. Voicu, “The kinetic gas universe,” Eur. Phys. J. C 80 (2020) no. 9, 809,
arXiv:2005.13561 [gr-qc].
[3199] M. Hohmann, C. Pfeifer, and N. Voicu, “Relativistic kinetic gases as direct sources of gravity,” Phys. Rev. D 101
(2020) no. 2, 024062, arXiv:1910.14044 [gr-qc].
[3200] S. Heefer, C. Pfeifer, A. Reggio, and A. Fuster, “A Cosmological unicorn solution to Finsler gravity,” Phys. Rev. D 108
(2023) no. 6, 064051, arXiv:2306.00722 [gr-qc].
[3201] A. F. Zakharov, P. Jovanovic, D. Borka, and V. B. Jovanovic, “Constraining the range of Yukawa gravity interaction
from S2 star orbits II: Bounds on graviton mass,” JCAP 05 (2016) 045, arXiv:1605.00913 [gr-qc].
368

[3202] P. Jovanović, V. B. Jovanović, D. Borka, and A. F. Zakharov, “Constraints on Yukawa gravity parameters from
observations of bright stars,” JCAP 03 (2023) 056, arXiv:2211.12951 [astro-ph.GA].
[3203] P. Jovanović, V. B. Jovanović, D. Borka, and A. F. Zakharov, “Improvement of graviton mass constraints using
GRAVITY’s detection of Schwarzschild precession in the orbit of S2 star around the Galactic Center,” Phys. Rev. D
109 (2024) no. 6, 064046, arXiv:2305.13448 [astro-ph.GA].
[3204] T. Harko and F. S. N. Lobo, “f(R,Lm ) gravity,” Eur. Phys. J. C 70 (2010) 373–379, arXiv:1008.4193 [gr-qc].
[3205] T. Harko, F. S. N. Lobo, S. Nojiri, and S. D. Odintsov, “f (R, T ) gravity,” Phys. Rev. D 84 (2011) 024020,
arXiv:1104.2669 [gr-qc].
[3206] N. Katırcı and M. Kavuk, “f (R, Tµν T µν ) gravity and Cardassian-like expansion as one of its consequences,” Eur. Phys.
J. Plus 129 (2014) 163, arXiv:1302.4300 [gr-qc].
[3207] M. Roshan and F. Shojai, “Energy-Momentum Squared Gravity,” Phys. Rev. D 94 (2016) no. 4, 044002,
arXiv:1607.06049 [gr-qc].
[3208] O. Akarsu, N. Katırcı, and S. Kumar, “Cosmic acceleration in a dust only universe via energy-momentum powered
gravity,” Phys. Rev. D 97 (2018) no. 2, 024011, arXiv:1709.02367 [gr-qc].
[3209] C. V. R. Board and J. D. Barrow, “Cosmological Models in Energy-Momentum-Squared Gravity,” Phys. Rev. D 96
(2017) no. 12, 123517, arXiv:1709.09501 [gr-qc]. [Erratum: Phys.Rev.D 98, 129902 (2018)].
[3210] Z. Haghani, T. Harko, F. S. N. Lobo, H. R. Sepangi, and S. Shahidi, “Further matters in space-time geometry:
f(R,T,RµνTµν) gravity,” Phys. Rev. D 88 (2013) no. 4, 044023, arXiv:1304.5957 [gr-qc].
[3211] I. Ayuso, J. Beltran Jimenez, and A. de la Cruz-Dombriz, “Consistency of universally nonminimally coupled
f (R, T, Rµν T µν ) theories,” Phys. Rev. D 91 (2015) no. 10, 104003, arXiv:1411.1636 [hep-th]. [Addendum:
Phys.Rev.D 93, 089901 (2016)].
[3212] T. Harko, F. S. N. Lobo, G. Otalora, and E. N. Saridakis, “Nonminimal torsion-matter coupling extension of f(T)
gravity,” Phys. Rev. D 89 (2014) 124036, arXiv:1404.6212 [gr-qc].
[3213] S. Carloni, F. S. N. Lobo, G. Otalora, and E. N. Saridakis, “Dynamical system analysis for a nonminimal
torsion-matter coupled gravity,” Phys. Rev. D 93 (2016) 024034, arXiv:1512.06996 [gr-qc].
[3214] T. Harko, F. S. N. Lobo, G. Otalora, and E. N. Saridakis, “f (T, T ) gravity and cosmology,” JCAP 12 (2014) 021,
arXiv:1405.0519 [gr-qc].
[3215] O. Akarsu, M. Bouhmadi-López, N. Katırcı, E. Nazari, M. Roshan, and N. M. Uzun, “Equivalence of matter-type
modified gravity theories to general relativity with nonminimal matter interaction,” Phys. Rev. D 109 (2024) no. 10,
104055, arXiv:2306.11717 [gr-qc].
[3216] F. G. Alvarenga, A. de la Cruz-Dombriz, M. J. S. Houndjo, M. E. Rodrigues, and D. Sáez-Gómez, “Dynamics of scalar
perturbations in f (R, T ) gravity,” Phys. Rev. D 87 (2013) no. 10, 103526, arXiv:1302.1866 [gr-qc]. [Erratum:
Phys.Rev.D 87, 129905 (2013)].
[3217] M. Asghari and A. Sheykhi, “Growth of cosmic perturbations in the modified f(R,T) gravity,” Phys. Dark Univ. 46
(2024) 101695, arXiv:2405.11840 [gr-qc].
[3218] S. Anand, P. Chaubal, A. Mazumdar, S. Mohanty, and P. Parashari, “Bounds on Neutrino Mass in Viscous
Cosmology,” JCAP 05 (2018) 031, arXiv:1712.01254 [astro-ph.CO].
[3219] S. Anand, P. Chaubal, A. Mazumdar, and S. Mohanty, “Cosmic viscosity as a remedy for tension between PLANCK
and LSS data,” JCAP 11 (2017) 005, arXiv:1708.07030 [astro-ph.CO].
[3220] P. Parashari, S. Anand, P. Chaubal, G. Lambiase, S. Mohanty, A. Mazumdar, and A. Narang, “Status of σ8 Tension in
Different Cosmological Models,” Springer Proc. Phys. 261 (2021) 907–912.
[3221] S. Mohanty, S. Anand, P. Chaubal, A. Mazumdar, and P. Parashari, “σ8 Discrepancy and its solutions,” J. Astrophys.
Astron. 39 (2018) no. 4, 46.
[3222] J. Khoury and A. Weltman, “Chameleon fields: Awaiting surprises for tests of gravity in space,” Phys. Rev. Lett. 93
(2004) 171104, arXiv:astro-ph/0309300.
[3223] J. Khoury and A. Weltman, “Chameleon cosmology,” Phys. Rev. D 69 (2004) 044026, arXiv:astro-ph/0309411.
[3224] B. Li and J. D. Barrow, “N-Body Simulations for Coupled Scalar Field Cosmology,” Phys. Rev. D 83 (2011) 024007,
arXiv:1005.4231 [astro-ph.CO].
[3225] A. Paliathanasis, “Dynamical Analysis in Chameleon Dark Energy,” Fortsch. Phys. 71 (2023) no. 8-9, 2300088,
arXiv:2306.03880 [gr-qc].
[3226] H. Farajollahi, A. Salehi, F. Tayebi, and A. Ravanpak, “Stability Analysis in Tachyonic Potential Chameleon
cosmology,” JCAP 05 (2011) 017, arXiv:1105.4045 [gr-qc].
[3227] A. Paliathanasis, “Dynamics in Interacting Scalar-Torsion Cosmology,” Universe 7 (2021) no. 7, 244,
arXiv:2107.05880 [gr-qc].
[3228] A. Paliathanasis, “Chameleon mechanism in scalar nonmetricity cosmology,” Annals Phys. 468 (2024) 169724,
arXiv:2407.05042 [gr-qc].
[3229] A. Paliathanasis, “4D Einstein–Gauss–Bonnet cosmology with Chameleon mechanism,” Gen. Rel. Grav. 56 (2024)
no. 7, 84.
[3230] A. Paliathanasis, G. Leon, and J. D. Barrow, “Einstein-aether theory in Weyl integrable geometry,” Eur. Phys. J. C 80
(2020) no. 12, 1099, arXiv:2007.06435 [gr-qc].
[3231] A. Borowiec and M. Postolak, “Is it possible to separate baryonic from dark matter within the Λ-CDM formalism?,”
Phys. Lett. B 860 (2025) 139176, arXiv:2309.10364 [gr-qc].
[3232] S. Vagnozzi, L. Visinelli, P. Brax, A.-C. Davis, and J. Sakstein, “Direct detection of dark energy: The XENON1T
excess and future prospects,” Phys. Rev. D 104 (2021) no. 6, 063023, arXiv:2103.15834 [hep-ph].
369

[3233] T. O’Shea, A.-C. Davis, M. Giannotti, S. Vagnozzi, L. Visinelli, and J. K. Vogel, “Solar chameleons: Novel channels,”
Phys. Rev. D 110 (2024) no. 6, 063027, arXiv:2406.01691 [hep-ph].
[3234] K. Hinterbichler and J. Khoury, “Symmetron Fields: Screening Long-Range Forces Through Local Symmetry
Restoration,” Phys. Rev. Lett. 104 (2010) 231301, arXiv:1001.4525 [hep-th].
[3235] M. Högås and E. Mörtsell, “Impact of symmetron screening on the Hubble tension: New constraints using cosmic
distance ladder data,” Phys. Rev. D 108 (2023) no. 2, 024007, arXiv:2303.12827 [astro-ph.CO].
[3236] E. Babichev and C. Deffayet, “An introduction to the Vainshtein mechanism,” Class. Quant. Grav. 30 (2013) 184001,
arXiv:1304.7240 [gr-qc].
[3237] H. Akaike, “A new look at the statistical model identification,” IEEE Trans. Automatic Control 19 (1974) no. 6,
716–723.
[3238] G. Schwarz, “Estimating the Dimension of a Model,” Annals Statist. 6 (1978) 461–464.
[3239] E. Belgacem, Y. Dirian, S. Foffa, and M. Maggiore, “Nonlocal gravity. Conceptual aspects and cosmological
predictions,” JCAP 03 (2018) 002, arXiv:1712.07066 [hep-th].
[3240] E. Di Valentino, E. Saridakis, and A. Riess, “Cosmological tensions in the birthplace of the heliocentric model,” Nature
Astronomy 6 (2022) 1353, arXiv:2211.05248 [astro-ph.CO].
[3241] M. Braglia, M. Ballardini, F. Finelli, and K. Koyama, “Early modified gravity in light of the H0 tension and LSS
data,” Phys. Rev. D 103 (2021) no. 4, 043528, arXiv:2011.12934 [astro-ph.CO].
[3242] G. Benevento, J. A. Kable, G. E. Addison, and C. L. Bennett, “An Exploration of an Early Gravity Transition in Light
of Cosmological Tensions,” Astrophys. J. 935 (2022) no. 2, 156, arXiv:2202.09356 [astro-ph.CO].
[3243] S. H.-S. Alexander, M. E. Peskin, and M. M. Sheikh-Jabbari, “Leptogenesis from gravity waves in models of inflation,”
Phys. Rev. Lett. 96 (2006) 081301, arXiv:hep-th/0403069.
[3244] D. H. Lyth, C. Quimbay, and Y. Rodriguez, “Leptogenesis and tensor polarisation from a gravitational Chern-Simons
term,” JHEP 03 (2005) 016, arXiv:hep-th/0501153.
[3245] J. Alexandre, N. Houston, and N. E. Mavromatos, “Dynamical Supergravity Breaking via the Super-Higgs Effect
Revisited,” Phys. Rev. D 88 (2013) 125017, arXiv:1310.4122 [hep-th].
[3246] J. Alexandre, N. Houston, and N. E. Mavromatos, “Inflation via Gravitino Condensation in Dynamically Broken
Supergravity,” Int. J. Mod. Phys. D 24 (2015) no. 04, 1541004, arXiv:1409.3183 [gr-qc].
[3247] O. Bertolami and C. Gomes, “Nonminimally coupled Boltzmann equation: Foundations,” Phys. Rev. D 102 (2020)
no. 8, 084051, arXiv:2002.08184 [gr-qc].
[3248] M. P. L. P. Ramos and J. Páramos, “Baryogenesis in Nonminimally Coupled f (R) Theories,” Phys. Rev. D 96 (2017)
no. 10, 104024, arXiv:1709.04442 [gr-qc].
[3249] T. Złośnik, F. Urban, L. Marzola, and T. Koivisto, “Spacetime and dark matter from spontaneous breaking of Lorentz
symmetry,” Class. Quant. Grav. 35 (2018) no. 23, 235003, arXiv:1807.01100 [gr-qc].
[3250] M. Nikjoo and T. Zlosnik, “Hamiltonian formulation of gravity as a spontaneously-broken gauge theory of the Lorentz
group,” Class. Quant. Grav. 41 (2024) no. 4, 045005, arXiv:2308.01108 [gr-qc].
[3251] P. Gallagher, T. S. Koivisto, L. Marzola, L. Varrin, and T. Zlosnik, “Consistent first-order action functional for gauge
theories,” Phys. Rev. D 109 (2024) no. 6, L061503, arXiv:2311.07464 [hep-th].
[3252] P. Gallagher and T. Koivisto, “The Λ and the CDM as Integration Constants,” Symmetry 13 (2021) no. 11, 2076,
arXiv:2103.05435 [gr-qc].
[3253] T. Koivisto, “Cosmology in the Lorentz gauge theory,” Int. J. Geom. Meth. Mod. Phys. 20 (2023) no. Supp01,
2450040, arXiv:2306.00963 [gr-qc].
[3254] N. J. Popławski, “Non-particle dark matter from Hubble parameter,” Eur. Phys. J. C 79 (2019) no. 9, 734,
arXiv:1906.03947 [physics.gen-ph].
[3255] F. Izaurieta, S. Lepe, and O. Valdivia, “The Spin Tensor of Dark Matter and the Hubble Parameter Tension,” Phys.
Dark Univ. 30 (2020) 100662, arXiv:2004.13163 [gr-qc].
[3256] S. Akhshabi and S. Zamani, “Cosmological distances and hubble tension in Einstein–Cartan theory,” Gen. Rel. Grav.
55 (2023) no. 9, 102, arXiv:2305.00415 [gr-qc].
[3257] T. S. Koivisto and T. Zlosnik, “Paths to gravitation via the gauging of parametrized field theories,” Phys. Rev. D 107
(2023) no. 12, 124013, arXiv:2212.04562 [gr-qc].
[3258] D. Iosifidis, E. Jensko, and T. S. Koivisto, “Relativistic interacting fluids in cosmology,” JCAP 11 (2024) 043,
arXiv:2406.01412 [gr-qc].
[3259] D. Benisty, E. I. Guendelman, D. Vasak, J. Struckmeier, and H. Stoecker, “Quadratic curvature theories formulated as
Covariant Canonical Gauge theories of Gravity,” Phys. Rev. D 98 (2018) no. 10, 106021, arXiv:1809.10447 [gr-qc].
[3260] R. Ferraro and F. Fiorini, “On Born-Infeld Gravity in Weitzenbock spacetime,” Phys. Rev. D 78 (2008) 124019,
arXiv:0812.1981 [gr-qc].
[3261] F. Fiorini and R. Ferraro, “A Type of Born-Infeld regular gravity and its cosmological consequences,” Int. J. Mod.
Phys. A 24 (2009) 1686–1689, arXiv:0904.1767 [gr-qc].
[3262] S. Jana, “Cosmology in a reduced Born-Infeld f (T ) theory of gravity,” Phys. Rev. D 90 (2014) 124007,
arXiv:1410.7117 [gr-qc].
[3263] S. Nesseris and L. Perivolaropoulos, “A Comparison of cosmological models using recent supernova data,” Phys. Rev. D
70 (2004) 043531, arXiv:astro-ph/0401556.
[3264] F. Fiorini, “Nonsingular Promises from Born-Infeld Gravity,” Phys. Rev. Lett. 111 (2013) 041104, arXiv:1306.4392
[gr-qc].
[3265] M. Bouhmadi-Lopez, C.-Y. Chen, and P. Chen, “Cosmological singularities in Born-Infeld determinantal gravity,”
370

Phys. Rev. D 90 (2014) 123518, arXiv:1407.5114 [gr-qc].


[3266] F. Fiorini, “Primordial brusque bounce in Born-Infeld determinantal gravity,” Phys. Rev. D 94 (2016) no. 2, 024030,
arXiv:1511.03227 [gr-qc].
[3267] E. Wilson-Ewing, “The Matter Bounce Scenario in Loop Quantum Cosmology,” JCAP 03 (2013) 026,
arXiv:1211.6269 [gr-qc].
[3268] Y.-F. Cai, S.-H. Chen, J. B. Dent, S. Dutta, and E. N. Saridakis, “Matter Bounce Cosmology with the f(T) Gravity,”
Class. Quant. Grav. 28 (2011) 215011, arXiv:1104.4349 [astro-ph.CO].
[3269] K. Bamba, J. de Haro, and S. D. Odintsov, “Future Singularities and Teleparallelism in Loop Quantum Cosmology,”
JCAP 02 (2013) 008, arXiv:1211.2968 [gr-qc].
[3270] J. Amorós, J. de Haro, and S. D. Odintsov, “Bouncing loop quantum cosmology from F (T ) gravity,” Phys. Rev. D 87
(2013) 104037, arXiv:1305.2344 [gr-qc].
[3271] A. Casalino, B. Sanna, L. Sebastiani, and S. Zerbini, “Bounce Models within Teleparallel modified gravity,” Phys. Rev.
D 103 (2021) no. 2, 023514, arXiv:2010.07609 [gr-qc].
[3272] J. Haro and J. Amoros, “Viability of the matter bounce scenario in F (T ) gravity and Loop Quantum Cosmology for
general potentials,” JCAP 12 (2014) 031, arXiv:1406.0369 [gr-qc].
[3273] J. de Haro and J. Amorós, “Viability of the Matter Bounce Scenario,” J. Phys. Conf. Ser. 600 (2015) no. 1, 012024,
arXiv:1411.7611 [gr-qc].
[3274] J. Haro and J. Amorós, “Matter Bounce Scenario in F (T ) gravity,” PoS FFP14 (2016) 163, arXiv:1501.06270
[gr-qc].
[3275] J. de Haro and Y.-F. Cai, “An Extended Matter Bounce Scenario: current status and challenges,” Gen. Rel. Grav. 47
(2015) no. 8, 95, arXiv:1502.03230 [gr-qc].
[3276] S. Raatikainen and S. Rasanen, “Higgs inflation and teleparallel gravity,” JCAP 12 (2019) 021, arXiv:1910.03488
[gr-qc].
[3277] A. Ashtekar, M. Bojowald, and J. Lewandowski, “Mathematical structure of loop quantum cosmology,” Adv. Theor.
Math. Phys. 7 (2003) no. 2, 233–268, arXiv:gr-qc/0304074.
[3278] A. Ashtekar, T. Pawlowski, and P. Singh, “Quantum nature of the big bang,” Phys. Rev. Lett. 96 (2006) 141301,
arXiv:gr-qc/0602086.
[3279] A. Ashtekar, T. Pawlowski, and P. Singh, “Quantum Nature of the Big Bang: Improved dynamics,” Phys. Rev. D 74
(2006) 084003, arXiv:gr-qc/0607039.
[3280] A. Ashtekar, A. Corichi, and P. Singh, “Robustness of key features of loop quantum cosmology,” Phys. Rev. D 77
(2008) 024046, arXiv:0710.3565 [gr-qc].
[3281] V. Taveras, “Corrections to the Friedmann Equations from LQG for a Universe with a Free Scalar Field,” Phys. Rev. D
78 (2008) 064072, arXiv:0807.3325 [gr-qc].
[3282] P. Diener, B. Gupt, M. Megevand, and P. Singh, “Numerical evolution of squeezed and non-Gaussian states in loop
quantum cosmology,” Class. Quant. Grav. 31 (2014) 165006, arXiv:1406.1486 [gr-qc].
[3283] J. Yang, Y. Ding, and Y. Ma, “Alternative quantization of the Hamiltonian in loop quantum cosmology II: Including
the Lorentz term,” Phys. Lett. B 682 (2009) 1–7, arXiv:0904.4379 [gr-qc].
[3284] A. Dapor and K. Liegener, “Cosmological Effective Hamiltonian from full Loop Quantum Gravity Dynamics,” Phys.
Lett. B 785 (2018) 506–510, arXiv:1706.09833 [gr-qc].
[3285] B.-F. Li, P. Singh, and A. Wang, “Towards Cosmological Dynamics from Loop Quantum Gravity,” Phys. Rev. D 97
(2018) no. 8, 084029, arXiv:1801.07313 [gr-qc].
[3286] B.-F. Li, P. Singh, and A. Wang, “Qualitative dynamics and inflationary attractors in loop cosmology,” Phys. Rev. D
98 (2018) no. 6, 066016, arXiv:1807.05236 [gr-qc].
[3287] M. Assanioussi, A. Dapor, K. Liegener, and T. Pawłowski, “Emergent de Sitter epoch of the Loop Quantum Cosmos: a
detailed analysis,” Phys. Rev. D 100 (2019) no. 8, 084003, arXiv:1906.05315 [gr-qc].
[3288] A. Delhom, G. J. Olmo, and P. Singh, “A diffeomorphism invariant family of metric-affine actions for loop
cosmologies,” JCAP 06 (2023) 059, arXiv:2302.04285 [gr-qc].
[3289] G. J. Olmo, “Palatini Approach to Modified Gravity: f(R) Theories and Beyond,” Int. J. Mod. Phys. D 20 (2011)
413–462, arXiv:1101.3864 [gr-qc].
[3290] G. J. Olmo and P. Singh, “Effective Action for Loop Quantum Cosmology a la Palatini,” JCAP 01 (2009) 030,
arXiv:0806.2783 [gr-qc].
[3291] F. Bombacigno, F. Cianfrani, and G. Montani, “Big-Bounce cosmology in the presence of Immirzi field,” Phys. Rev. D
94 (2016) no. 6, 064021, arXiv:1607.00910 [gr-qc].
[3292] F. Bombacigno and G. Montani, “Big bounce cosmology for Palatini R2 gravity with a Nieh–Yan term,” Eur. Phys. J.
C 79 (2019) no. 5, 405, arXiv:1809.07563 [gr-qc].
[3293] F. Bombacigno, S. Boudet, G. J. Olmo, and G. Montani, “Big bounce and future time singularity resolution in Bianchi
I cosmologies: The projective invariant Nieh-Yan case,” Phys. Rev. D 103 (2021) no. 12, 124031, arXiv:2105.06870
[gr-qc].
[3294] S. Boudet, F. Bombacigno, F. Moretti, and G. J. Olmo, “Torsional birefringence in metric-affine Chern-Simons gravity:
gravitational waves in late-time cosmology,” JCAP 01 (2023) 026, arXiv:2209.14394 [gr-qc].
[3295] S. Boudet, F. Bombacigno, G. J. Olmo, and P. J. Porfirio, “Quasinormal modes of Schwarzschild black holes in
projective invariant Chern-Simons modified gravity,” JCAP 05 (2022) no. 05, 032, arXiv:2203.04000 [gr-qc].
[3296] F. Bombacigno, F. Moretti, S. Boudet, and G. J. Olmo, “Landau damping for gravitational waves in parity-violating
theories,” JCAP 02 (2023) 009, arXiv:2210.07673 [gr-qc].
371

[3297] K. Choi, “String or M theory axion as a quintessence,” Phys. Rev. D 62 (2000) 043509, arXiv:hep-ph/9902292.
[3298] J. E. Kim and H. P. Nilles, “A Quintessential axion,” Phys. Lett. B 553 (2003) 1–6, arXiv:hep-ph/0210402.
[3299] A. Arvanitaki, S. Dimopoulos, S. Dubovsky, N. Kaloper, and J. March-Russell, “String Axiverse,” Phys. Rev. D 81
(2010) 123530, arXiv:0905.4720 [hep-th].
[3300] S. Chakraborty, E. González, G. Leon, and B. Wang, “Time-averaging axion-like interacting scalar fields models,” Eur.
Phys. J. C 81 (2021) no. 11, 1039, arXiv:2107.04651 [gr-qc].
[3301] J. E. Kim and H. P. Nilles, “Axionic dark energy and a composite QCD axion,” JCAP 05 (2009) 010,
arXiv:0902.3610 [hep-th].
[3302] A. Chatzistavrakidis, E. Erfani, H. P. Nilles, and I. Zavala, “Axiology,” JCAP 09 (2012) 006, arXiv:1207.1128
[hep-ph].
[3303] A. Chatzistavrakidis, G. Karagiannis, and P. Schupp, “Torsion-induced gravitational θ term and
gravitoelectromagnetism,” Eur. Phys. J. C 80 (2020) no. 11, 1034, arXiv:2007.06632 [gr-qc].
[3304] F. Zhang, J.-X. Feng, and X. Gao, “Scalar induced gravitational waves in symmetric teleparallel gravity with a
parity-violating term,” Phys. Rev. D 108 (2023) no. 6, 063513, arXiv:2307.00330 [gr-qc].
[3305] M. Lattanzi and S. Mercuri, “A solution of the strong CP problem via the Peccei-Quinn mechanism through the
Nieh-Yan modified gravity and cosmological implications,” Phys. Rev. D 81 (2010) 125015, arXiv:0911.2698 [gr-qc].
[3306] M. Li, H. Rao, and D. Zhao, “A simple parity violating gravity model without ghost instability,” JCAP 11 (2020) 023,
arXiv:2007.08038 [gr-qc].
[3307] A. Chatzistavrakidis, G. Karagiannis, G. Manolakos, and P. Schupp, “Axion gravitodynamics, Lense-Thirring effect,
and gravitational waves,” Phys. Rev. D 105 (2022) no. 10, 104029, arXiv:2111.04388 [gr-qc].
[3308] M. Lagos, L. Jenks, M. Isi, K. Hotokezaka, B. D. Metzger, E. Burns, W. M. Farr, S. Perkins, K. W. K. Wong, and
N. Yunes, “Birefringence tests of gravity with multimessenger binaries,” Phys. Rev. D 109 (2024) no. 12, 124003,
arXiv:2402.05316 [gr-qc].
[3309] J. Su, T. Harko, and S.-D. Liang, “Irreversible thermodynamic description of dark matter and radiation creation
during inflationary reheating,” Adv. High Energy Phys. 2017 (2017) 7650238, arXiv:1708.08004 [gr-qc].
[3310] T. Harko and H. Sheikhahmadi, “Warm inflation with non-comoving scalar field and radiation fluid,” Eur. Phys. J. C
81 (2021) no. 2, 165, arXiv:2102.04728 [gr-qc].
[3311] T. Matei, T. Harko, and G. Mocanu, “Dark matter and radiation production during warm inflation in a curved
Universe-an irreversible thermodynamic approach,” arXiv:2303.02464 [gr-qc].
[3312] I. Prigogine, J. Geheniau, E. Gunzig, and P. Nardone, “Thermodynamics of cosmological matter creation,” Proc. Nat.
Acad. Sci. 85 (1988) no. 20, 7428.
[3313] I. Prigogine, J. Geheniau, E. Gunzig, and P. Nardone, “Thermodynamics and cosmology,” Gen. Rel. Grav. 21 (1989)
767–776.
[3314] T. Harko, “Thermodynamic interpretation of the generalized gravity models with geometry - matter coupling,” Phys.
Rev. D 90 (2014) no. 4, 044067, arXiv:1408.3465 [gr-qc].
[3315] T. Harko, F. S. N. Lobo, J. P. Mimoso, and D. Pavón, “Gravitational induced particle production through a
nonminimal curvature–matter coupling,” Eur. Phys. J. C 75 (2015) 386, arXiv:1508.02511 [gr-qc].
[3316] M. A. S. Pinto, T. Harko, and F. S. N. Lobo, “Gravitationally induced particle production in scalar-tensor f(R,T)
gravity,” Phys. Rev. D 106 (2022) no. 4, 044043, arXiv:2205.12545 [gr-qc].
[3317] R. A. C. Cipriano, T. Harko, F. S. N. Lobo, M. A. S. Pinto, and J. a. L. Rosa, “Gravitationally induced matter
creation in scalar–tensor f(R,TµνTµν) gravity,” Phys. Dark Univ. 44 (2024) 101463, arXiv:2310.15018 [gr-qc].
[3318] M. A. S. Pinto, T. Harko, and F. S. N. Lobo, “Irreversible Geometrothermodynamics of Open Systems in Modified
Gravity,” Entropy 25 (2023) no. 6, 944, arXiv:2306.13912 [gr-qc].
[3319] S. Basilakos, S. Das, and E. C. Vagenas, “Quantum Gravity Corrections and Entropy at the Planck time,” JCAP 09
(2010) 027, arXiv:1009.0365 [hep-th].
[3320] S. Das, M. Fridman, G. Lambiase, and E. C. Vagenas, “Baryon asymmetry and minimum length,” in 16th Marcel
Grossmann Meeting on Recent Developments in Theoretical and Experimental General Relativity, Astrophysics and
Relativistic Field Theories. 2021. arXiv:2111.01278 [gr-qc].
[3321] S. Das, M. Fridman, G. Lambiase, and E. C. Vagenas, “Baryon asymmetry from the generalized uncertainty principle,”
Phys. Lett. B 824 (2022) 136841, arXiv:2107.02077 [gr-qc].
[3322] L. A. Escamilla, D. Fiorucci, G. Montani, and E. Di Valentino, “Exploring the Hubble tension with a late time
Modified Gravity scenario,” Phys. Dark Univ. 46 (2024) 101652, arXiv:2408.04354 [astro-ph.CO].
[3323] M. V. Battisti and G. Montani, “The Big bang singularity in the framework of a generalized uncertainty principle,”
Phys. Lett. B 656 (2007) 96–101, arXiv:gr-qc/0703025.
[3324] S. Kouwn, “Implications of Minimum and Maximum Length Scales in Cosmology,” Phys. Dark Univ. 21 (2018) 76–81,
arXiv:1805.07278 [astro-ph.CO].
[3325] F. Fragomeno, D. M. Gingrich, S. Hergott, S. Rastgoo, and E. Vienneau, “A generalized uncertainty-inspired quantum
black hole,” Phys. Rev. D 111 (2025) no. 2, 024048, arXiv:2406.03909 [gr-qc].
[3326] M. Reuter, “Nonperturbative evolution equation for quantum gravity,” Phys. Rev. D 57 (1998) 971–985,
arXiv:hep-th/9605030.
[3327] M. Reuter and F. Saueressig, “From big bang to asymptotic de Sitter: Complete cosmologies in a quantum gravity
framework,” JCAP 09 (2005) 012, arXiv:hep-th/0507167.
[3328] A. Bonanno and M. Reuter, “Entropy signature of the running cosmological constant,” JCAP 08 (2007) 024,
arXiv:0706.0174 [hep-th].
372

[3329] S. Weinberg, “Asymptotically Safe Inflation,” Phys. Rev. D 81 (2010) 083535, arXiv:0911.3165 [hep-th].
[3330] A. Bonanno and A. Platania, “Asymptotically safe inflation from quadratic gravity,” Phys. Lett. B 750 (2015)
638–642, arXiv:1507.03375 [gr-qc].
[3331] IceCube Collaboration, M. G. Aartsen et al., “Evidence for Astrophysical Muon Neutrinos from the Northern Sky
with IceCube,” Phys. Rev. Lett. 115 (2015) no. 8, 081102, arXiv:1507.04005 [astro-ph.HE].
[3332] IceCube Collaboration, M. G. Aartsen et al., “Determining neutrino oscillation parameters from atmospheric muon
neutrino disappearance with three years of IceCube DeepCore data,” Phys. Rev. D 91 (2015) no. 7, 072004,
arXiv:1410.7227 [hep-ex].
[3333] M. Chianese and A. Merle, “A Consistent Theory of Decaying Dark Matter Connecting IceCube to the Sesame Street,”
JCAP 04 (2017) 017, arXiv:1607.05283 [hep-ph].
[3334] G. Lambiase, S. Mohanty, and A. Stabile, “PeV IceCube signals and Dark Matter relic abundance in modified
cosmologies,” Eur. Phys. J. C 78 (2018) no. 4, 350, arXiv:1804.07369 [astro-ph.CO].
[3335] P. Jizba and G. Lambiase, “Tsallis cosmology and its applications in dark matter physics with focus on IceCube
high-energy neutrino data,” Eur. Phys. J. C 82 (2022) no. 12, 1123, arXiv:2206.12910 [hep-th].
[3336] P. Jizba and G. Lambiase, “Constraints on Tsallis Cosmology from Big Bang Nucleosynthesis and the Relic Abundance
of Cold Dark Matter Particles,” Entropy 25 (2023) no. 11, 1495, arXiv:2310.19045 [gr-qc].
[3337] K. Lundmark, “Über die Bestimmung der Entfernungen, Dimensionen, Massen und Dichtigkeit fur die nächstgelegenen
anagalacktischen Sternsysteme.,” Meddelanden fran Lunds Astronomiska Observatorium Serie I 125 (1930) 1–13.
[3338] F. Zwicky, “Die Rotverschiebung von extragalaktischen Nebeln,” Helv. Phys. Acta 6 (1933) 110–127.
[3339] F. Zwicky, “On the Masses of Nebulae and of Clusters of Nebulae,” Astrophys. J. 86 (1937) 217–246.
[3340] V. C. Rubin and W. K. Ford, Jr., “Rotation of the Andromeda Nebula from a Spectroscopic Survey of Emission
Regions,” Astrophys. J. 159 (1970) 379–403.
[3341] K. C. Freeman, “On the disks of spiral and SO Galaxies,” Astrophys. J. 160 (1970) 811.
[3342] J. de Swart, G. Bertone, and J. van Dongen, “How Dark Matter Came to Matter,” Nature Astron. 1 (2017) 0059,
arXiv:1703.00013 [astro-ph.CO].
[3343] G. Jungman, M. Kamionkowski, and K. Griest, “Supersymmetric dark matter,” Phys. Rept. 267 (1996) 195–373,
arXiv:hep-ph/9506380.
[3344] G. Bertone, D. Hooper, and J. Silk, “Particle dark matter: Evidence, candidates and constraints,” Phys. Rept. 405
(2005) 279–390, arXiv:hep-ph/0404175.
[3345] F. Iocco, M. Pato, and G. Bertone, “Evidence for dark matter in the inner Milky Way,” Nature Phys. 11 (2015)
245–248, arXiv:1502.03821 [astro-ph.GA].
[3346] I. de Martino, S. S. Chakrabarty, V. Cesare, A. Gallo, L. Ostorero, and A. Diaferio, “Dark matters on the scale of
galaxies,” Universe 6 (2020) no. 8, 107, arXiv:2007.15539 [astro-ph.CO].
[3347] M. Cirelli, A. Strumia, and J. Zupan, “Dark Matter,” arXiv:2406.01705 [hep-ph].
[3348] M. Vogelsberger, F. Marinacci, P. Torrey, and E. Puchwein, “Cosmological Simulations of Galaxy Formation,” Nature
Rev. Phys. 2 (2020) no. 1, 42–66, arXiv:1909.07976 [astro-ph.GA].
[3349] R. Davé, D. Anglés-Alcázar, D. Narayanan, Q. Li, M. H. Rafieferantsoa, and S. Appleby, “Simba: Cosmological
Simulations with Black Hole Growth and Feedback,” Mon. Not. Roy. Astron. Soc. 486 (2019) no. 2, 2827–2849,
arXiv:1901.10203 [astro-ph.GA].
[3350] L. Roszkowski, “Particle dark matter: A Theorist’s perspective,” Pramana 62 (2004) 389–401, arXiv:hep-ph/0404052.
[3351] J. E. Kim and G. Carosi, “Axions and the Strong CP Problem,” Rev. Mod. Phys. 82 (2010) 557–602, arXiv:0807.3125
[hep-ph]. [Erratum: Rev.Mod.Phys. 91, 049902 (2019)].
[3352] H. Baer, K.-Y. Choi, J. E. Kim, and L. Roszkowski, “Dark matter production in the early Universe: beyond the
thermal WIMP paradigm,” Phys. Rept. 555 (2015) 1–60, arXiv:1407.0017 [hep-ph].
[3353] B. Paczynski, “Gravitational microlensing by the galactic halo,” Astrophys. J. 304 (1986) 1–5.
[3354] K. Griest, “Galactic Microlensing as a Method of Detecting Massive Compact Halo Objects,” Astrophys. J. 366 (1991)
412–421.
[3355] MACHO Collaboration, C. Alcock et al., “The MACHO project: Microlensing results from 5.7 years of LMC
observations,” Astrophys. J. 542 (2000) 281–307, arXiv:astro-ph/0001272.
[3356] EROS-2 Collaboration, P. Tisserand et al., “Limits on the Macho Content of the Galactic Halo from the EROS-2
Survey of the Magellanic Clouds,” Astron. Astrophys. 469 (2007) 387–404, arXiv:astro-ph/0607207.
[3357] L. Wyrzykowski et al., “The OGLE View of Microlensing towards the Magellanic Clouds. III. Ruling out sub-solar
MACHOs with the OGLE-III LMC data,” Mon. Not. Roy. Astron. Soc. 413 (2011) 493, arXiv:1012.1154
[astro-ph.GA].
[3358] H. Niikura et al., “Microlensing constraints on primordial black holes with Subaru/HSC Andromeda observations,”
Nature Astron. 3 (2019) no. 6, 524–534, arXiv:1701.02151 [astro-ph.CO].
[3359] D. N. Page, “Particle Emission Rates from a Black Hole: Massless Particles from an Uncharged, Nonrotating Hole,”
Phys. Rev. D 13 (1976) 198–206.
[3360] M. Ricotti, J. P. Ostriker, and K. J. Mack, “Effect of Primordial Black Holes on the Cosmic Microwave Background
and Cosmological Parameter Estimates,” Astrophys. J. 680 (2008) 829, arXiv:0709.0524 [astro-ph].
[3361] D. Gaggero, G. Bertone, F. Calore, R. M. T. Connors, M. Lovell, S. Markoff, and E. Storm, “Searching for Primordial
Black Holes in the radio and X-ray sky,” Phys. Rev. Lett. 118 (2017) no. 24, 241101, arXiv:1612.00457
[astro-ph.HE].
[3362] S. K. Acharya and R. Khatri, “CMB and BBN constraints on evaporating primordial black holes revisited,” JCAP 06
373

(2020) 018, arXiv:2002.00898 [astro-ph.CO].


[3363] M. Korwar and S. Profumo, “Updated constraints on primordial black hole evaporation,” JCAP 05 (2023) 054,
arXiv:2302.04408 [hep-ph].
[3364] S. Tremaine and J. E. Gunn, “Dynamical Role of Light Neutral Leptons in Cosmology,” Phys. Rev. Lett. 42 (1979)
407–410.
[3365] J. Alvey, N. Sabti, V. Tiki, D. Blas, K. Bondarenko, A. Boyarsky, M. Escudero, M. Fairbairn, M. Orkney, and J. I.
Read, “New constraints on the mass of fermionic dark matter from dwarf spheroidal galaxies,” Mon. Not. Roy. Astron.
Soc. 501 (2021) no. 1, 1188–1201, arXiv:2010.03572 [hep-ph].
[3366] G. Chauhan, P. S. B. Dev, I. Dubovyk, B. Dziewit, W. Flieger, K. Grzanka, J. Gluza, B. Karmakar, and S. Zięba,
“Phenomenology of lepton masses and mixing with discrete flavor symmetries,” Prog. Part. Nucl. Phys. 138 (2024)
104126, arXiv:2310.20681 [hep-ph].
[3367] V. Keus, S. F. King, and S. Moretti, “Three-Higgs-doublet models: symmetries, potentials and Higgs boson masses,”
JHEP 01 (2014) 052, arXiv:1310.8253 [hep-ph].
[3368] S. Centelles Chuliá, E. Ma, R. Srivastava, and J. W. F. Valle, “Dirac Neutrinos and Dark Matter Stability from Lepton
Quarticity,” Phys. Lett. B 767 (2017) 209–213, arXiv:1606.04543 [hep-ph].
[3369] S. Weinberg, “Baryon and Lepton Nonconserving Processes,” Phys. Rev. Lett. 43 (1979) 1566–1570.
[3370] F. Wilczek and A. Zee, “Operator Analysis of Nucleon Decay,” Phys. Rev. Lett. 43 (1979) 1571–1573.
[3371] S. Weinberg, “Varieties of Baryon and Lepton Nonconservation,” Phys. Rev. D 22 (1980) 1694.
[3372] H. A. Weldon and A. Zee, “Operator Analysis of New Physics,” Nucl. Phys. B 173 (1980) 269–290.
[3373] J. R. Ellis, J. S. Hagelin, D. V. Nanopoulos, K. A. Olive, and M. Srednicki, “Supersymmetric Relics from the Big
Bang,” Nucl. Phys. B 238 (1984) 453–476.
[3374] M. Drees and M. M. Nojiri, “The Neutralino relic density in minimal N = 1 supergravity,” Phys. Rev. D 47 (1993)
376–408, arXiv:hep-ph/9207234.
[3375] H. Baer and M. Brhlik, “Cosmological relic density from minimal supergravity with implications for collider physics,”
Phys. Rev. D 53 (1996) 597–605, arXiv:hep-ph/9508321.
[3376] V. D. Barger and C. Kao, “Relic density of neutralino dark matter in supergravity models,” Phys. Rev. D 57 (1998)
3131–3139, arXiv:hep-ph/9704403.
[3377] A. B. Lahanas, D. V. Nanopoulos, and V. C. Spanos, “Neutralino relic density in a universe with nonvanishing
cosmological constant,” Phys. Rev. D 62 (2000) 023515, arXiv:hep-ph/9909497.
[3378] R. H. Cyburt, J. Ellis, B. D. Fields, F. Luo, K. A. Olive, and V. C. Spanos, “Nucleosynthesis Constraints on a Massive
Gravitino in Neutralino Dark Matter Scenarios,” JCAP 10 (2009) 021, arXiv:0907.5003 [astro-ph.CO].
[3379] T. Han, Z. Liu, and S. Su, “Light Neutralino Dark Matter: Direct/Indirect Detection and Collider Searches,” JHEP 08
(2014) 093, arXiv:1406.1181 [hep-ph].
[3380] L. Roszkowski, E. M. Sessolo, and S. Trojanowski, “WIMP dark matter candidates and searches—current status and
future prospects,” Rept. Prog. Phys. 81 (2018) no. 6, 066201, arXiv:1707.06277 [hep-ph].
[3381] T. Falk, K. A. Olive, and M. Srednicki, “Heavy sneutrinos as dark matter,” Phys. Lett. B 339 (1994) 248–251,
arXiv:hep-ph/9409270.
[3382] C. Arina and N. Fornengo, “Sneutrino cold dark matter, a new analysis: Relic abundance and detection rates,” JHEP
11 (2007) 029, arXiv:0709.4477 [hep-ph].
[3383] S. Weinberg, “Cosmological Constraints on the Scale of Supersymmetry Breaking,” Phys. Rev. Lett. 48 (1982) 1303.
[3384] M. Y. Khlopov and A. D. Linde, “Is It Easy to Save the Gravitino?,” Phys. Lett. B 138 (1984) 265–268.
[3385] J. R. Ellis, J. E. Kim, and D. V. Nanopoulos, “Cosmological Gravitino Regeneration and Decay,” Phys. Lett. B 145
(1984) 181–186.
[3386] M. Bolz, A. Brandenburg, and W. Buchmuller, “Thermal production of gravitinos,” Nucl. Phys. B 606 (2001) 518–544,
arXiv:hep-ph/0012052. [Erratum: Nucl.Phys.B 790, 336–337 (2008)].
[3387] M. Kawasaki, K. Kohri, T. Moroi, and A. Yotsuyanagi, “Big-Bang Nucleosynthesis and Gravitino,” Phys. Rev. D 78
(2008) 065011, arXiv:0804.3745 [hep-ph].
[3388] J. Pradler and F. D. Steffen, “Thermal gravitino production and collider tests of leptogenesis,” Phys. Rev. D 75 (2007)
023509, arXiv:hep-ph/0608344.
[3389] V. S. Rychkov and A. Strumia, “Thermal production of gravitinos,” Phys. Rev. D 75 (2007) 075011,
arXiv:hep-ph/0701104.
[3390] J. Ellis, M. A. G. Garcia, D. V. Nanopoulos, K. A. Olive, and M. Peloso, “Post-Inflationary Gravitino Production
Revisited,” JCAP 03 (2016) 008, arXiv:1512.05701 [astro-ph.CO].
[3391] E. Dudas, Y. Mambrini, and K. Olive, “Case for an EeV Gravitino,” Phys. Rev. Lett. 119 (2017) no. 5, 051801,
arXiv:1704.03008 [hep-ph].
[3392] K. Kaneta, Y. Mambrini, and K. A. Olive, “Radiative production of nonthermal dark matter,” Phys. Rev. D 99 (2019)
no. 6, 063508, arXiv:1901.04449 [hep-ph].
[3393] H. Eberl, I. D. Gialamas, and V. C. Spanos, “Gravitino thermal production revisited,” Phys. Rev. D 103 (2021) no. 7,
075025, arXiv:2010.14621 [hep-ph].
[3394] L. Covi, J. E. Kim, and L. Roszkowski, “Axinos as cold dark matter,” Phys. Rev. Lett. 82 (1999) 4180–4183,
arXiv:hep-ph/9905212.
[3395] L. Covi, H.-B. Kim, J. E. Kim, and L. Roszkowski, “Axinos as dark matter,” JHEP 05 (2001) 033,
arXiv:hep-ph/0101009.
[3396] K. Aoki and K.-i. Maeda, “Dark matter in ghost-free bigravity theory: From a galaxy scale to the universe,” Phys. Rev.
374

D 90 (2014) 124089, arXiv:1409.0202 [gr-qc].


[3397] K. Aoki and S. Mukohyama, “Massive gravitons as dark matter and gravitational waves,” Phys. Rev. D 94 (2016)
no. 2, 024001, arXiv:1604.06704 [hep-th].
[3398] E. Babichev, L. Marzola, M. Raidal, A. Schmidt-May, F. Urban, H. Veermäe, and M. von Strauss, “Bigravitational
origin of dark matter,” Phys. Rev. D 94 (2016) no. 8, 084055, arXiv:1604.08564 [hep-ph].
[3399] E. Babichev, L. Marzola, M. Raidal, A. Schmidt-May, F. Urban, H. Veermäe, and M. von Strauss, “Heavy spin-2 Dark
Matter,” JCAP 09 (2016) 016, arXiv:1607.03497 [hep-th].
[3400] E. W. Kolb, S. Ling, A. J. Long, and R. A. Rosen, “Cosmological gravitational particle production of massive spin-2
particles,” JHEP 05 (2023) 181, arXiv:2302.04390 [astro-ph.CO].
[3401] H.-C. Cheng, J. L. Feng, and K. T. Matchev, “Kaluza-Klein dark matter,” Phys. Rev. Lett. 89 (2002) 211301,
arXiv:hep-ph/0207125.
[3402] G. Servant and T. M. P. Tait, “Is the lightest Kaluza-Klein particle a viable dark matter candidate?,” Nucl. Phys. B
650 (2003) 391–419, arXiv:hep-ph/0206071.
[3403] D. Hooper and S. Profumo, “Dark Matter and Collider Phenomenology of Universal Extra Dimensions,” Phys. Rept.
453 (2007) 29–115, arXiv:hep-ph/0701197.
[3404] A. Cordero-Cid, J. Hernández-Sánchez, V. Keus, S. F. King, S. Moretti, D. Rojas, and D. Sokołowska, “CP violating
scalar Dark Matter,” JHEP 12 (2016) 014, arXiv:1608.01673 [hep-ph].
[3405] J. Hernandez-Sanchez, V. Keus, S. Moretti, D. Rojas-Ciofalo, and D. Sokolowska, “Complementary Probes of
Two-component Dark Matter,” arXiv:2012.11621 [hep-ph].
[3406] T. Alanne, M. Heikinheimo, V. Keus, N. Koivunen, and K. Tuominen, “Direct and indirect probes of Goldstone dark
matter,” Phys. Rev. D 99 (2019) no. 7, 075028, arXiv:1812.05996 [hep-ph].
[3407] V. Keus, “Dark CP-violation through the Z-portal,” Phys. Rev. D 101 (2020) no. 7, 073007, arXiv:1909.09234
[hep-ph].
[3408] K. Griest and D. Seckel, “Three exceptions in the calculation of relic abundances,” Phys. Rev. D 43 (1991) 3191–3203.
[3409] J. Edsjo and P. Gondolo, “Neutralino relic density including coannihilations,” Phys. Rev. D 56 (1997) 1879–1894,
arXiv:hep-ph/9704361.
[3410] P. Gondolo and G. Gelmini, “Cosmic abundances of stable particles: Improved analysis,” Nucl. Phys. B 360 (1991)
145–179.
[3411] J. Hisano, S. Matsumoto, M. M. Nojiri, and O. Saito, “Direct detection of the Wino and Higgsino-like neutralino dark
matters at one-loop level,” Phys. Rev. D 71 (2005) 015007, arXiv:hep-ph/0407168.
[3412] N. Arkani-Hamed, D. P. Finkbeiner, T. R. Slatyer, and N. Weiner, “A Theory of Dark Matter,” Phys. Rev. D 79
(2009) 015014, arXiv:0810.0713 [hep-ph].
[3413] C. Boehm and P. Fayet, “Scalar dark matter candidates,” Nucl. Phys. B 683 (2004) 219–263, arXiv:hep-ph/0305261.
[3414] M. Pospelov, A. Ritz, and M. B. Voloshin, “Secluded WIMP Dark Matter,” Phys. Lett. B 662 (2008) 53–61,
arXiv:0711.4866 [hep-ph].
[3415] J. L. Feng and J. Kumar, “The WIMPless Miracle: Dark-Matter Particles without Weak-Scale Masses or Weak
Interactions,” Phys. Rev. Lett. 101 (2008) 231301, arXiv:0803.4196 [hep-ph].
[3416] K. M. Zurek, “Dark Matter Candidates of a Very Low Mass,” Ann. Rev. Nucl. Part. Sci. 74 (2024) 287–319,
arXiv:2401.03025 [hep-ph].
[3417] R. Essig et al., “Snowmass2021 Cosmic Frontier: The landscape of low-threshold dark matter direct detection in the
next decade,” in Snowmass 2021. 2022. arXiv:2203.08297 [hep-ph].
[3418] G. Krnjaic et al., “A Snowmass Whitepaper: Dark Matter Production at Intensity-Frontier Experiments,”
arXiv:2207.00597 [hep-ph].
[3419] T. Yanagida, “Horizontal Symmetry and Mass of the Top Quark,” Phys. Rev. D 20 (1979) 2986.
[3420] T. Yanagida, “Horizontal Symmetry and Masses of Neutrinos,” Prog. Theor. Phys. 64 (1980) 1103.
[3421] R. N. Mohapatra and G. Senjanovic, “Neutrino Mass and Spontaneous Parity Nonconservation,” Phys. Rev. Lett. 44
(1980) 912.
[3422] J. Schechter and J. W. F. Valle, “Neutrino Masses in SU(2) x U(1) Theories,” Phys. Rev. D 22 (1980) 2227.
[3423] M. Fukugita and T. Yanagida, “Baryogenesis Without Grand Unification,” Phys. Lett. B 174 (1986) 45–47.
[3424] P. B. Pal and L. Wolfenstein, “Radiative Decays of Massive Neutrinos,” Phys. Rev. D 25 (1982) 766.
[3425] S. Dodelson and L. M. Widrow, “Sterile-neutrinos as dark matter,” Phys. Rev. Lett. 72 (1994) 17–20,
arXiv:hep-ph/9303287.
[3426] X.-D. Shi and G. M. Fuller, “A New dark matter candidate: Nonthermal sterile neutrinos,” Phys. Rev. Lett. 82 (1999)
2832–2835, arXiv:astro-ph/9810076.
[3427] R. S. L. Hansen and S. Vogl, “Thermalizing sterile neutrino dark matter,” Phys. Rev. Lett. 119 (2017) no. 25, 251305,
arXiv:1706.02707 [hep-ph].
[3428] A. De Gouvêa, M. Sen, W. Tangarife, and Y. Zhang, “Dodelson-Widrow Mechanism in the Presence of Self-Interacting
Neutrinos,” Phys. Rev. Lett. 124 (2020) no. 8, 081802, arXiv:1910.04901 [hep-ph].
[3429] T. Bringmann, P. F. Depta, M. Hufnagel, J. Kersten, J. T. Ruderman, and K. Schmidt-Hoberg, “Minimal sterile
neutrino dark matter,” Phys. Rev. D 107 (2023) no. 7, L071702, arXiv:2206.10630 [hep-ph].
[3430] F. Wilczek, “Problem of Strong P and T Invariance in the Presence of Instantons,” Phys. Rev. Lett. 40 (1978) 279–282.
[3431] S. Weinberg, “A New Light Boson?,” Phys. Rev. Lett. 40 (1978) 223–226.
[3432] R. D. Peccei and H. R. Quinn, “CP Conservation in the Presence of Instantons,” Phys. Rev. Lett. 38 (1977) 1440–1443.
[3433] J. E. Kim, “Weak Interaction Singlet and Strong CP Invariance,” Phys. Rev. Lett. 43 (1979) 103.
375

[3434] M. A. Shifman, A. I. Vainshtein, and V. I. Zakharov, “Can Confinement Ensure Natural CP Invariance of Strong
Interactions?,” Nucl. Phys. B 166 (1980) 493–506.
[3435] A. R. Zhitnitsky, “On Possible Suppression of the Axion Hadron Interactions. (In Russian),” Sov. J. Nucl. Phys. 31
(1980) 260.
[3436] M. Dine, W. Fischler, and M. Srednicki, “A Simple Solution to the Strong CP Problem with a Harmless Axion,” Phys.
Lett. B 104 (1981) 199–202.
[3437] J. Preskill, M. B. Wise, and F. Wilczek, “Cosmology of the Invisible Axion,” Phys. Lett. B 120 (1983) 127–132.
[3438] L. F. Abbott and P. Sikivie, “A Cosmological Bound on the Invisible Axion,” Phys. Lett. B 120 (1983) 133–136.
[3439] M. Dine and W. Fischler, “The Not So Harmless Axion,” Phys. Lett. B 120 (1983) 137–141.
[3440] A. D. Linde, “Axions in inflationary cosmology,” Phys. Lett. B 259 (1991) 38–47.
[3441] M. P. Hertzberg, M. Tegmark, and F. Wilczek, “Axion Cosmology and the Energy Scale of Inflation,” Phys. Rev. D 78
(2008) 083507, arXiv:0807.1726 [astro-ph].
[3442] L. Visinelli and P. Gondolo, “Dark Matter Axions Revisited,” Phys. Rev. D 80 (2009) 035024, arXiv:0903.4377
[astro-ph.CO].
[3443] G. Lazarides, R. K. Schaefer, D. Seckel, and Q. Shafi, “Dilution of Cosmological Axions by Entropy Production,” Nucl.
Phys. B 346 (1990) 193–212.
[3444] L. Visinelli and P. Gondolo, “Axion cold dark matter in non-standard cosmologies,” Phys. Rev. D 81 (2010) 063508,
arXiv:0912.0015 [astro-ph.CO].
[3445] L. Di Luzio, F. Mescia, and E. Nardi, “Redefining the Axion Window,” Phys. Rev. Lett. 118 (2017) no. 3, 031801,
arXiv:1610.07593 [hep-ph].
[3446] L. Di Luzio, F. Mescia, and E. Nardi, “Window for preferred axion models,” Phys. Rev. D 96 (2017) no. 7, 075003,
arXiv:1705.05370 [hep-ph].
[3447] S. M. Barr and D. Seckel, “Planck scale corrections to axion models,” Phys. Rev. D 46 (1992) 539–549.
[3448] L. Di Luzio, M. Giannotti, E. Nardi, and L. Visinelli, “The landscape of QCD axion models,” Phys. Rept. 870 (2020)
1–117, arXiv:2003.01100 [hep-ph].
[3449] I. G. Irastorza and J. Redondo, “New experimental approaches in the search for axion-like particles,” Prog. Part. Nucl.
Phys. 102 (2018) 89–159, arXiv:1801.08127 [hep-ph].
[3450] L. Amendola and R. Barbieri, “Dark matter from an ultra-light pseudo-Goldsone-boson,” Phys. Lett. B 642 (2006)
192–196, arXiv:hep-ph/0509257.
[3451] D. J. E. Marsh and J. Silk, “A Model For Halo Formation With Axion Mixed Dark Matter,” Mon. Not. Roy. Astron.
Soc. 437 (2014) no. 3, 2652–2663, arXiv:1307.1705 [astro-ph.CO].
[3452] D. J. E. Marsh, D. Grin, R. Hlozek, and P. G. Ferreira, “Axiverse cosmology and the energy scale of inflation,” Phys.
Rev. D 87 (2013) 121701, arXiv:1303.3008 [astro-ph.CO].
[3453] L. Hui, J. P. Ostriker, S. Tremaine, and E. Witten, “Ultralight scalars as cosmological dark matter,” Phys. Rev. D 95
(2017) no. 4, 043541, arXiv:1610.08297 [astro-ph.CO].
[3454] L. Visinelli, “Light axion-like dark matter must be present during inflation,” Phys. Rev. D 96 (2017) no. 2, 023013,
arXiv:1703.08798 [astro-ph.CO].
[3455] L. Visinelli and S. Vagnozzi, “Cosmological window onto the string axiverse and the supersymmetry breaking scale,”
Phys. Rev. D 99 (2019) no. 6, 063517, arXiv:1809.06382 [hep-ph].
[3456] H.-Y. Schive, M.-H. Liao, T.-P. Woo, S.-K. Wong, T. Chiueh, T. Broadhurst, and W. Y. P. Hwang, “Understanding
the Core-Halo Relation of Quantum Wave Dark Matter from 3D Simulations,” Phys. Rev. Lett. 113 (2014) no. 26,
261302, arXiv:1407.7762 [astro-ph.GA].
[3457] M. Nori and M. Baldi, “AX-GADGET: a new code for cosmological simulations of Fuzzy Dark Matter and Axion
models,” Mon. Not. Roy. Astron. Soc. 478 (2018) no. 3, 3935–3951, arXiv:1801.08144 [astro-ph.CO].
[3458] P. Mocz et al., “First star-forming structures in fuzzy cosmic filaments,” Phys. Rev. Lett. 123 (2019) no. 14, 141301,
arXiv:1910.01653 [astro-ph.GA].
[3459] J. Veltmaat, B. Schwabe, and J. C. Niemeyer, “Baryon-driven growth of solitonic cores in fuzzy dark matter halos,”
Phys. Rev. D 101 (2020) no. 8, 083518, arXiv:1911.09614 [astro-ph.CO].
[3460] S. May and V. Springel, “The halo mass function and filaments in full cosmological simulations with fuzzy dark
matter,” Mon. Not. Roy. Astron. Soc. 524 (2023) no. 3, 4256–4274, arXiv:2209.14886 [astro-ph.CO].
[3461] T. Zimmermann, J. Alvey, D. J. E. Marsh, M. Fairbairn, and J. I. Read, “Dwarf galaxies imply dark matter is heavier
than 2.2 × 10−21 eV,” arXiv:2405.20374 [astro-ph.CO].
[3462] I. De Martino, T. Broadhurst, S. H. Henry Tye, T. Chiueh, H.-Y. Schive, and R. Lazkoz, “Recognizing Axionic Dark
Matter by Compton and de Broglie Scale Modulation of Pulsar Timing,” Phys. Rev. Lett. 119 (2017) no. 22, 221103,
arXiv:1705.04367 [astro-ph.CO].
[3463] R. Hlozek, D. J. E. Marsh, and D. Grin, “Using the Full Power of the Cosmic Microwave Background to Probe Axion
Dark Matter,” Mon. Not. Roy. Astron. Soc. 476 (2018) no. 3, 3063–3085, arXiv:1708.05681 [astro-ph.CO].
[3464] A. Pozo, T. Broadhurst, I. De Martino, H. N. Luu, G. F. Smoot, J. Lim, and M. Neyrinck, “Wave dark matter and
ultra-diffuse galaxies,” Mon. Not. Roy. Astron. Soc. 504 (2021) no. 2, 2868–2876, arXiv:2003.08313 [astro-ph.GA].
[3465] J. H. H. Chan, H.-Y. Schive, S.-K. Wong, T. Chiueh, and T. Broadhurst, “Multiple Images and Flux Ratio Anomaly of
Fuzzy Gravitational Lenses,” Phys. Rev. Lett. 125 (2020) no. 11, 111102, arXiv:2002.10473 [astro-ph.GA].
[3466] N. Dalal and A. Kravtsov, “Excluding fuzzy dark matter with sizes and stellar kinematics of ultrafaint dwarf galaxies,”
Phys. Rev. D 106 (2022) no. 6, 063517, arXiv:2203.05750 [astro-ph.CO].
[3467] I. De Martino, “Constraining ultralight bosons in dwarf spheroidal galaxies with a radially varying anisotropy,” Phys.
376

Rev. D 108 (2023) no. 12, 123044, arXiv:2312.07217 [astro-ph.GA].


[3468] R. Della Monica and I. de Martino, “Bounding the mass of ultralight bosonic dark matter particles with the motion of
the S2 star around Sgr A*,” Phys. Rev. D 108 (2023) no. 10, L101303, arXiv:2305.10242 [gr-qc].
[3469] A. Burkert, “Fuzzy Dark Matter and Dark Matter Halo Cores,” Astrophys. J. 904 (2020) no. 2, 161,
arXiv:2006.11111 [astro-ph.GA].
[3470] P. Ullio, M. Kamionkowski, and P. Vogel, “Spin dependent WIMPs in DAMA?,” JHEP 07 (2001) 044,
arXiv:hep-ph/0010036.
[3471] XENON Collaboration, E. Aprile et al., “First Dark Matter Search with Nuclear Recoils from the XENONnT
Experiment,” Phys. Rev. Lett. 131 (2023) no. 4, 041003, arXiv:2303.14729 [hep-ex].
[3472] LZ Collaboration, J. Aalbers et al., “First Dark Matter Search Results from the LUX-ZEPLIN (LZ) Experiment,”
Phys. Rev. Lett. 131 (2023) no. 4, 041002, arXiv:2207.03764 [hep-ex].
[3473] PandaX Collaboration, Z. Huang et al., “Constraints on the axial-vector and pseudo-scalar mediated WIMP-nucleus
interactions from PandaX-4T experiment,” Phys. Lett. B 834 (2022) 137487, arXiv:2208.03626 [hep-ex].
[3474] DarkSide-50 Collaboration, P. Agnes et al., “Search for low-mass dark matter WIMPs with 12 ton-day exposure of
DarkSide-50,” Phys. Rev. D 107 (2023) no. 6, 063001, arXiv:2207.11966 [hep-ex].
[3475] DarkSide-20k Collaboration, C. E. Aalseth et al., “DarkSide-20k: A 20 tonne two-phase LAr TPC for direct dark
matter detection at LNGS,” Eur. Phys. J. Plus 133 (2018) 131, arXiv:1707.08145 [physics.ins-det].
[3476] ArDM Collaboration, J. Calvo et al., “Commissioning of the ArDM experiment at the Canfranc underground
laboratory: first steps towards a tonne-scale liquid argon time projection chamber for Dark Matter searches,” JCAP 03
(2017) 003, arXiv:1612.06375 [physics.ins-det].
[3477] P. Sikivie, “Experimental Tests of the Invisible Axion,” Phys. Rev. Lett. 51 (1983) 1415–1417. [Erratum:
Phys.Rev.Lett. 52, 695 (1984)].
[3478] P. Sikivie, “Detection Rates for ’Invisible’ Axion Searches,” Phys. Rev. D 32 (1985) 2988. [Erratum: Phys.Rev.D 36,
974 (1987)].
[3479] OSQAR Collaboration, P. Pugnat et al., “First results from the OSQAR photon regeneration experiment: No light
shining through a wall,” Phys. Rev. D 78 (2008) 092003, arXiv:0712.3362 [hep-ex].
[3480] ALPS Collaboration, K. Ehret et al., “Resonant laser power build-up in ALPS: A ’Light-shining-through-walls’
experiment,” Nucl. Instrum. Meth. A 612 (2009) 83–96, arXiv:0905.4159 [physics.ins-det].
[3481] R. Bähre et al., “Any light particle search II —Technical Design Report,” JINST 8 (2013) T09001, arXiv:1302.5647
[physics.ins-det].
[3482] OSQAR Collaboration, R. Ballou et al., “New exclusion limits on scalar and pseudoscalar axionlike particles from
light shining through a wall,” Phys. Rev. D 92 (2015) no. 9, 092002, arXiv:1506.08082 [hep-ex].
[3483] D. Horns, J. Jaeckel, A. Lindner, A. Lobanov, J. Redondo, and A. Ringwald, “Searching for WISPy Cold Dark Matter
with a Dish Antenna,” JCAP 04 (2013) 016, arXiv:1212.2970 [hep-ph].
[3484] P. Sikivie, N. Sullivan, and D. B. Tanner, “Proposal for Axion Dark Matter Detection Using an LC Circuit,” Phys.
Rev. Lett. 112 (2014) no. 13, 131301, arXiv:1310.8545 [hep-ph].
[3485] ADMX Collaboration, S. J. Asztalos et al., “Large scale microwave cavity search for dark matter axions,” Phys. Rev.
D 64 (2001) 092003.
[3486] B. M. Brubaker et al., “First results from a microwave cavity axion search at 24 µeV,” Phys. Rev. Lett. 118 (2017)
no. 6, 061302, arXiv:1610.02580 [astro-ph.CO].
[3487] R. Barbieri, C. Braggio, G. Carugno, C. S. Gallo, A. Lombardi, A. Ortolan, R. Pengo, G. Ruoso, and C. C. Speake,
“Searching for galactic axions through magnetized media: the QUAX proposal,” Phys. Dark Univ. 15 (2017) 135–141,
arXiv:1606.02201 [hep-ph].
[3488] D. Alesini et al., “Search for invisible axion dark matter of mass ma = 43 µeV with the QUAX–aγ experiment,” Phys.
Rev. D 103 (2021) no. 10, 102004, arXiv:2012.09498 [hep-ex].
[3489] C. M. Adair et al., “Search for Dark Matter Axions with CAST-CAPP,” Nature Commun. 13 (2022) no. 1, 6180,
arXiv:2211.02902 [hep-ex].
[3490] D. Alesini et al., “The future search for low-frequency axions and new physics with the FLASH resonant cavity
experiment at Frascati National Laboratories,” Phys. Dark Univ. 42 (2023) 101370, arXiv:2309.00351
[physics.ins-det].
[3491] S. Arguedas Cuendis et al., “The 3 Cavity Prototypes of RADES: An Axion Detector Using Microwave Filters at
CAST,” Springer Proc. Phys. 245 (2020) 45–51, arXiv:1903.04323 [physics.ins-det].
[3492] J. L. Ouellet et al., “First Results from ABRACADABRA-10 cm: A Search for Sub-µeV Axion Dark Matter,” Phys.
Rev. Lett. 122 (2019) no. 12, 121802, arXiv:1810.12257 [hep-ex].
[3493] M. A. Fedderke, P. W. Graham, D. F. J. Kimball, and S. Kalia, “Earth as a transducer for dark-photon dark-matter
detection,” Phys. Rev. D 104 (2021) no. 7, 075023, arXiv:2106.00022 [hep-ph].
[3494] CAST Collaboration, V. Anastassopoulos et al., “New CAST Limit on the Axion-Photon Interaction,” Nature Phys.
13 (2017) 584–590, arXiv:1705.02290 [hep-ex].
[3495] E. Armengaud et al., “Conceptual Design of the International Axion Observatory (IAXO),” JINST 9 (2014) T05002,
arXiv:1401.3233 [physics.ins-det].
[3496] J. E. Gunn, B. W. Lee, I. Lerche, D. N. Schramm, and G. Steigman, “Some Astrophysical Consequences of the
Existence of a Heavy Stable Neutral Lepton,” Astrophys. J. 223 (1978) 1015–1031.
[3497] F. W. Stecker, “The Cosmic Gamma-Ray Background from the Annihilation of Primordial Stable Neutral Heavy
Leptons,” Astrophys. J. 223 (1978) 1032–1036.
377

[3498] L. M. Krauss, K. Freese, W. Press, and D. Spergel, “Cold dark matter candidates and the solar neutrino problem,”
Astrophys. J. 299 (1985) 1001.
[3499] K. Freese, “Can Scalar Neutrinos Or Massive Dirac Neutrinos Be the Missing Mass?,” Phys. Lett. B 167 (1986)
295–300.
[3500] T. K. Gaisser, G. Steigman, and S. Tilav, “Limits on Cold Dark Matter Candidates from Deep Underground
Detectors,” Phys. Rev. D 34 (1986) 2206.
[3501] J. Silk and M. Srednicki, “Cosmic Ray anti-Protons as a Probe of a Photino Dominated Universe,” Phys. Rev. Lett. 53
(1984) 624.
[3502] F. W. Stecker, S. Rudaz, and T. F. Walsh, “Galactic Anti-protons From Photinos,” Phys. Rev. Lett. 55 (1985)
2622–2625.
[3503] J. R. Ellis, R. A. Flores, K. Freese, S. Ritz, D. Seckel, and J. Silk, “Cosmic Ray Constraints on the Annihilations of
Relic Particles in the Galactic Halo,” Phys. Lett. B 214 (1988) 403–412.
[3504] M. Cirelli, G. Corcella, A. Hektor, G. Hutsi, M. Kadastik, P. Panci, M. Raidal, F. Sala, and A. Strumia, “PPPC 4 DM
ID: A Poor Particle Physicist Cookbook for Dark Matter Indirect Detection,” JCAP 03 (2011) 051, arXiv:1012.4515
[hep-ph]. [Erratum: JCAP 10, E01 (2012)].
[3505] L. Pieri, J. Lavalle, G. Bertone, and E. Branchini, “Implications of High-Resolution Simulations on Indirect Dark
Matter Searches,” Phys. Rev. D 83 (2011) 023518, arXiv:0908.0195 [astro-ph.HE].
[3506] M. Ackermann et al., “Constraints on Dark Matter Annihilation in Clusters of Galaxies with the Fermi Large Area
Telescope,” JCAP 05 (2010) 025, arXiv:1002.2239 [astro-ph.CO].
[3507] Fermi-LAT Collaboration, M. Ackermann et al., “Searching for Dark Matter Annihilation from Milky Way Dwarf
Spheroidal Galaxies with Six Years of Fermi Large Area Telescope Data,” Phys. Rev. Lett. 115 (2015) no. 23, 231301,
arXiv:1503.02641 [astro-ph.HE].
[3508] M. Ajello et al., “The Origin of the Extragalactic Gamma-Ray Background and Implications for Dark-Matter
Annihilation,” Astrophys. J. Lett. 800 (2015) no. 2, L27, arXiv:1501.05301 [astro-ph.HE].
[3509] HAWC Collaboration, R. Alfaro et al., “Searching for TeV Dark Matter in Irregular Dwarf Galaxies with HAWC
Observatory,” Astrophys. J. 945 (2023) no. 1, 25, arXiv:2302.07929 [astro-ph.HE].
[3510] IceCube Collaboration, R. Abbasi et al., “Search for GeV-scale dark matter annihilation in the Sun with IceCube
DeepCore,” Phys. Rev. D 105 (2022) no. 6, 062004, arXiv:2111.09970 [astro-ph.HE].
[3511] IceCube Collaboration, R. Abbasi et al., “Search for neutrino lines from dark matter annihilation and decay with
IceCube,” Phys. Rev. D 108 (2023) no. 10, 102004, arXiv:2303.13663 [astro-ph.HE].
[3512] AMS Collaboration, M. Aguilar et al., “First Result from the Alpha Magnetic Spectrometer on the International
Space Station: Precision Measurement of the Positron Fraction in Primary Cosmic Rays of 0.5–350 GeV,” Phys. Rev.
Lett. 110 (2013) 141102.
[3513] Fermi-LAT Collaboration, M. Ackermann et al., “The Fermi Galactic Center GeV Excess and Implications for Dark
Matter,” Astrophys. J. 840 (2017) no. 1, 43, arXiv:1704.03910 [astro-ph.HE].
[3514] CTA Consortium Collaboration, M. Doro et al., “Dark Matter and Fundamental Physics with the Cherenkov
Telescope Array,” Astropart. Phys. 43 (2013) 189–214, arXiv:1208.5356 [astro-ph.IM].
[3515] CTA Collaboration, J. Carr et al., “Prospects for Indirect Dark Matter Searches with the Cherenkov Telescope Array
(CTA),” PoS ICRC2015 (2016) 1203, arXiv:1508.06128 [astro-ph.HE].
[3516] CTA Consortium Collaboration, B. S. Acharya et al., Science with the Cherenkov Telescope Array. WSP, 2018.
arXiv:1709.07997 [astro-ph.IM].
[3517] CTA Consortium Collaboration, A. Morselli, “Search for dark matter with IACTs and the Cherenkov Telescope
Array,” J. Phys. Conf. Ser. 2429 (2023) no. 1, 012019, arXiv:2302.11318 [astro-ph.HE].
[3518] CTAO Collaboration, S. Abe et al., “Dark matter line searches with the Cherenkov Telescope Array,” JCAP 07
(2024) 047, arXiv:2403.04857 [hep-ph].
[3519] D. Walsh, R. F. Carswell, and R. J. Weymann, “0957 + 561 A, B - Twin quasistellar objects or gravitational lens,”
Nature 279 (1979) 381–384.
[3520] M. V. Gorenstein, I. I. Shapiro, N. L. Cohen, B. E. Corey, E. E. Falco, J. M. Marcaide, A. E. E. Rogers, A. R.
Whitney, R. W. Porcas, R. A. Preston, and A. Rius, “Detection of a Compact Radio Source near the Center of a
Gravitational Lens: Quasar Image or Galactic Core?,” Science 219 (1983) no. 4580, 54–56.
[3521] A. Galan, G. Vernardos, Q. Minor, D. Sluse, L. Van de Vyvere, and M. Gomer, “Exploiting the diversity of modeling
methods to probe systematic biases in strong lensing analyses,” Astron. Astrophys. 692 (2024) A87, arXiv:2406.08484
[astro-ph.CO].
[3522] M. Meneghetti et al., “The Frontier Fields lens modelling comparison project,” Mon. Not. Roy. Astron. Soc. 472
(2017) no. 3, 3177–3216, arXiv:1606.04548 [astro-ph.CO].
[3523] M. Meneghetti, R. Argazzi, F. Pace, L. Moscardini, K. Dolag, M. Bartelmann, G. Li, and M. Oguri, “Arc sensitivity to
cluster ellipticity, asymmetries and substructures,” Astron. Astrophys. 461 (2007) 25–38, arXiv:astro-ph/0606006.
[3524] J. Wagner, “A model-independent characterisation of strong gravitational lensing by observables,” Universe 5 (2019)
177, arXiv:1906.05285 [astro-ph.CO].
[3525] R. E. Griffiths, M. Rudisel, J. Wagner, T. Hamilton, P.-C. Huang, and C. Villforth, “Hamilton’s Object – a clumpy
galaxy straddling the gravitational caustic of a galaxy cluster: constraints on dark matter clumping,” Mon. Not. Roy.
Astron. Soc. 506 (2021) no. 2, 1595–1608, arXiv:2105.04562 [astro-ph.CO].
[3526] J. Lin, J. Wagner, and R. E. Griffiths, “Generalized model-independent characterization of strong gravitational lenses
VIII. Automated multiband feature detection to constrain local lens properties,” Mon. Not. Roy. Astron. Soc. 517
378

(2022) no. 2, 1821–1836, arXiv:2207.01630 [astro-ph.CO].


[3527] M. Meneghetti et al., “A persistent excess of galaxy-galaxy strong lensing observed in galaxy clusters,” Astron.
Astrophys. 678 (2023) L2, arXiv:2309.05799 [astro-ph.CO].
[3528] S. Vegetti et al., “Strong Gravitational Lensing as a Probe of Dark Matter,” Space Sci. Rev. 220 (2024) no. 5, 58,
arXiv:2306.11781 [astro-ph.CO].
[3529] M. Castellano et al., “Constraints on photoionization feedback from number counts of ultra-faint high-redshift galaxies
in the Frontier Fields,” Astrophys. J. Lett. 823 (2016) no. 2, L40, arXiv:1605.01524 [astro-ph.GA].
[3530] B. Yue et al., “On the faint-end of the galaxy luminosity function in the Epoch of Reionization: updated constraints
from the HST Frontier Fields,” Astrophys. J. 868 (2018) no. 2, 115, arXiv:1711.05130 [astro-ph.GA].
[3531] P. Dayal, A. Mesinger, and F. Pacucci, “Early galaxy formation in warm dark matter cosmologies,” Astrophys. J. 806
(2015) no. 1, 67, arXiv:1408.1102 [astro-ph.GA].
[3532] N. Menci, A. Grazian, A. Lamastra, F. Calura, M. Castellano, and P. Santini, “Galaxy Formation in Sterile Neutrino
Dark Matter Models,” Astrophys. J. 854 (2018) no. 1, 1, arXiv:1801.03697 [astro-ph.CO].
[3533] N. Menci, A. Grazian, M. Castellano, and N. G. Sanchez, “A Stringent Limit on the Warm Dark Matter Particle
Masses from the Abundance of z=6 Galaxies in the Hubble Frontier Fields,” Astrophys. J. Lett. 825 (2016) no. 1, L1,
arXiv:1606.02530 [astro-ph.CO].
[3534] R. J. Bouwens et al., “UV Luminosity Functions at redshifts z ∼4 to z ∼10: 10000 Galaxies from HST Legacy Fields,”
Astrophys. J. 803 (2015) no. 1, 34, arXiv:1403.4295 [astro-ph.CO].
[3535] R. J. Bouwens, G. D. Illingworth, P. A. Oesch, J. Caruana, B. Holwerda, R. Smit, and S. Wilkins, “Reionization after
Planck: The Derived Growth of the Cosmic Ionizing Emissivity now matches the Growth of the Galaxy UV
Luminosity Density,” Astrophys. J. 811 (2015) no. 2, 140, arXiv:1503.08228 [astro-ph.CO].
[3536] B. E. Robertson, R. S. Ellis, S. R. Furlanetto, and J. S. Dunlop, “Cosmic Reionization and Early Star-forming
Galaxies: a Joint Analysis of new Constraints From Planck and the Hubble Space Telescope,” Astrophys. J. Lett. 802
(2015) no. 2, L19, arXiv:1502.02024 [astro-ph.CO].
[3537] S. L. Finkelstein, A. D’Aloisio, J.-P. Paardekooper, R. Ryan, P. Behroozi, K. Finlator, R. Livermore, P. R. U.
Sanderbeck, C. D. Vecchia, and S. Khochfar, “Conditions for Reionizing the Universe with A Low Galaxy Ionizing
Photon Escape Fraction,” Astrophys. J. 879 (2019) no. 1, 36, arXiv:1902.02792 [astro-ph.CO].
[3538] I. P. Carucci and P.-S. Corasaniti, “Cosmic Reionization History and Dark Matter Scenarios,” Phys. Rev. D 99 (2019)
no. 2, 023518, arXiv:1811.07904 [astro-ph.CO].
[3539] M. Romanello, N. Menci, and M. Castellano, “The Epoch of Reionization in Warm Dark Matter Scenarios,” Universe 7
(2021) no. 10, 365, arXiv:2110.05262 [astro-ph.CO].
[3540] P. S. Corasaniti, S. Agarwal, D. J. E. Marsh, and S. Das, “Constraints on dark matter scenarios from measurements of
the galaxy luminosity function at high redshifts,” Phys. Rev. D 95 (2017) no. 8, 083512, arXiv:1611.05892
[astro-ph.CO].
[3541] A. Rudakovskyi, A. Mesinger, D. Savchenko, and N. Gillet, “Constraints on warm dark matter from UV luminosity
functions of high-z galaxies with Bayesian model comparison,” Mon. Not. Roy. Astron. Soc. 507 (2021) no. 2,
3046–3056, arXiv:2104.04481 [astro-ph.CO].
[3542] A. Lapi, T. Ronconi, L. Boco, F. Shankar, N. Krachmalnicoff, C. Baccigalupi, and L. Danese, “Astroparticle
Constraints from Cosmic Reionization and Primordial Galaxy Formation,” Universe 8 (2022) no. 9, 476,
arXiv:2205.09474 [astro-ph.CO].
[3543] A. Lapi and L. Danese, “Cold or Warm? Constraining Dark Matter with Primeval Galaxies and Cosmic Reionization
after Planck,” JCAP 09 (2015) 003, arXiv:1508.02147 [astro-ph.CO].
[3544] A. Garzilli, A. Magalich, O. Ruchayskiy, and A. Boyarsky, “How to constrain warm dark matter with the Lyman-α
forest,” Mon. Not. Roy. Astron. Soc. 502 (2021) no. 2, 2356–2363, arXiv:1912.09397 [astro-ph.CO].
[3545] M. Drewes et al., “A White Paper on keV Sterile Neutrino Dark Matter,” JCAP 01 (2017) 025, arXiv:1602.04816
[hep-ph].
[3546] B. Dasgupta and J. Kopp, “Sterile Neutrinos,” Phys. Rept. 928 (2021) 1–63, arXiv:2106.05913 [hep-ph].
[3547] K. N. Abazajian, “Sterile neutrinos in cosmology,” Phys. Rept. 711-712 (2017) 1–28, arXiv:1705.01837 [hep-ph].
[3548] K. J. Kelly, M. Sen, W. Tangarife, and Y. Zhang, “Origin of sterile neutrino dark matter via secret neutrino
interactions with vector bosons,” Phys. Rev. D 101 (2020) no. 11, 115031, arXiv:2005.03681 [hep-ph].
[3549] M. D. Astros and S. Vogl, “Boosting the production of sterile neutrino dark matter with self-interactions,” JHEP 03
(2024) 032, arXiv:2307.15565 [hep-ph].
[3550] M. Shaposhnikov and I. Tkachev, “The nuMSM, inflation, and dark matter,” Phys. Lett. B 639 (2006) 414–417,
arXiv:hep-ph/0604236.
[3551] A. Kusenko, “Sterile neutrinos, dark matter, and the pulsar velocities in models with a Higgs singlet,” Phys. Rev. Lett.
97 (2006) 241301, arXiv:hep-ph/0609081.
[3552] K. Petraki and A. Kusenko, “Dark-matter sterile neutrinos in models with a gauge singlet in the Higgs sector,” Phys.
Rev. D 77 (2008) 065014, arXiv:0711.4646 [hep-ph].
[3553] F. Bezrukov, H. Hettmansperger, and M. Lindner, “keV sterile neutrino Dark Matter in gauge extensions of the
Standard Model,” Phys. Rev. D 81 (2010) 085032, arXiv:0912.4415 [hep-ph].
[3554] A. Kusenko, F. Takahashi, and T. T. Yanagida, “Dark Matter from Split Seesaw,” Phys. Lett. B 693 (2010) 144–148,
arXiv:1006.1731 [hep-ph].
[3555] J. A. Dror, D. Dunsky, L. J. Hall, and K. Harigaya, “Sterile Neutrino Dark Matter in Left-Right Theories,” JHEP 07
(2020) 168, arXiv:2004.09511 [hep-ph].
379

[3556] M. Heikinheimo, K. Huitu, V. Keus, and N. Koivunen, “Cosmological constraints on light flavons,” JHEP 06 (2019)
065, arXiv:1812.10963 [hep-ph].
[3557] S. May and V. Springel, “Structure formation in large-volume cosmological simulations of fuzzy dark matter: impact of
the non-linear dynamics,” Mon. Not. Roy. Astron. Soc. 506 (2021) no. 2, 2603–2618, arXiv:2101.01828
[astro-ph.CO].
[3558] A. Gough and C. Uhlemann, “When to interfere with dark matter? The impact of wave dynamics on statistics,” Open
J. Astrophys. 7 (2024) 2024, arXiv:2405.15852 [astro-ph.CO].
[3559] K. K. Rogers, R. Hložek, A. Laguë, M. M. Ivanov, O. H. E. Philcox, G. Cabass, K. Akitsu, and D. J. E. Marsh,
“Ultra-light axions and the S 8 tension: joint constraints from the cosmic microwave background and galaxy
clustering,” JCAP 06 (2023) 023, arXiv:2301.08361 [astro-ph.CO].
[3560] R. C. Pantig and A. Övgün, “Black Hole in Quantum Wave Dark Matter,” Fortsch. Phys. 71 (2023) no. 1, 2200164,
arXiv:2210.00523 [gr-qc].
[3561] V. Cardoso, T. Ikeda, R. Vicente, and M. Zilhão, “Parasitic black holes: The swallowing of a fuzzy dark matter
soliton,” Phys. Rev. D 106 (2022) no. 12, L121302, arXiv:2207.09469 [gr-qc].
[3562] G. Mustafa, S. K. Maurya, S. Ray, and F. Javed, “Construction of thin-shell around new wormhole solutions via
solitonic quantum wave dark matter,” Annals Phys. 460 (2024) 169551.
[3563] K. K. Boddy, V. Gluscevic, V. Poulin, E. D. Kovetz, M. Kamionkowski, and R. Barkana, “Critical assessment of CMB
limits on dark matter-baryon scattering: New treatment of the relative bulk velocity,” Phys. Rev. D 98 (2018) no. 12,
123506, arXiv:1808.00001 [astro-ph.CO].
[3564] Z. Li et al., “The Atacama Cosmology Telescope: limits on dark matter-baryon interactions from DR4 power spectra,”
JCAP 02 (2023) 046, arXiv:2208.08985 [astro-ph.CO].
[3565] R. Barkana, “Possible interaction between baryons and dark-matter particles revealed by the first stars,” Nature 555
(2018) no. 7694, 71–74, arXiv:1803.06698 [astro-ph.CO].
[3566] V. Gluscevic and K. K. Boddy, “Constraints on Scattering of keV–TeV Dark Matter with Protons in the Early
Universe,” Phys. Rev. Lett. 121 (2018) no. 8, 081301, arXiv:1712.07133 [astro-ph.CO].
[3567] M. Postolak, “Did the Big Bang and cosmic inflation really happen? (A tale of alternative cosmological models),”
arXiv:2404.18503 [physics.pop-ph].
[3568] J. Sakstein, H. Desmond, and B. Jain, “Screened Fifth Forces Mediated by Dark Matter–Baryon Interactions: Theory
and Astrophysical Probes,” Phys. Rev. D 100 (2019) no. 10, 104035, arXiv:1907.03775 [astro-ph.CO].
[3569] M. A. Buen-Abad, R. Essig, D. McKeen, and Y.-M. Zhong, “Cosmological constraints on dark matter interactions with
ordinary matter,” Phys. Rept. 961 (2022) 1–35, arXiv:2107.12377 [astro-ph.CO].
[3570] C. Dvorkin, K. Blum, and M. Kamionkowski, “Constraining Dark Matter-Baryon Scattering with Linear Cosmology,”
Phys. Rev. D 89 (2014) no. 2, 023519, arXiv:1311.2937 [astro-ph.CO].
[3571] B. Moore, “Evidence against dissipationless dark matter from observations of galaxy haloes,” Nature 370 (1994) 629.
[3572] S.-H. Oh, W. J. G. de Blok, F. Walter, E. Brinks, and R. C. Kennicutt, Jr, “High-resolution dark matter density
profiles of THINGS dwarf galaxies: Correcting for non-circular motions,” Astron. J. 136 (2008) 2761,
arXiv:0810.2119 [astro-ph].
[3573] J. Kormendy and K. C. Freeman, “Scaling laws for dark matter halos in late-type and dwarf spheroidal galaxies,” IAU
Symp. 220 (2004) 377, arXiv:astro-ph/0407321.
[3574] F. Donato and P. Salucci, “Cores of dark matter halos correlate with disk scale lengths,” Mon. Not. Roy. Astron. Soc.
353 (2004) L17–L22, arXiv:astro-ph/0403206.
[3575] F. Donato, G. Gentile, P. Salucci, C. F. Martins, M. I. Wilkinson, G. Gilmore, E. K. Grebel, A. Koch, and R. Wyse,
“A constant dark matter halo surface density in galaxies,” Mon. Not. Roy. Astron. Soc. 397 (2009) 1169–1176,
arXiv:0904.4054 [astro-ph.CO].
[3576] G. Gentile, B. Famaey, H. Zhao, and P. Salucci, “Universality of galactic surface densities within one dark halo
scale-length,” Nature 461 (2009) 627, arXiv:0909.5203 [astro-ph.CO].
[3577] C. Di Paolo, P. Salucci, and A. Erkurt, “The universal rotation curve of low surface brightness galaxies – IV. The
interrelation between dark and luminous matter,” Mon. Not. Roy. Astron. Soc. 490 (2019) no. 4, 5451–5477,
arXiv:1805.07165.
[3578] G. Sharma, P. Salucci, and G. van de Ven, “Observational evidence of evolving dark matter profiles at z ≤ 1,” Astron.
Astrophys. 659 (2022) A40, arXiv:2109.14224 [astro-ph.GA].
[3579] P. Salucci, N. Turini, and C. Di Paolo, “Paradigms and Scenarios for the Dark Matter Phenomenon,” Universe 6
(2020) no. 8, 118, arXiv:2008.04052 [astro-ph.CO].
[3580] Y. Shoji, E. Kuflik, Y. Birnboim, and N. C. Stone, “Heating galaxy clusters with interacting dark matter,” Mon. Not.
Roy. Astron. Soc. 528 (2024) no. 3, 4082–4091, arXiv:2306.08679 [astro-ph.CO].
[3581] G. Choi, T. T. Yanagida, and N. Yokozaki, “A model of interacting dark matter and dark radiation for H0 and σ8
tensions,” JHEP 01 (2021) 127, arXiv:2010.06892 [hep-ph].
[3582] M. G. Yengejeh, S. Fakhry, J. T. Firouzjaee, and H. Fathi, “The integrated Sachs–Wolfe effect in interacting dark
matter–dark energy models,” Phys. Dark Univ. 39 (2023) 101144, arXiv:2206.01030 [astro-ph.CO].
[3583] Y. Wu, S. Baum, K. Freese, L. Visinelli, and H.-B. Yu, “Dark stars powered by self-interacting dark matter,” Phys.
Rev. D 106 (2022) no. 4, 043028, arXiv:2205.10904 [hep-ph].
[3584] N. F. Bell, G. Busoni, M. E. Ramirez-Quezada, S. Robles, and M. Virgato, “Improved treatment of dark matter
capture in white dwarfs,” JCAP 10 (2021) 083, arXiv:2104.14367 [hep-ph].
[3585] K.-L. Leung, M.-c. Chu, and L.-M. Lin, “Tidal deformability of dark matter admixed neutron stars,” Phys. Rev. D 105
380

(2022) no. 12, 123010, arXiv:2207.02433 [astro-ph.HE].


[3586] O. Akarsu, N. Katirci, S. Kumar, R. C. Nunes, and M. Sami, “Cosmological implications of scale-independent
energy-momentum squared gravity: Pseudo nonminimal interactions in dark matter and relativistic relics,” Phys. Rev.
D 98 (2018) no. 6, 063522, arXiv:1807.01588 [gr-qc].
[3587] Z. Zhou, G. Liu, Y. Mu, and L. Xu, “Limit on the dark matter mass from its interaction with photons,” Phys. Rev. D
105 (2022) no. 10, 103509, arXiv:2205.08070 [astro-ph.CO].
[3588] J. Stadler and C. Bœhm, “Constraints on γ-CDM interactions matching the Planck data precision,” JCAP 10 (2018)
009, arXiv:1802.06589 [astro-ph.CO].
[3589] N. Becker, D. C. Hooper, F. Kahlhoefer, J. Lesgourgues, and N. Schöneberg, “Cosmological constraints on
multi-interacting dark matter,” JCAP 02 (2021) 019, arXiv:2010.04074 [astro-ph.CO].
[3590] Y. Ali-Haïmoud, “Testing dark matter interactions with CMB spectral distortions,” Phys. Rev. D 103 (2021) no. 4,
043541, arXiv:2101.04070 [astro-ph.CO].
[3591] Z. Xu, X. Hou, X. Gong, and J. Wang, “Black Hole Space-time In Dark Matter Halo,” JCAP 09 (2018) 038,
arXiv:1803.00767 [gr-qc].
[3592] R. A. Konoplya and A. Zhidenko, “Solutions of the Einstein Equations for a Black Hole Surrounded by a Galactic
Halo,” Astrophys. J. 933 (2022) no. 2, 166, arXiv:2202.02205 [gr-qc].
[3593] X. Hou, Z. Xu, and J. Wang, “Rotating Black Hole Shadow in Perfect Fluid Dark Matter,” JCAP 12 (2018) 040,
arXiv:1810.06381 [gr-qc].
[3594] X. Hou, Z. Xu, M. Zhou, and J. Wang, “Black hole shadow of Sgr A∗ in dark matter halo,” JCAP 07 (2018) 015,
arXiv:1804.08110 [gr-qc].
[3595] S. Haroon, M. Jamil, K. Jusufi, K. Lin, and R. B. Mann, “Shadow and Deflection Angle of Rotating Black Holes in
Perfect Fluid Dark Matter with a Cosmological Constant,” Phys. Rev. D 99 (2019) no. 4, 044015, arXiv:1810.04103
[gr-qc].
[3596] Z. Xu, X. Gong, and S.-N. Zhang, “Black hole immersed dark matter halo,” Phys. Rev. D 101 (2020) no. 2, 024029.
[3597] K. Jusufi, M. Jamil, and T. Zhu, “Shadows of Sgr A∗ black hole surrounded by superfluid dark matter halo,” Eur.
Phys. J. C 80 (2020) no. 5, 354, arXiv:2005.05299 [gr-qc].
[3598] Z. Xu, J. Wang, and M. Tang, “Deformed black hole immersed in dark matter spike,” JCAP 09 (2021) 007,
arXiv:2104.13158 [gr-qc].
[3599] S. Nampalliwar, S. Kumar, K. Jusufi, Q. Wu, M. Jamil, and P. Salucci, “Modeling the Sgr A* Black Hole Immersed in
a Dark Matter Spike,” Astrophys. J. 916 (2021) no. 2, 116, arXiv:2103.12439 [astro-ph.HE].
[3600] K. Jusufi and Saurabh, “Black hole shadows in Verlinde’s emergent gravity,” Mon. Not. Roy. Astron. Soc. 503 (2021)
no. 1, 1310–1318, arXiv:2010.15870 [gr-qc].
[3601] R. A. Konoplya, “Black holes in galactic centers: Quasinormal ringing, grey-body factors and Unruh temperature,”
Phys. Lett. B 823 (2021) 136734, arXiv:2109.01640 [gr-qc].
[3602] K. Saurabh and K. Jusufi, “Imprints of dark matter on black hole shadows using spherical accretions,” Eur. Phys. J. C
81 (2021) no. 6, 490, arXiv:2009.10599 [gr-qc].
[3603] R. C. Pantig and A. Övgün, “Dark matter effect on the weak deflection angle by black holes at the center of Milky
Way and M87 galaxies,” Eur. Phys. J. C 82 (2022) no. 5, 391, arXiv:2201.03365 [gr-qc].
[3604] R. C. Pantig and A. Övgün, “Dehnen halo effect on a black hole in an ultra-faint dwarf galaxy,” JCAP 08 (2022)
no. 08, 056, arXiv:2202.07404 [astro-ph.GA].
[3605] F. Atamurotov, U. Papnoi, and K. Jusufi, “Shadow and deflection angle of charged rotating black hole surrounded by
perfect fluid dark matter,” Class. Quant. Grav. 39 (2022) no. 2, 025014, arXiv:2104.14898 [gr-qc].
[3606] K. Jusufi, “Black holes surrounded by Einstein clusters as models of dark matter fluid,” Eur. Phys. J. C 83 (2023)
no. 2, 103, arXiv:2202.00010 [gr-qc].
[3607] R. C. Pantig, P. K. Yu, E. T. Rodulfo, and A. Övgün, “Shadow and weak deflection angle of extended uncertainty
principle black hole surrounded with dark matter,” Annals Phys. 436 (2022) 168722, arXiv:2104.04304 [gr-qc].
[3608] D. Liu, Y. Yang, A. Övgün, Z.-W. Long, and Z. Xu, “Gravitational ringing and superradiant instabilities of the
Kerr-like black holes in a dark matter halo,” Eur. Phys. J. C 83 (2023) no. 7, 565, arXiv:2204.11563 [gr-qc].
[3609] A. Anjum, M. Afrin, and S. G. Ghosh, “Investigating effects of dark matter on photon orbits and black hole shadows,”
Phys. Dark Univ. 40 (2023) 101195, arXiv:2301.06373 [gr-qc].
[3610] A. Övgün, L. J. F. Sese, and R. C. Pantig, “Constraints via the Event Horizon Telescope for Black Hole Solutions with
Dark Matter under the Generalized Uncertainty Principle Minimal Length Scale Effect,” Annalen Phys. 536 (2024)
no. 4, 2300390, arXiv:2309.07442 [gr-qc].
[3611] A. Errehymy, S. K. Maurya, G. Mustafa, S. Hansraj, H. I. Alrebdi, and A.-H. Abdel-Aty, “Black Hole Solutions with
Dark Matter Halos in the Four-Dimensional Einstein-Gauss-Bonnet Gravity,” Fortsch. Phys. 71 (2023) no. 10-11,
2300052.
[3612] C.-K. Qiao and M. Zhou, “Gravitational lensing of Schwarzschild and charged black holes immersed in perfect fluid
dark matter halo,” JCAP 12 (2023) 005, arXiv:2212.13311 [gr-qc].
[3613] X. Zhou, Y. Xue, B. Mu, and J. Tao, “Temporal and spatial chaos of RN-AdS black holes immersed in Perfect Fluid
Dark Matter,” Phys. Dark Univ. 39 (2023) 101168, arXiv:2209.03612 [gr-qc].
[3614] S. Capozziello, S. Zare, D. F. Mota, and H. Hassanabadi, “Dark matter spike around Bumblebee black holes,” JCAP
2023 (2023) no. 2305, 027, arXiv:2303.13554 [gr-qc].
[3615] S. Capozziello, S. Zare, and H. Hassanabadi, “Testing bumblebee gravity with global monopoles in a dark matter spike
by EHT observations from M87 and Sgr A,” arXiv:2311.12896 [gr-qc].
381

[3616] Y.-G. Liu, C.-K. Qiao, and J. Tao, “Gravitational lensing of spherically symmetric black holes in dark matter halos,”
JCAP 10 (2024) 075, arXiv:2312.15760 [gr-qc].
[3617] Y. Yang, D. Liu, A. Övgün, G. Lambiase, and Z.-W. Long, “Black hole surrounded by the pseudo-isothermal dark
matter halo,” Eur. Phys. J. C 84 (2024) no. 1, 63, arXiv:2308.05544 [gr-qc].
[3618] G. Gómez and P. Valageas, “Constraining self-interacting scalar field dark matter from the black hole shadow of the
Event Horizon Telescope,” Phys. Rev. D 109 (2024) no. 10, 103038, arXiv:2403.08988 [astro-ph.CO].
[3619] C.-K. Qiao and P. Su, “Time delay of light in the gravitational lensing of supermassive black holes in dark matter
halos,” Eur. Phys. J. C 84 (2024) no. 10, 1032, arXiv:2403.05682 [gr-qc].
[3620] C. F. B. Macedo, J. a. L. Rosa, and D. Rubiera-Garcia, “Optical appearance of black holes surrounded by a dark
matter halo,” JCAP 07 (2024) 046, arXiv:2402.13047 [gr-qc].
[3621] S. R. Wu, B. Q. Wang, Z. W. Long, and H. Chen, “Rotating black holes surrounded by a dark matter halo in the
galactic center of M87 and Sgr A∗,” Phys. Dark Univ. 44 (2024) 101455.
[3622] R. C. Pantig, “Apparent and emergent dark matter around a Schwarzschild black hole,” Phys. Dark Univ. 45 (2024)
101550, arXiv:2405.07531 [gr-qc].
[3623] R. A. Konoplya, “Shadow of a black hole surrounded by dark matter,” Phys. Lett. B 795 (2019) 1–6,
arXiv:1905.00064 [gr-qc].
[3624] M. A. Buen-Abad, G. Marques-Tavares, and M. Schmaltz, “Non-Abelian dark matter and dark radiation,” Phys. Rev.
D 92 (2015) no. 2, 023531, arXiv:1505.03542 [hep-ph].
[3625] J. Lesgourgues, G. Marques-Tavares, and M. Schmaltz, “Evidence for dark matter interactions in cosmological
precision data?,” JCAP 02 (2016) 037, arXiv:1507.04351 [astro-ph.CO].
[3626] Z. Chacko, Y. Cui, S. Hong, T. Okui, and Y. Tsai, “Partially Acoustic Dark Matter, Interacting Dark Radiation, and
Large Scale Structure,” JHEP 12 (2016) 108, arXiv:1609.03569 [astro-ph.CO].
[3627] M. A. Buen-Abad, M. Schmaltz, J. Lesgourgues, and T. Brinckmann, “Interacting Dark Sector and Precision
Cosmology,” JCAP 01 (2018) 008, arXiv:1708.09406 [astro-ph.CO].
[3628] M. A. Buen-Abad, R. Emami, and M. Schmaltz, “Cannibal Dark Matter and Large Scale Structure,” Phys. Rev. D 98
(2018) no. 8, 083517, arXiv:1803.08062 [hep-ph].
[3629] M. A. Buen-Abad, Z. Chacko, C. Kilic, G. Marques-Tavares, and T. Youn, “Stepped partially acoustic dark matter,
large scale structure, and the Hubble tension,” JHEP 06 (2023) 012, arXiv:2208.05984 [hep-ph].
[3630] N. Schöneberg, G. Franco Abellán, T. Simon, A. Bartlett, Y. Patel, and T. L. Smith, “Comparative analysis of
interacting stepped dark radiation,” Phys. Rev. D 108 (2023) no. 12, 123513, arXiv:2306.12469 [astro-ph.CO].
[3631] M. A. Buen-Abad, Z. Chacko, C. Kilic, G. Marques-Tavares, and T. Youn, “Stepped partially acoustic dark matter:
likelihood analysis and cosmological tensions,” JCAP 11 (2023) 005, arXiv:2306.01844 [astro-ph.CO].
[3632] L. G. van den Aarssen, T. Bringmann, and C. Pfrommer, “Is dark matter with long-range interactions a solution to all
small-scale problems of ΛCDM cosmology?,” Phys. Rev. Lett. 109 (2012) 231301, arXiv:1205.5809 [astro-ph.CO].
[3633] T. Bringmann, J. Hasenkamp, and J. Kersten, “Tight bonds between sterile neutrinos and dark matter,” JCAP 07
(2014) 042, arXiv:1312.4947 [hep-ph].
[3634] T. Bringmann, H. T. Ihle, J. Kersten, and P. Walia, “Suppressing structure formation at dwarf galaxy scales and
below: late kinetic decoupling as a compelling alternative to warm dark matter,” Phys. Rev. D 94 (2016) no. 10,
103529, arXiv:1603.04884 [hep-ph].
[3635] G. Poulot, E. M. Teixeira, C. van de Bruck, and N. J. Nunes, “Scalar field dark matter with time-varying equation of
state,” arXiv:2404.10524 [astro-ph.CO].
[3636] C. van de Bruck, G. Poulot, and E. M. Teixeira, “Scalar field dark matter and dark energy: a hybrid model for the
dark sector,” JCAP 07 (2023) 019, arXiv:2211.13653 [hep-th].
[3637] E. M. Teixeira, G. Poulot, C. van de Bruck, E. Di Valentino, and V. Poulin, “Alleviating cosmological tensions with a
hybrid dark sector,” arXiv:2412.14139 [astro-ph.CO].
[3638] A. Aboubrahim and P. Nath, “Interacting ultralight dark matter and dark energy and fits to cosmological data in a
field theory approach,” JCAP 09 (2024) 076, arXiv:2406.19284 [astro-ph.CO].
[3639] BOSS Collaboration, C. P. Ahn et al., “The Ninth Data Release of the Sloan Digital Sky Survey: First Spectroscopic
Data from the SDSS-III Baryon Oscillation Spectroscopic Survey,” Astrophys. J. Suppl. 203 (2012) 21,
arXiv:1207.7137 [astro-ph.IM].
[3640] C. Howlett, A. Ross, L. Samushia, W. Percival, and M. Manera, “The clustering of the SDSS main galaxy sample – II.
Mock galaxy catalogues and a measurement of the growth of structure from redshift space distortions at z = 0.15,”
Mon. Not. Roy. Astron. Soc. 449 (2015) no. 1, 848–866, arXiv:1409.3238 [astro-ph.CO].
[3641] J. L. Menestrina and R. J. Scherrer, “Dark Radiation from Particle Decays during Big Bang Nucleosynthesis,” Phys.
Rev. D 85 (2012) 047301, arXiv:1111.0605 [astro-ph.CO].
[3642] M. C. Gonzalez-Garcia, V. Niro, and J. Salvado, “Dark Radiation and Decaying Matter,” JHEP 04 (2013) 052,
arXiv:1212.1472 [hep-ph].
[3643] Z. Berezhiani, A. D. Dolgov, and I. I. Tkachev, “Reconciling Planck results with low redshift astronomical
measurements,” Phys. Rev. D 92 (2015) no. 6, 061303, arXiv:1505.03644 [astro-ph.CO].
[3644] K. Vattis, S. M. Koushiappas, and A. Loeb, “Dark matter decaying in the late Universe can relieve the H0 tension,”
Phys. Rev. D 99 (2019) no. 12, 121302, arXiv:1903.06220 [astro-ph.CO].
[3645] K. Enqvist, S. Nadathur, T. Sekiguchi, and T. Takahashi, “Decaying dark matter and the tension in σ8 ,” JCAP 09
(2015) 067, arXiv:1505.05511 [astro-ph.CO].
[3646] G. Franco Abellán, R. Murgia, V. Poulin, and J. Lavalle, “Implications of the S8 tension for decaying dark matter with
382

warm decay products,” Phys. Rev. D 105 (2022) no. 6, 063525, arXiv:2008.09615 [astro-ph.CO].
[3647] G. Franco Abellán, R. Murgia, and V. Poulin, “Linear cosmological constraints on two-body decaying dark matter
scenarios and the S8 tension,” Phys. Rev. D 104 (2021) no. 12, 123533, arXiv:2102.12498 [astro-ph.CO].
[3648] W. Liu, L. A. Anchordoqui, E. Di Valentino, S. Pan, Y. Wu, and W. Yang, “Constraints from high-precision
measurements of the cosmic microwave background: the case of disintegrating dark matter with Λ or dynamical dark
energy,” JCAP 02 (2022) no. 02, 012, arXiv:2108.04188 [astro-ph.CO].
[3649] L. A. Anchordoqui, V. Barger, H. Goldberg, X. Huang, D. Marfatia, L. H. M. da Silva, and T. J. Weiler, “IceCube
neutrinos, decaying dark matter, and the Hubble constant,” Phys. Rev. D 92 (2015) no. 6, 061301, arXiv:1506.08788
[hep-ph]. [Erratum: Phys.Rev.D 94, 069901 (2016)].
[3650] L. A. Anchordoqui, V. Barger, D. Marfatia, M. H. Reno, and T. J. Weiler, “Oscillations of sterile neutrinos from dark
matter decay eliminates the IceCube-Fermi tension,” Phys. Rev. D 103 (2021) no. 7, 075022, arXiv:2101.09559
[astro-ph.HE].
[3651] A. Chudaykin, D. Gorbunov, and I. Tkachev, “Dark matter component decaying after recombination: Lensing
constraints with Planck data,” Phys. Rev. D 94 (2016) 023528, arXiv:1602.08121 [astro-ph.CO].
[3652] V. Poulin, P. D. Serpico, and J. Lesgourgues, “A fresh look at linear cosmological constraints on a decaying dark
matter component,” JCAP 08 (2016) 036, arXiv:1606.02073 [astro-ph.CO].
[3653] S. J. Clark, K. Vattis, and S. M. Koushiappas, “Cosmological constraints on late-universe decaying dark matter as a
solution to the H0 tension,” Phys. Rev. D 103 (2021) no. 4, 043014, arXiv:2006.03678 [astro-ph.CO].
[3654] A. Chudaykin, D. Gorbunov, and I. Tkachev, “Dark matter component decaying after recombination: Sensitivity to
baryon acoustic oscillation and redshift space distortion probes,” Phys. Rev. D 97 (2018) no. 8, 083508,
arXiv:1711.06738 [astro-ph.CO].
[3655] A. Nygaard, T. Tram, and S. Hannestad, “Updated constraints on decaying cold dark matter,” JCAP 05 (2021) 017,
arXiv:2011.01632 [astro-ph.CO].
[3656] L. A. Anchordoqui, “Decaying dark matter, the H0 tension, and the lithium problem,” Phys. Rev. D 103 (2021) no. 3,
035025, arXiv:2010.09715 [hep-ph].
[3657] Z. Davari and N. Khosravi, “Can decaying dark matter scenarios alleviate both H0 and σ8 tensions?,” Mon. Not. Roy.
Astron. Soc. 516 (2022) no. 3, 4373–4382, arXiv:2203.09439 [astro-ph.CO].
[3658] T. Simon, G. Franco Abellán, P. Du, V. Poulin, and Y. Tsai, “Constraining decaying dark matter with BOSS data and
the effective field theory of large-scale structures,” Phys. Rev. D 106 (2022) no. 2, 023516, arXiv:2203.07440
[astro-ph.CO].
[3659] J. Bucko, S. K. Giri, F. H. Peters, and A. Schneider, “Probing the two-body decaying dark matter scenario with weak
lensing and the cosmic microwave background,” Astron. Astrophys. 683 (2024) A152, arXiv:2307.03222
[astro-ph.CO].
[3660] K. Sigurdson and M. Kamionkowski, “Charged - particle decay and suppression of small - scale power,” Phys. Rev.
Lett. 92 (2004) 171302, arXiv:astro-ph/0311486.
[3661] J. A. R. Cembranos, J. L. Feng, A. Rajaraman, and F. Takayama, “SuperWIMP solutions to small scale structure
problems,” Phys. Rev. Lett. 95 (2005) 181301, arXiv:hep-ph/0507150.
[3662] M. Kaplinghat, “Dark matter from early decays,” Phys. Rev. D 72 (2005) 063510, arXiv:astro-ph/0507300.
[3663] L. E. Strigari, M. Kaplinghat, and J. S. Bullock, “Dark Matter Halos with Cores from Hierarchical Structure
Formation,” Phys. Rev. D 75 (2007) 061303, arXiv:astro-ph/0606281.
[3664] J. A. R. Cembranos, J. L. Feng, and L. E. Strigari, “Resolving Cosmic Gamma Ray Anomalies with Dark Matter
Decaying Now,” Phys. Rev. Lett. 99 (2007) 191301, arXiv:0704.1658 [astro-ph].
[3665] K. R. Dienes and B. Thomas, “Dynamical Dark Matter: I. Theoretical Overview,” Phys. Rev. D 85 (2012) 083523,
arXiv:1106.4546 [hep-ph].
[3666] K. R. Dienes and B. Thomas, “Dynamical Dark Matter: II. An Explicit Model,” Phys. Rev. D 85 (2012) 083524,
arXiv:1107.0721 [hep-ph].
[3667] K. R. Dienes, F. Huang, J. Kost, S. Su, and B. Thomas, “Deciphering the archaeological record: Cosmological imprints
of nonminimal dark sectors,” Phys. Rev. D 101 (2020) no. 12, 123511, arXiv:2001.02193 [astro-ph.CO].
[3668] L. A. Anchordoqui, I. Antoniadis, and D. Lust, “Aspects of the dark dimension in cosmology,” Phys. Rev. D 107
(2023) no. 8, 083530, arXiv:2212.08527 [hep-ph].
[3669] G. Obied, C. Dvorkin, E. Gonzalo, and C. Vafa, “Dark dimension and decaying dark matter gravitons,” Phys. Rev. D
109 (2024) no. 6, 063540, arXiv:2311.05318 [astro-ph.CO].
[3670] C. Vafa, “The String landscape and the swampland,” arXiv:hep-th/0509212.
[3671] E. Gonzalo, M. Montero, G. Obied, and C. Vafa, “Dark dimension gravitons as dark matter,” JHEP 11 (2023) 109,
arXiv:2209.09249 [hep-ph].
[3672] A. Desai, K. R. Dienes, and B. Thomas, “Constraining Dark-Matter Ensembles with Supernova Data,” Phys. Rev. D
101 (2020) no. 3, 035031, arXiv:1909.07981 [astro-ph.CO].
[3673] E. N. Saridakis, “Do we need soft cosmology?,” Phys. Lett. B 822 (2021) 136649, arXiv:2105.08646 [astro-ph.CO].
[3674] E. N. Saridakis, W. Yang, S. Pan, F. K. Anagnostopoulos, and S. Basilakos, “Observational constraints on soft dark
energy and soft dark matter: Challenging ΛCDM cosmology,” Nucl. Phys. B 986 (2023) 116042, arXiv:2112.08330
[astro-ph.CO].
[3675] L. M. C. Sagis, “Dynamic properties of interfaces in soft matter: Experiments and theory,” Reviews of Modern Physics
83 (2011) no. 4, 1367–1403.
[3676] Z. Davari, A. Ashoorioon, and K. Rezazadeh, “Spherical collapse approach for non-standard dark matter models and
383

enhanced early galaxy formation in JWST,” Mon. Not. Roy. Astron. Soc. 534 (2024) no. 3, 2848–2857,
arXiv:2311.15083 [astro-ph.CO].
[3677] V. Poulin, J. Lesgourgues, and P. D. Serpico, “Cosmological constraints on exotic injection of electromagnetic energy,”
JCAP 03 (2017) 043, arXiv:1610.10051 [astro-ph.CO].
[3678] T. R. Slatyer and C.-L. Wu, “General Constraints on Dark Matter Decay from the Cosmic Microwave Background,”
Phys. Rev. D 95 (2017) no. 2, 023010, arXiv:1610.06933 [astro-ph.CO].
[3679] S. Colafrancesco, M. Regis, P. Marchegiani, G. Beck, R. Beck, H. Zechlin, A. Lobanov, and D. Horns, “Probing the
nature of Dark Matter with the SKA,” PoS AASKA14 (2015) 100, arXiv:1502.03738 [astro-ph.HE].
[3680] K. Dutta, A. Ghosh, A. Kar, and B. Mukhopadhyaya, “A general study of decaying scalar dark matter: existing limits
and projected radio signals at the SKA,” JCAP 09 (2022) 005, arXiv:2204.06024 [hep-ph].
[3681] S. F. King, R. Roshan, X. Wang, G. White, and M. Yamazaki, “Quantum gravity effects on dark matter and
gravitational waves,” Phys. Rev. D 109 (2024) no. 2, 024057, arXiv:2308.03724 [hep-ph].
[3682] S. F. King, R. Roshan, X. Wang, G. White, and M. Yamazaki, “Quantum gravity effects on fermionic dark matter and
gravitational waves,” JCAP 05 (2024) 071, arXiv:2311.12487 [hep-ph].
[3683] T. Asaka and M. Shaposhnikov, “The νMSM, dark matter and baryon asymmetry of the universe,” Phys. Lett. B 620
(2005) 17–26, arXiv:hep-ph/0505013.
[3684] A. Datta, R. Roshan, and A. Sil, “Imprint of the Seesaw Mechanism on Feebly Interacting Dark Matter and the
Baryon Asymmetry,” Phys. Rev. Lett. 127 (2021) no. 23, 231801, arXiv:2104.02030 [hep-ph].
[3685] I. H. Redmount, “Dynamics of a void-dominated universe: cell-lattice models.,” MNRAS 235 (1988) 1301–1312.
[3686] E. Yusofi, M. Khanpour, B. Khanpour, M. A. Ramzanpour, and M. Mohsenzadeh, “Surface tension of cosmic voids as
a possible source for dark energy,” Mon. Not. Roy. Astron. Soc. 511 (2022) no. 1, L82–L86, arXiv:1907.12418
[astro-ph.CO].
[3687] S. Mohammadi, E. Yusofi, M. Mohsenzadeh, and M. K. Salem, “A possible role for the merger of clusters/voids in the
cosmological expansion,” Mon. Not. Roy. Astron. Soc. 525 (2023) no. 3, 3274–3280, arXiv:2309.07826
[astro-ph.CO].
[3688] H. Moshafi, A. Talebian, E. Yusofi, and E. Di Valentino, “Observational constraints on the dark energy with a
quadratic equation of state,” Phys. Dark Univ. 45 (2024) 101524, arXiv:2403.02000 [astro-ph.CO].
[3689] S. Ahmadi, E. Yusofi, and M. A. Ramzanpour, “Incorporating the cosmological constant in a modified uncertainty
principle,” Mod. Phys. Lett. A 39 (2024) no. 27n28, 2450125, arXiv:2401.16126 [gr-qc].
[3690] A. Shahriar, M. Abbasiyan-Motlaq, M. Mohsenzadeh, and E. Yusofi, “Hubble Expansion and Entropy Rates in a
Cosmological Model with Merging Clusters and Voids,” arXiv:2412.05917 [astro-ph.CO].
[3691] R. van de Weygaert and E. Platen, “Cosmic Voids: structure, dynamics and galaxies,” Int. J. Mod. Phys. Conf. Ser.
01 (2011) 41–66, arXiv:0912.2997 [astro-ph.CO].
[3692] H.-Y. Wu and D. Huterer, “Sample variance in the local measurements of the Hubble constant,” Mon. Not. Roy.
Astron. Soc. 471 (2017) no. 4, 4946–4955, arXiv:1706.09723 [astro-ph.CO].
[3693] D. Camarena and V. Marra, “Impact of the cosmic variance on H0 on cosmological analyses,” Phys. Rev. D 98 (2018)
no. 2, 023537, arXiv:1805.09900 [astro-ph.CO].
[3694] R. C. Keenan, A. J. Barger, and L. L. Cowie, “Evidence for a ~300 Megaparsec Scale Under-density in the Local
Galaxy Distribution,” Astrophys. J. 775 (2013) 62, arXiv:1304.2884 [astro-ph.CO].
[3695] H. Böhringer, G. Chon, M. Bristow, and C. A. Collins, “The extended ROSAT-ESO Flux-Limited X-ray Galaxy
Cluster Survey (REFLEX II): V. Exploring a local underdensity in the Southern Sky,” Astron. Astrophys. 574 (2015)
A26, arXiv:1410.2172 [astro-ph.CO].
[3696] H. Böhringer, G. Chon, and C. A. Collins, “Observational evidence for a local underdensity in the Universe and its
effect on the measurement of the Hubble Constant,” Astron. Astrophys. 633 (2020) A19, arXiv:1907.12402
[astro-ph.CO].
[3697] S. J. Maddox, G. Efstathiou, W. J. Sutherland, and J. Loveday, “Galaxy correlations on large scales.,” MNRAS 242
(1990) 43.
[3698] T. Shanks, “Galaxy Count Models and the Extragalactic Background Light,” in The Galactic and Extragalactic
Background Radiation, S. Bowyer and C. Leinert, eds., vol. 139 of IAU Symposium, p. 269. 1990.
[3699] J. S. Huang, L. L. Cowie, J. P. Gardner, E. M. Hu, A. Songaila, and R. J. Wainscoat, “The hawaii k-band galaxy
survey. 2. Bright k-band imaging,” Astrophys. J. 476 (1997) 12, arXiv:astro-ph/9610084.
[3700] G. S. Busswell, T. Shanks, P. J. Outram, W. J. Frith, N. Metcalfe, and R. Fong, “The local hole in the galaxy
distribution: New optical evidence,” Mon. Not. Roy. Astron. Soc. 354 (2004) 991, arXiv:astro-ph/0302330.
[3701] W. J. Frith, G. S. Busswell, R. Fong, N. Metcalfe, and T. Shanks, “The local hole in the galaxy distribution: Evidence
from 2MASS,” Mon. Not. Roy. Astron. Soc. 345 (2003) 1049, arXiv:astro-ph/0302331.
[3702] W. J. Frith, T. Shanks, and P. J. Outram, “2MASS constraints on the local large-scale structure: A Challenge to
lambda-CDM?,” Mon. Not. Roy. Astron. Soc. 361 (2005) 701–709, arXiv:astro-ph/0411204.
[3703] W. J. Frith, N. Metcalfe, and T. Shanks, “New h-band galaxy number counts: a large local hole in the galaxy
distribution?,” Mon. Not. Roy. Astron. Soc. 371 (2006) 1601–1609, arXiv:astro-ph/0509875.
[3704] J. R. Whitbourn and T. Shanks, “The Local Hole revealed by galaxy counts and redshifts,” Mon. Not. Roy. Astron.
Soc. 437 (2014) 2146–2162, arXiv:1307.4405 [astro-ph.CO].
[3705] Extragalactic Astronomy Group, Durham University Collaboration, J. R. Whitbourn and T. Shanks, “The
galaxy luminosity function and the Local Hole,” Mon. Not. Roy. Astron. Soc. 459 (2016) no. 1, 496–507,
arXiv:1603.02322 [astro-ph.CO].
384

[3706] J. H. W. Wong, T. Shanks, N. Metcalfe, and J. R. Whitbourn, “The local hole: a galaxy underdensity covering
90 per cent of sky to ≈200 Mpc,” Mon. Not. Roy. Astron. Soc. 511 (2022) no. 4, 5742–5755, arXiv:2107.08505
[astro-ph.CO].
[3707] M. Rubart, D. Bacon, and D. J. Schwarz, “Impact of local structure on the cosmic radio dipole,” Astron. Astrophys.
565 (2014) A111, arXiv:1402.0376 [astro-ph.CO].
[3708] R. E. Angulo, V. Springel, S. D. M. White, A. Jenkins, C. M. Baugh, and C. S. Frenk, “Scaling relations for galaxy
clusters in the Millennium-XXL simulation,” Mon. Not. Roy. Astron. Soc. 426 (2012) 2046, arXiv:1203.3216
[astro-ph.CO].
[3709] M. Haslbauer, I. Banik, and P. Kroupa, “The KBC void and Hubble tension contradict ΛCDM on a Gpc scale −
Milgromian dynamics as a possible solution,” Mon. Not. Roy. Astron. Soc. 499 (2020) no. 2, 2845–2883,
arXiv:2009.11292 [astro-ph.CO].
[3710] R. C. Keenan, A. J. Barger, and L. L. Cowie, “Local Large-Scale Structure and the Assumption of Homogeneity,” IAU
Symp. 308 (2014) 295–298, arXiv:1409.8458 [astro-ph.CO].
[3711] T. Shanks, L. Hogarth, and N. Metcalfe, “Gaia Cepheid parallaxes and ’Local Hole’ relieve H0 tension,” Mon. Not.
Roy. Astron. Soc. 484 (2019) no. 1, L64–L68, arXiv:1810.02595 [astro-ph.CO].
[3712] T. Shanks, L. M. Hogarth, N. Metcalfe, and J. Whitbourn, “Local Hole revisited: evidence for bulk motions and
self-consistent outflow,” Mon. Not. Roy. Astron. Soc. 490 (2019) no. 4, 4715–4720, arXiv:1909.01878 [astro-ph.CO].
[3713] Q. Ding, T. Nakama, and Y. Wang, “A gigaparsec-scale local void and the Hubble tension,” Sci. China Phys. Mech.
Astron. 63 (2020) no. 9, 290403, arXiv:1912.12600 [astro-ph.CO].
[3714] M. S. Martín and C. Rubio, “Hubble tension and matter inhomogeneities: A theoretical perspective,” Annals Phys.
458 (2023) 169444, arXiv:2107.14377 [astro-ph.CO].
[3715] N. Kaiser, “Clustering in real space and in redshift space,” Mon. Not. Roy. Astron. Soc. 227 (1987) 1–27.
[3716] S. Mazurenko, I. Banik, P. Kroupa, and M. Haslbauer, “A simultaneous solution to the Hubble tension and observed
bulk flow within 250 h−1 Mpc,” Mon. Not. Roy. Astron. Soc. 527 (2024) no. 3, 4388–4396, arXiv:2311.17988
[astro-ph.CO].
[3717] H. Alnes and M. Amarzguioui, “CMB anisotropies seen by an off-center observer in a spherically symmetric
inhomogeneous Universe,” Phys. Rev. D 74 (2006) 103520, arXiv:astro-ph/0607334.
[3718] V. Nistane, G. Cusin, and M. Kunz, “CMB sky for an off-center observer in a local void. Part I. Framework for
forecasts,” JCAP 12 (2019) 038, arXiv:1908.05484 [astro-ph.CO].
[3719] A. Cimatti and M. Moresco, “Revisiting the Oldest Stars as Cosmological Probes: New Constraints on the Hubble
Constant,” Astrophys. J. 953 (2023) no. 2, 149, arXiv:2302.07899 [astro-ph.CO].
[3720] M. Xiang, H.-W. Rix, H. Yang, J. Liu, Y. Huang, and N. Frankel, “The formation and survival of the Milky Way’s
oldest stellar disk,” Nature Astron. 9 (2025) no. 1, 101–110, arXiv:2410.09705 [astro-ph.GA].
[3721] L. Perivolaropoulos, “Hubble tension or distance ladder crisis?,” Phys. Rev. D 110 (2024) no. 12, 123518,
arXiv:2408.11031 [astro-ph.CO].
[3722] C. Krishnan, E. O. Colgáin, Ruchika, A. A. Sen, M. M. Sheikh-Jabbari, and T. Yang, “Is there an early Universe
solution to Hubble tension?,” Phys. Rev. D 102 (2020) no. 10, 103525, arXiv:2002.06044 [astro-ph.CO].
[3723] C. Krishnan, E. O. Colgáin, M. M. Sheikh-Jabbari, and T. Yang, “Running Hubble Tension and a H0 Diagnostic,”
Phys. Rev. D 103 (2021) no. 10, 103509, arXiv:2011.02858 [astro-ph.CO].
[3724] X. D. Jia, J. P. Hu, and F. Y. Wang, “Evidence of a decreasing trend for the Hubble constant,” Astron. Astrophys. 674
(2023) A45, arXiv:2212.00238 [astro-ph.CO].
[3725] X. D. Jia, J. P. Hu, S. X. Yi, and F. Y. Wang, “Uncorrelated Estimations of H0 Redshift Evolution from DESI Baryon
Acoustic Oscillation Observations,” Astrophys. J. Lett. 979 (2025) no. 2, L34, arXiv:2406.02019 [astro-ph.CO].
[3726] S. Mazurenko, I. Banik, and P. Kroupa, “The redshift dependence of the inferred H0 in a local void solution to the
Hubble tension,” MNRAS 536 (2025) no. 4, 3232–3241, arXiv:2412.12245 [astro-ph.CO].
[3727] I. Banik and V. Kalaitzidis, “Testing the local void solution to the Hubble tension using baryon acoustic oscillation
measurements over the last twenty years,” arXiv:2501.17934 [astro-ph.CO].
[3728] K. Rezazadeh, A. Ashoorioon, and D. Grin, “Cascading Dark Energy,” Astrophys. J. 975 (2024) no. 1, 137,
arXiv:2208.07631 [astro-ph.CO].
[3729] M. Khanpour, E. Yusofi, and B. Khanpour, “Gravitational Merging as a Possible Source for the Cosmological
Accelerating,” arXiv:1709.08612 [astro-ph.CO].
[3730] W. D. Kenworthy, D. Scolnic, and A. Riess, “The Local Perspective on the Hubble Tension: Local Structure Does Not
Impact Measurement of the Hubble Constant,” Astrophys. J. 875 (2019) no. 2, 145, arXiv:1901.08681
[astro-ph.CO].
[3731] M. M. Phillips, “The absolute magnitudes of Type IA supernovae,” Astrophys. J. Lett. 413 (1993) L105–L108.
[3732] A. Friedman, “On the Curvature of space,” Z. Phys. 10 (1922) 377–386.
[3733] A. Friedmann, “Über die Möglichkeit einer Welt mit konstanter negativer Krümmung des Raumes,” Zeitschrift fur
Physik 21 (1924) no. 1, 326–332.
[3734] M. C. March, R. Trotta, P. Berkes, G. D. Starkman, and P. M. Vaudrevange, “Improved constraints on cosmological
parameters from SNIa data,” Mon. Not. Roy. Astron. Soc. 418 (2011) 2308–2329, arXiv:1102.3237 [astro-ph.CO].
[3735] SDSS Collaboration, J. Marriner, J. P. Bernstein, R. Kessler, H. Lampeitl, R. Miquel, J. Mosher, R. C. Nichol,
M. Sako, and M. Smith, “A More General Model for the Intrinsic Scatter in Type Ia Supernova Distance Moduli,”
Astrophys. J. 740 (2011) 72, arXiv:1107.4631 [astro-ph.CO].
[3736] R. Kessler and D. Scolnic, “Correcting Type Ia Supernova Distances for Selection Biases and Contamination in
385

Photometrically Identified Samples,” Astrophys. J. 836 (2017) no. 1, 56, arXiv:1610.04677 [astro-ph.CO].
[3737] S. Castello, M. Högås, and E. Mörtsell, “A cosmological underdensity does not solve the Hubble tension,” JCAP 07
(2022) 003, arXiv:2110.04226 [astro-ph.CO]. [Erratum: JCAP 09, E01 (2022)].
[3738] D. Camarena, V. Marra, Z. Sakr, and C. Clarkson, “The Copernican principle in light of the latest cosmological data,”
Mon. Not. Roy. Astron. Soc. 509 (2021) no. 1, 1291–1302, arXiv:2107.02296 [astro-ph.CO].
[3739] D. Camarena, V. Marra, Z. Sakr, and C. Clarkson, “A void in the Hubble tension? The end of the line for the Hubble
bubble,” Class. Quant. Grav. 39 (2022) no. 18, 184001, arXiv:2205.05422 [astro-ph.CO].
[3740] Z. G. Lane, A. Seifert, R. Ridden-Harper, and D. L. Wiltshire, “Cosmological foundations revisited with Pantheon+,”
Mon. Not. Roy. Astron. Soc. 536 (2025) no. 2, 1752–1777, arXiv:2311.01438 [astro-ph.CO].
[3741] A. Seifert, Z. G. Lane, M. Galoppo, R. Ridden-Harper, and D. L. Wiltshire, “Supernovae evidence for foundational
change to cosmological models,” Mon. Not. Roy. Astron. Soc. 537 (2025) no. 1, L55–L60, arXiv:2412.15143
[astro-ph.CO].
[3742] DES Collaboration, P. Wiseman et al., “Rates and delay times of Type Ia supernovae in the Dark Energy Survey,”
Mon. Not. Roy. Astron. Soc. 506 (2021) no. 3, 3330–3348, arXiv:2105.11954 [astro-ph.GA].
[3743] N. Nicolas, M. Rigault, Y. Copin, R. Graziani, G. Aldering, M. Briday, Y. L. Kim, J. Nordin, S. Perlmutter, and
M. Smith, “Redshift evolution of the underlying type Ia supernova stretch distribution,” Astron. Astrophys. 649 (2021)
A74, arXiv:2005.09441 [astro-ph.CO].
[3744] P. Wiseman, M. Sullivan, M. Smith, and B. Popovic, “Further evidence that galaxy age drives observed Type Ia
supernova luminosity differences,” Mon. Not. Roy. Astron. Soc. 520 (2023) no. 4, 6214–6222, arXiv:2302.05341
[astro-ph.GA].
[3745] R. E. Keeley, A. Shafieloo, and B. L’Huillier, “An Analysis of Variance of the Pantheon+ Dataset: Systematics in the
Covariance Matrix?,” Universe 10 (2024) no. 12, 439, arXiv:2212.07917 [astro-ph.CO].
[3746] K. S. Mandel, D. Scolnic, H. Shariff, R. J. Foley, and R. P. Kirshner, “The Type Ia Supernova Color–Magnitude
Relation and Host Galaxy Dust: A Simple Hierarchical Bayesian Model,” Astrophys. J. 842 (2017) no. 2, 93,
arXiv:1609.04470 [astro-ph.CO].
[3747] V. V. Luković, B. S. Haridasu, and N. Vittorio, “Exploring the evidence for a large local void with supernovae Ia data,”
Mon. Not. Roy. Astron. Soc. 491 (2020) no. 2, 2075–2087, arXiv:1907.11219 [astro-ph.CO].
[3748] E. Asencio, I. Banik, and P. Kroupa, “A massive blow for ΛCDM − the high redshift, mass, and collision velocity of
the interacting galaxy cluster El Gordo contradicts concordance cosmology,” Mon. Not. Roy. Astron. Soc. 500 (2020)
no. 4, 5249–5267, arXiv:2012.03950 [astro-ph.CO].
[3749] E. Asencio, I. Banik, and P. Kroupa, “The El Gordo Galaxy Cluster Challenges ΛCDM for Any Plausible Collision
Velocity,” Astrophys. J. 954 (2023) no. 2, 162, arXiv:2308.00744 [astro-ph.CO].
[3750] J. P. Hu, X. D. Jia, J. Hu, and F. Y. Wang, “Hints of New Physics for the Hubble Tension: Violation of Cosmological
Principle,” Astrophys. J. Lett. 975 (2024) no. 2, L36, arXiv:2410.06450 [astro-ph.CO].
[3751] A. Sah, M. Rameez, S. Sarkar, and C. Tsagas, “Anisotropy in Pantheon+ supernovae,” arXiv:2411.10838
[astro-ph.CO].
[3752] R. Durrer and A. Neronov, “Cosmological Magnetic Fields: Their Generation, Evolution and Observation,” Astron.
Astrophys. Rev. 21 (2013) 62, arXiv:1303.7121 [astro-ph.CO].
[3753] K. Subramanian, “The origin, evolution and signatures of primordial magnetic fields,” Rept. Prog. Phys. 79 (2016)
no. 7, 076901, arXiv:1504.02311 [astro-ph.CO].
[3754] T. Vachaspati, “Progress on cosmological magnetic fields,” Rept. Prog. Phys. 84 (2021) no. 7, 074901,
arXiv:2010.10525 [astro-ph.CO].
[3755] A. Elyiv, A. Neronov, and D. V. Semikoz, “Gamma-ray induced cascades and magnetic fields in intergalactic medium,”
Phys. Rev. D 80 (2009) 023010, arXiv:0903.3649 [astro-ph.CO].
[3756] A. Neronov and I. Vovk, “Evidence for strong extragalactic magnetic fields from Fermi observations of TeV blazars,”
Science 328 (2010) 73–75, arXiv:1006.3504 [astro-ph.HE].
[3757] F. Tavecchio, G. Ghisellini, L. Foschini, G. Bonnoli, G. Ghirlanda, and P. Coppi, “The intergalactic magnetic field
constrained by Fermi/LAT observations of the TeV blazar 1ES 0229+200,” Mon. Not. Roy. Astron. Soc. 406 (2010)
L70–L74, arXiv:1004.1329 [astro-ph.CO].
[3758] F. Tavecchio, G. Ghisellini, G. Bonnoli, and L. Foschini, “Extreme TeV blazars and the intergalactic magnetic field,”
Mon. Not. Roy. Astron. Soc. 414 (2011) 3566, arXiv:1009.1048 [astro-ph.HE].
[3759] A. M. Taylor, I. Vovk, and A. Neronov, “Extragalactic magnetic fields constraints from simultaneous GeV-TeV
observations of blazars,” Astron. Astrophys. 529 (2011) A144, arXiv:1101.0932 [astro-ph.HE].
[3760] I. Vovk, A. M. Taylor, D. Semikoz, and A. Neronov, “Fermi/LAT observations of 1ES 0229+200: implications for
extragalactic magnetic fields and background light,” Astrophys. J. Lett. 747 (2012) L14, arXiv:1112.2534
[astro-ph.CO].
[3761] K. Dolag, M. Kachelriess, S. Ostapchenko, and R. Tomas, “Lower limit on the strength and filling factor of
extragalactic magnetic fields,” Astrophys. J. Lett. 727 (2011) L4, arXiv:1009.1782 [astro-ph.HE].
[3762] K. Subramanian and J. D. Barrow, “Microwave background signals from tangled magnetic fields,” Phys. Rev. Lett. 81
(1998) 3575–3578, arXiv:astro-ph/9803261.
[3763] K. Jedamzik, V. Katalinic, and A. V. Olinto, “A Limit on primordial small scale magnetic fields from CMB
distortions,” Phys. Rev. Lett. 85 (2000) 700–703, arXiv:astro-ph/9911100.
[3764] R. Durrer, P. G. Ferreira, and T. Kahniashvili, “Tensor microwave anisotropies from a stochastic magnetic field,” Phys.
Rev. D 61 (2000) 043001, arXiv:astro-ph/9911040.
386

[3765] T. R. Seshadri and K. Subramanian, “CMBR polarization signals from tangled magnetic fields,” Phys. Rev. Lett. 87
(2001) 101301, arXiv:astro-ph/0012056.
[3766] A. Mack, T. Kahniashvili, and A. Kosowsky, “Microwave background signatures of a primordial stochastic magnetic
field,” Phys. Rev. D 65 (2002) 123004, arXiv:astro-ph/0105504.
[3767] K. Subramanian and J. D. Barrow, “Small-scale microwave background anisotropies due to tangled primordial
magnetic fields,” Mon. Not. Roy. Astron. Soc. 335 (2002) L57, arXiv:astro-ph/0205312.
[3768] K. Subramanian, T. R. Seshadri, and J. D. Barrow, “Small - scale CMB polarization anisotropies due to tangled
primordial magnetic fields,” Mon. Not. Roy. Astron. Soc. 344 (2003) L31, arXiv:astro-ph/0303014.
[3769] S. Mollerach, D. Harari, and S. Matarrese, “CMB polarization from secondary vector and tensor modes,” Phys. Rev. D
69 (2004) 063002, arXiv:astro-ph/0310711.
[3770] A. Lewis, “Observable primordial vector modes,” Phys. Rev. D 70 (2004) 043518, arXiv:astro-ph/0403583.
[3771] C. Scoccola, D. Harari, and S. Mollerach, “B polarization of the CMB from Faraday rotation,” Phys. Rev. D 70 (2004)
063003, arXiv:astro-ph/0405396.
[3772] S. K. Sethi and K. Subramanian, “Primordial magnetic fields in the post-recombination era and early reionization,”
Mon. Not. Roy. Astron. Soc. 356 (2005) 778–788, arXiv:astro-ph/0405413.
[3773] A. Kosowsky, T. Kahniashvili, G. Lavrelashvili, and B. Ratra, “Faraday rotation of the Cosmic Microwave Background
polarization by a stochastic magnetic field,” Phys. Rev. D 71 (2005) 043006, arXiv:astro-ph/0409767.
[3774] T. Kahniashvili and B. Ratra, “Effects of Cosmological Magnetic Helicity on the Cosmic Microwave Background,”
Phys. Rev. D 71 (2005) 103006, arXiv:astro-ph/0503709.
[3775] I. Brown and R. Crittenden, “Non-Gaussianity from cosmic magnetic fields,” Phys. Rev. D 72 (2005) 063002,
arXiv:astro-ph/0506570.
[3776] A. Zizzo and C. Burigana, “On the effect of cyclotron emission on the spectral distortions of the cosmic microwave
background,” New Astron. 11 (2005) 1–16, arXiv:astro-ph/0505259.
[3777] G. Chen, P. Mukherjee, T. Kahniashvili, B. Ratra, and Y. Wang, “Looking for cosmological Alfven waves in WMAP
data,” Astrophys. J. 611 (2004) 655–659, arXiv:astro-ph/0403695.
[3778] A. Lewis, “CMB anisotropies from primordial inhomogeneous magnetic fields,” Phys. Rev. D 70 (2004) 043011,
arXiv:astro-ph/0406096.
[3779] H. Tashiro, N. Sugiyama, and R. Banerjee, “Nonlinear evolution of cosmic magnetic fields and cosmic microwave
background anisotropies,” Phys. Rev. D 73 (2006) 023002, arXiv:astro-ph/0509220.
[3780] D. Yamazaki, K. Ichiki, T. Kajino, and G. J. Mathews, “Constraints on the evolution of the primordial magnetic field
from the small scale cmb angular anisotropy,” Astrophys. J. 646 (2006) 719–729, arXiv:astro-ph/0602224.
[3781] T. Kahniashvili and B. Ratra, “CMB anisotropies due to cosmological magnetosonic waves,” Phys. Rev. D 75 (2007)
023002, arXiv:astro-ph/0611247.
[3782] M. Giovannini and K. E. Kunze, “Magnetized CMB observables: A Dedicated numerical approach,” Phys. Rev. D 77
(2008) 063003, arXiv:0712.3483 [astro-ph].
[3783] T. R. Seshadri and K. Subramanian, “CMB bispectrum from primordial magnetic fields on large angular scales,” Phys.
Rev. Lett. 103 (2009) 081303, arXiv:0902.4066 [astro-ph.CO].
[3784] C. Caprini, F. Finelli, D. Paoletti, and A. Riotto, “The cosmic microwave background temperature bispectrum from
scalar perturbations induced by primordial magnetic fields,” JCAP 06 (2009) 021, arXiv:0903.1420 [astro-ph.CO].
[3785] R.-G. Cai, B. Hu, and H.-B. Zhang, “Acoustic signatures in the Cosmic Microwave Background bispectrum from
primordial magnetic fields,” JCAP 08 (2010) 025, arXiv:1006.2985 [astro-ph.CO].
[3786] P. Trivedi, K. Subramanian, and T. R. Seshadri, “Primordial Magnetic Field Limits from Cosmic Microwave
Background Bispectrum of Magnetic Passive Scalar Modes,” Phys. Rev. D 82 (2010) 123006, arXiv:1009.2724
[astro-ph.CO].
[3787] I. A. Brown, “Intrinsic Bispectra of Cosmic Magnetic Fields,” Astrophys. J. 733 (2011) 83, arXiv:1012.2892
[astro-ph.CO].
[3788] M. Shiraishi, D. Nitta, S. Yokoyama, K. Ichiki, and K. Takahashi, “Cosmic microwave background bispectrum of vector
modes induced from primordial magnetic fields,” Phys. Rev. D 82 (2010) 121302, arXiv:1009.3632 [astro-ph.CO].
[Erratum: Phys.Rev.D 83, 029901 (2011)].
[3789] M. Shiraishi, D. Nitta, S. Yokoyama, K. Ichiki, and K. Takahashi, “Cosmic microwave background bispectrum of tensor
passive modes induced from primordial magnetic fields,” Phys. Rev. D 83 (2011) 123003, arXiv:1103.4103
[astro-ph.CO].
[3790] P. Trivedi, T. R. Seshadri, and K. Subramanian, “Cosmic Microwave Background Trispectrum and Primordial
Magnetic Field Limits,” Phys. Rev. Lett. 108 (2012) 231301, arXiv:1111.0744 [astro-ph.CO].
[3791] D. G. Yamazaki, K. Ichiki, T. Kajino, and G. J. Mathews, “New Constraints on the Primordial Magnetic Field,” Phys.
Rev. D 81 (2010) 023008, arXiv:1001.2012 [astro-ph.CO].
[3792] D. Paoletti and F. Finelli, “CMB Constraints on a Stochastic Background of Primordial Magnetic Fields,” Phys. Rev.
D 83 (2011) 123533, arXiv:1005.0148 [astro-ph.CO].
[3793] J. R. Shaw and A. Lewis, “Constraining Primordial Magnetism,” Phys. Rev. D 86 (2012) 043510, arXiv:1006.4242
[astro-ph.CO].
[3794] K. E. Kunze, “CMB anisotropies in the presence of a stochastic magnetic field,” Phys. Rev. D 83 (2011) 023006,
arXiv:1007.3163 [astro-ph.CO].
[3795] L. Pogosian, T. Vachaspati, and A. Yadav, “Primordial Magnetism in CMB B-modes,” Can. J. Phys. 91 (2013)
451–454, arXiv:1210.0308 [astro-ph.CO].
387

[3796] D. Paoletti and F. Finelli, “Constraints on a Stochastic Background of Primordial Magnetic Fields with WMAP and
South Pole Telescope data,” Phys. Lett. B 726 (2013) 45–49, arXiv:1208.2625 [astro-ph.CO].
[3797] K. E. Kunze and E. Komatsu, “Constraining primordial magnetic fields with distortions of the black-body spectrum of
the cosmic microwave background: pre- and post-decoupling contributions,” JCAP 01 (2014) 009, arXiv:1309.7994
[astro-ph.CO].
[3798] M. Shiraishi and T. Sekiguchi, “First observational constraints on tensor non-Gaussianity sourced by primordial
magnetic fields from cosmic microwave background,” Phys. Rev. D 90 (2014) no. 10, 103002, arXiv:1304.7277
[astro-ph.CO].
[3799] P. Trivedi, K. Subramanian, and T. R. Seshadri, “Primordial magnetic field limits from the CMB trispectrum: Scalar
modes and Planck constraints,” Phys. Rev. D 89 (2014) no. 4, 043523, arXiv:1312.5308 [astro-ph.CO].
[3800] M. Ballardini, F. Finelli, and D. Paoletti, “CMB anisotropies generated by a stochastic background of primordial
magnetic fields with non-zero helicity,” JCAP 10 (2015) 031, arXiv:1412.1836 [astro-ph.CO].
[3801] T. Kahniashvili, Y. Maravin, G. Lavrelashvili, and A. Kosowsky, “Primordial Magnetic Helicity Constraints from
WMAP Nine-Year Data,” Phys. Rev. D 90 (2014) no. 8, 083004, arXiv:1408.0351 [astro-ph.CO].
[3802] K. E. Kunze and E. Komatsu, “Constraints on primordial magnetic fields from the optical depth of the cosmic
microwave background,” JCAP 06 (2015) 027, arXiv:1501.00142 [astro-ph.CO].
[3803] Planck Collaboration, P. A. R. Ade et al., “Planck 2015 results. XIX. Constraints on primordial magnetic fields,”
Astron. Astrophys. 594 (2016) A19, arXiv:1502.01594 [astro-ph.CO].
[3804] J. Ganc and M. S. Sloth, “Probing correlations of early magnetic fields using mu-distortion,” JCAP 08 (2014) 018,
arXiv:1404.5957 [astro-ph.CO].
[3805] J. Chluba, D. Paoletti, F. Finelli, and J.-A. Rubiño Martín, “Effect of primordial magnetic fields on the ionization
history,” Mon. Not. Roy. Astron. Soc. 451 (2015) no. 2, 2244–2250, arXiv:1503.04827 [astro-ph.CO].
[3806] A. Zucca, Y. Li, and L. Pogosian, “Constraints on Primordial Magnetic Fields from Planck combined with the South
Pole Telescope CMB B-mode polarization measurements,” Phys. Rev. D 95 (2017) no. 6, 063506, arXiv:1611.00757
[astro-ph.CO].
[3807] D. R. Sutton, C. Feng, and C. L. Reichardt, “Current and Future Constraints on Primordial Magnetic Fields,”
Astrophys. J. 846 (2017) no. 2, 164, arXiv:1702.01871 [astro-ph.CO].
[3808] T. Minoda, K. Ichiki, and H. Tashiro, “Small-scale CMB anisotropies induced by the primordial magnetic fields,”
JCAP 03 (2021) 093, arXiv:2012.12542 [astro-ph.CO].
[3809] R. Banerjee and K. Jedamzik, “Are cluster magnetic fields primordial?,” Phys. Rev. Lett. 91 (2003) 251301,
arXiv:astro-ph/0306211. [Erratum: Phys.Rev.Lett. 93, 179901 (2004)].
[3810] K. Jedamzik and T. Abel, “Weak Primordial Magnetic Fields and Anisotropies in the Cosmic Microwave Background
Radiation,” arXiv:1108.2517 [astro-ph.CO].
[3811] K. Jedamzik and T. Abel, “Small-scale primordial magnetic fields and anisotropies in the cosmic microwave
background radiation,” JCAP 10 (2013) 050.
[3812] K. Jedamzik and A. Saveliev, “Stringent Limit on Primordial Magnetic Fields from the Cosmic Microwave Background
Radiation,” Phys. Rev. Lett. 123 (2019) no. 2, 021301, arXiv:1804.06115 [astro-ph.CO].
[3813] K. Jedamzik and L. Pogosian, “Relieving the Hubble tension with primordial magnetic fields,” Phys. Rev. Lett. 125
(2020) no. 18, 181302, arXiv:2004.09487 [astro-ph.CO].
[3814] L. Thiele, Y. Guan, J. C. Hill, A. Kosowsky, and D. N. Spergel, “Can small-scale baryon inhomogeneities resolve the
Hubble tension? An investigation with ACT DR4,” Phys. Rev. D 104 (2021) no. 6, 063535, arXiv:2105.03003
[astro-ph.CO].
[3815] M. Rashkovetskyi, J. B. Muñoz, D. J. Eisenstein, and C. Dvorkin, “Small-scale clumping at recombination and the
Hubble tension,” Phys. Rev. D 104 (2021) no. 10, 103517, arXiv:2108.02747 [astro-ph.CO].
[3816] S. Galli, L. Pogosian, K. Jedamzik, and L. Balkenhol, “Consistency of Planck, ACT, and SPT constraints on
magnetically assisted recombination and forecasts for future experiments,” Phys. Rev. D 105 (2022) no. 2, 023513,
arXiv:2109.03816 [astro-ph.CO].
[3817] K. Jedamzik, T. Abel, and Y. Ali-Haimoud, “Cosmic recombination in the presence of primordial magnetic fields,”
JCAP 03 (2025) 012, arXiv:2312.11448 [astro-ph.CO].
[3818] A. Korochkin, O. Kalashev, A. Neronov, and D. Semikoz, “Sensitivity reach of gamma-ray measurements for strong
cosmological magnetic fields,” Astrophys. J. 906 (2021) no. 2, 116, arXiv:2007.14331 [astro-ph.CO].
[3819] A. H. Guth, “The Inflationary Universe: A Possible Solution to the Horizon and Flatness Problems,” Phys. Rev. D 23
(1981) 347–356.
[3820] A. R. Liddle and D. H. Lyth, “The end for extended inflation?,” Annals N. Y. Acad. Sci. 688 (1993) 653,
arXiv:astro-ph/9302010.
[3821] A. D. Linde, “Fast roll inflation,” JHEP 11 (2001) 052, arXiv:hep-th/0110195.
[3822] H. Motohashi, A. A. Starobinsky, and J. Yokoyama, “Inflation with a constant rate of roll,” JCAP 09 (2015) 018,
arXiv:1411.5021 [astro-ph.CO].
[3823] M. Guerrero, D. Rubiera-Garcia, and D. Saez-Chillon Gomez, “Constant roll inflation in multifield models,” Phys. Rev.
D 102 (2020) 123528, arXiv:2008.07260 [gr-qc].
[3824] T. Bjorkmo, R. Z. Ferreira, and M. C. D. Marsh, “Mild Non-Gaussianities under Perturbative Control from
Rapid-Turn Inflation Models,” JCAP 12 (2019) 036, arXiv:1908.11316 [hep-th].
[3825] L. Anguelova and C. I. Lazaroiu, “Dynamical consistency conditions for rapid turn inflation,” JCAP 05 (2023) 020,
arXiv:2210.00031 [hep-th].
388

[3826] C. Gomes, J. a. G. Rosa, and O. Bertolami, “Inflation in non-minimal matter-curvature coupling theories,” JCAP 06
(2017) 021, arXiv:1611.02124 [gr-qc].
[3827] I. I. Çimdiker, “Starobinsky inflation in emergent gravity,” Phys. Dark Univ. 30 (2020) 100736.
[3828] N. Bostan, C. Karahan, and O. Sargın, “Inflation in symmergent metric-Palatini gravity,” JCAP 02 (2024) 028,
arXiv:2308.04507 [astro-ph.CO].
[3829] A. Maleknejad and M. M. Sheikh-Jabbari, “Non-Abelian Gauge Field Inflation,” Phys. Rev. D 84 (2011) 043515,
arXiv:1102.1932 [hep-ph].
[3830] P. Adshead and M. Wyman, “Chromo-Natural Inflation: Natural inflation on a steep potential with classical
non-Abelian gauge fields,” Phys. Rev. Lett. 108 (2012) 261302, arXiv:1202.2366 [hep-th].
[3831] W. D. Garretson, G. B. Field, and S. M. Carroll, “Primordial magnetic fields from pseudoGoldstone bosons,” Phys.
Rev. D 46 (1992) 5346–5351, arXiv:hep-ph/9209238.
[3832] M. M. Anber and L. Sorbo, “N-flationary magnetic fields,” JCAP 10 (2006) 018, arXiv:astro-ph/0606534.
[3833] N. Barnaby and M. Peloso, “Large Nongaussianity in Axion Inflation,” Phys. Rev. Lett. 106 (2011) 181301,
arXiv:1011.1500 [hep-ph].
[3834] J. L. Cook and L. Sorbo, “Particle production during inflation and gravitational waves detectable by ground-based
interferometers,” Phys. Rev. D 85 (2012) 023534, arXiv:1109.0022 [astro-ph.CO]. [Erratum: Phys.Rev.D 86, 069901
(2012)].
[3835] L. Sorbo, “Parity violation in the Cosmic Microwave Background from a pseudoscalar inflaton,” JCAP 06 (2011) 003,
arXiv:1101.1525 [astro-ph.CO].
[3836] M. M. Anber and L. Sorbo, “Non-Gaussianities and chiral gravitational waves in natural steep inflation,” Phys. Rev. D
85 (2012) 123537, arXiv:1203.5849 [astro-ph.CO].
[3837] E. Dimastrogiovanni and M. Peloso, “Stability analysis of chromo-natural inflation and possible evasion of Lyth’s
bound,” Phys. Rev. D 87 (2013) no. 10, 103501, arXiv:1212.5184 [astro-ph.CO].
[3838] P. Adshead, E. Martinec, E. I. Sfakianakis, and M. Wyman, “Higgsed Chromo-Natural Inflation,” JHEP 12 (2016) 137,
arXiv:1609.04025 [hep-th].
[3839] E. Dimastrogiovanni, M. Fasiello, and T. Fujita, “Primordial Gravitational Waves from Axion-Gauge Fields
Dynamics,” JCAP 01 (2017) 019, arXiv:1608.04216 [astro-ph.CO].
[3840] A. Agrawal, T. Fujita, and E. Komatsu, “Large tensor non-Gaussianity from axion-gauge field dynamics,” Phys. Rev. D
97 (2018) no. 10, 103526, arXiv:1707.03023 [astro-ph.CO].
[3841] R. R. Caldwell and C. Devulder, “Axion Gauge Field Inflation and Gravitational Leptogenesis: A Lower Bound on B
Modes from the Matter-Antimatter Asymmetry of the Universe,” Phys. Rev. D 97 (2018) no. 2, 023532,
arXiv:1706.03765 [astro-ph.CO].
[3842] B. Thorne, T. Fujita, M. Hazumi, N. Katayama, E. Komatsu, and M. Shiraishi, “Finding the chiral gravitational wave
background of an axion-SU(2) inflationary model using CMB observations and laser interferometers,” Phys. Rev. D 97
(2018) no. 4, 043506, arXiv:1707.03240 [astro-ph.CO].
[3843] E. Dimastrogiovanni, M. Fasiello, R. J. Hardwick, H. Assadullahi, K. Koyama, and D. Wands, “Non-Gaussianity from
Axion-Gauge Fields Interactions during Inflation,” JCAP 11 (2018) 029, arXiv:1806.05474 [astro-ph.CO].
[3844] T. Fujita, R. Namba, and I. Obata, “Mixed Non-Gaussianity from Axion-Gauge Field Dynamics,” JCAP 04 (2019)
044, arXiv:1811.12371 [astro-ph.CO].
[3845] V. Domcke, B. Mares, F. Muia, and M. Pieroni, “Emerging chromo-natural inflation,” JCAP 04 (2019) 034,
arXiv:1807.03358 [hep-ph].
[3846] K. D. Lozanov, A. Maleknejad, and E. Komatsu, “Schwinger Effect by an SU (2) Gauge Field during Inflation,” JHEP
02 (2019) 041, arXiv:1805.09318 [hep-th].
[3847] Y. Watanabe and E. Komatsu, “Gravitational Wave from Axion-SU(2) Gauge Fields: Effective Field Theory for
Kinetically Driven Inflation,” arXiv:2004.04350 [hep-th].
[3848] J. Holland, I. Zavala, and G. Tasinato, “On chromonatural inflation in string theory,” JCAP 12 (2020) 026,
arXiv:2009.00653 [hep-th].
[3849] V. Domcke, V. Guidetti, Y. Welling, and A. Westphal, “Resonant backreaction in axion inflation,” JCAP 09 (2020)
009, arXiv:2002.02952 [astro-ph.CO].
[3850] O. Iarygina, E. I. Sfakianakis, R. Sharma, and A. Brandenburg, “Backreaction of axion-SU(2) dynamics during
inflation,” JCAP 04 (2024) 018, arXiv:2311.07557 [astro-ph.CO].
[3851] K. Ishiwata, E. Komatsu, and I. Obata, “Axion-gauge field dynamics with backreaction,” JCAP 03 (2022) no. 03, 010,
arXiv:2111.14429 [hep-ph].
[3852] R. Durrer, R. von Eckardstein, D. Garg, K. Schmitz, O. Sobol, and S. Vilchinskii, “Scalar perturbations from inflation
in the presence of gauge fields,” Phys. Rev. D 110 (2024) no. 4, 043533, arXiv:2404.19694 [astro-ph.CO].
[3853] E. Dimastrogiovanni, M. Fasiello, and A. Papageorgiou, “Novel primordial black hole production mechanism from
non-Abelian gauge fields during inflation,” Phys. Rev. D 110 (2024) no. 10, 103542, arXiv:2403.13581 [astro-ph.CO].
[3854] L. A. Anchordoqui, V. Barger, H. Goldberg, X. Huang, and D. Marfatia, “S-dual Inflation: BICEP2 data without
unlikeliness,” Phys. Lett. B 734 (2014) 134–136, arXiv:1403.4578 [hep-ph].
[3855] L. A. Anchordoqui, I. Antoniadis, D. Lüst, and J. F. Soriano, “S-dual inflation and the string swampland,” Phys. Rev.
D 103 (2021) no. 12, 123537, arXiv:2103.07982 [hep-th].
[3856] A. Övgün, “Inflation and Acceleration of the Universe by Nonlinear Magnetic Monopole Fields,” Eur. Phys. J. C 77
(2017) no. 2, 105, arXiv:1604.01837 [gr-qc].
[3857] H. B. Benaoum, G. Leon, A. Ovgun, and H. Quevedo, “Inflation driven by non-linear electrodynamics,” Eur. Phys. J.
389

C 83 (2023) no. 5, 367, arXiv:2206.13157 [gr-qc].


[3858] G. Otalora, A. Övgün, J. Saavedra, and N. Videla, “Inflation from a nonlinear magnetic monopole field nonminimally
coupled to curvature,” JCAP 06 (2018) 003, arXiv:1803.11358 [gr-qc].
[3859] A. Övgün, G. Leon, J. Magaña, and K. Jusufi, “Falsifying cosmological models based on a non-linear electrodynamics,”
Eur. Phys. J. C 78 (2018) no. 6, 462, arXiv:1709.09794 [gr-qc].
[3860] S. I. Kruglov, “Universe acceleration and nonlinear electrodynamics,” Phys. Rev. D 92 (2015) no. 12, 123523,
arXiv:1601.06309 [gr-qc].
[3861] V. A. De Lorenci, R. Klippert, M. Novello, and J. M. Salim, “Nonlinear electrodynamics and FRW cosmology,” Phys.
Rev. D 65 (2002) 063501.
[3862] M. Novello, S. E. Perez Bergliaffa, and J. Salim, “Non-linear electrodynamics and the acceleration of the universe,”
Phys. Rev. D 69 (2004) 127301, arXiv:astro-ph/0312093.
[3863] M. Novello, E. Goulart, J. M. Salim, and S. E. Perez Bergliaffa, “Cosmological Effects of Nonlinear Electrodynamics,”
Class. Quant. Grav. 24 (2007) 3021–3036, arXiv:gr-qc/0610043.
[3864] D. N. Vollick, “Homogeneous and isotropic cosmologies with nonlinear electromagnetic radiation,” Phys. Rev. D 78
(2008) 063524, arXiv:0807.0448 [gr-qc].
[3865] M. LoVerde, A. Miller, S. Shandera, and L. Verde, “Effects of Scale-Dependent Non-Gaussianity on Cosmological
Structures,” JCAP 04 (2008) 014, arXiv:0711.4126 [astro-ph].
[3866] C. Stahl, B. Famaey, R. Ibata, O. Hahn, N. Martinet, and T. Montandon, “Scale-dependent local primordial
non-Gaussianity as a solution to the S8 tension,” Phys. Rev. D 110 (2024) no. 6, 063501, arXiv:2404.03244
[astro-ph.CO].
[3867] F. L. Bezrukov and M. Shaposhnikov, “The Standard Model Higgs boson as the inflaton,” Phys. Lett. B 659 (2008)
703–706, arXiv:0710.3755 [hep-th].
[3868] J. Rubio, “Higgs inflation,” Front. Astron. Space Sci. 5 (2019) 50, arXiv:1807.02376 [hep-ph].
[3869] F. Bezrukov, D. Gorbunov, and M. Shaposhnikov, “On initial conditions for the Hot Big Bang,” JCAP 06 (2009) 029,
arXiv:0812.3622 [hep-ph].
[3870] J. Garcia-Bellido, D. G. Figueroa, and J. Rubio, “Preheating in the Standard Model with the Higgs-Inflaton coupled to
gravity,” Phys. Rev. D 79 (2009) 063531, arXiv:0812.4624 [hep-ph].
[3871] J. Repond and J. Rubio, “Combined Preheating on the lattice with applications to Higgs inflation,” JCAP 07 (2016)
043, arXiv:1604.08238 [astro-ph.CO].
[3872] F. Bezrukov, J. Rubio, and M. Shaposhnikov, “Living beyond the edge: Higgs inflation and vacuum metastability,”
Phys. Rev. D 92 (2015) no. 8, 083512, arXiv:1412.3811 [hep-ph].
[3873] F. Bezrukov, M. Pauly, and J. Rubio, “On the robustness of the primordial power spectrum in renormalized Higgs
inflation,” JCAP 02 (2018) 040, arXiv:1706.05007 [hep-ph].
[3874] J. G. Rodrigues, M. Benetti, and J. S. Alcaniz, “Possible discrepancies between cosmological and electroweak
observables in Higgs Inflation,” JHEP 11 (2021) 091, arXiv:2105.07009 [hep-ph].
[3875] J. G. Rodrigues, M. Benetti, R. de Souza, and J. Alcaniz, “Higgs inflation: Constraining the top quark mass and
breaking the H0-σ8 correlation,” Phys. Lett. B 852 (2024) 138607, arXiv:2301.11788 [astro-ph.CO].
[3876] M. Shaposhnikov and D. Zenhausern, “Scale invariance, unimodular gravity and dark energy,” Phys. Lett. B 671
(2009) 187–192, arXiv:0809.3395 [hep-th].
[3877] J. Garcia-Bellido, J. Rubio, M. Shaposhnikov, and D. Zenhausern, “Higgs-Dilaton Cosmology: From the Early to the
Late Universe,” Phys. Rev. D 84 (2011) 123504, arXiv:1107.2163 [hep-ph].
[3878] C. Germani and A. Kehagias, “New Model of Inflation with Non-minimal Derivative Coupling of Standard Model
Higgs Boson to Gravity,” Phys. Rev. Lett. 105 (2010) 011302, arXiv:1003.2635 [hep-ph].
[3879] F. Bauer and D. A. Demir, “Higgs-Palatini Inflation and Unitarity,” Phys. Lett. B 698 (2011) 425–429,
arXiv:1012.2900 [hep-ph].
[3880] S. Rasanen, “Higgs inflation in the Palatini formulation with kinetic terms for the metric,” Open J. Astrophys. 2 (2019)
no. 1, 1, arXiv:1811.09514 [gr-qc].
[3881] J. Rubio and E. S. Tomberg, “Preheating in Palatini Higgs inflation,” JCAP 04 (2019) 021, arXiv:1902.10148
[hep-ph].
[3882] M. Shaposhnikov, A. Shkerin, I. Timiryasov, and S. Zell, “Higgs inflation in Einstein-Cartan gravity,” JCAP 02 (2021)
008, arXiv:2007.14978 [hep-ph]. [Erratum: JCAP 10, E01 (2021)].
[3883] M. Piani and J. Rubio, “Higgs-Dilaton inflation in Einstein-Cartan gravity,” JCAP 05 (2022) no. 05, 009,
arXiv:2202.04665 [gr-qc].
[3884] S. Casas, M. Pauly, and J. Rubio, “Higgs-dilaton cosmology: An inflation–dark-energy connection and forecasts for
future galaxy surveys,” Phys. Rev. D 97 (2018) no. 4, 043520, arXiv:1712.04956 [astro-ph.CO].
[3885] M. Trashorras, S. Nesseris, and J. Garcia-Bellido, “Cosmological Constraints on Higgs-Dilaton Inflation,” Phys. Rev. D
94 (2016) no. 6, 063511, arXiv:1604.06760 [astro-ph.CO].
[3886] S. Casas, G. K. Karananas, M. Pauly, and J. Rubio, “Scale-invariant alternatives to general relativity. III. The
inflation-dark energy connection,” Phys. Rev. D 99 (2019) no. 6, 063512, arXiv:1811.05984 [astro-ph.CO].
[3887] Planck Collaboration, Y. Akrami et al., “Planck 2018 results. X. Constraints on inflation,” Astron. Astrophys. 641
(2020) A10, arXiv:1807.06211 [astro-ph.CO].
[3888] BICEP, Keck Collaboration, P. A. R. Ade et al., “Improved Constraints on Primordial Gravitational Waves using
Planck, WMAP, and BICEP/Keck Observations through the 2018 Observing Season,” Phys. Rev. Lett. 127 (2021)
no. 15, 151301, arXiv:2110.00483 [astro-ph.CO].
390

[3889] W. Giarè, “Inflation, the Hubble tension, and early dark energy: An alternative overview,” Phys. Rev. D 109 (2024)
no. 12, 123545, arXiv:2404.12779 [astro-ph.CO].
[3890] E. Silverstein and D. Tong, “Scalar speed limits and cosmology: Acceleration from D-cceleration,” Phys. Rev. D 70
(2004) 103505, arXiv:hep-th/0310221.
[3891] M. Alishahiha, E. Silverstein, and D. Tong, “DBI in the sky,” Phys. Rev. D 70 (2004) 123505, arXiv:hep-th/0404084.
[3892] X. Chen, “Inflation from warped space,” JHEP 08 (2005) 045, arXiv:hep-th/0501184.
[3893] X. Chen, “Running non-Gaussianities in DBI inflation,” Phys. Rev. D 72 (2005) 123518, arXiv:astro-ph/0507053.
[3894] H. V. Peiris, D. Baumann, B. Friedman, and A. Cooray, “Phenomenology of D-Brane Inflation with General Speed of
Sound,” Phys. Rev. D 76 (2007) 103517, arXiv:0706.1240 [astro-ph].
[3895] A. Ashoorioon, H. Firouzjahi, and M. M. Sheikh-Jabbari, “M-flation: Inflation From Matrix Valued Scalar Fields,”
JCAP 06 (2009) 018, arXiv:0903.1481 [hep-th].
[3896] D. D. Dimitrijevic, N. Bilić, G. S. Djordjevic, M. Milosevic, and M. Stojanovic, “Tachyon scalar field in a braneworld
cosmology,” Int. J. Mod. Phys. A 33 (2018) no. 34, 1845017.
[3897] N. Bilic, S. Domazet, and G. S. Djordjevic, “Particle creation and reheating in a braneworld inflationary scenario,”
Phys. Rev. D 96 (2017) no. 8, 083518, arXiv:1707.06023 [hep-th].
[3898] N. Bilic, D. Dimitrijevic, G. Djordjevic, and M. Milosevic, “Tachyon inflation in an AdS braneworld with
backreaction,” Int. J. Mod. Phys. A 32 (2017) no. 05, 1750039, arXiv:1607.04524 [gr-qc].
[3899] C.-M. Lin, “D-term inflation in braneworld models: Consistency with cosmic-string bounds and early-time Hubble
tension resolving models,” Phys. Rev. D 106 (2022) no. 10, 103511, arXiv:2204.10475 [hep-th].
[3900] L. Lombriser, “On the cosmological constant problem,” Phys. Lett. B 797 (2019) 134804, arXiv:1901.08588 [gr-qc].
[3901] J. Frieman, M. Turner, and D. Huterer, “Dark Energy and the Accelerating Universe,” Ann. Rev. Astron. Astrophys.
46 (2008) 385–432, arXiv:0803.0982 [astro-ph].
[3902] M. Park, M. Raveri, and B. Jain, “Reconstructing Quintessence,” Phys. Rev. D 103 (2021) no. 10, 103530,
arXiv:2101.04666 [astro-ph.CO].
[3903] S. Goldstein, M. Park, M. Raveri, B. Jain, and L. Samushia, “Beyond dark energy Fisher forecasts: How the Dark
Energy Spectroscopic Instrument will constrain LCDM and quintessence models,” Phys. Rev. D 107 (2023) no. 6,
063530, arXiv:2207.01612 [astro-ph.CO].
[3904] L. A. Anchordoqui, I. Antoniadis, and D. Lust, “S-dual Quintessence, the Swampland, and the DESI DR2 Results,”
arXiv:2503.19428 [hep-th].
[3905] P. J. E. Peebles and A. Vilenkin, “Quintessential inflation,” Phys. Rev. D 59 (1999) 063505, arXiv:astro-ph/9810509.
[3906] M. Wali Hossain, R. Myrzakulov, M. Sami, and E. N. Saridakis, “Unification of inflation and dark energy à la
quintessential inflation,” Int. J. Mod. Phys. D 24 (2015) no. 05, 1530014, arXiv:1410.6100 [gr-qc].
[3907] J. de Haro, J. Amorós, and S. Pan, “Simple inflationary quintessential model,” Phys. Rev. D 93 (2016) no. 8, 084018,
arXiv:1601.08175 [gr-qc].
[3908] J. de Haro, J. Amorós, and S. Pan, “Simple inflationary quintessential model II: Power law potentials,” Phys. Rev. D
94 (2016) no. 6, 064060, arXiv:1607.06726 [gr-qc].
[3909] J. Haro, W. Yang, and S. Pan, “Reheating in quintessential inflation via gravitational production of heavy massive
particles: A detailed analysis,” JCAP 01 (2019) 023, arXiv:1811.07371 [gr-qc].
[3910] J. Haro, J. Amorós, and S. Pan, “The Peebles - Vilenkin quintessential inflation model revisited,” Eur. Phys. J. C 79
(2019) no. 6, 505, arXiv:1901.00167 [gr-qc].
[3911] J. Haro, J. Amorós, and S. Pan, “Scaling solutions in quintessential inflation,” Eur. Phys. J. C 80 (2020) no. 5, 404,
arXiv:1908.01516 [gr-qc].
[3912] K. Dimopoulos and J. W. F. Valle, “Modeling quintessential inflation,” Astropart. Phys. 18 (2002) 287–306,
arXiv:astro-ph/0111417.
[3913] C.-Q. Geng, M. W. Hossain, R. Myrzakulov, M. Sami, and E. N. Saridakis, “Quintessential inflation with canonical and
noncanonical scalar fields and Planck 2015 results,” Phys. Rev. D 92 (2015) no. 2, 023522, arXiv:1502.03597 [gr-qc].
[3914] K. Dimopoulos and C. Owen, “Quintessential Inflation with α-attractors,” JCAP 06 (2017) 027, arXiv:1703.00305
[gr-qc].
[3915] C.-Q. Geng, C.-C. Lee, M. Sami, E. N. Saridakis, and A. A. Starobinsky, “Observational constraints on successful
model of quintessential Inflation,” JCAP 06 (2017) 011, arXiv:1705.01329 [gr-qc].
[3916] M. W. Hossain, R. Myrzakulov, M. Sami, and E. N. Saridakis, “Variable gravity: A suitable framework for
quintessential inflation,” Phys. Rev. D 90 (2014) no. 2, 023512, arXiv:1402.6661 [gr-qc].
[3917] M. W. Hossain, R. Myrzakulov, M. Sami, and E. N. Saridakis, “Class of quintessential inflation models with parameter
space consistent with BICEP2,” Phys. Rev. D 89 (2014) no. 12, 123513, arXiv:1404.1445 [gr-qc].
[3918] J. Rubio and C. Wetterich, “Emergent scale symmetry: Connecting inflation and dark energy,” Phys. Rev. D 96 (2017)
no. 6, 063509, arXiv:1705.00552 [gr-qc].
[3919] D. Bettoni and J. Rubio, “Quintessential Inflation: A Tale of Emergent and Broken Symmetries,” Galaxies 10 (2022)
no. 1, 22, arXiv:2112.11948 [astro-ph.CO].
[3920] Y. Akrami, S. Casas, S. Deng, and V. Vardanyan, “Quintessential α-attractor inflation: forecasts for Stage IV galaxy
surveys,” JCAP 04 (2021) 006, arXiv:2010.15822 [astro-ph.CO].
[3921] W. Giarè, E. Di Valentino, E. V. Linder, and E. Specogna, “Testing α-attractor quintessential inflation against CMB
and low-redshift data,” Phys. Dark Univ. 46 (2024) 101713, arXiv:2402.01560 [astro-ph.CO].
[3922] D. Benisty and E. I. Guendelman, “Lorentzian Quintessential Inflation,” Int. J. Mod. Phys. D 29 (2020) no. 14,
2042002, arXiv:2004.00339 [astro-ph.CO].
391

[3923] D. Benisty and E. I. Guendelman, “Quintessential Inflation from Lorentzian Slow Roll,” Eur. Phys. J. C 80 (2020)
no. 6, 577, arXiv:2006.04129 [astro-ph.CO].
[3924] L. Aresté Saló, D. Benisty, E. I. Guendelman, and J. d. Haro, “Quintessential inflation and cosmological seesaw
mechanism: reheating and observational constraints,” JCAP 07 (2021) 007, arXiv:2102.09514 [astro-ph.CO].
[3925] L. Aresté Saló, D. Benisty, E. I. Guendelman, and J. de Haro, “α-attractors in quintessential inflation motivated by
supergravity,” Phys. Rev. D 103 (2021) no. 12, 123535, arXiv:2103.07892 [astro-ph.CO].
[3926] A. D. Rendall, “Dynamics of k-essence,” Class. Quant. Grav. 23 (2006) 1557–1570, arXiv:gr-qc/0511158.
[3927] D. Langlois and S. Renaux-Petel, “Perturbations in generalized multi-field inflation,” JCAP 04 (2008) 017,
arXiv:0801.1085 [hep-th].
[3928] A. E. Romano, S. A. Vallejo-Peña, and K. Turzyński, “Model-independent approach to effective sound speed in
multi-field inflation,” Eur. Phys. J. C 82 (2022) no. 8, 767, arXiv:2006.00969 [gr-qc].
[3929] S. A. Hosseini Mansoori and H. Moshafi, “Alleviating H 0 and S 8 Tensions Simultaneously in K-essence Cosmology,”
Astrophys. J. 975 (2024) no. 2, 275, arXiv:2405.05843 [astro-ph.CO].
[3930] S. X. Tian and Z.-H. Zhu, “Early dark energy in k-essence,” Phys. Rev. D 103 (2021) no. 4, 043518, arXiv:2102.06399
[gr-qc].
[3931] A. Jawad, S. Rani, A. M. Sultan, and K. Embreen, “k-Essence Inflation Evading Swampland Conjectures and
Inflationary Parameters,” Universe 8 (2022) no. 10, 532.
[3932] S. Hussain, S. Nelleri, and K. Bhattacharya, “Comprehensive study of k-essence model: dynamical system analysis and
observational constraints from latest Type Ia supernova and BAO observations,” JCAP 03 (2025) 025,
arXiv:2406.07179 [astro-ph.CO].
[3933] W. Giarè, M. De Angelis, C. van de Bruck, and E. Di Valentino, “Tracking the multifield dynamics with cosmological
data: a Monte Carlo approach,” JCAP 12 (2023) 014, arXiv:2306.12414 [astro-ph.CO].
[3934] D. Staicova, “Special cases of the multi-measure model – Understanding the prolonged inflation,” JHEAp 36 (2022)
120–127, arXiv:2011.02967 [gr-qc].
[3935] D. Staicova and M. Stoilov, “Electromagnetic Waves in Cosmological Space-Time II. Luminosity Distance,” Universe
11 (2025) 50, arXiv:2502.11634 [gr-qc].
[3936] I. Antoniadis, J. Cunat, and A. Guillen, “Cosmological perturbations from five-dimensional inflation,” JHEP 05 (2024)
290, arXiv:2311.17680 [hep-ph].
[3937] C. Petretti, M. Braglia, X. Chen, D. K. Hazra, and S. Paban, “Investigating the Origin of CMB Large-Scale Features
Using LiteBIRD and CMB-S4,” arXiv:2411.03459 [astro-ph.CO].
[3938] L. A. Anchordoqui and I. Antoniadis, “Primordial Power Spectrum of Five Dimensional Uniform Inflation,”
arXiv:2412.19213 [astro-ph.CO].
[3939] T. Hirose, “Analysis of inflationary models in higher-dimensional uniform inflation,” arXiv:2501.13581 [hep-ph].
[3940] C. Moreno-Pulido and J. Sola Peracaula, “Equation of state of the running vacuum,” Eur. Phys. J. C 82 (2022) no. 12,
1137, arXiv:2207.07111 [gr-qc].
[3941] J. A. S. Lima, S. Basilakos, and J. Solà, “Thermodynamical aspects of running vacuum models,” Eur. Phys. J. C 76
(2016) no. 4, 228, arXiv:1509.00163 [gr-qc].
[3942] M. B. Green, J. H. Schwarz, and E. Witten, Superstring Theory Vol. 1: 25th Anniversary Edition. Cambridge
Monographs on Mathematical Physics. Cambridge University Press, 2012.
[3943] M. B. Green and J. H. Schwarz, “Anomaly Cancellation in Supersymmetric D=10 Gauge Theory and Superstring
Theory,” Phys. Lett. B 149 (1984) 117–122.
[3944] S. Basilakos, N. E. Mavromatos, and J. Solà, “Starobinsky-like inflation and running vacuum in the context of
Supergravity,” Universe 2 (2016) no. 3, 14, arXiv:1505.04434 [gr-qc].
[3945] R. Brandenberger, L. L. Graef, G. Marozzi, and G. P. Vacca, “Backreaction of super-Hubble cosmological
perturbations beyond perturbation theory,” Phys. Rev. D 98 (2018) no. 10, 103523, arXiv:1807.07494 [hep-th].
[3946] V. Comeau and R. Brandenberger, “Back-reaction of long-wavelength cosmological fluctuations as measured by a clock
field,” Eur. Phys. J. C 84 (2024) no. 3, 272, arXiv:2302.05873 [gr-qc].
[3947] D. W. Hogg, D. J. Eisenstein, M. R. Blanton, N. A. Bahcall, J. Brinkmann, J. E. Gunn, and D. P. Schneider, “Cosmic
homogeneity demonstrated with luminous red galaxies,” Astrophys. J. 624 (2005) 54–58, arXiv:astro-ph/0411197.
[3948] J. Yadav, S. Bharadwaj, B. Pandey, and T. R. Seshadri, “Testing homogeneity on large scales in the Sloan Digital Sky
Survey Data Release One,” Mon. Not. Roy. Astron. Soc. 364 (2005) 601–606, arXiv:astro-ph/0504315.
[3949] M. Scrimgeour et al., “The WiggleZ Dark Energy Survey: the transition to large-scale cosmic homogeneity,” Mon. Not.
Roy. Astron. Soc. 425 (2012) 116–134, arXiv:1205.6812 [astro-ph.CO].
[3950] J. K. Yadav, J. S. Bagla, and N. Khandai, “Fractal Dimension as a measure of the scale of Homogeneity,” Mon. Not.
Roy. Astron. Soc. 405 (2010) 2009, arXiv:1001.0617 [astro-ph.CO].
[3951] I. Horvath, J. Hakkila, and Z. Bagoly, “Possible structure in the GRB sky distribution at redshift two,” Astron.
Astrophys. 561 (2014) L12, arXiv:1401.0533 [astro-ph.CO].
[3952] L. G. Balazs, Z. Bagoly, J. E. Hakkila, I. Horvath, J. Kobori, I. Racz, and L. V. Toth, “A giant ring-like structure at
0.78 < z < 0.86 displayed by GRBs,” Mon. Not. Roy. Astron. Soc. 452 (2015) no. 3, 2236–2246, arXiv:1507.00675
[astro-ph.CO].
[3953] A. M. Lopez, R. G. Clowes, and G. M. Williger, “A Giant Arc on the Sky,” Mon. Not. Roy. Astron. Soc. 516 (2022)
no. 2, 1557–1572, arXiv:2201.06875 [astro-ph.CO].
[3954] A. M. Lopez, R. G. Clowes, and G. M. Williger, “A Big Ring on the sky,” JCAP 07 (2024) 055, arXiv:2402.07591
[astro-ph.CO].
392

[3955] S. Nadathur, “Seeing patterns in noise: Gigaparsec-scale ‘structures’ that do not violate homogeneity,” Mon. Not. Roy.
Astron. Soc. 434 (2013) 398–406, arXiv:1306.1700 [astro-ph.CO].
[3956] T. N. Ukwatta and P. R. Wozniak, “Investigation of Redshift- and Duration-Dependent Clustering of Gamma-ray
Bursts,” Mon. Not. Roy. Astron. Soc. 455 (2016) no. 1, 703–711, arXiv:1507.07117 [astro-ph.HE].
[3957] T. Sawala, M. Teeriaho, C. S. Frenk, J. Helly, A. Jenkins, G. Racz, M. Schaller, and J. Schaye, “The Emperor’s New
Arc: gigaparsec patterns abound in a ΛCDM universe,” arXiv e-prints (2025) arXiv:2502.03515, arXiv:2502.03515
[astro-ph.CO].
[3958] M. L. McClure and C. C. Dyer, “Anisotropy in the Hubble constant as observed in the HST Extragalactic Distance
Scale Key Project results,” New Astron. 12 (2007) 533–543, arXiv:astro-ph/0703556.
[3959] K. Migkas, F. Pacaud, G. Schellenberger, J. Erler, N. T. Nguyen-Dang, T. H. Reiprich, M. E. Ramos-Ceja, and
L. Lovisari, “Cosmological implications of the anisotropy of ten galaxy cluster scaling relations,” Astron. Astrophys.
649 (2021) A151, arXiv:2103.13904 [astro-ph.CO].
[3960] H. Bondi and T. Gold, “The Steady-State Theory of the Expanding Universe,” Mon. Not. Roy. Astron. Soc. 108
(1948) 252.
[3961] F. Hoyle, “A New Model for the Expanding Universe,” Mon. Not. Roy. Astron. Soc. 108 (1948) 372–382.
[3962] D. Benisty and E. I. Guendelman, “Cosmological Principle in Newtonian Dynamics,” Mod. Phys. Lett. A 35 (2020)
no. 16, 2050131, arXiv:1902.06511 [gr-qc].
[3963] E. I. Guendelman, E. Zamlung, and D. Benisty, “Noether symmetry in Newtonian dynamics and cosmology,” Gen. Rel.
Grav. 53 (2021) no. 11, 99, arXiv:2010.06448 [gr-qc].
[3964] K. Migkas, G. Schellenberger, T. H. Reiprich, F. Pacaud, M. E. Ramos-Ceja, and L. Lovisari, “Probing cosmic isotropy
with a new X-ray galaxy cluster sample through the LX − T scaling relation,” Astron. Astrophys. 636 (2020) A15,
arXiv:2004.03305 [astro-ph.CO].
[3965] A. Pandya, K. Migkas, T. H. Reiprich, A. Stanford, F. Pacaud, G. Schellenberger, L. Lovisari, M. E. Ramos-Ceja,
N. T. Nguyen-Dang, and S. Park, “Examining the local Universe isotropy with galaxy cluster velocity dispersion
scaling relations,” Astron. Astrophys. 691 (2024) A355, arXiv:2408.00726 [astro-ph.CO].
[3966] Z. Zhai and W. J. Percival, “Sample variance for supernovae distance measurements and the Hubble tension,” Phys.
Rev. D 106 (2022) no. 10, 103527, arXiv:2207.02373 [astro-ph.CO].
[3967] R. Cooke and D. Lynden-Bell, “Does the Universe Accelerate Equally in all Directions?,” Mon. Not. Roy. Astron. Soc.
401 (2010) 1409–1414, arXiv:0909.3861 [astro-ph.CO].
[3968] F. Sorrenti, R. Durrer, and M. Kunz, “The low multipoles in the Pantheon+SH0ES data,” arXiv:2403.17741
[astro-ph.CO].
[3969] F. Sorrenti, R. Durrer, and M. Kunz, “A local infall from a cosmographic analysis of Pantheon+,” JCAP 12 (2024)
003, arXiv:2407.07002 [astro-ph.CO].
[3970] P. Boubel, M. Colless, K. Said, and L. Staveley-Smith, “Testing anisotropic Hubble expansion,” arXiv e-prints (2024)
arXiv:2412.14607, arXiv:2412.14607 [astro-ph.CO].
[3971] V. Yadav, “Measuring Hubble constant in an anisotropic extension of ΛCDM model,” Phys. Dark Univ. 42 (2023)
101365, arXiv:2306.16135 [astro-ph.CO].
[3972] Planck Collaboration, P. A. R. Ade et al., “Planck intermediate results. XIII. Constraints on peculiar velocities,”
Astron. Astrophys. 561 (2014) A97, arXiv:1303.5090 [astro-ph.CO].
[3973] A. Kashlinsky, F. Atrio-Barandela, D. Kocevski, and H. Ebeling, “A measurement of large-scale peculiar velocities of
clusters of galaxies: technical details,” Astrophys. J. 691 (2009) 1479–1493, arXiv:0809.3733 [astro-ph].
[3974] T. Westmeier, N. Deg, K. Spekkens, T. N. Reynolds, A. X. Shen, S. Gaudet, S. Goliath, M. T. Huynh,
P. Venkataraman, X. Lin, T. O’Beirne, B. Catinella, L. Cortese, H. Dénes, A. Elagali, B. Q. For, G. I. G. Józsa,
C. Howlett, J. M. van der Hulst, R. J. Jurek, P. Kamphuis, V. A. Kilborn, D. Kleiner, B. S. Koribalski,
K. Lee-Waddell, C. Murugeshan, J. Rhee, P. Serra, L. Shao, L. Staveley-Smith, J. Wang, O. I. Wong, M. A. Zwaan,
J. R. Allison, C. S. Anderson, L. Ball, D. C. J. Bock, D. Brodrick, J. D. Bunton, F. R. Cooray, N. Gupta, D. B.
Hayman, E. K. Mahony, V. A. Moss, A. Ng, S. E. Pearce, W. Raja, D. N. Roxby, M. A. Voronkov, K. A. Warhurst,
H. M. Courtois, and K. Said, “WALLABY pilot survey: Public release of H I data for almost 600 galaxies from phase 1
of ASKAP pilot observations,” Publ. Astron. Soc. Aust. 39 (2022) e058, arXiv:2211.07094 [astro-ph.GA].
[3975] C. Murugeshan, N. Deg, T. Westmeier, A. X. Shen, B. Q. For, K. Spekkens, O. I. Wong, L. Staveley-Smith,
B. Catinella, K. Lee-Waddell, H. Dénes, J. Rhee, L. Cortese, S. Goliath, R. Halloran, J. M. van der Hulst,
P. Kamphuis, B. S. Koribalski, R. C. Kraan-Korteweg, F. Lelli, P. Venkataraman, L. Verdes-Montenegro, and N. Yu,
“WALLABY Pilot Survey: Public data release of ∼ 1800 H I sources and high-resolution cut-outs from Pilot Survey
Phase 2,” Publ. Astron. Soc. Aust. 41 (2024) e088, arXiv:2409.13130 [astro-ph.GA].
[3976] H. M. Courtois, J. Mould, A. M. Hollinger, A. Dupuy, and C.-P. Zhang, “In search for the Local Universe dynamical
homogeneity scale with CF4++ peculiar velocities,” arXiv:2502.01308 [astro-ph.CO].
[3977] M. Rameez, “Concerns about the reliability of publicly available SNe Ia data,” arXiv:1905.00221 [astro-ph.CO].
[3978] M. Rameez and S. Sarkar, “Is there really a Hubble tension?,” Class. Quant. Grav. 38 (2021) no. 15, 154005,
arXiv:1911.06456 [astro-ph.CO].
[3979] C. A. P. Bengaly, R. Maartens, and M. G. Santos, “Probing the Cosmological Principle in the counts of radio galaxies
at different frequencies,” JCAP 04 (2018) 031, arXiv:1710.08804 [astro-ph.CO].
[3980] A. K. Singal, “Discordance of dipole asymmetries seen in recent large radio surveys with the cosmological principle,”
Mon. Not. Roy. Astron. Soc. 524 (2023) no. 3, 3636–3646, arXiv:2303.05141 [astro-ph.CO].
[3981] J. Darling, “The Universe is Brighter in the Direction of Our Motion: Galaxy Counts and Fluxes are Consistent with
393

the CMB Dipole,” Astrophys. J. Lett. 931 (2022) no. 2, L14, arXiv:2205.06880 [astro-ph.CO].
[3982] J. D. Wagenveld et al., “The MeerKAT Absorption Line Survey Data Release 2: Wideband continuum catalogues and
a measurement of the cosmic radio dipole,” Astron. Astrophys. 690 (2024) A163, arXiv:2408.16619 [astro-ph.CO].
[3983] P. Tiwari and A. Nusser, “Revisiting the NVSS number count dipole,” JCAP 03 (2016) 062, arXiv:1509.02532
[astro-ph.CO].
[3984] C. Dalang and C. Bonvin, “On the kinematic cosmic dipole tension,” Mon. Not. Roy. Astron. Soc. 512 (2022) no. 3,
3895–3905, arXiv:2111.03616 [astro-ph.CO].
[3985] S. von Hausegger, “The expected kinematic matter dipole is robust against source evolution,” Mon. Not. Roy. Astron.
Soc. 535 (2024) no. 1, L49–L53, arXiv:2404.07929 [astro-ph.CO].
[3986] L. Giani, C. Howlett, K. Said, T. Davis, and S. Vagnozzi, “An effective description of Laniakea: impact on cosmology
and the local determination of the Hubble constant,” JCAP 01 (2024) 071, arXiv:2311.00215 [astro-ph.CO].
[3987] H. Boehringer, G. Chon, J. Truemper, R. C. Kraan-Korteweg, and N. Schartel, “Unveiling the largest structures in the
nearby Universe: Discovery of the Quipu superstructure,” arXiv e-prints (2025) arXiv:2501.19236, arXiv:2501.19236
[astro-ph.CO].
[3988] B. Javanmardi, C. Porciani, P. Kroupa, and J. Pflamm-Altenburg, “Probing the isotropy of cosmic acceleration traced
by Type Ia supernovae,” Astrophys. J. 810 (2015) no. 1, 47, arXiv:1507.07560 [astro-ph.CO].
[3989] J. Hu, J. Hu, X. Jia, B. Gao, and F. Wang, “Testing cosmic anisotropy with Pade approximation and Pantheon+
sample,” Astron. Astrophys. 689 (2024) A215, arXiv:2406.14827 [astro-ph.CO].
[3990] Planck Collaboration, N. Aghanim et al., “Planck 2013 results. XXVII. Doppler boosting of the CMB: Eppur si
muove,” Astron. Astrophys. 571 (2014) A27, arXiv:1303.5087 [astro-ph.CO].
[3991] P. d. S. Ferreira and M. Quartin, “Disentangling Doppler modulation, aberration and the temperature dipole in the
CMB,” Phys. Rev. D 104 (2021) no. 6, 063503, arXiv:2107.10846 [astro-ph.CO].
[3992] S. Saha, S. Shaikh, S. Mukherjee, T. Souradeep, and B. D. Wandelt, “Bayesian estimation of our local motion from the
Planck-2018 CMB temperature map,” JCAP 10 (2021) 072, arXiv:2106.07666 [astro-ph.CO].
[3993] T. R. Jaffe, A. J. Banday, H. K. Eriksen, K. M. Gorski, and F. K. Hansen, “Evidence of vorticity and shear at large
angular scales in the WMAP data: A Violation of cosmological isotropy?,” Astrophys. J. Lett. 629 (2005) L1–L4,
arXiv:astro-ph/0503213.
[3994] T. R. Jaffe, S. Hervik, A. J. Banday, and K. M. Gorski, “On the viability of Bianchi type viih models with dark
energy,” Astrophys. J. 644 (2006) 701–708, arXiv:astro-ph/0512433.
[3995] M. Bridges, J. D. McEwen, A. N. Lasenby, and M. P. Hobson, “Markov chain Monte Carlo analysis of Bianchi VII(h)
models,” Mon. Not. Roy. Astron. Soc. 377 (2007) 1473–1480, arXiv:astro-ph/0605325.
[3996] Planck Collaboration, P. A. R. Ade et al., “Planck 2013 results. XXVI. Background geometry and topology of the
Universe,” Astron. Astrophys. 571 (2014) A26, arXiv:1303.5086 [astro-ph.CO].
[3997] L. Bianchi, “Sugli spazi a tre dimensioni che ammettono un gruppo continuo di movimenti.,” Soc. Ital. Sci. Mem. di
Mat. 11 (1898) 267–352.
[3998] L. Bianchi, “On the Three-Dimensional Spaces Which Admit a Continuous Group of Motions.,” General Relativity and
Gravitation 33 (2001) 2171–2253.
[3999] O. Akarsu, S. Kumar, S. Sharma, and L. Tedesco, “Constraints on a Bianchi type I spacetime extension of the
standard ΛCDM model,” Phys. Rev. D 100 (2019) no. 2, 023532, arXiv:1905.06949 [astro-ph.CO].
[4000] H. Stephani, D. Kramer, M. A. H. MacCallum, C. Hoenselaers, and E. Herlt, Exact solutions of Einstein’s field
equations. Cambridge Monographs on Mathematical Physics. Cambridge Univ. Press, Cambridge, 2003.
[4001] A. Pontzen and A. Challinor, “Bianchi Model CMB Polarization and its Implications for CMB Anomalies,” Mon. Not.
Roy. Astron. Soc. 380 (2007) 1387–1398, arXiv:0706.2075 [astro-ph].
[4002] E. Russell, C. B. Kılınç, and O. K. Pashaev, “Bianchi I model: an alternative way to model the present-day Universe,”
Mon. Not. Roy. Astron. Soc. 442 (2014) no. 3, 2331–2341, arXiv:1312.3502 [astro-ph.CO].
[4003] G. Lemaitre, “The expanding universe,” Annales Soc. Sci. Bruxelles A 53 (1933) 51–85.
[4004] R. C. Tolman, “Effect of imhomogeneity on cosmological models,” Proc. Nat. Acad. Sci. 20 (1934) 169–176.
[4005] H. Bondi, “Spherically symmetrical models in general relativity,” Mon. Not. Roy. Astron. Soc. 107 (1947) 410–425.
[4006] V. Marra, T. Castro, D. Camarena, S. Borgani, and A. Ragagnin, “The BEHOMO project: Λ Lemaître-Tolman-Bondi
N-body simulations,” Astron. Astrophys. 664 (2022) A179, arXiv:2203.04009 [astro-ph.CO].
[4007] M. Barriola and A. Vilenkin, “Gravitational Field of a Global Monopole,” Phys. Rev. Lett. 63 (1989) 341.
[4008] L. Perivolaropoulos, “Six Puzzles for LCDM Cosmology,” arXiv:0811.4684 [astro-ph].
[4009] J. C. B. Sanchez and L. Perivolaropoulos, “Evolution of Dark Energy Perturbations in Scalar-Tensor Cosmologies,”
Phys. Rev. D 81 (2010) 103505, arXiv:1002.2042 [astro-ph.CO].
[4010] L. Perivolaropoulos, “Large Scale Cosmological Anomalies and Inhomogeneous Dark Energy,” Galaxies 2 (2014) 22–61,
arXiv:1401.5044 [astro-ph.CO].
[4011] G. F. R. Ellis, “The Bianchi models: Then and now,” Gen. Rel. Grav. 38 (2006) 1003–1015.
[4012] T. Schucker, A. Tilquin, and G. Valent, “Bianchi I meets the Hubble diagram,” Mon. Not. Roy. Astron. Soc. 444
(2014) no. 3, 2820–2836, arXiv:1405.6523 [astro-ph.CO].
[4013] G. Valent, “Bianchi type II,III and V diagonal Einstein metrics re-visited,” Gen. Rel. Grav. 41 (2009) 2433–2459,
arXiv:1002.1454 [math-ph].
[4014] D. H. King, “Gravity wave insights to Bianchi type IX universes,” Phys. Rev. D 44 (1991) 2356–2368.
[4015] H. Ringstrom, “The Bianchi IX attractor,” Annales Henri Poincare 2 (2001) 405–500, arXiv:gr-qc/0006035.
[4016] A. Ashtekar and J. Samuel, “Bianchi cosmologies: The Role of spatial topology,” Class. Quant. Grav. 8 (2011) 2191.
394

[4017] G. F. R. Ellis and M. A. H. MacCallum, “A Class of homogeneous cosmological models,” Commun. Math. Phys. 12
(1969) 108–141.
[4018] P. Cea, “The Ellipsoidal Universe in the Planck Satellite Era,” Mon. Not. Roy. Astron. Soc. 441 (2014) no. 2,
1646–1661, arXiv:1401.5627 [astro-ph.CO].
[4019] T. Koivisto and D. F. Mota, “Accelerating Cosmologies with an Anisotropic Equation of State,” Astrophys. J. 679
(2008) 1–5, arXiv:0707.0279 [astro-ph].
[4020] G. C. McVittie, “The mass-particle in an expanding universe,” Mon. Not. Roy. Astron. Soc. 93 (1933) 325–339.
[4021] N. Kaloper, M. Kleban, and D. Martin, “McVittie’s Legacy: Black Holes in an Expanding Universe,” Phys. Rev. D 81
(2010) 104044, arXiv:1003.4777 [hep-th].
[4022] M. Sereno and P. Jetzer, “Evolution of gravitational orbits in the expanding universe,” Phys. Rev. D 75 (2007) 064031,
arXiv:astro-ph/0703121.
[4023] V. Faraoni and A. Jacques, “Cosmological expansion and local physics,” Phys. Rev. D 76 (2007) 063510,
arXiv:0707.1350 [gr-qc].
[4024] R. Nandra, A. N. Lasenby, and M. P. Hobson, “The effect of a massive object on an expanding universe,” Mon. Not.
Roy. Astron. Soc. 422 (2012) 2931–2944, arXiv:1104.4447 [gr-qc].
[4025] D. Benisty, M. M. Chaichian, and A. Tureanu, “Galaxy groups in the presence of cosmological constant: Increasing the
masses of groups,” Phys. Lett. B 858 (2024) 139033, arXiv:2405.14944 [astro-ph.GA].
[4026] S. Peirani and J. A. de Freitas Pacheco, “Mass determination of groups of galaxies: effects of the cosmological
constant,” New Astron. 11 (2006) 325–330, arXiv:astro-ph/0508614.
[4027] S. Peirani and J. A. D. F. Pacheco, “Dynamics of Nearby Groups of Galaxies: the role of the cosmological constant,”
Astron. Astrophys. 488 (2008) 845–851, arXiv:0806.4245 [astro-ph].
[4028] I. D. Karachentsev, O. G. Kashibadze, D. I. Makarov, and R. B. Tully, “The Hubble flow around the Local Group,”
Mon. Not. Roy. Astron. Soc. 393 (2009) 1265, arXiv:0811.4610 [astro-ph].
[4029] J. Peñarrubia, Y.-Z. Ma, M. G. Walker, and A. McConnachie, “A dynamical model of the local cosmic expansion,”
Mon. Not. Roy. Astron. Soc. 443 (2014) no. 3, 2204–2222, arXiv:1405.0306 [astro-ph.GA].
[4030] P. Teerikorpi and A. D. Chernin, “The Hubble diagram for a system within dark energy: the location of the
zero-gravity radius and the global Hubble rate,” Astron. Astrophys. 516 (2010) A93, arXiv:1006.0066 [astro-ph.CO].
[4031] A. Del Popolo, M. Deliyergiyev, and M. H. Chan, “Improved Lemaitre–Tolman model and the mass and turn-around
radius in group of galaxies,” Phys. Dark Univ. 31 (2021) 100780, arXiv:2103.12714 [astro-ph.CO].
[4032] A. Del Popolo and M. H. Chan, “Improved Lemaitre–Tolman Model and the Mass and Turn-around Radius in Group
of Galaxies. II. The Role of Dark Energy,” Astrophys. J. 926 (2022) no. 2, 156, arXiv:2210.10397 [astro-ph.CO].
[4033] P. Szekeres, “Quasispherical Gravitational Collapse,” Phys. Rev. D 12 (1975) 2941.
[4034] M.-N. Célérier, “Precision cosmology with exact inhomogeneous solutions of General Relativity: the Szekeres models,”
arXiv e-prints (2024) arXiv:2407.04452, arXiv:2407.04452 [gr-qc].
[4035] J. Kristian and R. K. Sachs, “Observations in cosmology,” Astrophys. J. 143 (1966) 379–399.
[4036] C. A. Clarkson, “On the observational characteristics of inhomogeneous cosmologies: Undermining the cosmological
principle or have cosmologists put all their EGS in one basket?,” other thesis, Glasgow U., 1999.
[4037] T. Nadolny, R. Durrer, M. Kunz, and H. Padmanabhan, “A new way to test the Cosmological Principle: measuring our
peculiar velocity and the large-scale anisotropy independently,” JCAP 11 (2021) 009, arXiv:2106.05284
[astro-ph.CO].
[4038] D. L. Wiltshire, “Cosmic clocks, cosmic variance and cosmic averages,” New J. Phys. 9 (2007) 377,
arXiv:gr-qc/0702082.
[4039] D. L. Wiltshire, “Average observational quantities in the timescape cosmology,” Phys. Rev. D 80 (2009) 123512,
arXiv:0909.0749 [astro-ph.CO].
[4040] J. A. G. Duley, M. A. Nazer, and D. L. Wiltshire, “Timescape cosmology with radiation fluid,” Class. Quant. Grav. 30
(2013) 175006, arXiv:1306.3208 [astro-ph.CO].
[4041] D. L. Wiltshire, “Cosmic structure, averaging and dark energy,” in Cosmology and Gravitation: XVth Brazilian School
of Cosmology and Gravitation. Cambridge Scientific Publishers, 2014. arXiv:1311.3787 [astro-ph.CO].
[4042] T. Buchert, “On average properties of inhomogeneous fluids in general relativity. 1. Dust cosmologies,” Gen. Rel. Grav.
32 (2000) 105–125, arXiv:gr-qc/9906015.
[4043] T. Buchert, “On average properties of inhomogeneous fluids in general relativity: Perfect fluid cosmologies,” Gen. Rel.
Grav. 33 (2001) 1381–1405, arXiv:gr-qc/0102049.
[4044] T. Buchert, P. Mourier, and X. Roy, “On average properties of inhomogeneous fluids in general relativity III: general
fluid cosmologies,” Gen. Rel. Grav. 52 (2020) no. 3, 27, arXiv:1912.04213 [gr-qc].
[4045] A. Ishibashi and R. M. Wald, “Can the acceleration of our universe be explained by the effects of inhomogeneities?,”
Class. Quant. Grav. 23 (2006) 235–250, arXiv:gr-qc/0509108.
[4046] D. L. Wiltshire, “Cosmological equivalence principle and the weak-field limit,” Phys. Rev. D 78 (2008) 084032,
arXiv:0809.1183 [gr-qc].
[4047] K. Bolejko, M. A. Nazer, and D. L. Wiltshire, “Differential cosmic expansion and the Hubble flow anisotropy,” JCAP
06 (2016) 035, arXiv:1512.07364 [astro-ph.CO].
[4048] D. Sapone, E. Majerotto, and S. Nesseris, “Curvature versus distances: Testing the FLRW cosmology,” Phys. Rev. D
90 (2014) no. 2, 023012, arXiv:1402.2236 [astro-ph.CO].
[4049] C. Clarkson, B. Bassett, and T. H.-C. Lu, “A general test of the Copernican Principle,” Phys. Rev. Lett. 101 (2008)
011301, arXiv:0712.3457 [astro-ph].
395

[4050] DES Collaboration, R. Camilleri et al., “The dark energy survey supernova program: investigating beyond-ΛCDM,”
Mon. Not. Roy. Astron. Soc. 533 (2024) no. 3, 2615–2639, arXiv:2406.05048 [astro-ph.CO].
[4051] A. Heinesen, C. Blake, Y.-Z. Li, and D. L. Wiltshire, “Baryon acoustic oscillation methods for generic curvature:
application to the SDSS-III Baryon Oscillation Spectroscopic Survey,” JCAP 03 (2019) 003, arXiv:1811.11963
[astro-ph.CO].
[4052] A. S. Eddington, “On the instability of Einstein’s spherical world,” MNRAS 90 (1930) 668–678.
[4053] V. M. Slipher, “Spectrographic Observations of Nebulae,” Popular Astronomy 23 (1915) 21–24.
[4054] H. S. Leavitt, “1777 variables in the Magellanic Clouds,” Harvard Obs. Annals 60 (1908) 87–108.
[4055] G. Lemaitre, “A Homogeneous Universe of Constant Mass and Growing Radius Accounting for the Radial Velocity of
Extragalactic Nebulae,” Annales Soc. Sci. Bruxelles A 47 (1927) 49–59.
[4056] E. Hubble, “A relation between distance and radial velocity among extra-galactic nebulae,” Proc. Nat. Acad. Sci. 15
(1929) 168–173.
[4057] D. J. Schwarz, D. Bacon, S. Chen, C. Clarkson, D. Huterer, M. Kunz, R. Maartens, A. Raccanelli, M. Rubart, and
J.-L. Starck, “Testing foundations of modern cosmology with SKA all-sky surveys,” PoS AASKA14 (2015) 032,
arXiv:1501.03820 [astro-ph.CO].
[4058] P. Horava, “Quantum Gravity at a Lifshitz Point,” Phys. Rev. D 79 (2009) 084008, arXiv:0901.3775 [hep-th].
[4059] P. Horava, “Spectral Dimension of the Universe in Quantum Gravity at a Lifshitz Point,” Phys. Rev. Lett. 102 (2009)
161301, arXiv:0902.3657 [hep-th].
[4060] H. Lu, J. Mei, and C. N. Pope, “Solutions to Horava Gravity,” Phys. Rev. Lett. 103 (2009) 091301, arXiv:0904.1595
[hep-th].
[4061] G. Calcagni, “Cosmology of the Lifshitz universe,” JHEP 09 (2009) 112, arXiv:0904.0829 [hep-th].
[4062] C. Charmousis, G. Niz, A. Padilla, and P. M. Saffin, “Strong coupling in Horava gravity,” JHEP 08 (2009) 070,
arXiv:0905.2579 [hep-th].
[4063] R. Brandenberger, “Matter Bounce in Horava-Lifshitz Cosmology,” Phys. Rev. D 80 (2009) 043516, arXiv:0904.2835
[hep-th].
[4064] T. P. Sotiriou, M. Visser, and S. Weinfurtner, “Quantum gravity without Lorentz invariance,” JHEP 10 (2009) 033,
arXiv:0905.2798 [hep-th].
[4065] R.-G. Cai, L.-M. Cao, and N. Ohta, “Topological Black Holes in Horava-Lifshitz Gravity,” Phys. Rev. D 80 (2009)
024003, arXiv:0904.3670 [hep-th].
[4066] G. Panotopoulos, D. Vernieri, and I. Lopes, “Quark stars with isotropic matter in Hořava gravity and Einstein–æther
theory,” Eur. Phys. J. C 80 (2020) no. 6, 537, arXiv:2006.07652 [gr-qc].
[4067] D. Vernieri, “Anisotropic fluid spheres in Hořava gravity and Einstein-æther theory with a nonstatic æther,” Phys. Rev.
D 100 (2019) no. 10, 104021, arXiv:1906.07738 [gr-qc].
[4068] T. P. Sotiriou, I. Vega, and D. Vernieri, “Rotating black holes in three-dimensional Hořava gravity,” Phys. Rev. D 90
(2014) no. 4, 044046, arXiv:1405.3715 [gr-qc].
[4069] G. Leon and A. Paliathanasis, “Extended phase-space analysis of the Hořava–Lifshitz cosmology,” Eur. Phys. J. C 79
(2019) no. 9, 746, arXiv:1902.09961 [gr-qc].
[4070] E. Di Valentino, N. A. Nilsson, and M.-I. Park, “A new test of dynamical dark energy models and cosmic tensions in
Hořava gravity,” Mon. Not. Roy. Astron. Soc. 519 (2023) no. 4, 5043–5058, arXiv:2212.07683 [astro-ph.CO].
[4071] N. A. Nilsson, “Preferred-frame effects, the H0 tension, and probes of Hořava–Lifshitz gravity,” Eur. Phys. J. Plus 135
(2020) no. 4, 361, arXiv:1910.14414 [gr-qc].
[4072] S. M. Carroll and E. A. Lim, “Lorentz-violating vector fields slow the universe down,” Phys. Rev. D 70 (2004) 123525,
arXiv:hep-th/0407149.
[4073] R. Alves Batista et al., “EuCAPT White Paper: Opportunities and Challenges for Theoretical Astroparticle Physics in
the Next Decade,” arXiv:2110.10074 [astro-ph.HE].
[4074] D. Amati, M. Ciafaloni, and G. Veneziano, “Can Space-Time Be Probed Below the String Size?,” Phys. Lett. B 216
(1989) 41–47.
[4075] M. Maggiore, “A Generalized uncertainty principle in quantum gravity,” Phys. Lett. B 304 (1993) 65–69,
arXiv:hep-th/9301067.
[4076] C. Quesne and V. M. Tkachuk, “Lorentz-covariant deformed algebra with minimal length,” Czech. J. Phys. 56 (2006)
1269–1274, arXiv:quant-ph/0612093.
[4077] C. Quesne and V. M. Tkachuk, “Lorentz-covariant deformed algebra with minimal length and application to the
1+1-dimensional Dirac oscillator,” J. Phys. A 39 (2006) 10909–10922, arXiv:quant-ph/0604118.
[4078] G. Lambiase and F. Scardigli, “Lorentz violation and generalized uncertainty principle,” Phys. Rev. D 97 (2018) no. 7,
075003, arXiv:1709.00637 [hep-th].
[4079] Q. G. Bailey and V. A. Kostelecky, “Signals for Lorentz violation in post-Newtonian gravity,” Phys. Rev. D 74 (2006)
045001, arXiv:gr-qc/0603030.
[4080] S. Capozziello, M. Benetti, and A. D. A. M. Spallicci, “Addressing the cosmological H0 tension by the Heisenberg
uncertainty,” Found. Phys. 50 (2020) no. 9, 893–899, arXiv:2007.00462 [gr-qc].
[4081] H. Moradpour, S. Aghababaei, C. Corda, and N. Sadeghnezhad, “H0 tension and uncertainty principles,” Phys. Scripta
97 (2022) no. 5, 055008.
[4082] K. N. Abazajian et al., “Inflation Physics from the Cosmic Microwave Background and Large Scale Structure,”
Astropart. Phys. 63 (2015) 55–65, arXiv:1309.5381 [astro-ph.CO].
[4083] A. Kaya, “The imprint of primordial gravitational waves on the CMB intensity profile,” Phys. Lett. B 817 (2021)
396

136353, arXiv:2105.02236 [astro-ph.CO].


[4084] S. M. M. Rasouli and P. Vargas Moniz, “Noncommutative minisuperspace, gravity-driven acceleration, and kinetic
inflation,” Phys. Rev. D 90 (2014) no. 8, 083533, arXiv:1411.1346 [gr-qc].
[4085] F. G. Alvarenga, J. C. Fabris, N. A. Lemos, and G. A. Monerat, “Quantum cosmological perfect fluid models,” Gen.
Rel. Grav. 34 (2002) 651–663, arXiv:gr-qc/0106051.
[4086] A. Kempf, G. Mangano, and R. B. Mann, “Hilbert space representation of the minimal length uncertainty relation,”
Phys. Rev. D 52 (1995) 1108–1118, arXiv:hep-th/9412167.
[4087] S. Das and E. C. Vagenas, “Universality of Quantum Gravity Corrections,” Phys. Rev. Lett. 101 (2008) 221301,
arXiv:0810.5333 [hep-th].
[4088] A. F. Ali, S. Das, and E. C. Vagenas, “Discreteness of Space from the Generalized Uncertainty Principle,” Phys. Lett.
B 678 (2009) 497–499, arXiv:0906.5396 [hep-th].
[4089] A. Das, S. Das, and E. C. Vagenas, “Discreteness of Space from GUP in Strong Gravitational Fields,” Phys. Lett. B
809 (2020) 135772, arXiv:2006.05781 [gr-qc].
[4090] S. Aghababaei, H. Moradpour, and E. C. Vagenas, “Hubble tension bounds the GUP and EUP parameters,” Eur.
Phys. J. Plus 136 (2021) no. 10, 997, arXiv:2109.14826 [gr-qc].
[4091] T. Jacobson, S. Liberati, and D. Mattingly, “Lorentz violation at high energy: Concepts, phenomena and astrophysical
constraints,” Annals Phys. 321 (2006) 150–196, arXiv:astro-ph/0505267.
[4092] G. Amelino-Camelia, “Quantum-Spacetime Phenomenology,” Living Rev. Rel. 16 (2013) 5, arXiv:0806.0339 [gr-qc].
[4093] M. Arzano, G. Gubitosi, and J. J. Relancio, “Deformed Relativistic Symmetry Principles,” Lect. Notes Phys. 1017
(2023) 49–103, arXiv:2211.11684 [hep-th].
[4094] G. Amelino-Camelia, J. R. Ellis, N. E. Mavromatos, D. V. Nanopoulos, and S. Sarkar, “Tests of quantum gravity from
observations of gamma-ray bursts,” Nature 393 (1998) 763–765, arXiv:astro-ph/9712103.
[4095] J. R. Ellis, K. Farakos, N. E. Mavromatos, V. A. Mitsou, and D. V. Nanopoulos, “Astrophysical probes of the
constancy of the velocity of light,” Astrophys. J. 535 (2000) 139–151, arXiv:astro-ph/9907340.
[4096] J. R. Ellis, N. E. Mavromatos, D. V. Nanopoulos, and A. S. Sakharov, “Quantum-gravity analysis of gamma-ray bursts
using wavelets,” Astron. Astrophys. 402 (2003) 409–424, arXiv:astro-ph/0210124.
[4097] MAGIC, Other Contributors Collaboration, J. Albert et al., “Probing Quantum Gravity using Photons from a
flare of the active galactic nucleus Markarian 501 Observed by the MAGIC telescope,” Phys. Lett. B 668 (2008)
253–257, arXiv:0708.2889 [astro-ph].
[4098] C. Pfeifer, “Redshift and lateshift from homogeneous and isotropic modified dispersion relations,” Phys. Lett. B 780
(2018) 246–250, arXiv:1802.00058 [gr-qc].
[4099] J. R. Ellis, N. E. Mavromatos, and D. V. Nanopoulos, “Quantum gravitational diffusion and stochastic fluctuations in
the velocity of light,” Gen. Rel. Grav. 32 (2000) 127–144, arXiv:gr-qc/9904068.
[4100] J. R. Ellis, N. E. Mavromatos, and D. V. Nanopoulos, “Derivation of a Vacuum Refractive Index in a Stringy
Space-Time Foam Model,” Phys. Lett. B 665 (2008) 412–417, arXiv:0804.3566 [hep-th].
[4101] J. Ellis, N. E. Mavromatos, and D. V. Nanopoulos, “D-Foam Phenomenology: Dark Energy, the Velocity of Light and
a Possible D-Void,” Int. J. Mod. Phys. A 26 (2011) 2243–2262, arXiv:0912.3428 [astro-ph.CO].
[4102] U. Jacob and T. Piran, “Lorentz-violation-induced arrival delays of cosmological particles,” JCAP 01 (2008) 031,
arXiv:0712.2170 [astro-ph].
[4103] L. Barcaroli, L. K. Brunkhorst, G. Gubitosi, N. Loret, and C. Pfeifer, “Planck-scale-modified dispersion relations in
homogeneous and isotropic spacetimes,” Phys. Rev. D 95 (2017) no. 2, 024036, arXiv:1612.01390 [gr-qc].
[4104] G. Amelino-Camelia, D. Frattulillo, G. Gubitosi, G. Rosati, and S. Bedić, “Phenomenology of DSR-relativistic in-vacuo
dispersion in FLRW spacetime,” JCAP 01 (2024) 070, arXiv:2307.05428 [gr-qc].
[4105] N. E. Mavromatos, V. A. Mitsou, S. Sarkar, and A. Vergou, “Implications of a Stochastic Microscopic Finsler
Cosmology,” Eur. Phys. J. C 72 (2012) 1956, arXiv:1012.4094 [hep-ph].
[4106] N. E. Mavromatos and V. A. Mitsou, “Observational Evidence for Negative-Energy Dust in Late-Times Cosmology,”
Astropart. Phys. 29 (2008) 442–452, arXiv:0707.4671 [astro-ph].
[4107] S. Basilakos, N. E. Mavromatos, V. A. Mitsou, and M. Plionis, “Dynamics and constraints of the Dissipative Liouville
Cosmology,” Astropart. Phys. 36 (2012) 7–17, arXiv:1107.3532 [astro-ph.CO].
[4108] R. J. Protheroe and H. Meyer, “An Infrared background TeV gamma-ray crisis?,” Phys. Lett. B 493 (2000) 1–6,
arXiv:astro-ph/0005349.
[4109] G. Amelino-Camelia and T. Piran, “Cosmic rays and TeV photons as probes of quantum properties of space-time,”
Phys. Lett. B 497 (2001) 265–270, arXiv:hep-ph/0006210.
[4110] J.-P. Uzan, “The Fundamental Constants and Their Variation: Observational Status and Theoretical Motivations,”
Rev. Mod. Phys. 75 (2003) 403, arXiv:hep-ph/0205340.
[4111] J.-P. Uzan, “Varying Constants, Gravitation and Cosmology,” Living Rev. Rel. 14 (2011) 2, arXiv:1009.5514
[astro-ph.CO].
[4112] C. J. A. P. Martins, “The status of varying constants: a review of the physics, searches and implications,”
arXiv:1709.02923 [astro-ph.CO].
[4113] J. D. Bekenstein, “Fine Structure Constant: Is It Really a Constant?,” Phys. Rev. D 25 (1982) 1527–1539.
[4114] C. J. A. P. Martins, P. E. Vielzeuf, M. Martinelli, E. Calabrese, and S. Pandolfi, “Evolution of the fine-structure
constant in runaway dilaton models,” Phys. Lett. B 743 (2015) 377–382, arXiv:1503.05068 [astro-ph.CO].
[4115] J. Rich, “Which fundamental constants for cosmic microwave background and baryon-acoustic oscillation?,” Astron.
Astrophys. 584 (2015) A69, arXiv:1503.06012 [astro-ph.CO].
397

[4116] S. Galli, C. J. A. P. Martins, A. Melchiorri, and E. Menegoni, “Testing the Variation of Fundamental Constants with
the CMB,” Astrophys. Space Sci. Proc. (2011) 59–67.
[4117] B. Lamine, Y. Ozdalkiran, L. Mirouze, F. Erdogan, S. Ilic, I. Tutusaus, R. Kou, and A. Blanchard, “Cosmological
measurement of the gravitational constant G using the CMB, the BAO and the BBN,” arXiv e-prints (2024)
arXiv:2407.15553, arXiv:2407.15553 [astro-ph.CO].
[4118] M. H. van Putten, “Entropic constraint on cosmic variation of planck mass and the boltzmann constant,” Results in
Physics 57 (2024) 107425. https://www.sciencedirect.com/science/article/pii/S2211379724001074.
[4119] I. Banik, H. Desmond, and N. Samaras, “Strong constraints on a sharp change in G as a solution to the Hubble
tension,” arXiv:2411.15301 [astro-ph.CO].
[4120] M. Kaplinghat, R. J. Scherrer, and M. S. Turner, “Constraining variations in the fine structure constant with the
cosmic microwave background,” Phys. Rev. D 60 (1999) 023516, arXiv:astro-ph/9810133.
[4121] P. P. Avelino, C. J. A. P. Martins, G. Rocha, and P. T. P. Viana, “Looking for a varying alpha in the cosmic microwave
background,” Phys. Rev. D 62 (2000) 123508, arXiv:astro-ph/0008446.
[4122] R. A. Battye, R. Crittenden, and J. Weller, “Cosmic concordance and the fine structure constant,” Phys. Rev. D 63
(2001) 043505, arXiv:astro-ph/0008265.
[4123] P. P. Avelino, S. Esposito, G. Mangano, C. J. A. P. Martins, A. Melchiorri, G. Miele, O. Pisanti, G. Rocha, and
P. T. P. Viana, “Early universe constraints on a time varying fine structure constant,” Phys. Rev. D 64 (2001) 103505,
arXiv:astro-ph/0102144.
[4124] G. Rocha, R. Trotta, C. J. A. P. Martins, A. Melchiorri, P. P. Avelino, R. Bean, and P. T. P. Viana, “Measuring alpha
in the early universe: cmb polarization, reionization and the fisher matrix analysis,” Mon. Not. Roy. Astron. Soc. 352
(2004) 20, arXiv:astro-ph/0309211.
[4125] C. J. A. P. Martins, A. Melchiorri, G. Rocha, R. Trotta, P. P. Avelino, and P. T. P. Viana, “Wmap constraints on
varying alpha and the promise of reionization,” Phys. Lett. B 585 (2004) 29–34, arXiv:astro-ph/0302295.
[4126] C. G. Scoccola, S. J. Landau, and H. Vucetich, “WMAP 5-year constraints on α and me ,” Mem. Soc. Ast. It. 80 (2009)
no. 4, 814–819, arXiv:0910.1083 [astro-ph.CO].
[4127] E. Menegoni, M. Archidiacono, E. Calabrese, S. Galli, C. J. A. P. Martins, and A. Melchiorri, “The Fine Structure
Constant and the CMB Damping Scale,” Phys. Rev. D 85 (2012) 107301, arXiv:1202.1476 [astro-ph.CO].
[4128] L. Hart and J. Chluba, “New constraints on time-dependent variations of fundamental constants using Planck data,”
Mon. Not. Roy. Astron. Soc. 474 (2018) no. 2, 1850–1861, arXiv:1705.03925 [astro-ph.CO].
[4129] J. Chluba and R. M. Thomas, “Towards a complete treatment of the cosmological recombination problem,” Mon. Not.
Roy. Astron. Soc. 412 (2011) 748, arXiv:1010.3631 [astro-ph.CO].
[4130] L. Hart and J. Chluba, “Updated fundamental constant constraints from Planck 2018 data and possible relations to
the Hubble tension,” Mon. Not. Roy. Astron. Soc. 493 (2020) no. 3, 3255–3263, arXiv:1912.03986 [astro-ph.CO].
[4131] J. Chluba and L. Hart, “Varying fundamental constants meet Hubble,” arXiv:2309.12083 [astro-ph.CO].
[4132] Planck Collaboration, P. A. R. Ade et al., “Planck intermediate results - XXIV. Constraints on variations in
fundamental constants,” Astron. Astrophys. 580 (2015) A22, arXiv:1406.7482 [astro-ph.CO].
[4133] S. Bize et al., “Testing the stability of fundamental constants with the Hg-199+ single-ion optical clock,” Phys. Rev.
Lett. 90 (2003) 150802, arXiv:physics/0212109.
[4134] T. Rosenband et al., “Frequency Ratio of Al+ and Hg+ Single-Ion Optical Clocks; Metrology at the 17th Decimal
Place,” Science 319 (2008) no. 5871, 1154622.
[4135] P. Bonifacio et al., “Fundamental constants and high resolution spectroscopy,” Astron. Nachr. 335 (2014) 83,
arXiv:1310.6280 [astro-ph.CO].
[4136] S. M. Kotuš, M. T. Murphy, and R. F. Carswell, “High-precision limit on variation in the fine-structure constant from
a single quasar absorption system,” Mon. Not. Roy. Astron. Soc. 464 (2017) no. 3, 3679–3703, arXiv:1609.03860
[astro-ph.CO].
[4137] O. Seto and Y. Toda, “Big bang nucleosynthesis constraints on varying electron mass solution to the Hubble tension,”
Phys. Rev. D 107 (2023) no. 8, 083512, arXiv:2206.13209 [astro-ph.CO].
[4138] T. L. Smith, D. Grin, D. Robinson, and D. Qi, “Probing spatial variation of the fine-structure constant using the
CMB,” Phys. Rev. D 99 (2019) no. 4, 043531, arXiv:1808.07486 [astro-ph.CO].
[4139] M. Lucca, J. Chluba, and A. Rotti, “CRRfast: an emulator for the cosmological recombination radiation with effects
from inhomogeneous recombination,” Mon. Not. Roy. Astron. Soc. 530 (2024) no. 1, 668–683, arXiv:2306.08085
[astro-ph.CO].
[4140] T. Sekiguchi and T. Takahashi, “Early recombination as a solution to the H0 tension,” Phys. Rev. D 103 (2021) no. 8,
083507, arXiv:2007.03381 [astro-ph.CO].
[4141] L. Hart and J. Chluba, “Varying fundamental constants principal component analysis: additional hints about the
Hubble tension,” Mon. Not. Roy. Astron. Soc. 510 (2022) no. 2, 2206–2227, arXiv:2107.12465 [astro-ph.CO].
[4142] N. Lee, Y. Ali-Haïmoud, N. Schöneberg, and V. Poulin, “What It Takes to Solve the Hubble Tension through
Modifications of Cosmological Recombination,” Phys. Rev. Lett. 130 (2023) no. 16, 161003, arXiv:2212.04494
[astro-ph.CO].
[4143] K. Hoshiya and Y. Toda, “Electron mass variation from dark sector interactions and compatibility with cosmological
observations,” Phys. Rev. D 107 (2023) no. 4, 043505, arXiv:2202.07714 [astro-ph.CO].
[4144] T. Damour and A. M. Polyakov, “The String dilaton and a least coupling principle,” Nucl. Phys. B 423 (1994)
532–558, arXiv:hep-th/9401069.
[4145] T. Chiba, T. Kobayashi, M. Yamaguchi, and J. Yokoyama, “Time variation of proton-electron mass ratio and fine
398

structure constant with runaway dilaton,” Phys. Rev. D 75 (2007) 043516, arXiv:hep-ph/0610027.
[4146] V. da Fonseca et al., “Fundamental physics with ESPRESSO: Constraining a simple parametrisation for varying α,”
Astron. Astrophys. 666 (2022) A57, arXiv:2204.02930 [astro-ph.CO].
[4147] B. J. Barros and V. da Fonseca, “Coupling quintessence kinetics to electromagnetism,” JCAP 06 (2023) 048,
arXiv:2209.12189 [astro-ph.CO].
[4148] L. Vacher, J. a. F. Dias, N. Schöneberg, C. J. A. P. Martins, S. Vinzl, S. Nesseris, G. Cañas Herrera, and
M. Martinelli, “Constraints on extended Bekenstein models from cosmological, astrophysical, and local data,” Phys.
Rev. D 106 (2022) no. 8, 083522, arXiv:2207.03258 [astro-ph.CO].
[4149] H. M. Tohfa, J. Crump, E. Baker, L. Hart, D. Grin, M. Brosius, and J. Chluba, “Cosmic microwave background search
for fine-structure constant evolution,” Phys. Rev. D 109 (2024) no. 10, 103529, arXiv:2307.06768 [astro-ph.CO].
[4150] L. Vacher, N. Schöneberg, J. D. F. Dias, C. J. A. P. Martins, and F. Pimenta, “Runaway dilaton models: Improved
constraints from the full cosmological evolution,” Phys. Rev. D 107 (2023) no. 10, 104002, arXiv:2301.13500
[astro-ph.CO].
[4151] L. Vacher and N. Schöneberg, “Incompatibility of fine-structure constant variations at recombination with local
observations,” Phys. Rev. D 109 (2024) no. 10, 103520, arXiv:2403.02256 [astro-ph.CO].
[4152] G. P. Lynch, L. Knox, and J. Chluba, “Reconstructing the recombination history by combining early and late
cosmological probes,” Phys. Rev. D 110 (2024) no. 6, 063518, arXiv:2404.05715 [astro-ph.CO].
[4153] O. Seto and Y. Toda, “DESI constraints on the varying electron mass model and axionlike early dark energy,” Phys.
Rev. D 110 (2024) no. 8, 083501, arXiv:2405.11869 [astro-ph.CO].
[4154] G. P. Lynch, L. Knox, and J. Chluba, “DESI observations and the Hubble tension in light of modified recombination,”
Phys. Rev. D 110 (2024) no. 8, 083538, arXiv:2406.10202 [astro-ph.CO].
[4155] R. A. Sunyaev and Y. B. Zeldovich, “Small scale entropy and adiabatic density perturbations? Antimatter in the
Universe,” Astrophys. Space Sci. 9 (1970) no. 3, 368–382.
[4156] W. Hu and N. Sugiyama, “Anisotropies in the cosmic microwave background: An Analytic approach,” Astrophys. J.
444 (1995) 489–506, arXiv:astro-ph/9407093.
[4157] W. Hu, D. Scott, N. Sugiyama, and M. J. White, “The Effect of physical assumptions on the calculation of microwave
background anisotropies,” Phys. Rev. D 52 (1995) 5498–5515, arXiv:astro-ph/9505043.
[4158] J. Chluba and R. A. Sunyaev, “Induced two-photon decay of the 2s level and the rate of cosmological hydrogen
recombination,” Astron. Astrophys. 446 (2006) 39–42, arXiv:astro-ph/0508144.
[4159] A. Lewis, J. Weller, and R. Battye, “The Cosmic Microwave Background and the Ionization History of the Universe,”
Mon. Not. Roy. Astron. Soc. 373 (2006) 561–570, arXiv:astro-ph/0606552.
[4160] J. R. Shaw and J. Chluba, “Precise cosmological parameter estimation using CosmoRec,” Mon. Not. Roy. Astron. Soc.
415 (2011) 1343, arXiv:1102.3683 [astro-ph.CO].
[4161] J. Chluba and R. A. Sunyaev, “Is there need and another way to measure the Cosmic Microwave Background
temperature more accurately?,” Astron. Astrophys. 478 (2008) L27, arXiv:0707.0188 [astro-ph].
[4162] R. A. Sunyaev and J. Chluba, “Signals From the Epoch of Cosmological Recombination,” Astron. Nachr. 330 (2009)
657–674, arXiv:0908.0435 [astro-ph.CO].
[4163] L. Hart, A. Rotti, and J. Chluba, “Sensitivity forecasts for the cosmological recombination radiation in the presence of
foregrounds,” Mon. Not. Roy. Astron. Soc. 497 (2020) no. 4, 4535–4548, arXiv:2006.04826 [astro-ph.CO].
[4164] J. Chluba et al., “New horizons in cosmology with spectral distortions of the cosmic microwave background,” Exper.
Astron. 51 (2021) no. 3, 1515–1554, arXiv:1909.01593 [astro-ph.CO].
[4165] H. Desmond, B. Jain, and J. Sakstein, “Local resolution of the Hubble tension: The impact of screened fifth forces on
the cosmic distance ladder,” Phys. Rev. D 100 (2019) no. 4, 043537, arXiv:1907.03778 [astro-ph.CO]. [Erratum:
Phys.Rev.D 101, 069904 (2020), Erratum: Phys.Rev.D 101, 129901 (2020)].
[4166] C. Burrage and J. Sakstein, “A Compendium of Chameleon Constraints,” JCAP 11 (2016) 045, arXiv:1609.01192
[astro-ph.CO].
[4167] C. Burrage and J. Sakstein, “Tests of Chameleon Gravity,” Living Rev. Rel. 21 (2018) no. 1, 1, arXiv:1709.09071
[astro-ph.CO].
[4168] T. Baker et al., “Novel Probes Project: Tests of gravity on astrophysical scales,” Rev. Mod. Phys. 93 (2021) no. 1,
015003, arXiv:1908.03430 [astro-ph.CO].
[4169] J. Sakstein, “Astrophysical tests of screened modified gravity,” Int. J. Mod. Phys. D 27 (2018) no. 15, 1848008,
arXiv:2002.04194 [astro-ph.CO].
[4170] P. Brax, S. Casas, H. Desmond, and B. Elder, “Testing Screened Modified Gravity,” Universe 8 (2021) no. 1, 11,
arXiv:2201.10817 [gr-qc].
[4171] L. Berezhiani, J. Khoury, and J. Wang, “Universe without dark energy: Cosmic acceleration from dark matter-baryon
interactions,” Phys. Rev. D 95 (2017) no. 12, 123530, arXiv:1612.00453 [hep-th].
[4172] H. Desmond and J. Sakstein, “Screened fifth forces lower the TRGB-calibrated Hubble constant too,” Phys. Rev. D
102 (2020) no. 2, 023007, arXiv:2003.12876 [astro-ph.CO].
[4173] M. Högås and E. Mörtsell, “Hubble tension and fifth forces,” Phys. Rev. D 108 (2023) no. 12, 124050,
arXiv:2309.01744 [astro-ph.CO].
[4174] Ruchika, H. Rathore, S. Roy Choudhury, and V. Rentala, “A gravitational constant transition within cepheids as
supernovae calibrators can solve the Hubble tension,” JCAP 06 (2024) 056, arXiv:2306.05450 [astro-ph.CO].
[4175] L. Amendola, P. S. Corasaniti, and F. Occhionero, “Time variability of the gravitational constant and type Ia
supernovae,” arXiv:astro-ph/9907222.
399

[4176] E. Garcia-Berro, E. Gaztanaga, J. Isern, O. Benvenuto, and L. Althaus, “On the evolution of cosmological type ia
supernovae and the gravitational constant,” arXiv:astro-ph/9907440.
[4177] E. Gaztanaga, E. Garcia-Berro, J. Isern, E. Bravo, and I. Dominguez, “Bounds on the possible evolution of the
gravitational constant from cosmological type Ia supernovae,” Phys. Rev. D 65 (2002) 023506,
arXiv:astro-ph/0109299.
[4178] B. S. Wright and B. Li, “Type Ia supernovae, standardizable candles, and gravity,” Phys. Rev. D 97 (2018) no. 8,
083505, arXiv:1710.07018 [astro-ph.CO].
[4179] W. Zhao, B. S. Wright, and B. Li, “Constraining the time variation of Newton’s constant G with gravitational-wave
standard sirens and supernovae,” JCAP 10 (2018) 052, arXiv:1804.03066 [astro-ph.CO].
[4180] I. Goldman, “Neutron Stars constraints on a late G transition,” Phys. Lett. B 858 (2024) 139084, arXiv:2402.09859
[astro-ph.CO].
[4181] R. P. Gupta, “Constraining Co-Varying Coupling Constants from Globular Cluster Age,” Universe 9 (2023) no. 2, 70,
arXiv:2302.00552 [astro-ph.CO].
[4182] R. P. Gupta, “Constraining Coupling Constants’ Variation with Supernovae, Quasars, and GRBs,” Symmetry 15
(2023) no. 2, 259, arXiv:2301.09795 [astro-ph.CO].
[4183] R. P. Gupta, “Effect of evolving physical constants on type Ia supernova luminosity,” Mon. Not. Roy. Astron. Soc. 511
(2022) no. 3, 4238–4250, arXiv:2112.10654 [gr-qc].
[4184] J. W. Moffat, “Scalar-tensor-vector gravity theory,” JCAP 03 (2006) 004, arXiv:gr-qc/0506021.
[4185] J. D. Barrow and J. Magueijo, “Varying alpha theories and solutions to the cosmological problems,” Phys. Lett. B 443
(1998) 104–110, arXiv:astro-ph/9811072.
[4186] A. J. Anderson, P. Barry, A. N. Bender, B. A. Benson, L. E. Bleem, J. E. Carlstrom, T. W. Cecil, C. L. Chang, T. M.
Crawford, K. R. Dibert, M. A. Dobbs, K. Fichman, N. W. Halverson, W. L. Holzapfel, A. Hryciuk, K. S. Karkare,
J. Li, M. Lisovenko, D. Marrone, J. McMahon, J. Montgomery, T. Natoli, Z. Pan, S. Raghunathan, C. L. Reichardt,
M. Rouble, E. Shirokoff, G. Smecher, A. A. Stark, J. D. Vieira, and M. R. Young, “SPT-3G+: mapping the
high-frequency cosmic microwave background using kinetic inductance detectors,” in Millimeter, Submillimeter, and
Far-Infrared Detectors and Instrumentation for Astronomy XI, J. Zmuidzinas and J.-R. Gao, eds., vol. 12190 of Society
of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, p. 1219003. 2022. arXiv:2208.08559
[astro-ph.IM].
[4187] ACT Collaboration, W. Coulton et al., “Atacama Cosmology Telescope: High-resolution component-separated maps
across one third of the sky,” Phys. Rev. D 109 (2024) no. 6, 063530, arXiv:2307.01258 [astro-ph.CO].
[4188] POLARBEAR Collaboration, A. Suzuki et al., “The POLARBEAR-2 and the Simons Array Experiment,” J. Low
Temp. Phys. 184 (2016) no. 3-4, 805–810, arXiv:1512.07299 [astro-ph.IM].
[4189] H. Li et al., “Probing Primordial Gravitational Waves: Ali CMB Polarization Telescope,” Natl. Sci. Rev. 6 (2019)
no. 1, 145–154, arXiv:1710.03047 [astro-ph.CO].
[4190] QUBIC Collaboration, J. C. Hamilton et al., “QUBIC I: Overview and science program,” JCAP 04 (2022) no. 04,
034, arXiv:2011.02213 [astro-ph.IM].
[4191] CLASS Collaboration, R. Datta et al., “Cosmology Large Angular Scale Surveyor (CLASS): 90 GHz Telescope
Pointing, Beam Profile, Window Function, and Polarization Performance,” Astrophys. J. Suppl. 273 (2024) no. 2, 26,
arXiv:2308.13309 [astro-ph.IM].
[4192] GroundBIRD Collaboration, K. Lee et al., “GroundBIRD: A CMB Polarization Experiment with MKID Arrays,” J.
Low Temp. Phys. 200 (2020) no. 5-6, 384–391, arXiv:2011.07705 [astro-ph.IM].
[4193] H. Hui et al., “BICEP Array: a multi-frequency degree-scale CMB polarimeter,” Proc. SPIE Int. Soc. Opt. Eng. 10708
(2018) 1070807, arXiv:1808.00568 [astro-ph.IM].
[4194] LiteBIRD Collaboration, T. Namikawa et al., “LiteBIRD science goals and forecasts: improving sensitivity to
inflationary gravitational waves with multitracer delensing,” JCAP 06 (2024) 010, arXiv:2312.05194 [astro-ph.CO].
[4195] LiteBIRD Collaboration, D. Paoletti et al., “LiteBIRD science goals and forecasts: primordial magnetic fields,” JCAP
07 (2024) 086, arXiv:2403.16763 [astro-ph.CO].
[4196] LiteBIRD Collaboration, M. Remazeilles et al., “LiteBIRD science goals and forecasts. Mapping the hot gas in the
Universe,” JCAP 12 (2024) 026, arXiv:2407.17555 [astro-ph.CO].
[4197] LiteBIRD Collaboration, E. de la Hoz et al., “LiteBIRD Science Goals and Forecasts: constraining isotropic cosmic
birefringence,” arXiv:2503.22322 [astro-ph.CO].
[4198] N. Sehgal et al., “CMB-HD: An Ultra-Deep, High-Resolution Millimeter-Wave Survey Over Half the Sky,”
arXiv:1906.10134 [astro-ph.CO].
[4199] CMB-HD Collaboration, S. Aiola et al., “Snowmass2021 CMB-HD White Paper,” arXiv:2203.05728 [astro-ph.CO].
[4200] S. Hanany et al., “PICO: Probe of Inflation and Cosmic Origins,” arXiv e-prints (2019) arXiv:1902.10541,
arXiv:1902.10541 [astro-ph.IM].
[4201] PRISM Collaboration, P. André et al., “PRISM (Polarized Radiation Imaging and Spectroscopy Mission): An
Extended White Paper,” JCAP 02 (2014) 006, arXiv:1310.1554 [astro-ph.CO].
[4202] CHIME Collaboration, M. Amiri et al., “An Overview of CHIME, the Canadian Hydrogen Intensity Mapping
Experiment,” Astrophys. J. Supp. 261 (2022) no. 2, 29, arXiv:2201.07869 [astro-ph.IM].
[4203] E. Abdalla et al., “The BINGO project - I. Baryon acoustic oscillations from integrated neutral gas observations,”
Astron. Astrophys. 664 (2022) A14, arXiv:2107.01633 [astro-ph.CO].
[4204] L. Amendola et al., “Cosmology and fundamental physics with the Euclid satellite,” Living Rev. Rel. 21 (2018) no. 1,
2, arXiv:1606.00180 [astro-ph.CO].
400

[4205] M. Takada, R. S. Ellis, M. Chiba, J. E. Greene, H. Aihara, N. Arimoto, K. Bundy, J. Cohen, O. Doré, G. Graves, J. E.
Gunn, T. Heckman, C. M. Hirata, P. Ho, J.-P. Kneib, O. Le Fèvre, L. Lin, S. More, H. Murayama, T. Nagao,
M. Ouchi, M. Seiffert, J. D. Silverman, L. Sodré, D. N. Spergel, M. A. Strauss, H. Sugai, Y. Suto, H. Takami, and
R. Wyse, “Extragalactic science, cosmology, and Galactic archaeology with the Subaru Prime Focus Spectrograph,”
Publ. Astron. Soc. Jpn. 66 (2014) no. 1, R1, arXiv:1206.0737 [astro-ph.CO].
[4206] N. Tamura et al., “Prime Focus Spectrograph (PFS) for the Subaru Telescope: Overview, recent progress, and future
perspectives,” Proc. SPIE Int. Soc. Opt. Eng. 9908 (2016) 99081M, arXiv:1608.01075 [astro-ph.IM].
[4207] B. Blum et al., “Snowmass2021 Cosmic Frontier White Paper: Rubin Observatory after LSST,” in Snowmass 2021.
2022. arXiv:2203.07220 [astro-ph.CO].
[4208] R. Braun, T. L. Bourke, J. A. Green, E. Keane, and J. Wagg, “Advancing Astrophysics with the Square Kilometre
Array,” PoS AASKA14 (2015) 174.
[4209] DESI Collaboration, D. J. Schlegel et al., “A Spectroscopic Road Map for Cosmic Frontier: DESI, DESI-II, Stage-5,”
arXiv:2209.03585 [astro-ph.CO].
[4210] R. S. de Jong et al., “4MOST: 4-metre multi-object spectroscopic telescope,” in Ground-based and Airborne
Instrumentation for Astronomy IV, I. S. McLean, S. K. Ramsay, and H. Takami, eds., vol. 8446 of Society of
Photo-Optical Instrumentation Engineers (SPIE) Conference Series, p. 84460T. 2012. arXiv:1206.6885
[astro-ph.IM].
[4211] SPHEREx Collaboration, O. Doré et al., “Cosmology with the SPHEREX All-Sky Spectral Survey,”
arXiv:1412.4872 [astro-ph.CO].
[4212] K. A. Cleary et al., “COMAP Early Science: I. Overview,” arXiv:2111.05927 [astro-ph.CO].
[4213] PUMA Collaboration, A. Slosar et al., “Packed Ultra-wideband Mapping Array (PUMA): A Radio Telescope for
Cosmology and Transients,” Bull. Am. Astron. Soc. 51 (2019) 53, arXiv:1907.12559 [astro-ph.IM].
[4214] DES Collaboration, T. Abbott et al., “The Dark Energy Survey: more than dark energy – an overview,” Mon. Not.
Roy. Astron. Soc. 460 (2016) no. 2, 1270–1299, arXiv:1601.00329 [astro-ph.CO].
[4215] L. J. Romualdez et al., “Overview, design, and flight results from SuperBIT: a high-resolution, wide-field,
visible-to-near-UV balloon-borne astronomical telescope,” in Ground-based and Airborne Instrumentation for
Astronomy VII, C. J. Evans, L. Simard, and H. Takami, eds., vol. 10702 of Society of Photo-Optical Instrumentation
Engineers (SPIE) Conference Series, p. 107020R. 2018. arXiv:1807.02887 [astro-ph.IM].
[4216] B. Sathyaprakash et al., “Scientific Objectives of Einstein Telescope,” Class. Quant. Grav. 29 (2012) 124013,
arXiv:1206.0331 [gr-qc]. [Erratum: Class.Quant.Grav. 30, 079501 (2013)].
[4217] A. Abac et al., “The Science of the Einstein Telescope,” arXiv:2503.12263 [gr-qc].
[4218] M. Punturo et al., “The Einstein Telescope: A third-generation gravitational wave observatory,” Class. Quant. Grav.
27 (2010) 194002.
[4219] A. Utina et al., “ETpathfinder: a cryogenic testbed for interferometric gravitational-wave detectors,” Class. Quant.
Grav. 39 (2022) no. 21, 215008, arXiv:2206.04905 [astro-ph.IM].
[4220] S. Borhanian and B. S. Sathyaprakash, “Listening to the Universe with next generation ground-based
gravitational-wave detectors,” Phys. Rev. D 110 (2024) no. 8, 083040, arXiv:2202.11048 [gr-qc].
[4221] I. Gupta et al., “Characterizing gravitational wave detector networks: from A♯ to cosmic explorer,” Class. Quant.
Grav. 41 (2024) no. 24, 245001, arXiv:2307.10421 [gr-qc].
[4222] TianQin Collaboration, J. Mei et al., “The TianQin project: current progress on science and technology,” PTEP 2021
(2021) no. 5, 05A107, arXiv:2008.10332 [gr-qc].
[4223] J. Luo et al., “Fundamental Physics and Cosmology with TianQin,” arXiv:2502.20138 [gr-qc].
[4224] Z. Ren, T. Zhao, Z. Cao, Z.-K. Guo, W.-B. Han, H.-B. Jin, and Y.-L. Wu, “Taiji data challenge for exploring
gravitational wave universe,” Front. Phys. (Beijing) 18 (2023) no. 6, 64302, arXiv:2301.02967 [gr-qc].
[4225] S. Kawamura et al., “Current status of space gravitational wave antenna DECIGO and B-DECIGO,” PTEP 2021
(2021) no. 5, 05A105, arXiv:2006.13545 [gr-qc].
[4226] NANOGrav Collaboration, G. Agazie et al., “The NANOGrav 15 yr Data Set: Evidence for a Gravitational-wave
Background,” Astrophys. J. Lett. 951 (2023) no. 1, L8, arXiv:2306.16213 [astro-ph.HE].
[4227] R. N. Manchester, G. Hobbs, M. Bailes, W. A. Coles, W. van Straten, M. J. Keith, R. M. Shannon, N. D. R. Bhat,
A. Brown, S. G. Burke-Spolaor, D. J. Champion, A. Chaudhary, R. T. Edwards, G. Hampson, A. W. Hotan,
A. Jameson, F. A. Jenet, M. J. Kesteven, J. Khoo, J. Kocz, K. Maciesiak, S. Oslowski, V. Ravi, J. R. Reynolds, J. M.
Sarkissian, J. P. W. Verbiest, Z. L. Wen, W. E. Wilson, D. Yardley, W. M. Yan, and X. P. You, “The Parkes Pulsar
Timing Array Project,” Publ. Astron. Soc. Aust. 30 (2013) e017, arXiv:1210.6130 [astro-ph.IM].
[4228] R. N. Manchester and IPTA, “The International Pulsar Timing Array,” Classical and Quantum Gravity 30 (2013)
no. 22, 224010, arXiv:1309.7392 [astro-ph.IM].
[4229] A. P. Beardsley et al., “Science with the Murchison Widefield Array: Phase I results and Phase II opportunities,” Publ.
Astron. Soc. Aust. 36 (2019) e050, arXiv:1910.02895 [astro-ph.IM].
[4230] J. Cumner et al., “Radio Antenna Design for Sky-Averaged 21cm Cosmology Experiments: The REACH Case,” J.
Astron. Inst. 11 (2022) no. 01, 2250001, arXiv:2109.10098 [astro-ph.IM].
[4231] H. T. Intema, P. Jagannathan, K. P. Mooley, and D. A. Frail, “The GMRT 150 MHz All-sky Radio Survey: First
Alternative Data Release TGSS ADR1,” Astron. Astrophys. 598 (2017) A78, arXiv:1603.04368 [astro-ph.CO].
[4232] R. Nan, D. Li, C. Jin, Q. Wang, L. Zhu, W. Zhu, H. Zhang, Y. Yue, and L. Qian, “The Five-Hundred Aperture
Spherical Radio Telescope (fast) Project,” International Journal of Modern Physics D 20 (2011) no. 6, 989–1024,
arXiv:1105.3794 [astro-ph.IM].
[4233] M. J. Graham et al., “The Zwicky Transient Facility: Science Objectives,” Publ. Astron. Soc. Pac. 131 (2019)
no. 1001, 078001, arXiv:1902.01945 [astro-ph.IM].
[4234] W. M. Wood-Vasey et al., “The nearby supernova factory,” New Astron. Rev. 48 (2004) 637–640,
arXiv:astro-ph/0401513.
[4235] Y. Wang et al., “ATLAS Probe: Breakthrough Science of Galaxy Evolution, Cosmology, Milky Way, and the Solar
System,” arXiv:1909.00070 [astro-ph.IM].
[4236] I. Hook, “The Science Case for the European ELT,” in Science with the VLT in the ELT Era, A. Moorwood, ed., vol. 9
of Astrophysics and Space Science Proceedings, p. 225. 2009.
[4237] M. Colless, “Key early science with MANIFEST on GMT,” arXiv e-prints (2018) arXiv:1809.05804, arXiv:1809.05804
[astro-ph.IM].
[4238] TMT International Science Development Teams & TMT Science Advisory Committee Collaboration,
W. Skidmore et al., “Thirty Meter Telescope Detailed Science Case: 2015,” Res. Astron. Astrophys. 15 (2015) no. 12,
1945–2140, arXiv:1505.01195 [astro-ph.IM].
[4239] E. V. L. Telescope, The VLT White Book. European Organisation for Astronomical Research in the Southern
Hemisphere, 1988. https://www.eso.org/public/products/books/book_0004/.
[4240] R. Bustos, M. Rubio, A. Otárola, and N. Nagar, “Parque Astronómico de Atacama: An Ideal Site for Millimeter,
Submillimeter, and Mid-Infrared Astronomy,” PASP 126 (2014) no. 946, 1126, arXiv:1410.2451 [astro-ph.IM].
[4241] M. Davis et al., “Science objectives and early results of the DEEP2 redshift survey,” Proc. SPIE Int. Soc. Opt. Eng.
4834 (2003) 161–172, arXiv:astro-ph/0209419.
[4242] S. Johnston and J. Wall, “Science with ASKAP - the Australian Square Kilometre Array Pathfinder,” Exper. Astron.
22 (2008) 151, arXiv:0810.5187 [astro-ph].
[4243] M. Bailes, A. Jameson, C. Flynn, T. Bateman, E. D. Barr, S. Bhandari, J. D. Bunton, M. Caleb, D. Campbell-Wilson,
W. Farah, B. Gaensler, A. J. Green, R. W. Hunstead, F. Jankowski, E. F. Keane, V. V. Krishnan, T. Murphy,
M. O’Neill, S. Osłowski, A. Parthasarathy, V. Ravi, P. Rosado, and D. Temby, “The UTMOST: A Hybrid Digital
Signal Processor Transforms the Molonglo Observatory Synthesis Telescope,” Publ. Astron. Soc. Aust. 34 (2017) e045,
arXiv:1708.09619 [astro-ph.IM].
[4244] J. Jonas and MeerKAT Team, “The MeerKAT Radio Telescope,” in MeerKAT Science: On the Pathway to the SKA,
p. 1. 2016.
[4245] C. J. Law et al., “Deep Synoptic Array Science: First FRB and Host Galaxy Catalog,” Astrophys. J. 967 (2024) no. 1,
29, arXiv:2307.03344 [astro-ph.HE].
[4246] T. Clarke, W. Peters, W. Brisken, S. Giacintucci, N. Kassim, E. Polisensky, J. Helmboldt, E. E. Richards, A. Erickson,
P. S. Ray, M. T. Kerr, J. Deneva, W. Coburn, R. Huber, and J. Long, “The VLA Low-band Ionosphere and Transient
Experiment (VLITE),” in American Astronomical Society Meeting Abstracts #231, vol. 231 of American Astronomical
Society Meeting Abstracts, p. 354.11. 2018.
[4247] G. Hallinan and DSA-2000 collaboration, “The DSA-2000: the future of radio survey science,” in American
Astronomical Society Meeting Abstracts, vol. 241 of American Astronomical Society Meeting Abstracts, p. 239.07. 2023.
[4248] B. Adebahr et al., “Apercal-The Apertif calibration pipeline,” Astronomy and Computing 38 (2022) 100514,
arXiv:2112.03722 [astro-ph.IM].
[4249] K. Vanderlinde, A. Liu, B. Gaensler, D. Bond, G. Hinshaw, C. Ng, C. Chiang, I. Stairs, J.-A. Brown, J. Sievers,
J. Mena, K. Smith, K. Bandura, K. Masui, K. Spekkens, L. Belostotski, M. Dobbs, N. Turok, P. Boyle, M. Rupen,
T. Landecker, U.-L. Pen, and V. Kaspi, “The Canadian Hydrogen Observatory and Radio-transient Detector
(CHORD),” in Canadian Long Range Plan for Astronomy and Astrophysics White Papers, vol. 2020, p. 28. 2019.
arXiv:1911.01777 [astro-ph.IM].
[4250] G. J. Hill et al., “The Hobby-Eberly Telescope Dark Energy Experiment (HETDEX): Description and Early Pilot
Survey Results,” ASP Conf. Ser. 399 (2008) 115–118, arXiv:0806.0183 [astro-ph].
[4251] A. Fasano et al., “CONCERTO: Instrument and status,” EPJ Web Conf. 293 (2024) 00018, arXiv:2311.04704
[astro-ph.IM].
[4252] V. L. Butler et al., “TIME: the Tomographic Ionized-carbon Mapping Experiment: an update on design,
characterization, and data from the 2022 commissioning observations,” in Millimeter, Submillimeter, and Far-Infrared
Detectors and Instrumentation for Astronomy XII, J. Zmuidzinas and J.-R. Gao, eds., vol. 13102 of Society of
Photo-Optical Instrumentation Engineers (SPIE) Conference Series, p. 131022G. 2024.
[4253] A. Cooray, T.-C. Chang, S. Unwin, M. Zemcov, A. Coffey, P. Morrissey, N. Raouf, S. Lipscy, M. Shannon, G. Wu,
R. Cen, R. R. Chary, O. Doré, X. Fan, G. G. Fazio, S. L. Finkelstein, C. Heneka, B. Lee, P. Linden, H. Nayyeri,
J. Rhodes, R. Sadoun, M. B. Silva, H. Trac, H.-Y. Wu, and Z. Zheng, “Cosmic Dawn Intensity Mapper,” in Bulletin of
the American Astronomical Society, vol. 51, p. 23. 2019. arXiv:1903.03144 [astro-ph.GA].
[4254] G. J. Stacey et al., “CCAT-prime: Science with an Ultra-widefield Submillimeter Observatory at Cerro Chajnantor,” in
Ground-based and Airborne Telescopes VII. 2018. arXiv:1807.04354 [astro-ph.GA].
[4255] W. Sutherland, J. Emerson, G. Dalton, E. Atad-Ettedgui, S. Beard, R. Bennett, N. Bezawada, A. Born, M. Caldwell,
P. Clark, S. Craig, D. Henry, P. Jeffers, B. Little, A. McPherson, J. Murray, M. Stewart, B. Stobie, D. Terrett,
K. Ward, M. Whalley, and G. Woodhouse, “The Visible and Infrared Survey Telescope for Astronomy (VISTA):
Design, technical overview, and performance,” A&A 575 (2015) A25, arXiv:1409.4780 [astro-ph.IM].
[4256] J. Silk, “The limits of cosmology: role of the Moon,” Phil. Trans. A. Math. Phys. Eng. Sci. 379 (2021) 20190561,
arXiv:2011.04671 [astro-ph.CO].
402

White Paper Authors


Eleonora Di Valentino,1 Jackson Levi Said,2, 3 Adam Riess,4, 5 Agnieszka Pollo,6 Vivian Poulin,7 Adrià
Gómez-Valent,8 Amanda Weltman,9 Antonella Palmese,10 Caroline D. Huang,11 Carsten van de Bruck,12 Chandra
Shekhar Saraf,13 Cheng-Yu Kuo,14 Cora Uhlemann,15, 16 Daniela Grandón,17 Dante Paz,18 Dominique Eckert,19 Elsa
M. Teixeira,20 Emmanuel N. Saridakis,21, 22, 23 Eoin Ó Colgáin,24 Florian Beutler,25 Florian Niedermann,26
Francesco Bajardi,27, 28 Gabriela Barenboim,29 Giulia Gubitosi,30, 28 Ilaria Musella,31 Indranil Banik,32 Istvan
Szapudi,33 Jack Singal,34 Jaume Haro Cases,35 Jens Chluba,36 Jesús Torrado,37 Jurgen Mifsud,2, 3 Karsten
Jedamzik,38 Khaled Said,39 Konstantinos Dialektopoulos,40, 2 Laura Herold,4 Leandros Perivolaropoulos,41 Lei Zu,6
Lluís Galbany,42, 43 Louise Breuval,44 Luca Visinelli,45, 46 Luis A. Escamilla,1 Luis A. Anchordoqui,47 M.M.
Sheikh-Jabbari,48 Margherita Lembo,49, 50, 51 Maria Giovanna Dainotti,52 Maria Vincenzi,53 Marika Asgari,16
Martina Gerbino,54 Matteo Forconi,55 Michele Cantiello,56 Michele Moresco,57, 58 Micol Benetti,27, 28 Nils
Schöneberg,59, 60 Özgür Akarsu,61 Rafael C. Nunes,62, 63 Reginald Christian Bernardo,64 Ricardo Chávez,65, 66
Richard I. Anderson,67 Richard Watkins,68 Salvatore Capozziello,69, 27, 70 Siyang Li,4 Sunny Vagnozzi,71, 72 Supriya
Pan,73 Tommaso Treu,74 Vid Irsic,75, 76 Will Handley,77, 78 William Giarè,1 Yukei Murakami,4 Adèle Poudou,79 Alan
Heavens,80 Alan Kogut,81 Alba Domi,82 Aleksander Łukasz Lenart,83 Alessandro Melchiorri,84 Alessandro
Vadalà,85, 86, 87 Alexandra Amon,88, 89 Alexander Bonilla Rivera,90 Alexander Reeves,91 Alexander Zhuk,92, 93, 94
Alfio Bonanno,95 Ali Övgün,96 Alice Pisani,97, 98 Alireza Talebian,99 Amare Abebe,100 Amin Aboubrahim,101
Ana Luisa González Morán,102 András Kovács,103, 104 Andreas Lymperis,105 Andreas Papatriantafyllou,106
Andrew R. Liddle,107 Andronikos Paliathanasis,108, 109, 110 Andrzej Borowiec,111 Anil Kumar Yadav,112 Anita
Yadav,113 Anjan Ananda Sen,114 Anjitha John William Mini Latha,115 Anne Christine Davis,116 Anowar J.
Shajib,117, 118, 119 Anthony Walters,120, 121, 9 Anto Idicherian Lonappan,122 Anton Chudaykin,123 Antonio
Capodagli,124 Antonio da Silva,107 Antonio De Felice,125 Antonio Racioppi,126 Araceli Soler Oficial,127 Ariadna
Montiel,128 Arianna Favale,129, 8 Armando Bernui,130 Arrianne Crystal Velasco,131, 132 Asta Heinesen,133 Athanasios
Bakopoulos,134, 106 Athanasios Chatzistavrakidis,135 Bahman Khanpour,136 Bangalore S. Sathyaprakash,137 Bartek
Zgirski,138 Benjamin L’Huillier,139 Benoit Famaey,140 Bhuvnesh Jain,141 Biesiada Marek,6 Bing Zhang,142
Biswajit Karmakar,143 Branko Dragovich,144 Brooks Thomas,145 Carlos Correa,146 Carlos G. Boiza,147 Catarina
Marques,148, 149 Celia Escamilla-Rivera,150 Charalampos Tzerefos,151, 152 Chi Zhang,153, 154, 155 Chiara De Leo,156
Christian Pfeifer,157 Christine Lee,103 Christo Venter,158 Cláudio Gomes,159, 160 Clecio Roque De bom,161 Cristian
Moreno-Pulido,162 Damianos Iosifidis,163 Dan Grin,164 Daniel Blixt,27 Dan Scolnic,165 Daniele Oriti,166 Daria
Dobrycheva,167 Dario Bettoni,168, 169 David Benisty,170 David Fernández-Arenas,171 David L. Wiltshire,172 David
Sanchez Cid,173, 174 David Tamayo,175, 149 David Valls-Gabaud,115 Davide Pedrotti,71 Deng Wang,176 Denitsa
Staicova,177 Despoina Totolou,178 Diego Rubiera-Garcia,179 Dinko Milaković,180, 181 Dominic W. Pesce,182, 183
Dominique Sluse,184 Duško Borka,185 Ebrahim Yusofi,186, 187 Elena Giusarma,188 Elena Terlevich,189, 78 Elena
Tomasetti,190, 58 Elias C. Vagenas,191 Elisa Fazzari,192, 193, 194 Elisa G. M. Ferreira,195 Elvis Barakovic,196 Emanuela
Dimastrogiovanni,197 Emil Brinch Holm,198 Emil Mottola,199 Emre Özülker,200 Enrico Specogna,200 Enzo
Brocato,201, 202 Erik Jensko,203 Erika Antonette Enriquez,131 Esha Bhatia,204 Fabio Bresolin,205 Felipe Avila,130
Filippo Bouchè,206, 207 Flavio Bombacigno,208 Fotios K. Anagnostopoulos,209 Francesco Pace,210, 211, 212 Francesco
Sorrenti,213 Francisco S. N. Lobo,107, 214 Frédéric Courbin,215, 216 Frode K. Hansen,217 Greg Sloan,5, 218 Gabriel
Farrugia,2, 219 Gabriel Lynch,220 Gabriela Garcia-Arroyo,221 Gabriella Raimondo,56 Gaetano Lambiase,45, 46
Gagandeep S. Anand,5 Gaspard Poulot,200 Genly Leon,222, 223 Gerasimos Kouniatalis,224, 152 Germano Nardini,225
Géza Csörnyei,226 Giacomo Galloni,227, 228 Giada Bargiacchi,229 Giannis Papagiannopoulos,230 Giovanni
Montani,231, 232 Giovanni Otalora,233 Giulia De Somma,234, 235 Giuliana Fiorentino,236 Giuseppe Fanizza,237
Giuseppe Gaetano Luciano,238 Giuseppe Sarracino,239 Gonzalo J. Olmo,208, 240 Goran S. Djordjević,241 Guadalupe
Cañas-Herrera,242 Hanyu Cheng,243, 244, 200 Harry Desmond,32 Hassan Abdalla,109 Houzun Chen,245 Hsu-Wen
Chiang,246 Hume A. Feldman,247 Hussain Gohar,248 Ido Ben-Dayan,249 Ignacio Sevilla-Noarbe,173 Ignatios
Antoniadis,250 Ilim Cimdiker,248 Inês S. Albuquerque,107 Ioannis D. Gialamas,126 Ippocratis Saltas,251 Iryna
Vavilova,252 Isidro Gómez-Vargas,253 Ismael Ayuso,254 Ismailov Nariman Zeynalabdi,255 Ivan De Martino,169 Ivonne
Zavala Carrasco,256 J. Alberto Vázquez,257 Jacobo Asorey,258 Janusz Gluza,259 Javier Rubio,179 Jenny G.
Sorce,260, 261 Jenny Wagner,262, 263, 264 Jeremy Sakstein,265 Jessica Santiago,266 Jim Braatz,267 Joan Solà Peracaula,8
John Blakeslee,268 John Webb,78 Jose A. R. Cembranos,269 José Pedro Mimoso,107 Joseph Jensen,270 Juan
García-Bellido,271 Judit Prat,272 Kathleen Sammut,2 Kay Lehnert,273 Keith R. Dienes,274, 275 Kishan Deka,6 Konrad
Kuijken,276 Krishna Naidoo,277 László Árpád Gergely,278, 279 Laur Järv,280 Laura Mersini-Houghton,281 Leila L.
Graef,282 Léo Vacher,283 Levon Pogosian,284 Lilia Anguelova,285 Lindita Hamolli,286 Lu Yin,287, 288 Luca
403

Caloni,227, 228, 289 Luca Izzo,290 Lucas Macri,291 Luis E. Padilla,292 Luz Ángela García,293 Maciej Bilicki,294 Mahdi
Najafi,192, 295 Manolis Plionis,21, 296, 297 Manuel Gonzalez-Espinoza,298, 299 Manuel Hohmann,280 Marcel A. van der
Westhuizen,100 Marcella Marconi,31 Marcin Postolak,300 Marco de Cesare,301, 302 Marco Regis,303 Marek Biesiada,6
Maret Einasto,304 Margus Saal,305 Maria Caruana,2 Maria Petronikolou,306, 152 Mariam Bouhmadi-López,307, 308
Mariana Melo,148, 149 Mariaveronica De Angelis,200 Marie-Noëlle Célérier,309 Marina Cortês,107 Mark Reid,182
Markus Michael Rau,16, 310 Martin S. Sloth,311 Martti Raidal,126 Masahiro Takada,312 Masoume Reyhani,313, 295
Massimiliano Romanello,314 Massimo Marengo,124 Mathias Garny,315 Matías Leizerovich,316, 317 Matteo
Martinelli,318, 319 Matteo Tagliazucchi,320, 58 Mehmet Demirci,321 Miguel A. S. Pinto,107, 214 Miguel A. Sabogal,62
Miguel A. García-Aspeitia,322 Milan Milošević,323 Mina Ghodsi,324, 103 Mustapha Ishak,325 Nelson J. Nunes,107
Nick Samaras,326 Nico Hamaus,327 Nico Schuster,97, 327 Nicola Borghi,320, 58, 328 Nicola Deiosso,173 Nicola
Tamanini,329 Nicolao Fornengo,330 Nihan Katırcı,331 Nikolaos E. Mavromatos,106, 332 Nikolaos Petropoulos,333
Nikolina Šarčević,334 Nils A. Nilsson,335, 336 Noemi Frusciante,337 Octavian Postavaru,338 Oem Trivedi,339, 340
Oleksii Sokoliuk,341, 342, 343 Olga Mena,176 Paloma Morilla,344 Paolo Campeti,345, 346 Paolo Salucci,283 Paula
Boubel,347 Paweł Bielewicz,6 Pekka Heinämäki,348 Petar Suman,349 Petros Asimakis,350 Pierros Ntelis,97 Pran
Nath,351 Predrag Jovanović,352 Purba Mukherjee,114, 353 Radosław Wojtak,354 Rafaela Gsponer,355 Rafid
H. Dejrah,356 Rahul Shah,353 Rasmi Hajjar,176 Rebecca Briffa,2 Rebecca Habas,357 Reggie C. Pantig,358
Renier Mendoza,131, 132 Riccardo Della Monica,169 Richard Stiskalek,359 Rishav Roshan,360 Rita B. Neves,1
Roberto Molinaro,239 Roberto Terlevich,361, 78, 362 Rocco D’Agostino,85, 363 Rodrigo Sandoval-Orozco,364
Ronaldo C. Batista,365 Ruchika Kaushik,169 Ruth Lazkoz,366 Saeed Rastgoo,367, 368, 369 Sahar Mohammadi,370
Salvatore Samuele Sirletti,371, 372, 373 Sandeep Haridasu,374, 375 Sanjay Mandal,376 Saurya Das,377 Sebastian
Bahamonde,378, 379 Sebastian Grandis,380 Sebastian Trojanowski,6 Sergei D. Odintsov,381, 382 Sergij Mazurenko,383
Shahab Joudaki,173, 384 Sherry H. Suyu,385, 386 Shouvik Roy Choudhury,387 Shruti Bhatporia,9 Shun-Sheng Li,388, 389
Simeon Bird,390 Simon Birrer,391 Simone Paradiso,392 Simony Santos da Costa,393, 394 Sofia Contarini,395
Sophie Henrot-Versillé,396 Spyros Basilakos,21, 397, 398 Stefano Casertano,5 Stefano Gariazzo,399 Stylianos A.
Tsilioukas,400, 21 Surajit Kalita,401, 9 Suresh Kumar,402 Susana J. Landau,403 Sveva Castello,404 Swayamtrupta
Panda,405, 406 Tanja Petrushevska,407 Thanasis Karakasis,106 Thejs Brinckmann,408, 409 Tiago B. Gonçalves,107, 214
Tiziano Schiavone,410, 411 Tom Abel,412 Tomi Koivisto,163, 413 Torsten Bringmann,414 Umut Demirbozan,415 Utkarsh
Kumar,416, 417 Valerio Marra,418, 180, 181 Maurice H.P.M. van Putten,419 Vasileios Kalaitzidis,420 Vasiliki A.
Mitsou,421 Vasilios Zarikas,422 Vedad Pasic,423 Venus Keus,424, 425 Verónica Motta,426 Vesna Borka Jovanović,185
Víctor H. Cárdenas,426 Vincenzo Ripepi,239 Vincenzo Salzano,248 Violetta Impellizzeri,427 Vitor da Fonseca,107
Vittorio Ghirardini,428, 429 Vladas Vansevičius,430 Weiqiang Yang,431 Wojciech Hellwing,294 Xin Ren,432, 433 Yu-Min
Hu,432, 434 Yuejia Zhai,1 Abdul Malik Sultan,435 Adrienn Pataki,436 Alessandro Santoni,437, 438 Aliya Batool,435
Aneta Wojnar,439, 440 Arman Tursunov,441, 442 Avik De,443 Ayush Hazarika,444 Baojiu Li,445 Benjamin Bose,446, 447
Bivudutta Mishra,448 Bobomurat Ahmedov,449, 450 Chandra Shekhar Saraf,13 Claudia Scóccola,451 Crescenzo
Tortora,452 D’Arcy Kenworthy,453 Daniel E. Holz,454 David F. Mota,455 David S. Pereira,107 Devon M. Williams,456
Dillon Brout,457 Dong Ha Lee,1 Eduardo Guendelman,458 Edward Olex,459 Emanuelly Silva,460 Emre Onur
Kahya,461 Enzo Brocato,462, 463 Eva-Maria Mueller,464 Felipe Andrade-Oliveira,465 Feven Markos Hunde,294 F. R.
Joaquim,466, 467 Florian Pacaud,468 Francis-Yan Cyr-Racine,469 Pozo Nuñez, F,470 Gábor Rácz,471 Gene Carlo
Belinario,472 Geraint F. Lewis,473 Gergely Dálya,474, 475 Giorgio Laverda,476 Guido Risaliti,477, 478 Guillermo
Franco-Abellán,479 Hayden Zammit,2 Hayley Camilleri,2 Helene M. Courtois,480 Hooman Moradpour,481 Igor de
Oliveira Cardoso Pedreira,482 Ilídio Lopes,483 István Csabai,484 James W. Rohlf,485 Jana Bogdanoska,486 Javier
de Cruz Pérez,487 Joan Bachs-Esteban,476 Joseph Sultana,488 Julien Lesgourgues,489 Jun-Qian Jiang,490
Karem Peñaló Castillo,47 Lavinia Heisenberg,491 Laxmipriya Pati,163 Léon V.E. Koopmans,492 Lokesh kumar
Duchaniya,493 Lucas Lombriser,494 María Pérez Garrote,169 Mariano Domínguez,495 Marine Samsonyan,496 Mark
Pace,497, 2 Martin Krššák,498 Masroor C. Pookkillath,499 Matteo Peronaci,129 Matteo Piani,476 Matthildi
Raftogianni,134 Meet J. Vyas,339 Melina Michalopoulou,134 Merab Gogberashvili,500 Michael Klasen,501 Michele
Cicoli,502, 503 Michele Moresco,320, 58 Miguel Quartin,504, 505 Miguel Zumalacárregui,506 Milan S. Dimitrijević,507
Milos Dordevic,508 Mindaugas Karčiauskas,509 Morgan Le Delliou,510, 511, 512 Nastassia Grimm,404 Nicolás
Augusto Kozameh,495 Nicoleta Voicu,40 Nicolina Pop,513 Nikos Chatzifotis,106 Oliver Fabio Piattella,514, 515, 516
Paula Boubel,347 Pedro da Silveira Ferreira,517 Péter Raffai,436, 518 Peter Schupp,519 Pilar Ruiz-Lapuente,520
Pradyumn Kumar Sahoo,493 Roberto V. Maluf,521 Ruth Durrer,522 S. A. Kadam,523 Sabino Matarrese,524 Samuel
Brieden,525 Santiago González-Gaitán,526 Santosh V. Lohakare,493 Scott Watson,527, 528 Shao-Jiang Wang,529, 530
Simão Marques Nunes,107 Soumya Chakrabarti,531 Suvodip Mukherjee,532 Tajron Jurić,533 Tessa Baker,534
404

Theodoros Nakas,335 Tiago Barreiro,107, 535 Upala Mukhopadhyay,536 Veljko Vujčić,537 Violetta Sagun,538
Vladimir A. Srećković,539 Wangzheng Zhang,540 Yo Toda,541 Yun-Song Piao,490, 542, 543 and Zahra Davari544
1
School of Mathematical and Physical Sciences, University of Sheffield,
Hounsfield Road, Sheffield S3 7RH, United Kingdom
2
Institute of Space Sciences and Astronomy, University of Malta, Malta
3
Department of Physics, University of Malta, Malta
4
Department of Physics and Astronomy, Johns Hopkins University, Baltimore, MD 21218, USA
5
Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
6
National Centre for Nuclear Research, Pasteura 7, 02-093, Warszawa, Poland
7
Laboratoire Univers & Particules de Montpellier (LUPM),
CNRS & Université de Montpellier (UMR-5299),
Place Eugène Bataillon, F-34095 Montpellier Cedex 05, France
8
Departament de Física Quàntica i Astrofísica and Institut de Ciències del Cosmos,
Universitat de Barcelona, Av. Diagonal 647, E-08020 Barcelona, Catalonia, Spain
9
High Energy Physics, Cosmology and Astrophysics Theory (HEPCAT) Group,
Department of Mathematics and Applied Mathematics,
University of Cape Town, Cape Town 7700, South Africa
10
McWilliams Center for Cosmology and Astrophysics,
Department of Physics, Carnegie Mellon University,
5000 Forbes Avenue, Pittsburgh, PA 15213, USA
11
Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cambridge, MA 02138, USA
12
School of Mathematical and Physical Sciences, The University of Sheffield,
Hounsfield Road, S3 7RH Sheffield, United Kingdom
13
Korea Astronomy and Space Science Institute, 776 Daedeok-daero, Yuseong-gu, Daejeon, Republic of Korea
14
Physics Department, National Sun Yat-Sen University,
No. 70, Lien-Hai Rd, Kaosiung City 80424, Taiwan, R.O.C
15
Fakultät für Physik, Universität Bielefeld, Postfach 100131, 33501 Bielefeld, Germany
16
School of Mathematics, Statistics and Physics, Newcastle University,
Herschel Building, NE1 7RU, Newcastle-upon-Tyne, UK
17
Mathematical Institute, Leiden University, Gorlaeus Gebouw,
Einsteinweg 55, NL-2333 CC Leiden, The Netherlands
18
Instituto de Astronomía Teórica y Experimental,
UNC-Conicet, Laprida 854, Ciudad de Córdoba, Argentina
19
Department of Astronomy, University of Geneva, Ch. d’Ecogia 16, CH-1290 Versoix, Switzerland
20
Laboratoire Univers & Particules de Montpellier,
CNRS & Université de Montpellier (UMR-5299), 34095 Montpellier, France
21
National Observatory of Athens, Lofos Nymfon 11852, Greece
22
CAS Key Laboratory for Researches in Galaxies and Cosmology, School of Astronomy and Space Science,
University of Science and Technology of China, Hefei 230026, China
23
Departamento de Matemáticas, Universidad Católica del Norte,
Avda. Angamos 0610, Casilla 1280, Antofagasta, Chile
24
Atlantic Technological University, Ash Lane, Sligo, Ireland
25
Institute for Astronomy, University of Edinburgh Royal Observatory Edinburgh, Blackford Hill, Edinburgh, EH9 3HJ, UK
26
Nordita, KTH Royal Institute of Technology and Stockholm University,
Hannes Alfvéns väg 12, SE-106 91 Stockholm, Sweden
27
Scuola Superiore Meridionale, Largo S. Marcellino 10, I-80138 Napoli, Italy
28
Istituto Nazionale di Fisica Nucleare, Sezione di Napoli,
Complesso Univ. Monte S. Angelo, I-80126 Napoli, Italy
29
Departament de Física Teórica and IFIC, Universitat de València-CSIC, E-46100, Burjassot, Spain
30
Dipartimento di Fisica Ettore Pancini, Universit‘a di Napoli “Federico II”,
Complesso Univ. Monte S. Angelo, I-80126 Napoli, Italy
31
INAF-OACN Osservatorio Astronomico di Capodimonte, Salita Moiariello 16, 80131, Napoli, Italy
32
Institute of Cosmology & Gravitation, University of Portsmouth,
Dennis Sciama Building, Burnaby Road, Portsmouth PO1 3FX, UK
33
Institute for Astronomy, University of Hawaii,
2680 Woodlawn Drive, Honolulu, HI 96822, USA
34
Physics Department, University of Richmond, Richmond, VA 23173 USA
35
Department of Mathematics, Universitat Poliècnica de Catalunya, Carrer Colom 15, 08222 Terrassa, Spain
36
Jodrell Bank Centre for Astrophysics, School of Physics and Astronomy,
The University of Manchester, Manchester M13 9PL, UK
37
Instituto de Estructura de la Materia, CSIC, Serrano 121, 28006 Madrid, Spain
38
University of Montpellier, France
39
School of Mathematics and Physics, University of Queensland, 4072, Australia
405

40
Department of Mathematics and Computer Science, Transilvania University of Brasov, Romania
41
Department of Physics, University of Ioannina, 45110 Ioannina, Greece
42
Institute of Space Sciences (ICE-CSIC), Campus UAB,
Carrer de Can Magrans, s/n, E-08193 Barcelona, Spain
43
Institut d’Estudis Espacials de Catalunya (IEEC), 08860 Castelldefels (Barcelona), Spain
44
European Space Agency (ESA), ESA Office, Space Telescope Science Institute,
3700 San Martin Drive, Baltimore, MD 21218, USA
45
Dipartimento di Fisica “E.R. Caianiello”, Università degli Studi di Salerno, Via G. Paolo II, 84084, Fisciano (SA), Italy
46
INFN - Gruppo Collegato di Salerno, Via G.Paolo II, 84084 Fisciano (SA), Italy
47
Department of Physics and Astronomy, Lehman College, City University of New York, NY 10468, USA
48
School of Physics, Inst. for Research in Fundamental Science (IPM), P.O.Box: 19395-5531, Tehran, Iran
49
Dipartimento di Fisica e Scienze della Terra, Università degli Studi di Ferrara, via Saragat 1, I-44122 Ferrara, Italy
50
Istituto Nazionale di Fisica Nucleare, Sezione di Ferrara, Via G. Saragat 1, I-44122 Ferrara, Italy
51
Sorbonne Université, CNRS, UMR 7095, Institut d’Astrophysique de Paris, 98 bis bd Arago, 75014 Paris, France
52
National Astronomical Observatory of Japan, 2-21-1 Osawa,
Mitaka, Tokyo 181-8588, Japan and Nevada Center for Astrophysics,
University of Nevada, Las Vegas, 4505 Maryland Parkway, Las Vegas, 89154
53
Department of Physics, University of Oxford, Denys Wilkinson Building,
Keble Road, Oxford OX1 3RH, United Kingdom
54
Istituto Nazionale di Fisica Nucleare, Sezione di Ferrara, Via Giuseppe Saragat, 1, 44122, Ferrara, Italy
55
Physics Department and INFN sezione di Ferrara,
Università degli Studi di Ferrara, via Saragat 1, I-44122 Ferrara, Italy
56
INAF, Osservatorio Astronomico d’Abruzzo, Via Maggini snc, I-64100 Teramo, Italy
57
Università di Bologna, Dipartimento di Fisica e Astronomia “Augusto Righi”, via Piero Gobetti 93/2, I-40129 Bologna, Italy
58
INAF - Osservatorio di Astrofisica e Scienza dello Spazio di Bologna, via Piero Gobetti 93/3, I-40129 Bologna, Italy
59
University Observatory, Faculty of Physics, Ludwig-Maximilians-Universität, Scheinerstrasse 1, 81677 Munich, Germany
60
Excellence Cluster ORIGINS, Boltzmannstr 2, 85748 Garching, Germany
61
Department of Physics, Istanbul Technical University, Maslak 34469 Istanbul, Türkiye
62
Instituto de Física, Universidade Federal do Rio Grande do Sul, 91501-970 Porto Alegre RS, Brazil
63
Divisão de Astrofísica, Instituto Nacional de Pesquisas Espaciais,
Avenida dos Astronautas 1758, São José dos Campos, 12227-010, São Paulo, Brazil
64
Asia Pacific Center for Theoretical Physics, Pohang 37673, Korea
65
Universidad Nacional Autónoma de México, Instituto de Radioastronomía y Astrofísica, 58090, Morelia, Michoacán, México
66
Secretaría de Ciencia, Humanidades, Tecnología e Innovación,
Av. Insurgentes Sur 1582, 03940, Ciudad de México, México
67
Institute of Physics, École Polytechnique Fédérale de Lausanne (EPFL),
Observatoire de Sauverny, 1290 Versoix, Switzerland
68
Department of Physics, Willamette University, Salem, OR 97301, USA
69
Dipartimento di Fisica Ettore Pancini, Università di Napoli “Federico II”,
Complesso Univ. Monte S. Angelo, I-80126 Napoli, Italy
70
Istituto Nazionale di Fisica Nucleare, Sezione di Napoli,
Complesso Universitario di Monte S. Angelo, Via Cintia Edificio 6, 80126 Napoli, Italy
71
Department of Physics, University of Trento, Via Sommarive 14, 38123 Povo (TN), Italy
72
Trento Institute for Fundamental Physics and Applications-INFN, Via Sommarive 14, 38123 Povo (TN), Italy
73
Department of Mathematics, Presidency University, 86/1 College Street, Kolkata 700073, India
74
Department of Physics and Astronomy University of California USA
75
Kavli Institute for Cosmology (KICC), Madingley Road, CB3 0HA, Cambridge, United Kingdom
76
Centre for Astrophysics Research, Department of Physics,
Astronomy and Mathematics, University of Hertfordshire,
College Lane, Hatfield, AL10 9AB, United Kingdom
77
Kavli Institute for Cosmology (KICC), Madingley Road, CB3 0HA, Cambridge, UK
78
Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 9AL, UK
79
Laboratoire Univers et Particules de Montpellier (LUPM),
CNRS et Université de Montpellier (UMR-5299),
Place Eugène Bataillon, F-34095 Montpellier Cedex 05, France
80
Imperial Centre for Inference and Cosmology (ICIC), Department of Physics,
Blackett Laboratory, Imperial College London, Prince Consort Road, London SW7 2AZ, UK
81
NASA Goddard Space Flight Center, 8800 Greenbelt Road, Greenbelt, MD 20771, USA
82
Erlangen Centre for Astroparticle Physics, Friedrich-Alexander-Universität Erlangen-Nürnberg,
Department of Physics, Nikolaus-Fiebiger-Straße 2, Erlangen, 91058, Germany
83
Astronomical Observatory, Jagiellonian University in Kraków, ul. Orla 171, 30-244 Kraków, Poland
84
Dipartimento di Fisica “G. Marconi”, Università di Roma Sapienza, Ple Aldo Moro 2, 00185, Rome, Italy
85
INAF - Osservatorio Astronomico di Roma, Via Frascati 33, 00078 Monte Porzio Catone, Italy
406

86
INFN - Sezione di Roma, Piazzale Aldo Moro,
2 - c/o Dipartimento di Fisica, 00185 Roma, Italy
87
Dipartimento di Fisica, Università degli Studi di Roma “Tor Vergata”,
via della Ricerca Scientifica 1,I-00133 Rome, Italy
88
Department of Astrophysical Sciences, Princeton University, Peyton Hall, Princeton, NJ 08544, USA
89
Kavli Institute for Cosmology (KICC), University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK
90
Instituto de Física, Universidade Federal Fluminense, 24210-346 Niterói, RJ, Brazil
91
Institute for Particle Physics and Astrophysics, ETH Zürich,
Wolfgang-Pauli-Strasse 27, CH8093 Zürich, Switzerland
92
Astronomical Observatory, Odessa I.I. Mechnikov National University, Dvoryanskaya St. 2, Odessa 65082, Ukraine
93
Center for Advanced Systems Understanding, Untermarkt 20, 02826 Görlitz, Germany
94
Helmholtz-Zentrum Dresden-Rossendorf, Bautzner Landstraße 400, 01328 Dresden, Germany
95
INAF - Osservatorio Astrofisico di Catania, Via S.Sofia 78 I-95123 Catania,
INFN, Sezione di Catania, Via. S.Sofia 64, I-95123 Catania, italy
96
Physics Department, Eastern Mediterranean University, Famagusta, Cyprus
97
Aix-Marseille Université, CNRS/IN2P3, CPPM, Marseille, France
98
Department of Astrophysical Sciences, Peyton Hall,
Princeton University, Princeton, NJ 08544, USA
99
School of Astronomy, Institute for Research in Fundamental Sciences (IPM), P. O. Box 19395-5531, Tehran, Iran
100
Centre for Space Research, North-West University, Potchefstroom, South Africa
101
Department of Physics and Astronomy, Union College,
807 Union Street, Schenectady, NY 12308, USA
102
Instituto Nacional de Astrofísica, Óptica y Electrónica, Puebla, México
103
MTA-CSFK Lendület Large-scale Structure Research Group,
H-1121 Budapest, Konkoly Thege Miklós út 15-17, Hungary
104
Konkoly Observatory, HUN-REN CSFK, MTA Centre of Excellence,
Budapest, Konkoly Thege Miklós út 15-17. H-1121 Hungary
105
Department of Physics, University of Patras, 26500 Patras, Greece
106
National Technical University of Athens, School of Applied Mathematics
and Physical Sciences, Physics Division, Athens GR15780, Greece
107
Instituto de Astrofísica e Ciências do Espaço,
Faculdade de Ciências da Universidade de Lisboa,
Edifício C8, Campo Grande, P-1749-016 Lisbon, Portugal
108
School for Data Science and Computational Thinking and Department of Mathematical Sciences,
Stellenbosch University, Stellenbosch, 7602, South Africa
109
Centre for Space Research, North-West University, Potchefstroom 2520, South Africa
110
Departamento de Matemáticas, Universidad Católica del Norte,
Avda. Angamos 0610, Casilla 1280 Antofagasta, Chile
111
Institute of Theoretical Physics, University of Wroclaw, pl. M. Borna 9, 50-204 Wroclaw, Poland
112
Department of Physics, United College of Engineering and Research, Greater Noida 201310, India
113
Department of Mathematics, Indira Gandhi University, Meerpur, Haryana 122502, India
114
Centre For Theoretical Physics, Jamia Millia Islamia, New Delhi, India
115
LUX, CNRS, Observatoire de Paris, France
116
DAMTP, Centre for Mathematical Sciences, University of Cambridge, Cambridge, CB3 0WA, UK
117
Department of Astronomy & Astrophysics, University of Chicago, Chicago, IL 60637, USA
118
Kavli Institute for Cosmological Physics, University of Chicago, Chicago, IL 60637, USA
119
Center for Astronomy, Space Science and Astrophysics,
Independent University, Bangladesh, Dhaka 1229, Bangladesh
120
Astrophysics Research Centre, University of KwaZulu-Natal, Westville Campus, Durban 4041, South Africa
121
School of Mathematics, Statistics and Computer Science,
University of KwaZulu-Natal, Westville Campus, Durban 4041, South Africa
122
Department of Physics, 9500 Gilman Drive, San Diego, CA 92122, USA
123
University of Geneva, 51 chemin de Pegasi, 1290 Versoix, Switzerland
124
Florida State University, Department of Physics,
77 Chieftain Way, Tallahassee. FL 32306, USA
125
Center for Gravitational Physics and Quantum Information,
Yukawa Institute for Theoretical Physics, Kyoto University, 606-8502, Kyoto, Japan
126
National Institute of Chemical Physics and Biophysics, Rävala 10, 10143 Tallinn, Estonia
127
Department of Physics and EHU Quantum Center,
University of the Basque Country UPV/EHU, Barrio Sarriena s/n, 48940 Leioa, Spain
128
Physics Department, Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional (Cinvestav),
PO. Box 14-740, Av. Instituto Politécnico Nacional 2508, Mexico City, Mexico
129
Dipartimento di Fisica and INFN Sezione di Roma 2,
Università di Roma Tor Vergata, via della Ricerca Scientifica 1, 00133 Rome, Italy
407

130
Observatório Nacional, Rua General José Cristino 77,
São Cristóvão, 20921-400 Rio de Janeiro, RJ, Brasil
131
Institute of Mathematics, University of the Philippines Diliman, Quezon City, Metro Manila, Philippines
132
Computational Science Research Center, University of the Philippines Diliman, Quezon City, Metro Manila, Philippines
133
Niels Bohr Institute, Blegdamsvej 17, DK-2100 Copenhagen, Denmark
134
Division of Applied Analysis, Department of Mathematics,
University of Patras, Rio Patras GR-26504, Greece
135
Division of Theoretical Physics, Rudjer Bošković Institute, Bijenička 54, 10000 Zagreb, Croatia
136
Department of Electrical Engineering, Mazandaran University of Science and Technology, Babol, Iran
137
Institute for Gravitation and the Cosmos, Department of Physics and Department of Astronomy and Astrophysics,
Pennsylvania State University, University Park, PA 16802, USA
138
Universidad de Concepción, Departamento de Astronomía, Casilla 160-C, Concepción, Chile
139
Department of Physics and Astronomy, Sejong University, Seoul 05006, Korea
140
Université de Strasbourg, CNRS UMR 7550, Observatoire astronomique de Strasbourg,
11 rue de l’Université, 67000 Strasbourg, France
141
Center for Particle Cosmology, Department of Physics and Astronomy,
University of Pennsylvania, Philadelphia, PA 19104, USA
142
Nevada Center for Astrophysics and Department of Physics and Astronomy,
University of Nevada, Las Vegas, Las Vegas, NV 89154, USA
143
Institute of Physics, University of Silesia in Katowice, Poland
144
Mathematical Institute of the Serbian Academy of Sciences and Arts, Kneza Mihaila 36, 11000 Belgrade, Serbia
145
Department of Physics, Lafayette College, Easton, PA 18042, USA
146
Max Planck Institute for Extraterrestrial Physics, Garching, Germany
147
Department of Physics, University of the Basque Country UPV/EHU, P.O. Box 644, 48080 Bilbao, Spain
148
Faculdade de Ciências, Universidade do Porto,
Rua do Campo Alegre, 4150-007 Porto, Portugal
149
Instituto de Astrofísica e Ciências do Espaço,
Universidade do Porto, Rua das Estrelas, 4150-762 Porto, Portugal
150
Universidad Nacional Autónoma de México, Cuernavaca, México
151
Department of Physics, National & Kapodistrian University of Athens, Zografou Campus GR 157 73, Athens, Greece
152
National Observatory of Athens, Lofos Nymfon, 11852 Athens, Greece
153
Key Laboratory of Dark Matter and Space Astronomy, Purple Mountain Observatory,
Chines Academy of Sciences, Nanjing 210023, People’s Republic of China
154
School of Astronomy and Space Science, University of Science
and Technology of China, Hefei 230026, People’s Republic of China
155
SISSA - International School for Advanced Studies, Via Bonomea 265, 34136 Trieste, Italy
156
Sapienza University,Piazzale Aldo Moro, c/o Dipartimento di Fisica, Edificio E. Fermi, Roma, Italy
157
Center of applied space technology and microgravity (ZARM),
University of Bremen, Am Fallturm 2, 28359 Bremen, Germany
158
Centre for Space Research, North-West University,
Private Bag X6001, Potchefstroom, South Africa
159
Centro de Física das Universidades do Minho e do Porto,
Faculdade de Ciências da Universidade do Porto, Rua do Campo Alegre s/n, 4169-007 Porto
160
Universidade dos Açores Instituto de Investigação em Ciências do Mar - OKEANOS,
Campus da Horta, Rua Professor Doutor Frederico Machado 4, 9900-140 Horta, Portugal
161
Centro Brasileiro de Pesquisas Físicas, Rua Dr. Xavier Sigaud 150, CEP 22290-180, Rio de Janeiro, RJ, Brazil
162
Departament d’Informàtica, Matemàtica Aplicada i Estadística,
Universitat de Girona, Campus Montilivi, 17003 Girona, Spain
163
Laboratory of Theoretical Physics, Institute of Physics,
University of Tartu, W. Ostwaldi 1, 50411 Tartu, Estonia
164
Department of Physics and Astronomy, Haverford College, PA 19041, USA
165
Department of Physics, Duke University, Durham, NC 27708, USA
166
Depto. de Física Teórica, Facultad de Ciencias Físicas,
Universidad Complutense de Madrid, Plaza de las Ciencias 1, 28040 Madrid, Spain
167
Main Astronomical Observatory of National Academy of Sciences of Ukraine,
27 Akademik Zabolotnyi St. 03143, Kyiv, Ukraine
168
Departamento de Matemáticas, Universidad de León, Campus de Vegazana, s/n 24071 León, Spain
169
Departamento de Física Fundamental and Instituto Universitario de Física Fundamental y Matemáticas (IUFFyM),
Universidad de Salamanca, Plaza de la Merced, s/n, E-37008 Salamanca, Spain
170
Leibniz-Institut für Astrophysik Potsdam (AIP),
An der Sternwarte 16, 14482 Potsdam, Germany
171
Canada-France-Hawaii Telescope, 65-1238 Mamalahoa Hwy, Waimea, HI 96743, USA
172
School of Physical and Chemical Sciences, University of Canterbury,
Private Bag 4800, Christchurch 8140, New Zealand
408

173
Centro de Investigaciones Energéticas Medioambientales y Tecnológicas CIEMAT, Av Complutense 40, 28040 Madrid, Spain
174
Physik-Institut, University of Zürich, Winterthurerstrasse 190, CH-8057 Zürich, Switzerland
175
Instituto Tecnológico de Piedras Negras, Mexico
176
Instituto de Física Corpuscular (IFIC), University of Valencia-CSIC,
Parc Científic UV, c/ Catedrático José Beltrán 2, E-46980 Paterna, Spain
177
Bulgarian Academy of Sciences, Institute for Nuclear Research and Nuclear Energy,
72 Tszarigradsko Chaussee, Sofia 1784, Bulgaria
178
Department of Physics, Aristotle University of Thessaloniki, A.U.Th. Campus, 54635, Thessaloniki, Greece
179
Departamento de Física Teórica and IPARCOS,
Universidad Complutense de Madrid, E-28040 Madrid, Spain
180
INAF - Osservatorio Astronomico di Trieste, via Tiepolo 11, 34131 Trieste, Italy
181
IFPU - Institute for Fundamental Physics of the Universe, via Beirut 2, 34151, Trieste, Italy
182
Center for Astrophysics | Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138, USA
183
Black Hole Initiative, Harvard University, 20 Garden Street, Cambridge, MA 02138, USA
184
STAR Institute, University of Liège, Quartier Agora, Allée du six Août 19c, 4000 Liège, Belgium
185
Department of Theoretical Physics and Condensed Matter Physics (020),
Vinča Institute of Nuclear Sciences - National Institute of the Republic of Serbia,
University of Belgrade, P.O. Box 522, 11001 Belgrade, Serbia
186
School of Astronomy, Institute for Research in Fundamental Sciences (IPM), P.O. Box 19395-5531, Tehran, Iran
187
Innovation and Management Research Center, Ayatollah Amoli Branch,
Islamic Azad University, Amol, Mazandaran, Iran
188
Physics Department, Michigan Technological University, 1400 Townsend Dr, Houghton, MI 49931, USA
189
Instituto Nacional de Astrofisica, Optica y Electronica, L.E.Erro N.1, Tonantzintla, Puebla, Mexico
190
Dipartimento di Fisica e Astronomia “Augusto Righi” - Università di Bologna, via Piero Gobetti 93/2, I-40129 Bologna, Italy
191
Department of Physics, College of Science, Kuwait University,
Sabah Al Salem University City, P.O. Box 2544, Safat 1320, Kuwait
192
Physics Department, Sapienza University of Rome, P.le A. Moro 5, 00185 Roma, Italy
193
Istituto Nazionale di Fisica Nucleare (INFN),
Sezione di Roma, P.le A. Moro 5, I-00185, Roma, Italy
194
Physics Department, Tor Vergata University of Rome,
Via della Ricerca Scientifica 1, 00133 Roma, Italy
195
Kavli Institute for the Physics and Mathematics of the Universe (WPI),
UTIAS, The University of Tokyo, Chiba 277-8583, Japan
196
Faculty of Natural Sciences and Mathematics,
Department of Mathematics, University of Tuzla,
Ul. Urfeta Vejzagića br. 4, 75000 Tuzla, Bosnia and Herzegovina
197
Van Swinderen Institute for Particle Physics and Gravity,
University of Groningen, Nijenborgh 3, 9747 AG Groningen, The Netherlands
198
Department of Physics and Astronomy, Aarhus University, DK-8000 Aarhus C, Denmark
199
Department of Physics and Astronomy, University of New Mexico Albuquerque NM 87131, USA
200
School of Mathematical and Physical Sciences,
University of Sheffield, Hounsfield Road, Sheffield S3 7RH, UK
201
INAF - Osservatorio Astronomico d’Abruzzo, Teramo, Italy
202
INAF - Osservatorio Astronomico di Roma, via Frascati 33, 00078 Monte Porzio Catone (RM), Italy
203
Department of Mathematics, University College London, Gower Street, London WC1E 6BT, UK
204
Department of Physics, Indian Institute of Technology, Guwahati 781039, India
205
Institute for Astronomy, University of Hawaii,
2680 Woodlawn Drive, 96822 Honolulu, HI, USA
206
Scuola Superiore Meridionale, Via Mezzocannone 4, 80134 Napoli, Italy
207
Istituto Nazionale di Fisica Nucleare, Sez. di Napoli, Via Cinthia 21, 80126 Napoli, Italy
208
Departamento de Física Teórica and IFIC, Centro Mixto Universitat de
València - CSIC. Universitat de València, Burjassot-46100, Valencia, Spain
209
Department of Informatics and Telecommunications,
University of Peloponnese, Karaiskaki 70, 22100, Tripoli, Greece
210
Dipartimento di Fisica, Università degli Studi di Torino, Via P. Giuria 1, I-10125, Torino, Italy
211
INFN - Sezione di Torino, Via P. Giuria 1, I-10125, Torino, Italy
212
INAF - Istituto Nazionale di Astrofisica, Osservatorio Astrofisico di Torino,
strada Osservatorio 20, 10025, Pino torinese, Italy
213
Département de Physique Theorique and Center for Astroparticle Physics,
Université de Genève, 24 quai Ernest Ansermet, 1211 Genève 4, Switzerland
214
Departamento de Física, Faculdade de Ciências da Universidade de Lisboa,
Edifício C8, Campo Grande, P-1749-016 Lisbon, Portugal
215
Institut de Ciències del Cosmos, Universitat de Barcelona,
Martí i Franquès, 1, E-08028 Barcelona, Spain
409

216
ICREA, Pg. Lluís Companys 23, Barcelona, E-08010, Spain
217
Institute of Theoretical Astrophysics, University of Oslo, PO Box 1029 Blindern, 0315 Oslo, Norway
218
Department of Physics and Astronomy, University of North Carolina, Chapel Hill, NC 27599-3255, USA
219
Department of Physics, University of Malta, Msida, Malta
220
Department of Physics and Astronomy, University of California, Davis, CA, USA
221
Instituto de Ciencias Fı́sicas, Universidad Nacional Autónoma de México, 62210, Cuernavaca, Morelos, México
222
Departamento de Matemáticas, Universidad Católica del Norte,
Avda. Angamos 0610, Casilla 1280 Antofagasta, Chile
223
Institute of Systems Science, Durban University of Technology, PO Box 1334, Durban 4000, South Africa
224
Physics Department, National Technical University of Athens, 15780 Zografou Campus, Athens, Greece
225
Faculty of Science and Technology, University of Stavanger, 4036 Stavanger, Norway
226
European Southern Observatory, Karl-Schwarzschild str. 2., Garching 85748, Germany
227
Dipartimento di Fisica e Scienze della Terra, Università degli Studi di Ferrara, Via Giuseppe Saragat 1, I-44122 Ferrara, Italy
228
Istituto Nazionale di Fisica Nucleare, Sezione di Ferrara, Via Giuseppe Saragat 1, I-44122 Ferrara, Italy
229
INFN - Laboratori Nazionali di Frascati (LNF), Via E. Fermi 54, Frascati, Roma, 00044, Italy
230
National & Kapodistrian University of Athens,
Department of Physics, Zografou Campus GR 157 73, Athens, Greece
231
ENEA, Fusion and Nuclear Safety Department,
C.R. Frascati, Via E. Fermi 45, Frascati, 00044, Italy
232
Physics Department, “Sapienza” University of Rome, P.le Aldo Moro 5, Rome, 00185, Italy
233
Departamento de Física, Facultad de Ciencias,
Universidad de Tarapacá, Casilla 7-D, Arica, Chile
234
INAF - Osservatorio Astronomico di Capodimonte, Salita Moiariello 16, 80131, Napoli, Italy
235
INFN-Naples Section, Complesso Universitario di Monte S.Angelo, Via Cinthia, Ed. NI-80126, Napoli, Italy
236
INAF - Osservatorio Astronomico di Roma, Via Frascati 33, I-00040 Monte Porzio Catone, Roma, Italy
237
Dipartimento di Ingegneria, Università LUM, S.S. 100 km 18 70010, Casamassima, Bari, Italy
238
Department of Chemistry, Physics and Environmental and Soil Sciences,
Escola Politècnica Superior, Universidad de Lleida, Av. Jaume II, 69, 25001 Lleida, Spain
239
INAF-Osservatorio Astronomico di Capodimonte, Salita Moiariello 16, 80131, Napoli, Italy
240
Universidade Federal do Ceará (UFC), Departamento de Física,
Campus do Pici, Fortaleza - CE, C.P. 6030, 60455-760, Brazil
241
Department of Physics, University of Nis, Visegradska 33,
Nis, Serbia. SEENET-MTP Centre, Nis, Serbia
242
ESTEC - European Space Agency, Keplerlaan 1, 2201 AZ Noordwijk, The Netherlands
243
Tsung-Dao Lee Institute (TDLI), No. 1 Lisuo Road, 201210 Shanghai, China
244
School of Physics and Astronomy, Shanghai Jiao Tong University, Dongchuan Road 800, 201240 Shanghai, China
245
Institute for Astronomy, the School of Physics,
Zhejiang University, Hangzhou 310027, People’s Republic of China
246
Department of Physics, Southern University of Science and Technology, Shenzhen 518055, China
247
Department of Physics & Astronomy, University of Kansas,
1251 Wescoe Hall Drive, Lawrence KS, 66045, USA
248
Institute of Physics, University of Szczecin, Wielkopolska 15, 70-451 Szczecin, Poland
249
Physics Department, Ariel University, Ariel
250
High Energy Physics Research Unit, Faculty of Science,
Chulalongkorn University, Bangkok 10330, Thailand
251
CEICO, Institute of Physics, Czech Academy of Sciences,
Na Slovance 2, 182 21 Praha 8, Czech Republic
252
Main Astronomical Observatory of the National Academy of
Sciences of Ukraine, Akademik Zabolotnyi 27, Kyiv, 03143, Ukraine
253
Department of Astronomy of the University of Geneva, 51 chemin de Pegasi, 1290 Versoix, Switzerland
254
Department of Theoretical Physics, University of the Basque Country UPV/EHU, P.O. Box 644, 48080 Bilbao, Spain
255
Shamakhy Astrophysical Observatory, 5626, Shamakhy, Azerbaijan
256
Physics Department, Swansea University, UK
257
Instituto de Ciencias Físicas, Universidad Nacional Autónoma de México, Cuernavaca, México
258
Departamento de Física Teórica, Centro de Astropartículas y Física de Altas Energías,
Universidad de Zaragoza, 50009 Zaragoza, Spain
259
Institute of Physics, University of Silesia, Katowice, Poland
260
Univ. Lille, CNRS, Centrale Lille, UMR 9189 CRIStAL, F-59000 Lille, France
261
Université Paris-Saclay, CNRS, Institut d’Astrophysique Spatiale, 91405, Orsay, France
262
Helsinki Institute of Physics, P.O. Box 64, FI-00014 University of Helsinki, Finland
263
Academia Sinica Institute of Astronomy and Astrophysics,
11F of AS/NTU Astronomy-Mathematics Building, Roosevelt Rd, Taipei 106216, Taiwan, R.O.C
264
Bahamas Advanced Study Institute and Conferences, 4A Ocean Heights,
Hill View Circle, Stella Maris, Long Island, The Bahamas
410

265
Department of Physics & Astronomy, University of Hawai’i at Manoa,
Watanabe Hall, 2505 Correa Road, Honolulu, HI, 96822, USA
266
Leung Center for Cosmology & Particle Astrophysics,
National Taiwan University, Taipei 10617, Taiwan
267
National Radio Astronomy Observatory, 520 Edgemont Road, Charlottesville, VA 22903, USA
268
NSF NOIRLab, 950 N. Cherry Avenue, Tucson, AZ 85719, USA
269
Departamento de Física Teórica and IPARCOS, Facultad de Ciencias Físicas,
Universidad Complutense de Madrid, E-28040 Madrid, Spain
270
Utah Valley University, Orem, Utah, USA
271
Instituto de Fisica Teórica UAM/CSIC, Universidad Autonoma de Madrid, Cantoblanco 28049 Madrid, Spain
272
Nordita, KTH Institute of Technology and Stockholm University, SE-106 91 Stockholm
273
Department of Physics, National University of Ireland, Maynooth, Ireland
274
Department of Physics, University of Arizona, Tucson, AZ 85721, USA
275
Department of Physics, University of Maryland, College Park, MD 20742, USA
276
Leiden Observatory, Leiden University, P.O. Box 9513, 2300 RA Leiden, The Netherlands
277
Department of Physics and Astronomy, University College London, Gower Street, London WC1E 6BT, UK
278
Department of Theoretical Physics, University of Szeged,
Tisza Lajos krt. 84-86, H-6720 Szeged, Hungary
279
Department of Theoretical Physics, HUN-REN Wigner Research Centre for Physics,
Konkoly-Thege Miklós út 29-33, H-1121 Budapest, Hungary
280
Institute of Physics, University of Tartu, Estonia
281
Department of Physics and Astronomy, UNC-Chapel Hill, USA
282
Instituto de Física, Universidade Federal Fluminense, 24210-346 Niteroi, RJ, Brazil
283
SISSA, Scuola Internazionale Superiore di Studi Avanzati - via Bonomea, 265 - 34136 Trieste, Italy
284
Department of Physics, Simon Fraser University, Burnaby, BC V5A 1S6, Canada
285
INRNE, Bulgarian Academy of Sciences, Sofia 1784, Bulgaria
286
Department of Physics, University of Tirana, Boulevard “Zogu I”, Tirana, Albania
287
Department of Physics, Shanghai University, Shanghai, 200444, China
288
Asia Pacific Center for Theoretical Physics, Pohang, 37673, Korea
289
CFisUC, Department of Physics, University of Coimbra, P-3004 - 516 Coimbra, Portugal
290
INAF, Osservatorio Astronomico di Capodimonte, Salita Moiariello 16,
I-80131 Napoli, Italy & DARK, Niels Bohr Institute,
University of Copenhagen, Jagtvej 128, 2200 Copenhagen, Denmark
291
NSF NOIRLab, 950 N Cherry Ave, Tucson AZ 85719 USA
292
Astronomy Unit, Queen Mary University of London, Mile End Road, London, E1 4NS, UK
293
Universidad ECCI, Cra. 19 No. 49-20, Bogotá, Colombia, Código Postal 111311
294
Center for Theoretical Physics, Polish Academy of Sciences, al. Lotników 32/46, 02-668 Warsaw, Poland
295
PDAT Laboratory, Department of Physics, K.N. Toosi University of Technology, P.O. Box 15875-4416, Tehran, Iran
296
Department of Physics, Aristotle University of Thessaloniki, 54124, Thessaloniki, Greece
297
CERIDES- Excellence in Innovation and Technology,
European University of Cyprus, 1516, Nicosia, Cyprus
298
Laboratorio de investigación de Cómputo de Física, Facultad de Ciencias Naturales y Exactas,
Universidad de Playa Ancha, Subida Leopoldo Carvallo 270, Valparaíso, Chile
299
Laboratorio de Didáctica de la Física, Departamento de Matemática,
Física y Computación, Facultad de Ciencias Naturales y Exactas,
Universidad de Playa Ancha, Subida Leopoldo Carvallo 270, Valparaíso, Chile
300
University of Wrocław, Institute of Theoretical Physics, pl. Maxa Borna 9, 50-206 Wrocław, Poland
301
Scuola Superiore Meridionale, Largo San Marcellino 10, 80138 Napoli, Italy
302
INFN - Sezione di Napoli, Complesso Universitario Monte S. Angelo, 80126 Napoli, Italy
303
University of Torino and INFN, via P. Giuria 1, 10125 Torino, Italy
304
Tartu Observatory, University of Tartu, Observatooriumi 1, 61602 Tõravere, Estonia
305
Institute of Physics, University of Tartu, W. Ostwaldi 1, 50411 Tartu, Estonia
306
Physics Department, School of Applied Mathematical and Physical Sciences,
National Technical University of Athens, 15780 Zografou Campus, Athens, Greece
307
IKERBASQUE, Basque Foundation for Science, 48011, Bilbao, Spain
308
Department of Physics & EHU Quantum Center,
University of the Basque Country UPV/EHU, P.O. Box 644, 48080 Bilbao, Spain
309
Laboratoire d’étude de l’Univers et des phénomènes eXtrêmes,
Observatoire de Paris, UMR 8262 CNRS, Sorbonne Université
310
High Energy Physics Division, Argonne National Laboratory, Lemont, IL 60439, USA
311
Universe-Origins, University of Southern Denmark, Campusvej 55, 5230 Odense M, Denmark
312
Kavli Institute for the Physics and Mathematics of the Universe (WPI),
The University of Tokyo Institutes for Advanced Study (UTIAS),
The University of Tokyo, Chiba 277-8583, Japan
411

313
Department of Physics, K.N. Toosi University of Technology, P.O. Box 15875-4416, Tehran, Iran
314
Dipartimento di Fisica e Astronomia “A. Righi” - Alma Mater Studiorum Università di Bologna,
via Piero Gobetti 93/2, 40129 Bologna, Italy and INAF - Osservatorio di Astrofisica
e Scienza dello Spazio di Bologna, via Piero Gobetti 93/3, 40129 Bologna, Italy
315
Technical University Munich, School of Natural Sciences,
Physik Department T31, James-Franck Str. 1, 85748 Garching, Germany
316
Universidad de Buenos Aires, Facultad de Ciencias Exactas y Naturales,
Departamento de Física. Buenos Aires, Argentina
317
CONICET - Universidad de Buenos Aires, Instituto de Física de Buenos Aires (IFIBA). Buenos Aires, Argentina
318
INAF-Osservatorio Astronomico di Roma, Via Frascati 33, 00078 Monteporzio Catone, Italy
319
INFN-Sezione di Roma, Piazzale Aldo Moro, 2 - c/o Dipartimento di Fisica, Edificio G. Marconi, 00185 Roma, Italy
320
Dipartimento di Fisica e Astronomia “Augusto Righi” - Università di Bologna, via Piero Gobetti 93/2, I-40129 Bologna, Italy
321
Department of Physics, Karadeniz Technical University, Trabzon, TR61080, Türkiye
322
Depto. de Física y Matemáticas, Universidad Iberoamericana Ciudad de México,
Prolongación Paseo de la Reforma 880, México D. F. 01219, Mxico
323
Faculty of Sciences and Mathematics, University on Nis, Serbia
324
Konkoly Observatory, HUN-REN Research Centre of Astronomy and Earth Sciences (CSFK),
MTA Centre of Excellence, Budapest, Konkoly Thege Miklós út 15-17. H-1121 Hungary
325
Department of Physics, The University of Texas at Dallas, Richardson, TX 75080, USA
326
Astronomical Institute, Faculty of Mathematics and Physics,
Charles University, V Holešovičkách 2, CZ-180 00 Praha 8, Czech Republic
327
Universitäts-Sternwarte München, Fakultät für Physik,
Ludwig-Maximilians-Universität München, Scheinerstr. 1, D-81679 München, Germany
328
INFN-Sezione di Bologna, Viale Berti Pichat 6/2, 40127 Bologna, Italy
329
Laboratoire des 2 Infinis - Toulouse (L2IT-IN2P3),
Université de Toulouse, CNRS, F-31062 Toulouse Cedex 9, France
330
University of Torino and INFN/Torino, via P. Giuria 1, 10125 Torino, Italy
331
Department of Electrical and Electronics Engineering Doğuş University Ümraniye, 34775 Istanbul, Türkiye
332
Physics Department, King’s College London, Strand, London WC2R 2LS, UK
333
Department of Physics, Faculty of Science, University of Thessaly, GR 35100, Greece
334
Duke University, Durham, NC 27708, USA
335
Cosmology, Gravity and Astroparticle Physics Group,
Center for Theoretical Physics of the Universe, Institute for Basic Science, Daejeon 34126, Korea
336
LTE, Observatoire de Paris, Université PSL, CNRS, LNE,
Sorbonne Université, 61 avenue de l’Observatoire, 75 014 Paris, France
337
Dipartimento di Fisica “E. Pancini”, Università degli Studi di Napoli “Federico II",
Compl. Univ. di Monte S. Angelo, Edificio G, Via Cinthia, I-80126, Napoli, Italy
338
Center for Research and Training in Innovative Techniques of Applied Mathematics in Engineering,
University Politehnica of Bucharest, 060042 Bucharest, Romania
339
International Centre for Space and Cosmology,
Ahmedabad University, Ahmedabad 380009, India
340
Department of Physics and Astronomy, Vanderbilt University, Nashville, TN, 37235, USA
341
Main Astronomical Observatory of the National Academy of Sciences of Ukraine,
27 Akademik Zabolotny St., Kyiv, 03143, Ukraine
342
Astronomical Observatory, Taras Shevchenko National University of Kyiv, 3 Observatorna St., 04053 Kyiv, Ukraine
343
Department of Physics, University of Aberdeen, Aberdeen AB24 3UE, UK
344
Department of Physics, University of the Basque Country, Spain
345
INFN Sezione di Ferrara, Via Saragat 1, 44122 Ferrara, Italy
346
ICSC, Centro Nazionale “High Performance Computing, Big Data and Quantum Computing”, Italy
347
Research School of Astronomy and Astrophysics, The Australian National University,
Mount Stromlo Observatory, Canberra, ACT 2611, Australia
348
Department of Physics and Astronomy, Vesilinnantie 5, University of Turku, 20014 Turku, Finland
349
Centre for Theoretical Cosmology, Department of Applied Mathematics and Theoretical Physics,
University of Cambridge, Wilberforce Road, Cambridge, CB3 0WA, UK
350
Department of Physics, School of Applied Mathematical and Physical Sciences,
National Technical University of Athens, 9 Iroon Polytechniou Str., Zografou Campus GR 157 80, Athens, Greece
351
Department of Physics, Northeastern University, Boston, MA 02115, USA
352
Astronomical Observatory, Volgina 7, P.O. Box 74, 11060 Belgrade, Serbia
353
Physics and Applied Mathematics Unit, Indian Statistical Institute, 203 B.T. Road, Kolkata 700 108, India
354
DARK, Niels Bohr Institute, University of Copenhagen, Jagtvej 155, 2200 Copenhagen, Denmark
355
Institute of Physics, Laboratory of Astrophysics,
École Polytechnique Fédérale de Lausanne (EPFL),
Observatoire de Sauverny, CH-1290 Versoix, Switzerland
356
Department of Physics, Ankara University, Faculty of Sciences, 06100, Ankara, Turkiye
412

357
INAF - Osservatorio Astronomico d’Abruzzo, Via Maggini, 64100, Teramo, Italy
358
Physics Department, Mapúa University, 658 Muralla St., Intramuros, Manila 1002, Philippines
359
Astrophysics, University of Oxford, Denys Wilkinson Building, Keble Road, Oxford, OX1 3RH, UK
360
School of Physics and Astronomy, University of Southampton, Southampton, UK
361
Instituto Nacional de Astrofísica, Óptica y Electrónica,Tonantzintla, Puebla, México
362
Facultad de Astronomía y Geofísica, Universidad de La Plata, La Plata, Argentina
363
INFN - Sezione di Roma 1, P.le Aldo Moro 2, 00185 Roma, Italy
364
Instituto de Astronomía, Universidad Nacional Autónoma de México, Av. Universidad 3000,
Col. Universidad Nacional Autónoma de México, C.P. 04510, Ciudad de México
365
Escola de Ciências e Tecnologia, Universidade Federal do Rio Grande do Norte,
Campus Universitário Lagoa Nova, 59078-970, Natal, RN, Brazil
366
Physics Department, University of the Basque Country UPV/EHU, 644 PO Box, 48080 Bilbao, Spain
367
Department of Physics, University of Alberta, Edmonton, Alberta T6G 2E1, Canada
368
Department of Mathematical and Statistical Sciences,
University of Alberta, Edmonton, Alberta T6G 2G1, Canada
369
Theoretical Physics Institute, University of Alberta, Edmonton, Alberta T6G 2E1, Canada
370
Islamic Azad University, Science and Research Branch, Tehran, Iran
371
Dipartimento di Fisica e Scienze della Terra, Università degli Studi di Ferrara, Ferrara, 44122, Italy
372
Dipartimento di Fisica, Università di Trento, Trento, 38123, Italy
373
Department of Physics, Columbia University, New York, NY, USA
374
SISSA, Via Bonomea 265, 34136 Trieste, Italy
375
Institute for Fundamental Physics of the Universe (IFPU), Via Beirut 2, 34014 Trieste, Italy
376
Faculty of Symbiotic Systems Science, Fukushima University, Fukushima 960-1296, Japan
377
Theoretical Physics Group and Quantum Alberta,
Department of Physics and Astronomy, University of Lethbridge,
4401 University Drive, Lethbridge, Alberta T1K 7Z2, Canada
378
Kavli Institute for the Physics and Mathematics of the Universe (WPI),
The University of Tokyo Institutes for Advanced Study (UTIAS),
The University of Tokyo, Kashiwa, Chiba 277-8583, Japan
379
Cosmology, Gravity, and Astroparticle Physics Group, Center for Theoretical Physics of the Universe,
Institute for Basic Science (IBS), Daejeon, 34126, Korea
380
Universität Innsbruck, Institut für Astro- und Teilchenphysik, Technikerstrasse 25, 6020 Innsbruck, Austria
381
Institute of Space Sciences (ICE, CSIC) C. Can Magrans s/n, 08193 Barcelona, Spain
382
ICREA, Passeig Lluis Companys, 23, 08010 Barcelona, Spain
383
Universität Bonn, Regina-Pacis-Weg 3, 53113 Bonn, Germany
384
Institute of Cosmology & Gravitation, Dennis Sciama Building,
University of Portsmouth, Portsmouth PO1 3FX, UK
385
Technical University of Munich, TUM School of Natural Sciences,
Physics Department, James-Franck-Straße 1, 85748 Garching, Germany
386
Max-Planck-Institut für Astrophysik, Karl-Schwarzschild Straße 1, 85748 Garching, Germany
387
Institute of Astronomy and Astrophysics, Academia Sinica,
No. 1, Section 4, Roosevelt Road, Taipei 106319, Taiwan
388
Ruhr University Bochum, Faculty of Physics and Astronomy, Astronomical Institute (AIRUB),
German Centre for Cosmological Lensing, 44780 Bochum, Germany
389
Leiden Observatory, Leiden University, Einsteinweg 55, 2333 CC Leiden, The Netherlands
390
University of California, Riverside, CA 92507, USA
391
Department of Physics and Astronomy, Stony Brook University, Stony Brook, NY 11794, USA
392
INAF - Osservatorio di astrofisica e scienza dello spazio, via Gobetti 93/3, 40129, Bologna, Italy
393
Dipartimento di Fisica, Universit‘a di Trento, Via Sommarive 14, 38123 Povo, Trento, Italy
394
Trento Institute for Fundamental Physics and Applications (TIFPA), Via Sommarive 14, 38123 Povo, Trento, Italy
395
Max Planck Institute for Extraterrestrial Physics, Giessenbachstrasse 1, 85748 Garching, Germany
396
Universite Paris-Saclay, CNRS/IN2P3, IJCLab, 91405 Orsay, France
397
Academy of Athens, Research Center for Astronomy and Applied Mathematics, Soranou Efesiou 4, 11527, Athens, Greece
398
School of Sciences, European University Cyprus, Diogenes Street, Engomi, 1516 Nicosia, Cyprus
399
Università di Torino, Physics department, via P. Giuria 1, 10125 Turin, Italy
400
Department of Physics, University of Thessaly, 35100 Lamia, Greece
401
Astronomical Observatory, University of Warsaw,
Al. Ujazdowskie 4, PL-00478 Warszawa, Poland
402
Data Science Institute, Plaksha University, Mohali, Punjab-140306, India
403
CONICET-Universidad de Buenos Aires-Instituto de Física de Buenos Aires,
Av. Intendente Cantilo S/N 1428, CABA, Argentina
404
Département de Physique Théorique and Center for Astroparticle Physics,
Université de Genève, Quai E. Ansermet 24, CH-1211 Geneva 4, Switzerland
405
International Gemini Observatory, NSF NOIRLab, Casilla 603, La Serena, Chile
413

406
Laboratorio Nacional de Astrofisica - MCTI, Itajuba - MG, 37504-364, Brazil
407
Center for Astrophysics and Cosmology, University of Nova Gorica, Vipavska 11c, 5270 Ajdovščina, Slovenia
408
Dipartimento di Fisica e Scienze della Terra, Università degli Studi di Ferrara, Via G. Saragat 1, I-44122 Ferrara, Italy
409
Istituto Nazionale di Fisica Nucleare (INFN),
Sezione di Ferrara, Via G. Saragat 1, I-44122 Ferrara, Italy
410
Institute of Space Sciences (ICE,CSIC), C. Can Magrans s/n, 08193 Barcelona, Spain
411
Galileo Galilei Institute for Theoretical Physics, Largo Enrico Fermi 2, I-50125 Firenze, Italy
412
Department of Physics, Simon Fraser University, Burnaby, BC, V5A 1S6, Canada
413
National Institute of Chemical Physics and Biophysics, Rävala pst. 10, 10143 Tallinn, Estonia
414
Department of Physics, University of Oslo, Box 1048, N-0316 Oslo, Norway
415
Institut de Física d’Altes Energies (IFAE), The Barcelona Institute of Science and Technology,
Campus UAB, 08193 Bellaterra (Barcelona) Spain
416
Department of Physics, University of Ottawa, 75 Laurier Ave E, Ottawa, Canada
417
Astrophysics Research Center of the Open University, The Open University of Israel, Ra’anana, Israel
418
Departamento de Física, Universidade Federal do Espírito Santo, 29075-910, Vitória, ES, Brazil
419
Department of Physics and Astronomy, Sejong Universe, Seoul 05006,
South Korea, and INAF-OAS, Bologna, via P. Gobetti 101 I-40129, Italy
420
Scottish Universities Physics Alliance, University of Saint Andrews,
North Haugh, Saint Andrews, Fife, KY16 9SS, UK
421
Instituto de Física Corpuscular (IFIC), CSIC – Universitat de València,
C/ Catedrático José Beltrán 2, 46980 Paterna (Valencia), Spain
422
Department of Mathematics, University of Thessaly, GR-35100, Greece
423
Department of Mathematics, Faculty of Natural and Mathematical Sciences,
University of Tuzla, Urfeta Vejzagica 4, 75000 Tuzla, Bosnia and Herzegovina
424
School of Theoretical Physics, Dublin Institute for Advanced Studies (DIAS), 10 Burlington Road, D04 C932, Dublin, Ireland
425
Department of Physics and Helsinki Institute of Physics,
Gustaf Hallstromin katu 2, FIN-00014, University of Helsinki, Finland
426
Instituto de Física y Astronomía, Universidad de Valparaíso, Avda. Gran Bretaña 1111, Valparaíso, Chile
427
Leiden Observatory, Leiden University, PO Box 9513, 2300 RA Leiden, The Netherlands
428
Max-Planck-Institut für extraterrestrische Physik, Gießenbachstraße 1, 85748 Garching, Germany
429
INAF, Osservatorio di Astrofisica e Scienza dello Spazio, via Piero Gobetti 93/3, 40129 Bologna, Italy
430
Center for Physical Sciences and Technology, Saulėtekio av. 3, 10257 Vilnius, Lithuania
431
Department of Physics, Liaoning Normal University, Dalian, 116029, P. R. China
432
Department of Astronomy, School of Physical Sciences,
University of Science and Technology of China, Hefei 230026, China
433
Department of Physics, Institute of Science Tokyo, Tokyo 152-8551, Japan
434
CAS Key Laboratory for Research in Galaxies and Cosmology, School of Astronomy and Space Science,
University of Science and Technology of China, Hefei 230026, China
435
Department of Mathematics, University of Okara, Okara-56300 Pakistan
436
Institute of Physics and Astronomy, ELTE Eötvös Loránd University, 1117 Budapest, Hungary
437
Facultad de Física, Pontificia Universidad Católica de Chile, Vicuña Mackenna 4860, Santiago, Chile
438
Institut für Theoretische Physik and Atominstitut, Technische Universität Wien,
Wiedner Hauptstrasse 8–10, A-1040 Vienna, Austria
439
University of Wroclaw, Poland
440
Complutense University of Madrid
441
Max Planck Institute for Radio Astronomy, Auf dem Huegel 69, D-53121 Bonn, Germany
442
Institute of Physics, Silesian University in Opava, CZ-74601 Opava, Czech Republic
443
Universiti Malaya, 50603 Kuala Lumpur, Wilayah Persekutuan Kuala Lumpur
444
Department of Physics, Tezpur University, Napaam, Tezpur, 784028, Assam, India
445
Institute for Computational Cosmology, Department of Physics,
Durham University, South Road, Durham DH1 3LE, UK
446
Institute for Astronomy, University of Edinburgh,
Royal Observatory, Blackford Hill, Edinburgh, EH9 3HJ, UK
447
Higgs Centre for Theoretical Physics, School of Physics and Astronomy, Edinburgh, EH9 3FD, UK
448
Department of Mathematics, BITS-Pilani Hyderabad Campus, India
449
Institute for Advanced Studies, New Uzbekistan University,
Movarounnahr str. 1, Tashkent 100000, Uzbekistan
450
Institute of Theoretical Physics, National University of Uzbekistan, Tashkent 100174, Uzbekistan
451
Departamento de Física, Facultad de Ciencias Físicas y Matemáticas, Universidad de Chile
452
INAF - Osservatorio Astronomico di Capodimonte, Salita Moiariello 16, 80131 - Napoli, Italy
453
The Oskar Klein Centre, Department of Physics,
Stockholm University, SE - 106 91 Stockholm, Sweden
454
Department of Physics, Department of Astronomy & Astrophysics,
KICP, and EFI, University of Chicago, Chicago, IL 60637, USA
414

455
Institute of Theoretical Astrophysics, University of Oslo, 0315 Oslo, Norway
456
Department of Physics and Astronomy, PAB, 430 Portola Plaza,
Box 951547, Los Angeles, CA 90095-1547, USA
457
Departments of Astronomy and Physics, Boston University, Boston MA, 02215, USA
458
Ben Gurion University of the Negev, Beer Sheva, Israel
459
Departamento de Física Teórica, Módulo 15, Facultad de Ciencias,
Universidad Autónoma de Madrid, 28049 Madrid, Spain
460
Instituto de Física, Universidade Federal do Rio Grande do Sul, 91501-970 Porto Alegre RS, Brazil
461
Department of Physics, Faculty of Arts and Sciences, Istanbul Technical University, Istanbul, Turkey
462
INAF – Osservatorio Astronomico d’Abruzzo, via Mentore Maggini snc I-64100 Teramo, Italy
463
INAF - Osservatorio Astronomioco di Roma (OAR),
via Frascati 33, 00078 Monte Porzio Catone (RM), Italy
464
Department of Physics and Astronomy, University of Sussex, Brighton BN1 9QH, UK
465
University of Zurich, Winterthurerstrasse 190, 8057, Zurich, Switzerland
466
Instituto Superior Tecnico - IST, Universidade de Lisboa, 1049-001 Lisboa, Portugal
467
Laboratorio de Instrumenta¸cao e Fısica Experimental de Partıculas,
1649-003 Lisboa and 3004-516 Coimbra, Portugal
468
Argelander-Institut für Astronomie, Universität Bonn, Auf dem Hügel 71, 53121 Bonn, Germany
469
Department of Physics and Astronomy, University of New Mexico, Albuqerque, NM 87106, USA
470
Heidelberg Institute for Theoretical Studies (HITS),
Schloss-Wolfsbrunnenweg 35, 69118 Heidelberg, Germany
471
Department of Physics, University of Helsinki,
Gustaf Hällströmin katu 2, FI-00014 Helsinki, Finland
472
National Institute of Physics, University of the Philippines, Quezon City, Philippines
473
Sydney Institute for Astronomy, School of Physics,
A28, The University of Sydney, NSW 2006, Australia
474
L2IT, Laboratoire des 2 Infinis - Toulouse, CNRS/IN2P3, UPS, F-31062 Toulouse Cedex 9, France
475
MTA-ELTE Astrophysics Research Group, 1117 Budapest, Hungary
476
Centro de Astrofísica e Gravitação - CENTRA,
Departamento de Física, Instituto Superior Técnico - IST,
Universidade de Lisboa - UL, Av. Rovisco Pais 1, 1049-001 Lisboa, Portugal
477
Dipartimento di Fisica e Astronomia, Università di Firenze,
via G. Sansone 1, 50019 Sesto Fiorentino, Firenze, Italy
478
INAF – Osservatorio Astrofisico di Arcetri, Largo Enrico Fermi 5, I-50125 Firenze, Italy
479
GRAPPA Institute, Institute for Theoretical Physics Amsterdam,
University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands
480
Université Claude Bernard Lyon 1, IUF, IP2I Lyon,
4 rue Enrico Fermi, 69622 Villeurbanne, France
481
Research Institute for Astronomy and Astrophysics of Maragha (RIAAM), P.O. Box 55134-441, Maragha, Iran
482
Instituto de Física, Universidade Federal Fluminense,
Avenida General Milton Tavares de Souza s/n, Gragoat´a, 24210-346 Niteroi, Rio de Janeiro, Brazil
483
Centro de Astrofísica e Gravitação - CENTRA,
Departamento de Física, Instituto Superior Técnico - IST,
Universidade de Lisboa - UL, Av. Rovisco Pais 1, 1049-001 Lisboa, Portugal
484
Department of Physics of Complex Systems, ELTE Eötvös Loránd University, 1117 Budapest, Hungary
485
Department of Physics, Boston University, Boston, Massachusetts, 02215 USA
486
Institute of Physics, Ss Cyril and Methodius University in Skopje, North Macedonia
487
Departamento de Física, Universidad de Córdoba, E-14071, Córdoba, Spain
488
Department of Mathematics, Faculty of Science, University of Malta
489
Institute for Theoretical Particle Physics and Cosmology (TTK),
RWTH Aachen University, 52056 Aachen, Germany
490
School of Physical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China
491
Institute for Theoretical Physics, Universität Heidelberg, Philosophenweg 16, 69120 Heidelberg, Germany
492
Kapteyn Astronomical Insitute, University of Groningen,
P.O.Box 800, 9700AV Groningen, The Netherlands
493
Department of Mathematics, Birla Institute of Technology and Science-Pilani, Hyderabad Campus, Hyderabad-500078, India
494
Département de Physique Théorique, Université de Genève,
24 quai Ernest Ansermet, 1211 Genève 4, Switzerland
495
IATE-OAC-UNC and CONICET, Laprida 854,
Barrio Observatorio, Cordoba 5000, Cordoba, Argentina
496
Center for Cosmology and Astrophysics, Alikhanian National Laboratory and Yerevan State University, Yerevan, Armenia
497
Institute of Applied Sciences, The Malta College of Arts, Science and Technology, Poala, Malta
498
Department of Theoretical Physics, Faculty of Mathematics,
Physics and Informatics, Comenius University in Bratislava, 84248, Slovak Republic
415

499
Centre for Theoretical Physics and Natural Philosophy, Mahidol University,
Nakhonsawan Campus, Phayuha Khiri, Nakhonsawan 60130, Thailand
500
Javakhishvili Tbilisi State University, 3 Chavchavadze Avenue, Tbilisi 0179, Georgia
501
Institut für Theoretische Physik, Universität Münster, 48149 Münster, Germany
502
Dipartimento di Fisica e Astronomia, Università di Bologna, via Irnerio 46, 40126 Bologna, Italy
503
INFN, Sezione di Bologna, viale Berti Pichat 6/2, 40127 Bologna, Italy
504
Instituto de Física & Observatório do Valongo,
Universidade Federal do Rio de Janeiro, Rio de Janeiro, RJ, Brazil
505
PPG Cosmo, Universidade Federal do Espírito Santo, 29075-910, Vitória, ES, Brazil
506
Max Planck Institute for Gravitational Physics (Albert Einstein Institute),
Am Muhlenberg 1, D-14476 Potsdam-Golm, Germany
507
Astronomical Observatory, Volgina 7, 11060 Belgrade, Serbia
508
Vinca Institute of Nuclear Sciences - National Institute of the Republic of Serbia, University of Belgrade
509
Center for Physical Sciences and Technology, Sauletekio av. 3, 10257 Vilnius, Lithuania
510
Institute of Theoretical Physics & Research Center of Gravitation; Key Laboratory of
Quantum Theory and Applications of MoE; and Lanzhou Center for Theoretical Physics & Key
Laboratory of Theoretical Physics of Gansu Province, Lanzhou University, Lanzhou 730000, China
511
Instituto de Astrofísica e Ciências do Espaço, Universidade de Lisboa,
Faculdade de Ciências, Ed. C8, Campo Grande, 1769-016 Lisboa, Portugal
512
Université de Paris-Cité, APC-Astroparticule et Cosmologie (UMR-CNRS 7164), F-75006 Paris, France
513
Department of Physical Foundations of Engineering,
Politehnica University of Timisoara, 300223 Timisoara, Romania
514
Dipartimento di Scienza e Alta Tecnologia (DiSAT),
Università degli Studi dell’Insubria, via Valleggio 11, 22100, Como, Italy
515
Istituto Nazionale di Fisica Nucleare (INFN),
Sezione di MIlano, via Celoria 16, 20126, Milano, Italy
516
Como Lake centre for AstroPhysics (CLAP), DiSAT,
Università dell’Insubria, via Valleggio 11, 22100 Como, Italy
517
Centro Brasileiro de Pesquisas Físicas, Rua Dr. Xavier Sigaud 150, 22290-180 Rio de Janeiro, RJ, Brazil
518
HUN-REN–ELTE Extragalactic Astrophysics Research Group, 1117 Budapest, Hungary
519
School of Science, Constructor University, 28759 Bremen, Germany
520
Instituto de Física Fundamental (CSIC), Serrano 123, 28006 Madrid,
Spain and Institute of Cosmos Science (ICCUB), Martí i Franques 1, 08028 Barcelona, Spain
521
Universidade Federal do Ceará (UFC), Departamento de Física,
Campus do Pici, Fortaleza - CE, C.P. 6030, 60455-760 - Brazil
522
Département de Physique Thèorique, Université de Genève, Quai E. Ansemet 24, 1211 Geneve, Switzerland
523
Department of Mathematics, School of Computer Science and Artificial Intellegence,
S R Univrsity, Warangal, 506371, Telangana India
524
Dipartimento di Fisica e Astronomia G. Galilei, Università degli Studi di Padova, Padova, Italy
525
Institute for Astronomy, University of Edinburgh,
Royal Observatory Edinburgh, Blackford Hill, Edinburgh, EH9 3HJ, UK
526
Instituto de Astrofísica e Ciências do Espaço,
Faculdade de Ciências da Universidade de Lisboa, Campo Grande, PT1749-016 Lisboa, Portugal
527
Department of Physics, Syracuse University, Syracuse, NY 13244, USA
528
Department of Physics and Astronomy, University of South Carolina, Columbia, SC 29208, USA
529
Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing 100190, China
530
Asia Pacific Center for Theoretical Physics (APCTP), Pohang 37673, Korea
531
Department of Physics, School of Advanced Sciences, Vellore Institute of Technology,
Vellore, Tiruvalam Rd, Katpadi, Tamil Nadu 632014, India
532
Department of Astronomy & Astrophysics, Tata Institute of Fundamental Research,
1, Homi Bhabha Road, Colaba, Mumbai 400005, India
533
Ruđer Bošković Institute, Zagreb, Croatia
534
Institute of Cosmology and Gravitation, University of Portsmouth
535
ECEO, Universidade Lusófona, Campo Grande 376, 1749-024 Lisboa, Portugal
536
Department of Physics and Materials Science,
University of Luxembourg, L-1511 Luxembourg City, Luxembourg
537
Astronomical Observatory Belgrade, Volgina 7, 11060 Belgrade, Serbia
538
Mathematical Sciences and STAG Research Centre,
University of Southampton, Southampton SO17 1BJ, United Kingdom
539
Institute of physics Belgrade, Belgrade University, Pregrevica 118, Belgrade, Serbia
540
Department of Physics, the Chinese University of Hong Kong, Sha Tin, NT, Hong Kong
541
Department of Physics, Hokkaido University, Sapporo 060-0810, Japan
542
School of Fundamental Physics and Mathematical Sciences,
Hangzhou Institute for Advanced Study, UCAS, Hangzhou 310024, China
416

543
International Center for Theoretical Physics Asia-Pacific, Beijing/Hangzhou, China
544
School of Physics, Korea Institute for Advanced Study (KIAS),
85 Hoegiro, Dongdaemun-gu, Seoul, 02455, Korea
(April 3, 2025)

You might also like