arxiv: v3 [astro-ph.co] 6 Nov 2018

Similar documents
Cooking with Strong Lenses and Other Ingredients

arxiv: v1 [astro-ph.co] 3 Apr 2019

CAULDRON: dynamics meets gravitational lensing. Matteo Barnabè

DA(z) from Strong Lenses. Eiichiro Komatsu, Max-Planck-Institut für Astrophysik Inaugural MIAPP Workshop on Extragalactic Distance Scale May 26, 2014

Matteo Barnabè & Léon Koopmans Kapteyn Astronomical Institute - Groningen (NL)

arxiv: v1 [astro-ph.co] 25 Sep 2015

Integral-Field Spectroscopy of SLACS Lenses. Oliver Czoske Kapteyn Institute, Groningen, NL

Physics of the Universe: Gravitational Lensing. Chris Fassnacht UC Davis

Discovery of eight lensing clusters of galaxies

arxiv: v1 [astro-ph.co] 11 Sep 2013

Cosmological Constraints from the double source plane lens SDSSJ

arxiv:astro-ph/ v1 16 Jan 1997

arxiv:astro-ph/ v1 23 Jan 2005

Probabilistic Catalogs and beyond...

Brief update (3 mins/2 slides) on astrophysics behind final project

Gravitational lensing constraint on the cosmic equation of state

arxiv: v1 [astro-ph.co] 17 Jul 2013

4. Structure of Dark Matter halos. Hence the halo mass, virial radius, and virial velocity are related by

Strong Gravitational-Lensing by Galaxies: 30 years later...

An accurate determination of the Hubble constant from baryon acoustic oscillation datasets

For β = 0, we have from our previous equations: d ln ν d ln r + d ln v2 r

Connecting observations to simulations arxiv: Joe Wolf (UC Irvine)

SUPPLEMENTARY INFORMATION

Black Hole and Host Galaxy Mass Estimates

Statistical Searches in Astrophysics and Cosmology

4. Structure of Dark Matter halos. Hence the halo mass, virial radius, and virial velocity are related by

Evidence for/constraints on dark matter in galaxies and clusters

The Stellar Initial Mass Function of Massive Galaxies

arxiv: v2 [astro-ph.co] 26 Nov 2018

SLACS Spectroscopy. Observations, Kinematics & Stellar Populations. Oliver Czoske Kapteyn Institute, Groningen, NL

How many arcmin-separation lenses are expected in the 2dF QSO survey?

arxiv:astro-ph/ v3 31 Mar 2006

A Spectroscopically-Confirmed Double Source Plane Lens in the HSC SSP Tanaka, Wong, et al. 2016, ApJ, 826, L19

Future precision cosmology and neutrinos

Probing Dark Matter Halos with Satellite Kinematics & Weak Lensing

Gravitational Lensing: Strong, Weak and Micro

Milky Way s Mass and Stellar Halo Velocity Dispersion Profiles

AS1001:Extra-Galactic Astronomy

Gravitational Lensing. A Brief History, Theory, and Applications

GALAXY-SCALE STRONG-LENSING TESTS OF GRAVITY AND GEOMETRIC COSMOLOGY: CONSTRAINTS AND SYSTEMATIC LIMITATIONS

arxiv: v2 [astro-ph.co] 1 Dec 2009

5.1 Circular Velocities and Rotation Curves

Source plane reconstruction of the giant gravitational arc in Abell 2667: a condidate Wolf-Rayet galaxy at z 1

Modified Gravity (MOG) and Dark Matter: Can Dark Matter be Detected in the Present Universe?

Baryon Acoustic Oscillations (BAO) in the Sloan Digital Sky Survey Data Release 7 Galaxy Sample

Summary So Far! M87van der Maerl! NGC4342! van den Bosch! rotation velocity!

Clusters: Observations

arxiv:astro-ph/ v1 10 Nov 1999

Clusters, lensing and CFHT reprocessing

A tool to test galaxy formation theories. Joe Wolf (UC Irvine)

Astro 242. The Physics of Galaxies and the Universe: Lecture Notes Wayne Hu

The Galaxy Dark Matter Connection

arxiv: v1 [astro-ph.ga] 12 Dec 2018

Cosmology and Strongly Lensed QSOs

SUPPLEMENTARY INFORMATION

Modern Image Processing Techniques in Astronomical Sky Surveys

Gravitational Lensing

THE ROAD TO DARK ENERGY

Cosmological Constraints from a Combined Analysis of Clustering & Galaxy-Galaxy Lensing in the SDSS. Frank van den Bosch.

arxiv:astro-ph/ v1 30 Nov 2004

Clusters: Observations

The θ-z s relation for gravitational lenses as a cosmological test

The mass of a halo. M. White

arxiv:astro-ph/ v1 27 Nov 2000

Sample variance in the local measurements of H 0. Heidi Wu (Caltech OSU) with Dragan Huterer (U. Michigan) arxiv: , MNRAS accepted

2. Lens population. 3. Lens model. 4. Cosmology. z=0.5. z=0

Weak Gravitational Lensing

BAO and Lyman-α with BOSS

The cosmic distance scale

The Degeneracy of Dark Energy and Curvature

PHY323:Lecture 7 Dark Matter with Gravitational Lensing

From quasars to dark energy Adventures with the clustering of luminous red galaxies

Connecting observations to simulations arxiv: Joe Wolf (UC Irvine)

AST1100 Lecture Notes

Elliptical galaxies as gravitational lenses

arxiv: v1 [astro-ph] 3 Dec 2007

Homework 9 due Nov. 26 (after Thanksgiving)

Hubble s Law and the Cosmic Distance Scale

Galaxy mass assembly in KiDS: dynamics and gravitational lensing

Time Delay in Swiss Cheese Gravitational Lensing

Dark matter annihilation and decay factors in the Milky Way s dwarf spheroidal galaxies

Dependence of low redshift Type Ia Supernovae luminosities on host galaxies

The Effective Cross-sections of a Lensing-galaxy: Singular Isothermal Sphere with External Shear

Inclination Effects in Spiral Galaxy Gravitational Lensing

To Lambda or not to Lambda?

Dark Matter Detection Using Pulsar Timing

Infrared Mass-to-Light Profile Throughout the Infall Region of the Coma Cluster

Survey of Astrophysics A110

The Local Group Timing Argument

The SINFONI Nearby Elliptical Lens Locator Survey (SNELLS)

arxiv: v1 [astro-ph] 31 Jul 2007

AST1100 Lecture Notes

arxiv: v1 [astro-ph.co] 18 Apr 2018

A very bright (i = 16.44) quasar in the redshift desert discovered by the Guoshoujing Telescope (LAMOST)

Cosmology on small scales: Emulating galaxy clustering and galaxy-galaxy lensing into the deeply nonlinear regime

The structures of distant galaxies IV. A new empirical measurement of the time-scale for galaxy mergers implications for the merger history

Strong gravitational lensing

Testing gravity on Large Scales

Constraints on Holographic Cosmological Models from Gamma Ray Bursts

Galaxies 626. Lecture 3: From the CMBR to the first star

Transcription:

Mon. Not. R. Astron. Soc., 1?? () Printed 7 November 218 (MN LATEX style file v2.2) arxiv:189.9845v3 [astro-ph.co] 6 Nov 218 Assessing the effect of lens mass model in cosmological application with updated galaxy-scale strong gravitational lensing sample Yun Chen 1, Ran Li 1,2, Yiping Shu 3,4 1 Key Laboratory for Computational Astrophysics, National Astronomical Observatories, Chinese Academy of Sciences, Beijing, 112, China; 2 University of Chinese Academy of Sciences, 19 A Yuquan Rd, Shijingshan District, Beijing, 149, China; 3 Purple Mountain Observatory, Chinese Academy of Sciences, 2 West Beijing Road, Nanjing 218, China; 4 Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 HA, UK 7 November 218 1 INTRODUCTION In the last two decades, owing to the advent of powerful new space and ground-based telescopes for imaging and spectroscopic observations, many new strong gravitational lensing (SGL) systems have been discovered. Thus it has become achievable and significant to extract the information of lens objects and cosmological parameters from SGL data. Since the number of observed galaxy-scale SGL systems is much more than that of galaxy cluster-scale SGL systems, most statistical analyses have utilized the galaxy-scale SGL sample. In practice, several different quantities can be adopted as statistical quantities with galaxy-scale SGL sample, including the distribution of image angular separations (see, e.g., Turner et al. 1984; Dyer 1984; Chiba & Yoshii 1999; Dev et al 24; Cao & Zhu 212), the distribution of lens redshifts (see, e.g., Turner et al. 1984; Kochanek 1992; Ofek et al. 23; Mitchell et al. 25; Cao et al. 212a), the time delay between images ( τ) (see, e.g., Refsdal 1964; Treu & Marshall 216; Bonvin et al. 217), and the velocity dispersion (σ) of lenses (see, e.g., Futamase & Yoshida 21; Biesiada 26; Grillo et al. 28; Schwab et al. 21; Cao c RAS ABSTRACT By comparing the dynamical mass and lensing mass of early type lens galaxies, one can constrain both the cosmological parameters and the density profiles of galaxies. In this paper, we explore the constraining power of this method on cosmological parameters, using a compiled sample of 157 galaxy-scale strong lensing systems, which is currently the largest sample with both high resolution imaging and stellar dynamical data. These selected lenses are all early-type galaxies with E or S morphologies and without significant substructures or close companion galaxies. We assume a power-law mass model for the lenses, and consider three different parameterizations for the slope (γ ) of the total mass density profile (ρ(r) r γ ) to include the effect of the dependence of lens mass model on redshift and surface mass density. By fitting simultaneously the cosmological model and the lens mass model, we find that the cosmological parameters are not very well constrained with current data, and the posterior distribution on Ω m is heavily dependent on the choice of parameterization for the lens mass model. By employing the Bayesian information criterion, we find the lens mass model including dependencies on both redshift and surface mass density is favored. We observe the dependencies of γ on the redshift and the surface mass density at >3σ and >5σ levels, respectively. We conclude that unless the significant dependencies of γ on both the redshift and the surface mass density are properly taken into account, the strong gravitational lensing systems under consideration cannot served as a kind of promising cosmological probes. Key words: (cosmology:) cosmological parameters - cosmology: observations - gravitational lensing: strong - galaxies: structure et al. 217). The major disadvantage of using the distributions of image angular separations and lens redshifts as statistical quantities is that the theoretically predicted values of these two are dependent not only on the lens mass model but also on the lens luminosity function. While the theoretical predictions of τ and σ are dependent only on the lens mass model but not on the luminosity function. Thus the methods of using τ and σ as statistical quantities are more popular at present. According to the theoretical analysis, one can see that τ is more sensitive to the cosmological parameters than σ (Paraficz & Hjorth 29; Wei & Wu 217). The fact also proves that the measurements of τ are very powerful at constraints on the cosmological parameters and especially sensitive to the Hubble constant H (Bonvin et al. 217; Suyu et al. 217; Liao et al. 217). However, the related cosmological application is mainly limited by the uncertainty in the lens modeling and the small number of lenses with measured time delays. In contrast, the number of lenses with measured σ are much larger. Although the measurements of σ are weak at confining the cosmological parameters (see, e.g., Biesiada 26; Cao et al. 212b; Wang & Xu

2 Chen et al. 213; Chen et al. 215; Cao et al. 215; An et al. 216; Xia et al. 217; Cui et al. 217; Li et al. 218a), they are useful for investigating the lens mass models if the priors on cosmological parameters are given (see, e.g., Koopmans et al. 29; Sonnenfeld et al. 213a; Cao et al. 216; Holanda et al. 217). By combining the observations of SGL and stellar dynamics in elliptical galaxies, one can use the lens velocity dispersion (VD) as statistical quantity to put constrains on both the cosmological parameters and the density profiles of galaxies. The core idea of this method is that the gravitational mass M E grl and the dynamical mass M E dyn enclosed within the disk defined by the so-called Einstein ring should be equivalent, namely, M E grl = ME dyn. Further, ME grl inferring from the strong lensing data depends on cosmological parameters, and M E dyn inferring from the stellar VD depends on the lens mass model, so one can relate the VD with the model parameters including both cosmological and lens mass model parameters. This method can be traced back to Futamase & Yoshida (21), but at that time there were no available observational data of lens VD. Grillo et al. (28) first applied this method to constrain cosmological parameters with observational data, wherein the sample included 2 SGL systems from the Lens Structure and Dynamics (LSD) survey and the Sloan Lens ACS (SLACS) survey. In the literature, a recent compiled sample which can be used in this method includes 118 galaxy-scale SGL systems (Cao et al. 215, hereafter C15) from the SLACS survey, the Baryon Oscillation Spectroscopic Survey (BOSS) emission-line lens survey (BELLS), the LSD survey, and the Strong Lensing Legacy Survey (SL2S). In this paper, we update the sample with definite criteria by taking advantage of new observational data, and then explore the effect of lens mass model on constraining cosmologcial parameters, as well as evaluate several different lens mass models. The rest of the paper is organized as follows. In Section 2, we demonstrate the methodology of using the lens velocity dispersion as statistical quantity to constrain model parameters. Then, in Section 3 the SGL data sample used in our analysis is introduced. In Section 4, we first investigate the sensitivity of the sample under consideration to cosmological parameters, and diagnose whether our lens sample has any evolutionary signal via the qualitative and semi-quantitative analysis; and then carry out observational constraints on parameters of cosmology and lens mass models; and finally turn to compare several different lens mass models. In the last section, the main conclusions are summarized. 2 METHODOLOGY As discussed in the last section, the method of using the galaxy lens VD as statistical quantity has some special merits. However, in this method, besides the imaging data of the SGL systems, one also has to possess the spectroscopic data of the systems and measure the central velocity dispersion of the lens galaxies from the spectroscopy. On the basis of various recent lensing surveys which have carried out both imaging and spectroscopic observations, this method has become feasible. In this method, the main idea is that the projected gravitational mass M E grl and the projected dynamical mass ME dyn within the Einstein radius should be equivalent, i.e., M E grl = ME dyn. (1) From the theory of gravitational lensing, the projected gravitational mass within the Einstein radius is M E grl = Σ cr πr 2 E. The Einstein radius R E is determined by R E = θ E D l, wherein θ E is the Einstein angle, and D l is the angular diameter distance between observer and lens. The critical surface mass density Σ cr is defined by Σ cr = c 2 D s 4πG D l D ls, where D ls is the angular diameter distance between lens and source, and D s is that between observer and source. Thus, one can further figure out M E grl = c2 D s D l θ 2 4G D E, (2) ls wherein the distances D s, D l and D ls are dependent on the cosmological model. To estimate the projected dynamical mass M E dyn from the lens galaxy VD, one must first suppose the mass distribution model for the lens galaxy. Here we choose a general mass model (Koopmans 26) for the lens galaxies which are early-type galaxies (ETGs) with E or S morphologies: ρ(r) = ρ r γ ν(r) = ν r δ β(r) = 1 σ 2 θ /σ2 r where ρ(r) is the total (i.e. luminous plus dark-matter) mass density distribution, and ν(r) is the luminosity density of stars. The parameter β(r) denotes the anisotropy of the stellar velocity dispersion, and is also called as the stellar orbital anisotropy, where σ θ and σ r are the tangential and radial velocity dispersions, respectively. Based on the assumption that the relationship between stellar number density n(r) and stellar luminosity density ν(r) is spatially constant, an assumption unlikely to be violated appreciably within the effective radius of the early-type lens galaxies under consideration, the radial Jeans equation in Spherical Coordinate can be written as where (3) d dr [ν(r)σ2 r ] + 2β r ν(r)σ2 r = ν(r) dφ dr, (4) dφ dr = GM(r) r 2, (5) and M(r) is the total mass inside a sphere with radius r. By substituting Eq. (5) into Eq. (4), one can get the expression for σ 2 r, σ 2 r (r) = G r dr r 2β 2 ν(r )M(r ), (6) r 2β ν(r) By defining r to be the spherical radial coordinate from the lens center, Z to be the axis along the line of sight (LOS), and R to be the cylindrical radius which is perpendicular to the LOS, then one has r 2 = R 2 +Z 2. The projected dynamical mass M dyn contained within a cylinder of radius equal to the Einstein radius R E can be calculated with where Σ(R) = = = RE M E dyn = dr2πr Σ(R ), (7) ρ(r)dz dz ρ (Z 2 + R 2 ) γ/2 r γ Γ ( γ 1 (π)r 1 γ 2 ρ Γ(γ/2) r γ By substituting Eq.(8) into Eq.(7), one can have M E R3 γ Γ ( ) γ 1 E 2 ρ dyn = 2π3/2 3 γ Γ(γ/2) ) r γ (8). (9)

3 The total mass contained within a sphere with radius r is M(r) = r dr 4πr 2 ρ(r ) = 4π ρ r 3 γ r γ 3 γ. (1) By combining Eqs. (9) and (1), one can further have M(r) = 2 ( ) 3 γ Γ(γ/2) r π Γ( γ 1 ) M E dyn R. (11) E 2 By substituting Eqs. (11) and (3) into Eq. (6), one reads σ 2 r (r) = 2 GM E ( ) 2 γ dyn 1 Γ(γ/2) r π R E ξ 2β Γ( γ 1 ), (12) R E 2 where ξ = γ + δ 2, and β is assumed to be independent of the radius r. The actual velocity dispersion of the lens galaxy measured by the observation is the component of luminosity-weighted average along the LOS and over the effective spectroscopic aperture R A, that can be expressed mathematically σ 2 ( R A) = RA RA dr 2πR dz σ2 los ν(r) dr 2πR dz ν(r) (13) where σ 2 los is the LOS velocity dispersion, which is a combination of the radial (σ 2 r ) and tangential (σ 2 t ) velocity dispersions. Using θ to indicate the angle between the LOS (Z-axis) and the radial direction (r-axis), then one reads σ 2 los = (σ r cos θ) 2 + (σ t sin θ) 2 = σ 2 r 2 R 2 r + σ 2 R 2 r 2 t r 2 = σ 2 r (1 R2 r ) + (1 R 2 2 β)σ2 r r 2 = σ 2 r (1 β R2 r 2 ) (14) By substituting Eq.(14) into Eq.(13), one can read σ 2 ( R A) = RA dr 2πR dz σ2 r (r)(1 β R2 )ν(r) r 2 RA dr 2πR (15) dz ν(r) Further, by substituting Eq.(12) and (3) into Eq.(15), one obtains [ σ 2 ( R A) = 2 GM dyn E 3 δ Γ[(ξ 1)/2] π R E β Γ[(ξ+1)/2] ] (ξ 2β)(3 ξ) Γ(ξ/2) Γ(γ/2)Γ(δ/2) Γ[(γ 1)/2]Γ[(δ 1)/2] Γ[(ξ+2)/2] ( RA R E ) 2 γ. (16) Finally, with the relation expressed in Eq.(1), the above formula can be rewritten as [ σ 2 ( R A) = c2 2 D s 3 δ Γ[(ξ 1)/2] π D ls θ E β Γ[(ξ+1)/2] ] (ξ 2β)(3 ξ) Γ(ξ/2) Γ(γ/2)Γ(δ/2) Γ[(γ 1)/2]Γ[(δ 1)/2] Γ[(ξ+2)/2] ( θa θ E ) 2 γ, (17) where R A = θ A D l. From the spectroscopic data, one can measure the velocity dispersion σ ap inside the circular aperture with the angular radius θ ap. In practice, if the σ ap are measured within rectangular apertures, one usually derives the equivalent circular apertures with the angular radii θ ap following Jørgensen et al. (1995), θ ap 1.25 (θ x θ y /π), (18) where θ x and θ y are the angular sizes of width and length of the rectangular aperture. More precisely, σ ap is the projected, luminosity weighted average of the radially dependent velocity dispersion profile of the lensing galaxy inside θ ap. For a fair comparison and in consideration of the effect of the aperture size on the measurements of velocity dispersions, all velocity dispersions σ ap measured within apertures of arbitrary sizes, are normalized to a typical physical aperture, σ e2, with the radius R eff /2, where R eff is the half-light radius of the lens galaxy, and R eff /2 is a quantity that is well-matched to the typical Einstein radius. Following the prescription, one can use the aperture correction formula, σ obs σ e2 = σ ap [θ eff /(2θ ap )] η, (19) where θ eff = R eff /D l. The best-fitting values of the correction factor η are different when using different observational samples. For example, the best-fitting values of η are.4,.6 and.66 ±.35 found by Jørgensen et al. (1995), Mehlert et al. (23) and Cappellari et al. (26), respectively. Compared with the values of σ e2 calculated with η =.4, the relative differences of σ e2 in the cases of η =.6 and η =.66 are smaller than 5% for the lenses under consideration. In practice, we first calculate the mean value and the statistical error of σ e2 with η =.4 following Jørgensen et al. (1995), where the the statistical error, σ stat e2, is propagated from the measurement error of σ ap. And on this basis, we conservatively assume an uncertainty of 5% on σ e2 due to the aperture correction, which is marked as σ AC e2. In order to compare the observational values of the VD with the corresponding model-predicted ones, one needs to calculate the theoretical value of the VD within the radius R eff /2 from Eq. (17), σ ( θeff /2) = c 2 2 π D s 3 δ θ E D ls (ξ 2β)(3 ξ) ( ) (2 γ) θeff F(γ, δ, β), 2θ E (2) where [ [ ] Γ (ξ 1)/2 F = β Γ [ (ξ + 1)/2 ] ] Γ(ξ/2) Γ [ (ξ + 2)/2 ] Γ(γ/2)Γ(δ/2) Γ [ (γ 1)/2 ] Γ [(δ 1)/2]. (21) In the case of γ = δ = 2 and β =, the mass model is reduced to the well-known Singular Isothermal Sphere(SIS) model, and the predicted value of the VD is recovered to σ SIS = c 2 4π D s D ls θ E. (22) In our analysis, the likelihood is assumed to be χ 2 is constructed as χ 2 = L e χ2 /2. (23) N i=1 σ th,i σobs,i σ obs,i is the observa- ) 2 = ( σ stat e2 )2 + from Eqs. (19) and (2), re- where N is the number of the data points, σ obs,i tional error of σ obs,i, which is calculated with ( σobs ( σ AC e2 )2. One can obtain σ obs,i spectively. 3 DATA SAMPLE and σ th,i 2, (24) According to the analysis in the last section, one can learn that the method under consideration requires the following information from observations, including the lens redshift z l, the source redshift

4 Chen et al. z s, the Einstein angle θ E, the central VD of the lens galaxy σ ap, the spectroscopic aperture angular radius θ ap, and the half-light angular radius of the lens galaxy θ eff. Additionally, to ensure the validity of the assumption of spherical symmetry on the lens galaxy, the selected lens galaxies should satisfy the following conditions:(i) the lens galaxy should be ETGs with E or S morphologies; and (ii) the lens galaxy should not have significant substructure or close massive companion. Some lens galaxies from C15 do not satisfy the condition (i). Here we assemble a sample including 157 definite galaxy-scale SGL systems which meet all the requirements mentioned above, where 5 systems from the LSD survey 1 (Koopmans & Treu 22; Treu & Koopmans 22; Koopmans & Treu 23; Treu & Koopmans 24), 26 from the SL2S (Ruff et al. 211; Sonnenfeld et al. 213a,b; Sonnenfeld et al. 215), 53 from the SLACS (Bolton et al. 28; Auger et al. 29, 21), 38 from the an extension of the SLACS survey known as SLACS for the Masses (hereafter S4TM, Shu et al. 215; Shu et al. 217), 21 from the BELLS (Brownstein et al. 212), and 14 from the BELLS for GALaxy-Lyα EmitteR systemsgallery (hereafter BELLS GALLERY, Shu et al. 216a,b). The useful information of these 157 systems is listed in Table 1. The velocity dispersions of the lenses from LSD and SL2S surveys, which are measured within rectangular slits, are transformed into velocity dispersions σ e2, within a circular aperture with radius R eff /2 based on Eqs.(18) and (19). The SLACS and S4TM surveys select candidates from Sloan Digital Sky Survey I (SDSS-I, Eisenstein et al. 21; Strauss et al. 22) data, in which the velocity dispersions of the lenses are measured within the 1.5 radius fibers. The lens candidates of the BELLS and BELLS GALLERY surveys are spectroscopically selected from the BOSS (Dawson et al. 213) of the Sloan Digital Sky Survey-III (SDSS-III, Eisenstein et al. 211), in which the VD of the lenses are measured within the 1 radius fibers. These velocity dispersions measured with fibers are corrected to σ e2 based on Eq.(19). The distribution of the whole SGL sample is shown in Fig. 1. From the upper panels of Fig. 1, one can see that 3% of the lenses are located at z l.2, and only 5% located at z l >.75. The lower panels of Fig. 1 show that 8% of the lenses possess the velocity dispersions 18 km s 1 < σ e2 < 3 km s 1. 4 ANALYSIS AND RESULTS 4.1 Qualitative analysis 4.1.1 Sensitivity to cosmological parameters From Eq. (2) one can see that the cosmological model enters into the theoretical observable σ th not through a distance measure directly, but rather through a distance ratio D s D ls = zs dz E(z;p) zs dz E(z;p) z l, (25) where in the framework of the flat FLRW metric the theoretical values of D s and D ls can be obtained by D s (z s ; p, H ) = c H (1 + z s ) 1 http://web.physics.ucsb.edu/ tt/lsd/ zs dz E(z; p), (26) and D ls (z l, z s ; p, H ) = c zs dz H (1 + z s ) z l E(z; p), (27) respectively, where p denotes the parameter space of the considered cosmological model, and E = H/H is the dimensionless Hubble parameter, and c is the velocity of light. Hence, the advantage and disadvantage are both remarkable in this method. The positive side is that theoretical observable is independent of the Hubble constant H which gets canceled in the distance ratio. Currently, the trouble with H is that the values measured by different methods are significantly in tension (see, e.g., Bernal et al. 216; Chen et al. 217; Freedman 217). The most representative example is that the indirect (model-dependent) measurement of H from Planck mission (Planck Collaboration: Ade, P. A. R., et al. 214, 216) when ΛCDM model is assumed is much smaller than and in tension (at > 2σ confidence level) with the direct measurement from Hubble Space Telescope (HST) key project (Riess et al. 211, 216). Thus, it is a very tough problem to set a prior on H. In this situation, the significant advantage of the method under consideration is that it directly avoids the effect of H. On the other side, the main disadvantage is that the power of estimating cosmological parameters is poor. The distance ratio D s /D ls is a ratio of two integrals which have the same integrand (i.e., 1/E(z; p)) and differ only by the limits of integration, so the theoretical observable σ th D s /D ls is insensitive to the cosmological parameters involved in the integrand (Biesiada et al. 21). In Fig. 2, we show the impact of the matter-density parameter Ω m on the distance ratio by taking a spatially flat ΛCDM model with Ω m =.3 as a fiducial cosmological model. The three panels of Fig. 2 display the evolution of D s /D ls with respect to the source redshift z s along with variety of Ω m, corresponding to the cases of the lens redshift z l =.1,.5, and1 from left to right. The general trend is that the sensitivity of D s /D ls to Ω m increases with the increase of z l. In Fig. 2, the shadows denote the cases that the relative uncertainties of D s /D ls are 1% and 2%, respectively, with respect to the fiducial value. One can see that an individual lens with z l =.1 cannot put any constraint on Ω m even when D s /D ls only has 1% uncertainty. An individual lens with z l =.5 can bound on Ω m with 8% 16% relative uncertainty when D s /D ls only has 1% uncertainty, but cannot put any constraint on Ω m when the uncertainty of D s /D ls increases to 2%. Unfortunately, with regard to the SGL sample under consideration, the typical values of the relative uncertainties of D s /D 2 ls are approximately 1% and 2% at z l.1 and.5, respectively. It means that most lenses with z l <.5 in our sample do not contribute to the limit on Ω m. An individual lens with z l = 1 can put a limit on Ω m with 5% 1% ( 8% 2%)relative uncertainty, corresponding to D s /D ls with 1%(2%) uncertainty. In our sample, there is only one system with z l > 1, that is MG216+112 with z l = 1.4 from LSD survey. In general, one is not able to make a reliable estimate on Ω m with the sample under consideration. After repeating similar analyses for other cosmological parameters (i.e., the equation of state parameter of dark energy, and the curvature parameter), we find out that the current sample is really weak at confining these cosmological parameters. 2 The uncertainty on D s /D ls is mainly propagated from that on σ e2. The relative uncertainty on D s /D ls is about 2 times of that on σ e2 because of D s /D ls σ 2 e2.

5 4.1.2 Diagnosing the evolution of the lens sample s mass-density profile In order to diagnose whether there is any evolutionary signal among our lens sample, we divide the entire sample into several different sub-samples, and then check whether significant differences exist among the ranges of lens mass model parameters constrained from different sub-samples. In practice, we divide the entire sample into four sub-samples according to the velocity dispersion: σ e2 21 (n = 4), 21 < σ e2 243 (n = 39), 243 < σ e2 276 (n = 39), and σ e2 > 276 (n = 39), where σ e2 is in unit of km/s, and n is the size of each sub-sample. Note that the breakpoints are chosen to insure that the sub-sample sizes are nearly equal. We carry out observational constraints on all parameters of the lens mass model formulized by Eq. (3) from four sub-samples, respectively, by keeping the cosmological parameters fixed with Ω m =.3 in the framework of a spatially flat ΛCDM model. In our analysis, the likelihood is computed with Eqs. (23) and (24). We derive the posterior probability distributions of parameters through an affine invariant Markov chain Monte Carlo (MCMC) Ensemble sampler (emcee; Foreman-Mackey et al. 213). The onedimensional (1D) probability distributions and two-dimensional (2D) contours for the model parameters are displayed in Fig. 3. It turns out that the limits on parameter γ from different sub-samples have significant distinctions. While the ranges of δ and β obtained from different sub-samples are consistent at 68% confidence level. In addition, the limits on these two parameters are much weaker than those on γ. It is reasonable because, theoretically, the deflection of light due to gravitational lensing is more sensitive to the total mass, so the lensing data are most sensitive to the parameter γ. In general, the parameter γ presents the most significant evolutionary signal. In other words, the dependence of γ on the properties of lenses should not be ignored in the statistical analysis. 4.2 Observational constraints 4.2.1 Constraints on parameters of cosmology and lens mass models We assume a kind of spherically symmetric mass distributions (i.e. Eq. (3)) for the lens galaxies in the kinematic analysis. As discussed above, the dependence of γ on the properties of lenes should be taken into account. In the literature, the dependence of γ on the redshift has been widely studied (see, e.g., Ruff et al. 211; Bolton et al 212; Cao et al. 215; Cao et al. 216; Cui et al. 217; Holanda et al. 217). Besides, Auger et al. (21) also found a significant correlation between γ and total mass surface density, that has also been confirmed by Dutton & Treu (214) and Sonnenfeld et al. (213a). In the light of these works, we specifically consider three lens mass models on the basis of Eq. (3), corresponding to three parameterizations for γ, namely: P 1 : γ = Constant, P 2 : γ = γ + γ z z l, P 3 : γ = γ + γ z z l + γ s log Σ, where γ is treated as an arbitrary constant in case P 1, and its dependence on the lens redshift z l is considered in case P 2. Besides, the dependence on both the redshift and the surface mass density is taken into account in case P 3. According to the virial theorem, the projected dynamical mass within the radius R eff /2 satisfies σ 2 e2 R eff (see, e.g., Auger et al. 21), so the corresponding surface mass density is Σ σ 2 e2 /R eff. Here, we use Σ to denote M dyn e2 the normalized surface mass density of the lens galaxy, which is expressed as Σ = (σ e2/1km s 1 ) 2 R eff /1h 1 kpc, (28) where the usual convention of writing the Hubble constant as H = 1h kms 1 Mpc 1 is adopted. Hereafter, Model 1, Model 2, and Model 3 denote the lens mass models in which γ is parameterized in the forms of P 1, P 2, and P 3, respectively. As mentioned above, the sample under consideration is quit weak at constraining cosmological parameters, so constraining too many cosmological parameters simultaneously would only distort the results. Thus, we only attempt to fit Ω m in the framework of flat ΛCDM model, where Ω m is the only free parameter of cosmology. We then conduct observational constraints on Ω m and lens mass model parameters in the schemes of Model 1, Model 2 and Model 3, respectively. The results are displayed in Figure 4, including the 1D probability distributions and 2D contours for the parameters of interest. The limits on Ω m at 68% (95%) confidence level are Ω m <.47(Ω m <.113), Ω m <.159(Ω m <.41) and.432 < Ω m <.88 (.261 < Ω m <.948) in the frameworks of three different lens mass models, respectively. The main tendencies can be summarized to three aspects. First, the limits on Ω m are significantly dependent on the lens mass model. The allowed range of Ω m in the third scenario is inconsistent with those obtained in the former two at 68% confidence level. Second, the constraints on Ω m are weak. In the first two scenarios, the lower limits on Ω m are unavailable. In the last scenario, the relative uncertainty of Ω m is about 3% at 68% confidence level. It is consistent with the qualitative analysis mentioned previously, which reveals that the sample under consideration is insensitivity to Ω m. Third, the estimations on Ω m are biased. The mean value of Ω m constrained from the standard cosmological probes is around.3 (see, e.g. Huterer & Shafer 218; Scolnic et al. 218; Alam et al. 217), such as Ω m =.315 ±.7 in the framework of flat ΛCDM model obtained from the recent Planck 218 result (Planck Collaboration: Aghanim, N., et al. 218). In the first scenario, the limit on Ω m is inconsistent with that from the Placnck result at 95% confidence level, and the allowed values of Ω m are especially low. In the second scenario, it is consistent with the Placnk result at 95% confidence level, but the mean value of Ω m is much lower. In the last scenario, it is consistent with the Placnk result at 95% confidence level, but the mean value is much higher. The bias in the estimation of Ω m must be related to the insensitivity of the distance ratio D s /D ls to the cosmological parameters as discussed previously. Besides, it may also be due to some unkown systematic errors. This is the first time to constrain the cosmological parameter in the scenario of considering the dependence of γ on both redshift and surface mass density. In C15, they constrained the equation of state (EoS) of dark energy (with other cosmological parameters fixed) from their sample with 118 systems in the scenarios of P 1 and P 2. From the results listed in Table 2 of C15, one can see that the constraints on the EoS of dark energy are also quit weak, the uncertainties are bigger than 3%. 4.2.2 Comparing lens mass models As discussed previously, one is not able to make reliable estimates for cosmological parameters based on the current lensing and kinematics data of the galaxy-scale lens sample. What is more, the limits on cosmological parameters are quite dependent on the lens mass model. So it is necessary to compare the lens mass models and

6 Chen et al. select the most competitive one, that can supply helpful reference for future studies on selecting the lens mass model. And fortunately, the key cosmological parameters have been determined precisely by the current standard cosmological probes. Given these things, we choose to compare the lens mass models by using priors on cosmological parameters from the Planck 218 results (Planck Collaboration: Aghanim, N., et al. 218), i.e., assuming a flat ΛCDM model with Ω m =.315 ±.7 as a fiducial cosmological model. Then, we put observational constraints on the lens mass models. Our results are summarized in Table 2 and Figure 5. The 1D probability distributions and 2D contours for the parameters of interest are shown in Figure 5. The mean values with 68% confidence limits for the parameters listed in Table 2 are derived from the corresponding 1D marginalized distribution probabilities. In the framework of Model 1 (i.e., γ = Const.), one can find out that the limit on γ is much tighter than those on δ and β, where the relative uncertainties of γ, δ and β are 2%, 25% and 1%, respectively. It agrees with the theory of gravitational lensing (see, e.g., Wambsganss 1998), which indicates that the lensing data are most sensitive to the parameter γ as mentioned previously. Comparing with the result in the Model 1, we observe a minor shift in the mean value and a significant reduction in the uncertainty for δ in the other two models. More specifically, the relative uncertainty of δ is 25% in the Model 1, which is reduced to 15% in the Model 2, and further to 5% in the Model 3. In contrast, there is a significant shift in the mean value of β, but because of the large uncertainty on it, the limits on β are consistent at 68% confidence level in all three scenarios. And, the mean value of β in Model 3 is much closer to that from Koopmans et al. (29). In the framework of Model 2, γ z = is ruled out at 2.3σ level. Cao et al. (216) used the sample of C15 to constrain the lens mass model parameters in the case of Model 2 with cosmological parameters fixed, but without considering the dependence of γ on surface density. Thus, we just can compare our result with theirs in the case of Model 2. While they found the dependence on redshift at 1.5σ level from C15 s sample. It turns out that the dependence of γ on the redshift can be observed at higher confidence level based on our updated sample. Further, in the framework of Model 3, γ z = is ruled out at 3.5σ level, and γ s = is ruled out at 6σ level. Sonnenfeld et al. (213a) found the depdendence of γ on redshift and surface mass density at 3.1σ and 5.4σ levels, respectively, from the SL2S, SLACS and LSD lenses. It implies that the corresponding signals are extracted from our sample at a bit higher levels. Based on the previous analysis, we conclude that dependencies of γ on both the redshift and the surface mass density are very significant. Furthermore, we employ the Bayesian information criterion (BIC) to compare the three models. The BIC (Schwarz 1978) is defined as BIC = 2 ln L max + k ln N, (29) where L max is the maximum likelihood (satisfying 2 ln L max = χ 2 min under the Gaussian assumption), k is the number of the parameters of the considered model, and N is the number of data points used in the fitting. The BIC is widely used in a cosmological context(see, e.g., Liddle 24; Godłowski & Szydłowski 25; Magueijo & Sorkin 27; Mukherjee et al. 26; Biesiada 27; Davis et al. 27; Li et al. 213; Wen et al. 218). This statistic prefers models that give a good fit with fewer parameters. The favorite model is the one with the minimum BIC value. The BIC values for the three models are 344.924, 334.259, and 194.84, respectively. So, the most competitive one is Model 3, that is consistent with the previous conclusion. In order to verify the robustness of the above conclusion, we then turn to consider the impacts of the prior on Ω m and the systematic error in the velocity dispersion. In previous analyses, we have adopted a prior value on Ω m from the Planck 218 result. When changing the prior value to Ω m =.38 ±.12 from the Planck 215 result (Planck Collaboration: Ade et al. 216), we find that the relative differences in the mean values of the lens model parameters are less than 3%, which brings in an insignificant impact on our conclusion. In addition, the systematic error in the velocity dispersion has not yet been taken into account. While Jiang & Kochanek (27) considered five sources of systematic error in the velocity dispersion and found that the fractional systematic error is approximately 8%. When combining the 8% systematic error with the observational error in quadrature, we still can observe the dependence of γ on the redshift and the surface mass density at 3σ and 5σ levels, respectively. 5 SUMMARY AND CONCLUSIONS We have compiled a galaxy-scale strong gravitational lensing sample including 157 systems with the gravitational lensing and stellar velocity dispersion measurements, which are selected with strict criteria to satisfy the assumption of spherical symmetry on the lens mass model. Actually, the selected lenses are all early-type galaxies with E or S morphologies. A kind of spherically symmetric mass distributions formulized by Eq.(3) is assumed for the lens galaxies throughout this paper. After carrying out the qualitative and semiquantitative analysis, we find that the current sample is weak at confining cosmological parameters, and the lens mass model parameter γ (i.e., the slope of the total mass density profile) presents a significant evolutionary signal. Given this, we specifically consider three lens mass models on the basis of Eq. (3), corresponding to three parameterizations for the slope γ. The slope γ is treated as an arbitrary constant without considering any dependency in the first scenario (namely Model 1 ). And its dependence on the lens redshift is considered in the second scenario (namely Model 2 ). Further, its dependencies on both the redshift and the surface mass density of the lens are taken into account in the last scenario (namely Model 3 ). We first try to constrain both the cosmological and lens mass model parameters. However, it turns out that the limit on the cosmological paramerer, Ω m, is quit weak and biased, as well as quite dependent on the lens mass model. So, we turn to compare and select the lens mass models by keeping the cosmological parameters fixed. The result shows that the Model 3 is most preferred by the current sample, and the dependencies of γ on the redshift and on the surface mass density are observed at >3σ and >5σ levels, respectively. And, as discussed in Sec. 4.2.2, the corresponding signals are observed at higher levels from our updated sample, compared to the previous studies. The consequence still holds after considering the impacts of the prior on Ω m and the systematics error in the velocity dispersion. In addition, we have treated both the slope (δ) of the luminous matter density profile (ν(r) r δ ) and the orbit anisotropy parameter β as arbitrary constants in our observational constraints. However, Xu et al. (217) found the dependencies of δ and β on stellar velocity dispersion and redshift at different confidence levels, based on the Illustris simulation. Even so, our results are still not in conflict with theirs, because the uncertainties on these two parameters obtained from our constraints can mimic the intrinsic spreads in these quantities.

7 Consequently, the dependencies of γ on both the redshift and the surface mass density are very significant, that should be taken into account in the cases of employing the strong gravitational lensing systems as cosmological probes. Moreover, the slope γ has a positive correlation with the surface mass density, and a negative correlation with the redshift. It is worth noting that the dependency of γ on redshift and surface mass density does not represent that the ETGs change their mass density profile over the lifetime (Sonnenfeld et al. 213a), that mainly denotes the statistical feature of the population included in our sample. The overall trends show that, at a given redshift, the galaxies with high density also have steeper slopes; and, at fixed surface mass density, the galaxies at a lower redshift have steeper slopes. These trends are consistent with those obtained in the previous studies (e.g., Auger et al. 21; Ruff et al. 211; Bolton et al 212; Holanda et al. 217; Sonnenfeld et al. 213a; Li et al. 218b). Understanding the deep meanings of these trends on the evolution of individual ETGs requires more other observational data, that is beyond the scope of this paper. Finally, we point out that besides the dependence of γ on redshift and surface mass density considered in this work, other important dependencies may also be found in future, that can lead to a more accurate phenomenological model for lens galaxies. In addition, although the measurements of the velocity dispersions (σ) of lens galaxies alone are insufficient to make reliable estimates on the cosmological parameters, measurements of time delays ( τ) and the joint measurements of the former two ( τ/σ 2 ) are both proved to be more sensitive to the cosmological parameters (Paraficz & Hjorth 29; Wei & Wu 217). At present, the cosmological implementation is limited by the uncertainty in the lens modeling and the small number of lenses with measured time delays. One can anticipate significant improvements in these two aspects in view of the future observational facilities and hydrodynamic simulations of galaxy formation (see, e.g., Oguri & Marshall 21; Suyu et al. 217; Shu et al. 218; Treu et al. 218). ACKNOWLEDGMENTS We would like to thank Shuo Cao, Marek Biesiada, Lixin Xu, Guojian Wang, Qiao Wang, Ming Li, Jie Wang and Liang Gao for helpful discussions. YC has been supported by the National Natural Science Foundation of China (Nos. 117334 and 1157331), and the NAOC Nebula Talents Program. RL has been supported by the National Key Program for Science and Technology Research and Development of China (217YFB233), the National Natural Science Foundation of China (Nos. 1177332 and 118513), and the NAOC Nebula Talents Program. YS has been supported by the National Natural Science Foundation of China (Nos. 116332 and 113338), the 973 program (No. 215CB8573), and the Royal Society K.C. Wong International Fellowship (NF17995). REFERENCES Alam, S., et al., 217, MNRAS, 47, 2617. An, J., Chang, B.-R., & Xu, L.-X., 216, ChPhL, 33, 981 Auger, M. W., et al., 29, ApJ, 75, 199. Auger, M. W., et al., 21, ApJ, 724, 511. Bernal, J. L., Verde, L., & Riess, A. G., 216, JCAP, 1, 19. Biesiada, M. 26, PhRvD, 73, 36. Biesiada, M. 27, JCAP, 2, 3. Biesiada, M., Piorkowska, A., & Malec, B., 21, MNRAS, 46, 155. Bolton, A. S., et al., 28, ApJ, 682, 964. Bolton, A. S., et al., 212, ApJ, 757, 82. Bonvin, V., et al., 217, MNRAS, 465, 4914 Brownstein, J. R., et al., 212, ApJ, 744, 41. Cao, S., Covone, G., & Zhu, Z.-H., 212a, ApJ, 755 (212) 31. Cao, S., et al., 212b, JCAP, 3, 16. Cao, S. & Zhu, Z.-H., 212, A& A, 538, 43. Cao, S., et al., 215, ApJ, 86, 185 (C15) Cao, S., et al., 216, MNRAS, 461, 2192. Cao, S., et al., 217, ApJ, 835, 92. Cappellari, M., et al., 26, MNRAS, 366, 1126 Chen, Y., et al., 215, JCAP, 2, 1. Chen, Y., Kumar, S., & Ratra, B., 217, ApJ, 835, 86. Chiba, M., & Yoshii, Y.,1999, ApJ, 51, 42. Cui, J.-L., Li, H.-L. & Zhang, X., 217, Sci. China-Phys. Mech. Astron., 6, 8411. Davis, T. M., et al., 27, ApJ, 666, 716. Dawson, K. S., et al., 213, AJ, 145, 1. Dev, A., Jain, D., & Mahajan, S., 24, IJMPD, 13, 15. Dutton, A. A. & Treu, T., 214, MNRAS, 438, 3594. Dyer, C. C., 1984, ApJ, 287, 26. Eisenstein, D. J., et al., 21, AJ, 122, 2267. Eisenstein, D. J., et al., 211, AJ, 142, 72. Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J., 213, PASP, 125, 36. Freedman, W. L., 217, Nature Astronomy, 1, 169. Futamase, T., & Yoshida, S., 21, PThPh, 15, 887. Godłowski, W., & Szydłowski, M. 25, Physics Letters B, 623, 1. Grillo, C., Lombardi, M., & Bertin, G., 28, A&A, 477, 397. Holanda, R. F. L., Pereira, S. H., & Jain, D., 217, MNRAS, 471, 379 Huterer, D., & Shafer, D. L., 218, Reports on Progress in Physics, 81, 1691. Jiang, G. & Kochanek, C. S., 27, ApJ, 671, 1568. Jørgensen, I., Franx, M., & Kjaergaard, P., 1995, MNRAS, 276, 1341. Kochanek, C. S., 1992, ApJ, 384, 1. Koopmans, L. V. E. & Treu, T., 22, ApJ, 568, L5. Koopmans, L. V. E. & Treu, T., 23, ApJ, 583, 66. Koopmans, L. V. E., 26, EAS Publications Series, 2, 161. Koopmans, L. V. E., et al., 29, ApJ, 73, L51. Li, M., Li, X.-D., Wang, S., & Wang, Yi, 213, Frontiers of Physics, 8, 828. Li, Z., et al., 218a, ApJ, 854, 146. Li, R., Shu, Y., & Wang, J., 218b, MNRAS, 48, 431. Liao, K., et al., 217, Nature Communications, 8, 1148. Liddle, A. R., 24, MNRAS, 351, L49. Magueijo, J., & Sorkin, R. D. 27, MNRAS, 377, L39. Mehlert, D., Thomas, D., Saglia, R.P., Bender, R. & Wegner, G. 23, A&A, 47, 423. Mitchell, J. L., et al., 25, ApJ, 622, 81. Mukherjee, P., Parkinson, D., Corasaniti, P. S., Liddle, A. R., & Kunz, M. 26, MNRAS, 369, 1725. Ofek, E. O., Rix, H.-W., & Maoz, D., 23, MNRAS, 343, 639. Oguri, M., & Marshall, P. J., 218, MNRAS, 45, 2579. Paraficz, D. & Hjorth J., 29, A&A, 57, L49 Planck Collaboration, Ade, P. A. R., et al., 214, A&A, 571, A16. Planck Collaboration: Ade, P. A. R., et al., 216, A&A, 594, A13.

8 Chen et al. Planck Collaboration: Aghanim, N., et al, 218, preprint, (arxiv:187.629). Refsdal, S., 1964, MNRAS, 128, 37. Riess, A. G., et al., 211, ApJ, 73, 119. Riess, A. G., et al., 216, ApJ, 826, 56. Ruff, A. J., et al., 211, ApJ, 727, 96. Schwarz, G., 1978, The Annals of Statistics, 6, 461. Schwab, J., Bolton, A. S. & Rappaport, S. A., 21, ApJ, 78, 75. Scolnic, D. M., et al., 218, ApJ, 859, 11. Shu, Y., et al., 215, ApJ, 83, 71. Shu, Y., et al., 216a, ApJ, 824, 86. Shu, Y., et al., 216b, ApJ, 833, 264. Shu, Y., et al., et al., 217, ApJ, 851, 48. Shu, Y., et al., 218, preprint, (arxiv:183.7569) Sonnenfeld, A., et al., 213a, ApJ, 777, 98. Sonnenfeld, A., Gavazzi, R., Suyu, S. H., Treu, T., & Marshall, P. J., 213b, ApJ, 777, 97. Sonnenfeld, A., et al., 215, ApJ, 8, 94. Strauss, M. A., et al., 22, AJ, 124, 181. Suyu, S. H., et al., 217, MNRAS, 468, 259. Treu, T. & Koopmans, L. V. E., 22, ApJ, 575, 87. Treu, T. & Koopmans, L. V. E., 24, ApJ, 611, 739. Treu, T., & Marshall, P. J., 216, A&Arv, 24, 11. Treu, T, et al., 218, MNRAS, 481, 141. Turner, E. L., Ostriker, J. P., & Gott, J. R., III, 1984, ApJ, 284, 1. Wambsganss, J., 1998, Living Reviews in Relativity, 1, 12. Wang, N. & Xu, L., 213, MPLA, 28, 13557. Wei, J.-J. & Wu, X.-F., 217, MNRAS 472, 296. Wen, S., Wang, S., & Luo, X., JCAP, 7, 11. Xia, J.-Q., et al., 217, ApJ, 834, 75. Xu, D., et al., 217, MNRAS, 469, 1824.

9 Table 1: Observational data of the selected galaxy-scale SGL systems Lens Name z l z s θ E [ ] θ eff [ ] slit[ ] Fiber radius [ ] θ ap [ ] σ ap [km/s] Survey Name MG216+112 1.4 3.263 1.56.31 1. 1.25....65 34 ± 27 LSD 47 281.485 3.595 1.34.82.4 1.25....41 219 ± 12 LSD CFRS3.177.938 2.941 1.24 1.6.5 1.25....46 256 ± 19 LSD HST14176+5226.81 3.399 1.41 1.6.32 1.25....37 212 ± 18 LSD HSTT15433+5352.497 2.92.36.41.3 1.25....35 18 ± 14 LSD SL2SJ25 932.557 1.33.76.75.9 1.6....69 276 ± 37 SL2S SL2SJ212 555.75 2.74 1.27 1.22.9 1.6....69 273 ± 22 SL2S SL2SJ213 743.717 3.48 2.39 1.97 1. 1.68....75 293 ± 34 SL2S SL2SJ214 45.69 1.88 1.41 1.21 1. 1.88....79 287 ± 47 SL2S SL2SJ217 513.646 1.85 1.27.73 1.5 1.68....92 239 ± 27 SL2S SL2SJ218 82.884 2.6 1. 1.2.9 1.6....69 246 ± 48 SL2S SL2SJ2192 829.389 2.15 1.3.95 1. 1.68....75 289 ± 23 SL2S SL2SJ22 949.572 2.61 1..53 1. 1.9....8 254 ± 29 SL2S SL2SJ225 454.238 1.2 1.76 2.12 1..81....52 234 ± 21 SL2S SL2SJ226 42.494 1.23 1.19.84 1. 1.62....74 263 ± 24 SL2S SL2SJ232 48.352 2.34 1.4 1.14 1. 1.68....75 281 ± 26 SL2S SL2SJ848 351.682 1.55.85.45.9 1.6....69 197 ± 21 SL2S SL2SJ849 412.722 1.54 1.1.46.9 1.6....69 32 ± 24 SL2S SL2SJ849 251.274 2.9 1.16 1.34.9 1.6....69 276 ± 35 SL2S SL2SJ855 147.365 3.39 1.3.69.7 1.62....62 222 ± 25 SL2S SL2SJ94 59.611 2.36 1.4 2..9 1.6....69 183 ± 21 SL2S SL2SJ959+26.552 3.35.74.46.9 1.6....69 188 ± 22 SL2S SL2SJ1359+5535.783 2.77 1.14 1.13 1. 1.62....74 228 ± 29 SL2S SL2SJ144+52.456 1.59 2.55 2.3 1. 1.62....74 342 ± 2 SL2S SL2SJ145+5243.526 3.1 1.51.83 1. 1.62....74 284 ± 21 SL2S SL2SJ146+5226.716 1.47.94.8 1. 1.62....74 253 ± 19 SL2S SL2SJ1411+5651.322 1.42.93.85 1. 1.62....74 214 ± 23 SL2S SL2SJ142+563.483 3.12 1.4 1.62 1. 1.62....74 228 ± 19 SL2S SL2SJ223+25.4 2.15 1.95.99 1. 1.62....74 213 ± 21 SL2S SL2SJ225+147.476 2.53 1.66.66.9 1.6....69 317 ± 3 SL2S SL2SJ2221+115.325 2.35 1.4 1.12 1. 1.88....79 222 ± 23 SL2S SDSSJ37 942.195.632 1.53 2.3... 1.5 1.5 279 ± 1 SLACS SDSSJ44+113.12.197.8 2.83... 1.5 1.5 266 ± 13 SLACS SDSSJ216 813.332.523 1.16 2.76... 1.5 1.5 333 ± 23 SLACS SDSSJ252+39.28.982 1.4 1.34... 1.5 1.5 164 ± 12 SLACS SDSSJ33 2.351 1.71 1.1 1.26... 1.5 1.5 212 ± 21 SLACS SDSSJ728+3835.26.688 1.25 1.74... 1.5 1.5 214 ± 11 SLACS SDSSJ822+2652.241.594 1.17 2.1... 1.5 1.5 259 ± 15 SLACS SDSSJ912+29.164.324 1.63 4.15... 1.5 1.5 326 ± 12 SLACS SDSSJ936+913.19.588 1.9 2.21... 1.5 1.5 243 ± 11 SLACS SDSSJ946+16.222.69 1.38 2.54... 1.5 1.5 263 ± 21 SLACS SDSSJ956+51.241.47 1.33 2.25... 1.5 1.5 334 ± 15 SLACS SDSSJ959+41.126.535.99 1.48... 1.5 1.5 197 ± 13 SLACS SDSSJ959+4416.237.531.96 1.93... 1.5 1.5 244 ± 19 SLACS SDSSJ116+3859.168.439 1.9 1.53... 1.5 1.5 247 ± 13 SLACS SDSSJ12+1122.282.553 1.2 1.46... 1.5 1.5 282 ± 18 SLACS SDSSJ123+423.191.696 1.41 1.88... 1.5 1.5 242 ± 15 SLACS SDSSJ129+42.14.615 1.1 1.58... 1.5 1.5 21 ± 9 SLACS SDSSJ116+5228.95.47 1.23 2.5... 1.5 1.5 262 ± 9 SLACS SDSSJ1112+826.273.629 1.49 1.55... 1.5 1.5 32 ± 2 SLACS SDSSJ1134+627.153.474 1.1 1.98... 1.5 1.5 239 ± 11 SLACS SDSSJ1142+11.222.54.98 1.95... 1.5 1.5 221 ± 22 SLACS SDSSJ1143 144.16.42 1.68 5.28... 1.5 1.5 269 ± 5 SLACS SDSSJ1153+4612.18.875 1.5 1.32... 1.5 1.5 226 ± 15 SLACS SDSSJ124+358.164.631 1.31 1.63... 1.5 1.5 267 ± 17 SLACS SDSSJ125+491.215.481 1.22 2.59... 1.5 1.5 281 ± 13 SLACS SDSSJ1218+83.135.717 1.45 3.18... 1.5 1.5 219 ± 1 SLACS SDSSJ125+523.232.795 1.13 1.86... 1.5 1.5 252 ± 14 SLACS

1 Chen et al. Table 1 (continued) Lens Name z l z s θ E [ ] θ eff [ ] slit[ ] Fiber radius [ ] θ ap [ ] σ ap [km/s] Survey Name SDSSJ136+6.173.472 1.32 2.8... 1.5 1.5 237 ± 17 SLACS SDSSJ1313+4615.185.514 1.37 2.1... 1.5 1.5 263 ± 18 SLACS SDSSJ1318 313.24 1.3 1.58 3.7... 1.5 1.5 213 ± 18 SLACS SDSSJ133 148.81.711.86.91... 1.5 1.5 185 ± 9 SLACS SDSSJ142+6321.25.481 1.35 2.65... 1.5 1.5 267 ± 17 SLACS SDSSJ143+6.189.473.83 1.62... 1.5 1.5 213 ± 17 SLACS SDSSJ1416+5136.299.811 1.37 1.33... 1.5 1.5 24 ± 25 SLACS SDSSJ142+619.63.535 1.4 2.11... 1.5 1.5 25 ± 4 SLACS SDSSJ143+415.285.575 1.52 2.42... 1.5 1.5 322 ± 32 SLACS SDSSJ1436.285.85 1.12 2.41... 1.5 1.5 224 ± 17 SLACS SDSSJ1443+34.134.419.81 1.11... 1.5 1.5 29 ± 11 SLACS SDSSJ1451 239.125.52 1.4 2.6... 1.5 1.5 223 ± 14 SLACS SDSSJ1525+3327.358.717 1.31 2.82... 1.5 1.5 264 ± 26 SLACS SDSSJ1531 15.16.744 1.71 2.73... 1.5 1.5 279 ± 12 SLACS SDSSJ1538+5817.143.531 1. 1.45... 1.5 1.5 189 ± 12 SLACS SDSSJ1621+3931.245.62 1.29 2.3... 1.5 1.5 236 ± 2 SLACS SDSSJ1627 53.28.524 1.23 2.2... 1.5 1.5 29 ± 14 SLACS SDSSJ163+452.248.793 1.78 2.1... 1.5 1.5 276 ± 16 SLACS SDSSJ1636+477.228.675 1.8 1.62... 1.5 1.5 231 ± 15 SLACS SDSSJ1644+2625.137.61 1.27 1.85... 1.5 1.5 229 ± 12 SLACS SDSSJ1719+2939.181.578 1.28 1.42... 1.5 1.5 286 ± 15 SLACS SDSSJ2238 754.137.713 1.27 2.39... 1.5 1.5 198 ± 11 SLACS SDSSJ23+22.228.463 1.24 1.88... 1.5 1.5 279 ± 17 SLACS SDSSJ233+1422.155.517 1.62 3.46... 1.5 1.5 255 ± 16 SLACS SDSSJ2321 939.82.532 1.6 4.22... 1.5 1.5 249 ± 8 SLACS SDSSJ2341+.186.87 1.44 3.4... 1.5 1.5 27 ± 13 SLACS SDSSJ143 16.221 1.15 1.23 2.18... 1.5 1.5 216 ± 15 S4TM SDSSJ159 6.158.748.92.67... 1.5 1.5 233 ± 13 S4TM SDSSJ324 11.446.624.63 1.21... 1.5 1.5 235 ± 26 S4TM SDSSJ324+45.321.92.55 1.24... 1.5 1.5 145 ± 18 S4TM SDSSJ753+3416.137.963 1.24 1.84... 1.5 1.5 22 ± 9 S4TM SDSSJ754+1927.153.74 1.5 2.71... 1.5 1.5 187 ± 12 S4TM SDSSJ757+1956.121.833 1.92 1.52... 1.5 1.5 234 ± 9 S4TM SDSSJ826+563.132 1.291 1.2 1.75... 1.5 1.5 151 ± 6 S4TM SDSSJ847+2348.155.533.96 1.34... 1.5 1.5 2 ± 12 S4TM SDSSJ851+55.128.637.9 1.27... 1.5 1.5 175 ± 9 S4TM SDSSJ92+328.288.392.7 1.72... 1.5 1.5 38 ± 13 S4TM SDSSJ955+314.321.467.53 2.7... 1.5 1.5 273 ± 23 S4TM SDSSJ956+5539.196.848 1.17 2.41... 1.5 1.5 196 ± 8 S4TM SDSSJ11+3124.167.425 1.13 2.96... 1.5 1.5 235 ± 9 S4TM SDSSJ141+112.11.217.6 2.13... 1.5 1.5 197 ± 6 S4TM SDSSJ148+1313.133.668 1.2 2.97... 1.5 1.5 22 ± 7 S4TM SDSSJ151+4439.163.538.99 2.73... 1.5 1.5 28 ± 12 S4TM SDSSJ156+4141.134.832.73 1.1... 1.5 1.5 145 ± 8 S4TM SDSSJ111+1523.178.517 1.18.74... 1.5 1.5 295 ± 12 S4TM SDSSJ1116+729.17.686.82 4.72... 1.5 1.5 177 ± 9 S4TM SDSSJ1127+2312.13.361 1.26 4.49... 1.5 1.5 244 ± 7 S4TM SDSSJ1137+1818.124.463 1.29 1.57... 1.5 1.5 223 ± 6 S4TM SDSSJ1142+259.164.659.79 1.6... 1.5 1.5 158 ± 7 S4TM SDSSJ1144+436.14.255.76 1.75... 1.5 1.5 24 ± 12 S4TM SDSSJ1213+293.91.595 1.35 1.92... 1.5 1.5 236 ± 6 S4TM SDSSJ131+834.9.533 1. 4.74... 1.5 1.5 185 ± 6 S4TM SDSSJ133+175.27.372 1.2 1.9... 1.5 1.5 263 ± 9 S4TM SDSSJ143+339.62.772 1.3 3.55... 1.5 1.5 27 ± 5 S4TM SDSSJ143+614.169.654 1. 5.87... 1.5 1.5 23 ± 13 S4TM SDSSJ1433+2835.91.411 1.53 2.7... 1.5 1.5 238 ± 5 S4TM SDSSJ1541+3642.141.739 1.15 1.41... 1.5 1.5 183 ± 8 S4TM SDSSJ1543+222.268.397.78 1.71... 1.5 1.5 294 ± 14 S4TM