Skip to main content

ORIGINAL RESEARCH article

Front. Phys., 23 May 2022
Sec. Cosmology
Volume 10 - 2022 | https://doi.org/10.3389/fphy.2022.826188

Cosmological Redshift and Cosmic Time Dilation in the FLRW Metric

  • Institute of Geophysics, Czech Academy of Sciences, Prague, Czechia

The paper shows that the commonly used Friedmann-Lemaitre-Robertson-Walker (FLRW) metric describing the expanding Universe must be modified to properly predict the cosmological redshift. It is proved that the change in the frequency of redshifted photons is always connected with time dilation, similarly as for the gravitational redshift. Therefore, the cosmic time runs differently at high redshifts than at present. Consequently, the cosmological time must be identified with the conformal time and the standard FLRW metric must be substituted by its conformal version. The correctness of the proposed conformal metric is convincingly confirmed by Type Ia supernovae (SNe Ia) observations. The standard FLRW metric produces essential discrepancy with the SNe Ia observations called the ‘supernova dimming’, and dark energy has to be introduced to comply theoretical predictions with data. By contrast, the conformal FLRW metric fits data well with no need to introduce any new free parameter. Hence, the discovery of the supernova dimming actually revealed a failure of the FLRW metric and introducing dark energy was just an unsuccessful attempt to cope with the problem within this false metric. Obviously, adopting the conformal FLRW metric for describing the evolution of the Universe has many fundamental cosmological consequences.

1 Introduction

Friedmann [1] applied the Einstein equations of General Relativity (GR) for describing the Universe and firstly showed that the space filled by uniformly distributed matter might evolve in time. The possibility that the Universe is really dynamic but not static was later supported by Lemaitre [2] and Hubble [3], who observed a systematic redshift of nearby galaxies, which was roughly proportional to their distance. This observation (called the Hubble-Lemaitre law) was interpreted as the Doppler effect produced by galaxies moving away from the Earth due to the Universe expansion.

However, the intuitive idea of the redshift as the Doppler effect was later abandoned. At present, the Universe is described by the so-called Friedmann-Lemaitre-Robertson-Walker (FLRW) metric [48], which introduces the scale factor a(t) for describing the space expansion. The redshift is not related to the speed of the expansion as for the Doppler effect but to the ratio between sizes of the space, in which the photons were emitted and received [9, 10].

1+z=arae(1)

where z is the redshift, and a(e) and a(r) are the scale factors for the emitter and receiver, respectively. Hence, the redshift of distant galaxies would be observed even in the case, when the Universe is not expanding anymore at the present epoch.

In contrast to the space coordinates, the time coordinate is assumed to be invariable during the Universe history. This is somewhat strange and surprising, because other solutions in GR such as the well-known Schwarzschild solution [1113] involve distortions in space and time together. Therefore, some authors pointed out to other alternative theories admissible in GR and introduced more general metrics for describing isotropic homogeneous Universe evolving in time [1416]. In this case, another function is considered in the metric tensor gαβ, which describes the evolution of the time component g00.

Among many possibilities how to define this function, the simplest way is to assume that the time and scale factors are defined by the same function a(t). This option has a clear advantage, because the cosmological redshift will be defined by the same formula as the gravitational redshift

1+z=g00rg00e(2)

where g00(e) and g00(r) are the time components of the metric tensor gαβ for the emitter and receiver, respectively.

Introducing the same scale factor for time and space coordinates has also other advantages. Firstly, this metric evolves in time according to the so-called conformal transformation, properties of which are intensively studied in GR in recent years [1719]. The new time coordinate is called the conformal time and the metric utilizing this time is called the conformal metric [1416]. This metric is particularly interesting, because it leaves the Maxwell’s equations unchanged from their form in the Minkowski spacetime [2022]. The conformal metrics have also other exceptional properties and open space for new cosmological models as the Conformally Flat Space-Time Cosmology [14, 15, 23], Conformal Gravity [17, 24] or the Conformal Cyclic Cosmology [19, 2527].

Nevertheless, introducing the conformal time into the FLRW metric is commonly viewed as a mathematical concept different from the physical cosmic time [16]. Otherwise, we have to admit a variable coordinate speed of light dependent on the scale factor a(t). Although, theories of variable speed of light (VSL) exist [28, 29], they are not paid much attention, because they are against a deeply rooted concept of the speed of light as a nature constant. Nevertheless, Dicke [30] argues in his pioneering work on gravity that VSL is physically admissible. Also Dirac [31] states that “The laws may be changing, and in particular quantities which are considered to be constants of nature may be varying with cosmological time.”

In this paper, the problem of cosmic time dilation and cosmological redshift in the standard FLRW metric is revisited. It is shown that time dilation and redshift observations are, actually, inconsistent with the original FLRW metric. Instead, the conformal FLRW metric should be used for describing the Universe evolution, because it predicts time dilation and redshift correctly. Cosmological consequences of this correction are discussed.

2 Theory

2.1 FLRW Metric

The space filled by a homogenous and isotropic matter is described by the following general metric [12, 16, 22, 32]:

ds2=A2tc2dt2+B2tdΣ2,(3)

where ds = cdτ is the spacetime element, c is the speed of light, τ is the proper time, t is the coordinate time, Σ is the 3-dimensional coordinate in space of uniform curvature, and A(t) and B(t) are arbitrary functions describing time evolution of time dilation and space expansion, respectively.

The standard FLRW metric is based on the assumption of the space expansion described by the scale factor a(t) = B(t) and with no time dilation A(t) = 1. Hence, the metric reads in the spherical coordinate system as [9, 10, 33].

ds2=c2dt2+a2tdr21kr2+r2dΩ2,dΩ2=dΘ2+sin2Θdϕ2,(4)

k is the curvature index of the space, r is the comoving distance, and Θ and ϕ are the spherical angles.

An alternative to Eq. 4 is the so-called conformal form of the FLRW metric [16], which assumes the same factor a(t) for time dilation and space expansion, A(t) = B(t) = a(t),

ds2=a2tc2dt2+dr21kr2+r2dΩ2,(5)

where time t has a different physical meaning than in Eq. 4 being often denoted as η.

Obviously, Einstein’s equations do not constrain functions A(t) and B(t) in Eq. 3 and they do not give us any preference between Eq. 4 for the standard FLRW metric and Eq. 5 for the conformal FLRW metric. Both metrics are based on the assumption of perfect isotropy and homogeneity and they satisfy the GR equations.

2.2 Coordinate Freedom of Choosing Time

We can see that Eq. 5 is obtained from Eq. 4 by a simple transformation

dt=atdη,(6)

where η is called the conformal (comoving) time and t is the proper time. Commonly, the conformal time η is considered as a mathematical concept different from the physical coordinate time. In this case, Eqs 4, 5 are physically equivalent, because we applied just rescaling of time using Eq. 6 and the Einstein equations are coordinate invariant [12, 34].

However, we should be aware that the coordinate invariance of the Einstein equations does not mean that we can rescale time and space coordinates arbitrarily with no physical consequences. The physically meaningful coordinates should be identified with the “cosmological coordinate system,” in which all fundamental bodies are in rest [14, 15, 20, 21]. Also, we cannot mix comoving and proper coordinates in the metric. If we ignore this condition and do not distinguish between comoving and proper coordinates, Eqs 4, 5 can possibly describe the static Universe, provided distance r is substituted by the proper distance R as

dr=dRat.(7)

Hence, the key for understanding Eqs 4, 5 is to define, which quantities are physical (being related to the cosmological coordinate system) and which quantities describe just an arbitrary coordinate with no physical meaning. If r is the comoving distance, Eqs 4, 5 do not describe the static Universe but the expanding Universe.

Similarly, if the conformal time η is the comoving time measured by clocks in the cosmological coordinate system, then Eqs 4, 5 define two physically different Universe models. This is obvious, because Eq. 4 assumes the cosmic time being invariant of the space expansion, while Eq. 5 assumes the cosmic time being dependent on the space expansion. Consequently, the coordinate speed of light is invariant in Eq. 4 but it depends on a(t) in Eq. 5, see Appendix A. Since both equations are admissible in GR, the correct form of the metric of the cosmological coordinate system must be found from observations. Primarily, the correct metric should satisfactorily explain observations of the cosmological redshift.

2.3 Cosmological Redshift Inconsistency

The cosmological redshift in the standard FLRW metric is commonly explained as the change of the photon wavelength due to the space expansion [9, 10, 33, 35, 36]. The common derivation in textbooks is as follows. Light travels along the null geodesic, ds = cdτ = 0, hence

c2dt2=a2tdl2,(8)

where dl is the element of the comoving distance. Consequently,

cdtat=dl.(9)

Suppose the distant galaxy emits photons at constant rate Δte and with wavelength λe. The photons are observed at rate Δtr and with wavelength λr. The first photon is emitted at time te and received at time tr. Taking into account that the comoving distance between the galaxy and the observer is the same for the two successive photons

tetrcdtat=te+Δtetr+Δtrcdtat(10)

and subtracting the integral

te+Δtetrcdtat(11)

we get

tete+Δtecdtat=trtr+Δtrcdtat(12)

Since the scale factor a(t) varies slowly and does not change much during emission and observation of the two successive photons, we write

1atetete+Δtecdt=1atrtrtr+Δtrcdt.(13)

Hence,

deate=dratr(14)

where de = cΔte and dr = cΔtr are the distances between two successive photons at the emitter and the receiver, respectively. Subsequently, we can conclude that the wavelengths of photons λe and λr obey the same relation

λeate=λratr(15)

This derivation is not, however, correct. Using Eq. 13, we can also obtain the following equation

Δteate=Δtratr(16)

which indicates that the coordinate time depends on the scale factor a(t). Obviously, Eq. 16 is inconsistent with the standard FLRW metric described by Eq. 4, where the coordinate time is invariant. Alternatively, we can keep the coordinate time independent of the scale factor, but then we have to assume that the light speed c depends on the scale factor a(t) and we have to distinguish between the light speed in the emitter, ce, and in the receiver, cr. This is again inconsistent with Eq. 4.

The basic difficulty with the above derivation of redshift-dependent wavelengths of photons lies in an incorrect definition of the wavelength as distance between two different spacetime events, see Appendices B, C. Obviously, the distance must be measured at one coordinate system, but not as a distance between points in two different coordinate systems connected with two photons measured at different times. Once we consider two photons travelling along the same ray path with distance d between them at the same coordinate time, the effect of increasing the distance between photons during the space expansion disappears. After any time t, both photons travel the same distance along the same ray, and consequently the distance between them keeps time independent, see Appendix B.

Mathematically, we modify Eq. 10, in which we do not assume the equality of the comoving distance but the equality of the light travel distance of the photons propagating along the same raypath from the emitter to the receiver:

tetrcdt=te+Δttr+Δtcdt.(17)

Using the same logic as above, we obtain that if time and speed of light is not changing, the wavelength of photons does not change. Hence, two successive photons travelling along the same raypath keep their mutual proper distance constant and independent of redshift. However, the proper distance between two photons travelling along two parallel rays at the same time depends on redshift and increases with the space expansion. This is because the comoving distance between two photons moving along parallel raypaths is constant, hence the proper distance must increase with the space expansion, see Appendix C. Only the proper distance between two successive photons travelling along the same ray does not change, see Appendix B.

The above derivation proves that the standard FLRW metric cannot be applied to the Universe, because it does not predict the cosmological redshift. The cosmological redshift can be observed only if the cosmic time depends on the scale factor a(t) and it runs differently at high redshift than at present. Therefore, the cosmological redshift is not a consequence of the space expansion but of time dilation. A disputable character of the original FLRW metric is also indicated by comparing this metric with other solutions in GR, where the expansion/contraction of space is tightly connected with time dilation. If we insist on no time dilation, no redshift will be observed.

The variability of the cosmic time during the Universe evolution would be supported by the fact that the mass density in the Universe is time dependent. At previous epochs, the Universe was much denser and the gravitational field much stronger. Going back in time to high redshifts is analogous to the case, when an observer moves towards the black hole. According to the Schwarzschild solution, the coordinate time for the observer close to the black hole runs differently than for the observer far from the black hole. Similarly, the coordinate time must run differently at the high redshift Universe than at the present epoch. Consequently, assuming that the Universe expands but the cosmic time is invariant is physically unjustified.

Hence, the correct metric is the conformal form of the FLRW metric described by Eq. 5 and the cosmological redshift obeys the same formula as the gravitational redshift:

νeνr=1+z=g00rg00e(18)

where z is the redshift, νe and νr are the frequencies of the photon at the emitter and receiver, and g00(e) and g00(r) are the time components of the metric tensor gαβ at the emitter and receiver, respectively.

2.4 Properties of the Conformal FLRW Metric

The conformal FLRW metric is essentially different from the original FLRW metric with fundamental physical consequences:

Eq. 5 implies that the comoving speed of light is constant but the proper speed of light depends on redshift. Hence, the volume of the Universe and distance between galaxies were smaller at high redshift, but photons emitted by a galaxy reach a neighbouring galaxy after the same time at high redshift as well as at the present epoch. In other words, this Universe model is conformal with the static Universe.

• The frequency νe of photons emitted at redshift z is higher than the frequency νr of photons received as:

νeνr=1+z.(19)

• Not only the frequency ν of photons but also the rate of photons increases with redshift as (1 + z).

• The proper speed of light c in the cosmological coordinate system decreases with redshift as (1 + z)−1.

• The wavelength λe of photons emitted at redshift z is shorter than the wavelength λr of photons received as:

λeλr=1+z2.(20)

This includes a decrease of frequency ν and an increase of the speed of light c with cosmic time.

2.5 Friedmann Equations Revisited

If the expansion of the Universe is described by the conformal FLRW metric, the Friedmann equations must be modified. The standard Friedmann equations for the pressureless fluid read [10, 33].

ȧa2=8πG3ρkc2a2+13Λc2,(21)
äa=4πG3ρ+13Λc2,(22)

where a=1+z1 is the scale factor, G is the gravitational constant, ρ is the mean mass density, k/a2 is the spatial curvature of the Universe, and Λ is the cosmological constant.

In order to express the Friedmann equations for the conformal FLRW metric, we have to substitute time t by the conformal time η and time derivative ȧ=da/dt by a=da/dη=aȧ. Hence, the conformal Friedmann equations read

aa2=8πG3ρa2kc2,(23)
aa=4πG3ρa2,(24)

where we omitted the cosmological constant, because it was inserted into Eqs 21 and 22 artificially in order to fit Type Ia supernova observations. Considering the matter-dominated Universe, we get

8πG3ρ=H02Ωma3(25)

and Eq. 23 reads

H2a=H02Ωma1+Ωk(26)

with the condition

Ωm+Ωk=1,(27)

where H(a) = a′/a is the Hubble parameter, H0 is the Hubble constant, Ωm is the normalized matter density, and Ωk is the normalized space curvature. Since this model is basically the Einstein-de Sitter (EdS) model but applied to the conformal FLRW metric, it will be called as the “conformal EdS model” in contrast to the standard EdS model based on the original FLRW metric.

3 Supernovae Observations

The correctness of Eq. 26 for the time evolution of the Universe can be checked by Type Ia supernova (SNe Ia) observations, which provide the most accurate measurements of cosmological distances and of the expansion history of the Universe. A discrepancy between the supernova observations and the predictions of the standard EdS model was called the “supernovae dimming” [37, 38], and led to reintroducing the cosmological constant Λ into the Einstein and Friedmann equations. The observation of the unexpected SNe Ia dimming motivated large-scale systematic searches for SNe Ia and resulted in a rapid extension of supernovae compilations.

The current supernovae compilations Union2.1 [3944], and Pantheon [45, 46] comprise of hundreds of SNe Ia discovered and spectroscopically confirmed. The Pantheon dataset is the most accurate SNe Ia compilation at present. Every SN Ia is described by its apparent rest-frame B-band magnitude mB, the absolute B-band magnitude MB, the stretch parameter x1, and the colour parameter c. These parameters are used in the Tripp formula [47, 48] for calculating the redshift-dependent distance modulus μ(z), which serves for testing the cosmological models,

μ=mBMB+αx1βc(28)

where coefficients α and β are the global nuisance parameters to be determined when seeking an optimum cosmological model. The expansion history is calculated from μ using the following equations,

μ=25+5log10dL,dL=1+z0zcdzHz(29)

where dL is the luminosity distance expressed for the flat Universe. The Hubble function H(z) is expressed for the flat Universe described by the standard ΛCDM model as

H2z=H02Ωm1+z3+ΩΛ,(30)

by the standard EdS model as

H2z=H02Ωm1+z3+Ωk1+z2,(31)

and by the conformal EdS model as

H2z=H02Ωm1+z+Ωk.(32)

While the ΛCDM model contains dark energy ΩΛ as a free parameter, which must be adjusted by fitting with the SNe Ia observations, the conformal EdS model requires no free parameter for the flat Universe, and the curvature parameter Ωk is needed for a curved Universe. Since the Universe is nearly flat, this parameter should be close to zero and can be determined from other independent observations. Model-independent methods for estimating Ωk are based on reconstructing the comoving distances by Hubble parameter data and comparing with the luminosity distances [4951], on the angular diameter distances [52], on strongly gravitational lensed SNe Ia [53] or on gravitational waves [54]. The authors report the curvature term Ωk ranging between −0.3 and −0.1 indicating that the Universe is nearly flat and closed.

Figure 1 shows a comparison of the SNe Ia measurements with predictions of the ΛCDM model and the standard and conformal EdS models. The standard EdS model is in a visible disagreement with the SNe Ia measurements and this disagreement led to developing the ΛCDM model by introducing the normalized density of dark energy ΩΛ into Eq. 30 to get a satisfactory fit. Strikingly, the conformal EdS model defined by Eq. 32 fits data equally well as the ΛCDM model with no assumption on dark energy (see Figure 2). This confirms that the solution of the puzzle with the supernovae dimming does not lie in introducing dark energy but in correcting the metric used in the Friedmann equations.

FIGURE 1
www.frontiersin.org

FIGURE 1. The Hubble diagram with Type Ia supernovae observations. Blue dots show measurements of the SNe Pantheon compilation [45, 46]. The red line in (A) shows the ΛCDM model described by Eq. 30 with Ωm = 0.3 and ΩΛ = 0.7. The red line in (B) shows the conformal EdS model described by Eq. 32 with Ωm = 1.2 and Ωk = −0.2. The black line in (A,B) shows the standard EdS model described by Eq. 31 with Ωm = 1.0 and Ωk = 0. The Hubble constant is H0 = 69.8 km s−1 Mpc−1, obtained from observations of the SNe Ia data with a red giant calibration [55].

FIGURE 2
www.frontiersin.org

FIGURE 2. Residual Hubble plots for (A,B) the individual SNe Ia data and (C,D) the binned SNe Ia data. (A,C) The flat ΛCDM model, (B,D) the conformal EdS model. For parameters of the models, see caption of Figure 1. The error bars in (C,D) show the 99% confidence intervals. Data are taken from the SNe Pantheon compilation [45, 46].

4 Discussion

The Friedmann equations introduce the expansion of the Universe and form fundamentals of modern cosmology. Intuitively, the space expansion can explain the cosmological redshift, because the distant galaxies are moving away due to the expansion and we observe their light distorted by the Doppler effect. This was probably the motivation for describing the Universe by the standard FLRW metric. The problem is, however, more involved, and we know that the cosmological redshift is not due to the Doppler effect but due to distortion of the spacetime described by GR. The redshift of distant galaxies would be observed even for a non-expanding Universe at the present epoch. From this point of view, there is no clear argument, why the standard FLRW metric introduces just the space expansion with no time dilation.

In fact, it is surprising to assume distortion of space only, because other solutions in GR such as the well-known Schwarzschild solution involve distortions in space and time together. At previous epochs, the Universe was much denser and the gravitational field much stronger, hence going back in time to high redshifts is analogous to an observer moving towards the black hole. Since the coordinate time runs differently close to and far from the black hole, we can expect to observe a similar effect when comparing clocks at the high redshift Universe and at the present epoch.

In addition, the assumption of no time dilation during the Universe evolution is not strange only from the theoretical point of view but it is also in contradiction with astronomical observations. The existence of cosmic time dilation and its real physical nature is supported by observations of gamma ray-bursts [5658] and Type Ia supernovae light curves [59, 60]. For example, Zhang [61] studied a sample of 139 SWIFT long gamma-ray bursts (GRBs) with redshift z ≤ 8.2 and obtained a significant correlation between their duration and redshift. Similarly, Littlejohns and Butler [62] analysed 232 GRBs detected by the Swift/Burst Alert Telescope (BAT) and revealed that the observed durations are consistent with cosmic time dilation. As regards supernovae, the SNe Ia display rather uniform light curves and thus they can be used as local clocks. The spectral evolution of the light curves and stretching of time in the observer frame was disclosed by many authors [59, 6365], and corrections for time dilation are now routinely applied to the SNe Ia data [60, 66].

The re-examination of light propagation in space defined by the standard FLRW metric reveals another severe contradiction with observations: this metric actually does not predict the cosmological redshift. This is surprising and against the common opinion that the standard FLRW metric produces the cosmological redshift. However, it is shown that the mathematical derivation originally proposed by Lemaitre [2] and repeated in textbooks is not correct. Lemaitre [2] analysed the change of the wavelength of photons propagating in expanding space and he came to a wrong conclusion that the wavelength of photons must increase, similarly as the proper distance between objects in rest. An increasing wavelength of photons is then transformed into the change of their frequency under the assumption of the constant speed of light. Since this derivation gave intuitively acceptable results, there was no reason to critically check its correctness by other cosmologists.

A correct analysis shows, however, that the wavelength of photons does not increase and the frequency of photons is constant during the space expansion defined by the standard FLRW metric. The change in the frequency of photons is always connected with time dilation and with a variation of the time metric g00 in GR, similarly as for the gravitational redshift. Therefore, the standard FLRW metric must be substituted by the conformal FLRW metric that predicts the cosmic time dilation and the cosmological redshift properly. Consequently, the cosmic time should be identified with the conformal time and the space-time evolution of the Universe should be described by the conformal FLRW metric only.

Obviously, we can ask a question: why atoms radiate photons with the same (rest-frame) frequency at all redshifts and why this frequency is not affected by time dilation? The answer is straightforward: the frequency of emitted photons is independent of redshift, because it depends on quantized energy levels of electrons in atoms and these energy levels are redshift independent. Once the photon is emitted, its frequency decreases due to time dilation when photon propagates along the ray path from the emitter to the receiver. Since the comoving speed of light is constant, the proper speed of light must be variable. In this way, the emitted photons with frequency ν have shorter proper wavelengths at high redshift than the photons with the same frequency ν but emitted at the present epoch.

The correctness of the conformal FLRW metric is convincingly confirmed by SNe Ia observations. In fact, observations of the SNe Ia were originally proposed by Riess et al. [37] and Perlmutter et al. [38] for testifying the existing cosmological model and the SNe Ia observations surprisingly revealed essential discrepancy between theoretical predictions and measurements. However, instead of questioning the validity of the standard FLRW metric and the Friedmann equations, Riess et al. [37] and Perlmutter et al. [38] introduced a free parameter into the Friedmann equations to comply them with data. In this way, the model is capable to fit the SNe Ia observations, but at the cost of introducing a physically controversial concept of dark energy. By contrast, the EdS model based on the conformal FLRW metric fits the SNe Ia data with no need to introduce any new free parameter.

An argument that dark energy is not physical, but originates in the applied standard FLRW metric is used also by other authors [6770]. For example, the accelerated expansion could be an artefact of neglecting inhomogeneity of the Universe [7175] as proposed in the Swiss-cheese cosmology [7678] or in the timescape cosmology [7981]. The SNe Ia dimming can partly be a result of cosmic opacity neglected in interpretations of the SNe Ia luminosity [8285]. By contrast, here I show that the essential difficulty with the standard FLRW metric is not in the oversimplification of the model by assuming perfect homogeneity and isotropy of the Universe, but in false neglecting time dilation during the Universe history. The results indicate that anisotropy, heterogeneity and opacity of the Universe produce probably only the second-order effects in observations.

5 Conclusion

In summary, we conclude that the conformal FLRW metric is the only correct metric for describing the evolution of the Universe, which can predict the cosmological redshift and time dilation properly. If the time rate is independent of the expansion of the Universe as in the standard FLRW metric, the frequency of photons cannot change during the expansion. Therefore, the variable rate of time during the expansion is inevitable and implies the following fundamental consequences:

(1) The gravitational and cosmological redshifts are calculated by the same formula and describe the same physical process. Both redshifts reflect a distortion of time produced by changes in the gravitational field. While the gravitational redshift originates in spatial variations of the gravitational field, the cosmological redshift originates in temporal variations of the gravitational field.

(2) The metric describing the evolution of the Universe is conformal with the static model. This metric leaves the Maxwell’s equations unchanged from their form in the Minkowski spacetime [2022].

(3) The conformal FLRW metric predicts correctly the cosmological redshift: the frequency of photons increases with redshift as (1 + z). Not only the frequency of photons but also the rate of photons increases with redshift as (1 + z) due to time dilation. The real physical nature of cosmic time dilation is supported by observations of gamma ray-bursts [5658] and Type Ia supernovae light curves [59, 60, 66].

(4) The comoving speed of light is constant. The proper speed of light decreases with redshift as (1 + z)−1. Hence, the speed of light is not a nature constant but it varies being dependent on the scale factor a(t) [28, 86]. Consequently, distance between galaxies changes with redshift, but photons emitted by a galaxy reach a neighbouring galaxy after the same time at high redshift as well as at the present epoch. The wavelength of photons does not decrease with redshift as (1 + z)−1 as assumed in the standard FLRW metric but it decreases with redshift as (1 + z)−2.

(5) The conformal FLRW metric fits the SN Ia observations with no need to introduce dark energy into the Einstein and Friedmann equations. The dark energy is an artefact of the erroneous metric used for describing the evolution of the Universe. Consequently, no repulsive forces produced by dark energy and acting against gravity are present in the corrected Friedmann equations. Since the only force considered in the Friedmann equations is gravity, the expansion of the Universe is decelerating at the present epoch.

Data Availability Statement

Publicly available datasets were analyzed in this study. This data can be found here: https://archive.stsci.edu/prepds/ps1cosmo/.

Author Contributions

VV is the only author of all presented results.

Conflict of Interest

The author declares that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Acknowledgments

I thank reviewers for their detailed and helpful reviews.

References

1. Friedman A. Uber die Krummung des Raumes. Z Physik (1922) 10:377–86. doi:10.1007/BF01332580

CrossRef Full Text | Google Scholar

2. Lemaître G. Un Univers homogène de masse constante et de rayon croissant rendant compte de la vitesse radiale des nébuleuses extra- galactiques. Ann de la Société Scientifique de Bruxelles (1927) 47:49–59.

Google Scholar

3. Hubble E. A Relation between Distance and Radial Velocity Among Extra-galactic Nebulae. Proc Natl Acad Sci U.S.A (1929) 15:168–73. doi:10.1073/pnas.15.3.168

PubMed Abstract | CrossRef Full Text | Google Scholar

4. Friedman A. On the Curvature of Space. Gen Relativity Gravitation (1999) 31:1991–2000. doi:10.1023/a:1026751225741

CrossRef Full Text | Google Scholar

5. Robertson HP. Kinematics and World-Structure. ApJ (1935) 82:284. doi:10.1086/143681

CrossRef Full Text | Google Scholar

6. Robertson HP. Kinematics and World-Structure III. ApJ (1936) 83:257. doi:10.1086/143726

CrossRef Full Text | Google Scholar

7. Walker AG, Milne EA. On the Formal Comparison of Milne's Kinematical System with the Systems of General Relativity. Monthly Notices R Astronomical Soc (1935) 95:263–9. doi:10.1093/mnras/95.3.263

CrossRef Full Text | Google Scholar

8. Walker AG. On Milne's Theory of World-Structure*. Proc Lond Math Soc (1937) s2-42:90–127. doi:10.1112/plms/s2-42.1.90

CrossRef Full Text | Google Scholar

9. Peebles PJE. Principles of Physical Cosmology (1993). p. 736.

Google Scholar

10. Peacock JA. Cosmological Physics. Cambridge University Press (1999). p. 704.

Google Scholar

11. Misner CW, Thorne KS, Wheeler JA. Gravitation. Princeton University Press (1973).

Google Scholar

12. Weinberg S. Gravitation and Cosmology: Principles and Applications of the General Theory of Relativity. Wiley (1972). p. 657.

Google Scholar

13. Carroll SM. Spacetime and Geometry. An Introduction to General Relativity (2004). doi:10.1017/9781108770385

CrossRef Full Text | Google Scholar

14. Endean G. Redshift and the Hubble Constant in Conformally Flat Spacetime. ApJ (1994) 434:397. doi:10.1086/174741

CrossRef Full Text | Google Scholar

15. Endean G. Cosmology in Conformally Flat Spacetime. ApJ (1997) 479:40–5. doi:10.1086/303862

CrossRef Full Text | Google Scholar

16. Grøn Ø, Johannesen S. FRW Universe Models in Conformally Flat-Spacetime Coordinates I: General Formalism. Eur Phys J Plus (2011) 126:28. doi:10.1140/epjp/i2011-11028-6

CrossRef Full Text | Google Scholar

17. Mannheim P. Alternatives to Dark Matter and Dark Energy. Prog Part Nucl Phys (2006) 56:340–445. doi:10.1016/j.ppnp.2005.08.001

CrossRef Full Text | Google Scholar

18. Capozziello S, de Laurentis M. Extended Theories of Gravity. Phys Rep (2011) 509:167–321. doi:10.1016/j.physrep.2011.09.003

CrossRef Full Text | Google Scholar

19. Penrose R. Republication of: Conformal Treatment of Infinity. Gen Relativ Gravit (2011) 43:901–22. doi:10.1007/s10714-010-1110-5

CrossRef Full Text | Google Scholar

20. Infeld L, Schild A. A New Approach to Kinematic Cosmology. Phys Rev (1945) 68:250–72. doi:10.1103/physrev.68.250

CrossRef Full Text | Google Scholar

21. Infeld L, Schild AE. A New Approach to Kinematic Cosmology-(B). Phys Rev (1946) 70:410–25. doi:10.1103/physrev.70.410

CrossRef Full Text | Google Scholar

22. Ibison M. On the Conformal Forms of the Robertson-Walker Metric. J Math Phys (2007) 48:122501. doi:10.1063/1.2815811

CrossRef Full Text | Google Scholar

23. Barut AO, Budinich P, Niederle J, Raçzka R. Conformal Space-Times-The Arenas of Physics and Cosmology. Found Phys (1994) 24:1461–94. doi:10.1007/BF02054779

CrossRef Full Text | Google Scholar

24. Mannheim PD. Making the Case for Conformal Gravity. Found Phys (2012) 42:388–420. doi:10.1007/s10701-011-9608-6

CrossRef Full Text | Google Scholar

25. Penrose R. On Cosmological Mass with Positive Λ. Gen Relativ Gravit (2011) 43:3355–66. doi:10.1007/s10714-011-1255-x

CrossRef Full Text | Google Scholar

26. Penrose R. The Big Bang and its Dark-Matter Content: Whence, Whither, and Wherefore. Found Phys (2018) 48:1177–90. doi:10.1007/s10701-018-0162-3

CrossRef Full Text | Google Scholar

27. Tod P. The Equations of Conformal Cyclic Cosmology. Gen Relativ Gravit (2015) 47:17. doi:10.1007/s10714-015-1859-7

CrossRef Full Text | Google Scholar

28. Magueijo J. New Varying Speed of Light Theories. Rep Prog Phys (2003) 66:2025–68. doi:10.1088/0034-4885/66/11/r04

CrossRef Full Text | Google Scholar

29. Ellis GFR. Note on Varying Speed of Light Cosmologies. Gen Relativ Gravit (2007) 39:511–20. doi:10.1007/s10714-007-0396-4

CrossRef Full Text | Google Scholar

30. Dicke RH. Gravitation without a Principle of Equivalence. Rev Mod Phys (1957) 29:363–76. doi:10.1103/revmodphys.29.363

CrossRef Full Text | Google Scholar

31. Dirac P. On Methods in Theoretical Physics (1968). Lecture in ICTP, Trieste.

Google Scholar

32. Harada T, Carr BJ, Igata T. Complete Conformal Classification of the Friedmann-Lemaître-Robertson-Walker Solutions with a Linear Equation of State. Class Quan Grav. (2018) 35:105011. doi:10.1088/1361-6382/aab99f

CrossRef Full Text | Google Scholar

33. Ryden B. Introduction to Cosmology (2016).

Google Scholar

34. Mitra A. Deriving Friedmann Robertson Walker Metric and Hubble's Law from Gravitational Collapse Formalism. Results Phys (2012) 2:45–9. doi:10.1016/j.rinp.2012.04.002

CrossRef Full Text | Google Scholar

35. Mukhanov V. Physical Foundations of Cosmology. Cosmology (2005). doi:10.1017/cbo9780511790553

CrossRef Full Text | Google Scholar

36. Matravers D. Steven Weinberg: Cosmology. Gen Relativ Gravit (2008) 41:1455–8. doi:10.1007/s10714-008-0728-z

CrossRef Full Text | Google Scholar

37. Riess AG, Filippenko AV, Challis P, Clocchiatti A, Diercks A, Garnavich PM, et al. Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant. Astronomical J (1998) 116:1009–38. doi:10.1086/300499

CrossRef Full Text | Google Scholar

38. Perlmutter S, Aldering G, Goldhaber G, Knop RA, Nugent P, Castro PG, et al. Measurements of Ω and Λ from 42 High‐Redshift Supernovae. ApJ (1999) 517:565–86. doi:10.1086/307221

CrossRef Full Text | Google Scholar

39. Sullivan M, Guy J, Conley A, Regnault N, Astier P, Balland C, et al. SNLS3: Constraints on Dark Energy Combining the Supernova Legacy Survey Three-Year Data with Other Probes. ApJ (2011) 737:102. doi:10.1088/0004-637X/737/2/102

CrossRef Full Text | Google Scholar

40. Suzuki N, Rubin D, Lidman C, Aldering G, Amanullah R, Barbary K, et al. Thehubble Space Telescopecluster Supernova Survey. V. Improving the Dark-Energy Constraints Abovez> 1 and Building an Early-Type-Hosted Supernova Sample. ApJ (2012) 746:85. doi:10.1088/0004-637X/746/1/85

CrossRef Full Text | Google Scholar

41. Campbell H, D’Andrea CB, Nichol RC, Sako M, Smith M, Lampeitl H, et al. Cosmology with Photometrically Classified Type Ia Supernovae from the SDSS-II Supernova Survey. ApJ (2013) 763:88. doi:10.1088/0004-637x/763/2/88

CrossRef Full Text | Google Scholar

42. Betoule M, Kessler R, Guy J, Mosher J, Hardin D, Biswas R, et al. Improved Cosmological Constraints from a Joint Analysis of the SDSS-II and SNLS Supernova Samples. A&A (2014) 568:A22. doi:10.1051/0004-6361/201423413

CrossRef Full Text | Google Scholar

43. Rest A, Scolnic D, Foley RJ, Huber ME, Chornock R, Narayan G, et al. Cosmological Constraints from Measurements of Type Ia Supernovae Discovered during the First 1.5 Yr of the Pan-STARRS1 Survey. ApJ (2014) 795:44. doi:10.1088/0004-637X/795/1/44

CrossRef Full Text | Google Scholar

44. Riess AG, Casertano S, Yuan W, Macri L, Bucciarelli B, Lattanzi MG, et al. Milky Way Cepheid Standards for Measuring Cosmic Distances and Application to Gaia DR2: Implications for the Hubble Constant. ApJ (2018) 861:126. doi:10.3847/1538-4357/aac82e

CrossRef Full Text | Google Scholar

45. Scolnic DM, Jones DO, Rest A, Pan YC, Chornock R, Foley RJ, et al. The Complete Light-Curve Sample of Spectroscopically Confirmed SNe Ia from Pan-STARRS1 and Cosmological Constraints from the Combined Pantheon Sample. ApJ (2018) 859:101. doi:10.3847/1538-4357/aab9bb

CrossRef Full Text | Google Scholar

46. Jones DO, Scolnic DM, Riess AG, Rest A, Kirshner RP, Berger E, et al. Measuring Dark Energy Properties with Photometrically Classified Pan-STARRS Supernovae. II. Cosmological Parameters. ApJ (2018) 857:51. doi:10.3847/1538-4357/aab6b1

CrossRef Full Text | Google Scholar

47. Tripp R. A Two-Parameter Luminosity Correction for Type IA Supernovae. Astron Astrophysics (1998) 331:815–20.

Google Scholar

48. Guy J, Astier P, Baumont S, Hardin D, Pain R, Regnault N, et al. SALT2: Using Distant Supernovae to Improve the Use of Type Ia Supernovae as Distance Indicators. A&A (2007) 466:11–21. doi:10.1051/0004-6361:20066930

CrossRef Full Text | Google Scholar

49. Clarkson C, Cortês M, Bassett B. Dynamical Dark Energy or Simply Cosmic Curvature? J Cosmol Astropart Phys (2007) 2007:011. doi:10.1088/1475-7516/2007/08/011

CrossRef Full Text | Google Scholar

50. Li Z, Wang G-J, Liao K, Zhu Z-H. Model-independent Estimations for the Curvature from Standard Candles and Clocks. ApJ (2016) 833:240. doi:10.3847/1538-4357/833/2/240

CrossRef Full Text | Google Scholar

51. Wei J-J, Wu X-F. An Improved Method to Measure the Cosmic Curvature. ApJ (2017) 838:160. doi:10.3847/1538-4357/aa674b

CrossRef Full Text | Google Scholar

52. Yu H, Wang FY. New Model-independent Method to Test the Curvature of the Universe. ApJ (2016) 828:85. doi:10.3847/0004-637X/828/2/85

CrossRef Full Text | Google Scholar

53. Qi J-Z, Cao S, Pan Y, Li J. Cosmic Opacity: Cosmological-model-independent Tests from Gravitational Waves and Type Ia Supernova. Phys Dark Universe (2019) 26:100338. doi:10.1016/j.dark.2019.100338

CrossRef Full Text | Google Scholar

54. Liao K. The Cosmic Distance Duality Relation with Strong Lensing and Gravitational Waves: An Opacity-free Test. ApJ (2019) 885:70. doi:10.3847/1538-4357/ab4819

CrossRef Full Text | Google Scholar

55. Freedman WL, Madore BF, Hatt D, Hoyt TJ, Jang IS, Beaton RL, et al. The Carnegie-Chicago Hubble Program. VIII. An Independent Determination of the Hubble Constant Based on the Tip of the Red Giant Branch*. ApJ (2019) 882:34. doi:10.3847/1538-4357/ab2f73

CrossRef Full Text | Google Scholar

56. Norris JP. Gamma-Ray Bursts: The Time Domain. Astrophys Space Sci (1995) 231:95–102. doi:10.1007/bf00658595

CrossRef Full Text | Google Scholar

57. Lee A, Bloom ED, Petrosian V. On the Intrinsic and Cosmological Signatures in Gamma‐Ray Burst Time Profiles: Time Dilation. Astrophys J Suppl S (2000) 131:21–38. doi:10.1086/317365

CrossRef Full Text | Google Scholar

58. Chang H-Y. Fourier Analysis of Gamma-Ray Burst Light Curves: Searching for a Direct Signature of Cosmological Time Dilation. Astrophysical J Lett (2001) 557:L85–L88. doi:10.1086/323331

CrossRef Full Text | Google Scholar

59. Leibundgut B, Schommer R, Phillips M, Riess A, Schmidt B, Spyromilio J, et al. Time Dilation in the Light Curve of the Distant Type IA Supernova SN 1995K. Astrophysical J Lett (1996) 466:L21–L24. doi:10.1086/310164

CrossRef Full Text | Google Scholar

60. Leibundgut B. Cosmological Implications from Observations of Type Ia Supernovae. Annu Rev Astron Astrophys (2001) 39:67–98. doi:10.1146/annurev.astro.39.1.67

CrossRef Full Text | Google Scholar

61. Zhang F-W, Fan Y-Z, Shao L, Wei D-M. Cosmological Time Dilation in Durations of Swift Long Gamma-Ray Bursts. ApJ (2013) 778:L11. doi:10.1088/2041-8205/778/1/L11

CrossRef Full Text | Google Scholar

62. Littlejohns OM, Butler NR. Investigating Signatures of Cosmological Time Dilation in Duration Measures of Prompt Gamma-ray Burst Light Curves. Monthly Notices R Astronomical Soc (2014) 444:3948–60. doi:10.1093/mnras/stu1767

CrossRef Full Text | Google Scholar

63. Goldhaber G, Deustua S, Gabi S, Groom D, Hook I, Kim A, et al. Observation of Cosmological Time Dilation Using Type Ia Supernovae as Clocks. In: P Ruiz-Lapuente, R Canal, and J Isern, editors. Thermonuclear Supernovae (1997), 486. NATO Advanced Study Institute (ASI) Series C. Kluwer Academic Publishers, London (1997). p. 777–84. doi:10.1007/978-94-011-5710-0_48

CrossRef Full Text | Google Scholar

64. Goldhaber G, Groom DE, Kim A, Aldering G, Astier P, Conley A, et al. Timescale Stretch Parameterization of Type Ia SupernovaB‐Band Light Curves. ApJ (2001) 558:359–68. doi:10.1086/322460

CrossRef Full Text | Google Scholar

65. Phillips MM, Lira P, Suntzeff NB, Schommer RA, Hamuy M, Maza J. The Reddening-free Decline Rate versus Luminosity Relationship for Type [CLC]Ia[/CLC] Supernovae. Astronomical J (1999) 118:1766–76. doi:10.1086/301032

CrossRef Full Text | Google Scholar

66. Goobar A, Leibundgut B. Supernova Cosmology: Legacy and Future. Annu Rev Nucl Part Sci (2011) 61:251–79. doi:10.1146/annurev-nucl-102010-130434

CrossRef Full Text | Google Scholar

67. Moffat JW. Cosmic Microwave Background, Accelerating Universe and Inhomogeneous Cosmology. J Cosmol Astropart Phys (2005) 2005:012. doi:10.1088/1475-7516/2005/10/012

CrossRef Full Text | Google Scholar

68. Křížek M, Somer L. Antigravity-Its Manifestations and Origin. Ijaa (2013) 03:227–35. doi:10.4236/ijaa.2013.33027

CrossRef Full Text | Google Scholar

69. Visser M. Conformally Friedmann-Lemaître-Robertson-Walker Cosmologies. Class Quan Grav. (2015) 32:135007. doi:10.1088/0264-9381/32/13/135007

CrossRef Full Text | Google Scholar

70. Křížek M, Somer L. Excessive Extrapolations in Cosmology. Gravit Cosmol (2016) 22:270–80. doi:10.1134/S0202289316030105

CrossRef Full Text | Google Scholar

71. Bolejko K, Célérier M-N, Krasiński A. Inhomogeneous Cosmological Models: Exact Solutions and Their Applications. Class Quan Grav. (2011) 28:164002. doi:10.1088/0264-9381/28/16/164002

CrossRef Full Text | Google Scholar

72. Biswas T, Notari A. 'Swiss-cheese' Inhomogeneous Cosmology and the Dark Energy Problem. J Cosmol Astropart Phys (2008) 2008:021. doi:10.1088/1475-7516/2008/06/021

CrossRef Full Text | Google Scholar

73. Mitra A. Interpretational conflicts between the static and non-static forms of the de Sitter metric. Sci Rep (2012) 2:923. doi:10.1038/srep00923

PubMed Abstract | CrossRef Full Text | Google Scholar

74. Mitra A, Bhattacharyya S, Bhatt N. Λcdm Cosmology through the Lens of Einstein's Static Universe, the Mother of Λ. Int J Mod Phys D (2013) 22:1350012. doi:10.1142/s0218271813500120

CrossRef Full Text | Google Scholar

75. Mitra A. Energy of Einstein's Static Universe and its Implications for the ΛCDM Cosmology. J Cosmol Astropart Phys (2013) 2013:007. doi:10.1088/1475-7516/2013/03/007

CrossRef Full Text | Google Scholar

76. Marra V, Kolb EW, Matarrese S, Riotto A. Cosmological Observables in a Swiss-cheese Universe. Phys Rev D (2007) 76:123004. doi:10.1103/PhysRevD.76.123004

CrossRef Full Text | Google Scholar

77. Vanderveld RA, Flanagan ÉÉ, Wasserman I. Luminosity Distance in “Swiss Cheese” Cosmology with Randomized Voids. I. Single Void Size. Phys Rev D (2008) 78:083511. doi:10.1103/PhysRevD.78.083511

CrossRef Full Text | Google Scholar

78. Flanagan ÉÉ, Kumar N, Wasserman I, Vanderveld RA. Luminosity Distance in “Swiss Cheese” Cosmology with Randomized Voids. II. Magnification Probability Distributions. Phys Rev D (2012) 85:023510. doi:10.1103/physrevd.85.023510

CrossRef Full Text | Google Scholar

79. Wiltshire DL. Exact Solution to the Averaging Problem in Cosmology. Phys Rev Lett (2007) 99:251101. doi:10.1103/PhysRevLett.99.251101

PubMed Abstract | CrossRef Full Text | Google Scholar

80. Wiltshire DL. Average Observational Quantities in the Timescape Cosmology. Phys Rev D (2009) 80:123512. doi:10.1103/PhysRevD.80.123512

CrossRef Full Text | Google Scholar

81. Smale PR, Wiltshire DL. Supernova Tests of the Timescape Cosmology. Monthly Notices R Astronomical Soc (2011) 413:367–85. doi:10.1111/j.1365-2966.2010.18142.x

CrossRef Full Text | Google Scholar

82. Aguirre A. Intergalactic Dust and Observations of Type IA Supernovae. ApJ (1999) 525:583–93. doi:10.1086/307945

CrossRef Full Text | Google Scholar

83. Aguirre AN. Dust versus Cosmic Acceleration. Astrophysical J (1999) 512:L19–L22. doi:10.1086/311862

CrossRef Full Text | Google Scholar

84. Ménard B, Kilbinger M, Scranton R. On the Impact of Intergalactic Dust on Cosmology with Type Ia Supernovae. Monthly Notices R Astronomical Soc (2010) 406:no. doi:10.1111/j.1365-2966.2010.16464.x

CrossRef Full Text | Google Scholar

85. Vavryčuk V. Universe Opacity and Type Ia Supernova Dimming. Monthly Notices R Astronomical Soc (2019) 489:L63–L68. doi:10.1093/mnrasl/slz128

CrossRef Full Text | Google Scholar

86. Ellis GFR. On the Definition of Distance in General Relativity: I. M. H. Etherington (Philosophical Magazine Ser. 7, Vol. 15, 761 (1933)). Gen Relativ Gravit (2007) 39:1047–52. doi:10.1007/s10714-006-0355-5

CrossRef Full Text | Google Scholar

Appendix A: Coordinate Speed of Light in the Standard and Conformal FLRW Metrics

Let us assume light propagating in the space described by the standard FLRW metric, see Eq. 4. The equation of the null geodesics for photons, ds2 = 0, yields

cdt=atdl,(A1)

where dl is the element of the comoving distance. The comoving speed of light v reads

v=dldt=cat,(A2)

and the proper speed of light ṽ is

ṽ=vivi=vivigii=atv=c.(A3)

If light propagates in the space described by the conformal FLRW metric described by Eq. 5, the equation of the null geodesics for photons, ds2 = 0, yields

cdt=dl.(A4)

Hence, the comoving speed of light v is

v=dldt=c,(A5)

and the proper speed of light ṽ is

ṽ=vivi=vivigii=atv=atc.(A6)

The dependence of ṽ on the scale factor a(t) in Eq. A6 is a trivial consequence of Eq. A5 expressing that the speed of light is constant in the comoving coordinates. Since the proper speed of light is the actually measured speed of light, Eqs A3, A6 predict essentially different behaviour of light in the standard and conformal FLRW metrics.

Appendix B: Distance Between Two Successive Photons Travelling Along the Same Raypath

Let us assume two photons propagating in the space described by the standard FLRW metric, see Eq. 4. We will consider the case of two successive photons travelling along the same raypath with time delay Δt between them. The photons are emitted by a common source situated at the origin of coordinates and they travel in the space along the x-axis for time T to reach a receiver. The equations of the null geodesics for the photons, ds2 = 0, yield

cdt=atdx,cdt=atdx,(B1)

where t′ = t + Δt. The initial comoving coordinates of photons at the initial time t0 are taken as

x0=t0t0+Δtcdtat,y0=0,z0=0,(B2)
x0=0,y0=0,z0=0,(B3)

and the comoving distance d0 between the photons at time t0 reads

d0=x0x0=t0t0+Δtcdtat=d̃0a0(B4)

where d̃0 is the proper distance between the photons at time t0 defined as

d̃0=t0t0+Δtcdt=cΔt(B5)

and we assumed in Eq. B4 that the scale factor a(t) does not change much during the time interval Δt. Once the second photon reaches the receiver, we get

dT=xTxT=t0+Tt0+T+Δtcdtat=TT+Δtcdtat=d̃TaT(B6)

where aT is the scale factor at time t0 + T and d̃T is the proper distance between the photons at time t0 + T

d̃T=t0+Tt0+T+Δtcdt=cΔt.(B7)

Comparing Eqs B5, B7, we see that the proper distance between two successive photons is constant and independent of the scale factor a(t). Consequently, the wavelength of photons cannot change with the scale factor a(t) in the standard FLRW metric.

Appendix C: Distance Between Two Photons Travelling Along Parallel Raypaths

Let us assume two photons propagating in the space described by the standard FLRW metric, see Eq. 4. We will consider the case of two photons emitted at the same time by two different sources and travelling along two parallel rays. The photons travel in the space along the x-axis and need time T to reach their receivers. The equations of the null geodesics for the photons, ds2 = 0, yield

cdt=atdx,cdt=atdx.(C1)

The initial comoving coordinates of photons at the initial time t0 are taken as

x0=0,y0=d0,z0=0,(C2)
x0=0,y0=0,z0=0.(C3)

Hence, the initial comoving distance between the two photons is d0. After elapsing time T, we get

xT=t0t0+Δtcdtat,y0=d0,z0=0,(C4)
xT=t0t0+Δtcdtat,y0=0,z0=0,(C5)

and the comoving distance dT between the photons at time t0 + T reads

dT=d0.(C6)

Consequently, the proper distances d̃0 and d̃T between the two photons at times t0 and t0 + T read

d̃0=a0d0,d̃T=aTdT,(C7)

implying that the proper distance between the photons linearly increases with the increasing scale factor a(t).

Keywords: expansion of the Universe, dark energy, cosmic time dilation, Type Ia supernovae, early Universe

Citation: Vavryčuk V (2022) Cosmological Redshift and Cosmic Time Dilation in the FLRW Metric. Front. Phys. 10:826188. doi: 10.3389/fphy.2022.826188

Received: 30 November 2021; Accepted: 07 April 2022;
Published: 23 May 2022.

Edited by:

Pradyumn Kumar Sahoo, Birla Institute of Technology and Science, India

Reviewed by:

Michal Křížek, Czech Academy of Sciences, Czechia
Ravi Kant Mishra, Sant Longowal Institute of Engineering and Technology, India

Copyright © 2022 Vavryčuk. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Václav Vavryčuk, vv@ig.cas.cz

Download