Next Article in Journal
Neutrino Dynamics in a Quantum-Corrected Schwarzschild Spacetime
Next Article in Special Issue
Antistars or Antimatter Cores in Mirror Neutron Stars?
Previous Article in Journal
D-Branes in Para-Hermitian Geometries
Previous Article in Special Issue
A Dark Matter WIMP That Can Be Detected and Definitively Identified with Currently Planned Experiments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modified Equations of State for Dark Energy and Observational Limitations

by
German S. Sharov
1,2,* and
Vasily E. Myachin
1
1
Mathematics Department, Tver State University, Sadovyjper. 35, 170002 Tver, Russia
2
International Laboratory for Theoretical Cosmology, Tomsk State University of Control Systems and Radioelectronics (TUSUR), 634050 Tomsk, Russia
*
Author to whom correspondence should be addressed.
Universe 2022, 8(4), 201; https://doi.org/10.3390/universe8040201
Submission received: 7 February 2022 / Revised: 16 March 2022 / Accepted: 19 March 2022 / Published: 24 March 2022
(This article belongs to the Special Issue Focus on Dark Matter)

Abstract

:
Cosmological models with variable and modified equations of state for dark energy are confronted with observational data, including Type Ia supernovae, Hubble parameter data H ( z ) from different sources, and observational manifestations of cosmic microwave background radiation (CMB). We consider scenarios generalizing the Λ CDM, wCDM, and Chevallier–Polarski–Linder (CPL) models with nonzero curvature and compare their predictions. The most successful model with the dark energy equation of state w = w 0 + w 1 ( 1 a 2 ) / 2 was studied in detail. These models are interesting in possibly alleviating the Hubble constant H 0 tension, but they achieved a modest success in this direction with the considered observational data.

1. Introduction

In modern cosmology based on recent observational data, our Universe includes dominating fractions of dark energy and dark matter, whereas all kinds of visible matter fill about 4% in total energy balance nowadays. The latest estimations of Planck collaboration [1,2] predict about 70% fraction of dark energy, if we apply the standard Λ CDM model, where dark energy may be represented as the cosmological constant Λ or as a matter with density ρ x and pressure p x = ρ x . Almost all remaining part of matter in this model is cold dark matter with close to zero pressure. Because of the last property, it is convenient to consider cold dark matter together with visible baryonic matter, where the unified density is ρ m = ρ c + ρ b . One should also add the radiation component including relativistic species (neutrinos) with p r = 1 3 ρ r , which was sufficient before and during the recombination era, but is is almost negligible now.
The Λ CDM model successfully describes numerous observations, including Type Ia supernovae (SNe Ia) data, estimates of the Hubble parameter H ( z ) , manifestations of baryonic acoustic oscillations (BAO), cosmic microwave background radiation (CMB), and other data [1,2]. However this model does not explain the nature of dark energy, the small observable value of the phenomenological constant Λ and the approximate equality ρ x and ρ m now (although these densities evolve differently).
Another essential problem in the Λ CDM model is the tension between Planck estimations of the Hubble constant H 0 = 67.37 ± 0.54 km /(s·Mpc) [2] (2018) and measurements of SH0ES group in the Hubble Space Telescope H 0 = 73.2 ± 1.3 km s 1 Mpc 1 [3] (2020) or H 0 = 73.3 ± 1.04 km s 1 Mpc 1 [4] (2021).
The Planck estimations [1,2] are based on the Λ CDM model and the Planck satellite measurements of the CMB anisotropy and power spectra related to the early Universe at redshifts z 1000 , whereas the SH0ES method uses local distance ladder measurements of Cepheids in our Galaxy [3] and in nearest galaxies, in particular, in the Large Magellanic Cloud [5], that implies z close to 0 (the late Universe). This H 0 tension has not diminished during the last years, and now it exceeds 4 σ .
Cosmologists suggested numerous scenarios for solving the mentioned problems with dark energy and the H 0 tension; they include modifications of early or late dark energy, dark energy with extra degrees of freedom, models with interaction in dark sector, models with extra relativistic species, viscosity, modified gravity including F ( R ) theories, and other models (see reviews [6,7,8,9,10] and papers [11,12,13,14,15,16,17,18,19,20,21]).
In this paper, we consider cosmological scenarios with modified equation of state (EoS) for dark energy; they generalize the Λ CDM model and its simplest extensions: the wCDM model with EoS
p x = w ρ x , w = const
and the models with variable EoS
p x = w ( a ) ρ x ,
where w depends on the scale factor a. The class (2) includes the following well-known dark energy equations of state: the linear model [22]
w = w 0 + w 1 ( a 1 1 ) = w 0 + w 1 z ;
Chevallier–Polarski–Linder (CPL) parametrization [23,24]
w = w 0 + w 1 ( 1 a ) = w 0 + w 1 z 1 + z
and (their generalization) Barboza–Alcaniz–Zhu–Silva (BAZS) EoS [25]
w = w 0 + w 1 1 a β β = w 0 + w 1 1 ( 1 + z ) β β .
Here, the scale factor a is normalized, so a ( t 0 ) = 1 at the present time t 0 ; a is connected with redshift z: a = ( 1 + z ) 1 . Obviously, BAZS parametrization (5) transforms into CPL EoS (4) at β = 1 , into the linear EoS (3) if β = 1 and into the logarithmic EoS w = w 0 + w 1 log ( 1 + z ) , if β 0 .
For all mentioned equations of state, one can integrate the continuity equation for non-interacting dark energy:
ρ ˙ x = 3 H ( ρ x + p x ) .
In this paper, we explore the above scenarios and suggest the following generalization of BAZS EoS:
w = w 0 + w 1 1 a β β a γ .
We test these models, confronting them with the following observational data: Type Ia supernovae data (SNe Ia) from the Pantheon sample survey [26], data extracted from Planck 2018 observations [2,27] of cosmic microwave background radiation (CMB), and estimations of the Hubble parameter H ( z ) = a ˙ / a for different redshifts z (the dot means d d t ). We use H ( z ) data from two sources: (a) from cosmic chronometers, that is, measured from differential ages of galaxies, and (b) estimates of H ( z ) obtained from line-of-sight baryonic acoustic oscillations (BAO) data.
This paper is organized as follows. In the next section, we describe H ( z ) , SNe Ia, and CMB observational data analyzed here. Section 3 is devoted to dynamics and free model parameters for scenarios (1)–(7). In Section 4 we analyze the results of our calculations for these models, and estimate values of model parameters including the Hubble constant H 0 , and in Section 5 we discuss the results and their possible applications for alleviating the Hubble constant tension problem.

2. Observational Data

Observational data should be described by the considered cosmological models. For each model, we calculate the best fit for its free parameters from the abovementioned data sources: (a) Type Ia supernovae (SNe Ia) data from Pantheon sample [26], (b) CMB data from Planck 2018 [2,27], and (c) estimates of the Hubble parameter H ( z ) from cosmic chronometers and line-of-sight BAO data.
The Pantheon sample database [26] for SNe Ia contains N S N = 1048 data points of distance moduli μ i obs at redshifts z i in the range 0 < z i < 1.92 . We compare them with theoretical values by minimizing the χ 2 function:
χ SN 2 ( θ 1 , ) = min H 0 i , j = 1 N SN Δ μ i C SN 1 i j Δ μ j , Δ μ i = μ th ( z i , θ 1 , ) μ i obs ,
Here, θ j are free model parameters, C SN is the covariance matrix [26], and the distance moduli μ th are expressed via the luminosity distance D L depending on the spacial curvature fraction Ω k and the Hubble parameter H ( z ) :
μ th ( z ) = m B ( z ) M B = 5 log 10 D L ( z ) 10 p c , D L ( z ) = ( 1 + z ) D M ,
D M ( z ) = c H 0 S k H 0 0 z d z ˜ H ( z ˜ ) , S k ( x ) = sinh x Ω k / Ω k , Ω k > 0 , x , Ω k = 0 , sin x | Ω k | / | Ω k | . Ω k < 0 ,
Here, m B is a supernova apparent magnitude, and M B is its absolute magnitude. The distance moduli μ th are not Hubble-free and depend on the Hubble constant H 0 = H ( t 0 ) (via the summand 5 log 10 H 0 ). On the other hand, μ i obs is essentially connected with the absolute magnitude M B and calculated with corrections coming from deviations of lightcurve shape, SN Ia color, and mass of a host galaxy [26,28]. In the Pantheon sample, these corrections and the connected pair ( H 0 , M B ) were considered as nuisance parameters, in particular: “Using only SNe, there is no constraint on H 0 since H 0 and M B are degenerate” [26]. We cannot divide uncertainties in the Hubble constant H 0 and possible uncertainties in M B [29,30,31].
Due to these reasons, we have to consider H 0 in Equation (8) as a nuisance parameter, its estimations cannot be obtained from χ SN 2 , and we minimize this function over H 0 [17,18,19,20,21]. However, SHe Ia data in χ SN 2 is important for fitting other model parameters.
Unlike SNe Ia Pantheon data, the CMB observations are related to the photon-decoupling epoch near z * 1090 . We use the following parameters extracted from Planck 2018 CMB observations [2,20,27]:
x = R , A , ω b , R = Ω m 0 H 0 D M ( z * ) c , A = π D M ( z * ) r s ( z * ) , ω b = Ω b 0 h 2
and their estimations for the non-flat Λ CDM + Ω k model [27]:
x Pl = R Pl , A Pl , ω b Pl = 1.7429 ± 0.0051 , 301.409 ± 0.091 , 0.0226 ± 0.00017 .
Considering the flat case ( Ω k = 0 ) of these models, we use the flat wCDM data [27]. The comoving sound horizon r s at z * is calculated as the integral
r s ( z * ) = z * c s ( z ˜ ) H ( z ˜ ) d z ˜ = 1 3 0 1 / ( 1 + z * ) d a a 2 H ( a ) 1 + 3 Ω b 0 / ( 4 Ω γ 0 ) a
with the fitting formula from Refs. [27,32] for the value z * . The resulting χ 2 function is
χ CMB 2 = min ω b Δ x · C CMB 1 Δ x T , Δ x = x x Pl ,
where we minimize over the normalized baryon fraction ω b 0 to diminish the effective number N p of free model parameters. The covariance matrix C CMB = C ˜ i j σ i σ j and other details are described in papers [17,27].
In this paper, we use the Hubble parameter data H ( z ) obtained from two different sources [17,18,19,20,21,33]. The first one is the cosmic chronometers (CC), in other words, estimations of H ( z ) via differences of ages Δ t for galaxies with close redshifts Δ z and the formula
H ( z ) = a ˙ a 1 1 + z Δ z Δ t .
Here, we include 31 CC H ( z ) data points from Refs. [34,35,36,37,38,39,40] used earlier in papers [17,18,19,20] and the recent estimate from Ref. [41]; they are shown in Table 1.
These 32 CC data points need a covariance matrix of systematic uncertainties connected with a choice of initial mass function, metallicity, star formation history, stellar population synthesis models, and other factors [42,43].
We describe these uncertainties as corrections Δ C H to the diagonal covariance matrix C H d = diag { σ i 2 } (from Table 1) taking into account their diagonal terms in the form [42]
( Δ C H ) i i = η ( z i ) H ( z i ) 2 .
Here, η ( z ) is a mean percentage bias depending on redshift z. We consider “the best-case scenario” from the paper [42] for η ( z ) and include these contributions of stellar population synthesis and metallicity omitting the non-diagonal terms Δ ( C H ) i j (they are negligible for metallicity [43]).
The second source of H ( z ) estimates is the baryon acoustic oscillation (BAO) data along the line-of-sight direction. We use here 36 H BAO ( z ) data points from Refs. [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58] (see Table 1). They were considered earlier in Ref. [21].
Some of the H ( z ) measurements [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58] in Table 1 used the same or overlapping large-scale structure data, so these H estimates for close redshifts z may be in duplicate. It concerns, for example, measurements of Delubac et al. [51], Font-Ribera et al. [50], uses quasars with Lyman- α forest from Data Release 11 SDSS-III survey; estimates of Alam et al. [54], Wang et al. [53], Bautista et al. [55], and Bourboux et al. [56] were made with data from or DR12 of SDSS-III etc. To avoid this doubling, we multiply the errors σ i by 2 for H BAO estimates in Table 1 with close z, data, and methods.
Note that H BAO estimates in Table 1 should be multiplied by the factor r d fid / r d , where fiducial values r d fid of the sound horizon size r d = r s ( z d ) at the drag epoch vary from 147.33 Mpc to 157.2 Mpc for different authors [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58]. We include this correction to errors σ i quadratically, comparing deviations of r d fid with r s ( z d ) calculated with Formula (11) for a considered model.
For any cosmological model we calculate the χ 2 function
χ H 2 ( θ 1 , ) = j = 1 N H H ( z j , θ 1 , ) H o b s ( z j ) σ j 2
by using (a) only CC H ( z ) data and (b) the full set CC + H BAO data. Note that H BAO data points are correlated with BAO angular distances considered in the previous papers [18,19,20]). Thus, here, we do not use data with BAO angular distances, to avoid any correlation.
Table 1. H ( z ) data from cosmic chronometers (CC) [34,35,36,37,38,39,40,41] and line-of-sight BAO [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58].
Table 1. H ( z ) data from cosmic chronometers (CC) [34,35,36,37,38,39,40,41] and line-of-sight BAO [44,45,46,47,48,49,50,51,52,53,54,55,56,57,58].
CC Data H BAO Data
z H ( z ) σ Refs z H ( z ) σ Refs
0.0706919.6Zhang 140.24079.692.992Gaztañaga 09
0.0906912Simon 050.3081.76.22Oka 14
0.12068.626.2Zhang 140.3178.184.74Wang 17
0.170838Simon 050.3483.83.66Gaztañaga 09
0.1791754Moresco 120.35082.79.13ChuangW 13
0.1993755Moresco 120.3679.943.38Wang 17
0.20072.929.6Zhang 140.3881.51.9Alam 17
0.2707714Simon 050.40082.042.03Wang 17
0.28088.836.6Zhang 140.43086.453.974Gaztañaga 09
0.35198314Moresco 120.4482.67.8Blake 12
0.38028313.5Moresco 160.4484.811.83Wang 17
0.4009517Simon 050.4887.792.03Wang 17
0.40047710.2Moresco 160.5190.41.9Alam 17
0.424787.111.2Moresco 160.5294.352.64Wang 17
0.44592.812.9Moresco 160.5693.342.3Wang 17
0.4708934Ratsimbazafy0.5787.67.83Chuang 13
0.478380.99Moresco 160.5796.83.4Anderson 14
0.489762Stern 100.5998.483.18Wang 17
0.592910413Moresco 120.60087.96.1Blake 12
0.6797928Moresco 120.6197.32.1Alam 17
0.7598.833.6Borghi 210.6498.822.98Wang 17
0.781210512Moresco 120.73097.37.0Blake 12
0.875412517Moresco 120.8106.94.9Zhu 18
0.8809040Stern 100.978113.7214.63Zhao 19
0.90011723Simon 051.0120.77.3Zhu 18
1.03715420.17Moresco 121.230131.4412.42Zhao 19
1.30016817Simon 051.5161.430.9Zhu 18
1.36316033.6Moresco 151.526148.1112.75Zhao 19
1.43017718Simon 051.944172.6314.79Zhao 19
1.53014014Simon 052.0189.932.9Zhu 18
1.75020240Simon 052.2232.554.6;Zhu 18
1.965186.550.4Moresco 152.3002248.57Buska 13
2.3302248.0Bautista 17
2.3402228.515Delubac 15
2.3602269.33Font-Ribera 14
2.40227.69.10Bourboux 17

3. Models

We explore all considered models in a homogeneous isotropic universe with the Friedmann–Lemaître–Robertson–Walker metric
d s 2 = d t 2 + a 2 ( t ) d r 2 1 k r 2 + r 2 d θ 2 + sin 2 θ d ϕ 2 ,
where k is the sign of spatial curvature. In this case, the Einstein equations are reduced to the system of the Friedmann equation
3 a ˙ 2 + k a 2 = 8 π G ρ
and the continuity equation
ρ ˙ + 3 H ( ρ + p ) = 0 .
Here, the total density ρ includes densities of the abovementioned cold pressureless matter (dark matter unified with baryonic matter), radiation, and dark energy:
ρ = ρ m + ρ r + ρ x , ρ m = ρ c + ρ b .
We suppose here that dark energy and the mentioned components do not interact in the form [13,14,15,16] and independently satisfy the continuity Equation (6) or (15). We integrate this equation for cold and relativistic matter:
ρ m = ρ m 0 a 3 , ρ r = ρ r 0 a 4
(the index “0” corresponds to the present time t 0 ) and substitute these relations into the Friedmann Equation (14) that can be rewritten as
H 2 = H 0 2 Ω m 0 a 3 + Ω r 0 a 4 + Ω k a 2 + Ω x ( a )
or
H ( z ) = H 0 Ω m 0 ( 1 + z ) 3 + Ω r 0 ( 1 + z ) 4 + Ω k ( 1 + z ) 2 + Ω x ( z ) .
Here,
Ω j 0 = 8 π G ρ j 0 3 H 0 2 , j = m , r , Ω k = k a 0 2 H 0 2 , Ω x ( a ) = 8 π G ρ x ( a ) 3 H 0 2 .
The dark energy fraction Ω x ( a ) results from the continuity Equation (6) ρ ˙ x + 3 H ( ρ x + p x ) = 0 , that, for the variable EoS (2) x = w ( a ) ρ x , is reduced to the form
log Ω x ( a ) = 3 1 + w ( a ) a 1 d a .
In particular, for BAZS parametrization (5) w = w 0 + w 1 ( 1 a β ) / β [25], the expression (20) is
Ω x ( a ) = Ω x 0 a 3 ( 1 + w 0 + w 1 / β ) exp 3 w 1 ( a β 1 ) β 2 .
One should substitute it into Equation (18).
In the above–mentioned particular cases, the BAZS Formula (21) takes the form
Ω x ( a ) = Ω x 0 a 3 ( 1 + w 0 w 1 ) e 3 w 1 z ,
for the linear model (3) w = w 0 + w 1 z (if β = 1 ) and
Ω x ( a ) = Ω x 0 a 3 ( 1 + w 0 + w 1 ) e 3 w 1 ( a 1 ) .
for CPL EoS (4) w = w 0 + w 1 ( 1 a ) (if β = 1 ). In the case w 1 = 0 (and w 0 w ), both models (22) and (23) transform into the wCDM model with
Ω x ( a ) = Ω x 0 a 3 ( 1 + w ) ,
Its particular case at w = 1 is the Λ CDM model where Ω x = Ω x 0 = Ω Λ = const.
For the generalization (7) (with the factor γ ) of BAZS parametrization (5), the expression (20) takes the form
Ω x ( a ) = Ω x 0 a 3 ( 1 + w 0 ) exp 3 w 1 β 1 a γ γ + a β + γ 1 β + γ .
If γ 0 it transforms into Equation (21), and in the case β 0 we have
Ω x ( a ) = Ω x 0 a 3 ( 1 + w 0 w 1 a γ / γ ) exp 3 w 1 1 a γ γ 2 .
For all considered models, the dark energy fraction Ω x 0 and other Ω j satisfy the equality
Ω m 0 + Ω r 0 + Ω k + Ω x 0 = 1 ,
resulting from Equation (18) or (19). Further, if we fix the ratio [18,19,20,59]
X r = ρ r 0 ρ m 0 = Ω r 0 Ω m 0 = 2.9656 · 10 4
to diminish the number N p of free model parameters. We will work with the following five free parameters in the linear (22) and CPL (23) models:
Ω m 0 , Ω k , H 0 , w 0 , w 1 .
Here, we consider ω b in χ CMB 2 (12) as a nuisance parameter. In the wCDM model (24), the number of parameters N p = 4 with w 0 w and the Λ CDM model N p = 3 ( Ω m 0 , Ω k , H 0 ). Otherwise, in the generalized model (7), (25) we have N p = 7 free parameters with additional β and γ to the set (28).
In the next sections, we compare predictions of these models with the observational data from Section 2.

4. Results

We evaluate how the considered models fit the observations taking into account the χ 2 functions for SNe Ia (8), CMB (12), and H ( z ) data (13) in the form
χ tot 2 = χ SN 2 + χ H 2 + χ CMB 2 .
For the Hubble parameter data, we separately use (a) only 32 data points with cosmic chronometers (CC) H ( z ) data and (b) the full set CC + H BAO data (see Table 1).
When we compare two models with different number N p of free parameters, we can expect that the model with larger N p achieves more success in minimizing χ tot 2 . However, some models are not successful in this sense. In particular, the linear model (3), (22) with N p = 5 parameters (28), with our set of observational data, yields the same minimal value min χ tot 2 1092.34 (for CC + H BAO ) as the wCDM model (24) with N p = 4 . The reason is the following: the best fitted value w 1 for the linear model (3) is very close to zero ( w 1 0.0006 ). In this case, the linear model works as the wCDM model (24) and yields the same min χ tot 2 .
Such a behavior of the model (3) is shown in Figure 1, where 1 σ and 2 σ filled contour plots in the left panels correspond to the full set of H ( z ) data (here and below, “all” denotes CC + H BAO ). The 1 σ , 2 σ contours for the wCDM model (red lines) in the Ω m 0 Ω k plane behave similarly and closely; positions of χ tot 2 minima point (shown as the star and the circle) practically coincide.
Here, the contours are drawn for χ 2 functions minimized over all other parameters, in particular, for the linear model (3) in the top-left panel:
χ tot 2 ( Ω m 0 , Ω k ) = min H 0 , w 0 , w 1 χ tot 2 .
A similar picture also takes place for only CC H ( z ) data, where min χ tot 2 1074.73 for both models. Equality of these minima for only CC and all H ( z ) data is illustrated with one-parameter distributions χ tot 2 ( Ω m 0 ) in the top-right panel. In one-parameter distributions, we also minimize over all other parameters.
In the Ω m 0 w 1 plane (the bottom-left panel) we see that for both models, minima of χ tot 2 are achieved near w 1 = 0 . It is also shown in the bottom-right panel, where the likelihood functions
L ( w 1 ) exp χ tot 2 ( w 1 ) / 2
are depicted for the linear model.
For the same observational data we can observe in Figure 2 more successful behavior of the Chevallier–Polarski–Linder (CPL) model (23) [23,24]. Here, 1 σ and 2 σ contours are drawn for χ tot 2 ( θ i , θ j ) of the type (30) in all planes of two parameters ( θ i θ j ) in notation of Figure 1. We consider four cases: for two models (CPL and wCDM) we calculate χ tot 2 for CC and all H ( z ) data, positions of all χ tot 2 minima points are shown. Naturally, in the panels with w 1 , only the CPL model is presented.
The likelihood functions L ( θ i ) of the type (31) are shown in Figure 2 for all five model parameters (28). They are used for estimating the best fits and 1 σ errors for these parameters, summarized below in Table 2.
The CPL model achieves lower values of min χ tot 2 in comparison with the wCDM and Λ CDM models. It can be seen in Table 2 and in the top-right panel of Figure 2, where one-parameter distributions χ tot 2 ( H 0 ) of these models are compared. The graphs χ tot 2 ( H 0 ) and the correspondent likelihoods L ( H 0 ) in the top-left panel demonstrate that the best fitted values H 0 are very close for the wCDM and CPL models and differ more essentially when we compare CC and all H ( z ) data.
The best fits of Ω m 0 depend stronger on the chosen model and vary from Ω m 0 0.206 for CPL, all H ( z ) to Ω m 0 0.289 for wCDM, CC. The best fits of Ω k behave similarly, but with the maximal estimate for CPL, all H ( z ) . The best CPL fits for w 1 are close to 0.66 for both variants of H ( z ) data. This value is far from zero; in other words, the CPL model with the considered observational data behaves differently to the wCDM model.
The CPL model achieves good results, but it is not the most successful scenario in the frameworks of the generalized model (7), (25). However, this generalized model has N p = 7 free parameters, including β and γ in addition to five parameters (28). This large number is the serious disadvantage of the generalized model (25) if we keep in mind informational criteria [19,20,21] and difficulties in calculations.
Following these reasons, we calculated the χ tot 2 function (29) for the generalized model (18), (25) with all H ( z ) data searching its minimum in the β γ plane. The results are presented in Figure 3 as one-parameter distributions χ tot 2 ( β ) and χ tot 2 ( γ ) (minimized over all other parameters).
We see in Figure 3 that the absolute minimum for the generalized model (25) min χ tot 2 1084.0 is achieved near the point β = 2 , γ = 0 . One may conclude that the Barboza–Alcaniz–Zhu–Silva (BAZS) model [25] (corresponding to γ = 0 ) with Ω x ( a ) (21) and β = 2 appeared to be the most successful for the considered observational data Ia + CMB + CC + H BAO . When we fix β = 2 , the BAZS model with
w = w 0 + w 1 ( 1 a 2 ) / 2
will have five free model parameters (28). We investigate this model (denoted below as “ β = 2 ”) in detail; the results are presented in Table 2 and in Figure 4.
Figure 4 illustrates the BAZS model with β = 2 (32): 1 σ and 2 σ contour plots are shown for all H ( z ) data (filled contours) and for only CC data. They are compared with the corresponding 1 σ contours of the CPL model (shown also in Figure 2). One can compare the related one-parameter distributions χ tot 2 ( H 0 ) in the top-right panel and likelihood functions (31) L ( H 0 ) , L ( Ω m 0 ) , L ( Ω k ) , etc.
Figure 4 and Table 2 demonstrate that the β = 2 BAZS model (32) is more successful in minimizing χ tot 2 if we compare it with the CPL model (and, naturally, with the wCDM and Λ CDM models). For only CC data and for all H ( z ) data, the best fit values of χ tot 2 for the β = 2 model (32) are achieved at lower values of Ω m 0 ( Ω m 0 = 0 . 177 0.072 + 0.113 for CC and 0 . 19 0.072 + 0.071 for all H data) and at larger values of Ω k in comparison with CPL model. However, the best fit values of H 0 approximately coincide for all considered models; they depend on a chosen dataset: only CC data or all H ( z ) data.
The best fitted values of w 0 are close for the β = 2 , CPL, and wCDM models in the CC case and slightly differ for all H ( z ) data. The optimal values of w 1 are larger for the β = 2 model ( w 1 = 1 . 33 1.56 + 0.42 for CC and 1 . 12 1.09 + 0.40 for all H data) if we compare with the CPL model; however, one should take into account the factor 1 2 in EoS (32): w = w 0 + w 1 ( 1 a 2 ) / 2 .
To compare models with different number N p of free model parameters, we use here the Akaike information criterion [19,20,60]:
AIC = min χ t o t 2 + 2 N p .
This criterion emphasizes the advantage of models with small number of N p . It can be seen in Table 3 for the mentioned models, where, for only CC H ( z ) data, the minimal Akaike values (33) are achieved for the Λ CDM ( AIC = 1079.81 ) and the wCDM ( AIC = 1080.97 ) models. However, for all H ( z ) data, the β = 2 model (32) with N p = 5 parameters appeared to be more successful than Λ CDM not only in min χ tot 2 1084.03 , but also with Akaike information (33): AIC = 1094.03 . The lowest AIC = 1093.19 is achieved here for the wCDM model.
On the right side of the Table 3, we present the predictions of the same models on their flat case k = 0 ( Ω k = 0 ). In this case, the values min χ tot 2 and AIC appeared to be larger, and the β = 2 model (32) loses its advantage in comparison with CPL scenario for all H ( z ) data.

5. Discussion

We considered different cosmological models with variable equations of state (EoS) for dark energy of the type (2) p x = w ( a ) ρ x , more precisely, models with EoS (7),
w ( a ) = w 0 + w 1 1 a β β a γ ,
generalizing the Λ CDM, wCDM, Chevallier–Polarski–Linder (CPL), and Barboza–Alcaniz–Zhu–Silva (BAZS) models [23,25]. These scenarios with nonzero spatial curvature and with Ω k = 0 were confronted with observational data described in Section 2 and including SNe Ia, CMB data, and two classes of the Hubble parameter estimates H ( z ) : from cosmic chronometers (CC) and from line-of-sight baryonic acoustic oscillations ( H BAO ) data.
The results of our calculations for different models are presented in Table 2 and Table 3, including minima of χ tot 2 , Akaike information criterion (33), and the best fitted values with 1 σ estimates of model parameters. We also investigated the linear model (3), (22) with N p = 5 parameters (28), however it appeared to be unsuccessful because it achieved the best fitted value of χ tot 2 at w 1 very close to zero (see Figure 1). In other words, the linear model (22) with the considered observational data was reduced to the wCDM model with only N p = 4 parameters, but both models have the same min χ tot 2 .
Unlike the linear model (22) CPL scenario (4), (23) with the same N p = 5 parameters (28) appeared to be more successful, in particular, for all H ( z ) data, the CPL model yields min χ tot 2 1084.28 in comparison with 1085.19 for the wCDM and 1089.42 for the Λ CDM models. In this case, the best fitted value w 1 0.62 is far from zero, so CPL is not reduced to the wCDM model.
We should remember that a large number N p of free model parameters is a drawback of any model, and when we use, in Table 3, the Akaike information criterion (33), the wCDM model with AIC = 1093.19 will have the advantage over CPL with AIC = 1094.28 (and more essential advantage over the Λ CDM model with 1095.42 ). If we consider the generalized model (7), (25) with N p = 7 and additional parameters β and γ , the Akaike expression (33) becomes too large and the model looks worse in comparison with others.
However, our analysis and Figure 3 showed that the minimum min χ tot 2 1084.03 of the generalized model (25) may be achieved if we fix β = 2 , γ = 0 ; the resulting model “ β = 2 ” (32) has the same N p = 5 parameters (28) and absolutely minimal min χ tot 2 for all H ( z ) data (its AIC = 1094.03 is behind only the wCDM AIC). The behavior of one-parameter distributions χ tot 2 ( H 0 ) and their minima for all models with all H ( z ) data are shown in Figure 5.
As mentioned above, if we use only CC H ( z ) data, we observe smaller differences between minima of χ tot 2 for the considered models in Table 2. In this case, the Akaike criterion (33) gives advantage to the Λ CDM model with minimal N p .
Table 2 and Figure 2 and Figure 4 demonstrate that the success of the CPL scenario is achieved at lower (best fitted) values of Ω m 0 and at larger values of Ω k , if we compare these results with the wCDM model. This tendency is strengthened for the β = 2 model (32). It looks natural, because the CPL model (23) is the particular case of the BAZS model (21) with β = 1 .
One can see in Table 2 and Table 3 and Figure 2, Figure 4, and Figure 5 that the best fitted values of the Hubble parameter H 0 are very close for models wCDM, CPL, and β = 2 in general and spatially flat cases. However, the predictions of these models diverge with that of the Λ CDM model for all H ( z ) data. The last result is illustrated in Figure 5; it becomes more clear if we look at Figure 2, keeping in mind that the Λ CDM model is the particular case of the wCDM model when w = 1 . One should add the observed difference between H 0 estimates of all models when we compare CC and all H ( z ) data.
In the left panels of Figure 5 we draw one-parameter distributions χ tot 2 ( H 0 ) and likelihood functions (31) L ( H 0 ) for the considered models with all H ( z ) data to clarify their best fits of H 0 . For the considered four models we demonstrate here the results for the flat cases k = 0 ( Ω k = 0 ) as dotted lines.
One can see that in the flat case k = 0 , the mentioned models yield appreciably larger minima min χ tot 2 , but the best fitted values of H 0 change unessentially for the wCDM, CPL, and β = 2 models. For the flat Λ CDM model, the graph χ tot 2 ( H 0 ) is inscribed between the correspondent graphs of the general Λ CDM and flat wCDM models; hence, the H 0 estimate for the flat Λ CDM model lies between two estimates of the mentioned scenarios.
All H 0 predictions of the wCDM, CPL, and β = 2 models with all H ( z ) data fit the Planck 2018 estimation of the Hubble constant [2] (see Figure 5), but they are far from the SH0ES 2021 estimation [4]. All H 0 predictions with CC H ( z ) data are very close for the four considered models and their flat variants: they are larger but overlap the Planck 2018 value; however, they cannot describe the tension with the SH0ES data.
We can see from Table 3 that the best values of AIC (33) for all H ( z ) data demonstrate the wCDM and β = 2 (32) models, but for CC H ( z ) data, the best AIC has the flat Λ CDM model.

Author Contributions

Conceptualization, G.S.S.; methodology, G.S.S.; software, G.S.S. and V.E.M.; formal analysis, G.S.S. and V.E.M.; investigation, G.S.S. and V.E.M.; writing—original draft preparation, G.S.S. and V.E.M.; writing—review and editing, G.S.S.; visualization, G.S.S. and V.E.M.; supervision, G.S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

G.S.S. is grateful to Sergei D. Odintsov for useful discussions and support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ade, P.A.; Aghanim, N.; Arnaud, M.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Banday, A.J.; Barreiro, R.B.; Bartlett, J.G.; Bartolo, N.; et al. Planck 2015 results. XIII. Cosmological parameters. Astron. Astrophys. 2016, 594, A13. [Google Scholar]
  2. Aghanim, N.; Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Banday, A.J.; Barreiro, R.B.; Bartolo, N.; Basak, S.; et al. Planck 2018 results. VI. Cosmological parameters. Astron. Astrophys. 2020, 641, A6. [Google Scholar]
  3. 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. Astrophys. J. Lett. 2021, 908, L6. [Google Scholar] [CrossRef]
  4. Riess, A.G.; Yuan, W.; Macri, L.M.; Scolnic, D. A Comprehensive Measurement of the Local Value of the Hubble Constant with 1 km/s/Mpc Uncertainty from the Hubble Space Telescope and the SH0ES Team. Astrophys. J. Lett. 2021, 908, L6. [Google Scholar]
  5. 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 LambdaCDM. Astrophys J. 2019, 876, 85. [Google Scholar] [CrossRef]
  6. Bamba, K.; Capozziello, S.; Nojiri, S.; Odintsov, S.D. Dark energy cosmology: The equivalent description via different theoretical models and cosmography tests. Astrophys. Space Sci. 2012, 342, 155–228. [Google Scholar] [CrossRef] [Green Version]
  7. Nojiri, S.; Odintsov, S.D. Unified cosmic history in modified gravity: From F(R) theory to Lorentz non-invariant models. Phys. Rept. 2011, 505, 59–144. [Google Scholar] [CrossRef] [Green Version]
  8. Odintsov, S.D.; Oikonomou, V.K. Modified Gravity Theories on a Nutshell: Inflation, Bounce and Late-time Evolution. Phys. Rept. 2017, 692, 1–104. [Google Scholar]
  9. Di Valentino, E.; Anchordoqui, L.A.; Ali-Haimoud, Y.; Amendola, L.; Arendse, N.; Asgari, M.; Ballardini, M.; Battistelli, E.; Benetti, M.; Birrer, S. Cosmology Intertwined II: The Hubble Constant Tension. Astropart. Phys. 2021, 131, 102605. [Google Scholar] [CrossRef]
  10. 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. Quantum Grav. 2021, 38, 153001. [Google Scholar] [CrossRef]
  11. Kumar, S.; Nunes, R.C. Probing the interaction between dark matter and dark energy in the presence of massive neutrinos. Phys. Rev. D 2016, 94, 123511. [Google Scholar] [CrossRef] [Green Version]
  12. Kumar, S.; Nunes, R.C. Echo of interactions in the dark sector. Phys. Rev. D 2017, 96, 103511. [Google Scholar] [CrossRef] [Green Version]
  13. 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] [Green Version]
  14. 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 Univ. 2020, 30, 100666. [Google Scholar] [CrossRef]
  15. Pan, S.; Sharov, G.S. A model with interaction of dark components and recent observational data. Mon. Not. R. Astron. Soc. 2017, 472, 4736. [Google Scholar] [CrossRef]
  16. Pan, S.; Sharov, G.S.; Yang, W. Field theoretic interpretations of interacting dark energy scenarios and recent observations. Phys. Rev. D 2020, 101, 103533. [Google Scholar] [CrossRef]
  17. Odintsov, S.D.; Saez-Gomez, D.; Sharov, G.S. Testing the equation of state for viscous dark energy. Phys. Rev. D 2020, 101, 044010. [Google Scholar] [CrossRef] [Green Version]
  18. Odintsov, S.D.; Saez-Gomez, D.; Sharov, G.S. Is exponential gravity a viable description for the whole cosmological history? Eur. Phys. J. C 2017, 77, 862. [Google Scholar] [CrossRef] [Green Version]
  19. Odintsov, S.D.; Saez-Gomez, D.; Sharov, G.S. Testing logarithmic corrections to R2-exponential gravity by observational data. Phys. Rev. D 2019, 99, 024003. [Google Scholar] [CrossRef] [Green Version]
  20. Nojiri, S.; Odintsov, S.D.; Saez-Gomez, D.; Sharov, G.S. Modelling and testing the equation of state for (Early) dark energy. Phys. Dark Univ. 2021, 32, 100837. [Google Scholar] [CrossRef]
  21. Odintsov, S.D.; Saez-Gomez, D.; Sharov, G.S. Analyzing the H0 tension in F(R) gravity models. Nucl. Phys. B 2021, 966, 115377. [Google Scholar] [CrossRef]
  22. Cooray, A.R.; Huterer, D. Gravitational Lensing as a Probe of Quintessence. Astrophys. J. 1999, 513, L95–L98. [Google Scholar] [CrossRef] [Green Version]
  23. Chevallier, M.; Polarski, D. Accelerating Universes with Scaling Dark Matter. Int. J. Mod. Phys. D 2001, 10, 213–224. [Google Scholar] [CrossRef] [Green Version]
  24. Linder, E.V. Exploring the expansion history of the universe. Phys. Rev. Lett. 2003, 90, 091301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Barboza, E.M.; Alcaniz, J.S.; Zhu, Z.-H.; Silva, R. A generalized equation of state for dark energy. Phys. Rev. D 2009, 80, 043521. [Google Scholar] [CrossRef]
  26. 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 Type Ia Supernovae from Pan-STARRS1 and Cosmological Constraints from The Combined Pantheon Sample. Astrophys. J. 2018, 859, 101. [Google Scholar] [CrossRef]
  27. Chen, L.; Huang, Q.-G.; Wang, K. Distance priors from Planck final release. J. Cosmol. Astropart. Phys. 2019, 1902, 028. [Google Scholar] [CrossRef] [Green Version]
  28. Nesseris, S.; Perivolaropoulos, L. Comparison of the legacy and gold SN Ia dataset constraints on dark energy models. Phys. Rev. D 2005, 72, 123519. [Google Scholar] [CrossRef] [Green Version]
  29. Efstathiou, G. To H0 or not to H0? Mon. Not. R. Astron. Soc. 2021, 505, 3866–3872. [Google Scholar] [CrossRef]
  30. Camarena, D.; Marra, V. On the use of the local prior on the absolute magnitude of Type Ia supernovae in cosmological inference. Mon. Not. R. Astron. Soc. 2021, 504, 5164–5171. [Google Scholar] [CrossRef]
  31. Nunes, R.C.; Di Valentino, E. Dark sector interaction and the supernova absolute magnitude tension. Phys. Rev. D 2021, 104, 063529. [Google Scholar] [CrossRef]
  32. Hu, W.; Sugiyama, N. Small Scale Cosmological Perturbations: An Analytic Approach. Astrophys. J. 1996, 471, 542–570. [Google Scholar] [CrossRef] [Green Version]
  33. Sharov, G.S.; Sinyakov, E.S. Cosmological models, observational data and tension in Hubble constant. Math. Model. Geom. 2020, 8, 1–20. [Google Scholar] [CrossRef]
  34. Simon, J.; Verde, L.; Jimenez, R. Constraints on the redshift dependence of the dark energy potential. Phys. Rev. D. 2005, 71, 123001. [Google Scholar] [CrossRef] [Green Version]
  35. Stern, D.; Jimenez, R.; Verde, L.; Kamionkowski, M.; Stanford, S.A. Cosmic Chronometers: Constraining the Equation of State of Dark Energy. I: H(z) Measurements. JCAP 2010, 1002, 008. [Google Scholar] [CrossRef] [Green Version]
  36. Moresco, M.; Cimatti, A.; Jimenez, R.; Pozzetti, L.; Zamorani, G.; Bolzonella, M.; Dunlop, J.; Lamareille, F.; Mignoli, M.; Pearce, H.; et al. Improved constraints on the expansion rate of the Universe up to z∼1.1 from the spectroscopic evolution of cosmic chronometers. JCAP 2012, 1208, 006. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, C.; Zhang, H.; Yuan, S.; Zhang, T.-J.; Sun, Y.-C. Four New Observational H(z) Data From Luminous Red Galaxies Sloan Digital Sky Survey Data Release Seven. Res. Astron. Astrophys. 2014, 14, 1221. [Google Scholar] [CrossRef] [Green Version]
  38. Moresco, M. Raising the bar: New constraints on the Hubble parameter with cosmic chronometers at z∼2. Mon. Not. R. Astron. Soc. 2015, 450, L16. [Google Scholar] [CrossRef] [Green Version]
  39. Moresco, M.; Pozzetti, L.; Cimatti, A.; Jimenez, R.; Maraston, C.; Verde, L.; Thomas, D.; Citro, A.; Tojeiro, R.; Wilkinson, D. A 6% measurement of the Hubble parameter at z∼0.45: Direct evidence of the epoch of cosmic re-acceleration. JCAP 2016, 1605, 014. [Google Scholar] [CrossRef] [Green Version]
  40. Ratsimbazafy, A.L.; Loubser, S.I.; Crawford, S.M.; Cress, C.M.; Bassett, B.A.; Nichol, R.C.; Väisänen, P. Age-dating Luminous Red Galaxies observed with the Southern African Large Telescope. Mon. Not. R. Astron. Soc. 2017, 467, 3239. [Google Scholar] [CrossRef] [Green Version]
  41. Borghi, N.; Moresco, M.; Cimatti, A. Towards a Better Understanding of Cosmic Chronometers: A new measurement of H(z) at z=0.7. arXiv 2021, arXiv:2110.04304. [Google Scholar] [CrossRef]
  42. Moresco, M.; Jimenez, R.; Verde, L.; Cimatti, A.; Lucia Pozzetti, L. Setting the Stage for Cosmic Chronometers. II. Impact of Stellar Population Synthesis Models Systematics and Full Covariance Matrix. Astrophys. J. 2020, 898, 82. [Google Scholar] [CrossRef]
  43. Moresco, M.; Amati, L.; Amendola, L.; Birrer, S.; Blakeslee, J.P.; Cantiello, M.; Cimatti, A.; Darling, J.; Valle, M.D.; Fishbach, M.; et al. Unveiling the Universe with Emerging Cosmological Probes. Astrophys. J. 2020, 898, 82. [Google Scholar] [CrossRef]
  44. Gaztañaga, E.; Cabre, A.; Hui, L. Clustering of Luminous Red Galaxies IV: Baryon Acoustic Peak in the Line-of-Sight Direction and a Direct Measurement H(z). Mon. Not. Roy. Astron. Soc. 2009, 399, 1663. [Google Scholar] [CrossRef] [Green Version]
  45. Blake, C.; Brough, S.; Colless, M.; Contreras, C.; Couch, W.; Croom, S.; Croton, D.; Davis, T.M.; Drinkwater, M.J.; Forster, K.; et al. The WiggleZ Dark Energy Survey: Joint measurements of the expansion and growth history at z<1. Mon. Not. R. Astron. Soc. 2012, 425, 405. [Google Scholar]
  46. Chuang, C.-H.; Wang, Y. Modeling the Anisotropic Two-Point Galaxy Correlation Function on Small Scales and Improved Measurements of H(z), DA(z), and f(z)σ8(z) from the Sloan Digital Sky Survey DR7 Luminous Red Galaxies. Mon. Not. R. Astron. Soc. 2013, 435, 255. [Google Scholar] [CrossRef] [Green Version]
  47. Chuang, C.-H.; Prada, F.; Cuesta, A.J.; Eisenstein, D.J.; Kazin, E.; Padmanabhan, N.; Sánchez, A.G.; Xu, X.; Beutler, F.; Manera, M.; et al. The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic Survey: Single-probe measurements and the strong power of f(z)σ8(z) on constraining dark energy. Mon. Not. R. Astron. Soc. 2013, 433, 3559. [Google Scholar] [CrossRef]
  48. Busca, N.G.; Delubac, T.; Rich, J.; Bailey, S.; Font-Ribera, A.; Kirkby, D.; Le Goff, J.-M.; Pieri, M.M.; Slosar, A.; Aubourg, E.; et al. Baryon Acoustic Oscillations in the Lyα forest of BOSS quasars. Astron. Astrophys. 2013, 552, A96. [Google Scholar] [CrossRef] [Green Version]
  49. Oka, A.; Saito, S.; Nishimichi, T.; Taruya, A.; Yamamoto, K. Simultaneous constraints on the growth of structure and cosmic expansion from the multipole power spectra of the SDSS DR7 LRG sample. Mon. Not. R. Astron. Soc. 2014, 439, 2515–2530. [Google Scholar] [CrossRef]
  50. Font-Ribera, A.; Kirkby, D.; Busca, N.; Miralda-Escudé, J.; Ross, N.P.; Slosar, A.; Rich, J.; Aubourg, E.; Bailey, S.; Bhardwaj, V.; et al. Quasar-Lyman α Forest Cross-Correlation from BOSS DR11: Baryon Acoustic Oscillations. J. Cosmol. Astropart. Phys. 2014, 2014, 027. [Google Scholar] [CrossRef]
  51. Delubac, T.; Bautista, J.E.; Busca, N.G.; Rich, J.; Kirkby, D.; Bailey, S.; Font-Ribera, A.; Slosar, A.; Lee, K.-G.; Pieri, M.M.; et al. Baryon Acoustic Oscillations in the Lyα forest of BOSS DR11 quasars. Astron. Astrophys. 2015, 574, A59. [Google Scholar] [CrossRef] [Green Version]
  52. Anderson, L.; Aubourg, E.; Bailey, S.; Beutler, F.; Bhardwaj, V.; Blanton, M.; Bolton, A.S.; Brinkmann, J.; Brownstein, J.R.; Burden, A.; et al. The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic Survey: Baryon Acoustic Oscillations in the Data Release 10 and 11 Galaxy Samples. Mon. Not. R. Astron. Soc. 2014, 441, 24. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, Y.; Zhao, G.-B.; Chuang, C.-H.; Ross, A.J.; Percival, W.J.; Gil-Martín, H.; Cuesta, A.J.; Kitaura, F.-S.; Rodriguez-Torres, S.; Brownstein, J.R.; et al. The clustering of galaxies in the completed SDSS-III Baryon Oscillation Spectroscopic Survey: Tomographic BAO analysis of DR12 combined sample in configuration space. Mon. Not. R. Astron. Soc. 2017, 469, 3762. [Google Scholar] [CrossRef]
  54. 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. [Google Scholar] [CrossRef] [Green Version]
  55. Bautista, J.E.; Busca, N.G.; Guy, J.; Rich, J.; Blomqvist, M.; Bourboux, H.M.; Pieri, M.M.; Font-Ribera, A.; Bailey, S.; Delubac, T.; et al. Measurement of baryon acoustic oscillation correlations at z=2.3 with SDSS DR12 Lyα-Forests. Astron. Astrophys. 2017, 603, A12. [Google Scholar] [CrossRef] [Green Version]
  56. Bourboux, H.; Le Goff, J.-M.; Blomqvist, M.; Busca, N.G.; Guy, J.; Rich, J.; Yèche, C.; Bautista, J.E.; Burtin, E.; Dawson, K.S.; et al. Baryon acoustic oscillations from the complete SDSS-III Lyα-quasar cross-correlation function at z=2.4. Astron. Astrophys. 2017, 608, A130. [Google Scholar] [CrossRef] [Green Version]
  57. Zhu, F.; Padmanabhan, N.; Ross, A.J.; White, M.; Percival, W.J.; Ruggeri, R.; Zhao, G.-B.; Wang, D.; Mueller, E.-M.; Burtin, E.; et al. The clustering of the SDSS-IV extended Baryon Oscillation Spectroscopic Survey DR14 quasar sample: Measuring the anisotropic Baryon Acoustic Oscillations with redshift weights. Mon. Not. R. Astron. Soc. 2018, 480, 1096. [Google Scholar] [CrossRef] [Green Version]
  58. Zhao, G.-B.; Wang, Y.; Saito, S.; Gil-Martín, H.; Percival, W.J.; Wang, D.; Chuang, C.-H.; Ruggeri, R.; Mueller, E.-M.; Zhu, F.; et al. The clustering of the SDSS-IV extended Baryon Oscillation Spectroscopic Survey DR14 quasar sample: Tomographic measurement of cosmic structure growth and expansion rate based on optimal redshift weights. Mon. Not. R. Astron. Soc. 2019, 482, 3497. [Google Scholar] [CrossRef] [Green Version]
  59. Ade, P.A.; Aghanim, N.; Armitage-Caplan, C.; Arnaud, M.; Ashdown, M.; Atrio-Barandela, F.; Aumont, J.; Baccigalupi, C.; Banday, A.J.; Barreiro, R.B.; et al. Planck 2013 results. XVI. Cosmological parameters. Astron. Astrophys. 2014, 571, A16. [Google Scholar]
  60. Akaike, H. A New Look at the Statistical Model Identification. IEEE Trans. Autom. Control 1974, 19, 716–723. [Google Scholar] [CrossRef]
Figure 1. For the linear model (3), 1 σ , 2 σ contour plots in the left panels (“all” means all H ( z ) data); one-parameter distributions χ tot 2 ( Ω m 0 ) and likelihood functions L ( w 1 ) in the right panels are compared with the wCDM model. The stars and circles denote positions of χ tot 2 minima points.
Figure 1. For the linear model (3), 1 σ , 2 σ contour plots in the left panels (“all” means all H ( z ) data); one-parameter distributions χ tot 2 ( Ω m 0 ) and likelihood functions L ( w 1 ) in the right panels are compared with the wCDM model. The stars and circles denote positions of χ tot 2 minima points.
Universe 08 00201 g001
Figure 2. CPL model (23) in comparison with the wCDM model: 1 σ , 2 σ contours, likelihoods L ( θ j ) , and one-parameter distributions χ tot 2 ( H 0 ) .
Figure 2. CPL model (23) in comparison with the wCDM model: 1 σ , 2 σ contours, likelihoods L ( θ j ) , and one-parameter distributions χ tot 2 ( H 0 ) .
Universe 08 00201 g002
Figure 3. The generalized model (25) with all H ( z ) data: one-parameter distributions for β and γ .
Figure 3. The generalized model (25) with all H ( z ) data: one-parameter distributions for β and γ .
Universe 08 00201 g003
Figure 4. BAZS model β = 2 (32) with 1 σ , 2 σ contours in comparison with CPL model ( 1 σ contours), likelihoods L ( θ j ) , and one-parameter distributions χ tot 2 ( H 0 ) .
Figure 4. BAZS model β = 2 (32) with 1 σ , 2 σ contours in comparison with CPL model ( 1 σ contours), likelihoods L ( θ j ) , and one-parameter distributions χ tot 2 ( H 0 ) .
Universe 08 00201 g004
Figure 5. One-parameter distributions χ tot 2 ( H 0 ) and likelihood functions for models with all H ( z ) data (the flat cases k = 0 are shown with dotted lines); H 0 estimates in the right panel are drawn as whisker plots.
Figure 5. One-parameter distributions χ tot 2 ( H 0 ) and likelihood functions for models with all H ( z ) data (the flat cases k = 0 are shown with dotted lines); H 0 estimates in the right panel are drawn as whisker plots.
Universe 08 00201 g005
Table 2. The best fit values with 1 σ estimates of model parameters and min χ tot 2 for SNe Ia + CMB and CC or all H ( z ) data for the models: β = 2 (32), CPL (23), wCDM, and Λ CDM.
Table 2. The best fit values with 1 σ estimates of model parameters and min χ tot 2 for SNe Ia + CMB and CC or all H ( z ) data for the models: β = 2 (32), CPL (23), wCDM, and Λ CDM.
ModelData min χ tot 2 / d . o . f H 0 Ω m 0 Ω k w 0 w w 1
β = 2 CC 1072.30 / 1079 68 . 40 1.91 + 1.86 0 . 177 0.072 + 0.113 0 . 080 0.076 + 0.067 0 . 967 0.14 + 0.12 1 . 33 1.56 + 0.42
CPLCC 1072.45 / 1079 68 . 53 1.88 + 1.84 0 . 218 0.074 + 0.102 0 . 044 0.059 + 0.047 0 . 955 0.14 + 0.14 0 . 65 1.04 + 0.25
w CDMCC 1072.97 / 1080 68 . 58 1.89 + 1.85 0 . 283 0.058 + 0.055 0 . 003 0.021 + 0.024 0 . 946 0.14 + 0.14 -
Λ CDMCC 1073.81 / 1081 68 . 52 1.94 + 1.89 0 . 305 0.029 + 0.029 0 . 0052 0.0018 + 0.0018 1 -
β = 2 all 1084.03 / 1115 67 . 65 1.26 + 1.28 0 . 190 0.072 + 0.071 0 . 068 0.053 + 0.072 0 . 96 0.067 + 0.087 1 . 12 1.09 + 0.40
CPLall 1084.28 / 1115 67 . 76 1.01 + 1.15 0 . 206 0.025 + 0.054 0 . 052 0.036 + 0.027 0 . 922 0.06 + 0.07 0 . 62 0.52 + 0.18
w CDMall 1085.19 / 1116 67 . 98 1.11 + 1.11 0 . 257 0.019 + 0.019 0 . 017 0.008 + 0.008 0 . 89 0.05 + 0.05 -
Λ CDMall 1089.42 / 1117 69 . 02 1.06 + 1.04 0 . 272 0.012 + 0.013 0 . 004 0.0017 + 0.0017 1 -
Table 3. The best fits of min χ tot 2 , AIC, and H 0 for the models β = 2 (32), CPL (23), wCDM, Λ CDM, and these models in the spatially flat case ( k = 0 ).
Table 3. The best fits of min χ tot 2 , AIC, and H 0 for the models β = 2 (32), CPL (23), wCDM, Λ CDM, and these models in the spatially flat case ( k = 0 ).
ModelData min χ tot 2 AIC H 0 Flat Model min χ tot 2 / d . o . f AIC H 0
β = 2 CC 1072.30 1082.30 68 . 40 1.91 + 1.86 flat β = 2 1072.99 / 1080 1080.99 68 . 46 1.81 + 1.82
CPLCC 1072.45 1082.45 68 . 53 1.88 + 1.84 flat CPL 1072.98 / 1080 1080.98 68 . 47 1.80 + 1.81
w CDMCC 1072.97 1080.97 68 . 58 1.89 + 1.85 flat wCDM 1073.01 / 1081 1079.01 68 . 50 1.79 + 1.81
Λ CDMCC 1073.81 1079.81 68 . 52 1.94 + 1.89 flat Λ CDM 1073.93 / 1082 1077.93 69 . 21 1.91 + 1.92
β = 2 all 1084.03 1094.03 67 . 65 1.26 + 1.28 flat β = 2 1087.84 / 1116 1095.84 67 . 81 1.06 + 1.12
CPLall 1084.28 1094.28 67 . 76 1.01 + 1.15 flat CPL 1087.56 / 1116 1095.56 67 . 99 1.16 + 1.11
w CDMall 1085.19 1093.19 67 . 98 1.11 + 1.11 flat wCDM 1089.90 / 1117 1095.90 68 . 16 1.14 + 1.12
Λ CDMall 1089.42 1095.42 69 . 02 1.06 + 1.04 flat Λ CDM 1089.96 / 1118 1093.96 68 . 35 0.55 + 0.56
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sharov, G.S.; Myachin, V.E. Modified Equations of State for Dark Energy and Observational Limitations. Universe 2022, 8, 201. https://doi.org/10.3390/universe8040201

AMA Style

Sharov GS, Myachin VE. Modified Equations of State for Dark Energy and Observational Limitations. Universe. 2022; 8(4):201. https://doi.org/10.3390/universe8040201

Chicago/Turabian Style

Sharov, German S., and Vasily E. Myachin. 2022. "Modified Equations of State for Dark Energy and Observational Limitations" Universe 8, no. 4: 201. https://doi.org/10.3390/universe8040201

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