Next Article in Journal
Graviton Physics: A Concise Tutorial on the Quantum Field Theory of Gravitons, Graviton Noise, and Gravitational Decoherence
Next Article in Special Issue
Gravitational Particle Production and the Hubble Tension
Previous Article in Journal
Direct and Indirect Measurements of the 19F(p,α)16O Reaction at Astrophysical Energies Using the LHASA Detector and the Trojan Horse Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

ΛCDM Tensions: Localising Missing Physics through Consistency Checks

1
Department of Physics, Istanbul Technical University, Maslak, Istanbul 34469, Turkey
2
Atlantic Technological University, Ash Lane, F91 YW50 Sligo, Ireland
3
Centre for Theoretical Physics, Jamia Millia Islamia, New Delhi 110025, India
4
School of Physics, Institute for Research in Fundamental Sciences (IPM), Tehran P.O. Box 19395-5531, Iran
*
Author to whom correspondence should be addressed.
Submission received: 10 June 2024 / Revised: 16 July 2024 / Accepted: 19 July 2024 / Published: 23 July 2024
(This article belongs to the Special Issue Current Status of the Hubble Tension)

Abstract

:
Λ CDM tensions are by definition model-dependent; one sees anomalies through the prism of Λ CDM. Thus, progress towards tension resolution necessitates checking the consistency of the Λ CDM model to localise missing physics either in redshift or scale. Since the universe is dynamical and redshift is a proxy for time, it is imperative to first perform consistency checks involving redshift, then consistency checks involving scale as the next steps to settle the “systematics versus new physics” debate and foster informed model building. We present a review of the hierarchy of assumptions underlying the Λ CDM cosmological model and comment on whether relaxing them can address the tensions. We focus on the lowest lying fruit of identifying missing physics through the identification of redshift-dependent Λ CDM model fitting parameters. We highlight the recent progress made on S 8 : = σ 8 Ω m / 0.3 tension and elucidate how similar progress can be made on H 0 tension. Our discussions indicate that H 0 tension, equivalently a redshift-dependent H 0 , and a redshift-dependent S 8 imply a problem with the background Λ CDM cosmology.

“Will you walk into my parlour?” said a spider to a fly;
“Tis the prettiest little parlour that ever you did spy.
The way into my parlour is up a winding stair,
And I have many pretty things to show when you are there.”
“Oh no, no!" said the little fly, “to ask me is in vain,
For who goes up your winding stair can ne’er come down again.”—The Spider and the Fly, Mary Howitt (1829)

1. Introduction

In the absence of firm theoretical guidance, cosmology makes progress through Occam’s razor and Bayesian model comparison [1,2,3]1. To play this game, one identifies a cosmological model that has (i) a minimal degree of complexity, as quantified by counting the free parameters, and (ii) a physics backstory and narrative. The standard model of cosmology today, the Lambda Cold Dark Matter ( Λ CDM) model, represents the minimal model that explains Cosmic Microwave Background (CMB) data [6]. Both cold dark matter and late-time accelerated expansion driven by dark energy (DE) are observationally well-motivated in a spatially homogeneous and isotropic universe, while the cosmological constant Λ has theoretical pedigree since it can be added to the Einstein field equations for free. Recently, in addition to the long-standing, notoriously challenging theoretical issues related to Λ [7,8,9], observational discrepancies and tensions involving Λ CDM model parameters have emerged between data sets from different cosmic epochs [10,11,12]. If not due to systematics, the tensions point to missing physics not incorporated by the Λ CDM model.
The prevailing responses of the cosmology community to Λ CDM tensions have either involved studying systematics or the exploration of models with new fundamental physics [10,13]. Nevertheless, less effort has been made to localise or narrow down the effects of missing physics either in redshift, a proxy for cosmic time, or scale, a proxy for cosmic distances. On the contrary, one of the recognised front-running proposals to resolve the tension in the Hubble constant H 0 [6,14,15,16,17,18] that invokes a new physics pre-recombination [19,20,21,22,23,24,25] effectively delocalises the missing physics in epochs with at best indirect observational constraints. One turns the H 0 tension from an apparent problem in the redshift range 0 z 1100 2 to a problem in the redshift range 1100 z < , more generally 0 z < ; see [10,11,12] for a comprehensive account of the models injecting new physics at different epochs, and [26] for arguments saying that one must introduce new physics in both the pre- and post-recombination universe. Despite delocalising physics, from the perspective of Bayesian model comparison, these proposals are par for the course. They introduce a finite number of additional degrees of freedom, thereby making the proposals both competitive and testable while also providing a requisite physics backstory. This modelling game is without end.
Against this backdrop, this letter has two objectives. The first is to provide a review of the assumptions underlying the Λ CDM cosmology and comment on whether relaxing them promises to alleviate the Λ CDM tensions. The second objective is to highlight the low lying fruit of better localising the missing physics within the Λ CDM model. This entails conducting consistency checks of the Λ CDM model at different redshifts (or scales),3 where one looks for signatures that the fitting parameters of the model are changing; the redshift or scale dependence of the Λ CDM cosmological parameters signal an inconsistency and breakdown of the model. If signatures of drift or evolution of the fitting parameters can be found, this localises the missing physics to the redshift ranges under study. If they cannot be found, the Λ CDM model performs well in the range. Furthermore, if one succeeds in localising the missing physics to a set of measure zero, this swings the tensions debate back to systematics.
For cosmologists familiar with Bayesian model comparison, the proposed checks of the Λ CDM model may sound unfamiliar. Admittedly, there are key differences. Notably, there is no second model, and one is performing a consistency check of a single model to show that the model fitted to data at different redshift ranges (or scales) leads to consistent fitting parameters. Intuitively, if different subsamples of the data prefer different fitting parameters, this worsens the goodness of fit, so the methodology complements the Bayesian approach. To fully appreciate this, note that, if one fits the Λ CDM model to data in different redshift bins and one finds the same best fit parameters in all the bins, there is little hope of finding a model that beats Λ CDM in model comparison. Only when the redshift evolution of the Λ CDM fitting parameters is established does an alternative model become competitive. For this reason, it should be quicker to diagnose the missing physics through redshift (or scale)-dependent fitting parameters than to construct a new model—necessitated by a Bayesian framework—with additional degrees of freedom that address the evolution. Bluntly put, in any race to diagnose Λ CDM tensions, Bayesian methods are expected to finish second best.

2. Hierarchy of Assumptions

In this section, we provide a review of the assumptions that culminate in the Λ CDM cosmology. We comment on the prospects of relaxing these assumptions to resolve H 0 tension [6,14,15,16,17,18] and S 8 tension [27,28,29,30,31,32]. Our comments overlap with earlier observations in [33], but we expand them to include greater and more recent observational scope.

2.1. Cosmological Bedrock: GR + FLRW

Could relaxing GR resolve ΛCDM tensions? Modern cosmology rests upon General Relativity (GR) and the Einstein field equations,
R μ ν 1 2 R g μ ν + Λ g μ ν = 8 π G c 4 T μ ν ,
where g μ ν is the spacetime metric, R μ ν and R = g μ ν R μ ν are the Ricci (curvature) tensor and scalar, respectively, both derived from g μ ν , Λ is the cosmological constant, T μ ν the total stress-energy tensor, G Newton’s constant, and c the speed of light. Λ is special because it may be consistently added to the left hand side of the Einstein field equations.4 As with any theory, one does not expect GR to hold universally. Notably, the formidable task remains to reconcile GR with quantum physics, the other key pillar of 20th century physics, and curvature singularities within black hole solutions point to regimes where the theory may break down. These observations have motivated extensive theoretical speculation in the cosmology literature [34,35]. In contrast, observationally, GR has passed all tests with flying colours [36]. GR is well-tested to μm scales in the laboratory [37,38], in the weak-field setting in the solar system [39], the Milky Way [40], and at extragalactic scales [41]. Moreover, GR has been tested in the strong-field regime with binary pulsars [42,43,44] and gravitational waves [45,46,47]. It is imperative to continue to test the predictions of GR, but the absence of deviations makes the assumption (1) remarkably solid [48]. Despite investigations of the potential of modified gravity to resolve Λ CDM tensions [49,50,51,52], the difficulties are well-documented [53,54,55,56].
Once one comes to terms with the fact that there is no observational motivation5 for deviating from (1), the next most fundamental assumption in modern cosmology is the cosmological principle. Succinctly put, the universe is a gravitating system, thus necessitating an assumption for g μ ν . The consensus holds that, in accordance with the cosmological principle, at sufficiently large scales, the average evolution of the universe is precisely described by the Friedmann–Lemaître–Robertson–Walker (FLRW) metric,
g μ ν d x μ d x ν = c 2 d t 2 + a ( t ) 2 d Σ ( 3 ) 2 ,
where t is the cosmic time, Σ ( 3 ) is a maximally symmetric 3D space, and time dependence only enters through the scale factor a = a ( t ) . Note, Σ ( 3 ) need not be flat and it can have constant curvature, as hinted at by a spatially closed universe in Planck data [57,58]. However, such an outcome drives H 0 to lower values, H 0 50 km s 1 Mpc 1 , and sits uneasy with the inflationary paradigm, a key theoretical input in modern cosmology. In fact, inflation is not necessarily in conflict with a non-flat space, but spatially closed inflationary models are significantly fine-tuned [59,60].6 For these reasons, we will focus on spatially flat Σ ( 3 ) .7 Once one inserts (2) in (1), H 0 emerges mathematically as an integration constant when one solves the Einstein → Friedmann equations [66]. Relaxing FLRW removes the requirement that the universe’s rate of expansion, as measured by the distance ladder through averages of the recession velocity over the sky, be a constant.8
Could relaxing FLRW resolve ΛCDM tensions? In principle, yes,9 but this outcome would be most compelling if one succeeded in localising missing physics to the local universe z 0.1 because of the corroborating evidence for anisotropies in the same redshift range [74,75,76,77,78,79]. Of course, if one relaxes the FLRW setup, the whole standard model of cosmology and its tensions should be redefined from scratch. That being said, null tests of FLRW have been performed using radio galaxy and quasar (QSO) samples [80,81,82,83,84,85], arriving at findings at odds with FLRW, so one cannot preclude a violation of FLRW at distances beyond 500 Mpc ( z 0.1 ). In contrast to GR, where there are no anomalies worth reporting, recently, some statistically significant deviations from FLRW behaviour have prompted debate [74,75,76,77,78,79,80,81,82,83,84,85,86,87]. Moreover, the independent anomalies appear synergistic, and a preliminary science case has been built [88]. Separately, leveraging longstanding CMB anomalies [65], it has been argued that a statistically anisotropic universe cannot be dismissed [89].10 A plausible and conservative interpretation of some of these results is that FLRW is not a valid description of the local universe out to at least 500 Mpc. This is enough to impact all distance ladder anchor galaxies, and sociologically may be a palatable resolution for cosmologists, allowing one a potential resolution to H 0 tension if the local universe is not FLRW. To circumvent this, one could just push the domain of FLRW cosmology to higher redshifts. Alternatively, one could attempt to model the local universe, but early indications suggest this could make H 0 tension worse [93].
However, one can caution against this idea. We observe that hemisphere decompositions of both the CMB and Type Ia SNe samples under standard assumptions typically do not lead to pronounced enough angular variations of Λ CDM parameters to resolve tensions [94,95,96,97,98]. The key point here is that, while angular variations are evident, angular variations in Planck data centre on H 0 67 km s 1 Mpc 1 , whereas angular variations in Cepheid-SNe centre on H 0 73 km s 1 Mpc 1 . That is, the reported values of H 0 may be viewed as a value averaged over the sky. Furthermore, even allowing for systematics, or alternatively the potential impact of foregrounds on the CMB [99,100,101] in Planck data, terrestrial CMB experiments clearly agree on a lower H 0 < 70 km s 1 Mpc 1 value [102,103]. In summary, while a violation of FLRW in the local universe z 0.1 seems compelling [88], and can impact the distance ladder [104,105,106,107], the prospect of resolving H 0 through relaxing FLRW currently seems distant.
Moving along to S 8 tension, it is worth stressing that, since weak lensing sensitivity peaks at z 0.4 (see Figure 1 from [63]), a local z 0.1 FLRW violation is not expected to address the S 8 tension in weak lensing. Nevertheless, it may explain lower values of S 8 from peculiar velocities [108,109,110] and potentially Sunyaev–Zeldovich (SZ) cluster counts [111] since samples are biased to lower redshifts. Moving beyond the local universe, lower values of S 8 observed in redshift space distortions (RSD) [112,113,114,115,116,117,118,119,120], especially evolution of S 8 or σ 8 with redshift [121,122]11, and evolution in the Weyl potential with DES data [125] point to an S 8 problem at cosmological scales no less than z 0.5 , necessitating an unfathomably large scale violation of FLRW. However, if it turns out that the RSD and Weyl potential results are false, and one succeeds in showing no evolution in the S 8 parameter down to redshifts of z 0.2 , as recently claimed [126]12, one may be able to localise both H 0 and S 8 tension in the relatively local universe. That being said, it is clear that one can currently draw different conclusions from different observables, so systematics are clearly an issue.
The importance of localising missing physics should now be evident to the reader. One could contemplate absorbing both H 0 and S 8 tensions into a local violation of FLRW. However, if tensions cannot be localised enough at low redshifts, this idea does not work unless FLRW is violated at large scales ( z > 0.1 ). As we see it, while radio galaxy and QSO samples used in cosmic dipole studies [80,81,82,83,84,85] must be at high redshift to avoid a spurious clustering dipole [129], they are simply null tests of FLRW. Despite allowing FLRW to be violated at redshifts up to z 1 , they do not narrow down a relevant redshift range.

2.2. Matter, Radiation, and Λ

We now turn our attention to the additional assumptions defining the Λ CDM cosmology. A solution to (1) subject to the spacetime assumption (2), and the further assumption of vanishing spatial curvature, require one to specify the stress-energy tensor T μ ν . Therein lies the physics. We note that one can reconstruct the Hubble parameter H : = a ˙ / a (where a dot denotes d / d t ) directly from observational Hubble data (OHD) [130,131,132,133,134] or other observational data using Gaussian Process regression [135,136,137,138,139], more general machine learning algorithms [140,141,142,143,144,145,146,147,148], and even wavelets anchored on the comoving distance to the last scattering surface [149], but, without a solution to the Einstein field equations, there is no physics. The utility of data reconstructions is that one may diagnose departures from a specific model, e.g., the Λ CDM model, but care is required with quantifying degrees of freedom (see, for example, [150,151,152,153,154]) and confidence intervals [155].13
Returning to specifying T μ ν , we require matter (humans exist) and radiation (the sun shines). Observationally, matter and radiation are beyond question, and we possess a deep theoretical understanding of both. It is a reasonable additional assumption that matter is pressureless at large scales. That being said, one should still test this assumption by checking that any high-redshift Hubble parameter scales as H ( z ) 100 Ω m h 2 ( 1 + z ) 3 2 , (with h : = H 0 / 100 km s 1 Mpc 1 being the reduced Hubble parameter) in the matter-dominated regime. Note, if Ω m h 2 changes with effective redshift, the pressureless matter assumption is rejected. See [161] for a recent study in this direction.
Beyond this point, one runs into stronger assumptions based on weaker supports that are either theoretically or observationally motivated. Theoretical preconceptions about a big bang and inflationary epoch have led to the assumption of purely adiabatic scalar primordial perturbations with a nearly scale-invariant spectrum. Observationally, the matter sector is further decomposed into baryonic and dark matter (DM) components, and the discovery of a late-time accelerated expansion [162,163] necessitates an additional DE sector, which is modelled through Λ . This brings us to the flat Λ CDM Hubble parameter:
H ( z ) = H 0 Ω Λ + Ω m ( 1 + z ) 3 + Ω r ( 1 + z ) 4 , Ω Λ = 1 Ω m Ω r , Ω m = Ω c + Ω b ,
where H 0 , Ω b , Ω c , Ω r , and Ω m are parameters representing the present-day ( z = 0 ) values of the Hubble expansion rate and the density parameters for baryonic matter, (cold) DM, radiation, and pressureless matter (baryons+DM). These parameters are mathematically constants of integration of Friedmann and continuity equations when these sectors are only gravitationally interacting [66]. While integration constants from the mathematical perspective, these are fitting parameters of the model when confronted with data. Their values read from different data sets at different redshifts may not turn out to be redshift-independent constants. Ω r is fixed by the CMB monopole temperature, while constraints on H 0 , Ω b , and Ω c are derived from Planck CMB data, thereby defining the Planck- Λ CDM model. The mathematical consistency of flat Λ CDM model then implies/requires constancy of the observationally inferred values of these fitting parameters [164,165].
For theoretical physicists, the presence of Ω Λ in (3) is deeply perplexing. There are two interpretations in the literature. The most nailed down (testable) interpretation is that Λ is the cosmological constant from GR (1). If this is indeed the case, one expects to see traces of Λ at all scales where gravity is relevant, and, in principle, this opens up tests of Λ in the solar system [166,167,168] or local group [169,170,171]. It also begets the longstanding and unsolved cosmological constant problem [7,8], which demands a physical explanation to the observed smallness of the cosmological constant in the face of theoretical expectations from quantum physics that it should be much larger. The second interpretation is that Ω Λ is purely a phenomenological term in the flat Λ CDM model. This then allows one to introduce other phenomenological parametrisations, such as the Chevallier–Polarski–Linder (CPL) model [172,173], to test whether DE density is a constant or not. In cosmology, there is a tendency to flit between these two interpretations. Progress here has been slow, and, 25 years after the discovery of DE [162,163], little progress has been made on whether Λ is the cosmological constant from GR or not.
Nevertheless, this may be changing with recent large SNe compilations from Pantheon+ [174], Union3 [175], and DES [176], showing a preference for an evolving DE equation of state. Here, two points are worth stressing. First, CPL is not a model and it is merely a diagnostic for dynamical dark energy because all one is doing is performing a Taylor expansion in ( 1 a ) in the DE equation of state. As commented upon earlier, one could equally apply a CPL-like ansatz to the matter sector to test if matter is pressureless. Secondly, as remarked in [177], CPL is an expansion in a smaller redshift-dependent parameter, 1 a = z / ( 1 + z ) , which makes CPL less sensitive to low-redshift data. In other words, care is required to check which redshift ranges are driving deviations from Λ in the CPL model [175,176]. It cannot be naively assumed that data in the traditional DE-dominated regime z 0.7 are responsible for the deviation.
Returning to CMB, the flat Λ CDM fitting parameters are fixed up to sub-percent errors once one fits Planck data [6]. The effective redshift of the CMB is z 1100 . Thus, bearing in mind that redshift is a proxy for time, from the perspective of physics, one is in effect fixing all the degrees of freedom of the Λ CDM model using data on a single time-slice. As a result, the Λ CDM model is superpredictive; the model either leads to consistent results on independent time-slices or it does not. In particular, while late universe determinations of Ω m from Type Ia SNe and baryon acoustic oscillations (BAO) are largely consistent with Planck [174,178], the problem is that neither H 0 nor S 8 may be consistent. Neglecting systematics, this is enough to rule out the model because the Planck- Λ CDM model is a dynamical model of the universe with no free degrees of freedom.

3. Testable Λ CDM Subsectors

In physics, and science more generally, given a dynamical model with constant fitting parameters and time series data, there is an inevitable question about the range of validity of the model.
A valid, self-consistent model is a model that returns the same fitting parameters in a given time domain or epoch. If the model does not, and the fitting parameters evolve outside of the errors, then the model is not predictive and is meaningless in a physics context.
Note, the time domain where the model is valid may not cover all the time series data, in which case the model is an effective model that is only valid in a restricted regime. Any putative model of the observable universe is expected to return consistent values of the fitting parameters from recombination at z 1100 down to z 0 today. Systematics aside, Λ CDM tensions caution that this may not be true. Obviously, this is at odds with Bayesian model comparison, where it is inherently assumed that model fitting parameters are constant.14

3.1. H 0 Tension Subsector

The standard Λ CDM model is a 6-parameter model. This leads to complicated degeneracies between the parameters. Nevertheless, one can identify two subsectors of the flat Λ CDM model where one can safely decouple additional parameters.
First, the modelling can be greatly simplified by decoupling the radiation sector from (3) and removing perturbative physics. The latter is justified in the H 0 tension context as H 0 is defined through (1) and (2). Alternatively put, perturbations impact peculiar velocities, but the rate of expansion of the universe must not depend on peculiar velocities. The radiation sector is easily decoupled since Ω r Ω m , and as a result the Ω r ( 1 + z ) 4 term in (3) can be ignored if one works at lower redshifts, where z 30 is conservatively low redshift. This reduces the Λ CDM model at the background level to a 2-parameter model:
H ( z ) = H 0 1 Ω m + Ω m ( 1 + z ) 3 ,
parametrised exclusively through H 0 and matter density Ω m . What is important here is that the universe’s look-back time in gigayears (Gyr) at redshift z is
t ( z ) = 977.8 0 z d z ( 1 + z ) H ( z ) Gyr .
Since H ( z ) is a strictly increasing function of redshift, this means that the Λ CDM model describes about 13 billion years of evolution at the level of equations of motion with only two parameters ( H 0 , Ω m ) . Moreover, H 0 is simply an overall scale, as dictated by the FLRW assumption, so it has no bearing on any evolution in time or its proxy redshift.
The current quality of astronomical data aside, there is good reason to believe the evolution of the universe cannot be described by (4) over 13 billion years. Here is the argument. Consider a generic Taylor expansion of the Hubble parameter in ( 1 a ) : = y = z / ( 1 + z ) about a = 1 ,
H ( z ) = H 0 n = 0 N b n y n , y = z 1 + z ,
where we set b 0 = 1 so that H ( z = 0 ) = H 0 . As explained by Cattoën and Visser [179], in contrast to Taylor expansion in z, the validity of the expansion in y is not confined to z 1 . The only assumption being made in (6) is that H ( z ) is a continuous function, but otherwise one is looking at a model with N degrees of freedom. Nevertheless, in the Λ CDM model (4), there is only one degree of freedom Ω m . As a result, if the Λ CDM model accurately describes the universe, the N parameters b n , 1 n N must all depend on a single parameter Ω m .
In the limit N , and in the absence of a theoretical reason and framework,15 the ΛCDM model becomes a set of measure zero in the space of all possible Hubble parameters H ( z ) .
This is a bet that no discerning scientist should take, but what keeps the Λ CDM model relevant is the relatively poor quality of cosmological data. Thus, the task remains to show that the b n parameters do not converge to their Λ CDM expectations. While this is not a feasible exercise with a large number of b n parameters, one can expedite the process by simply fitting the model (4) to observational data in redshift bins and looking for (likely) evolution of the fitting parameter Ω m with redshift. Note, if Λ CDM is falsified, one cannot guarantee that minimal extensions of the Λ CDM model will turn out to be valid models.16

3.2. S 8 Tension Subsector

While the Λ CDM subsector (4) is apt for studying H 0 tension, one can define a comparable subsector for S 8 tension. To that end, we re-emphasise that redshift is more fundamental than scale. This is a simple observation, but it is routinely overlooked. In short, one can only define density perturbations δ ( r , t ) once one specifies a Hubble parameter H ( z ) , i.e., a solution to the Einstein field Equation (1), and adopts redshift as a proxy for time, a : = ( 1 + z ) 1 . Scale k only enters once one defines a perturbation δ in matter density, ρ m = ρ ¯ m ( 1 + δ ) , where ρ ¯ m denotes the average matter density and we assume the perturbation is small, δ 1 . A standard analysis in perturbation theory that combines the continuity and Euler and Poisson equations (Newtonian gravity) then leads to
2 t 2 + 2 H t c s 2 a 2 2 4 π G ρ ¯ m δ = 0 ,
where c s is the speed of sound. The scale k then enters when one decomposes δ in Fourier modes, 2 δ = k 2 δ . However, provided one works on large scales, k 4 π G ρ ¯ m ( a / c s ) , one can safely neglect k. In this limit, (7) reduces to
δ ¨ + 2 H δ ˙ 4 π G ρ ¯ m δ = 0 ,
where dots denote derivatives with respect to time, and H and ρ ¯ m : = 3 H 0 2 Ω m / ( 8 π G ) a 3 are defined through the scale factor a at the background level. The key point is that, provided one works at linear scales, typically 0.001 h   Mpc 1 k 0.1 h   Mpc 1 , then scale k is irrelevant. Moreover, at these scales, DE is smooth, so that in the Λ CDM model the same parameter Ω m determines both background and perturbed evolution. This reinforces the importance of Ω m across billions years of cosmological evolution. Unfortunately, the reality is that most cosmological observables probe physics beyond these linear scales (see [182] for a deeper discussion), so great care is required to make sure that one is working in a strict linear regime.

4. Localising Λ CDM Tensions

The basic approach to Λ CDM tensions, in particular H 0 and S 8 tension, advocated in this letter, is to reduce the Λ CDM model to subsectors with a minimal number of fitting parameters. Once this is accomplished, one can bin data by effective redshift and confront them to simplified settings (4) and (8).
We stress that, since redshift is defined at the background level but scale is defined at the level of perturbations, one needs to study the redshift evolution of fitting parameters before their scale evolution. Moreover, redshift evolution implies a scale problem, but the converse is not true.
Naturally, we understand why the cosmology community may like to focus on scale as it is less disruptive to the Λ CDM model, but nature can be fickle. The goal then is to identify redshift ranges where the Λ CDM model returns consistent fitting parameters and redshift ranges where the parameters are no longer consistent and evolve or drift with redshift. The latter redshifts are then candidates for missing physics, i.e., physics not incorporated by the Λ CDM model. We review the current status of these consistency checks of the model. We begin with S 8 tension because the progress is more tangible, and, from our perspective, S 8 tension is more appealing since it has some semblance of a problem that can be addressed with novel scale-dependent physics, e.g., baryonic feedback [183,184]. This encourages cosmologists to engage with the S 8 problem, while it is not yet as statistically significant as H 0 tension. Nevertheless, as stressed, H 0 tension is not an obvious scale problem.

4.1. Localising S 8 Tension

Over the last year, noticeable progress has been made on S 8 tension. While S 8 tension is primarily viewed as a 2 3 σ discrepancy between lower S 8 values from weak lensing [27,28,29,30,31,32] and a larger Planck value [6], ACT CMB lensing recently localised the problem (if there is a problem) to the late universe [63,185,186]. The key point is that, despite CMB lensing being sensitive to larger redshifts (see Figure 1 of [63]) and larger scales than weak lensing, ACT independently recovers the Planck S 8 value.
One can now define S 8 tension as a putative problem in the redshift range z 2 .
This interpretation is supported by independent observations, including lower S 8 inferences from peculiar velocities in the local universe [108,109,110], SZ cluster counts [111], and RSD [112,113,114,115,116,117,118,119,120]. Moreover, separately, it has been noted that S 8 , or more accurately σ 8 , may evolve from lower to higher values between low and high redshifts in these observables [121,122,125]. If true, this provides independent corroborating evidence for the discrepancy between weak and CMB lensing. In contrast, a recent study [126] of cross-correlations of unWISE galaxies and CMB lensing is in conflict with these observations and recovers the Planck value at redshifts z 0.2 . More recently, a study of cross-correlations of DESI luminous red galaxies with CMB lensing [127,128] has shown that (i) S 8 or σ 8 exhibits a mild increasing trend with bin redshift, (ii) S 8 or σ 8 is robust to scale cuts, and (iii) baryonic feedback cannot explain the results since the study is primarily sensitive to the linear regime. Note, the S 8 literature is extremely confusing in the sense that there is no clear pattern emerging from the results. S 8 whisker plots are everywhere, but there is a focus on scale-dependent physics, e.g., baryonic feedback [183,184], whereas our arguments here point to missing redshift-dependent physics, in essence because one cannot resolve H 0 tension with scale-dependent physics. Interestingly, at redshifts beyond z 2 , one finds reports of lower values of σ 8 relative to Planck, which are ostensibly at odds with the ACT results [187] (see also [188]). Evidently, the systematics are not fully under control.
The relevant question now is whether the evolution of S 8 or σ 8 with scale k could be mimicking the evolution of S 8 or σ 8 with redshift. While it is true that weak lensing probes smaller scales than CMB lensing, peculiar velocities lead to lower S 8 values while still probing larger scales [108,109,110]. Moreover, both the galaxy cluster number counts and Lyman- α spectra, analysed in [121], are expected to be probing similar scales. In addition, the evolution of σ 8 with redshift is evident across the f σ 8 ( z ) constraints from RSD, and all the data are expected to be probing linear scales [122].17 Finally, in [125], the Weyl potential is probed through an observable that is expected to be independent of scale; nevertheless, once specialised to the Λ CDM setting, the differences in σ 8 are reported across weak lensing bins in DES data [190].18 Given these results, while it may be appealing to the cosmology community if S 8 increased with increasing scale, this is not borne out in the data. That being said, it is reasonable to expect that some S 8 constraints in the literature are scale-dependent.
At no more than 2 3 σ statistical significance, and potentially even less significant [191], the jury is still out on whether S 8 tension is a definite problem; however, the fact that we see a lower S 8 or σ 8 in at least four independent observables, i.e., weak lensing, the peculiar velocities, RSD, and cluster counts, which probe lower redshifts, point to an obvious problem. The fact that higher redshift probes recover the Planck S 8 value offers support to the idea that S 8 tension is due to missing physics in the late universe, which conservatively can be demarcated as redshifts z 2 . This begets an interesting question. Assuming there is missing physics at the perturbative level, and given that H 0 tension is defined at the background level, how is one sure that there is no missing physics at the background level? Note, while the evolution of S 8 with scale can be addressed with new perturbative physics, the evolution of S 8 with redshift is a different kettle of fish, forcing one to also change the background. This nicely brings us to H 0 tension and the consistency tests investigating the constancy of Λ CDM fitting parameter H 0 from (4).

4.2. Localising H 0 Tension

Here, we focus on how progress can be made on H 0 tension. First, we need to confirm that local (cosmological model-independent) H 0 determinations consistently return larger H 0 > 70 km s 1 Mpc 1 values. Combined with the robustness of the early universe determinations across CMB experiments [6,102,103] and BBN+BAO [192,193,194,195], this demonstrates that there must be missing physics in the Λ CDM model, not just at the perturbative level, as outlined above, but also at the background level.
Recent observations suggest that we are on our way to achieving H 0 > 70 km s 1 Mpc 1 local measurements. In particular, Cepheid-SNe H 0 inferences have exhibited stable central values of H 0 72 74 km s 1 Mpc 1 over the last decade [14,196,197,198,199]. Importantly, independent groups agree on the result, and JWST has thrown up no major surprises on Cepheid calibration [200]. Moreover, a host of independent distance indicators, including megamasers [16], surface brightness fluctuations [18], Tully–Fisher relations [17,201], and Type II SN [202,203], have independently led to H 0 > 70 km s 1 Mpc 1 .19 Tip of the Red Giant Branch (TRGB) calibration has regularly returned lower values in this window, leading to considerable debate [15,206,207,208,209], but there appears to be a consensus now that H 0 > 70 km s 1 Mpc 1 [210]. The only bone of contention is whether 1 % errors are achievable.20
In some sense, we appear to have already confirmed locally that H 0 is in the 70–77 km s 1 Mpc 1 range, and all that remains is to debate the errors. The next exercise of interest is to study the H 0 determinations at cosmological scales where one must assume a model. Invariably, the assumption made is that the Λ CDM model is correct, which alarmingly is only the case if there is no H 0 tension!
If the H 0 tension is physical and the ΛCDM model has broken down, it is reasonable to expect the H 0 inferences at the cosmological scale to be biased if they assume ΛCDM. Not allowing for this possibility runs the risk of circular logic.
For this reason, one expects H 0 determinations from strong lensing time delay [211,212,213,214,215,216,217] and gravitational wave dark sirens [218,219,220] to be potentially impacted by model breakdown if the model is breaking down in the late universe. Whether the former converges to an H 0 value consistent with local or early universe H 0 determinations depends on the assumptions being made [215]. Recently, a strongly lensed Type Ia supernova has led to an H 0 value [221] that favours an H 0 value consistent with early universe determinations, so, evidently, lensed systems have not been able to arbitrate H 0 tension.21 On the other hand, dark siren constraints are not yet strong enough to distinguish between local and early universe H 0 determinations.
One situation where we may have observed a hint of model breakdown is H0LiCOW and TDCOSMO reported a descending trend in H 0 with lens redshift, admittedly at low statistical significance, 1.7 σ [211,212,213]. What matters here is that this low significance trend is not an obvious systematic [213]. Moreover, the lensed system driving this trend at the lowest redshift probed was recently revisited by analysing the spatially resolved stellar kinematics [217], and, while the errors have inflated, the central value H 0 77 km s 1 Mpc 1 is unchanged from [211]. Given that TDCOSMO plans to release 40 odd H 0 determinations from different lensed QSOs with different lens redshifts in the future, if there is a trend, this will hopefully be evident.
While each strongly lensed QSO or supernova system constitutes its own redshift bin, one can look for corroborating evidence by binning independent observables by the effective redshift. Concretely, one assumes that model (4) is correct, bins observables by redshift, and studies the fitting parameter H 0 with varying effective redshift. This consistency test of a constant H 0 fitting parameter was initiated in [226] with a combination of local and cosmological data, including megamasers, SNe, cosmic chronometer, and BAO data. An independent descending trend with statistical significance 2 σ was reported. It was subsequently noted in [227,228] that this trend in a combination of data could be recovered through an ansatz when one isolated SNe and BAO data on their own. The result has since been confirmed by a number of different groups with different methodologies and perspectives [164,229,230,231,232] (see [233] for further discussion).
The result is curious for at least three reasons. First, we are now looking at distinct decreasing H 0 trends with the effective redshift assuming the correctness of the Λ CDM model across different observables that share no obvious systematic, e.g., strongly lensed QSOs, SNe, and observational Hubble data (BAO). Secondly, bearing in mind that local universe H 0 measurements [14,15,16,17,18] are biased to larger values than early universe cosmological H 0 determinations [6], a decreasing H 0 trend with the effective redshift is consistent with what one observes in the H 0 tension problem. Thirdly, just as the case in strongly lensed QSOs [213], one may struggle to find a systematic in Type Ia SNe that could cause H 0 to evolve with the effective redshift, especially given that SNe data as distance indicators have been studied since the early 1990s [234].
Nevertheless, a decreasing H 0 with an increasing effective redshift trend may only be half the story. The reason is that Ω m is anti-correlated with H 0 in the model (4). Thus, one expects any decreasing H 0 trend with effective redshift to be complemented by an increasing trend of Ω m with effective redshift. Note, it is possible that the combination Ω m h 2 remains a constant, but this is not observationally required; as explained, changes in Ω m h 2 imply a problem with the pressureless matter assumption (see [161]). Indeed, building on the observations of larger-than-expected matter density values in standardisable QSOs [235,236,237] (see also [238]), it was shown in [230] that standardisable QSOs and Type Ia SNe share two key traits. First, both probes recover a Planck value for Ω m at lower redshifts z 0.7 , but an increasing trend of Ω m with redshift is observed at higher redshifts. Upgrading the Pantheon SNe sample [239] to the Pantheon+ SNe sample [174] does not eliminate the evolution [240]22. Seemingly consistent results with SNe and/or QSOs were reported elsewhere [242,243]. It is worth stressing that, despite this agreement, the standardisability of QSOs is open to debate [244,245,246,247,248,249]. Throughout these studies, it is observed that changes in Ω m are noticeable at higher redshifts, either at or beyond the deceleration–acceleration transition redshift z 0.7 . This highlights a potential problem with the Λ CDM model, but this problem is only evident at redshifts where the pressureless matter assumption is important. This implies missing physics in the matter sector of the Λ CDM model.
Nevertheless, there are other angles worth mentioning. JWST has reported high-redshift galaxies [250,251,252,253,254] that appear anomalous if the Planck- Λ CDM model is correct. Nevertheless, it is easy to show that increasing Ω m h 2 in the Λ CDM model would make the preliminary JWST results less anomalous [255]. From Table 1 of [230], one can convince oneself that Ω m h 2 is increasing with the effective redshift in the QSO sample. This coincidence is intriguing and warrants further study. Moreover, both Union3 and DES now have a preference for larger-than-expected Ω m determinations assuming the Λ CDM model [175,176], thereby throwing the constancy of Ω m into question. At Ω m = 0 . 356 0.026 + 0.028 [175] and Ω m = 0.352 ± 0.017 [176], respectively, one is looking at 1.5 σ and 2 σ differences with Planck, Ω m = 0.315 ± 0.007 [6]. It will be interesting to see if Ω m is constant across these large SNe samples when the data are binned by redshift. Finally, since we see a decreasing H 0 /increasing Ω m trend with effective redshift in multiple independent observables, it is relatively easy to combine the observables to find a statistically significant result 3 σ [164]. Physically, one may try to interpret the results as motivation for an Mpc-scale [256,257,258,259,260,261] or Gpc-scale void [262,263], but, if a model is breaking down, this can be due to a lack of complexity, and there may be no deep physical reason.
It should be stressed that a decrease in H 0 compensated by an increase in Ω m is often misinterpreted as an artefact of the degeneracy or anti-correlation between the two parameters. Indeed, for exclusively high redshift observational Hubble data constraints H ( z ) or angular diameter/luminosity distance constraints D A ( z ) / D L ( z ) , one expects 2D Markov Chain Monte Carlo (MCMC) posteriors corresponding to banana-shaped contours in the ( H 0 , Ω m ) -plane. Note, the Λ CDM parameters are confined to H 0 2 Ω m = const . curves in the ( H 0 , Ω m ) -plane when one fits exclusively high redshift data [165], so the observation of banana-shaped 2D MCMC posteriors is expected.23 These contours are interpreted as the fit to the data from any point in ( H 0 , Ω m ) parameter space is more or less the same. Translated into a frequentist framework, if this is true, one expects to see a curve of almost constant χ 2 that includes the global minimum of the χ 2 . Nevertheless, by studying profile distributions [264,265], a variant of profile likelihoods [266,267,268], it is possible to show that the points in parameter space that are equivalent from the perspective of MCMC posteriors are no longer equivalent when one analyses the χ 2 . In other words, the data distinguish between the points in the model parameter space, but this distinction is not evident in MCMC posteriors. We observe that, if these trends in the fitting parameters are established, there are evidently physics missing from the Λ CDM model in the late universe.

5. Outlook

Given the H 0 and S 8 discrepancies in the Λ CDM model [11,12], it is scientifically prudent to conduct consistency checks of the Λ CDM model to either confirm model breakdown or assure that unexplored systematics are at play. Given that redshift is more fundamental than scale, since redshift is a proxy for time and the universe is dynamical, the most relevant consistency checks probe redshift. Simply put, one bins observations by effective redshift, fits the Λ CDM model, and identifies the flat Λ CDM fitting parameters preferred by the model in each bin. If these fitting parameters are the same at all redshifts within 2 σ , as judged by the errors, then the model passes the consistency check. On the contrary, if the fitting parameters evolve or drift outside of 2 σ , and one sees the same feature in multiple observables,24 then the model has broken down. Note, the arguments being run are largely targeting modeling 101; namely, the fitting parameters should not evolve. The Λ CDM model may also be targeted using more physical arguments, e.g., [269], but here our arguments are more mathematical than physical.
Note, in contrast to Bayesian model comparison, there is no second model to compare. Moreover, there are no additional parameters. If the fitting parameters evolve, this worsens the fit to the data, and, as the data improve, one will eventually be able to construct a new model that incorporates the evolution. This is how one eventually connects to Bayesian methods. The key point here is that cosmology purports to be a sub-branch of physics; however, physics demands that models are predictive. In our context, then, the fitting parameters should not evolve. We remark that this discussion applies to all cosmological models. They are not specific to the Λ CDM model. For any model of the universe to be physically meaningful, it is imperative that the fitting parameters do not change with redshift or scale.
In this work, we reviewed the assumptions underlying the Λ CDM model. We concluded that the foundations based on GR and FLRW do not appear to offer a compelling direction to resolve the tensions. We dispelled the first possibility on the grounds that there are simply no observed deviations from GR in controlled environments, where it should be stressed that little in cosmology is under control, hence the need for multiple probes. In contrast, FLRW looks shaky [88], especially at lower redshifts z 0.1 , but the angular variations observed in the distance ladder and CMB (they exist!) are simply too small to explain a 10 %   H 0 discrepancy [94,95,96,97,98]. In short, we believe that the FLRW assumption is worth studying, but we do not see it coming to the rescue on Λ CDM tensions, other than relaxing FLRW leads to the Λ CDM fitting parameters, e.g., H 0 , Ω m , and S 8 , being ill-defined.
In the latter part of this work, we outlined simplified subsectors of the flat Λ CDM model where one can decouple additional parameters. The H 0 tension subsector (4) allows one to probe the constancy of fitting parameters H 0 and Ω m across 13 billion years of evolution of the Hubble parameter H ( z ) . Since H 0 is a scale, this is 13 billion years modelled by a single parameter Ω m . As argued earlier, (4) is a set of measure zero in all the continuous functions H ( z ) that one can construct, thereby making it likely that a deviation will be found as the data quality improves. Union3 and DES may already be hinting at this result [175,176]. On the other hand, performing consistency checks of S 8 involves restricting observables to linear scales 0.001 h Mpc 1 k 0.1 h Mpc 1 , where the perturbations in matter density are independent of k (8). Nevertheless, observationally, one needs to make sure that non-linear physics is not impacting the observables, and this is admittedly tricky.
Over the last year, progress appears to have been made in terms of localising the S 8 tension to the late universe, but this may have gone under the radar. A comparison between weak [27,28,29,30,31,32] and CMB lensing [63,185,186] now restricts the deviations of S 8 from the Planck- Λ CDM value to z 2 . This raises the possibility that S 8 either increases with redshift or scale. Separately, peculiar velocities [108,109,110], galaxy cluster counts [111], and RSD favour lower values of S 8 [112,113,114,115,116,117,118,119,120] that are consistent with weak lensing. A scale-dependent S 8 [183,184] can plausibly explain the discrepancies between weak and CMB lensing, but it would struggle to explain the lower S 8 values from peculiar velocities as the determinations are made at large scales. Moreover, three independent studies [121,122,125] incorporating galaxy cluster counts, RSD, and weak lensing data now point to variations in σ 8 with redshift. While this is countered elsewhere [126] (however, see [127,128]), it behooves us to once again remind the reader that redshift is more fundamental than scale. In addition, the more serious H 0 tension problem can only be due to redshift evolution if the Λ CDM model is breaking down [66]. The simplicity of 2 + 2 = 4 cannot be denied. For this reason, a redshift-dependent S 8 or σ 8 is a natural and plausible signature of Λ CDM model breakdown.
This brings us nicely to H 0 tension [14,15,16,17,18]. As explained in [66], a prerequisite to the tension not being due to systematics is an H 0 fitting parameter that evolves with redshift in the Λ CDM model. Mathematics demands that this must be observed, although the redshift range is uncertain and requires localisation. To that end, H 0 inferences from strong lensing time delay, Type Ia SNe, standardisable QSOs, and OHD all point to a decreasing trend in H 0 with increasing effective redshift [164,211,213,226,227,228,229,230,231,232]. The statistical significance in each observable may be low 2 σ , but combining the results increases the significance quickly; it has been estimated as no less than 3 σ [164]. On the flip side, as argued earlier, one should not expect to be able to model 13 billion years of background evolution with a single parameter Ω m dialled to Ω m 0.3 . Nevertheless, since H 0 and Ω m are anti-correlated and one typically encounters banana-shaped contours, i.e., H 0 2 Ω m = const . curves in the ( H 0 , Ω m ) -plane in high-redshift bins, because the Λ CDM model only constrains combinations of H 0 and Ω m well at high redshifts, great care is required in estimating errors and assigning statistical significance. This is a technical problem that needs to be overcome. What is safe to say currently is that the banana-shaped 2D MCMC posteriors from the Bayesian approach do not imply curves of constant χ 2 in the frequentist approach [265]. Thus, what appear to be equivalent points in the ( H 0 , Ω m ) -parameter space for Bayesians may no longer be equivalent points in the parameter space for frequentists. These conceptual differences will need to be overcome, otherwise frequentists will make discoveries well before the results are adopted by Bayesians.
At a conceptual level, what is being said is simple. If in response to Λ CDM tensions one builds a replacement model by injecting new physics at a given cosmic epoch, this is only well-motivated if it correlates with the missing physics in the Λ CDM model. Thus, whether one confronts the new model with data or the original Λ CDM model with the same data, one should agree on the redshift ranges or scales where new physics are required. The problem here is that EDE [20,21,22,23,24,25] injects new physics in epochs where there are no data enabling direct tests of the Λ CDM model. Nevertheless, if the evolution of Λ CDM parameters H 0 and S 8 is established in the late universe, one can attempt to correlate this with coupled DE-matter models [270,271,272,273,274], sign-switching cosmological constant ( Λ s CDM) models [275,276,277,278], etc.
In summary, H 0 tension is a serious anomaly. Given that we only see anomalous results in the local universe, interpreting it cosmologically is challenging. Nevertheless, the less significant S 8 tension promises to be insightful since we appear to have localised the problem to the late universe z 2 . From there, it all depends on whether S 8 or σ 8 varies with redshift or scale, but one must preclude redshift first as it is more fundamental. A number of results now favour the redshift evolution of S 8 or σ 8 when assuming the Λ CDM model [121,122,125]. If so, this is a problem that can only be addressed by changing the cosmology at the background level. The same is true for the H 0 tension problem. Going forward, establishing that the H 0 and Ω m   Λ CDM fitting parameters evolve with redshift in the late universe should be enough to settle the issue provided we see similar evolution across independent cosmological probes. Preliminary results in that direction exist.

Author Contributions

Conceptualization, Ö.A., E.Ó.C., A.A.S. and M.M.S.-J.; Writing-original draft, Ö.A., E.Ó.C., A.A.S. and M.M.S.-J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Turkish Academy of Sciences Outstanding Young Scientist Award (TÜBA-GEBİP), Iran National Science Foundation (INSF) grant no. 4026712, SERB, Govt. of India, grant no. CRG/2020/004347 and the EU COST Action CA21136—“Addressing observational tensions in cosmology with systematics and fundamental physics (CosmoVerse)”, supported by COST (European Cooperation in Science and Technology).

Data Availability Statement

We made no use of data in this work.

Acknowledgments

We would like to thank Yashar Akrami, Luca Amendola, Adam Riess, István Szapudi, and Sunny Vagnozzi for fruitful discussions. ÖA acknowledges the support of the Turkish Academy of Sciences in the scheme of the Outstanding Young Scientist Award (TÜBA-GEBİP). The work of MMShJ is supported in part by INSF grant no. 4026712. AAS acknowledges the funding from SERB, Govt. of India under the research grant no. CRG/2020/004347.

Conflicts of Interest

The authors declare no conflicts of interest.

Notes

1
See [4] for a sharp discussion on the arbitrariness inherent to ratios of Bayesian evidences. In particular, when there is poor theoretical guidance on parameter space, it seems possible to relax the priors so that the minimal model, e.g., Λ CDM, in any nested model is preferred [5].
2
Combining supernovae (SNe) and BAO constraints, one may be able to whittle this down to 1 z 1100 .
3
In cosmology, redshift is more fundamental than scale, so one must eliminate redshift first before turning one’s attention to scale. For this reason, we focus largely on redshift.
4
The consistency condition of (1) is that both sides of this equation are divergence-free, i.e., μ ( R μ ν 1 2 R g μ ν ) = 0 by virtue of Bianchi identity, and μ ( Λ g μ ν ) = 0 as long as Λ is a constant, implying that μ T μ ν = 0 . Note overall consistency of (1) only implies the total stress-energy tensor to be divergence-free. Requiring individual components of the cosmic fluid, i.e., DE, pressureless matter, and radiation, to have divergence-free stress tensors is an additional assumption.
5
The slogan here is simple: theoretical yes, observational no!
6
Modern cosmology blends data analysis with story telling in settings where observations do not exist. One could rewrite the inflationary backstory to allow for curved spacetimes, e.g., [61]. We note that curvature can also be exploited to test FLRW [62].
7
Recently, the ACT collaboration analysing the CMB lensing data resolved the so-called A L lensing anomaly of the Planck collaboration, providing further support for flat cosmology [63]. See also [64] for a recent investigation of the A L anomaly that unearths a connection to ecliptic latitude. This is curious because coincidences involving the ecliptic also appear in CMB anomalies [65], but the origin could easily be a systematic effect in Planck data.
8
It is well-known that, in the simplest extensions of FLRW setting that allow for anisotropic expansion, the cosmic shear contributes as a 6 in the Friedmann equations, and hence expansion anisotropies are expected to be theoretically more relevant in the early universe, thereby delocalising physics and demanding a complete rewrite for the universe’s evolution [67,68]. However, one should note that, besides the shear, an anisotropy in the metric, an anisotropy can enter in the T μ ν part of cosmological models through the notion of the tilt [69]. A minimalistic, maximally Copernican beyond-FLRW model within the titled cosmology setting has been formulated as “dipole cosmology”, which, as the name suggests, accommodates cosmic dipoles in various matter sectors even in the late universe [70,71,72].
9
One can argue that resolutions to Hubble tension are strongly constrained by FLRW, the age of the universe, and the assumption of constant matter density [73].
10
Note that the breakdown of cosmological principle (isotropy) can happen due to topology of constant time (spatial) slices while the metric is still FLRW, i.e., while the universe is locally described by FLRW metric; e.g., see [90,91,92]. In particular, isotropy may be broken due to violation of parity, as reported in [89]. Topology effects induce various relations among 2-point function correlations on the sky, which are in principle detectable in the CMB or distributions of other large structures over the sky.
11
There is an independent anomaly in the late-time integrated Sachs–Wolfe (ISW) effect from supervoids [123,124]. In particular, an excess ISW effect is reported at lower redshifts [123], whereby A ISW 5 , and a deficit at higher redshifts [124] with A ISW 5 . A ISW varies with redshift, and one encounters the expected Λ CDM value A ISW = 1 at intermediate redshifts. Translated into the growth parameter f [124], this implies weaker and stronger growth of structures than Planck expectations at lower and higher redshifts, respectively. Given the overlap with f σ 8 ( z ) constraints from RSD, it is plausible that the ISW anomaly and S 8 evolution in RSD [122] are symptoms of the Planck- Λ CDM cosmology being simply an approximation that underestimates or overestimates quantities when one performs a tomographic analysis but on average obtains the right result.
12
Note this is contradicted by an increasing trend of S 8 / σ 8 in Figure 10 of [127] and Figure 16 of [128], admittedly at low statistical significance.
13
Interestingly, extrapolations of H ( z ) reconstructions based on cosmic chronometer data [156], a cosmological model that is agnostically observable, have favoured lower H 0 < 70 km s 1 Mpc 1 values [157,158,159]. However, if systematics are properly propagated, the errors are too large to exclude local H 0 determinations [160], so they do not currently arbitrate on H 0 tension.
14
This becomes a little more puzzling in the context of H 0 tension, an apparent serious > 3 σ discrepancy. In order to compare model A, i.e., Λ CDM, to model B, a potential replacement that resolves/alleviates H 0 tension, one has to combine data sets that are inconsistent when confronted with model A.
15
Unlike the firm theoretical guiding principle for particle physics, such as Wilsonian effective field theory, the notions of relevant, marginal, and irrelevant operators, decoupling of scales, and cluster decomposition, cosmology lacks such guiding principles. These guiding principles make it possible to effectively verify the theoretically important notion of “stability of description”. That is, only a small number of deformations yield relevant perturbations around a given model, which is a point in the space of all possible models. In the absence of such principles, there is no systematical way to classify deformations relevant to certain features and argue for or verify stability of description in cosmology.
16
For example, consider the wCDM model as a 1-parameter extension of the Λ CDM model, with a constant DE equation of state w that can be viewed as a deformation of Λ CDM. In keeping with the general thrust of the paper, there is of course no guarantee that w is a constant at different redshifts, e.g., [180]. However, even in the same redshift ranges, one may find contradictory results. For example, Union3 and DES SNe have a preference for w > 1 [175,176], whereas SPT Clusters with DES/HST weak lensing prefer w < 1 [181]. Systematics aside, this says that w is not a good deformation parameter around Λ CDM. Obviously, if the fitting parameters evolve, none of these models are good physical models.
17
In recent years, RSD has been extended to non-linear scales in order to extract better constraints from the data. This leads to lower values of f σ 8 ( z ) , as is evident from Figure 7 [189], so scale certainly impacts RSD constraints.
18
As demonstrated [125], scale cuts inflate errors but have little bearing on the central values. One can of course trivially resolve the S 8 tension by throwing information away.
19
We note that a gravitational wave standard siren H 0 determination exists [204,205], but, with only one event, the errors are not currently competitive.
20
Ref. [210] originally pointed to larger systematic errors, but the errors have since been revised downwards. One eventually expects to reach a precision where hemisphere decompositions, or equivalent, of distance indicators lead to differences in H 0 values on the sky that exceed the errors [95,96], which is expected if the universe at lower redshifts is not FLRW.
21
A real concern here is that, if there is a large spread in H 0 values from strong lensing with constrained errors assuming the Λ CDM model, one arrives at the conclusion that strong lensing cannot consistently determine H 0 . Ultimately, it is plausible that lens degeneracies, most notably the mass sheet transformation [222], make it difficult to determine H 0 uniquely [223,224]. Nevertheless, if one succeeds in precluding systematics, such an outcome could rule out the Λ CDM model [225].
22
The same increasing Ω m trend with effective redshift can be demonstrated in DES SNe [241], where there is no overlap with Pantheon+ SNe at higher redshifts [176].
23
Observational Hubble data constrain best fits to the curve Ω m h 2 = c , whereas angular diameter/luminosity distance constraints confine best fit parameters to the curve ( 1 Ω m ) h 2 = c , where c , c are data-dependent constants. c , c may also evolve with effective redshift and its constancy assuming the Λ CDM model needs to be observationally checked.
24
This is a necessity in cosmology as it would be foolhardy to trust any single observable. To put this in context, one only trusts the existence of DE because we see it in multiple observables.

References

  1. Trotta, R. Applications of Bayesian model selection to cosmological parameters. Mon. Not. Roy. Astron. Soc. 2007, 378, 72–82. [Google Scholar] [CrossRef]
  2. Hobson, M.P.; Liddle, A.H.J.A.H.A.R.; Mukherjee, P.; Parkinson, D. Bayesian Methods in Cosmology; Cambridge University Press: Cambridge, UK, 2009. [Google Scholar]
  3. Trotta, R. Bayesian Methods in Cosmology. arXiv 2017, arXiv:1701.01467. [Google Scholar]
  4. Amendola, L.; Patel, V.; Sakr, Z.; Sellentin, E.; Wolz, K. The distribution of Bayes’ ratio. arXiv 2024, arXiv:2404.00744. [Google Scholar]
  5. Patel, V.; Amendola, L. Comments on the prior dependence of the DESI results. arXiv 2024, arXiv:2407.06586. [Google Scholar]
  6. Aghanim, N. et al. [Planck] Planck 2018 results. VI. Cosmological parameters. Astron. Astrophys. 2020, 641, A6, Erratum in Astron. Astrophys. 2021, 652, C4. [Google Scholar] [CrossRef]
  7. Weinberg, S. The Cosmological Constant Problem. Rev. Mod. Phys. 1989, 61, 1–23. [Google Scholar] [CrossRef]
  8. Weinberg, S. The Cosmological constant problems. arXiv 2000, arXiv:astro-ph/0005265. [Google Scholar]
  9. Peebles, P.J.E.; Ratra, B. The Cosmological Constant and Dark Energy. Rev. Mod. Phys. 2003, 75, 559–606. [Google Scholar] [CrossRef]
  10. Valentino, E.D.; Mena, O.; Pan, S.; Visinelli, L.; Yang, W.; Melchiorri, A.; Mota, D.F.; Riess, A.G.; Silk, J. In the realm of the Hubble tension—A review of solutions. Class. Quant. Grav. 2021, 38, 153001. [Google Scholar] [CrossRef]
  11. Perivolaropoulos, L.; Skara, F. Challenges for ΛCDM: An update. New Astron. Rev. 2022, 95, 101659. [Google Scholar] [CrossRef]
  12. Abdalla, E.; Abellán, G.F.; Aboubrahim, A.; Agnello, A.; Akarsu, O.; Akrami, Y.; Alestas, G.; Aloni, D.; Amendola, L.; Anchordoqui, L.A.; et al. Cosmology intertwined: A review of the particle physics, astrophysics, and cosmology associated with the cosmological tensions and anomalies. J. High Energy Astrophys. 2022, 34, 49–211. [Google Scholar] [CrossRef]
  13. Schöneberg, N.; Abellán, G.F.; Sánchez, A.P.; Witte, S.J.; Poulin, V.; Lesgourgues, J. The H0 Olympics: A fair ranking of proposed models. Phys. Rept. 2022, 984, 1–55. [Google Scholar] [CrossRef]
  14. Riess, A.G.; Yuan, W.; Macri, L.M.; Scolnic, D.; Brout, D.; Casertano, S.; Jones, D.O.; Murakami, Y.; Breuval, L.; Brink, T.G.; et al. A Comprehensive Measurement of the Local Value of the Hubble Constant with 1 km s−1 Mpc−1 Uncertainty from the Hubble Space Telescope and the SH0ES Team. arXiv 2021, arXiv:2112.04510. [Google Scholar] [CrossRef]
  15. Freedman, W.L. Measurements of the Hubble Constant: Tensions in Perspective. Astrophys. J. 2021, 919, 16. [Google Scholar] [CrossRef]
  16. Pesce, D.W.; Braatz, J.A.; Reid, M.J.; Riess, A.G.; Scolnic, D.; Condon, J.J.; Gao, F.; Henkel, C.; Impellizzeri, C.M.V.; Kuo, C.Y.; et al. The Megamaser Cosmology Project. XIII. Combined Hubble constant constraints. Astrophys. J. Lett. 2020, 891, L1. [Google Scholar] [CrossRef]
  17. Kourkchi, E.; Tully, R.B.; Anand, G.S.; Courtois, H.M.; Dupuy, A.; Neill, J.D.; Rizzi, L.; Seibert, M. Cosmicflows-4: The Calibration of Optical and Infrared Tully—Fisher Relations. Astrophys. J. 2020, 896, 3. [Google Scholar] [CrossRef]
  18. Blakeslee, J.P.; Jensen, J.B.; Ma, C.P.; Milne, P.A.; Greene, J.E. The Hubble Constant from Infrared Surface Brightness Fluctuation Distances. Astrophys. J. 2021, 911, 65. [Google Scholar] [CrossRef]
  19. Knox, L.; Millea, M. Hubble constant hunter’s guide. Phys. Rev. D 2020, 101, 043533. [Google Scholar] [CrossRef]
  20. Poulin, V.; Smith, T.L.; Karwal, T.; Kamionkowski, M. Early Dark Energy Can Resolve The Hubble Tension. Phys. Rev. Lett. 2019, 122, 221301. [Google Scholar] [CrossRef]
  21. Agrawal, P.; Cyr-Racine, F.Y.; Pinner, D.; Randall, L. Rock ‘n’ roll solutions to the Hubble tension. Phys. Dark Univ. 2023, 42, 101347. [Google Scholar] [CrossRef]
  22. Lin, M.X.; Benevento, G.; Hu, W.; Raveri, M. Acoustic Dark Energy: Potential Conversion of the Hubble Tension. Phys. Rev. D 2019, 100, 063542. [Google Scholar] [CrossRef]
  23. Niedermann, F.; Sloth, M.S. New early dark energy. Phys. Rev. D 2021, 103, L041303. [Google Scholar] [CrossRef]
  24. Ye, G.; Piao, Y.S. Is the Hubble tension a hint of AdS phase around recombination? Phys. Rev. D 2020, 101, 083507. [Google Scholar] [CrossRef]
  25. Poulin, V.; Smith, T.L.; Karwal, T. 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. 2023, 42, 101348. [Google Scholar] [CrossRef]
  26. Vagnozzi, S. Seven Hints That Early-Time New Physics Alone Is Not Sufficient to Solve the Hubble Tension. Universe 2023, 9, 393. [Google Scholar] [CrossRef]
  27. Heymans, C.; Grocutt, E.; Heavens, A.; Kilbinger, M.; Kitching, T.D.; Simpson, F.; Benjamin, J.; Erben, T.; Hildebrandt, H.; Hoekstra, H.; et al. CFHTLenS tomographic weak lensing cosmological parameter constraints: Mitigating the impact of intrinsic galaxy alignments. Mon. Not. Roy. Astron. Soc. 2013, 432, 2433. [Google Scholar] [CrossRef]
  28. Joudaki, S.; Blake, C.; Heymans, C.; Choi, A.; Harnois-Deraps, J.; Hildebrandt, H.; Joachimi, B.; Johnson, A.; Mead, A.; Parkinson, D.; et al. CFHTLenS revisited: Assessing concordance with Planck including astrophysical systematics. Mon. Not. Roy. Astron. Soc. 2017, 465, 2033–2052. [Google Scholar] [CrossRef]
  29. Troxel, M.A. et al. [DES] Dark Energy Survey Year 1 results: Cosmological constraints from cosmic shear. Phys. Rev. D 2018, 98, 043528. [Google Scholar] [CrossRef]
  30. Hikage, C. et al. [HSC] Cosmology from cosmic shear power spectra with Subaru Hyper Suprime-Cam first-year data. Publ. Astron. Soc. Jap. 2019, 71, 43. [Google Scholar] [CrossRef]
  31. Asgari, M. et al. [KiDS] KiDS-1000 Cosmology: Cosmic shear constraints and comparison between two point statistics. Astron. Astrophys. 2021, 645, A104. [Google Scholar] [CrossRef]
  32. Abbott, T.M.C. et al. [DES] Dark Energy Survey Year 3 results: Cosmological constraints from galaxy clustering and weak lensing. Phys. Rev. D 2022, 105, 023520. [Google Scholar] [CrossRef]
  33. Beenakker, W.; Venhoek, D. A structured analysis of Hubble tension. arXiv 2021, arXiv:2101.01372. [Google Scholar]
  34. Clifton, T.; Ferreira, P.G.; Padilla, A.; Skordis, C. Modified Gravity and Cosmology. Phys. Rept. 2012, 513, 1–189. [Google Scholar] [CrossRef]
  35. Bamba, K.; Capozziello, S.; Nojiri, S.; Odintsov, S.D. Dark energy cosmology: The equivalent description via different theoretical models and cosmography tests. Astrophys. Space Sci. 2012, 342, 155–228. [Google Scholar] [CrossRef]
  36. Will, C.M. The Confrontation between General Relativity and Experiment. Living Rev. Rel. 2014, 17, 4. [Google Scholar] [CrossRef]
  37. Kapner, D.J.; Cook, T.S.; Adelberger, E.G.; Gundlach, J.H.; Heckel, B.R.; Hoyle, C.D.; Swanson, H.E. Tests of the gravitational inverse-square law below the dark-energy length scale. Phys. Rev. Lett. 2007, 98, 021101. [Google Scholar] [CrossRef] [PubMed]
  38. Lee, J.G.; Adelberger, E.G.; Cook, T.S.; Fleischer, S.M.; Heckel, B.R. New Test of the Gravitational 1/r2 Law at Separations down to 52 μm. Phys. Rev. Lett. 2020, 124, 101101. [Google Scholar] [CrossRef]
  39. Bertotti, B.; Iess, L.; Tortora, P. A test of general relativity using radio links with the Cassini spacecraft. Nature 2003, 425, 374–376. [Google Scholar] [CrossRef]
  40. Abuter, R. et al. [GRAVITY] Detection of the Schwarzschild precession in the orbit of the star S2 near the Galactic centre massive black hole. Astron. Astrophys. 2020, 636, L5. [Google Scholar] [CrossRef]
  41. Collett, T.E.; Oldham, L.J.; Smith, R.J.; Auger, M.W.; Westfall, K.B.; Bacon, D.; Nichol, R.C.; Masters, K.L.; Koyama, K.; Bosch, R.v. A precise extragalactic test of General Relativity. Science 2018, 360, 1342. [Google Scholar] [CrossRef] [PubMed]
  42. Taylor, J.H.; Fowler, L.A.; McCulloch, P.M. Measurements of general relativistic effects in the binary pulsar PSR 1913+16. Nature 1979, 277, 437–440. [Google Scholar] [CrossRef]
  43. Taylor, J.H.; Weisberg, J.M. A new test of general relativity: Gravitational radiation and the binary pulsar PS R 1913+16. Astrophys. J. 1982, 253, 908–920. [Google Scholar] [CrossRef]
  44. Kramer, M.; Stairs, I.H.; Manchester, R.N.; Wex, N.; Deller, A.T.; Coles, W.A.; Ali, M.; Burgay, M.; Camilo, F.; Cognard, I.; et al. Strong-Field Gravity Tests with the Double Pulsar. Phys. Rev. X 2021, 11, 041050. [Google Scholar] [CrossRef]
  45. Abbott, B.P. et al. [LIGO Scientific and Virgo] Tests of General Relativity with the Binary Black Hole Signals from the LIGO-Virgo Catalog GWTC-1. Phys. Rev. D 2019, 100, 104036. [Google Scholar] [CrossRef]
  46. Abbott, R. et al. [LIGO Scientific and Virgo] Tests of general relativity with binary black holes from the second LIGO-Virgo gravitational-wave transient catalog. Phys. Rev. D 2021, 103, 122002. [Google Scholar] [CrossRef]
  47. Abbott, R. et al. [LIGO Scientific, VIRGO and KAGRA] Tests of General Relativity with GWTC-3. arXiv 2022, arXiv:2112.06861. [Google Scholar]
  48. Clifton, T.; Barrow, J.D. The Power of General Relativity. Phys. Rev. D 2005, 72, 103005, Erratum in Phys. Rev. D 2014, 90, 029902. [Google Scholar] [CrossRef]
  49. Rossi, M.; Ballardini, M.; Braglia, M.; Finelli, F.; Paoletti, D.; Starobinsky, A.A.; Umiltà, C. Cosmological constraints on post-Newtonian parameters in effectively massless scalar-tensor theories of gravity. Phys. Rev. D 2019, 100, 103524. [Google Scholar] [CrossRef]
  50. Braglia, M.; Ballardini, M.; Emond, W.T.; Finelli, F.; Gumrukcuoglu, A.E.; Koyama, K.; Paoletti, D. Larger value for H0 by an evolving gravitational constant. Phys. Rev. D 2020, 102, 023529. [Google Scholar] [CrossRef]
  51. Ballesteros, G.; Notari, A.; Rompineve, F. The H0 tension: ΔGN vs. ΔNeff. J. Cosmol. Astropart. Phys. 2020, 11, 024. [Google Scholar] [CrossRef]
  52. Nguyen, N.M.; Huterer, D.; Wen, Y. Evidence for Suppression of Structure Growth in the Concordance Cosmological Model. Phys. Rev. Lett. 2023, 131, 111001. [Google Scholar] [CrossRef] [PubMed]
  53. Sakr, Z.; Sapone, D. Can varying the gravitational constant alleviate the tensions? J. Cosmol. Astropart. Phys. 2022, 03, 034. [Google Scholar] [CrossRef]
  54. Heisenberg, L.; Villarrubia-Rojo, H.; Zosso, J. Simultaneously solving the H0 and σ8 tensions with late dark energy. Phys. Dark Univ. 2023, 39, 101163. [Google Scholar] [CrossRef]
  55. Heisenberg, L.; Villarrubia-Rojo, H.; Zosso, J. Can late-time extensions solve the H0 and σ8 tensions? Phys. Rev. D 2022, 106, 043503. [Google Scholar] [CrossRef]
  56. Lee, B.H.; Lee, W.; Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Thakur, S. Is local H 0 at odds with dark energy EFT? J. Cosmol. Astropart. Phys. 2022, 04, 004. [Google Scholar] [CrossRef]
  57. Handley, W. Curvature tension: Evidence for a closed universe. Phys. Rev. D 2021, 103, L041301. [Google Scholar] [CrossRef]
  58. Valentino, E.D.; Melchiorri, A.; Silk, J. Planck evidence for a closed Universe and a possible crisis for cosmology. Nat. Astron. 2019, 4, 196–203. [Google Scholar] [CrossRef]
  59. Linde, A.D. Inflation with variable Omega. Phys. Lett. B 1995, 351, 99–104. [Google Scholar] [CrossRef]
  60. Linde, A.D. Can we have inflation with Omega > 1? J. Cosmol. Astropart. Phys. 2003, 5, 002. [Google Scholar] [CrossRef]
  61. Park, C.G.; Ratra, B. 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. 2019, 882, 158. [Google Scholar] [CrossRef]
  62. Clarkson, C.; Bassett, B.; Lu, T.H.C. A general test of the Copernican Principle. Phys. Rev. Lett. 2008, 101, 011301. [Google Scholar] [CrossRef]
  63. Madhavacheril, M.S. et al. [ACT] The Atacama Cosmology Telescope: DR6 Gravitational Lensing Map and Cosmological Parameters. arXiv 2023, arXiv:2304.05203. [Google Scholar]
  64. Addison, G.E.; Bennett, C.L.; Halpern, M.; Hinshaw, G.; Weiland, J.L. Revisiting the AL Lensing Anomaly in Planck 2018 Temperature Data. arXiv 2023, arXiv:2310.03127. [Google Scholar]
  65. Schwarz, D.J.; Copi, C.J.; Huterer, D.; Starkman, G.D. CMB Anomalies after Planck. Class. Quant. Grav. 2016, 33, 184001. [Google Scholar] [CrossRef]
  66. Krishnan, C.; Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Yang, T. Running Hubble Tension and a H0 Diagnostic. Phys. Rev. D 2021, 103, 103509. [Google Scholar] [CrossRef]
  67. Akarsu, Ö.; Kumar, S.; Sharma, S.; Tedesco, L. Constraints on a Bianchi type I spacetime extension of the standard ΛCDM model. Phys. Rev. D 2019, 100, 023532. [Google Scholar] [CrossRef]
  68. Akarsu, O.; Valentino, E.D.; Kumar, S.; Ozyigit, M.; Sharma, S. Testing spatial curvature and anisotropic expansion on top of the ΛCDM model. Phys. Dark Univ. 2023, 39, 101162. [Google Scholar] [CrossRef]
  69. King, A.R.; Ellis, G.F.R. Tilted homogeneous cosmological models. Commun. Math. Phys. 1973, 31, 209–242. [Google Scholar] [CrossRef]
  70. Krishnan, C.; Mondol, R.; Sheikh-Jabbari, M.M. Dipole cosmology: The Copernican paradigm beyond FLRW. J. Cosmol. Astropart. Phys. 2023, 07, 020. [Google Scholar] [CrossRef]
  71. Ebrahimian, E.; Krishnan, C.; Mondol, R.; Sheikh-Jabbari, M.M. Towards A Realistic Dipole Cosmology: The Dipole ΛCDM Model. arXiv 2023, arXiv:2305.16177. [Google Scholar] [CrossRef]
  72. Allahyari, A.; Ebrahimian, E.; Mondol, R.; Sheikh-Jabbari, M.M. Big Bang in Dipole Cosmology. arXiv 2023, arXiv:2307.15791. [Google Scholar]
  73. Krishnan, C.; Mohayaee, R.; Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Yin, L. Does Hubble tension signal a breakdown in FLRW cosmology? Class. Quant. Grav. 2021, 38, 184001. [Google Scholar] [CrossRef]
  74. Kashlinsky, A.; Atrio-Barandela, F.; Kocevski, D.; Ebeling, H. A measurement of large-scale peculiar velocities of clusters of galaxies: Results and cosmological implications. Astrophys. J. Lett. 2009, 686, L49–L52. [Google Scholar] [CrossRef]
  75. Migkas, K.; Schellenberger, G.; Reiprich, T.H.; Pacaud, F.; Ramos-Ceja, M.E.; Lovisari, L. Probing cosmic isotropy with a new X-ray galaxy cluster sample through the LXT scaling relation. Astron. Astrophys. 2020, 636, A15. [Google Scholar] [CrossRef]
  76. Migkas, K.; Pacaud, F.; Schellenberger, G.; Erler, J.; Nguyen-Dang, N.T.; Reiprich, T.H.; Ramos-Ceja, M.E.; Lovisari, L. Cosmological implications of the anisotropy of ten galaxy cluster scaling relations. Astron. Astrophys. 2021, 649, A151. [Google Scholar] [CrossRef]
  77. Watkins, R.; Allen, T.; Bradford, C.J.; Ramon, A.; Walker, A.; Feldman, H.A.; Cionitti, R.; Al-Shorman, Y.; Kourkchi, E.; Tully, R.B. Analysing the large-scale bulk flow using cosmicflows4: Increasing tension with the standard cosmological model. Mon. Not. Roy. Astron. Soc. 2023, 524, 1885–1892. [Google Scholar] [CrossRef]
  78. Whitford, A.M.; Howlett, C.; Davis, T.M. Evaluating bulk flow estimators for CosmicFlows–4 measurements. Mon. Not. Roy. Astron. Soc. 2023, 526, 3051–3071. [Google Scholar] [CrossRef]
  79. Hoffman, Y.; Valade, A.; Libeskind, N.I.; Sorce, J.G.; Tully, R.B.; Pfeifer, S.; Gottlöber, S.; Pomaréde, D. The large scale velocity field from the Cosmicflows-4 data. arXiv 2023, arXiv:2311.01340. [Google Scholar] [CrossRef]
  80. Singal, A.K. Large peculiar motion of the solar system from the dipole anisotropy in sky brightness due to distant radio sources. Astrophys. J. Lett. 2011, 742, L23. [Google Scholar] [CrossRef]
  81. Rubart, M.; Schwarz, D.J. Cosmic radio dipole from NVSS and WENSS. Astron. Astrophys. 2013, 555, A117. [Google Scholar] [CrossRef]
  82. Singal, A.K. Large disparity in cosmic reference frames determined from the sky distributions of radio sources and the microwave background radiation. Phys. Rev. D 2019, 100, 063501. [Google Scholar] [CrossRef]
  83. Secrest, N.J.; von Hausegger, S.; Rameez, M.; Mohayaee, R.; Sarkar, S.; Colin, J. A Test of the Cosmological Principle with Quasars. Astrophys. J. Lett. 2021, 908, L51. [Google Scholar] [CrossRef]
  84. Siewert, T.M.; Schmidt-Rubart, M.; Schwarz, D.J. Cosmic radio dipole: Estimators and frequency dependence. Astron. Astrophys. 2021, 653, A9. [Google Scholar] [CrossRef]
  85. Secrest, N.J.; von Hausegger, S.; Rameez, M.; Mohayaee, R.; Sarkar, S. A Challenge to the Standard Cosmological Model. Astrophys. J. Lett. 2022, 937, L31. [Google Scholar] [CrossRef]
  86. Pranav, P.; Buchert, T. Homology reveals significant anisotropy in the cosmic microwave background. arXiv 2023, arXiv:2308.10738. [Google Scholar]
  87. Mittal, V.; Oayda, O.T.; Lewis, G.F. The Cosmic Dipole in the Quaia Sample of Quasars: A Bayesian Analysis. arXiv 2023, arXiv:2311.14938. [Google Scholar] [CrossRef]
  88. Aluri, P.K.; Cea, P.; Chingangbam, P.; Chu, M.C.; Clowes, R.G.; Hutsemékers, D.; Kochappan, J.P.; Lopez, A.M.; Liu, L.; Martens, N.C.M.; et al. Is the observable Universe consistent with the cosmological principle? Class. Quant. Grav. 2023, 40, 094001. [Google Scholar] [CrossRef]
  89. Jones, J.; Copi, C.J.; Starkman, G.D.; Akrami, Y. The Universe is not statistically isotropic. arXiv 2023, arXiv:2310.12859. [Google Scholar]
  90. Akrami, Y. et al. [COMPACT] The Search for the Topology of the Universe Has Just Begun. arXiv 2022, arXiv:2210.11426. [Google Scholar]
  91. Petersen, P. et al. [COMPACT] Cosmic topology. Part I. Limits on orientable Euclidean manifolds from circle searches. J. Cosmol. Astropart. Phys. 2023, 1, 30. [Google Scholar] [CrossRef]
  92. Eskilt, J.R. et al. [COMPACT] Cosmic topology. Part II. Eigenmodes, correlation matrices, and detectability of orientable Euclidean manifolds. arXiv 2023, arXiv:2306.17112. [Google Scholar]
  93. Giani, L.; Howlett, C.; Said, K.; Davis, T.; Vagnozzi, S. An effective description of Laniakea: Impact on cosmology and the local determination of the Hubble constant. J. Cosmol. Astropart. Phys. 2024, 1, 71. [Google Scholar] [CrossRef]
  94. Krishnan, C.; Mohayaee, R.; Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Yin, L. Hints of FLRW breakdown from supernovae. Phys. Rev. D 2022, 105, 063514. [Google Scholar] [CrossRef]
  95. Zhai, Z.; Percival, W.J. Sample variance for supernovae distance measurements and the Hubble tension. Phys. Rev. D 2022, 106, 103527. [Google Scholar] [CrossRef]
  96. Conville, R.M.; Ó Colgáin, E. Anisotropic distance ladder in Pantheon+supernovae. Phys. Rev. D 2023, 108, 123533. [Google Scholar] [CrossRef]
  97. Fosalba, P.; Gaztanaga, E. Explaining Cosmological Anisotropy: Evidence for Causal Horizons from CMB data. arXiv 2020, arXiv:2011.00910. [Google Scholar] [CrossRef]
  98. Yeung, S.; Chu, M.C. Directional variations of cosmological parameters from the Planck CMB data. Phys. Rev. D 2022, 105, 083508. [Google Scholar] [CrossRef]
  99. Kovács, A.; García-Bellido, J. Cosmic troublemakers: The Cold Spot, the Eridanus Supervoid, and the Great Walls. Mon. Not. Roy. Astron. Soc. 2016, 462, 1882–1893. [Google Scholar] [CrossRef]
  100. Kovács, A. et al. [DES] The DES view of the Eridanus supervoid and the CMB cold spot. Mon. Not. Roy. Astron. Soc. 2022, 510, 216–229. [Google Scholar] [CrossRef]
  101. Lambas, D.G.; Hansen, F.K.; Toscano, F.; Luparello, H.E.; Boero, E.F. The CMB Cold Spot as predicted by foregrounds around nearby galaxies. Astron. Astrophys. 2024, 681, A2. [Google Scholar] [CrossRef]
  102. Aiola, S. et al. [ACT] The Atacama Cosmology Telescope: DR4 Maps and Cosmological Parameters. J. Cosmol. Astropart. Phys. 2020, 12, 047. [Google Scholar] [CrossRef]
  103. Balkenhol, L. et al. [SPT-3G] Measurement of the CMB temperature power spectrum and constraints on cosmology from the SPT-3G 2018 TT, TE, and EE dataset. Phys. Rev. D 2023, 108, 023510. [Google Scholar] [CrossRef]
  104. Mortsell, E.; Goobar, A.; Johansson, J.; Dhawan, S. The Hubble Tension Revisited: Additional Local Distance Ladder Uncertainties. Astrophys. J. 2022, 935, 58. [Google Scholar] [CrossRef]
  105. Perivolaropoulos, L.; Skara, F. Hubble tension or a transition of the Cepheid SnIa calibrator parameters? Phys. Rev. D 2021, 104, 123511. [Google Scholar] [CrossRef]
  106. Perivolaropoulos, L.; Skara, F. On the homogeneity of SnIa absolute magnitude in the Pantheon+ sample. Mon. Not. Roy. Astron. Soc. 2023, 520, 5110–5125. [Google Scholar] [CrossRef]
  107. Lane, Z.G.; Seifert, A.; Ridden-Harper, R.; Wagner, J.; Wiltshire, D.L. Cosmological foundations revisited with Pantheon+. arXiv 2023, arXiv:2311.01438. [Google Scholar]
  108. Boruah, S.S.; Hudson, M.J.; Lavaux, G. Cosmic flows in the nearby Universe: New peculiar velocities from SNe and cosmological constraints. Mon. Not. Roy. Astron. Soc. 2020, 498, 2703–2718. [Google Scholar] [CrossRef]
  109. Said, K.; Colless, M.; Magoulas, C.; Lucey, J.R.; Hudson, M.J. Joint analysis of 6dFGS and SDSS peculiar velocities for the growth rate of cosmic structure and tests of gravity. Mon. Not. Roy. Astron. Soc. 2020, 497, 1275–1293. [Google Scholar] [CrossRef]
  110. Hollinger, A.M.; Hudson, M.J. Cosmological parameters estimated from velocity—Density comparisons: Calibrating 2M++. arXiv 2023, arXiv:2312.03904. [Google Scholar] [CrossRef]
  111. Ade, P.A.R. et al. [Planck] Planck 2013 results. XX. Cosmology from Sunyaev–Zeldovich cluster counts. Astron. Astrophys. 2014, 571, A20. [Google Scholar] [CrossRef]
  112. Macaulay, E.; Wehus, I.K.; Eriksen, H.K. Lower Growth Rate from Recent Redshift Space Distortion Measurements than Expected from Planck. Phys. Rev. Lett. 2013, 111, 161301. [Google Scholar] [CrossRef]
  113. Battye, R.A.; Charnock, T.; Moss, A. Tension between the power spectrum of density perturbations measured on large and small scales. Phys. Rev. D 2015, 91, 103508. [Google Scholar] [CrossRef]
  114. Nesseris, S.; Pantazis, G.; Perivolaropoulos, L. Tension and constraints on modified gravity parametrizations of Geff(z) from growth rate and Planck data. Phys. Rev. D 2017, 96, 023542. [Google Scholar] [CrossRef]
  115. Kazantzidis, L.; Perivolaropoulos, L. Evolution of the 8 tension with the Planck15/ΛCDM determination and implications for modified gravity theories. Phys. Rev. D 2018, 97, 103503. [Google Scholar] [CrossRef]
  116. Skara, F.; Perivolaropoulos, L. Tension of the EG statistic and redshift space distortion data with the Planck-ΛCDM model and implications for weakening gravity. Phys. Rev. D 2020, 101, 063521. [Google Scholar] [CrossRef]
  117. Quelle, A.; Maroto, A.L. On the tension between growth rate and CMB data. Eur. Phys. J. C 2020, 80, 369. [Google Scholar] [CrossRef]
  118. Li, E.K.; Du, M.; Zhou, Z.H.; Zhang, H.; Xu, L. Testing the effect of H0 on 8 tension using a Gaussian process method. Mon. Not. Roy. Astron. Soc. 2021, 501, 4452–4463. [Google Scholar] [CrossRef]
  119. Benisty, D. Quantifying the S8 tension with the Redshift Space Distortion data set. Phys. Dark Univ. 2021, 31, 100766. [Google Scholar] [CrossRef]
  120. Nunes, R.C.; Vagnozzi, S. Arbitrating the S8 discrepancy with growth rate measurements from redshift-space distortions. Mon. Not. Roy. Astron. Soc. 2021, 505, 5427–5437. [Google Scholar] [CrossRef]
  121. Esposito, M.; Iršič, V.; Costanzi, M.; Borgani, S.; Saro, A.; Viel, M. Weighing cosmic structures with clusters of galaxies and the intergalactic medium. Mon. Not. Roy. Astron. Soc. 2022, 515, 857–870. [Google Scholar] [CrossRef]
  122. Adil, S.A.; Akarsu, Ö.; Malekjani, M.; Ó Colgáin, E.; Pourojaghi, S.; Sen, A.A.; Sheikh-Jabbari, M.M. S8 increases with effective redshift in ΛCDM cosmology. arXiv 2023, arXiv:2303.06928. [Google Scholar]
  123. Kovács, A. et al. [DES] More out of less: An excess integrated Sachs-Wolfe signal from supervoids mapped out by the Dark Energy Survey. Mon. Not. Roy. Astron. Soc. 2019, 484, 5267–5277. [Google Scholar] [CrossRef]
  124. Kovács, A.; Beck, R.; Smith, A.; Rácz, G.; Csabai, I.; Szapudi, I. Evidence for a high-z ISW signal from supervoids in the distribution of eBOSS quasars. Mon. Not. Roy. Astron. Soc. 2022, 513, 15–26. [Google Scholar] [CrossRef]
  125. Tutusaus, I.; Bonvin, C.; Grimm, N. First measurement of the Weyl potential evolution from the Year 3 Dark Energy Survey data: Localising the σ8 tension. arXiv 2023, arXiv:2312.06434. [Google Scholar]
  126. Farren, G.S. et al. [ACT] The Atacama Cosmology Telescope: Cosmology from cross-correlations of unWISE galaxies and ACT DR6 CMB lensing. arXiv 2023, arXiv:2309.05659. [Google Scholar]
  127. Kim, J. et al. [ACT and DESI] The Atacama Cosmology Telescope DR6 and DESI: Structure formation over cosmic time with a measurement of the cross-correlation of CMB Lensing and Luminous Red Galaxies. arXiv 2024, arXiv:2407.04606. [Google Scholar]
  128. Sailer, N.; Kim, J.; Ferraro, S.; Madhavacheril, M.S.; White, M.; Abril-Cabezas, I.; Aguilar, J.N.; Ahlen, S.; Bond, J.R.; Brooks, D.; et al. Cosmological constraints from the cross-correlation of DESI Luminous Red Galaxies with CMB lensing from Planck PR4 and ACT DR6. arXiv 2024, arXiv:2407.04607. [Google Scholar]
  129. Tiwari, P.; Nusser, A. Revisiting the NVSS number count dipole. J. Cosmol. Astropart. Phys. 2016, 03, 062. [Google Scholar] [CrossRef]
  130. Ma, C.; Zhang, T.J. Power of Observational Hubble Parameter Data: A Figure of Merit Exploration. Astrophys. J. 2011, 730, 74. [Google Scholar] [CrossRef]
  131. Moresco, M. Raising the bar: New constraints on the Hubble parameter with cosmic chronometers at z ∼ 2. Mon. Not. Roy. Astron. Soc. 2015, 450, L16–L20. [Google Scholar] [CrossRef]
  132. Moresco, M.; Jimenez, R.; Verde, L.; Pozzetti, L.; Cimatti, A.; Citro, A. Setting the Stage for Cosmic Chronometers. I. Assessing the Impact of Young Stellar Populations on Hubble Parameter Measurements. Astrophys. J. 2018, 868, 84. [Google Scholar] [CrossRef]
  133. Vagnozzi, S.; Loeb, A.; Moresco, M. Eppur è piatto? The Cosmic Chronometers Take on Spatial Curvature and Cosmic Concordance. Astrophys. J. 2021, 908, 84. [Google Scholar] [CrossRef]
  134. Jiao, K.; Borghi, N.; Moresco, M.; Zhang, T.J. New Observational H(z) Data from Full-spectrum Fitting of Cosmic Chronometers in the LEGA-C Survey. Astrophys. J. Suppl. 2023, 265, 48. [Google Scholar] [CrossRef]
  135. Holsclaw, T.; Alam, U.; Sanso, B.; Lee, H.; Heitmann, K.; Habib, S.; Higdon, D. Nonparametric Reconstruction of the Dark Energy Equation of State. Phys. Rev. D 2010, 82, 103502. [Google Scholar] [CrossRef]
  136. Holsclaw, T.; Alam, U.; Sanso, B.; Lee, H.; Heitmann, K.; Habib, S.; Higdon, D. Nonparametric Dark Energy Reconstruction from Supernova Data. Phys. Rev. Lett. 2010, 105, 241302. [Google Scholar] [CrossRef]
  137. Shafieloo, A.; Kim, A.G.; Linder, E.V. Gaussian Process Cosmography. Phys. Rev. D 2012, 85, 123530. [Google Scholar] [CrossRef]
  138. Keeley, R.E.; Shafieloo, A.; Zhao, G.B.; Vazquez, J.A.; Koo, H. Reconstructing the Universe: Testing the Mutual Consistency of the Pantheon and SDSS/eBOSS BAO Data Sets with Gaussian Processes. Astron. J. 2021, 161, 151. [Google Scholar] [CrossRef]
  139. Seikel, M.; Clarkson, C.; Smith, M. Reconstruction of dark energy and expansion dynamics using Gaussian processes. J. Cosmol. Astropart. Phys. 2012, 6, 36. [Google Scholar] [CrossRef]
  140. Nesseris, S.; Garcia-Bellido, J. A new perspective on Dark Energy modeling via Genetic Algorithms. J. Cosmol. Astropart. Phys. 2012, 11, 33. [Google Scholar] [CrossRef]
  141. Arjona, R.; Nesseris, S. What can Machine Learning tell us about the background expansion of the Universe? Phys. Rev. D 2020, 101, 123525. [Google Scholar] [CrossRef]
  142. Mukherjee, P.; Said, J.L.; Mifsud, J. Neural network reconstruction of H’(z) and its application in teleparallel gravity. J. Cosmol. Astropart. Phys. 2022, 12, 29. [Google Scholar] [CrossRef]
  143. Bengaly, C.; Dantas, M.A.; Casarini, L.; Alcaniz, J. Measuring the Hubble constant with cosmic chronometers: A machine learning approach. Eur. Phys. J. C 2023, 83, 548. [Google Scholar] [CrossRef]
  144. Giambagli, L.; Fanelli, D.; Risaliti, G.; Signorini, M. Nonparametric analysis of the Hubble diagram with neural networks. Astron. Astrophys. 2023, 678, A13. [Google Scholar] [CrossRef]
  145. Gómez-Vargas, I.; Andrade, J.B.; Vázquez, J.A. Neural networks optimized by genetic algorithms in cosmology. Phys. Rev. D 2023, 107, 043509. [Google Scholar] [CrossRef]
  146. Dialektopoulos, K.F.; Mukherjee, P.; Said, J.L.; Mifsud, J. Neural network reconstruction of cosmology using the Pantheon compilation. Eur. Phys. J. C 2023, 83, 956. [Google Scholar] [CrossRef]
  147. Medel-Esquivel, R.; Gómez-Vargas, I.; Sánchez, A.A.M.; García-Salcedo, R.; Vázquez, J.A. Cosmological parameter estimation with Genetic Algorithms. Universe 2024, 10, 11. [Google Scholar] [CrossRef]
  148. Gómez-Vargas, I.; Esquivel, R.M.; García-Salcedo, R.; Vázquez, J.A. Neural network reconstructions for the Hubble parameter, growth rate and distance modulus. Eur. Phys. J. C 2023, 83, 304. [Google Scholar] [CrossRef]
  149. Akarsu, O.; Ó Colgáin, E.; Özulker, E.; Thakur, S.; Yin, L. Inevitable manifestation of wiggles in the expansion of the late Universe. Phys. Rev. D 2023, 107, 123526. [Google Scholar] [CrossRef]
  150. Zhao, G.B.; Raveri, M.; Pogosian, L.; Wang, Y.; Crittenden, R.G.; Handley, W.J.; Percival, W.J.; Beutler, F.; Brinkmann, J.; Chuang, C.H.; et al. Dynamical dark energy in light of the latest observations. Nat. Astron. 2017, 1, 627–632. [Google Scholar] [CrossRef]
  151. Padilla, L.E.; Tellez, L.O.; Escamilla, L.A.; Vazquez, J.A. Cosmological Parameter Inference with Bayesian Statistics. Universe 2021, 7, 213. [Google Scholar] [CrossRef]
  152. Wang, Y.; Pogosian, L.; Zhao, G.B.; Zucca, A. Evolution of dark energy reconstructed from the latest observations. Astrophys. J. Lett. 2018, 869, L8. [Google Scholar] [CrossRef]
  153. Escamilla, L.A.; Vazquez, J.A. Model selection applied to reconstructions of the Dark Energy. Eur. Phys. J. C 2023, 83, 251. [Google Scholar] [CrossRef]
  154. Escamilla, L.A.; Akarsu, O.; Valentino, E.D.; Vazquez, J.A. Model-independent reconstruction of the interacting dark energy kernel: Binned and Gaussian process. J. Cosmol. Astropart. Phys. 2023, 11, 051. [Google Scholar] [CrossRef]
  155. Ó Colgáin, E.; Sheikh-Jabbari, M.M. Elucidating cosmological model dependence with H0. Eur. Phys. J. C 2021, 81, 892. [Google Scholar] [CrossRef]
  156. Jimenez, R.; Loeb, A. Constraining cosmological parameters based on relative galaxy ages. Astrophys. J. 2002, 573, 37–42. [Google Scholar] [CrossRef]
  157. Busti, V.C.; Clarkson, C.; Seikel, M. Evidence for a Lower Value for H0 from Cosmic Chronometers Data? Mon. Not. Roy. Astron. Soc. 2014, 441, 11. [Google Scholar] [CrossRef]
  158. Gómez-Valent, A.; Amendola, L. H0 from cosmic chronometers and Type Ia supernovae, with Gaussian Processes and the novel Weighted Polynomial Regression method. J. Cosmol. Astropart. Phys. 2018, 04, 051. [Google Scholar] [CrossRef]
  159. Haridasu, B.S.; Luković, V.V.; Moresco, M.; Vittorio, N. An improved model-independent assessment of the late-time cosmic expansion. J. Cosmol. Astropart. Phys. 2018, 10, 015. [Google Scholar] [CrossRef]
  160. Moresco, M. Addressing the Hubble tension with cosmic chronometers. arXiv 2023, arXiv:2307.09501. [Google Scholar]
  161. Liu, G.; Wang, Y.; Zhao, W. Testing the consistency of early and late cosmological parameters with BAO and CMB data. arXiv 2024, arXiv:2401.10571. [Google Scholar] [CrossRef]
  162. Riess, A.G. et al. [Supernova Search Team] Observational evidence from supernovae for an accelerating universe and a cosmological constant. Astron. J. 1998, 116, 1009–1038. [Google Scholar] [CrossRef]
  163. Perlmutter, S. et al. [Supernova Cosmology Project] Measurements of Ω and Λ from 42 high redshift supernovae. Astrophys. J. 1999, 517, 565–586. [Google Scholar] [CrossRef]
  164. Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Solomon, R.; Dainotti, M.G.; Stojkovic, D. Putting Flat ΛCDM In The (Redshift) Bin. arXiv 2022, arXiv:2206.11447. [Google Scholar]
  165. Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Solomon, R. High redshift ΛCDM cosmology: To bin or not to bin? Phys. Dark Univ. 2023, 40, 101216. [Google Scholar] [CrossRef]
  166. Kerr, A.W.; Hauck, J.C.; Mashhoon, B. Standard clocks, orbital precession and the cosmological constant. Class. Quant. Grav. 2003, 20, 2727. [Google Scholar] [CrossRef]
  167. Arakida, H.; Kasai, M. Effect of the cosmological constant on the bending of light and the cosmological lens equation. Phys. Rev. D 2012, 85, 023006. [Google Scholar] [CrossRef]
  168. Arakida, H. Note on the perihelion/periastron advance due to cosmological constant. Int. J. Theor. Phys. 2013, 52, 1408–1414. [Google Scholar] [CrossRef]
  169. Benisty, D.; Davis, A.C.; Evans, N.W. Constraining Dark Energy from the Local Group Dynamics. Astrophys. J. Lett. 2023, 953, L2. [Google Scholar] [CrossRef]
  170. Benisty, D.; Wagner, J.; Staicova, D. Dark Energy as a Critical Period in Binary Motion: Bounds from Multi-scale Binaries. arXiv 2023, arXiv:2310.11488. [Google Scholar] [CrossRef]
  171. Benisty, D. Dark energy Constraints from different Local Group Histories. arXiv 2024, arXiv:2401.09546. [Google Scholar]
  172. Chevallier, M.; Polarski, D. Accelerating universes with scaling dark matter. Int. J. Mod. Phys. D 2001, 10, 213–224. [Google Scholar] [CrossRef]
  173. Linder, E.V. Exploring the expansion history of the universe. Phys. Rev. Lett. 2003, 90, 091301. [Google Scholar] [CrossRef]
  174. Brout, D.; Scolnic, D.; Popovic, B.; Riess, A.G.; Zuntz, J.; Kessler, R.; Carr, A.; Davis, T.M.; Hinton, S.; Jones, D.; et al. The Pantheon+ Analysis: Cosmological Constraints. Astrophys. J. 2022, 938, 110. [Google Scholar] [CrossRef]
  175. Rubin, D.; Aldering, G.; Betoule, M.; Fruchter, A.; Huang, X.; Kim, A.G.; Lidman, C.; Linder, E.; Perlmutter, S.; Ruiz-Lapuente, P.; et al. Union Through UNITY: Cosmology with 2000 SNe Using a Unified Bayesian Framework. arXiv 2023, arXiv:2311.12098. [Google Scholar]
  176. Abbott, T.M.C. et al. [DES] The Dark Energy Survey: Cosmology Results with ~1500 New High-redshift Type Ia Supernovae Using The Full 5-year Dataset. arXiv 2024, arXiv:2401.02929. [Google Scholar]
  177. Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Yin, L. Can dark energy be dynamical? Phys. Rev. D 2021, 104, 023510. [Google Scholar] [CrossRef]
  178. Alam, S. et al. [eBOSS] Completed SDSS-IV extended Baryon Oscillation Spectroscopic Survey: Cosmological implications from two decades of spectroscopic surveys at the Apache Point Observatory. Phys. Rev. D 2021, 103, 083533. [Google Scholar] [CrossRef]
  179. Cattoën, C.; Visser, M. The Hubble series: Convergence properties and redshift variables. Class. Quant. Grav. 2007, 24, 5985–5998. [Google Scholar] [CrossRef]
  180. Dong, F.; Park, C.; Hong, S.E.; Kim, J.; Hwang, H.S.; Park, H.; Appleby, S. Tomographic Alcock–Paczyński Test with Redshift-space Correlation Function: Evidence for the Dark Energy Equation-of-state Parameter w > −1. Astrophys. J. 2023, 953, 98. [Google Scholar] [CrossRef]
  181. Bocquet, S. et al. [DES and SPT] SPT Clusters with DES and HST Weak Lensing. II. Cosmological Constraints from the Abundance of Massive Halos. arXiv 2024, arXiv:2401.02075. [Google Scholar]
  182. Huterer, D. Growth of cosmic structure. Astron. Astrophys. Rev. 2023, 31, 2. [Google Scholar] [CrossRef]
  183. Amon, A.; Efstathiou, G. A non-linear solution to the S8 tension? arXiv 2022, arXiv:2206.11794. [Google Scholar] [CrossRef]
  184. Preston, C.; Amon, A.; Efstathiou, G. A non-linear solution to the S8 tension—II. Analysis of DES Year 3 cosmic shear. Mon. Not. Roy. Astron. Soc. 2023, 525, 5554–5564. [Google Scholar] [CrossRef]
  185. Qu, F.J. et al. [ACT] The Atacama Cosmology Telescope: A Measurement of the DR6 CMB Lensing Power Spectrum and its Implications for Structure Growth. arXiv 2023, arXiv:2304.05202. [Google Scholar]
  186. Marques, G.A. et al. [ACT and DES] Cosmological constraints from the tomography of DES-Y3 galaxies with CMB lensing from ACT DR4. arXiv 2023, arXiv:2306.17268. [Google Scholar]
  187. Miyatake, H.; Harikane, Y.; Ouchi, M.; Ono, Y.; Yamamoto, N.; Nishizawa, A.J.; Bahcall, N.; Miyazaki, S.; Malagón, A.A.P. First Identification of a CMB Lensing Signal Produced by 1.5 Million Galaxies at z∼4: Constraints on Matter Density Fluctuations at High Redshift. Phys. Rev. Lett. 2022, 129, 061301. [Google Scholar] [CrossRef] [PubMed]
  188. Alonso, D.; Fabbian, G.; Storey-Fisher, K.; Eilers, A.C.; García-García, C.; Hogg, D.W.; Rix, H.W. Constraining cosmology with the Gaia-unWISE Quasar Catalog and CMB lensing: Structure growth. J. Cosmol. Astropart. Phys. 2023, 11, 043. [Google Scholar] [CrossRef]
  189. Chapman, M.J.; Zhai, Z.; Percival, W.J. Isolating the linear signal when making redshift space distortion measurements. Mon. Not. Roy. Astron. Soc. 2023, 525, 2135–2153. [Google Scholar] [CrossRef]
  190. Porredon, A. et al. [DES] Dark Energy Survey Year 3 results: Cosmological constraints from galaxy clustering and galaxy-galaxy lensing using the MagLim lens sample. Phys. Rev. D 2022, 106, 103530. [Google Scholar] [CrossRef]
  191. Abbott, T.M.C. et al. [Kilo-Degree Survey and Dark Energy Survey] DES Y3 + KiDS-1000: Consistent cosmology combining cosmic shear surveys. Open J. Astrophys. 2023, 6, 2305.17173. [Google Scholar] [CrossRef]
  192. Addison, G.E.; Watts, D.J.; Bennett, C.L.; Halpern, M.; Hinshaw, G.; Weiland, J.L. Elucidating ΛCDM: Impact of Baryon Acoustic Oscillation Measurements on the Hubble Constant Discrepancy. Astrophys. J. 2018, 853, 119. [Google Scholar] [CrossRef]
  193. Cuceu, A.; Farr, J.; Lemos, P.; Font-Ribera, A. Baryon Acoustic Oscillations and the Hubble Constant: Past, Present and Future. J. Cosmol. Astropart. Phys. 2019, 10, 044. [Google Scholar] [CrossRef]
  194. Schöneberg, N.; Lesgourgues, J.; Hooper, D.C. The BAO+BBN take on the Hubble tension. J. Cosmol. Astropart. Phys. 2019, 10, 029. [Google Scholar] [CrossRef]
  195. Schöneberg, N.; Verde, L.; Gil-Marín, H.; Brieden, S. BAO+BBN revisited—Growing the Hubble tension with a 0.7 km s−1 Mpc−1 constraint. J. Cosmol. Astropart. Phys. 2022, 11, 039. [Google Scholar] [CrossRef]
  196. Freedman, W.L. et al. [HST] Final results from the Hubble Space Telescope key project to measure the Hubble constant. Astrophys. J. 2001, 553, 47–72. [Google Scholar] [CrossRef]
  197. Freedman, W.L.; Madore, B.F.; Scowcroft, V.; Burns, C.; Monson, A.; Persson, S.E.; Seibert, M.; Rigby, J. Carnegie Hubble Program: A Mid-Infrared Calibration of the Hubble Constant. Astrophys. J. 2012, 758, 24. [Google Scholar] [CrossRef]
  198. Cardona, W.; Kunz, M.; Pettorino, V. Determining H0 with Bayesian hyper-parameters. J. Cosmol. Astropart. Phys. 2017, 03, 056. [Google Scholar] [CrossRef]
  199. Follin, B.; Knox, L. Insensitivity of the distance ladder Hubble constant determination to Cepheid calibration modelling choices. Mon. Not. Roy. Astron. Soc. 2018, 477, 4534–4542. [Google Scholar] [CrossRef]
  200. Riess, A.G.; Anand, G.S.; Yuan, W.; Casertano, S.; Dolphin, A.; Macri, L.M.; Breuval, L.; Scolnic, D.; Perrin, M.; Anderson, R.I. Crowded No More: The Accuracy of the Hubble Constant Tested with High-resolution Observations of Cepheids by JWST. Astrophys. J. Lett. 2023, 956, L18. [Google Scholar] [CrossRef]
  201. Schombert, J.; McGaugh, S.; Lelli, F. Using the Baryonic Tully–Fisher Relation to Measure Ho. Astron. J. 2020, 160, 71. [Google Scholar] [CrossRef]
  202. de Jaeger, T.; Stahl, B.E.; Zheng, W.; Filippenko, A.V.; Riess, A.G.; Galbany, L. A measurement of the Hubble constant from Type II supernovae. Mon. Not. Roy. Astron. Soc. 2020, 496, 3402–3411. [Google Scholar] [CrossRef]
  203. de Jaeger, T.; Galbany, L.; Riess, A.G.; Stahl, B.E.; Shappee, B.J.; Filippenko, A.V.; Zheng, W. A 5 per cent measurement of the Hubble–Lemaître constant from Type II supernovae. Mon. Not. Roy. Astron. Soc. 2022, 514, 4620–4628. [Google Scholar] [CrossRef]
  204. Abbott, B.P. et al. [LIGO Scientific, Virgo, 1M2H, Dark Energy Camera GW-E, DES, DLT40, Las Cumbres Observatory, VINROUGE and MASTER] A gravitational-wave standard siren measurement of the Hubble constant. Nature 2017, 551, 85–88. [Google Scholar] [CrossRef]
  205. Nicolaou, C.; Lahav, O.; Lemos, P.; Hartley, W.; Braden, J. The Impact of Peculiar Velocities on the Estimation of the Hubble Constant from Gravitational Wave Standard Sirens. Mon. Not. Roy. Astron. Soc. 2020, 495, 90–97. [Google Scholar] [CrossRef]
  206. Reid, M.J.; Pesce, D.W.; Riess, A.G. An Improved Distance to NGC 4258 and its Implications for the Hubble Constant. Astrophys. J. Lett. 2019, 886, L27. [Google Scholar] [CrossRef]
  207. Freedman, W.L.; Madore, B.F.; Hoyt, T.; Jang, I.S.; Beaton, R.; Lee, M.G.; Monson, A.; Neeley, J.; Rich, J. Calibration of the Tip of the Red Giant Branch (TRGB). arXiv 2020, arXiv:2002.01550. [Google Scholar]
  208. Anderson, R.I.; Koblischke, N.W.; Eyer, L. Reconciling astronomical distance scales with variable red giant stars. arXiv 2023, arXiv:2303.04790. [Google Scholar]
  209. Scolnic, D.; Riess, A.G.; Wu, J.; Li, S.; Anand, G.S.; Beaton, R.; Casertano, S.; Anderson, R.I.; Dhawan, S.; Ke, X. CATS: The Hubble Constant from Standardized TRGB and Type Ia Supernova Measurements. Astrophys. J. Lett. 2023, 954, L31. [Google Scholar] [CrossRef]
  210. Uddin, S.A.; Burns, C.R.; Phillips, M.M.; Suntzeff, N.B.; Freedman, W.L.; Brown, P.J.; Morrell, N.; Hamuy, M.; Krisciunas, K.; Wang, L.; et al. Carnegie Supernova Project-I and -II: Measurements of H0 using Cepheid, TRGB, and SBF Distance Calibration to Type Ia Supernovae. arXiv 2023, arXiv:2308.01875. [Google Scholar]
  211. Wong, K.C.; Suyu, S.H.; Chen, G.C.F.; Rusu, C.E.; Millon, M.; Sluse, D.; Bonvin, V.; Fassnacht, C.D.; Taubenberger, S.; Auger, M.W.; et al. H0LiCOW—XIII. A 2.4 per cent measurement of H0 from lensed quasars: 5.3σ tension between early- and late-Universe probes. Mon. Not. Roy. Astron. Soc. 2020, 498, 1420–1439. [Google Scholar] [CrossRef]
  212. Shajib, A.J. et al. [DES] STRIDES: A 3.9 per cent measurement of the Hubble constant from the strong lens system DES J0408−5354. Mon. Not. Roy. Astron. Soc. 2020, 494, 6072–6102. [Google Scholar] [CrossRef]
  213. Millon, M.; Galan, A.; Courbin, F.; Treu, T.; Suyu, S.H.; Ding, X.; Birrer, S.; Chen, G.C.F.; Shajib, A.J.; Sluse, D.; et al. TDCOSMO. I. An exploration of systematic uncertainties in the inference of H0 from time-delay cosmography. Astron. Astrophys. 2020, 639, A101. [Google Scholar] [CrossRef]
  214. Yang, T.; Birrer, S.; Hu, B. The first simultaneous measurement of Hubble constant and post-Newtonian parameter from Time-Delay Strong Lensing. Mon. Not. Roy. Astron. Soc. 2020, 497, L56–L61. [Google Scholar] [CrossRef]
  215. Birrer, S.; Shajib, A.J.; Galan, A.; Millon, M.; Treu, T.; Agnello, A.; Auger, M.; Chen, G.C.F.; Christensen, L.; Collett, T.; et al. TDCOSMO—IV. Hierarchical time-delay cosmography—Joint inference of the Hubble constant and galaxy density profiles. Astron. Astrophys. 2020, 643, A165. [Google Scholar] [CrossRef]
  216. Denzel, P.; Coles, J.P.; Saha, P.; Williams, L.L.R. The Hubble constant from eight time-delay galaxy lenses. Mon. Not. Roy. Astron. Soc. 2021, 501, 784–801. [Google Scholar] [CrossRef]
  217. Shajib, A.J.; Mozumdar, P.; Chen, G.C.F.; Treu, T.; Cappellari, M.; Knabel, S.; Suyu, S.H.; Bennert, V.N.; Frieman, J.A.; Sluse, D.; et al. TDCOSMO. XII. Improved Hubble constant measurement from lensing time delays using spatially resolved stellar kinematics of the lens galaxy. Astron. Astrophys. 2023, 673, A9. [Google Scholar] [CrossRef]
  218. Palmese, A. et al. [DES] 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. 2020, 900, L33. [Google Scholar] [CrossRef]
  219. Palmese, A.; Bom, C.R.; Mucesh, S.; Hartley, W.G. 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. Astrophys. J. 2023, 943, 56. [Google Scholar] [CrossRef]
  220. Ballard, W. et al. [DESI] A Dark Siren Measurement of the Hubble Constant with the LIGO/Virgo Gravitational Wave Event GW190412 and DESI Galaxies. Res. Notes AAS 2023, 7, 250. [Google Scholar] [CrossRef]
  221. Kelly, P.L.; Rodney, S.; Treu, T.; Oguri, M.; Chen, W.; Zitrin, A.; Birrer, S.; Bonvin, V.; Dessart, L.; Diego, J.M.; et al. Constraints on the Hubble constant from supernova Refsdal’s reappearance. Science 2023, 380, abh1322. [Google Scholar] [CrossRef]
  222. Falco, E.E.; Gorenstein, M.V.; Shapiro, I.I. On model-dependent bounds on H (0) from gravitational images Application of Q0957+ 561A, B. Astrophys. J. 1985, 289, L1. [Google Scholar] [CrossRef]
  223. Wagner, J. Generalised model-independent characterisation of strong gravitational lenses IV: Formalism-intrinsic degeneracies. Astron. Astrophys. 2018, 620, A86. [Google Scholar] [CrossRef]
  224. Wagner, J. Generalised model-independent characterisation of strong gravitational lenses—VI. The origin of the formalism intrinsic degeneracies and their influence on H0. Mon. Not. Roy. Astron. Soc. 2019, 487, 4492–4503. [Google Scholar] [CrossRef]
  225. Li, X.; Liao, K. Determining Cosmological-model-independent H0 with Gravitationally Lensed Supernova Refsdal. arXiv 2024, arXiv:2401.12052. [Google Scholar]
  226. Krishnan, C.; Ó Colgáin, E.; Ruchika; Sen, A.A.; Sheikh-Jabbari, M.M.; Yang, T. Is there an early Universe solution to Hubble tension? Phys. Rev. D 2020, 102, 103525. [Google Scholar] [CrossRef]
  227. Dainotti, M.G.; Simone, B.D.; Schiavone, T.; Montani, G.; Rinaldi, E.; Lambiase, G. On the Hubble constant tension in the SNe Ia Pantheon sample. Astrophys. J. 2021, 912, 150. [Google Scholar] [CrossRef]
  228. Dainotti, M.G.; Simone, B.D.; Schiavone, T.; Montani, G.; Rinaldi, E.; Lambiase, G.; Bogdan, M.; Ugale, S. 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 2022, 10, 24. [Google Scholar] [CrossRef]
  229. Hu, J.P.; Wang, F.Y. Revealing the late-time transition of H0: Relieve the Hubble crisis. Mon. Not. Roy. Astron. Soc. 2022, 517, 576–581. [Google Scholar] [CrossRef]
  230. Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Solomon, R.; Bargiacchi, G.; Capozziello, S.; Dainotti, M.G.; Stojkovic, D. Revealing intrinsic flat ΛCDM biases with standardizable candles. Phys. Rev. D 2022, 106, L041301. [Google Scholar] [CrossRef]
  231. Jia, X.D.; Hu, J.P.; Wang, F.Y. Evidence of a decreasing trend for the Hubble constant. Astron. Astrophys. 2023, 674, A45. [Google Scholar] [CrossRef]
  232. Dainotti, M.; Simone, B.D.; Montani, G.; Schiavone, T.; Lambiase, G. The Hubble constant tension: Current status and future perspectives through new cosmological probes. arXiv 2023, arXiv:2301.10572. [Google Scholar]
  233. Wagner, J. Solving the Hubble tension à la Ellis & Stoeger 1987. PoS CORFU 2023, 2022, 267. [Google Scholar] [CrossRef]
  234. Phillips, M.M. The absolute magnitudes of Type IA supernovae. Astrophys. J. Lett. 1993, 413, L105–L108. [Google Scholar] [CrossRef]
  235. Risaliti, G.; Lusso, E. A Hubble Diagram for Quasars. Astrophys. J. 2015, 815, 33. [Google Scholar] [CrossRef]
  236. Risaliti, G.; Lusso, E. Cosmological constraints from the Hubble diagram of quasars at high redshifts. Nat. Astron. 2019, 3, 272–277. [Google Scholar] [CrossRef]
  237. Lusso, E.; Risaliti, G.; Nardini, E.; Bargiacchi, G.; Benetti, M.; Bisogni, S.; Capozziello, S.; Civano, F.; Eggleston, L.; Elvis, M.; et al. Quasars as standard candles III. Validation of a new sample for cosmological studies. Astron. Astrophys. 2020, 642, A150. [Google Scholar] [CrossRef]
  238. Yang, T.; Banerjee, A.; Ó Colgáin, E. Cosmography and flat ΛCDM tensions at high redshift. Phys. Rev. D 2020, 102, 123532. [Google Scholar] [CrossRef]
  239. Scolnic, D.M. et al. [Pan-STARRS1] The Complete Light-curve Sample of Spectroscopically Confirmed SNe Ia from Pan-STARRS1 and Cosmological Constraints from the Combined Pantheon Sample. Astrophys. J. 2018, 859, 101. [Google Scholar] [CrossRef]
  240. Malekjani, M.; Conville, R.M.; Ó Colgáin, E.; Pourojaghi, S.; Sheikh-Jabbari, M.M. On redshift evolution and negative dark energy density in Pantheon + Supernovae. Eur. Phys. J. C 2024, 84, 317. [Google Scholar] [CrossRef]
  241. Ó Colgáin, E.; Pourojaghi, S.; Sheikh-Jabbari, M.M. Implications of DES 5YR SNe Dataset for ΛCDM. arXiv 2024, arXiv:2406.06389. [Google Scholar]
  242. Pourojaghi, S.; Zabihi, N.F.; Malekjani, M. Can high-redshift Hubble diagrams rule out the standard model of cosmology in the context of cosmography? Phys. Rev. D 2022, 106, 123523. [Google Scholar] [CrossRef]
  243. Pastén, E.; Cárdenas, V. Testing ΛCDM cosmology in a binned universe: Anomalies in the deceleration parameter. arXiv 2023, arXiv:2301.10740. [Google Scholar] [CrossRef]
  244. Khadka, N.; Ratra, B. Using quasar X-ray and UV flux measurements to constrain cosmological model parameters. Mon. Not. Roy. Astron. Soc. 2020, 497, 263–278. [Google Scholar] [CrossRef]
  245. Khadka, N.; Ratra, B. Determining the range of validity of quasar X-ray and UV flux measurements for constraining cosmological model parameters. Mon. Not. Roy. Astron. Soc. 2021, 502, 6140–6156. [Google Scholar] [CrossRef]
  246. Dainotti, M.G.; Bardiacchi, G.; Lenart, A.L.; Capozziello, S.; Ó Colgáin, E.; Solomon, R.; Stojkovic, D.; Sheikh-Jabbari, M.M. Quasar Standardization: Overcoming Selection Biases and Redshift Evolution. Astrophys. J. 2022, 931, 106. [Google Scholar] [CrossRef]
  247. Singal, J.; Mutchnick, S.; Petrosian, V. The X-Ray Luminosity Function Evolution of Quasars and the Correlation between the X-Ray and Ultraviolet Luminosities. Astrophys. J. 2022, 932, 111. [Google Scholar] [CrossRef]
  248. Petrosian, V.; Singal, J.; Mutchnick, S. Can the Distance-Redshift Relation be Determined from Correlations between Luminosities? Astrophys. J. Lett. 2022, 935, L19. [Google Scholar] [CrossRef]
  249. Zajaček, M.; Czerny, B.; Khadka, N.; Martínez-Aldama, M.L.; Prince, R.; Panda, S.; Ratra, B. Effect of extinction on quasar luminosity distances determined from UV and X-ray flux measurements. arXiv 2023, arXiv:2305.08179. [Google Scholar] [CrossRef]
  250. Adams, N.J.; Conselice, C.J.; Ferreira, L.; Austin, D.; Trussler, J.A.A.; Juodžbalis, I.; Vijayan, A.P. Discovery and properties of ultra-high redshift galaxies (9 < z < 12) in the JWST ERO SMACS 0723 Field. Mon. Not. Roy. Astron. Soc. 2022, 518, 4755. [Google Scholar]
  251. Labbe, I.; van Dokkum, P.; Nelson, E.; Bezanson, R.; Suess, K.A.; Leja, J.; Brammer, G.; Whitaker, K.; Mathews, E.; Stefanon, M.; et al. A population of red candidate massive galaxies ~600 Myr after the Big Bang. Nature 2023, 616, 266–269. [Google Scholar] [CrossRef]
  252. Castellano, M.; Fontana, A.; Treu, T.; Santini, P.; Merlin, E.; Leethochawalit, N.; Yang, L. Early results from GLASS-JWST. III. Galaxy candidates at z 9–15. Astrophys. J. Lett. 2022, 938, L15. [Google Scholar] [CrossRef]
  253. Naidu, R.P.; Oesch, P.A.; van Dokkum, P.; Nelson, E.J.; Suess, K.A.; Brammer, G.; Weibel, A. Two remarkably luminous galaxy candidates at z ≈ 10–12 revealed by JWST. Astrophys. J. Lett. 2022, 940, L14. [Google Scholar] [CrossRef]
  254. Xiao, M.; Oesch, P.; Elbaz, D.; Bing, L.; Nelson, E.; Weibel, A.; Wyithe, J.S. Massive Optically Dark Galaxies Unveiled by JWST Challenge Galaxy Formation Models. arXiv 2023, arXiv:2309.02492. [Google Scholar]
  255. Boylan-Kolchin, M. Stress testing ΛCDM with high-redshift galaxy candidates. Nat. Astron. 2023, 7, 731–735. [Google Scholar] [CrossRef]
  256. Shanks, T.; Hogarth, L.; Metcalfe, N. Gaia Cepheid parallaxes and ‘Local Hole’ relieve H0 tension. Mon. Not. Roy. Astron. Soc. 2019, 484, L64–L68. [Google Scholar] [CrossRef]
  257. Kenworthy, W.D.; Scolnic, D.; Riess, A. The Local Perspective on the Hubble Tension: Local Structure Does Not Impact Measurement of the Hubble Constant. Astrophys. J. 2019, 875, 145. [Google Scholar] [CrossRef]
  258. Haslbauer, M.; Banik, I.; Kroupa, P. The KBC void and Hubble tension contradict ΛCDM on a Gpc scale—Milgromian dynamics as a possible solution. Mon. Not. Roy. Astron. Soc. 2020, 499, 2845–2883. [Google Scholar] [CrossRef]
  259. Cai, R.G.; Ding, J.F.; Guo, Z.K.; Wang, S.J.; Yu, W.W. Do the observational data favor a local void? Phys. Rev. D 2021, 103, 123539. [Google Scholar] [CrossRef]
  260. Camarena, D.; Marra, V.; Sakr, Z.; Clarkson, C. A void in the Hubble tension? The end of the line for the Hubble bubble. Class. Quant. Grav. 2022, 39, 184001. [Google Scholar] [CrossRef]
  261. Mazurenko, S.; Banik, I.; Kroupa, P.; Haslbauer, M. A simultaneous solution to the Hubble tension and observed bulk flow within 250 h−1 Mpc. Mon. Not. Roy. Astron. Soc. 2024, 527, 4388–4396. [Google Scholar] [CrossRef]
  262. Ding, Q.; Nakama, T.; Wang, Y. A gigaparsec-scale local void and the Hubble tension. Sci. China Phys. Mech. Astron. 2020, 63, 290403. [Google Scholar] [CrossRef]
  263. Haslbauer, M.; Kroupa, P.; Jerabkova, T. The cosmological star formation history from the Local Cosmological Volume of galaxies and constraints on the matter homogeneity. Mon. Not. Roy. Astron. Soc. 2023, 524, 3252–3262. [Google Scholar] [CrossRef]
  264. Gómez-Valent, A. Fast test to assess the impact of marginalization in Monte Carlo analyses and its application to cosmology. Phys. Rev. D 2022, 106, 063506. [Google Scholar] [CrossRef]
  265. Ó Colgáin, E.; Pourojaghi, S.; Sheikh-Jabbari, M.M.; Sherwin, D. MCMC Marginalisation Bias and ΛCDM tensions. arXiv 2023, arXiv:2307.16349. [Google Scholar]
  266. Herold, L.; Ferreira, E.G.M.; Komatsu, E. New Constraint on Early Dark Energy from Planck and BOSS Data Using the Profile Likelihood. Astrophys. J. Lett. 2022, 929, L16. [Google Scholar] [CrossRef]
  267. Holm, E.B.; Herold, L.; Simon, T.; Ferreira, E.G.M.; Hannestad, S.; Poulin, V.; Tram, T. Bayesian and frequentist investigation of prior effects in EFT of LSS analyses of full-shape BOSS and eBOSS data. Phys. Rev. D 2023, 108, 123514. [Google Scholar] [CrossRef]
  268. Holm, E.B.; Nygaard, A.; Dakin, J.; Hannestad, S.; Tram, T. PROSPECT: A profile likelihood code for frequentist cosmological parameter inference. arXiv 2023, arXiv:2312.02972. [Google Scholar]
  269. Kroupa, P.; Gjergo, E.; Asencio, E.; Haslbauer, M.; Pflamm-Altenburg, J.; Wittenburg, N.; Samaras, N.; Thies, I.; Oehm, W. The many tensions with dark-matter based models and implications on the nature of the Universe. PoS CORFU 2023, 2022, 231. [Google Scholar] [CrossRef]
  270. Wang, B.; Abdalla, E.; Atrio-Barandela, F.; Pavon, D. Dark Matter and Dark Energy Interactions: Theoretical Challenges, Cosmological Implications and Observational Signatures. Rept. Prog. Phys. 2016, 79, 096901. [Google Scholar] [CrossRef]
  271. Valentino, E.D.; Melchiorri, A.; Mena, O. Can interacting dark energy solve the H0 tension? Phys. Rev. D 2017, 96, 043503. [Google Scholar] [CrossRef]
  272. Valentino, E.D.; Melchiorri, A.; Mena, O.; Vagnozzi, S. Interacting dark energy in the early 2020s: A promising solution to the H0 and cosmic shear tensions. Phys. Dark Univ. 2020, 30, 100666. [Google Scholar] [CrossRef]
  273. Gómez-Valent, A.; Pettorino, V.; Amendola, L. Update on coupled dark energy and the H0 tension. Phys. Rev. D 2020, 101, 123513. [Google Scholar] [CrossRef]
  274. Wang, B.; Abdalla, E.; Atrio-Barandela, F.; Pavón, D. Further understanding the interaction between dark energy and dark matter: Current status and future directions. arXiv 2024, arXiv:2402.00819. [Google Scholar] [CrossRef]
  275. Akarsu, Ö.; Barrow, J.D.; Escamilla, L.A.; Vazquez, J.A. Graduated dark energy: Observational hints of a spontaneous sign switch in the cosmological constant. Phys. Rev. D 2020, 101, 063528. [Google Scholar] [CrossRef]
  276. Akarsu, Ö.; Kumar, S.; Özülker, E.; Vazquez, J.A. Relaxing cosmological tensions with a sign switching cosmological constant. Phys. Rev. D 2021, 104, 123512. [Google Scholar] [CrossRef]
  277. Akarsu, O.; Kumar, S.; Özülker, E.; Vazquez, J.A.; Yadav, A. Relaxing cosmological tensions with a sign switching cosmological constant: Improved results with Planck, BAO, and Pantheon data. Phys. Rev. D 2023, 108, 023513. [Google Scholar] [CrossRef]
  278. Akarsu, O.; Valentino, E.D.; Kumar, S.; Nunes, R.C.; Vazquez, J.A.; Yadav, A. ΛsCDM model: A promising scenario for alleviation of cosmological tensions. arXiv 2023, arXiv:2307.10899. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Akarsu, Ö.; Ó Colgáin, E.; Sen, A.A.; Sheikh-Jabbari, M.M. ΛCDM Tensions: Localising Missing Physics through Consistency Checks. Universe 2024, 10, 305. https://doi.org/10.3390/universe10080305

AMA Style

Akarsu Ö, Ó Colgáin E, Sen AA, Sheikh-Jabbari MM. ΛCDM Tensions: Localising Missing Physics through Consistency Checks. Universe. 2024; 10(8):305. https://doi.org/10.3390/universe10080305

Chicago/Turabian Style

Akarsu, Özgür, Eoin Ó Colgáin, Anjan A. Sen, and M. M. Sheikh-Jabbari. 2024. "ΛCDM Tensions: Localising Missing Physics through Consistency Checks" Universe 10, no. 8: 305. https://doi.org/10.3390/universe10080305

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop