Next Article in Journal
Damping Scenarios of Kink Oscillations of Solar Coronal Loops
Next Article in Special Issue
Search for DA White Dwarf Binary Candidates from LAMOST DR7
Previous Article in Journal
Geminga SNR: Possible Candidate of Local Cosmic-Ray Factory (II)
Previous Article in Special Issue
Jet Cloud–Star Interaction as an Interpretation of Neutrino Outburst from the Blazar TXS 0506+056
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hubble Tension: The Evidence of New Physics

1
School of Astronomy and Space Science, Nanjing University, Nanjing 210093, China
2
Key Laboratory of Modern Astronomy and Astrophysics (Nanjing University), Ministry of Education, Nanjing 210093, China
*
Author to whom correspondence should be addressed.
Universe 2023, 9(2), 94; https://doi.org/10.3390/universe9020094
Submission received: 13 January 2023 / Revised: 21 January 2023 / Accepted: 6 February 2023 / Published: 10 February 2023
(This article belongs to the Special Issue Advances in Astrophysics and Cosmology – in Memory of Prof. Tan Lu)

Abstract

:
The Λ CDM model provides a good fit to most astronomical observations but harbors large areas of phenomenology and ignorance. With the improvements in the precision and number of observations, discrepancies between key cosmological parameters of this model have emerged. Among them, the most notable tension is the 4 σ to 6 σ deviation between the Hubble constant ( H 0 ) estimations measured by the local distance ladder and the cosmic microwave background (CMB) measurement. In this review, we revisit the H 0 tension based on the latest research and sort out evidence from solutions to this tension that might imply new physics beyond the Λ CDM model. The evidence leans more towards modifying the late-time universe.

1. Introduction

The cosmological constant ( Λ ) cold dark matter model ( Λ CDM) is the simplest cosmological model and consistent with the most astronomical observations [1,2,3,4,5,6,7,8,9,10,11,12,13,14]. In the past 30 years, it has been the best model, and no better one has yet been presented to replace it. Despite its remarkable successes, the validity of the Λ CDM model is currently under intense investigation [13,15,16,17] (for reviews, see [18,19,20,21,22,23]). The fine tuning and coincidence problems are the most important theoretical difficulties [24,25,26,27,28]. The fundamental problem with the former is that there is a large discrepancy between the observations and theoretical expectations of Λ [24,26,29,30]. The latter is related to the observed vacuum energy density Ω Λ and the matter energy density Ω m , which are now nearly equal despite their dramatically different evolutionary properties. The anthropic principle, as a possible solution to these problems, states that these “coincidences” results from a selection bias towards the existence of human life in the context of a multiverse [31,32]. In addition to the above theoretical challenges, the two main components in the Λ CDM model, dark matter (DM) and dark energy (DE), are poorly understood. Moreover, there are also some tensions between the cosmological and astrophysical observations and the Λ CDM model, which include the Hubble tension [3,33,34,35,36,37,38,39,40] (5 σ ), growth tension [41,42,43] (2–3 σ ), CMB anisotropy anomalies [22,44,45,46,47,48,49,50,51,52,53] (2–3 σ ), cosmic dipoles [23,54,55,56,57,58,59,60,61,62,63] (2–5 σ ), Baryon Acoustic Oscillation (BAO) curiosities [64,65,66] (2.5–3 σ ), parity violating rotation of CMB linear polarization [67,68,69,70], small-scale curiosities [71,72,73,74], age of the universe [75], the Lithium problem [76] (2–4 σ ), the quasar Hubble diagram [77,78,79,80,81] (∼4 σ ), oscillating signals in short range gravity experiments [82,83], anomalously low baryon temperature [84] (∼3.8 σ ), colliding clusters with high velocity [85,86] (∼6 σ ), etc. More detailed information of these tensions can be found from Perivolaropoulos and Skara [21]. The H 0 tension emerged with the first release of Planck results [87], and has grown in significance in the past few years [3,33,34,35,36,37,38,39,40].
There have been many studies dedicated to finding out what causes the Hubble tension, but so far there is no convincing explanation. The reanalyses of the Planck observations and Hubble Space Telescope (HST) measurements demonstrate that the serious discrepancy of H 0 may not be caused by systematics (including photometric biases, environmental effects, calibration error, lens mass modelling biases, CMB foreground effects and so on) [35,40,46,88,89,90,91,92,93,94]. Hence, many researchers prefer to believe that the H 0 tension might be caused by new physics beyond the Λ CDM model [36]. So far, there has been a large number of modified models which adopted to resolve or relieve the H 0 tension, see [19,21] for a review of H 0 tension solutions. Although many extensions of Λ CDM can alleviate the H 0 tension, none are supported by the majority of observations [95,96]. There are many international conferences held on the H 0 tension, such as “Beyond Λ CDM”, “Hubble Tension Headache”, “Tensions between the Early and the Late Universe”, etc. [97]. “Beyond Λ CDM” in Oslo in 2015, a poll at the conference showed that 69% of participants believe that new physics is the most likely explanation. On the contrary, more than 50% of the participants of the “Hubble Tension Headache” conference held by the University of Southampton supported the explanation that there were still systematics unknown to us in the observational data. Theoretical Physics workshop (“Tensions between the Early and the Late Universe”) in July 2019 directed attention to the Hubble constant discrepancy. The workshop pointed out that “New results showed that it does not appear to depend on the use of any one method, team or source” [98]. To streamline the interaction between these different communities and promote the transparent transfer of information, participants also gave some reasonable suggestions. More details about the workshop are available at website: https://www.kitp.ucsb.edu/activities/enervac-c19. In a word, the research and discussion on the H 0 tension is still going on.
This review is organized as follows. In Section 2, we briefly introduce the H 0 tension, and then detail two methods for constraining H 0 , the CMB measurements and the local distance ladder. In Section 3, we discuss recent methods for estimating H 0 using otherwise independent observations. We also discuss a taxonomy of solutions to the H 0 tension in Section 4. Afterwards, from all the solutions to the H 0 tension, we sort out evidence that might imply new physics beyond the Λ CDM model and briefly introduce some of the proposed scenarios in Section 5. Finally, we give a brief conclusion and future prospects.

2. H 0 Tension

The local expansion rate of the universe H 0 is a fundamental value. It also determines the age of the universe; thus, it is important to determine it accurately. The accuracy of H 0 measurements has been improved as the number of probes has increased. A review of most well-established probes can be found in literature [99]. In general, the H 0 measurements can be estimated from the cosmological model utilizing the early universe measurements, or more directly measured from the local universe. Interestingly, the H 0 values measured by these two approaches are inconsistent. The discrepancy between the H 0 values measured from the local distance ladder and from the CMB is the most serious challenge to the standard Λ CDM model. This H 0 discrepancy is also known as “Hubble Tension” [87].

2.1. Constraints of H 0 from the CMB Measurements

The estimation of H 0 from the CMB data proceeds in three steps [100,101]: (1) determine the baryon density and matter density to allow for the calculation of the sound horizon size ( r s ); (2) infer θ s from the spacing between the acoustic peaks to determine the comoving angular diameter distance to last scattering D A = r s / θ s ; (3) adjust the only remaining free density parameter in the model that D A gives this inferred distance. With this last step complete, we now have H ( z ) determined for all z, including H 0 (z = 0).
In 2013, Bennett et al. [100] provided H 0 = 70.00 ± 2.20 km/s/Mpc through analysing the nine-year Wilkinson Microwave Anisotropy Probe (WMAP) observations. Meanwhile, the first data release of the Planck space observatory, which was operated by the European Space Agency (ESA), gave a precise result H 0 = 67.40 ± 1.40 km/s/Mpc [87]. After that, a more accurate H 0 = 67.40 ± 0.50 km/s/Mpc yielded by the Planck final data release is also in line with the Planck2013 results [3]. Researchers also consider adding the other observational data to constrain H 0 , Planck2018+lensing 67.36 ± 0.54 km/s/Mpc and Planck2018+lensing+BAO 67.66 ± 0.42 km/s/Mpc [3]. The main H 0 measurements are shown in Figure 1. In addition, there exists a lot of H 0 predictions adopting other CMB experiments from the ground, including the South Pole Telescope (SPTPol) [66,102] and Atacama Cosmology Telescope (ACT) [103]. These H 0 predictions are all consistent with the Planck2018 result [3].

2.2. Constrain H 0 from the Local Distance Ladder

At present, there has been a lot of methods in the local universe to estimate the Hubble constant based on the distance-redshift relation. These methods are usually undertaken by building a “local distance ladder”; the most common approach is to adopt geometry to calibrate the luminosity of specific star types. Cepheid variables are often employed to determine the distance of 10–40 Mpc [104,105]. Measuring longer distances requires other standard candles, for example SNe Ia, whose maximum redshift can reach 2.36 [2,7]. Quasars [106,107,108,109,110,111,112,113,114,115] and gamma radio bursts (GRBs) [5,10,14,116,117,118,119,120,121,122,123,124,125,126,127,128] offer the prospect of extending the Hubble diagram up to higher redshifts.
The first H 0 estimation, using the Cepheid variables and SNe Ia provided by the HST project, was 72 ± 8 km/s/Mpc [104]. A improved result, 74.3 ± 2.1 km/s/Mpc, was yielded by using a modified distance calibration [105]. After that, the SH0ES project, which started in 2005, also produces many H 0 results [34,35,39,129]. Based on the SH0ES data, H 0 results from numerous reanalyses using different formalisms, statistical methods of inference, or replacement of parts of the data-set are both in line with the previous results [130,131,132,133,134,135,136]. The latest H 0 result provided by the SH0ES collaboration shows that H 0 = 73.04 ± 1.04 km/s/Mpc [40].
During the last few decades there has been remarkable progress in measuring the Hubble constant. The available technology and measurement methods determine the accuracy of this quantity. The main H 0 results up to 2022 obtained from the CMB measurements and the local distance ladder, as a function of the publication year, is shown in Figure 1. The uncertainties in these values have been decreasing for both methods and the recent measurements disagree by 5 σ [40]. There is an obvious H 0 tension without physical explanations.
Figure 1. H 0 measurements from the early-time and the late-time observations. Blue points denote H 0 estimations from analyses of CMB data, including first year WMAP [137], three year WMAP [138], five year WMAP [139], seven-year WMAP [140], nine-year WMAP [100], Planck 2013 [87], Planck 2015 [33], Planck 2018 [3] and BAO [65]. Red points denote H 0 values measured by utilizing the local distance ladder including the Cepheid distance scale [104], Carnegie Hubble Program [105] and SH0ES [34,35,39,40,129,141,142]. (Source: Figure 12 in Perivolaropoulos and Skara [21]).
Figure 1. H 0 measurements from the early-time and the late-time observations. Blue points denote H 0 estimations from analyses of CMB data, including first year WMAP [137], three year WMAP [138], five year WMAP [139], seven-year WMAP [140], nine-year WMAP [100], Planck 2013 [87], Planck 2015 [33], Planck 2018 [3] and BAO [65]. Red points denote H 0 values measured by utilizing the local distance ladder including the Cepheid distance scale [104], Carnegie Hubble Program [105] and SH0ES [34,35,39,40,129,141,142]. (Source: Figure 12 in Perivolaropoulos and Skara [21]).
Universe 09 00094 g001

3. H 0 Arbitration

The Planck CMB measurements and the local ladder measurements both provide very precise constraints on cosmological parameters. However, as with any experimental measurement, they are not free from systematic errors. Hence, possible systematics in the Planck observations and the HST measurements are suspected to be responsible for the Hubble discrepancy in the early and late measurements. However, this possibility has been largely ruled out [35,40,46,88,89,90,91,92,94]. Since reanalyses of the Planck and local ladder measurements could not find a satisfactory answer to the H 0 tension, hope was pinned on other observations, such as quasar lensing [37,143], Megamaser [144,145,146,147,148], gravitational waves (GW) [149,150,151], fast radio bursts (FRBs) [152,153,154,155,156], tip of the red giant branch (TRGs) [157,158,159], BAOs [4,65], Type II supernovae [94], Ages of Old Astrophysical Objects [160,161], etc. [162]. These observations do not assume a cosmological model and are independent of the CMB and distance ladder measurements.

3.1. Quasar Lensing

Quasar lensing as an independent method can be used to estimate H 0 . With observed time delay Δ τ o b s and lens mass model, H 0 can be inferred. The observed time delay is owing to the geometrical path length difference caused by the gravitational potential of the lens, which is associated with the path of the light rays from the vicinity of the lens to the observer [163]. The time delay distance D Δ t inferred from Δ τ o b s is actually a combination of angular diameter distances [37]:
D Δ t ( 1 + z l e n s ) D d D s D d s ,
where z l e n s is the lens redshift, D d is the angular diameter distance to the lens, D s is the angular diameter distance to the source, and D d s is the angular diameter distance between the source and the lens. In the flat Λ CDM model, D d s can be given as [164]
D d s = c H 0 ( 1 + z s ) z l e n s z s ( Ω m ( 1 + z ) 3 + Ω Λ ) 1 / 2 d z ,
here, z s is the redshift of source. Substitute z l e n s and z s into the following formula:
D i = c H 0 ( 1 + z i ) 0 z i Ω m ( 1 + z ) 3 + Ω Λ d z ,
the expressions of D d and D s can be obtained. The time delay distance is primarily sensitive to H 0 , with weak dependence on other cosmological parameters. There were seven systems which provided the estimations of H 0 , as shown in the following Table 1. In addition, Wong et al. [37] and Millon et al. [143] utilize the multiple lens systems to constrain H 0 ; the former provides a result of 73 . 3 1.8 + 1.7 km/s/Mpc and the latter provides a result of 74 . 0 1.8 + 1.7 km/s/Mpc. Recently, Shajib et al. [165] updated the H 0 measurement ( 78 . 3 3.3 + 3.4 km/s/Mpc) derived from the lens RXJ1131-1231, and provided a new H 0 measurement ( 74 . 2 1.6 + 1.6 km/s/Mpc) for seven lenses. Their new measurement is in excellent agreement with those obtained in the past using standard simply parametrized mass profiles.

3.2. Megamaser

Water megamasers residing in the accretion disks around supermassive black holes (SMBHs) in active galactic nuclei (AGNs) provide a unique way to bypass the distance ladder and make one-step, geometric distance measurements to their host galaxies [148]. The archetypal AGN accretion disk megamaser system is located in the nearby (7.6 Mpc) [147,175] Seyfert 2 galaxy NGC 4258 [176,177,178]. The Megamaser Cosmology Project (MCP) is a multi-year campaign to find, monitor, and map AGN accretion disk megamaser systems [179,180]. The primary goal of the MCP is to constrain H 0 to a precision of several percent through making geometric distance measurements to megamaser galaxies in the Hubble flow [144,145,146,181]. Distance measurements using the megamaser technique do not rely on the CMB or distance ladders measurements. Therefore, megamaser measurements provide an independent handle on H 0 estimates.
The MCP has so far published distances to six galaxies including UGC 3789 [145], NGC 6264 [144], NGC 6323 [146], NGC 5765b [182], NGC 4258 [147] and CGCG 074-064 [183]. The latest details of these six galaxies are shown in Table 2. In 2018, Braatz et al. [184] updated the published value to the UGC 3789 and obtained the H 0 estimation by using the first four galaxies, H 0 = 69.3 ± 4.2 km/s/Mpc. Recently, Pesce et al. [148] applied an improved approach for fitting maser data and obtained better distance estimates for the first four galaxies. Combining all the distance measurements of galaxies and assuming a fixed velocity uncertainty of 250 km/s in connection with peculiar motions, they provided the H 0 estimation, H 0 = 73.9 ± 3.0 km/s/Mpc, independent of the CMB and distance ladders measurements.

3.3. Gravitational Wave

On 17 August 2017, the Advanced Laser Interferometer Gravitational-wave Observatory (LIGO) [185] and Virgo [186] detectors observed GW170817, a strong signal from the merger of a binary neutron-star system. Less than 2 s after the merger, a gamma-ray burst (GRB 170817A) was detected within a region of the sky consistent with the LIGO-Virgo-derived location of the GW source [187,188,189]. The detection of GW170817 in both gravitational waves and the electromagnetic (EM) waves heralds the age of the gravitational-wave multi-messenger astronomy.
With the luminosity distance fitted from the GW waveform and the redshift information from the host galaxy, GW can be treated as standard sirens to conduct research in cosmology. The GW amplitude depends on the chirp mass and luminosity distance of the GW source. The mass can be precisely determined by the phase measurement of the GW signal. Therefore, as long as the amplitude and phase information of the GW source are obtained at the same time, the corresponding luminosity distance can be given. The heliocentric redshift measurement, z h e l i o = 0.009783, was obtained from the optical identification of the host galaxy NGC 4993 [190]. The joint analysis of the GW signal from GW170817 and its EM localization led to the first H 0 estimation, H 0 = 74 8 + 16 km/s/Mpc (median and symmetric 68% credible interval) [149]. In this analysis, the degeneracy in the GW signal between the source distance and the observing angle dominated the uncertainty of the H 0 measurement. Tight constraints on the observing angle using high angular resolution imaging of the radio counterpart of GW170817 have been obtained [150]. Hotokezaka et al. [151] reported an improved measurement of H 0 = 70 . 30 5.0 + 5.3 km/s/Mpc by using these new radio observations, combined with the previous GW and EM data. Recently, using 47 GW sources from the Third LIGO-Virgo-KAGRA Gravitational-Wave Transient Catalog (GWTC-3), The LIGO Scientific Collaboration et al. [191] presented H 0 = 68 7 + 12 km/s/Mpc (68% credible interval) when combined with the H 0 measurement from GW170817 and its EM counterpart. Moreover, combining the GWTC-3 with the H 0 measurement from GW170817, Mukherjee et al. [192] provided a more compact H 0 result of 67 3.8 + 6.3 km/s/Mpc.

3.4. Fast Radio Burst

FRBs are millisecond-duration pulses occurring at cosmological distances [193,194,195]. The total dispersion measure ( D M o b s ) of FRBs can provide a distance estimation to the source. The expanded form of D M o b s is as follows:
D M o b s ( z ) = D M M W + D M I G M ( z ) + D M h o s t ( z ) 1 + z ,
where D M I G M represents the contribution of the intergalactic medium (IGM), D M M W are contributed by the interstellar medium (ISM) and the halo of the Milky Way, and the D M h o s t is contributed by the host galaxy. Considering a flat universe, the averaged value of DM IGM is [196]
DM IGM ( z ) = A Ω b H 0 2 H 0 0 z FRB f IGM ( z ) f e ( z ) ( 1 + z ) Ω m ( 1 + z ) 3 + 1 Ω m d z ,
where A = 3 c 8 π G m p and m p is the proton mass. The electron fraction is f e ( z ) = Y H X e , H ( z ) + 1 2 Y H e X e , H e ( z ) , with hydrogen fraction Y H = 0.75 and helium fraction Y H e = 0.25 . Hydrogen and helium are completely ionized at z < 3 , which implies the ionization fractions of intergalactic hydrogen and helium X e , H = X e , H e = 1 . The cosmological parameters Ω m and Ω b are the the density of matter and the density of baryons, respectively. At present, there is no observation that can provide the evolution of the fraction of baryon in the IGM f IGM with redshift. Shull et al. [197] provided an estimation of f IGM 0.83 [153]. Then, the dispersion measure-redshift relation allows FRBs to be used as cosmological probes. However, the degeneracy between D M I G M and D M h o s t is the main obstacle for the cosmological application of FRBs. A reasonable method is to consider the probability distributions of D M I G M [198,199] and D M h o s t [200].
Hagstotz et al. [152] presented the first H 0 estimation, H 0 = 62.3 ± 9.1 km/s/Mpc, using the nine then available FRBs. Employing the probability distributions of D M I G M [199] and D M h o s t [200] from the IllustrisTNG simulation, a more compact result, H 0 = 68 . 81 4.33 + 4.99 km/s/Mpc, was given by Wu et al. [153] using eighteen localized FRBs. These two H 0 estimations seem to favor the smaller H 0 value. After that, James et al. [154] show H 0 = 73 8 + 12 km/s/Mpc employing an updated sample of 16 Australian Square Kilometre Array Pathfinder (ASKAP) FRBs detected by the Commensal Real-time ASKAP Fast Transients (CRAFT) Survey and localised to their host galaxies, and 60 unlocalised FRBs from Parkes and ASKAP. Compared to previous FRB-based estimates, uncertainties in FRB energetics and D M h o s t produce larger uncertainties in the inferred value of H 0 . Furthermore, Hagstotz et al. [152] performed a forecast using a realistic mock sample to demonstrate that a high-precision measurement of the expansion rate is possible without relying on other cosmological probes. Another H 0 measurement, H 0 = 70.60 ± 2.11 km/s/Mpc, was given by Liu et al. [155], employing a cosmological-model independent method. Recently, Zhao et al. [156] provided the first statistical H 0 measurements using unlocalized FRBs. They provided two H 0 measurements, H 0 = 71.7 7.4 + 8.8 km/s/Mpc and H 0 = 71.5 8.1 + 10.0 km/s/Mpc, which were obtained from the simulation-based case and the observation-based case, respectively. They also proposed that in the next few years, a 3% precision on the random error of the Hubble constant could be achieved using thousands of FRBs.

3.5. Tip of the Red Giant Branch

Compared to other standard candles, the tip of the red giant branch (TRGB) offers many advantages; for example, in the K-band they are ∼1.6 magnitudes brighter than Cepheids. They have nearly exhausted the hydrogen in their cores and have just begun helium burning. Employing parallax methods, their brightness can be standardized. They thus can be regarded as the standard candles visible in the local universe. Instead of the Cepheid, the TRGB can be used as calibrators of SNe Ia. Freedman et al. [157] gave the Hubble constant result H 0 = 69.8 ± 0.8 km/s/Mpc by measuring TRGB in nine SNe Ia hosts and calibrating TRGB in the large magellanic cloud. A consistent result, H 0 = 69.6 ± 0.8 km/s/Mpc, was also given by Freedman et al. [158] using their revised measurement of the Large Magellanic Cloud TRGB extinction. After that, Freedman [159] combined several recent calibrations of the TRGB method, and are internally self-consistent at the 1% level. The updated TRGB calibration applied to a sample of SNe Ia from the Carnegie Supernova Project results in a value of H 0 = 69.8 ± 0.8 km/s/Mpc.
We display H 0 measurements obtained from the recent observations in Table 3, and describe the correlation between the H 0 values and the published year in Figure 2. Combining Table 3 and Figure 2, it is easy to find that the measurements of the quasar lensing are tending to the SH0ES results as a whole. The H 0 measurements obtained from Megamaser, GW and FRB are more diffuse and have a large error which both covered the SH0ES results and CMB results. It is interesting that the H 0 values measured by TRGB have small errors and are between the SH0ES results and the CMB results. Collectively, these independent observations are currently unable to arbitrate the H 0 tension.

4. Solutions for the H 0 Tension

The case for an observational discrepancy between the early and late universe appears strong, is hard to dismiss, and merits an explanation. The analyses of possible systematics did not lead us to the cause of the Hubble tension [35,40,46,88,89,90,91,92,94]. The arbitration of H 0 given by otherwise independent observations also did not produce consistent results in favor of one side [37,143,149,151,152,153,159,184]. Hence, many researchers prefer to believe that the Hubble tension may be caused by new physics beyond the Λ CDM model [201]. Until now, a large number of theoretical solutions have been proposed to solve or relieve the H 0 tension. The detailed information can be found from these references [21,97,202].

4.1. Classification of Solutions to H 0 Tension

In the review article for H 0 tension, Di Valentino et al. [202] gave a detailed classification for the solutions of the H 0 tension. They divided all schemes proposed for solving the H 0 tension into 11 major categories with 123 subcategories, and some subcategories include several different schemes. The detailed classification results are as follows:
(a)
Early dark energy [203,204,205]:
(1) Anharmonic oscillations [206];
(2) Ultra-light axions [95,96,207,208,209,210,211];
(2.1) Dissipative axion [212];
(2.2) Axion interacting with a dilaton [213];
(3) Power-law potential [214];
(4) Rock ‘n’ roll [215];
(5) New early dark energy [216,217];
(6) Chain early dark energy [218];
(7) Anti-de Sitter phase [219,220];
(8) Gradusted dark energy [221];
(9) Acoustic dark energy [222,223];
(9.1) Exponential acoustic dark energy [223];
(10) Early dark energy in α attractors [224].
(b)
Late dark energy [225,226,227]:
(1) wCDM model [201,228,229,230];
(2) w 0 w a CDM or CPL parameterization [228,229,231];
(3) Dark energy in extended parameter spaces [232];
(4) Dynamical dark energy parameterizations with two free parameters [233,234];
(5) Dynamical dark energy parameterizations with one free parameter [235];
(6) Matastable dark energy [236,237,238];
(7) Phantom crossing [239];
(8) Late dark energy transition [240,241];
(9) Running vacuum model [242,243];
(10) Transitional dark energy model [244];
(11) Negative dark energy [245];
(12) Bulk viscous models [246,247,248];
(13) Holographic dark energy [249,250,251,252];
(13.1) Tsallis holographic dark energy [253];
(14) Swampland conjectures [254,255,256,257,258];
(15) Late time transitions in the quintessence field [259,260];
(16) Phantom braneworld dark energy [261];
(17) Frame-dependent dark energy [262];
(18) Chameleon dark energy [263,263,264].
(c)
Dark energy models with degrees of freedom and their extensions:
(1) Phenomenologically emergent dark energy [265,266,267];
(1.1) Generalized emergent dark energy [268,269];
(1.2) Modified emergent dark energy [270];
(2) Vacuum metamorphosis [271,272,273];
(2.1) Elaborated vacuum metamorphosis [271,272,273].
(d)
Models with extra relativistic degrees of freedom:
(1) Sterile neutrinos [274,275];
(2) Neutrino asymmetries [276];
(3) Thermal axions [277,278];
(4) Decaying dark matter [279,280,281,282,283,284];
(4.1) Self-interacting dark matter [285,286];
(4.2) Two-body dark matter decays [287,288];
(4.3) Light gravitino scenarios [289,290,291];
(4.4) Decaying ‘Z’ [292];
(4.5) Dynamical dark matter [293];
(4.6) Degenerate decaying fermion dark matter [294];
(5) Neutrino–dark matter interactions [295,296,297];
(5.1) Neutrino–Majoron interactions [298,299,300];
(5.2) Feebly interacting massive particles (FIMPs) decay into neutrinos [301];
(6) Interacting dark radiation [301];
(7) Coupled DM–dark radiation scenarios [302,303];
(8) Cannibal dark matter [304];
(9) Decaying ultralight scalar [304,305];
(10) Ultralight dark photon [306,307];
(11) Primordial black holes [308,309,310];
(12) Unparticles [311].
(e)
Models with extra interactions:
(1) Interacting dark energy (IDE) [312];
(1.1) Interacting vacuum energy [313,314,315,316,317];
(1.2) Coupled scalar field [318];
(1.3) IDE with a constant DE equation of state [319,320,321,322];
(1.4) IDE with variable DE equation of state [273,323];
(1.5) Interacting vacuum scenario and IDE with variable coupling [321,324];
(1.6) IDE with sign-changing interaction [325];
(1.7) Anisotropic stress in IDE [326];
(1.8) Interaction in the anisotropic universe [327];
(1.9) Metastable interacting dark energy [236,238];
(1.10) Quantum field cosmology [316,328];
(1.11) Interacting quintom dark energy [329];
(2) Interacting dark matter [330,331];
(2.1) DM–photon coupling [332,333];
(2.2) DM–baryon coupling [334,335];
(3) DE–baryon coupling [336,337];
(4) Interacting neutrinos [338,339,340];
(4.1) Self-interacting neutrinos: [338,339,341,342,343];
(4.2) Self-interacting sterile neutrino model [344];
(4.3) Dark neutrino interactions [345,346].
(f)
Unified cosmologies:
(1) Generalized Chaplygin gas model [347];
(2) A new unified model [348];
(3) Λ (t)CDM model [349];
(4) Λ -gravity [350,351].
(g)
Modified gravity [352]:
(1) f( R ) gravity theory; [353,354,355,356];
(2) f( T ) gravity theory [357,358,359,360];
(3) f( T , B ) gravity theory [361,362];
(4) f( Q ) gravity theory [363,364];
(5) Jordan–Brans–Dicke (JBD) gravity [365];
(5.1) Brans and Dicke- Λ CDM [366,367];
(6) Scalar-tensor theories of gravity [368,369];
(6.1) Early modified gravity [370,371];
(6.2) Screened fifth forces [372,373];
(7) Űber-gravity [374,375];
(8) Galileon gravity [376,377];
(9) Nonlocal gravity [378,379];
(10) Unimodular gravity [380];
(11) Scale-dependent scenario of gravity [381];
(12) VCDM [382,383].
(h)
Inflationary models [384]:
(1) Exponential inflation [385,386];
(2) Reconstructed primordial power spectrum [387,388];
(3) Lorentzian quintessential inflation [389];
(4) Harrison–Zel’dovich spectrum [390].
(i)
Modified recombination history [391]:
(1) Effective electron rest mass [392,393];
(2) Time varying electron mass [394];
(3) Axi–Higgs model [395];
(4) Primordial magnetic fields [396,397].
(j)
Physics of the critical phenomena:
(1) Double- Λ CDM [398];
(2) Ginzburg–Landau theory of phase transition [321];
(3) Critically emergent dark energy [399].
(k)
Alternative proposals:
(1) Local inhomogeneity [400,401];
(2) Bianchi type I spacetime [402];
(3) Scaling solutions [403,404,405];
(4) CMB monopole temperature T 0 [406];
(4.1) Open and hotter universe [407,408];
(5) Super- Λ CDM [409];
(6) Heisenberg uncertainty [410]
(7) Diffusion [411];
(8) Casimir cosmology [412];
(9) Surface forces [413];
(10) Milne model [414,415];
(11) Running Hubble tension [416,417];
(12) Rapid transition in the effective gravitational constant [418];
(13) Causal horizons [419,420];
(14) Milgromian dynamics [421];
(15) Charged dark matter [422,423].
It can be observed that Di Valentino et al. [202]’s classification of schemes to alleviate the H 0 tension is very detailed. With help of this classification, we can quickly find out the solutions and corresponding articles we need. Recently, Perivolaropoulos and Skara [21] updated and optimized the classification scheme based on the latest research work. The new classification scheme is more concise than before. They divided all solutions into 5 major categories, each of which contained several sub-categories, for a total of 19 sub-categories. The detailed classification is as follows:
(i)
Late time deformations of the Hubble expansion rate H(z):
(1) Phantom dark energy [424];
(2) Running vacuum model [425];
(3) Phenomenologically emergent dark energy [426,427,428];
(4) Vacuum phase transition [429];
(5) Phase transition in dark energy [430].
(ii)
Deformations of the Hubble expansion rate H(z) with additional interactions/degrees of freedom:
(1) Interacting dark energy [431,432,433,434,435,436];
(2) Decaying dark matter [437].
(iii)
Deformations of the Hubble expansion rate H(z) with inhomogeneous or anisotropic modifications:
(1) Chameleon dark energy [438,439];
(2) Cosmic voids [440,441];
(3) Inhomogeneous causal horizons [419].
(iv)
Late time modifications: Transition or recalibration of the SNe Ia absolute luminosity:
(1) Gravity and evolution of the SNe Ia intrinsic luminosity [442];
(2) Transition of the SNe Ia absolute magnitude M at a redshift z ≃ 0.01 [443,444,445];
(3) Late (low-redshift) w M phantom transition [446,447,448].
(v)
Early time modifications of sound horizon:
(1) Early dark energy [449,450,451,452,453,454,455,456,457];
(2) Dark radiation [458,459,460,461,462,463];
(3) Neutrino self-interactions [340,464,465,466];
(4) Large primordial non-Gaussianities [409];
(5) Heisenberg’s uncertainty principle [410];
(6) Early modified gravity [368,369,370].
Of course, the above classifications may not completely cover all the solutions to relive H 0 tension. The proposal of new schemes never stops, such as the Weyl invariant gravity [467], Horndeski gravity [468], Λ S CDM model [469,470], Early Integrated Sachs–Wolfe (eISW) effect [471], realistic model of dark atoms [472], information dark energy [473], U(1) L μ L τ model with Majoron [474], etc.
Classification is a summary of previous research work on relieving the H 0 tension, which helps to find out the physical origin that causes the discrepancy in H 0 measurements. For the classification of solutions to the H 0 tension schemes, not all schemes fall perfectly into one of these categories [475]. At present, the proposed models and theories are usually divided into three categories: early-time model, late-time model and modified gravity. The boundary between the early-time models and late-time models is the recombination redshift ( z 1100 ). Detailed introduction of each scheme has been made in previous H 0 review literature [202] and Perivolaropoulos and Skara [21]. We will not repeat it here. In addition to categorizing so many solutions, there are several works that provide comparative analysis of solutions for H 0 tension, and discussion on the H 0 tension [388,475,476,477,478,479,480,481,482,483,484,485]. Schöneberg et al. [475] organized H 0 Olympic-like games for the relative success of seventeen models which have been proposed to resolve the H 0 tension, and gave a ranking. Finally, the early dark energy model, new early dark energy model, early modified gravity model and varying m e + Ω K model are the most successful of the models studied in the H 0 Olympic-like games.
According to the research and discussions on solving the H 0 tension, we tend to divide all of the solutions into two categories: one proposes a new cosmological model first, and then combines the existing early-time (for example, CMB) or late-time (for example, Cepheid) observational data-sets to constrain the cosmological parameters (including H 0 ). A higher H 0 value which is consistent with the SH0ES results can be given by utilizing the recent available observations. This category is the most common used to relieve the H 0 tension. We define such schemes as the sequential scheme. There is one thing to note here. Considering that the new model needs to introduce additional parameters, and the increase in the volume of the parameter space will make the final H 0 result incompact, this would amplify the extent to which the new model mitigates the H 0 tension. Moreover, it is diffcult to estimate the contribution of additional parameters to the degree of mitigation of the H 0 tension. The other is to use the Λ CDM model or model-independent methods, combined with the existing late-time observations center to find the H 0 singular behaviors which can be used to resolve or alleviate the H 0 tension. Compared to the sequential scheme, such a scheme should be called a reverse-order scheme. These H 0 singular behaviors might hint to new physics beyond the Λ CDM model, and still require new cosmological models to explain them. The proposal of the new cosmological model no longer directly alleviates the H 0 tension, but explains the H 0 singular behaviors. It seems that there is no need to worry about the problem of increasing the parameter space with the additional parameters. The former scheme has been elaborated in many review articles [21,202] and will not be repeated here. We will revisit the latter scenario, which might hint at new physics beyond the Λ CDM model, in the next section.

5. Evidence of New Physics beyond the Λ CDM

According to previous analyses of the quasar lensing, Wong et al. [37] found that the inferred value of H 0 , which was estimated using the strongly-lensed quasar time delay (H0LiCOW), decreases with the lens redshift. The H0LiCOW H 0 descending trend is of low statistical significance at 1.9 σ . Adding a new H0LiCOW H 0 result [174] (DES J0408-5354) reduces the statistical significance to 1.7 σ [143], as shown in Figure 3. After that, the TDCOSMO IV re-analysis lowers the H 0 estimates and increases the error bar [486]. Hence, the significance of the H 0 descending trend may be lower, and it is not clear that the H 0 descending trend is not a systematically driven-by-analysis choice in its rather complicated pipeline [486]. Even so, this still provides a new diagnostic for the H 0 tension.
Motivated by the H0LiCOW results [37], Krishnan et al. [487] focused their attention mainly on late-time observations. They estimated the cosmological parameters in different redshift ranges by binning the observational data-sets (z < 0.7) comprising megamasers, CCs, SNe Ia and BAOs according their redshifts. The total data-set is divided into six parts. Then, they constrained the H 0 value for each of the bin, and using ( z ¯ , H 0 ) present the final result. The form of z ¯ is written as [487]
z ¯ = n N i z n ( σ n ) 2 n N i ( σ n ) 2 ,
where σ k denotes the error in the observable at redshift z n . Finally, they found a similar H 0 descending trend in the Λ CDM model with a low statistical significance (2.1 σ ), as shown in the left panel of Figure 4. This result obtained by using the observation data completely independent of H0LiCOW is consistent with the H0LiCOW results to a certain extent.
Based on the flat Λ CDM model and flat w 0 w a CDM model, a similar H 0 descending trend is also found by Dainotti et al. [417] using the Pantheon sample only. In this analysis, they set the absolute magnitude of SNe Ia so that H 0 = 73.5 km/s/Mpc, and they fix fiducial values for Ω 0 m Λ C D M = 0.298 and Ω 0 m Λ C D M = 0.308. The g ( z ) function was used to describe the behavior of the H 0 descending. Its form is as follows:
g ( z ) = H 0 ( z ) = H 0 ˜ ( 1 + z ) α ,
where H 0 ˜ and α are free parameters, and α indicates the evolutionary trend. The g ( z ) function is the standard for characterizing the evolution of many astrophysical sources and is widely used for GRBs and quasars [488,489,490]. In addition, they also considered four kinds of binning methods: 3 bins, 4 bins, 20 bins and 40 bins. Finally, they reduced the H 0 tension in the range of (54%, 72%) for both cosmological models and pointed out that the H 0 descending trend is independent of the cosmological model and number of bins. A more detailed result can be found from Table 1 in Dainotti et al. [417].
Here, we demonstrate their fitting results obtained from the 4 bins method in the flat Λ CDM model. The results of H 0 ˜ and α are 73.493 ± 0.144 km/s/Mpc and 0.008 ± 0.006, respectively. Adopting the g ( z ) function, they obtained H 0 ( z = 1100 ) = 69.271 ± 2.815 km/s/Mpc, which can be used to reduce the H 0 tension effectively, as shown in the right panel of Figure 4. The reduction of the H 0 tension is 66%. A similar descending trend (i.e., H 0 decreasing as z m i n increases) was also found by Horstmann et al. [61] and Ó Colgáin et al. [491] from the Pantheon sample. The constraints of H 0 ( z m i n ) are obtained by using the SNe Ia data larger than z m i n .
After that, Dainotti et al. [492] made a further analysis for the H 0 evolution, which was described by Equation (7) combining the Pantheon sample and BAO data. The final results demonstrate that a descending trend with α 10 2 is still visible in the combined sample. The α coefficient reaches zero in 2.0 σ and 5.8 σ for the Λ CDM model and w 0 w a CDM model, respectively. In addition, Colgáin et al. [493] performed an independent investigation in the Λ CDM model, adopting the composite sample consisting of H(z), SNe Ia and quasars. In the end, they confirmed that the observations exhibit an increasing Ω m (decreasing H 0 ) trend with an increasing bin redshift and that such behaviour can arise randomly within the flat Λ CDM model with a lower probability p = 0.0021 (3.1 σ ).
At the same time, Perivolaropoulos and Skara [447] proposed a physically motivated new degree of freedom in the Cepheid calibrator analysis, allowing for a transition in one of the Cepheid modeling parameters R w (Cepheid Wesenheit color-luminosity parameter) or M w (Cepheid Wesenheit H-band absolute magnitude). This is mildly favored by the observational data and can yield a lower H 0 value. Their results may imply the presence of a fundamental physics transition taking place at a time more recent than 100 M y r s ago. The transition magnitude is consistent with the magnitude required for the resolution of the H 0 tension in the context of a fundamental gravitational transition occurring by a sudden increase in the strength of the gravitational interactions G e f f by about 10% [418] at a redshift z t ≲ 0.01. Such a transition would abruptly increase the absolute magnitude of SNe Ia by Δ M B ⋍ 0.2 [241,418].
Using the publicly available SH0ES data described in Riess et al. [39], the extended analyses were given by Perivolaropoulos and Skara [444] in a detailed and comprehensive manner. They found that when an absolute magnitude M B transition of the SNe Ia at D c ≃ 50 Mpc (about 160 Myrs ago) can drop, the H 0 constraint drops from 73.04 ± 1.04 km/s/Mpc to 67.32 ± 4.64 km/s/Mpc, which is in full consistency with the Planck results. When the inverse distance ladder constraint on M B > is included in the analyses, the uncertainties for H 0 reduce dramatically ( H 0 = 68.2 ± 0.8 km/s/Mpc) and the M B > transition model is strongly preferred over the baseline SH0ES model in terms of the Akaike Information Criterion (AIC) [494] and the Bayesian Information Criterion (BIC) [495] model selection criteria. Similar hints for a transition behavior is found for the other three main parameters of the analysis ( b W , M W and Z W ) at the same critical distance D c ≃ 50 Mpc, even though in that case the H 0 estimation is not significantly affected [444]. In addition, Wojtak and Hjorth [496] also reanalysed the SNe Ia and Cepheids’ observations and found that the H 0 local measurements become dependent on the choice of SN reference colour. These recent investigations hint towards the need of more detailed Cepheid + SNe Ia calibrating data at distances D c ≳ 50Mpc, i.e., at the high end of rung 2 on the distance ladder.
In a recent analysis, Krishnan et al. [416] construct the H 0 diagnostic H 0 ( z ) :
H 0 ( z ) = H ( z ) 1 Ω m 0 + Ω m 0 ( 1 + z ) 3 ,
to specify the possible deviations from the flat Λ CDM model using the Gaussian process (GP) method [497]. The Gaussian process has been extensively used for cosmological applications, such as the constraint on H 0 [498,499,500,501,502] and the comparison of cosmological models [503]. A more detailed explanation can be discovered from the literature [504,505,506]. As shown in Figure 5, we are given the reconstructed result of 36 H(z) data (31 CCs + 5 BAOs) using the GP method. From this figure, it can be learned that combining the H(z) data and GP method allows us to obtain a continuous function f(z) to represent the discrete H(z) data. Utilizing the function f(z), the H(z) value at any redshift within a certain range can be obtained, including H 0 at z = 0. The effective range mainly depends on the highest redshift of the H(z) data used. Finally, H(z) follows from the continuous function f(z) and the Ω m 0 is the adopted Planck values [3]. Based on the H(z) data and GP method, they found a running H 0 with redshift z. The main result can be found from Figure 2 in Krishnan et al. [416].
Also employing the H ( z ) data [499] and the GP method, Hu and Wang [507] reported a late-time transition of H 0 , i.e., H 0 changes from a low value to a high one from an early to late cosmic time that can be used to relieve the H 0 tension. Unlike previous studies [417,487], they processed the H(z) data using a cumulative binning method. An introduction to this method can be found in Figure 1 and Section 2 in Hu and Wang [507]. They found that the redshift of the H 0 transition occurs at z 0.49 . Without proposing a new cosmological model, their finding can be used to relieve the H 0 tension with a mitigation level of around 70 percent, and is consistent with the H0LiCOW results in the 1 σ range. They also tested the influence of BAOs on the result, and concluded that removing the BAOs data had no substantial effect, i.e., did not make the H 0 transition disappear. Their final results are shown in Figure 6.
Recently, utilizing the latest SNe Ia sample (Pantheon+ sample) and H(z) data, Jia et al. [508] presented a novel non-parametric method to estimate H 0 as a function of the redshift. They found a descending trend of H 0 , z with the statistical significance of 3.6 σ and 5.1 σ , corresponding to the equal-number and equal-width binning methods, respectively. Here, H 0 , z defined as the value of H 0 are derived from the cosmic observations at redshift z. The main results are presented in Figures 1 and 2 of Jia et al. [508]. The evolution of H 0 , z can effectively relieve the Hubble tension without the early-time modifications. Moreover, the results of the AIC and BIC demonstrate that the observational data favor the H 0 , z model over the Λ CDM model. Recently, utilizing a different approach than Jia et al. [508], Malekjani et al. [509] also found a similar H 0 descending trend from the Pantheon+ sample.
The statistical signification of the H 0 descending trend found from the quasar lensing is not high, at only 1.7 σ . Moreover, it is not clear that the H 0 descending trend is not caused by systematics. This still provides a new diagnostic for the H 0 tension. The descending trend of H 0 has also been discovered by utilizing the different data-sets and methods, most of which are based on an explicit model ( Λ CDM or w 0 w a CDM model) [37,61,416,417,487,491,492,493,508], except for Hu and Wang [507]. The H 0 descending trend can effectively alleviate the H 0 tension. If this trend is substantiated going forward, a late-time modification consistent with most observations is required. However, some studies are not in favor of modifying the late-time universe [510,511]. Among many late-time solutions, local void [512,513,514,515], modified gravity [410,516] and modified cosmological models [260] might be considered as competitive candidates. The local void model has been disfavored by the SNe Ia data [517,518,519], but can not completely be ruled out. Of course, there is also some evidence supporting the existence of the local void model [520,521,522]. The reasons for the transition of the Cepheid parameters and the M B transition of SNe Ia are still unclear. Such a transition may be attributed to either new physics or to unknown systematics hidden in the data [447]. If the source of the demonstrated transitions are physical, it could lead to new cosmological physics beyond the Λ CDM model. In previous studies, there are precedents, e.g., the gravitational constant in the context of a recent first-order phase transition [523,524,525] to a new vacuum of a scalar-tensor theory, or in the context of a generalization of the symmetron screening mechanism [445]. A similar first-order transition was implemented in early dark energy models [216], attempting to change the last scattering sound horizon scale without affecting other well-constrained cosmological observables. Thus, even though no relevant detailed analysis has been performed so far, there are physical mechanisms that could potentially induce the SNe Ia luminosity transition degree of freedom. In any case, in the face of these evidences that may be new physics beyond the Λ CDM model, one should not just be skeptical and do nothing.

6. Conclusions and Future Prospects

The Λ CDM model as the current standard cosmological model is consistent with almost all of the observational probes available until the present. However, it is not perfect, and there are still many theoretical difficulties and tensions. The significant discrepancy between the H 0 values measured from the local distance ladder and from the cosmic microwave background, i.e., Hubble tension, is the most serious challenge to the standard Λ CDM model. In this review, we have revisited this as the hottest issue, incorporating the latest research.
Until now, there has been a 4–6 σ discrepancy in H 0 measured by these two approaches, and the discrepancy is still increasing (see Section 1 and Section 2 and the reference therein). Initially, possible systematics in the Planck observations and the HST measurements were thought to be responsible for the H 0 tension. However, this possibility has been largely ruled out [35,40,46,88,89,90,91,92,94]. The current arbitration results given by other independent observations (including quasar lensing, Megamaser, GW, FRB, TRGB, etc.) cannot effectively arbitrate the H 0 tension. See Section 3 for details. Many researchers therefore choose to believe that the Hubble tension may be caused by new physics beyond the Λ CDM model [201]. So far, there have been a lot of schemes proposed to solve the H 0 tension (see Perivolaropoulos and Skara [21], Shah et al. [97], Di Valentino et al. [202] for a review). We have reviewed the classifications of solving schemes for the H 0 tension based on previous work by Di Valentino et al. [202] and Shah et al. [97] as well as newer studies (see Section 4.1). According to the research and discussions about solving the H 0 tension, we tend to divide all of the solutions into two categories: sequential and reverse-order schemes. In Section 5, we mainly review the reverse-order schemes, which might hint at new physics beyond the Λ CDM. Some of these schemes have discovered the late-time H 0 descending trend [37,61,416,417,487,491,492,493,508] or the late-time H 0 transition [507] that can be used to alleviate the H 0 tension through different independent observations and different methods. The remaining schemes found that considering the new degrees of freedom (parameter transition) can also effectively alleviate the H 0 tension when analyzing the Cepheid and SNe Ia data [444,447,496].
Looking to the future, two new CMB Stage 4 telescopes will be operational in Chile and the South Pole around 2030, which will further extend the spectral resolution. The depth of these surveys will be able to support or rule out many precombination modifications based on the Λ CDM model. The Zwicky Transient Facility and Foundation surveys of nearby SNe Ia will effectively reduce the potential calibration issues of the local distance ladder by resolving the underlying population characteristics, having cleaner selection functions, and providing more galaxies. The H 0 arbitration of independent observations will also be improved. The James Webb Space Telescope (JWST) will greatly expand the range of TRGB observation, and provide continuity in the case of any further degradation of the ageing HST. The VIRGO detector in Italy, and recently the KAGRA detector in Japan will provide more frequent event detections and better sky localization. They will provide a 2% measurement of H 0 within this decade by combining with an improved instrumental calibration. Many thousands of lenses will be detected by the wide-field surveys, such as the Vera Rubin Observatory, Euclid and the Nancy Grace Roman Observatory, and hundreds of which will have accurate time delay measurements [526,527,528]. Thus, there is a strong incentive to resolve the remaining systematics in the modelling and speed up the analysis pipeline, then clear the relationship between the H 0 descending trend and systematics. The ASKAP and Very Large Array will provide a large number of positioned FRBs in the future, which will provide higher-precision H 0 measurements. In addition, the e-ROSITA all-sky survey, French–Chinese satellite space-based multi-band astronomical variable objects monitor (SVOM) [529], Einstein Probe (EP) [530], and Transient High-Energy Sky and Early Universe Surveyor (THESEUS) [531] space missions together with ground- and space-based multi-messenger facilities will allow us to investigate H 0 tension in the poorly explored high-redshift universe.

Author Contributions

Conceptualization, J.-P.H. and F.-Y.W.; investigation, J.-P.H.; writing—original draft preparation, J.-P.H.; writing—review and editing, J.-P.H. and F.-Y.W.; supervision, F.-Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (grant No. U1831207), the China Manned Spaced Project (CMS-CSST-2021-A12), National Natural Science Foundation of China (grant No. 12273009), Jiangsu Funding Program for Excellent Postdoctoral Talent (20220ZB59) and Project funded by China Postdoctoral Science Foundation (2022M721561).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Alam, S.; Ata, M.; Bailey, S.; Beutler, F.; Bizyaev, D.; Blazek, J.A.; Bolton, A.S.; Brownstein, J.R.; Burden, A.; Chuang, C.H.; et al. The clustering of galaxies in the completed SDSS-III Baryon Oscillation Spectroscopic Survey: Cosmological analysis of the DR12 galaxy sample. Mon. Not. R. Astron. Soc. 2017, 470, 2617–2652. [Google Scholar] [CrossRef]
  2. Scolnic, D.M.; Jones, D.O.; Rest, A.; Pan, Y.C.; Chornock, R.; Foley, R.J.; Huber, M.E.; Kessler, R.; Narayan, G.; Riess, A.G.; et al. The Complete Light-curve Sample of Spectroscopically Confirmed SNe Ia from Pan-STARRS1 and Cosmological Constraints from the Combined Pantheon Sample. Astroph. J. 2018, 859, 101. [Google Scholar] [CrossRef]
  3. Planck Collaboration. Planck 2018 results. VI. Cosmological parameters. Astron. Astrophys. 2020, 641, A6. [Google Scholar] [CrossRef]
  4. Benisty, D.; Staicova, D. Testing late-time cosmic acceleration with uncorrelated baryon acoustic oscillation dataset. Astron. Astrophys. 2021, 647, A38. [Google Scholar] [CrossRef]
  5. Hu, J.P.; Wang, F.Y.; Dai, Z.G. Measuring cosmological parameters with a luminosity-time correlation of gamma-ray bursts. Mon. Not. R. Astron. Soc. 2021, 507, 730–742. [Google Scholar] [CrossRef]
  6. DES Collaboration. The first Hubble diagram and cosmological constraints using superluminous supernovae. Mon. Not. R. Astron. Soc. 2021, 504, 2535–2549. [Google Scholar] [CrossRef]
  7. Brout, D.; Scolnic, D.; Popovic, B.; Riess, A.G.; Carr, A.; Zuntz, J.; Kessler, R.; Davis, T.M.; Hinton, S.; Jones, D.; et al. The Pantheon+ Analysis: Cosmological Constraints. Astroph. J. 2022, 938, 110. [Google Scholar] [CrossRef]
  8. DES Collaboration. Dark Energy Survey Year 3 results: Calibration of lens sample redshift distributions using clustering redshifts with BOSS/eBOSS. Mon. Not. R. Astron. Soc. 2022, 513, 5517–5539. [Google Scholar] [CrossRef]
  9. Dainotti, M.G.; Nielson, V.; Sarracino, G.; Rinaldi, E.; Nagataki, S.; Capozziello, S.; Gnedin, O.Y.; Bargiacchi, G. Optical and X-ray GRB Fundamental Planes as cosmological distance indicators. Mon. Not. R. Astron. Soc. 2022, 514, 1828–1856. [Google Scholar] [CrossRef]
  10. Cao, S.; Ryan, J.; Ratra, B. Cosmological constraints from H II starburst galaxy, quasar angular size, and other measurements. Mon. Not. R. Astron. Soc. 2022, 509, 4745–4757. [Google Scholar] [CrossRef]
  11. de Cruz Perez, J.; Park, C.G.; Ratra, B. Current data are consistent with flat spatial hypersurfaces in the ΛCDM cosmological model but favor more lensing than the model predicts. arXiv 2022, arXiv:2211.04268. [Google Scholar]
  12. Liu, Y.; Liang, N.; Xie, X.; Yuan, Z.; Yu, H.; Wu, P. Gamma-Ray Burst Constraints on Cosmological Models from the Improved Amati Correlation. Astroph. J. 2022, 935, 7. [Google Scholar] [CrossRef]
  13. 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]
  14. Wang, F.Y.; Hu, J.P.; Zhang, G.Q.; Dai, Z.G. Standardized Long Gamma-Ray Bursts as a Cosmic Distance Indicator. Astroph. J. 2022, 924, 97. [Google Scholar] [CrossRef]
  15. Blanchard, A.; Héloret, J.Y.; Ilić, S.; Lamine, B.; Tutusaus, I. ΛCDM is alive and well. arXiv 2022, arXiv:2205.05017. [Google Scholar]
  16. Berti, E.; Cardoso, V.; Haiman, Z.; Holz, D.E.; Mottola, E.; Mukherjee, S.; Sathyaprakash, B.; Siemens, X.; Yunes, N. Snowmass2021 Cosmic Frontier White Paper: Fundamental Physics and Beyond the Standard Model. arXiv 2022, arXiv:2203.06240. [Google Scholar]
  17. Schmitz, K. Modern Cosmology, an Amuse-Gueule. arXiv 2022, arXiv:2203.04757. [Google Scholar]
  18. Buchert, T.; Coley, A.A.; Kleinert, H.; Roukema, B.F.; Wiltshire, D.L. Observational challenges for the standard FLRW model. Int. J. Mod. Phys. D 2016, 25, 1630007. [Google Scholar] [CrossRef]
  19. 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 Astrop. 2022, 34, 49–211. [Google Scholar] [CrossRef]
  20. Di Valentino, E.D. Challenges of the Standard Cosmological Model. Universe 2022, 8, 399. [Google Scholar] [CrossRef]
  21. Perivolaropoulos, L.; Skara, F. Challenges for ΛCDM: An update. New Astron. Rev. 2022, 95, 101659. [Google Scholar] [CrossRef]
  22. Aluri, P.K.; Cea, P.; Chingangbam, P.; Chu, M.C.; Clowes, R.G.; Hutsemékers, D.; Kochappan, J.P.; Krasiński, A.; Lopez, A.M.; Liu, L.; et al. Is the Observable Universe Consistent with the Cosmological Principle? arXiv 2022, arXiv:2207.05765. [Google Scholar]
  23. Krishnan, C.; Mondol, R.; Sheikh-Jabbari, M.M. Dipole Cosmology: The Copernican Paradigm Beyond FLRW. arXiv 2022, arXiv:2209.14918. [Google Scholar]
  24. Weinberg, S. The cosmological constant problem. Rev. Mod. Phys. 1989, 61, 1–23. [Google Scholar] [CrossRef]
  25. Courbin, V.L.; Marlow, D.R.; Dementi, A.E. Critical Problems in Physics; Princeton University Press: Princeton, NJ, USA, 1997. [Google Scholar]
  26. Martin, J. Everything you always wanted to know about the cosmological constant problem (but were afraid to ask). Comptes Rendus Phys. 2012, 13, 566–665. [Google Scholar] [CrossRef]
  27. Burgess, C.P. The Cosmological Constant Problem: Why it’s hard to get Dark Energy from Micro-physics. arXiv 2013, arXiv:1309.4133. [Google Scholar]
  28. Velten, H.E.S.; vom Marttens, R.F.; Zimdahl, W. Aspects of the cosmological “coincidence problem”. Eur. Phys. J. C 2014, 74, 3160. [Google Scholar] [CrossRef] [Green Version]
  29. Copeland, E.J.; Sami, M.; Tsujikawa, S. Dynamics of Dark Energy. Int. J. Mod. Phys. D 2006, 15, 1753–1935. [Google Scholar] [CrossRef]
  30. Solà, J. Cosmological constant and vacuum energy: Old and new ideas. J. Phys. Conf. Ser. 2013, 453, 012015. [Google Scholar] [CrossRef]
  31. Weinberg, S. Anthropic bound on the cosmological constant. Phys. Rev. Lett. 1987, 59, 2607–2610. [Google Scholar] [CrossRef]
  32. Susskind, L. The Anthropic Landscape of String Theory. In Proceedings of the Davis Meeting ON Cosmic Inflation, Davis, CA, USA, 22–23 March 2003; p. 26. [Google Scholar]
  33. Planck Collaboration. Planck 2015 results. XIII. Cosmological parameters. Astron. Astrophys. 2016, 594, A13. [Google Scholar] [CrossRef]
  34. Riess, A.G.; Casertano, S.; Yuan, W.; Macri, L.; Anderson, J.; MacKenty, J.W.; Bowers, J.B.; Clubb, K.I.; Filippenko, A.V.; Jones, D.O.; et al. New Parallaxes of Galactic Cepheids from Spatially Scanning the Hubble Space Telescope: Implications for the Hubble Constant. Astroph. J. 2018, 855, 136. [Google Scholar] [CrossRef]
  35. Riess, A.G.; Casertano, S.; Yuan, W.; Macri, L.M.; Scolnic, D. Large Magellanic Cloud Cepheid Standards Provide a 1% Foundation for the Determination of the Hubble Constant and Stronger Evidence for Physics beyond ΛCDM. Astroph. J. 2019, 876, 85. [Google Scholar] [CrossRef]
  36. Riess, A.G. The expansion of the Universe is faster than expected. Nat. Rev. Phys. 2020, 2, 10–12. [Google Scholar] [CrossRef]
  37. 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. R. Astron. Soc. 2020, 498, 1420–1439. [Google Scholar] [CrossRef]
  38. Di Valentino, E. A combined analysis of the H0 late time direct measurements and the impact on the Dark Energy sector. Mon. Not. R. Astron. Soc. 2021, 502, 2065–2073. [Google Scholar] [CrossRef]
  39. Riess, A.G.; Casertano, S.; Yuan, W.; Bowers, J.B.; Macri, L.; Zinn, J.C.; Scolnic, D. Cosmic Distances Calibrated to 1% Precision with Gaia EDR3 Parallaxes and Hubble Space Telescope Photometry of 75 Milky Way Cepheids Confirm Tension with ΛCDM. Astroph. J. Lett. 2021, 908, L6. [Google Scholar] [CrossRef]
  40. Riess, A.G.; Yuan, W.; Macri, L.M.; Scolnic, D.; Brout, D.; Casertano, S.; Jones, D.O.; Murakami, Y.; Anand, G.S.; Breuval, L.; 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. Astroph. J. Lett. 2022, 934, L7. [Google Scholar] [CrossRef]
  41. Basilakos, S.; Nesseris, S. Conjoined constraints on modified gravity from the expansion history and cosmic growth. Phys. Rev. D 2017, 96, 063517. [Google Scholar] [CrossRef]
  42. DES Collaboration. Dark Energy Survey year 1 results: Cosmological constraints from galaxy clustering and weak lensing. Phys. Rev. D 2018, 98, 043526. [Google Scholar] [CrossRef]
  43. Joudaki, S.; Blake, C.; Johnson, A.; Amon, A.; Asgari, M.; Choi, A.; Erben, T.; Glazebrook, K.; Harnois-Déraps, J.; Heymans, C.; et al. KiDS-450 + 2dFLenS: Cosmological parameter constraints from weak gravitational lensing tomography and overlapping redshift-space galaxy clustering. Mon. Not. R. Astron. Soc. 2018, 474, 4894–4924. [Google Scholar] [CrossRef]
  44. Akarsu, O.; Barrow, J.D.; Uzun, N.M. Screening anisotropy via energy-momentum squared gravity: Λ CDM model with hidden anisotropy. Phys. Rev. D 2020, 102, 124059. [Google Scholar] [CrossRef]
  45. Hu, J.P.; Wang, Y.Y.; Wang, F.Y. Testing cosmic anisotropy with Pantheon sample and quasars at high redshifts. Astron. Astrophys. 2020, 643, A93. [Google Scholar] [CrossRef]
  46. Planck Collaboration. Planck 2018 results. VII. Isotropy and statistics of the CMB. Astron. Astrophys. 2020, 641, A7. [Google Scholar] [CrossRef]
  47. 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]
  48. Secrest, N.J.; von Hausegger, S.; Rameez, M.; Mohayaee, R.; Sarkar, S.; Colin, J. A Test of the Cosmological Principle with Quasars. Astroph. J. Lett. 2021, 908, L51. [Google Scholar] [CrossRef]
  49. Zhao, D.; Xia, J.Q. Constraining the anisotropy of the Universe with the X-ray and UV fluxes of quasars. Eur. Phys. J. C 2021, 81, 694. [Google Scholar] [CrossRef]
  50. Akarsu, O.; Dereli, T.; Katırcı, N. ACDM cosmology with a quiescent anisotropy in a higher dimensional steady state universe. J. Phys. Conf. Ser. 2022, 2191, 012001. [Google Scholar] [CrossRef]
  51. Kalbouneh, B.; Marinoni, C.; Bel, J. The multipole expansion of the local expansion rate. arXiv 2022, arXiv:2210.11333. [Google Scholar] [CrossRef]
  52. Zhao, D.; Xia, J.Q. Testing cosmic anisotropy with the Ep-Eiso (‘Amati’) correlation of GRBs. Mon. Not. R. Astron. Soc. 2022, 511, 5661–5671. [Google Scholar] [CrossRef]
  53. Akarsu, O.; Di Valentino, E.; Kumar, S.; Ozyiğit, M.; Sharma, S. Testing spatial curvature and anisotropic expansion on top of the ΛCDM model. Phys. Dark Universe 2023, 39, 101162. [Google Scholar] [CrossRef]
  54. Kashlinsky, A.; Atrio-Barandela, F.; Kocevski, D.; Ebeling, H. A Measurement of Large-Scale Peculiar Velocities of Clusters of Galaxies: Results and Cosmological Implications. Astroph. J. Lett. 2008, 686, L49. [Google Scholar] [CrossRef]
  55. Watkins, R.; Feldman, H.A.; Hudson, M.J. Consistently large cosmic flows on scales of 100h−1Mpc: A challenge for the standard ΛCDM cosmology. Mon. Not. R. Astron. Soc. 2009, 392, 743–756. [Google Scholar] [CrossRef]
  56. Webb, J.K.; King, J.A.; Murphy, M.T.; Flambaum, V.V.; Carswell, R.F.; Bainbridge, M.B. Indications of a Spatial Variation of the Fine Structure Constant. Phys. Rev. Lett. 2011, 107, 191101. [Google Scholar] [CrossRef] [PubMed]
  57. King, J.A.; Webb, J.K.; Murphy, M.T.; Flambaum, V.V.; Carswell, R.F.; Bainbridge, M.B.; Wilczynska, M.R.; Koch, F.E. Spatial variation in the fine-structure constant—New results from VLT/UVES. Mon. Not. R. Astron. Soc. 2012, 422, 3370–3414. [Google Scholar] [CrossRef]
  58. Wiltshire, D.L.; Smale, P.R.; Mattsson, T.; Watkins, R. Hubble flow variance and the cosmic rest frame. Phys. Rev. D 2013, 88, 083529. [Google Scholar] [CrossRef]
  59. Bengaly, C.A.P.; Maartens, R.; Santos, M.G. Probing the Cosmological Principle in the counts of radio galaxies at different frequencies. J. Cosmol. Astropart. Phys. 2018, 2018, 031. [Google Scholar] [CrossRef] [Green Version]
  60. Zhao, D.; Xia, J.Q. A tomographic test of cosmic anisotropy with the recently-released quasar sample. Eur. Phys. J. C 2021, 81, 948. [Google Scholar] [CrossRef]
  61. Horstmann, N.; Pietschke, Y.; Schwarz, D.J. Inference of the cosmic rest-frame from supernovae Ia. Astron. Astrophys. 2022, 668, A34. [Google Scholar] [CrossRef]
  62. Luongo, O.; Muccino, M.; Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Yin, L. Larger H0 values in the CMB dipole direction. Phys. Rev. D 2022, 105, 103510. [Google Scholar] [CrossRef]
  63. Guandalin, C.; Piat, J.; Clarkson, C.; Maartens, R. Theoretical systematics in testing the Cosmological Principle with the kinematic quasar dipole. arXiv 2022, arXiv:2212.04925. [Google Scholar]
  64. Evslin, J. Isolating the Lyman alpha forest BAO anomaly. J. Cosmol. Astropart. Phys. 2017, 2017, 024. [Google Scholar] [CrossRef]
  65. 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. Astroph. J. 2018, 853, 119. [Google Scholar] [CrossRef]
  66. 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, 2019, 044. [Google Scholar] [CrossRef]
  67. Minami, Y.; Ochi, H.; Ichiki, K.; Katayama, N.; Komatsu, E.; Matsumura, T. Simultaneous determination of the cosmic birefringence and miscalibrated polarisation angles from CMB experiments. arXiv 2019, arXiv:1904.12440. [Google Scholar]
  68. Minami, Y. Determination of miscalibrated polarization angles from observed cosmic microwave background and foreground EB power spectra: Application to partial-sky observation. Prog. Theor. Exp. Phys. 2020, 2020, 063E01. [Google Scholar] [CrossRef]
  69. Minami, Y.; Komatsu, E. New Extraction of the Cosmic Birefringence from the Planck 2018 Polarization Data. Phys. Rev. Lett. 2020, 125, 221301. [Google Scholar] [CrossRef]
  70. Minami, Y.; Komatsu, E. Simultaneous determination of the cosmic birefringence and miscalibrated polarization angles II: Including cross-frequency spectra. Prog. Theor. Exp. Phys. 2020, 2020, 103E02. [Google Scholar] [CrossRef]
  71. Bullock, J.S.; Boylan-Kolchin, M. Small-Scale Challenges to the ΛCDM Paradigm. Annu. Rev. Astron. Astr. 2017, 55, 343–387. [Google Scholar] [CrossRef] [Green Version]
  72. Del Popolo, A.; Le Delliou, M. Small Scale Problems of the ΛCDM Model: A Short Review. Galaxies 2017, 5, 17. [Google Scholar] [CrossRef]
  73. Salucci, P. The distribution of dark matter in galaxies. Astron. Astrophys. Rev. 2019, 27, 2. [Google Scholar] [CrossRef]
  74. Di Paolo, C.; Salucci, P. Fundamental properties of the dark and the luminous matter from Low Surface Brightness discs. arXiv 2020, arXiv:2005.03520. [Google Scholar]
  75. Verde, L.; Protopapas, P.; Jimenez, R. Planck and the local Universe: Quantifying the tension. Phys. Dark Universe 2013, 2, 166–175. [Google Scholar] [CrossRef]
  76. Fields, B.D. The Primordial Lithium Problem. Annu. Rev. Nucl. Part. Sci. 2011, 61, 47–68. [Google Scholar] [CrossRef]
  77. Lusso, E.; Piedipalumbo, E.; Risaliti, G.; Paolillo, M.; Bisogni, S.; Nardini, E.; Amati, L. Tension with the flat ΛCDM model from a high-redshift Hubble diagram of supernovae, quasars, and gamma-ray bursts. Astron. Astrophys. 2019, 628, L4. [Google Scholar] [CrossRef]
  78. Risaliti, G.; Lusso, E. Cosmological Constraints from the Hubble Diagram of Quasars at High Redshifts. Nat. Astron. 2019, 3, 272–277. [Google Scholar] [CrossRef]
  79. Yang, T.; Banerjee, A.; Ó Colgáin, E. Cosmography and flat Λ CDM tensions at high redshift. Phys. Rev. D 2020, 102, 123532. [Google Scholar] [CrossRef]
  80. Banerjee, A.; Ó Colgáin, E.; Sasaki, M.; Sheikh-Jabbari, M.M.; Yang, T. On problems with cosmography in cosmic dark ages. Phys. Lett. B 2021, 818, 136366. [Google Scholar] [CrossRef]
  81. Hu, J.P.; Wang, F.Y. High-redshift cosmography: Application and comparison with different methods. Astron. Astrophys. 2022, 661, A71. [Google Scholar] [CrossRef]
  82. Antoniou, I.; Perivolaropoulos, L. Constraints on spatially oscillating sub-mm forces from the Stanford Optically Levitated Microsphere Experiment data. Phys. Rev. D 2017, 96, 104002. [Google Scholar] [CrossRef] [Green Version]
  83. Perivolaropoulos, L. Submillimeter spatial oscillations of Newton’s constant: Theoretical models and laboratory tests. Phys. Rev. D 2017, 95, 084050. [Google Scholar] [CrossRef]
  84. Bowman, J.D.; Rogers, A.E.E.; Monsalve, R.A.; Mozdzen, T.J.; Mahesh, N. An absorption profile centred at 78 megahertz in the sky-averaged spectrum. Nature 2018, 555, 67–70. [Google Scholar] [CrossRef] [PubMed]
  85. Kraljic, D.; Sarkar, S. How rare is the Bullet Cluster (in a ΛCDM universe)? J. Cosmol. Astropart. Phys. 2015, 2015, 050. [Google Scholar] [CrossRef]
  86. Asencio, E.; Banik, I.; Kroupa, P. A massive blow for ΛCDM—The high redshift, mass, and collision velocity of the interacting galaxy cluster El Gordo contradicts concordance cosmology. Mon. Not. R. Astron. Soc. 2021, 500, 5249–5267. [Google Scholar] [CrossRef]
  87. Planck Collaboration. Planck 2013 results. XVI. Cosmological parameters. Astron. Astrophys. 2014, 571, A16. [Google Scholar] [CrossRef]
  88. Planck Collaboration. Planck intermediate results. LI. Features in the cosmic microwave background temperature power spectrum and shifts in cosmological parameters. Astron. Astrophys. 2017, 607, A95. [Google Scholar] [CrossRef]
  89. Jones, D.O.; Riess, A.G.; Scolnic, D.M.; Pan, Y.C.; Johnson, E.; Coulter, D.A.; Dettman, K.G.; Foley, M.M.; Foley, R.J.; Huber, M.E.; et al. Should Type Ia Supernova Distances Be Corrected for Their Local Environments? Astroph. J. 2018, 867, 108. [Google Scholar] [CrossRef]
  90. Shanks, T.; Hogarth, L.M.; Metcalfe, N. Gaia Cepheid parallaxes and ’Local Hole’ relieve H0 tension. Mon. Not. R. Astron. Soc. 2019, 484, L64–L68. [Google Scholar] [CrossRef]
  91. Planck Collaboration. Planck 2018 results. I. Overview and the cosmological legacy of Planck. Astron. Astrophys. 2020, 641, A1. [Google Scholar] [CrossRef]
  92. Rigault, M.; Brinnel, V.; Aldering, G.; Antilogus, P.; Aragon, C.; Bailey, S.; Baltay, C.; Barbary, K.; Bongard, S.; Boone, K.; et al. Strong dependence of Type Ia supernova standardization on the local specific star formation rate. Astron. Astrophys. 2020, 644, A176. [Google Scholar] [CrossRef]
  93. Carneiro, S.; Pigozzo, C.; Alcaniz, J.S. Redshift systematics and the H0 tension problem. Eur. Phys. J. Plus 2022, 137, 537. [Google Scholar] [CrossRef]
  94. 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. R. Astron. Soc. 2022, 514, 4620–4628. [Google Scholar] [CrossRef]
  95. Hill, J.C.; McDonough, E.; Toomey, M.W.; Alexander, S. Early dark energy does not restore cosmological concordance. Phys. Rev. D 2020, 102, 043507. [Google Scholar] [CrossRef]
  96. D’Amico, G.; Senatore, L.; Zhang, P.; Zheng, H. The Hubble tension in light of the Full-Shape analysis of Large-Scale Structure data. J. Cosmol. Astropart. Phys. 2021, 2021, 072. [Google Scholar] [CrossRef]
  97. Shah, P.; Lemos, P.; Lahav, O. A buyer’s guide to the Hubble constant. Astron. Astrophys. Rev. 2021, 29, 9. [Google Scholar] [CrossRef]
  98. Verde, L.; Treu, T.; Riess, A.G. Tensions between the early and late Universe. Nat. Astron. 2019, 3, 891–895. [Google Scholar] [CrossRef]
  99. Weinberg, D.H.; Mortonson, M.J.; Eisenstein, D.J.; Hirata, C.; Riess, A.G.; Rozo, E. Observational probes of cosmic acceleration. Phys. Rep. 2013, 530, 87–255. [Google Scholar] [CrossRef]
  100. Bennett, C.L.; Larson, D.; Weiland, J.L.; Jarosik, N.; Hinshaw, G.; Odegard, N.; Smith, K.M.; Hill, R.S.; Gold, B.; Halpern, M.; et al. Nine-year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Final Maps and Results. Astrophys. J. Suppl. Ser. 2013, 208, 20. [Google Scholar] [CrossRef]
  101. Knox, L.; Millea, M. Hubble constant hunter’s guide. Phys. Rev. D 2020, 101, 043533. [Google Scholar] [CrossRef]
  102. SPT-3G Collaboration. Measurements of the E -mode polarization and temperature-E -mode correlation of the CMB from SPT-3G 2018 data. Phys. Rev. D 2021, 104, 022003. [Google Scholar] [CrossRef]
  103. Aiola, S.; Calabrese, E.; Maurin, L.; Naess, S.; Schmitt, B.L.; Abitbol, M.H.; Addison, G.E.; Ade, P.A.R.; Alonso, D.; Amiri, M.; et al. The Atacama Cosmology Telescope: DR4 maps and cosmological parameters. J. Cosmol. Astropart. Phys. 2020, 2020, 047. [Google Scholar] [CrossRef]
  104. Freedman, W.L.; Madore, B.F.; Gibson, B.K.; Ferrarese, L.; Kelson, D.D.; Sakai, S.; Mould, J.R.; Kennicutt, R.C., Jr.; Ford, H.C.; Graham, J.A.; et al. Final Results from the Hubble Space Telescope Key Project to Measure the Hubble Constant. Astroph. J. 2001, 553, 47–72. [Google Scholar] [CrossRef]
  105. 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. Astroph. J. 2012, 758, 24. [Google Scholar] [CrossRef]
  106. Lusso, E.; Risaliti, G. The Tight Relation between X-Ray and Ultraviolet Luminosity of Quasars. Astroph. J. 2016, 819, 154. [Google Scholar] [CrossRef]
  107. Bisogni, S.; Risaliti, G.; Lusso, E. A Hubble Diagram for Quasars. Front. Astron. Space Sci. 2017, 4, 68. [Google Scholar] [CrossRef]
  108. Lusso, E.; Risaliti, G. Quasars as standard candles. I. The physical relation between disc and coronal emission. Astron. Astrophys. 2017, 602, A79. [Google Scholar] [CrossRef]
  109. Melia, F. Cosmological test using the Hubble diagram of high-z quasars. Mon. Not. R. Astron. Soc. 2019, 489, 517–523. [Google Scholar] [CrossRef]
  110. Khadka, N.; Ratra, B. Quasar X-ray and UV flux, baryon acoustic oscillation, and Hubble parameter measurement constraints on cosmological model parameters. Mon. Not. R. Astron. Soc. 2020, 492, 4456–4468. [Google Scholar] [CrossRef]
  111. Cao, S.; Zajaček, M.; Panda, S.; Martínez-Aldama, M.L.; Czerny, B.; Ratra, B. Standardizing reverberation-measured C IV time-lag quasars, and using them with standardized Mg II quasars to constrain cosmological parameters. Mon. Not. R. Astron. Soc. 2022, 516, 1721–1740. [Google Scholar] [CrossRef]
  112. Khadka, N.; Martínez-Aldama, M.L.; Zajaček, M.; Czerny, B.; Ratra, B. Do reverberation-measured Hβ quasars provide a useful test of cosmology? Mon. Not. R. Astron. Soc. 2022, 513, 1985–2005. [Google Scholar] [CrossRef]
  113. Khadka, N.; Zajaček, M.; Panda, S.; Martínez-Aldama, M.L.; Ratra, B. Consistency study of high- and low-accreting Mg II quasars: No significant effect of the Fe II to Mg II flux ratio on the radius-luminosity relation dispersion. Mon. Not. R. Astron. Soc. 2022, 515, 3729–3748. [Google Scholar] [CrossRef]
  114. Khadka, N.; Ratra, B. Do quasar X-ray and UV flux measurements provide a useful test of cosmological models? Mon. Not. R. Astron. Soc. 2022, 510, 2753–2772. [Google Scholar] [CrossRef]
  115. Wang, B.; Liu, Y.; Yuan, Z.; Liang, N.; Yu, H.; Wu, P. Redshift-evolutionary X-Ray and UV Luminosity Relation of Quasars from Gaussian Copula. Astroph. J. 2022, 940, 174. [Google Scholar] [CrossRef]
  116. Dainotti, M.G.; Cardone, V.F.; Capozziello, S. A time-luminosity correlation for γ-ray bursts in the X-rays. Mon. Not. R. Astron. Soc. 2008, 391, L79–L83. [Google Scholar] [CrossRef]
  117. Wang, F.Y.; Dai, Z.G.; Liang, E.W. Gamma-ray burst cosmology. New Astron. Rev. 2015, 67, 1–17. [Google Scholar] [CrossRef]
  118. Dainotti, M.G.; Del Vecchio, R. Gamma Ray Burst afterglow and prompt-afterglow relations: An overview. New Astron. Rev. 2017, 77, 23–61. [Google Scholar] [CrossRef]
  119. Cao, S.; Khadka, N.; Ratra, B. Standardizing Dainotti-correlated gamma-ray bursts, and using them with standardized Amati-correlated gamma-ray bursts to constrain cosmological model parameters. Mon. Not. R. Astron. Soc. 2022, 510, 2928–2947. [Google Scholar] [CrossRef]
  120. Cao, S.; Dainotti, M.; Ratra, B. Standardizing Platinum Dainotti-correlated gamma-ray bursts, and using them with standardized Amati-correlated gamma-ray bursts to constrain cosmological model parameters. Mon. Not. R. Astron. Soc. 2022, 512, 439–454. [Google Scholar] [CrossRef]
  121. Cao, S.; Ratra, B. Using lower redshift, non-CMB, data to constrain the Hubble constant and other cosmological parameters. Mon. Not. R. Astron. Soc. 2022, 513, 5686–5700. [Google Scholar] [CrossRef]
  122. Deng, C.; Huang, Y.F.; Xu, F. Pseudo Redshifts of Gamma-Ray Bursts Derived from the L-T-E Correlation. arXiv 2022, arXiv:2212.01990. [Google Scholar] [CrossRef]
  123. Jia, X.D.; Hu, J.P.; Yang, J.; Zhang, B.B.; Wang, F.Y. E iso-Ep correlation of gamma-ray bursts: Calibration and cosmological applications. Mon. Not. R. Astron. Soc. 2022, 516, 2575–2585. [Google Scholar] [CrossRef]
  124. Liu, Y.; Chen, F.; Liang, N.; Yuan, Z.; Yu, H.; Wu, P. The Improved Amati Correlations from Gaussian Copula. Astroph. J. 2022, 931, 50. [Google Scholar] [CrossRef]
  125. Liang, N.; Li, Z.; Xie, X.; Wu, P. Calibrating Gamma-Ray Bursts by Using a Gaussian Process with Type Ia Supernovae. Astroph. J. 2022, 941, 84. [Google Scholar] [CrossRef]
  126. Li, Z.; Zhang, B.; Liang, N. Constraints on Dark Energy Models with Gamma-Ray Bursts Calibrated from the Observational H(z) Data. arXiv 2022, arXiv:2212.14291. [Google Scholar]
  127. Muccino, M.; Luongo, O.; Jain, D. Constraints on the transition redshift from the calibrated Gamma-ray Burst Ep-Eiso correlation. arXiv 2022, arXiv:2208.13700. [Google Scholar]
  128. Luongo, O.; Muccino, M. Intermediate redshift calibration of gamma-ray bursts and cosmic constraints in non-flat cosmology. Mon. Not. R. Astron. Soc. 2023, 518, 2247–2255. [Google Scholar] [CrossRef]
  129. Riess, A.G.; Macri, L.M.; Hoffmann, S.L.; Scolnic, D.; Casertano, S.; Filippenko, A.V.; Tucker, B.E.; Reid, M.J.; Jones, D.O.; Silverman, J.M.; et al. A 2.4% Determination of the Local Value of the Hubble Constant. Astroph. J. 2016, 826, 56. [Google Scholar] [CrossRef]
  130. Cardona, W.; Kunz, M.; Pettorino, V. Determining H0 with Bayesian hyper-parameters. J. Cosmol. Astropart. Phys. 2017, 2017, 056. [Google Scholar] [CrossRef]
  131. Burns, C.R.; Parent, E.; Phillips, M.M.; Stritzinger, M.; Krisciunas, K.; Suntzeff, N.B.; Hsiao, E.Y.; Contreras, C.; Anais, J.; Boldt, L.; et al. The Carnegie Supernova Project: Absolute Calibration and the Hubble Constant. Astroph. J. 2018, 869, 56. [Google Scholar] [CrossRef]
  132. Feeney, S.M.; Mortlock, D.J.; Dalmasso, N. Clarifying the Hubble constant tension with a Bayesian hierarchical model of the local distance ladder. Mon. Not. R. Astron. Soc. 2018, 476, 3861–3882. [Google Scholar] [CrossRef]
  133. Follin, B.; Knox, L. Insensitivity of the distance ladder Hubble constant determination to Cepheid calibration modelling choices. Mon. Not. R. Astron. Soc. 2018, 477, 4534–4542. [Google Scholar] [CrossRef]
  134. Dhawan, S.; Jha, S.W.; Leibundgut, B. Measuring the Hubble constant with Type Ia supernovae as near-infrared standard candles. Astron. Astrophys. 2018, 609, A72. [Google Scholar] [CrossRef] [Green Version]
  135. Camarena, D.; Marra, V. Local determination of the Hubble constant and the deceleration parameter. Phys. Rev. Res. 2020, 2, 013028. [Google Scholar] [CrossRef]
  136. Javanmardi, B.; Mérand, A.; Kervella, P.; Breuval, L.; Gallenne, A.; Nardetto, N.; Gieren, W.; Pietrzyński, G.; Hocdé, V.; Borgniet, S. Inspecting the Cepheid Distance Ladder: The Hubble Space Telescope Distance to the SN Ia Host Galaxy NGC 5584. Astroph. J. 2021, 911, 12. [Google Scholar] [CrossRef]
  137. Spergel, D.N.; Verde, L.; Peiris, H.V.; Komatsu, E.; Nolta, M.R.; Bennett, C.L.; Halpern, M.; Hinshaw, G.; Jarosik, N.; Kogut, A.; et al. First-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Determination of Cosmological Parameters. Astrophys. J. Suppl. Ser. 2003, 148, 175–194. [Google Scholar] [CrossRef]
  138. Spergel, D.N.; Bean, R.; Doré, O.; Nolta, M.R.; Bennett, C.L.; Dunkley, J.; Hinshaw, G.; Jarosik, N.; Komatsu, E.; Page, L.; et al. Three-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Implications for Cosmology. Astrophys. J. Suppl. Ser. 2007, 170, 377–408. [Google Scholar] [CrossRef]
  139. Komatsu, E.; Dunkley, J.; Nolta, M.R.; Bennett, C.L.; Gold, B.; Hinshaw, G.; Jarosik, N.; Larson, D.; Limon, M.; Page, L.; et al. Five-Year Wilkinson Microwave Anisotropy Probe Observations: Cosmological Interpretation. Astrophys. J. Suppl. Ser. 2009, 180, 330–376. [Google Scholar] [CrossRef]
  140. Komatsu, E.; Smith, K.M.; Dunkley, J.; Bennett, C.L.; Gold, B.; Hinshaw, G.; Jarosik, N.; Larson, D.; Nolta, M.R.; Page, L.; et al. Seven-year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Cosmological Interpretation. Astrophys. J. Suppl. Ser. 2011, 192, 18. [Google Scholar] [CrossRef]
  141. Riess, A.G.; Macri, L.; Casertano, S.; Sosey, M.; Lampeitl, H.; Ferguson, H.C.; Filippenko, A.V.; Jha, S.W.; Li, W.; Chornock, R.; et al. A Redetermination of the Hubble Constant with the Hubble Space Telescope from a Differential Distance Ladder. Astroph. J. 2009, 699, 539–563. [Google Scholar] [CrossRef]
  142. Riess, A.G.; Macri, L.; Casertano, S.; Lampeitl, H.; Ferguson, H.C.; Filippenko, A.V.; Jha, S.W.; Li, W.; Chornock, R. A 3% Solution: Determination of the Hubble Constant with the Hubble Space Telescope and Wide Field Camera 3. Astroph. J. 2011, 730, 119. [Google Scholar] [CrossRef]
  143. 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]
  144. Kuo, C.Y.; Braatz, J.A.; Reid, M.J.; Lo, K.Y.; Condon, J.J.; Impellizzeri, C.M.V.; Henkel, C. The Megamaser Cosmology Project. V. An Angular-diameter Distance to NGC 6264 at 140 Mpc. Astroph. J. 2013, 767, 155. [Google Scholar] [CrossRef] [Green Version]
  145. Reid, M.J.; Braatz, J.A.; Condon, J.J.; Lo, K.Y.; Kuo, C.Y.; Impellizzeri, C.M.V.; Henkel, C. The Megamaser Cosmology Project. IV. A Direct Measurement of the Hubble Constant from UGC 3789. Astroph. J. 2013, 767, 154. [Google Scholar] [CrossRef]
  146. Kuo, C.Y.; Braatz, J.A.; Lo, K.Y.; Reid, M.J.; Suyu, S.H.; Pesce, D.W.; Condon, J.J.; Henkel, C.; Impellizzeri, C.M.V. The Megamaser Cosmology Project. VI. Observations of NGC 6323. Astroph. J. 2015, 800, 26. [Google Scholar] [CrossRef]
  147. Reid, M.J.; Pesce, D.W.; Riess, A.G. An Improved Distance to NGC 4258 and Its Implications for the Hubble Constant. Astroph. J. Lett. 2019, 886, L27. [Google Scholar] [CrossRef]
  148. 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. Astroph. J. Lett. 2020, 891, L1. [Google Scholar] [CrossRef]
  149. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; Adya, V.B.; et al. A gravitational-wave standard siren measurement of the Hubble constant. Nature 2017, 551, 85–88. [Google Scholar] [CrossRef]
  150. Mooley, K.P.; Deller, A.T.; Gottlieb, O.; Nakar, E.; Hallinan, G.; Bourke, S.; Frail, D.A.; Horesh, A.; Corsi, A.; Hotokezaka, K. Superluminal motion of a relativistic jet in the neutron-star merger GW170817. Nature 2018, 561, 355–359. [Google Scholar] [CrossRef]
  151. Hotokezaka, K.; Nakar, E.; Gottlieb, O.; Nissanke, S.; Masuda, K.; Hallinan, G.; Mooley, K.P.; Deller, A.T. A Hubble constant measurement from superluminal motion of the jet in GW170817. Nat. Astron. 2019, 3, 940–944. [Google Scholar] [CrossRef]
  152. Hagstotz, S.; Reischke, R.; Lilow, R. A new measurement of the Hubble constant using fast radio bursts. Mon. Not. R. Astron. Soc. 2022, 511, 662–667. [Google Scholar] [CrossRef]
  153. Wu, Q.; Zhang, G.Q.; Wang, F.Y. An 8 per cent determination of the Hubble constant from localized fast radio bursts. Mon. Not. R. Astron. Soc. 2022, 515, L1–L5. [Google Scholar] [CrossRef]
  154. James, C.W.; Ghosh, E.M.; Prochaska, J.X.; Bannister, K.W.; Bhandari, S.; Day, C.K.; Deller, A.T.; Glowacki, M.; Gordon, A.C.; Heintz, K.E.; et al. A measurement of Hubble’s Constant using Fast Radio Bursts. Mon. Not. R. Astron. Soc. 2022, 516, 4862–4881. [Google Scholar] [CrossRef]
  155. Liu, Y.; Yu, H.; Wu, P. Cosmological-model-independent determination of Hubble constant from fast radio bursts and Hubble parameter measurements. arXiv 2022, arXiv:2210.05202. [Google Scholar]
  156. Zhao, Z.W.; Zhang, J.G.; Li, Y.; Zou, J.M.; Zhang, J.F.; Zhang, X. First statistical measurement of the Hubble constant using unlocalized fast radio bursts. arXiv 2022, arXiv:2212.13433. [Google Scholar]
  157. Freedman, W.L.; Madore, B.F.; Hatt, D.; Hoyt, T.J.; Jang, I.S.; Beaton, R.L.; Burns, C.R.; Lee, M.G.; Monson, A.J.; Neeley, J.R.; et al. The Carnegie-Chicago Hubble Program. VIII. An Independent Determination of the Hubble Constant Based on the Tip of the Red Giant Branch. Astroph. J. 2019, 882, 34. [Google Scholar] [CrossRef]
  158. 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. Astroph. J. 2020, 891, 57. [Google Scholar] [CrossRef]
  159. Freedman, W.L. Measurements of the Hubble Constant: Tensions in Perspective. Astroph. J. 2021, 919, 16. [Google Scholar] [CrossRef]
  160. Vagnozzi, S.; Pacucci, F.; Loeb, A. Implications for the Hubble tension from the ages of the oldest astrophysical objects. J. High Energy Astrop. 2022, 36, 27–35. [Google Scholar] [CrossRef]
  161. Wei, J.J.; Melia, F. Exploring the Hubble Tension and Spatial Curvature from the Ages of Old Astrophysical Objects. Astroph. J. 2022, 928, 165. [Google Scholar] [CrossRef]
  162. Moresco, M.; Amati, L.; Amendola, L.; Birrer, S.; Blakeslee, J.P.; Cantiello, M.; Cimatti, A.; Darling, J.; Della Valle, M.; Fishbach, M.; et al. Unveiling the Universe with emerging cosmological probes. Living Rev. Relativ. 2022, 25, 6. [Google Scholar] [CrossRef]
  163. Courbin, F.; Minniti, D. Gravitational Lensing: An Astrophysical Tool; Springer: Berlin/Heidelberg, Germany, 2002; Volume 608. [Google Scholar]
  164. Suyu, S.H.; Chang, T.C.; Courbin, F.; Okumura, T. Cosmological Distance Indicators. Space Sci. Rev. 2018, 214, 91. [Google Scholar] [CrossRef]
  165. 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. XIII. Improved Hubble constant measurement from lensing time delays using spatially resolved stellar kinematics of the lens galaxy. arXiv 2023, arXiv:2301.02656. [Google Scholar]
  166. Suyu, S.H.; Halkola, A. The halos of satellite galaxies: The companion of the massive elliptical lens SL2S J08544-0121. Astron. Astrophys. 2010, 524, A94. [Google Scholar] [CrossRef]
  167. Jee, I.; Suyu, S.H.; Komatsu, E.; Fassnacht, C.D.; Hilbert, S.; Koopmans, L.V.E. A measurement of the Hubble constant from angular diameter distances to two gravitational lenses. Science 2019, 365, 1134–1138. [Google Scholar] [CrossRef] [PubMed]
  168. Suyu, S.H.; Treu, T.; Hilbert, S.; Sonnenfeld, A.; Auger, M.W.; Blandford, R.D.; Collett, T.; Courbin, F.; Fassnacht, C.D.; Koopmans, L.V.E.; et al. Cosmology from Gravitational Lens Time Delays and Planck Data. Astroph. J. Lett. 2014, 788, L35. [Google Scholar] [CrossRef]
  169. Chen, G.C.F.; Fassnacht, C.D.; Suyu, S.H.; Rusu, C.E.; Chan, J.H.H.; Wong, K.C.; Auger, M.W.; Hilbert, S.; Bonvin, V.; Birrer, S.; et al. A SHARP view of H0LiCOW: H0 from three time-delay gravitational lens systems with adaptive optics imaging. Mon. Not. R. Astron. Soc. 2019, 490, 1743–1773. [Google Scholar] [CrossRef]
  170. Wong, K.C.; Suyu, S.H.; Auger, M.W.; Bonvin, V.; Courbin, F.; Fassnacht, C.D.; Halkola, A.; Rusu, C.E.; Sluse, D.; Sonnenfeld, A.; et al. H0LiCOW-IV. Lens mass model of HE 0435-1223 and blind measurement of its time-delay distance for cosmology. Mon. Not. R. Astron. Soc. 2017, 465, 4895–4913. [Google Scholar] [CrossRef]
  171. Birrer, S.; Treu, T.; Rusu, C.E.; Bonvin, V.; Fassnacht, C.D.; Chan, J.H.H.; Agnello, A.; Shajib, A.J.; Chen, G.C.F.; Auger, M.; et al. H0LiCOW-IX. Cosmographic analysis of the doubly imaged quasar SDSS 1206+4332 and a new measurement of the Hubble constant. Mon. Not. R. Astron. Soc. 2019, 484, 4726–4753. [Google Scholar] [CrossRef]
  172. Rusu, C.E.; Wong, K.C.; Bonvin, V.; Sluse, D.; Suyu, S.H.; Fassnacht, C.D.; Chan, J.H.H.; Hilbert, S.; Auger, M.W.; Sonnenfeld, A.; et al. H0LiCOW XII. Lens mass model of WFI2033-4723 and blind measurement of its time-delay distance and H0. Mon. Not. R. Astron. Soc. 2020, 498, 1440–1468. [Google Scholar] [CrossRef]
  173. DES Collaboration. Discovery of the Lensed Quasar System DES J0408-5354. Astroph. J. Lett. 2017, 838, L15. [Google Scholar] [CrossRef]
  174. Shajib, A.J.; Birrer, S.; Treu, T.; Agnello, A.; Buckley-Geer, E.J.; Chan, J.H.H.; Christensen, L.; Lemon, C.; Lin, H.; Millon, M.; et al. STRIDES: A 3.9 per cent measurement of the Hubble constant from the strong lens system DES J0408-5354. Mon. Not. R. Astron. Soc. 2020, 494, 6072–6102. [Google Scholar] [CrossRef]
  175. Humphreys, E.M.L.; Reid, M.J.; Moran, J.M.; Greenhill, L.J.; Argon, A.L. Toward a New Geometric Distance to the Active Galaxy NGC 4258. III. Final Results and the Hubble Constant. Astroph. J. 2013, 775, 13. [Google Scholar] [CrossRef]
  176. Claussen, M.J.; Heiligman, G.M.; Lo, K.Y. Water-vapour maser emission from galactic nuclei. Nature 1984, 310, 298–300. [Google Scholar] [CrossRef]
  177. Nakai, N.; Inoue, M.; Miyoshi, M. Extremely-high-velocity H20 maser emission in the galaxy NGC4258. Nature 1993, 361, 45–47. [Google Scholar] [CrossRef]
  178. Herrnstein, J.R.; Moran, J.M.; Greenhill, L.J.; Diamond, P.J.; Inoue, M.; Nakai, N.; Miyoshi, M.; Henkel, C.; Riess, A. A geometric distance to the galaxy NGC4258 from orbital motions in a nuclear gas disk. Nature 1999, 400, 539–541. [Google Scholar] [CrossRef] [Green Version]
  179. Braatz, J.; Greenhill, L.; Reid, M.; Condon, J.; Henkel, C.; Lo, K.Y. Precision cosmology with H2O megamasers: Progress in measuring distances to galaxies in the Hubble flow. In Proceedings of the Astrophysical Masers and Their Environments; Chapman, J.M., Baan, W.A., Eds.; Cambridge University Press: Cambridge, UK, 2007; Volume 242, pp. 399–401. [Google Scholar] [CrossRef]
  180. Braatz, J.A.; Gugliucci, N.E. The Discovery of Water Maser Emission from Eight Nearby Galaxies. Astroph. J. 2008, 678, 96–101. [Google Scholar] [CrossRef]
  181. Gao, F.; Braatz, J.A.; Reid, M.J.; Condon, J.J.; Greene, J.E.; Henkel, C.; Impellizzeri, C.M.V.; Lo, K.Y.; Kuo, C.Y.; Pesce, D.W.; et al. The Megamaser Cosmology Project. IX. Black Hole Masses for Three Maser Galaxies. Astroph. J. 2017, 834, 52. [Google Scholar] [CrossRef]
  182. Gao, F.; Braatz, J.A.; Reid, M.J.; Lo, K.Y.; Condon, J.J.; Henkel, C.; Kuo, C.Y.; Impellizzeri, C.M.V.; Pesce, D.W.; Zhao, W. The Megamaser Cosmology Project. VIII. A Geometric Distance to NGC 5765b. Astroph. J. 2016, 817, 128. [Google Scholar] [CrossRef]
  183. Pesce, D.W.; Braatz, J.A.; Reid, M.J.; Condon, J.J.; Gao, F.; Henkel, C.; Kuo, C.Y.; Lo, K.Y.; Zhao, W. The Megamaser Cosmology Project. XI. A Geometric Distance to CGCG 074-064. Astroph. J. 2020, 890, 118. [Google Scholar] [CrossRef]
  184. Braatz, J.; Condon, J.; Henkel, C.; Greene, J.; Lo, F.; Reid, M.; Pesce, D.; Gao, F.; Impellizzeri, V.; Kuo, C.Y.; et al. A Measurement of the Hubble Constant by the Megamaser Cosmology Project. In Proceedings of the Astrophysical Masers: Unlocking the Mysteries of the Universe; Tarchi, A., Reid, M.J., Castangia, P., Eds.; Cambridge University Press: Cambridge, UK, 2018; Volume 336, pp. 86–91. [Google Scholar] [CrossRef]
  185. LIGO Scientific Collaboration. Advanced LIGO. Class. Quant. Grav. 2015, 32, 074001. [Google Scholar] [CrossRef]
  186. Acernese, F.; Agathos, M.; Agatsuma, K.; Aisa, D.; Allemandou, N.; Allocca, A.; Amarni, J.; Astone, P.; Balestri, G.; Ballardin, G.; et al. Advanced Virgo: A second-generation interferometric gravitational wave detector. Class. Quant. Grav. 2015, 32, 024001. [Google Scholar] [CrossRef]
  187. LIGO Scientific Collaboration; Virgo Collaboration. Gravitational Waves and Gamma-Rays from a Binary Neutron Star Merger: GW170817 and GRB 170817A. Astroph. J. Lett. 2017, 848, L13. [Google Scholar] [CrossRef]
  188. Goldstein, A.; Veres, P.; Burns, E.; Briggs, M.S.; Hamburg, R.; Kocevski, D.; Wilson-Hodge, C.A.; Preece, R.D.; Poolakkil, S.; Roberts, O.J.; et al. An Ordinary Short Gamma-Ray Burst with Extraordinary Implications: Fermi-GBM Detection of GRB 170817A. Astroph. J. Lett. 2017, 848, L14. [Google Scholar] [CrossRef] [Green Version]
  189. Savchenko, V.; Ferrigno, C.; Kuulkers, E.; Bazzano, A.; Bozzo, E.; Brandt, S.; Chenevez, J.; Courvoisier, T.J.L.; Diehl, R.; Domingo, A.; et al. INTEGRAL Detection of the First Prompt Gamma-Ray Signal Coincident with the Gravitational-wave Event GW170817. Astroph. J. Lett. 2017, 848, L15. [Google Scholar] [CrossRef]
  190. LIGO Scientific Collaboration; Virgo Collaboration. Multi-messenger Observations of a Binary Neutron Star Merger. Astroph. J. Lett. 2017, 848, L12. [Google Scholar] [CrossRef]
  191. LIGO Scientific Collaboration; Virgo Collaboration; KAGRA Collaboration. Constraints on the cosmic expansion history from GWTC-3. arXiv 2021, arXiv:2111.03604. [Google Scholar]
  192. Mukherjee, S.; Krolewski, A.; Wandelt, B.D.; Silk, J. Cross-correlating dark sirens and galaxies: Measurement of H0 from GWTC-3 of LIGO-Virgo-KAGRA. arXiv 2022, arXiv:2203.03643. [Google Scholar]
  193. Lorimer, D.R.; Bailes, M.; McLaughlin, M.A.; Narkevic, D.J.; Crawford, F. A Bright Millisecond Radio Burst of Extragalactic Origin. Science 2007, 318, 777. [Google Scholar] [CrossRef]
  194. Xiao, D.; Wang, F.; Dai, Z. The physics of fast radio bursts. Science China Physics, Mechanics, and Astronomy 2021, 64, 249501. [Google Scholar] [CrossRef]
  195. Zhang, B. The Physics of Fast Radio Bursts. arXiv 2022, arXiv:2212.03972. [Google Scholar]
  196. Deng, W.; Zhang, B. Cosmological Implications of Fast Radio Burst/Gamma-Ray Burst Associations. Astroph. J. Lett. 2014, 783, L35. [Google Scholar] [CrossRef]
  197. Shull, J.M.; Smith, B.D.; Danforth, C.W. The Baryon Census in a Multiphase Intergalactic Medium: 30% of the Baryons May Still be Missing. Astroph. J. 2012, 759, 23. [Google Scholar] [CrossRef]
  198. McQuinn, M. Locating the “Missing” Baryons with Extragalactic Dispersion Measure Estimates. Astroph. J. Lett. 2014, 780, L33. [Google Scholar] [CrossRef]
  199. Zhang, Z.J.; Yan, K.; Li, C.M.; Zhang, G.Q.; Wang, F.Y. Intergalactic Medium Dispersion Measures of Fast Radio Bursts Estimated from IllustrisTNG Simulation and Their Cosmological Applications. Astroph. J. 2021, 906, 49. [Google Scholar] [CrossRef]
  200. Zhang, G.Q.; Yu, H.; He, J.H.; Wang, F.Y. Dispersion Measures of Fast Radio Burst Host Galaxies Derived from IllustrisTNG Simulation. Astroph. J. 2020, 900, 170. [Google Scholar] [CrossRef]
  201. Vagnozzi, S. New physics in light of the H0 tension: An alternative view. Phys. Rev. D 2020, 102, 023518. [Google Scholar] [CrossRef]
  202. Di Valentino, E.; 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]
  203. Sabla, V.I.; Caldwell, R.R. No H0 assistance from assisted quintessence. Phys. Rev. D 2021, 103, 103506. [Google Scholar] [CrossRef]
  204. Kamionkowski, M.; Riess, A.G. The Hubble Tension and Early Dark Energy. arXiv 2022, arXiv:2211.04492. [Google Scholar]
  205. Herold, L.; Ferreira, E.G.M. Resolving the Hubble tension with Early Dark Energy. arXiv 2022, arXiv:2210.16296. [Google Scholar]
  206. 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] [PubMed]
  207. Kaloper, N. Dark energy, H0 and weak gravity conjecture. Int. J. Mod. Phys. D 2019, 28, 1944017. [Google Scholar] [CrossRef]
  208. Lucca, M. The role of CMB spectral distortions in the Hubble tension: A proof of principle. Phys. Lett. B 2020, 810, 135791. [Google Scholar] [CrossRef]
  209. Chudaykin, A.; Gorbunov, D.; Nedelko, N. Exploring an early dark energy solution to the Hubble tension with Planck and SPTPol data. Phys. Rev. D 2021, 103, 043529. [Google Scholar] [CrossRef]
  210. Haridasu, B.S.; Viel, M.; Vittorio, N. Sources of H0-tension in dark energy scenarios. Phys. Rev. D 2021, 103, 063539. [Google Scholar] [CrossRef]
  211. Murgia, R.; Abellán, G.F.; Poulin, V. Early dark energy resolution to the Hubble tension in light of weak lensing surveys and lensing anomalies. Phys. Rev. D 2021, 103, 063502. [Google Scholar] [CrossRef]
  212. Berghaus, K.V.; Karwal, T. Thermal friction as a solution to the Hubble tension. Phys. Rev. D 2020, 101, 083537. [Google Scholar] [CrossRef]
  213. Alexander, S.; McDonough, E. Axion-dilaton destabilization and the Hubble tension. Phys. Lett. B 2019, 797, 134830. [Google Scholar] [CrossRef]
  214. Chudaykin, A.; Gorbunov, D.; Nedelko, N. Combined analysis of Planck and SPTPol data favors the early dark energy models. J. Cosmol. Astropart. Phys. 2020, 2020, 013. [Google Scholar] [CrossRef]
  215. Agrawal, P.; Cyr-Racine, F.Y.; Pinner, D.; Randall, L. Rock ‘n’ Roll Solutions to the Hubble Tension. arXiv 2019, arXiv:1904.01016. [Google Scholar]
  216. Niedermann, F.; Sloth, M.S. Resolving the Hubble tension with new early dark energy. Phys. Rev. D 2020, 102, 063527. [Google Scholar] [CrossRef]
  217. Niedermann, F.; Sloth, M.S. New early dark energy. Phys. Rev. D 2021, 103, L041303. [Google Scholar] [CrossRef]
  218. Freese, K.; Winkler, M.W. Chain early dark energy: A Proposal for solving the Hubble tension and explaining today’s dark energy. Phys. Rev. D 2021, 104, 083533. [Google Scholar] [CrossRef]
  219. 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]
  220. Ong, Y.C. An Effective Sign Switching Dark Energy: Lotka-Volterra Model of Two Interacting Fluids. arXiv 2022, arXiv:2212.04429. [Google Scholar]
  221. 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]
  222. 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]
  223. Yin, L. Reducing the H0 tension with exponential acoustic dark energy. Eur. Phys. J. C 2022, 82, 78. [Google Scholar] [CrossRef]
  224. Braglia, M.; Ballardini, M.; Emond, W.T.; Finelli, F.; Gümrükçüoǧlu, A.E.; Koyama, K.; Paoletti, D. Larger value for H0 by an evolving gravitational constant. Phys. Rev. D 2020, 102, 023529. [Google Scholar] [CrossRef]
  225. 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] [Green Version]
  226. Batista, R.C. A Short Review on Clustering Dark Energy. Universe 2021, 8, 22. [Google Scholar] [CrossRef]
  227. Heisenberg, L.; Villarrubia-Rojo, H.; Zosso, J. Simultaneously solving the H0 and σ8 tensions with late dark energy. arXiv 2022, arXiv:2201.11623. [Google Scholar] [CrossRef]
  228. Martinelli, M.; Tutusaus, I. CMB Tensions with Low-Redshift H0 and S8 Measurements: Impact of a Redshift-Dependent Type-Ia Supernovae Intrinsic Luminosity. Symmetry 2019, 11, 986. [Google Scholar] [CrossRef]
  229. Alestas, G.; Kazantzidis, L.; Perivolaropoulos, L. H0 tension, phantom dark energy, and cosmological parameter degeneracies. Phys. Rev. D 2020, 101, 123516. [Google Scholar] [CrossRef]
  230. D’Amico, G.; Senatore, L.; Zhang, P. Limits on wCDM from the EFTofLSS with the PyBird code. J. Cosmol. Astropart. Phys. 2021, 2021, 006. [Google Scholar] [CrossRef]
  231. Yang, W.; Pan, S.; Valentino, E.D.; Saridakis, E.N. Observational Constraints on Dynamical Dark Energy with Pivoting Redshift. Universe 2019, 5, 219. [Google Scholar] [CrossRef]
  232. Di Valentino, E.; Melchiorri, A.; Silk, J. Cosmological constraints in extended parameter space from the Planck 2018 Legacy release. J. Cosmol. Astropart. Phys. 2020, 2020, 013. [Google Scholar] [CrossRef]
  233. Vagnozzi, S.; Dhawan, S.; Gerbino, M.; Freese, K.; Goobar, A.; Mena, O. Constraints on the sum of the neutrino masses in dynamical dark energy models with w (z) ≥ − 1 are tighter than those obtained in Λ CDM. Phys. Rev. D 2018, 98, 083501. [Google Scholar] [CrossRef]
  234. Du, M.; Yang, W.; Xu, L.; Pan, S.; Mota, D.F. Future constraints on dynamical dark-energy using gravitational-wave standard sirens. Phys. Rev. D 2019, 100, 043535. [Google Scholar] [CrossRef]
  235. Yang, W.; Pan, S.; Di Valentino, E.; Saridakis, E.N.; Chakraborty, S. Observational constraints on one-parameter dynamical dark-energy parametrizations and the H0 tension. Phys. Rev. D 2019, 99, 043543. [Google Scholar] [CrossRef] [Green Version]
  236. Li, X.; Shafieloo, A.; Sahni, V.; Starobinsky, A.A. Revisiting Metastable Dark Energy and Tensions in the Estimation of Cosmological Parameters. Astroph. J. 2019, 887, 153. [Google Scholar] [CrossRef]
  237. Szydłowski, M.; Stachowski, A.; Urbanowski, K. The evolution of the FRW universe with decaying metastable dark energy—A dynamical system analysis. J. Cosmol. Astropart. Phys. 2020, 2020, 029. [Google Scholar] [CrossRef]
  238. Yang, W.; Di Valentino, E.; Pan, S.; Basilakos, S.; Paliathanasis, A. Metastable dark energy models in light of Planck 2018 data: Alleviating the H0 tension. Phys. Rev. D 2020, 102, 063503. [Google Scholar] [CrossRef]
  239. Di Valentino, E.; Mukherjee, A.; Sen, A.A. Dark Energy with Phantom Crossing and the H0 Tension. Entropy 2021, 23, 404. [Google Scholar] [CrossRef] [PubMed]
  240. Benevento, G.; Hu, W.; Raveri, M. Can late dark energy transitions raise the Hubble constant? Phys. Rev. D 2020, 101, 103517. [Google Scholar] [CrossRef]
  241. Alestas, G.; Kazantzidis, L.; Perivolaropoulos, L. w -M phantom transition at zt < 0.1 as a resolution of the Hubble tension. Phys. Rev. D 2021, 103, 083517. [Google Scholar] [CrossRef]
  242. Solà Peracaula, J.; Gómez-Valent, A.; de Cruz Pérez, J.; Moreno-Pulido, C. Running vacuum against the H0 and σ8 tensions. Europhys. Lett. 2021, 134, 19001. [Google Scholar] [CrossRef]
  243. Mavromatos, N.E.; Solà Peracaula, J. Stringy-running-vacuum-model inflation: From primordial gravitational waves and stiff axion matter to dynamical dark energy. Eur. Phys. J. Spec. Top. 2021, 230, 2077–2110. [Google Scholar] [CrossRef]
  244. Keeley, R.E.; Joudaki, S.; Kaplinghat, M.; Kirkby, D. Implications of a transition in the dark energy equation of state for the H0 and σ8 tensions. J. Cosmol. Astropart. Phys. 2019, 2019, 035. [Google Scholar] [CrossRef]
  245. Dutta, K.; Roy, A.; Ruchika; Sen, A.A.; Sheikh-Jabbari, M.M. Beyond Λ CDM with low and high redshift data: Implications for dark energy. Gen. Relat. Gravit. 2020, 52, 15. [Google Scholar] [CrossRef] [Green Version]
  246. Yang, W.; Pan, S.; Di Valentino, E.; Paliathanasis, A.; Lu, J. Challenging bulk viscous unified scenarios with cosmological observations. Phys. Rev. D 2019, 100, 103518. [Google Scholar] [CrossRef]
  247. Elizalde, E.; Khurshudyan, M.; Odintsov, S.D.; Myrzakulov, R. Analysis of the H0 tension problem in the Universe with viscous dark fluid. Phys. Rev. D 2020, 102, 123501. [Google Scholar] [CrossRef]
  248. da Silva, W.J.C.; Silva, R. Growth of matter perturbations in the extended viscous dark energy models. Eur. Phys. J. C 2021, 81, 403. [Google Scholar] [CrossRef]
  249. Guo, R.Y.; Zhang, J.F.; Zhang, X. Can the H0 tension be resolved in extensions to ΛCDM cosmology? J. Cosmol. Astropart. Phys. 2019, 2019, 054. [Google Scholar] [CrossRef]
  250. van Putten, M.H.P.M. Alleviating tension in ΛCDM and the local distance ladder from first principles with no free parameters. Mon. Not. R. Astron. Soc. 2020, 491, L6–L10. [Google Scholar] [CrossRef]
  251. Dai, W.M.; Ma, Y.Z.; He, H.J. Reconciling Hubble constant discrepancy from holographic dark energy. Phys. Rev. D 2020, 102, 121302. [Google Scholar] [CrossRef]
  252. Ó Colgáin, E.; Sheikh-Jabbari, M.M. A critique of holographic dark energy. Class. Quant. Grav. 2021, 38, 177001. [Google Scholar] [CrossRef]
  253. da Silva, W.J.C.; Silva, R. Cosmological perturbations in the Tsallis holographic dark energy scenarios. Eur. Phys. J. Plus 2021, 136, 543. [Google Scholar] [CrossRef]
  254. Ó Colgáin, E.; van Putten, M.H.P.M.; Yavartanoo, H. de Sitter Swampland, H0 tension & observation. Phys. Lett. B 2019, 793, 126–129. [Google Scholar] [CrossRef]
  255. Ó Colgáin, E.; Yavartanoo, H. Testing the Swampland: H0 tension. Phys. Lett. B 2019, 797, 134907. [Google Scholar] [CrossRef]
  256. Anchordoqui, L.A.; Antoniadis, I.; Lüst, D.; Soriano, J.F.; Taylor, T.R. H0 tension and the string swampland. Phys. Rev. D 2020, 101, 083532. [Google Scholar] [CrossRef]
  257. Agrawal, P.; Obied, G.; Vafa, C. H0 tension, swampland conjectures, and the epoch of fading dark matter. Phys. Rev. D 2021, 103, 043523. [Google Scholar] [CrossRef]
  258. Banerjee, A.; Cai, H.; Heisenberg, L.; Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Yang, T. Hubble sinks in the low-redshift swampland. Phys. Rev. D 2021, 103, L081305. [Google Scholar] [CrossRef]
  259. Miao, H.; Huang, Z. The H 0 Tension in Non-flat QCDM Cosmology. Astroph. J. 2018, 868, 20. [Google Scholar] [CrossRef]
  260. Di Valentino, E.; Ferreira, R.Z.; Visinelli, L.; Danielsson, U. Late time transitions in the quintessence field and the H0 tension. Phys. Dark Universe 2019, 26, 100385. [Google Scholar] [CrossRef]
  261. Bag, S.; Sahni, V.; Shafieloo, A.; Shtanov, Y. Phantom Braneworld and the Hubble Tension. Astroph. J. 2021, 923, 212. [Google Scholar] [CrossRef]
  262. Adler, S.L. Implications of a frame dependent dark energy for the spacetime metric, cosmography, and effective Hubble constant. Phys. Rev. D 2019, 100, 123503. [Google Scholar] [CrossRef]
  263. Cai, R.G.; Guo, Z.K.; Li, L.; Wang, S.J.; Yu, W.W. Chameleon dark energy can resolve the Hubble tension. Phys. Rev. D 2021, 103, L121302. [Google Scholar] [CrossRef]
  264. Karwal, T.; Raveri, M.; Jain, B.; Khoury, J.; Trodden, M. Chameleon early dark energy and the Hubble tension. Phys. Rev. D 2022, 105, 063535. [Google Scholar] [CrossRef]
  265. Li, X.; Shafieloo, A. A Simple Phenomenological Emergent Dark Energy Model can Resolve the Hubble Tension. Astroph. J. Lett. 2019, 883, L3. [Google Scholar] [CrossRef] [Green Version]
  266. Hernández-Almada, A.; Leon, G.; Magaña, J.; García-Aspeitia, M.A.; Motta, V. Generalized emergent dark energy: Observational Hubble data constraints and stability analysis. Mon. Not. R. Astron. Soc. 2020, 497, 1590–1602. [Google Scholar] [CrossRef]
  267. Yang, W.; Di Valentino, E.; Pan, S.; Mena, O. Emergent Dark Energy, neutrinos and cosmological tensions. Phys. Dark Universe 2021, 31, 100762. [Google Scholar] [CrossRef]
  268. Li, X.; Shafieloo, A. Evidence for Emergent Dark Energy. Astroph. J. 2020, 902, 58. [Google Scholar] [CrossRef]
  269. Yang, W.; Di Valentino, E.; Pan, S.; Shafieloo, A.; Li, X. Generalized emergent dark energy model and the Hubble constant tension. Phys. Rev. D 2021, 104, 063521. [Google Scholar] [CrossRef]
  270. Benaoum, H.B.; Yang, W.; Pan, S.; Di Valentino, E. Modified Emergent Dark Energy and its Astronomical Constraints. arXiv 2020, arXiv:2008.09098. [Google Scholar] [CrossRef]
  271. Di Valentino, E.; Linder, E.V.; Melchiorri, A. Vacuum phase transition solves the H0 tension. Phys. Rev. D 2018, 97, 043528. [Google Scholar] [CrossRef]
  272. Di Valentino, E.; Linder, E.V.; Melchiorri, A. H0 ex machina: Vacuum metamorphosis and beyond H0. Phys. Dark Universe 2020, 30, 100733. [Google Scholar] [CrossRef]
  273. Di Valentino, E.; Pan, S.; Yang, W.; Anchordoqui, L.A. Touch of neutrinos on the vacuum metamorphosis: Is the H0 solution back? Phys. Rev. D 2021, 103, 123527. [Google Scholar] [CrossRef]
  274. Carneiro, S.; de Holanda, P.C.; Pigozzo, C.; Sobreira, F. Is the H0 tension suggesting a fourth neutrino generation? Phys. Rev. D 2019, 100, 023505. [Google Scholar] [CrossRef]
  275. Gelmini, G.B.; Kawasaki, M.; Kusenko, A.; Murai, K.; Takhistov, V. Big bang nucleosynthesis constraints on sterile neutrino and lepton asymmetry of the Universe. J. Cosmol. Astropart. Phys. 2020, 2020, 051. [Google Scholar] [CrossRef]
  276. Barenboim, G.; Kinney, W.H.; Park, W.I. Flavor versus mass eigenstates in neutrino asymmetries: Implications for cosmology. Eur. Phys. J. C 2017, 77, 590. [Google Scholar] [CrossRef]
  277. D’Eramo, F.; Ferreira, R.Z.; Notari, A.; Bernal, J.L. Hot axions and the H0 tension. J. Cosmol. Astropart. Phys. 2018, 2018, 014. [Google Scholar] [CrossRef]
  278. Di Luzio, L.; Giannotti, M.; Nardi, E.; Visinelli, L. Corrigendum to “The landscape of QCD axion models”. Phys. Rep. 2020, 870, 1–117. [Google Scholar] [CrossRef]
  279. Xiao, L.; Zhang, L.; An, R.; Feng, C.; Wang, B. Fractional Dark Matter decay: Cosmological imprints and observational constraints. J. Cosmol. Astropart. Phys. 2020, 2020, 045. [Google Scholar] [CrossRef]
  280. Blinov, N.; Keith, C.; Hooper, D. Warm decaying dark matter and the hubble tension. J. Cosmol. Astropart. Phys. 2020, 2020, 005. [Google Scholar] [CrossRef]
  281. Haridasu, B.S.; Viel, M. Late-time decaying dark matter: Constraints and implications for the H0-tension. Mon. Not. R. Astron. Soc. 2020, 497, 1757–1764. [Google Scholar] [CrossRef]
  282. Pandey, K.L.; Karwal, T.; Das, S. Alleviating the H0 and σ8 anomalies with a decaying dark matter model. J. Cosmol. Astropart. Phys. 2020, 2020, 026. [Google Scholar] [CrossRef]
  283. Anchordoqui, L.A. Decaying dark matter, the H0 tension, and the lithium problem. Phys. Rev. D 2021, 103, 035025. [Google Scholar] [CrossRef]
  284. Davari, Z.; Khosravi, N. Can decaying dark matter scenarios alleviate both H0 and σ8 tensions? Mon. Not. R. Astron. Soc. 2022, 516, 4373–4382. [Google Scholar] [CrossRef]
  285. Hryczuk, A.; Jodłowski, K. Self-interacting dark matter from late decays and the H0 tension. Phys. Rev. D 2020, 102, 043024. [Google Scholar] [CrossRef]
  286. Jodlowski, K. Self-interacting dark matter from late decays and the H0 tension. In Proceedings of the European Physical Society Conference on High Energy Physics, Jointly organized by Universität Hamburg and the Research Center DESY, Online Conference, 23–30 July 2021; p. 115. [Google Scholar] [CrossRef]
  287. Vattis, K.; Koushiappas, S.M.; Loeb, A. Dark matter decaying in the late Universe can relieve the H0 tension. Phys. Rev. D 2019, 99, 121302. [Google Scholar] [CrossRef]
  288. Clark, S.J.; Vattis, K.; Koushiappas, S.M. Cosmological constraints on late-Universe decaying dark matter as a solution to the H0 tension. Phys. Rev. D 2021, 103, 043014. [Google Scholar] [CrossRef]
  289. Choi, G.; Suzuki, M.; Yanagida, T.T. Quintessence axion dark energy and a solution to the hubble tension. Phys. Lett. B 2020, 805, 135408. [Google Scholar] [CrossRef]
  290. Gu, Y.; Khlopov, M.; Wu, L.; Yang, J.M.; Zhu, B. Light gravitino dark matter: LHC searches and the Hubble tension. Phys. Rev. D 2020, 102, 115005. [Google Scholar] [CrossRef]
  291. Alcaniz, J.; Bernal, N.; Masiero, A.; Queiroz, F.S. Light dark matter: A common solution to the lithium and H0 problems. Phys. Lett. B 2021, 812, 136008. [Google Scholar] [CrossRef]
  292. Escudero, M.; Hooper, D.; Krnjaic, G.; Pierre, M. Cosmology With a Very Light Lμ-Lτ Gauge Boson. arXiv 2019, arXiv:1901.02010. [Google Scholar] [CrossRef]
  293. Desai, A.; Dienes, K.R.; Thomas, B. Constraining dark-matter ensembles with supernova data. Phys. Rev. D 2020, 101, 035031. [Google Scholar] [CrossRef]
  294. Choi, G.; Suzuki, M.; Yanagida, T.T. Degenerate Sub-keV fermion dark matter from a solution to the Hubble tension. Phys. Rev. D 2020, 101, 075031. [Google Scholar] [CrossRef]
  295. Di Valentino, E.; Bœhm, C.; Hivon, E.; Bouchet, F.R. Reducing the H0 and σ8 tensions with dark matter-neutrino interactions. Phys. Rev. D 2018, 97, 043513. [Google Scholar] [CrossRef] [Green Version]
  296. Stadler, J.; Bœhm, C.; Mena, O. Comprehensive study of neutrino-dark matter mixed damping. J. Cosmol. Astropart. Phys. 2019, 2019, 014. [Google Scholar] [CrossRef]
  297. Mosbech, M.R.; Boehm, C.; Hannestad, S.; Mena, O.; Stadler, J.; Wong, Y.Y.Y. The full Boltzmann hierarchy for dark matter-massive neutrino interactions. J. Cosmol. Astropart. Phys. 2021, 2021, 066. [Google Scholar] [CrossRef]
  298. Arias-Aragón, F.; Fernández-Martínez, E.; González-López, M.; Merlo, L. Neutrino masses and Hubble tension via a Majoron in MFV. Eur. Phys. J. C 2021, 81, 28. [Google Scholar] [CrossRef]
  299. Escudero, M.; Witte, S.J. The hubble tension as a hint of leptogenesis and neutrino mass generation. Eur. Phys. J. C 2021, 81, 515. [Google Scholar] [CrossRef]
  300. Huang, G.y.; Rodejohann, W. Solving the Hubble tension without spoiling big bang nucleosynthesis. Phys. Rev. D 2021, 103, 123007. [Google Scholar] [CrossRef]
  301. Boyarsky, A.; Ovchynnikov, M.; Sabti, N.; Syvolap, V. When feebly interacting massive particles decay into neutrinos: The Neff story. Phys. Rev. D 2021, 104, 035006. [Google Scholar] [CrossRef]
  302. Archidiacono, M.; Hooper, D.C.; Murgia, R.; Bohr, S.; Lesgourgues, J.; Viel, M. Constraining Dark Matter-Dark Radiation interactions with CMB, BAO, and Lyman-α. J. Cosmol. Astropart. Phys. 2019, 2019, 055. [Google Scholar] [CrossRef]
  303. Becker, N.; Hooper, D.C.; Kahlhoefer, F.; Lesgourgues, J.; Schöneberg, N. Cosmological constraints on multi-interacting dark matter. J. Cosmol. Astropart. Phys. 2021, 2021, 019. [Google Scholar] [CrossRef]
  304. Buen-Abad, M.A.; Emami, R.; Schmaltz, M. Cannibal dark matter and large scale structure. Phys. Rev. D 2018, 98, 083517. [Google Scholar] [CrossRef]
  305. Freese, K.; Sfakianakis, E.I.; Stengel, P.; Visinelli, L. The Higgs boson can delay reheating after inflation. J. Cosmol. Astropart. Phys. 2018, 2018, 067. [Google Scholar] [CrossRef] [Green Version]
  306. Anchordoqui, L.A.; Bergliaffa, S.E.P. Hot thermal universe endowed with massive dark vector fields and the Hubble tension. Phys. Rev. D 2019, 100, 123525. [Google Scholar] [CrossRef]
  307. Flambaum, V.V.; Samsonov, I.B. Ultralight dark photon as a model for early Universe dark matter. Phys. Rev. D 2019, 100, 063541. [Google Scholar] [CrossRef]
  308. Carr, B.; Kühnel, F. Primordial Black Holes as Dark Matter: Recent Developments. Annu. Rev. Nucl. Part. S. 2020, 70, 355–394. [Google Scholar] [CrossRef]
  309. Flores, M.M.; Kusenko, A. Primordial Black Holes from Long-Range Scalar Forces and Scalar Radiative Cooling. Phys. Rev. Lett. 2021, 126, 041101. [Google Scholar] [CrossRef] [PubMed]
  310. Green, A.M.; Kavanagh, B.J. Primordial black holes as a dark matter candidate. J. Phys. Nucl. Phys. 2021, 48, 043001. [Google Scholar] [CrossRef]
  311. Artymowski, M.; Ben-Dayan, I.; Kumar, U. Emergent dark energy from unparticles. Phys. Rev. D 2021, 103, L121303. [Google Scholar] [CrossRef]
  312. Yao, Y.H.; Meng, X.H. Can interacting dark energy with dynamical coupling resolve the Hubble tension. arXiv 2022, arXiv:2207.05955. [Google Scholar] [CrossRef]
  313. Kumar, S.; Nunes, R.C.; Yadav, S.K. Dark sector interaction: A remedy of the tensions between CMB and LSS data. Eur. Phys. J. C 2019, 79, 576. [Google Scholar] [CrossRef]
  314. Yang, W.; Pan, S.; Nunes, R.C.; Mota, D.F. Dark calling dark: Interaction in the dark sector in presence of neutrino properties after Planck CMB final release. J. Cosmol. Astropart. Phys. 2020, 2020, 008. [Google Scholar] [CrossRef]
  315. Di Valentino, E.; 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 Universe 2020, 30, 100666. [Google Scholar] [CrossRef]
  316. Gao, L.Y.; Zhao, Z.W.; Xue, S.S.; Zhang, X. Relieving the H0 tension with a new interacting dark energy model. J. Cosmol. Astropart. Phys. 2021, 2021, 005. [Google Scholar] [CrossRef]
  317. Wang, L.F.; Zhang, J.H.; He, D.Z.; Zhang, J.F.; Zhang, X. Constraints on interacting dark energy models from time-delay cosmography with seven lensed quasars. Mon. Not. R. Astron. Soc. 2022, 514, 1433–1440. [Google Scholar] [CrossRef]
  318. 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]
  319. Yang, W.; Pan, S.; Di Valentino, E.; Nunes, R.C.; Vagnozzi, S.; Mota, D.F. Tale of stable interacting dark energy, observational signatures, and the H0 tension. J. Cosmol. Astropart. Phys. 2018, 2018, 019. [Google Scholar] [CrossRef]
  320. Di Valentino, E.; Melchiorri, A.; Mena, O.; Vagnozzi, S. Nonminimal dark sector physics and cosmological tensions. Phys. Rev. D 2020, 101, 063502. [Google Scholar] [CrossRef]
  321. Banihashemi, A.; Khosravi, N.; Shirazi, A.H. Ginzburg-Landau theory of dark energy: A framework to study both temporal and spatial cosmological tensions simultaneously. Phys. Rev. D 2019, 99, 083509. [Google Scholar] [CrossRef]
  322. Pan, S.; Yang, W.; Paliathanasis, A. Non-linear interacting cosmological models after Planck 2018 legacy release and the H0 tension. Mon. Not. R. Astron. Soc. 2020, 493, 3114–3131. [Google Scholar] [CrossRef]
  323. Pan, S.; Yang, W.; Di Valentino, E.; Saridakis, E.N.; Chakraborty, S. Interacting scenarios with dynamical dark energy: Observational constraints and alleviation of the H0 tension. Phys. Rev. D 2019, 100, 103520. [Google Scholar] [CrossRef]
  324. Yang, W.; Di Valentino, E.; Mena, O.; Pan, S. Dynamical dark sectors and neutrino masses and abundances. Phys. Rev. D 2020, 102, 023535. [Google Scholar] [CrossRef]
  325. Pan, S.; Yang, W.; Singha, C.; Saridakis, E.N. Observational constraints on sign-changeable interaction models and alleviation of the H0 tension. Phys. Rev. D 2019, 100, 083539. [Google Scholar] [CrossRef] [Green Version]
  326. Yang, W.; Pan, S.; Xu, L.; Mota, D.F. Effects of anisotropic stress in interacting dark matter—Dark energy scenarios. Mon. Not. R. Astron. Soc. 2019, 482, 1858–1871. [Google Scholar] [CrossRef]
  327. Amirhashchi, H.; Yadav, A.K. Interacting Dark Sectors in Anisotropic Universe: Observational Constraints and H0 Tension. arXiv 2020, arXiv:2001.03775. [Google Scholar] [CrossRef]
  328. Bégué, D.; Stahl, C.; Xue, S.S. A model of interacting dark fluids tested with supernovae and Baryon Acoustic Oscillations data. Nucl. Phys. B 2019, 940, 312–320. [Google Scholar] [CrossRef]
  329. Panpanich, S.; Burikham, P.; Ponglertsakul, S.; Tannukij, L. Resolving Hubble tension with quintom dark energy model. Chin. Phys. C 2021, 45, 015108. [Google Scholar] [CrossRef]
  330. Jesus, J.F.; Escobal, A.A.; Benndorf, D.; Pereira, S.H. Can dark matter-dark energy interaction alleviate the Cosmic Coincidence Problem? arXiv 2020, arXiv:2012.07494. [Google Scholar] [CrossRef]
  331. Harko, T.; Asadi, K.; Moshafi, H.; Sheikhahmadi, H. Observational constraints on the interacting dark energy—Dark matter (IDM) cosmological models. Phys. Dark Universe 2022, 38, 101131. [Google Scholar] [CrossRef]
  332. Stadler, J.; Bœhm, C. Constraints on γ-CDM interactions matching the Planck data precision. J. Cosmol. Astropart. Phys. 2018, 2018, 009. [Google Scholar] [CrossRef]
  333. Yadav, S.K. Constraints on dark matter-photon coupling in the presence of time-varying dark energy. Mod. Phys. Lett. A 2020, 35, 1950358. [Google Scholar] [CrossRef]
  334. Barkana, R. Possible interaction between baryons and dark-matter particles revealed by the first stars. Nature 2018, 555, 71–74. [Google Scholar] [CrossRef]
  335. Slatyer, T.R.; Wu, C.L. Early-Universe constraints on dark matter-baryon scattering and their implications for a global 21 cm signal. Phys. Rev. D 2018, 98, 023013. [Google Scholar] [CrossRef] [Green Version]
  336. Beltrán Jiménez, J.; Bettoni, D.; Figueruelo, D.; Teppa Pannia, F.A. On cosmological signatures of baryons-dark energy elastic couplings. J. Cosmol. Astropart. Phys. 2020, 2020, 020. [Google Scholar] [CrossRef]
  337. Vagnozzi, S.; Visinelli, L.; Mena, O.; Mota, D.F. Do we have any hope of detecting scattering between dark energy and baryons through cosmology? Mon. Not. R. Astron. Soc. 2020, 493, 1139–1152. [Google Scholar] [CrossRef]
  338. Blinov, N.; Kelly, K.J.; Krnjaic, G.; McDermott, S.D. Constraining the Self-Interacting Neutrino Interpretation of the Hubble Tension. Phys. Rev. Lett. 2019, 123, 191102. [Google Scholar] [CrossRef] [PubMed]
  339. He, H.J.; Ma, Y.Z.; Zheng, J. Resolving Hubble tension by self-interacting neutrinos with Dirac seesaw. J. Cosmol. Astropart. Phys. 2020, 2020, 003. [Google Scholar] [CrossRef]
  340. Lyu, K.F.; Stamou, E.; Wang, L.T. Self-interacting neutrinos: Solution to Hubble tension versus experimental constraints. Phys. Rev. D 2021, 103, 015004. [Google Scholar] [CrossRef]
  341. Kreisch, C.D.; Cyr-Racine, F.Y.; Doré, O. Neutrino puzzle: Anomalies, interactions, and cosmological tensions. Phys. Rev. D 2020, 101, 123505. [Google Scholar] [CrossRef]
  342. Brinckmann, T.; Chang, J.H.; LoVerde, M. Self-interacting neutrinos, the Hubble parameter tension, and the cosmic microwave background. Phys. Rev. D 2021, 104, 063523. [Google Scholar] [CrossRef]
  343. Berryman, J.M.; Blinov, N.; Brdar, V.; Brinckmann, T.; Bustamante, M.; Cyr-Racine, F.Y.; Das, A.; de Gouvêa, A.; Denton, P.B.; Bhupal Dev, P.S.; et al. Neutrino Self-Interactions: A White Paper. arXiv 2022, arXiv:2203.01955. [Google Scholar]
  344. Archidiacono, M.; Gariazzo, S.; Giunti, C.; Hannestad, S.; Tram, T. Sterile neutrino self-interactions: H0 tension and short-baseline anomalies. J. Cosmol. Astropart. Phys. 2020, 2020, 029. [Google Scholar] [CrossRef]
  345. Ghosh, S.; Khatri, R.; Roy, T.S. Dark neutrino interactions make gravitational waves blue. Phys. Rev. D 2018, 97, 063529. [Google Scholar] [CrossRef]
  346. Ghosh, S.; Khatri, R.; Roy, T.S. Can dark neutrino interactions phase out the Hubble tension? Phys. Rev. D 2020, 102, 123544. [Google Scholar] [CrossRef]
  347. Yang, W.; Pan, S.; Vagnozzi, S.; Di Valentino, E.; Mota, D.F.; Capozziello, S. Dawn of the dark: Unified dark sectors and the EDGES Cosmic Dawn 21-cm signal. J. Cosmol. Astropart. Phys. 2019, 2019, 044. [Google Scholar] [CrossRef]
  348. Yang, W.; Pan, S.; Paliathanasis, A.; Ghosh, S.; Wu, Y. Observational constraints of a new unified dark fluid and the H0 tension. Mon. Not. R. Astron. Soc. 2019, 490, 2071–2085. [Google Scholar] [CrossRef]
  349. Benetti, M.; Borges, H.; Pigozzo, C.; Carneiro, S.; Alcaniz, J. Dark sector interactions and the curvature of the Universe in light of Planck’s 2018 data. arXiv 2021, arXiv:2102.10123. [Google Scholar] [CrossRef]
  350. Gurzadyan, V.G.; Stepanian, A. H0 tension: Clue to common nature of dark sector? Eur. Phys. J. C 2019, 79, 568. [Google Scholar] [CrossRef]
  351. Gurzadyan, V.G.; Stepanian, A. Hubble tension vs two flows. Eur. Phys. J. Plus 2021, 136, 235. [Google Scholar] [CrossRef]
  352. Quiros, I. Selected topics in scalar-tensor theories and beyond. Int. J. Mod. Phys. D 2019, 28, 1930012. [Google Scholar] [CrossRef]
  353. D’Agostino, R.; Nunes, R.C. Measurements of H0 in modified gravity theories: The role of lensed quasars in the late-time Universe. Phys. Rev. D 2020, 101, 103505. [Google Scholar] [CrossRef]
  354. Odintsov, S.D.; Sáez-Chillón Gómez, D.; Sharov, G.S. Analyzing the H0 tension in F(R) gravity models. Nuclear Physics B 2021, 966, 115377. [Google Scholar] [CrossRef]
  355. Wang, D. Can f(R) gravity relieve H0 and σ8 tensions? Eur. Phys. J. C 2021, 81, 482. [Google Scholar] [CrossRef]
  356. Schiavone, T.; Montani, G. f(R) gravity in the Jordan Frame as a Paradigm for the Hubble Tension. arXiv 2022, arXiv:2211.16737. [Google Scholar]
  357. Nunes, R.C. Structure formation in f(T) gravity and a solution for H0 tension. J. Cosmol. Astropart. Phys. 2018, 2018, 052. [Google Scholar] [CrossRef] [Green Version]
  358. Yan, S.F.; Zhang, P.; Chen, J.W.; Zhang, X.Z.; Cai, Y.F.; Saridakis, E.N. Interpreting cosmological tensions from the effective field theory of torsional gravity. Phys. Rev. D 2020, 101, 121301. [Google Scholar] [CrossRef]
  359. Wang, D.; Mota, D. Can f (T) gravity resolve the H0 tension? Phys. Rev. D 2020, 102, 063530. [Google Scholar] [CrossRef]
  360. Aljaf, M.; Elizalde, E.; Khurshudyan, M.; Myrzakulov, K.; Zhadyranova, A. Solving the H0 tension in f(T) Gravity through Bayesian Machine Learning. arXiv 2022, arXiv:2205.06252. [Google Scholar]
  361. Escamilla-Rivera, C.; Said, J.L. Cosmological viable models in f(T, B) theory as solutions to the H0 tension. Class. Quant. Grav. 2020, 37, 165002. [Google Scholar] [CrossRef]
  362. Paliathanasis, A. Minisuperspace Quantization of f(T, B) Cosmology. Universe 2021, 7, 150. [Google Scholar] [CrossRef]
  363. Albuquerque, I.S.; Frusciante, N. A designer approach to f(Q) gravity and cosmological implications. Phys. Dark Universe 2022, 35, 100980. [Google Scholar] [CrossRef]
  364. Koussour, M.; Pacif, S.K.J.; Bennai, M.; Sahoo, P.K. A new parametrization of Hubble parameter in f(Q) gravity. arXiv 2022, arXiv:2208.04723. [Google Scholar]
  365. Joudaki, S.; Ferreira, P.G.; Lima, N.A.; Winther, H.A. Testing gravity on cosmic scales: A case study of Jordan-Brans-Dicke theory. Phys. Rev. D 2022, 105, 043522. [Google Scholar] [CrossRef]
  366. Solà Peracaula, J.; Gómez-Valent, A.; de Cruz Pérez, J.; Moreno-Pulido, C. Brans-Dicke Gravity with a Cosmological Constant Smoothes Out ΛCDM Tensions. Astroph. J. Lett. 2019, 886, L6. [Google Scholar] [CrossRef]
  367. Peracaula, J.S.; Gómez-Valent, A.; de Cruz Pérez, J.; Moreno-Pulido, C. Brans-Dicke cosmology with a Λ-term: A possible solution to ΛCDM tensions. Class. Quant. Grav. 2020, 37, 245003. [Google Scholar] [CrossRef]
  368. Ballardini, M.; Braglia, M.; Finelli, F.; Paoletti, D.; Starobinsky, A.A.; Umiltà, C. Scalar-tensor theories of gravity, neutrino physics, and the H0 tension. J. Cosmol. Astropart. Phys. 2020, 2020, 044. [Google Scholar] [CrossRef]
  369. Ballesteros, G.; Notari, A.; Rompineve, F. Δ GN vs. Δ Neff. J. Cosmol. Astropart. Phys. 2020, 2020, 024. [Google Scholar] [CrossRef]
  370. Abadi, T.; Kovetz, E.D. Can conformally coupled modified gravity solve the Hubble tension? Phys. Rev. D 2021, 103, 023530. [Google Scholar] [CrossRef]
  371. Braglia, M.; Ballardini, M.; Finelli, F.; Koyama, K. Early modified gravity in light of the H0 tension and LSS data. Phys. Rev. D 2021, 103, 043528. [Google Scholar] [CrossRef]
  372. Desmond, H.; Jain, B.; Sakstein, J. Erratum: Local resolution of the Hubble tension: The impact of screened fifth forces on the cosmic distance ladder [Phys. Rev. D 100, 043537 (2019)]. Phys. Rev. D 2020, 101, 129901. [Google Scholar] [CrossRef]
  373. Desmond, H.; Sakstein, J. Screened fifth forces lower the TRGB-calibrated Hubble constant too. Phys. Rev. D 2020, 102, 023007. [Google Scholar] [CrossRef]
  374. Khosravi, N. Über-gravity and the cosmological constant problem. Phys. Dark Universe 2018, 21, 21. [Google Scholar] [CrossRef]
  375. Khosravi, N.; Baghram, S.; Afshordi, N.; Altamirano, N. H0 tension as a hint for a transition in gravitational theory. Phys. Rev. D 2019, 99, 103526. [Google Scholar] [CrossRef]
  376. Zumalacárregui, M. Gravity in the era of equality: Towards solutions to the Hubble problem without fine-tuned initial conditions. Phys. Rev. D 2020, 102, 023523. [Google Scholar] [CrossRef]
  377. Heisenberg, L.; Villarrubia-Rojo, H. Proca in the sky. J. Cosmol. Astropart. Phys. 2021, 2021, 032. [Google Scholar] [CrossRef]
  378. Belgacem, E.; Dirian, Y.; Foffa, S.; Maggiore, M. Nonlocal gravity. Conceptual aspects and cosmological predictions. J. Cosmol. Astropart. Phys. 2018, 2018, 002. [Google Scholar] [CrossRef]
  379. Belgacem, E.; Dirian, Y.; Finke, A.; Foffa, S.; Maggiore, M. Gravity in the infrared and effective nonlocal models. J. Cosmol. Astropart. Phys. 2020, 2020, 010. [Google Scholar] [CrossRef]
  380. Linares Cedeño, F.X.; Nucamendi, U. Revisiting cosmological diffusion models in Unimodular Gravity and the H0 tension. Phys. Dark Universe 2021, 32, 100807. [Google Scholar] [CrossRef]
  381. Alvarez, P.D.; Koch, B.; Laporte, C.; Rincón, Á. Can scale-dependent cosmology alleviate the H0 tension? J. Cosmol. Astropart. Phys. 2021, 2021, 019. [Google Scholar] [CrossRef]
  382. De Felice, A.; Mukohyama, S.; Pookkillath, M.C. Addressing H0 tension by means of VCDM. Phys. Lett. B 2021, 816, 136201. [Google Scholar] [CrossRef]
  383. Ganz, A.; Martens, P.; Mukohyama, S.; Namba, R. Bouncing Cosmology in VCDM. arXiv 2022, arXiv:2212.13561. [Google Scholar]
  384. Ratra, B. Tilted spatially nonflat inflation. Phys. Rev. D 2022, 106, 123524. [Google Scholar] [CrossRef]
  385. Di Valentino, E.; Mersini-Houghton, L. Testing predictions of the quantum landscape multiverse 2: The exponential inflationary potential. J. Cosmol. Astropart. Phys. 2017, 2017, 020. [Google Scholar] [CrossRef]
  386. Guo, R.Y.; Zhang, X. Constraints on inflation revisited: An analysis including the latest local measurement of the Hubble constant. Eur. Phys. J. C 2017, 77, 882. [Google Scholar] [CrossRef]
  387. Keeley, R.E.; Shafieloo, A.; Hazra, D.K.; Souradeep, T. Inflation wars: A new hope. J. Cosmol. Astropart. Phys. 2020, 2020, 055. [Google Scholar] [CrossRef]
  388. Liu, M.; Huang, Z. Band-limited Features in the Primordial Power Spectrum Do Not Resolve the Hubble Tension. Astroph. J. 2020, 897, 166. [Google Scholar] [CrossRef]
  389. Aresté Saló, L.; Benisty, D.; Guendelman, E.I.; Haro, J.d. Quintessential inflation and cosmological seesaw mechanism: Reheating and observational constraints. J. Cosmol. Astropart. Phys. 2021, 2021, 007. [Google Scholar] [CrossRef]
  390. Di Valentino, E.; Melchiorri, A.; Fantaye, Y.; Heavens, A. Bayesian evidence against the Harrison-Zel’dovich spectrum in tensions with cosmological data sets. Phys. Rev. D 2018, 98, 063508. [Google Scholar] [CrossRef] [Green Version]
  391. Chiang, C.T.; Slosar, A. Inferences of H0 in presence of a non-standard recombination. arXiv 2018, arXiv:1811.03624. [Google Scholar]
  392. Hart, L.; Chluba, J. New constraints on time-dependent variations of fundamental constants using Planck data. Mon. Not. R. Astron. Soc. 2018, 474, 1850–1861. [Google Scholar] [CrossRef]
  393. Hart, L.; Chluba, J. Updated fundamental constant constraints from Planck 2018 data and possible relations to the Hubble tension. Mon. Not. R. Astron. Soc. 2020, 493, 3255–3263. [Google Scholar] [CrossRef]
  394. Sekiguchi, T.; Takahashi, T. Early recombination as a solution to the H0 tension. Phys. Rev. D 2021, 103, 083507. [Google Scholar] [CrossRef]
  395. Fung, L.W.H.; Li, L.; Liu, T.; Luu, H.N.; Qiu, Y.C.; Tye, S.H.H. Axi-Higgs cosmology. J. Cosmol. Astropart. Phys. 2021, 2021, 057. [Google Scholar] [CrossRef]
  396. Jedamzik, K.; Saveliev, A. Stringent Limit on Primordial Magnetic Fields from the Cosmic Microwave Background Radiation. Phys. Rev. Lett. 2019, 123, 021301. [Google Scholar] [CrossRef]
  397. Jedamzik, K.; Pogosian, L. Relieving the Hubble Tension with Primordial Magnetic Fields. Phys. Rev. Lett. 2020, 125, 181302. [Google Scholar] [CrossRef] [PubMed]
  398. Banihashemi, A.; Khosravi, N.; Shirazi, A.H. Phase transition in the dark sector as a proposal to lessen cosmological tensions. Phys. Rev. D 2020, 101, 123521. [Google Scholar] [CrossRef]
  399. Banihashemi, A.; Khosravi, N.; Shafieloo, A. Dark energy as a critical phenomenon: A hint from Hubble tension. J. Cosmol. Astropart. Phys. 2021, 2021, 003. [Google Scholar] [CrossRef]
  400. Kasai, M.; Futamase, T. A possible solution to the Hubble constant discrepancy: Cosmology where the local volume expansion is driven by the domain average density. Prog. Theor. Exp. Phys. 2019, 2019, 073E01. [Google Scholar] [CrossRef]
  401. Yusofi, E.; Ramzanpour, M.A. Cosmological Constant Problem and H0 Tension in Void-dominated Cosmology. arXiv 2022, arXiv:2204.12180. [Google Scholar]
  402. Akarsu, Ö.; Kumar, S.; Sharma, S.; Tedesco, L. Constraints on a Bianchi type I spacetime extension of the standard Λ CDM model. Phys. Rev. D 2019, 100, 023532. [Google Scholar] [CrossRef] [Green Version]
  403. Bolejko, K. Emerging spatial curvature can resolve the tension between high-redshift CMB and low-redshift distance ladder measurements of the Hubble constant. Phys. Rev. D 2018, 97, 103529. [Google Scholar] [CrossRef]
  404. Macpherson, H.J.; Lasky, P.D.; Price, D.J. The Trouble with Hubble: Local versus Global Expansion Rates in Inhomogeneous Cosmological Simulations with Numerical Relativity. Astroph. J. Lett. 2018, 865, L4. [Google Scholar] [CrossRef]
  405. Heinesen, A.; Buchert, T. Solving the curvature and Hubble parameter inconsistencies through structure formation-induced curvature. Class. Quant. Grav. 2020, 37, 164001. [Google Scholar] [CrossRef]
  406. Ivanov, M.M.; Ali-Haïmoud, Y.; Lesgourgues, J. H0 tension or T0 tension? Phys. Rev. D 2020, 102, 063515. [Google Scholar] [CrossRef]
  407. Bengaly, C.A.P.; Gonzalez, J.E.; Alcaniz, J.S. Is there evidence for a hotter Universe? Eur. Phys. J. C 2020, 80, 936. [Google Scholar] [CrossRef]
  408. Bose, B.; Lombriser, L. Easing cosmic tensions with an open and hotter universe. Phys. Rev. D 2021, 103, L081304. [Google Scholar] [CrossRef]
  409. Adhikari, S.; Huterer, D. Super-CMB fluctuations and the Hubble tension. Phys. Dark Universe 2020, 28, 100539. [Google Scholar] [CrossRef]
  410. Capozziello, S.; Benetti, M.; Spallicci, A.D.A.M. Addressing the cosmological H0 tension by the Heisenberg uncertainty. Found. Phys. 2020, 50, 893–899. [Google Scholar] [CrossRef]
  411. Perez, A.; Sudarsky, D.; Wilson-Ewing, E. Resolving the H0 tension with diffusion. Gen. Relat. Gravit. 2021, 53, 7. [Google Scholar] [CrossRef]
  412. Berechya, D.; Leonhardt, U. Lifshitz cosmology: Quantum vacuum and Hubble tension. Mon. Not. R. Astron. Soc. 2021, 507, 3473–3485. [Google Scholar] [CrossRef]
  413. Ortiz, C. Surface tension: Accelerated expansion, coincidence problem & Hubble tension. Int. J. Mod. Phys. D 2020, 29, 2050115. [Google Scholar] [CrossRef]
  414. Vishwakarma, R.G. Resolving Hubble tension with the Milne model. Int. J. Mod. Phys. D 2020, 29, 2043025. [Google Scholar] [CrossRef]
  415. Milne, E.A. Relativity, gravitation and world-structure. Nature 1935, 135, 635–636. [Google Scholar]
  416. 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]
  417. Dainotti, M.G.; De Simone, B.; Schiavone, T.; Montani, G.; Rinaldi, E.; Lambiase, G. On the Hubble Constant Tension in the SNe Ia Pantheon Sample. Astroph. J. 2021, 912, 150. [Google Scholar] [CrossRef]
  418. Marra, V.; Perivolaropoulos, L. Rapid transition of Geff at zt≃0.01 as a possible solution of the Hubble and growth tensions. Phys. Rev. D 2021, 104, L021303. [Google Scholar] [CrossRef]
  419. Fosalba, P.; Gaztañaga, E. Explaining cosmological anisotropy: Evidence for causal horizons from CMB data. Mon. Not. R. Astron. Soc. 2021, 504, 5840–5862. [Google Scholar] [CrossRef]
  420. Gaztañaga, E. The size of our causal Universe. Mon. Not. R. Astron. Soc. 2020, 494, 2766–2772. [Google Scholar] [CrossRef]
  421. 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. R. Astron. Soc. 2020, 499, 2845–2883. [Google Scholar] [CrossRef]
  422. Beltrán Jiménez, J.; Bettoni, D.; Brax, P. Screening away the H0 tension. Int. J. Mod. Phys. D 2020, 29, 2043010. [Google Scholar] [CrossRef]
  423. Jiménez, J.B.; Bettoni, D.; Brax, P. Charged dark matter and the H0 tension. Phys. Rev. D 2021, 103, 103505. [Google Scholar] [CrossRef]
  424. Cruz, M.; Lepe, S.; Soto, G.E. Phantom cosmologies from QCD ghost dark energy. Phys. Rev. D 2022, 106, 103508. [Google Scholar] [CrossRef]
  425. Sola, J.; Gomez-Valent, A.; de Cruz Perez, J.; Moreno-Pulido, C. Running vacuum against the H0 and σ8 tensions. arXiv 2021, arXiv:2102.12758. [Google Scholar]
  426. Pan, S.; Yang, W.; Di Valentino, E.; Shafieloo, A.; Chakraborty, S. Reconciling H0 tension in a six parameter space? J. Cosmol. Astropart. Phys. 2020, 2020, 062. [Google Scholar] [CrossRef]
  427. Liu, Z.; Miao, H. Update constraints on neutrino mass and mass hierarchy in light of dark energy models. Int. J. Mod. Phys. D 2020, 29, 2050088. [Google Scholar] [CrossRef]
  428. Di Valentino, E.; Gariazzo, S.; Giunti, C.; Mena, O.; Pan, S.; Yang, W. Minimal dark energy: Key to sterile neutrino and Hubble constant tensions? Phys. Rev. D 2022, 105, 103511. [Google Scholar] [CrossRef]
  429. Moshafi, H.; Firouzjahi, H.; Talebian, A. Multiple Transitions in Vacuum Dark Energy and H 0 Tension. Astroph. J. 2022, 940, 121. [Google Scholar] [CrossRef]
  430. Moshafi, H.; Baghram, S.; Khosravi, N. CMB lensing in a modified Λ CDM model in light of the H0 tension. Phys. Rev. D 2021, 104, 063506. [Google Scholar] [CrossRef]
  431. Yang, W.; Mukherjee, A.; Di Valentino, E.; Pan, S. Interacting dark energy with time varying equation of state and the H0 tension. Phys. Rev. D 2018, 98, 123527. [Google Scholar] [CrossRef]
  432. Carrilho, P.; Moretti, C.; Bose, B.; Markovič, K.; Pourtsidou, A. Interacting dark energy from redshift-space galaxy clustering. J. Cosmol. Astropart. Phys. 2021, 2021, 004. [Google Scholar] [CrossRef]
  433. Guo, R.Y.; Feng, L.; Yao, T.Y.; Chen, X.Y. Exploration of interacting dynamical dark energy model with interaction term including the equation-of-state parameter: Alleviation of the H0 tension. J. Cosmol. Astropart. Phys. 2021, 2021, 036. [Google Scholar] [CrossRef]
  434. Nunes, R.C.; Di Valentino, E. Dark sector interaction and the supernova absolute magnitude tension. Phys. Rev. D 2021, 104, 063529. [Google Scholar] [CrossRef]
  435. Chatzidakis, S.; Giacomini, A.; Leach, P.G.L.; Leon, G.; Paliathanasis, A.; Pan, S. Interacting dark energy in curved FLRW spacetime from Weyl Integrable Spacetime. J. High Energy Astrop. 2022, 36, 141–151. [Google Scholar] [CrossRef]
  436. Gariazzo, S.; Di Valentino, E.; Mena, O.; Nunes, R.C. Late-time interacting cosmologies and the Hubble constant tension. Phys. Rev. D 2022, 106, 023530. [Google Scholar] [CrossRef]
  437. Anchordoqui, L.A.; Barger, V.; Marfatia, D.; Soriano, J.F. Decay of multiple dark matter particles to dark radiation in different epochs does not alleviate the Hubble tension. Phys. Rev. D 2022, 105, 103512. [Google Scholar] [CrossRef]
  438. Vagnozzi, S.; Visinelli, L.; Brax, P.; Davis, A.C.; Sakstein, J. Direct detection of dark energy: The XENON1T excess and future prospects. Phys. Rev. D 2021, 104, 063023. [Google Scholar] [CrossRef]
  439. Benisty, D.; Davis, A.C. Dark energy interactions near the Galactic Center. Phys. Rev. D 2022, 105, 024052. [Google Scholar] [CrossRef]
  440. Lombriser, L. Consistency of the local Hubble constant with the cosmic microwave background. Phys. Lett. B 2020, 803, 135303. [Google Scholar] [CrossRef]
  441. Contarini, S.; Pisani, A.; Hamaus, N.; Marulli, F.; Moscardini, L.; Baldi, M. Voids fill us in on rising cosmology tensions. arXiv 2022, arXiv:2212.07438. [Google Scholar]
  442. Kazantzidis, L.; Perivolaropoulos, L. Hints of a local matter underdensity or modified gravity in the low z Pantheon data. Phys. Rev. D 2020, 102, 023520. [Google Scholar] [CrossRef]
  443. Alestas, G.; Perivolaropoulos, L. Late-time approaches to the Hubble tension deforming H(z), worsen the growth tension. Mon. Not. R. Astron. Soc. 2021, 504, 3956–3962. [Google Scholar] [CrossRef]
  444. Perivolaropoulos, L.; Skara, F. A Reanalysis of the Latest SH0ES Data for H0: Effects of New Degrees of Freedom on the Hubble Tension. Universe 2022, 8, 502. [Google Scholar] [CrossRef]
  445. Perivolaropoulos, L.; Skara, F. Gravitational transitions via the explicitly broken symmetron screening mechanism. Phys. Rev. D 2022, 106, 043528. [Google Scholar] [CrossRef]
  446. Dhawan, S.; Brout, D.; Scolnic, D.; Goobar, A.; Riess, A.G.; Miranda, V. Cosmological Model Insensitivity of Local H0 from the Cepheid Distance Ladder. Astroph. J. 2020, 894, 54. [Google Scholar] [CrossRef]
  447. Perivolaropoulos, L.; Skara, F. Hubble tension or a transition of the Cepheid SnIa calibrator parameters? Phys. Rev. D 2021, 104, 123511. [Google Scholar] [CrossRef]
  448. Mörtsell, E.; Goobar, A.; Johansson, J.; Dhawan, S. The Hubble Tension Revisited: Additional Local Distance Ladder Uncertainties. Astroph. J. 2022, 935, 58. [Google Scholar] [CrossRef]
  449. Smith, T.L.; Poulin, V.; Amin, M.A. Oscillating scalar fields and the Hubble tension: A resolution with novel signatures. Phys. Rev. D 2020, 101, 063523. [Google Scholar] [CrossRef]
  450. Sakstein, J.; Trodden, M. Early Dark Energy from Massive Neutrinos as a Natural Resolution of the Hubble Tension. Phys. Rev. Lett. 2020, 124, 161301. [Google Scholar] [CrossRef]
  451. Gogoi, A.; Kumar Sharma, R.; Chanda, P.; Das, S. Early Mass-varying Neutrino Dark Energy: Nugget Formation and Hubble Anomaly. Astroph. J. 2021, 915, 132. [Google Scholar] [CrossRef]
  452. Poulin, V.; Smith, T.L.; Bartlett, A. Dark energy at early times and ACT data: A larger Hubble constant without late-time priors. Phys. Rev. D 2021, 104, 123550. [Google Scholar] [CrossRef]
  453. Seto, O.; Toda, Y. Comparing early dark energy and extra radiation solutions to the Hubble tension with BBN. Phys. Rev. D 2021, 103, 123501. [Google Scholar] [CrossRef]
  454. Ye, G.; Zhang, J.; Piao, Y.S. Resolving both H0 and S8 tensions with AdS early dark energy and ultralight axion. arXiv, 2021; arXiv:2107.13391. [Google Scholar]
  455. Jiang, J.Q.; Piao, Y.S. Testing AdS early dark energy with Planck, SPTpol, and LSS data. Phys. Rev. D 2021, 104, 103524. [Google Scholar] [CrossRef]
  456. Niedermann, F.; Sloth, M.S. Hot new early dark energy: Towards a unified dark sector of neutrinos, dark energy and dark matter. Phys. Lett. B 2022, 835, 137555. [Google Scholar] [CrossRef]
  457. Smith, T.L.; Lucca, M.; Poulin, V.; Abellan, G.F.; Balkenhol, L.; Benabed, K.; Galli, S.; Murgia, R. Hints of early dark energy in Planck, SPT, and ACT data: New physics or systematics? Phys. Rev. D 2022, 106, 043526. [Google Scholar] [CrossRef]
  458. Fernandez-Martinez, E.; Pierre, M.; Pinsard, E.; Rosauro-Alcaraz, S. Inverse Seesaw, dark matter and the Hubble tension. Eur. Phys. J. C 2021, 81, 954. [Google Scholar] [CrossRef]
  459. Seto, O.; Toda, Y. Hubble tension in lepton asymmetric cosmology with an extra radiation. Phys. Rev. D 2021, 104, 063019. [Google Scholar] [CrossRef]
  460. Aboubrahim, A.; Klasen, M.; Nath, P. Analyzing the Hubble tension through hidden sector dynamics in the early universe. J. Cosmol. Astropart. Phys. 2022, 2022, 042. [Google Scholar] [CrossRef]
  461. Gu, Y.; Wu, L.; Zhu, B. Axion dark radiation: Hubble tension and the Hyper-Kamiokande neutrino experiment. Phys. Rev. D 2022, 105, 095008. [Google Scholar] [CrossRef]
  462. Aloni, D.; Berlin, A.; Joseph, M.; Schmaltz, M.; Weiner, N. A Step in understanding the Hubble tension. Phys. Rev. D 2022, 105, 123516. [Google Scholar] [CrossRef]
  463. Ghosh, S.; Kumar, S.; Tsai, Y. Free-streaming and coupled dark radiation isocurvature perturbations: Constraints and application to the Hubble tension. J. Cosmol. Astropart. Phys. 2022, 2022, 014. [Google Scholar] [CrossRef]
  464. Berbig, M.; Jana, S.; Trautner, A. The Hubble tension and a renormalizable model of gauged neutrino self-interactions. Phys. Rev. D 2020, 102, 115008. [Google Scholar] [CrossRef]
  465. Choudhury, S.R.; Hannestad, S.; Tram, T. Updated constraints on massive neutrino self-interactions from cosmology in light of the H0 tension. J. Cosmol. Astropart. Phys. 2021, 2021, 084. [Google Scholar] [CrossRef]
  466. Mazumdar, A.; Mohanty, S.; Parashari, P. Flavour specific neutrino self-interaction: H0 tension and IceCube. J. Cosmol. Astropart. Phys. 2022, 2022, 011. [Google Scholar] [CrossRef]
  467. Shimon, M. Possible resolution of the Hubble tension with Weyl invariant gravity. J. Cosmol. Astropart. Phys. 2022, 2022, 048. [Google Scholar] [CrossRef]
  468. Petronikolou, M.; Basilakos, S.; Saridakis, E.N. Alleviating H0 tension in Horndeski gravity. Phys. Rev. D 2022, 106, 124051. [Google Scholar] [CrossRef]
  469. 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]
  470. Akarsu, O.; Kumar, S.; Ozulker, E.; Vazquez, J.A.; Yadav, A. Relaxing cosmological tensions with a sign switching cosmological constant: Improved results with Planck, BAO and Pantheon data. arXiv 2022, arXiv:2211.05742. [Google Scholar]
  471. Vagnozzi, S. Consistency tests of Λ CDM from the early integrated Sachs-Wolfe effect: Implications for early-time new physics and the Hubble tension. Phys. Rev. D 2021, 104, 063524. [Google Scholar] [CrossRef]
  472. Blinov, N.; Krnjaic, G.; Li, S.W. Realistic model of dark atoms to resolve the Hubble tension. Phys. Rev. D 2022, 105, 095005. [Google Scholar] [CrossRef]
  473. Gough, M.P. Information Dark Energy Can Resolve the Hubble Tension and Is Falsifiable by Experiment. Entropy 2022, 24, 385. [Google Scholar] [CrossRef] [PubMed]
  474. Araki, T.; Asai, K.; Honda, K.; Kasuya, R.; Sato, J.; Shimomura, T.; Yang, M.J.S. Resolving the Hubble tension in a U(1)Lμ-Lτ model with the Majoron. Prog. Theor. Exp. Phys. 2021, 2021, 103B05. [Google Scholar] [CrossRef]
  475. 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. Rep. 2022, 984, 1–55. [Google Scholar] [CrossRef]
  476. Desmond, H.; Jain, B.; Sakstein, J. Local resolution of the Hubble tension: The impact of screened fifth forces on the cosmic distance ladder. Phys. Rev. D 2019, 100, 043537. [Google Scholar] [CrossRef]
  477. Poulin, V. How to Resolve the Hubble tension. In Proceedings of the H0 2020: Assessing Uncertainties in Hubble’s Constant across the Universe, Online Conference, 22–26 June 2020; p. 22. [Google Scholar] [CrossRef]
  478. Addison, G.E. High H0 Values from CMB E-mode Data: A Clue for Resolving the Hubble Tension? Astroph. J. Lett. 2021, 912, L1. [Google Scholar] [CrossRef]
  479. Thiele, L.; Guan, Y.; Hill, J.C.; Kosowsky, A.; Spergel, D.N. Can small-scale baryon inhomogeneities resolve the Hubble tension? An investigation with ACT DR4. Phys. Rev. D 2021, 104, 063535. [Google Scholar] [CrossRef]
  480. Jedamzik, K.; Pogosian, L.; Zhao, G.B. Why reducing the cosmic sound horizon alone can not fully resolve the Hubble tension. Commun. Phys. 2021, 4, 123. [Google Scholar] [CrossRef]
  481. Cai, R.G.; Guo, Z.K.; Wang, S.J.; Yu, W.W.; Zhou, Y. No-go guide for the Hubble tension: Late-time solutions. Phys. Rev. D 2022, 105, L021301. [Google Scholar] [CrossRef]
  482. Cai, R.G.; Guo, Z.K.; Wang, S.J.; Yu, W.W.; Zhou, Y. No-go guide for late-time solutions to the Hubble tension: Matter perturbations. Phys. Rev. D 2022, 106, 063519. [Google Scholar] [CrossRef]
  483. Escudero, H.G.; Kuo, J.L.; Keeley, R.E.; Abazajian, K.N. Early or phantom dark energy, self-interacting, extra, or massive neutrinos, primordial magnetic fields, or a curved universe: An exploration of possible solutions to the H0 and σ8 problems. Phys. Rev. D 2022, 106, 103517. [Google Scholar] [CrossRef]
  484. Wang, Y.Y.; Tang, S.P.; Li, X.Y.; Jin, Z.P.; Fan, Y.Z. Prospects of calibrating afterglow modeling of short GRBs with gravitational wave inclination angle measurements and resolving the Hubble tension with a GW-GRB association event. Phys. Rev. D 2022, 106, 023011. [Google Scholar] [CrossRef]
  485. Khodadi, M.; Schreck, M. Hubble tension as a guide for refining the early Universe: Cosmologies with explicit local Lorentz and diffeomorphism violation. arXiv 2023, arXiv:2301.03883. [Google Scholar] [CrossRef]
  486. 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]
  487. 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]
  488. Lloyd, N.M.; Petrosian, V. Synchrotron Radiation as the Source of Gamma-Ray Burst Spectra. Astroph. J. 2000, 543, 722–732. [Google Scholar] [CrossRef]
  489. Singal, J.; Petrosian, V.; Lawrence, A.; Stawarz, Ł. On the Radio and Optical Luminosity Evolution of Quasars. Astroph. J. 2011, 743, 104. [Google Scholar] [CrossRef]
  490. Dainotti, M.G.; Lenart, A.; Sarracino, G.; Nagataki, S.; Capozziello, S.; Fraija, N. The X-Ray Fundamental Plane of the Platinum Sample, the Kilonovae, and the SNe Ib/c Associated with GRBs. Astroph. J. 2020, 904, 97. [Google Scholar] [CrossRef]
  491. Ó Colgáin, E.; Sheikh-Jabbari, M.M.; Solomon, R.; Bargiacchi, G.; Capozziello, S.; Dainotti, M.G.; Stojkovic, D. Revealing intrinsic flat Λ CDM biases with standardizable candles. Phys. Rev. D 2022, 106, L041301. [Google Scholar] [CrossRef]
  492. Dainotti, M.G.; De 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]
  493. 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]
  494. Schwarz, G. Estimating the Dimension of a Model. Ann. Stat. 1978, 6, 461–464. [Google Scholar] [CrossRef]
  495. Akaike, H. A new look at the statistical model identification. IEEE Trans. Autom. Control 1974, 19, 716–723. [Google Scholar] [CrossRef]
  496. Wojtak, R.; Hjorth, J. Intrinsic tension in the supernova sector of the local Hubble constant measurement and its implications. Mon. Not. R. Astron. Soc. 2022, 515, 2790–2799. [Google Scholar] [CrossRef]
  497. Pedregosa, F.; Varoquaux, G.; Gramfort, A.; Michel, V.; Thirion, B.; Grisel, O.; Blondel, M.; Prettenhofer, P.; Weiss, R.; Dubourg, V.; et al. Scikit-learn: Machine Learning in Python. J. Mach. Learn. Res. 2011, 12, 2825–2830. [Google Scholar]
  498. 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, 2018, 051. [Google Scholar] [CrossRef]
  499. Yu, H.; Ratra, B.; Wang, F.Y. Hubble Parameter and Baryon Acoustic Oscillation Measurement Constraints on the Hubble Constant, the Deviation from the Spatially Flat ΛCDM Model, the Deceleration-Acceleration Transition Redshift, and Spatial Curvature. Astroph. J. 2018, 856, 3. [Google Scholar] [CrossRef] [Green Version]
  500. Liao, K.; Shafieloo, A.; Keeley, R.E.; Linder, E.V. A Model-independent Determination of the Hubble Constant from Lensed Quasars and Supernovae Using Gaussian Process Regression. Astroph. J. Lett. 2019, 886, L23. [Google Scholar] [CrossRef]
  501. Liao, K.; Shafieloo, A.; Keeley, R.E.; Linder, E.V. Determining Model-independent H0 and Consistency Tests. Astroph. J. Lett. 2020, 895, L29. [Google Scholar] [CrossRef]
  502. Salti, M.; Ciger, E.; Kangal, E.E.; Zengin, B. Data-driven predictive modeling of Hubble parameter. Phys. Scripta 2022, 97, 085011. [Google Scholar] [CrossRef]
  503. Melia, F.; Yennapureddy, M.K. Model selection using cosmic chronometers with Gaussian Processes. J. Cosmol. Astropart. Phys. 2018, 2018, 034. [Google Scholar] [CrossRef]
  504. Rasmussen, C.E.; Williams, C.K.I. Gaussian Processes for Machine Learning; MIT Press: Cambridge, MA, USA, 2006. [Google Scholar]
  505. Frazier, P.I. A Tutorial on Bayesian Optimization. arXiv 2018, arXiv:1807.02811. [Google Scholar]
  506. Schulz, E.; Speekenbrink, M.; Krause, A. A tutorial on Gaussian process regression: Modelling, exploring, and exploiting functions. J. Math. Psychol. 2018, 85, 1–16. [Google Scholar] [CrossRef]
  507. Hu, J.P.; Wang, F.Y. Revealing the late-time transition of H0: Relieve the Hubble crisis. Mon. Not. R. Astron. Soc. 2022, 517, 576–581. [Google Scholar] [CrossRef]
  508. Jia, X.D.; Hu, J.P.; Wang, F.Y. The evidence for a decreasing trend of Hubble constant. arXiv 2022, arXiv:2212.00238. [Google Scholar]
  509. Malekjani, M.; Mc Conville, R.; Ó Colgáin, E.; Pourojaghi, S.; Sheikh-Jabbari, M.M. Negative Dark Energy Density from High Redshift Pantheon+ Supernovae. arXiv 2023, arXiv:2301.12725. [Google Scholar]
  510. 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]
  511. Keeley, R.E.; Shafieloo, A. Ruling Out New Physics at Low Redshift as a solution to the H0 Tension. arXiv 2022, arXiv:2206.08440. [Google Scholar]
  512. Keenan, R.C.; Barger, A.J.; Cowie, L.L. Evidence for a ~300 Megaparsec Scale Under-density in the Local Galaxy Distribution. Astroph. J. 2013, 775, 62. [Google Scholar] [CrossRef]
  513. Wang, F.Y.; Dai, Z.G. Testing the local-void alternative to dark energy using galaxy pairs. Mon. Not. R. Astron. Soc. 2013, 432, 3025–3029. [Google Scholar] [CrossRef]
  514. 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]
  515. Yusofi, E.; Khanpour, M.; Khanpour, B.; Ramzanpour, M.A.; Mohsenzadeh, M. Surface tension of cosmic voids as a possible source for dark energy. Mon. Not. R. Astron. Soc. 2022, 511, L82–L86. [Google Scholar] [CrossRef]
  516. Benetti, M.; Capozziello, S.; Lambiase, G. Updating constraints on f(T) teleparallel cosmology and the consistency with big bang nucleosynthesis. Mon. Not. R. Astron. Soc. 2021, 500, 1795–1805. [Google Scholar] [CrossRef]
  517. Kenworthy, W.D.; Scolnic, D.; Riess, A. The Local Perspective on the Hubble Tension: Local Structure Does Not Impact Measurement of the Hubble Constant. Astroph. J. 2019, 875, 145. [Google Scholar] [CrossRef]
  518. Luković, V.V.; Haridasu, B.S.; Vittorio, N. Exploring the evidence for a large local void with supernovae Ia data. Mon. Not. R. Astron. Soc. 2020, 491, 2075–2087. [Google Scholar] [CrossRef]
  519. 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]
  520. Böhringer, H.; Chon, G.; Collins, C.A. Observational evidence for a local underdensity in the Universe and its effect on the measurement of the Hubble constant. Astron. Astrophys. 2020, 633, A19. [Google Scholar] [CrossRef]
  521. DES Collaboration. Dark Energy Survey Year 3 results: Imprints of cosmic voids and superclusters in the Planck CMB lensing map. Mon. Not. R. Astron. Soc. 2022, 515, 4417–4429. [Google Scholar] [CrossRef]
  522. Krishnan, C.; Mondol, R. H0 as a Universal FLRW Diagnostic. arXiv 2022, arXiv:2201.13384. [Google Scholar]
  523. Coleman, S. Fate of the false vacuum: Semiclassical theory. Phys. Rev. D 1977, 15, 2929–2936. [Google Scholar] [CrossRef]
  524. Callan, C.G., Jr.; Coleman, S. Fate of the false vacuum. II. First quantum corrections. Phys. Rev. D 1977, 16, 1762–1768. [Google Scholar] [CrossRef]
  525. Patwardhan, A.V.; Fuller, G.M. Late-time vacuum phase transitions: Connecting sub-eV scale physics with cosmological structure formation. Phys. Rev. D 2014, 90, 063009. [Google Scholar] [CrossRef] [Green Version]
  526. Oguri, M.; Marshall, P.J. Gravitationally lensed quasars and supernovae in future wide-field optical imaging surveys. Mon. Not. R. Astron. Soc. 2010, 405, 2579–2593. [Google Scholar] [CrossRef]
  527. Collett, T.E. The Population of Galaxy-Galaxy Strong Lenses in Forthcoming Optical Imaging Surveys. Astroph. J. 2015, 811, 20. [Google Scholar] [CrossRef]
  528. LSST Dark Energy Science Collaboration. Strongly lensed SNe Ia in the era of LSST: Observing cadence for lens discoveries and time-delay measurements. Astron. Astrophys. 2019, 631, A161. [Google Scholar] [CrossRef]
  529. Wei, J.; Cordier, B.; Antier, S.; Antilogus, P.; Atteia, J.L.; Bajat, A.; Basa, S.; Beckmann, V.; Bernardini, M.G.; Boissier, S.; et al. The Deep and Transient Universe in the SVOM Era: New Challenges and Opportunities—Scientific prospects of the SVOM mission. arXiv 2016, arXiv:1610.06892. [Google Scholar]
  530. Yuan, W.; Zhang, C.; Feng, H.; Zhang, S.N.; Ling, Z.X.; Zhao, D.; Deng, J.; Qiu, Y.; Osborne, J.P.; O’Brien, P.; et al. Einstein Probe—A small mission to monitor and explore the dynamic X-ray Universe. arXiv 2015, arXiv:1506.07735. [Google Scholar]
  531. Amati, L.; O’Brien, P.; Götz, D.; Bozzo, E.; Tenzer, C.; Frontera, F.; Ghirlanda, G.; Labanti, C.; Osborne, J.P.; Stratta, G.; et al. The THESEUS space mission concept: Science case, design and expected performances. Adv. Space Res. 2018, 62, 191–244. [Google Scholar] [CrossRef] [Green Version]
Figure 2. H 0 measurements in the late-time universe derived from other independent observations which include quasar lensing [37,143,165,166,167,168,169,170,171,172], FRB [152,153,156], GW [149,151,191,192], TRGB [157,158,159] and Megamaser [184]. The purple and blue regions correspond to the results of SH0ES [40] and Planck collaborations [3], respectively.
Figure 2. H 0 measurements in the late-time universe derived from other independent observations which include quasar lensing [37,143,165,166,167,168,169,170,171,172], FRB [152,153,156], GW [149,151,191,192], TRGB [157,158,159] and Megamaser [184]. The purple and blue regions correspond to the results of SH0ES [40] and Planck collaborations [3], respectively.
Universe 09 00094 g002
Figure 3. H 0 measurements from H0LiCOW. The trend of smaller H 0 value with increasing lens redshift has significance levels of 1.7 σ . (Source: Figure 5 in Millon et al. [143]).
Figure 3. H 0 measurements from H0LiCOW. The trend of smaller H 0 value with increasing lens redshift has significance levels of 1.7 σ . (Source: Figure 5 in Millon et al. [143]).
Universe 09 00094 g003
Figure 4. A similar H 0 descending trend obtained from the combined sample (including megamasers, CCs, SNe Ia and BAOs) and Pantheon sample. Left panel shows the result obtained from the combined sample. The statistical significance of descending trend is 2.1 σ . Right panel shows the result of Pantheon sample which can reduce the H 0 tension by 66%. (Source: left panel, Figure 2 in Krishnan et al. [487]; right panel, Figure 5 in Dainotti et al. [417]).
Figure 4. A similar H 0 descending trend obtained from the combined sample (including megamasers, CCs, SNe Ia and BAOs) and Pantheon sample. Left panel shows the result obtained from the combined sample. The statistical significance of descending trend is 2.1 σ . Right panel shows the result of Pantheon sample which can reduce the H 0 tension by 66%. (Source: left panel, Figure 2 in Krishnan et al. [487]; right panel, Figure 5 in Dainotti et al. [417]).
Universe 09 00094 g004
Figure 5. Smoothed H(z) function (blue solid line) with 2 σ errors (gray regions) obtained from the 36 H(z) data (31 CCs + 5 BAOs) employing GP method. GP regression is implemented by employing the package scikit-learn (https://scikit-learn.org) [497] in the Python environment. (Source: Figure 4 in Hu et al. [5].)
Figure 5. Smoothed H(z) function (blue solid line) with 2 σ errors (gray regions) obtained from the 36 H(z) data (31 CCs + 5 BAOs) employing GP method. GP regression is implemented by employing the package scikit-learn (https://scikit-learn.org) [497] in the Python environment. (Source: Figure 4 in Hu et al. [5].)
Universe 09 00094 g005
Figure 6. Predictions of H 0 ( z m a x ) adopting the Matérn kernel from the 36 H ( z ) data (31 CCs + 5 BAOs) binned by the cumulative method. H 0 ( z m a x ) is the H 0 value derived from a data-set with maximal redshift z max . Red points are the predictions of H 0 ( z m a x ) based on our analyses. The gray and purple regions correspond to the results of SH0ES and Planck collaborations. Blue dotted line (z = 0.49) is the transition redshift. We also show the H 0 results derived from quasar lens observations with green points in ( z l e n s , H 0 ( z l e n s ) ) coordinates. (Source: Figures 3 and 4 in Hu and Wang [507].)
Figure 6. Predictions of H 0 ( z m a x ) adopting the Matérn kernel from the 36 H ( z ) data (31 CCs + 5 BAOs) binned by the cumulative method. H 0 ( z m a x ) is the H 0 value derived from a data-set with maximal redshift z max . Red points are the predictions of H 0 ( z m a x ) based on our analyses. The gray and purple regions correspond to the results of SH0ES and Planck collaborations. Blue dotted line (z = 0.49) is the transition redshift. We also show the H 0 results derived from quasar lens observations with green points in ( z l e n s , H 0 ( z l e n s ) ) coordinates. (Source: Figures 3 and 4 in Hu and Wang [507].)
Universe 09 00094 g006
Table 1. Summary of the H 0 estimations derived from the quasar lensing systems.
Table 1. Summary of the H 0 estimations derived from the quasar lensing systems.
Lens Name z d z S H 0 (km/s/Mpc)Reference
B1608+6560.63041.394 71 . 0 3.3 + 2.9 [166,167]
RXJ1131-12310.2950.654 78 . 3 3.3 + 3.4 [165,168,169]
HE0435-12230.45461.693 71 . 7 4.5 + 4.8 [169,170]
SDSS 1206+43320.7451.789 68 . 9 5.1 + 5.4 [171]
WFI2033-47230.65751.662 71 . 6 4.9 + 3.8 [172]
PG1115+0800.3111.722 81 . 1 7.1 + 8.0 [169]
DES J0408-53540.5972.375 74 . 2 3.0 + 2.7 [173,174]
Table 2. The details of six galaxies. The H 0 estimations are obtained by assuming a fixed velocity uncertainty of 250 km/s. (Source: Tables 1 and 2 in Pesce et al. [148].)
Table 2. The details of six galaxies. The H 0 estimations are obtained by assuming a fixed velocity uncertainty of 250 km/s. (Source: Tables 1 and 2 in Pesce et al. [148].)
Galaxy NameDistance (Mpc)Velocity (km/s) H 0 (km/s/Mpc)Reference
UGC 3789 51 . 5 4.0 + 4.5 3319.9 ± 0.8 75 . 8 3.3 + 3.4 [145]
NGC 6264 132 . 1 17.0 + 21.0 10192.6 ± 0.873.8 3.2 + 3.2 [144]
NGC 6323 109 . 4 23.0 + 34.0 7801.5 ± 1.5 73 . 8 3.0 + 3.1 [146]
NGC 5765b 112 . 2 5.1 + 5.4 8525.7 ± 0.7 74 . 1 4.4 + 4.5 [182]
NGC 4258 7.58 ± 0.11 679.3 ± 0.4 73 . 6 3.0 + 3.1 [147]
CGCG 074-064 87 . 6 7.2 + 7.9 7172.2 ± 1.9 72 . 5 3.2 + 3.4 [183]
Table 3. H 0 measurements with the 68% confidence level derived from the recent observations.
Table 3. H 0 measurements with the 68% confidence level derived from the recent observations.
Observation H 0 (km/s/Mpc)ReferenceObservation H 0 (km/s/Mpc)Reference
Quasar lens 73 . 3 1.8 + 1.7 [37]FRB 62.3 ± 9.1 [152]
Quasar lens 74 . 0 1.8 + 1.7 [143]FRB 68 . 81 4.33 + 4.99 [153]
Quasar lens 74 . 2 1.6 + 1.6 [165]FRB 73 . 0 8.0 + 12.0 [154]
Megamaser 69.3 ± 4.2 [184]FRB70.6 ± 2.11[155]
Megamaser 73.9 ± 3.0 [148]FRB71.5 8.1 + 10.0 [156]
GW 74 . 0 8.0 + 16.0 [149]TRGB69.8 ± 0.8[157]
GW + EM 70 . 3 5.0 + 5.3 [151]TRGB69.6 ± 0.8[158]
GW 68 . 0 7.0 + 12.0 [191]TRGB69.8 ± 0.8[159]
GW 67 . 0 3.8 + 6.3 [192]
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

Hu, J.-P.; Wang, F.-Y. Hubble Tension: The Evidence of New Physics. Universe 2023, 9, 94. https://doi.org/10.3390/universe9020094

AMA Style

Hu J-P, Wang F-Y. Hubble Tension: The Evidence of New Physics. Universe. 2023; 9(2):94. https://doi.org/10.3390/universe9020094

Chicago/Turabian Style

Hu, Jian-Ping, and Fa-Yin Wang. 2023. "Hubble Tension: The Evidence of New Physics" Universe 9, no. 2: 94. https://doi.org/10.3390/universe9020094

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