MNRAS 000,1–15 (2020) Preprint 23 December 2020 Compiled using MNRAS LATEX style file v3.0

Testing General Relativity on cosmological scales at 푧 ∼ 1.5 with quasar and CMB lensing

Yucheng Zhang ,1★ Anthony R. Pullen,1,2 Shadab Alam ,3 Sukhdeep Singh,4 Etienne Burtin,5 Chia-Hsun Chuang,6 Jiamin Hou,7 Brad W. Lyke,8 Adam D. Myers,8 Richard Neveux,5 Ashley J. Ross,9 Graziano Rossi,10 Cheng Zhao11

1 Center for Cosmology and Particle Physics, Department of Physics, New York University, 726 Broadway, New York, NY 10003, USA 2 Center for Computational Astrophysics, Flatiron Institute, New York, NY 10010, USA 3 Institute for Astronomy, University of Edinburgh, Royal Observatory, Blackford Hill, Edinburgh, EH9 3HJ, UK 4 Berkeley Center for Cosmological Physics, University of California, Berkeley, CA 94720, USA 5 IRFU,CEA, Université Paris-Saclay, F-91191 Gif-sur-Yvette, France 6 Kavli Institute for Particle Astrophysics and Cosmology, Stanford University, 452 Lomita Mall, Stanford, CA 94305, USA 7 Max-Planck-Institut für Extraterrestrische Physik, Postfach 1312, Giessenbachstrasse 1, 85748 Garching bei München, Germany 8 University of Wyoming, 1000 E. University Ave., Laramie, WY 82071, USA 9 Center for Cosmology and Astro-Particle Physics, Ohio State University, Columbus, Ohio, USA 10 Department of Physics and Astronomy, Sejong University, Seoul, 143-747, Korea 11 Institute of Physics, Laboratory of Astrophysics, École Polytechnique Fédérale de Lausanne (EPFL), Observatoire de Sauverny, 1290 Versoix, Switzerland

Accepted XXX. Received YYY; in original form ZZZ

ABSTRACT We test general relativity (GR) at the effective redshift 푧¯ ∼ 1.5 by estimating the statistic 퐸퐺, a probe of gravity, on cosmological scales 19 − 190 ℎ−1Mpc. This is the highest redshift and largest scale estimation of 퐸퐺 so far. We use the quasar sample with 0.8 < 푧 < 2.2 from Sloan Digital Sky Survey IV extended Baryon Oscillation Spectroscopic Survey Data Release 16 as the large-scale structure (LSS) tracer, for which the angular power spectrum 푞푞 퐶ℓ and the redshift-space distortion parameter 훽 are estimated. By cross-correlating with the Planck 2018 cosmic microwave background (CMB) lensing map, we detect the angular cross- 휅푞 power spectrum 퐶ℓ signal at 12 휎 significance. Both jackknife resampling and simulations are used to estimate the covariance matrix (CM) of 퐸퐺 at five bins covering different scales, with the later preferred for its better constraints on the covariances. We find 퐸퐺 estimates agree with the GR prediction at 1 휎 level over all these scales. With the CM estimated with 300 simulations, we report a best-fitting scale-averaged estimate of 퐸퐺 (푧¯) = 0.30 ± 0.05, which is GR ( ) in line with the GR prediction 퐸퐺 푧¯ = 0.33 with Planck 2018 CMB+BAO matter density fraction Ωm = 0.31. The statistical errors of 퐸퐺 with future LSS surveys at similar redshifts will be reduced by an order of magnitude, which makes it possible to constrain modified gravity models. Key words: cosmology: theory – cosmology: observations – large-scale structure of Universe – gravitation – gravitational lensing: weak – cosmic background radiation arXiv:2007.12607v2 [astro-ph.CO] 22 Dec 2020

1 INTRODUCTION explain the cosmic expansion and acceleration (see e.g. Silvestri & Trodden 2009, and references therein), among which Λ-cold dark The expansion of the universe was first discovered by measuring matter (ΛCDM) has been regarded as the standard model for its the redshifts and relative distances of (Hubble 1929). One simplicity and success in explaining a wide range of cosmological of the milestones in cosmology in the past decades has been the observations, including the CMB (cosmic microwave background) detection of a negative deceleration parameter and hence the ac- surveys (e.g. Planck Collaboration VI 2018) and redshift sur- celerated expansion of the Universe at late times from supernovae veys (e.g. Alam et al. 2017b). ΛCDM takes general relativity (GR) observations (Riess et al. 1998; Perlmutter et al. 1999). Many the- as the true theory for gravity on both galactic and cosmological oretical models of cosmology and gravity have been proposed to scales, and assumes the existence of the cosmological constant (Λ), a special form of dark energy (DE) whose spatially uniform energy density does not evolve with cosmic expansion, and CDM, along ★ E-mail: [email protected]

© 2020 The Authors 2 Y. Zhang et al. with ordinary (baryonic) matter. Although the expansion history can order to do certain estimations and generate the simulations needed. be well described by ΛCDM-GR by fine-tuning the relative density So for now it is very difficult to design one blind test for various ratios of the energy components, the nature of dark matter (DM) and gravity models. To do a rigorous estimation of 퐸퐺 based on other DE are not well understood and their properties are hard to detect MG models, the corresponding changes have to be made for either with observations. On the other hand, some modified gravity (MG) simulations or analytic calculations (see e.g. Hojjati et al. 2011). models (see e.g. Carroll et al. 2005; Sotiriou & Faraoni 2010; Dvali The paper is organized as follows. In Section2, we review the et al. 2000), which can predict the same expansion history of the 퐸퐺 theory and describe the estimator we use. The quasar and CMB universe as ΛCDM-GR with or completely without assuming the data, simulations and jackknife resampling for the estimation of existence of DE, have been developed to challenge GR as the true covariance matrices are described in Section3. Section4 includes theory for gravity on cosmological scales. There have been some analytic models, estimators, systematics and calibrations for the great reviews of the two approaches, see e.g. Peebles & Ratra(2003) angular power spectra. Section5 describes our estimation of the for the cosmological constant and DE, Clifton et al.(2012) for MG, quasar 2-point correlation function and the maximum likelihood and Joyce et al.(2016) for a comparison. fitting of the redshift-space distortion (RSD) parameter. We present Despite the degeneracy between ΛCDM-GR and MG models all the estimates and our final results in Section6 and conclude in in explaining the cosmic expansion, their predictions of the growth Section7. of the DM large scale structure (LSS) are usually distinguishable. For our self-consistency test of GR, wherever needed, we Combining the gravitational lensing ∇2 (Ψ − Φ) and the divergence assume a flat ΛCDM fiducial cosmology with Planck 2018 of the peculiar velocity 휃, Zhang et al.(2007) proposed a statistic CMB+BAO parameters (Planck Collaboration VI 2018): Ωm = 2 2 퐸퐺 as a function of redshift and scale, to probe gravity on cosmo- 0.3111 ± 0.0056, Ωcℎ = 0.11933 ± 0.00091, Ωbℎ = 0.02242 ± logical scales. Lensing is related to the underlying matter overden- 0.00014, 푛s = 0.9665 ± 0.0038, 퐻0 = 67.66 ± 0.42, and 휎8 = sity 훿 through the Poisson equation which depends on the gravity 0.8102 ± 0.0060. model (see e.g. Hojjati et al. 2011). On linear scales, 휃 = − 푓 훿, where 푓 is the linear growth rate. In real surveys, instead of the DM field, the direct observables are the LSS tracers, e.g. galaxies or quasars. The distribution of these tracers is connected to the un- 2 퐸 FORMALISM AND ESTIMATOR derlying matter perturbation field with the clustering bias 푏, which 퐺 varies with the physical properties of the tracers that are targeted in In this section, we briefly review the 퐸퐺 theory and describe the a particular survey. Defined as the ratio between ∇2 (Ψ − Φ) and 휃, estimator used in this work. We assume a flat Universe described 퐸퐺 has the advantage of being independent of 푏 and the variance by the perturbed Friedmann-Robertson-Walker (FRW) metric in of the matter density field 휎8. Interested readers may refer to Ishak conformal Newtonian gauge, (2019) for a comprehensive review on various cosmological tests of 2 h 2 2i GR, including the 퐸퐺 statistic. 푑푠 = 푎(휏) (1 + 2Ψ)푑휏 − (1 + 2Φ)푑푥 , (1) The estimation of 퐸퐺 requires data from both gravitational lensing and redshift surveys. Accurate estimates of tracers’ red- where Ψ and Φ are the scalar perturbations to the time and spatial shifts are necessary in order to do the 3-D clustering analysis, from components of the metric. The statistic 퐸퐺 is defined in Fourier which the growth of the structure can be probed. Thus spectro- Space (Zhang et al. 2007) as scopic redshift surveys are usually preferred. For photometric sur- " 2 # veys, Giannantonio et al.(2016) proposed a statistic 퐷 , which does ∇ (Ψ − Φ) 퐺 퐸퐺 (푘, 푧) = not require the estimation of the growth rate. However, this quan- −3퐻2 (1 + 푧)휃 0 푘 tity cannot be directly used to discriminate GR and MG models. (2) 푘2 (Ψ − Φ) Using galaxy-galaxy lensing and luminous red galaxies (LRGs), = , −1 3퐻2 (1 + 푧)휃 퐸퐺 has been measured over scales . 70 ℎ Mpc at redshifts in 0 0.2 < 푧 < 0.6 (Reyes et al. 2010; Blake et al. 2015; de la Torre et al. where 퐻 is the Hubble constant and 휃 = ∇·풗/퐻(푧) is the divergence 2017; Alam et al. 2017a; Amon et al. 2018; Singh et al. 2018; Blake 0 of the comoving peculiar velocity field. In linear perturbation theory, et al. 2020). Besides tracing the lensing signal with background 휃 = − 푓 훿, where 푓 is the linear growth rate and 훿 is the matter galaxies, Pullen et al.(2015) proposed to use the cosmic microwave perturbation. For GR, assuming no anisotropic stress (Φ = −Ψ) and background (CMB) lensing map, which allows the estimation of 2 2 using Poisson equation ∇ Ψ = 4휋퐺푎 휌푚훿, we have 퐸퐺 at higher redshifts and larger scales (Pullen et al. 2016; Singh et al. 2018). Ω푚,0 퐸GR (푧) = , (3) In this work, using quasars and CMB lensing, we test ΛCDM- 퐺 푓 (푧) GR on cosmological scales 19−190 ℎ−1Mpc at the effective redshift / 푧¯ ∼ 1.5, which is the highest redshift and largest scale 퐸퐺 estima- where Ω푚,0 = 휌푚,0 휌crit,0 is the fraction of matter density today = 2/ ( )' Ω ( )훾 ' tion so far. Quasars, also known as quasi-stellar objects (QSOs), with 휌crit,0 3퐻0 8휋퐺, and 푓 푧 푚 푧 with 훾 0.55 and are active galactic nuclei (AGN) with very high luminosity, which 3 makes them good candidates to trace LSS at higher redshifts (e.g. Ω푚,0 (1 + 푧) Ω (푧) = (4) 푚 3 1 < 푧 < 2). As part of the primary motivation of constraining 퐸퐺, Ω푚,0 (1 + 푧) + (1 − Ω푚,0) we also investigate the reliability of quasars as a tracer of the DM GR ( ) in both auto- and cross-clustering analyses. The redshift range of at late times. Notice that 퐸퐺 푧 is scale-independent and only the quasar targets is very close to the peak of CMB lensing kernel relies on the relative fraction of matter density in the Universe. at 푧 ∼ 2, so we should expect a promising cross-correlation signal, For different MG models, 퐸퐺 can have different amplitudes or be which is usually harder to be detected than the auto-correlation. As- scale-dependent (Zhang et al. 2007; Pullen et al. 2015). sumptions of the cosmology and gravity models have to be made in The angular estimator for 퐸퐺 at the effective redshift 푧¯ can be

MNRAS 000,1–15 (2020) Testing GR with quasar and CMB lensing 3

constructed as (Pullen et al. 2015) NGC 260 , 50 N 휅푞 ◦ ◦ 푐2 퐶 120◦, 60◦ N ˆ ( ) ℓ 퐸퐺 ℓ 푧¯ = 3퐻2 퐶 휃푞 0 ℓ 푧¯ (5) 180◦, 60◦ N 퐶휅푞 ' Γ(푧¯) ℓ , ( ) 푞푞 240◦, 30◦ N 훽 푧¯ 퐶ℓ 120 , 20 N where 푐 is the speed of light, 휅 and 푞 denote the CMB lensing ◦ ◦ convergence and quasar overdensity maps respectively. 퐶 ’s are the ℓ 180◦, 30◦ N angular power spectra, 훽 = 푓 /푏 is the RSD parameter given by the ratio of the linear growth rate and the clustering bias, and Γ is an analytic factor, 1 56 2푐 퐻(푧¯) 푓푞 (푧¯) Γ(푧¯) = , (6) SGC 3퐻2 (1 + 푧¯)푊 (푧¯) 0 0◦, 40◦ N

where 푓푞 (푧¯) is the normalized redshift distribution of the quasar 320◦, 30◦ N 40◦, 30◦ N sample at the effective redshift and 푊 (푧) is the CMB lensing kernel. 푓푞 (푧) and 푊 (푧) work as the radial projection kernels for 푞 and 휅 fields when we transform the 3-D power spectra 푃(푘, 푧) into 휃푞 angular 퐶ℓ ’s, as shown in Eq. 17 and Eq. 19. To convert 퐶ℓ to the 푞푞 directly measurable 퐶ℓ , the approximation made in Eq.5 which includes the substitution of a certain redshift-dependent factor with

the effective value at 푧¯ is not perfect. This can cause a systematic 320◦, 10◦ S 40◦, 10◦ S 0◦, 10◦ S bias around 5% to our 퐸퐺 estimation. Following Pullen et al.(2016) and assuming a scale-independent linear bias 푏(푧), we introduce the calibration factor 1 35 푚푞 푐 푊 (푧¯)(1 + 푧¯) 퐶ℓ 퐶Γ = 푚푞 , (7) 2 퐻(푧¯) 푓푞 (푧¯) 푄 ℓ Figure 1. The overlapped sky coverage of Planck 2018 CMB lensing and where eBOSS DR16 quasar NGC (upper) and SGC (lower) clustering catalogs. 2 ∫ 푧2   NGC (SGC) covers about 2929 (1815) deg . The orientation of the regions 푚푞 −2 퐻(푧) 2 ℓ + 1/2 퐶 ≡ 푑푧휒 (푧) 푓 (푧)푏(푧)푃푚 , 푧 , (8) are shown in J2000 coordinates. For jackknife resampling, NGC and SGC ℓ 푐 푞 휒(푧) 푧1 are divided into 56 and 35 equally-weighted regions respectively. and 1 ∫ 푧2  ℓ + 1/2  푄푚푞 ≡ 푑푧(1 + 푧) 휒−2 (푧)푊 (푧) 푓 (푧)푏(푧)푃 , 푧 , 3 DATA AND COVARIANCES ℓ 푞 푚 ( ) 2 푧1 휒 푧 (9) In this section, we describe the quasar and CMB lensing data used in this work. We also discuss the simulations and the jackknife where 휒(푧) is the radial comoving distance at redshift 푧, 푃푚 is the resampling method used to estimate the covariance matrices. matter power spectrum and Limber approximation 푘 휒 ' ℓ +1/2 has been used. Due to the limited size of the quasar sample, it is hard to study the redshift evolution of the bias by cutting the redshift range 3.1 Quasar catalogs into a few smaller bins. Here we just take a constant bias at the effective redshift, i.e. 푏(푧)' 푏(푧¯). We also tried an eBOSS quasar We use the quasar sample for clustering analysis from the fourth bias model presented in Laurent et al.(2017), and the difference is phase of the Sloan Digital Sky Survey (SDSS-IV) (Blanton negligible considering that the systematic bias calibrated by 퐶Γ is et al. 2017) extended Baryon Oscillation Spectroscopic Survey only around 5% of the 퐸퐺 signal. Another systematic bias concern (eBOSS) (Dawson et al. 2016) Data Release 16 (DR16) (Ahumada is the non-linear quasar bias and the imperfect connection between et al. 2019), which is observed with the Sloan Foundation 2.5-meter quasars and the matter field at small scales, which is hard to model Telescope located at the Apache Point Observatory (Gunn et al. and needs to be corrected with N-body simulations. However, for 2006) with double-armed spectrographs (Smee et al. 2013). The the scales (≥ 19 ℎ−1Mpc) we are considering, this systematic bias construction of these eBOSS DR16 clustering catalogs for quasars should be negligible (Pullen et al. 2016; Singh et al. 2018). from the complete SDSS DR16 quasar (DR16Q) catalog (Lyke et al. The correspondence between multipoles ℓ and linear scales 2020) is described in Ross et al.(2020), along with the catalogs for 휒⊥ at a certain redshift is given by 휒⊥ = 2휋휒(푧)/ℓ. With Eq.5, luminous red galaxies (LRGs) and emission line galaxies (ELGs). we can estimate 퐸ˆ퐺 (ℓ) for a range of multipoles. These multipoles The quasar sample comprises the north galactic cap (NGC) and the are binned into a few bandpowers in practice, with more details south galactic cap (SGC), which correspond to two separate regions discussed in the estimation of 퐶ℓ ’s (Section 4.2). In the end, we find on the sky. Since jackknife resampling is used for covariance es- the best-fit (denoted as 퐸¯퐺) of the overall amplitude of 퐸ˆ퐺 (ℓ) to timation (see Section 3.3), we only use the sky region covered by compare to the GR prediction. To make the discussion coherent, we both the quasar and CMB lensing surveys (Fig.1). The sky cov- present our fitting method along with our estimates of the covariance erage fraction and number of quasars are shown in Table1. This matrix for 퐸퐺 (ℓ) in Section 6.4. overlapped coverage masks out around 3.4 % quasars in NGC and

MNRAS 000,1–15 (2020) 4 Y. Zhang et al.

Table 1. The overlapped sky coverage fraction of eBOSS DR16 quasar which is proposed for the measurement of the 2-point correlation catalogs and Planck 2018 CMB lensing, and the corresponding number of function and the summation is conducted over pairs with separation − quasars. The (weighted) mean and median redshifts agree with each other distance 25 ≤ 푠 ≤ 120 Mpc 1ℎ. With this definition, Hou et al. (see text), denoted as 푧¯. The last column shows the number of quasars which (2020) find 푧eff ' 1.48 for the full clustering quasar sample. For are in the original eBOSS DR16 catalogs but not covered by the Planck both NGC and SGC quasar samples used in this work, we find that CMB lensing footprint, and hence these targets are not included in the data Í Í the mean, weighted mean ( 푖 푤푖 푧푖/ 푖 푤푖) and median redshifts analysis. agree with each other, with the value shown as 푧¯ in Table1. Although the overlapped mask removes some quasars, these redshift values Cap 푓sky (%) # quasars 푧¯ # masked almost remain the same. The tiny difference in the definitions of NGC 7.1 210 881 1.51 7 328 the effective redshift is completely negligible compared with the SGC 4.4 116 249 1.52 9 250 statistical accuracy. Thus for simplicity, in this work, we take the effective redshift at 푧¯ = 1.5 for both NGC and SGC.

0.8 NGC 3.2 CMB lensing map 2 SGC

deg The gravitational lensing convergence (휅) map used is the minimum- 0.6 /

) variance estimate with CMB temperature and polarization measure- z ments (Planck Collaboration VIII 2018), reconstructed and provided 01 . as part of the Planck 2018 data release (Planck Collaboration I

(0 0.4 / 2018). The map covers about 70 percent of the sky and is provided in spherical harmonics 휅ℓ푚’s up to ℓ = 4096. However, in this work, 0.2 we only use the multipoles in 8 ≤ ℓ ≤ 2048. We do not use the multipoles ℓ > 2048 due to the significant reconstruction noise at # quasars those very small scales. Since we are only considering well-defined 0.0 linear scales 100 ≤ ℓ ≤ 1000 for the angular cross-power spectra 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 z estimation of quasar and CMB lensing, contributions from those much smaller and non-linear scales should be negligible compared with the statistical errors. Figure 2. Number density redshift distribution of eBOSS DR16 quasar clustering catalogs (with the overlapped sky coverage with CMB lensing applied, see text). NGC has a higher number density than SGC, which 3.3 Covariance matrices results in lower shot noise. Covariance matrices (CMs) are needed for constructing the likeli- hood functions used in the posterior distribution sampling of the 7.4 % quasars in SGC. Even without jackknife resampling, using parameters, e.g. RSD parameters and scale-averaged 퐸¯퐺. Like any this total mask is still reasonable since the removed quasars do not other statistics, an accurate estimation of the CM relies on a large have the corresponding lensing signal anyway. number of samples. In this work, we estimate the CMs in two ways. Using the HEALPix (Górski et al. 2005) pixelization, we con- One is using simulations, and the other is jackknife resampling, struct the quasar overdensity map with which only depends on the data itself. 푛 For simulations, we run all of them through the same data 훿 = 푖 − 1 , (10) 푖 푛¯ analysis pipeline as we do for the real data, with which we can then construct the CM for any statistical quantity in the procedure. We where 푖 is the pixel index, 푛 = Í 푤 is the weighted num- 푖 푞∈푖 푞 use 300 simulated 휅 maps coming with Planck 2018 CMB lens- ber count of quasars for each pixel and 푛¯ is the the average ing analysis (Planck Collaboration VIII 2018), in which the lensing over all covered pixels. The weight for each quasar is given by reconstruction noise is included. For the eBOSS quasar sample, 푤 = 푤 · 푤 · 푤 , where 푤 · 푤 corrects for the spectro- 푞 sys cp noz cp noz Zhao et al.(2020) generated 1000 effective Zel’dovich approxima- scopic completeness due to close pairs and redshift failures across tion mock catalogs (EZ mocks, Chuang et al. 2015). The fiducial fibers, and 푤 accounts for the imaging systematics. Additionally, sys cosmology for generating the mocks is flat ΛCDM with parame- for the estimation of the correlation function, 푤 is also applied FKP ters: Ω = 0.307115, Ω = 0.048206, ℎ = 0.6777, 휎 = 0.8225, to optimize the clustering statistics (Feldman et al. 1994). The de- m b 8 and 푛 = 0.9611. These are slightly different from the Planck 2018 termination of all these weights is described in detail in Ross et al. s CMB+BAO parameters we are assuming, but the influence on the (2020). CMs should be negligible. Combining these simulated 휅 maps and The redshift distribution of the two catalogs are shown in Fig.2, EZ mocks, we have 300 sets of independent simulations for our where we see that NGC has a higher number density than SGC. So 퐸 analysis. These lensing maps and quasar mocks are not corre- the shot noise due to the Poisson distribution of the quasars, which is 퐺 lated, which results in zero mean signal and lower error estimates inversely proportional to the number density, is lower for NGC than 휅푞 (Eq. 20) for the cross correlation 퐶 . As discussed in Section 6.2, SGC. The quasars are observed in redshift bin 0.8 < 푧 < 2.2, for ℓ the contribution of 퐶휅푞 signal to the error distribution of itself is which we need to determine the effective redshift for our angular ℓ negligible compared with the noise level in the auto correlation of analysis. The recommended definition of the effective redshift in the current surveys. However, the 퐶휅푞 signal is important in the eBOSS DR16 clustering analysis is given by ℓ CM estimation for functions of it like 퐸 , which can be seen from Í 퐺 푖, 푗 푤푖푤 푗 (푧푖 + 푧 푗 )/2 the Gaussian error propagation. We discuss our approach to fix this 푧eff = Í , (11) 푖, 푗 푤푖푤 푗 issue in Section 6.4, where the estimates of CMs for 퐸퐺 (ℓ) are

MNRAS 000,1–15 (2020) Testing GR with quasar and CMB lensing 5 presented. Although using realistic simulations is a reliable way to It should be noticed that the above corrections are derived for inde- estimate CMs since we can run as many simulations as needed (with pendent samples like the simulations, which may not be the proper enough computing resources), it should still be reiterated that sim- solution for jackknife samples (Taylor et al. 2013). However, more ulations depend on the fiducial model, where extra consideration is detailed discussion is out of the scope of this paper, which we leave necessary for the purpose of testing different models on the data. for future work. Specifically in this work, two CMs are used for fit- Another CM estimation method which only relies on the data ting purposes. One is for the 2-point correlation function of quasars sample is jackknife resampling. In this work, we divide the over- in RSD fitting and the other is for the fitting of 퐸퐺 (ℓ) over scales. lapped sky coverage of quasar and CMB lensing into 푁 equally weighted regions and make leave-one-out jackknife samples by tak- ing one region out each time. This process leaves us 푁 correlated re-samples of the original full data. We do the analysis for each of 4 ANGULAR POWER SPECTRA these jackknife samples, with each result denoted as a vector 풙, e.g. In this section, we describe the theoretical models and estimators for the correlation function or power spectrum. Then the covariance the angular power spectra. We also discuss the influence of possible matrix of 풙 is given by systematics and the corresponding calibrations applied. 푁 푁 − 1 ∑︁  ( )   ( )  Cov(푥 , 푥 ) = 푥 푘 − 푥¯ 푥 푘 − 푥¯ , (12) 푖 푗 푁 푖 푖 푗 푗 푘=1 4.1 Theory where 풙¯ is the mean of all the jackknife estimates, which are labeled The analytic expressions for the angular power spectra can be de- with index 푘. Compared with the normal unbiased sample CM es- rived by integrating the 3-D power spectra 푃(푘, 푧) over the wave 2 timation, a factor of (푁 − 1) /푁 is multiplied, which corresponds number 푘, with proper radial projection kernels 퐹(휒) applied. At to the fact that the jackknife samples are not independent. Jackknife high ℓ’s (e.g. ℓ > 10 is good enough for the wide redshift bin of resampling has the advantage of being dependent only on the data, the quasar sample), the spherical Bessel functions 푗ℓ (푘 휒) vary fast which hence naturally includes all the systematics and noise in the compared with 퐹(휒), which picks out the scale 푘 '(ℓ +1/2)/휒(푧). observations. However, the maximum number of jackknife samples Based on this, the Limber approximation replaces the 푗ℓ with the is limited by the largest scale to be probed. In this work, by requiring Dirac delta function, which significantly speed up the numerical the linear scale of each region to be at least two times the largest evaluation of the integral. In what follows, this approximation is scale we are interested in, we are able to use 56 (35) jackknives always applied. for NGC (SGC) (Fig.1). We make sure that jackknife resampling The CMB lensing × quasar cross-power spectrum reads is unbiased by comparing the mean of all the jackknife estimates ∫ 푧2   with the estimate using the full data sample. It turns out that for 휅푞 1 −2 ℓ + 1/2 퐶 = 푑푧휒 (푧)푊 (푧) 푓 (푧)푃 2 , 푧 , the statistics in this work, they are always consistent. However, the ℓ 푞 ∇ (Ψ−Φ)푞 ( ) 2 푧1 휒 푧 number of jackknives used may not be enough to give us accurate (17) estimates of the CMs, especially the off-diagonal terms (i.e. cross correlations between different scales), whose relative strength com- where 휒(푧) is the radial comoving distance at redshift 푧, 푊 (푧) = pared with variances (diagonal terms) can be quantified with the 휒(푧) [1 − 휒(푧)/휒 (푧CMB)] is the CMB lensing kernel with 푧CMB ' 1 푑푁 correlation matrix, 1100, 푓푞 (푧) = 푁 푑푧 is the normalized quasar redshift distribution ( ) C푖 푗 and 푃∇2 (Ψ−Φ)푞 푘, 푧 is the 3-D cross-power spectrum of the two Corr (C)푖 푗 = , (13) fields. Assuming GR and using the Poisson equation to replace the √︁C C 푖푖 푗 푗 lensing convergence with matter perturbation, Eq. 17 can be written where C denotes the CM. as ˆ The estimated C for a multivariate Gaussian vector with a 2 3Ω 퐻 ∫ 푧2 limited number of samples follows the Wishart distribution, which 휅푞 푚,0 0 −2 퐶 = 푑푧휒 (푧)(1 + 푧)푊 (푧) 푓푞 (푧) ˆ −1 ℓ 2푐2 is an unbiased estimate of the true CM, C. However, C , the 푧1 (18) inverse of Cˆ , which obeys the inverse Wishart distribution, is a  ℓ + 1/2  −1 ˆ × 푃푚푞 , 푧 . biased estimate of C due to the error in C. This can be corrected 휒(푧) with a simple factor (Hartlap et al. 2007), Similarly, the quasar auto-power spectrum is given by  푁 + 1  ˆ −1 = − 푑 ˆ −1 Cunbiased 1 C , (14) ∫ 푧2 퐻(푧)  ℓ + 1/2  푁푠 − 1 푞푞 = −2 ( ) 2 ( ) 퐶ℓ 푑푧휒 푧 푓푞 푧 푃푞푞 , 푧 , (19) 푧 푐 휒(푧) where 푁푑 is the size of the data vector and 푁푠 is the number of 1 samples. Furthermore, the error in Cˆ propagates to the CM of the where 퐻(푧) is the Hubble parameter at redshift 푧. On linear scales, model parameters in the maximum likelihood fitting (Dodelson & the quasar overdensity field 훿푞 (푘, 푧) is connected to the under- Schneider 2013). This can be corrected by multiplying the factor lying matter perturbation 훿푚 (푘, 푧) with a local bias, 훿푞 (푘, 푧) = 1 + 퐵(푁푑 − 푁푝) 푏(푧)훿푚 (푘, 푧), where 푏(푧) is the linear bias of the quasar sample at 푀 = (15) redshift 푧. In general, 푏(푧) could also be scale-dependent due to pri- 1 + 퐴 + 퐵(푁푝 + 1) mordial non-Gaussianity (PNG, Dalal et al. 2008), which introduces to the CM of the parameters (Percival et al. 2014), where 푁푝 is the an additional scale-dependent part whose amplitude is parameter- number of parameters and ized by 푓NL. This scale dependence for local PNG decreases rapidly 2 (∝ 푘−2) with scale, and given the latest constraint 푓 = −0.9 ± 5.1 퐴 = , NL (푁 − 푁 − 1)(푁 − 푁 − 4) reported by Planck Collaboration IX(2020), we should be allowed 푠 푑 푠 푑 (16) 푁 − 푁 − 2 to assume a scale-independent bias. It is worth mentioning that our 퐵 = 푠 푑 . (푁푠 − 푁푑 − 1)(푁푠 − 푁푑 − 4) estimators of the angular power spectra (Section 4.2 below) do not

MNRAS 000,1–15 (2020) 6 Y. Zhang et al.

directly rely on these theoretical predictions, and a redshift depen- NAMASTER 2 (Alonso et al. 2019) to do the computation. For the dent 푏(푧) model may only matter in the 퐶Γ calibration, as discussed maps used in this work, the results of this more complicated PCL below Eq.7. estimator are consistent with the results given by the simpler version 휅푞 휅푞 휅푞 Later in our analysis it will be useful to have analytic expres- 퐶ˆ푝 ' 퐷ˆ 푝 / 푓 , where the couplings between modes due to the 휅푞 푞푞 sky sions for statistical errors of 퐶ℓ and 퐶ℓ . Assuming 휅 and 푞 to be geometry of the masks are ignored. 휅푞 Gaussian fields, the sample variance of 퐶ℓ can be approximated For the cross correlation, the noise in the two maps from sep- as arate surveys are usually uncorrelated and only contributes to the statistical error without causing systematic bias. However, the situ- 1 h i 휎2 (퐶휅푞) = (퐶휅푞)2 + (퐶휅 휅 + 푁 휅 휅 )(퐶푞푞 + 푁푞푞) ation is more complicated for the auto correlation because the noise ℓ ( + ) 휅푞 ℓ ℓ ℓ ℓ ℓ 2ℓ 1 푓sky may not be correlated to the signal but is obviously correlated with (푟2 + 1) itself and hence can significantly bias the signal. Thus for the es- ℓ 휅 휅 휅 휅 푞푞 푞푞 푞푞 = (퐶 + 푁 )(퐶 + 푁 ) , timation of 퐶푝 , instead of the PCL estimator, we use the optimal ( + ) 휅푞 ℓ ℓ ℓ ℓ 2ℓ 1 푓sky quadratic minimum variance (QMV) estimator which marginalizes (20) over the noise (Tegmark 1997). We denote the pixelated quasar over- density map with an 1-D vector 풙 and the corresponding covariance ≡ 휅푞/[( 휅 휅 + 휅 휅 )( 푞푞 + 푞푞)]1/2 where 푟ℓ 퐶ℓ 퐶ℓ 푁ℓ 퐶ℓ 푁ℓ is known as matrix with C. Defining the quadratic vector 휅 휅 the cross-correlation coefficient, 푁ℓ is the lensing reconstruction 푞푞 휅푞 noise, 푁 is the shot noise and 푓 is the overlapped sky coverage 1 † −1 휕C −1 ℓ sky 푄ˆ 푝 = 풙 C C 풙 , (25) 푞푞 2 휕퐶푝 fraction of the surveys. Similarly, we have the variance for 퐶ℓ , and the Fisher matrix 2 푞푞 2 푞푞 푞푞 2 휎 (퐶 ) = 푞 (퐶 + 푁 ) . (21)   ℓ (2ℓ + 1) 푓 ℓ ℓ 1 −1 휕C −1 휕C sky F푝 푝0 = Tr C C , (26) 2 휕퐶푝 휕퐶푝0 When the multipoles ℓ’s are averaged into bandpowers 푝’s (as dis- cussed below) weighted by inverse variance (i.e. minimum variance the estimator can be constructed as average of Gaussian random vectors assuming no covariances), the 푞푞 ∑︁ h −1i 퐶ˆ푝 = F 푄ˆ 푝0 . (27) uncertainty for the binned signal is given by 푝 푝0 푝0 −1/2   The shot noise is properly fitted and marginalized in the estimation. ∑︁ −2  휎(퐶푝) =  휎 (퐶 ) . (22) 6  ℓ  In this work, 풙 includes ∼ 10 pixels, which makes it computation- ℓ ∈푝    ally impossible to invert C directly. We use the conjugate gradient −1 These analytic uncertainties, which include the well known lens- method to iteratively evaluate C 풙 and the trace for the Fisher ma- ing reconstruction and shot noise, have been widely used in doing trix. This optimal QMV estimator has been used in previous CMB forecasts. So it would be useful to have them as references and and galaxy power spectra analysis, and we refer our readers to the compared to the statistical errors estimated with simulations and references (Padmanabhan et al. 2001, 2003; Padmanabhan et al. jackknife resampling. 2007; Hirata et al. 2004, 2008; Ho et al. 2008) for more details.

4.2 Estimators 4.3 Systematics & calibrations Due to the noise and computational complexity, it is neither neces- Compared with the theoretical predictions in Eq. 18 and 19, the sary nor possible to estimate 퐶ℓ for each multipole. Thus we bin estimated power spectra can be biased due to several aspects, most the multipoles into bandpowers, denoted with subscript 푝 1. Here 휅푞 푞푞 of which are hard to be corrected in the estimators above and hence we briefly describe the estimators we use for 퐶푝 and 퐶푝 . 휅푞 extra calibrations might be needed. For the quasars, the observed We estimate 퐶푝 with the Pseudo-퐶ℓ (PCL) estimator, flux and measured redshift from the photometric and spectroscopic ∑︁ h i surveys are distorted due to the foreground density perturbation 퐶ˆ휅푞 = M−1 퐷ˆ 휅푞 , (23) 푝 0 푝0 and RSD. There can also be bias due to redshift smearing. For 0 푝 푝 푝 the CMB survey, the temperature map can be contaminated by where M is the binned mode coupling matrix computed with the foregrounds, e.g. dust emission and point sources. Here we mainly 휅푞 masks of the two fields and 퐷ˆ 푝 is the binned cross-power spectrum focus on the bias caused by the distortion of quasar catalogs. The of the masked full sky maps, possible systematics due to contamination in CMB are discussed in AppendixA, where it is shown that the bias to 퐶ˆ휅푞 is negligible. ℓ ! ℓ ∑︁ 1 ∑︁ Our observed targets are distorted by the gravitational lens- 퐷ˆ 휅푞 = 푤 휅∗ 푞 , (24) 푝 ℓ 2ℓ + 1 ℓ푚 ℓ푚 ing of the foreground density perturbations. First, compared with ℓ ∈푝 푚=−ℓ the intrinsic value, the flux of an individual source can be either where 푤ℓ is the normalized weight of each multipole inside the bin, increased or decreased by lensing. This can cause bias to the flux 휅ℓ푚 and 푞ℓ푚 are the harmonics of the masked (i.e. with pixel values or magnitude based target selection of the clustering catalog. Also, set to 0 if not covered) 휅 and 푞 maps. We use the fast implementation the observed angular distribution of the targets can be magnified. These effects can be quantified by the magnification bias 푠 (Liu et al.

1 To be clear, we use 푝 in this subsection. But for simplicity and consistency with the theory, we still use subscript ℓ for the bandpowers in the following sections. 2 https://github.com/LSSTDESC/NaMaster

MNRAS 000,1–15 (2020) Testing GR with quasar and CMB lensing 7

2014; Hui et al. 2007), which can be measured with the slope of the since the redshift bin 0.8 < 푧 < 2.2 for the quasar sample is cumulative apparent magnitude function, wide, while the distortion only happens around the edges of the bin. Similarly, the bias due to redshift smearing error from the redshift 푑 log10 푛푞 (푚 < 푚∗) 푠 = , (28) fitting pipeline should also be negligible for these angular power 푑푚 푚=푚∗ spectra. where 푚∗ is the faint magnitude limit of the survey and 푛푞 (푚 < 푚∗) is the number of quasars that are apparently brighter than the sur- vey limit. Depending on the value of 푠 and the linear bias 푏, our 5 REDSHIFT-SPACE DISTORTION estimates of the power spectra can be more or less biased (Dizgah & Durrer 2016). Following Yang & Pullen(2018) (see Section 2 We estimate the RSD parameter 훽 of the quasar sample at the and Appendix A therein for the expressions using Limber approx- effective redshift by fitting an analytic model to the monopole and imation), we do the calibration by adding correction terms Δ퐶ˆ휅푞 quadrupole of the configuration space 2-point correlation function ℓ (2PCF). Δ ˆ푞푞 and 퐶ℓ to our estimates from Eq. 23 and Eq. 27. Besides 푠 and 푏, these corrections also depend on the measured CMB lensing ˆ휅 휅 auto-power spectrum 퐶ℓ . The target selection for eBOSS quasars 5.1 Two point correlation function includes the magnitude cutoff for two frequency bands, 푔 < 22 or 푟 < 22 (Myers et al. 2015). Looking into the apparent point spread The 2PCF is estimated with the standard Landy & Szalay estima- function (PSF) magnitudes, we find that the overall cutoff is mainly tor (Landy & Szalay 1993), dominated by the 푟 band. With Eq. 28, we get 푠 ' 0.1 for both NGC h퐷퐷i − 2h퐷푅i + h푅푅i 휅푞 푞푞 휉 = , (33) and SGC. The corrections are around 13% for 퐶ℓ and 7% for 퐶ℓ , h푅푅i which results in about 5% calibrations in 퐸퐺 (ℓ). RSD describes the distortion in the observed radial positions of where 퐷 is the data catalog, 푅 is the random catalog and h i denotes the targets due to peculiar velocities. Although the redshift details the normalized pair count between two catalogs. We sample the are mostly erased in the projection of 2D angular maps and 3D pair counts in (푠, 휇) bins, where 푠 is the separation distance and 휇 is the cosine of the angle between the line-of-sight (LOS) and clustering is usually used in RSD analysis (as discussed in the next −1 section), RSD could still bias the angular power spectra mainly due separation vectors. We use a bin size of 5 ℎ Mpc for 푠 and 0.01 for to the flow at the cutoff boundaries of the redshift bin. Starting from 휇. The multipoles are extracted by expanding the 2PCF in Legendre the additional RSD component in the window function (see Eq. 27 polynomials, in Padmanabhan et al. 2007) and using Limber approximation, the ∑︁ 휅푞 휉(푠, 휇) = 휉ℓ (푠)퐿ℓ (휇) , (34) bias on 퐶ℓ due to RSD can be described with a higher-order term, ℓ where 0 2 푧 3Ω푚퐻 ∫ 2 푐푑푧  ℓ + 1/2  ∫ 1 퐶휅푞,푟 = 0 퐾(ℓ, 휒) 푓 (푧)휒−2푊 (푧)푃 , 푧 , 2ℓ + 1 ℓ 2 퐻(푧) 푚 휒 휉ℓ (푠) = 휉(휇, 푠)퐿ℓ (휇)푑휇 . (35) 2푐 푧1 2 −1 (29) The two lowest order even multipoles, monopole 휉0 (푠) and where 푓 (푧) is the linear growth rate and quadrupole 휉2 (푠), are used in the following fitting process. The non-zero quadrupole results from the peculiar velocity due to grav- 2ℓ2 + 2ℓ − 1 퐾(ℓ, 휒) ≡ 휙(휒) ity and contains the information about the growth of the structure (2ℓ − 1)(2ℓ + 3) and RSD. For jackknife resampling, the pair count process is opti- ℓ(ℓ − 1)  ℓ − 3/2  mized to get the results for all samples in one run 3. For the large − √ √ 휙 휒 (30) 4 2ℓ − 3(2ℓ − 1) 2ℓ + 1 ℓ + 1/2 number of mocks, we count the pairs with Corrfunc (Sinha & (ℓ + 1)(ℓ + 2)  ℓ + 5/2  Garrison 2020). − √ √ 휙 휒 , 2ℓ + 1(2ℓ + 3) 2ℓ + 5 ℓ + 1/2 퐻 (푧) 5.2 CLPT-GS model where 휙(휒) = 푓푞 (푧) 푐 is the normalized quasar redshift distri- bution as a function of the comoving distance. Similarly, the bias The analytic model of the 2PCF we use is a combination (Wang 푞푞 et al. 2013) of the Convolution Lagrangian Perturbation Theory on 퐶ℓ can be described with two extra terms, (CLPT) (Carlson et al. 2012) and the Gaussian Streaming (GS) ∫ 푧2 푐푑푧 퐻(푧)  ℓ + 1/2  model (Reid & White 2011). In the GS model, the correlation 퐶푞푞,푟 = 퐾(ℓ, 휒) 휒−2 푓 (푧)푏(푧)푃 , 푧 ℓ ( ) 푚 푧1 퐻 푧 푐 휒 function in redshift space is given by (31) ∫ ( 2 ) 1 + 휉(푟) [푠휇 − 푦 − 휇푣12 (푟)] and 1 + 휉(푠, 휇) = 푑푦 exp − , √︃ 2 2 ( ) 2휋휎 (푟, 휇) 2휎12 푟, 휇 ∫ 푧2 푐푑푧 퐻(푧)  ℓ + 1/2  12 퐶푞푞,푟푟 = 퐾2 (ℓ, 휒) 휒−2 푓 2 (푧)푃 , 푧 . ℓ ( ) 푚 (36) 푧1 퐻 푧 푐 휒 (32) where the real space correlation function 휉(푟), pairwise velocity 푣 (푟) and velocity dispersion 휎2 (푟, 휇) are outputs from CLPT For the multipoles we are considering, 100 ≤ ℓ ≤ 1000, we find 12 12 휅푞,푟 / 휅푞 −4 푞푞,푟 / 푞푞 −4 푞푞,푟푟 / 푞푞 −8 퐶ℓ 퐶ℓ < 10 , 퐶ℓ 퐶ℓ < 10 and 퐶ℓ 퐶ℓ < 10 , 휅푞 푞푞 3 where 퐶ℓ and 퐶ℓ are the true power spectra in Eq. 17 and Eq. 19. https://gitlab.com/shadaba/CorrelationFunction So the bias due to RSD is completely negligible. This is expected 4 https://github.com/manodeep/Corrfunc

MNRAS 000,1–15 (2020) 8 Y. Zhang et al.

modified by the growth rate 푓 and the first and second-order La- NGC, jackknife NGC, mock +1.00 grangian bias, 퐹0 and 퐹00. On linear scales, the (Eulerian) bias and RSD parameter are given by 푏 = 1+퐹0 and 훽 = 푓 /푏 respectively. 퐹0 and 퐹00 can also be constrained with a single overdensity parameter +0.75 휈 through peak-background split (White 2014), +0.50 1  2푝  퐹0 = 푎휈2 − 1 + , 훿푐 1 + (푎휈2) 푝 (37) +0.25 1  2푝(2푎휈2 + 2푝 − 1)  퐹00 = 푎2휈4 − 3푎휈2 + , 2 2 푝 훿푐 1 + (푎휈 ) SGC, jackknife SGC, mock 0.00 where 푎 = 0.707 and 푝 = 0.3 with the Sheth-Tormen mass func- 0.25 tion (Sheth & Tormen 1999), and 훿푐 = 1.686 is the linear critical − overdensity of spherical collapse. To account for the finger-of-god 2 (FoG) effect and redshift smearing error, 휎 is modified by adding 0.50 12 − 2 = 2 + 2 a nuisance term 휎tot 휎FoG 휎푧 , where 휎FoG and 휎푧 are degener- 0.75 ate in this model. CLPT takes the matter power spectrum as a input, − which is calculated using CAMB 5 (Lewis et al. 2000) with our fidu-

cial cosmological parameters. This CLPT-GS model has been used 1.00 − in the RSD analysis of CMASS galaxies in 0.43 < 푧 < 0.7 (Alam 1.5 et al. 2015), BOSS DR12 galaxies in 0.2 < 푧 < 0.7 (Satpathy et al. 2017), and eBOSS DR14 quasars in 0.8 < 푧 < 2.2 (Zarrouk et al. mock 1.0 /σ jk

2018). σ NGC SGC 0.5

5.3 Parameters distribution sampling Figure 3. Correlation matrices (Eq. 13) of 휉0,2, a 1-D vector consists of Given the data and analytic model of the 2PCF, we construct a monopole and quadrupole, estimated with jackknife resampling (left) and multivariate Gaussian likelihood function, 1000 EZ mocks (right) for NGC (upper) and SGC (middle). 휉0 and 휉2   sample points with separation distances 30 ≤ 푠 ≤ 135 are included (Fig.6), 1 푇 − L(휃|휉ˆ) ∝ exp − 휉(휃) − 휉ˆ Cˆ 1 휉(휃) − 휉ˆ , (38) which gives 42 data points in total. The lower panel shows the ratio of the 2 1 휎 errors estimated with the two methods, i.e. 휎jackknife/휎mock.

where 휉ˆ is the data vector consists of 휉ˆ0 and 휉ˆ2, Cˆ is the estimated covariance matrix of 휉, and 휉(휃) is the output of the CLPT-GS model described above with the set of free parameters denoted as 6.1 Combination of NGC and SGC 휃. We estimate Cˆ with 1000 EZ mocks and do the correction as described in Section 3.3. As a comparison, Cˆ is also estimated As mentioned in Section 3.1, the quasar sample comprises two with jackknife resampling. We plot the correlation matrices of Cˆ catalogs, which correspond to two separate regions on the sky, with both methods and the ratio of the statistical errors in Fig.3. namely NGC and SGC. A proper combination of the two caps, We can see that compared with the mocks, jackknife resampling which we denote as NS, should give us better constrained estimates. tends to overestimate the statistical errors and the relative strength Throughout the data analysis pipeline in this work, this process can of the covariances (i.e. the off-diagonal terms). Considering that the be conducted at several stages. number of jackknives we are using is not very large, which hence First, at the raw data level, the simplest approach is to put may not be able to give us well-constrained estimates, here we use the two caps together before doing any estimation. For the quasar the simulated Cˆ in the likelihood function. The set of free parameters overdensity map, we may simply use all the quasars in the two 휃 includes the RSD parameter 훽, the overdensity parameter 휈 and a catalogs to make one map or merge the two overdensity maps into nuisance velocity dispersion term 휎tot. Flat priors are used for these one. For the estimation of the correlation function, we may combine parameters and the posterior distribution is sampled using Markov the pair counts in the Landy & Szalay estimator. However, we do Chain Monte Carlo (MCMC) with emcee 6 (Foreman-Mackey et al. not do the combination at this data level since NGC and SGC are 2013). observed with different photometric calibrations and have different number densities (Fig.2), which result in different shot noise and 푞푞 other possible systematics. For the estimation of 퐶ℓ , where the shot noise contributes much more than the signal at smaller scales, 6 RESULTS this simple combination of two maps with different shot noise is not In this section, we first discuss the methods of combining NGC and optimal. SGC. Then we present our estimates of the angular power spectra Instead of combining the data of the two caps directly, we mea- 휅푞 푞푞 sure 퐶ℓ ’s and 훽 separately for the two caps, which are then averaged 퐶ℓ and 퐶ℓ , and the RSD parameter 훽. These are then combined into 퐸퐺 (ℓ) at the 5 bandpowers, with which we find the best-fit to get the estimates for NS. This process is conducted for the full scale-independent 퐸¯퐺 estimate. data sample, simulations and jackknife samples. Assuming no cross correlation between the two caps, the average is weighted with in- verse variances, which are estimated with the 300 simulations. For 5 https://camb.info jackknife resampling, 91 = 56 + 35 jackknife estimates for NS are 6 https://github.com/dfm/emcee constructed by averaging each of the jackknife estimates from one

MNRAS 000,1–15 (2020) Testing GR with quasar and CMB lensing 9

1 cap with the full estimate from the other cap. It is worth mentioning χ (h− Mpc) that the jackknife estimates for NS are not constructed by simply 190 120 75⊥ 47 30 19 10 6 stacking NGC and SGC estimates together, since the 91 jackknives − should make up a complete sample from which one jackknife re- gion is left out each time. This also requires that we are using equal 7 weights when making jackknife regions for NGC and SGC sepa- 10− rately in order to make sure that they are statistically equivalent. κq `

Adhering to the advantage of being dependent only on the data, the C variances used in these averages are also estimates from jackknife 10 8 resampling instead of simulations or analytic uncertainties. These − variances for jackknives should also be rescaled with the ratio of Model NGC 푓sky’s between the leave-one-out jackknife mask and the full mask, NS SGC 9 while the difference is negligible. 10− At last, we may also do the average with estimates of 퐸 (ℓ) 100 158 251 398 630 1000 퐺 ` or 퐸¯퐺 for the two caps. As long as the error distributions of 퐶ℓ ’s, 훽 and 퐸퐺 are approximately Gaussian, this should be consistent with 1.4 NS, jackknife NS, simulation the method above. analytic 1.2

σ/σ 1.0 6.2 퐶휅푞 and 퐶푞푞 15 ℓ ℓ Individual Cumulative We consider the multipoles 100 ≤ ℓ ≤ 1000 for our analysis of the 10 angular power spectra and 퐸퐺 (ℓ). This corresponds to the linear SNR 5 −1 scales 19 < 휒⊥ < 190 ℎ Mpc with the radial comoving distance 0 휒(푧 = 1.5) = 3029 ℎ−1Mpc given our fiducial cosmology. We do not consider smaller scales since 퐸퐺 is well defined only on lin- ear scales. The largest scale that can be probed is limited by the Figure 4. CMB lensing convergence 휅 × quasar overdensity 푞 angular spatial size of the sample and the trade-off between the number of cross-power spectra. The crosses are estimates using Planck 2018 CMB jackknives. These multipoles are binned into 5 evenly spaced band- lensing map and eBOSS DR16 quasar clustering catalogs. We shifted the powers on the log scale, and thus ℓ indices for the estimates denote data points of NGC and SGC horizontally in the plot for reading conve- nience. For reference, we also plot the analytic model (Eq. 18) with a linear the bandpowers. We do not use narrower bins because the low SNR bias 푏 = 2.32 fitted from 퐶푞푞/퐶 휅푞. The statistical 1 휎 errors are esti- 퐶푞푞 ℓ ℓ of ℓ bandpowers at small scales could result in outliers in the mated with 300 simulations. The middle panel includes the comparison of 퐸퐺 (ℓ) estimates with the 300 simulations, whose error distribution error estimates for NS using simulations and jackknife resampling with the would no longer be appropriate for estimating the Gaussian covari- analytic uncertainties (Eq. 20). Individual and cumulative SNRs for NS over ance matrix. Also, 퐸퐺 as a ratio of noisy quantities can be biased, the bandpowers are shown in the lower panel, where the cumulative SNR hence using wider bins with smaller errors is preferable. starts from the smallest scale (i.e. highest ℓ) and the covariances between 휅푞 푞푞 scales are included (Eq. 39). The estimates of 퐶ℓ and 퐶ℓ are shown in the upper panels in Fig.4 and Fig.5 respectively, with the statistical 1 휎 errors given by simulations. For reference, we also plot the analytic models 푞푞/ 휅푞 above the summation runs over the 5 bins for the estimates. The in- discussed in Section 4.1, with a linear bias 푏 fitted from 퐶ℓ 퐶ℓ . 푞푞 dividual SNR for each bandpower and the cumulative SNRs starting It is worth noticing that for 퐶 estimates with quasar mocks, NGC 휅푞 ℓ from the highest-ℓ band are shown in the lower panels. For 퐶 , both and SGC are not very well consistent on the largest-scale bandpower. ℓ methods give similar errors and hence comparable SNRs. While for This might be caused by observational effects like completeness 푞푞 퐶 , jackknife resampling errors are higher than that from sim- levels or systematic weights that are different for the two caps. ℓ ulations. We are not doing any fittings with these angular power A better understanding requires more simulations with different spectra, so more comparisons between the two methods are dis- possible systematics applied, which we leave for future work. We cussed in Section 6.4, where the covariance matrices for 퐸 (ℓ) are do not have 퐶휅푞 signals with simulations since as mentioned, our 퐺 ℓ presented. With the simulated covariance matrices, we get overall simulated 휅 maps and quasar mocks are not correlated. Besides 휅푞 푞푞 SNR(퐶 ) = 12.5 and SNR(퐶 ) = 14.0 for NS. Although the simulations, the statistical errors are also estimated with jackknife ℓ ℓ SNR for each band depends on our binning scheme, the overall resampling. In the middle panels, we show the comparison of the value should remain roughly the same. error estimates from both methods with the analytic uncertainties in Eq. 20 and 21, where the quasar bias and shot noise are derived from data. As expected, the error estimates are mostly higher than 6.3 RSD parameter the analytic uncertainties where only the lensing reconstruction and shot noise are considered. Even though our simulated 휅 maps and We show the estimated monopole and quadrupole of the 2PCF quasar mocks are not correlated, the underestimation in 휎(퐶휅푞) is of the quasar catalogs in Fig.6, along with the best-fit CLPT-GS ℓ model and the 1000 EZ mocks. Data points with separation distances negligible due to the low cross correlation coefficient 푟ℓ < 0.2 (see 30 ≤ 푠 ≤ 135 ℎ−1Mpc are included in the RSD fitting. We do not Eq. 20). We measure the marginalized SNR over scales with the full −1 covariance matrix use smaller scales 푠 < 30 ℎ Mpc where the CLPT-GS model has not been validated. We apply larger scale cutoff to optimize 1/2 ∑︁ −1  ( ) 0 the model calculation (Alam et al. 2015) and remove any very large SNR 퐶ℓ = 0 퐶ℓ Cℓℓ0퐶ℓ (39) ℓ,ℓ scale systematic in the QSO sample (Castorina et al. 2019). Also, the to quantify the overall strength of the signal, where as mentioned contribution to our RSD fitting for the scale-independent parameters

MNRAS 000,1–15 (2020) 10 Y. Zhang et al.

1 χ (h− Mpc) 100 190 120 75⊥ 47 30 19 NGC 5 10−

2 50 Mpc)

6 1

10− − 0 h )( qq ` s ( C

ξ 50 2 − 7 s 10−

SGC - mock Model NGC 100 NGC - mock NS SGC − NS - mock mocks Monopole Quadrupole model 8 10− 100 158 251 398 630 1000 100 ` SGC

NS, jackknife NS, simulation 1.5 2 50 analytic Mpc)

1.0 1 σ/σ

− 0 h 15 )(

Individual Cumulative s (

10 ξ 50 2 − s

SNR 5

0 100 − 25 50 75 100 125 150 175 1 Figure 5. Quasar overdensity angular auto-power spectra, with similar in- s (h− Mpc) formation as Fig.4. The shaded area denotes the average and 1 휎 error bar of estimates from 300 EZ mocks. The analytic model in Eq. 19 is plotted for reference, with the same bias used in Fig.4. The analytic uncertainty is Figure 6. The monopole and quadrupole of the 2PCF with the best-fit computed with Eq. 21. CLPT-GS model. The crosses are estimates using eBOSS DR16 quasar NGC (upper) and SGC (lower) clustering catalogs. The grey shaded area denotes the mean and 1 휎 error of the 1000 EZ mocks used to estimate the covariance matrix in Eq. 38. Notice that the overall sky mask with Table 2. The maximum likelihood estimates of the RSD parameters with CMB lensing has been applied on both the data and EZ mock catalogs. The flat priors, where the 68.3% confidence intervals are quoted with L( 휃 ) = − two vertical dashed lines enclose the data points used in RSD fitting, with L( 휃 ). + separation distances 30 ≤ 푠 ≤ 135 ℎ−1Mpc.

Parameter 훽 휈 휎tot Prior [0, 1][1, 3][0, 16] distributions are quoted for reference, these are not propagated to +0.091 +0.056 +2.1 the error estimation of 퐸퐺 (ℓ). As mentioned in Section 3.3, to esti- NGC 0.449−0.063 2.047−0.072 7.2−4.3 mate the full covariance matrix for 퐸퐺 (ℓ), we also need to run all the +0.093 +0.081 +3.89 SGC 0.474−0.082 1.991−0.089 0.63−0.58 simulations through the data analysis pipeline, including the RSD fitting process. For the 300 EZ mocks, the average along with the standard deviation of the best-fit estimates are 푓 휎8 = 0.380 ± 0.055 from these larger scales should be negligible due to the large errors. for NGC and 푓 휎8 = 0.366 ± 0.067 for SGC, which are consistent 2 The goodness of fitting is given as 휒 /dof = 35/40 (45/40) for with the fiducial value 푓 휎8 = 0.381 given the cosmological pa- NGC (SGC). rameters used in the EZ mock simulation. The analysis of MCMC The posterior distributions with flat priors (i.e. likelihood func- chains including the plots and statistics is conducted with the usage tions) of the RSD parameter 훽, the overdensity parameter 휈 and the of ChainConsumer 7 (Hinton 2016). nuisance velocity dispersion parameter 휎tot are shown in Fig.7. For For our consistency test of ΛCDM-GR on the data, we are 훽 and 휈, while slight skewness is observed, the distributions are allowed to fix the fiducial cosmological parameters in this RSD fit- approximately Gaussian around the maximum likelihood estimates. ting process since the Planck 2018 results are measured to very high This skewness might be caused by the strong cross correlation with accuracy, and a flat prior based on this will not really change the 휎tot at large values, as we can tell from the banana-shaped con- marginalized distribution of the RSD parameters given the statisti- tours. These covariances with velocity dispersion might be better cal accuracy. If the true parameters are statistically different from constrained with an optimized modelling that breaks the degener- Planck results or ΛCDM-GR is not a proper model, we should be acy between the FoG effect and the redshift smearing error. For able to see the deviation of 퐸퐺 (ℓ) estimates from the ΛCDM-GR 휎FoG, a scale-dependent analytic model would be more accurate. prediction with Planck parameters. From 훽 and 휈, we can also in- The constraint on 휎푧 could also be improved by constructing an fer the posterior distribution of the linear growth rate, which gives = +0.064 = +0.058 informative prior based on redshifts measured with different meth- 푓 휎8 0.424−0.047 for NGC and 푓 휎8 0.430−0.057 for SGC. Our ods. The best-fit estimates for the marginalized distribution of each parameter along with the confidence intervals are summarized in Table2. Though the confidence intervals inferred from posterior 7 https://github.com/samreay/ChainConsumer

MNRAS 000,1–15 (2020) Testing GR with quasar and CMB lensing 11

1 SGC χ (h− Mpc) 190 120 75⊥ 47 30 19 NGC 1.0 ΛCDM-GR NGC 0.8 NS SGC

0.6

G 0.4 E 10 2. 0.2

ν 95 1.

80 0.0 1.

65 1. 0.2 − 12 100 158 251 398 630 1000 ` 9 3

tot NS NGC SGC σ 6

sim 2

3 /σ

jk 1 σ 0 30 45 60 75 90 65 80 95 10 3 6 9 12 0. 0. 0. 0. 0. 1. 1. 1. 2. 15 σtot β ν jk - indv. jk - cum. 10 sim - indv. sim - cum.

SNR 5 Figure 7. Posterior distributions of the parameters in RSD fitting, sampled 0 with MCMC. The properties of the marginalized distributions of individual parameters are summarized in Table2.

Figure 8. 퐸퐺 estimates using Planck 2018 CMB lensing map and eBOSS DR16 quasar clustering catalogs. The data points for NGC and estimates are consistent with the eBOSS DR16 consensus result SGC are shifted horizontally in the plot for reading convenience. The c ( = ) = ± of the quasar sample, 푓 휎8 푧eff 1.48 0.462 0.045, which green solid line and shaded area is the ΛCDM-GR prediction using the is a combination of the configuration space (Hou et al. 2020) and Planck 2018 CMB+BAO matter density parameter and 1 휎 uncertainty, Fourier space (Neveux et al. 2020) analysis. The possible sources Ωm = 0.3111 ± 0.0056. The 1 휎 error bars are estimated using simulations, of difference include the overlapped mask with CMB lensing, fixed with the comparison to errors given by jackknife resampling shown in the Alcock-Paczynski (AP; Alcock & Paczynski 1979) parameters and middle panel as 휎jackknife/휎simulation. As in Fig.4 and5, the individual and a different analytic model used in this work. The combination of cumulative SNRs of the bandpowers are shown in the lower panel. 휉(푠) and 푃(푘) analysis could also help reduce the systematics in the consensus result (Smith et al. 2020). A more detailed discussion of models and systematics in RSD fitting is out of the scope of this work, thus we refer our readers to the series of papers presenting comparison between theory and observations more straightforward. the eBOSS final data release (eBOSS Collaboration et al. 2020). In Since the RSD parameter 훽 is assumed to be scale-independent, the RSD analysis of eBOSS DR14 quasar catalog using the same the fluctuations of 퐸퐺 (ℓ) estimates are mainly determined by the CLPT-GS model (Zarrouk et al. 2018), a shift on the linear bias ratio 퐶휅푞/퐶푞푞 (see Fig. B1). As discussed in Section 6.1, to get Δ푏휎 = 0.037 was observed when 퐹00 was set free instead of fixed. ℓ ℓ 8 퐸퐺 (ℓ) estimates for NS, we can combine NGC and SGC at either So besides the main analysis using 휈 and peak-background split, the {퐶 0s, 훽} level or 퐸 (ℓ) level. The NS signals shown in Fig.8 0 00 ℓ 퐺 we also do a test by running the RSD fitting with free 퐹 and 퐹 are derived using the first method, which are consistent with that parameters on the data sample. For the RSD parameter we are in- using the second method. For the scale-averaged 퐸¯퐺 discussed be- terested in, we get 훽 = 0.445+0.090 for NGC and 훽 = 0.458+0.101 −0.059 −0.069 low, besides fitting 퐸퐺 (ℓ) of NS, we can also do the fitting for NGC for SGC, which are consistent with the values in Table2. and SGC separately and then combine the results to get 퐸¯퐺 for NS. We have tried all these methods, and the results are consistent within 3%, which is expected as for all the statistical quantities, 6.4 퐸 estimates 퐺 the error distributions are approximately Gaussian and the two spa- 휅푞 푞푞 ( ) tially separated caps should not be correlated for the scales we are We combine our estimates of 퐶ℓ , 퐶ℓ and 훽 into 퐸퐺 ℓ following Eq.5, with the calibration in Eq.7 applied, which shifts the 퐸퐺 (ℓ) considering. signals lower for about 5 %. We find the factor Γ(푧¯ = 1.5)' 0.74 Given the consistency between the 퐸퐺 estimates at all the for both caps and NS. The 퐸퐺 (ℓ) estimates for the bandpowers are 5 scale bins and the scale-independent ΛCDM-GR prediction, we shown in Fig.8, where the 1 휎 statistical errors for individual bins further improve the constraint on 퐸퐺 by fitting a constant 퐸¯퐺 over are determined using simulations. These errors are also estimated the 5 bins. We infer the best-fit value of 퐸¯퐺 by maximizing the using jackknife resampling, with the comparison shown in the mid- multivariate Gaussian likelihood function, dle panel. We see that 퐸 (ℓ) estimates at all the 5 bandpowers 퐺   agree with the GR prediction at 1 휎 level, and we could not see an 1  푇 −1   L(퐸¯퐺) ∝ exp − 퐸ˆ퐺 (ℓ) − 퐸¯퐺 Cˆ 퐸ˆ퐺 (ℓ) − 퐸¯퐺 , (40) obvious scale-dependence pattern. Unlike the power spectra (Fig.4 2 and5), the theoretical 퐸퐺 model does not depend on the cluster- ing bias and is thus independent of the estimates, which makes the where Cˆ is the estimated covariance matrix of 퐸퐺 (ℓ). For this linear

MNRAS 000,1–15 (2020) 12 Y. Zhang et al.

NGC, jackknife SGC, jackknife NS, jackknife Table 3. 퐸퐺 estimates at the effective redshift 푧¯ = 1.5 averaged over 1 0.24 0.21 0.29 0.07 1 0.23 0.31 0.25 0.18 1 0.21 0.26 0.38 0.14 −1 scales 19 ≤ 휒⊥ ≤ 190 ℎ Mpc with Planck 2018 CMB lensing map and

0.24 1 -0.01 -0.06 0.23 0.23 1 0.34 0.18 0.09 0.21 1 0.12 0.09 0.15 eBOSS DR16 quasar clustering catalogs. Best-fit results for NGC, SGC and the combination NS with simulated Cˆ are quoted with 1 휎 statistical 0.21 -0.01 1 0.08 0.06 0.31 0.34 1 0.04 -0.00 0.26 0.12 1 0.08 0.08 errors. The deviations from ΛCDM-GR prediction 퐸퐺 (푧 = 1.5) = 0.3346

0.29 -0.06 0.08 1 -0.05 0.25 0.18 0.04 1 0.21 0.38 0.09 0.08 1 0.13 with Ω푚,0 = 0.3111 are also presented. The last row includes the best-fit estimates using Cˆ from jackknife resampling, which are not reported as our 0.07 0.23 0.06 -0.05 1 0.18 0.09 -0.00 0.21 1 0.14 0.15 0.08 0.13 1 final results due to the possible poor constraints on the covariance matrices NGC, simulation SGC, simulation NS, simulation (see text).

1 0.15 0.20 0.21 0.01 1 0.10 0.20 0.09 0.21 1 0.12 0.26 0.13 0.18 Cap NS NGC SGC 0.15 1 0.08 0.23 0.12 0.10 1 0.12 0.20 0.05 0.12 1 0.07 0.15 0.14 퐸퐺 0.295 ± 0.054 0.309 ± 0.068 0.272 ± 0.087 0.20 0.08 1 0.11 0.14 0.20 0.12 1 0.16 0.13 0.26 0.07 1 0.16 0.20 Deviation 0.74 휎 0.38 휎 0.72 휎

0.21 0.23 0.11 1 -0.04 0.09 0.20 0.16 1 0.06 0.13 0.15 0.16 1 0.02 퐸퐺 with Cˆ jk 0.253 ± 0.050 0.283 ± 0.066 0.214 ± 0.076

0.01 0.12 0.14 -0.04 1 0.21 0.05 0.13 0.06 1 0.18 0.14 0.20 0.02 1

Figure 9. Estimated correlation matrices (Eq. 13) of 퐸퐺 (ℓ) with jackknife Ωm,0 resampling (upper) and 300 simulations (lower) for NGC, SGC and the 0.1 0.2 0.3 0.4 0.5 combination, NS. The number of jackknife samples is 56 for NGC, 35 for 1.0 SGC and 91 for NS. ΛCDM-GR NGC 0.8 SGC fitting model, the max-L point can be analytically written as NS Í ˆ −1 ˆ 0 0.6 ℓ,ℓ0 C 0 퐸퐺 (ℓ ) 퐸¯ = ℓℓ , (41) 퐺 Í ˆ −1 L ℓ,ℓ0 C 0 ℓℓ 0.4 with the statistical error −1/2  ∑︁ −1  0.2 휎 퐸¯ = 푀 × Cˆ 0 , (42) 퐺 ℓ,ℓ0 ℓℓ where the summation runs over the 5 bandpowers for the estimates, 0.0 Cˆ −1 is the ℓ, ℓ0 element of Cˆ inverse with the correction in Eq. 14 0.0 0.1 0.2 0.3 0.4 0.5 0.6 ℓℓ0 E applied, and 푀 is the calibration factor in Eq. 15. As discussed G in Section 3.3, we estimate Cˆ with both jackknife resampling and simulations. One defect of the simulations is that the 휅 maps and quasar mocks are not correlated. While the impact on 휎(퐶휅푞) is Figure 10. Likelihood functions of scale-averaged 퐸퐺, with covariance ℓ matrices estimated using simulations. The green line with shaded area cor- negligible compared with the current lensing reconstruction and responds to the ΛCDM-GR prediction with the Planck 2018 CMB+BAO shot noise (see Section 6.2), the 퐶휅푞 signal does matter in the co- ℓ matter density and 1 휎 uncertainty, Ω푚,0 = 0.3111 ± 0.0056. variance matrix of 퐸퐺 (ℓ). To fix this issue, we shift the center of 휅푞 the error distribution of the 300 simulated 퐶ℓ ’s from zero to the expected signal with a fiducial quasar bias measured from the data. 휅푞 By doing this, the distribution of the simulated 퐶ℓ ’s should be roughly equivalent to what we would get if the simulations were correlated. The correlation matrices (Eq. 13) of Cˆ from both meth- different, especially for SGC. One reason might be that the numbers ods are shown in Fig.9, and the square root ratios of the diagonal of jackknives, with only 35 samples for SGC, are not enough to get terms are shown in the middle panel in Fig.8. We see that Cˆ ’s converged estimates. Also, for the two caps, the observational sys- given by both methods include non-negligible cross correlations tematics in the imaging used to target quasars are different, which between scales. This can be caused by the fact that we are using may result in different unknown bias. The poor constraint on Cˆ for one scale-independent 훽 estimate for all bandpowers, which intro- either or both of 퐸퐺 (ℓ) and 푅ℓ can then bias our fitting for 퐸¯퐺. duces the same variation for all of them and hence contributes to We summarize our best-fit estimates of the scale-averaged 퐸¯퐺 the covariances. To test if Cˆ ’s given by jackknife or simulations are in Table3. Considering the result of the test above and the fact that well constrained, we take another approach of estimating 퐸¯퐺 by the simulations we are using are designed to be as realistic as possi- ≡ 휅푞/ 푞푞 fitting the ratio 푅ℓ 퐶ℓ 퐶ℓ over the 5 scale bins first, which ble, i.e. including all the known systematics, we take the estimates is then divided by the scale-independent 훽. In that case, CM for with simulated Cˆ as our primary results. Although the signals are 푅ℓ instead of 퐸퐺 (ℓ) is estimated but final 퐸¯퐺 estimates should be different, the statistical errors given by the two methods are almost consistent if CMs in both approaches are well constrained. More the same. We report a best-fit 퐸¯퐺 (푧 ' 1.5) = 0.295±0.054 estimate details are included in AppendixB. It is shown that the covariances for NS, which is about 0.74 휎 lower than the ΛCDM-GR predic- 0 of 푅ℓ are much weaker (Fig. B2) than that of 퐸퐺 (ℓ) (Fig.9), which tion with Planck 2018 CMB+BAO Ω푚. For the two separate caps, is expected without the same 훽 variation for all bins. The two ap- they agree with each other and NGC is more consistent with the proaches give consistent final results with Cˆ ’s given by simulations. GR prediction with a 0.38 휎 deviation. For reference, the likelihood While with jackknife resampling, the final 퐸¯퐺 estimates are more functions of 퐸¯퐺 are shown in Fig. 10.

MNRAS 000,1–15 (2020) Testing GR with quasar and CMB lensing 13

0.8 lensing maps and galaxy/quasar mocks that are truly correlated for Reyes et al. 2010 Alam et al. 2017 NGC future surveys where lensing reconstruction and shot noise will 0.7 Blake et al. 2015 Amon et al. 2018 SGC be lower and the contribution to the covariance matrix from cross Pullen et al. 2016 Singh et al. 2018 NS 0.6 de la Torre et al. 2017 Blake et al. 2020 correlation will no longer be negligible. Planck has been a very successful CMB survey which gives ΛCDM-GR, Ωm,0 = 0.3111 0.0056 0.5 ± the best constraints on the cosmological parameters so far. The next

G 0.4 stage CMB surveys, e.g. CMB-S4 (CMB-S4 Collaboration 2016) E and Simons Observatory (SO; SO Collaboration 2019), will produce 0.3 even more accurate maps with higher resolution and lower noise. 0.2 BOSS and eBOSS in SDSS has made the largest catalogs of LSS tracers in the Universe. While DR16 is the last data release of the 0.1 series, more and larger LSS surveys are in progress. In the coming 0.0 few years, the Dark Energy Spectroscopic Instrument (DESI; DESI 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 Collaboration 2016) survey will target about 17 million ELGs in z the redshift range 0.6 ≤ 푧 ≤ 1.6, which will be able to fill the gap in Fig. 11. Redshifts of 1.7 million quasars with 푧 < 2.1 as LSS tracers will also be measured over a sky area of 14 000 deg2, which Figure 11. Some previous 퐸퐺 estimates and the results of this work. For reading convenience, some results are slightly shifted horizontally. For the corresponds to 푓sky ' 34 %. Compared with the eBOSS sample results in this work, the NS is plotted at the effective redshift 푧 = 1.5. The used in this work, the sky coverage and angular number density data points with white marker face color are estimated using CMB lensing are increased by a factor of 3 and 1.7 respectively. Some analytic while others are estimated with galaxy-galaxy lensing. The solid line is the forecasts of constraining 퐸퐺 with future CMB and LSS surveys are ΛCDM-GR prediction (Eq.3) with Ωm,0 from Planck 2018 CMB+BAO discussed in Pullen et al.(2015), where we can see that the SNR in cosmological parameters. this work can be improved by an order of magnitude with the DESI quasar sample. With all these promising future surveys, modern cosmology will be able to explore the origin and evolution of the 7 CONCLUSIONS Universe with higher and higher precision. 퐸퐺 is a promising probe of gravity on cosmological scales by combining gravitational lensing and LSS, with the advantage of being independent of the tracer bias and 휎8. In this work, we estimate −1 ACKNOWLEDGEMENTS 퐸퐺 at the effective redshift 푧 ∼ 1.5 over scales 19 − 190 ℎ Mpc with the Planck 2018 CMB lensing convergence map and SDSS We thank Jeremy Tinker and Michael Blanton for helpful discus- eBOSS DR16 quasar clustering catalogs. This is the highest redshift sions. We also thank eBOSS QGC working group for comments and largest scale where 퐸퐺 has been estimated so far. We show that on the RSD fitting analysis. ARP was supported by NASA un- quasars are promising DM LSS tracers for both auto correlation der award numbers 80NSSC18K1014 and NNH17ZDA001N. ARP clustering analysis and cross correlation with the weak gravitational was also supported by the Simons Foundation. SA is supported by lensing signal reconstructed from CMB. Our results are in line with the European Research Council through the COSFORM Research the ΛCDM-GR prediction within 1 휎 confidence interval. Some Grant (#670193). ADM was supported by the U.S. Department of previous estimates of 퐸퐺 at lower redshifts and results in this work Energy, Office of Science, Office of High Energy Physics, under are summarized in Fig. 11. The statistical errors are still too large to Award Number DE-SC0019022. GR acknowledges support from discriminate between different gravity models. This work extends the National Research Foundation of Korea (NRF) through Grants the redshift baseline of testing GR with 퐸퐺, while there is still a No. 2017R1E1A1A01077508 and No. 2020R1A2C1005655 funded gap between 푧 ∼ 0.6 and 푧 ∼ 1.5, where 퐸퐺 has not been explored by the Korean Ministry of Education, Science and Technology mainly due to the lack of promising LSS tracers considering the (MoEST), and from the faculty research fund of Sejong Univer- drop in the CMB lensing kernel. sity. There are still a few concerns which can be improved in the This work is based on observations obtained with Planck future with larger data samples. First, the redshift range 0.8 < 푧 < (http://www.esa.int/Planck), an ESA science mission with in- 2.2 of the quasar sample in this work is wide, and the effective struments and contributions directly funded by ESA Member States, redshift description may not be perfect. We tried to split the sample NASA, and Canada. into smaller redshift bins, and study the redshift evolution of all the Funding for the Sloan Digital Sky Survey IV has been pro- quantities. However, limited by the size of the sample, the SNRs are vided by the Alfred P. Sloan Foundation, the U.S. Department of too low to give us reliable estimates. Second, we used both jackknife Energy Office of Science, and the Participating Institutions. SDSS- resampling and simulations to estimate the covariance matrix for IV acknowledges support and resources from the Center for High- 퐸퐺 (ℓ), with the latter taken for the final result reported. However, Performance Computing at the University of Utah. The SDSS web we know that both these two methods have limitations. Although the site is www.sdss.org. simulations are designed to be realistic, it is still possible that there SDSS-IV is managed by the Astrophysical Research Consor- are unknown systematics that contribute to the covariances. For tium for the Participating Institutions of the SDSS Collaboration future surveys with a larger sky area, a larger number of jackknives including the Brazilian Participation Group, the Carnegie Institu- would serve as a reliable comparison. At last, so far the statistical tion for Science, Carnegie Mellon University, the Chilean Participa- error bars are still very large, which make it difficult to do a selection tion Group, the French Participation Group, Harvard-Smithsonian of different gravity models. Also, a rigorous self-consistency test of Center for Astrophysics, Instituto de Astrofísica de Canarias, The any gravity model requires the corresponding fiducial cosmology Johns Hopkins University, Kavli Institute for the Physics and Math- and simulations. Besides, it is necessary to have simulated CMB ematics of the Universe (IPMU) / University of Tokyo, the Ko-

MNRAS 000,1–15 (2020) 14 Y. Zhang et al. rean Participation Group, Lawrence Berkeley National Laboratory, Gunn J. E., et al., 2006,AJ, 131, 2332 Leibniz Institut für Astrophysik Potsdam (AIP), Max-Planck-Institut Han J., Ferraro S., Giusarma E., Ho S., 2019, MNRAS, 485, 1720 für Astronomie (MPIA Heidelberg), Max-Planck-Institut für Astro- Hartlap J., Simon P., Schneider P., 2007, A&A, 464, 399 physik (MPA Garching), Max-Planck-Institut für Extraterrestrische Hinton S. R., 2016, The Journal of Open Source Software, 1, 00045 Physik (MPE), National Astronomical Observatories of China, New Hirata C. M., Padmanabhan N., Seljak U. c. v., Schlegel D., Brinkmann J., 2004, Phys. Rev. D, 70, 103501 Mexico State University, New York University, University of Notre Hirata C. M., Ho S., Padmanabhan N., Seljak U. c. v., Bahcall N. A., 2008, Dame, Observatário Nacional / MCTI, The Ohio State University, Phys. Rev. D, 78, 043520 Pennsylvania State University, Shanghai Astronomical Observa- Ho S., Hirata C., Padmanabhan N., Seljak U., Bahcall N., 2008, Phys. Rev. D, tory, United Kingdom Participation Group, Universidad Nacional 78, 043519 Autónoma de México, University of Arizona, University of Col- Hojjati A., Pogosian L., Zhao G.-B., 2011, J. Cosmology Astropart. Phys., orado Boulder, University of Oxford, University of Portsmouth, 2011, 005 University of Utah, University of Virginia, University of Wash- Hou J., et al., 2020, arXiv e-prints, p. arXiv:2007.08998 ington, University of Wisconsin, Vanderbilt University, and Yale Hubble E., 1929, PNAS, 15, 168 University. Hui L., Gaztañaga E., Loverde M., 2007, Phys. Rev. D, 76, 103502 Ishak M., 2019, Living Reviews in Relativity, 22, 1 Joyce A., Lombriser L., Schmidt F., 2016, ANNU REV NUCL PART S, 66, 95 DATA AVAILABILITY Landy S. D., Szalay A. S., 1993, ApJ, 412, 64 Laurent P., et al., 2017, J. Cosmology Astropart. Phys., 2017, 017 The eBOSS DR16 quasar clustering catalogs are available in Lewis A., Challinor A., Lasenby A., 2000, ApJ, 538, 473 the SDSS-IV Science Archive Server at https://data.sdss. Liu J., Haiman Z., Hui L., Kratochvil J. M., May M., 2014, Phys. Rev. D, org/sas/dr16/eboss/lss/catalogs/DR16/. The Planck 2018 89, 023515 8 CMB lensing data and simulations were accessed from Planck Lyke B. W., et al., 2020, arXiv e-prints, p. arXiv:2007.09001 Legacy Archive at https://pla.esac.esa.int/. The derived Myers A. D., et al., 2015, ApJS, 221, 27 data generated in this research will be shared on reasonable request Neveux R., et al., 2020, arXiv e-prints, p. arXiv:2007.08999 to the corresponding author. Padmanabhan N., Tegmark M., Hamilton A. J. S., 2001, ApJ, 550, 52 Padmanabhan N., Seljak U., Pen U., 2003, New Astron., 8, 581 Padmanabhan N., et al., 2007, MNRAS, 378, 852 Peebles P. J. E., Ratra B., 2003, Rev. Mod. Phys., 75, 559 REFERENCES Percival W. J., et al., 2014, MNRAS, 439, 2531 Ahumada R., et al., 2019, arXiv e-prints, p. arXiv:1912.02905 Perlmutter S., et al., 1999, ApJ, 517, 565 Alam S., Ho S., Vargas-Magaña M., Schneider D. P., 2015, MNRAS, 453, Planck Collaboration I 2018, arXiv e-prints, p. arXiv:1807.06205 1754 Planck Collaboration IX 2020, A&A, 641, A9 Alam S., Miyatake H., More S., Ho S., Mandelbaum R., 2017a, MNRAS, Planck Collaboration VI 2018, arXiv e-prints, p. arXiv:1807.06209 465, 4853 Planck Collaboration VIII 2018, arXiv e-prints, p. arXiv:1807.06210 Alam S., et al., 2017b, MNRAS, 470, 2617 Planck Collaboration XXVI 2016, A&A, 594, A26 Alcock C., Paczynski B., 1979, Nature, 281, 358 Planck Collaboration XXVII 2016, A&A, 594, A27 Alonso D., Sanchez J., Slosar A., Collaboration L. D. E. S., 2019, MNRAS, Planck Collaboration XXVIII 2016, A&A, 594, A28 484, 4127 Pullen A. R., Alam S., Ho S., 2015, MNRAS, 449, 4326 Amon A., et al., 2018, MNRAS, 479, 3422 Pullen A. R., Alam S., He S., Ho S., 2016, MNRAS, 460, 4098 Blake C., et al., 2015, MNRAS, 456, 2806 Reid B. A., White M., 2011, MNRAS, 417, 1913 Blake C., et al., 2020, A&A, 642, A158 Reyes R., Mandelbaum R., Seljak U., Baldauf T., Gunn J. E., Lombriser L., Blanton M. R., et al., 2017,AJ, 154, 28 Smith R. E., 2010, Nature, 464, 256 CMB-S4 Collaboration 2016, arXiv e-prints, p. arXiv:1610.02743 Riess A. G., et al., 1998,AJ, 116, 1009 Carlson J., Reid B., White M., 2012, MNRAS, 429, 1674 Ross A. J., et al., 2020, arXiv e-prints, p. arXiv:2007.09000 Carroll S. M., de Felice A., Duvvuri V., Easson D. A., Trodden M., Turner SO Collaboration 2019, J. Cosmology Astropart. Phys., 2019, 056 M. S., 2005, Phys. Rev. D, 71, 063513 Satpathy S., et al., 2017, MNRAS, 469, 1369 Castorina E., et al., 2019, J. Cosmology Astropart. Phys., 2019, 010 Schlegel D. J., Finkbeiner D. P., Davis M., 1998, ApJ, 500, 525 Chuang C.-H., Kitaura F.-S., Prada F., Zhao C., Yepes G., 2015, MNRAS, Sheth R. K., Tormen G., 1999, MNRAS, 308, 119 446, 2621 Silvestri A., Trodden M., 2009, Rep. Prog. Phys., 72, 096901 Clifton T., Ferreira P. G., Padilla A., Skordis C., 2012, Phys. Rep., 513, 1 Singh S., Alam S., Mandelbaum R., Seljak U., Rodriguez-Torres S., Ho S., DESI Collaboration 2016, arXiv e-prints, p. arXiv:1611.00036 2018, MNRAS, 482, 785 Dalal N., Doré O., Huterer D., Shirokov A., 2008, Phys. Rev. D, 77, 123514 Sinha M., Garrison L. H., 2020, MNRAS, 491, 3022 Dawson K. S., et al., 2016,AJ, 151, 44 Smee S. A., et al., 2013,AJ, 146, 32 Dizgah A. M., Durrer R., 2016, J. Cosmology Astropart. Phys., 2016, 035 Smith A., et al., 2020, arXiv e-prints, p. arXiv:2007.09003 Dodelson S., Schneider M. D., 2013, Phys. Rev. D, 88, 063537 Sotiriou T. P., Faraoni V., 2010, Rev. Mod. Phy., 82, 451 Dvali G., Gabadadze G., Porrati M., 2000, Phys. Rev. B, 485, 208 Sunyaev R. A., Zeldovich I. B., 1980, ARA&A, 18, 537 Feldman H. A., Kaiser N., Peacock J. A., 1994, ApJ, 426, 23 Taylor A., Joachimi B., Kitching T., 2013, MNRAS, 432, 1928 Foreman-Mackey D., Hogg D. W., Lang D., Goodman J., 2013, PASP, 125, Tegmark M., 1997, Phys. Rev. D, 55, 5895 306 Wang L., Reid B., White M., 2013, MNRAS, 437, 588 Giannantonio T., et al., 2016, MNRAS, 456, 3213 White M., 2014, MNRAS, 439, 3630 Górski K. M., Hivon E., Banday A. J., Wand elt B. D., Hansen F. K., Yang S., Pullen A. R., 2018, MNRAS, 481, 1441 Reinecke M., Bartelmann M., 2005, ApJ, 622, 759 Zarrouk P., et al., 2018, MNRAS, 477, 1639 Zhang P., Liguori M., Bean R., Dodelson S., 2007, Phys. Rev. Lett., 99, 141302 8 https://wiki.cosmos.esa.int/planck-legacy-archive/ Zhao C., et al., 2020, arXiv e-prints, p. arXiv:2007.08997 index.php/Lensing de la Torre S., et al., 2017, A&A, 608, A44

MNRAS 000,1–15 (2020) Testing GR with quasar and CMB lensing 15 eBOSS Collaboration et al., 2020, arXiv e-prints, p. arXiv:2007.08991 Table B1. Scale-averaged 퐸퐺 estimates, similar as Table3 but with the approach discussed in AppendixB.

APPENDIX A: CMB CONTAMINATION SYSTEMATIC Cap NS NGC SGC ˆ BIAS 퐸퐺 with Csim 0.294 ± 0.057 0.308 ± 0.073 0.272 ± 0.092 퐸퐺 with Cˆ jk 0.267 ± 0.045 0.291 ± 0.062 0.240 ± 0.066 In this section, we consider the foregrounds, including dust and point sources, that might contaminate the CMB temperature maps. These sources can leave signatures in the CMB lensing map and second approach in Table B1. Compared with the results in Table3, hence bias our cross-correlation signal with quasars. the estimates with simulated CMs are well consistent while those We use the galactic dust emission map constructed in Schlegel with jackknife resampling CMs differ by 3% for NGC and 12% for et al.(1998). For point sources, we make angular maps for several SGC. This disagreement in jackknife resampling can be caused by Planck point source catalogs, including galactic cold clumps (GCC; the small number of samples, which may not be enough to give us Planck Collaboration XXVIII 2016), Sunyaev-Zel’dovich (SZ; Sun- accurate CMs for either or both of 퐸퐺 (ℓ) and 푅ℓ . On the other yaev & Zeldovich 1980) sources (Planck Collaboration XXVII hand, for simulations, the consistency between the two approaches 2016) and compact sources (CS; Planck Collaboration XXVI 2016) indicates that the CMs should be well constrained. at 100, 143 and 217 GHz. Following Pullen et al.(2016), we estimate 휅푞 ˆ This paper has been typeset from a T X/LAT X file prepared by the author. the biases to 퐶ℓ by these possible sources with E E 퐶ˆ휅푐퐶ˆ푞푐 Δ퐶ˆ휅푞 = ℓ ℓ , (A1) ℓ ˆ푐푐 퐶ℓ where 푐 is any of the contamination maps, and the errors are given by " 푞푐 #    2 휎2 (퐶ˆ휅푐) 휎2 (퐶ˆ ) 휎2 Δ퐶ˆ휅푞 = Δ퐶ˆ휅푞 ℓ + ℓ . (A2) ℓ ℓ ( ˆ휅푐)2 ( ˆ푞푐)2 퐶ℓ 퐶ℓ The estimates are shown in Fig. A1. We find that the biases are consistent with zero, with statistical errors that are much lower than ˆ휅푞 our 퐶ℓ signal (Fig.4). This is expected since the most foreground- contaminated area of the dust emission map, i.e. the Galactic plane, and the sky regions of many point sources have already been masked out in the Planck maps. Compared with the previous analysis for cross-correlating CMASS galaxies (Pullen et al. 2016) with Planck 2015 CMB lensing map, the removal of the contamination has been improved for Planck 2018 data release. A similar analysis has also been conducted for eBOSS DR14 quasars (Han et al. 2019).

APPENDIX B: TEST ON FITTING 퐸퐺 (ℓ) OVER SCALES

Here we take a slightly different approach on fitting 퐸퐺 (ℓ) over scales (i.e. the 5 bandpowers) for the scale-averaged 퐸¯퐺, which also serves as a test on the reliability of estimating the covariance matrices (CMs) with simulations and jackknife resampling. With our 퐸퐺 (ℓ) estimator given by Eq.5 and assuming a scale- independent RSD parameter 훽, 퐸퐺 (ℓ) could be scale-dependent only through the ratio of the angular power spectra, ≡ 휅푞/ 푞푞 푅ℓ 퐶ℓ 퐶ℓ . (B1)

Thus fitting 퐸퐺 (ℓ) as discussed in Section 6.4 should be equivalent to fitting 푅ℓ over scales first, whose best-fit estimate is then com- bined with 훽 into 퐸¯퐺. The Gaussian likelihood function and best-fit value are in the same form as that for 퐸퐺 (ℓ) (Eq. 40 and 41), with 퐸퐺 (ℓ) replaced by 푅ℓ . The key point is that the corresponding Cˆ is now the CM for 푅ℓ . We present the estimates and the correlation matrices of 푅ℓ in Fig. B1 and B2. Compared with that for 퐸퐺 (ℓ) (Fig.9), the cross correlations between scales are weaker, which is expected since using the same scale-independent 훽 value for all bins of 퐸퐺 (ℓ) introduces covariances. With the 퐶Γ (Eq.7, which are almost the same value for the 5 bins) calibration factor applied, we summarize the final scale-independent 퐸¯퐺 estimates with this

MNRAS 000,1–15 (2020) 16 Y. Zhang et al.

9 8 8 10− 10− 10− 2 × × × E(B-V) GCC SZ 1 1 2

0 0 0

1 − 1 2 NGC SGC − − 2 − 8 8 8 10− 10− 10− 2 × × × CS-100 1 CS-143 CS-217 1 1

0 0 0

1 − 1 1 − 2 − −100 158 251 398 630 1000 100 158 251 398 630 1000 100 158 251 398 630 1000 ` ` `

ˆ 휅푞 × Figure A1. The estimated bias Δ퐶ℓ to the CMB lensing quasar cross-power spectrum caused by possible contamination sources in the CMB temperature map.

1 NGC, jackknife SGC, jackknife NS, jackknife χ (h− Mpc) ⊥ 190 120 75 47 30 19 1 0.01 0.09 0.12 -0.23 1 -0.03 0.12 0.23 0.05 1 -0.04 0.07 0.24 -0.11 0.5 NS NGC SGC 0.01 1 -0.09 -0.24 0.03 -0.03 1 0.11 0.16 -0.09 -0.04 1 -0.04 -0.07 -0.04 0.4 0.09 -0.09 1 -0.02 -0.08 0.12 0.11 1 -0.01 -0.16 0.07 -0.04 1 -0.06 -0.08

0.3 0.12 -0.24 -0.02 1 -0.25 0.23 0.16 -0.01 1 0.18 0.24 -0.07 -0.06 1 -0.02

` 0.2 -0.23 0.03 -0.08 -0.25 1 0.05 -0.09 -0.16 0.18 1 -0.11 -0.04 -0.08 -0.02 1 R NGC, simulation SGC, simulation NS, simulation . 0 1 1 -0.06 0.00 0.02 -0.09 1 -0.08 0.02 -0.06 0.14 1 -0.04 0.10 -0.01 0.06

0.0 -0.06 1 -0.13 0.06 0.02 -0.08 1 -0.07 0.08 -0.07 -0.04 1 -0.15 0.02 0.01

0.00 -0.13 1 -0.04 0.06 0.02 -0.07 1 0.01 0.04 0.10 -0.15 1 0.01 0.09 0.1 − 100 158 251 398 630 1000 ` 0.02 0.06 -0.04 1 -0.09 -0.06 0.08 0.01 1 -0.07 -0.01 0.02 0.01 1 -0.07 3 -0.09 0.02 0.06 -0.09 1 0.14 -0.07 0.04 -0.07 1 0.06 0.01 0.09 -0.07 1 NS NGC SGC

sim 2 /σ

jk 1 σ 0 Figure B2. Estimated correlation matrices, similar as Fig.9 but for 푅ℓ (Eq. B1). 15 jk - indv. jk - cum. 10 sim - indv. sim - cum.

SNR 5

0

Figure B1. 푅ℓ (Eq. B1) estimates, with similar information as in Fig.8. The shaded area in the upper panel is the best-fit value over scales with 1 휎 error, for each of the two caps and the combination.

MNRAS 000,1–15 (2020)