Next Article in Journal
Modeling the Covariance of Financial Assets Using Neutrosophic Fuzzy Numbers
Next Article in Special Issue
Stability Properties of Self-Similar Solutions in Symmetric Teleparallel f(Q)-Cosmology
Previous Article in Journal
Closed-Form Formulas for the nth Derivative of the Power-Exponential Function xx
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Observational Constraints on F(T,TG) Gravity with Hubble’s Parametrization

by
Salim Harun Shekh
1,*,
Nurgissa Myrzakulov
2,3,†,
Anirudh Pradhan
4 and
Assem Mussatayeva
2,5,†
1
Department of Mathematics, S. P. M. Science and Gilani Arts, Commerce College, Ghatanji, Dist. Yavatmal, Maharashtra 445301, India
2
Department of General and Theoretical Physics, L N Gumilyov Eurasian National University, Astana 010008, Kazakhstan
3
Ratbay Myrzakulov Eurasian International Centre for Theoretical Physics, Astana 010009, Kazakhstan
4
Centre for Cosmology, Astrophysics and Space Science (CCASS), GLA University, Mathura 281406, India
5
Department of Physics and Chemistry, S. Seifullin Kazakh Agrotechnical University, Astana 010011, Kazakhstan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Symmetry 2023, 15(2), 321; https://doi.org/10.3390/sym15020321
Submission received: 9 December 2022 / Revised: 6 January 2023 / Accepted: 18 January 2023 / Published: 23 January 2023
(This article belongs to the Special Issue Symmetry and Asymmetry in Gravity Research)

Abstract

:
Any new gravitational theories can be built with the help of a gauge theory with local Poincare symmetry. This local Poincare symmetry can set up a space-time with torsion. In the present study, the authors working on the parametrization approach towards Hubble’s parameter in the frame of modified teleparallel Gauss-Bonnet gravity which is established on the torsion invariant T and the teleparallel equivalent of the Gauss-Bonnet term T G , say F ( T , T G ) gravity. In particular, gravity is responsible for an integrated explanation of the cosmological history from early-time inflation to late-time acceleration expansion, by lacking the addition of a cosmological constant. The domino effect acquired is reliable with recent cosmological outcomes. A transition scenario from a decelerating phase to an accelerating phase of cosmic evolution has been detected. Using the combined datasets (SNe-Ia+BAO+CMB+ H ( z ) ), we have constrained the transition redshift z t (at which the universe transit from a decelerating phase to an accelerating) and established the best fit value of z t . Next, we paralleled the renovated results of q ( z ) and ω ( z ) and found that the outcomes are well-suited with a Λ CDM universe.

1. Introduction

Astronomical data and theoretical influences put forward that the cosmos passed through an early inflationary era and brought about an accelerated era at a late time [1,2,3,4,5,6,7]. The confirmations that have been specified to explain this stage can commonly be put into two categories. One way is to revise the universe content by bringing together additional fields such as canonical and non-canonical scalars, fermionic, etc., that make known the conceptions of the inflation and/or the dark energy (DE), which can be protracted in an enormous class of models see [8,9,10]. Moreover, DE is classified by the equation of state parameter ( ω ) which is defined as ω = p ρ . The equation of state parameter is not certainly constant but the inconstant time dependant equation of state parameter is responsible for more physically feasible and realistic DE models explored by the Chevallier-Polarski-Linder (CPL) parametrization [11,12]. If ω is slightly upper and lower than 1 then it would be respectively equal to the quintessence and phantom DE cosmology whereas the possibility ω < < 1 is ruled out by current cosmological data. Together with several additional restrictions of the equation of state parameter are obtained from SNe-Ia data and a combination of SNe-Ia data with CMB anisotropy and Galaxy clustering statistics are 1.66 < ω < 0.62 and 1.33 < ω < 0.79 correspondingly. The modern outcome obtained from cosmological data sets coming from CMB anisotropy, luminosity distances of high red-shift SNe-Ia, and galaxy clustering constrain the range of ω lies in 1.44 < ω < 0.92 . Further, the CMB revealed that DE subsidizes 68% to the entire energy at the ease of the Universe. The cosmological constant Λ is the supreme up-front DE candidate with the equation of state ω = 1 . The Λ CDM model, containing the both cosmological constant ( Λ ) and cold dark matter (CDM), is the regular model that constantly fits with the current observational data sets. On the other hand, the Λ suffers from fine-tuning and coincidence matters. The other one is to transform the gravitational sector (Hilbert-Einstein action) instead which resulted in modified gravity theories. An extension of Hilbert-Einstein action leads to the f ( R ) gravity [13,14,15,16,17,18,19,20,21,22] which represents a class of theories defined as arbitrary functions of R. An approach to transforming the gravitational Hilbert- Einstein action is to spread out the work of the corresponding torsion formulation, called Teleparallel Equivalence of General Relativity (TEGR) [23,24,25,26]. In this class of gravity, the Weitzenbock connection is used instead of the torsion-less Levi-Civita connection and the torsion scalar T can be achieved after the contraction of the torsion tensor. Torsional formulation of gravity plays an essential role in the so-called Poincare gauge gravity, where the Poincare invariance plays a fundamental role. Moreover, torsional gravities can be obtained starting from the local Poincare symmetry. In the Poincare approach, torsion contributes to the overall dynamics where internal degrees of freedom (spins) and external degrees of freedom (space-time). Furthermore, based on a nonlinear realization of the local conformal-affine group of symmetry transformations was formulated. A number of significant studies on the astrophysical as well as cosmological aspects of f ( T ) gravity are ended in the references [27,28,29,30,31,32,33,34,35,36,37]. Other alternatives are f ( R , T ) gravity [38,39,40,41,42,43], f ( R , L m ) gravity [44,45,46,47,48,49]. Another, in the teleparallel formulation of gravity, the complex curvature rectifications can be acquainted with the Gauss-Bonnet combination G in which the torsion invariant term T G has been pulling out without commanding the Weitzenbock connection [50,51,52,53,54,55,56,57,58,59], this lead to an alternative remarkable class of modified gravity, known as F ( T , T G ) gravity [60,61]. These F ( T , T G ) gravity have been broadly considered in several contexts and acquired an exciting domino effect on multiple scales [62,63,64]. By adding the matter sector along with the gravitational one, the total action for F ( T , T G ) gravity is defined as [62],
S t o t = 1 2 κ 2 d 4 x e F ( T , T G ) + S m ,
where S m relates to a matter energy-momentum tensor and κ 2 = 8 π G be the four-dimensional Newton’s constant.
Any model that just reduces the sound horizon recombination can never fully resolve the Hubble tension while still being consistent with other cosmological datasets, as shown by Jedamzik et al. [65] in their study. They have explicitly demonstrated that models with larger Ω m h 2 develop tension with galaxy-weak lensing data, whereas models with smaller Ω m h 2 develop tension with observations of baryon acoustic oscillations. This is because models with lower matter density Ω m h 2 achieve a higher Hubble constant.
Inspiring by the effective cosmological domino effect of f ( T ) and the extension of f ( T ) gravity, in this study, the author study a gravitational action of the torsion scalar and the Gauss-Bonnet component which omits the F ( T , T G ) theory is considered. In particular, the author shall make an effort to the behavior of the Universe at the late time of its evolution.
The article is organized as follows: In Section 2. given basic formalism of F ( T , T G ) theory. In Section 3. presented parametrization of the Hubble parameter. Basic observational constraints are shown in Section 4. and results of tests summarised Section 5. Interesting behavior of physical parameters graphically presented in Section 6. Conclusions in detail given in Section 7.

2. Cosmology with F ( T , T G ) Gravity

In order to inspect the astrophysical and cosmological implications of the action (1), here the author thinks through a spatially flat cosmological ansatz of the form
d s 2 = N 2 ( t ) d t 2 + a 2 ( t ) d x 2 + d y 2 + d z 2 ,
where a ( t ) indicates the scale factor and N ( t ) be the lapse function. This metric omits the diagonal vierbein and its determinant as
e μ α = d i a g ( N , a , a , a ) , and e = N a 3
As usual, consider N ( t ) = 1 . Using the vierbein (3) of the considered cosmological ansatz (2), we found
T = 6 H 2 and T G = 24 H 2 ( H ˙ + H 2 ) ,
where H = a ˙ / a represents the Hubble’s parameter and the overhead dot denotes the differentiation with respect to cosmic time t.
Furthermore, varying the action (1), the equations of motion using diagonal vierbein (3) are formed as [60,61],
2 κ 2 ρ = F 12 H 2 F T T G F T G + 24 H 3 F ˙ T G
2 κ 2 p = F 4 ( H ˙ + 3 H 2 ) F T 4 H F ˙ T T G F T G + 2 T G F ˙ T G 3 H + 8 H 2 F ¨ T G
here F = F ( T , T G ) , F T = F ( T , T G ) T   and   F T G = F ( T , T G ) T G .
To analyze the above components of the equation of motion, it is necessarily required to consider the form of F = F ( T , T G ) . Hence go through the utmost essential and non-trivial model of the form (which does not introduce a new mass scale into the problem),
F ( T , T G ) = T + ϵ 1 f ( T , T G ) ,
where f ( T , T G ) = T 2 + ϵ 2 T G , ϵ 1 and ϵ 2 are dimensionless coupling model parameters. Also, the model is forecast to be necessary in late times. As, this model can yield significant cosmic behavior validating the advantages, opportunities, and original features of F ( T , T G ) cosmology [66].
In view of the F ( T , T G ) the model given in (7), and the components of equation of motion provided in Equations (5) and (6) become,
κ 2 ρ = 6 H 2 2 ϵ 1 2 T G ϵ 2 T + T ˙ G ϵ 2 2 + 2 T 3 T G ϵ 2 + T 2 3 / 2 + ϵ 1 T G ϵ 2 + 2 T 2 2 T G ϵ 2 + T 2 T
κ 2 p = 3 H 2 ϵ 1 4 T T G ϵ 2 + T 2 2 T ˙ G 2 ϵ 2 3 T G ϵ 2 + T 2 5 / 2 4 + T G T ˙ G ϵ 1 ϵ 2 2 6 H T G ϵ 2 + T 2 3 / 2 + 4 T G H ϵ 1 ϵ 2 T ˙ T G ϵ 2 + T 2 3 / 2 + H ˙ 4 ϵ 1 T T G ϵ 2 + T 2 4 ϵ 1 T G ϵ 2 + 2 T 2 2 T G λ 2 + T 2 + T

3. Isotropization

To examine the changing aspects of models/dark energy models, there are several somatic arguments and inspirations for the model are present in the literature. In this section, the authors keep an eye on the similar idea of parametrization of cosmological models and explain the gravitational equations openly and study the dynamics in an altered phase of the universe’s evolution. To define the phase transition of cosmological models from early-time inflation and deceleration to acceleration in late time, various cosmologists have well-thought-out parametrization of cosmological parameters and constrained them through observational data. Most of them are equations of state, deceleration, Hubble’s, and some well-known CPL parametrizations [67,68]. Hence, examine the parametrization of the Hubble parameter as
H ( z ) = H 0 1 + γ 1 + γ a 1 / β ,
where a is the average scale factor, H 0 be the present value of H ( z ) , γ and β both are non-negative free model parameters. Using ( 1 + z ) = 1 a , one can represent the above Equation (10) in z as
H ( z ) = H 0 1 + γ 1 + γ ( 1 + z ) 1 / β .

4. Observational Constraints

In the previous sections, we have briefly described the F ( T , T G ) gravity and solved the field equation with a new parametrization of the Hubble parameter. In order to extract the best fit values, the considered form of H ( z ) was constrained by SNIa from Pantheon, CMB from Planck 2018, BAO, and 36 data points from Hubble. In what follows, we describe in detail the methodology adopted and data used in our analysis in the following subsections. The results of our study are shown in Table 1 and Table 2 along with the contour plots (two-dimensional) with 1 σ and 2 σ errors.

4.1. Supernovae Type Ia (SNe-Ia)

Here, we use supernovae from Pantheon compilation made of 1048 spectroscopically confirmed Type Ia Supernovae distributed in the Redshift range 0.01 < z < 2.26 [69]. So for the Pantheon compilation, the chi-square equation is expressed as
χ S N 2 = ( μ o b s μ t h ) T . C P a n 1 . ( μ o b s μ t h ) ,
where μ t h = 5 log 10 c D L H 0 M p c + 25 , μ o b s be the observed distance modulus and C P a n is the covariance matrix of Pantheon data.

4.2. Cosmic Microwave Background

The χ 2 for cosmic microwave background is expressed as follows [70]
χ CMB 2 ( β , γ , h ) = X CMB T ( β , γ , h ) . C CMB 1 . X CMB ( β , γ , h ) ,
where X CMB T ( β , γ , h ) = ( R ( β , γ , h ) 1.74963 , l a ( β , γ , h ) 301.80845 , Ω b h 2 ( β , γ , h ) 0.02237 ) and C CMB be the cosmic microwave background covariance matrix.

4.3. Baryon Acoustic Oscillation

In this paper, we use correlated BAO data (6dFGS, WiggleZ and Lya) and uncorrelated ones (SDSS DR7 MGS, BOSS-LOWZ, BOSS-DR12, BOSS-CMASS and DES) [71,72]. Hence, the total chi-square for BAO, χ B A O 2 , is expressed as
χ B A O 2 ( β , γ , h ) = χ 6 d F G S 2 + χ S D S S 2 + χ B O S S L O W Z 2 + χ B O S S C M A S S 2 + χ W i g g l e Z 2 + χ B O S S D R 12 2 .
The peak positions of Baryon Acoustic Oscillation are in general given in terms of D V ( z ) / r s ( z ) , D A ( z ) / r s ( z ) and H ( z ) / r s ( z ) measured at the drag epoch z d r a g i.e., where baryons were released from photons.

4.4. Hubble’s Data

For the constraints, we also make use of the Hubble measurements H ( z ) . This observation can be regarded as cosmic chronometers, and we use a sample covering the redshift range 0 < z < 2.34 . For, these measurement the estimator χ 2 equation is expressed as:
χ H ( z ) 2 ( β , γ , h ) = i = 1 36 H o b s , i H ( z i , β , γ , h ) σ H , i 2 .
where H o b s be the observed value of H ( z ) and σ H ( z i ) represents the observational errors on the measured values H o b s ( z i ) .

4.5. Monte Carlo Markov Chain (MCMC)

We can test the predictions of our theory with the available data by implementing the MCMC process. Hence, the total χ 2 defined as
χ 2 = χ S N 2 + χ CMB 2 + χ B A O 2 + χ H ( z ) 2

5. Results on Observational Tests

After successfully implementing the MCMC process, we got the best fit and mean values of the cosmological free parameters β , γ that appear in Table 1. Further, an exciting feature is perceived in the significance of the Hubble constant, well-defined as H 0 = 100 h. With respect to the values acquired by [73] and the one forecast by Planck, our end result for H 0 perfectly resembles the value of Planck. The constrained values of the free cosmological parameters γ and β are found to be as γ = 0 . 62501 000062138 + 0.000062138 , β = 0 . 67364 0.0036638 + 0.0036638 (See Figure 1. Also compare the resultant values of χ t o t 2 , A I C c and Δ A I C c with the standard Λ CDM model, and observed that our model is supported and perfectly consistent with the observations as Δ A I C c < 2 (See Table 2).

6. Physical Parameters

6.1. Deceleration Parameter

The deceleration parameter (DP) describes the rate of acceleration of the Universe and is defined as
q ( z ) = 1 H ˙ H 2
The Universe is in a decelerating phase for optimistic q > 0 , whereas an accelerating phase can be seen for unfavorable q < 0 . Here, the model parameters γ and β both are cast-off to analyze the DP q. Figure 2, gives details of the expansion of transition from the past (positive) deceleration to the present (negative) acceleration for redshift z. At present, the DP is observed as q 0 = 0.45015 for the SNe-Ia+CMB+BAO+ H ( z ) (SCBH) dataset. Because of this, q ( z ) at the current cosmic epoch is fairly compatible with the range q 0 = 0 . 528 0.088 + 0.092 as found by a recent observation [66,74].

6.2. Equation of State Parameter

When used to define the physical relevance of energy sources in the evolution of the universe, the equation of state (EoS) parameter is produced as
ω = p ρ = T 3 T ˙ G 2 H 2 ϵ 1 ϵ 2 3 G ϵ 2 + T 2 5 / 2 + T G T ˙ G ϵ 1 ϵ 2 2 6 H T G ϵ 2 + T 2 3 / 2 + 4 3 H 2 + H ˙ ϵ 1 T T G ϵ 2 + T 2 1 + 4 H T ˙ ϵ 1 T G ϵ 2 + T 2 ϵ 1 T 2 G ϵ 2 + T 2 3 / 2 + T G ϵ 1 ϵ 2 2 T G ϵ 2 + T 2 ϵ 1 T G ϵ 2 + T 2 6 T ˙ G H 2 ϵ 1 ϵ 2 2 T G ϵ 2 + T 2 3 / 2 12 H 2 ϵ 1 T T G ϵ 2 + T 2 1 + ϵ 1 T G ϵ 2 + T 2 T G ϵ 1 ϵ 2 2 T G ϵ 2 + T 2 T
The EoS parameter is characterized as follows:
  • “For the dust phase the EoS parameter, ω = 0 ,
  • in the radiation-dominated phase, ω = 1 3 ,
  • in the vacuum energy or Λ CDM model, the EoS parameter is recovered by ω = 1 .
  • in the quintessence phase, EoS parameter lies in range ( 1 < ω < 0 ) ,
  • in the phantom regime ( ω < 1 ) ”.
In our investigation, the redshift-dependent behavior of the EoS parameter from Equation (13) is depicted in Figure 3 and Figure 4 by taking into account the change in the values of ϵ 1 ( 0.6681 ϵ 1 0.6881 ) and ϵ 2 ( 0.9851 ϵ 2 0.9951 ). This picture indicates that the model approaches the Λ CDM limit at z 0 for all ϵ 2 and ϵ 1 = 0.6781 ; this Λ CDM deviation also requires attention. It is usually convenient to have a Λ CDM limit to ensure that a model fits the data. The EoS parameter is constrained to have a range of 1.02 ω 0.986 at the 1 σ and 2 σ confidence level for the collective dataset of the observations, whereas the best fit value is close to ω = 1.0024 at the present epoch (see red color line in Figure 3) and it approaches 1 at late times (see Figure 3 and Figure 4). As a result, we determine that the end result is almost reliable with the recent observational constraints on ω obtained by Wood-Vasey et al. [75] and Davis et al. [76] at z 0 and also in good agreement with the cosmological data Amanullah et al. [77] at z 1 . Hence, we fix ϵ 1 = 0.6781 and ϵ 2 = 0.9901 throughout the analysis.

6.3. Energy Density and Pressure

With the best fit constraints value of β and γ which obtained from SCBH data and considered fix values of ϵ 1 and ϵ 2 the behavior of energy density and pressure is given in Figure 5 and Figure 6.
The Figure 5 and Figure 6 reveal that the redshift progression of energy density and isotropic pressure derived here in the framework of F ( T , T G ) gravity is completely reliable with the outcomes resulting in more than a few works mentioned in the collected analysis [78,79,80]. In detail, the cosmic energy density is non-negative and increases with the redshift, despite the fact an isotropic pressure is negative at present and in the future. As a result, negative pressure is liable for the accelerating expansion of the Universe at present and in the future.

6.4. Energy Conditions

The energy conditions (EC) consent us to consider the behavior of gravitating systems without stipulating the detailed behavior of the matter. The well-known Raychaudhuri equation, which deals with appealing gravity, has shown to be highly helpful in describing energy conditions. Strong energy condition (SEC), weak energy condition (WEC), null energy condition (NEC), and dominant energy condition (DEC) are the four point-wise energy conditions that are limits on the stress-energy-momentum tensor of the matter that are most frequently utilized. In generally known that WEC is the most spontaneous of the energy conditions. The WEC implies ρ 0 for a perfect fluid. Furthermore, the pressure of the universe can’t be so negative it takes over the energy density, or ρ + p 0 . Next, the second law of black hole thermodynamics is determined by the null energy condition (NEC) [81,82,83]. Conferring to the meaning, we can also understand that if the NEC is violated, then the WEC and DEC on the other cannot be satisfied. In physical these energy conditions are defined as [84],
  • “Null energy condition (NEC) ⇔ ρ + p 0 ,
  • Weak energy condition (WEC) ⇔ ρ + p 0 and ρ 0 ,
  • Strong energy condition (SEC) ⇔ ρ + p 0 and ρ + 3 p 0 ,
  • Dominant energy condition (DEC) ⇔ ρ | p | 0 and ρ 0 ”.
The energy conditions in several modified theories of gravity are studied such as Capozziello [85] and Alvarenga [86] studied using the power law in f ( R ) and f ( R , T ) gravity, Liu [87] studied it with exponential as well as Born-Infeld f ( T ) gravity, in f ( G ) gravity by Garcia [88] and Bamba [89] also Atazadeh [90] investigated in f ( R , G ) gravity. The energy conditions in cosmological models with variable anisotropic parameters in f ( R , T ) gravity have been examined by Mishra et al. [91].
The plots of the energy conditions are presented in Figure 7. We can see from the figure that the model’s NEC and DEC hold true, but the SEC is violated, which clearly indicates that the universe has accelerated expansion.

7. Conclusions

In the present work the investigation of observational constraints towards a modified gravity based on the torsion scalar ( T ) and the teleparallel equivalent of the Gauss-Bonnet combination ( T G ) say F ( T , T G ) gravity is presented. Firstly, the author takes out the general Friedmann equations and then, choosing specific F ( T , T G ) ansatzes we executed a complete study of various observable, such as energy density, the equation of state parameters, etc. The consequential cosmology leads to motivating behaviors.
  • A geometrical Hubble’s parameter H parametrization has been deliberated using the best fit constraints values of free parameters β and γ obtained from SCBH data. This parametrization generates a time-dependent q and gives details the current accelerated expansion of the Universe, i.e., q < 0 with a prior deceleration, i.e., q > 0 . Additionally, it is noted that the model deviates from the typical big bang scenario. The model with q as time-dependent has a signature flipping behavior with evolution. So, the feature of an early deceleration to the late acceleration of the model is appropriate for structure formation in the early stage of evolution and accelerated expansion in the later stage of the evolution.
  • Also discovered that the free cosmological parameters that are included in H ( z ) more specifically, could be associated in some way with the background parameters Ω m and Ω r . To end with, an exciting result here to comment on is that the value achieved for the Hubble parameter H 0 , lies close to the Planck estimation.
  • The behavior of energy density and pressure can be completely predicted using the best fit constraints values of β and γ , which were obtained from SCBH data and taken into consideration fix values of ϵ 1 and ϵ 2 . The results have been reported in a number of works. In detail, the cosmic energy density is non-negative and increases with the redshift, despite the fact an isotropic pressure is negative at present and in the future. As a result, negative pressure is liable for the accelerating expansion of the Universe at present and in the future.
  • In the analysis it is detected that the NEC and DEC both energy conditions are held, meanwhile, the SEC violates, the violation of SEC direct leads to the accelerating expansion of the Universe. Also, the WEC is non-negative from the early to late time phase of the Universe. Hence, the model reveals quintessential behavior. Simultaneously, the SEC was violated at a late time and satisfied at the early times (Figure 7).

Author Contributions

S.H.S.: Conceptualization, writing—review & editing, Programming, Investigation. N.M.: Writing—original draft, writing-review, editing. A.P.: Writing-original draft, Writing—review, editing, programming. A.M.: writing—review, editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Science Committee of the Ministry of Education and Science of the Republic of Kazakhstan (Grant No. AP09058240).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful for the anonymous referees’ insightful criticism, which helped to improve the manuscript in its current form. The authors would like to acknowledge the RM Eurasian International Centre for Theoretical Physics (RM EICTP), Kazakhstan for partially funding this work. During a visit when a part of this article was completed, the author (A. Pradhan) is appreciative of the assistance and resources provided by the University of Zululand, South Africa.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Perlmutter, S.; Aldering, G.; Goldhaber, G.; Knop, R.A.; Nugent, P.; Castro, P.G.; Deustua, S.; Fabbro, S.; Goobar, A.; Groom, D.E.; et al. Measurements of Ω and Λ from 42 high-redshift supernovae. Astrophys. J. 1999, 517, 565–586. [Google Scholar]
  2. Perlmutter, S.; Aldering, G.; Valle, M.D.; Deustua, S.; Ellis, R.S.; Fabbro, S.; Fruchter, A.; Goldhaber, G.; Groom, D.E.; Hook, I.M.; et al. Discovery of a supernova explosion at half the age of the Universe. Nature 1998, 91, 51. [Google Scholar]
  3. Riess, A.G. The case for an accelerating universe from supernovae. Astron. Soci. Pac. 2000, 112, 1284. [Google Scholar]
  4. Tonry, J.L.; Schmidt, B.P.; Barris, B.; Candia, P.; Challis, P.; Clocchiatti, A.; Coil, A.L.; Filippenko, A.V.; Garnavich, P.; Hogan, C.; et al. Cosmological results from high-z supernovae. Astrophys. J. 2003, 94, 1. [Google Scholar]
  5. de Bernardis, P.; Ade, P.A.; Bock, J.J.; Bond, J.R.; Borrill, J.; Boscaleri, A.; Coble, K.; Crill, B.P.; De Gasperis, G.; Farese, P.C.; et al. A flat Universe from high-resolution maps of the cosmic microwave background radiation. Nature 2000, 404, 955–959. [Google Scholar] [CrossRef] [Green Version]
  6. 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] [Green Version]
  7. Tegmark, M.; Strauss, M.A.; Blanton, M.R.; Abazajian, K.; Dodelson, S.; Sandvik, H.; Wang, X.; Weinberg, D.H.; Zehavi, I.; Bahcall, N.A.; et al. Cosmological parameters from SDSS and WMAP. Phys. Rev. D 2004, 69, 103501. [Google Scholar]
  8. Bassett, B.A.; Tsujikawa, S.; Wands, D. Inflation dynamics and reheating. Rev. Mod. Phys. 2006, 78, 537. [Google Scholar] [CrossRef] [Green Version]
  9. Copeland, E.J.; Sami, M.; Tsujikawa, S. Dynamics of dark energy. Int. J. Phys. D 2006, 15, 1753–1935. [Google Scholar]
  10. Cai, Y.F.; Saridakis, E.N.; Setare, M.R.; Xia, J.Q. Quintom cosmology: Theoretical implications and observations. Phys. Rep. 2010, 493, 1–60. [Google Scholar] [CrossRef] [Green Version]
  11. Chevallier, M.; Polarski, D. Accelerating universes with scaling dark matter. Int. J. Mod. Phys. D 2001, 10, 213–223. [Google Scholar] [CrossRef] [Green Version]
  12. Linder, E.V. Exploring the expansion history of the universe. Phys. Rev. Lett. 2003, 90, 091301. [Google Scholar] [CrossRef] [Green Version]
  13. Nojiri, S.I.; Odintsov, S.D. Modified f (R) gravity unifying Rm inflation with the ΛCDM epoch. Phys. Rev. D 2008, 77, 026007. [Google Scholar] [CrossRef] [Green Version]
  14. Inagaki, T.; Taniguchi, M. Cartan F(R) gravity and equivalent scalar-tensor theory. Symmetry 2022, 14, 1830. [Google Scholar] [CrossRef]
  15. Ali, S.; Saif, M.; Khan, K.A.; Shah, N.A. A note on varying G and Λ in Chern-Simons modified gravity. Symmetry 2022, 14, 1430. [Google Scholar] [CrossRef]
  16. Bekov, S.; Myrzakulov, K.; Myrzakulov, R.; Gomez, S.-C. General slow-roll inflation in f(R) gravity under the Palatini approach. Symmetry 2020, 12, 1958. [Google Scholar] [CrossRef]
  17. Granda, L. Unified inflation and late-time accelerated expansion with exponential and R2 corrections in modified gravity. Symmetry 2020, 12, 749. [Google Scholar] [CrossRef]
  18. Godani, N. Thin-shell wormhole solution in f(R) gravity. New Astron. 2023, 98, 101941. [Google Scholar]
  19. Nojiri, S.I.; 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]
  20. Capozziello, S.; De Laurentis, M. Extended theories of gravity. Phys. Rep. 2011, 509, 167–332. [Google Scholar]
  21. Odintsov, S.D.; Oikonomou, V.K. Early-time cosmology with stiff era from modified gravity. Phys. Rev. D 2017, 6, 104059. [Google Scholar]
  22. Odintsov, S.D.; Oikonomou, V.K. Reconstruction of slow-roll F (R) gravity inflation from the observational indices. Ann. Phys. 2018, 388, 267–275. [Google Scholar] [CrossRef] [Green Version]
  23. Einstein, A. Riemann-Geometrie mit Aufrechterhaltung des Begriffes des Fernparallelismus, Neue Möglichkeit für eine einheitliche Feldtheorie von Gravitation und Elektrizität. Sitzungsber. Preuss. Akad. Wiss. Berl. Phys. Math. 1928, Kl, 224. [Google Scholar]
  24. Arcos, H.I.; Pereira, J.G. Torsion gravity: A reappraisal. Int. J. Mod. Phys. D 2004, 13, 2193–2240. [Google Scholar] [CrossRef] [Green Version]
  25. Maluf, J.W. The teleparallel equivalent of general relativity. Ann. Phys. 2013, 525, 339–357. [Google Scholar]
  26. Aldrovandi, R.; Pereira, J.G. Teleparallel Gravity: An Introduction; Springer: Dordrecht, The Netherlands, 2013. [Google Scholar]
  27. Capozziello, S.; Cardone, V.F.; Farajollahi, H.; Ravanpak, A. Cosmography in f (T) gravity. Phys. Rev. D 2011, 84, 043527. [Google Scholar]
  28. Myrzakulov, R. Accelerating universe from F (T) gravity. Eur. Phys. J. C 2011, 71, 1752. [Google Scholar]
  29. Jeon, I.; Lee, K.; Park, J.H. Differential geometry with a projection: Application to double field theory. J. High Energy Phys. 2011, 2011, 14. [Google Scholar] [CrossRef] [Green Version]
  30. Tamanini, N.; Boehmer, C.G. Good and bad tetrads in f (T) gravity. Phys. Rev. D 2012, 86, 044009. [Google Scholar]
  31. Cai, Y.F.; Capozziello, S.; De Laurentis, M.; Saridakis, E.N. f (T) teleparallel gravity and cosmology. Rep. Prog. Phys. 2016, 9, 106901. [Google Scholar]
  32. Anagnostopoulos, F.K.; Basilakos, S.; Saridakis, E.N. Bayesian analysis of f (T) gravity using 8 data. Phys. Rev. D 2019, 100, 083517. [Google Scholar]
  33. Nair, K.K.; Arun, M.T. Kalb-Ramond field-induced cosmological bounce in generalized teleparallel gravity. Phys. Rev. D 2022, 105, 103505. [Google Scholar] [CrossRef]
  34. Shekh, S.H.; Chirde, V.R. Accelerating Bianchi type dark energy cosmological model with cosmic string in f(T) gravity. Astrophys. Space Sci. 2020, 365, 1–10. [Google Scholar] [CrossRef]
  35. Chirde, V.R.; Shekh, S.H. Analysis of general relativistic hydrodynamic cosmological models with stability factor in theories of gravitation. Gen. Relativ. Gravit. 2019, 51, 87. [Google Scholar]
  36. Chirde, V.R.; Shekh, S.H. Dynamic minimally interacting holographic dark energy cosmological model in f(T) gravity. Indian J. Phys. 2018, 92, 1485. [Google Scholar]
  37. Bhoyar, S.R.; Chirde, V.R.; Shekh, S.H. Stability of accelerating universe with a linear equation of state in f(T) gravity using hybrid expansion law. Astrophysics 2017, 60, 259–272. [Google Scholar] [CrossRef]
  38. Zubair, M.; Zeeshan, M.; Hasan, S.S.; Oikonomou, V.K. Impact of Collisional Matter on the Late-time Dynamics of f(R,T) gravity. Symmetry 2018, 10, 463. [Google Scholar] [CrossRef] [Green Version]
  39. Hulke, N.; Singh, G.P.; Bishi, B.K.; Singh, A. Variable chaplygin gas cosmologies in f(R,T) gravity with particle creation. New Astron. 2020, 77, 101357. [Google Scholar] [CrossRef] [Green Version]
  40. Sharma, U.K.; Kumar, M.; Varshney, G. Scalar field for Barrow holographic dark energy in f(R,T) gravity. Universe 2022, 8, 642. [Google Scholar]
  41. Mishra, A.K.; Sharma, U.K. Wormhole models in R2-fravity for f(R,T) theory with a hybrid shape function. Can. J. Phys. 2021, 99, 481–489. [Google Scholar]
  42. Pretel, J.M.Z.; Tangphati, T.; Banerjee, A.; Pradhan, A. Charged Quark Stars in f(R,T) Gravity. Chin. Phys. C 2022, 46, 115103. [Google Scholar] [CrossRef]
  43. Tangphati, T.; Hansraj, S.; Banerjee, A.; Pradhan, A. Quark stars in f(R,T) gravity with an interacting quark equation of state. Phys. Dark Univ. 2022, 35, 100990. [Google Scholar] [CrossRef]
  44. Harko, T.; Lobo, F.S.N. f(R,Lm) gravity. Eur. Phys. J. C 2010, 70, 373–379. [Google Scholar] [CrossRef]
  45. Faraoni, V. Cosmology in Scalar-Tensor Gravity; Kluwer Academic: Dordrecht, The Netherlands, 2004. [Google Scholar]
  46. Bertolami, O.; Piramos, J.; Turyshev, S. General Theory of Relativity: Will It Survive the Next Decade? Springer: Berlin/Heidelberg, Germany, 2008; pp. 27–74. [Google Scholar]
  47. Wang, J.; Liao, K. Energy conditions in f(R,Lm) gravity. Class. Quantum Gravity 2012, 29, 215016. [Google Scholar]
  48. Pradhan, A.; Maurya, D.C.; Goswami, G.K.; Beesham, A. Modeling transit dark energy in f(R,Lm) gravity. arXiv 2022, arXiv:2209.14269. [Google Scholar] [CrossRef]
  49. Lakhan, V.J.; Solanki, R.; Mandal, S.; Sahoo, P.K. Cosmology in f(R,Lm) gravity. Phys. Lett. B 2022, 831, 137148. [Google Scholar]
  50. Boulware, D.G.; Deser, S. String-generated gravity models. Phys. Rev. Lett. 1985, 55, 2656. [Google Scholar] [CrossRef] [Green Version]
  51. Nojiri, S.I.; Odintsov, S.D.; Sasaki, M. Gauss-Bonnet dark energy. Phys. Rev. D 2005, 71, 123509. [Google Scholar]
  52. Rodrigues, M.E.; Houndjo, M.J.S.; Momeni, D.; Myrzakulov, R. A type of Levi–Civita solution in modified Gauss-Bonnet gravity. Can. J. Phys. 2014, 92, 173. [Google Scholar] [CrossRef] [Green Version]
  53. Tangphati, T.; Pradhan, A.; Errehymy, A.; Banerjee, A. Quark Stars in the Einstein-Gauss-Bonnet theory: A New Branch of Stellar Configurations. Ann. Phys. 2021, 430, 168498. [Google Scholar]
  54. Tangphati, T.; Pradhan, A.; Frrehymy, A.; Banerjee, A. Anisotropic quark stars in Einstein-Gauss-Bonnet theory. Phys. Lett. B 2021, 819, 136423. [Google Scholar] [CrossRef]
  55. Tangphati, T.; Pradhan, A.; Banerjee, A.; Panotopoulos, G. Anisotropic Stars in 4D Einstein-Gauss-Bonnet Gravity. Phys. Dark Univ. 2021, 33, 100877. [Google Scholar]
  56. Panotopoulos, G.; Pradhan, A.; Tangphati, T.; Banerjee, A. Charged Polytropic Compact Stars in 4D Einstein-Gauss-Bonnet Gravity. Chin. J. Phys. 2022, 77, 2106–2114. [Google Scholar]
  57. Naicker, S.; Maharaj, S.D.; Brasel, B.P. Isotropic perfect fluids in modified gravity. Universe 2023, 9, 47. [Google Scholar] [CrossRef]
  58. Shekh, S.H.; Katore, S.D.; Chirde, V.R.; Raut, S.V. Signature flipping of isotropic homogeneous space-time with holographic dark energy in f(G) gravity. New Astron. 2020, 84, 101535. [Google Scholar]
  59. Shekh, S.H. Dynamical analysis with thermodynamic aspects of anisotropic dark energy bounce cosmological model in f(R,G) gravity. New Astron. 2021, 83, 101464. [Google Scholar]
  60. Kofinas, G.; Saridakis, E.N. Teleparallel equivalent of Gauss-Bonnet gravity and its modifications. Phys. Rev. D 2014, 90, 084044. [Google Scholar] [CrossRef] [Green Version]
  61. Kofinas, G.; Saridakis, E.N. Cosmological applications of F(T,TG) gravity. Phys. Rev. D 2014, 90, 084045. [Google Scholar] [CrossRef] [Green Version]
  62. Kofinas, G.; Leon, G.; Saridakis, E.N. Dynamical behavior in f(T,TG) cosmology. Class. Quantum Grav. 2014, 31, 175011. [Google Scholar]
  63. Chattopadhyay, S.; Jawad, A.; Momeni, D.; Myrzakulov, R. Pilgrim dark energy in f(T,TG) cosmology. Astrophys. Space Sci. 2014, 53, 279–292. [Google Scholar]
  64. Capozziello, S.; De Laurentis, M.; Dialektopoulos, K.F. Noether symmetries in Gauss-Bonnet-teleparallel cosmology. Eur. Phys. J. C 2016, 76, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. 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]
  66. Lohakare, S.V.; Mishra, B.; Maurya, S.K.; Singh, K. Constraining the cosmological parameters of modified Teleparallel-Gauss-Bonnet model. arXiv 2022, arXiv:2209.13197. [Google Scholar]
  67. Linder, E.V. Mapping the dark energy equation of state. In Symposium-International Astronomical Union; Cambridge University Press: Cambridge, UK, 2005; Volume 216, pp. 59–66. [Google Scholar]
  68. Cunha, J.V.; Lima, J.A.S. Transition redshift: New kinematic constraints from supernovae. Mon. Not. R. Astron. Soc. 2008, 390, 210–217. [Google Scholar]
  69. 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. Astrophys. J. 2018, 859, 101. [Google Scholar]
  70. Zhai, Z.; Wang, Y. Robust and model-independent cosmological constraints from distance measurements. J. Cosmol. Astropart. Phys. 2019, 1907, 005. [Google Scholar] [CrossRef] [Green Version]
  71. Anderson, L.; Aubourg, E.; Bailey, S.; Beutler, F.; Bhardwaj, V.; Blanton, M.; Bolton, A.S.; Brinkmann, J.; Brownstein, J.R.; Burden, A.; et al. [BOSS Collaboration], The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic Survey: Baryon acoustic oscillations in the Data Releases 10 and 11 Galaxy samples. Mon. Not. R. Astron. Soc. 2014, 441, 24–62. [Google Scholar] [CrossRef] [Green Version]
  72. Alam, U.; Bag, S.; Sahni, V. Constraining the Cosmology of the Phantom Brane using Distance Measures. Phys. Rev. D 2017, 95, 023524. [Google Scholar]
  73. 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. Astrophys. J. 2019, 876, 85. [Google Scholar]
  74. Gruber, C.; Luongo, O. Cosmographic analysis of the equation of state of the universe through Pade approximations. Phys. Rev. D 2014, 89, 103506. [Google Scholar]
  75. Wood-Vasey, W.M.; Miknaitis, G.; Stubbs, C.W.; Jha, S.; Riess, A.G.; Garnavich, P.M.; Kirshner, R.P.; Aguilera, C.; Becker, A.C.; Blackman, J.W.; et al. Observational constraints on the nature of dark energy: First cosmological results from the essence supernova survey. Astrophys. J. 2007, 666, 694. [Google Scholar]
  76. Davis, T.M.; Mortsell, E.; Sollerman, J.; Becker, A.C.; Blondin, S.; Challis, P.; Clocchiatti, A.; Filippenko, A.V.; Foley, R.J.; Garnavich, P.M.; et al. Scrutinizing exotic cosmological models using ESSENCE supernova data combined with other cosmological probes. Astrophys. J. 2007, 66, 716. [Google Scholar] [CrossRef]
  77. Amanullah, R.; Lidman, C.; Rubin, D.; Aldering, G.; Astier, P.; Barbary, K.; Burns, M.S.; Conley, A.; Dawson, K.S.; Deustua, S.E.; et al. Spectra and Hubble Space Telescope light curves of six type Ia supernovae at 0.511 < z < 1.12 and the Union2 compilation. Astrophys. J. Lett. 2010, 716, 712. [Google Scholar]
  78. Shekh, S.H. Models of holographic dark energy in f (Q) gravity. Phys. Dark Univ. 2013, 33, 100850. [Google Scholar]
  79. Koussour, M.; Filali, H.; Shekh, S.H.; Bennai, M. Holographic dark energy in Gauss-Bonnet gravity with Granda-Oliveros cut-off. Nucl. Phys. B 2022, 978, 115738. [Google Scholar] [CrossRef]
  80. Shekh, S.H.; Moraes, P.H.; Sahoo, P.K. Physical acceptability of the renyi, tsallis and sharma-mittal holographic dark energy models in the f (t, b) gravity under hubble’s cutoff. Universe 2021, 7, 67. [Google Scholar]
  81. Wald, R.M. General Relativity; University of Chicago Press: Chicago, IL, USA, 1984. [Google Scholar]
  82. Santos, J.; Alcaniz, J.S.; Reboucas, M.J.; Carvalho, F.C. Energy conditions in f (R) gravity. Phys. Rev. D 2007, 76, 083513. [Google Scholar]
  83. Xu, Y.; Harko, T.; Shahidi, S.; Liang, S.D. Weyl type f (Q, T) gravity, and its cosmological implications. Eur. Phys. J. C 2020, 80, 1–22. [Google Scholar] [CrossRef]
  84. Bouhmadi-Lopez, M.; Errahmani, A.; Martin-Moruno, P.; Ouali, T.; Tavakoli, Y. The little sibling of the big rip singularity. Int. J. Mod. Phys. D 2015, 24, 1550078. [Google Scholar] [CrossRef] [Green Version]
  85. Capozziello, S.; Nojiri, S.I.; Odintsov, S.D. The role of energy conditions in f (R) cosmology. Phys. Lett. B 2018, 781, 99–106. [Google Scholar]
  86. Alvarenga, F.G.; Houndjo, M.J.S.; Monwanou, A.V.; Orou, J.B.C. f (R, T) gravity from null energy condition. Int. J. Mod. Phys. 2013, 4, 130–139. [Google Scholar]
  87. Liu, D.; Reboucas, M.J. Energy conditions bounds on f (T) gravity. Phys. Rev. D 2012, 86, 083515. [Google Scholar]
  88. Garcia, N.M.; Harko, T.; Lobo, F.S.; Mimoso, J.P. Energy conditions in modified Gauss-Bonnet gravity. Phys. Rev. D 2011, 83, 104032. [Google Scholar] [CrossRef] [Green Version]
  89. Bamba, K.; Ilyas, M.; Bhatti, M.Z.; Yousaf, Z. Energy conditions in modified f (G) gravity. Gen. Relativ. Gravit. 2017, 49, 1–17. [Google Scholar]
  90. Atazadeh, K.; Darabi, F. Energy conditions in f(R,G) gravity. Gen. Relativ. Gravit. 2014, 46, 1–14. [Google Scholar] [CrossRef] [Green Version]
  91. Mishra, B.; Esmeili, F.M.; Ray, S. Cosmological Models with Variable Anisotropic Parameter in f(R,T) Gravity. Indian J. Phys. 2021, 95, 2245–2254. [Google Scholar] [CrossRef]
Figure 1. The figure shows the MCMC 1 σ and 2 σ confidence contours plot obtained from the combine SCBH dataset.
Figure 1. The figure shows the MCMC 1 σ and 2 σ confidence contours plot obtained from the combine SCBH dataset.
Symmetry 15 00321 g001
Figure 2. The DP with the best fit constraints value obtained from combined SCBH data versus redshift z.
Figure 2. The DP with the best fit constraints value obtained from combined SCBH data versus redshift z.
Symmetry 15 00321 g002
Figure 3. EoS parameter with the best fit constraints value obtained from SCBH data ϵ 1 = 0.6681 , ϵ 1 = 0.6781 and ϵ 1 = 0.6881 versus redshift z.
Figure 3. EoS parameter with the best fit constraints value obtained from SCBH data ϵ 1 = 0.6681 , ϵ 1 = 0.6781 and ϵ 1 = 0.6881 versus redshift z.
Symmetry 15 00321 g003
Figure 4. EoS parameter with the best fit constraints value obtained from SCBH data ϵ 2 = 0.9851 , ϵ 2 = 0.9901 and ϵ 2 = 0.9951 versus redshift z.
Figure 4. EoS parameter with the best fit constraints value obtained from SCBH data ϵ 2 = 0.9851 , ϵ 2 = 0.9901 and ϵ 2 = 0.9951 versus redshift z.
Symmetry 15 00321 g004
Figure 5. Energy density of the Universe with the best fit constraints value obtained from SCBH data ϵ 1 = 0.6681 , ϵ 1 = 0.6781 and ϵ 1 = 0.6881 versus redshift z.
Figure 5. Energy density of the Universe with the best fit constraints value obtained from SCBH data ϵ 1 = 0.6681 , ϵ 1 = 0.6781 and ϵ 1 = 0.6881 versus redshift z.
Symmetry 15 00321 g005
Figure 6. Isotropic pressure of the Universe with the best fit constraints value obtained from SCBH data ϵ 2 = 0.9851 , ϵ 2 = 0.9901 and ϵ 2 = 0.9951 versus redshift z.
Figure 6. Isotropic pressure of the Universe with the best fit constraints value obtained from SCBH data ϵ 2 = 0.9851 , ϵ 2 = 0.9901 and ϵ 2 = 0.9951 versus redshift z.
Symmetry 15 00321 g006
Figure 7. EC of the Universe with the best fit constraints value obtained from SCBH data ϵ 1 = 0.6681 , ϵ 1 = 0.6781 and ϵ 1 = 0.6881 versus redshift z.
Figure 7. EC of the Universe with the best fit constraints value obtained from SCBH data ϵ 1 = 0.6681 , ϵ 1 = 0.6781 and ϵ 1 = 0.6881 versus redshift z.
Symmetry 15 00321 g007
Table 1. The table displays the mean values of the free cosmological parameters as well as their summary and best fit.
Table 1. The table displays the mean values of the free cosmological parameters as well as their summary and best fit.
ModelParPriorBest FitMean
Λ CDM Ω m [0.001, 1] 0 . 311408 0.00589651 + 0.00589651 0 . 311568 0.0058866 + 0.0058866
Ω b h 2 [0.001, 1] 0 . 0223949 0.000131606 + 0.000131606 0 . 0224038 0.000131345 + 0.000131345
h[0.4, 1] 0 . 679136 0.00435014 + 0.00435014 0 . 679113 0.0043422 + 0.0043422
Model β [0, 1] 0 . 673644 0.00366388 + 0.00366388 0 . 673728 0.00366858 + 0.00366858
γ [0, 1] 0 . 62501 0.0000621386 + 0.0000621386 0 . 625011 0.00006221 + 0.00006221
h[0.4, 1] 0 . 657403 0.00408987 + 0.00408987 0 . 657422 0.00409376 + 0.00409376
Table 2. The table shows the summary of χ t o t 2 m i n , A I C c and Δ A I C c .
Table 2. The table shows the summary of χ t o t 2 m i n , A I C c and Δ A I C c .
Model χ tot 2 min AIC c Δ AIC c
Λ CDM1102.26111108.280
Model1108.28111109.45861.1771
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

Shekh, S.H.; Myrzakulov, N.; Pradhan, A.; Mussatayeva, A. Observational Constraints on F(T,TG) Gravity with Hubble’s Parametrization. Symmetry 2023, 15, 321. https://doi.org/10.3390/sym15020321

AMA Style

Shekh SH, Myrzakulov N, Pradhan A, Mussatayeva A. Observational Constraints on F(T,TG) Gravity with Hubble’s Parametrization. Symmetry. 2023; 15(2):321. https://doi.org/10.3390/sym15020321

Chicago/Turabian Style

Shekh, Salim Harun, Nurgissa Myrzakulov, Anirudh Pradhan, and Assem Mussatayeva. 2023. "Observational Constraints on F(T,TG) Gravity with Hubble’s Parametrization" Symmetry 15, no. 2: 321. https://doi.org/10.3390/sym15020321

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