Next Article in Journal
Highlights from NA61/SHINE: Proton Intermittency Analysis
Previous Article in Journal
Polarized Baryon Production in Heavy Ion Collisions: An Analytic Hydrodynamical Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Weighing Cosmological Models with SNe Ia and Gamma Ray Burst Redshift Data

by
Rajendra P. Gupta
Macronix Research Corporation, 9 Veery Lane, Ottawa, ON K1J 8X4, Canada
Universe 2019, 5(5), 102; https://doi.org/10.3390/universe5050102
Submission received: 12 March 2019 / Revised: 26 April 2019 / Accepted: 28 April 2019 / Published: 3 May 2019

Abstract

:
Many models have been proposed to explain the intergalactic redshift using different observational data and different criteria for the goodness-of-fit of a model to the data. The purpose of this paper is to examine several suggested models using the same supernovae Ia data and gamma-ray burst (GRB) data with the same goodness-of-fit criterion and weigh them against the standard Lambda cold dark matter model (ΛCDM). We have used the redshift—distance modulus (zμ) data for 580 supernovae Ia with 0.015 ≤ z ≤ 1.414 to determine the parameters for each model and then use these model parameter to see how each model fits the sole SNe Ia data at z = 1.914 and the GRB data up to z = 8.1. For the goodness-of-fit criterion, we have used the chi-square probability determined from the weighted least square sum through non-linear regression fit to the data relative to the values predicted by each model. We find that the standard ΛCDM model gives the highest chi-square probability in all cases albeit with a rather small margin over the next best model—the recently introduced nonadiabatic Einstein de Sitter model. We have made (zμ) projections up to z = 1096 for the best four models. The best two models differ in μ only by 0.328 at z = 1096, a tiny fraction of the measurement errors that are in the high redshift datasets.

1. Introduction

The universe has been modeled in different ways by physicists, especially since the observation of redshift by Hubble in the early part of the last century. Not satisfied with Doppler effect as the cause of the extragalactic redshift and consequently with the expansion of the universe, alternative models of the universe explaining the redshift were developed, the tired light being the most prominent among those suggested by the Hubble contemporaries [1]. Since the microwave background radiation discovery by Penzias and Wilson in 1964 [2] is most easily explained by the big-bang expansion model of the universe, most cosmologists consider any other model of the universe to be not worth pursuing with any seriousness. And perhaps rightly so, since on close scrutiny none of the examined models explain the observables as well as the standard big-bang model. One common problem that has plagued alternative models is that they have not been measured with the same yardstick and compared.
The redshift of the extragalactic objects supernovae Ia (SNe Ia) is considered the gold standard of cosmic observations that are used for modeling the universe. Louis Marmet [3] has compiled a summary of various mechanisms used for explaining the spectral redshift of astronomical objects. This work is quite comprehensive and does show theoretical predictions of various models but it has not attempted to compare them against the observed data. Recently, Vishwakarma and Narlikar [4] have studied some variations of the Lambda cold dark matter (ΛCDM) model and Milne model with the same methodology and several data sets and established that it is the variation in methodology and presumptions about a model that are used by most to determine the confidence level of various model parameters, rather than the validity of a model against a standard yardstick.
Our attempt in this paper is to consider several cosmological models and perform for each of these the χ2 test using exactly the same data and same methodology. The models considered in this work are described briefly in the next section. We will also attempt to see how each model is able to fit a significantly higher redshift data of supernova Ia at z = 1.914, which is not in the database used for fitting the model and the GRB analyzed zμ data provided by several researchers. We will extend our calculations for selected models to very high redshift region, say to z ≈ 1000 to check (a) how much the models diverge when extrapolated to such outrageous values, (b) to provide a base to the teams working for very high redshift measurements, such as by observing the redshifted 21-cm (1420 MHz) line of hydrogen atom using radio telescopes [5,6]. A low level of divergence among models based on substantially different theoretical considerations will provide confidence in the basic physics of the expansion of the universe. Consequently, astronomers can use the extrapolated curves or tables to estimate the distance of the redshifted 21-cm line source. Bowman et al. have recently reported observation of an absorption profile of this line centered at 78 MHz in the sky averaged spectrum [7]. This translates into z ≈ 17 spanning over 15 < z < 20.
The intent here is not to be comprehensive since there are several excellent reviews on the subject, for example, by López-Corredoira [8] and Orlov and Raikov [9]. In addition, we do not intend to discuss the dark matter and dark energy issues and other problems of the ΛCDM model since they have been extensively discussed and reviewed in the literature, e.g., [10,11,12,13].
In Section 2 we briefly review the theory and describe the various models tested. Section 3 describes the test methodology and Section 4 the test results. The findings are discussed in Section 5 followed by our projections in Section 6. Finally, conclusions are presented in Section 7.

2. A Brief Description of the Tested Models

Rather than being comprehensive, we have considered only selected but diverse models so as to keep this work manageable and demonstrate its usefulness in weighing models against each other without bias. The methodology used for this comparative study is fairly straightforward and can be easily applied to any model not considered here1.
The proper distance d of the light emitting source is determined from the measurement of its bolometric flux f and comparing it with a known luminosity L. The luminosity distance dL is defined as
d L = L / 4 π f
In a flat universe the measured flux could be related to the to the luminosity L with an inverse square relation f = L/(4πd2). However, this relation needs to be modified to take into account the flux losses due to the expansion of the universe, the redshift and all other phenomena which can result in the loss of flux. Generally accepted flux loss phenomena are as follows [14]:
(a)
Increase in the wavelength causes the flux loss proportional to 1/(1 + z), and
(b)
In an expanding universe, an Increase in detection time between two consecutive photons emitted from a source leads to a reduction of flux, also proportional to 1/(1 + z).
Therefore, in an expanding universe the necessary flux correction required is proportional to 1/(1 + z)2, whereas in a non-expanding universe the correction required is proportional only to 1/(1 + z). Any other flux loss may similarly be written as proportional to 1/(1 + z)b, where the constant b is ideally determined analytically by the model or by fitting the data to the model. The measured bolometric flux fB and the luminosity distance dL may thus be written as:
f B = L / [ 4 π d 2 ( 1 + z ) 2 ( 1 + z ) b ] , or
d L = d ( 1 + z ) ( 1 + z ) b / 2 ,
for expanding universe and for non-expanding universe as:
f B = L / [ 4 π d 2 ( 1 + z ) ( 1 + z ) b ] , or
d L = d ( 1 + z ) 1 / 2 ( 1 + z ) b / 2 ,
Since the observed quantity is distance modulus μ rather than the luminosity distance dL, we will use the relation
μ = 5 log ( d L ) + 25 .
All distances are in Mpc unless otherwise specified. We will set the constant b = 0 if the model under consideration already has two or more parameters to be determined by fitting the data.

2.1. ΛCDM Standard Model

This model is the most accepted model for explaining cosmological phenomena and thus may be considered the reference models for all the other models considered in this work. Ignoring the contribution of radiation density at the current epoch and all the times at the highest redshift considered in this paper, we may write the distance modulus μ for redshift z in a flat universe as follows [14]:
μ = 5 log [ R 0 0 z d u / Ω m , 0 ( 1 + u ) 3 + 1 Ω m , 0 ] + 5 log ( 1 + z ) + 25 .
Here R0c/H0 with c the speed of light and H0 the Hubble constant; Ω0,m is the current matter density relative to critical density and 1 − Ωm,0 ≡ ΩΛ,0 is the current dark energy density relative to critical density. The constant parameter b is set to zero since the parameter ΩM,0 and through it the dark energy, may be deemed to cause the other flux loss represented by b.

2.2. Einstein de Sitter (EdeS) Model

Einstein de Sitter model represents ΛCDM model with Ωm,0 = 1. With this simplification, the integral in Equation (7) may be determined analytically. The simplified equation has only one parameter to be determined for fitting the data. We will therefore restore parameter b to estimate the unknown luminosity flux loss and write the distance modulus as:
μ = 5 log [ 2 R 0 ( 1 ( 1 + z ) 1 2 ) ] + 5 log ( 1 + z ) + 25 + 2.5 b log ( 1 + z ) .

2.3. Empty Universe (Milne) Model

Empty expanding universe was considered by Milne in the 1930s. Such a universe can be considered as a mathematical curiosity only since for it to exist the universe density, including all the elements of the universe (radiation, matter, dark energy, etc.), must be very small compared to the critical density of the universe. The proper distance of a galaxy in such a universe may be written as [14] d = R0ln(1 + z) and the distance modulus as:
μ = 5 log [ R 0 ln ( 1 + z ) ] + 5 log ( 1 + z ) + 25 + 2.5 b log ( 1 + z ) .
Here parameter b is introduced as in other models to estimate the luminosity flux loss not accounted for by the model.

2.4. Tired Light Model

This model ignores the expansion of the universe and thus the luminosity flux loss is represented by Equation (4) and the luminosity distance by Equation (5). Since proper distance is given by d = R0ln(1 + z) [15], we may write the distance modulus as:
μ = 5 log [ R 0 ln ( 1 + z ) ] + 2.5 log ( 1 + z ) + 25 + 2.5 b log ( 1 + z ) .
It should be noted that empty expanding universe of Milne has the same proper distance solution as for tired light. However, the luminosity flux loss equations for the two are different; that is, the second term on the right hand side of Equations (9) and (10).

2.5. Hybrid Tired Light + Einstein de Sitter (EDSM) Model

In this model, it is assumed that the observed redshift results jointly from the expansion of the universe and the tired light phenomenon. Since the proper distance of the light emitting source is the same whether it is determined by expansion of the universe or by the tired light, this fact can be used to determine the ratio in which the two causes share the redshift. The distance modulus is written as follows [16]:
μ = 5 log [ 2 R 0 ( z / z X ) ( 1 ( 1 + z X ) 1 2 ) ] + 2.5 log [ ( 1 + z ) ( 1 + z X ) ] + 25 + 2.5 b log ( 1 + z ) .
Here zX is the component of z due to the expansion of the universe and the balance zM is the component due to the tired light phenomenon, the two being related through 1 + z = (1 + zX)(1 + zM). The component zX in terms of z is obtained by equating the two proper distances and solving the resulting equation:
R M ln ( 1 + z M ) = 2 R X ( 1 1 1 + z X ) , or R 0 z z M ln [ 1 + z 1 + z X ] = 2 R 0 z z X ( 1 1 1 + z X ) .
Here R0z = RMzM = RXzX due to the fact that all expressions determining the proper distance must reduce to the Hubble’s law d = R0z in the limit of very small z. Here we have assumed that the expansion of the universe is defined by the Einstein de Sitter model.

2.6. Crawford’s Curvature Cosmology (Crawford) Model

This model is based on a theory comprising two hypotheses [17]. The first hypothesis is that the redshift is due to an interaction of photons with curved spacetime where they lose energy to other very low energy photons in a uniform high temperature plasma at a constant density. The second hypothesis is that there is a pressure (curvature pressure) that acts to stabilize expansion and provides a static stable universe. This hypothesis leads to modified Friedmann equations which have a simple solution for a uniform cosmic gas. It is essentially a modified tired light model with distance modulus defined as follows:
μ = 5 log [ R 0 3 sin ( ln ( 1 + z ) 3 ) ] + 2.5 log ( 1 + z ) + 25 + 2.5 b log ( 1 + z ) .
Here we have slightly modified Crawford’s expression to be consistent with other expressions we are using for distance modulus and thus added the term with parameter b. If Crawford’s expression (without this term) is a good fit to the data then b should turn out to be close to zero when we fit observed data to this model.

2.7. Marosi’s Power Law (Marosi) Model

Marosi [18] has shown that the observational data for 280 supernovae fits admirably well with the power law expression:
μ = a z b .
However, this one is not really a model as it does not have any physics or phenomenology behind it. Nevertheless, it may be used to compare with other models rather than fitting each model with the data.

2.8. Vishwakarma’s Scale Invariant (Vishwa) Model

This model is derived from the scale invariant theory that unifies the Machian theory of gravitation and electrodynamics [19]. The luminosity distance and distance modulus are written as:
d L = c H 0 1 ( 1 + z ) sinh [ ln ( 1 + z ) ] = c H 0 1 z ( 2 + z ) / 2 , and
μ = 5 log [ R 0 z ( 2 + z ) 2 ] + 25 + 2.5 b log ( 1 + z ) .
Here we have added the last term with parameter b that is not derivable from Equation (15). If Vishwakarma’s expression (without this term) is a good fit to the data then b should turn out to be close to zero.

2.9. Plasma Cosmology Model (Plasma)

This is a modified tired light approach mediated by plasma cosmology [20,21,22]. The expression for distance modulus is essentially the same as for the tired light model, Equation (10), with parameter β determining the plasma cosmology contribution.
μ = 5 log [ R 0 ln ( 1 + z ) ] + 2.5 log ( 1 + z ) + 2.5 β log ( 1 + z ) + 25 .
Thus, we should expect the results to be the same as for the tired light case with flux loss factor b there to be equal to β here.
In the plasma model, β may be seen as representing Compton scattering as a possible source of absorption but could also be considered as a regulating parameter of an unknown scattering mechanism. Zaninetti [20] limits the value of β to be 1 when only tired light is considered and 3 when Compton scattering is included. He has determined β = 2.37 for his modified tired light (MTL) model [21]. He has provided comparison of the MTL model with some other models. It should be mentioned that Zaninetti’s plasma cosmology has been developed in a Eucledean 3D space and does not require expansion of the universe.

2.10. Einstein de Sitter (EdeS-NA) Model in Nonadibatic Universe

This model is developed by relaxing the constraint of adiabatic universe used in all the models we have considered above [23]. As per the first law of thermodynamic [14,24]:
d Q = d E + d W
where dQ is the thermal energy transfer into the system, dE is the change in the internal energy of the system and dW = PdV is the work done on the system having pressure P to increase its volume by dV. dQ is normally set to zero on the grounds that the universe is perfectly homogeneous and that there can therefore be no bulk flow of thermal energy. However, if the energy loss of a particle, such as that of a photon through tired light phenomenon, is equally shared by all the particles of the universe (or by the ‘fabric’ of the universe) then dQ can be nonzero while conserving the homogeneity of the universe. A reverse argument is also true if it is determined that there is a gain in energy rather than a loss.
Thus, by abandoning the assumption that dQ = 0 of the adiabatic universe, Equation (18) may be written as:
E ˙ + P V ˙ = Q ˙
By assuming that the energy loss Q ˙ is proportional to the internal energy E of the sphere
Q ˙ = β E ,
with β as the proportionality constant (not the same as in plasma cosmology—Equation (17)), the proper distance of the redshifted source has been shown to be [23]:
d P ( z ) = ( 3 c β ) 0 z d u [ 1 ( 1 + u ) 3 2 A ] 1 .
with A ( 1 e β t 0 / 2 ) , e β t 0 2 = 1 + ( 1 H 0 ) ( β 3 ) and the deceleration parameter q 0 = 1 + ( 3 2 ) e β t 0 2 in a flat, matter only universe. Here t0 is the age of the universe. We may then write μ in terms of A ≡ −1/D as follows:
d P ( z ) = R 0 1 + D 0 z d u [ 1 + D ( 1 + u ) 3 2 ] 1 .
We may also write this expression where D is expressed in terms of q0 since D = 2 ( 1 + q 0 ) 1 2 q 0 :
d P ( z ) = R 0 ( 1 2 q 0 ) 3 0 z d u [ 1 + ( 2 ( 1 + q 0 ) 1 2 q 0 ) ( 1 + u ) 3 2 ] 1 .
The distance modulus μ may be written as:
μ = 5 log [ R 0 ( 1 2 q 0 ) 3 0 z d u [ 1 + ( 2 ( 1 + q 0 ) 1 2 q 0 ) ( 1 + u ) 3 2 ] 1 ] + 5 log ( 1 + z ) + 25 + 2.5 b log ( 1 + z ) .
The last term has been added to see how much flux correction is required if q0 is fixed to a value determined analytically or by alternative approaches.

2.11. Hybrid Tired Light + Einstein de Sitter (EDSM-NA) Model in Nonadiabatic Universe

We can include tired light contribution to the redshift following the approach in Section 2.5 above and in reference [23] and recalculate the distance modulus μ. Using subscript M for tired light and X for expansion effect and equating the proper distance expressions for the two and since 1 + z = (1 + zM)(1 + zX) and R0z = RMzM = RXzM, we may write
R M ln ( 1 + z M ) = ( R X 1 + D ) 0 z X d u [ 1 + D ( 1 + u ) 3 2 ] 1 , or
( R 0 z z M ) ln ( 1 + z M ) = ( R 0 z ( 1 + D ) z X ) 0 z X d u [ 1 + D ( 1 + u ) 3 2 ] 1 , or
( R 0 z ( 1 + z X ) z z X ) ln ( ( 1 + z ) / ( 1 + z X ) ) = ( R 0 z ( 1 + D ) z X ) 0 z X d u [ 1 + D ( 1 + u ) 3 2 ] 1 .
It is not possible to express analytically zX (or zM) in terms of z and write μ directly in terms of z. Nevertheless, Equation (27) can be numerically solved for zX for any value of z and distance modulus calculated to include tired light effect as well as expansion effect using the expression
μ = 5 log [ R 0 ( 1 2 q 0 ) 3 ( z z X ) 0 z X d u ( 1 + ( 2 ( 1 + q 0 ) 1 2 q 0 ) ( 1 + u ) 3 2 ) 1 ] + 2.5 log ( 1 + z X ) ( 1 + z ) + 25 + 2.5 b log ( 1 + z ) .
In order to be consistent with other models, we have added the last term on the right hand side of Equation (28) to see how much of the luminosity loss factor is unaccounted for when q0 has been fixed to a value determined by alternative methods. For example, q0 = −0.4 is determined analytically in the adiabatic hybrid model [16]. When this value of q0 is substituted in Equation (28), it becomes
μ = 5 log [ R 0 0.6 ( z z X ) 0 z X d u ( 1 + 2 3 ( 1 + u ) 3 2 ) 1 ] + 2.5 log ( 1 + z X ) ( 1 + z ) + 25 + 2.5 b log ( 1 + z ) .
As we will see below, when fitted to the SNe Ia data with q0 = −0.4, it yields b = 0.1371 ± 0.06786 meaning that a small luminosity flux remains unaccounted for by Equation (29). It will result in the luminosity distance correction factor equal to (1 + z)0.069±0.034. Therefore, for b = 0, q0 has to be −0.5767 ± 0.0743, about the same as for the ΛCDM model.

3. Test Methodology

We have used Matlab curve fitting tool to fit the data to each of the above model by minimizing χ2, the weighted summed square of residual of μ, through nonlinear regression [25]:
χ 2 = i = 1 N w i [ μ ( z i ; R 0 , p 1 , p 2 ) μ o b s , i ] 2 .
Here N is the number of data points, wi is the weight of the ith data point μobs,i determined from the measurement error σ μ O b s , i in the observed distance modulus μobs,i using the relation w i = 1 / σ μ O b s , i 2 and μ(zi; R0, p1, p2 …) is the model calculated distance modulus dependent on parameters R0 and all other model dependent parameter p1, p2 and so forth. For example, in the case of ΛCDM model considered here p1 ≡ Ωm,0 and there is no other unknown parameter.
We can quantify the goodness-of-fit of a model by calculating the χ2 probability for a model whose χ2 has been determined by fitting the observed data with known measurement error as above. This probability P is given for a χ2 distribution with n degrees of freedom (DOF), the latter being the number of data points less the number of fitted parameters, by:
P ( χ 2 , n ) = ( 1 Γ ( n 2 ) ) χ 2 2 e u u n 2 1 d u .
Here Γ is the well know gamma function that is generalization of the factorial function to complex and non-integer numbers. If the data fits the model reasonably well then χ2 determined using Equation (30) should be about the same as the degrees of freedom since the fit variance for each data point should not be significantly more than the measurement error for each data point, possibly less. Lower the value of χ2 better is the fit but the real test of the goodness-of-fit is the χ2 probability P; higher the value of P for a model, better is the model’s fit to the data. There are several on line calculators available to determine P from the input of χ2 and DOF [26].
Apart from the above statistical criterion, we also have to consider what each parameter represents. Is the parameter part of the model itself or is it used to estimate the unaccounted loss (or gain) of the luminosity flux? If it is the former, it can have any acceptable value, such as Ωm,0 for ΛCDM model. If it is the latter, then it should have as small a value as possible for the model to be considered a good fit to the data. For the latter case, we represent it with the parameter b such as in many of the models above.
One problem we faced in comparing the various models, already parameterized with SNe Ia data, for the goodness-of-fit with zμ GRB data provided by various researchers, was that the χ2 is strongly depended on the error bars on μ. If the error bars are too tight then none of the model would give an acceptable value of the χ2 probability P and if the error bars are too lax then all the models would yield very high value of P close to 100%. Since μ and its error bars for GRB are dependent on the assumption made in determining them from observed afterglow luminosity, we had to develop a normalization method to make the comparison meaningful. The method we developed is as follows:
(a)
We assumed that the error bars represented by the variance σ have errors in the same proportion for all data points in a dataset and thus the error in estimating χ2 using Equation (30) is affected in the same proportion for all models.
(b)
We further assumed that the standard ΛCDM model gives P = 50% and calculated the corresponding χ2 for the degree of freedom for the GRB dataset being analysed.
(c)
We then compared the above χ2 value with that actually found using the already parameterized ΛCDM model with the GRB dataset. The ratio of the two values F was then used as a multiplier to normalize values of χ2 of all the models for the dataset.
(d)
The normalized values of χ2 were then used to determine the χ2 probability P for each model. Those models giving higher P value than 50% can then be considered better than the ΛCDM model for the data set used and vice versa.

4. Test Results

The database used for parameterizing the models in this study is for 580 supernovae Ia (SNe Ia) data points with redshifts ranging from 0.015 to 1.414 as compiled in the Union2 zμ database [27] updated to 2017.
The model fit results are presented in Table 1. Results are placed in three classes for easy comparison. In class A are the models with two adjustable model parameters to fit the data: the first one is R0(≡c/H0), which is the adjustable parameter for all the models; and the second one for example Ωm,0 for the ΛCDM model.
Class B comprises the models that have some fixed model parameter based on the physics of the model or otherwise, such as q0 determined analytically. In this class we determine the parameter b discussed above (as well as R0) by fitting the data to estimate the unaccounted luminosity flux loss in the form (1 + z)b. If b is large then the model is incomplete in fitting the observed data even if the P value is high.
Class C has some of the models in class B but with no adjustable parameter other than R0 to fit the SNe Ia 580 dataset.
In each class the models are arranged in a descending order of χ2 probability P. We have not included in class C those models of class B which yield P less than 5%. If parameter b has a value 0.3 or higher then setting b = 0 for class C yields χ2 value too high to give P ≥ 5%. For example, the Milne model has b = 0.3691 in class B. When we set b = 0, the data fitted χ2 and P values using Equation (9) for the model turn out to be 680.1 and 0.23% respectively.
The table does not include Marosi’s power law model [18], which he has reported to yield μ = azb = 44.109769z0.059883 for the 280 supernovae zμ data he used after removing four outliers with a standard deviation greater than 3σ. For the 580 supernovae data we are using, the fit power law parameters are very close to Marosi’s: a = 44.11 ± 0.02 and b = 0.06126 ± 0.00031 with χ2 = 562.9 and P = 66.6%. The exclusion is because the power law μ = azb does not represent any cosmological physics as Marosi himself has pointed out in his papers.
The parameterized models obtained by fitting the observed SNe Ia data can now be tested by determining the χ2 values and corresponding probability for each parameterized model when higher z supernovae data [28], not used in parameterizing the model, is included. Currently, there is only one observed supernova Ia at higher z value (SN UDS10Wil at z = 1.914 and μ = 45.54 ± 0.39) [28]. When we recalculate the χ2 values for each parametrized model and determine corresponding new probabilities, we notice a decrease in the probability for all but two models. The probability change is shown in the last column of Table 1 as ΔP. For example, the new probability for the ΛCDM model is 67.5% and the old one is 67.3%, hence ΔP = 0.2%.
The parameterized models can also be tested using the zμ numbers determined by many researchers from the observation made on gamma-ray bursts (GRBs). The distance modulus determination for GRBs is not as reliable as for Ia supernovae, nevertheless we should still get some idea about the comparative goodness-of-fit for different models by calculating χ2 and corresponding probability for each model with GRB datasets reported by several researchers [29,30,31,32]. The results are presented in Table 2. The order of models in the rows is the same as in Table 1. The χ2 values have been normalized as discussed near the end of Section 3 above and probabilities determined for the normalized χ2 values. The LLC12 dataset in Table 2 gives the χ2 and P trends that are opposite to the other 3 datasets. This dataset is small comprising only 12 GRBs and data points are in rather a small range of z. It may therefore be considered an outlier.

5. Discussion

Merit of any model is not only in how well it fits the existing observation but in predicting the future observations. Without being comprehensive, we have considered here several cosmological models to explore how well they fit the supernovae Ia observations, which are considered as gold standard for the cosmological model and presented them in Table 1. All models are able to fit the data very well with two adjustable parameter. First parameter indeed is the Hubble constant H0(≡c/R0) and the second parameter is the model dependent parameter, such as Ωm,0 in the ΛCDM model or the luminosity flux correction parameter, such as b in the tired light model. The χ2 probability P for all these models ranges from a low of 43.9% to a high of 67.3% with most models yielding P values greater than 60%. Statistically, therefore none of the model can be rejected, although the highest P value model is none other than the ΛCDM model. Worth noticing is the EdeS-NA model in Class C in the table that yields over 60% χ2 probability with a single fit parameter fit (H0 = 69.05 ± 0.44). This model is also the only model in Table 1, other than the ΛCDM model, that yields an increase in the χ2 probability when the sole high redshift supernova at z = 1.914 is added to calculate χ2 while constraining the model parameter to those already determined using the SNe Ia 580 points dataset. This model therefore may be worth taking seriously when analysing GRB data as well as for predicting future higher redshift supernovae observation with the forthcoming thirty meter telescope (TMT) [33].
The analysis summarized in Table 2 is derived from the gamma-ray burst observation based zμ data and the models whose parameters are already fixed as per R0 and p1 values in Table 1. Models considered are the same as in Table 1 and are in the same order. The LLC12 dataset gives the χ2 and P trends that are opposite to the other 3 datasets. The LLC12 dataset is rather small comprising only 12 GRBs and the data points are in a small range of z(1.48 to 3.8). This dataset may therefore be considered an outlier. We will therefore focus our discussion on the numbers corresponding to the other three datasets in the table. The normalization process described at the end of Section 3 is the reason why all P value in the table for ΛCDM model are 50%.
We notice that all but one of the models with extension b, that had very high value of χ2 probability in Table 1 by adjusting the luminosity flux correction factor b, have less than 25% value, that is, half of the value of 50% for ΛCDM model. This means the luminosity flux correction is redshift dependent and cannot be considered a model constant or parameter. The only b model with higher than 25% value is EdeS-NA.b model. But this model in Table 2 has a higher P value without the b parameter, that is, EdeS-NA model has an even higher P value. So, we see no need to consider EdeS-NA.b model any further. Additionally, EdeS-NA model yields P value close to the ΛCDM value and is the same model which stands out in Table 1 when considering z = 1.914 SNe Ia data. In fact, for the dataset LZ42, the EdeS-NA model yields a slightly higher value of P at 51.71% than the ΛCDM model at normalized 50%. It may be recalled that EdeS-NA model is the Einstein de Sitter model in a nonadiabatic universe [23] with deceleration parameter determined analytically (q0 = −0.4). The third model in Table 2, that is not an EdeS-NA derived model and that yields average P value higher than 30% is the single parameter EDSM-NA model with H0 = 68.35 ± 0.45 and analytically determined deceleration parameter q0 = −0.4.
For completeness sake we have included the Marosi’s power law model with parameters fixed from fitting the SNe Ia 580 dataset. This physics independent model consistently yields fairly high P values.

6. Projections

In this section we present the zμ projections numerically in Table 3 and graphically in Figure 1 using ΛCDM, EdeS-NA, EDSM-NA and Marosi models. In the figure, the lines indicate the four models constrained using the corresponding parameters in Table 1. SNe Ia data points (580 + 1) as well as the LW79 data points (79) have been shown with error bars in the figure. The redshift in the figure is up to z = 15 and in the table up to z = 1096. As the radiation density effects become significant for very high redshifts, extrapolating the models to such high values of the redshift may be considered academic. Nevertheless, the same is true for all the models and thus the comparison may still be meaningful. In addition, the radio telescope measurements of the redshift of the 21-cm (1420 MHz) hydrogen atom line at z ≥ 15 have started to be reported [7]. Radio astronomers can estimate the distances with some confidence using the graphs and tables produced here. Bowman et al. have recently reported observation of an absorption profile of this line centered at 78 MHz in the sky averaged spectrum [7]. This translates into z ≈ 17 spanning over 15 < z < 20. One could see from Table 3 that it corresponds to μ ≈ 51.5.
It should be mentioned that here we have considered only the redshift of the 21-cm absorption line. We do not consider in this paper the anomaly in the measured intensity of this line as it is a topic in itself and has been extensively discussed in literature since the publication of paper by Bowman et al. in Nature in 2018 [7], mostly on the ground that it might suggest new physics in cosmology. For example, Fraser et al. [34] propose soft photon emission by light dark matter as a natural solution to this anomaly; McGaugh [35] predicts that a universe devoid of cold dark matter will exhibit twice as much absorption as in the ΛCDM; Kovetz et al. [36] investigate the hypothesis that Coulomb-type interaction between dark matter and baryons to explain the anomaly; Xiao et al. [37] allow conformal and disformal coupling between dark matter and dark energy in a cosmological model to investigate the anomaly; Houston et al. [38] explain the anomaly using axions and axion like particles in the standard model, as they have the ability to mediate the required cooling process; Venumadhav et al. [39] show that HI spins and Lyman-α photons act as mediators between the radio background and the random thermal motions of the HI atoms, and hence cause extra heating of the intergalactic medium during cosmic dawn, causing higher than normal absorption; Lawson and Zhitnitsky [40] argue that dark matter in the form of macroscopically large nuggets of standard model quarks and antiquarks can explain the anomaly; Pospelov et al. [41] have suggested the enhancement of the Rayleigh-Jeans tail of the cosmic microwave background and show that the resonant oscillation of dark photons into regular photons in the redshift interval 20 < z < 1700 can be invoked as an explanation of the anomalous absorption; and D’Amico et al. [42] provide bounds on the dark-matter annihilations from the analysis of the 21-cm data.
We notice that the calculated distance modulus differences among the selected models for any particular redshift are way less than the μ error bars in the data. The difference in μ values between the ΛCDM model and the EdeS-NA model even at the highest redshift of 1096 in Table 3 is only 0.328. Until the error bars can be reduced significantly, it would be difficult to say which model or models should be rejected. One may wonder if the thirty meter telescope (TMT), due for commissioning in the year 2027, will be able to observe supernovae with high redshifts comparable to the highest GRB redshift with high enough precision to positively decide which model is better [33]!

7. Conclusions

The analysis of various models has shown that while most models can be made to fit the observed data rather satisfactorily with fairly high value of χ2 probability, they tend to fizzle out when they are tested for their predictive power. We have used the redshift—distance modulus (zμ) data for 580 supernovae Ia with 0.015 ≤ z ≤ 1.414 to determine the parameters for each model and then use the parameterized models to see how well each model fits the sole SNe Ia data at z = 1.914 and the GRB data up to z = 8.1. This essentially shows the predictive capability of a model parameterized with z ≤ 1.414 for data obtained at z > 1.414. We find that the standard ΛCDM model gives the highest χ2 probability in all cases but one, albeit with a rather small margin over the next best model—the EdeS-NA model. We have made (zμ) projections up to z = 1096 for the best four models just to show that how little the best two models, based on entirely different assumptions, differ even at such high redshift value which are indeed unmeasurable. The best two models differ in μ only by 0.328 at z = 1096, a tiny fraction of the measurement errors that are in the high redshift datasets. The third best model, the EDSM-NA model, has somewhat higher μ, with a difference of 3.828 at z = 1096, well within the expected measurement errors for such a high redshift objects. It appears unlikely that the measurement errors could be reduced to level where we could say for sure which model predicts the observations better. Other attributes of the model may then decide which model should be preferred.
The ΛCDM model is based on the ad hoc cosmological constant Λ introduced by Einstein in his field equations to prevent the universe from collapsing. There is no physics behind it, and it leads to energy entering the universe from nowhere to keep the energy density corresponding to Λ constant in time. In other words, the presence of Λ makes the universe implicitly nonadiabatic2. The EdeS-NA and EDSM-NA models are based on explicitly nonadiabatic assumption about the universe, that is, the energy leaves or enters a volume of the universe proportional to the energy contained in that volume [23]. This leads to time dependence of the nonadiabatic component of the energy density in an expanding or contracting universe when its energy density is changing. Another attribute of the EdeS-NA and EDSM-NA models is that they are parameterized by just one parameter, the Hubble constant. The second model parameter is not a fit parameter as it is determined analytically in the model. We therefore conclude that the EdeS-NA and EDSM-NA models are viable alternative to the ΛCDM model, even at very high redshift such as those encountered in radio astronomy.

Funding

This research received no external funding.

Acknowledgments

The author wishes to express his sincere gratitude to Professor Martin López-Corredoira and Professor Ethan Vishniac for their critical comments on the author’s previous research which led to the work presented in this paper. He is indebted to five reviewers who provided invaluable comments to improve the quality of the paper.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Zwicky, F. On the red shift of spectral lines through interstellar space. Proc. Natl Acad. Sci. USA 1929, 15, 773–779. [Google Scholar] [CrossRef]
  2. Penzias, A.A.; Wilson, R.W. A measurement of excess antenna temperature at 4080 Mc/s. Astrophys. J. 1965, 142, 419–421. [Google Scholar] [CrossRef]
  3. Marmet, L. On the Interpretation of Spectral Red-Shift in Astrophysics: A Survey of Red-Shift Mechanisms-II. arXiv 2018, arXiv:1801.07582. [Google Scholar]
  4. Vishwakarma, R.G.; Narlikar, J.V. Is it no longer necessary to test cosmologies with type Ia supernovae? Universe 2018, 4, 73. [Google Scholar] [CrossRef]
  5. Prichard, J.R.; Loeb, A. 21-cm cosmology. Rep. Prog. Phys. 2012, 75, 086901. [Google Scholar] [CrossRef] [PubMed]
  6. Boyarsky, A.; Lakubovskyi, D.; Ruchayskiy, O.; Rudakovskyi, A.; Valkenburg, W. 21-cm observation and warm dark matter models. arXiv 2019, arXiv:1904-03097. [Google Scholar]
  7. Bowman, J.D.; Rogers, A.E.E.; Mousalve, R.A.; Mozdzen, T.J.; Mahesh, N. An absorption profile centred at 78 MHz in the sky-averaged spectrum. Nature 2018, 555, 67–70. [Google Scholar] [CrossRef]
  8. López-Corredoira, M. Test and problems of the standard model in cosmology. Found. Phys. 2017, 47, 711–768. [Google Scholar] [CrossRef]
  9. Orlov, V.V.; Raikov, A.A. Cosmological tests and evolution of extragalactic objects. Astron. Rep. 2016, 60, 477–485. [Google Scholar] [CrossRef]
  10. Berton, G.; Hooper, D. Particle dark matter: Evidence. Phys. Rept. 2005, 405, 279–390. [Google Scholar] [CrossRef]
  11. Del Popolo, A. Nonbaryonic dark matter in cosmology. Int. J. Mod. Phys. D 2014, 23, 1430005. [Google Scholar] [CrossRef]
  12. Del Popolo, A.; Le Delliou, M. Small scale problems of the ΛCDM model: A short review. Galaxies 2017, 5, 17. [Google Scholar] [CrossRef]
  13. Mortonson, M.J.; Weinberg, D.H.; While, M. Dark energy: A short review. arXiv 2013, arXiv:1401.0046. [Google Scholar]
  14. Ryden, B. Introduction to Cosmology; Cambridge University Press: Cambridge, UK, 2017. [Google Scholar]
  15. Gupta, R.P. Mass of the universe and the redshift. Int. J. Astron. Astrophys. 2018, 8, 68–78. [Google Scholar] [CrossRef]
  16. Gupta, R.P. Static and dynamic components of the redshift. Int. J. Astron. Astrophys. 2018, 8, 219–229. [Google Scholar] [CrossRef]
  17. Crawford, D. A problem with the analysis of type Ia Supernovae. Open Astron. 2017, 26, 111–119. [Google Scholar] [CrossRef]
  18. Marosi, L.A. Hubble diagram test of 280 supernovae redshift data. J. Mod. Phys. 2014, 5, 29–33. [Google Scholar] [CrossRef]
  19. Vishwakarma, R.G. A scale-invariant, Machian theory of gravitation and electrodynamics unified. Int. J. Geom. Meth. Mod. Phys. 2018, 15, 1850178. [Google Scholar] [CrossRef]
  20. Zaninetti, L. On the number of galaxies at high redshift. Galaxie 2015, 3, 129–155. [Google Scholar] [CrossRef]
  21. Zaninetti, L. Padé Approximant and Minimax Rational Approximation in Standard Cosmology. Galaxies 2016, 4, 4–24. [Google Scholar] [CrossRef]
  22. Brynjolfsson, A. Redshift of photons penetrating a hot plasma. arXiv 2004, arXiv:astro-ph/0401420. [Google Scholar]
  23. Gupta, R.P. SNe Ia redshift in a nonadiabatic universe. Universe 2018, 4, 104. [Google Scholar] [CrossRef]
  24. Peebles, P.J.E. Principles of Physical Cosmology; Princeton University Press: Princeton, NJ, USA, 1993. [Google Scholar]
  25. Press, W.H.; Teukolsky, S.A.; Vetterling, W.T.; Flannery, B.P. Numerical Recipes in C—The Art of Scientific Computing, 2nd ed.; Cambridge University Press: Cambridge, UK, 1992. [Google Scholar]
  26. Walker, J. Chi-Square Calculator. 2018. Available online: https: //www.fourmilab.ch/rpkp/experiments/analysis/chiCalc.html (accessed on 30 April 2019).
  27. Amanullah, R.; Lidman, C.; Rubin, D.; Aldering, G.; Astier, P.; Barbary, K.; Burns, M.S.; Conley, A.; Dawson, K.S.; Deustua, S.E.; et al. Spectra and Hubble space telescope light curves of six type Ia supernovae at 0.511 < z < 1.12 and the UNION2 Compilation. Astrophys. J. 2010, 716, 712–738. [Google Scholar]
  28. Jones, D.O.; Rodney, S.A.; Riess, A.G.; Mobasher, B.; Dahlen, T.; McCully, C.; Frederiksen, T.F.; Casertano, S.; Hjorth, J.; Keeton, C.R.; et al. The discovery of the most distant known Type Ia supernova at redshift 1.914. Astrophys. J. 2013, 768, 166. [Google Scholar] [CrossRef]
  29. Liu, J.; Wei, H. Cosmological models and gamma-ray bursts calibrated by using Padé method. Gen. Relativ. Gravit. 2015, 47, 141. [Google Scholar] [CrossRef]
  30. Cardone, V.F.; Capozziello, S.; Dainotti, M.G. An updated gamma-ray bursts Hubble Diagram. Mon. Not. R. Astron. Soc. 2009, 400, 775–790. [Google Scholar] [CrossRef]
  31. Liang, N.; Zhang, S.N. Cosmology-independent distance moduli of 42 gamma-ray bursts between redshift of 1.44 and 6.60. AIP Conf. Proc. 2008, 1065, 367. [Google Scholar]
  32. Lin, H.-N.; Li, X.; Chang, Z. Model independent distance calibration of high-redshift gamma-ray bursts and constrain on ΛCDM model. Mon. Not. R. Astron. Soc. 2016, 455, 2131–2138. [Google Scholar] [CrossRef]
  33. Pandey, S.B. Core-collapse supernovae and gamma-ray bursts in TMT era. J. Astrophys. Astron. 2013, 34, 157–173. [Google Scholar] [CrossRef]
  34. Fraser, S.; Hector, A.; Hutsi, G.; Kannike, K.; Marzo, C.; Marzola, L.; Racioppi, A.; Raidal, M.; Spethmann, C.; Vaskonen, V.; et al. The EDGES 21 cm Anomaly and Properties of Dark Matter. Phys. Letts. B 2018, 785, 159–164. [Google Scholar] [CrossRef]
  35. McGaugh, S.S. Predictions for the sky averaged depth of the 21 cm absorption signal at high redshift in cosmologies with and without nonbaryonic cold dark matter. Phys. Rev. Lett. 2018, 121, 081305. [Google Scholar] [CrossRef]
  36. Kovetz, E.D.; Poulin, V.; Gluscevic, V.; Boddy, K.K.; Barkana, R.; Kamionkowski, M. Tighter limits on dark matter explanations on the anomalous EDGES 21 cm signal. Phys. Rev. D 2018, 98, 103529. [Google Scholar] [CrossRef]
  37. Xiao, L.; An, R.; Zhang, L.; Yue, B.; Xu, Y.; Wang, B. Can conformal and disformal coupling between dark sectors explain the EDGES 21-cm anomaly? Phys. Rev. D 2018, 99, 023528. [Google Scholar] [CrossRef]
  38. Houston, N.; Li, C.; Li, T.; Yang, Q.; Zhang, X. Natural explanation for 21 cam absorption signal via axion-induced cooling. Phys. Rev. Lett. 2018, 121, 111301. [Google Scholar] [CrossRef]
  39. Venumadhav, T.; Dai, L.; Kaurov, A.; Zaldarriaga, M. Heating of the intergalactic medium by the cosmic microwave background during cosmic dawn. Phys. Rev. D 2018, 98, 103513. [Google Scholar] [CrossRef]
  40. Lawson, K.; Zhitnitsky, A.R. The 21 cm absorption line and the axion quark nugget dark matter model. Phys. Dark Universe 2018, 24, 100295. [Google Scholar] [CrossRef]
  41. Pospelov, M.; Pradler, J.; Ruderman, J.T.; Urbano, A. Room for new physics in the Rayleigh-Jeans tail of the cosmic microwave background. Phys. Rev. Lett. 2018, 121, 031103. [Google Scholar] [CrossRef]
  42. D’Amico, G.; Panci, P.; Strumia, A. Bounds on dark-matter annihilation from 21-cm data. Phys. Rev. Lett. 2018, 121, 011103. [Google Scholar] [CrossRef]
1
Readers with their own models may contact the author to include their models in his study and communicate back the results and possibly include the same in his future research.
2
This view may be in conflict with the thermodynamic notion developed, for example, in the following papers: arXiv:1603.08299; arXiv:1808.03825; arXiv:1812.03540 and arXiv:1902.06651.
Figure 1. Four selected models fitted to the SNe Ia 580 dataset with projections and gamma-ray burst data.
Figure 1. Four selected models fitted to the SNe Ia 580 dataset with projections and gamma-ray burst data.
Universe 05 00102 g001
Table 1. SNe Ia 580 dataset points fit for different models. R0 is in Mpc and H0(≡c/R0), is in km/s/Mpc. R0 is the first parameter and p1, explicitly shown in the last but one column, is the second parameter determined by fitting the data. There is no second parameter for the models in the last 3 rows. DOF is the degrees of freedom, P is the χ2 probability and q0 is the deceleration parameter either derived from the fitted parameters (first 3 model rows) or determined analytically. ΔP is the increase or decrease in the χ2 probability when the supernova with z = 1.914 is added to the dataset without any change in the model parameters determined by the SNe Ia 580 dataset.
Table 1. SNe Ia 580 dataset points fit for different models. R0 is in Mpc and H0(≡c/R0), is in km/s/Mpc. R0 is the first parameter and p1, explicitly shown in the last but one column, is the second parameter determined by fitting the data. There is no second parameter for the models in the last 3 rows. DOF is the degrees of freedom, P is the χ2 probability and q0 is the deceleration parameter either derived from the fitted parameters (first 3 model rows) or determined analytically. ΔP is the increase or decrease in the χ2 probability when the supernova with z = 1.914 is added to the dataset without any change in the model parameters determined by the SNe Ia 580 dataset.
ModelClassR0 ± 95% CLH0 ± 95% CLp1 ± 95% CLχ2DOFP (%)q0Eqn.p1ΔP (%)
ΛCDMA4283 ± 4070.00 ± 0.650.2776 ± 0.0377562.257867.3−0.58367Ωm,00.2
EdeS-NA.q0A4304 ± 4169.65 ± 0.66−0.4776 ± 0.0616563.757865.7−0.4824q0−0.2
EDSM-NA.q0A4317 ± 4369.45 ± 0.68−0.5767 ± 0.0743566.657862.5−0.5828q0−0.8
PlasmaA4324 ± 4269.33 ± 0.671.369 ± 0.068568.057860.9017β−0.9
EdeS-NA.bB4302 ± 4269.69 ± 0.670.08415 ± 0.06761563.657865.8−0.424b−0.3
Crawford.bB4311 ± 4269.54 ± 0.671.44 ± 0.069565.057864.3013b−0.6
EDSM-NA.bB4321 ± 4269.38 ± 0.670.1371 ± 0.06786567.357861.7−0.429b−0.9
Milne.bB4324 ± 4269.33 ± 0.660.3691 ± 0.0679568.057860.909b−0.9
Tired.bB4324 ± 4269.33 ± 0.661.369 ± 0.068568.057860.9010b−0.9
EdeS.bB4327 ± 4269.29 ± 0.660.8558 ± 0.0679568.657860.20.58b−1.2
EDSM.bB4333 ± 4269.19 ± 0.670.525 ± 0.0680570.357858.2−0.411b−1.4
Vishwa.bB4364 ± 4368.70 ± 0.670.1589 ± 0.0687582.657843.9016b−2.7
EdeS-NAC4342 ± 2869.05 ± 0.44None569.557960.3−0.424NA0.2
EDSM-NAC4386 ± 2968.35 ± 0.45None582.7057944.9−0.429NA−0.1
VishwaC4439 ± 2967.54 ± 0.44None603.457923.4016NA−1
Table 2. χ2 test of the parameterized models with four datasets. Models considered are the same as in Table 1 and are in the same order. The LLC12 dataset gives the χ2 and P trends that are opposite to the other 3 datasets. The dataset is small comprising only 12 gamma ray bursts (GRBs) and data points are in rather a small range of z. It may therefore be considered an outlier.
Table 2. χ2 test of the parameterized models with four datasets. Models considered are the same as in Table 1 and are in the same order. The LLC12 dataset gives the χ2 and P trends that are opposite to the other 3 datasets. The dataset is small comprising only 12 gamma ray bursts (GRBs) and data points are in rather a small range of z. It may therefore be considered an outlier.
Model DatasetLW79CCD69LZ42LLC12
Dataset Reference[29] [30][31][32]
Redshift Range1.44 ≤ z ≤ 8.10.17 ≤ z ≤ 6.61.44 ≤ z ≤ 6.61.48 ≤ z ≤ 3.8
GRB Data Points79694212
Normalization Factor2.24248820.49763311.11085280.2639186
Normalized Parametersχ2P in %χ2P in %χ2P in %χ2P in %
ΛCDM76.334350.0066.334550.0039.335350.009.341850.00
EdeS-NA.q081.7833.3273.4027.6442.4036.798.79855.13
EDSM-NA.q092.4411.0785.206.6151.6710.228.40658.92
Plasma95.697.3288.633.6954.736.038.35059.46
EdeS-NA.b82.1032.4373.7526.7042.6635.748.78055.31
Crawford.b85.9522.7178.1816.5145.7724.508.60057.04
EDSM-NA.b94.548.5187.384.7953.657.308.37059.27
Milne.b95.697.3288.633.6954.746.028.35059.46
Tired.b95.697.3288.633.6954.746.028.35059.46
EdeS.b97.795.5190.822.8656.784.128.33059.66
EDSM.b102.822.6395.991.1661.721.528.27960.16
Vishwa.b135.780.00128.890.0095.930.008.43258.670
EdeS-NA77.7748.6067.6848.8139.9551.719.09461.32
EDSM-NA83.3331.8973.7529.5744.0634.348.64665.45
Vishwa112.280.66103.310.3771.570.218.13170.15
Marosi79.6339.6269.7738.4841.3241.278.90054.16
Table 3. Calculated values of distance moduli versus redshift for the four selected models parameterized using the SNe Ia 580 dataset.
Table 3. Calculated values of distance moduli versus redshift for the four selected models parameterized using the SNe Ia 580 dataset.
zµ-ΛCDMµ-EdeS-NAµ-EDSM-NAµ-Marosizµ-ΛCDMµ-EdeS-NAµ-EDSM-NAµ-Marosi
0.10038.31938.33138.33938.30710.96550.37050.56851.17951.080
0.11038.53338.54438.55038.52412.02350.59650.80251.45951.369
0.12038.74938.75738.76338.74213.18350.82151.03551.73851.659
0.13238.96538.97238.97638.96114.45451.04651.26652.01751.952
0.14539.18339.18939.19139.18115.84951.26951.49652.29652.246
0.15839.40339.40639.40739.40317.37851.49251.72452.57552.541
0.17439.62439.62539.62539.62619.05551.71451.95252.85352.838
0.19139.84639.84539.84339.85020.89351.93552.17853.13153.137
0.20940.07040.06740.06340.07622.90952.15552.40453.40953.438
0.22940.29640.29040.28540.30225.11952.37452.62853.68753.740
0.25140.52340.51540.50840.53027.54252.59352.85153.96454.044
0.27540.75240.74240.73340.76030.20052.81153.07454.24154.350
0.30240.98340.97040.96040.99033.11353.02853.29554.51754.658
0.33141.21641.20041.18841.22236.30853.24553.51654.79454.967
0.36341.45041.43241.41941.45639.81153.46153.73555.07055.278
0.39841.68641.66641.65141.69043.65253.67753.95455.34555.591
0.43741.92441.90241.88641.92647.86353.89154.17255.62055.905
0.47942.16442.14042.12242.16352.48154.10554.38955.89556.222
0.52542.40542.37942.36142.40257.54454.31954.60656.17056.540
0.57542.64842.62142.60242.64263.09654.53254.82156.44456.860
0.63142.89342.86442.84542.88369.18354.74555.03656.71757.181
0.69243.13843.10943.09043.12675.85854.95755.25156.99157.505
0.75943.38543.35643.33843.37083.17655.16855.46557.26457.830
0.83243.63343.60443.58743.61591.20155.37955.67857.53658.158
0.91243.88143.85343.83943.862100.00055.59055.89057.80958.487
1.00044.13044.10444.09344.110109.64855.80056.10358.08158.818
1.09644.37944.35644.35044.360120.22656.01056.31458.35259.150
1.20244.62844.61044.60844.611131.82656.21956.52558.62459.485
1.31844.87844.86344.86844.863144.54456.42856.73658.89459.822
1.44545.12745.11845.13045.117158.48956.63656.94659.16560.160
1.58545.37545.37345.39445.372173.78056.84457.15559.43560.501
1.73845.62345.62845.66045.629190.54657.05257.36559.70560.843
1.90545.87145.88345.92745.887208.93057.26057.57359.97561.187
2.08946.11746.13846.19646.147229.08757.46757.78260.24461.533
2.29146.36346.39346.46746.408251.18957.67457.99060.51361.882
2.51246.60746.64746.73846.670275.42357.88058.19860.78262.232
2.75446.85146.90147.01146.934301.99558.08758.40561.05062.584
3.02047.09347.15447.28547.200331.13158.29358.61261.31862.938
3.31147.33447.40647.56047.467363.07858.49858.81961.58663.294
3.63147.57447.65747.83547.736398.10758.70459.02561.85463.652
3.98147.81347.90748.11248.006436.51658.90959.23162.12164.012
4.36548.05148.15648.38948.277478.63059.11459.43762.38864.374
4.78648.28848.40348.66748.551524.80759.31959.64362.65564.739
5.24848.52348.64948.94548.825575.44059.52459.84862.92165.105
5.75448.75848.89449.22449.101630.95759.72860.05363.18765.473
6.31048.99149.13749.50349.379691.83159.93260.25863.45365.844
6.91849.22449.37949.78249.659758.57860.13660.46363.71966.216
7.58649.45549.62050.06149.940831.76460.34060.66763.98466.591
8.31849.68549.85950.34150.222912.01160.54460.87164.24966.968
9.12049.91450.09750.62050.5061000.00060.74761.07664.51467.347
10.00050.14250.33350.90050.7921096.47860.95161.27964.77967.728

Share and Cite

MDPI and ACS Style

Gupta, R.P. Weighing Cosmological Models with SNe Ia and Gamma Ray Burst Redshift Data. Universe 2019, 5, 102. https://doi.org/10.3390/universe5050102

AMA Style

Gupta RP. Weighing Cosmological Models with SNe Ia and Gamma Ray Burst Redshift Data. Universe. 2019; 5(5):102. https://doi.org/10.3390/universe5050102

Chicago/Turabian Style

Gupta, Rajendra P. 2019. "Weighing Cosmological Models with SNe Ia and Gamma Ray Burst Redshift Data" Universe 5, no. 5: 102. https://doi.org/10.3390/universe5050102

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

Article Metrics

Back to TopTop