This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The constrained cosmological model in Lyra geometry

J. K. Singh jksingh@nsut.ac.in Department of Mathematics, Netaji Subhas University of Technology, New Delhi-110078, India    Shaily shailytyagi.iitkgp@gmail.com Department of Mathematics, Netaji Subhas University of Technology, New Delhi-110078, India    Shri Ram srmathitbhu@rediffmail.com Department of Mathematical Sciences, Indian Institute of Technology (BHU), Varanasi 221 005, India    Joao R. L. Santos joaorafael@df.ufcg.edu.br UFCG-Universidade Federal de Campina Grande-Unidade Academica de Fisica, 58429-900 Campina Grande, PB, Brazil    Jéferson A. S. Fortunato jeferson.fortunato@edu.ufes.br PPGCosmo, CCE, Universidade Federal do Espírito Santo (UFES), 29.075-910, Vitória, ES, Brazil
Abstract

In this article, we study a flat homogeneous FLRW model in Lyra geometry which is described by a time-dependent displacement vector. We consider an appropriate parametrization of the energy density of scalar field ρϕ\rho_{\phi} in terms of the cosmic scale factor. The result shows two transitions from deceleration to acceleration. Furthermore, we constrain the model parameter α\alpha and the displacement field vector β\beta using the recent supernovae data, Hubble data set of 77 points and their joint data which predicts the accelerated expanding phase of the universe in late times. The effective equation of state parameter ωeff\omega_{eff} speculate Λ\LambdaCDM in late times. Finally, we use the statefinder diagnostic to differentiate our model from the various dark energy models.

PACS numbers: 04.20.-q, 04.50.Kd, 98.80.Es
Keywords: Lyra geometry, Observational constraints, Dark energy, FLRW space-time

I Introduction

From various observations, the accelerating behavior of the Universe is confirmed. However, figuring out the reason for acceleration is still a big challenge in cosmology. In standard cosmology, dark energy models are supported as a resolution to this fundamental question. Here, the cosmological constant Λ\Lambda, which can be associated with the vacuum quantum energy, is most prominent to describe the dark energy [1, 2]. Another reassuring way to understand the recent probes for cosmic expansion is Einstein theory of general relativity [3, 4]. The Einstein’s GR models seize up at large cosmic scales and a more generic action characterizes the gravitational field. To generalize the Einstein-Hilbert action of GR, there are various models where standard action is replaced by the generic function f(R)f(R), where RR is the Ricci scalar [5, 6, 7, 8, 9, 10, 11, 12]. Using these f(R)f(R) models, the early inflation of the universe and the concept of dark energy are discussed. In the extension of f(R)f(R) gravity theory, an explicit coupling of matter Lagrangian density LmL_{m} is included with f(R)f(R) [13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23]. For a proper description of the early universe, homogeneous and isotropic cosmological models play a significant role.

It is noteworthy that among all of the alternative theories of gravitation, Lyra geometry is one of the important theory, which is proposed by Lyra [24]. Lyra geometry comes with a modification of Riemannian geometry and may be considered as a modification of Weyl’s geometry. In Weyl’s geometry, the change in the length of vectors under parallel transport is non-zero and depends on a new vector quantity ζi\zeta_{i}, which is used for electromagnetic potential. In 1951, Lyra proposed a different way to maintain the vector length integrability by adopting a gauge function β(t)\beta(t) as an intrinsic part of the manifold’s geometric structure [25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35, 36]. Further, Sen [37] formulate a static cosmological model where he obtained the EFE through the variational principle. Singh and Shri Ram [38] discussed the spatially homogeneous Bianchi type-I metric in the normal gauge for Lyra’s geometry.

In the literature, a reconstructing technique or model-independent technique has been employed. Nowadays, this model-independent technique is of great interest to configure of some DE candidates. Initially, this approach is discussed by Starobinsky [16]. Many authors proposed this parametrization scheme on density, pressure, deceleration parameter, Hubble parameter, and scale factor. This method has two categories: (1) non-parametric and (2) parametric. The non-parametric method has certain restrictions because it could not be able to explain all the incidents of the present Universe while the parametric method includes a specific parametrization of the distinct types of cosmological parameters which are used to explore the cosmological models [39].

The present work is organized as follows: In Sec. II, we present the model field equations using Lyra’s geometry frame transformation for flat FLRW space-time metric. We parametrize the energy density for the scalar field and obtain the analytical solutions of the Hubble parameter and other cosmological parameters. In Sec. III, we discuss a brief introduction of Type Ia supernova data and observational Hubble data to derive the constraints on the model parameters. In Sec. IV, we study the evolution of various quantities of the universe. Finally, we discuss and conclude our findings in Sec. V.

II Field Equations in Lyra geometry

The gravitational action proposed by Sen [37] is defined as

A=(β4R+Lϕ+Lm)gd4x,A=\int(\beta^{4}R+L_{\phi}+L_{m})\sqrt{-g}d^{4}x, (1)

Using Lyra’s reference frame transformation the field equations obtained as:

Rγδ12Rgγδ+32ζγζδ34gγδζλζλ=κTγδ,R_{\gamma\delta}-\frac{1}{2}Rg_{\gamma\delta}+\frac{3}{2}\zeta_{\gamma}\zeta_{\delta}-\frac{3}{4}g_{\gamma\delta}\zeta_{\lambda}\zeta^{\lambda}=\kappa T_{\gamma\delta}, (2)

where κ=8πG\kappa=8\pi G which is normalized to 11 for further calculation, RγδR_{\gamma\delta} is Ricci tensor, RR is scalar curvature, and ζγ\zeta_{\gamma} is the time-like displacement vector, defined as (β(t),0,0,0)(\beta(t),0,0,0).

The flat FLRW space-time metric is given by

ds2=a2(t)(dx2+dy2+dz2)+dt2,ds^{2}=-a^{2}(t)(dx^{2}+dy^{2}+dz^{2})+dt^{2}, (3)

where a(t)a(t) is the scale factor of the universe. From Eq. (2), we get

3H23β24=ρeff=ρϕ+ρm,3H^{2}-\frac{3\beta^{2}}{4}=\rho_{eff}=\rho_{\phi}+\rho_{m}, (4)
2H˙+3H2+3β24=peff=pϕ.2\dot{H}+3H^{2}+\frac{3\beta^{2}}{4}=-p_{eff}=p_{\phi}. (5)

Overhead dot is used for derivative w.r.t. time. ρeff\rho_{eff} and peffp_{eff} signifies the total amount of energy density and pressure contained in the universe. Also ρϕ\rho_{\phi}, ρm\rho_{m}, pϕp_{\phi} depict the energy density of the scalar field, energy density of matter, and pressure of scalar field respectively. According to the present Universe is filled with two types of fluid, one is corresponding to the scalar field and another is pressureless cold matter.

The action of the scalar field is defined as:

Lϕ=[12νϕνϕV(ϕ)]gd4x.L_{\phi}=\int[\frac{1}{2}\partial_{\nu}{\phi}\partial^{\nu}{\phi}-V(\phi)]\sqrt{-g}d^{4}x. (6)

In the study regarding an isotropy and homogeneity of the model, we require a dominant unit of the scalar field ϕ\phi. As dependency of scalar field is on time therefore this may assume as perfect fluid with ρϕ\rho_{\phi} and ϕ{\phi}. And for FLRW cosmology energy density and pressure for scalar field can be obtained as

ρϕ=ϕ˙22+V(ϕ),\rho_{\phi}=\frac{\dot{\phi}^{2}}{2}+V(\phi), (7)
pϕ=ϕ˙22V(ϕ),p_{\phi}=\frac{\dot{\phi}^{2}}{2}-V(\phi), (8)

where V(ϕ)V(\phi) and 12ϕ˙2\frac{1}{2}\dot{\phi}^{2} are potential energy and kinetic energy depending on scalar field ϕ\phi. Now the conservation equations for matter field and scalar field read as

ρm˙+3Hρm=0,\dot{\rho_{m}}+3H\rho_{m}=0, (9)
ρϕ˙+3H(ρϕ+pϕ)=0.\dot{\rho_{\phi}}+3H(\rho_{\phi}+p_{\phi})=0. (10)

As EoS parameter ωϕ=pϕρϕ\omega_{\phi}=\frac{p_{\phi}}{\rho_{\phi}}, thus Eq. (10) leads

ρϕ˙+3H(1+ωϕ)ρϕ=0,\dot{\rho_{\phi}}+3H(1+\omega_{\phi})\rho_{\phi}=0, (11)

or

ωϕ=(1+a13ρϕdρϕda),\omega_{\phi}=-(1+a\frac{1}{3\rho_{\phi}}\frac{d\rho_{\phi}}{da}), (12)

Since we have four unknowns in three independent Eqs. (9), (10), (12), thus in order to solve the system of independent equations we need one more constraint equation. Therefore, we consider an appropriate parametrization of ρϕ\rho_{\phi} as

ρϕ(a)=eαatan1(aα).\rho_{\phi}(a)=e^{-\alpha a}tan^{-1}(a^{-\alpha}). (13)

This kind of approach is also been used by Singh et al. [39]. In this expression α(0,0.2)\alpha\in(0,0.2) is the model parameter and will be constrained from observational datasets. The value of ρϕ\rho_{\phi} can be written in terms of redshift by using the relation a=a0(1+z)a=\frac{a_{0}}{(1+z)}, where a0=1a_{0}=1 is the present value of scale factor, as

ρϕ(z)=eα1+ztan1(1+z)α,\rho_{\phi}(z)=e^{\frac{-\alpha}{1+z}}tan^{-1}(1+z)^{\alpha}, (14)

and the present value of energy density i.e. at z=0,

ρϕ0=π4eα.\rho_{\phi_{0}}=\frac{\pi}{4}e^{-\alpha}. (15)

Eq. (14) and (15) yield

ρϕ(z)=4πρϕ0eαz1+ztan1(1+z)α,\rho_{\phi}(z)=\frac{4}{\pi}\rho_{\phi_{0}}e^{\frac{\alpha z}{1+z}}tan^{-1}(1+z)^{\alpha}, (16)

Also using Eq. (9), energy density of matter field ρm\rho_{m} can be calculated in terms of redshift zz as

ρm(z)=ρm0(1+z)3\rho_{m}(z)=\rho_{m_{0}}(1+z)^{3} (17)

Together Eqs. (4), (16) and (17), we have

3H2=3β24+ρm0(1+z)3+4πρϕ0eαz1+ztan1(1+z)α,3H^{2}=\frac{3\beta^{2}}{4}+\rho_{m_{0}}(1+z)^{3}+\frac{4}{\pi}\rho_{\phi_{0}}e^{\frac{\alpha z}{1+z}}tan^{-1}(1+z)^{\alpha}, (18)

Now, by taking the gauge function β2=3β0H02a2α\beta^{2}=3\beta_{0}H_{0}^{2}a^{-2\alpha} and considering the density parameter Ω=ρρc\Omega=\frac{\rho}{\rho_{c}} which plays the key role to explain the whole content of the Universe. Here ρc=3H28πG\rho_{c}=\frac{3H^{2}}{8\pi G} is the critical density of the Universe.

Eq. (18) can be written in terms of density parameter of scalar field and matter,

H=14β0(1+z)2α+43πΩϕ0eαzz+1tan1(z+1)α+13Ωm0(z+1)3,H=\sqrt{\frac{1}{4}\beta_{0}(1+z)^{2\alpha}+\frac{4}{3\pi}\Omega_{\phi_{0}}e^{\frac{\alpha z}{z+1}}\tan^{-1}(z+1)^{\alpha}+\frac{1}{3}{\Omega_{m_{0}}}(z+1)^{3}}, (19)

where H0H_{0}, Ωm0\Omega_{m_{0}} and Ωϕ0\Omega_{\phi_{0}} are the current values of Hubble parameter, density parameters of matter and scalar field respectively. We consider the Hubble parameter HH as dimensionless quantity by dividing Eq. (18) by 3H0\sqrt{3}H_{0}, which yields Eq. (19).

The expression for deceleration parameter qq given by

q=aa¨a˙2=1+1+zHdHdz.q=-\frac{a\ddot{a}}{\dot{a}^{2}}=-1+\frac{1+z}{H}\frac{dH}{dz}. (20)

Thus, the deceleration parameter qq and EoS parameter for scalar field ωϕ\omega_{\phi} can be evaluated as

q(z)=12αβ0(1z+1)2α14αΩϕ0eα(11z+1)(1z+1)α13π((1z+1)2α+1)4αΩϕ0eα(11z+1)tan1((1z+1)α)3πΩm0(z+1)42(z+1)(14β0(1z+1)2α+4Ωϕ0eα(11z+1)tan1((1z+1)α)3π+13Ωm0(z+1)3)1,q(z)=-\frac{-\frac{1}{2}\alpha\beta_{0}(\frac{1}{z+1})^{-2\alpha-1}-\frac{4\alpha\Omega_{\phi_{0}}e^{\alpha(1-\frac{1}{z+1})}(\frac{1}{z+1})^{-\alpha-1}}{3\pi((\frac{1}{z+1})^{-2\alpha}+1)}-\frac{4\alpha\Omega_{\phi_{0}}e^{\alpha(1-\frac{1}{z+1})}\tan^{-1}((\frac{1}{z+1})^{-\alpha})}{3\pi}-\Omega_{m_{0}}(z+1)^{4}}{2(z+1)(\frac{1}{4}\beta_{0}(\frac{1}{z+1})^{-2\alpha}+\frac{4\Omega_{\phi_{0}}e^{\alpha(1-\frac{1}{z+1})}\tan^{-1}((\frac{1}{z+1})^{-\alpha})}{3\pi}+\frac{1}{3}\Omega_{m_{0}}(z+1)^{3})}-1, (21)
ωϕ(z)=πeα(1z+11)(4αΩϕeα(11z+1)(1z+1)α1π((1z+1)2α+1)4αΩϕ0eα(11z+1)tan1((1z+1)α)π)12Ωϕ0(z+1)2tan1((1z+1)α)\displaystyle\omega_{\phi}(z)=-\frac{\pi e^{\alpha(\frac{1}{z+1}-1)}(-\frac{4\alpha\Omega\phi e^{\alpha(1-\frac{1}{z+1})}(\frac{1}{z+1})^{-\alpha-1}}{\pi((\frac{1}{z+1})^{-2\alpha}+1)}-\frac{4\alpha\Omega_{\phi_{0}}e^{\alpha(1-\frac{1}{z+1})}\tan^{-1}((\frac{1}{z+1})^{-\alpha})}{\pi})}{12\Omega_{\phi_{0}}(z+1)^{2}\tan^{-1}((\frac{1}{z+1})^{-\alpha})}
×14β0(1z+1t)2α+4Ωϕeα(11z+1)tan1((1z+1)α)3π+13Ωm0(z+1)312Ωϕ0(z+1)2tan1((1z+1)α)1.\displaystyle\times\frac{\sqrt{\frac{1}{4}\beta_{0}(\frac{1}{z+1}t)^{-2\alpha}+\frac{4\Omega\phi e^{\alpha(1-\frac{1}{z+1})}\tan^{-1}((\frac{1}{z+1})^{-\alpha})}{3\pi}+\frac{1}{3}\Omega_{m_{0}}(z+1)^{3}}}{12\Omega_{\phi_{0}}(z+1)^{2}\tan^{-1}((\frac{1}{z+1})^{-\alpha})}-1. (22)

In the next section, we constrain the model parameter α\alpha and β0\beta_{0} using different observational datasets and obtain their best fit value. Thereafter use these values to examine the behaviours of various physical parameters.

III Statistical observation for model parameters

Observational cosmology plays a major role to explore the evolution of the Universe. Therefore, in the current section, we are using the expression for the Hubble parameter from Eq. (19) to perform statistical analysis with three observational datasets namely the Hubble dataset (77 points) [40], Type Ia Supernovae dataset [41] and their joint datasets and to acquire the present best fit value of the model parameters α\alpha and β0\beta_{0}. Here, for the likelihoods minimization of χ2\chi^{2} for the goodness of fit of the model, we use the Markov chain Monte Carlo (MCMC) method to investigate the parameter space from the python package emcee [42], which is extensively used in observational cosmology.

In this process, the present value of Hubble parameter, density parameters of matter, and scalar field are taken as 67.4KmSec1Mpc167.4KmSec^{-1}Mpc^{-1}, 0.315, 0.685 respectively, which are given in recent Planck data [43].

Refer to caption
(a)
Refer to caption
(b)
Refer to caption
(c)
Figure 1: Figs. (a), (b) and (c) illustrate the constraints on the model for the data OHDOHD, SNeIaSNeIa, and joint dataset OHD+SNeIaOHD+SNeIa respectively corresponding to 1σ1\sigma and 2σ2\sigma confidence regions.

III.1 Supernovae type Ia Data

We use SNeIaSNeIa Union 2.12.1 compilation dataset and constrained both the model parameters. For flat universe Chi-square function χSN2(α,β0,H0)\chi_{SN}^{2}(\alpha,\beta_{0},H_{0}) can be stated as

χSN2(α,β0,H0)=i=1580[μth(α,β0,H0,zi)μobs(zi)σμ(zi)]2.\chi_{SN}^{2}(\alpha,\beta_{0},H_{0})=\sum\limits_{i=1}^{580}\left[\frac{\mu_{th}(\alpha,\beta_{0},H_{0},z_{i})-\mu_{obs}(z_{i})}{\sigma_{\mu(z_{i})}}\right]^{2}. (23)

where SNSN is stand for the observational SNeIaSNeIa dataset. μth\mu_{th} and μobs\mu_{obs} represents theoretical and observed distance modulus of the model. The standard error in the observed value is denoted by σμ(zi)\sigma_{\mu(z_{i})}. Also, the theoretical distance modulus μ(z)\mu(z) is defined by

μ(z)=mM=5LogDl(z)+μ0,\mu(z)=m-M=5LogD_{l}(z)+\mu_{0}, (24)

where mm and MM are used for the apparent and absolute magnitudes of a standard candle respectively. The luminosity distance Dl(z)D_{l}(z) for flat Universe and the nuisance parameter μ0\mu_{0} are given by

Dl(z)=(1+z)c0z1H(z)𝑑z,D_{l}(z)=(1+z)c\int_{0}^{z}\frac{1}{H(z^{*})}dz^{*}, (25)

and

μ0=5Log(H011Mpc)+25,\mu_{0}=5Log\Big{(}\frac{H_{0}^{-1}}{1Mpc}\Big{)}+25, (26)

respectively. Here, we perform a global fitting to dictate the model parameters using the Markov chain Monte Carlo method and Python implementation of the ensemble sampler for the MCMC method with the EMCEE library, introduced by Foreman-Mackey et al. to show the best-fit value of model parameters written in Table-I.

III.2 Hubble Dataset

Since Hubble data is directly based on differential ages of the galaxies and related to the expansion history of the universe thus this is very useful to understand the dark section i.e. dark energy, dark matter, and dark ages of the universe. The value HH can be written in terms of zz as

H(z)=(1+z)1dzdt.H(z)=-(1+z)^{-1}\frac{dz}{dt}. (27)

The best-fit values of the model parameters are determined by minimizing the Chi-square value

χHD2(α,β0)=i=177[H(α,β0,zi)Hobs(zi)]2σzi2,\chi_{HD}^{2}(\alpha,\beta_{0})=\sum\limits_{i=1}^{77}\frac{[H(\alpha,\beta_{0},z_{i})-H_{obs}(z_{i})]^{2}}{\sigma_{z_{i}}^{2}}, (28)

where H(α,β0,zi)H(\alpha,\beta_{0},z_{i}) and HobsH_{obs} represent the theoretical and observed values respectively and σzi\sigma_{z_{i}} indicates the standard deviation for every H(zi)H(z_{i}). For H(z)H(z) 77 data points, the best fit values of α\alpha and β0\beta_{0} are estimated in Table-I.

Table 1: Best fit values of model parameters.
Dataset       α\alpha       β0\beta_{0}
SNeIaSNeIa       0.6350.060+0.0600.635^{+0.060}_{-0.060}       0.19350.0083+0.00830.1935^{+0.0083}_{-0.0083}
H(z)H(z)       0.6870.0025+0.0130.687^{+0.013}_{-0.0025}       0.1570.0098+0.0440.157^{+0.044}_{-0.0098}
H(z)H(z) + SNeIaSNeIa      0.6860.0027+0.0140.686^{+0.014}_{-0.0027}       0.18680.0031+0.00250.1868^{+0.0025}_{-0.0031}

The Gelman-Rubin convergence test [44] is a widely used statistical tool in Bayesian inference that allows to access the convergence of the chains. The test is based on the idea that multiple MCMC chains with different starting points should converge to the same posterior distribution if they have been run for long enough, that is, after a number of steps. Basically, this test measures, for each parameter of the considered model, the quantity called potential scale reduction R^\hat{R}, which is the ratio between the variance W\rm W within a chain and the variance Var(θ)\rm Var(\theta) among the chains:

R^=V^ar(θ)W.\hat{R}=\sqrt{\frac{\rm\hat{V}ar(\theta)}{W}}. (29)

The value of R^\hat{R} for a perfectly converged chain should be 11, so R^1.1\hat{R}\approx 1.1 corresponds to the maximum allowed value for this parameter for which the convergence of the chain is achieved.

We then apply the Gelman-Rubin convergence test to chains of the fits of α\alpha and β0\beta_{0} from Hubble data alone. The test yielded R^=1.03008001\hat{R}=1.03008001, for α\alpha and R^=1.02498708\hat{R}=1.02498708 for β0\beta_{0}, respectively. As 10001000 steps with 100100 steps of burn-in were used to run the MCMC chains, in Fig. 2 we verify that convergence, for both parameters, is reached just after approximately 250250 steps when crossing the horizontal line that represents the convergence criterion.

Refer to caption
Figure 2: The evolution of the potential scale reduction R^\hat{R} with the chain length for α\alpha and β\beta.

III.3 Joint Datasets (H(z)+SNeIaH(z)+SNeIa)

The χ2\chi^{2} function for joint analysis is given by

χHS2=χHD2+χSN2.\chi_{HS}^{2}=\chi_{HD}^{2}+\chi_{SN}^{2}. (30)

Using this joint statistical analysis, the stronger constraints of the model parameter can be obtained. For joint datasets, the best fit values of α\alpha and β0\beta_{0}, can be seen in Fig. 1(c) and Table-I. Now to compare our model with ΛCDM\Lambda CDM, we use error bar plots for Hubble datasets and Supernovae datasets. The values of model parameters are constrained from observational datasets. In Fig. 3, the error bar plots of the Hubble and Type Ia Supernovae dataset show the best fit plots.

Refer to caption
(a)
Refer to caption
(b)
Figure 3: The error bar plots in respect of H(z)H(z) and Type Ia Supernovae datasets show the similarity between our model and Λ\LambdaCDM.

IV Evolution of the Universe

In this section, we study the evolution of various cosmological quantities using the best-fit values of the model parameters obtained from different observational datasets. Energy density for the scalar field is decreasing with respect to redshift zz. In Fig. 4(a), It is clearly visible that ρϕ\rho_{\phi} monotonically decreases and approaches zero in late times, which indicates that amount of dark energy density of the scalar field diminished in the future. In Fig. 4(b), the isotropic pressure of the scalar field is negative for all values of zz, which indicates the presence of dark energy up to late times.

Refer to caption
(a)
Refer to caption
(b)
Figure 4: The graphs of ρϕ\rho_{\phi} and pϕp_{\phi} w.r.t. zz.

Also the plots of EoS parameter ωϕ\omega_{\phi} and ωeff=pϕρϕ+ρm\omega_{eff}=\frac{p_{\phi}}{\rho_{\phi}+\rho_{m}} show the quintessence model in late times. Finally, our model shows the quintessence model in late times as it transits from the quintessence region to a perfect fluid state and again converges to the quintessence region in the future (see Fig. 5).

Table 2: The present values of the cosmological parameters.
Dataset       HH         qq       ωeff\omega_{eff}
SNeIaSNeIa        70.80270.802         0.1961-0.1961       0.5283-0.5283
H(z)H(z)        70.220270.2202         0.1604-0.1604       0.5384-0.5384
H(z)H(z) + SNeIaSNeIa       70.639670.6396        0.1640-0.1640       0.5270-0.5270
Refer to caption
(a)
Refer to caption
(b)
Figure 5: The graphs of ωϕ\omega_{\phi} and ωeff\omega_{eff}  vs. zz for distinct observational data.

Fig. 6(a) shows the evolution of the deceleration parameter for different observational datasets. In this model, we observe that the Universe has three transit points during the evolution of each observational data. In early time universe transited from deceleration to acceleration, at present, it shows the acceleration phase. In late time again the universe is having a transition from acceleration to deceleration and afterwords deceleration to acceleration. These transition points are shown in Table II. In our model, we observe that the present value of qq is not very near to 0.5 and it is relevant to point out that q0.5q\approx-0.5 stands for Λ\LambdaCDM model and constraints over this parameter using the CMB data. Therefore, new theories of gravity may admit different values for the deceleration parameter, as we can see in the following works [45, 46]. Moreover, the crises over H0H_{0} in different regimes of redshift may also open the possibility of different values for q0q_{0} and ω0\omega_{0}, as we can see in [47, 48].

Table 3: Transition points of deceleration parameter for different datasets.
Dataset       ztr1z_{tr_{1}}         ztr2z_{tr_{2}}       ztr3z_{tr_{3}}
SNeIaSNeIa        0.75780.7578         0.5815-0.5815       0.8824-0.8824
H(z)H(z)        0.70930.7093         0.4942-0.4942       0.8824-0.8824
H(z)H(z) + SNeIaSNeIa       0.71890.7189        0.5136-0.5136       0.8824-0.8824
Refer to caption
(a)
Refer to caption
(b)
Figure 6: The plot of qq and jerk parameter w.r.t. zz.

To understand the evolution of the expanding universe in a more significant way, we explore some cosmographic parameters which contain the higher-order derivatives of scale factor aa. Here, we intend to discuss a kinematic quantity that contains the third-order derivative and is known as the jerk parameter. The Jerk parameter jj is defined as

j=a˙˙˙aH3=q+2q(1+q)+(1+z)dqdz.j=\frac{\dddot{a}}{aH^{3}}=-q+2q(1+q)+(1+z)\frac{dq}{dz}. (31)

The standard value of the Jerk parameter jj for Λ\LambdaCDM model is one, thus digression from j=1j=1 explores the evolution of different kinds of dark energy models. Fig. 6(b) highlights that at the early time our model is similar to Λ\LambdaCDM model and afterward this deviates from Λ\LambdaCDM till the late times.

To differentiate our model with Λ\LambdaCDM, we also use statefinder diagnostic technique, which is established by Sahni et al. [49, 50]. The two geometrical diagnostic parameters (s,r)(s,r), known as statefinder parameters are defined as

r=a˙˙˙aH3s=r13(q12),q12r=\frac{\dddot{a}}{aH^{3}}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}~{}s=\frac{r-1}{3(q-\frac{1}{2})},~{}q\neq\frac{1}{2} (32)

Using statefinder diagnostic parameters, we can compare the goodness of various dark energy models with Λ\LambdaCDM. In srs-r plot, the points (0,1)(0,1) and (1,1)(1,1) represent Λ\LambdaCDM model and SCDMSCDM respectively. From Eqs. (21), (31) and (32), rr and ss can be calculated, and by using the best-fit values of model parameters we draw these trajectories (see Fig. 7(a)). The arrows on the curves show the direction of evolution. In our model r<1r<1 and s>0s>0 at an early time, which corresponds to the quintessence DE model. The present values of (s,r){(s,r)} are evaluated (0.5214,0.0330),(0.4729,0.0125),(0.5190,0.0338){(0.5214,-0.0330),(0.4729,-0.0125),(0.5190,-0.0338)} from OHDOHD, SNeIaSNeIa and their joint datasets respectively, which show the deviations from Λ\LambdaCDM at present.

In Fig. 7(b), it is highlighted that the qrq-r curves start from SCDMSCDM (1,12)(1,\frac{1}{2}) and enter into the acceleration zone for all datasets in the future. In this plot, the dotted horizontal line r=1r=1 denotes Λ\LambdaCDM. Here, we observe that the trajectories are crossing the transition line thrice from early evolution up to late times. As qq is negative in late times, therefore this model shows an acceleration phase of the universe in the future.

Refer to caption
(a)
Refer to caption
(b)
Figure 7: The Statefinder plots srs-r and qrq-r

V Conclusion

In this model, we have explored the late-time cosmic expansion of the universe in the framework of a flat FLRW space-time metric with time-dependent displacement vectors based on Lyra’s geometry. We have derived EEF from the principal action (1) and find solutions. Using the recent observational datasets OHDOHD, SNeIaSNeIa and OHD+SNeIaOHD+SNeIa, we have obtained the best-fit values of model parameters α\alpha, β\beta by applying MCMC method with emcee package in Python. Furthermore, we investigated the behavior of cosmological parameters for the constrained values of model parameters, which are evaluated in Tables I, and II.

The energy density of scalar field ρϕ\rho_{\phi} decreases monotonically from high redshift to low redshift and ρϕ0\rho_{\phi}\rightarrow 0 as z1z\rightarrow-1, \forall observations. The pressure pϕp_{\phi} is having negative finite value throughout the range, which indicates the accelerating behavior of the Universe (see Fig. 4). The EoS parameter of the scalar field ωϕ\omega_{\phi} depicts that our model is a quintessence dark energy model as ωϕ\omega_{\phi} is having a value between -1 to 0 for the large range of zz. Its value becomes positive for a while, and at the end, it converges to -1 for all datasets. In addition, it clearly highlighted the evolution profile of the deceleration parameter involving three transitions for each dataset which are shown in Table III. The jerk parameter shows that our model is similar to Λ\LambdaCDM in the early universe and it deviates in the future (see Fig. 5,6). The statefinder diagnostic graphs show that for all datasets, srs-r trajectories are looking in the form of loops in our model and passing through the Λ\LambdaCDM model. Also, three qrq-r trajectories, which are also loops for the present model, starts from SCDM and show acceleration at its final phase (see Fig. 7). Thus, in this paper, we conclude that this model is an accelerated expanding model having the characteristics of the quintessence model studied in Lyra geometry, which is different from Λ\LambdaCDM model. It is relevant to point that the approach here presented could be applied in other theories of gravity, such as f(R)f(R) [51], f(R,T)f(R,T) [52], f(Q)f(Q) [53, 54], and f(Q,T)f(Q,T) [55, 56]. Such applications could give us extra parameters to test the viability of Lyra geometry with respect to experimental data. We hope to report these extra contributions in near future.

Acknowledgements.
JRLS would like to thank CNPq (Grant no. 309494/2021-4), and PRONEX/CNPq/FAPESQ-PB (Grant nos. 165/2018, and 0015/2019) for financial support. JASF thanks FAPES for financial support.

References

  • [1] S. Weinberg, Rev. Mod. Phys. 61, 1-23 (1989).
  • [2] E. J. Copeland, M. Sami and S. Tsujikawa, Int. J. Mod. Phys. D 15, 1753-1936 (2006).
  • [3] H. A. Buchdahl, Mon. Not. Roy. Astron. Soc. 150, 1 (1970)
  • [4] R. Kerner, Gen. Rel. Grav. 14, 453-469 (1982).
  • [5] H. Kleinert and H. J. Schmidt, Gen. Rel. Grav. 34, 1295-1318 (2002).
  • [6] S. M. Carroll, V. Duvvuri, M. Trodden and M. S. Turner, Phys. Rev. D 70, 043528 (2004).
  • [7] S. Capozziello, S. Nojiri, S. D. Odintsov and A. Troisi, Phys. Lett. B 639, 135-143 (2006).
  • [8] L. Amendola, D. Polarski and S. Tsujikawa, Phys. Rev. Lett. 98, 131302 (2007).
  • [9] S. Nojiri and S. D. Odintsov, Phys. Rev. D 68, 123512 (2003).
  • [10] V. Faraoni, Phys. Rev. D 74, 023529 (2006).
  • [11] P. J. Zhang, Phys. Rev. D 76, 024007 (2007).
  • [12] L. Amendola and S. Tsujikawa, Phys. Lett. B 660, 125-132 (2008).
  • [13] S. Tsujikawa, Phys. Rev. D 77, 023507 (2008).
  • [14] T. Liu, X. Zhang and W. Zhao, Phys. Lett. B 777, 286-293 (2018).
  • [15] S. Capozziello and S. Tsujikawa, Phys. Rev. D 77, 107501 (2008).
  • [16] A. A. Starobinsky, JETP Lett. 86, 157-163 (2007).
  • [17] S. Nojiri and S. D. Odintsov, Phys. Lett. B 657, 238-245 (2007).
  • [18] S. Nojiri and S. D. Odintsov, Phys. Rev. D 77, 026007 (2008).
  • [19] G. Cognola, E. Elizalde, S. Nojiri, S. D. Odintsov, L. Sebastiani and S. Zerbini, Phys. Rev. D 77, 046009 (2008).
  • [20] J. Santos, J. S. Alcaniz, M. J. Reboucas and F. C. Carvalho, Phys. Rev. D 76, 083513 (2007).
  • [21] S. Capozziello, V. F. Cardone and V. Salzano, Phys. Rev. D 78, 063504 (2008).
  • [22] O. Bertolami, C. G. Boehmer, T. Harko and F. S. N. Lobo, Phys. Rev. D 75, 104016 (2007).
  • [23] T. Harko, Phys. Lett. B 669, 376-379 (2008).
  • [24] G. Lyra, Math. Z. 54, 52 (1951).
  • [25] R. R. Cuzinatto, E. M. De Morais and B. M. Pimentel, Phys. Rev. D 103 (2021) no.12, 124002.
  • [26] A. A. Penzias and R. W. Wilson, Astrophys. J. 142, 419-421 (1965).
  • [27] K. Bhamra, Australian Journal of Physics 27, 541 (1974).
  • [28] A. Beesham, Astrophysics and Space Science 127, 355 (1986).
  • [29] W. D. Halford, Australian Journal of Physics 23, 863 (1970).
  • [30] D. R. Reddy and R. Venkateswarlu, Astrophysics and Space Science 149, 287 (1988).
  • [31] T. Singh and G. P. Singh, J. Math. Phys. 32, 2456-2458 (1991).
  • [32] T. Singh and G. P. Singh, Nuovo Cimento, B106, 617 (1991).
  • [33] T. Singh and G. P. Singh, Int. J. Theor. Phys. 31, 1433-1446 (1992).
  • [34] D. Sen, Canadian Mathematical Bulletin 3, 255 (1960).
  • [35] D. K. Sen and K. A. Dunn, J. Math. Phys. 12, no.4, 578-586 (1971).
  • [36] G. P. Singh and D. Kalyani, Pramana 49, 205-212 (1997).
  • [37] D. K. Sen, Zeitschrift f¨ur Physik 149, 311 (1957).
  • [38] J. K. Singh, Int. J. Theor. Phys. 48, 905-912 (2009).
  • [39] J. K. Singh and R. Nagpal, Eur. Phys. J. C 80, no.4, 295 (2020).
  • [40] Shaily, M. Zeyauddin and J. K. Singh, [arXiv:2207.05076 [gr-qc]].
  • [41] D. M. Scolnic et al. [Pan-STARRS1], Astrophys. J. 859 (2018) no.2, 101.
  • [42] D. Foreman-Mackey, D. W. Hogg, D. Lang and J. Goodman, Publ. Astron. Soc. Pac. 125 (2013), 306-312.
  • [43] N. Aghanim et al. [Planck], Astron. Astrophys. 641 (2020), A6 [erratum: Astron. Astrophys. 652 (2021), C4].
  • [44] A. Gelman and D. B. Rubin, Statist. Sci. 7 (1992), 457-472
  • [45] C. P. Singh and M. Srivastava, Eur. Phys. J. C 78, no.3, 190 (2018)
  • [46] S. K. Sahu, S. K. Tripathy, P. K. Sahoo and A. Nath, Chin. J. Phys. 55, 862-869 (2017)
  • [47] P. D. Alvarez, B. Koch, C. Laporte, F. Canales and A. Rincon, JCAP 10, 071 (2022)
  • [48] D. Camarena and V. Marra, Phys. Rev. Res. 2, no.1, 013028 (2020)
  • [49] U. Alam, V. Sahni, T. D. Saini and A. A. Starobinsky, Mon. Not. Roy. Astron. Soc. 344 (2003), 1057.
  • [50] V. Sahni, T. D. Saini, A. A. Starobinsky, U. Alam, JETP Lett. 77 201 (2003).
  • [51] M. Rinaldi, G. Cognola, L. Vanzo and S. Zerbini, JCAP 08, 015 (2014).
  • [52] P. H. R. S. Moraes and J. R. L. Santos, Eur. Phys. J. C 76, 60 (2016).
  • [53] J. Beltrán Jiménez, L. Heisenberg and T. Koivisto, Phys. Rev. D 98, no.4, 044048 (2018).
  • [54] S. Mandal, P. K. Sahoo and J. R. L. Santos, Phys. Rev. D 102, no.2, 024057 (2020).
  • [55] Y. Xu, G. Li, T. Harko and S. D. Liang, Eur. Phys. J. C 79, no.8, 708 (2019).
  • [56] S. Arora, J. R. L. Santos and P. K. Sahoo, Phys. Dark Univ. 31, 100790 (2021).