The following article is Open access

A Distance-deviation Consistency and Model-independent Method to Test the Cosmic Distance–Duality Relation

, , , , and

Published 2021 March 10 © 2021. The Author(s). Published by the American Astronomical Society.
, , Citation Chichun Zhou et al 2021 ApJ 909 118 DOI 10.3847/1538-4357/abc9bf

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/909/2/118

Abstract

A distance-deviation consistency and model-independent method to test the cosmic distance–duality relation (CDDR) is provided. This method is worthy of attention for two reasons. First, a distance-deviation consistency method is used to pair subsamples: instead of pairing subsamples with redshift deviations smaller than a value of, say, $\left|{\rm{\Delta }}z\right|\lt 0.005$. The redshift deviation between subsamples decreases with the redshift to ensure the distance deviation stays the same. The method selects more subsamples at high redshift, up to z = 2.16, and provides 120 subsample pairs. Second, the model-independent method involves the latest data set of 1048 SNe Ia and 205 strong gravitational lensing systems (SGLS), which are used to obtain the luminosity distances DL and the ratio of angular diameter distance DA, respectively. With the model-independent method, parameters of the CDDR, the SNe Ia light curve, and the SGLS are fitted simultaneously. The result shows that $\eta ={0.047}_{-0.151}^{+0.190}$ and CDDR is validated at the 1σ confidence level in the form of $\tfrac{{D}_{L}}{{D}_{A}}{\left(1+z\right)}^{-2}=1+\eta z$.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The cosmic distance duality relation (CDDR), also called Etherington's reciprocity relation (Etherington 1933), plays an important role in modern cosmology, especially in galaxy observations (Cunha et al. 2007; Mantz et al. 2010), cosmic microwave background (CMB) radiation observations (Komatsu et al. 2011), and the gravitational lensing (Ellis 2007). The CDDR reads

Equation (1)

where DL is the luminosity distance, DA is the angular diameter distance, and z is the redshift. The CDDR is valid for all cosmological models based on Riemannian geometry. The basis of this relation is that the number of photons is conservative and photons travel along the null geodesic in a Riemannian spacetime (Ellis 2007).

The validity of the CDDR has been explored widely for the past decades, because any deviation of CDDR may trigger new physics. Uzan et al. (2004) investigates the possible deviation from the CDDR by analyzing the measurements of SZE and X-ray emission data of galaxy clusters and reports that the parameter $\eta ={0.89}_{-0.03}^{+0.04}$ and is at the 1σ confidence level. Holanda et al. (2011) takes more parameterized forms of η and found no departure from the CDDR. Nair et al. (2011), Basset & Kunz (2004), Holanda et al. (2010, 2011), Cao & Liang (2011), and Meng et al. (2012) used DL directly from SNe Ia to test the CDDR. Hu & Wang (2018) and Melia (2018) used the DL form SNe Ia with the Rh = ct cosmology model to test the CDDR and compared the model with the ΛCDM model. Basset & Kunz (2004) found a 2σ violation of the CDDR using the luminosity distances DL from SNe Ia and the angular diameter distance DA from FRIIb radio galaxies. Räsänen et al. (2016) used CMB anisotropy to test the CDDR. The CDDR is also important in studying cosmic opacity (Lv & Xia 2016; Hu et al. 2017).

To test the CDDR, many DL and DA pairs at the same redshift z need to be provided simultaneously. In principle, the two distances should neither be correlated nor based on any cosmology models. That is, a model-independent method and a quality and quantity collection of sample pairs are important. Conventionally, in determining DL , the method of Standard Candles (e.g., SNe Ia and GRB; Wang & Dai 2006; Wang et al. 2007; Wang & Dai 2006; Wang et al. 2009, 2015; Wang & Dai 2009; Tu & Wang 2018; Wang & Wang 2019) plays a prominent part. However, the method of Standard Candles is model-dependent, i.e., a special cosmology model is used in calibrating the light-curve parameters. For example, Suzuki et al. (2012) used the cold dark matter (CDM), wCDM, and owCDM models to fit the parameters of Union2.1 SNe Ia and to constrain the cosmology parameters. In determining DA , the method of using the Sunyaev–Zel'dovich effect (SZE) and X-ray observations (Cavaliere & Fusco-Fermiano 2011; Sunyaev & Zel'dovich 1972; Bonamente et al. 2006) is important in finding DA from galaxy clusters. DA can also be obtained from ultra-compact radio sources (Li & Lin 2018) and baryon acoustic oscillations (BAOs; Wu et al. 2015).

There are model-independent methods. Liao et al. (2016) introduced a new method to test the CDDR based on strong gravitational lensing systems and SNe Ia. In their work, they constrain η, the parameter of the SNe Ia light curve, and the parameter of the SGLS simultaneously. Räsänen et al. (2016) uses the temperature–redshift relation of CMB to test the CDDR, in their work, a flat FRW universe is assumed. Ruan et al. (2018) use a similar model-independent method with the SGLS, the SNe Ia, and the HII galaxy Hubble diagram to test the CDDR. To avoid the effect of the cosmic opacity, Liao (2019) uses the DL from the gravitational wave signals and the ratio of DA from the SGLS (for details, see below) with a model-independent method, which is proposed in Liao et al. (2016). These model-independent methods show that the CDDR is valid in the given redshift range, say, z < 1.0. Recently, Renzi et al. (2020) proposed the use of the strongly lensed supernovae as a self-sufficient tool to test the CDD relation.

In this paper, we provide a distance-deviation consistency and model-independent method to test the CDDR. The distance-deviation consistency method pairs subsamples with redshift deviation that decreases with the redshift to ensure the distance deviation stays the same. Because the distance grows nonlinearly with redshift, the larger the redshift is, the smaller the redshift deviation is between two sources with the same distance deviation. The latest data set of SNe Ia with 1048 samples and strong gravitational lensing system (SGLS) with 205 samples are involved and the distance-deviation consistency method enables us to take full advantage of the data: the method selects more subsamples at high redshift, up to z = 2.16, and provide 121 subsample pairs. With the model-independent method, parameters of the CDDR, the SNe Ia light curve, and the SGLS are fitted simultaneously.

This paper is organized as follows. In Section 2, we introduce the latest data set of the SNe Ia with 1048 samples and the SGLS with 205 samples. In Section 3, we explain a distance-deviation consistency method to pair subsamples and describe the method of statistical analysis. The numerical results are shown. Conclusions and discussions are given in Section 4.

2. Data

In this section, we describe two sets of data suitable for testing the CDDR, one based on the redshift and the light curve of the SNe Ia, from which we can obtain DL (z), and the other base on the observational velocity dispersion of the lens galaxy and redshifts of the strong gravitational lensing system (SGLS), from which we can obtain the ratio of DA (z).

2.1. The Pantheon SNe Ia Sample

In this section, we introduce the contents of the Pantheon sample (Scolnic et al. 2018). The Pantheon sample consisting of a total of 1048 SNe Ia in the range of 0.01 < z < 2.3 is constructed by a subset including 279 SNe Ia (0.03 < z < 0.68) from the Pan-STARRS1 (PS1) Medium Deep Survey, the Sloan Digital Sky Survey (SDSS), the SNLS, the various low-z, and Hubble Space Telescope (HST) samples.

On the one hand, the luminosity distances DL (z) can be determined accurately by multiple light-curve fitters (e.g., Jha et al. 2007; Guy et al. 2010; Burns et al. 2011; Mandel et al. 2011). From the phenomenological point of view, the distance modulus, μ, of an SNe Ia can be extracted from its light curve. In a modified version of the Tripp formula (Tripp 1998), the SALT2 light-curve fit parameters are transformed into distances:

Equation (2)

where mB is the apparent magnitude, M is the absolute B-band magnitude of a fiducial SNe Ia with x1 = 0 and c = 0, ΔM and ΔB are distance corrections based on the mass of the host galaxy of the SNe Ia and predicted biases from simulations respectively. α and β are light-curve parameters of relations between the luminosity and the stretch and between the luminosity and the color, respectively. Moreover, ΔM in the Equation (2) can be written in the form (Scolnic et al. 2018)

Equation (3)

where γ, mstep, and τ are coefficients to be determined (Scolnic et al. 2018). On the other hand, one assumes that the SNe Ia with identical color, shape, and galactic environment have on average the same intrinsic luminosity for all redshifts (Betoule et al. 2014). According to the definition of the distance modulus, it can be written as

Equation (4)

By using Equations (2) and (4), we can obtain the luminosity distances DL (z) for the Pantheon SNe Ia Sample.

2.2. The Strong Gravitational Lensing System Sample

For the new SGLS sample, we used Amante et al. (2020), which contains 205 SGLS. This sample was generated from several survey projects (e.g., the SLACS, the CASTLES survey, the BELLS, the LSD, etc.). In an SGLS, the light is bent by massive bodies (e.g., galaxy or galaxy cluster) which is predicted by the general theory of relativity. The SGLS is a powerful astrophysical tool, which has been rapidly developed in recent years, to explore the universe and galaxy, especially dark energy (Biesiada 2006; Biesiada et al. 2010; Jullo et al. 2010; Cao et al. 2012, 2015; Magaña et al. 2015, 2018), the CDDR (Liao et al. 2016; Liao 2019), the cosmic acceleration (Tu et al. 2019), calibrating the standard candles (Wen & Liao 2020), and cosmological model comparison (Melia et al. 2015; Leaf & Melia 2018; Tu et al. 2019). In an SGLS, a single galaxy acting as the lens, the Einstein radius depends on three parameters: the angular distance to the source and between the lens and the source, and the mass distribution within the lensing galaxy. A singular isothermal sphere (SIS) model (Ratnatunga et al. 1999) is used to describe the lens galaxy's mass distribution. The ratio of the angular diameter distances between lens and source and between observer and source can be obtained from a special physical model (e.g., SIS model). Because these distances depend on the cosmological metric, the ratio can be used to constrain cosmological parameters.

In an SIS model of the SGLS, the distance ratio RA (zl , zs ) (${D}_{l}^{A}s/{D}_{s}^{A}$) is related to observables in the following way (Biesiada et al. 2010),

Equation (5)

where c is the speed of light, θE is the Einstein radius, and σSIS is the velocity dispersion of the star in the periphery of the lens galaxy due to the lens mass distribution in the SIS model. In general, σSIS does not equal the observed stellar velocity dispersion σ0 (White & Davis 1996). To express the difference, researchers use a phenomenological free parameter fe defined by the relation σSIS = fe σ0 (Kochanek 1992; Ofek et al. 2003; Cao et al. 2012), where (0.8)1/2 < fe < (1.2)1/2. In this case, the systematic error is caused by σ0 as σSIS, the deviation of the SIS model, the effects of secondary lenses (nearby galaxies), the line-of-sight contamination (Ofek et al. 2003), etc. The uncertainty of Equation (5) can be written by (Liao et al. 2016)

Equation (6)

In Equation (6), ${\delta }_{{\sigma }_{{sis}}}$ and ${\delta }_{{\theta }_{E}}$ are the fractional uncertainties of σsis and θE , respectively. To test the CDDR, the left term of Equation (ref5), RA (zl , zs ), should be expressed as the ratio of luminosity distance, ${D}_{{ls}}^{A}/{D}_{s}^{A}$. We transform RA (zl , zs ) into the ratio of comoving distance (DC ) or dimensionless distance H0 DC /c. In a flat space, the dimensionless distance satisfies

Equation (7)

By using the equation,

Equation (8)

RA (zl , zs ) can be written as

Equation (9)

In a nonflat space, the expression of RA (zl , zs ) is more complicated; one can refer to Räsänen et al. (2015). Fortunately, most cosmological tests today support a flat cosmic space (Planck Collaboration et al. 2016). In this work, we test the CDD relation in a case where we assume that spacetime is flat.

To test the CDDR, Equation (1) is rewritten by the parameterization of the deviation:

Equation (10)

Combining Equations (9) and (10), RA (zl , zs ) can be written as

Equation (11)

By using Equation (4), $\tfrac{{d}_{L}({z}_{l})}{{d}_{L}({z}_{s})}$ of Equation (11) can be rewritten as

Equation (12)

where the absolute magnitude of the SNe Ia subsample is offset, zl is the redshift of the lens, and zs is the redshift of the sources.

3. Method and Results

In this section, we introduce a distance-deviation consistency method of data selection and show the result.

3.1. Method of Data Selection: A Distance-deviation Consistency Method

To test the CDDR, DL and DA at the same redshift z need to be provided simultaneously. However, redshifts of subsamples from the SGLS and the SNe Ia are different. To take full advantage of the data, one needs to pair the subsamples efficiently. In this section, we provide a distance-deviation consistency method to pair subsamples, which outperforms the conventional method.

The redshift-difference of subsample pairs in this work is not fixed, and it decreases with the redshifts to ensure the distance deviation of the sources stays the same. The relation between Δz and coordinate distance with a cosmology model reads

Equation (13)

where Rh = ct and flat ΛCDM model with Ωm = 0.31 (Planck Collaboration et al. 2016) are used. By setting Δdc /dc equal to 5% and combining Equation (13), Δz(z) can be calculated:

Equation (14)

where the distance formulas of the two cosmology model, ΛCDM and Rh = ct are used. The Rh = ct cosmological model was proposed in Melia (2007). In the Rh = ct universe, the luminosity distance DL can be written by

Equation (15)

This model is a Friedmann–Robertson–Walker (FRW) cosmological model, which is obeying the cosmological principle and Weyl's postulate (Melia 2007; Melia & Shevchuk 2012). In an Rh = ct universe, space expands at a constant rate, rather than an accelerating rate. Theoretically, there are some controversies with this model that focus on the zero active mass condition ρ+3p = 0 (for more details, see Kim et al. 2016; Melia 2016, 2017). In terms of the fitting of observational data, the model performs relatively well. The creator of this model himself and his collaborators have done extensive work comparing it with the standard model using many different types of observations, and they have found that the Rh = ct model is better than the Standard Model (e.g., Melia 2013; Wei et al. 2015; Fatuzzo & Melia 2017; Melia et al. 2018; Melia 2019). Additional work also illustrates this point (Yu & Wang 2014; Yuan & Wang 2015). Though there is also a lot of work that argues against this model (e.g., Tutusaus et al. 2016; Shafer 2015). Despite the theoretical controversy, this model has gained some support for the data, and we can use this model jointly with the standard model to select the data.

For a given SGLS subsample, the redshift–dimensionless distance relation, Equation (14), is used to acquire a suitable redshift deviation interval. Then, the subsamples of the SNe Ia with redshifts within the interval are selected as the candidates. Finally, the subsample with the smallest redshift deviation is selected.

Our method outperforms the conventional method in two ways: (1) the information of subsamples at high redshifts is conserved. Pairing subsamples with a slight difference of redshift is a simple and commonly used method, for example, Δz = 0.005 (Holanda et al. 2010, 2012, 2016; Li et al. 2011; Nair et al. 2011), Δz = 0.006 (Goncalves et al. 2012), and Δz = 0.003 (Liao 2019). Gaussian process (GP) reconstruction is also a usable method (Ruan et al. 2018; Zhang 2019), and the linear interpolation method has been chosen by some researchers as well (Liang et al. 2013; Hu & Wang 2018). These methods reduce the systematic error to an extent but do not consider the distance-deviation consistency. The relation between the distance and the redshift is nonlinear: the same redshift deviation at higher redshift has a smaller distance deviation. For example, some researchers set Δz = 0.005, with the uncertainty of dimensionless distance being 5% at z ∼ 0.1, which is the minimum redshift of the subsample we select, but smaller than 1% at z ∼ 1 for selecting data with a general cosmology model; thus, the selecting uncertainty is not the consistency of distance deviation. The selecting uncertainty of their methods are all ignored, and if they did take them into account in their fitting, they had to introduce a cosmological model so that their methods were no longer model-independent. In our approach, although two cosmological models are introduced, they are used to jointly pick the data, breaking the dependence on a single cosmological model when picking the data, and are not introduced into the χ2 function. In other words, the final parameter fit results are independent of the cosmological model. Cao et al. (2017) used a similar method for selecting the data with a ΛCDM model to fit the parameters of the ultra-compact radio quasars. At high redshift, in general, subsamples are sparse and matching pairs of subsamples is difficult if a fixed Δz is used. The distance-deviation consistency method selects more subsamples at high redshifts.

(2) More subsample pairs are selected. For example, in the work of Liao et al. (2016), they select only 60 pairs of samples to test the CDDR. According to Figure 1, at z = 0.11, the previously used Δz = 0.005 and the curve of Δd/d in this method intersect. So Δd/d = 5% is the maximum allowable uncertainty of dimensionless distance. In other words, if the error of selecting exceeds 5%, the statistical error of our results must be higher than that of previous work. If the error is significantly lower than 5%, the number of data pairs we choose will not be significantly improved. With the method of fixed Δd/d = 5%, the number of pairs can reach 68. The data utilization increases by 13.3%. In this work, we have obtained 120 pairs of samples with redshifts from 0.11 to 2.16.

Figure 1.

Figure 1. The relation between Δd/d and z, the red line and the blue line, indicate the relationship between the relative error of d and the redshift in Rh = ct model and ΛCDM model, respectively. The black line indicates the scheme we used.

Standard image High-resolution image

A comparison between our method with a fixed Δd/d and the conventional method with a fixed Δz are shown in Figures 1 and 2. In Figure 1, the deviation, Δd/d, falls rapidly with the increase of redshift with Δz fixed. The main advantage of a fixed Δd/d is shown in Figure 2. To conclude, on the one hand, our method maintains more information about subsamples at high redshifts. On the other hand, our method selects more subsample pairs and thus reduces the systematic error.

Figure 2.

Figure 2. The relation between Δz and z.

Standard image High-resolution image

3.2. Method of Statistical Analysis

To determine the parameters, we minimize the χ2 function. By using Equations (5) and (11), the χ2 function can be written as

Equation (16)

where ${\sigma }_{{R}^{A}({z}_{l},{z}_{s})}$ is the uncertainty of the SGLS with the SIS model, ${\sigma }_{{R}_{0}^{A}({z}_{l},{z}_{s})}$ is the error caused by the uncertainty of the distance modulus of SNe Ia, which is related to the uncertainty of observed data (e.g., mB ), and σsel is the uncertainty of data selection, which is related to artificial selection. By using Equation (11), the uncertainty of data selection can be written as

Equation (17)

where $\tfrac{{\rm{\Delta }}{d}_{l}}{{d}_{l}}=\tfrac{{\rm{\Delta }}{d}_{s}}{{d}_{s}}=\tfrac{{\rm{\Delta }}d}{d}=5 \% $.

We use the Markov Chain Monte Carlo (MCMC) method to constrain the parameters in Equation (17). The EMCEE (Foreman-Mackey et al. 2013) python package is used. In order to execute the MCMC process, we need to provide the prior values first. In Scolnic et al. (2018), the best value and the 1σ confidence level of the parameters are shown. But in our work, we take an SIS model of SGLS to calibrating these parameters, which will differ. We take the prior interval that completely covers the range of the values from Scolnic et al. (2018). The prior probability for parameters P(α, β, fe , η, γ, mstep, τ) is the product of the prior probability of each parameter. The prior probability is assumed to be uniform distributions: P(α) = U[−0.2, 0.2], P(β) = U[2, 6], P(fe ) = U[0.5, 1.5], P(η) = U[−0.5, 1.5], P(γ) = U[0, 0.3], P(Mstep) = U[5, 15], P(τ) = U[0.001, 1]. In Pantheon samples, the errors include both statistic and systematic deviation. The systematic error is relative to all data points and appears as a huge covariance matrix. In this work, some of the SNe Ia are selected, but only the statistic error is considered.

3.3. Results

The result is shown in Figure 3 and Table 1. Triangle contours are plotted by using the open-source python package Getdist. One can see from Figure 3 that the best-fitted center value is $\eta ={0.047}_{-0.151}^{+0.190}$, which is at 1σ confidence level. The result indicates that the CDDR is in agreement with the observations and there are no signs of violation in light of SN Ia and SL data.

Figure 3.

Figure 3. The 2D regions and 1D marginalized distributions with 1σ and 2σ contours for the parameters α, β, fe , γ, mstep, τ, and η using the Pantheon sample and the SGLS sample.

Standard image High-resolution image

Table 1. Constraints on the Coefficients of Light-curve Parameters and η at the 1σ Confidence Levels

ParameterValue
α ${0.001}_{-0.061}^{+0.061}$
β ${5.283}_{-0.464}^{+0.417}$
fe ${1.046}_{-0.019}^{+0.020}$
η ${0.047}_{-0.151}^{+0.190}$
γ ${0.141}_{-0.070}^{+0.080}$
mstep ${10.055}_{-0.148}^{+0.177}$
τ ${0.134}_{-0.073}^{+0.048}$
χ2 117.430
χ2/d. o. f 117.430/113

Download table as:  ASCIITypeset image

4. Discussion and Conclusions

The validation of the CDDR is a crucial topic in cosmology. Any violation of the CDDR may generate a new theory of physics. In recent years, to compare the DL derived from SNe la and the DA measured using galaxy clusters is the common method to test the CDDR. To use this method, a specific cosmology model with some parameters (e.g., the matter density parameter Ωm , the cosmic equation of state, and the Hubble constant) must be assumed. Such results are hardly convincing.

In testing the CDDR, using a model-independent method is necessary. Moreover, obtaining a data sample that contains a large number of DL and DA pairs is also important. However, the number of useful subsample pairs is limited by the observed data, and pair subsamples with redshift deviation smaller than a constant will lose the subsamples at high redshift, which leads to systematic errors.

In this paper, we provide a distance-deviation consistency and a model-independent method to test the CDDR. By applying the distance-deviation consistency method to the latest data set of SNe Ia with 1048 samples and a strong gravitational lensing system (SGLS) with 205 samples, we obtain a collection of subsample pairs that not only contains more subsamples but also maintains the information of subsamples at high redshift, up to z = 2.16. By applying a model-independent method the SGLS model is used to replace the cosmology model in SNe Ia light-curve fitting, the result shows that $\eta ={0.047}_{-0.151}^{+0.190}$ and CDDR is validated at the 1σ confidence level for the form $\tfrac{{D}_{L}}{{D}_{A}}{(1+z)}^{-2}=1+\eta z$.

We thank the anonymous referee for constructive comments. This work is supported by Yunnan Applied Basic Research Projects (2019FB007), Yunnan Youth Basic Research Projects (202001AU070020 and 202001AU070013), and Doctoral Programs of Dali University (KYBS201910 and KYBS201912).

Please wait… references are loading.
10.3847/1538-4357/abc9bf