Clustering of LRGs in the DECaLS DR8 Footprint: Distance Constraints from Baryon Acoustic Oscillations Using Photometric Redshifts

, , , , , , , , , and

Published 2020 November 23 © 2020. The American Astronomical Society. All rights reserved.
, , Citation Srivatsan Sridhar et al 2020 ApJ 904 69 DOI 10.3847/1538-4357/abc0f0

Download Article PDF
DownloadArticle ePub

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

0004-637X/904/1/69

Abstract

A photometric redshift sample of luminous red galaxies (LRGs) obtained from the DECam Legacy Survey (DECaLS) is analyzed to probe cosmic distances by exploiting the wedge approach of the two-point correlation function. Although the cosmological information is highly contaminated by the uncertainties existing in the photometric redshifts from the galaxy map, an angular diameter distance can be probed at the perpendicular configuration in which the measured correlation function is minimally contaminated. An ensemble of wedged correlation functions selected up to a given threshold based on having the least contamination was studied in previous work (Sridhar & Song 2019) using simulations, and the extracted cosmological information was unbiased within this threshold. We apply the same methodology for analyzing the LRG sample from DECaLS, which will provide the optical imaging for targeting two-thirds of the Dark Energy Spectroscopic Instrument footprint and measure the angular diameter distances at z = 0.69 and z = 0.87 to be ${D}_{A}(0.697)=(1529\pm 73\,\mathrm{Mpc})({r}_{d}/{r}_{d,\mathrm{fid}})$ and ${D}_{A}(0.874)=(1674\pm 102\,\mathrm{Mpc})({r}_{d}/{r}_{d,\mathrm{fid}})$ with a fractional error of 4.77% and 6.09%, respectively. We obtain a value of H0 = 66.58 ± 5.31 km s−1 Mpc−1, which supports the H0 measured by all other baryon acoustic oscillation results and is consistent with the ΛCDM model.

Export citation and abstract BibTeX RIS

1. Introduction

Measuring the expansion history of the universe is of paramount importance in the field of modern cosmology. It can be revealed by diverse cosmic distance measures in tomographic redshift space, such as cosmic parallax (Benedict et al. 1999), standard candles (Fernie 1969), or standard rulers (Eisenstein et al. 1998, 2005). To date, the best constraints come from the distance–redshift relation and imply that the expansion rate has changed from a decelerating phase to an accelerated one (Riess et al. 1998; Perlmutter et al. 1999). Most ongoing observations support the ΛCDM model with the presence of the cosmological constant, but confirming it with high precision or possibly finding any deviation from it still remains an interesting observational mission. One of the most robust methods for measuring the distance–redshift relation is to use the baryon acoustic oscillation (BAO) feature that is observed as a bump in the two-point correlation function or as wiggles in the power spectrum. The tension between gravitational infall and radiative pressure caused by the baryon-photon fluid in the early universe gave rise to an acoustic peak structure that was imprinted on the last-scattering surface (Peebles & Yu 1970). The BAO feature has been measured through the correlation function (Eisenstein et al. 2005), and the most successful measurements in the clustering of large-scale structure at low redshifts have been obtained using data from the Sloan Digital Sky Survey (SDSS; Eisenstein et al. 2005; Estrada et al. 2009; Hong et al. 2012; Padmanabhan et al. 2012; Veropalumbo et al. 2014, 2016; Alam et al. 2017). The Dark Energy Spectroscopic Instrument (DESI) is an upcoming survey (DESI Collaboration et al. 2016) that will be launched to probe the earlier expansion history with greater precision using spectroscopic redshifts. However, the photometric footprint for DESI has already been completed by the Legacy Imaging Surveys (Dey et al. 2019). Photometric surveys provide more observed galaxies compared to a spectroscopic survey even at deeper redshifts (Euclid Collaboration et al. 2019), but the uncertainty on the redshift obtained from photometric surveys is larger than the uncertainty on the redshift obtained from spectroscopic surveys. Although these photometric redshifts are measured with a much poorer resolution, and an unpredictable damping of clustering at small scales and a smearing of the BAO peak are caused by the photo-z uncertainty (Estrada et al. 2009), possible BAO signatures that have not been washed out by the redshift uncertainty might still be present.

We investigate the optimized methodology to extract the cosmic distance information from the photometric data sets and provide a precursor of cosmic distance information that will be revealed by the follow-up spectroscopy experiment much later on. We apply the wedge approach (Kazin et al. 2013; Sánchez et al. 2013, 2014; Sabiu & Song 2016; Ross et al. 2017; Sánchez et al. 2017; Sridhar & Song 2019) to probe the uncontaminated BAO feature by binning the angular direction from the perpendicular to radial directions, and we recover the residual BAO peak that has survived and get constraints on the angular diameter distance DA and H−1. It has also been shown recently by Ross et al. (2017) that the statistics obtained using the wedge correlation function are about 6% more accurate than that from the angular correlation function. Thus, using ${\xi }_{w}(s,\mu )$ not only adds more information compared to w(θ), but also overcomes the above disadvantages.

Recently, some improved methodologies have measured the Hubble constant with great precision, which reveals a tension among measurements. This tension draws attention from the community as a possible presence of new physics or unknown systematic uncertainties that need to be fixed. The Hubble constant is indirectly measured by the highest-resolution cosmic microwave background maps provided by the Planck satellite experiment (The Planck Collaboration 2006), and they find it to be ${H}_{0}=67.4\pm 0.5\,\mathrm{km}\,{{\rm{s}}}^{-1}\,{\mathrm{Mpc}}^{-1}$ (Planck Collaboration et al. 2020). The Hubble constant can also be directly probed by the classical distance ladder using type Ia supernova samples (Scolnic et al. 2019). The latest value from Riess et al. (2019) gives us a constraint of ${H}_{0}=74.03\pm 1.42\,\mathrm{km}\,{{\rm{s}}}^{-1}\,{\mathrm{Mpc}}^{-1}$ with a few percent marginal error. Both efforts leave a huge discrepancy in the H0 measurement, with the values being 4σ apart, which needs to be resolved.

While the current analyses of most cosmological observations at low redshift support the H0 measured by Planck, ;next-generation survey programs such as DESI will be launched in the near future. DESI will probe the earlier expansion history with greater precision using spectroscopic redshifts. However, by using the DECam Legacy Survey (DECaLS) data, which provides us constraints on the angular diameter distance, and by using the information of the sound horizon from Planck, we get constraints on H0. Our analyses use a fiducial cosmological model with the following parameters: Ωm  = 0.31, Ωb  = 0.049, $h\equiv {H}_{0}/(100\,\mathrm{km}\ {{\rm{s}}}^{-1}{\mathrm{Mpc}}^{-1})=0.676$, ns  = 0.96, and σ8 =0.8. The paper is organized as follows. In Section 2, we describe the DECaLS DR8 data, including the magnitude cuts we employ for our sample. Section 3 describes the clustering measurements and the fitting procedure used. We present our cosmic distance constraints obtained in Section 4 and discuss our overall results and conclusions in Section 5.

2. The Data

In this section, we describe the DESI Legacy Imaging Survey DR8 data used in this paper along with the Dark Sky simulation data used to test and validate our results.

2.1. DECaLS DR8 Data

The DESI Legacy Imaging Surveys will provide the target catalog for the upcoming DESI survey. One among the three imaging projects conducted for the Legacy Survey is DECaLS (Dey et al. 2019), which covers the South Galactic Cap region at DEC $\leqslant $ 34°. The data makes use of three optical bands (g, r, and z) to a depth of at least g = 24.0, r = 23.4, and z = 22.5, which is 1–2 magnitudes deeper than SDSS. We use the DECaLS data from the Legacy Surveys eighth data release (DR8), which is the first release to include images and catalogs from all three of the Legacy Surveys in a single release. The Legacy Surveys also processed some of the imaging data from the Dark Energy Survey (DES; Dark Energy Survey Collaboration 2005), and we include the DES imaging with DEC ≥  −30° in our analysis. In addition to the optical imaging, 4 yr of Wide-Field Infrared Survey Explorer (WISE; Wright et al. 2010; Meisner et al. 2017) data in the W1 and W2 bands are also included, which provide additional color information.

For the parent luminous red galaxy (LRG) sample in this study, we use a nonstellar cut of $(z-{\rm{W}}1)-0.8* (r-z)\,\gt -0.6$, a faint limit of z < 20.41, a color cut of $0.75\,\lt (r-z)\lt 2.45$, and a sliding magnitude–color cut of $(z-17.18)/2\lt (r-z)$. These selection cuts are motivated by the current DESI LRG target selection cuts (DESI Collaboration et al. 2016).

We also apply masks to get the final footprint for our parent sample using the "MASKBITS" column in the DR8 catalog. 10 Objects (and randoms) with the following bits are removed: 1 (Tycho-2 and GAIA bright stars), 8 (WISE W1 bright stars), 9 (WISE W2 bright stars), 11 (fainter GAIA stars), 12 (large galaxies), and 13 (globular clusters). Imaging data sets often suffer from systematic effects, and one such major contribution toward the systematic contamination comes from correlation with stellar density (Rezaie et al. 2020). A more detailed test of this effect on the large-scale structure correlation is explained in Appendix A.

After applying the magnitude cuts and masking scheme, we use random-forest-based (Breiman 2001) photo-z's from Zhou et al. (2020) to obtain the final photometric redshifts. The dispersion on the redshift is usually approximated by

Equation (1)

where σ0 denotes the dispersion at redshift z = 0, and ztrue is the true redshift or the spectroscopic redshift. The mean redshift uncertainty for the DECaLS DR8 sample within the range $0.3\lt {z}_{\mathrm{phot}}\lt 1.2$ is σ0 = 0.0264. The angular distribution of the parent LRG sample after applying the masking and selection cuts is plotted in the left panel of Figure 1, with higher-density regions denoted by a darker shade. The redshift distribution of the parent sample within the range $0.3\lt {z}_{\mathrm{phot}}\lt 1.2$ is plotted as a blue filled histogram in the right panel of Figure 1. For comparison, we also overplot the forecasted dN/dz LRG redshift distribution from DESI for a sky coverage of 9000 deg2 (given by the green solid line) and 14,000 deg2 (given by the orange solid line). The DECaLS DR8 sample has a sky coverage of ≈9500 deg2, which is more than two-thirds of the 14,000 deg2 DESI footprint.

Figure 1.

Figure 1. Left panel: angular distribution of the LRGs from the DECaLS DR8 photometric sample after applying the masking scheme and selection cuts. The density variations are shown using a normalized "gray-scale" color map, with darker regions denoting the highly dense regions. Right panel: photometric redshift distribution of the DECaLS DR8 parent sample (blue filled histogram) obtained using the random forest method. The green and orange solid lines are the forecasted dN/dz for LRG galaxies achievable by DESI (DESI Collaboration et al. 2016) for a sky coverage of 9000 and 14,000 deg2, respectively.

Standard image High-resolution image

2.2. Dark Sky and EZmock Simulation Data

In order to test and validate the results we obtain, we also need to analyze the clustering from a realistic mock catalog based on numerical simulations that calculate the nonlinear evolution of structure and predict the dependence of survey observables on cosmological parameters. We use the publicly available Dark Sky simulation set as described in Skillman et al. (2014) for this purpose. The simulation has been generated using particle numbers varying from 20483 to 10,2403 and in a comoving cosmological volume varying from 100 h−1 Mpc to 8 h−1 Gpc on a side. The objects have been placed in the simulation using a simple (time-evolving) halo occupation distribution (HOD) assuming spherical, Navarro–Frenk–White (Navarro et al. 1996) halos for the satellite. To identify dark matter halos and substructures, the ROCKSTAR halo finder (Behroozi et al. 2013) has been used. The halo-finding approach is based on an adaptive hierarchical refinement of friends-of-friends groups in both position and velocity. From the set of simulations (with varying particle numbers), we make use of the ds14_a simulation, which has 10,2403 particles and a particle mass of $3.9\times {10}^{10}{h}^{-1}{M}_{\odot }$. These specific mocks contain only R.A., decl., and true redshift information for objects preidentified to be LRGs and do not contain color or luminosity information.

To validate our results on photometric redshifts, we generate a set of photo-z's using the true redshift information available. In reality, the statistical nature of the photo-z error is too complicated to be specified with any known distribution function, but it is assumed that the error propagation of photo-z uncertainty into cosmological information is mainly caused by the dispersion length (Arnalte-Mur et al. 2009). Thus a simple Gaussian function of statistical distribution is usually chosen to generate the photo-z uncertainty distribution, and the photo-z error dispersion σz as defined in Equation (1) (a more detailed description is given in Section 3.4) is applied. In reality, the precision is dependent on many factors such as magnitude and spectral type, but here only the redshift factor is counted in Equation (1), in which the coherent statistical property determined only by z is applied for all types of galaxies in the simulation.

In this paper, rather than assuming a true Gaussian distribution for the photo-z errors, we use the spectroscopic redshifts of galaxies from our training set along with their photo-z's and σz 's to get the $({z}_{\mathrm{phot}}-{z}_{\mathrm{spec}})/{\sigma }_{z}$ distribution. We use this distribution to create our photometric redshifts, which we believe is a more accurate representation than using a true Gaussian distribution. The detailed description of generating the photo-z's for our simulations is given in Section 2.3, and the comparison of a true Gaussian distribution and the $({z}_{\mathrm{phot}}-{z}_{\mathrm{spec}})/{\sigma }_{z}$ distribution is given in Appendix B and plotted in Figure A2.

Table 1 summarizes the number density information from both the DECaLS data and the Dark Sky data for the entire redshift range and for the two redshift cuts used in this paper. It can be seen that the number density in all three redshift ranges for the mock is smaller than in the DECaLS sample. Thus, the comparison in the clustering between the two samples should only be looked into as a consistency check rather than a precise validation of our results.

Table 1. Number of LRGs, Volume for the Sample within the Range 0.3 < zphot < 1.2 (Also for the Two Redshift Cut Samples Used in This Paper), and the Mean Redshift Uncertainty within the Redshift Range (${\sigma }_{0}$) for the DECaLS, Dark Sky Mock, and EZmock Samples (Average Values from the Realizations)

 Redshift Range Ngals V (Gpc3) σ0
 0.3 < zphot < 1.251930787.640.0263
DECaLS0.6 < zphot < 0.820833941.740.0262
 0.8 < zphot < 1.010749162.250.0352
 0.3 < zphot < 1.227818966.510.0284
Dark Sky0.6 < zphot < 0.812716701.650.0282
 0.8 < zphot < 1.05607492.160.0341
EZmock0.6 < zphot < 0.820589061.650.0263
 0.8 < zphot < 1.09636732.160.0342

Note. The Ngals quoted is the total number of LRGs used in the large-scale clustering analysis.

Download table as:  ASCIITypeset image

To compute the error on ${\xi }_{w}(s,{\mu }_{i})$, we make use of 100 EZmock (Chuang et al. 2015) simulations, all of which have the DESI expected sky coverage. To match the DECaLS DR8 footprint, we cut the EZmock samples within $-30^\circ \lt \mathrm{DEC}\lt $ 34°. The EZmock sample after the DEC cut has an area of ≈9300 deg2, which is similar to the area of the DECaLS DR8 parent sample that we use in this paper. The mocks contain R.A., decl., zcosmo, and dzrsd information. Thus, we need to generate photometric redshifts for the mocks so that they can be used to obtain the covariance matrix. The methods of generating photometric redshifts for the EZmock samples are explained in detail in Section 2.3.

2.3. Photometric Redshifts for Simulations

To generate the photo-z's for our simulations, we obtain ${\sigma }_{z}$'s randomly from the parent DECaLS sample, along with random points from the $({z}_{\mathrm{phot}}-{z}_{\mathrm{spec}})/{\sigma }_{z}$ distribution. To select the random σz 's, we restrict them to galaxies of similar redshifts that are within a redshift range of ±0.01 zphot, which ensures that the dependence of errors on redshift is included. For example, for N number of EZmock galaxies that are within $0.60\lt {z}_{\mathrm{cosmo}}\lt 0.61$, N σz 's from the DECaLS data within 0.60 < zphot < 0.61 are randomly selected. This process is repeated over the entire redshift range. Once we generate the photo-z's, they are diluted according to the DECaLS N(z) to make sure that they are consistent. The number of galaxies, volume within the redshift range, and the σ0 for the two redshift cuts are listed in Table 1. Our detailed analysis comparing ${\xi }_{w}(s,\mu )$ between the EZmock sample and the DECaLS sample is explained in Appendix B and shown in Figure A1.

3. Methodology

In this paper, we follow the same methodology and formulation that were applied in Sridhar & Song (2019) to simulated photometric galaxy catalogs to get cosmological distance constraints. We explain in detail the clustering measurements obtained from the wedge correlation function and the comparison between the DECaLS and Dark Sky mock data.

3.1. Clustering Measurements and Fitting Procedure

The excess probability of finding two objects relative to a Poisson distribution at volumes dV1 and dV2 separated by a vector distance  r is given by the two-point correlation function ξ(r) (Totsuji & Kihara 1969; Davis & Peebles 1983). The galaxy distribution seen in redshift space exhibits an anisotropic feature distorting ξ(r) into $\xi (\sigma ,\pi )$ along the line of sight (LOS), where σ and π denote the transverse and radial components of the separation vector  r . Acoustic fluctuations of the baryon–radiation plasma of the primordial universe leave a signature on the density perturbation of baryons. This standard ruler length scale, set by the acoustic wave, propagates until it is frozen at the decoupling epoch to remain in the large-scale structure of the universe. The threshold length scale of the acoustic wave is called the sound horizon, which is given by

Equation (2)

where cs is the sound speed of the plasma. This scale is imprinted on the correlation function as a peak and is imprinted on the matter power spectrum as a series of waves. Assuming standard matter and radiation content in the universe, the Planck Collaboration et al. (2020) measurements of the matter and baryon density determine the sound horizon to 0.2%. By measuring the BAO feature using an anisotropic analysis, one can separately measure DA (z) and ${H}^{-1}(z)$. But adjustments to the cosmological parameters or changes to the prerecombination energy density can alter the value of rd (Alam et al. 2017). So, the BAO measurements constrain the combinations ${D}_{A}(z)/{r}_{d}$ and ${H}^{-1}(z){r}_{d}$. The sound horizon for this fiducial model is ${r}_{d,\mathrm{fid}}=147.21$ Mpc as obtained from Planck Collaboration et al. (2020). The scalings of rd with cosmological parameters can be found in detail in Aubourg et al. (2015). The distance constraints quoted in this paper are in units of megaparsecs and with a scaling factor, for example, ${D}_{A}(z)\times ({r}_{d,\mathrm{fid}}/{r}_{d})$, so that the numbers provided are independent of the fiducial cosmological parameters used.

The Landy & Szalay estimator (hereafter LS) in (s, μ) coordinates is best suited to calculate the two-point correlation function (Farrow et al. 2015; Sridhar et al. 2017) and extract BAO information from photometric redshift galaxy maps (Sridhar & Song 2019). The radius to shell s and the observed cosine of the angle the galaxy pair makes with respect to the LOS μ are given by ${s}^{2}={\sigma }^{2}+{\pi }^{2}$ and μ = π/s, respectively, where σ and π denote the transverse and radial separation between the galaxy pairs.

It is common practice to separate the random sample distributions into the angular and redshift components separately. We make use of the random catalog provided in the DR8 data release 11 by the Legacy Survey, which gives us the angular component. These randoms have been downsampled to the surface density of 10,000/deg2 and require +2 exposures in the g, r, and z bands. The number of objects is usually twice or more than the data catalog to avoid shot noise effects. In our case, the random catalog has five times more objects than the data catalog. For the redshift component, we extract redshifts randomly from the data catalog within the chosen redshift range (see Ross et al. 2012, 2017; Veropalumbo et al. 2016; Sridhar & Song 2019, for more info). The same number of exposure requirements, footprint cuts, and bright star masks are applied on the randoms as used in constructing the LRG sample. We use the publicly available KSTAT (KD-tree Statistics Package) code (Sabiu 2018) to calculate all of our correlation functions.

We pay attention to the usefulness of exploiting the wedge correlation function to separate the radial contamination from the BAO signal imprinted on perpendicular configuration pairs. The wedge correlation function ξw is given by

Equation (3)

where μi is the mean μ in each bin (we will refer to the mean value of the μ bin using $\bar{\mu }$ hereafter), ${\mu }_{i}^{\min }$ and ${\mu }_{i}^{\max }$ are the minimum and maximum values of μ, and W is a window function within the chosen minimum and maximum limits of μ. Using too many μ bins will complicate the covariance matrix, and by using very few μ bins we will not be able to separate the error propagation along the LOS clearly (Sabiu & Song 2016). Thus, we choose six bins in the μ direction with Δμ = 0.17 between μ = 0 and 1, with $\mu \to $ 0 corresponding to the transverse plane and $\mu \to $ 1 corresponding to the LOS plane.

In the case of photometric redshift samples, the noise on the pairs increases along the radial configuration and thus causes a smearing of the BAO peak (Estrada et al. 2009). This smearing not only increases with increasing photometric uncertainty but also increases along the LOS for a given σ0. These noisy pairs can be removed using a cutoff $\bar{\mu }$. It has been shown in Sridhar & Song (2019) that using a cutoff $\bar{\mu }=0.42$ for photometric redshift samples can remove most of the contaminated pairs, so we use cutoff $\bar{\mu }=0.42$ in this paper. The empirical model that we use to fit the correlation function and obtain the BAO peak location is similar to the one proposed by Sánchez et al. (2011, 2012) and is given by

Equation (4)

where B takes into account a possible negative correlation at very large scales, s0 is the correlation length (the scale at which the correlation function ≃1), and γ denotes the slope. The remaining three parameters, N, σ, and sm , are the parameters of the Gaussian function that model the BAO feature, and, in particular, sm represents the estimate of the BAO peak position. This empirical model can be used to accurately extract the BAO peak position (Sánchez et al. 2011; Veropalumbo et al. 2016; Sridhar & Song 2019) when the correlation function is provided.

The likelihood on sm from previous BAO studies is either obtained by using a 1D grid on sm where the χ2 is minimized at each grid point or from the marginalized posterior from a Monte Carlo Markov Chain (MCMC) analysis. In this study, the fitting is performed by applying the MCMC technique (we make use of the emcee Python package, Foreman-Mackey et al. 2013), using the full covariance matrix obtained using Equation (17). The fitting parameter space is given by

Equation (5)

and we place flat, wide priors on all six parameters. This five-parameter space is used for all of our μ bins when we perform the fitting. For the first three parameters, the ranges of the priors are $0.0\lt B\lt 1.0$, 0.0 < s0 < 3.0, and 0.0 < γ < 3.0, and for the remaining three parameters of the Gaussian function, the ranges of the priors are 0.0 < N < 1.0, 85.0 < sm  < 130.0, and $0.0\lt \sigma \lt 35.0$. Several variations of the range of these priors were tested, especially the range of the prior on sm and σ. A smaller range for the sm affects the posterior distribution, and we miss most of the information at high $s({h}^{-1}\,\mathrm{Mpc})$. A similar effect is seen when we use a smaller range for the σ prior, which eventually amplifies the BAO peak. We have also tested several ranges within which to perform the fit, and we fit the correlation function within the range $30.0\lt s({h}^{-1}\,\mathrm{Mpc})\,\lt 130.0$ after experimenting with other ranges. We find that there is a maximum shift of 1% in the BAO peak when we vary the range of the fit. This 1% shift is negligible compared to the error on the BAO peak point we obtain (as discussed in Section 3.4), and thus we believe it is subdominant. The constraints on the BAO peak sm for the wedge correlation function are obtained after fully marginalizing all other parameters in xp . We adopt a standard likelihood, ${\mathscr{L}}\,\propto \exp (-{\chi }^{2}/2)$, where the function χ2 is defined as

Equation (6)

where ${\xi }_{\mathrm{mod}}(s)$ is the model correlation function as given by Equation (4), ${\xi }_{w}(s,{\mu }_{i})$ is the observed correlation function for the ith $\bar{\mu }$ bin, and ${C}^{-1}$ is the inverse covariance matrix.

3.2. Theoretical Model for the Correlation Function

We compute the theoretical correlation function ξth(s, μ) in redshift space, exploiting the improved power spectrum based upon the Taruya, Nishimichi, and Saito (TNS) model as

Equation (7)

with ${ \mathcal P }$ being the Legendre polynomials. Here, we define $\mu =\pi /s$ and $s={\left({\sigma }^{2}+{\pi }^{2}\right)}^{1/2}$. The moments of the correlation function, ξ (s), are defined by

Equation (8)

The multipole power spectra ${\tilde{P}}_{{\ell }}(k)$ are explicitly given by

Equation (9)

where we define the function pm (k):

Equation (10)

with $\kappa ={k}^{2}{\sigma }_{p}^{2}$. The function γ is the incomplete gamma function of the first kind:

Equation (11)

The Q2n is explained below.

The observed power spectrum in redshift space $\tilde{P}(k,\mu )$ is written in the following form:

Equation (12)

where the velocity dispersion σp is set to be a free parameter for the Finger of God (FoG) effect, and the function Q2n is given by

Equation (13)

where Cn includes the higher-order polynomials caused by the correlation between density and velocity fluctuations, and PXY (k) denotes the power spectrum in real space. The standard perturbation model exhibits the ill-behaved expansion leading to the bad UV behavior. In this manuscript, we use the resummed perturbation theory RegPT, which is regularized by introducing a UV cutoff (Taruya et al. 2012). The auto and cross spectra of PXY (k) are computed up to first order, and higher-order polynomials are computed up to zeroth order, which are consistent in the perturbative order.

Although cosmic distances are estimated using the BAO at linear regimes, there are smearing effects at small scales that need to be computed. These small-scale corrections are included to make the final precise constraints on the BAO (Taruya et al. 2010). We make use of an improved model of the redshift-space power spectrum (Taruya et al. 2010), in which the coupling between the density and velocity fields associated with the Kaiser and the FoG effects is perturbatively incorporated into the power spectrum expression. The result includes nonlinear corrections consisting of higher-order polynomials (Taruya et al. 2010):

Equation (14)

Here, the A(k, μ) and B(k, μ) terms are the nonlinear corrections and are expanded as a power series of μ. Those spectra are computed using the fiducial cosmological parameters. The FoG effect GFoG is given by a simple Gaussian function that is written as

Equation (15)

where ${\sigma }_{p}$ denotes the one-dimensional velocity dispersion.

Thus the theoretical correlation function ${\xi }_{\mathrm{th}}(s,\mu )$ is parameterized by

Equation (16)

wherein Gb and GΘ are the normalized density and coherent motion growth functions. When working with photo-z samples, the effect of the photo-z error on the correlation function is incoherent. Thus, an extra parameter is needed for the theoretical template to model ξth(s, μ) as a function of the photo-z error, but it is not well understood. So, we use Equation (4) instead to fit our observed ${\xi }_{w}(s,\mu )$. This functional form only assumes a power law at small scales and a Gaussian function to fit the BAO peak at large scales and seems to model ${\xi }_{w}(s,\mu )$ quite well, as we can see from Figure 2.

Figure 2.

Figure 2. Correlation function ${\xi }_{w}(s,\mu )$ (multiplied by s2) calculated by splitting into wedges of μ for the 0.6 < zphot < 0.8 sample (given by blue dots) and for the $0.8\lt {z}_{\mathrm{phot}}\lt 1.0$ sample (given by red dots). The first, second, and third columns in the figure represent the $\bar{\mu }=0.08$, 0.25, and 0.42 bins, respectively. The dashed black lines in each plot show the best fit (maximum likelihood) obtained from the empirical model by applying the MCMC technique. The error bars plotted are the square root of the diagonal elements of the full covariance matrix, as mentioned in Equation (17).

Standard image High-resolution image

In our previous work (Sridhar & Song 2019), we have verified that the BAO feature from the theoretical correlation function is weakly dependent on the growth functions and ${\sigma }_{p}$. We have verified that changing the value of σp by ±10% does alter the location of the BAO peak, but only by less than 0.1% for samples at $\bar{\mu }$ close to 0 and less than 1% for samples at $\bar{\mu }$ close to 1, which is negligible. Thus, when we fit the cosmic distances, we fix ${\sigma }_{p}$. To find the best-fit σp , we vary it within $3.0\lt {\sigma }_{p}({\text{}}{h}^{-1}\ \mathrm{Mpc})\lt 6.0$ (by fixing DA and ${H}^{-1}$ to their fiducial values), compute the theoretical correlation function ${\xi }_{\mathrm{th}}(s,\mu )$, and use the ${\sigma }_{p}$ for which ${\xi }_{\mathrm{th}}(s,\mu )$ best matches our measured ${\xi }_{w}(s,\mu )$. The best-fit σp used in this work is ${\sigma }_{p}=4.8\,{\text{}}{h}^{-1}\ \mathrm{Mpc}$, and we fix it for both of the redshift ranges.

Note that we apply the TNS model to compute the theoretical ${\xi }_{w}(s,\mu )$, from which we extract the BAO peak information using the fitting procedure mentioned in Section 3.1. The theoretical BAO peaks in the fiducial cosmology can also be transformed into a new cosmology according to simple coordinate projections wherein we can vary the DA and H−1 from the fiducial values by a small fraction. But it has been shown in Sridhar & Song (2019; see Figure 6) that the location shift of the BAO peak with varying DA and H−1 is not completely consistent with this coordinate transformation. The difference is exceeding the detectability limit by about 5%, and thus the TNS model is adopted to determine the theoretical BAO points.

3.3. Covariance Matrix from EZmock Samples

A random catalog that is approximately 20 times the mock data is provided for the EZmock samples, but it contains only the angular component (R.A., decl.). Since the density of the DECaLS randoms is only five times the data catalog, we dilute the EZmock randoms to the same density to ensure consistency in our results. For the redshift component, we follow the same method of randomly extracting σz from the data catalog.

We calculate the covariance matrix, which is given by

Equation (17)

where the total number of simulations is given by N. The ${\xi }_{w}^{n}({\vec{x}}_{i})$ represents the value of the wedge correlation function of the ith bin of ${\vec{x}}_{i}$ in the nth realization, and ${\overline{\xi }}_{w}({\vec{x}}_{i})$ is the mean value of ${\xi }_{w}^{n}({\vec{x}}_{i})$ over all realizations. Due to the limited number of mock samples (100) that we have, we use 12 bins in s. We obtain the correlation matrix as

Equation (18)

which is plotted in Figure 3.

Figure 3.

Figure 3. Correlation matrix for the 0.6 < zphot < 0.8 sample using 12 bins in s and six bins in μ computed using Equation (18).

Standard image High-resolution image

The inverse of Cij is well defined and thus does not require any denoising procedures such as singular value decomposition. Additionally, we also count the offset caused by the finite number of realizations (Hartlap et al. 2007) as

Equation (19)

where Nbins denotes the total number of i bins. As mentioned in Section 3.1, we fit the correlation function within the range $30.0\lt s({h}^{-1}\mathrm{Mpc})\lt 130.0$. Thus, the number of s bins in this range is nine. As the final three μ bins along the LOS do not contain any BAO information, we perform the fitting by ignoring them. Thus, the shape of the matrix is 27 × 27, and for 100 mock realizations, the factor in Equation (19) becomes 0.71. Apart from the above correction factor, an additional correction to the inverse covariance matrix is proposed by Percival et al. (2014), which is given by

Equation (20)

where nb is the number of bins used for the two-point correlation measurements, np is the number of parameters measured, and the A and B terms are given by

Equation (21a)

Equation (21b)

where ns is the number of simulations used for the covariance matrix calculations. Applying the square root of this expression to the measured standard deviation should take care of the extra correction. For our binning scheme as mentioned above (with 100 mock realizations), we get $\sqrt{{m}_{1}}=0.89$, which is significantly smaller than the correction factor we already apply. The BAO peak position sm of the wedge correlation function is found by fitting the phenomenological model by considering the full covariance using Cij .

3.4. DECaLS and Dark Sky Mock Acoustic-scale Measurements from the Wedge Correlation Function

From the parent DECaLS DR8 sample within the redshift range $0.3\lt {z}_{\mathrm{phot}}\lt 1.2$, we choose two redshift cuts between $0.6\lt {z}_{\mathrm{phot}}\lt 0.8$ and 0.8 < zphot < 1.0 for our analysis. The reason for choosing these two redshift ranges is that the redshift distribution of the sample peaks at ${z}_{\mathrm{phot}}\approx 0.7$, as can be seen from Figure 1. Thus, we expect to have the maximum number of galaxies around this redshift range. The redshift uncertainty scales with redshift, so we quote the mean values for our two redshift cut samples in Table 1. The wedge correlation functions are calculated using Equation (3) by using six μ bins of thickness Δμ = 0.17. The ${\xi }_{w}(s,{\mu }_{i})$ calculated from the first three μ bins for the two redshift samples is shown in Figure 2. We fit ${\xi }_{w}(s,\mu )$ using Equation (4) by following the MCMC procedure described in Section 3.1, and the values of the BAO peak sm from the fit are provided in Table 2.

Table 2. Results of Fitting the Correlation Function (Plotted Using the Dotted Lines in Figure 2) for the Two Redshift Samples and in the Three $\bar{\mu }$ Bins using Equation (4)

Redshift Range $\bar{\mu }$ Bin sm (h−1 Mpc)
 0.08 ${111.6}_{-3.2}^{+3.0}$
$0.6\lt {z}_{\mathrm{phot}}\lt 0.8$ 0.25 ${106.8}_{-5.6}^{+4.8}$
 0.42 ${111.9}_{-8.0}^{+5.7}$
 0.08 ${114.2}_{-5.1}^{+4.7}$
$0.8\lt {z}_{\mathrm{phot}}\lt 1.0$ 0.25 ${109.4}_{-10.0}^{+10.9}$
 0.42 ${110.5}_{-6.0}^{+5.5}$

Note. The sm is the BAO peak point obtained from the fit, and the units are in h−1 Mpc.

Download table as:  ASCIITypeset image

One can observe that for both redshift samples, the BAO signal is diluted as μ increases and is also more clearly visible for the first redshift sample than for the second redshift sample. This is because the photometric redshift errors scale proportional to ${\sigma }_{0}\times (1+z)$, and thus the isotropy along the LOS is destroyed more strongly for the high-redshift sample than for the low-redshift one. The errors on sm gradually increase with increasing $\bar{\mu }$, as expected.

For both redshift ranges and all three $\bar{\mu }$ bins used, we fit the correlation function within the range $30.0\lt s({h}^{-1}\mathrm{Mpc})\,\lt 130.0$. When using log binning instead of linear binning, we note that the BAO peak results slightly shift to higher values. However, the effect is of the order of 1%, which is well below the estimated accuracy of the BAO peak position, which is between 4% and 10% for our samples. We use the ${\chi }^{2}/{dof}$ goodness-of-fit indicator to validate the performance of our empirical model to fit the correlation function. We have nine data points within the range $30.0\lt s({h}^{-1}\mathrm{Mpc})\lt 130.0$, and we fit for six parameters in each of the three $\bar{\mu }$ bins. The overall fit to the $0.6\lt {z}_{\mathrm{phot}}\lt 0.8$ sample yields a χ2/dof = 10/9, including all cross-covariances between $\bar{\mu }$ bins. We obtain a χ2/dof = 8/9 for the $0.8\lt {z}_{\mathrm{phot}}\lt 1.0$ sample.

It has been shown from previous studies (Ross et al. 2012, 2017) that data from the different μ bins are expected to be correlated and that the results from splitting the clustering by μ show a slight decrease in the BAO information content with increasing μ. Thus, we perform the fit using the full data vector including all of the μ bins. By performing the fit using the full covariance matrix, we make sure that the correlations between the different μ bins that exist are taken into account. The one- and two-dimensional projections of the posterior probability distribution of the sm parameter from the MCMC chains for the two redshift samples are shown in Figure 4. The marginalized distribution for each sm value from each $\bar{\mu }$ bin is shown independently in the histograms along the diagonal and the marginalized two-dimensional distributions in the other panels. We find that the correlation between the sm values obtained at the different μ bins for both redshift samples is minimal.

Figure 4.

Figure 4. Left panel: marginalized posterior distribution of the peak point sm obtained from the MCMC analysis using Equation (4) for all three μ bins in the redshift range $0.6\lt {z}_{\mathrm{phot}}\lt 0.8$. The covariances between the different sm values are shown in the contour plots, and the orange circle encompasses all points within the 1σ region. The marginalized distribution for each independent sm is shown in the histograms along the diagonal. Right panel: the same for the $0.8\lt {z}_{\mathrm{phot}}\lt 1.0$ sample.

Standard image High-resolution image

To validate our clustering results obtained on the DECaLS data, we use the LRGs from the Dark Sky mock catalog. To mimic the photometric redshifts from the DECaLS data, we follow the methodology as explained in Section 2.3. To compare the correlation function results obtained from the two redshift samples of the DECaLS catalog, we compute ${\xi }_{w}(s,\mu )$ from the Dark Sky photometric redshift catalogs with the same redshift cuts and use the same binning scheme. We find that by using the true values of ${\xi }_{w}(s,\mu )$ for the two photometric redshift samples from the Dark Sky mocks, the amplitudes of ${\xi }_{w}(s,\mu )$ do not match with the ${\xi }_{w}(s,\mu )$ from the DECaLS data. There are a few reasons for this behavior, and one of them is the poor estimation of the survey volume (due to systematics). Apart from systematics, there can also be other potential sources for this behavior. If the photo-z's were overestimated, then when applied to the mock catalogs they can cause a lower amplitude of the correlation function. The mocks could also have a lower galaxy bias compared to the observed galaxies, which can also result in a lower amplitude of the correlation function. However, by adding a constant value of 0.0005 and 0.0010 to the ${\xi }_{w}(s,\mu )$ obtained from the Dark Sky mocks for the 0.6 < zphot < 0.8 and 0.8 < zphot < 1.0 samples, we see that the amplitudes match well. The results are presented in Figure 5.

Figure 5.

Figure 5. Top panel: ${\xi }_{w}(s,\mu )$ (multiplied by s2) calculated for the DECaLS data (dotted points in red, same as in Figure 2) compared with ${\xi }_{w}(s,\mu )$ calculated for the Dark Sky simulation (solid lines in blue) within the redshift range $0.6\lt {z}_{\mathrm{phot}}\lt 0.8$ for the three $\mu =0.08,0.25,0.42$ bins from left to right. Bottom panel: same plot for the samples within the redshift range $0.8\lt {z}_{\mathrm{phot}}\lt 1.0$. A constant value of 0.0005 and 0.0010 has been added to the ${\xi }_{w}(s,\mu )$ from the Dark Sky mocks for the 0.6 < zphot < 0.8 and 0.8 < zphot < 1.0 samples, respectively, so that the amplitudes match the ${\xi }_{w}(s,\mu )$ from the DECaLS data. The error bars are the square roots of the diagonal elements of the full covariance matrix as mentioned in Equation (17). The error bars for the Dark Sky sample are represented by the shaded blue region.

Standard image High-resolution image

We are interested in two key features when comparing the data and the mock catalog. One is the amplitude of ${\xi }_{w}(s,\mu )$, and the other is the location of the BAO peak. We find that the amplitude of the clustering measurements from the Dark Sky mock catalog (given by the solid blue line) matches the amplitude of the clustering measurements from the DECaLS sample (given by the red scatter points) for both redshift samples in all three $\bar{\mu }$ bins after adding the constant values to our redshift samples as described above. We repeat the MCMC procedure to obtain the BAO peak for the Dark Sky sample and find that the location of the BAO peak from the Dark Sky samples for both redshift ranges agrees with the DECaLS sample, at least within 1σ. A similar result has been obtained by Ross et al. (2017) by doing a comparison between mock samples and model curves using mock photometric data.

We also verify the internal consistency of the BAO peaks obtained from the Dark Sky photometric catalog by comparing them with the BAO peaks obtained from the true redshift (zpec, cosmological redshift with peculiar velocity added) catalog for the same $\bar{\mu }$ bins. The wedge correlation functions from the three $\bar{\mu }$ bins are calculated using Equation (3), and the sm values obtained from the two samples are plotted in Figure 6. For all three $\bar{\mu }$ bins, it can be seen that the sm values from the photometric samples are within 1σ compared to the sm values from the zpec sample, with the 1σ errors on sm being larger for the zphot sample.

Figure 6.

Figure 6. The x-axis denotes the three $\bar{\mu }$ bins we have used, and the y-axis denotes the value of sm obtained from the empirical fit for the Dark Sky zpec sample (black dots) and for the Dark Sky zphot sample (blue dots) within the redshift range 0.6 < zphot < 0.8. The $\bar{\mu }$ bins for the zphot sample have been shifted by 0.01 for better visualization.

Standard image High-resolution image

4. Measured Cosmic Distances

In the previous section, we measured the BAO peak position sm for our DECaLS sample at different $\bar{\mu }$ bins. In this section, we explain the theoretical model that we use to obtain the theoretical correlation function ξth(s, μ). The ξth is a function of both s and μ, which can then be used to translate our measured BAO peak positions to physical distances.

The volume distance at a redshift is given by

Equation (22)

and is measured through the BAO by exploiting the monopole correlation function (ξ0). It has been shown from previous studies (Estrada et al. 2009; Sridhar & Song 2019) that for photometric redshift samples, the BAO peak is smeared out in ξ0. Thus, using the wedge approach, both the transverse and radial cosmic distances can be separately measured.

In Section 3.4, we obtained the sm values for our two redshift samples from the DECaLS data. In this section, we explain how we use a two-step process to go from sm to measured cosmic distances DA and H−1. As mentioned in the previous section, ${\xi }_{\mathrm{th}}$ is a function of s and μ and depends on five parameters as mentioned in Equation (16). Since ξth(s, μ) is weakly dependent on the growth functions (Gb and GΘ) and σp , we vary the tangential and radial distance measures from the fiducial values of DA and H−1 for the two redshift samples when we fit the cosmic distances. We use a 13 × 13 grid to vary DA and H−1, and both DA and H−1 are sampled within $0.6({{param}}^{\mathrm{fid}})\lt {{param}}^{\mathrm{fid}}\lt 1.4({{param}}^{\mathrm{fid}})$, where param is either DA or H−1. For each of these parameter sets, we obtain a ${\xi }_{\mathrm{th}}(s,\mu )$ for the given μ. We then fit each of our ${\xi }_{\mathrm{th}}(s,\mu )$ functions using Equation (4) to obtain the theoretical sm values. We then compare the ${\xi }_{\mathrm{th}}(s,\mu )$ with our measured ${\xi }_{w}(s,\mu )$ and compute the ${\chi }^{2}$ values using the sm values from the DECaLS data and the theoretical templates for the first three $\bar{\mu }$ bins (as we find that the BAO peak is washed out for the last three $\bar{\mu }$ bins), taking into account the covariance between the sm values between the different $\bar{\mu }$ bins that exist.

The uncertainty on the redshift determination prevents us from accessing the radial cosmic distance, and thus the BAO peak is not clearly visible for the $\xi (s,\bar{\mu }\gt 0.5)$ correlation functions. Thus, we do not get tight constraints on H−1(z). However, even after fully marginalizing over H−1(z), the transverse cosmic distance ${D}_{A}(z)\times ({r}_{d,\mathrm{fid}}/{r}_{d})$ is measured with good precision for both redshift samples. The fiducial values of DA fid for our cosmology at the two mean redshifts are ${D}_{A}^{\mathrm{fid}}(\bar{z}=0.697)=1514$ Mpc and ${D}_{A}^{\mathrm{fid}}(\bar{z}=0.874)=1638$ Mpc, and the measured values are

Equation (23)

Equation (24)

These values correspond to distance measures of 4.77% and 6.09% precision for the two redshift samples, respectively. The 0.2% statistical error on rd based on the Planck Collaboration et al. (2020) measurements only makes a negligible contribution when added in the above equations. We compare our results with previous studies in Figure 7. The constraints using four spectroscopic redshift surveys, that is, Blake et al. (2011; WiggleZ), Alam et al. (2017; DR12 BOSS), Chuang et al. (2017; DR12 CMASS), and Bautista et al. (2018; DR14 eBOSS) are plotted in yellow, green, blue, and black, respectively. The constraints using the DES photometric redshift survey (The Dark Energy Survey Collaboration et al. 2019) are plotted in brown. The solid black line corresponds to the theoretical predictions as a function of redshift obtained using the cosmological parameters from Planck Collaboration et al. (2020).

Figure 7.

Figure 7. Left panel: The ${D}_{A}(z)\times ({r}_{d,\mathrm{fid}}/{r}_{d})$ measurements obtained from the two redshift samples are plotted in red along with the values of ${D}_{A}(z)\times ({r}_{d,\mathrm{fid}}/{r}_{d})$ measured by other surveys (as color coded in the legend and mentioned in the text). The solid black line corresponds to the theoretical predictions for DA (z) as a function of redshift obtained using the cosmological parameters measured from Planck Collaboration et al. (2020). Units are in Mpc. The normalized likelihood ${\mathscr{L}}$ = exp $(-{\rm{\Delta }}{\chi }^{2}/2)$ for DA (z) obtained for the 0.6 < zphot < 0.8 sample (solid red line) and the 0.8 < zphot < 1.0 sample (dotted red line) is plotted in the inset plot. Right panel: the value of H0 obtained using the likelihoods from our DA measurements (on top) compared with H0 measurements from different probes with 1σ error bars. The 1σ error from the Planck Collaboration et al. (2020; constraints including BAO) is plotted in light green, and the 1σ error from Riess et al. (2019) is plotted in light blue. From bottom to top, enumerated on the vertical axis, we show Tröster et al. (2020; BOSS DR12 constraints from anisotropic clustering measurements), Colas et al. (2020; SDSS/BOSS DR12 constraints using effective field theory), Domínguez et al. (2019; γ-ray attenuation), Abbott et al. (2017; LIGO binary black hole merger GW170817), Fernández Arenas et al. (2018; HII galaxies), Yu et al. (2018; cosmic chronometers + BAO), Yuan et al. (2019; TRGB calibrated SNIa), Wong et al. (2020; H0LiCOW, gravitationally lensed quasars), and Abbott et al. (2018; DES clustering + weak lensing).

Standard image High-resolution image

The cosmological concordance model with the cosmological constant is assumed to be a cause of cosmic acceleration, with the Hubble constant unknown. The measured angular diameter distance at the two redshifts from the DECaLS sample and the prior information for the sound horizon size and ${w}_{m}\equiv {{\rm{\Omega }}}_{m}{h}^{2}=0.1430\pm 0.0011$ as determined by the Planck experiment (Planck Collaboration et al. 2020) are used to get constraints on H0. We fit this cosmological concordance model to the measured DA , applying the wm prior probed by Planck. We fit for the parameters (${{\rm{\Omega }}}_{m},{{\rm{\Omega }}}_{m}{h}^{2},{H}_{0}$) using flat, wide priors that extend well beyond the regions of high likelihood and have no effect on the cosmological fits, and we obtain a value of ${H}_{0}=66.58\pm 5.31\,\mathrm{km}\,{{\rm{s}}}^{-1}\,{\mathrm{Mpc}}^{-1}$.

We also make a comparison plot with H0 measurements obtained from recent works using different probes in the right panel of Figure 7. Our H0 value is measured with 8.1% precision, whereas some of the estimates from other probes plotted in Figure 7 have a better precision. To quote a few, the HII galaxy data (Fernández Arenas et al. 2018) deliver ${\sigma }_{{H}_{0}}/{H}_{0}=4.9 \% $, and the DES + BAO + BBN data deliver ${\sigma }_{{H}_{0}}/{H}_{0}=1.8 \% $. The reason for our conservative estimate is that we only use the likelihoods from our DA (z) measurements, which have been obtained from photometric redshift samples. We have used the prior information of the sound horizon scale and wm from Planck like other BAO studies. We believe that the photo-z error is subdominant compared to the error that we get from cosmic variance. The DESI catalog and the DECaLS catalog share the same footprint, so the cosmic variance will be minimal; however, due to the photo-z uncertainty, the error on the H0 value we obtain increases.

5. Discussion and Conclusions

We provide a statistical methodology to extract cosmic distance information using BAO peaks only from the DECaLS DR8 LRG photometric galaxy sample. To extract the BAO peak from photometric redshift catalogs, the common practice is to measure the incomplete angular correlation function. In this manuscript, we make use of the wedge correlation function, wherein we split the sample into small wedges and include the BAO information from all of the wedges in which they are still present ($\bar{\mu }\lt 0.5$, above which there is noticeable contamination). The transverse cosmic distance DA (z) is measured with good precision for both redshift samples, giving us values of ${D}_{A}(\bar{z}=0.69)=1529\pm 73\,\mathrm{Mpc}({r}_{d}/{r}_{d,\mathrm{fid}})$ and ${D}_{A}(\bar{z}=0.87)=1674\pm 102\,\mathrm{Mpc}({r}_{d}/{r}_{d,\mathrm{fid}})$ with a fractional error of 4.77% and 6.09%, respectively. The values that we obtain have been compared with the theoretical prediction for DA (z) as a function of redshift obtained using the cosmological parameters measured from Planck Collaboration et al. (2020) and are well within the 1σ region.

We have also compared our results with the results of DA (z) obtained from other similar surveys (both spectroscopic and photometric) and find them to be consistent with each other. Since most radial information is contained in the $\bar{\mu }\gt 0.5$ bins, which are contaminated by the photometric redshift uncertainty, we are not able to extract information on the radial cosmic distance. This is the first time that DA (z) is constrained at such a high redshift ($\bar{z}=0.87$) using LRGs.

Most of the recent works (Sánchez et al. 2011; Carnero et al. 2012; Seo et al. 2012) have used the angular correlation function w(θ) to get cosmic distance measures using several narrow redshift slices. Since radial binning blends data beyond what is induced by the photometric redshift error, the full information that is present is not utilized. Another important aspect that is often ignored when calculating w(θ) is the cross correlation between the different redshift bins used, along with the complications that it brings with calculating the covariance matrix; that is, the computing time increases with the number of bins in θ and the number of redshift slices used.

The full spectroscopy DESI galaxy catalog will be available around 2025 and will cover the footprint observed by DECaLS, but with higher precision. Here we try to probe the cosmological signature imprinted in this photometric footprint map. Most BAO measurements (Anderson et al. 2012; Alam et al. 2017; Chuang et al. 2017) at low redshifts have supported the H0 measurement by the Planck experiment, and it becomes interesting to see whether DESI will provide a similar result or not. By using the information obtained on the angular diameter distance from the DECaLS samples at the two median redshifts along with prior information of the sound horizon from Planck, we try to provide a precursor for the H0 value expected from DESI. Although precise information on H0 is not possible from photometric redshift catalogs, we obtain a value of H0 = 66.58 ± 5.31 km s−1 Mpc−1 with a fractional error of 7.97%. Our value of H0 supports the H0 measured by all other BAO results and is consistent with the ΛCDM model.

We would like to thank Alfonso Veropalumbo for providing us specific details on the empirical fitting procedure using the MCMC analysis. We would also like to thank Behzad Ansarinejad for general discussions on the BAO. Data analysis was performed using the high-performance computing cluster POLARIS at the Korea Astronomy and Space Science Institute. This research made use of TOPCAT and STIL: Starlink Table/VOTable Processing Software developed by Taylor (2005) and also the Code for Anisotropies in the Microwave Background (CAMB; Lewis et al. 2000; Howlett et al. 2012). Srivatsan Sridhar would also like to thank Sridhar Krishnan, Revathy Sridhar, and Madhumitha Srivatsan for their support and encouragement during this work.

The Photometric Redshifts for the Legacy Surveys (PRLS) catalog used in this paper was produced thanks to funding from the U.S. Department of Energy Office of Science, Office of High Energy Physics, via grant DE-SC0007914.

The Legacy Surveys consist of three individual and complementary projects: the Dark Energy Camera Legacy Survey (DECaLS; NOAO Proposal ID 2014B-0404; PIs: David Schlegel and Arjun Dey), the Beijing-Arizona Sky Survey (BASS; NOAO Proposal ID 2015A-0801; PIs: Zhou Xu and Xiaohui Fan), and the Mayall z-band Legacy Survey (MzLS; NOAO Proposal ID 2016A-0453; PI: Arjun Dey). DECaLS, BASS, and MzLS together include data obtained, respectively, at the Blanco telescope, Cerro Tololo Inter-American Observatory, the National Optical Astronomy Observatory (NOAO); the Bok telescope, Steward Observatory, University of Arizona; and the Mayall telescope, Kitt Peak National Observatory, NOAO. The Legacy Surveys project is honored to be permitted to conduct astronomical research on Iolkam Du'ag (Kitt Peak), a mountain with particular significance to the Tohono O'odham Nation.

NOAO is operated by the Association of Universities for Research in Astronomy (AURA) under a cooperative agreement with the National Science Foundation.

This project used data obtained with the Dark Energy Camera (DECam), which was constructed by the Dark Energy Survey (DES) collaboration. Funding for the DES Projects has been provided by the U.S. Department of Energy, the U.S. National Science Foundation, the Ministry of Science and Education of Spain, the Science and Technology Facilities Council of the United Kingdom, the Higher Education Funding Council for England, the National Center for Supercomputing Applications at the University of Illinois at Urbana-Champaign, the Kavli Institute of Cosmological Physics at the University of Chicago, the Center for Cosmology and AstroParticle Physics at the Ohio State University, the Mitchell Institute for Fundamental Physics and Astronomy at Texas A&M University, Financiadora de Estudos e Projetos, Fundacao Carlos Chagas Filho de Amparo, Financiadora de Estudos e Projetos, Fundacao Carlos Chagas Filho de Amparo a Pesquisa do Estado do Rio de Janeiro, Conselho Nacional de Desenvolvimento Cientifico e Tecnologico and the Ministerio da Ciencia, Tecnologia e Inovacao, the Deutsche Forschungsgemeinschaft and the Collaborating Institutions in the Dark Energy Survey. The Collaborating Institutions are Argonne National Laboratory, the University of California at Santa Cruz, the University of Cambridge, Centro de Investigaciones Energeticas, Medioambientales y Tecnologicas-Madrid, the University of Chicago, University College London, the DES-Brazil Consortium, the University of Edinburgh, the Eidgenossische Technische Hochschule (ETH) Zurich, Fermi National Accelerator Laboratory, the University of Illinois at Urbana-Champaign, the Institut de Ciencies de l'Espai (IEEC/CSIC), the Institut de Fisica d'Altes Energies, Lawrence Berkeley National Laboratory, the Ludwig-Maximilians Universitat Munchen and the associated Excellence Cluster Universe, the University of Michigan, the National Optical Astronomy Observatory, the University of Nottingham, the Ohio State University, the University of Pennsylvania, the University of Portsmouth, SLAC National Accelerator Laboratory, Stanford University, the University of Sussex, and Texas A&M University.

The Legacy Survey team makes use of data products from the Near-Earth Object Wide-field Infrared Survey Explorer (NEOWISE), which is a project of the Jet Propulsion Laboratory/California Institute of Technology. NEOWISE is funded by the National Aeronautics and Space Administration.

The Legacy Surveys imaging of the DESI footprint is supported by the Director, Office of Science, Office of High Energy Physics of the U.S. Department of Energy under Contract No. DE-AC02-05CH1123; by the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility under the same contract; and by the U.S. National Science Foundation, Division of Astronomical Sciences under Contract No. AST-0950945 to NOAO.

Software: astropy (Astropy Collaboration et al. 2013; Price-Whelan et al. 2018), TOPCAT (Taylor 2005), emcee (Foreman-Mackey et al. 2013).

Appendix A: Systematic Contamination from Stellar Density

Systematic effects are often present in an imaging data set such as the DECaLS data set, which can lead to spurious fluctuations in the target density and in turn to changes in the shape of the redshift distribution. One such major contribution toward systematic contamination in the data comes from correlations with stellar density (Rezaie et al. 2020). To check for this systematic effect, we compare our DECaLS LRG density and the density of the random catalog with the density of Gaia stars. First, we convert the sky coordinates (R.A. and decl.) from our data and random catalog into HEALPIX pixels using the same nside = 256 as used for the Gaia stellar density maps. We then use the Pearson correlation coefficient (PCC) to assess the linear correlation between the two data sets. For two variables X and Y, the PCC is defined as

Equation (A1)

where ${cov}(X,Y)$ is the covariance between X and Y across all pixels. We get a value of ${\rho }_{X,Y}=-0.0416$ for X and Y being the Gaia stellar density and DECaLS LRG density and ${\rho }_{X,Y}=-0.0443$ for X and Y being the Gaia stellar density and random catalog density, which shows that there is almost no strong positive or negative correlation between the two data sets separately.

Appendix B: Comparison of the Clustering Results between DECaLS and EZmock

As mentioned in Section 2.3, we generate photo-z's for the EZmock samples from random σz values obtained from the DECaLS data by restricting to galaxies of similar redshifts. We calculate ${\xi }_{w}(s,\mu )$ for the 100 EZmock samples separately using the same s and μ binning scheme as mentioned in Section 3.4, and the mean ${\xi }_{w}(s,\mu )$ for both of the redshift samples is plotted as the solid blue line in Figure A1. We find that by using the true values of ${\xi }_{w}(s,\mu )$ for the two redshift samples from the EZmock, the amplitudes of ${\xi }_{w}(s,\mu )$ do not match. However, by adding a constant value (0.0018 for the $0.6\lt {z}_{\mathrm{phot}}\lt 0.8$ sample and 0.0011 for the 0.8 < zphot < 1.0 sample) to the ${\xi }_{w}(s,\mu )$, we see that the mean ${\xi }_{w}(s,\mu )$ of the values obtained from the EZmock photo-z's created by randomly selecting ${\sigma }_{z}$'s from the DECaLS sample match closely the DECaLS ${\xi }_{w}(s,\mu )$, especially at the BAO scales as shown in Figure A1. When adding the constant, we assume that there is some bias in the data, but we do not expect that it would change the covariance matrix.

Figure A1.

Figure A1. First row: comparison of the correlation function ${\xi }_{w}(s,\mu )$ (multiplied by s2 ) for $\bar{\mu }=0.08,0.25,0.42$ (from the $0.6\lt {z}_{\mathrm{phot}}\lt 0.8$ sample) between the DECaLS data (given by red dots) and the mean ${\xi }_{w}(s,\mu )$ of the 100 EZmock samples (given by the solid blue line). The photo-z's for all of the EZmock samples have been created by extracting a random σz from the parent DECaLS sample. The ${\xi }_{w}(s,\mu )$ from all 100 samples are plotted as lighter blue lines in the background. Note: We have added a constant value of 0.0018 to the EZmock ${\xi }_{w}(s,\mu )$ for the $0.6\lt {z}_{\mathrm{phot}}\lt 0.8$ sample for the amplitudes to match. Second row: the same as above, but for the 0.8 < zphot < 1.0 sample. We have added a constant value of 0.0011 to the EZmock ${\xi }_{w}(s,\mu )$ for the $0.8\lt {z}_{\mathrm{phot}}\lt 1.0$ sample so the amplitudes match.

Standard image High-resolution image
Figure A2.

Figure A2. Distribution of a Gaussian function (green solid line) and the distribution of $({z}_{\mathrm{phot}}-{z}_{\mathrm{spec}})/{\sigma }_{z}$ (blue filled histogram) that we use to create the photometric redshifts for our mocks. Here, σz is the estimated photo-z error within the redshift range 0.6 < zphot < 0.8.

Standard image High-resolution image

As a further test, we fit the mean ${\xi }_{w}(s,\mu )$ from the 100 EZmock samples for the two redshift bins using the MCMC technique with the same fitting parameter space as given by Equation (5). We apply the same priors to all our samples as mentioned in Section 3.1. The values of sm obtained from the MCMC fit for the two samples are given in Table B1. First we compare the sm value obtained from the EZmock samples with the same obtained from the DECaLS data. The errors obtained on sm for the EZmock samples are similar to what we have obtained for the DECaLS sample in both redshift ranges, with the error on sm increasing with $\bar{\mu }$.

Table B1. Results of Fitting the EZmock Correlation Function for the Two Redshift Samples and in the Three $\bar{\mu }$ Bins using Equation (4)

Redshift Range $\bar{\mu }$ Bin sm (h−1 Mpc)
 0.08 ${104.1}_{-4.0}^{+3.6}$
$0.6\lt {z}_{{phot}}\lt 0.8$ 0.25 ${105.8}_{-3.7}^{+3.5}$
 0.42 ${113.2}_{-6.1}^{+4.5}$
 0.08 ${105.6}_{-4.8}^{+4.5}$
$0.8\lt {z}_{{phot}}\lt 1.0$ 0.25 ${110.3}_{-3.5}^{+3.1}$
 0.42 ${113.6}_{-6.5}^{+5.7}$

Note. The sm is the BAO peak point obtained from the fit, and the units are in h−1 Mpc.

Download table as:  ASCIITypeset image

The obtained sm values from the EZmock samples are converted to physical distances using the same methodology as explained in Section 4. For the $0.6\lt {z}_{\mathrm{phot}}\lt 0.8$ and 0.8 < zphot < 1.0 samples, we obtain values of ${D}_{A}(0.69)=1495\,\pm 63\,\mathrm{Mpc}({r}_{d}/{r}_{d,\mathrm{fid}})$ and ${D}_{A}(0.87)=1599\pm 75\,\mathrm{Mpc}({r}_{d}/{r}_{d,\mathrm{fid}})$. These values correspond to distance measures of 4.21% and 4.69% precision for the two redshift samples, respectively. It can also be noted that the recovered values of DA for the EZmock samples are a good match to the expected values DA fid from the fiducial cosmology we use. The fiducial values of DA fid for our cosmology at the two mean redshifts are ${D}_{A}^{\mathrm{fid}}(\bar{z}=0.697)=1514$ Mpc (1.3% higher than the recovered value) and ${D}_{A}^{\mathrm{fid}}(\bar{z}=0.874)=1638$ Mpc (2.3% higher than the recovered value). Given the overall precision, these values are well within the expected limits.

Footnotes

Please wait… references are loading.
10.3847/1538-4357/abc0f0