Next Article in Journal
Nonanalytic Relativistic r-Modes of Slowly Rotating Nonbarotropic Neutron Stars
Next Article in Special Issue
Quintom Fields from Chiral K-Essence Cosmology
Previous Article in Journal
On a Class of Harko-Kovacs-Lobo Wormholes
Previous Article in Special Issue
Metric Gravity in the Hamiltonian Form—Canonical Transformations—Dirac’s Modifications of the Hamilton Method and Integral Invariants of the Metric Gravity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sign Switching Dark Energy from a Running Barrow Entropy

1
Center for Gravitation and Cosmology, College of Physical Science and Technology, Yangzhou University, 180 Siwangting Road, Yangzhou 225002, China
2
Shanghai Frontier Science Center for Gravitational Wave Detection, School of Aeronautics and Astronautics, Shanghai Jiao Tong University, Shanghai 200240, China
*
Author to whom correspondence should be addressed.
Universe 2022, 8(10), 541; https://doi.org/10.3390/universe8100541
Submission received: 4 October 2022 / Revised: 13 October 2022 / Accepted: 14 October 2022 / Published: 19 October 2022
(This article belongs to the Collection Modified Theories of Gravity and Cosmological Applications)

Abstract

:
Barrow proposed that the area law of the entropy associated with a horizon might receive a “fractal correction” due to quantum gravitational effects—in place of S A , we have instead S A 1 + δ / 2 , where 0 δ 1 measures the deviation from the standard area law ( δ = 0 ). Based on black hole thermodynamics, we argue that the Barrow entropy should run (i.e., energy scale dependent), which is reasonable given that quantum gravitational corrections are expected to be important only in the high-energy regime. When applied to the Friedmann equation, we demonstrate the possibility that such a running Barrow entropy index could give rise to a dynamical effective dark energy, which is asymptotically positive and vanishing, but negative at the Big Bang. Such a sign switching dark energy could help to alleviate the Hubble tension. Other cosmological implications are discussed.

1. Introduction: Barrow Entropy and Quantum Gravity

One of the most striking properties of gravity is that black holes—and by extension, apparent horizons—have entropy that scales with the surface area [1,2]. Specifically, the Bekenstein–Hawking entropy is
S = A 4 k B c 3 G ħ ,
where k B , c, and G are the Boltzmann constant, the speed of light, and Newton’s gravitational constant, respectively.
In different approaches to quantum gravity, this expression receives different forms of correction (see Section 2). Recently, Barrow [3] proposed that, due to quantum gravity corrections, black holes might exist with extremely wrinkled surfaces, such that the event horizon is like, e.g., a Koch fractal surface. The entropy expression becomes
S = A A 0 1 + δ 2 ,
where 0 δ 1 is the parameter that governs how “fractalized” the surface has become, and we shall refer to it as the “Barrow entropy index” (BEI). It is not quite clear whether we should have S = k B 4 A p 2 1 + δ / 2 or S = k B A 4 p 2 1 + δ / 2 . Barrow worked with S A / 4 A from the beginning (before the correction), and thus, this issue was not discussed. However, the constant factor 1 / 4 , whether it is raised to the power of 1 + δ / 2 or not, does not affect the results much, except possibly when dealing with microscopic black holes, since the area would be very large anyway, so we use the form of Equation (2), in line with most literature, for easy comparison. However, for definiteness, we set A 0 = 4 p 2 , where p is the Planck length p = G ħ / c 3 . Hereinafter, we employ the Planck units, so that ħ = G = c = k B = 1 .
Cosmological and black hole shadow constraints (assuming a fixed δ ) have put an upper bound on δ , which is typically δ O ( 10 3 ) or O ( 10 4 ) [4,5,6,7,8,9].
In this work, we argue that δ has to be a function of the energy scale and study the implications of Barrow entropy for cosmology (by the expansion of the universe, δ is thus a function of cosmic time). Such a possibility was also raised in [7,10,11]. We expect that, as quantum gravitational effects become more pronounced near the Big Bang, δ 1 . As the universe expands and cools, δ 0 monotonically. Incidentally, in the context of an asymptotically flat Schwarzschild black hole, this would mean that, as the black hole evaporates and its temperature increases, δ would approach unity (the effect of a running BEI on Hawking evaporation will be studied elsewhere).
The running Barrow entropy index gives rise to an additional term in the modified Friedmann equation, which we propose to be identified with an effective dynamical dark energy, which is asymptotically zero (hence, small at late time). Surprisingly, the effective dark energy has a negative sign right about the Big Bang, i.e., the Universe started out as an anti-de Sitter-like spacetime. Interestingly, this may help to relax the Hubble tension—the mismatch between the locally measured expansion rate of the universe and the inferred rate from early universe via the cosmic microwave background (CMB) [12,13].
In addition, following a recent work of Sheykhi [14], which considered the fixed δ scenario, we shall identify the correction on the right-hand side (r.h.s.) of the Friedmann equation as an effective modification on the gravitational constant, instead of the matter field. This has the advantage that the resulting fixed δ cosmology satisfies the generalized second law (GSL), whereas if the matter sector is modified instead of the geometry, then the GSL fails [15]. The viability of the GSL in the context of the interacting Barrow holographic dark energy model was also investigated in [16].

2. Barrow Entropy Index Is Energy Scale Dependent

In a recent work [17], Chen et al. studied black holes in asymptotically safe gravity, in which the gravitational constant also runs [18] (see [19,20] for cosmological studies). They showed that the area law receives a well-known logarithmic correction in the quantum gravity literature:
S = A 4 + C ln ( A ) ;
see [21] for a review. This at least suggests that the reverse implication—a modification of the Bekenstein–Hawking entropy leading to a varying effective gravitational constant (as in the BEI case)—is perhaps not so surprising.
Here, we remark that, although the Barrow entropy is supposed to quantify a quantum gravitational correction, it is not the—arguably more understood—quantum gravity effect that gives rise to Equation (3). (In fact, the effects of having both corrections together were considered in [22].) The sign of the constant C in front of the logarithmic term in Equation (3) is usually negative [21], but positive in some works, such as [17,23]. In [24], Gour argued that a positive constant might be required if the area fluctuation is taken into account, whereas for a fixed area, the corrections due to the number of microstates (that describe the black hole) would give a negative constant. We have not much to add to this discussion. Instead, we would like to point out that the Barrow entropy, Equation (2), if expanded out as a series assuming small δ (when the quantum gravity correction is small), yields, up to the first order in δ ,
S = A 4 + A 8 ln A 4 δ .
Note the presence of the area as a coefficient of the logarithmic correction. Thus, there is a risk that the “correction” term can be of the same order as the leading term. In fact, for any fixed δ 1 , the expression will still deviate significantly from the Bekenstein–Hawking area law for a sufficiently large black hole, precisely when we expect quantum gravity corrections to be small. This provides a strong argument for a running BEI— δ should at least scale inversely proportional to A ln ( A ) to keep the subleading term small at O ( 1 ) . Such a running BEI would guarantee that quantum gravity correction is small when the black hole is large (i.e., when the energy scale is small, since the Hawking temperature is inversely proportional to the mass). Another option is to consider a δ that runs like 1 / A , then the Barrow entropy would just be the same as the standard logarithmic correction (at least up to the first order in δ ), but in this work, we take the former view, with the hope of uncovering new physics.

3. Dark Energy from a Running Barrow Entropy Index

Sheykhi derived the correction to the Friedmann equation due to the Barrow entropy (with a fixed BEI) in [14]:
H 2 + k a 2 1 δ 2 = 8 π 3 ρ 2 δ 2 + δ A 0 1 + δ 2 4 ( 4 π ) δ 2 = : 8 π ρ 3 G eff .
In other words, the matter field feels an effective gravitational constant1 G eff instead of G. This is one of the main results of [14].
This was obtained by applying the first law of thermodynamics to the energy flux through the apparent horizon of the Friedmann–Lemaître–Robertson–Walker (FLRW) universe, instead of the black hole horizon [23]. The expression for the horizon is
r AH = H 2 + k a 2 1 2 .
In the following, we assumed a spatially flat universe ( k = 0 ).
Allowing the Barrow entropy index to run, the change in entropy would be:
d S = d 4 π r 2 A 0 1 + δ 2 = 4 π r 2 A 0 1 + δ 2 ( 2 + δ ) r ˙ r + ln 4 π r 2 A 0 δ ˙ 2 d t .
Then, using the first law of thermodynamics and the continuity equation, we can show, following [14], that
d E = 1 2 π r 4 π r 2 A 0 1 + δ 2 ( 2 + δ ) r ˙ r + ln 4 π r 2 A 0 δ ˙ 2 d t = 4 π r 3 H ( ρ + p ) d t ,
and, consequently,
( 4 π ) δ 2 2 π A 0 1 + δ 2 ( 2 + δ ) r δ 3 + ln 4 π r 2 A 0 r δ 2 2 d δ d r d r = ρ ˙ 3 d t .
We note that r AH = 1 / H . In a sensible cosmology, H decreases with time, so r AH is an increasing function of time. In other words, a larger r corresponds to a later time. From the black hole thermodynamics argument as per Section 2, we want δ to at least decrease as δ const . / r 2 ln ( r 2 ) at large r. At early time, it should approach unity. For definiteness, we can take a function that is always smaller than the asymptotic behavior const . / r 2 ln ( r 2 ) , so a natural choice is δ ( r ) = e r .
The differential form of the modified Friedmann equation is equivalent to
( 2 δ ) r δ 3 d r + 2 δ 2 + δ ln 4 π r 2 A 0 r δ 2 2 d δ d r d r = 8 π 3 2 δ 2 + δ A 0 1 + δ 2 4 ( 4 π ) δ 2 = : G eff d ρ ,
where G eff is the effective gravitational constant, the same as defined in the fixed BEI case [14], as in Equation (5). We immediately notice that G eff is monotonically decreasing as δ 0 , with G eff ( δ = 1 ) = 1 / ( 3 π ) 0.189 near the Big Bang, which is about 20%, the percent value of unity (in Planck units). We defer the possible cosmological implications of this varying G eff to the Discussion Section.
For now, we note that, since δ is a function of H, the Friedmann equation in the case of a running BEI cannot be easily solved by direct integration. We can, however, still analyze the late time and early time behavior separately. At late time, we chose some r * sufficiently large such that 1 / r 2 ln ( r 2 ) 1 is as small as we wish. We have δ 0 (more precisely, we may say that δ is small enough that it is slowly varying compared to r, so we can take some small value of fixed δ when integrating with respect to r). Then, with A 0 = 4 in Planck units, the first term yields (with R > r * )
2 r ˜ R r 3 d r = 2 1 2 r 2 r * R = 1 r * 2 1 R 2 .
The r.h.s. of Equation (10) integrates to give the matter density
8 π 3 [ ρ ] ρ * ρ ( R ) .
Thus, in the absence of the second term in Equation (10), this will just give the usual Friedmann equation (as we took R ; note ρ ( R ) 0 in this limit, because matter density decreases as the universe expands):
H 2 = 8 π ρ 3
evaluated at the time that corresponds to r * .
Therefore, we can interpret the second term in Equation (10) as the differential form of the effective dark energy (so that Equation (10) yields the form2 H 2 Λ / 3 = 8 π ρ / 3 ):
d Λ eff 3 = 2 δ 2 + δ ln 4 π r 2 A 0 r δ 2 2 d δ d r d r .
If we use the ansatz 1 / r 2 ln ( r 2 ) for δ , then
d Λ eff 3 = 2 δ 2 + δ ln 4 π r 2 A 0 × r δ 2 2 2 r 3 ln ( r 2 ) + 2 r 3 ( ln ( r 2 ) ) 2 d r .
At late time, δ 0 , we have an upper bound:
Λ eff 3 Λ ( r * ) Λ ( R ) 1 2 r * R ln ( π r 2 ) r 2 2 r 3 ln ( r 2 ) + 2 r 3 ( ln ( r 2 ) ) 2 d r < 1 2 r * R ln ( π r 2 ) r 2 4 r 3 ln ( r 2 ) d r = 2 r * R 1 r 5 1 ln ( r 2 ) [ ln π + ln ( r 2 ) ] d r < 2 r * R 1 r 5 + 1 r 5 d r = 4 r * R 1 r 5 d r = 1 r 4 r * R = 1 r * 4 1 R 4 .
That is,
Λ eff ( r * ) 3 < 1 r * 4
as R . Since r * was chosen to be very large, Λ eff must be very small, and furthermore, it becomes smaller at later time. In fact, the integral Equation (16) yields
Λ eff ( r * ) 3 4 1 r * 4 ln ( π r * 2 ) ln ( r * 2 ) 2 1 ln ( π ) Ei 4 ln ( r * ) ,
where Ei is the exponential integral function. One can check numerically that it is positive and decreases towards 0 very rapidly. Such an asymptotically vanishing effective cosmological constant scenario was previously discussed in [26,27,28,29,30,31]. Next, we show that, although the universe is late time de Sitter (dS)-like, it is anti-de Sitter (AdS)-like in the very early universe. Such a behavior had been observed, for example, in non-commutative quantum field theory [32].
In the vicinity of the Big Bang, the effective dark energy term depends on the profile of δ (how δ approaches 1 / r 2 ln ( r 2 ) ). For example, if δ = 1 for some time after the Big Bang, then d δ / d r = 0 , and there is no effective dark energy. However, if we chose as we did previously δ = e r , then the effective dark energy at the Big Bang is (with δ 1 and 0 < ε 1 )
Λ eff 3 Λ eff ( ε ) Λ eff ( 2 ε ) 1 3 ε 2 ε ln ( π r 2 ) e r 2 r d r .
The integration interval is kept small near the Big Bang to ensure that δ = e r is nearly constant, so that we can approximate as follows:
Λ eff Λ eff ( ε ) Λ eff ( 2 ε ) ε 2 ε ln ( π r 2 ) e r 2 r d r ε ln ( π ε 2 ) e ε 2 ε .
That is,
Λ eff ( 2 ε ) Λ eff ( ε ) ε ln ( π ε 2 ) e ε 2 ε .
In the limit ε 0 , this yields d Λ eff / d ε . It follows that3 Λ eff at the Big Bang with a log divergence 4. It is worth commenting that [5] gives a strong constraint for the smallness of δ (assuming δ to be fixed) during the Big Bang nucleosynthesis (BBN) epoch. Thus, we know that δ cannot be slowly varying in the early universe if δ starts out near unity at the Big Bang. In fact, this is partly the reason why we chose δ = e r in our toy model, as it decays rather quickly at the beginning. Here, however, we used the mathematical fact that any function is nearly constant if the interval of its domain is sufficiently small.
An infinitely large negative cosmological “constant” is not good (the universe might immediately re-collapse), though we should also note that this is a logarithmic divergence O ( ln ε ) , which is still “smaller” in magnitude than the 1 / ε -divergence in the modified Hubble term 5 (from the first term in Equation (10)), which is proportional to
ε 2 ε r 2 d r ε 1 ε 2 = 1 ε .
Of course, we could in principle choose other functions to obtain a finite nonzero value. Regardless, the integral:
Λ eff ( ε ) ε 2 ε ln π r 2 1 2 r d δ d r d r .
is always negative because d δ / d r < 0 and ln ( π r 2 ) < 0 for small r, as long as δ is a strictly decreasing function. Therefore, Λ eff started out negative and crossed zero at some point in time (at r H = 1 / π 0.564 ), and at late time, it is small and positive. By continuity, this suggests that Λ eff reached a maximum value at some point in the past, before decreasing towards its current small value 10 122 . Another possibility is that δ contains a factor that cancels the ln ( π r 2 ) term; then, in principle, the transition for AdS to dS can happen at an even later time, depending on the zero of the function. Similar scenarios have been proposed in the literature [33,34,35,36,37,38,39,40], with [35,36], that the transition from the AdS phase to the dS phase occurred around the recombination epoch, while [37] argued that the transition occurred as late as at redshift z 2.32 , which triggered the late time acceleration. It is also worth mentioning that [35,41] proposed that the AdS–dS transition could help to alleviate the Hubble tension. In fact, [41] argued that the S 8 tension [42,43] (equivalently, the σ 8 tension) is also alleviated, among other improvements; see also [44]. In [45], the authors concluded that solving both the H 0 and σ 8 tensions would require that G eff < G at some redshift z, if the equation of state satisfies w ( z ) 1 (otherwise, w ( z ) must cross the phantom divide). Thus, it is worth mentioning that a running BEI also causes a varying gravitational constant, and furthermore, from Equation (5), our G eff , in the Planck units, always satisfies
1 3 π < G eff = 2 δ 2 + δ 1 π δ 2 < 1 = G .
for all values of admissible δ [ 0 , 1 ] . However, there are also arguments that state that this is unlikely to alleviate the Hubble tension by itself [46]; see also the subtleties discussed in the next section.

4. Discussion

In this work, we argued that if the Barrow entropy does encode some form of quantum gravitational correction, then in order to be consistent with the black hole thermodynamics of large black holes (for which quantum gravity correction should be small), the BEI must vary according to the energy scale. In terms of the horizon area, it should decay at least as fast as 1 / ( A ln A ) .
When applied to the cosmological context, we showed that the modified Friedmann equation has many surprising properties: the universe started out with an effective negative cosmological constant, but at late time, it became positive and small, decaying towards zero asymptotically. As already explored in the literature, such a scenario could help to alleviate the Hubble tension and the S 8 tension. See, however, in ref. [47,48].
Despite the fact that the BEI is, thus far, only a phenomenological parametrization of quantum gravity effects, it is interesting to see that it could potentially explain some cosmological mysteries. In fact, although the explicit form of δ can only be determined once we have a better understanding of how to derive the Barrow entropy from a theory of quantum gravity, the features we obtained are quite generic as long as δ is assumed to be monotonically decreasing with the energy scale. The shortcoming is that we still do not know the exact behavior of δ , especially at early time. Another caveat is that in deriving Equation (8), we used the first law as d E = T d S = ρ ˙ V d t , in which the volume change was neglected, which amounts to the assumption that r ˙ H r . This holds at late time with exponential late time expansion, but may not hold near the Big Bang; this will depend on the explicit solution of the scale factor, which in turn depends on the profile of δ .
Lastly, we comment on the varying effective gravitational constant. Whenever there is a varying G, one might worry that the continuity equation (the one employed in Equation (8) is the standard one ρ ˙ = 3 H ( ρ + p ) that follows from μ T μ ν = 0 ) may be modified since μ ( G T μ ν ) may no longer imply μ ( T μ ν ) = 0 . However, the “bare” gravitational constant G that appears at the level of the action is not the same as the effective G eff that appears in the modified Friedmann Equation (10). Furthermore, the Barrow entropy itself contains A 0 = 4 G , which is the “bare” G. The fact that G eff is smaller in the early universe could suggest that we may be able to ameliorate the arrow of time problem [49,50,51,52,53,54,55,56,57,58] following the same line of thought as Greene et al. [59], as well as Sloan and Ellis [60]. (Note that to ameliorate the H 0 tension, one usually requires that G eff < G at late time. However, in our case, this condition always holds, even in the early universe.) However, this would require a very careful examination of the structure formation process by studying the perturbation equation. Typically, the effective gravitational constant that governs structure formation may not be the same as the G eff that appears in the background Friedmann equation, which may also differ from the “bare” G. A concrete example was provided in [61]. Structure formation in the case of a fixed BEI was recently studied in [10]. In fact, a varying gravitational constant would cause many issues, and early time cosmology would need to be re-examined in close detail [62] (for example, the CMB angular power spectrum would be modified [46,62]).
In conclusion, the effects of a running BEI requires a deeper investigation. There is also a need to better understand how Barrow entropy can possibly arise from a theory of quantum gravity. Does the fractalized geometry only hold for horizons or any spacetime hypersurface in general? What about the spacetime manifold itself? Barrow entropy is not the first proposal that involves fractal geometry in gravity. It has previously been proposed that the spacetime dimension becomes fractalized and decreases towards the Planck scale [63,64,65,66,67,68], which is a different modification (a two-sphere in such a fractalized geometry has a lower dimension [63], not higher, as per Barrow’s proposal), but perhaps, one could obtain some ideas about how to derive Barrow entropy from these different approaches. Connections with fractional quantum mechanics [69] and other notions of entropies in relation to holographic dark energy [11,70,71,72] should also be further investigated.

Author Contributions

Conceptualization, Y.C.O.; validation, Y.C.O.; formal analysis, S.D.G. and Y.C.O.; investigation, S.D.G.; writing–original draft preparation, Y.C.O.; writing—review and editing, S.D.G.; supervision, Y.C.O.; funding acquisition, Y.C.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China grant number 11922508.

Acknowledgments

Y.C.O. thanks Brett McInnes for useful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

Notes

1
We remark that [25], on the contrary, incorporated the extra term in a new energy density ρ DE , which is attributed to dark energy, instead of modifying the gravitational constant. We can check that the two are equivalent by taking Equation (2.12) of [25] and substituting their expressions for Λ and β . The difference between the two expressions is given by the fact that [25] considered only the case k = 0 and kept the integration constant C, which in [14] was set to 0 or considered as part of the total energy density ρ . In our work, we use the form of [14]. As we shall see below, when allowing the BEI to run, we would have an extra term that can be interpreted as a dynamical dark energy, different from the identification in [25] for the fixed index case.
2
1 3 Λ eff ( r * ) Λ eff ( R ) d Λ eff = 1 3 [ Λ eff ( R ) Λ eff ( r * ) ] . This means that the observed value of the dark energy would be Λ eff ( r * ) Λ eff ( ) in the limit R . Without loss of generality, we can take Λ eff ( ) 0 , so that Λ eff ( r * ) is the observed value of the effective cosmological constant.
3
As per Footnote 4, the observed value of the effective cosmological constant over the interval ( ε , 2 ε ) is Λ eff ( ε ) Λ eff ( 2 ε ) < 0 . Since they are both comparable in magnitude (negative and with the same divergence properties), we can just refer to Λ ( ε ) as the effective cosmological constant at the Big Bang.
4
Indeed, a function f ( ε ) = ln ( π ε 2 ) e ε < 0 for small ε , satisfies f ( 2 ε ) f ( ε ) and f ( ε ) . Another example to illustrate the same phenomenon, albeit with a different divergence, is to consider f ( x ) = 1 / x . This function satisfies f ( 2 ε ) f ( ε ) and f ( ε ) = 1 / ε 2 . Colloquially, f climbs out from at the origin with an infinite positive slope.
5
Since H , there is still a Big Bang singularity in this model.

References

  1. Wang, B.; Gong, Y.; Abdalla, E. Thermodynamics of an Accelerated Expanding Universe. Phys. Rev. D 2006, 74, 083520. [Google Scholar] [CrossRef] [Green Version]
  2. Gong, Y.; Wang, A. Friedmann Equations and Thermodynamics of Apparent Horizons. Phys. Rev. Lett. 2007, 99, 211301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Barrow, J.D. The Area of a Rough Black Hole. Phys. Lett. B 2020, 808, 135643. [Google Scholar] [CrossRef]
  4. Anagnostopoulos, F.K.; Basilakos, S.; Saridakis, E.N. Observational Constraints on Barrow Holographic Dark Energy. Eur. Phys. J. C 2020, 80, 826. [Google Scholar] [CrossRef]
  5. Barrow, J.D.; Basilakos, S.; Saridakis, E.N. Big Bang Nucleosynthesis Constraints on Barrow Entropy. Phys. Lett. B 2021, 815, 136134. [Google Scholar] [CrossRef]
  6. Leon, G.; Magaña, J.; Hernàndez-Almada, A.; García-Aspeitia, M.A.; Verdugo, T.; Motta, V. Barrow Entropy Cosmology: An Observational Approach With a Hint of Stability Analysis. J. Cosmol. Astropart. Phys. 2021, 12, 032. [Google Scholar] [CrossRef]
  7. Luciano, G.G.; Saridakis, E.N. Baryon Asymmetry From Barrow Entropy: Theoretical Predictions and Observational Constraints. arXiv 2022, arXiv:2203.12010. [Google Scholar] [CrossRef]
  8. Jusufi, K.; Azreg-Aïnou, M.; Jamil, M.; Saridakis, E.N. Constraints on Barrow Entropy From M87* And S2 Star Observations. Universe 2022, 8, 102. [Google Scholar] [CrossRef]
  9. Vagnozzi, S.; Roy, R.; Tsai, Y.; Visinelli, L. Horizon-Scale Tests of Gravity Theories and Fundamental Physics From the Event Horizon Telescope Image of Sagittarius A*. arXiv 2022, arXiv:2205.07787. [Google Scholar]
  10. Farsi, B.; Sheykhi, A. Growth of Perturbations in Tsallis and Barrow Cosmology. arXiv 2022, arXiv:2205.04138. [Google Scholar]
  11. Nojiri, S.; Odintsov, S.D.; Paul, T. Barrow Entropic Dark Energy: A Member of Generalized Holographic Dark Energy Family. Phys. Lett. B 2022, 825, 136844. [Google Scholar] [CrossRef]
  12. Verde, L.; Treu, T.; Riess, A.G. Tensions Between the Early and Late Universe. Nat. Astron. 2019, 3, 891. [Google Scholar] [CrossRef]
  13. 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. Quantum Grav. 2021, 38, 153001. [Google Scholar] [CrossRef]
  14. Sheykhi, A. Barrow Entropy Corrections to Friedmann Equations. Phys. Rev. D 2021, 103, 123503. [Google Scholar] [CrossRef]
  15. Saridakis, E.N.; Basilakos, S. The Generalized Second Law of Thermodynamics With Barrow Entropy. Eur. Phys. J. C 2021, 81, 7. [Google Scholar] [CrossRef]
  16. al Mamon, A.; Paliathanasis, A.; Saha, S. Dynamics of an Interacting Barrow Holographic Dark Energy Model and its Thermodynamic Implications. Eur. Phys. J. Plus 2021, 136, 134. [Google Scholar] [CrossRef]
  17. Chen, C.; Chen, Y.; Ishibashi, A.; Ohta, N.; Yamaguchi, D. Running Newton Coupling, Scale Identification and Black Hole Thermodynamics. arXiv 2022, arXiv:2204.09892. [Google Scholar] [CrossRef]
  18. Bonanno, A.; Reuter, M. Renormalization Group Improved Black Hole Spacetimes. Phys. Rev. D 2000, 62, 043008. [Google Scholar] [CrossRef] [Green Version]
  19. Cai, Y.-F.; Easson, D.A. Asymptotically Safe Gravity as a Scalar-Tensor Theory and Its Cosmological Implications. Phys. Rev. D 2011, 84, 103502. [Google Scholar] [CrossRef] [Green Version]
  20. Gubitosi, G.; Ooijer, R.; Ripken, C. Frank Saueressig, Consistent Early and Late Time Cosmology From the RG Flow of Gravity. J. Cosmol. Astropart. Phys. 2018, 12, 004. [Google Scholar] [CrossRef] [Green Version]
  21. Page, D.N. Hawking Radiation and Black Hole Thermodynamics. New J. Phys. 2005, 7, 203. [Google Scholar] [CrossRef]
  22. Abreu, E.M.C.; Neto, J.A. Thermal Features of Barrow Corrected-Entropy Black Hole Formulation. Eur. Phys. J. C 2020, 80, 776. [Google Scholar] [CrossRef]
  23. Cai, R.-G.; Cao, L.-M.; Hu, Y.-P. Corrected Entropy-Area Relation and Modified Friedmann Equations. J. High Energy Phys. 2008, 8, 090. [Google Scholar] [CrossRef] [Green Version]
  24. Gour, G. Algebraic Approach to Quantum Black Holes: Logarithmic Corrections to Black Hole Entropy. Phys. Rev. D 2002, 66, 104022. [Google Scholar] [CrossRef] [Green Version]
  25. Saridakis, E.N. Modified Cosmology Through Spacetime Thermodynamics and Barrow Horizon Entropy. J. Cosmol. Astropart. Phys. 2020, 7, 031. [Google Scholar] [CrossRef]
  26. Wetterich, C. An Asymptotically Vanishing Time-Dependent Cosmological ‘Constant’. Astron. Astrophys. 1995, 301, 321. [Google Scholar]
  27. Wetterich, C. The Cosmological Constant and Higher Dimensional Dilatation Symmetry. Phys. Rev. D 2010, 81, 103507. [Google Scholar] [CrossRef] [Green Version]
  28. Wetterich, C. Graviton Fluctuations Erase the Cosmological Constant. Phys. Lett. B 2017, 773, 6. [Google Scholar] [CrossRef]
  29. Overduin, J.M.; Wesson, P.S.; Mashhoon, B. Decaying Dark Energy in Higher-Dimensional Gravity. Astron. Astrophys. 2007, 473, 727. [Google Scholar] [CrossRef] [Green Version]
  30. Mashhoon, B.; Wesson, P.S. Gauge-Dependent Cosmological ‘Constant’. Class. Quant. Grav. 2004, 21, 3611. [Google Scholar] [CrossRef] [Green Version]
  31. Babourova, O.; Frolov, B. The Solution of the Cosmological Constant Problem: The Cosmological Constant Exponential Decrease in the Super-Early Universe. Universe 2020, 6, 230. [Google Scholar] [CrossRef]
  32. Available online: https://link.springer.com/article/10.1140/epjc/s10052-020-7918-6 (accessed on 13 October 2022).
  33. Prokopec, T. Negative Energy Cosmology and the Cosmological Constant. arXiv 2011, arXiv:1105.0078. [Google Scholar]
  34. Biswas, T.; Koivisto, T.; Mazumdar, A. Could Our Universe Have Begun With Negative Lambda? arXiv 2011, arXiv:1105.2636. [Google Scholar]
  35. Ye, G.; Piao, Y.-S. Is the Hubble Tension a Hint of AdS Phase Around Recombination? Phys. Rev. D 2020, 101, 083507. [Google Scholar] [CrossRef] [Green Version]
  36. Ye, G.; Piao, Y.-S. T0 Censorship of Early Dark Energy and AdS Vacua. Phys. Rev. D 2020, 102, 083523. [Google Scholar] [CrossRef]
  37. Akarsu, O.; Barrow, J.D.; Escamilla, L.A.; Vazquez, J.A. Graduated Dark Energy: Observational Hints of a Spontaneous Sign Switch in the Cosmological Constant. Phys. Rev. D 2020, 101, 063528. [Google Scholar] [CrossRef] [Green Version]
  38. Gupt, B.; Singh, P. Non-singular AdS-DS Transitions in a Landscape Scenario. Phys. Rev. D 2014, 89, 063520. [Google Scholar] [CrossRef] [Green Version]
  39. Dutta, K.; Ruchika; Roy, A.; Sen, A.A.; Sheikh-Jabbari, M.M. Beyond ΛCDM with Low and High Redshift Data: Implications for Dark Energy. Gen. Relativ. Gravit. 2020, 52, 15. [Google Scholar] [CrossRef] [Green Version]
  40. Sen, A.A.; Adil, S.A.; Sen, S. Do Cosmological Observations Allow a Negative Λ? arXiv 2021, arXiv:2112.10641. [Google Scholar] [CrossRef]
  41. Akarsu, O.; Kumar, S.; Ozulker, E.; Vazquez, J.A. Relaxing Cosmological Tensions With a Sign Switching Cosmological Constant. Phys. Rev. D 2021, 104, 123512. [Google Scholar] [CrossRef]
  42. Battye, R.A.; Charnock, T.; Moss, A. Tension Between the Power Spectrum of Density Perturbations Measured on Large and Small Scales. Phys. Rev. D 2015, 91, 103508. [Google Scholar] [CrossRef] [Green Version]
  43. Benisty, D. Quantifying the S8 Tension With the Redshift Space Distortion Data Set. Phys. Dark Univ. 2021, 1, 100766. [Google Scholar] [CrossRef]
  44. Benevento, G.; Hu, W.; Raveri, M. Can Late Dark Energy Transitions Raise the Hubble Constant. Phys. Rev. D 2020, 101, 103517. [Google Scholar] [CrossRef]
  45. Heisenberg, L.; Villarrubia-Rojo, H.; Zosso, J. Simultaneously Solving the H0 and σ8 Tensions With Late Dark Energy. arXiv 2022, arXiv:2201.11623. [Google Scholar]
  46. Sakr, Z.; Sapone, D. Can Varying the Gravitational Constant Alleviate the Tensions? J. Cosmol. Astropart. Phys. 2022, 3, 034. [Google Scholar] [CrossRef]
  47. Visinelli, L.; Vagnozzi, S.; Danielsson, U. Revisiting a Negative Cosmological Constant From Low-Redshift Data. Symmetry 2019, 11, 1035. [Google Scholar] [CrossRef]
  48. Available online: https://journals.aps.org/prd/abstract/10.1103/PhysRevD.106.L041301 (accessed on 13 October 2022).
  49. Feynman, R.; Sands, M.; Leighton, R.B. The Feynmann Lectures on Physics; Addison Wesley Publishing Co.: Reading, MA, USA, 1963; pp. 46-1–46-9. [Google Scholar]
  50. Price, H. Cosmology, Time’s Arrow, and That Old Double Standard. In Time’s Arrows Today; Savitt, S., Ed.; Cambridge University Press: Cambridge, UK, 1994. [Google Scholar]
  51. Price, H. The Thermodynamic Arrow: Puzzles and Pseudo-puzzles. arXiv 2004, arXiv:physics/0402040. [Google Scholar]
  52. Vaas, R. Time After Time—Big Bang Cosmology and the Arrows of Time. Fundam. Theor. Phys. 2012, 172, 5. [Google Scholar]
  53. Price, H. Time’s Arrow and Eddington’s Challenge. Prog. Math. Phys. 2013, 63, 187. [Google Scholar]
  54. Penrose, R. Singularities and Time-Asymmetry. In General Relativity, An Einstein Centenary Survey; Hawking, S.T., Israel, W., Eds.; Cambridge University Press: Cambridge, UK, 1979. [Google Scholar]
  55. McInnes, B. Arrow of Time in String Theory. Nucl. Phys. B 2007, 782, 1. [Google Scholar] [CrossRef] [Green Version]
  56. Penrose, R. Before the Big Bang: An Outrageous New Perspective and Its Implications for Particle Physics. Contribution to: 10th European Particle Accelerator Conference (EPAC 06). Conf. Proc. C 2006, 060626, 2759. [Google Scholar]
  57. McInnes, B. The Arrow Of Time In The Landscape. arXiv 2007, arXiv:0711.1656. [Google Scholar]
  58. Carroll, S.M. In What Sense Is the Early Universe Fine-Tuned? arXiv 2007, arXiv:1406.3057. [Google Scholar]
  59. Greene, B.; Hinterbichler, K.; Judes, S.; Parikh, M.K. Smooth Initial Conditions From Weak Gravity. Phys. Lett. B 2011, 97, 178. [Google Scholar] [CrossRef] [Green Version]
  60. Sloan, D.; Ellis, G. Solving the Cosmological Entropy Issue with a Higgs Dilaton. Phys. Rev. D 2019, 99, 063518. [Google Scholar] [CrossRef] [Green Version]
  61. Maeda, K.-i.; Panpanich, S. Cuscuta-Galileon Cosmology: Dynamics, Gravitational Constants and the Hubble Constant. Phys. Rev. D 2022, 105, 104022. [Google Scholar] [CrossRef]
  62. Uzan, J.-P. Varying Constants, Gravitation and Cosmology. Living Rev. Rel. 2011, 14, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Benedetti, D. Fractal Properties of Quantum Spacetime. Phys. Rev. Lett. 2009, 102, 111303. [Google Scholar] [CrossRef] [Green Version]
  64. Carlip, S. The Small Scale Structure of Spacetime. arXiv 2010, arXiv:1009.1136. [Google Scholar]
  65. Alencar, G.; Bezerra, V.B.; Cunha, M.S.; Muniz, C.R. On Effective Spacetime Dimension in the Hořava-Lifshitz Gravity. Phys. Lett. B 2015, 747, 536. [Google Scholar] [CrossRef] [Green Version]
  66. Calcagni, G. ABC of Multi-Fractal Spacetimes and Fractional Sea Turtles. Eur. Phys. J. C 2016, 76, 181. [Google Scholar] [CrossRef] [Green Version]
  67. Lok Hu, B.L. Fractal Spacetimes in Stochastic Gravity?—Views from Anomalous Diffusion and the Correlation Hierarchy. J. Phys. Conf. Ser. 2017, 880, 012004. [Google Scholar]
  68. Carlip, S. Dimension and Dimensional Reduction in Quantum Gravity. Universe 2019, 5, 83. [Google Scholar] [CrossRef] [Green Version]
  69. Jalalzadeh, S.; da Silva, F.R.; Moniz, P.V. Prospecting Black Hole Thermodynamics With Fractional Quantum Mechanics. Eur. Phys. J. C 2021, 81, 632. [Google Scholar] [CrossRef]
  70. Nojiri, S.; Odintsov, S.D.; Paul, T. Different Faces of Generalized Holographic Dark Energy. arXiv 2021, arXiv:2105.08438. [Google Scholar] [CrossRef]
  71. Nojiri, S.; Odintsov, S.D.; Paul, T. Early and Late Universe Holographic Cosmology From a New Generalized Entropy. arXiv 2022, arXiv:2205.08876. [Google Scholar] [CrossRef]
  72. Nojiri, S.; Odintsov, S.D.; Faraoni, V. How Fundamental Is Entropy? From Non-extensive Statistics and Black Hole Physics to the Holographic Dark Universe. arXiv 2022, arXiv:2201.02424. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Di Gennaro, S.; Ong, Y.C. Sign Switching Dark Energy from a Running Barrow Entropy. Universe 2022, 8, 541. https://doi.org/10.3390/universe8100541

AMA Style

Di Gennaro S, Ong YC. Sign Switching Dark Energy from a Running Barrow Entropy. Universe. 2022; 8(10):541. https://doi.org/10.3390/universe8100541

Chicago/Turabian Style

Di Gennaro, Sofia, and Yen Chin Ong. 2022. "Sign Switching Dark Energy from a Running Barrow Entropy" Universe 8, no. 10: 541. https://doi.org/10.3390/universe8100541

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