-
The Arnowitt-Desert-Misner metric decomposition is advantageous, and it can be formulated as follows [89−91]:
$ {\rm d}s^2=-N^2{\rm d}t^2+g_{i j} \left({\rm d}x^i+N^i {\rm d}t\right)\left({\rm d}x^j+N^j {\rm d}t\right). $
(1) In this context, N represents the lapse function,
$ N_i $ denotes the shift vector, and$ g_{ij} $ corresponds to the spatial metric. The coordinate scaling transformation is expressed as follows:$ t\rightarrow l^3 t $ and$ x^i\rightarrow l x^i $ . The HLG action comprises two components, specifically, the kinetic and potential terms, formulated as follows:$ S_g=S_k+S_v=\int {\rm d}t {\rm d}^3 x \sqrt{g} N\left(L_k + L_v\right), $
In this context, the kinetic term is defined as follows:
$ S_k=\int {\rm d}t {\rm d}^3 x \sqrt{g} N \left[\frac{2\left(K_{ij}K^{ij} -\lambda K^2\right)}{\kappa^2}\right], $
In this situation, the extrinsic curvature is defined as
$ K_{ij}=\frac{\dot{g}_{ij}-\Delta_i N_j-\Delta_j N_i}{2N}. $
The Lagrangian, represented by
$ L_v $ , exhibits a reduction in the number of invariants due to its inherent symmetric property, as previously discussed in [59, 60−92]. This symmetry, often termed detailed balance, plays a crucial role. Considering the implications of detailed balance, the expression for the action can be expanded as follows:$ S_g = \int {\rm d}t {\rm d}^3x \sqrt{g} N \left[\frac{2\left(K_{ij}K^{ij} -\lambda K^2\right)}{\kappa^2}+\frac{\kappa^2 C_{ij}C^{ij}}{2\omega^4} -\frac{\kappa^2 \mu \epsilon^{i j k } R_{i, j} \Delta_j R^l_k}{2\omega^2 \sqrt{g}}+\frac{\kappa^2 \mu^2 R_{ij} R^{ij}}{8} -\frac{\kappa^2 \mu^2}{8(3\lambda-1)}\left\{\frac{(1-4\lambda)R^2}{4} +\Lambda R -3 \Lambda^2 \right\}\right] $
In this context, the expression
$ C^{ij}=\dfrac{\epsilon^{ijk}\Delta_k\left(R_i^j-\dfrac{R}{4} \delta^j_i\right)}{\sqrt{g}} $ represents the Cotton tensor, with all covariant derivatives determined with respect to the spatial metric$ g_{ij} $ . The symbol$ \epsilon^{ijk} $ denotes a totally antisymmetric unit tensor, while$ \lambda $ is a dimensionless constant, and$ \kappa $ ,$ \omega $ , and$ \mu $ are constants. Horava derived a gravitational action under the assumption of temporal dependence solely on the lapse function (i.e.,$ N\equiv N(t) $ ). Utilizing the Friedmann Robertson Walker (FRW) metric, where N=1,$g_{ij}= a^2(t)\gamma_{ij}$ ,$ N^i=0 $ , and$ \gamma_{ij}{\rm d}x^i {\rm d}x^j=\frac{{\rm d}r^2}{1-kr^2}+r^2 {\rm d}\Omega_2^2 $
with k = –1, 1, 0 representing an open, closed, and flat universe, respectively, and considering variations of N and
$ g_{ij} $ , the resulting equations are identified as the Friedmann equations [93, 94]$ \begin{aligned}[b]H^2=\;&\frac{\kappa^2\rho}{6\left(3\lambda-1\right)} +\frac{\kappa^2}{6\left(3\lambda-1\right)}\Bigg[\frac{3\kappa^2\mu^2 k^2} {8\left(3\lambda-1\right)a^4}\\&+\frac{3\kappa^2\mu^2 \Lambda^2} {8\left(3\lambda-1\right)}\Bigg]-\frac{\kappa^4 \mu^2 \Lambda k}{8\left(3\lambda-1\right)^2a^2}, \end{aligned}$
(2) $ \begin{aligned}[b]\dot{H}+\frac{3H^2}{2}=\;&-\frac{\kappa^2 p}{4\left(3\lambda-1\right)} -\frac{\kappa^2} {4\left(3\lambda-1\right)}\\&\times\left[\frac{3\kappa^2\mu^2 k^2} {8\left(3\lambda-1\right)a^4}+\frac{3\kappa^2\mu^2 \Lambda^2} {8\left(3\lambda-1\right)}\right]-\frac{\kappa^4 \mu^2 \Lambda k}{8\left(3\lambda-1\right)^2a^2} \end{aligned}$
(3) The term proportional to
$ \dfrac{1}{a^4} $ constitutes a distinctive contribution within HLG, often interpreted as the "Dark radiation term" [90, 91]. In this framework, the constant term is associated with the cosmological constant. Here,$ H = \dfrac{\dot{a}}{a} $ represents the Hubble parameter, and the dot signifies a derivative with respect to cosmic time t. Considering the universe's composition with both dark matter (DM) and dark energy (DE), the total energy density$ \rho $ and total pressure p can be expressed as$ \rho = \rho_m + \rho_d $ and$ p = p_m + p_d $ , respectively. Assuming distinct conservation equations for DM and DE, we derive$ \dot{\rho}_m + 3H(\rho_m + p_m) = 0, $
(4) and
$ \dot{\rho}_d + 3H(\rho_d + p_d) = 0. $
(5) As dark matter exhibits pressurelessness, i.e.,
$ p_m = 0 $ , Eq. (4) leads to$ \rho_m = \rho_{m0}a^{-3} $ . Let the equation of state parameter be denoted as$ w(z)=p/\rho $ . Consequently, from Eq. (5), we derive$\rho_d=\rho_{d0}\; {\rm e}^{3\int \frac{1+w(z)}{1+z} {\rm d}z}$ . Here,$ \rho_{m0} $ and$ \rho_{d0} $ represent the present values of the energy densities of dark matter (DM) and dark energy (DE), respectively. Introducing$ G_{c}=\dfrac{\kappa^2}{16\pi \left(3\lambda-1\right)} $ under the condition$ \dfrac{\kappa^4 \mu^2 \Lambda}{8\left(3\lambda-1\right)}=1 $ and adhering to detailed balance, the above Friedmann equations can be expressed as$ H^2=\frac{8\pi G_c}{3}\left(\rho_m + \rho_{d}\right)+\left(\frac{k^2} {2\Lambda a^4}+\frac{\Lambda}{2}\right)-\frac{k}{a^2}, $
(6) $ \dot{H}+\frac{3}{2}H^2=-4\pi G_c p_d -\left(\frac{k^2}{4\Lambda a^4}+\frac{3\Lambda}{4}\right)-\frac{k}{2a^2}. $
(7) Using the dimensionless parameters
$ \Omega_{i0}\equiv\dfrac{8\pi G_c}{3H_0^2}\rho_{i0} $ ,$ \Omega_{k0}=-\dfrac{k}{H_0^2} $ , and$ \Omega_{\Lambda 0}=\dfrac{\Lambda}{2H_0^2} $ , we obtain$ \begin{aligned}[b]H^2(z)=\;&H_0^2\Bigg[\Omega_{m0}(1+z)^3+\Omega_{k0}(1+z)^2 +\Omega_{\Lambda 0}\\&+\frac{\Omega_{k0}^2(1+z)^4}{4\Omega_{\Lambda 0}}+\Omega_{d0}\; {\rm e}^{3\int \frac{1+w(z)}{1+z} {\rm d}z}\Bigg] \end{aligned}$
(8) with
$ \Omega_{m0}+\Omega_{d0}+\Omega_{k0}+\Omega_{\Lambda 0}+\frac{\Omega_{k0}^2}{4\Omega_{\Lambda 0}}=1 $
(9) The observational data analysis for linear, CPL, and JBP models in HLG have been studied in [95], where they used the Stern+BAO+CMB datasets. However, in the next sections, we will assume some other parametrizations, such as those of Wettrich, Efstathiou, and Ma-Zhang in the framework of HLG, and we will study the data analysis with the MCMC method.
-
The Wettrich parametrization is assumed as [44, 45]
$ w(z)=\frac{w_{0}}{1+w_{1}\; {\rm log}(1+z)} $
(10) The solution is
$ \rho_{d}=\rho_{d0}(1+z)^{3}\left[1+w_{1}\; {\rm log}(1+z)\right]^{\frac{3w_{0}}{w_{1}}} $
(11) Thus, from Eq. (8), we obtain
$ \begin{aligned}[b] H^2(z)=\; & H_0^2\left[\Omega_{m0}(1+z)^3+\Omega_{k0}(1+z)^2 +\Omega_{\Lambda 0}+\frac{\Omega_{k0}^2(1+z)^4}{4\Omega_{\Lambda 0}} \right. \\ &+\left. \left(1-\Omega_{m0}-\Omega_{k0}-\Omega_{\Lambda 0}-\frac{\Omega_{k0}^2}{4\Omega_{\Lambda 0}}\right)\; (1+z)^{3}\left(1+w_{1}\; {\rm log}(1+z)\right)^{\frac{3w_{0}}{w_{1}}}\right] \end{aligned} $
(12) -
The Efstathiou parametrization [46, 47] is assumed as
$ w(z)=w_{0}+w_{1}\; {\rm log}(1+z) $
(13) The solution is obtained as
$ \rho_{d}=\rho_{d0}(1+z)^{3(1+w_{0})}{\rm e}^{\frac{3}{2}w_{1}[{\rm log}(1+z)]^{2}} $
(14) Thus, from Eq. (8), we obtain
$ \begin{aligned}[b] {H^2(z)=}\; &{ H_0^2\left[\Omega_{m0}(1+z)^3+\Omega_{k0}(1+z)^2 +\Omega_{\Lambda 0}+\frac{\Omega_{k0}^2(1+z)^4}{4\Omega_{\Lambda 0}} \right. } \\ &{+\left. \left(1-\Omega_{m0}-\Omega_{k0}-\Omega_{\Lambda 0}-\frac{\Omega_{k0}^2}{4\Omega_{\Lambda 0}}\right)\; (1+z)^{3(1+w_{0})}{\rm e}^{\frac{3}{2}w_{1}[{\rm log}(1+z)]^{2}}\right]} \end{aligned} $
(15) -
The Ma-Zhang parametrization is assumed as [48, 49]
$ w(z)=w_{0}+w_{1}\left(\frac{{\rm log}(2+z)}{1+z}-{\rm log}\; 2 \right) $
(16) The solution is obtained as
$ \rho_{d}=\rho_{d0}\; 2^{6w_{1}}(1+z)^{3(1+w_{0}+w_{1}-w_{1}\; {\rm log}\; 2)}(2+z)^{-\frac{3w_{1}(2+z)}{1+z}} $
(17) Thus, from Eq. (8), we obtain
$ \begin{aligned}[b] {H^2(z)=} \;& {H_0^2\left[\Omega_{m0}(1+z)^3+\Omega_{k0}(1+z)^2 +\Omega_{\Lambda 0}+\frac{\Omega_{k0}^2(1+z)^4}{4\Omega_{\Lambda 0}} \right.} \\ &{+\left. \left(1-\Omega_{m0}-\Omega_{k0}-\Omega_{\Lambda 0}-\frac{\Omega_{k0}^2}{4\Omega_{\Lambda 0}}\right)\; 2^{6w_{1}}(1+z)^{3(1+w_{0}+w_{1}-w_{1}\; {\rm log}\; 2)}(2+z)^{-\frac{3w_{1}(2+z)}{1+z}}\right]} \end{aligned} $
(18) -
In our examination of the late-stage cosmic expansion of the universe, we utilized a comprehensive dataset containing the latest BAO measurements obtained from a variety of observational sources. These data points have been carefully selected from different sources, including the Sloan Digital Sky Survey (SDSS) [96−101]. Furthermore, our dataset incorporates measurements from other significant surveys, such as the Dark Energy Survey (DES) [101], Dark Energy Camera Legacy Survey (DECaLS) [102], and 6dF Galaxy Survey BAO (6dFGS BAO) [103]. It is important to note that the BAO dataset presents a notable challenge due to the potential presence of correlations among measurements stemming from different data releases. To address this challenge and evaluate systematic errors, it is common practice to generate mock datasets using N-body simulations with well-defined cosmological parameters. These mock datasets allow us to derive the appropriate covariance matrices, aiding in the precise analysis of the BAO measurements. Given that our analysis combines data from multiple experiments, it is essential to acknowledge that we lack the precise covariance matrix outlining their interconnections, as such information is not readily available. In lieu of this, we conducted a covariance analysis following the methodology proposed in [104]. For uncorrelated data points, the covariance matrix can be expressed as
$ C_{i i} = \sigma_{i}^{2} $ . To account for potential correlations among data points, we introduced a specific number of non-diagonal elements into the covariance matrix, ensuring its symmetry. This approach allows us to introduce positive correlations among up to twelve pairs of randomly selected data points, encompassing more than 66.6% of the dataset. The positions of these non-diagonal elements are determined randomly, and the magnitude of these selected covariance matrix elements, denoted as$ C_{i j} $ , is set to$ C_{i j} = 0.5 \sigma_{i} \sigma_{j} $ . Here,$ \sigma_{i} $ and$ \sigma_{j} $ represent the published 1$ \sigma $ errors associated with data points i and j', respectively. To refine our cosmological model parameters, we enhance the dataset by including 24 BAO data points, as detailed in Table 1, including the latest measurement of the Hubble constant,$ H_{0} = 73.04 \pm 1.04 $ (km/s)/Mpc, as an additional prior known as R22 [4]. In our analysis, we utilized a nested sampler implemented in the open-source package Polychord [105], in conjunction with the GetDist package [106] for presenting the results. Figures 1, 2, and 3 depict the 68% and 95% confidence levels for key cosmological parameters for the Wettrich, Efstathiou, Ma-Zhang, and ΛCDM models. Table 2 presents the best-fit values of each model's parameters along with the current Hubble constant H0.BAO Dataset $ z_{e f f} $ Observable Measurement Error Year Dataset Survey Reference 0.106 $ r_{d} / D_{V} $ 0.336 0.015 2011 6dFGS BAO [41] 0.15 $ D_{V} / r_{d} $ 4.47 0.17 2021 SDSS Main Galaxy Sample [107] 0.31 $ D_{A} / r_{d} $ 6.29 0.14 2017 SDSS-III BOSS-DR12 [108] 0.36 $ D_{A} / r_{d} $ 7.09 0.16 2017 SDSS-III BOSS-DR12 [108] 0.38 $ D_{H} / r_{d} $ 25.00 0.76 2021 SDSS BOSS Galaxy Sample [108] 0.40 $ D_{A} / r_{d} $ 7.70 0.16 2017 SDSS-III BOSS-DR12 [108] 0.44 $ D_{A} / r_{d} $ 8.20 0.13 2017 SDSS-III BOSS-DR12 [108] 0.48 $ D_{A} / r_{d} $ 8.64 0.11 2017 SDSS-III BOSS-DR12 [108] 0.51 $ D_{M} / r_{d} $ 13.36 0.21 2021 SDSS BOSS Galaxy Sample [107] 0.52 $ D_{A} / r_{d} $ 8.90 0.12 2017 SDSS-III BOSS-DR12 [108] 0.56 $ D_{A} / r_{d} $ 9.16 0.14 2017 SDSS-III BOSS-DR12 [108] 0.59 $ D_{A} / r_{d} $ 9.45 0.17 2017 SDSS-III BOSS-DR12 [108] 0.64 $ D_{A} / r_{d} $ 9.62 0.22 2017 SDSS-III BOSS-DR12 [108] 0.697 $ D_{A}\left(r_{d} / r_{d, f i d}\right) $ 1529 73 2020 DECaLS DR8 Footprint LRG [102] 0.698 $ D_{H} / r_{d} $ 19.77 0.47 2020 eBOSS DR16 LRG Sample [98] 0.698 $ D_{M} / r_{d} $ 17.65 0.30 2020 eBOSS DR16 LRG Sample [98] 0.70 $ D_{M} / r_{d} $ 17.96 0.51 2021 eBOSS DR16 ELG Sample [109] 0.835 $ D_{M} / r_{d} $ 18.92 0.51 2022 Dark Energy Survey Year 3 [110] 0.845 $ D_{H} / r_{d} $ 20.91 2.86 2021 eBOSS DR16 ELG Sample [109] 0.874 $ D_{A}\left(r_{d} / r_{d, f i d}\right) $ 1680 109 2020 DECaLS DR8 Footprint LRG [102] 1.48 $ D_{H} / r_{d} $ 13.23 0.47 2021 eBOSS DR16 Quasar Sample [100] 1.48 $ D_{M} / r_{d} $ 30.21 0.79 2021 eBOSS DR16 Quasar Sample [100] 2.33 $ D_{H} / r_{d} $ 8.99 0.19 2020 eBOSS DR16 Ly $ \alpha $ -Quasar[101] 2.33 $ D_{M} / r_{d} $ 37.5 1.1 2020 eBOSS DR16 Ly $ \alpha $ -Quasar[101] Table 1. 24 independent BAO data points that form the basis of our analytical investigation. The data primarily originate from the conclusive measurements of the SDSS-III BOSS-DR12 and SDSS-IV eBOSS-DR16 datasets, contributing to the reinforcement of our findings.
Figure 1. (color online) Posterior distribution of BAO dataset measurements using the Wettrich type model, highlighting the 1
$ \sigma $ and 2$ \sigma $ confidence levels.Figure 2. (color online) Posterior distribution of BAO dataset measurements using the Efstathiou type model, highlighting the 1
$ \sigma $ and 2$ \sigma $ confidence levels.Figure 3. (color online) Posterior distribution of BAO dataset measurements using the Ma-Zhang type model, highlighting the 1
$ \sigma $ and 2$ \sigma $ confidence levels.MCMC Results Model Parameters Priors BAO BA0 + R22 H0 [50, 100] $ 68.33_{-5.00}^{+7.03} $ $ 73.76_{-1.20}^{+2.26} $ ΛCDM Model $ \Omega_{m0} $ [0.00,1.00] $ 0.27_{-0.01}^{+0.04} $ $ 0.27_{-0.01}^{+0.03} $ rd (Mpc) [100.00,200.00] $ 149.90_{-10.63}^{+23.88} $ $ 138.39_{-2.25}^{+4.68} $ $ r_{d}/r_{fid} $ [0.90,1.10] $ 1.00_{-0.07}^{+0.09} $ $ 0.93_{-0.02}^{+0.03} $ H0 [50, 80] $ 66.08_{-4.46}^{+7.40} $ $ 73.24_{-1.14}^{+2.28} $ $ \Omega_{m0} $ [0.24,0.35] $ 0.27_{-0.03}^{+0.05} $ $ 0.28_{-0.01}^{+0.02} $ $ \Omega_{k0} $ [-0.10,0.10] $ 0.01_{-0.04}^{+0.08} $ $ -0.00_{-0.02}^{+0.05} $ Wettrich Model $ \Omega_{\Lambda0} $ [0.00,0.10] $ 0.08_{-0.03}^{+0.06} $ $ 0.06_{-0.03}^{+0.04} $ $ w_{0} $ [-1.20,-0.50] $ -0.78_{-0.10}^{+0.22} $ $ -0.84_{-0.09}^{+0.17} $ $ w_{1} $ [0.00,0.50] $ 0.12_{-0.17}^{+0.26} $ $ 0.21_{-0.14}^{+0.25} $ rd (Mpc) [100.00,200.00] $ 150.24_{-10.17}^{+15.29} $ $ 136.08_{-2.57}^{+5.00} $ $ r_{d}/r_{fid} $ [0.90,1.10] $ 1.00_{-0.06}^{+0.09} $ $ 0.92_{-0.01}^{+0.02} $ H0 [50, 80] $ 66.36_{-4.95}^{+7.77} $ $ 73.22_{-1.23}^{+2.30} $ $ \Omega_{m0} $ [0.24,0.35] $ 0.28_{-0.01}^{+0.02} $ $ 0.28_{-0.01}^{+0.02} $ $ \Omega_{k0} $ [-0.10,0.10] $ -0.00_{-0.02}^{+0.06} $ $ -0.00_{-0.02}^{+0.05} $ Efstathiou Model $ \Omega_{\Lambda0} $ [0.00,0.10] $ 0.06_{-0.03}^{+0.05} $ $ 0.06_{-0.02}^{+0.05} $ $ w_{0} $ [-1.20,-0.50] $ -0.78_{-0.09}^{+0.16} $ $ -0.83_{-0.08}^{+0.20} $ $ w_{1} $ [-0.10,0.10] $ 0.10_{-0.14}^{+0.25} $ $ 0.16_{-0.16}^{+0.29} $ rd (Mpc) [100.00,200.00] $ 149.49_{-11.53}^{+16.10} $ $ 136.07_{-2.32}^{+4.71} $ $ r_{d}/r_{fid} $ [0.90,1.10] $ 1.00_{-0.06}^{+0.08} $ $ 0.92_{-0.01}^{+0.01} $ H0 [50, 80] $ 66.21_{-4.52}^{+7.75} $ $ 73.32_{-1.38}^{+2.15} $ $ \Omega_{m0} $ [0.24,0.35] $ 0.26_{-0.01}^{+0.02} $ $ 0.27_{-0.01}^{+0.02} $ $ \Omega_{k0} $ [-0.10,0.10] $ -0.00_{-0.02}^{+0.04} $ $ 0.00_{-0.02}^{+0.04} $ Ma-Zhang Model $ \Omega_{\Lambda0} $ [0.00,0.10] $ 0.06_{-0.03}^{+0.05} $ $ 0.06_{-0.03}^{+0.05} $ $ w_{0} $ [-1.20,-0.50] $ -0.97_{-0.09}^{+0.20} $ $ -0.92_{-0.09}^{+0.19} $ $ w_{1} $ [-0.50,0.00] $ -0.15_{-0.17}^{+0.24} $ $ -0.16_{-0.17}^{+0.23} $ rd (Mpc) [100.00,200.00] $ 150.82_{-9.31}^{+15.58} $ $ 137.02_{-3.11}^{+5.76} $ $ r_{d}/r_{fid} $ [0.90,1.10] $ 1.00_{-0.05}^{+0.09} $ $ 0.93_{-0.02}^{+0.02} $ Table 2. At a 95% confidence level (CL), we provide constraints on cosmological parameters for the Wettrich, Efstathiou, Ma-Zhang, and ΛCDM models. BAOs and Hubble constant measurements are incorporated as Gaussian priors.
-
Cosmography analysis is a cornerstone in modern cosmology [111] and provides an invaluable lens through which we delve into the enigmatic story of the universe's expansion. This research article endeavors to harness the power of cosmography, enabling us to unlock profound insights into cosmic dynamics. Through the meticulous analysis of observational data, we navigate the cosmos, comparing our findings to a range of theoretical models, including the Wettrich, Efstathiou, and Ma-Zhang models, alongside the well-established ΛCDM paradigm. This approach unveils how the universe's evolution unfolds across various redshifts. Cosmography, a potent tool, allows us to traverse the annals of time, enhancing our comprehension of the universe's past, present, and future.
-
The deceleration parameter [112], symbolized as "q," was initially postulated by Edwin Hubble in the early 20th century, and it is a fundamental cosmological parameter that is necessary to study the dynamics of the universe's expansion. Its mathematical definition is as follows:
$ q = -\frac{a\ddot{a}}{\dot{a}^2}, $
(19) Here, "a(t)" signifies the universe's scale factor over time, "
$ \dot{a} $ " represents its first derivative, and "$ \ddot{a} $ " is the second derivative. Gaining crucial insights into the past and future of cosmic evolution depends critically on the deceleration parameter. A positive number indicates a slowing down of the expansion, similar to a time when gravity was the primary driver. When the deceleration parameter is zero, the expansion rate is constant and the universe is called a "critical universe." Meanwhile, a negative deceleration parameter means that expansion is speeding, which is related to the recent finding of dark energy in cosmology. In order to better comprehend the general geometry of the universe, this parameter is crucial in examining the nature of cosmic elements like dark matter and dark energy. It is also essential for determining the general geometry of the universe. -
Within cosmology, a deeper understanding of the universe's dynamics has led to the development of the "jerk parameter," denoted by j [112]. As a contrasting measure to the deceleration parameter, this offers a more nuanced perspective on cosmic acceleration. It is mathematically represented as
$ j = \frac{1}{a}\frac{{\rm d}^3a}{{\rm d}\tau^3}\left[\frac{1}{a}\frac{{\rm d}a}{{\rm d}\tau}\right]^{-3} = q(2q + 1) + (1 + z)\frac{{\rm d}q}{{\rm d}z}, $
(20) To improve our understanding of cosmic development, the jerk parameter is a critical quantity that represents the third time derivative of the universe's scale factor.
-
Adding further depth to our cosmological toolkit, the snap parameter [113], symbolized as "
$ s_0 $ ," encapsulates the fifth time derivative of the expansion factor. In the Taylor expansion of the scale factor, it embodies the fourth-order term:$ s = \frac{1}{a}\frac{{\rm d}^4a}{{\rm d}\tau^4}\left[\frac{1}{a}\frac{{\rm d}a}{{\rm d}\tau}\right]^{-4} = \frac{j - 1}{3\left(q - \frac{1}{2}\right)}. $
(21) The snap parameter unveils the curvature of the universe and its manner of expansion, serving as a critical piece of the cosmographic puzzle.
-
In our study, we employed a robust diagnostic tool for probing DE known as the Om diagnostic, initially introduced by [113]. This diagnostic method is particularly notable for its simplicity, relying solely on the directly measurable Hubble parameter
$ H(z) $ obtained from observational data. The Om diagnostic serves as a valuable instrument for discerning between cosmological scenarios, specifically distinguishing the cosmological constant indicative of a standard ΛCDM model from a dynamic model associated with a curved ΛCDM. This discrimination is facilitated by utilizing the priors of Om and$ \Omega_{m0} $ . The condition$ O_{m} = \Omega_{m0} $ signifies consistency with the ΛCDM model. In contrast, situations where$ O_{m} > \Omega_{m0} $ suggest a quintessence scenario, while$ O_{m} < \Omega_{m0} $ points to a phantom scenario [114]. This diagnostic test not only provides a robust approach to understanding DE but also presents a unique method for distinguishing between various cosmological models. Its reliance on observational measurements, particularly the Hubble parameter, enhances its practicality and reliability in revealing the fundamental dynamics of the universe. In a flat universe, the expression for Om is defined as$ O_{m}=\frac{\left( \dfrac{H(z)}{H_{0}}\right) ^{2}-1}{(1+z)^{3}-1}. $
(22) -
In our statistical analysis, we aim to determine the most suitable cosmological model by considering both the number of free parameters and the obtained
$ \chi_{\rm{min}}^{2} $ value. We recognize that choosing among various information criteria can be a complex task, but we opt for commonly used ones. One of these criteria is the Akaike Information Criterion (AIC) [120−123], which define as$ {\rm AIC} \equiv -2 \ln {\cal{L}}_{\rm{max}} + 2p_{\rm{tot}} = \chi_{\rm{min}}^{2} + 2p_{\rm{tot}}. $
(23) Here,
$ p_{\rm{tot}} $ represents the total number of free parameters in a specific model, and$ {\cal{L}}_{\rm{max}} $ denotes the maximum likelihood of the considered model. Additionally, we utilize the Bayesian Information Criterion (BIC), introduced by [120, 122, 123], defined as$ {\rm BIC} \equiv -2 \ln {\cal{L}}_{\rm{max}} + p_{\rm{tot}} \ln \left(N_{\rm{tot}}\right). $
(24) By computing the differences
$\triangle {\rm A I C}$ and$\triangle {\rm B I C}$ relative to the ΛCDM model under consideration, we assess the model's performance. Following the guidelines in [124], if$0 < |\triangle {\rm A I C}| \leq 2$ , it suggests that the compared models are compatible. Conversely, if$|\triangle {\rm A I C}| \geq 4$ , it implies that the model with the higher AIC value is unsupported by the data. Similarly, for$0 < |\triangle {\rm B I C}| \leq 2$ , the model with the higher BIC value is marginally less favored by the data. When$2 < |\Delta {\rm B I C}| \leq 6$ ($|\triangle {\rm B I C}| > 6$ ), the model with the higher BIC value is significantly (highly) less favored. In cosmological statistical analyses, terms such as "P-value (probability value)" [87, 88, 125] play crucial roles in evaluating the significance of observations and testing hypotheses. The P-value quantifies the evidence against a null hypothesis. It indicates the probability of observing data as extreme or more extreme than those given, assuming that the null hypothesis is true. Cosmologists use P-values to assess whether observed data align with the predictions of a particular cosmological model. A low P-value suggests data inconsistency with the model, while a high P-value indicates consistency. These statistical tools are fundamental in cosmological analyses, helping cosmologists make inferences about key parameters, such as dark matter density, dark energy properties, and the universe's geometry. They assist in determining the most suitable models for describing our universe's behavior. We provide specific distinctions among the studied cosmological models in Table 3.Model ${\chi^2}_{\rm tot, min}$ $ \chi_{\rm {red }}^2 $ $ {\cal{K}}_{\rm{f}} $ AIC $ \Delta $ AICBIC $ \Delta $ BICP-value ΛCDM Model 10.81 0.92 3 16.81 0 17.16 0 0.55 Wettrich Model 9.91 0.92 6 21.91 5.10 19.44 2.28 0.53 Efstathiou Model 9.72 0.91 6 21.72 4.91 19.25 2.08 0.52 Ma-Zhang Model 9.90 0.92 6 21.90 5.09 19.43 2.26 0.52 Table 3. Summary of
${\chi^2}_{\rm tot, min}$ ,$ \chi_{\rm{red }}^2 $ , AIC,$ \Delta $ AIC, BIC,$ \Delta $ BIC, and P-value for ΛCDM, Wettrich, Efstathiou, and Ma-Zhang models.
Cosmological test of dark energy parametrizations within the framework of Horava-Lifshitz gravity via baryon acoustic oscillation
- Received Date: 2024-06-09
- Available Online: 2024-11-15
Abstract: We conduct an investigation to explore late-time cosmic acceleration through various dark energy parametrizations (Wettrich, Efstathiou, and Ma-Zhang) within the Horava-Lifshitz gravity framework. As an alternative to general relativity, this theory introduces anisotropic scaling at ultraviolet scales. Our primary objective is to constrain the key cosmic parameters and baryon acoustic oscillation (BAO) scale, specifically the sound horizon (rd), by utilizing 24 uncorrelated measurements of BAOs derived from recent galaxy surveys spanning a redshift range from z = 0.106 to z = 2.33. Additionally, we integrate the most recent Hubble constant measurement by Riess in 2022 (denoted as R22) as an extra prior. For the parametrizations of Wettrich, Efstathiou, and Ma-Zhang, our analysis of BAO data yields sound horizon results of rd = 148.1560 ± 2.7688 Mpc, rd = 148.6168 ± 10.2469 Mpc, and rd = 147.9737 ± 10.6096 Mpc, respectively. Incorporating the R22 prior into the BAO dataset results in rd = 139.5806 ± 3.8522 Mpc, rd = 139.728025 ± 2.7858 Mpc, and rd = 139.6001 ± 2.7441 Mpc. These outcomes highlight a distinct inconsistency between early and late observational measurements, analogous to the H0 tension. A notable observation is that, when we do not include the R22 prior, the outcomes for rd tend to be in agreement with Planck and SDSS results. Following this, we conducted a cosmography test and comparative study of each parametrization within the Lambda Cold Dark Matter paradigm. Our diagnostic analyses demonstrate that all models fit seamlessly within the phantom region. All dark energy parametrizations predict an equation of state parameter close to