Teleparallel Gravity and Quintessence: The Role of Nonminimal Boundary Couplings
Teleparallel Gravity and Quintessence: The Role of Nonminimal Boundary Couplings
Teleparallel Gravity and Quintessence: The Role of Nonminimal Boundary Couplings
1 Department of Mathematics, Birla Institute of Technology and Science-Pilani, Hyderabad Campus, Hyderabad-500078, India
Abstract: In this paper, we have outlined the development of an autonomous dynamical system
within a general scalar-tensor gravity framework. This framework encompasses the overall structure
of the non-minimally coupled scalar field functions for both the torsion scalar (T) and the bound-
ary term (B). We have examined three well-motivated forms of potential functions and constrained
the model parameters through dynamical system analysis. This analysis has played a crucial role in
identifying cosmologically viable models. We have analysed the behaviour of dynamical paramet-
ers such as equation-of-state parameters for dark energy and the total, as well as all the standard
density parameters for radiation, matter, and dark energy to assess their compatibility with current
observational data. The phase space diagrams are presented to support the stability conditions of
arXiv:2408.03417v1 [gr-qc] 6 Aug 2024
the corresponding critical points. The Universe is apparent in its late-time cosmic acceleration phase
via the dark energy-dominated critical points. Additionally, we compare our findings with the most
prevailing ΛCDM model. The outcomes are further inspected using the cosmological data sets of
Supernovae Ia and the Hubble rate H (z).
Keywords: Teleparallel Gravity, Dynamical System Analysis, Numerical Solutions, Boundary Term
I. INTRODUCTION
Einstein’s General Relativity (GR) theory has successfully explained various astrophysical phenomena, from tests
within our solar system to studying strong field gravitational wave physics [1]. The observations of galaxies and
their dynamic structures suggest the existence of dark matter (DM), which interacts only through gravity and may
not fit into the standard model of particle physics [2, 3]. Moreover, the modification of GR has been significantly
influenced by the recent observation of the accelerating expansion of the Universe [4, 5], which is predicted to be
caused by an observational fact known as dark energy (DE). The ΛCDM model, one of the most promising models,
proposes that DE is represented by a cosmological constant with a constant EoS (ω = -1) [6]. Although favored by
many cosmological observations, this model faces several theoretical [7] and observational difficulties like the recent
findings from the Planck collaboration have shown an increasing discrepancy between local and global measure-
ments of H0 and f σ8 [8]. It is possible that the issues with ΛCDM theory may be resolved in the coming years, or it
may be necessary to modify ΛCDM in some way. Efforts have been made in recent decades to extend GR to address
certain aspects of these issues [9]. However, it may also be necessary to consider a new approach to address the
increasing demands for developing a viable theory of gravity. One such approach is to study the teleparallel gravity
theories [10–12].
In contrast to GR, the teleparallel equivalent of general relativity (TEGR) offers a different perspective on grav-
ity. TEGR uses a specific connection known as the Weizenböck connection, which characterizes a space-time with
non-zero torsion. TEGR considers torsion a force, while in GR, gravity is understood as the effect of curved space-
time. Similar to f ( R) extends in GR where R is the Ricci Scalar, TEGR has been further developed by incorporating
a function of the torsion scalar T, denoted as f ( T ), into its Lagrangian density [11, 13–15]. However, f ( R) and
f ( T ) theories lead to different dynamics, as equating f ( R) with f ( T ) and the total derivative is no longer possible.
Specifically, f ( R) gravity generally yields fourth-order field equations, while f ( T ) theories are somewhat less prob-
lematic as this modification only results in second-order field equations; however, the loss of local Lorentz invariance
is the price paid for this. Researchers in cosmology are studying f ( T ) gravity. The constraints on f ( T ) cosmology
with Pantheon+ show potential in explaining aspects of cosmic evolution [16]. The spherically symmetric solutions
∗ [email protected]
† [email protected]
‡ [email protected]
2
for f ( T ) gravity models can be obtained using the Noether Symmetry Approach [17]. Furthermore, f ( T ) theory’s
validity has been supported through the solar system tests [18].
In scalar-tensor theories based on GR, a common coupling function in the corresponding action is represented by
a term such as ξRϕ2 [19–21]. Analogous to this, the same coupling function replacing the R by T has been studied
in TEGR [22–25]. These considerations are well suited in describing the behaviour of the equation of state (EoS)
parameter around the value −1 at the present time [23]. In the TEGR, interesting scenarios are represented in a
dynamic model resulting from the non-minimal coupling between the quintessence field similar to the canonical
scalar field and T [26]. This action was further modified and analysed considering the non-minimal coupling to the
boundary term B in the teleparallel quintessence formalism [27]. This formalism can be generalised by replacing the
coupling coefficients of T and the boundary term B with a general function of the scalar field f (ϕ), g(ϕ), in which the
generalised second law of thermodynamics is studied in [28]. Also, the Lorentzian wormholes are constructed using
the Noether symmetry [29], which helps to establish some important cosmological solutions for the modified field
equations [30]. The linear stability technique is used to analyse the dynamical properties of the current tachyonic DE
model. This analysis involves different potentials and couplings, considering a generalized non-minimal coupling
of a tachyonic scalar field with the teleparallel boundary term [31]. To study the cosmological dynamics of tachyonic
teleparallel DE, one can refer to [32, 33].
In this work, we focus on constructing and studying an autonomous dynamical system in one of the most general
forms of scalar-tensor teleparallel formalism. The comprehensive study of the dynamical system analysis approach
in different modified teleparallel gravity formalisms has been presented in [34–42]. In this paper, we follow the
approach in which we can analyse the dynamics of this scalar tensor formalism in a possible general way. In this
formalism, the nonminimal coupling to T and boundary term B is demonstrated and is constructed in [27]. It is
studied in a more general way in [28–30]. The presented work establishes the dynamical system analysis formal-
ism with the case where the coupling functions of T and B are the general functions of the canonical scalar field ϕ.
Moreover, in this theory construction, we came across an important point that this formalism obeys the conservation
equation. In constructing the dynamical system, we adhere to the exponential coupling function of the scalar field
ϕ to the boundary term B in the absence of the nonminimal coupling function of T. The well-motivated different
forms of the potential function V (ϕ) have been analysed. The paper is organized as follows: In Sec. II, we introduce
the generalised teleparallel gravity with the non-minimal couplings with T, B and derive the corresponding evol-
ution equations. In Sec. III, we conduct a dynamical system analysis of the DE model and reveal the fundamental
discussion of the critical points. Moving on to Sec. III A to Sec. III C, we employ analysis of different scalar field
potentials. In each section, the numerical approach is used to study the evolution of the field equations for the model
and analyse the implications of T and B couplings for the study of DE EoS parameter. We have also analysed the
Hubble rate evolution as a function of redshift z and demonstrated its compatibility with the Hubble rate HΛCDM (z)
in the ΛCDM model, with the Hubble data points [43]. The ΛCDM model modulus function µΛCDM , 1048 pantheon
data points, and modulus function µ(z) are also taken into analysis. Finally, in Sec. IV, we present the conclusions
of the study and the final remarks.
In this section, we have presented the basic equations and formalism to construct the TEGR, as well as its scalar
µ
torsion formalism. In this theory, the tetrad eµa and its inverses Ea serve as the dynamical variable [11, 44, 45]. The
Greek indices represent spacetime indices, and the Latin indices denote the tangent space indices. The metric gµν is
expressed in terms of the tetrad as,
Where ηab = diag (−1, 1, 1, 1) represents the Minkowski space-time metric. Like to the metric, the tetrads meet
orthogonality conditions that are in the form of
e a µ Eb = δba , e a µ Ea ν = δµν .
µ
(2)
3
Similar to the Levi-Cività connection used in GR, the Weitzenböck connection (Γσνµ ) is chosen in TG [46, 47]. The
Weitzenböck connection is expressed in terms of the derivatives of the tetrad as,
Γσνµ := Eaσ ∂µ eνa + ω abµ eνb , (3)
The expression ω abµ represents the spin connection and is part of a linear affine connection that lacks curvature and
meets the metricity condition. To maintain the covariance of the field equations, the spin connection is explicitly
included in the Weitzenböck connection [47, 48]. The torsion scalar can be obtained by using the contractions of the
torsion tensor as follows,
1 α µν 1 νµ β νµ
T µν Tα + T αµν T α − T βµ Tν := T . (4)
4 2
In the context of exchanging the Levi-Civita connection with the teleparallel connection, it is observed that meas-
ures of curvature identically vanish, such as R ≡ 0. In teleparallel theories, T can be considered a counterpart to the
curvature R in GR. For more details about teleparallel theory, please see reference [10, 45]. In this scenario, we can
express the following relation for the curvature and torsion scalars [49, 50].
2
− T + ∂µ(eT µ ) = − T + B = R . (5)
e
√
This relation includes R, T, and B = 2e ∂µ(eT µ ) represents a total divergence term, and e = det eµa = − g stands
for the determinant of the tetrad. This ensures that the GR and TEGR actions will produce identical field equations.
Initially, to study the late time cosmic acceleration, the works are done in the quintessence scalar field where
the power law coupling potentials are considered [51, 52], for review in curvature formalism one can refer to [53].
The coupling coefficient ξϕ2 of T was considered in the teleparallel modifications and was analysed in [24, 54]. To
analyse the scaling attractors in the teleparallel gravity formalism, the ξϕ2 coupling coefficient is generalized to ξ f (ϕ)
in [21]. Moreover, in [27], this formalism has further modified the teleparallel approach by considering the inclusion
of the nonminimal coupling ξϕ2 , χϕ2 to the T, and B respectively. The more general form where the general scalar
field functions f (ϕ), g(ϕ) are incorporated to study the generalised second law of thermodynamics and the Noether
symmetry approach in [28, 30]. Moreover, the generalized non-minimal coupling of a tachyonic scalar field with
the teleparallel boundary term is studied in [31]. The Lorentzian wormholes are constructed in this formalism by
Noether symmetries [29]. During the literature study, we came across the scope of the study of dynamical system
analysis, which has not been studied previously in this general scalar-tensor formalism. Hence in this study, we aim
to construct an autonomous dynamical system to analyse the viability of the different well-motivated scalar field
potentials to analyse the different epochs of the Universe’s evolution. We consider the action formula in which the
scalar field is non-minimally coupled to both T and B as follows,
−T 1
Z
4
S= − f ( ϕ ) T + g ( ϕ ) B − ∂ µ ϕ∂ µ
ϕ − V ( ϕ ) + S m + S r ed x. (6)
2κ 2 2
In modified teleparallel gravity, the gravitational field equations and their solutions depend on the spin connection.
Therefore, it’s crucial to have a method for determining the corresponding spin connection for each tetrad field in
order to accurately solve the field equations. In the context of FLRW cosmologies, it has been demonstrated that the
diagonal tetrad is a suitable tetrad presented as follow,
This means that the appropriate spin connection associated with this tetrad is the vanishing spin connection, res-
ulting in physically meaningful outcomes [47, 48, 55]. This tetrad choice leads to the flat Friedmann-Lemaı̂tre-
Robertson-Walker (FLRW) metric as,
The tetrad field allows for expressing the T and B in terms of the scale factor and its time derivatives as,
T = 6H 2 , B = 6(3H 2 + Ḣ ) , (9)
where H ≡ aȧ is the Hubble parameter, a dot represents the derivative with respect to time. To obtain the contribution
of the DE for pressure and the energy density, we use the general Friedmann equations which can be written as,
3H 2 = κ 2 ρm + ρr + ρ DE ,
(10)
3H 2 + 2 Ḣ = −κ 2 pr + p DE .
(11)
Varying the above action Eq. (6), the field equations can be obtained as,
" #
2 2 2 ′ ϕ̇2
3H =κ ρm + ρr − 3H f (ϕ) + 3Hg (ϕ)ϕ̇ + V (ϕ) + , (12)
2
" #
2 2
2
′ ϕ̇2 2 ′′ ′
3H + 2 Ḣ = − κ pr − V (ϕ) + 3H + 2 Ḣ f (ϕ) + 2H f (ϕ)ϕ̇ + − ϕ̇ g (ϕ) − g (ϕ)ϕ̈ . (13)
2
Where prime denotes the differentiation with respect to the scalar field ϕ. With the same setting, the Klein-Gordon
equation can be obtained as,
B ′ T ′ ′
ϕ̈ + 3H ϕ̇ + g (ϕ) + f (ϕ) + V (ϕ) = 0 . (14)
2 2
Comparing Eqs. (12)-(13) and Eqs. (10)-(11), one can retrieve the field equations required to analyse the dynamics
of the DE as,
′ ϕ̇2
−3H 2 f (ϕ) + 3Hg (ϕ)ϕ̇ + V (ϕ) + = κ 2 ρ DE , (15)
2
′ ϕ̇2 ′′ ′
−V (ϕ) + 3H 2 + 2 Ḣ f (ϕ) + 2H f (ϕ)ϕ̇ + − ϕ̇2 g (ϕ) − g (ϕ)ϕ̈ = κ 2 p DE . (16)
2
One can easily verify that the equations Eqs. (15)-(16) are satisfying the standard conservation equation as stated
below,
ρ̇ DE + 3H ρ DE + p DE = 0 , (17)
which is in agreement with the energy conservation law and the fluid evolution equations,
ρ̇m + 3Hρm = 0 ,
ρ̇r + 4Hρr = 0 .
(18)
The standard density parameters for matter, radiation, and the DE can be written as,
κ 2 ρm κ 2 ρr κ 2 ρ DE
Ωm = , Ωr = , Ω DE = , (19)
3H 2 3H 2 3H 2
Ωm + Ωr + Ω DE = 1. (20)
5
Here the case f (ϕ) ̸= 0 and g(ϕ) = 0 have been already analysed in [21], the detailed analysis of the particular
case f (ϕ) = ξ of this form is analysed in detail in [24, 54]. Therefore, in this case, we will analyse the scalar field
model non-minimally coupled to the boundary term. Hence we will consider f (ϕ) = 0 and g(ϕ) ̸= 0. For this, we
will define the dimensionless variables as follows,
√ √ ′
κ ϕ̇ κ V ′ κ ρr −V ( ϕ )
x= √ , y= √ , u = κg (ϕ) , ρ = √ , λ= . (21)
6H 3H 3H κV (ϕ)
Subsequently, Eq. (10) can be written in terms of dynamical variables as,
√
1 = Ωm + ρ2 + 6xu + y2 + x2 , (22)
where,
√
Ωϕ = 6xu + y2 + x2 . (23)
To frame an autonomous dynamical system, we have to consider the exponential coupling function to the boundary
term. In this case, we will consider the case where the coupling function g(ϕ) = g0 e−αϕκ [31, 56], which will be the
same in all the three following different potential function cases.
The dynamical system can be obtained by differentiating the dimensionless variables with respect to N = log( a)
as follows,
!
dx 9 3λy2 3u Ḣ
= −3x − √ u + √ − x+ √ ,
dN 6 6 6 H2
r
dy 3 Ḣ
=− xλy − y 2 ,
dN 2 H
du √
= − 6αux ,
dN
dρ Ḣ
= −2ρ − ρ 2 ,
dN H
dλ √ 2
= − 6xλ (Γ − 1) .
dN
(24)
Where,
√
Ḣ ρ2 + 9u2 + 6αux2 + 3 6ux − 3y2 (λu + 1) + 3x2 + 3
=− , (25)
H2 3u2 + 2
Ḣ
The substitution of from Eq. (25) into Eq. (24), will generate the autonomous dynamical system as follows,
H2
√ √
2x ρ2 + 9u2 − 3λuy2 − 3y2 − 3 + 6x3 (2αu + 1) + 3 6ux2 (2αu + 3) + 6 u ρ2 − 3y2 − 3 + 2λy2
dx
= ,
dN 6u2 + 4
√
y ρ2 + 9u2 + 6αux2 + 3 6ux − 3y2 (λu + 1) + 3x2 + 3
r
dy 3
=− λxy + ,
dN 2 3u2 + 2
du √
= − 6αux ,
dN !
√
ρ ρ2 + 3 u2 + ux 2αx + 6 − λuy2 + x2 − 3y2 − 1
dρ
= ,
dN 3u2 + 2
dλ √
= − 6x f . (26)
dN
6
where f = λ2 (Γ − 1) this form can be written in a generalise way as f = α1 λ2 + α2 λ + α3 [57]. The potentials that
we are going to study are presented as follows,
P3 − λ2
V0 Sinh−η ( βϕ) − ηβ2 1
0 −ηβ2
[58] η η
In the above Table I, V0 , ξ, η, β are the constants. We have demonstrated the dynamical analysis for each of the
above cases in detail in the following sections.
A. Potential P1 V0 e−κϕ
The exponential potential is considered in the literature to study the teleparallel DE scalar field models [57, 59, 60].
This potential also can be considered to study the generalized teleparallel non-minimally coupled tachyonic models
in the presence of the boundary term B [31]. The critical points throughout this study are the points at which the
dx dy du dρ
autonomous dynamical system vanishes and can be obtained through dN = 0, dN = 0, dN = 0, dN = 0. In this
case, the value of f = 0; hence, the above autonomous dynamical system reduces to the four dimensions. The
critical points, along with the value of q, ωtot and the standard density parameters Ωr , Ωm , Ω DE are presented in the
following Table II.
7
C. P. { x, y, u, ρ} q ωtot Ωr Ωm Ω DE
1
AR {0, 0, 0, 1} 1 3 1 0 0
√
2 q
2
BR 3
, √2 , 0, 1− 4
λ2
1 1
3 1− 4
λ2
0 4
λ2
λ 3λ
1
CM {0, 0, 0, 0} 2 0 0 1 0
√ √3
3
1 3 3
DM λ
2
, λ
2
, 0, 0 2 0 0 1− λ2 λ2
q
λ2 λ2 λ2
EDE √λ , 1− 6 , 0, 0 −1 + 2 −1 + 3 0 0 1
6
n o
FDE 0, 1, λ
3, 0 −1 −1 0 0 1
GDE {0, 1, 0, 0} −1 −1 0 0 1
HD {1, 0, 0, 0} 2 1 0 0 1
Table II: Critical points with the corresponding values of q, ωtot , Ωr , Ωm and Ω DE for III A( P1 ).
exists in the study of standard quintessence [61] and in the teleparallel DE models [23]. The critical point FDE
corresponds to an accelerating Universe with complete DE domination with ωtot = −1. In this case, the DE
behaves like a cosmological constant. This critical point is a novel critical point that is not present in standard
quintessence [61], and also, in terms of the coordinates, it varies from [23]. Similar to FDE , the critical point
GDE is also behaving like a cosmological constant. This solution describes a standard DE-dominated era with
Ω DE = 1.
• Critical Point Representing the Stiff DE HD : In the stiff matter era, the pressure p is equal to the energy
p
density ρ of the Universe, moreover, the EoS parameter ω = ρ = 1. In this case, the energy density ρ evolves
as ρ ∝ a(t)−6 . This is a much more rapid decrease than for radiation (ρ ∝ a−4 ) or matter (ρ ∝ a−3 ) [62]. The
critical point HD is corresponding to a non-accelerating, DE-dominated Universe with a stiff DE. In this case,
the EoS parameter ωtot = 1. This critical point exists in studying the standard quintessence model and the
teleparallel DE model [23, 61].
points represent early-time matter-dominated epochs of the evolution of the Universe as expected.
Numerical Results:
In this study, we have analysed critical points representing different epochs of the evolution of Universe. Among
these critical points, A R , BR are the critical points describing the radiation-dominated epoch of the evolution of the
Universe. Moreover, these critical points show saddle point behaviour. The critical points C M , D M represent the
matter-dominated epoch of the Universe, and both show saddle point behaviour. The critical points EDE , FDE and
GDE are the DE- solutions. From the stability analysis, one can confirm that these critical points are the late point
stable attractors within the particular range of the model parameters. Now, we will analyse the numerical solution
using the autonomous system presented in Eq. (26). These numerical solutions are obtained using the ND-solve
command in Mathematica. We used the Hubble and Supernovae Ia (SNe Ia) observational data set and described in
Appendix-V A.
The evolution of the energy densities of radiation, DE, and dark matter are shown in Figure 1b. From these plots,
it can be analysed that the radiation occurred in the early Universe, followed by a short period of dominance of
DM and, finally, the cosmological constant. The contribution of dark matter and DE sector density parameters at
present is obtained as Ωm ≈ 0.3 and Ω DE ≈ 0.7, respectively. From Figure 1b, the time of matter-radiation equality
is around z ≈ 3387 and is pointed out using an arrow. In Figure 1a, we have plotted the EoS parameters for DE,
total, and to compare with ΛCDM, the behaviour of the EoS parameter of ΛCDM has also been incorporated. The
plots show that the ωtot (cyan) evolves from the radiation value of 13 , it reaches 0 during the matter-dominated era
and, at last, it reaches −1. Similarly both ω DE (blue) and ωΛCDM (red) tends to −1 at late-time. At present time, the
value of ω DE (z = 0) = −1, which is compatible with the observational studies presented by Planck Collaboration
results in ω DE (z = 0) = −1.026 ± 0.034 [8]. In Figure 2a, we illustrate the Hubble rate evolution with the Hubble
rate HΛCDM (z) in the ΛCDM model and the Hubble data points [43], H0 = 70 Km/(Mpc sec) [8]. The model is
close to the standard ΛCDM model. In Figure 2b, the evolutionary behaviour of the deceleration parameter can be
studied, and it shows the transient behaviour at z ≈ 0.66, which is compatible with the current observational data
[63]. The present value of the deceleration parameter is q(z = 0) ≈ −0.53 and is in agreement with the cosmological
observational studies made in [64]. In Figure 3, we have presented the evolution of the modulus function µ(z) and
observe that the model curve along with the ΛCDM model modulus function µΛCDM well within the error bars.
10
2 1.0
ωDE
1
ωtot 0.8 ΩDE
ωΛCDM ΩDE
Ωm
0.6 Ωr
Ω
0
ω
-1 0.2
0.0
-2 -2 0 2 4 6
-2 0 2 4 6
Log10(1+z) Log10(1+z)
(a) Evolution of EoS parameters (b) Evolution of Standard density parameters
Figure 1: In this case, the initial conditions are xC = 10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λ =
−0.01, α = −5.2.
250 H(z)
0.5
H(z)[Km/s/Mpc]
0.0
150
q
Transition point
100
-0.5
50 Acceleration phase (q<0)
q
qΛCDM
0 -1.0
0.0 0.5 1.0 1.5 2.0 2.5 -1 0 1 2 3 4 5 6
z z
(a) Evolution of Hubble parameter
(b) Evolution of deceleration parameter in redshift
in redshift
Figure 2: In this case, the initial conditions are xC = 10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λ =
−0.01, α = −5.2.
46
44
42
40
μ(z)
38
36 μ(z)
34 μ (ΛCDM) (z)
32
0.0 0.5 1.0 1.5 2.0 2.5
z
Figure 3: Evolution of distance modulus function µ(z) ( dashed blue) and ΛCDM model distance modulus function
µΛCDM (z) (plane orange) along with 1048 Supernovae Ia (SNe Ia) data points. with the initial conditions xC =
10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λ = −0.01, α = −5.2.
11
In Figure 4, we have plotted the 2-D phase space portrait between the dynamical variables x and y. The critical
points A R and EDE connect with the heteroclinic orbit (red line). For this, the initial condition and the model para-
meter value are the same as Figure 1. From the phase space trajectories behaviour, one can see that the solution
approaches from a saddle critical point (A R ) to a stable critical point (EDE , FDE ). Through the behaviour of the tra-
jectories, we can say that the critical points A R , HD , D M , and BR show saddle behaviour and critical points EDE
and FDE indicate stable behaviour. In this plot, we have defined the region for the physical conditions 0 < Ωm ≤ 1
and y > 0. The green/shaded region represents the part of the accelerating expansion of the Universe in particular,
describe the quintessence behaviour (−1 < ωtot < − 13 ) and the stable critical points EDE and FDE lie in this region.
These critical points defined the late-time cosmic acceleration phase of the Universe.
1.2
-1<ωtot<-1/3
EDE
1.0 FDE
0.8
DM
0.6 BR
y
0.4
0.2
0.0
AR HD
Figure 4: The 2D phase space portrait. The initial condition and model parameter value are the same as Figure 1.
The green/shaded region shows the accelerating (quintessence) phase of the Universe, (−1 < ωtot < − 13 ).
B. Potential P2 Cosh(ξϕ) − 1
This potential plays an important role in studying the dynamics of the phantom and the quintessence DE, DM
models [6, 57]. The critical points are presented in Table III. In this case, from Table I, the value of α1 = − 12 and
2
α3 = ξ2 , hence we will analyse this potential with the autonomous dynamical system having five independent
variables.
12
C. P. { x, y, u, ρ, λ} q ωtot Ωr Ωm Ω DE
aR 1
{0, 0, 0, 1, ξ } 1 3 1 0 0
√ √
2
2 ξ 2 −4
bR 3
, √2 , 0, ,ξ 1 1
3 1− 4
ξ2
0 4
ξ2
ξ 3ξ ξ
cM 1
{0, 0, 0, 0, ξ } 2 0 0 1 0
√ √3
3
1 3 3
dM ξ
2
, ξ
2
, 0, 0, ξ 2 0 0 1− ξ2 ξ2
q
e DE √ξ , ξ2 ξ2 ξ2
1− 6 , 0, 0, ξ −1 + 2 −1 + 3 0 0 1
6
n o
f DE 0, 1, 3ξ , 0, ξ −1 −1 0 0 1
gDE {0, 1, 0, 0, 0} −1 −1 0 0 1
hD {1, 0, 0, 0, ξ } 2 1 0 0 1
Table III: Critical points with the corresponding values of q, ωtot , Ωr , Ωm and Ω DE for III B( P2 ).
In this potential function, one thing can be noticed in most of the critical points, the dynamical variable λ equals
the potential parameter ξ. The detailed analysis of the critical point is presented as follows,
points f DE , gDE are the cosmological constant solutions with the value of the ωtot = −1. These (e DE , f DE , gDE ,)
critical points represent the standard DE era of the Universe evolution with Ω DE = 1.
• Critical Point Representing the Stiff DE h D : This critical point describes the stiff matter-dominated era.
During the stiff matter-dominated era, the Universe would expand and cool rapidly. The energy density of
stiff matter decreases rapidly with the Universe’s expansion compared to other forms of matter or radiation.
The value of the Ω DE = 1 at this critical point. The value of the ωtot = 1 as expected.
critical point is stable at (λ < 0 ∧ ξ > 0) ∨ (λ > 0 ∧ ξ > 0) the detailed formalism is presented in the Ap-
pendix-V B.
p
The eigenvalues at critical point gDE are ν1 = −3, ν2 = −2, ν3 = 0, ν4 = 21 − 9 − 6ξ 2 − 3 , ν5 =
p
1 2−3
2 9 − 6ξ . According to the CMT presented in Appendix-V B, this critical point shows stable be-
haviour for α > 0, the detailed calculations for CMT is presented in the Appendix-V B.
Numerical Results:
14
The behaviour of the EoS parameter can be observed from Figure 5a for DE, total, and the ΛCDM show the
dominance of radiation at the early epoch. Gradually, this dominance decreased, and the rise of the dark matter-
domination era occurred. Continuing this sequence at the tail end, the accelerating expansion of the Universe era
has been observed at present, and the late time with values of the EoS parameter approaches −1. Figure 5a explains
the EoS parameter ωtot (magenta), which began at 13 for radiation, falls to 0 during the matter-dominated period,
and ultimately reaches −1. Both ωΛCDM (cyan) and ω DE (blue) approach −1 at late time. We observed that the value
of ω DE (z = 0) = −1 at present time. At present, Ωm ≈ 0.3 and Ω DE ≈ 0.7, the matter-radiation equality observed at
z ≈ 3387 and can be observed from Figure 5b. The Hubble rate evolution as a function of redshift z, the Hubble rate
HΛCDM (z) in the ΛCDM model, and the Hubble data points [43] are displayed in Figure 6a. It has been observed
that the model presented here is closely analogous to the standard ΛCDM model. The deceleration parameters
Figure 6b show the transition from deceleration to acceleration, which happens at z ≈ 0.65 and the present value of
the deceleration parameter noted ≈ −0.56. The ΛCDM model modulus function µΛCDM , 1048 pantheon data points
and modulus function µ(z) are shown in Figure 7.
2 1.0
ωDE
1
ωtot 0.8 ΩDE
ωΛCDM ΩDE
Ωm
Ω 0.6 Ωr
0
ω
-1 0.2
0.0
-2 -2 0 2 4 6
-2 0 2 4 6
Log10(1+z) Log10(1+z)
(a) Evolution of EoS parameters (b) Evolution of Standard density parameters
Figure 5: In this case, the initial conditions are xC = 10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λc =
10−1.3 , α = −5.2, ξ = −0.02.
250 H(z)
0.5
H(z)[Km/s/Mpc]
0.0
150
q
Transition point
100
-0.5
50 Acceleration phase (q<0)
q
qΛCDM
0 -1.0
0.0 0.5 1.0 1.5 2.0 2.5 -1 0 1 2 3 4 5 6
z z
(a) Evolution of Hubble parameter
(b) Evolution of deceleration parameter in redshift
in redshift
Figure 6: In this case, the initial conditions are xC = 10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λc =
10−1.3 , α = −5.2, ξ = −0.02.
15
46
44
42
40
μ(z)
38
36 μ(z)
34 μ (ΛCDM) (z)
32
0.0 0.5 1.0 1.5 2.0 2.5
z
Figure 7: Evolution of distance modulus function µ(z) ( dashed blue) and ΛCDM model distance modulus function
µΛCDM (z) (plane orange) along with 1048 Supernovae Ia (SNe Ia) data points. with the initial conditions xC =
10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λc = 10−1.3 , α = −5.2, ξ = −0.02.
In Figure 8, we have presented the 2D phase space portrait, plotted between dynamical variables x and y. The
behaviour of the phase space trajectories, the critical points h D , a R , d M and bR are the saddle critical points and the
critical points e DE and f DE display stable behaviour which lie inside the region of the accelerating expansion phase of
the Universe (quintessence) (−1 < ωtot < − 13 ), which is shaded by the cyan color in the upside of the phase-portrait
region in the image.
1.2
-1<ωtot<-1/3
eDE
1.0 fDE
0.8
dM
0.6 bR
y
0.4
0.2
0.0
aR hD
Figure 8: The 2D phase space portrait. The initial condition and model parameter value are the same as Figure 5.
The cyan/shaded region shows the accelerating (quintessence) phase of the Universe, (−1 < ωtot < − 31 ).
This form of potential plays an important role in analysing the generalise form teleparallel tachyonic model [31]
and to analyse the value of cosmological constant Λ using type Ia Supernovae [58]. In this study, We consider this
potential form to check its physical viability to discuss the evolutionary epochs of the Universe in this generalized
teleparallel scalar tensor formalism. The critical points corresponding to this potential are presented in the following
Table IV. In this case, the value of the α1 = η1 and α2 = −ηβ2 , as presented in Table I, hence the autonomous
16
C. P. { x, y, u, ρ, λ} q ωtot Ωr Ωm Ω DE
1
AR {0, 0, 0, 1, λ} 1 3 1 0 0
√ √
2
2 β2 η 2 −4
BR 3
, √ 2 , 0, , βη 1 1
3 1− 4
β2 η 2
0 4
β2 η 2
βη 3βη βη
1
CM {0, 0, 0, 0, λ} 2 0 0 1 0
√ √3
3
1 3 3
DM βη
2
, 2
βη , 0, 0, βη 2 0 0 1− β2 η 2 β2 η 2
q
βη β2 η 2 β2 η 2 β2 η 2
E DE √ , 1− 6 , 0, 0, βη −1 + 2 −1 + 3 0 0 1
6
n o
βη
F DE 0, 1, 3 , 0, βη −1 −1 0 0 1
G DE {0, 1, 0, 0, 0} −1 −1 0 0 1
HD 1, 0, 0, 0, βη 2 1 0 0 1
Table IV: Critical points with the corresponding values of q, ωtot , Ωr , Ωm and Ω DE for III C( P3 ).
This matter-dominated critical point D M appeared in the study made in the standard quintessence model and
the teleparallel DE model [23, 61].
17
points F DE and GDE are the critical points at which the dark energy behaves like a cosmological constant.
These are the de-Sitter solutions and represent the standard DE-dominated era of the evolution of the Universe
with Ω DE = 1.
• Stability of Critical Points C M , D M : The eigenvalues at the critical point C M are ν1 = − 23 , ν2 = 32 , ν3 =
1
− 2 , ν4 = 0, ν5 = 0 , as there is a presence of positive and the negative eigenvalues, this is a saddle critical
√
3α 3 24β4 η 4 −7β6 η 6
point. The eigenvalues at the critical point D M are ν1 = − βη , ν2 = − η6 , ν3 = − 12 , ν4 = − 4β3 η 3
−
√ 4 4 6 6
3 3 24β η −7β η
4 , ν5 = 4β3 η 3
− 34 . This critical point shows stability within the range of the model parameters as
α ∈ R ∧ α ̸= 0. Moreover, this critical point is showing saddle point behaviour at α < 0.
• Stability of Critical Points E DE , F DE , G DE The eigenvalues at the critical point E DE are ν1 = −αβη, ν2 =
:
2 β2 η 2 β2 η 2 2 2
−2β η, ν3 = 2 − 3, ν4 = 2 − 2, ν5 = β η − 3 and it shows Stability at α < 0 ∧ β < 0 ∧ 0 < η <
√ q1 √ q1 √ q1
3 β2 ∨ α > 0 ∧ β > 0 ∧ 0 < η < 3 β2 , saddle at α < 0 ∧ β > 0 ∧ 0 < η < 3 β2 ∨ α > 0 ∧ β <
√ q √ q √ q
0 ∧ 0 < η < 3 β12 and is unstable at α < 0 ∧ β < 0 ∧ η < − 6 β12 ∨ α > 0 ∧ β > 0 ∧ η < − 6 β12 .
q
3 ( β2 η 2 +6)( βη (8α+ βη )+6)
The eigenvalues at F DE are ν1 = 0, ν2 = −3, ν3 = −2, ν4 = − − 23 ,
2 ( β2 η 2 +6 )
q 2 2
3 ( β η +6)( βη (8α+ βη )+6)
ν5 = 2 β2 η 2 +6
− 1 . The stability at this critical point is analysed using CMT. This critical
point shows stable behaviour within the range η ∈ R ∧ ((λ < 0 ∧ β < 0) ∨ (λ > 0 ∧ β < 0)), and the
detailed formalism is presented in the Appendix-V B. At the critical point G DE have the eigenvalues ν1 =
p p
−3, ν2 = −2, ν3 = 0, ν4 = 12 − 12β2 η + 9 − 3 , ν5 = 12 12β2 η + 9 − 3 . Based upon the CMT, this
critical point shows stable behaviour for α > 0, where u̇ is negative., and the detailed formalism is presented
in the Appendix-V B.
18
Numerical Results:
In this case, the plots for the EoS parameter and the standard density parameters are presented in Figure 9. From
these plots, one can analyse the dominance of radiation at the early epoch, then the dominance of dark matter
for a while, and finally, at present and at the late time, the dominance of the DE era. At present, Ωm ≈ 0.3 and
Ω DE ≈ 0.7, the matter-radiation equality observed at z ≈ 3387 is highlighted in the Figure 9b using an arrow. Figure
9a describes the behaviour of the EoS parameter ωtot (cyan), which began at 13 for radiation, approaches to 0 during
the matter-dominated period, and ultimately tends to −1. Both ωΛCDM (black) and ω DE (blue) approach −1 at late
time. From these figures, one can analyse that the current value of ω DE (z = 0) = −1. The deceleration parameter
(q) and for the qΛCDM is plotted in Figure 10b is capable of describing the transition phase from deceleration to
acceleration, and the transition took place at z ≈ 0.65. The present value of the deceleration parameter is observed to
be ≈ −0.57. The Hubble rate evolution as a function of redshift z, the Hubble rate HΛCDM (z) in the ΛCDM model,
and the Hubble data points [43] are displayed in Figure 10a. It has been observed that the model presented here is
in close agreement with the standard ΛCDM model. The ΛCDM model modulus function µΛCDM , 1048 pantheon
data points and modulus function µ(z) are shown in Figure 11.
2 1.0
ωDE
1
ωtot 0.8 ΩDE
ωΛCDM ΩDE
Ωm
0.6 Ωr
Ω
0
ω
-1 0.2
0.0
-2 -2 0 2 4 6
-2 0 2 4 6
Log10(1+z) Log10(1+z)
(a) Evolution of EoS parameters (b) Evolution of Standard density parameters
Figure 9: In this case, the initial conditions are xC = 10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λc =
10−1.3 , α = −5.2, η = −0.2, β = −0.21.
19
250 H(z)
0.5
H(z)[Km/s/Mpc]
Deceleration phase (q>0)
H (z)
200 (ΛCDM)
0.0
150
q
Transition point
100
-0.5
50 Acceleration phase (q<0)
q
qΛCDM
0 -1.0
0.0 0.5 1.0 1.5 2.0 2.5 -1 0 1 2 3 4 5 6
z z
(a) Evolution of Hubble parameter
(b) Evolution of deceleration parameter in redshift
in redshift
Figure 10: In this case, the initial conditions are xC = 10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λc =
10−1.3 , α = −5.2, η = −0.2, β = −0.21.
46
44
42
40
μ(z)
38
36 μ(z)
34 μ (ΛCDM) (z)
32
0.0 0.5 1.0 1.5 2.0 2.5
z
Figure 11: Evolution of distance modulus function µ(z) ( dashed blue) and ΛCDM model distance modulus function
µΛCDM (z) (plane orange) along with 1048 Supernovae Ia (SNe Ia) data points. with the initial conditions xC =
10−8.89 , yC = 10−2.89 , uC = 10−5.96 , ρC = 10−0.75 , λc = 10−1.3 , α = −5.2, η = −0.2, β = −0.21.
We have presented the 2d-phase space diagram in Figure 12, we plot the phase space portrait between dynamical
variables x and y. From this plot, one can study the behaviour of trajectories, which implies that the critical points
E DE , F DE show stable behaviour, while the critical points H D , A R , B R , and D M show the saddle behaviour. The
critical points E DE and F DE lie inside the region of the accelerating expansion (quintessence) phase of the Universe
where (−1 < ωtot < − 31 ), which is shaded by the blue colored region in the figure.
20
1.2
-1<ωtot<-1/3
ℰDE
1.0 ℱDE
0.8
M
0.6 ℬR
y
0.4
0.2
0.0
R ℋD
Figure 12: The 2D phase space portrait. The initial condition and model parameter value are the same as Figure 9.
The blue/shaded region shows the accelerating (quintessence) phase of the Universe, (−1 < ωtot < − 31 ).
A key approach to understand background cosmology is analyzing dynamical systems. Using this technique,
we can explore the critical points corresponding to a cosmological viable model and their respective characteristics.
These predictions are validated by observable physics and cosmology as the Universe expands. Such tests could
be performed on a modified gravity model. Additional proof can be obtained to validate or invalidate specific
models or parameter ranges within the selected models by connecting critical point analysis to stability and phase
space images. In the present work, we have taken into consideration coupling function in the exponential form
g(ϕ) = g0 e−αϕκ to the teleparallel boundary term B and three different forms of the potential function, which is
presented in Table I. It is also required for the matter and radiation-dominated eras to occur before the DE period
for any DE scenario to succeed. The matter and radiation solutions are explained within the context of dynamical
systems as critical points of the autonomous system that are unstable for radiation, or saddle for matter. We have
discovered novel scaling solutions for the scaling radiation and matter epochs and the crucial points characterizing
the standard radiation and matter eras for the DE model under study in this investigation.
All three potential functions have stable critical points that describe the late-time cosmic accelerated phase. Scal-
ing solutions have also been obtained for critical points. According to the critical points, non-standard matter and
radiation-dominated eras in the Universe have been observed. From our results, we can see that the results match
the quintessence model presented in [23, 61]. The density parameters of the matter and DE sectors at the current
time z = 0 are found to be Ω0m ≈ 0.3 and Ω0de ≈ 0.7. Additionally, the matter-radiation equality value is found at
z ≈ 3387. For all three scenarios, the EoS parameter at present and late time have an approach to −1 with the present
time values aligned with the recent observation [8]. Based on the behaviour of Figure 2b, we can claim that for
the exponential form of V (ϕ), the deceleration parameter displays the transition from deceleration to acceleration
phase at z = 0.66. Its current value is q(z = 0) = −0.53. We obtained the transition point at z = 0.65 for the
V (ϕ) = Cosh(ξϕ) − 1, and the current value of the deceleration parameter is q(z = 0) = −0.56 Figure 6b. We find
the transition point at z = 0.61 for the V (ϕ) = V0 Sinh−η ( βϕ), and the current value of the deceleration parameter
is q(z = 0) = −0.57 can be observed from Figure 10b. In each of the three scenarios, the deceleration parameter
value and transition point matched cosmological findings [63, 64]. We compared our results in all three potential
functions with the Hubble 31 data points [43] and the Supernovae Ia data 1048 data points [65]. We may conclude
that the outcomes of our model closely resemble those of the conventional ΛCDM model based on the behaviour
21
of the Figure [2a, 6a, 10a]. The modulus function of our models was shown alongside the 1048 Supernovae Ia data
points in Figure [3, 7, 11], using the conventional ΛCDM model modulus function. The outcomes closely align with
the ΛCDM model. The quintessence region is shaded and is pointed using an arrow in the 2-d phase space diagrams
presented in Figures 4, 8, and 12. The phase space trajectories move from the early-time decelerating phase to the
stable late-time DE solutions.
This work shows promise; more work in the cosmic setting, utilizing either perturbation theory or observational
constraint analysis, should be done to explore further. This might shed further light on this idea and its connections
to the Universe’s large-scale structure and the cosmic microwave background radiation power spectrum.
V. APPENDIX
A. Datasets
In this study, we will analyse 31 data points [43] to investigate the behaviour of the Hubble rate in our model.
Additionally, we will compare our model with the standard ΛCDM model. The standard ΛCDM model is defined
as
q
HΛCDM = H0 (1 + z)3 Ωm + (1 + z)4 Ωr + Ωde , (27)
Supernovae Ia
Another component of our baseline data set is the Pantheon compilation of 1048 SNIa distance measurements
spanning 0.01 < z < 2.3 redshifts [65]. This dataset incorporates observations from prominent programs such as
PanSTARRS1, Hubble Space Telescope (HST) survey, SNLS, and SDSS. By amalgamating data from diverse sources,
the Pantheon collection offers valuable insights into the properties and behaviors of Type Ia supernovae and their
cosmic implications. Furthermore, it demonstrates the use of stellar luminosity as a means of determining distances
in an expanding Universe, with the distance moduli function being a key component of this analysis, and is repres-
ented as,
where M and DL define the nuisance parameter and the luminosity distance, respectively. The luminosity distance
can be written as
Z z
i dz
D ( z i , Θ ) = c (1 + z i ) , (29)
0 H (z, Θ)
Central manifold theory is a branch of dynamical systems theory that deals with the behavior of systems near
fixed points. The mathematical framework of Central Manifold Theory (CMT) was explained by Perko [66]. The
linear stability theory fails to explain the stability of critical points if their eigenvalues include zero eigenvalues. An
analysis of stability is possible in a CMT because the system’s dimensionality is reduced near that point. When the
system passes through the critical point, it behaves in an invariant local center manifold W c . The central manifold W c
is an invariant manifold associated with eigenvalues having zero real parts. The dynamics on this manifold capture
the essential features of the system’s behavior near the equilibrium.
22
Consider f ∈ Cr ( E), where E is an open subset of Rn containing the origin and r ≥ 1. Suppose f (0) = 0 and
D f (0) have c eigenvalues with zero real parts and s eigenvalues with negative real parts, where c + s = n. In general,
the system can be reduced to the following form
ẋ = Ax + F ( x, y)
ẏ = By + G ( x, y) (30)
Here the ordered pair ( x, y) ∈ Rc × Rs , A is a square matrix with c eigenvalues having zero real parts, B is a square
matrix with s eigenvalues with negative real parts, and F (0) = G (0) = 0, DF (0) = DG (0) = 0. Moreover, there
exists a δ > 0 and a function h( x ) ∈ Cr ( Nδ (0)) that defines the local center manifold which satisfies
for | x | < δ; and the center manifold can be defined by the system of differential equations
ẋ = Ax + F ( x, h( x )) (32)
The Jacobian matrix at the critical point GDE for the autonomous system (26 ) can be written as follows:
q
3
−3 0 −3 2 0
J ( GDE ) = 0 −3
0 0
0 0 0 0
0 0 0 −2
The eigenvalues of Jacobian matrix GDE are ν1 = −2, ν2 = 0, ν3 = −3, ν4 = −3. The eigenvectors corresponding
to these eigenvalues are
q T
[0, 1, 0, 0] T [1, 0, 0, 0] T , [0, 0, 0, 1] T , − 32 , 0, 1, 0
Now, applying the center manifold theory, we examine the stability of the critical point GDE because of its non-
hyperbolic nature. To apply CMT to this critical point, we have shifted it to the origin using a shifting transformation.
we have followed these transformations as X = x, Y = 1 + y, Z = u, and R = ρ then we can write equations in the
new coordinate system as
Ẋ −3 0 0 0 X
non
Ẏ 0 −3 0 0 Y + linear
= (33)
Ṙ 0 0 −2 0 R
term
Ż 0 0 0 0 Z
Comparing this diagonal matrix with the general form (30). Here, we can see that here, X, Y, and R are the stable
variables, and Z is the central variable. At this critical point, the matrix A and B matrix appears as
−3 0 0 h i
A=
0 − 3 0
B= 0
0 0 −2
23
According to CMT, the manifold can be defined by a continuous differential function. we have assumed the
following functions for the stable variables X = h1 ( Z ), Y = h2 ( Z ), and R = h3 ( Z ). With the help of the equation
(31), we have obtained the following zeroth approximation of the manifold functions as follow,
√
3 6Z 9Z2
2
, N (h2 ( Z )) = − 2
N (h1 ( Z )) = , N (h3 ( Z )) = 0 . (34)
3Z + 2 3Z + 2
In this case, the center manifold is acquired by the following expression,
18αZ2
Ż = − + higher order term (35)
3Z2 + 2
According to the CMT, this critical point shows stable behaviour for Z ̸= 0, α > 0, where Ż is negative.
matrix at the critical point f DE for the autonomous system (26 ) isas follows:
The Jacobian
√ √ √
2 6ξ
− 2ξ 12
2 2ξ 2
− 2ξ62 6 0 2 6
2ξ 2
+ 4 + 4 3 +4 3 +4
√ 3 q 3
6ξ 3 2ξ 2 6 3ξ ξ
− 2 ξ − ξ2 − ξ2 0 − ξ2
ξ2 ξ2
3 +2q 3 +2 3 +2 3 +2 +2
3
J ( f DE ) =
− 3 αξ2
0 0 0 0
2
0 0 0 − 2ξ2 − 4
ξ2
0
3 +2
ξ
3 3 +2
0 0 0 0 0
q
3 (ξ 2 +6)(8αξ +ξ 2 +6)+ξ 2 +6
The eigenvalues of Jacobian matrix f DE are ν1 = 0, ν2 = −3, ν3 = −2, ν4 = − , ν5 =
2 ( ξ 2 +6 )
q !
3 (ξ 2 +6)(8αξ +ξ 2 +6)
2 ξ 2 +6
− 1 . The corresponding eigenvectors are
" q #T
h iT 3 3
3(2α−ξ )
0, 0, 1
3 , 0, 1 , αξ
2
, 2αξ , 1, 0, 0 , [0, 0, 0, 1, 0] T ,
q q q T
3 − 2 +6)(8αξ + ξ 2 +6) + ξ 2 +6 2 +ξ 2 +6)(8αξ + ξ 2 +6) +24α − ξ 3 −6ξ
3 2 ( ξ 3ξ 4αξ ( ξ
, , 1, 0, 0 and,
q
2αξ (ξ 2 +6)
2α(ξ 2 +6) (ξ 2 +6)(8αξ +ξ 2 +6)+3ξ 2 +6
q q q T
3 + ξ 2 +6 4αξ 2 −ξ +24α−ξ 3 −6ξ
3 2 ( ξ 2 +6 )( 8αξ +ξ 2 +6 ) 3ξ ( ξ 2 +6 )( 8αξ +ξ 2 +6 )
,− , 1, 0, 0
q
2αξ (ξ 2 +6)
2α(ξ 2 +6) (ξ 2 +6)(8αξ +ξ 2 +6)−3ξ 2 −6
Now, applying the center manifold theory, we examine the stability of the critical point GDE because of its non-
hyperbolic nature. To apply CMT to this critical point, we have shifted it to the origin using a shifting transformation.
we have followed these transformations as: X = x, Y = 1 + y, Z = u + 3ξ , R = ρ and L = λ + ξ then we can write
equations in the new coordinate system as
q
3 (ξ 2 +6)(8αξ +ξ 2 +6)+ξ 2 +6
− 0 0 X 0 0
Ẋ 2 ( ξ 2 +6 ) !
q
Ẏ Y non
3 (ξ +6)(8αξ +ξ +6)
2 2
Ż = 0 2 ξ 2 +6
−1 0 0 0
Z + linear
(36)
Ṙ 0 0 −3 0 0 R term
L̇ 0 0 0 −2 0 L
0 0 0 0 0
24
Comparing this diagonal matrix with the general form (30). After that, we can say that here X, Y, Z, R and are the
stable variables, and L is the central variable. At this critical point, the A and B matrix appears as
q
3 (ξ 2 +6)(8αξ +ξ 2 +6)+ξ 2 +6
− 0 0 0
2 ( ξ 2 +6 ) !
q
(ξ 2 +6)(8αξ +ξ 2 +6)
h i
A= 3 B= 0
0 −1 0 0
2 ξ 2 +6
0 0 −3 0
0 0 0 −2
According to CMT, the manifold can be defined by a continuous differential function. we have assumed the
following functions for the stable variables X = h1 ( L), Y = h2 ( L), Z = h3 ( L) and R = h4 ( L). With the help of the
equation (31), we have obtained the following zeroth approximation of the manifold functions
√
3 6L 3Lξ
N (h1 ( L)) = − 2 , N (h2 ( L)) = , N (h3 ( L)) = 0 , N (h4 ( L)) = 0 (37)
ξ +6 ξ2+6
With these, the central manifold can be obtained as
18L2 ξ
L̇ = − + higher order term (38)
ξ2 + 6
According to the CMT, this critical point shows stable behavior for ξ > 0, where L̇ is negative.
The Jacobian matrix at the critical point gDEfor the autonomous system (26 ) is as follows:
q q
−3 0 −3 32 0 3
2
0 −3 0 0 0
J ( gDE ) =
0 0 0 0 0
q0
0 0 −2 0
3 2
− 2ξ 0 0 0 0
p p
The eigenvalues at critical point gDE are ν1 = −3, ν2 = −2, ν3 = 0, ν4 = 12 − 9 − 6ξ 2 − 3 , ν5 = 1
2 9 − 6ξ 2 − 3 .
The eigenvectors corresponding to these eigenvalues are [0, 1, 0, 0, 0] T ,
iT √ √ √ T √ √ √ T
− 2 3−2ξ 2 − 6 2 3−2ξ 2 − 6
h
[0, 0, 0, 1, 0] T 0, 0, 13 , 0, 1 − 2ξ 2
, 0, 0, 0, 1 − 2ξ 2
, 0, 0, 0, 1
Now, applying the center manifold theory, we examine the stability of the critical point GDE because of its non-
hyperbolic nature. To apply CMT to this critical point, we have shifted it to the origin using a shifting transformation.
we have followed these transformations as: X = x, Y = 1 + y, Z = u, R = ρ and L = λ then we can write equations
in the new coordinate system as
p
1
− 9 − 6ξ 2 − 3
Ẋ 2 0 0 X 0 0
p
Ẏ 1 Y non
0 2
9 − 6ξ − 3 0 0 0
2
Ṙ =
0 0 −3 0 0 R + linear (39)
L̇ L term
0 0 0 0−2 0
Ż Z
0 0 0 0 0
Comparing this diagonal matrix with the general form (30). After that, we can say that here X, Y, R, L and are the
stable variables, and Z is the central variable. At this critical point, the A and B matrix appears as
25
p
1
2 − 9 − 6ξ 2 − 3 0 0 0
p
1
0 9 − 6ξ 2 − 3 0 0
h i
A= 2 B= 0
0 0 −3 0
0 0 0 0−2
According to CMT, the manifold can be defined by a continuous differential function. we have assumed the
following functions for the stable variables X = h1 ( Z ), Y = h2 ( Z ), R = h3 ( Z ) and L = h4 ( Z ). With the help of the
equation (31), we have obtained the following zeroth approximation of the manifold functions
√
3 6Z 9Z2
N (h1 ( Z )) = , N (h2 ( Z )) = − , N (h3 ( Z )) = 0 , N (h4 ( Z )) = 0 (40)
3Z2 + 2 3Z2 + 2
18αZ2
Ż = − + higher order term (41)
3Z2 + 2
According to the CMT, this critical point shows stable behaviour for Z ̸= 0, α > 0, where Ż is negative.
The Jacobian matrix at the critical point F DE for the autonomous system (26 ) is as follows:
√ √ √
2 6βη
− 2β212
η 2 2β 2 η2 − 2β62 η2 6 0 2β
2 6
2 η2
√ 3 q +4 3 +4 3 +4 3 +4
6βη 3 2β2 η 2 6 3βη
− 2 βη − β2 η2 − β2 η2 0 − β2 ηβη
β2 η 2 β2 η 2 2
3 +2 3 + 2 3 + 2 3 + 2 3 +2
q
J (F DE ) =
− 23 αβη 0 0 0 0
2 η2
2β 4
0 0 0 − β2 η 2 − β2 η 2 0
3 3 +2 3 +2
0 0 0 0 0
q
( β2 η 2 +6)( βη (8α+ βη )+6) 3 3
The eigenvalues at F DE are ν1 = 0, ν2 = −3, ν3 = −2, ν4 = − − 2,
2 ( β2 η 2 +6 )
q 2 2
( β η +6)( βη (8α+ βη )+6)
ν5 = 32 β2 η 2 +6
− 1 . The eigenvectors corresponding to these eigenvalues are
" q #T
h iT 3 3
2 3(2α − βη )
0, 0, 3 , 0, 1 , αβη , 2αβη , 1, 0, 0 [0, 0, 0, 1, 0] T
1
q q q T
3 − 2 η 2 +6 2 η 2 + βη 3 η 3 −6βη
3 2 ( β 2 η 2 + 6 )( 8αβη + β 2 η 2 + 6 ) + β 3βη 4αβ ( β 2 η 2 + 6 )( 8αβη + β 2 η 2 + 6 ) + 24α − β
, , 1, 0, 0
q
2αβη ( β2 η 2 +6)
2α( β2 η 2 +6) ( β2 η 2 +6)(8αβη + β2 η 2 +6)+3β2 η 2 +6
q q q T
3 + β2 η 2 +6 4αβ2 η 2 − βη +24α− β3 η 3 −6βη
3 2 ( β2 η 2 +6 )( 8αβη + β2 η 2 +6 ) 3βη ( β2 η 2 +6 )( 8αβη + β2 η 2 +6 )
,− , 1, 0, 0 Now, apply-
q
2αβη ( β2 η 2 +6 )
2α( β2 η 2 +6) ( β2 η 2 +6)(8αβη + β2 η 2 +6)−3β2 η 2 −6
ing the center manifold theory, we examine the stability of the critical point F DE because of its non-hyperbolic
nature. To apply CMT to this critical point, we have shifted it to the origin using a shifting transformation. we have
βη
followed these transformations as: X = x, Y = 1 + y, Z = u + 3 , R = ρ and L = λ + βη then we can write equations
in the new coordinate system as
26
q
3 ( β2 η 2 +6)( βη (8α+ βη )+6) 3
− − 0 0 0 X 0
Ẋ 2 ( β2 η 2 +6 ) 2
q
Ẏ ( β2 η 2 +6)( βη (8α+ βη )+6) Y non
3
Ż = 0 2 β2 η 2 +6
−1 0 0 0
Z + linear (42)
Ṙ
0 0 −3 0 0 R
term
L̇
0 0 0 0−2 0
L
0 0 0 0 0
Comparing this diagonal matrix with the general form (30). After that, we can say that here X, Y, Z, R and are the
stable variables, and L is the central variable. At this critical point, the A and B matrix appears as
q
3 ( β2 η 2 +6)( βη (8α+ βη )+6) 3
− − 2 0 0 0
2 ( β2 η 2 +6 )
q
3 ( β2 η 2 +6)( βη (8α+ βη )+6) h i
A=
0 2 β2 η 2 +6
−1 0 0
B= 0
0 0 −3 0
0 0 0 0−2
According to CMT, the manifold can be defined by a continuous differential function. we have assumed the
following functions for the stable variables X = h1 ( L), Y = h2 ( L), Z = h3 ( L) and R = h4 ( L). With the help of the
equation (31), we have obtained the following zeroth approximation of the manifold functions
√
3 6L 3βηL
N (h1 ( L)) = − , N (h2 ( L)) = , N (h3 ( L)) = 0 , N (h4 ( L)) = 0 (43)
β2 η 2 + 6 β2 η 2 + 6
36βL2
L̇ = + higher order term (44)
β2 η 2 + 6
According to the CMT, this critical point shows stable behaviour for η ∈ R and β < 0 where L̇ is negative.
The Jacobian matrix at the critical point G DE for the autonomous system (26 ) is as follows:
q q
3 3
−3 0 −3 2 0 2
0 −3 0 0 0
J (G DE ) =
0 0 0 0 0
√ 0 0 0 − 2 0
6β2 η 0 0 0 0
p
At the critical point G DE we have the eigenvalues ν1 =, ν1 = −3, ν2 = −2, ν3 = 0, ν4 = 12 − 12β2 η + 9 − 3 , ν5 =
p
1
12β 2 η + 9 − 3 The eigenvectors corresponding to these eigenvalues are [0, 1, 0, 0, 0] T ,
2
iT √ √ 2 √ T √ √ √ T
6− 2 4β2 η +3
h
T 1 2 4β η +3+ 6
[0, 0, 0, 1, 0] 0, 0, 3 , 0, 1 − 4β2 η
, 0, 0, 0, 1 − 4β2 η
, 0, 0, 0, 1
Now, applying the center manifold theory, we examine the stability of the critical point G DE because of its non-
hyperbolic nature. To apply CMT to this critical point, we have shifted it to the origin using a shifting transformation.
we have followed these transformations as: X = x, Y = 1 + y, Z = u, R = ρ and L = λ then we can write equations
in the new coordinate system as
27
p
1
− 12β2 η + 9 − 3
Ẋ 2 0 0
0 X 0
p
Ẏ 1 Y non
0 2
12β η + 9 − 3 0 0 0
2
Ṙ =
0 0 −3 0 0 R + linear (45)
L̇ L term
0 0 0 0−2 0
Ż Z
0 0 0 0 0
Comparing this diagonal matrix with the general form (30). After that, we can say that here X, Y, R, L and are the
stable variables, and Z is the central variable. At this critical point, the A and B matrix appears as
p
1
2 − 12β2 η + 9 − 3 0 0 0
p
1
0 12β2 η + 9 − 3 0 0
h i
A= 2 B= 0
0 0 −3 0
0 0 0 0−2
According to CMT, the manifold can be defined by a continuous differential function. we have assumed the
following functions for the stable variables X = h1 ( Z ), Y = h2 ( Z ), R = h3 ( Z ) and L = h4 ( Z ). With the help of the
equation (31), we have obtained the following zeroth approximation of the manifold functions
√
3 6Z 9Z2
N (h1 ( Z )) = , N (h2 ( Z )) = − , N (h3 ( Z )) = 0 , N (h4 ( Z )) = 0 (46)
3Z2 + 2 3Z2 + 2
With these, the central manifold can be obtained as
18αZ2
Ż = − + higher order term (47)
3Z2 + 2
According to the CMT, this critical point shows stable behaviour for Z ̸= 0, α > 0, where Ż is negative.
ACKNOWLEDGEMENTS
SAK and LKD acknowledges the financial support provided by the University Grants Commission (UGC) through
Senior Research Fellowship (UGC Ref. No.: 191620205335), and (UGC Ref. No.: 191620180688) respectively. BM
acknowledges SERB-DST to provide financial support under the MATRICS grant (MTR/2023/000371).
REFERENCES
[1] C. Misner, K. Thorne, and J. Wheeler, Gravitation. No. pt. 3 in Gravitation. W. H. Freeman, 1973.
https://books.google.com.mt/books?id=w4Gigq3tY1kC.
[2] L. Baudis, “Dark matter detection,” J. Phys. G 43 (2016) no. 4, 044001.
[3] G. Bertone, D. Hooper, and J. Silk, “Particle dark matter: Evidence, candidates and constraints,” Phys. Rept. 405 (2005)
279–390, arXiv:hep-ph/0404175 [hep-ph].
[4] Supernova Search Team Collaboration, A. G. Riess et al., “Observational evidence from supernovae for an accelerating
universe and a cosmological constant,” Astron. J. 116 (1998) 1009–1038, arXiv:astro-ph/9805201 [astro-ph].
28
[5] Supernova Cosmology Project Collaboration, S. Perlmutter et al., “Measurements of Ω and Λ from 42 high redshift
supernovae,” Astrophys. J. 517 (1999) 565–586, arXiv:astro-ph/9812133 [astro-ph].
[6] V. Sahni and L. Wang, “New cosmological model of quintessence and dark matter,” Phys. Rev. D 62 (2000) no. 10, ,
arXiv:astro-ph/9910097 [astro-ph].
[7] J. Martin, “Everything you always wanted to know about the cosmological constant problem (but were afraid to ask),”
Comptes Rendus Physique 13 (2012) no. 6-7, 566–665, arXiv:1205.3365v1 [astro-ph].
[8] Planck Collaboration, N. Aghanim et al., “Planck 2018 results. VI. Cosmological parameters,” Astron. Astrophys. 641 (2020)
A6, arXiv:1807.06209 [astro-ph.CO]. [Erratum: Astron.Astrophys. 652, C4 (2021)].
[9] S. Capozziello and M. De Laurentis, “Extended Theories of Gravity,” Phys. Rept. 509 (2011) 167–321, arXiv:1108.6266
[gr-qc].
[10] S. Bahamonde et al., “Teleparallel gravity: from theory to cosmology,” Rep. Prog. Phys. 86 (2023) 207pp,
arXiv:2106.13793 [gr-qc].
[11] Y.-F. Cai, S. Capozziello, M. De Laurentis, and E. N. Saridakis, “ f ( T ) teleparallel gravity and cosmology,” Rept. Prog. Phys.
79 (2016) no. 10, 106901, arXiv:1511.07586 [gr-qc].
[12] T. Clifton, P. G. Ferreira, A. Padilla, and C. Skordis, “Modified Gravity and Cosmology,” Phys. Rept. 513 (2012) 1–189,
arXiv:1106.2476 [astro-ph.CO].
[13] S. Basilakos, S. Nesseris, F. K. Anagnostopoulos, and E. N. Saridakis, “Updated constraints on f ( T ) models using direct and
indirect measurements of the Hubble parameter,” JCAP 08 (2018) 008, arXiv:1803.09278 [astro-ph.CO].
[14] R. Ferraro and F. Fiorini, “Modified teleparallel gravity: Inflation without inflaton,” Phys. Rev. D 75 (2007) 084031,
arXiv:gr-qc/0610067 [gr-qc].
[15] G. R. Bengochea and R. Ferraro, “Dark torsion as the cosmic speed-up,” Phys. Rev. D 79 (2009) 124019, arXiv:0812.1205
[astro-ph].
[16] R. Briffa, C. Escamilla-Rivera, J. L. Said, and J. Mifsud, “Constraints on f ( T ) cosmology with Pantheon+,” MNRAS 522
(2023) no. 4, 6024–6034, arXiv:2303.13840 [gr-qc].
[17] A. Paliathanasis, J. D. Barrow, and P. Leach, “Cosmological solutions of f ( T ) gravity,” Phys. Rev. D 94 (2016) no. 2, ,
arXiv:1606.00659 [gr-qc].
[18] G. Farrugia, J. Levi Said, and M. L. Ruggiero, “Solar System tests in f ( T ) gravity,” Phys. Rev. D 93 (2016) no. 10, 104034,
arXiv:1605.07614 [gr-qc].
[19] N. Chernikov and E. Tagirov, “Quantum theory of scalar field in de sitter space-time,” in Annales de l’institut Henri Poincaré.
Section A, Physique Théorique, vol. 9, pp. 109–141. 1968.
[20] N. D. Birrell and P. C. W. Davies, “Conformal-symmetry breaking and cosmological particle creation in λφ4 theory,” Phys.
Rev. D 22 (1980) .
[21] G. Otalora, “Scaling attractors in interacting teleparallel dark energy,” JCAP 2013 (2013) , arXiv:1305.0474 [gr-qc].
[22] C.-Q. Geng, C.-C. Lee, and E. N. Saridakis, “Observational constraints on teleparallel dark energy,” JCAP 2012 (2012) ,
arXiv:1110.0913 [astro-ph].
[23] C. Xu, E. N. Saridakis, and G. Leon, “Phase-space analysis of teleparallel dark energy,” JCAP 2012 (2012) ,
arXiv:1202.3781 [gr-qc].
[24] H. Wei, “Dynamics of teleparallel dark energy,” Phys. Lett. B 712 (2012) , arXiv:1109.6107 [gr-qc].
[25] Y. Kucukakca, “Scalar tensor teleparallel dark gravity via noether symmetry,” Eur. Phys. J. C 73 (2013) , arXiv:1404.7315
[gr-qc].
[26] C.-Q. Geng, C.-C. Lee, E. N. Saridakis, and Y.-P. Wu, ““teleparallel” dark energy,” Phys. Lett. B 704 (2011) ,
arXiv:1109.1092 [hep-th].
[27] S. Bahamonde and M. Wright, “Teleparallel quintessence with a nonminimal coupling to a boundary term,” Phys. Rev. D 92
(2015) , arXiv:1508.06580 [gr-qc].
[28] M. Zubair, S. Bahamonde, and M. Jamil, “Generalized second law of thermodynamic in modified teleparallel theory,” Eur.
Phys. J. C 77 (2017) no. 7, , arXiv:1604.02996 [gr-qc].
[29] S. Bahamonde, U. Camci, S. Capozziello, and M. Jamil, “Scalar-tensor teleparallel wormholes by noether symmetries,” Phys.
Rev. D 94 (2016) , arXiv:1608.03918 [gr-qc].
[30] G. Gecim and Y. Kucukakca, “Scalar–tensor teleparallel gravity with boundary term by noether symmetries,” Inte. J. Geom.
Meth. Mod. Phys. 15 (2018) , arXiv:1708.07430v1 [gr-qc].
[31] S. Bahamonde, M. Marciu, and J. Levi Said, “Generalized tachyonic teleparallel cosmology,” Eur. Phys. J. C 79 (2019) ,
arXiv:1901.04973 [gr-qc].
[32] G. Otalora, “Cosmological dynamics of tachyonic teleparallel dark energy,” Phys. Rev. D 88 (2013) no. 6, ,
arXiv:1305.5896v2 [gr-qc].
[33] A. Banijamali and B. Fazlpour, “Tachyonic teleparallel dark energy,” Astrophys. Sp. Sci. 342 (2012) , arXiv:1206.3580
[gr-qc].
29
[34] S. A. Kadam, B. Mishra, and J. Said Levi, “Teleparallel scalar-tensor gravity through cosmological dynamical systems,” Eur.
Phys. J. C 82 (2022) no. 8, 680, arXiv:2205.04231 [gr-qc].
[35] S. A. Kadam, S. V. Lohakare, and B. Mishra, “Dynamical complexity in teleparallel gauss–bonnet gravity,” Annals of Physics
460 (2024) 169563, arXiv:2303.16911 [gr-qc].
[36] S. A. Kadam, N. P. Thakkar, and B. Mishra, “Dynamical system analysis in teleparallel gravity with boundary term,” Eur.
Phys. J. C 83 (2023) no. 9, , arXiv:2306.06677 [gr-qc].
[37] L. K. Duchaniya, S. V. Lohakare, B. Mishra, and S. K. Tripathy, “Dynamical stability analysis of accelerating f ( T ) gravity
models,” Eur. Phys. J. C 82 (2022) no. 5, 448, arXiv:2202.08150 [gr-qc].
[38] L. K. Duchaniya, S. V. Lohakare, and B. Mishra, “Cosmological models in f ( T, T ) gravity and the dynamical system
analysis,” Phys. Dark Univ. 43 (2024) 101402, arXiv:2302.07132 [gr-qc].
[39] K. Bamba, S. Capozziello, S. Nojiri, and S. D. Odintsov, “Dark energy cosmology: the equivalent description via different
theoretical models and cosmography tests,” Astrophys. Space Sci. 342 (2012) 155–228, arXiv:1205.3421 [gr-qc].
[40] A. Paliathanasis, “Dynamics in Interacting Scalar-Torsion Cosmology,” Universe 7 (2021) no. 7, 244, arXiv:2107.05880
[gr-qc].
[41] G. Leon, A. Paliathanasis, E. N. Saridakis, and S. Basilakos, “Unified dark sectors in scalar-torsion theories of gravity,” Phys.
Rev. D 106 (2022) , arXiv:2203.14866 [gr-qc].
[42] A. Paliathanasis and G. Leon, “Cosmological evolution in f ( T, B) gravity,” Eur. Phys. J. P. 136 (2021) 1–14,
arXiv:2106.01137 [gr-qc].
[43] M. Moresco, , et al., “Unveiling the universe with emerging cosmological probes,” Living Reviews in Relativity 25 (2022) 6,
arXiv:2201.07241 [astro-ph].
[44] M. Krssak et al., “Teleparallel theories of gravity: illuminating a fully invariant approach,” Class. Quant. Grav. 36 (2019)
no. 18, 183001, arXiv:1810.12932 [gr-qc].
[45] R. Aldrovandi and J. G. Pereira, Teleparallel Gravity: An Introduction. Springer, 2013.
[46] R. Weitzenböock, ‘Invariantentheorie’. Noordhoff, Gronningen, 1923.
[47] M. Krššák and E. N. Saridakis, “The covariant formulation of f ( T ) gravity,” Class. Quant. Grav. 33 (2016) no. 11, 115009,
arXiv:1510.08432 [gr-qc].
[48] M. Hohmann, L. Järv, and U. Ualikhanova, “Covariant formulation of scalar-torsion gravity,” Physical Review D 97 (2018)
no. 10, 104011, arXiv:1801.05786 [gr-qc].
[49] S. Bahamonde, C. G. Böhmer, and M. Wright, “Modified teleparallel theories of gravity,” Phys. Rev. D 92 (2015) no. 10,
104042, arXiv:1508.05120 [gr-qc].
[50] G. Farrugia and J. Levi Said, “Stability of the flat FLRW metric in f ( T ) gravity,” Phys. Rev. D 94 (2016) no. 12, 124054,
arXiv:1701.00134 [gr-qc].
[51] C. Wetterich, “Cosmology and the fate of dilatation symmetry,” Nucl. Phys. B 302 (1988) 668–696, arXiv:1711.03844
[hep-th].
[52] B. Ratra and P. Peebles, “Cosmological consequences of a rolling homogeneous scalar field,” Phys. Rev. D 37 (1988)
3406–3427.
[53] E. J. Copeland, M. Sami, and S. Tsujikawa, “Dynamics of dark energy,” Int. J. Mod. Phys. D 15 (2006) 1753–1936,
arXiv:hep-th/0603057 [hep-th].
[54] C.-Q. Geng, C.-C. Lee, E. N. Saridakis, and Y.-P. Wu, “Teleparallel dark energy,” Phys. Lett. B 704 (2011) 384–387,
arXiv:1109.1092v2 [hep-th].
[55] M. Gonzalez-Espinoza and G. Otalora, “Generating primordial fluctuations from modified teleparallel gravity with local
lorentz-symmetry breaking,” Phys. Lett. B 809 (2020) 135696, arXiv:2005.03753 [gr-qc].
[56] I. Zlatev, L. Wang, and P. J. Steinhardt, “Quintessence, cosmic coincidence, and the cosmological constant,” Phys. Rev. Lett.
82 (1999) no. 5, 896–899, arXiv:astro-ph/9807002 [astro-ph].
[57] N. Roy and N. Bhadra, “Dynamical systems analysis of phantom dark energy models,” JCAP 2018 (2018) 002.
[58] V. Sahani and A. Starobinsky, “The Case for a Positive Cosmological Λ-Term,” Inter. J. Mod. Phys. D 09 (2000) no. 04,
373–443, arXiv:astro-ph/9904398 [astro-ph].
[59] M. Gonzalez-Espinoza and G. Otalora, “Cosmological dynamics of dark energy in scalar-torsion f ( T, ϕ) gravity,” Eur. Phys.
J. C 81 (2021) no. 5, 480, arXiv:2011.08377 [gr-qc].
[60] L. K. Duchaniya, S. A. Kadam, J. L. Said, and B. Mishra, “Dynamical systems analysis in f ( T, ϕ) gravity,” Eur. Phys. J. C 83
(2023) no. 1, 27, arXiv:2209.03414 [gr-qc].
[61] E. J. Copeland, A. R. Liddle, and D. Wands, “Exponential potentials and cosmological scaling solutions,” Phys. Rev. D 57
(1998) no. 8, 4686, arXiv:gr-qc/9711068 [gr-qc].
[62] P. H. Chavanis, “Cosmology with a stiff matter era,” Phys. Rev. D 92 (2015) no. 10, 103004, arXiv:1412.0743 [gr-qc].
[63] S. Capozziello, O. Farooq, O. Luongo, and B. Ratra, “Cosmographic bounds on the cosmological deceleration-acceleration
transition redshift in f (R) gravity,” Phys. Rev. D 90 (2014) 044016, arXiv:1403.1421 [gr-qc].
30
[64] D. Camarena and V. Marra, “Local determination of the hubble constant and the deceleration parameter,” Phys. Rev. Res. 2
(2020) 013028, arXiv:1906.11814 [astro-ph.CO].
[65] D. M. Scolnic et al., “The Complete Light-curve Sample of Spectroscopically Confirmed SNe Ia from Pan-STARRS1 and
Cosmological Constraints from the Combined Pantheon Sample,” The Astrophysical Journal 859 (2018) no. 2, 101,
arXiv:1710.00845 [astro-ph].
[66] L. Perko, Differential equations and Dynamical systems. Springer-Verlag, New York, 2001.