Home Inflation and the principle of equivalence
Article Open Access

Inflation and the principle of equivalence

  • Fulvio Melia EMAIL logo
Published/Copyright: November 26, 2024

Abstract

A formal, mathematical statement of the principle of equivalence in general relativity is that one must always be able to find – at each location within a curved spacetime – the local free-falling frame against which one can measure the acceleration-induced time dilation and degree of curvature relative to flat spacetime. In this article, we use this theorem to prove that a de Sitter expansion, required during cosmic inflation, does not satisfy this condition and is therefore inconsistent with the PoE. To emphasize the importance – and reality – of this outcome, we contrast it with the analogous derivation for the Schwarzschild metric, which instead satisfies this requirement completely. We point out that this failure by de Sitter results from its incorrect handling of the Friedmann–Lemaître–Robertson–Walker (FLRW) lapse function, g tt . Our conclusion calls into question whether a period of inflated expansion could have even been possible in the context of FLRW cosmologies, and is noteworthy in light of recent, high-precision measurements showing that inflation could not have solved the temperature horizon problem while simultaneously producing the observed primordial power spectrum.

1 Introduction

Inflation has become an indispensable component of the standard model [1,2] (see also previous studies [3,4]). Its implied accelerated expansion in the early Universe may have solved several inconsistencies plaguing big bang cosmology and, perhaps more importantly, may have also seeded the primordial fluctuation spectrum [5] that eventually led to the formation of large-scale structure.

In spite of the considerable effort expended in finding its underlying field and potential, however, it still lacks true predictive power. In addition, very little attention has been paid to its fundamental basis within the context of general relativity (GR), the theory that spawned it via one of the most influential solutions to Einstein’s equations.

In this article, we begin to address this deficiency by examining the consistency of such an accelerated expansion with the foundational ingredient in Einstein’s theory, i.e., the principle of equivalence (PoE), which requires that there exist at every spacetime point a local free-falling (inertial) frame relative to which one may measure the acceleration in the comoving frame. We shall demonstrate that the adoption of the Friedmann–Lemaître–Robertson–Walker (FLRW) metric – with its constrained lapse function g tt = 1 – to describe the inflationary spacetime is actually inconsistent with the requirements of the PoE. To emphasize de Sitter’s failure in this regard, we contrast it with the analogous derivation for the Schwarzschild metric, which instead completely satisfies the PoE self-consistently. We end with a brief discussion of the consequences of this conclusion.

2 PoE

We define x μ = ( c t , x 1 , x 2 , x 3 ) to be the coordinates in the comoving frame, g μ ν the metric coefficients, and Γ λ μ ν the corresponding Christoffel symbols. Most of the observations today appear to be telling us that the Universe is spatially flat [6], so we assume the spatial curvature constant, k , to be zero in the FLRW metric, which may therefore simply be written as

(1) d s 2 = c 2 d t 2 a ( t ) 2 [ ( d x 1 ) 2 + ( d x 2 ) 2 + ( d x 3 ) 2 ] ,

in terms of the expansion factor a ( t ) . In the following, we shall adopt its de Sitter form relevant during inflation, i.e., with the inclusion of

(2) a ( t ) = e H t ,

where the Hubble parameter H is constant. The metric coefficients are thus

(3) g μ ν = 1 0 0 0 0 a 2 0 0 0 0 a 2 0 0 0 0 a 2 ,

and one can easily show from this that the non-zero Christoffel symbols in the comoving frame must be

(4) Γ i i 0 = 1 c a a ˙ = H c e 2 H t , Γ i 0 i = Γ 0 i i = 1 c a ˙ a = H c .

A formal, mathematical expression of the PoE holds that there must exist at every spacetime point x 0 μ in x μ a local free-falling (inertial) frame ξ μ ( x ) that one may use to “measure” the spacetime curvature in x μ [7]. For convenience and without loss of generality, we shall adopt the cosmological principle, in which the Universe is homogeneous and isotropic (at least on scales larger than 300 Mpc today), and arrange the coordinates x μ and ξ μ to have a common origin.

In order for ξ μ to fulfill the role played by the local, free-falling frame, they must satisfy the equations (see Eq. 3.2.11 in Weinberg’s book [7])

(5) 2 ξ α x μ x ν = Γ λ μ ν ξ α x λ .

To identify these coordinates, we shall follow a procedure analogous to the transformation used to cast the Schwarzschild metric into its so-called Cartesian isotropic form [8], which we shall revisit shortly in the following section.

Of course, while velocity is “relative” in GR, acceleration is “absolute,” meaning that if one can measure an acceleration with respect to one inertial frame, the same acceleration would emerge relative to any other inertial frame. So the point of finding the coordinates ξ α is merely to identify at least one such inertial frame to test whether the metric (in this case Eq. (1)) is consistent with Eq. (5). The goal is thus to find a transformation from the frame represented by the coordinates x μ into a local frame falling freely, whose metric has a Minkowski form (represented by the ξ α coordinates), as we shall see shortly in Eq. (20). This procedure is facilitated by the recognition that de Sitter space is one of the six (and exactly six) applications of the FLRW metric corresponding to a constant spacetime curvature, i.e., a spacetime whose metric coefficients may be written independently of time [9,10].

The fact that de Sitter space has constant spacetime curvature is not obvious from Eq. (1), but we can demonstrate this using a transformation (based on fixed observer coordinates) that renders its metric coefficients independent of time [11]. To see this, we define

(6) c T = c t 1 2 R h ln Φ , R = a ( t ) r ,

where r 2 = ( x 1 ) 2 + ( x 2 ) 2 + ( x 3 ) 2 and the gravitational (or Hubble) radius is defined as

(7) R h c H .

In addition,

(8) Φ 1 R R h 2 .

With this transformation, the de Sitter metric may be written

(9) d s 2 = Φ c 2 d T 2 Φ 1 d R 2 R 2 d Ω 2 ,

where d Ω 2 d θ 2 + sin 2 θ d ϕ 2 . As we can easily see, all the metric coefficients in this equation are independent of the new time coordinate T and, for future reference, we define the Cartesian coordinates ( X 1 , X 2 , X 3 ) , such that

(10) X 1 = R sin θ cos ϕ , X 2 = R sin θ sin ϕ , X 3 = R cos θ .

Eq. (9), however, is not yet in Minkowki form, but we can now easily find a transformation from the coordinates ( c T , R , θ , ϕ ) into a free-falling frame. We do this by following the approach of finding the metric’s Cartesian isotropic form, in which we put

(11) R = σ 1 + σ σ h 2 1 ,

where

(12) σ h 2 R h .

We thus have

(13) d R = d σ P Q 2

and

(14) Φ = P 2 Q 2 ,

in terms of the quantities

(15) P 1 σ σ h 2 , Q 1 + σ σ h 2 .

The newly transformed de Sitter interval may thus be written

(16) d s 2 = P Q 2 c 2 d T 2 1 Q 2 ( d σ 2 + σ 2 d Ω 2 )

and, if we define the Cartesian coordinates corresponding to σ as

(17) χ 1 = σ sin θ cos ϕ , χ 2 = σ sin θ sin ϕ , χ 3 = σ cos θ ,

we arrive at its Cartesian isotropic form,

(18) d s 2 = P Q 2 c 2 d T 2 1 Q 2 [ ( d χ 1 ) 2 + ( d χ 2 ) 2 + ( d χ 3 ) 2 ] .

The de Sitter metric written in this way allows us to see immediately what the local free-falling frame written in Cartesian coordinates looks like.

As we now apply the requirements of the PoE to the metric in Eq. (18), we invoke the condition that the local inertial frame needs to be defined only in the vicinity of x 0 μ (i.e., the local inertial frame at one such point need not be the same as those elsewhere in the spacetime), so we may put R constant and σ constant wherever they appear in the metric coefficients. In other words, we may assume that P P ( x 0 μ ) and Q Q ( x 0 μ ) near x 0 μ , which finally allows us to identify the local inertial frame coordinates:

(19) ξ 0 = P ( x 0 μ ) Q ( x 0 μ ) c T , ξ 1 = 1 Q ( x 0 μ ) χ 1 , ξ 2 = 1 Q ( x 0 μ ) χ 2 , ξ 3 = 1 Q ( x 0 μ ) χ 3 ,

i.e.,

(20) d s 2 = ( d ξ 0 ) 2 ( d ξ 1 ) 2 ( d ξ 2 ) 2 ( d ξ 3 ) 2

at any given spacetime point x 0 μ .

It is now a simple exercise to examine whether the coordinates x μ and ξ μ satisfy Eq. (5). As we shall see, they do not. For example, one finds for the α = μ = ν = 0 component that

(21) P Q 2 R h R R h 2 Φ 2 = 0 ,

but this cannot be consistent for arbitrary values of R (or r ). Similarly, one finds for α = μ = ν = i that

(22) 0 = e 2 H t R h Q P X i R h ,

which, again, is inconsistent for arbitrary values of X i . A de Sitter expansion, such as that required during inflation, therefore does not satisfy the requirements of the PoE.

Having said this, a possible caveat to this conclusion could have been that an analysis of the cosmic microwave background (CMB) anisotropies clearly shows a fluctuation spectrum that is not exactly scale free (for which the scalar spectral index would then have been exactly 1). Its exact value is instead n s = 0.9649 ± 0.0042 [6], the slight difference implying some version of a slow-roll potential [5,12], so that the Hubble parameter H during inflation could not have remained perfectly constant. The expansion during inflation is therefore not perfectly de Sitter and would have proceeded almost according to Eq. (2), but not exactly. This slight difference, however, would have had only a minimal impact, if any, on the derivations in this section, introducing essentially only small variations in the constants P and Q in Eqs (19), (21), and (22). But these changes could not prevent the inconsistencies shown in the latter two expressions, which would still require R and/or r to be zero everywhere.

A better way to understand this is that the outcome discussed here emerges for all the FLRW metrics with an equation of state different from ρ + 3 p = 0 , not just for the case being considered here (i.e., ρ + p = 0 ), as formally demonstrated in the study of Melia [13]. Such previous publications have established the problem highlighted here with a forced lapse function g t t = 1 in a more general context, so even a variable H would not change the outcome of this discussion. The benefit of the streamlined approach we are using here is that the Christoffel symbols (Eq. 4) for a constant H are so simple that one can show the result quite robustly with just straightforward analytic derivations. In other words, the problem with the use of de Sitter in the FLRW metric is not the constant H , but the fact that inflation is necessarily an accelerated expansion, which conflicts with the zero time dilation implied by g t t = 1 . The conclusions drawn in this section do not rely on the specific choice of ρ + p = 0 for the equation of state, but are most easily demonstrated analytically for this case.

3 Comparison with the Schwarzschild metric

Before we discuss this crucial result, let us – for comparison and to emphasize the importance of the PoE in this analysis – carry out an analogous derivation for the Schwarzschild metric and show that the outcome in the latter case is completely different. In the next section, we shall point out that a distinguishing feature between these two spacetimes is the nature of their lapse function, g tt .

The Schwarzschild metric describes the (vacuum) spacetime external to a spherically symmetric mass M and is typically written in its standard form

(23) d s 2 = 1 r S r c 2 d t 2 1 r S r 1 d r 2 r 2 d Ω 2 ,

in terms of spherical coordinates ( r , θ , ϕ ) and the Schwarzschild radius r S 2 G M c 2 .

The most direct way of finding the local free-falling frame ξ s μ is to introduce a new radial coordinate, ϖ , such that

(24) r = ϖ 1 + ϖ S ϖ 2 ,

where ϖ S r S 4 . It is not difficult to show that

(25) d r = 1 ϖ S ϖ 1 + ϖ S ϖ d ϖ

and

(26) 1 r S r = 1 ϖ S ϖ 2 1 + ϖ S ϖ 2 .

Substituting these into Eq. (23), one therefore obtains

(27) d s 2 = 1 ϖ S ϖ 2 1 + ϖ S ϖ 2 c 2 d t 2 1 + ϖ S ϖ 4 × [ ( d y 1 ) 2 + ( d y 2 ) 2 + ( d y 3 ) 2 ] ,

where ϖ 2 ( y 1 ) 2 + ( y 2 ) 2 + ( y 3 ) 2 , i.e.,

(28) y 1 = ϖ sin θ cos ϕ , y 2 = ϖ sin θ sin ϕ , y 3 = ϖ cos θ ,

and

(29) ( d y 1 ) 2 + ( d y 2 ) 2 + ( d y 3 ) 2 = d ϖ 2 + ϖ 2 d Ω 2 .

Eq. (27) is the Schwarzschild metric written in terms of Cartesian isotropic coordinates [8].

In the vicinity of any given point x 0 μ = ( c t , r 0 , θ 0 , ϕ 0 ) , the r and ϖ coordinates within the metric coefficients of Eq. (27) are approximately constant. We define

(30) U 1 ϖ S ϖ 1 + ϖ S ϖ 1

and

(31) V 1 + ϖ S ϖ 2 ,

which are therefore also both approximately constant at x 0 μ , and thus,

(32) d s 2 U 2 c 2 d t 2 V 2 ( d ϖ 2 + ϖ 2 d Ω 2 ) .

So we may define another set of coordinates

(33) ξ s 0 U c t , ξ s 1 V ϖ sin θ cos ϕ , ξ s 2 V ϖ sin θ sin ϕ , ξ s 3 V ϖ cos θ .

These allow us to write the Schwarzschild interval as

(34) d s 2 = ( d ξ s 0 ) 2 ( d ξ s 1 ) 2 ( d ξ s 2 ) 2 ( d ξ s 3 ) 2 ,

the familiar Minkowski form in Cartesian coordinates, which therefore represents the metric as seen within the local free-falling frame at x 0 μ , given in terms of the coordinates ξ s μ = ( ξ s 0 , ξ s 1 , ξ s 2 , ξ s 3 ) . To be clear, this frame is the local free-falling frame only at that point, x 0 μ . It will be different elsewhere. Indeed, a sequence of these frames derived at progressively smaller r would reveal the infalling trajectory toward the origin, characterized by a variation in U and V with radius consistent with the free-fall velocity at each point.

It is now straightforward to show that the coordinates x μ = ( c t , r , θ , ϕ ) in Eq. (23) and ξ s μ in Eq. (33) satisfy Eq. (5), or the equivalent expression

(35) g μ ν = ξ s α x μ ξ s β x ν η α β ,

where η α β = diag ( 1 , 1 , 1 , 1 ) is the corresponding metric tensor in flat spacetime. To do this, we need the Christoffel symbols corresponding to the metric in Eq. (23), whose non-zero components are simply

(36) Γ r t t = 1 2 r S r 2 1 r S r Γ r r r = 1 2 r S r 2 1 r S r 1 , Γ t t r = 1 2 r S r 2 1 r S r 1 = Γ t r t , Γ θ r θ = 1 r = Γ θ θ r , Γ r θ θ = r 1 r S r , Γ ϕ r ϕ = 1 r = Γ ϕ ϕ r , Γ r ϕ ϕ = r sin 2 θ 1 r S r , Γ θ ϕ ϕ = sin θ cos θ , Γ ϕ θ ϕ = cot θ = Γ ϕ ϕ θ .

For example, in the α = μ = ν = 1 component of Eq. (5), one has

(37) ξ s 1 r 1 + ϖ S ϖ 1 ϖ S ϖ 1 sin θ cos ϕ ,

which yields

(38) 2 ξ s 1 r 2 1 2 r S r 2 1 r S r 3 2 sin θ cos ϕ .

By comparison,

(39) Γ λ 11 ξ s 1 x λ = 1 2 r S r 2 1 r S r 1 ξ s 1 r ,

which becomes

(40) Γ λ 11 ξ s 1 x λ = 1 2 r S r 2 1 r S r 3 2 sin θ cos ϕ ,

matching Eq. (38). Repeating this exercise for any of the other components in Eq. (5) produces a similarly matching result.

This level of consistency may also be demonstrated via Eq. (35). For example, the g 11 metric coefficient may be calculated as follows:

(41) g 11 = ξ s α r ξ s β r η α β ,

which yields

(42) g 11 = 1 + ϖ S ϖ 4 ϖ r 2 .

And using Eqs (25) and (26), one finally shows that

(43) g 11 = 1 r S r 1 ,

as it appears in Eq. (23).

So unlike the situation we found with de Sitter in Section 2, the Schwarzschild metric is completely consistent with the PoE, based on the fact that one can find – at any location x 0 μ within this spacetime – the local, free-falling frame ξ s μ , relative to which one may measure the spacetime curvature associated with the metric for the accelerated observer in Eq. (23).

4 Discussion

The obvious question is why the well-studied de Sitter solution to Einstein’s equations should be inconsistent with the PoE, which is, after all, the basis for the field equations themselves. Such an issue actually falls within the broader context of whether the simplifications used to derive the FLRW metric in the first place remain valid when one subsequently chooses an equation of state for the cosmic fluid in order to uniquely calculate the expansion factor a ( t ) . As discussed in previous publications, the short answer is generally no, except for one particular case corresponding to the zero active mass condition in GR, i.e., ρ + 3 p = 0 [13,14]. Thus, choosing an equation of state ρ + p = 0 , producing a de Sitter expansion, or slow-roll conditions producing an expansion very close to de Sitter, appears to be inconsistent with the free-fall requirements associated with g tt = 1 , given that both of these situations produce an accelerated Hubble frame. Ultimately, this is why we see the evident, unavoidable failure resulting in Eqs (21) and (22).

The more general examination of the validity of the FLRW metric in the context of the PoE has revealed that such contradictory results disappear when the metric coefficient g tt correctly accounts for the time dilation resulting from a curved spacetime, such as one finds in the Schwarzschild case. This spacetime includes the lapse function g tt = 1 r S r and, not surprisingly, is therefore able to fully satisfy the requirements of the PoE.

But in this article, we have not needed to address such deeper considerations. We have merely used the conventional form of the FLRW metric applied to inflationary cosmology. The PoE expressed in the form of Eq. (5) derives from one of the most fundamental tenets in GR, notably that spacetime curvature must always be measurable relative to the local free-falling frame. And we cannot avoid the conclusion that Eq. (1), with the expansion factor in Eq. (2), is simply inconsistent with this condition.

What are we to make of this result vis-à-vis the observations? Inflationary cosmology has been under development for over 40 years, and the observations have not yet ruled it out (or confirmed it). But therein lies part of the problem. Inflation is frustratingly flexible, easily adjustable to match new data – at least globally, if not in fine detail. Note, for example, how poorly its underlying field and potential are understood, even after four decades of development (see, e.g., a recent review in the study of Vazquez et al. [15]). Recent advances, however, have begun to pose some serious questions. For example, though inflation was introduced largely to fix the temperature horizon problem in the cosmic microwave background, it turns out that there are other, equally critical, horizon problems in cosmology, such as that associated with the electroweak phase transition at t 1 0 11  s [16]. So inflation is starting to look more like a customized process designed to address primarily one aspect of the early Universe, rather than a foundation for everything else that followed.

This is not to say that inflation no longer provides some ingredients necessary to establish the internal self-consistency of the standard model. A realistic discussion of its merits ought to include both the positives (see, e.g., the study of Mukhanov [5]) and its negatives (see, e.g., previous studies [17,18]), constituting an extensive discussion outside the scope of this article [19]. But to highlight some recent observational developments that should motivate the type of discussion we have presented in this article, we point to the detailed analysis of the latest Planck data release, which suggests that standard, slow-roll inflation could not have solved the temperature horizon problem while simultaneously producing the observed primordial power spectrum. Evidence is growing that the fluctuation spectrum seen in the CMB has a k min cutoff large enough to impact the time at which inflation could have started [2022], which, in turn, would have constrained the number of e-folds experienced by the Universe during its inflated expansion.

This implied delayed initiation to inflation ironically weakens an earlier argument that this rapid expansion would have violated the PoE. If inflation had started within the Planck domain, where quantum gravity effects were unavoidable, one would necessarily have had to include the existence of a minimal, measurable length of the order of the Planck length in the Hamiltonian, which would have violated the PoE [23]. In the post Planck satellite era, however, the energy of inflation is constrained to be well below the Planck energy and the “old” version of chaotic inflation has had to give way to a newer replacement with a much delayed initiation well past the Planck time [17,18]. This required delayed beginning to any de Sitter expansion would have postponed it well past a quantum beginning [24], so even this originally valid criticism could in principle be avoided now.

Nevertheless, in this “new” chaotic inflationary paradigm, the measured value of k min suggests that the inflation-induced expansion would have missed solving the temperature horizon problem by a factor 10 [12]. It is too early to tell whether such a result impacts all conceivable slow-roll inflationary potentials, or whether it creates a hurdle primarily for those proposed thus far. But the fact that inflationary theory has not yet converged to a unified, self-consistent picture after four decades of work may just be an indication that it could not have happened. In this context, the theoretical argument presented in this article does not stand in isolation. It would provide a foundation for the growing tension seen between the expectations of inflation and the new, high-precision measurements becoming available now.

5 Conclusion

Fortunately, this discussion will benefit from an influx of concrete evidence – one way or the other – in the coming years. Today, the standard model is faced with many inconsistencies and deficiencies, not solely the unresolved temperature horizon problem (see, e.g., the study of Melia [19]). The campaign to measure the real-time redshift drift of distant quasars (see, e.g.., the study of Melia [25]) is a particularly exciting prospect, because it will probe the underlying cosmology in a truly model-independent way, and its principal outcome will be a simple yes/no answer. The ELT-HIRES [26] and the SKA Phase 2 array [27] will attempt to measure the redshift derivatives of distant sources. A measurement of zero drift at all redshifts, contrasting with any variation at all, would argue against the use of the FLRW metric to describe the spacetime of models with an accelerated expansion, such as inflationary cosmology. These observational campaigns will be able to differentiate between zero and non-zero redshift drift at a confidence level of 3 σ after only 5 years of observation, and 5 σ over a lifetime of 20 years.

A determination that the contents of the Universe exhibit zero redshift drift would point to a cosmic expansion consistent with the zero active mass condition in GR [10,14], which produces an expansion factor a ( t ) = ( t t 0 ) . Such a Universe does not have any horizon problems [16,28], so the inconsistency we have highlighted in this article may simply become moot if it turns out that the Universe did not experience a phase of inflated expansion after all.

This is but one of several possible alternative scenarios in which the early Universe would not have needed to inflate in order to produce the expansion we see today. Other than an adherence to the zero active mass condition in GR, the Universe may have undergone a (perhaps infinite) sequence of cyclic bounces [29,30], avoiding all horizon problems due to an equilibration before each cycle. Arguments such as those we have made in this article can help to refine the discussion more consistently with fundamental theory.

Acknowledgments

I am grateful to Amherst College for its support through a John Woodruff Simpson Lectureship. I am also grateful to the anonymous referees for comments and suggestions that have led to an improvement in the presentation of this manuscript.

  1. Funding information: The author states no funding involved.

  2. Author contribution: The author has accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: The author states no conflict of interest.

References

[1] Guth AH. Inflationary universe: A possible solution to the horizon and flatness problems. PRD. 1981 Jan;23(2):347–56. 10.1103/PhysRevD.23.347Search in Google Scholar

[2] Linde AD. A new inflationary universe scenario: A possible solution of the horizon, flatness, homogeneity, isotropy and primordial monopole problems. Phys Lett B. 1982 Feb;108(6):389–93. 10.1016/0370-2693(82)91219-9Search in Google Scholar

[3] Starobinsky AA. Spectrum of relict gravitational radiation and the early state of the universe. ZhETF Pisma Redaktsiiu. 1979 Dec;30:719–23. Search in Google Scholar

[4] Kazanas D. Dynamics of the universe and spontaneous symmetry breaking. ApJ Letters. 1980 Oct;241:L59–63. 10.1086/183361Search in Google Scholar

[5] Mukhanov VF. Physical foundations of cosmology. Cambridge: Cambridge University Press; 2005. 10.1017/CBO9780511790553Search in Google Scholar

[6] Planck Collaboration, Aghanim N, Akrami Y, Ashdown M, Aumont J, Baccigalupi C, et al. Planck 2018 results. VI. Cosmological parameters. A&A. 2020 Sep;641:A6. Search in Google Scholar

[7] Weinberg S. Gravitation and cosmology: principles and applications of the general theory of relativity. New York: John Wiley and Sons; 1972. Search in Google Scholar

[8] Misner CW, Thorne KS, Wheeler JA. Gravitation. New York: Princeton University Press; 2017. Search in Google Scholar

[9] Florides PS. The Robertson-Walker metrics expressible in static form. General Relativity Gravitation. 1980 Jul;12(7):563–74. 10.1007/BF00756530Search in Google Scholar

[10] Melia F. Cosmological redshift in Friedmann-Robertson-Walker metrics with constant space-time curvature. MNRAS. 2012 May;422(2):1418–24. 10.1111/j.1365-2966.2012.20714.xSearch in Google Scholar

[11] Melia F. Fitting the Union2.1 Supernova Sample with the Rh = ct Universe. Astron J. 2012 Oct;144(4):110. 10.1088/0004-6256/144/4/110Search in Google Scholar

[12] Liu T, Cao S, Zhang J, Biesiada M, Liu Y, Lian Y. Testing the cosmic curvature at high redshifts: the combination of LSST strong lensing systems and quasars as new standard candles. MNRAS. 2020 Jul;496(1):708–17. 10.1093/mnras/staa1539Search in Google Scholar

[13] Melia F. The lapse function in Friedmann-Lemaiiiitre-Robertson-Walker cosmologies. Ann Phys. 2019 Dec;411:167997. 10.1016/j.aop.2019.167997Search in Google Scholar

[14] Melia F. The cosmic spacetime. Oxford: Taylor and Francis; 2020. 10.1201/9781003081029Search in Google Scholar

[15] Vazquez JA, Padilla LE, Matos T. Inflationary cosmology: from theory to observations. 2018 Oct. arXiv:1810.09934. Search in Google Scholar

[16] Melia F. A solution to the electroweak horizon problem in the Rh=ct universe. Europ Phys J C. 2018 Sep;78(9):739. 10.1140/epjc/s10052-018-6231-0Search in Google Scholar

[17] Ijjas A, Steinhardt PJ, Loeb A. Inflationary paradigm in trouble after Planck2013. Phys Lett B. 2013 Jun;723(4–5):261–6. 10.1016/j.physletb.2013.05.023Search in Google Scholar

[18] Ijjas A, Steinhardt PJ, Loeb A. Inflationary schism. Phys Lett B. 2014 Sep;736:142–6. 10.1016/j.physletb.2014.07.012Search in Google Scholar

[19] Melia F. A candid assessment of standard cosmology. Pub Astron Soc Pacific. 2022 Dec;134:121001. 10.1088/1538-3873/aca51fSearch in Google Scholar

[20] Melia F. The apparent (gravitational) horizon in cosmology. Am J Phys. 2018 Aug;86(8):585–93. 10.1119/1.5045333Search in Google Scholar

[21] Melia F, Ma Q, Wei JJ, Yu B. Hint of a truncated primordial spectrum from the CMB large-scale anomalies. A&A. 2021 Nov;655:A70. 10.1051/0004-6361/202141251Search in Google Scholar

[22] Sanchis-Lozano MA, Melia F, López-Corredoira M, Sanchis-Gual N. Missing large-angle correlations versus even-odd point-parity imbalance in the cosmic microwave background. A&A. 2022 Apr;660:A121. 10.1051/0004-6361/202142296Search in Google Scholar

[23] Gnatenko KP, Tkachuk VM. Weak equivalence principle in quantum space. Front Astron Space Sci. 2022 Aug;9:950468. 10.3389/fspas.2022.950468Search in Google Scholar

[24] Liu J, Melia F. Challenges to inflation in the post-Planck era. ApJ Suppl. 2024 Jun;967(2):109. 10.3847/1538-4357/ad4036Search in Google Scholar

[25] Melia F. Definitive test of the Rh = ct universe using redshift drift. MNRAS. 2016 Nov;463(1):L61–3. 10.1093/mnrasl/slw157Search in Google Scholar

[26] Liske J. Status of the European extremely large telescope. In: Dickinson M, Inami H, editors. Thirty meter telescope science forum. The American Astronomical Society; 2014. p. 52. Search in Google Scholar

[27] Kloeckner HR, Obreschkow D, Martins C, Raccanelli A, Champion D, Roy AL, et al. Real time cosmology - A direct measure of the expansion rate of the Universe with the SKA. In: Advancing Astrophysics with the Square Kilometre Array (AASKA14). Sissa Medialab srl Partita IVA; 2015. p. 27. 10.22323/1.215.0027Search in Google Scholar

[28] Melia F. The Rh = ct universe without inflation. A&A. 2013 May;553:A76. 10.1051/0004-6361/201220447Search in Google Scholar

[29] Ijjas A, Steinhardt PJ. A new kind of cyclic universe. Phys Lett B. 2019 Aug;795:666–72. 10.1016/j.physletb.2019.06.056Search in Google Scholar

[30] Vavryčuk V. Cosmological redshift and cosmic time dilation in the FLRW metric. Front Phys. 2022 May;10:826188. 10.3389/fphy.2022.826188Search in Google Scholar

Received: 2024-01-05
Revised: 2024-06-06
Accepted: 2024-08-03
Published Online: 2024-11-26

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Numerical study of flow and heat transfer in the channel of panel-type radiator with semi-detached inclined trapezoidal wing vortex generators
  3. Homogeneous–heterogeneous reactions in the colloidal investigation of Casson fluid
  4. High-speed mid-infrared Mach–Zehnder electro-optical modulators in lithium niobate thin film on sapphire
  5. Numerical analysis of dengue transmission model using Caputo–Fabrizio fractional derivative
  6. Mononuclear nanofluids undergoing convective heating across a stretching sheet and undergoing MHD flow in three dimensions: Potential industrial applications
  7. Heat transfer characteristics of cobalt ferrite nanoparticles scattered in sodium alginate-based non-Newtonian nanofluid over a stretching/shrinking horizontal plane surface
  8. The electrically conducting water-based nanofluid flow containing titanium and aluminum alloys over a rotating disk surface with nonlinear thermal radiation: A numerical analysis
  9. Growth, characterization, and anti-bacterial activity of l-methionine supplemented with sulphamic acid single crystals
  10. A numerical analysis of the blood-based Casson hybrid nanofluid flow past a convectively heated surface embedded in a porous medium
  11. Optoelectronic–thermomagnetic effect of a microelongated non-local rotating semiconductor heated by pulsed laser with varying thermal conductivity
  12. Thermal proficiency of magnetized and radiative cross-ternary hybrid nanofluid flow induced by a vertical cylinder
  13. Enhanced heat transfer and fluid motion in 3D nanofluid with anisotropic slip and magnetic field
  14. Numerical analysis of thermophoretic particle deposition on 3D Casson nanofluid: Artificial neural networks-based Levenberg–Marquardt algorithm
  15. Analyzing fuzzy fractional Degasperis–Procesi and Camassa–Holm equations with the Atangana–Baleanu operator
  16. Bayesian estimation of equipment reliability with normal-type life distribution based on multiple batch tests
  17. Chaotic control problem of BEC system based on Hartree–Fock mean field theory
  18. Optimized framework numerical solution for swirling hybrid nanofluid flow with silver/gold nanoparticles on a stretching cylinder with heat source/sink and reactive agents
  19. Stability analysis and numerical results for some schemes discretising 2D nonconstant coefficient advection–diffusion equations
  20. Convective flow of a magnetohydrodynamic second-grade fluid past a stretching surface with Cattaneo–Christov heat and mass flux model
  21. Analysis of the heat transfer enhancement in water-based micropolar hybrid nanofluid flow over a vertical flat surface
  22. Microscopic seepage simulation of gas and water in shale pores and slits based on VOF
  23. Model of conversion of flow from confined to unconfined aquifers with stochastic approach
  24. Study of fractional variable-order lymphatic filariasis infection model
  25. Soliton, quasi-soliton, and their interaction solutions of a nonlinear (2 + 1)-dimensional ZK–mZK–BBM equation for gravity waves
  26. Application of conserved quantities using the formal Lagrangian of a nonlinear integro partial differential equation through optimal system of one-dimensional subalgebras in physics and engineering
  27. Nonlinear fractional-order differential equations: New closed-form traveling-wave solutions
  28. Sixth-kind Chebyshev polynomials technique to numerically treat the dissipative viscoelastic fluid flow in the rheology of Cattaneo–Christov model
  29. Some transforms, Riemann–Liouville fractional operators, and applications of newly extended M–L (p, s, k) function
  30. Magnetohydrodynamic water-based hybrid nanofluid flow comprising diamond and copper nanoparticles on a stretching sheet with slips constraints
  31. Super-resolution reconstruction method of the optical synthetic aperture image using generative adversarial network
  32. A two-stage framework for predicting the remaining useful life of bearings
  33. Influence of variable fluid properties on mixed convective Darcy–Forchheimer flow relation over a surface with Soret and Dufour spectacle
  34. Inclined surface mixed convection flow of viscous fluid with porous medium and Soret effects
  35. Exact solutions to vorticity of the fractional nonuniform Poiseuille flows
  36. In silico modified UV spectrophotometric approaches to resolve overlapped spectra for quality control of rosuvastatin and teneligliptin formulation
  37. Numerical simulations for fractional Hirota–Satsuma coupled Korteweg–de Vries systems
  38. Substituent effect on the electronic and optical properties of newly designed pyrrole derivatives using density functional theory
  39. A comparative analysis of shielding effectiveness in glass and concrete containers
  40. Numerical analysis of the MHD Williamson nanofluid flow over a nonlinear stretching sheet through a Darcy porous medium: Modeling and simulation
  41. Analytical and numerical investigation for viscoelastic fluid with heat transfer analysis during rollover-web coating phenomena
  42. Influence of variable viscosity on existing sheet thickness in the calendering of non-isothermal viscoelastic materials
  43. Analysis of nonlinear fractional-order Fisher equation using two reliable techniques
  44. Comparison of plan quality and robustness using VMAT and IMRT for breast cancer
  45. Radiative nanofluid flow over a slender stretching Riga plate under the impact of exponential heat source/sink
  46. Numerical investigation of acoustic streaming vortices in cylindrical tube arrays
  47. Numerical study of blood-based MHD tangent hyperbolic hybrid nanofluid flow over a permeable stretching sheet with variable thermal conductivity and cross-diffusion
  48. Fractional view analytical analysis of generalized regularized long wave equation
  49. Dynamic simulation of non-Newtonian boundary layer flow: An enhanced exponential time integrator approach with spatially and temporally variable heat sources
  50. Inclined magnetized infinite shear rate viscosity of non-Newtonian tetra hybrid nanofluid in stenosed artery with non-uniform heat sink/source
  51. Estimation of monotone α-quantile of past lifetime function with application
  52. Numerical simulation for the slip impacts on the radiative nanofluid flow over a stretched surface with nonuniform heat generation and viscous dissipation
  53. Study of fractional telegraph equation via Shehu homotopy perturbation method
  54. An investigation into the impact of thermal radiation and chemical reactions on the flow through porous media of a Casson hybrid nanofluid including unstable mixed convection with stretched sheet in the presence of thermophoresis and Brownian motion
  55. Establishing breather and N-soliton solutions for conformable Klein–Gordon equation
  56. An electro-optic half subtractor from a silicon-based hybrid surface plasmon polariton waveguide
  57. CFD analysis of particle shape and Reynolds number on heat transfer characteristics of nanofluid in heated tube
  58. Abundant exact traveling wave solutions and modulation instability analysis to the generalized Hirota–Satsuma–Ito equation
  59. A short report on a probability-based interpretation of quantum mechanics
  60. Study on cavitation and pulsation characteristics of a novel rotor-radial groove hydrodynamic cavitation reactor
  61. Optimizing heat transport in a permeable cavity with an isothermal solid block: Influence of nanoparticles volume fraction and wall velocity ratio
  62. Linear instability of the vertical throughflow in a porous layer saturated by a power-law fluid with variable gravity effect
  63. Thermal analysis of generalized Cattaneo–Christov theories in Burgers nanofluid in the presence of thermo-diffusion effects and variable thermal conductivity
  64. A new benchmark for camouflaged object detection: RGB-D camouflaged object detection dataset
  65. Effect of electron temperature and concentration on production of hydroxyl radical and nitric oxide in atmospheric pressure low-temperature helium plasma jet: Swarm analysis and global model investigation
  66. Double diffusion convection of Maxwell–Cattaneo fluids in a vertical slot
  67. Thermal analysis of extended surfaces using deep neural networks
  68. Steady-state thermodynamic process in multilayered heterogeneous cylinder
  69. Multiresponse optimisation and process capability analysis of chemical vapour jet machining for the acrylonitrile butadiene styrene polymer: Unveiling the morphology
  70. Modeling monkeypox virus transmission: Stability analysis and comparison of analytical techniques
  71. Fourier spectral method for the fractional-in-space coupled Whitham–Broer–Kaup equations on unbounded domain
  72. The chaotic behavior and traveling wave solutions of the conformable extended Korteweg–de-Vries model
  73. Research on optimization of combustor liner structure based on arc-shaped slot hole
  74. Construction of M-shaped solitons for a modified regularized long-wave equation via Hirota's bilinear method
  75. Effectiveness of microwave ablation using two simultaneous antennas for liver malignancy treatment
  76. Discussion on optical solitons, sensitivity and qualitative analysis to a fractional model of ion sound and Langmuir waves with Atangana Baleanu derivatives
  77. Reliability of two-dimensional steady magnetized Jeffery fluid over shrinking sheet with chemical effect
  78. Generalized model of thermoelasticity associated with fractional time-derivative operators and its applications to non-simple elastic materials
  79. Migration of two rigid spheres translating within an infinite couple stress fluid under the impact of magnetic field
  80. A comparative investigation of neutron and gamma radiation interaction properties of zircaloy-2 and zircaloy-4 with consideration of mechanical properties
  81. New optical stochastic solutions for the Schrödinger equation with multiplicative Wiener process/random variable coefficients using two different methods
  82. Physical aspects of quantile residual lifetime sequence
  83. Synthesis, structure, IV characteristics, and optical properties of chromium oxide thin films for optoelectronic applications
  84. Smart mathematically filtered UV spectroscopic methods for quality assurance of rosuvastatin and valsartan from formulation
  85. A novel investigation into time-fractional multi-dimensional Navier–Stokes equations within Aboodh transform
  86. Homotopic dynamic solution of hydrodynamic nonlinear natural convection containing superhydrophobicity and isothermally heated parallel plate with hybrid nanoparticles
  87. A novel tetra hybrid bio-nanofluid model with stenosed artery
  88. Propagation of traveling wave solution of the strain wave equation in microcrystalline materials
  89. Innovative analysis to the time-fractional q-deformed tanh-Gordon equation via modified double Laplace transform method
  90. A new investigation of the extended Sakovich equation for abundant soliton solution in industrial engineering via two efficient techniques
  91. New soliton solutions of the conformable time fractional Drinfel'd–Sokolov–Wilson equation based on the complete discriminant system method
  92. Irradiation of hydrophilic acrylic intraocular lenses by a 365 nm UV lamp
  93. Inflation and the principle of equivalence
  94. The use of a supercontinuum light source for the characterization of passive fiber optic components
  95. Optical solitons to the fractional Kundu–Mukherjee–Naskar equation with time-dependent coefficients
  96. A promising photocathode for green hydrogen generation from sanitation water without external sacrificing agent: silver-silver oxide/poly(1H-pyrrole) dendritic nanocomposite seeded on poly-1H pyrrole film
  97. Photon balance in the fiber laser model
  98. Propagation of optical spatial solitons in nematic liquid crystals with quadruple power law of nonlinearity appears in fluid mechanics
  99. Theoretical investigation and sensitivity analysis of non-Newtonian fluid during roll coating process by response surface methodology
  100. Utilizing slip conditions on transport phenomena of heat energy with dust and tiny nanoparticles over a wedge
  101. Bismuthyl chloride/poly(m-toluidine) nanocomposite seeded on poly-1H pyrrole: Photocathode for green hydrogen generation
  102. Infrared thermography based fault diagnosis of diesel engines using convolutional neural network and image enhancement
  103. On some solitary wave solutions of the Estevez--Mansfield--Clarkson equation with conformable fractional derivatives in time
  104. Impact of permeability and fluid parameters in couple stress media on rotating eccentric spheres
  105. Review Article
  106. Transformer-based intelligent fault diagnosis methods of mechanical equipment: A survey
  107. Special Issue on Predicting pattern alterations in nature - Part II
  108. A comparative study of Bagley–Torvik equation under nonsingular kernel derivatives using Weeks method
  109. On the existence and numerical simulation of Cholera epidemic model
  110. Numerical solutions of generalized Atangana–Baleanu time-fractional FitzHugh–Nagumo equation using cubic B-spline functions
  111. Dynamic properties of the multimalware attacks in wireless sensor networks: Fractional derivative analysis of wireless sensor networks
  112. Prediction of COVID-19 spread with models in different patterns: A case study of Russia
  113. Study of chronic myeloid leukemia with T-cell under fractal-fractional order model
  114. Accumulation process in the environment for a generalized mass transport system
  115. Analysis of a generalized proportional fractional stochastic differential equation incorporating Carathéodory's approximation and applications
  116. Special Issue on Nanomaterial utilization and structural optimization - Part II
  117. Numerical study on flow and heat transfer performance of a spiral-wound heat exchanger for natural gas
  118. Study of ultrasonic influence on heat transfer and resistance performance of round tube with twisted belt
  119. Numerical study on bionic airfoil fins used in printed circuit plate heat exchanger
  120. Improving heat transfer efficiency via optimization and sensitivity assessment in hybrid nanofluid flow with variable magnetism using the Yamada–Ota model
  121. Special Issue on Nanofluids: Synthesis, Characterization, and Applications
  122. Exact solutions of a class of generalized nanofluidic models
  123. Stability enhancement of Al2O3, ZnO, and TiO2 binary nanofluids for heat transfer applications
  124. Thermal transport energy performance on tangent hyperbolic hybrid nanofluids and their implementation in concentrated solar aircraft wings
  125. Studying nonlinear vibration analysis of nanoelectro-mechanical resonators via analytical computational method
  126. Numerical analysis of non-linear radiative Casson fluids containing CNTs having length and radius over permeable moving plate
  127. Two-phase numerical simulation of thermal and solutal transport exploration of a non-Newtonian nanomaterial flow past a stretching surface with chemical reaction
  128. Natural convection and flow patterns of Cu–water nanofluids in hexagonal cavity: A novel thermal case study
  129. Solitonic solutions and study of nonlinear wave dynamics in a Murnaghan hyperelastic circular pipe
  130. Comparative study of couple stress fluid flow using OHAM and NIM
  131. Utilization of OHAM to investigate entropy generation with a temperature-dependent thermal conductivity model in hybrid nanofluid using the radiation phenomenon
  132. Slip effects on magnetized radiatively hybridized ferrofluid flow with acute magnetic force over shrinking/stretching surface
  133. Significance of 3D rectangular closed domain filled with charged particles and nanoparticles engaging finite element methodology
  134. Robustness and dynamical features of fractional difference spacecraft model with Mittag–Leffler stability
  135. Characterizing magnetohydrodynamic effects on developed nanofluid flow in an obstructed vertical duct under constant pressure gradient
  136. Study on dynamic and static tensile and puncture-resistant mechanical properties of impregnated STF multi-dimensional structure Kevlar fiber reinforced composites
  137. Thermosolutal Marangoni convective flow of MHD tangent hyperbolic hybrid nanofluids with elastic deformation and heat source
  138. Investigation of convective heat transport in a Carreau hybrid nanofluid between two stretchable rotatory disks
  139. Single-channel cooling system design by using perforated porous insert and modeling with POD for double conductive panel
  140. Special Issue on Fundamental Physics from Atoms to Cosmos - Part I
  141. Pulsed excitation of a quantum oscillator: A model accounting for damping
  142. Review of recent analytical advances in the spectroscopy of hydrogenic lines in plasmas
  143. Heavy mesons mass spectroscopy under a spin-dependent Cornell potential within the framework of the spinless Salpeter equation
  144. Coherent manipulation of bright and dark solitons of reflection and transmission pulses through sodium atomic medium
  145. Effect of the gravitational field strength on the rate of chemical reactions
  146. The kinetic relativity theory – hiding in plain sight
  147. Special Issue on Advanced Energy Materials - Part III
  148. Eco-friendly graphitic carbon nitride–poly(1H pyrrole) nanocomposite: A photocathode for green hydrogen production, paving the way for commercial applications
Downloaded on 7.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/phys-2024-0077/html
Scroll to top button