Next Article in Journal
Probing Modified Gravity Theories with Scalar Fields Using Black-Hole Images
Previous Article in Journal
Stability Analysis of the Inhomogeneous Perturbed Einstein Universe in Energy–Momentum Squared Gravity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The General Property of Tracking and Thawing Models and Their Observational Constraints

School of Physical Science and Technology, Southwest University, Chongqing 400715, China
*
Author to whom correspondence should be addressed.
Universe 2023, 9(3), 146; https://doi.org/10.3390/universe9030146
Submission received: 19 February 2023 / Revised: 27 February 2023 / Accepted: 9 March 2023 / Published: 11 March 2023
(This article belongs to the Section Cosmology)

Abstract

:
We study the general property of the evolution of a class of scalar fields with tracking and thawing behaviors. For the tracking solutions, we show explicitly with three different potentials that, independent of initial conditions, there exists a general relation between the equation of state w ϕ and the fractional energy density Ω ϕ , so that the scalar field follows the same w ϕ Ω ϕ trajectory during the evolution. The analytical approximations of the w ϕ Ω ϕ trajectories are derived even though the analytical expression depends upon the particular form of the potential. For thawing solutions, a universal w ϕ Ω ϕ relation exists and the relation is independent of both the particular form of the potential and the initial condition of the scalar field. Based on the derived w ϕ Ω ϕ relation for the thawing models, we derive a tighter upper limit on w ϕ = d w ϕ / d ln a . The observational data is also used to constrain the thawing potential with the help of the universal w ϕ Ω ϕ relation.

1. Introduction

The observations of type Ia supernova provided the evidence to support the late time cosmic acceleration [1,2,3]. To explain the late time cosmic acceleration, an exotic matter component with negative pressure dubbed as dark energy is usually introduced. The observations suggested that dark energy contributes about 72% to the total energy density in the universe. The simplest candidate for dark energy is the cosmological constant, but the theoretical estimation of the value of the cosmological constant is in discrepancy with the astronomical observations by about 120 orders of magnitude [4], and the cosmological constant as dark energy candidate faces fine tuning and coincidence problems. Based on the concordance Λ CDM model, the release of Planck first year data on the measurements of the cosmic microwave background anisotropy gave the value of the Hubble constant as H 0 = 67.3 ± 1.2 km/s/Mpc [5,6]. This value is in tension with the local measurement on the Hubble constant H 0 = 73.8 ± 2.4 km/s/Mpc by calibrating the magnitude–redshift relation of 253 type Ia supernovae with over 600 Cepheid variables [7]. With more data and more accurate measurements, the measurement accuracy of the Hubble constant was improved greatly [7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27] and the tension became more serious [28]. For example, Planck 2018 data gave the result H 0 = 67.4 ± 0.5 km/s/Mpc [25], the supernovae and H 0 for the equation of state (SH0ES) team measured the local Hubble constant as H 0 = 73.15 ± 0.97 km/s/Mpc from Type Ia supernovae (SNe Ia) data by recalibrating the extragalactic distance ladder using a sample of Milky Way Cepheids with the Hubble Space Telescope photometry and Gaia EDR3 parallaxes [13]. By using the tip of the red giant branch method, the local measurements of the Hubble constant gave H 0 = 69.8 ± 0.6 (stat) ± 1.6 (sys)km/s/Mpc [14]. Since the Planck results from the early Universe probe of cosmic microwave background anisotropy depend on the Λ CDM model and the local measurements from type Ia supernova standard candles suffer the zero-point calibration problem, we are still uncertain about the origin of the tension on the Hubble constant. Gravitational waves as standard sirens provide an independent method of measuring cosmological distance and the observations of gravitational waves and their electromagnetic counterparts may shed light on the origin of the tension on the Hubble constant [29,30,31]. The first gravitational wave event GW170817 from binary neutron star merger and its counterpart GRB 170817A gave H 0 = 70 . 0 8.0 + 12.0 km/s/Mpc [32]. Combining the galaxy catalog and 47 gravitational wave events from the third LIGO-Virgo-KAGRA Gravitational-Wave Transient Catalog (GWTC-3), LIGO-Virgo-KAGRA collaborations obtained H 0 = 68 6 + 8 km/s/Mpc [33].
Although more astronomical observations support the late time cosmic acceleration, the question whether the cosmological constant is dark energy remains unanswered. For these reasons, dark energy is usually tackled by a dynamical scalar field called quintessence [34,35,36,37]. As discussed in [38,39,40,41], the value of the Hubble constant H 0 can be lowered by quintessence models relative to the cosmological constant. Other possibilities for dark energy such as phantom [42], quintom [43,44,45], tachyon [46,47,48], and k-essence [49] were also considered. Alternatively, the late time cosmic acceleration approach can also be explained by modifying general relativity at the cosmological scale, such as the Dvali–Gabadadze–Porrati model [50], f(R) gravity [51,52,53,54], and dRGT ghost-free massive gravity [55,56]. For a review of dark energy and alternative theories of gravity, please see Refs. [57,58,59,60,61,62,63,64,65].
In general, to discuss the evolution of a scalar field, we need to impose initial conditions, so it is difficult to get the property of the quintessence field from observational data. Therefore, it is worthwhile to seek the general behavior of dynamical dark energy which is independent of initial conditions. Fortunately, there exist general w ϕ Ω ϕ trajectories for the thawing and tracking solutions [66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88]. For a wide class of quintessence potentials with the tracking solution [37,68], the equation of state parameter w ( z ) for the quintessence varies slowly. Motivated by this behavior, Efstathiou approximated w ( z ) as w ( z ) = w 0 α ln ( 1 + z ) in the redshift range z < 4 [89]. For a wide class of scalar fields, Chevallier–Polarski–Linder (CPL) parametrization with w ( a ) = w 0 + w a ( 1 a ) [90,91] was usually used as a model-independent parametrization. Combining the generic w ϕ Ω ϕ relations and the CPL parametrization, the degeneracy between the two parameters w 0 and w a in the CPL parametrization can be broken by an analytical relation [87,92,93]. In this paper, we study the general property of the evolution for the tracker and thawing fields.
The paper is organized as follows. In Section 2, we discuss the reason for the same w ϕ Ω ϕ trajectory in the tracker solution. In Section 3, we derive the w ϕ Ω ϕ trajectories for different potentials with tracking solutions. We derived the thawing solution and obtained the constraint by using observational data for different potentials in Section 4. The conclusion is drawn in Section 5.

2. The Tracking Solution

Using the fractional energy density Ω ϕ and the equation of state parameter γ = 1 + w ϕ , the evolution of a scalar field is determined by the following equations:
Ω ϕ = 3 ( γ b γ ϕ ) Ω ϕ ( 1 Ω ϕ ) ,
γ ϕ = ( 2 γ ϕ ) ( 3 γ ϕ + | λ | 3 γ ϕ Ω ϕ ) ,
λ = 3 γ ϕ Ω ϕ λ | λ | ( Γ 1 ) ,
where “′” represents the derivative with respect to ln a , w b = [ ( 1 / 3 ) a e q / a ] / [ 1 + a e q / a ] is the equation of state for the background, λ = V 1 d V / d ϕ is the roll parameter, and Γ = λ 2 V 1 d 2 V / d ϕ 2 is the tracker parameter. With these dynamical Equations (1)–(3), we first discuss the behavior of tracking solution. To get the tracking solution, the potential V ( ϕ ) must satisfy the condition Γ 1 [68] so that the absolute value of the roll parameter | λ | either decreases with time or remains a constant. When the tracking solution is reached, γ ϕ = 0 , the tracker condition [68]:
γ ϕ = 1 + w ϕ = 1 3 λ 2 Ω ϕ
is satisfied, the flow parameter F = γ ϕ / ( Ω ϕ λ 2 ) = 1 / 3 and β = 1 + ( 1 w ϕ ) / 12 F = γ ϕ / 2 . If γ ϕ 0 , then γ ϕ is nearly a constant [68,74]:
w ϕ = w b ( 1 Ω ϕ ) 2 ( Γ 1 ) 2 Γ 1 Ω ϕ .
It is clear that w ϕ is not a constant in general. To get a nearly constant w ϕ , we must require that Ω ϕ is negligibly small and the tracker parameter Γ 1 is nearly a constant for the tracking solution. The exponential potential has Γ = 1 , and the inverse power-law potential V ( ϕ ) = V 0 ϕ α has Γ = 1 + 1 / α , so they have the tracking behavior. If Ω ϕ 0 , then:
w ϕ w ϕ t r k = w b 2 ( Γ 1 ) 2 Γ 1 ,
and the roll parameter λ must be large so that the tracker condition can be satisfied. If Γ 1 is not a constant, then it must be either large or small, so that w ϕ 1 or w ϕ w b . For example, the potential V ( ϕ ) = V 0 exp [ ( 3 ϕ ) 2 / 3 / 2 ] has large Γ 1 = λ 2 and the potential V ( ϕ ) = V 0 exp ( 1 / ϕ ) has small Γ 1 = 2 / λ . In summary, the tracking solution satisfies the two conditions (4) and (6) initially, and the necessary conditions for the tracker behavior to occur are: when the tracker condition (4) is satisfied, the fractional energy density Ω ϕ should be negligible which requires that the initial value of | λ | should be big enough and the roll parameter | λ | either decreases with time or remains a constant.
Now let us discuss the dependence of the tracking solution on initial conditions. We consider two different initial conditions: (i) If the initial value of the fractional energy density Ω ϕ i is not too small or the initial value of the roll parameter λ i is large enough so that λ i 2 > 3 γ ϕ i / Ω ϕ i , then from Equation (2), we see that whatever the initial value of w ϕ , γ ϕ will always increase toward 2 until γ ϕ = 0 . If γ ϕ > γ b , then Equation (1) tells us that Ω ϕ will start to decrease because Ω ϕ < 0 . Once Ω ϕ decreases to a small value so that λ 2 < 3 γ ϕ / Ω ϕ , Equation (2) tells us that γ ϕ < 0 , then γ ϕ decreases toward to zero and Ω ϕ starts to increase if γ ϕ < γ b . Although λ decreases, it can be larger than 3 γ ϕ / Ω ϕ if γ ϕ 0 because Ω ϕ increases, so γ ϕ will increase again. When the equation of state parameter γ ϕ increases away from zero, it will satisfy the relation λ 2 < 3 γ ϕ / Ω ϕ and w ϕ will reach the tracker behavior (6), perhaps with several oscillations. However, if the roll parameter decreases too quickly, which can be seen in the potential V ( ϕ ) = V 0 exp [ ( 3 ϕ ) 2 / 3 / 2 ] with Γ = 1 + λ 2 and λ λ 4 , then there is no tracker solution with a nearly constant w ϕ because λ is too small when the tracker condition is satisfied. The reason is as follows: If λ i is small, then Ω ϕ i cannot be too small and Ω ϕ may reach 1 very quickly, so there is no tracker solution with nearly constant w ϕ . (ii) If the initial value Ω ϕ i or λ i is small, and the condition λ i 2 < 3 γ ϕ i / Ω ϕ i is satisfied initially, then independent of the initial value γ ϕ i , Equations (1) and (2) tell us that the equation of state parameter γ ϕ decreases toward 0 and the fractional energy density Ω ϕ starts to increase if γ ϕ < γ b , then the analysis is similar to that in the first case (i). If λ i is small, then Ω ϕ always increases although Ω ϕ i is small, and it will reach 1 soon, so there is no tracker solution with nearly constant w ϕ during the matter domination. However, the current values of Ω ϕ , λ , and w ϕ are determined by their initial values. These analyses tell us that there is no tracker solution with nearly constant w ϕ if the initial value of λ is small because Ω ϕ reaches 1 very quickly and the scalar field becomes dominant. When Ω ϕ reaches 1, then the evolution of Ω ϕ stops and the dynamical Equations (1)–(3) reduce to the equations for γ ϕ and λ . From Equations (2) and (3), we get a relation between w ϕ and λ . Since Ω ϕ 1 and w ϕ approaches 1 asymptotically, if Γ is a function of λ , then the relation between w ϕ and λ becomes universal which is independent of the initial conditions, so the relation between w ϕ and w ϕ is also universal when w ϕ 1 . For example, there exists the attractor solution Ω ϕ = 1 , w ϕ = 1 and λ = 0 for the potential V ( ϕ ) = V 0 exp [ ( 3 ϕ ) 2 / 3 / 2 ] .
Since the tracking conditions (4) and (6) are the initial conditions for the tracking solution, if the same initial value of Ω ϕ is given, then w ϕ , Ω ϕ , and λ will follow the same trajectory, and independent of the initial conditions we will have the same w ϕ Ω ϕ trajectory for the tracking solution. Although w ϕ Ω ϕ trajectory is the same, the exact values of w ϕ and Ω ϕ at a moment are still determined by the initial conditions.

3. The Analysis of Different Potentials

For the inverse power-law potential V ( ϕ ) = V 0 ϕ α , the tracker parameter Γ = 1 + α 1 , and the critical point ( Ω ϕ c , γ ϕ c , λ c ) = ( 1 , 0 , 0 ) is a stable point. As Ω ϕ 1 , both γ ϕ and λ decrease to zero asymptotically, γ ϕ λ 2 / 3 + 2 λ 4 / ( 9 α ) , λ 2 Ω ϕ 3 γ ϕ ( 1 γ ϕ / α ) and γ ϕ 3 ( 2 γ ϕ ) γ ϕ 2 / α . When the scalar field starts to catch up the background, the w ϕ Ω ϕ trajectory can be approximated with a linear relation [75,94]:
Ω ϕ = 1 2 w i + 4 w i 2 ( 1 w i 2 ) w i ( w ϕ w i ) ,
where w i = w ϕ t r k . Therefore:
λ 2 Ω ϕ 3 γ ϕ γ b γ i γ i γ ϕ 3 γ ϕ 2 γ ϕ 1 γ ϕ γ i .
Substitute the above relation (8) into Equations (1) and (2), we derive the approximate relation:
Ω ϕ = 1 ( 1 w i 2 ) w i 1 2 w i + 4 w i 2 1 γ i γ ϕ γ ϕ γ i 1 / 2 + γ i γ b / 4 ( γ b γ i ) exp γ i γ b γ i γ b 1 γ ϕ 1 γ i 1 .
The Ω ϕ w ϕ trajectories for the inverse power-law potential along with the analytical approximation (9) are shown in Figure 1. It is clear that the analytical result (9) approximates the Ω ϕ w ϕ trajectory well, especially for smaller α as shown in Figure 1, and the inverse power-law potential V ( ϕ ) = V 0 ϕ 2 is excluded by observational data.
For the potential V ( ϕ ) = V 0 exp ( 1 / ϕ ) , Γ = 1 + 2 / λ . The tracking solution starts with γ ϕ t r k = γ b / ( 1 + 4 / λ ) γ b because λ is large due to the tracker condition. As γ ϕ 0 and Ω ϕ 1 , we have λ Ω ϕ 3 γ ϕ [ 1 2 ( 3 γ ϕ ) 3 / 4 / 3 ] and γ ϕ 4 ( 3 γ ϕ ) 7 / 4 / 3 . The w ϕ Ω ϕ relation for the tracking solution can be approximated with:
Ω ϕ = 1 + 3 γ ϕ 1 / 4 ( 1 γ ϕ ) 15 / 2 exp [ ( 3 γ ϕ ) 1 / 4 ( 3 γ ϕ ) 3 / 4 ] 1 .
The Ω ϕ w ϕ trajectory and the analytical approximation (10) are shown in Figure 1. As seen from Figure 1, the analytical relation (10) approximates the trajectory well and the potential V ( ϕ ) = V 0 exp ( 1 / ϕ ) is consistent with observational data.
For the potential V ( ϕ ) = V 0 sinh β ( α ϕ ) , the tracker parameter Γ = 1 + 1 / β α 2 β / λ 2 . By setting Ω ϕ = γ ϕ = λ = 0 , we obtain the following critical points. Point C1 with ( Ω ϕ C 1 , γ ϕ C 1 , λ C 1 ) = ( 1 , 2 , α β ) : it exists for all parameters, the point is a saddle point when λ > 6 and unstable when λ < 6 . Point C2 with ( Ω ϕ C 2 , γ ϕ C 2 , λ C 2 ) = ( 0 , 2 , λ ) : it exists for all parameters and the point is an unstable point. Point C3 with ( Ω ϕ C 3 , γ ϕ C 3 , λ C 3 ) = ( 0 , 0 , λ ) : it exists for all parameters and the point is a saddle point for 1 < w < 1 . Point C4 with ( Ω ϕ C 4 , γ ϕ C 4 , λ C 4 ) = ( 1 , 0 , λ ) : it exists for all parameters and the point is a stable point only when λ = 0 . Point C5 with ( Ω ϕ C 5 , γ ϕ C 5 , λ C 5 ) = ( 3 γ b / λ 2 , γ b , α β ) : the existence condition is λ 2 > 3 ( 1 + w ϕ ) , the point is a stable point for 3 ( 1 + w ϕ ) < λ 2 < 24 ( 1 + w ϕ ) 2 / ( 7 + 9 w ϕ ) and saddle point when λ 2 > 24 ( 1 + w ϕ ) 2 / ( 7 + 9 w ϕ ) . Point C6 with ( Ω ϕ C 6 , γ ϕ C 6 , λ C 6 ) = ( 1 , α 2 β 2 / 3 , α β ) : the existence condition is λ 2 < 6 , the point is a stable point for λ 2 < 3 ( 1 + w ϕ ) and saddle point when 3 ( 1 + w ϕ ) < λ 2 < 6 . If we choose α = 0.1 and β = 2 , the w ϕ Ω ϕ relation for the tracking solution can be approximated with:
Ω ϕ = 1 28 × 3 3 / 2 γ ϕ 17 / 4 .
The Ω ϕ w ϕ trajectory and the analytical approximation (11) are shown in Figure 1. As seen from Figure 1, the analytical relation (11) approximates the trajectory well and the model V ( ϕ ) = V 0 sinh 2 ( 0.1 ϕ ) is excluded by observational data.
Although we derived analytical approximations for the Ω ϕ w ϕ trajectories, it seems to be different for different potentials, and the tracking behavior depends on the particular form of the potential, so it is difficult to reconstruct the general property of the tracker potential from observational data. However, the situation is different for the thawing solution.

4. The Thawing Solution

For the thawing solution, w ϕ starts with the initial value 1 and starts to increase. From Equation (3), it is easy to see that λ will keep to be a constant when w ϕ = 1 . Taking the approximation γ ϕ 1 and λ λ i , we get [88]:
d γ ϕ d Ω ϕ = 6 γ ϕ + 2 λ 3 γ ϕ Ω ϕ 3 γ b Ω ϕ ( 1 Ω ϕ ) , γ ϕ = λ i 2 3 1 + 1 2 γ b 2 Ω ϕ ( 1 Ω ϕ ) 2 / γ b 2 F 1 2 1 γ b + 1 2 , 1 γ b + 1 , 1 γ b + 3 2 ; Ω ϕ ,
where 2 F 1 ( a , b , c , x ) is the hypergeometric function. Note that this approximation is invalid when γ ϕ 1 . If we take γ b = 1 , then we recover the familiar w ϕ Ω ϕ relation [81,92]:
γ ϕ = λ i 2 3 1 Ω ϕ 1 Ω ϕ 1 ( tanh 1 Ω ϕ ) 2 = ( 1 + w 0 ) 1 Ω ϕ 1 Ω ϕ 1 tanh 1 ( Ω ϕ ) 2 × 1 Ω ϕ 0 ( Ω ϕ 0 1 1 ) tanh 1 Ω ϕ 0 2 .
Note that the last equality also holds for thawing phantom models [78,92]. For thawing phantom models with the potentials V ( ϕ ) ϕ 6 , V ( ϕ ) ϕ 2 , V ( ϕ ) ϕ 2 , and V ( ϕ ) exp ( λ ϕ ) , it was explicitly shown that the analytical result (13) gives the behavior of w ( a ) [78]. Therefore, we can use w ( a ) given by Equation (13) to approximate thawing scalar fields including quintessence and phantom. If γ ϕ is near zero, the scalar field is not much different from the cosmological constant, then we can approximate the dynamics of the thawing field by the SSLCPL model [87,92] with w ( a ) = w 0 + w a ( 1 a ) and:
w a = 6 ( 1 + w 0 ) ( Ω ϕ 0 1 1 ) [ Ω ϕ 0 tanh 1 ( Ω ϕ 0 ) ] Ω ϕ 0 1 / 2 ( Ω ϕ 0 1 1 ) tanh 1 ( Ω ϕ 0 ) .
Although we discuss quintessence field only, the SSLCPL model can also approximate the thawing phantom fields.
As Ω ϕ 0 and w ϕ 1 , γ ϕ λ 0 2 Ω ϕ / 3 ( 1 + γ b / 2 ) 2 , we get the flow parameter F = 1 / 3 ( 1 + γ b / 2 ) 2 and β = γ b / 2 , the results are the same as those found in [84]. If w ϕ starts to increase during the matter domination, γ b = 1 and F = 4 / 27 . Therefore, 4 / 27 F 1 / 3 for the thawing solution, a tighter upper limit w ϕ 3 ( 1 w ϕ 2 ) / 2 is then obtained. As discussed in [88], the approximation (13) works better for smaller λ i and the w ϕ Ω ϕ relation (13) approximates the trajectory for the power-law potential better than that for the pseudo-Nambu Goldstone boson potential.
With the approximation (14), we use the 740 spectroscopically confirmed type Ia supernovae data from three years of the SDSS-II supernovae survey and SNLS3 supernovae data [95], the combination of Planck 2013 results on the measurements of the cosmic microwave background power spectra with the nine-year WMAP polarization low-multipole likelihood data [5,6,96], the 8 baryon acoustic oscillation data from the 6 dFGS survey, the baryon oscillation spectroscopic survey, the SDSS survey and the WiggleZ dark energy survey [97,98,99,100,101], and the 21 Hubble parameter H ( z ) data [102,103,104,105] to constrain w ϕ 0 and Ω ϕ 0 , and the results are shown in Figure 2. As discussed above, the approximation (14) holds for both quintessence and phantom thawing fields, so the 2 σ constraint 1.23 < w ϕ 0 < 0.85 applies for both quintessence and phantom thawing fields. From Figure 2, we also see that the 2 σ bound on the roll parameter is | λ i | = 1.013 for the quintessence thawing field and | λ i | = 1.199 for the phantom thawing field. From the marginalized 95% contours of w ϕ 0 and Ω ϕ 0 obtained in Figure 2, we reconstruct the thawing potential V ( ϕ ) by using the following relations:
d ϕ d ln a 2 = 3 m p l 2 Ω ϕ ( a ) | 1 + w ( a ) | ,
V ( a ) = 1 2 ρ c r 0 ( 1 w ( a ) ) Ω ϕ ( a ) ( H ( a ) / H 0 ) 2 ,
where m p l = ( 8 π G ) 1 / 2 and the current critical density ρ c r 0 = 3 m p l 2 H 0 2 . The reconstructed potential is shown in Figure 3. From Figure 3, we see that the power-law potential V ( ϕ ) = V 0 ϕ 0.8 is consistent with the observational data. The power-law potential V ( ϕ ) = V 0 ϕ 0.8 as a inflationary model is also consistent with the measurements on the cosmic microwave background anisotropy. We also find that the potential V ( ϕ ) = V 0 ϕ 0.7 is consistent with the 2 σ upper bound and the potential V ( ϕ ) = V 0 ϕ 1.1 is consistent with the 2 σ lower bound.

5. Conclusions

To guarantee an almost constant w ϕ when the tracking solution begins, the fractional energy density Ω ϕ should be small and the roll parameter λ must be large enough so that the tracker condition can be satisfied, and | λ | does not increase with time. Therefore, a nearly constant tracker parameter with Γ > 1 and large initial value of the roll parameter λ are the necessary conditions for tracking solutions. Although the current value of Ω ϕ and w ϕ depend on the initial conditions for the tracking solutions, the w ϕ Ω ϕ trajectory is independent of the initial conditions. We derived the analytical approximations of the w ϕ Ω ϕ trajectory for the potentials V ( ϕ ) ϕ α and V ( ϕ ) exp ( 1 / ϕ ) . Unfortunately, the analytical approximations are different for different potentials.
For the thawing solutions, w ϕ 1 initially, the roll parameter changes very slowly and it can be approximated as a constant, a general w ϕ Ω ϕ relation (12) is then obtained. Based on the asymptotical behavior of the w ϕ Ω ϕ relation, we derive the tighter upper bound w ϕ 3 ( 1 w ϕ 2 ) / 2 . In general, Ω ϕ 0 and w ϕ 0 depend on the initial conditions, so it will be difficult to constrain general scalar fields by using the observational data. However, the w ϕ Ω ϕ trajectories for both the tracking and thawing solutions are independent of the initial conditions, so it will be easier to constrain the tracker and thawing potentials. We apply the observational data to reconstruct the thawing potential and we find that the power-law potential V ( ϕ ) ϕ 0.8 is consistent with the observational data. We also find that the potential V ( ϕ ) = V 0 ϕ 0.7 is consistent with the 2 σ upper bound and the potential V ( ϕ ) = V 0 ϕ 1.1 is consistent with the 2 σ lower bound.

Author Contributions

Conceptualization, Q.G.; methodology, Q.G. and Y.Y.; formal analysis, Y.Y. and Q.Q.; writing—original draft preparation, Q.G.; writing—review and editing, Q.G.; supervision, Y.Y.; funding acquisition, Q.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the National Natural Science Foundation of China (Grant No. 12175184), the Chongqing Natural Science Foundation (CSTB2022NSCQ-MSX1324).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Perlmutter, S.; Aldering, G.; Della Valle, M.; Deustua, S.; Ellis, R.S.; Fabbro, S.; Fruchter, A.; Goldhaber, G.; Groom, D.E.; Hook, I.M.; et al. Discovery of a supernova explosion at half the age of the Universe and its cosmological implications. Nature 1998, 391, 51–54. [Google Scholar] [CrossRef] [Green Version]
  2. Riess, A.G.; Filippenko, A.V.; Challis, P.; Clocchiatti, A.; Diercks, A.; Garnavich, P.M.; Gillil, R.L.; Hogan, C.J.; Jha, S.; Kirshner, R.P.; et al. Observational evidence from supernovae for an accelerating universe and a cosmological constant. Astron. J. 1998, 116, 1009–1038. [Google Scholar] [CrossRef] [Green Version]
  3. Perlmutter, S.; Aldering, G.; Goldhaber, G.; Knop, R.A.; Nugent, P.; Castro, P.G.; Deustua, S.; Fabbro, S.; Goobar, A.; Groom, D.E.; et al. Measurements of Ω and Λ from 42 high redshift supernovae. Astrophys. J. 1999, 517, 565–586. [Google Scholar] [CrossRef]
  4. Weinberg, S. The Cosmological Constant Problem. Rev. Mod. Phys. 1989, 61, 1–23. [Google Scholar] [CrossRef]
  5. Ade, P.A.R.; Aghanim, N.; Alves, M.I.R.; Armitage-Caplan, C.; Arnaud, M.; Ashdown, M.; Atrio-Barandela, F.; Aumont, J.; Aussel, H.; Baccigalupi, C.; et al. Planck 2013 results. I. Overview of products and scientific results. Astron. Astrophys. 2014, 571, A1. [Google Scholar] [CrossRef] [Green Version]
  6. Ade, P.A.R.; Aghanim, N.; Armitage-Caplan, C.; Arnaud, M.; Ashdown, M.; Atrio-Barandela, F.; Aumont, J.; Baccigalupi, C.; Banday, A.J.; Barreiro, R.B.; et al. Planck 2013 results. XVI. Cosmological parameters. Astron. Astrophys. 2014, 571, A16. [Google Scholar] [CrossRef] [Green Version]
  7. Riess, A.G.; Macri, L.; Casertano, S.; Lampeitl, H.; Ferguson, H.C.; Filippenko, A.V.; Jha, S.W.; Li, W.; Chornock, R. A 3% Solution: Determination of the Hubble Constant with the Hubble Space Telescope and Wide Field Camera 3. Astrophys. J. 2011, 730, 119, Erratum in Astrophys. J. 2011, 732, 129. [Google Scholar] [CrossRef] [Green Version]
  8. Riess, A.G.; Macri, L.; Casertano, S.; Sosey, M.; Lampeitl, H.; Ferguson, H.C.; Filippenko, A.V.; Jha, S.W.; Li, W.; Chornock, R.; et al. A Redetermination of the Hubble Constant with the Hubble Space Telescope from a Differential Distance Ladder. Astrophys. J. 2009, 699, 539–563. [Google Scholar] [CrossRef] [Green Version]
  9. Efstathiou, G. H0 Revisited. Mon. Not. R. Astron. Soc. 2014, 440, 1138–1152. [Google Scholar] [CrossRef] [Green Version]
  10. Riess, A.G.; Macri, L.M.; Hoffmann, S.L.; Scolnic, D.; Casertano, S.; Filippenko, A.V.; Tucker, B.E.; Reid, M.J.; Jones, D.O.; Silverman, J.M.; et al. A 2.4% Determination of the Local Value of the Hubble Constant. Astrophys. J. 2016, 826, 56. [Google Scholar] [CrossRef] [Green Version]
  11. Riess, A.G.; Casertano, S.; Yuan, W.; Macri, L.M.; Scolnic, D. Large Magellanic Cloud Cepheid Standards Provide a 1% Foundation for the Determination of the Hubble Constant and Stronger Evidence for Physics beyond ΛCDM. Astrophys. J. 2019, 876, 85. [Google Scholar] [CrossRef]
  12. Riess, A.G.; Casertano, S.; Yuan, W.; Bowers, J.B.; Macri, L.; Zinn, J.C.; Scolnic, D. Cosmic Distances Calibrated to 1% Precision with Gaia EDR3 Parallaxes and Hubble Space Telescope Photometry of 75 Milky Way Cepheids Confirm Tension with ΛCDM. Astrophys. J. Lett. 2021, 908, L6. [Google Scholar] [CrossRef]
  13. Riess, A.G.; Breuval, L.; Yuan, W.; Casertano, S.; Macri, L.M.; Bowers, J.B.; Scolnic, D.; Cantat-Gaudin, T.; Anderson, R.I.; Reyes, M.C. Cluster Cepheids with High Precision Gaia Parallaxes, Low Zero-point Uncertainties, and Hubble Space Telescope Photometry. Astrophys. J. 2022, 938, 36. [Google Scholar] [CrossRef]
  14. Freedman, W.L. Measurements of the Hubble Constant: Tensions in Perspective. Astrophys. J. 2021, 919, 16. [Google Scholar] [CrossRef]
  15. Suyu, S.H.; Bonvin, V.; Courbin, F.; Fassnacht, C.D.; Rusu, C.E.; Sluse, D.; Treu, T.; Wong, K.C.; Auger, M.W.; Ding, X.; et al. H0LiCOW—I. H0 Lenses in COSMOGRAIL’s Wellspring: Program overview. Mon. Not. R. Astron. Soc. 2017, 468, 2590–2604. [Google Scholar] [CrossRef] [Green Version]
  16. Wong, K.C.; Suyu, S.H.; Chen, G.C.; Rusu, C.E.; Millon, M.; Sluse, D.; Bonvin, V.; Fassnacht, C.D.; Taubenberger, S.; Auger, M.W. H0LiCOW—XIII. A 2.4 per cent measurement of H0 from lensed quasars: 5.3σ tension between early- and late-Universe probes. Mon. Not. R. Astron. Soc. 2020, 498, 1420–1439. [Google Scholar] [CrossRef]
  17. Birrer, S.; Dhawan, S.; Shajib, A.J. The Hubble Constant from Strongly Lensed Supernovae with Standardizable Magnifications. Astrophys. J. 2022, 924, 2. [Google Scholar] [CrossRef]
  18. Jimenez, R.; Cimatti, A.; Verde, L.; Moresco, M.; Wandelt, B. The local and distant Universe: Stellar ages and H0. JCAP 2019, 3, 43. [Google Scholar] [CrossRef] [Green Version]
  19. D’Amico, G.; Gleyzes, J.; Kokron, N.; Markovic, K.; Senatore, L.; Zhang, P.; Beutler, F.; Gil-Marín, H. The Cosmological Analysis of the SDSS/BOSS data from the Effective Field Theory of Large-Scale Structure. JCAP 2020, 5, 5. [Google Scholar] [CrossRef]
  20. Ivanov, M.M.; Simonović, M.; Zaldarriaga, M. Cosmological Parameters from the BOSS Galaxy Power Spectrum. JCAP 2020, 5, 42. [Google Scholar] [CrossRef]
  21. Colas, T.; D’amico, G.; Senatore, L.; Zhang, P.; Beutler, F. Efficient Cosmological Analysis of the SDSS/BOSS data from the Effective Field Theory of Large-Scale Structure. JCAP 2020, 6, 1. [Google Scholar] [CrossRef]
  22. Birrer, S.; Treu, T. Astrometric requirements for strong lensing time-delay cosmography. Mon. Not. R. Astron. Soc. 2019, 489, 2097–2103. [Google Scholar] [CrossRef]
  23. Camarena, D.; Marra, V. Local determination of the Hubble constant and the deceleration parameter. Phys. Rev. Res. 2020, 2, 013028. [Google Scholar] [CrossRef] [Green Version]
  24. Hinshaw, G.; Larson, D.; Komatsu, E.; Spergel, D.N.; Bennett, C.; Dunkley, J.; Nolta, M.R.; Halpern, M.; Hill, R.S.; Odegard, N.; et al. Nine-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Cosmological Parameter Results. Astrophys. J. Suppl. 2013, 208, 19. [Google Scholar] [CrossRef] [Green Version]
  25. Aghanim, N.; Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Banday, A.J.; Barreiro, R.B.; Bartolo, N.; Basak, S.; et al. Planck 2018 results. VI. Cosmological parameters. Astron. Astrophys. 2020, 641, A6, Erratum in Astron. Astrophys. 2021, 652, C4. [Google Scholar] [CrossRef] [Green Version]
  26. Dainotti, M.G.; De Simone, B.; Schiavone, T.; Montani, G.; Rinaldi, E.; Lambiase, G.; Bogdan, M.; Ugale, S. On the Evolution of the Hubble Constant with the SNe Ia Pantheon Sample and Baryon Acoustic Oscillations: A Feasibility Study for GRB-Cosmology in 2030. Galaxies 2022, 10, 24. [Google Scholar] [CrossRef]
  27. Dainotti, M.G.; De Simone, B.; Schiavone, T.; Montani, G.; Rinaldi, E.; Lambiase, G. On the Hubble constant tension in the SNe Ia Pantheon sample. Astrophys. J. 2021, 912, 150. [Google Scholar] [CrossRef]
  28. Di Valentino, E.; Mena, O.; Pan, S.; Visinelli, L.; Yang, W.; Melchiorri, A.; Mota, D.F.; Riess, A.G.; Silk, J. In the realm of the Hubble tension—A review of solutions. Class. Quant. Grav. 2021, 38, 153001. [Google Scholar] [CrossRef]
  29. Schutz, B.F. Determining the Hubble Constant from Gravitational Wave Observations. Nature 1986, 323, 310–311. [Google Scholar] [CrossRef] [Green Version]
  30. Holz, D.E.; Hughes, S.A. Using gravitational-wave standard sirens. Astrophys. J. 2005, 629, 15–22. [Google Scholar] [CrossRef]
  31. Kyutoku, K.; Seto, N. Gravitational-wave cosmography with LISA and the Hubble tension. Phys. Rev. D 2017, 95, 083525. [Google Scholar] [CrossRef] [Green Version]
  32. Abbott, B.P.; Abbott, R.; Abbott, T.D.; Acernese, F.; Ackley, K.; Adams, C.; Adams, T.; Addesso, P.; Adhikari, R.X.; Adya, V.B.; et al. A gravitational-wave standard siren measurement of the Hubble constant. Nature 2017, 551, 85–88. [Google Scholar] [CrossRef] [Green Version]
  33. Abbott, R.; Abe, H.; Acernese, F.; Ackley, K.; Adhikari, N.; Adhikari, R.X.; Adkins, V.K.; Adya, V.B.; Affeldt, C.; Agarwal, D.; et al. Constraints on the cosmic expansion history from GWTC-3. arXiv 2021, arXiv:2111.03604. [Google Scholar]
  34. Ratra, B.; Peebles, P.J.E. Cosmological Consequences of a Rolling Homogeneous Scalar Field. Phys. Rev. D 1988, 37, 3406. [Google Scholar] [CrossRef] [PubMed]
  35. Wetterich, C. Cosmology and the Fate of Dilatation Symmetry. Nucl. Phys. B 1988, 302, 668–696. [Google Scholar] [CrossRef] [Green Version]
  36. Caldwell, R.R.; Dave, R.; Steinhardt, P.J. Cosmological imprint of an energy component with general equation of state. Phys. Rev. Lett. 1998, 80, 1582–1585. [Google Scholar] [CrossRef] [Green Version]
  37. Zlatev, I.; Wang, L.M.; Steinhardt, P.J. Quintessence, cosmic coincidence, and the cosmological constant. Phys. Rev. Lett. 1999, 82, 896–899. [Google Scholar] [CrossRef] [Green Version]
  38. Banerjee, A.; Cai, H.; Heisenberg, L.; Colgáin, E.O.; Sheikh-Jabbari, M.M.; Yang, T. Hubble sinks in the low-redshift swampland. Phys. Rev. D 2021, 103, L081305. [Google Scholar] [CrossRef]
  39. Lee, B.H.; Lee, W.; Colgáin, E.O.; Sheikh-Jabbari, M.M.; Thakur, S. Is local H 0 at odds with dark energy EFT? JCAP 2022, 4, 4. [Google Scholar] [CrossRef]
  40. Heisenberg, L.; Villarrubia-Rojo, H.; Zosso, J. Simultaneously solving the H0 and σ8 tensions with late dark energy. Phys. Dark Univ. 2023, 39, 101163. [Google Scholar] [CrossRef]
  41. Heisenberg, L.; Villarrubia-Rojo, H.; Zosso, J. Can late-time extensions solve the H0 and σ8 tensions? Phys. Rev. D 2022, 106, 043503. [Google Scholar] [CrossRef]
  42. Caldwell, R.R. A Phantom menace? Phys. Lett. B 2002, 545, 23–29. [Google Scholar] [CrossRef] [Green Version]
  43. Feng, B.; Wang, X.L.; Zhang, X.M. Dark energy constraints from the cosmic age and supernova. Phys. Lett. B 2005, 607, 35–41. [Google Scholar] [CrossRef]
  44. Feng, B.; Li, M.; Piao, Y.S.; Zhang, X. Oscillating quintom and the recurrent universe. Phys. Lett. B 2006, 634, 101–105. [Google Scholar] [CrossRef] [Green Version]
  45. Guo, Z.K.; Piao, Y.S.; Zhang, X.M.; Zhang, Y.Z. Cosmological evolution of a quintom model of dark energy. Phys. Lett. B 2005, 608, 177–182. [Google Scholar] [CrossRef] [Green Version]
  46. Sen, A. Tachyon matter. JHEP 2002, 7, 65. [Google Scholar] [CrossRef] [Green Version]
  47. Sen, A. Rolling tachyon. JHEP 2002, 4, 48. [Google Scholar] [CrossRef]
  48. Padmanabhan, T. Accelerated expansion of the universe driven by tachyonic matter. Phys. Rev. D 2002, 66, 021301. [Google Scholar] [CrossRef] [Green Version]
  49. Armendariz-Picon, C.; Mukhanov, V.F.; Steinhardt, P.J. A Dynamical solution to the problem of a small cosmological constant and late time cosmic acceleration. Phys. Rev. Lett. 2000, 85, 4438–4441. [Google Scholar] [CrossRef] [Green Version]
  50. Dvali, G.R.; Gabadadze, G.; Porrati, M. 4-D gravity on a brane in 5-D Minkowski space. Phys. Lett. B 2000, 485, 208–214. [Google Scholar] [CrossRef] [Green Version]
  51. Carroll, S.M.; Duvvuri, V.; Trodden, M.; Turner, M.S. Is cosmic speed - up due to new gravitational physics? Phys. Rev. D 2004, 70, 043528. [Google Scholar] [CrossRef] [Green Version]
  52. Nojiri, S.; Odintsov, S.D. Modified gravity with negative and positive powers of the curvature: Unification of the inflation and of the cosmic acceleration. Phys. Rev. D 2003, 68, 123512. [Google Scholar] [CrossRef] [Green Version]
  53. Starobinsky, A.A. Disappearing cosmological constant in f(R) gravity. JETP Lett. 2007, 86, 157–163. [Google Scholar] [CrossRef] [Green Version]
  54. Hu, W.; Sawicki, I. Models of f(R) Cosmic Acceleration that Evade Solar-System Tests. Phys. Rev. D 2007, 76, 064004. [Google Scholar] [CrossRef] [Green Version]
  55. de Rham, C.; Gabadadze, G.; Tolley, A.J. Resummation of Massive Gravity. Phys. Rev. Lett. 2011, 106, 231101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Gong, Y. Cosmology in massive gravity. Commun. Theor. Phys. 2013, 59, 319–323. [Google Scholar] [CrossRef] [Green Version]
  57. Sahni, V.; Starobinsky, A.A. The Case for a positive cosmological Lambda term. Int. J. Mod. Phys. D 2000, 9, 373–444. [Google Scholar] [CrossRef]
  58. Copeland, E.J.; Sami, M.; Tsujikawa, S. Dynamics of dark energy. Int. J. Mod. Phys. D 2006, 15, 1753–1936. [Google Scholar] [CrossRef] [Green Version]
  59. Padmanabhan, T. Dark energy and gravity. Gen. Rel. Grav. 2008, 40, 529–564. [Google Scholar] [CrossRef] [Green Version]
  60. Li, M.; Li, X.D.; Wang, S.; Wang, Y. Dark Energy. Commun. Theor. Phys. 2011, 56, 525–604. [Google Scholar] [CrossRef] [Green Version]
  61. Nojiri, S.; Odintsov, S.D. Unified cosmic history in modified gravity: From F(R) theory to Lorentz non-invariant models. Phys. Rept. 2011, 505, 59–144. [Google Scholar] [CrossRef] [Green Version]
  62. Hinterbichler, K. Theoretical Aspects of Massive Gravity. Rev. Mod. Phys. 2012, 84, 671–710. [Google Scholar] [CrossRef] [Green Version]
  63. Clifton, T.; Ferreira, P.G.; Padilla, A.; Skordis, C. Modified Gravity and Cosmology. Phys. Rept. 2012, 513, 1–189. [Google Scholar] [CrossRef] [Green Version]
  64. de Rham, C. Massive Gravity. Living Rev. Rel. 2014, 17, 7. [Google Scholar] [CrossRef] [Green Version]
  65. Nojiri, S.; Odintsov, S.D.; Oikonomou, V.K. Modified Gravity Theories on a Nutshell: Inflation, Bounce and Late-time Evolution. Phys. Rept. 2017, 692, 1–104. [Google Scholar] [CrossRef] [Green Version]
  66. Copeland, E.J.; Liddle, A.R.; Wands, D. Exponential potentials and cosmological scaling solutions. Phys. Rev. D 1998, 57, 4686–4690. [Google Scholar] [CrossRef] [Green Version]
  67. Liddle, A.R.; Scherrer, R.J. A Classification of scalar field potentials with cosmological scaling solutions. Phys. Rev. D 1999, 59, 023509. [Google Scholar] [CrossRef] [Green Version]
  68. Steinhardt, P.J.; Wang, L.M.; Zlatev, I. Cosmological tracking solutions. Phys. Rev. D 1999, 59, 123504. [Google Scholar] [CrossRef] [Green Version]
  69. Brax, P.; Martin, J. The Robustness of quintessence. Phys. Rev. D 2000, 61, 103502. [Google Scholar] [CrossRef] [Green Version]
  70. Urena-Lopez, L.A.; Matos, T. A New cosmological tracker solution for quintessence. Phys. Rev. D 2000, 62, 081302. [Google Scholar] [CrossRef] [Green Version]
  71. Bludman, S.A.; Roos, M. Quintessence cosmology and the cosmic coincidence. Phys. Rev. D 2002, 65, 043503. [Google Scholar] [CrossRef] [Green Version]
  72. Dodelson, S.; Kaplinghat, M.; Stewart, E. Solving the Coincidence Problem: Tracking Oscillating Energy. Phys. Rev. Lett. 2000, 85, 5276–5279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Johri, V.B. Search for tracker potentials in quintessence theory. Class. Quant. Grav. 2002, 19, 5959–5968. [Google Scholar] [CrossRef] [Green Version]
  74. Rubano, C.; Scudellaro, P.; Piedipalumbo, E.; Capozziello, S.; Capone, M. Exponential potentials for tracker fields. Phys. Rev. D 2004, 69, 103510. [Google Scholar] [CrossRef] [Green Version]
  75. Watson, C.R.; Scherrer, R.J. The Evolution of inverse power law quintessence at low redshift. Phys. Rev. D 2003, 68, 123524. [Google Scholar] [CrossRef] [Green Version]
  76. Aguirregabiria, J.M.; Lazkoz, R. Tracking solutions in tachyon cosmology. Phys. Rev. D 2004, 69, 123502. [Google Scholar] [CrossRef] [Green Version]
  77. Fang, W.; Li, Y.; Zhang, K.; Lu, H.Q. Exact Analysis of Scaling and Dominant Attractors Beyond the Exponential Potential. Class. Quant. Grav. 2009, 26, 155005. [Google Scholar] [CrossRef]
  78. Scherrer, R.J.; Sen, A.A. Thawing quintessence with a nearly flat potential. Phys. Rev. D 2008, 77, 083515. [Google Scholar] [CrossRef] [Green Version]
  79. Scherrer, R.J.; Sen, A.A. Phantom Dark Energy Models with a Nearly Flat Potential. Phys. Rev. D 2008, 78, 067303. [Google Scholar] [CrossRef] [Green Version]
  80. Dutta, S.; Scherrer, R.J. Hilltop Quintessence. Phys. Rev. D 2008, 78, 123525. [Google Scholar] [CrossRef] [Green Version]
  81. Dutta, S.; Scherrer, R.J. Slow-roll freezing quintessence. Phys. Lett. B 2011, 704, 265–269. [Google Scholar] [CrossRef] [Green Version]
  82. Chiba, T.; Dutta, S.; Scherrer, R.J. Slow-roll k-essence. Phys. Rev. D 2009, 80, 043517. [Google Scholar] [CrossRef]
  83. Gupta, G.; Saridakis, E.N.; Sen, A.A. Non-minimal quintessence and phantom with nearly flat potentials. Phys. Rev. D 2009, 79, 123013. [Google Scholar] [CrossRef]
  84. Chiba, T. Slow-Roll Thawing Quintessence. Phys. Rev. D 2009, 79, 083517, Erratum in Phys. Rev. D 2009, 80, 109902. [Google Scholar] [CrossRef] [Green Version]
  85. Sen, S.; Sen, A.A.; Sami, M. The thawing dark energy dynamics: Can we detect it? Phys. Lett. B 2010, 686, 1–5. [Google Scholar] [CrossRef] [Green Version]
  86. del Campo, S.; Cardenas, V.H.; Herrera, R. The effect of curvature in thawing models. Phys. Lett. B 2011, 694, 279–283. [Google Scholar] [CrossRef] [Green Version]
  87. Gong, Y.; Gao, Q. On the effect of the degeneracy among dark energy parameters. Eur. Phys. J. C 2014, 74, 2729. [Google Scholar] [CrossRef]
  88. Gong, Y. The general property of dynamical quintessence field. Phys. Lett. B 2014, 731, 342–349. [Google Scholar] [CrossRef] [Green Version]
  89. Efstathiou, G. Constraining the equation of state of the universe from distant type Ia supernovae and cosmic microwave background anisotropies. Mon. Not. R. Astron. Soc. 1999, 310, 842–850. [Google Scholar] [CrossRef] [Green Version]
  90. Chevallier, M.; Polarski, D. Accelerating universes with scaling dark matter. Int. J. Mod. Phys. D 2001, 10, 213–224. [Google Scholar] [CrossRef] [Green Version]
  91. Linder, E.V. Exploring the expansion history of the universe. Phys. Rev. Lett. 2003, 90, 091301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Gao, Q.; Gong, Y. Constraints on slow-roll thawing models from fundamental constants. Int. J. Mod. Phys. D 2013, 22, 1350035. [Google Scholar] [CrossRef] [Green Version]
  93. Gao, Q.; Gong, Y. The tension on the cosmological parameters from different observational data. Class. Quant. Grav. 2014, 31, 105007. [Google Scholar] [CrossRef] [Green Version]
  94. Chiba, T. The Equation of State of Tracker Fields. Phys. Rev. D 2010, 81, 023515. [Google Scholar] [CrossRef] [Green Version]
  95. Betoule, M.; Kessler, R.; Guy, J.; Mosher, J.; Hardin, D.; Biswas, R.; Astier, P.; El-Hage, P.; Konig, M.; Kuhlmann, S.; et al. Improved cosmological constraints from a joint analysis of the SDSS-II and SNLS supernova samples. Astron. Astrophys. 2014, 568, A22. [Google Scholar] [CrossRef]
  96. Bennett, C.L.; Larson, D.; Weil, J.L.; Jarosik, N.; Hinshaw, G.; Odegard, N.; Smith, K.M.; Hill, R.S.; Gold, B.; Halpern, M.; et al. Nine-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Final Maps and Results. Astrophys. J. Suppl. 2013, 208, 20. [Google Scholar] [CrossRef] [Green Version]
  97. Beutler, F.; Blake, C.; Colless, M.; Jones, D.H.; Staveley-Smith, L.; Campbell, L.; Parker, Q.; Saunders, W.; Watson, F. The 6dF Galaxy Survey: Baryon Acoustic Oscillations and the Local Hubble Constant. Mon. Not. R. Astron. Soc. 2011, 416, 3017–3032. [Google Scholar] [CrossRef]
  98. Percival, W.J.; Reid, B.A.; Eisenstein, D.J.; Bahcall, N.A.; Budavari, T.; Frieman, J.A.; Fukugita, M.; Gunn, J.E.; Ivezić, Ž.; Knapp, G.R.; et al. Baryon Acoustic Oscillations in the Sloan Digital Sky Survey Data Release 7 Galaxy Sample. Mon. Not. R. Astron. Soc. 2010, 401, 2148–2168. [Google Scholar] [CrossRef] [Green Version]
  99. Anderson, L.; Aubourg, E.; Bailey, S.; Bizyaev, D.; Blanton, M.; Bolton, A.S.; Brinkmann, J.; Brownstein, J.R.; Burden, A.; Cuesta, A.J.; et al. The clustering of galaxies in the SDSS-III Baryon Oscillation Spectroscopic Survey: Baryon Acoustic Oscillations in the Data Release 9 Spectroscopic Galaxy Sample. Mon. Not. R. Astron. Soc. 2013, 427, 3435–3467. [Google Scholar] [CrossRef] [Green Version]
  100. Blake, C.; Kazin, E.A.; Beutler, F.; Davis, T.M.; Parkinson, D.; Brough, S.; Colless, M.; Contreras, C.; Couch, W.; Croom, S.; et al. The WiggleZ Dark Energy Survey: Mapping the distance-redshift relation with baryon acoustic oscillations. Mon. Not. R. Astron. Soc. 2011, 418, 1707–1724. [Google Scholar] [CrossRef]
  101. Busca, N.G.; Delubac, T.; Rich, J.; Bailey, S.; Font-Ribera, A.; Kirkby, D.; Le Goff, J.M.; Pieri, M.M.; Slosar, A.; Aubourg, É.; et al. Baryon Acoustic Oscillations in the Ly-α forest of BOSS quasars. Astron. Astrophys. 2013, 552, A96. [Google Scholar] [CrossRef] [Green Version]
  102. Simon, J.; Verde, L.; Jimenez, R. Constraints on the redshift dependence of the dark energy potential. Phys. Rev. D 2005, 71, 123001. [Google Scholar] [CrossRef] [Green Version]
  103. Gaztanaga, E.; Cabre, A.; Hui, L. Clustering of Luminous Red Galaxies IV: Baryon Acoustic Peak in the Line-of-Sight Direction and a Direct Measurement of H(z). Mon. Not. R. Astron. Soc. 2009, 399, 1663–1680. [Google Scholar] [CrossRef] [Green Version]
  104. Stern, D.; Jimenez, R.; Verde, L.; Kamionkowski, M.; Stanford, S.A. Cosmic Chronometers: Constraining the Equation of State of Dark Energy. I: H(z) Measurements. JCAP 2010, 2, 8. [Google Scholar] [CrossRef] [Green Version]
  105. Moresco, M.; Cimatti, A.; Jimenez, R.; Pozzetti, L.; Zamorani, G.; Bolzonella, M.; Dunlop, J.; Lamareille, F.; Mignoli, M.; Pearce, H.; et al. Improved constraints on the expansion rate of the Universe up to z~1.1 from the spectroscopic evolution of cosmic chronometers. JCAP 2012, 8, 6. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The w ϕ Ω ϕ trajectories for the inverse power-law potential V ( ϕ ) ϕ α , the potential V ( ϕ ) exp ( 1 / ϕ ) , and the potential V ( ϕ ) = V 0 sinh 2 ( 0.1 ϕ ) with the tracking behavior. The dashed lines denote the analytical approximations (9)–(11). The shaded region is the allowed region by astronomical observations, Ω ϕ 0 ( 0.6 , 0.8 ) and w ϕ 0 < 0.7 .
Figure 1. The w ϕ Ω ϕ trajectories for the inverse power-law potential V ( ϕ ) ϕ α , the potential V ( ϕ ) exp ( 1 / ϕ ) , and the potential V ( ϕ ) = V 0 sinh 2 ( 0.1 ϕ ) with the tracking behavior. The dashed lines denote the analytical approximations (9)–(11). The shaded region is the allowed region by astronomical observations, Ω ϕ 0 ( 0.6 , 0.8 ) and w ϕ 0 < 0.7 .
Universe 09 00146 g001
Figure 2. The marginalized 68% and 95% confidence contours of w ϕ 0 and Ω ϕ 0 from observational data. The solid lines denote the analytical relation (13).
Figure 2. The marginalized 68% and 95% confidence contours of w ϕ 0 and Ω ϕ 0 from observational data. The solid lines denote the analytical relation (13).
Universe 09 00146 g002
Figure 3. The 2 σ constraint on quintessence thawing potentials. The potential is in the unit of the current critical density ρ c r 0 and Δ ϕ is the difference between the scalar field ϕ and its current value ϕ 0 . The dashed lines denotes the power-law potential V ( ϕ ) ϕ 0.8 .
Figure 3. The 2 σ constraint on quintessence thawing potentials. The potential is in the unit of the current critical density ρ c r 0 and Δ ϕ is the difference between the scalar field ϕ and its current value ϕ 0 . The dashed lines denotes the power-law potential V ( ϕ ) ϕ 0.8 .
Universe 09 00146 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

You, Y.; Qiang, Q.; Gao, Q. The General Property of Tracking and Thawing Models and Their Observational Constraints. Universe 2023, 9, 146. https://doi.org/10.3390/universe9030146

AMA Style

You Y, Qiang Q, Gao Q. The General Property of Tracking and Thawing Models and Their Observational Constraints. Universe. 2023; 9(3):146. https://doi.org/10.3390/universe9030146

Chicago/Turabian Style

You, Yujie, Qichao Qiang, and Qing Gao. 2023. "The General Property of Tracking and Thawing Models and Their Observational Constraints" Universe 9, no. 3: 146. https://doi.org/10.3390/universe9030146

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop