Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: docmute
  • failed: docmute

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: CC BY 4.0
arXiv:2304.11177v2 [astro-ph.CO] 03 Jan 2024

An analytic surface density profile for 𝚲𝚲\Lambdabold_ΛCDM haloes and gravitational lensing studies

Alexandres Lazar1,212{}^{1,2}start_FLOATSUPERSCRIPT 1 , 2 end_FLOATSUPERSCRIPT, James S. Bullock11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT, Anna Nierenberg33{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPT, Leonidas A. Moustakas22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPT, and Michael Boylan-Kolchin44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPT
11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPTDepartment of Physics and Astronomy, University of California, Irvine, CA 92697 USA
22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPTJet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Dr., Pasadena CA 91109, USA
33{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPTDepartment of Physics, University of California Merced, 5200 North Lake Rd. Merced, CA 95343, USA
44{}^{4}start_FLOATSUPERSCRIPT 4 end_FLOATSUPERSCRIPTDepartment of Astronomy, The University of Texas at Austin, 2515 Speedway, Stop C1400, Austin, Texas 78712-1205, USA
aalazar@uci.edu
(Working Draft)
Abstract

We introduce an analytic surface density profile for dark matter haloes that accurately reproduces the structure of simulated haloes of mass Mvir=10711Msubscript𝑀virsuperscript10711subscript𝑀direct-productM_{\rm vir}=10^{7-11}\ M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 7 - 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, making it useful for modeling line-of-sight perturbers in strong gravitational lensing models. The two-parameter function has an analytic deflection potential and is more accurate than the projected Navarro, Frenk & White (NFW) profile commonly adopted at this mass scale for perturbers, especially at the small radii of most relevant for lensing perturbations. Using a characteristic radius, R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, where the log slope of surface density is equal to 11-1- 1, and an associated surface density, Σ1subscriptΣ1\Sigma_{-1}roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, we can represent the expected lensing signal from line-of-sight halos statistically, for an ensemble of halo orientations, using a distribution of projected concentration parameters, 𝒞vir:=rvir/R1assignsubscript𝒞virsubscript𝑟virsubscript𝑅1\mathcal{C}_{\rm vir}:=r_{\rm vir}/R_{-1}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT := italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT. Though an individual halo can have a projected concentration that varies with orientation with respect to the observer, the range of projected concentrations correlates with the usual three-dimensional halo concentration in a way that enables ease of use.

keywords:
cosmology:theory – dark matter – gravitational lensing: strong
pubyear: 2021pagerange: An analytic surface density profile for 𝚲𝚲\Lambdabold_ΛCDM haloes and gravitational lensing studiesC

1 Introduction

The ΛΛ\Lambdaroman_ΛCDM (the cosmological constant + cold dark matter) cosmogony has served as the benchmark model for decades. A major component of its success is in matching the large-scale structure of the Universe, which also places principal constraints on the Universe’s composition (e.g. Davis et al., 1985; Geller & Huchra, 1989; Bond et al., 1996; Tegmark et al., 2004; Sánchez et al., 2006; Weinberg et al., 2013). A key prediction of ΛΛ\Lambdaroman_ΛCDM is that the Universe is lavished with a high number density of very low-mass ( Mhalo109Mless-than-or-similar-tosubscript𝑀halosuperscript109subscript𝑀direct-productM_{\rm halo}\lesssim 10^{9}\ M_{\odot}italic_M start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) dark matter halos that can act as perturbers in lensing studies (Press & Schechter, 1974; Metcalf & Madau, 2001; Green et al., 2005; Diemand et al., 2007; Springel et al., 2008; Frenk & White, 2012). The existence of very low mass, nearly starless, dark halos of the kind and character predicted have yet to be confirmed by observations and may only be detectable via their gravitational effects e.g. on strong gravitational lenses.

While the cosmological model of ΛΛ\Lambdaroman_ΛCDM has been successful at matching large-scale observations, ΛΛ\Lambdaroman_ΛCDM still suffers from discrepancies found at small scales (see Bullock & Boylan-Kolchin 2017 for a comprehensive overview) and this has motivated alternative dark matter models. One such model is warm dark matter (WDM), which suppresses the matter power spectrum of initial density perturbations at scales smaller than the free-streaming length (Colín et al., 2000; Bode et al., 2001; Lovell et al., 2012; Schneider et al., 2012). As an example, a 7 keV sterile neutrino of the type that could be responsible for the observed 3.5 keV line in galaxy cluster X-ray spectra (e.g. Boyarsky et al., 2014; Bulbul et al., 2014), would produce a sharp cutoff in the abundance of halos smaller than Mhalo108Mless-than-or-similar-tosubscript𝑀halosuperscript108subscript𝑀direct-productM_{\rm halo}\lesssim 10^{8}\ M_{\odot}italic_M start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT ≲ 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. Demonstrating the existence of halos below this mass could rule out this class of WDM; conversely, the ability to rule out such a population would eliminate ΛΛ\Lambdaroman_ΛCDM as a complete model of cosmology. One other prediction for galaxies in a WDM cosmology is that they form with lower central densities compared to ΛΛ\Lambdaroman_ΛCDM (Lovell et al., 2014). It is also possible that dark matter particles interact among themselves within a hidden sector (Feng et al., 2009). If self-interaction cross-sections are high enough, low-mass dark matter halos will have lower central densities compared to those of ΛΛ\Lambdaroman_ΛCDM (Rocha et al., 2013; Elbert et al., 2015; Tulin & Yu, 2018). For dark matter halos of a given mass, all of these alternative cosmological paradigms make distinct predictions for both the abundance and central densities compared to expectations of ΛΛ\Lambdaroman_ΛCDM.

Regardless of the cosmological model, a common characteristic of low-mass dark matter halos is that they are extremely inefficient at forming stars and are nearly devoid of baryons (e.g. Sawala et al., 2016; Benitez-Llambay & Frenk, 2020). This results in these objects being too difficult to detect electromagnetically. A more promising route for their detection is via gravitational perturbations of strongly lensed images (Dalal & Kochanek, 2002; Moustakas & Metcalf, 2003; Koopmans, 2005; Vegetti & Koopmans, 2009a, b; Vegetti et al., 2010, 2012; Hezaveh et al., 2016; Nierenberg et al., 2020; Gilman et al., 2020b; Hsueh et al., 2020). To fully constrain the nature of dark matter using strong lensing, it is necessary to know the expected number of halos in ΛΛ\Lambdaroman_ΛCDM and other cosmology models. It is convenient to characterize low-mass halo perturbers as either subhalos embedded within the main lens halo or field halos that are not part of the main lens but rather lie in projection near the lens’ Einstein radius. In the strong lensing literature, this population of field halos are typically called “line-of-sight” (LOS) halos (or LOS perturbers). In Metcalf (2005), they showed that the LOS component contributes significantly to strongly lensed signals within ΛΛ\Lambdaroman_ΛCDM. Li et al. (2017) later found that that LOS halos dominated the subhalo signal by a factor of 3-4. A similar result was found in Despali et al. (2018), where the number of LOS perturbers compared to subhalo perturbers ranges from 3 to 10 times depending on the lens configuration. Similar results are found in Gilman et al. (2019), He et al. (2022b), and Lazar et al. (2021).

The internal structure of dark matter halo perturbers can greatly impact their lensing effect. Specifically, in order to properly constrain the mass function predicted by ΛΛ\Lambdaroman_ΛCDM, the surface density profiles of the dark matter halos of interest must be precisely known (e.g., Minor et al., 2017; Minor et al., 2021). A common way to characterize low-mass dark matter halo structure for perturbing, field (LOS) halos is to assume that they follow spherically-symmetric NFW profiles (Navarro et al., 1997; Nierenberg et al., 2017; Gilman et al., 2018; He et al., 2022a). With this assumption, the expected distribution of NFW concentration parameters at fixed mass becomes an important input for models (e.g. Gilman et al., 2022).

The spherical NFW profile is simpler analytically, which has an advantage that enables direct connection with theoretical predictions, especially for low-mass halos where feedback is believed to be less important in altering the density structure compared to dark-matter-only predictions (e.g. Lazar et al., 2020). Additionally, the NFW form also has a fairly easy-to-use surface-density profile for lensing-based analysis (Wright & Brainerd, 2000). However, there are some potential shortcomings in this approach. One is that dark matter halos are not perfectly spherical and tend to be more triaxial (e.g. Frenk et al., 1988; Cole & Lacey, 1996; Allgood et al., 2006). This means that, even for a halo with a spherically-averaged profile that is well described by an NFW fit, its surface density could well be different from the two-dimensional projection inferred by the spherical-average fit depending on its orientation. Second, it is well known that dark-matter-only simulations produce halos that are better modeled using the Einasto (1965) profile than NFW when the particle resolution is increased (Wang et al., 2020). However, transforming that equation into the projected two-dimensional density and the lensing signal results in a relatively complicated expression (Retana-Montenegro et al., 2012) which can be better approximated (Dhar & Williams, 2010; Dhar, 2021).

The aim of this work is to provide an easy-to-use surface-density profile that accurately reproduces the structure of simulated dark matter halos in any projection. Below we show with regression analysis that the projected density profiles of simulated dark matter halos deviate from the projected NFW profile inferred from their three-dimensional, spherical NFW fits (i.e., their best-fit fit concentrations and normalization) in ways that do not average out over all orientations. This is particularly true at projected radii smaller than the scale radius. Since the NFW form is commonly used for substructure lensing analysis, this motivates the exploration of an alternative approach that is just as easy to use and more accurate. The projected profile presented below has a scale radius and a corresponding “projected concentration.” This projected concentration can account both for the fact that individual halos have different projected density profiles depending on their orientation with respect to the observer and that, at fixed halo mass, there is an intrinsic halo-to-halo (spherical) concentration scatter. While this paper focuses on circularly-averaged projected densities on the sky, the functional form can be easily generalized to elliptically-averaged profiles with one additional parameter, which would be a natural extension of this work. Note, however, that the corresponding extension for the lensing potential, reflection, and shear will be less straightforward (see, e.g. Tessore & Metcalf, 2015; O’Riordan et al., 2021).

This paper is structured as follows. Section 2 introduces the suite of high-resolution simulations of dark matter halos used, provides a description of the selected sample of halos, and details the methods of constructing the radial dark matter profiles of each sampled halo. Section 3 provides the main results of this paper and introduces our simple fitting formula for LOS perturbers for lensing analysis. We discuss the implications for our fitting formula in Section 4 and demonstrate its impact to the commonly used LOS perturber formula, the NFW profile. Finally, we summarize out results in Section 5.

2 Methodology

We use “zoom-in” dark matter only (DMO) simulations to study halo surface density structure at high resolution, focusing on isolated dark matter halos within the mass range that is applicable to substructure lensing, Mhalo10711Msubscript𝑀halosuperscript10711subscript𝑀direct-productM_{\rm halo}\in 10^{7-11}\,M_{\odot}italic_M start_POSTSUBSCRIPT roman_halo end_POSTSUBSCRIPT ∈ 10 start_POSTSUPERSCRIPT 7 - 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, at a characteristic redshift of z=0.2𝑧0.2z=0.2italic_z = 0.2. We use these simulations to discover an accurate surface density profile shape and to compare the shape parameters to the three-dimensional profile parameters of the same halos. We plan to follow-up with an analysis that utilizes a high-resolution cosmological environment (i.e., a cosmological box) to put the results presented here in a proper statistical setting and to explore redshift dependence.

2.1 Numerical zoom-in simulations

All of our simulations use the multi-method code GIZMO (Hopkins, 2015). The initial conditions were calculated using MUSIC (Hahn & Abel, 2011) at a redshift z100𝑧100z\approx 100italic_z ≈ 100 following the methodology outlined in Oñorbe et al. (2014). This approach identifies a region spanning several virial radii around a main halo to resolve and produces a volume of uncontaminated halos around the main halo, which we use to study lower mass systems.

The first set of volumes we study come from the DMO versions of the following main halos from Fitts et al. (2017), which assume a WMAP year 7 cosmology: m10f, m10g, m10h, m10i, m10j, m10k, m10l, and m10m; these will be collectively referred to as the “m10” suite of simulations. These simulations have a dark matter particle mass of mdm=3000Msubscript𝑚dm3000subscript𝑀direct-productm_{\rm dm}=3000\ M_{\odot}italic_m start_POSTSUBSCRIPT roman_dm end_POSTSUBSCRIPT = 3000 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT with a physical force resolution of ϵdm=35pcsubscriptitalic-ϵdm35pc\epsilon_{\rm dm}=35\ \rm pcitalic_ϵ start_POSTSUBSCRIPT roman_dm end_POSTSUBSCRIPT = 35 roman_pc. At this resolution, we are able to also explore dark matter halos surrounding the main halo down to masses of 107Msuperscript107subscript𝑀direct-product10^{7}\ M_{\odot}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, which contains 104similar-toabsentsuperscript104{\sim}10^{4}∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT particles within the virial radius. In addition, we use several of the dark matter only volumes surrounding main halos mass of 1011Msimilar-toabsentsuperscript1011subscript𝑀direct-product\sim 10^{11}\,M_{\odot}∼ 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, which were first presented in Lazar et al. (2020): m11d, m11e, m11h, and m11i; these will be collectively referred to as the “m11” suite of simulations. The m11 simulations have a mass resolution similar-to\sim 14 times coarser than the m10 suite. Within the m11 volumes, we only explore the main halo with Mvir1011Msimilar-tosubscript𝑀virsuperscript1011subscript𝑀direct-productM_{\rm vir}\sim 10^{11}\,M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. These simulations use a Planck 2015 cosmology (Planck Collaboration et al., 2016).

Dark matter halos in all the simulations considered here are identified using the Amiga halo finder (AHF; Knollmann & Knebe 2009). The halo finder uses a recursively refined grid that determines the local overdensities found within the density field and identifies the density peaks of this field as the center of these halos.

2.2 Dark matter halo nomenclature

In this section we provide definitions of the global and physical quantities of dark matter halos we will be studying in this work. Throughout this paper, lower case r𝑟ritalic_r denotes the physical, de-projected, three-dimensional radius. The projected, two-dimensional radius is denoted by an upper case R𝑅Ritalic_R without a subscript.

2.2.1 Halo mass and radius

Throughout this paper, dark matter halos are defined to be spherical systems with virial mass, Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, with a virial radius, rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, inside of which the average density is equal to the critical density times a multiplicative factor Δvir(z)subscriptΔvir𝑧\Delta_{\rm vir}(z)roman_Δ start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ( italic_z ), i.e.,

Mvir=4π3rvir3Δvir(z)ρc(z).subscript𝑀vir4𝜋3superscriptsubscript𝑟vir3subscriptΔvir𝑧subscript𝜌c𝑧\displaystyle M_{\rm vir}=\frac{4\pi}{3}r_{\rm vir}^{3}\,\Delta_{\rm vir}(z)\,% \rho_{\rm c}(z)\,.italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = divide start_ARG 4 italic_π end_ARG start_ARG 3 end_ARG italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_Δ start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ( italic_z ) italic_ρ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ( italic_z ) . (1)

Here, ΔvirsubscriptΔvir\Delta_{\rm vir}roman_Δ start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT is redshift-dependent virial overdensity (Bryan & Norman, 1998) and ρc(z)=3H2(z)/8πGsubscript𝜌c𝑧3superscript𝐻2𝑧8𝜋𝐺\rho_{\rm c}(z)=3H^{2}(z)/8\pi Gitalic_ρ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ( italic_z ) = 3 italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_z ) / 8 italic_π italic_G is the critical density of the universe at redshift z𝑧zitalic_z. The Bryan & Norman (1998) definition is the primary spherical overdensity definition used for the AHF halo finding, but is computed using the gravitationally bound particles of the system. For the purposes of this analysis, Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT and rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT are recomputed using all dark matter particles (bound and unbound) of the spherical overdensity peak identified by AHF. It should be noted that isolated dark matter halos are also commonly defined with overdensity Δ=200Δ200\Delta=200roman_Δ = 200 with either a critical or mean-matter background densities. We do not explore such definitions other than Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT in this analysis, but will do so in a follow up paper.

2.2.2 Three-dimensional structure of dark matter halos

The three-dimensional density profiles of dark matter halos in ΛΛ\Lambdaroman_ΛCDM are commonly modeled with the Navarro, Frenk & White (1997) (NFW) profile,

ρNFW(r)=ρs(rrs)1(1+rrs)2,subscript𝜌NFW𝑟subscript𝜌𝑠superscript𝑟subscript𝑟𝑠1superscript1𝑟subscript𝑟𝑠2\displaystyle\rho_{\rm NFW}(r)=\rho_{s}\left(\frac{r}{r_{s}}\right)^{-1}\left(% 1+\frac{r}{r_{s}}\right)^{-2}\,,italic_ρ start_POSTSUBSCRIPT roman_NFW end_POSTSUBSCRIPT ( italic_r ) = italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( divide start_ARG italic_r end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( 1 + divide start_ARG italic_r end_ARG start_ARG italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT , (2)

where rssubscript𝑟𝑠r_{s}italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the scale radius and ρssubscript𝜌𝑠\rho_{s}italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is the characteristic density. The structure of a NFW halo is parameterized by the concentration (cNFWsubscript𝑐NFWc_{\rm NFW}italic_c start_POSTSUBSCRIPT roman_NFW end_POSTSUBSCRIPT) of a dark matter halo, which is formally defined as the ratio between the size of the halo, rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, and the scale radius, rssubscript𝑟𝑠r_{s}italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT:

cNFW:=rvir/rs.assignsubscript𝑐NFWsubscript𝑟virsubscript𝑟𝑠\displaystyle c_{\rm NFW}:=r_{\rm vir}/r_{s}\,.italic_c start_POSTSUBSCRIPT roman_NFW end_POSTSUBSCRIPT := italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT / italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT . (3)

In DMO simulations, dark matter halos are better described by the Einasto profile (Navarro et al., 2004, 2010), which provides a good description over 20 decades of halo masses (Wang et al., 2020):

ρϵ(r)subscript𝜌italic-ϵ𝑟\displaystyle\rho_{\epsilon}(r)italic_ρ start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT ( italic_r ) =ρ2exp{2αϵ[(rr2)αϵ1]}.absentsubscript𝜌22subscript𝛼italic-ϵdelimited-[]superscript𝑟subscript𝑟2subscript𝛼italic-ϵ1\displaystyle=\rho_{-2}\exp\Bigg{\{}-\frac{2}{\alpha_{\epsilon}}\Bigg{[}\Bigg{% (}\frac{r}{r_{-2}}\Bigg{)}^{\alpha_{\epsilon}}-1\Bigg{]}\Bigg{\}}\,.= italic_ρ start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT roman_exp { - divide start_ARG 2 end_ARG start_ARG italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT end_ARG [ ( divide start_ARG italic_r end_ARG start_ARG italic_r start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT - 1 ] } . (4)

Here, r2subscript𝑟2r_{-2}italic_r start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT is the scale radius where the log-slope is equal to 22-2- 2, the scale density is ρ2:=ρ(r2)assignsubscript𝜌2𝜌subscript𝑟2\rho_{-2}:=\rho(r_{-2})italic_ρ start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT := italic_ρ ( italic_r start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT ), and αϵsubscript𝛼italic-ϵ\alpha_{\epsilon}italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT is the shape parameter. Usually αϵsubscript𝛼italic-ϵ\alpha_{\epsilon}italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT is fixed to make this a two-parameter function, but it can be expressed in terms of peak height (e.g. Gao et al., 2008; Prada et al., 2012; Klypin et al., 2016; Child et al., 2018). We fix αϵ=0.17subscript𝛼italic-ϵ0.17\alpha_{\epsilon}=0.17italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = 0.17 for this analysis, which fits well for the range of halo masses we explore here and is close to commonly-adopted values (e.g. Wang et al., 2020). The internal structure from the Einasto profile is parameterized by the dimensionless concentration parameter:

cvir:=rvir/r2.assignsubscript𝑐virsubscript𝑟virsubscript𝑟2\displaystyle c_{\rm vir}:=r_{\rm vir}/r_{-2}\,.italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT := italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT / italic_r start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT . (5)

Note that rssubscript𝑟𝑠r_{s}italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT in the NFW profile is also the radius where the log slope of the NFW form is 22-2- 2, so typically cvircNFWsimilar-to-or-equalssubscript𝑐virsubscript𝑐NFWc_{\rm vir}\simeq c_{\rm NFW}italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ≃ italic_c start_POSTSUBSCRIPT roman_NFW end_POSTSUBSCRIPT for any individual halo.

2.2.3 Two-dimensional structure of dark matter halos

Dark matter halos in substructure lensing are modeled as two-dimensional surface density profiles, as these coincide with two-dimensional, face-on projections of observations viewed from the surface of the sky. For a given system that has a (intrinsic) spherically-averaged local density profile in three-dimensions, ρ(r)𝜌𝑟\rho(r)italic_ρ ( italic_r ), the cylindricaly-averaged, local surface density profile, Σρ(R)subscriptΣ𝜌𝑅\Sigma_{\rho}(R)roman_Σ start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT ( italic_R ), is quantified by integrating the along the line of sight, \ellroman_ℓ, for a path of some depth \mathcal{L}caligraphic_L

Σρ(R)subscriptΣ𝜌𝑅\displaystyle\Sigma_{\rho}(R)roman_Σ start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT ( italic_R ) =/2/2𝑑ρ(R,)absentsubscriptsuperscript22differential-d𝜌𝑅\displaystyle=\int^{\mathcal{L}/2}_{-\mathcal{L}/2}d\ell\,\rho(R,\ell)= ∫ start_POSTSUPERSCRIPT caligraphic_L / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - caligraphic_L / 2 end_POSTSUBSCRIPT italic_d roman_ℓ italic_ρ ( italic_R , roman_ℓ ) (6)
=2R/2𝑑rrρ(r)r2R2,absent2subscriptsuperscript2𝑅differential-d𝑟𝑟𝜌𝑟superscript𝑟2superscript𝑅2\displaystyle=2\int^{\mathcal{L}/2}_{R}dr\,\frac{r\rho(r)}{\sqrt{r^{2}-R^{2}}}\,,= 2 ∫ start_POSTSUPERSCRIPT caligraphic_L / 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_R end_POSTSUBSCRIPT italic_d italic_r divide start_ARG italic_r italic_ρ ( italic_r ) end_ARG start_ARG square-root start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG , (7)

where R𝑅Ritalic_R denotes the projected radius relative to the center of the halo and =r2R2superscript𝑟2superscript𝑅2\ell=\sqrt{r^{2}-R^{2}}roman_ℓ = square-root start_ARG italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG is a coordinate along the line of sight to the observer. In what follows we explore choices for the path length of integration in constructing numerical surface density profiles of the form =2ξrvir2𝜉subscript𝑟vir\mathcal{L}=2\,\xi\,r_{\rm vir}caligraphic_L = 2 italic_ξ italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. As described below, we settle on ξ=1.5𝜉1.5\xi=1.5italic_ξ = 1.5 as an optimal choice.

In lensing studies, it is common to use an NFW profile in the limit where =\mathcal{L}=\inftycaligraphic_L = ∞ via a forward Abel transformation. In this case, the surface density profile is analytic (Wright & Brainerd, 2000):

ΣNFW(x)={2ρsrs(x21)[12x21arctanx11+x]x>12ρsrs3x=12ρsrs(x21)[121x2arctanh1x1+x]x<1,subscriptΣNFW𝑥cases2subscript𝜌𝑠subscript𝑟𝑠superscript𝑥21delimited-[]12superscript𝑥21𝑥11𝑥𝑥12subscript𝜌𝑠subscript𝑟𝑠3𝑥12subscript𝜌𝑠subscript𝑟𝑠superscript𝑥21delimited-[]121superscript𝑥2arctanh1𝑥1𝑥𝑥1\displaystyle\Sigma_{\rm NFW}(x)=\begin{cases}\ \frac{2\,\rho_{s}r_{s}}{(x^{2}% -1)}\left[1-\frac{2}{\sqrt{x^{2}-1}}\arctan\sqrt{\frac{x-1}{1+x}}\right]&x>1\\ \ \frac{2\,\rho_{s}r_{s}}{3}&x=1\\ \ \frac{2\,\rho_{s}r_{s}}{(x^{2}-1)}\left[1-\frac{2}{\sqrt{1-x^{2}}}\mathrm{% arctanh}\sqrt{\frac{1-x}{1+x}}\right]&x<1\end{cases}\,,roman_Σ start_POSTSUBSCRIPT roman_NFW end_POSTSUBSCRIPT ( italic_x ) = { start_ROW start_CELL divide start_ARG 2 italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG ( italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) end_ARG [ 1 - divide start_ARG 2 end_ARG start_ARG square-root start_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 end_ARG end_ARG roman_arctan square-root start_ARG divide start_ARG italic_x - 1 end_ARG start_ARG 1 + italic_x end_ARG end_ARG ] end_CELL start_CELL italic_x > 1 end_CELL end_ROW start_ROW start_CELL divide start_ARG 2 italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG 3 end_ARG end_CELL start_CELL italic_x = 1 end_CELL end_ROW start_ROW start_CELL divide start_ARG 2 italic_ρ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG ( italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 ) end_ARG [ 1 - divide start_ARG 2 end_ARG start_ARG square-root start_ARG 1 - italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG roman_arctanh square-root start_ARG divide start_ARG 1 - italic_x end_ARG start_ARG 1 + italic_x end_ARG end_ARG ] end_CELL start_CELL italic_x < 1 end_CELL end_ROW , (8)

with x=R/rs𝑥𝑅subscript𝑟𝑠x=R/r_{s}italic_x = italic_R / italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. While the Einasto profile describes the spherically averaged distribution of high-resolution dark matter halos to better accuracy than NFW (see Figure 1 below), the analytical lens properties are significantly more complicated to work with (see Retana-Montenegro et al., 2012). For this reason, the NFW profile is most commonly adopted in lensing studies. In later sections we present an analytic surface density profile that describes projected dark matter halo structure with a similar accuracy as the Einasto profile does in three-dimensions, with the added feature that it has an easy-to-use, analytic lensing potential.

2.2.4 Non-spherical dark matter halos

Another important aspect of dark matter halos is their non-spherical shape. ΛΛ\Lambdaroman_ΛCDM halos are triaxial and tend to be more elongated at higher halo masses. Dark matter halos viewed along their major axis (i.e., the densest axis) could be misidentified as a higher mass halo or a halo of higher concentration. To explore the magnitude of this effect, we calculate dark matter halo shapes for our halos by computing the shape inertia tensor outlined in Allgood et al. (2006). This is done by solving the eigenvalues from all of the dark matter particles within a shell between 10 - 20% of rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. The resulting eigenvalues of the shape tensor are proportional to the square root of the principal axes of the dark matter distribution, which we will refer to as the “major”, “intermediate”, and “minor” axes throughout this paper.

Note that the triaxial shape of DMO halos can be taken into account in lensing studies by projecting a triaxial NFW profile directly (e.g., Feroz & Hobson, 2012). Doing so requires six parameters rather than two for each halo. In what follows, we present a direct fit to the projected, cylindricaly-averaged, profiles that provide improved accuracy over the projected NFW with the same number of free parameters.

Refer to caption
Figure 1: The spherically averaged density profiles, ρ(r)𝜌𝑟\rho(r)italic_ρ ( italic_r ), of the simulated dark matter halos at z=0.2𝑧0.2z=0.2italic_z = 0.2. The top panels present several density profiles of the isolated dark matter halos (dashed curves) in mass groups spanning Mvir=10711Msubscript𝑀virsuperscript10711subscript𝑀direct-productM_{\rm vir}=10^{7-11}\ M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 7 - 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. The density profiles are scaled by r2superscript𝑟2r^{2}italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to set apart the dynamic range and the x𝑥xitalic_x-axis is scaled by rssubscript𝑟𝑠r_{s}italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT of the best-fit NFW profile (left panels) and r2subscript𝑟2r_{-2}italic_r start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT of the best-fit Einasto profile (right panels). The simulated profiles are plotted from the convergence radius, rconvsubscript𝑟convr_{\rm conv}italic_r start_POSTSUBSCRIPT roman_conv end_POSTSUBSCRIPT out to rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. The thin solid lines illustrates the best-fit halo density profile of each halo for the NFW profile (left panel) and the Einasto profile (right panel). The bottom panels depicts the quality of fit based on the best-fit parameters for each individual halo using the NFW and Einasto density profiles. The gray dashed lines encapsulates fits within 20% accuracy.

2.3 Constructing and fitting radial profiles

2.3.1 Region of numerical convergence

One of the key components of our analysis focuses on the spherically (and cylindrically) averaged mass-density profiles. Before radial bins are constructed, dark matter particles are first shifted relative to the halo center determined by AHF. The innermost regions of N𝑁Nitalic_N-body simulated dark matter halos, to some extent, are impacted by numerical relaxation. The region of convergence, rconvsubscript𝑟convr_{\rm conv}italic_r start_POSTSUBSCRIPT roman_conv end_POSTSUBSCRIPT, is quantified using the method specified in Power et al. (2003), where the effective resolution of the simulations dictates the location of the radius where the two-body relaxation timescale, trelaxsubscript𝑡relaxt_{\rm relax}italic_t start_POSTSUBSCRIPT roman_relax end_POSTSUBSCRIPT, becomes shorter than the age of the universe, t0subscript𝑡0t_{0}italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, set by the criterion:

trelax(r)t0=2008N(<r)lnN(<r)[ρ¯(<r)ρc(z)]1/2.subscript𝑡relax𝑟subscript𝑡02008annotated𝑁absent𝑟annotated𝑁absent𝑟superscriptdelimited-[]annotated¯𝜌absent𝑟subscript𝜌c𝑧12\displaystyle\frac{t_{\rm relax}(r)}{t_{0}}=\frac{\sqrt{200}}{8}\frac{N(<r)}{% \ln N(<r)}\left[\frac{\bar{\rho}(<r)}{\rho_{\rm c}(z)}\right]^{-1/2}\,.divide start_ARG italic_t start_POSTSUBSCRIPT roman_relax end_POSTSUBSCRIPT ( italic_r ) end_ARG start_ARG italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG = divide start_ARG square-root start_ARG 200 end_ARG end_ARG start_ARG 8 end_ARG divide start_ARG italic_N ( < italic_r ) end_ARG start_ARG roman_ln italic_N ( < italic_r ) end_ARG [ divide start_ARG over¯ start_ARG italic_ρ end_ARG ( < italic_r ) end_ARG start_ARG italic_ρ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT ( italic_z ) end_ARG ] start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT . (9)

Here, N(<r)annotated𝑁absent𝑟N(<r)italic_N ( < italic_r ) is the cumulative number of particles within radius r𝑟ritalic_r and the cumulative density profile, ρ¯(<r)=3M(<r)/4πr3\bar{\rho}(<r)=3M(<r)/4\pi r^{3}over¯ start_ARG italic_ρ end_ARG ( < italic_r ) = 3 italic_M ( < italic_r ) / 4 italic_π italic_r start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, with M(<r)annotated𝑀absent𝑟M(<r)italic_M ( < italic_r ) being the cumulative mass. For N𝑁Nitalic_N-body simulations of our resolution, convergence is shown to be well resolved to the radius at which the criterion satisfies trelax>0.6t0subscript𝑡relax0.6subscript𝑡0t_{\rm relax}>0.6\,t_{0}italic_t start_POSTSUBSCRIPT roman_relax end_POSTSUBSCRIPT > 0.6 italic_t start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT with <1% deviations for isolated zoom runs (see Hopkins et al. 2018). The typical convergence radii within our sample at redshift z=0.2𝑧0.2z=0.2italic_z = 0.2 are resolved to regions relevant for lensing based analysis; in what follows, we present results grouped in five mass bins, set by five decades in halo virial mass: Mvir10711Msimilar-to-or-equalssubscript𝑀virsuperscript10711subscript𝑀direct-productM_{\rm vir}\simeq 10^{7-11}\,M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ≃ 10 start_POSTSUPERSCRIPT 7 - 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. The halo masses with bins of Mvir={107, 108, 109, 1010, 1011}Msubscript𝑀virsuperscript107superscript108superscript109superscript1010superscript1011subscript𝑀direct-productM_{\rm vir}=\{10^{7},\,10^{8},\,10^{9},\,10^{10},\,10^{11}\}\ M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = { 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT } italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT have convergence radii of rconv{300, 270, 240, 200, 500}pcsimilar-to-or-equalssubscript𝑟conv300270240200500pcr_{\rm conv}\simeq\{300,\,270,\,240,\,200,\,500\}\ \rm pcitalic_r start_POSTSUBSCRIPT roman_conv end_POSTSUBSCRIPT ≃ { 300 , 270 , 240 , 200 , 500 } roman_pc, in order from lowest to highest mass decade.

Refer to caption
Figure 2: The resulting mean surface density profiles of the simulated dark matter halos at z=0.2𝑧0.2z=0.2italic_z = 0.2. Similar in presentation as the previous figure, the top panel presents the same isolated dark matter halos, but now for their projected density profiles, Σ(R)Σ𝑅\Sigma(R)roman_Σ ( italic_R ), along the major (thick-dashed curve), intermediate (thin-dashed curve), and minor axis (dotted curve). The bottom panels depict the fit quality made to each density axis. Here, we scale the Σ(R)Σ𝑅\Sigma(R)roman_Σ ( italic_R ) with R𝑅Ritalic_R to emphasize the dynamic range and we also scale the projected radius (R𝑅Ritalic_R) by the respective, best-fit scale radii, where the simulated profiles in R𝑅Ritalic_R are plotted from the convergence radius, rconvsubscript𝑟convr_{\rm conv}italic_r start_POSTSUBSCRIPT roman_conv end_POSTSUBSCRIPT out to rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. Left: The thin solid line shows the expected surface-density profile of each halo from the previous figure according to the NFW model, which are based on their best-fit parameters for each individual three-dimensional density profile. The radii is also scaled by the corresponding scale radius, rssubscript𝑟𝑠r_{s}italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT). Right: The same simulated curves presented in the left panel, but instead, the best-fit curves from fitting the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profile (Equation 13 with β=0.3𝛽0.3\beta=0.3italic_β = 0.3) we introduced in the main text applied to each density projection. The projected radii for these panels are normalized by the radius at which the logarithmic slope is equal to 11-1- 1, R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, which is acquired by the curve fitting procedure. The Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profile is able to model each orientation to 20similar-toabsent20\sim 20∼ 20 % and better based off of the fit quality presented in the bottom panel; an improvement from the NFW model shown in the left panel.

2.3.2 Constructing spherical density profiles

For each halo, we construct density profiles, ρ(r)𝜌𝑟\rho(r)italic_ρ ( italic_r ), using the total (bound and unbound) dark matter mass found in radial shells divided by the volume of each shell. We fit NFW and Einasto profiles to each profile using 35 logarithmic-spaced radial bins from rconvsubscript𝑟convr_{\rm conv}italic_r start_POSTSUBSCRIPT roman_conv end_POSTSUBSCRIPT to rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. The best-fit parameters are determined by the following minimization of the figure-of-merit in a least-squares optimization:111The convention used here in this paper is loglogelnsubscript𝑒\log\equiv\log_{e}\equiv\lnroman_log ≡ roman_log start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ≡ roman_ln.

Qρ2=1NbinsiNbins[log(ri2ρi)log(ri2ρimodel)]2.superscriptsubscript𝑄𝜌21subscript𝑁binssuperscriptsubscript𝑖subscript𝑁binssuperscriptdelimited-[]superscriptsubscript𝑟𝑖2subscript𝜌𝑖superscriptsubscript𝑟𝑖2superscriptsubscript𝜌𝑖model2\displaystyle Q_{\rho}^{2}=\frac{1}{N_{\rm bins}}\sum_{i}^{N_{\rm bins}}\left[% \log\left(r_{i}^{2}\rho_{i}\right)-\log\left(r_{i}^{2}\rho_{i}^{\rm model}% \right)\right]^{2}\,.italic_Q start_POSTSUBSCRIPT italic_ρ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_bins end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_bins end_POSTSUBSCRIPT end_POSTSUPERSCRIPT [ roman_log ( italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - roman_log ( italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_model end_POSTSUPERSCRIPT ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (10)

Instead of ρ𝜌\rhoitalic_ρ being minimized, which has the numerical value decrease by many orders of magnitude between the inner and outer region of the profile, the r2ρsuperscript𝑟2𝜌r^{2}\rhoitalic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ρ merit provides a more balanced indicator of goodness-of-fit across the entire radial range.222The choice of a scaled density as the function to be minimized for the curve fitting, has been more often used (e.g. Navarro et al. 2004), since the scale density fits the entirety of the density profile without weighing the inner-density more, while a density (ρ𝜌\rhoitalic_ρ) minimization would bias this. Both minimization definitions have been compared quantitatively, and we find the scaled minimization performs better. From each individual fit, we record best-fit parameters for the NFW density profile and Einasto profile with fixed αϵ=0.17subscript𝛼italic-ϵ0.17\alpha_{\epsilon}=0.17italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = 0.17.

2.3.3 Constructing cylindrical surface-density profiles

Halo surface density profiles are constructed in a similar manner to the three-dimensional profile: we measure the local surface density profile in two-dimensions, Σ(R)Σ𝑅\Sigma(R)roman_Σ ( italic_R ), by counting all (bound and unbound) dark matter particle mass that exists along a line-of-sight of total depth \mathcal{L}caligraphic_L in a thin circular ring of average radius R𝑅Ritalic_R divided by the area of that ring. We use 35 logarithmic spaced bins of projected radius R𝑅Ritalic_R that span an inner radius R=rconv𝑅subscript𝑟convR=r_{\rm conv}italic_R = italic_r start_POSTSUBSCRIPT roman_conv end_POSTSUBSCRIPT to an outer cylindrical radius R=rvir𝑅subscript𝑟virR=r_{\rm vir}italic_R = italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT.

In order to define the surface density of a three-dimensional object, we must define a depth of interest, \mathcal{L}caligraphic_L (see Equation 6). In analytic investigations, it is common to chose =\mathcal{L}=\inftycaligraphic_L = ∞; however this is neither possible numerically nor physically meaningful since we would like to characterize the degree of perturbation caused by the halo over the background. At minimum, one should span the full diameter of the halo and use =2rvir2subscript𝑟vir\mathcal{L}=2\,r_{\rm vir}caligraphic_L = 2 italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. However, we expect that there is additional clustered matter outside rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. We have explored this question using a depth parameter ξ𝜉\xiitalic_ξ defined via =2ξrvir2𝜉subscript𝑟vir\mathcal{L}=2\,\xi\,r_{\rm vir}caligraphic_L = 2 italic_ξ italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. As discussed in Appendix B, we find that ξ=1.5𝜉1.5\xi=1.5italic_ξ = 1.5 provides stable results that appear to capture the relevant over-density adequately. We use ξ=1.5𝜉1.5\xi=1.5italic_ξ = 1.5 for the rest of this work.

In the following subsections we discuss fits to projected profiles. We do so by minimizing the figure-of-merit in a least squares optimization for surface density structures in a given projection:

QΣ2=1NbinsiNbins[log(RiΣi)log(RiΣimodel)]2.superscriptsubscript𝑄Σ21subscript𝑁binssuperscriptsubscript𝑖subscript𝑁binssuperscriptdelimited-[]subscript𝑅𝑖subscriptΣ𝑖subscript𝑅𝑖superscriptsubscriptΣ𝑖model2\displaystyle Q_{\Sigma}^{2}=\frac{1}{N_{\rm bins}}\sum_{i}^{N_{\rm bins}}% \left[\log\left(R_{i}\Sigma_{i}\right)-\log\left(R_{i}\Sigma_{i}^{\rm model}% \right)\right]^{2}\,.italic_Q start_POSTSUBSCRIPT roman_Σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT roman_bins end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT roman_bins end_POSTSUBSCRIPT end_POSTSUPERSCRIPT [ roman_log ( italic_R start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) - roman_log ( italic_R start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT roman_Σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_model end_POSTSUPERSCRIPT ) ] start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (11)

As with the figure of merit used for the three-dimensional density profile, the RΣ𝑅ΣR\Sigmaitalic_R roman_Σ merit enables a more balanced weighing across the radial range being fitted. We exclude the outer radii (R>0.5rvir)𝑅0.5subscript𝑟vir(R>0.5r_{\rm vir})( italic_R > 0.5 italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ) when fitting since we aim to model the inner, higher-density regions that are most relevant for strong-gravitational lensing studies.

3 The spherical and projected structure of CDM haloes

This section present the main results of this paper. We first start by contextualizing our work by comparing the accuracy of NFW and Einasto fits to the spherically-averaged density profiles for our sample of haloes (Section 3.1). We then present the circularly-averaged surface density profiles of the same dark matter halos and introduce an easy-to-use analytic profile that improves upon the projected NFW case, with an accuracy similar to that of an Einasto in three dimensions (Section 3.2). Finally, we introduce a projected concentration parameter as a convenient characterization of the surface density structure of haloes in Section 3.3.

3.1 Accuracy of NFW and Einasto profiles to simulations

The top panels of Figure 1 plot the local, spherically-averaged density profiles, ρ(r)𝜌𝑟\rho(r)italic_ρ ( italic_r ), for several simulated halos (colored dash curves) grouped by decades of mass at z=0.2𝑧0.2z=0.2italic_z = 0.2. We chose this redshift since it is typical of lens systems. The density is additionally scaled by r2superscript𝑟2r^{2}italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT to constrain the dynamic range of the horizontal axis. The halos with masses ranging from 1071010Msuperscript107superscript1010subscript𝑀direct-product10^{7}-10^{10}\ M_{\odot}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT are collected from the m10 suite of simulations while the m11 simulations only presents their 1011Msuperscript1011subscript𝑀direct-product10^{11}\ M_{\odot}10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT main halos. Each profile is plotted from the virial radius, rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, down to the innermost converged radius, rconvsubscript𝑟convr_{\rm conv}italic_r start_POSTSUBSCRIPT roman_conv end_POSTSUBSCRIPT.333We adopt this convention for most of the figures in this paper, unless otherwise specified. The thin solid lines show the best-fit NFW profiles (left plot) and Einasto profiles (right plot). The radii along the horizontal axis are all scaled by their best-fit scale radius, rssubscript𝑟𝑠r_{s}italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for the NFW profile fits and r2subscript𝑟2r_{-2}italic_r start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT for the Einasto profile fits. Residuals of the fits are shown in the bottom panels. We see that best-fit NFW profiles capture the normalization of the simulated profiles to within 40%similar-toabsentpercent40\sim 40\%∼ 40 %, though the residuals do show a systematic U-shape. The best-fit Einasto profiles show no systematic difference with radius, and are accurate to within 20%similar-toabsentpercent20\sim 20\%∼ 20 % except at large radius. It is this level of improvement that has motivated the community to utilize the Einasto profile in favor of the NFW profile when precise predictions are required. We present this comparison here to provide context for the new surface-density fit we propose. We note that the projected version of the Einasto profile is analytically cumbersome, and this is why most lensing searches for low-mass perturbers assume projected NFW forms.

The left panels of Figure 2 presents the same simulated halos shown in Figure 1, but now as circularly-averaged surface-density profiles, Σ(R)Σ𝑅\Sigma(R)roman_Σ ( italic_R ), along with the projected version of the best-fit NFW profile for each halo (Equation 8, thin solid lines). For each halo, we show the surface density profile along the three main halo density axes (as described in Section 2.2.4): the major axis (thick-dashed curves), intermediate axis (thin-dashed curves), and the minor axis (dotted curves). On the vertical axis we plot the surface density profile multiplied by the projected radius, R𝑅Ritalic_R, to limit the dynamic range. The horizontal axis shows the projected radius, R𝑅Ritalic_R, divided by the best-fit NFW scale radius determine from the spherically average density fits (left plot), rssubscript𝑟𝑠r_{s}italic_r start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The fit residuals are shown along the bottom. We see that, especially at small radii, the fits are systematically biased (either low or high depending on the halo orientation).

In the next section, we present a profile, Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT that captures the surface density profiles of cylindrical-averaged halos more accurately than the best-fit NFW average density profile.

3.2 An improved surface density profile for dark matter halos

Refer to caption
Figure 3: Top: The profile residuals for the eight 1010Msuperscript1010subscript𝑀direct-product10^{10}\ M_{\odot}10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT halos viewed along the major, minor, and intermediate density axes the NFW model (dashed magenta curves) and our Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT model. Residuals for the NFW fit often exceed 20%percent2020\%20 % while the new treatment is significantly better, especially at small radii . Bottom: The distribution of figure-of-merits (Equation 11) for each halo mass bin for the two fits. Here we evaluate the figure-of-merit using only bins with R<R1𝑅subscript𝑅1R<R_{-1}italic_R < italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT in order to focus on the high-density region of most interest for lensing studies.

In Figure 2 it is interesting to note that there is a near-constant turnover in the profiles as a function of projected radius. It is instructive then to see how the logarithmic slope of the projected density, dlogΣ/dlogRdΣd𝑅\mathrm{d}\log{\Sigma}/\mathrm{d}\log{R}roman_d roman_log roman_Σ / roman_d roman_log italic_R, behaves a function of projected radius, R𝑅Ritalic_R. We go into a much further discussion in Appendix C, where we find that the behavior of the logarithmic slope for the surface-density profile is captured by a radially-dependent power law profile that is simple in form:

dlogΣβdlogR=(RR1)β.dsubscriptΣ𝛽d𝑅superscript𝑅subscript𝑅1𝛽\displaystyle\frac{\mathrm{d}\log{\Sigma_{\beta}}}{\mathrm{d}\log{R}}=-\Bigg{(% }\frac{R}{R_{-1}}\Bigg{)}^{\beta}\,.divide start_ARG roman_d roman_log roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_ARG start_ARG roman_d roman_log italic_R end_ARG = - ( divide start_ARG italic_R end_ARG start_ARG italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT . (12)

Here, R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT is the projected radius where the logarithmic slope of the surface density is equal to 11-1- 1 and β𝛽\betaitalic_β is a shape parameter. For a fixed shape parameter, β𝛽\betaitalic_β, a given orientation can be parameterized by the projected scale radius, R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, which in-turn captures the variation and scatter expected from halo-to-halo. Figure 13 shows that this profile, with β𝛽\betaitalic_β fixed at 0.30.30.30.3, reproduces the observed behavior of our simulated halos along multiple axes quite accurately, and is consistent with a lack of convergence to a central power law.

Integrating Equation 12 gives us a generalized surface density profile:

Σβ(R)=Σ1exp{1β[(RR1)β1]},subscriptΣ𝛽𝑅subscriptΣ11𝛽delimited-[]superscript𝑅subscript𝑅1𝛽1\displaystyle\Sigma_{\beta}(R)=\Sigma_{-1}\exp\Bigg{\{}-\frac{1}{\beta}\Bigg{[% }\Bigg{(}\frac{R}{R_{-1}}\Bigg{)}^{\beta}-1\Bigg{]}\Bigg{\}}\,,roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R ) = roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT roman_exp { - divide start_ARG 1 end_ARG start_ARG italic_β end_ARG [ ( divide start_ARG italic_R end_ARG start_ARG italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT - 1 ] } , (13)

where the normalization is defined by Σ1:=Σ(R1)assignsubscriptΣ1Σsubscript𝑅1\Sigma_{-1}:=\Sigma(R_{-1})roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT := roman_Σ ( italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ). This has the same functional form as the Sersic (1968) and Einasto (1965) profiles. Note that at R=0𝑅0R=0italic_R = 0 there is a finite surface density: Σ(0)=Σ1exp(1/β)Σ0subscriptΣ11𝛽\Sigma(0)=\Sigma_{-1}\exp(1/\beta)roman_Σ ( 0 ) = roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT roman_exp ( 1 / italic_β ). As written, ΣβsubscriptΣ𝛽\Sigma_{\beta}roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT has three free parameters: Σ1subscriptΣ1\Sigma_{-1}roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, and β𝛽\betaitalic_β. Leaving all three parameters free to fit provides a means to describe halo density structure extremely accurately. Equation 17 in the appendix provides an analytic expression for the the mass within given projected radius R𝑅Ritalic_R for given these three parameters. Equation 20 provides an easy-to-use fitting function that relates the three parameters to the total projected mass within the virial radius (approximately equal to the three-dimensional virial mass, see Appendix B).

Usefully, for the mass ranges of ΛΛ\Lambdaroman_ΛCDM halos explored here, we find that β=0.3𝛽0.3\beta=0.3italic_β = 0.3 allows us to characterize our entire sample of halos with two free parameters quite well, as shown below. With β𝛽\betaitalic_β fixed to 0.30.30.30.3, we have a two-parameter function, Σ0.3(R)subscriptΣ0.3𝑅\Sigma_{0.3}(R)roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT ( italic_R ), that can be fully specified by Σ1subscriptΣ1\Sigma_{-1}roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT and R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT. As demonstrated in Section 3.3, we can also characterize the profile with a concentration parameter, in analogy with what is traditionally done with NFW profiles or Einasto profiles in three-dimensions.

Refer to caption
Figure 4: Best-fit values of the R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT scale radius versus halo virial mass for Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits to 10 random projections to each halo in our sample. The solid line and shaded bands reflect the median and ninety percentile bands of the distribution. The vertical pink band shows the region where fits are likely not reliable because the typical value of R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT is smaller than the typical convergence radius for halos. Dotted power-law fit shows the relation R1Mvir0.35proportional-tosubscript𝑅1superscriptsubscript𝑀vir0.35R_{-1}\propto M_{\rm vir}^{0.35}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0.35 end_POSTSUPERSCRIPT. Note that these simulations are not a representative cosmological sample, so this masss-radius relation may not be representative for the halo population in general.
Refer to caption
Figure 5: The quality of fit for inner-most projected density (at 600 pc) between the simulation results, ΣtruesubscriptΣtrue\Sigma_{\rm true}roman_Σ start_POSTSUBSCRIPT roman_true end_POSTSUBSCRIPT, and the profile fits, ΣfitsubscriptΣfit\Sigma_{\rm fit}roman_Σ start_POSTSUBSCRIPT roman_fit end_POSTSUBSCRIPT, for the ρNFWsubscript𝜌NFW\rho_{\rm NFW}italic_ρ start_POSTSUBSCRIPT roman_NFW end_POSTSUBSCRIPT and Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profiles (purple and blue curves, respectively), as a function of Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. The bands encapsulate the 2σ2𝜎2\sigma2 italic_σ scatter for ten random projections on each resolved isolated halo. The relative error in NFW fits is larger at larger masses because 600600600600 pc is “deeper" within the core (smaller compared to R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT) at higher masses (see Figure 4). The Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits provide the most improvement for radii smaller than the scale radius R<R1𝑅subscript𝑅1R<R_{-1}italic_R < italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT.

The quality of the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits can be seen in the right-hand panels of Figure 2. Here we use the same simulated halos presented earlier in the left-most panel, but instead of using NFW fits for each halo, we now use the best-fit Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fit for the individual density-axis projections (the solid-thin lines). The projected radius along the horizontal axis is normalized by the best-fit projected scale radius, R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT. The residuals of the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits are shown in the lower-right panel, where the inner-most region is captured to better than 20similar-toabsent20\sim 20∼ 20% in all cases. This a clear improvement to the NFW model shown on the left, and similar in quality to the Einasto profile fits of halos within three dimensions shown in Figure 1. The improvement with respect to NFW is most significant at small radii RR1less-than-or-similar-to𝑅subscript𝑅1R\lesssim R_{-1}italic_R ≲ italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, where the surface densities are the highest.

To further emphasize the improvement of the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits compared to NFW, we present quality-of-fit parameters in Figure 3. The top panel shows residuals for the eight 1010Msuperscript1010subscript𝑀direct-product10^{10}\ M_{\odot}10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT halos in our sample for the NFW model (dashed magenta curves) and our Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT model (solid cyan curves). We use the projected profiles for each halo projected along its major, minor, intermediate density axes. We see that the NFW model can be off by as much as 50%similar-toabsentpercent50\sim 50\%∼ 50 % at small radius, and that the sign of the offset is systematic, depending on orientation. The new model captures the profiles to within 20% in all cases. At small radii, R2less-than-or-similar-to𝑅2R\lesssim 2italic_R ≲ 2 kpc, which corresponds to RR1less-than-or-similar-to𝑅subscript𝑅1R\lesssim R_{-1}italic_R ≲ italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT in this mass range, we again see that the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profile is able to capture the true projected density structure more accurately than the inferred NFW shape.

Figure 4 shows the relationship between best-fit R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT values and halo Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT for our full sample. We have fit Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profiles to ten random projections of each halo. The thick magenta line and associated band show the median and 90 percentile range as a function of virial mass. The dotted cyan line shows a power-law fit, which has a best-fit slope R1Mvirγproportional-tosubscript𝑅1superscriptsubscript𝑀vir𝛾R_{-1}\propto M_{\rm vir}^{\gamma}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ∝ italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT, where γ=0.35𝛾0.35\gamma=0.35italic_γ = 0.35, from only the sample up to Mvir=1010Msubscript𝑀virsuperscript1010subscript𝑀direct-productM_{\rm vir}=10^{10}\ M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. The vertical pink band shows the region (Mvir<108Msubscript𝑀virsuperscript108subscript𝑀direct-productM_{\rm vir}<10^{8}\,M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT < 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT) where typical values of R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT are smaller than the typical convergence radii for the simulated halos. Fits in this region rely on extrapolation to infer the value of R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT and this likely gives rise to some non-physical dispersion in characteristic radius.

Refer to caption
Figure 6: The projected mass within 1 kpc, (<R=1kpc)\mathcal{M}(<R=1{\rm kpc})caligraphic_M ( < italic_R = 1 roman_k roman_p roman_c ), as a function of virial mass, Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, for our sample of halos, each viewed along 10 random orientations. The points are color coded by the projected concentrations associated with Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits along each orientation. We see that the projected concentration correlates with the density of the halo along a given projection at fixed Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, and that this accounts for as much as a factor of 2similar-toabsent2\sim 2∼ 2 variation in the relationship between virial mass and projected mass at this radius.
Refer to caption
Figure 7: Relationship between the best-fit projected concentration using Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profiles, 𝒞virsubscript𝒞vir\mathcal{C}_{\rm vir}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, and the standard three-dimensional halo concentration, cvirsubscript𝑐virc_{\rm vir}italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, determined for the same halos’ best-fit Einasto profiles. The gray points correspond to 10 projected random orientations of all halos in our sample. The thick, solid black is a power-law fit to all of light-gray points while the dashed lines reflect the 2σ2𝜎2-\sigma2 - italic_σ scatter at fixed cvirsubscript𝑐virc_{\rm vir}italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. The colored solid lines are power-law fits to the projected fits along the major, intermediate, and minor axis of each halo.

Another way to understand the utility of Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT parameterization is compare its accuracy to the typical NFW approach at a fixed radius. Figure 5 shows the relative error (simulated vs. fit) of the two approaches at a projected radius R=600pc𝑅600pcR=600\ \rm pcitalic_R = 600 roman_pc, which corresponds to 0.2similar-toabsent0.2\sim 0.2∼ 0.2 arcsec at z=0.2𝑧0.2z=0.2italic_z = 0.2 as a function of halo virial mass. In this exercise, we have viewed each simulated halo along 10 random orientations and used Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits associated with each orientation compared with the implied spherical NFW fit in each case. Relative errors for the best-fit ρNFWsubscript𝜌NFW\rho_{\rm NFW}italic_ρ start_POSTSUBSCRIPT roman_NFW end_POSTSUBSCRIPT and Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits are shown in magenta and cyan, respectively. The solid curve shows the median as a function of Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, while the bands encapsulates the 2σ2𝜎2\sigma2 italic_σ range. We see that the relative error in NFW fits is larger at larger masses. This reflects the fact that the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits show the most improvement for R<R1𝑅subscript𝑅1R<R_{-1}italic_R < italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT. At higher masses, 600600600600 pc is “deeper" within the core, so that the relative error coming from the NFW assumption increases.

3.3 The projected concentration parameter

The spherically-averaged, three-dimensional density profile of a dark matter halo is often quantified by an NFW profile using two parameters: a halo mass and a concentration parameter (e.g. Bullock et al., 2001). Similarly, we can quantify the circularly-averaged, projected density profile of a dark matter halo along any orientation with a projected concentration, defined as

𝒞vir:=rvir/R1.assignsubscript𝒞virsubscript𝑟virsubscript𝑅1\displaystyle\mathcal{C}_{\rm vir}:=r_{\rm vir}/R_{-1}\,.caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT := italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT . (14)

Figure 6 provides some physical insight on the meaning of these parameters. We plot the projected mass within 1 kpc, (<R=1kpc)\mathcal{M}(<R=1\ \rm kpc)caligraphic_M ( < italic_R = 1 roman_kpc ), as a function of halo mass. For reference, 1 kpc has an angular size of 0.3similar-toabsent0.3\sim 0.3∼ 0.3 arcsec at z=0.2𝑧0.2z=0.2italic_z = 0.2. Each halo in our sample is viewed along 10 random orientations, which can be seen as vertical stacking at fixed virial mass at high masses. The color bar maps to the best-fit projected concentration 𝒞virsubscript𝒞vir\mathcal{C}_{\rm vir}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT for each halo/orientation. Lower-density projections for the same halo have lower projected concentrations, and this provides a way to quantify the scatter associated with halo shape. As one might expect, there is a correlation between (1kpc)1kpc\mathcal{M}(1\ \rm kpc)caligraphic_M ( 1 roman_kpc ) and Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. For virial masses below 1010Msuperscript1010subscript𝑀direct-product10^{10}M_{\odot}10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, where R11less-than-or-similar-tosubscript𝑅11R_{-1}\lesssim 1italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ≲ 1 kpc, the median relation scales as (<1kpc)Mvir1/2similar-toannotatedabsent1kpcsuperscriptsubscript𝑀vir12\mathcal{M}(<1{\rm kpc})\sim M_{\rm vir}^{1/2}caligraphic_M ( < 1 roman_k roman_p roman_c ) ∼ italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT. At higher masses, R11greater-than-or-equivalent-tosubscript𝑅11R_{-1}\gtrsim 1italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ≳ 1kpc, and the relationship becomes even flatter (<1kpc)Mvir1/3similar-toannotatedabsent1kpcsuperscriptsubscript𝑀vir13\mathcal{M}(<1{\rm kpc})\sim M_{\rm vir}^{1/3}caligraphic_M ( < 1 roman_k roman_p roman_c ) ∼ italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 / 3 end_POSTSUPERSCRIPT. The weak power-law dependence makes inferring the halo mass from a projected mas somewhat challenging; a 20% error in surface density would map to a 4060%similar-toabsent40percent60\sim 40-60\%∼ 40 - 60 % error in interpreted virial mass on average. The factor of 2similar-toabsent2\sim 2∼ 2 scatter in projected mass at fixed virial mass is roughly captured by the scatter in projected concentration. Note that this level of variation at fixed virial mass is considerable; in the median, a factor of 2222 increase in projected mass at 1111 kpc is equivalent to a factor of 4similar-toabsent4\sim 4∼ 4 (8888) increase in Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT at low (high) virial mass.

The light gray points in Figure 7 show the relationship between the projected concentration, 𝒞virsubscript𝒞vir\mathcal{C_{\rm vir}}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, and the standard three-dimensional halo concentration, cvirsubscript𝑐virc_{\rm vir}italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, for individual halos. Each halo has a fixed cvirsubscript𝑐virc_{\rm vir}italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, but has 10 values of 𝒞virsubscript𝒞vir\mathcal{C_{\rm vir}}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT derived from Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits from 10 random orientations. We only include halos more massive than Mvir=109Msubscript𝑀virsuperscript109subscript𝑀direct-productM_{\rm vir}=10^{9}\,M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT in this analysis in order to ensure accurate fits. The thick, solid black line shows a power-law fit to all the points, which provides a good representation of the average relation: 𝒞vircvirδproportional-tosubscript𝒞virsuperscriptsubscript𝑐vir𝛿\mathcal{C_{\rm vir}}\propto c_{\rm vir}^{\delta}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ∝ italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_δ end_POSTSUPERSCRIPT. The best-fit function is found to be 𝒞vir=2.3cvir0.87subscript𝒞vir2.3superscriptsubscript𝑐vir0.87\mathcal{C_{\rm vir}}=2.3\,c_{\rm vir}^{0.87}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = 2.3 italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0.87 end_POSTSUPERSCRIPT. The dashed lines show the 2σ2𝜎2\sigma2 italic_σ distribution at fixed three-dimensional halo concentration, cvirsubscript𝑐virc_{\rm vir}italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. For reference, we also plot power-law fits for the same relation using only their major-, intermediate-, and minor-axis projections as pink, orange, and cyan lines, respectively. The intermediate-axis is roughly consistent with the average over all orientations, while the averages of the two other orientations track the 1.5σsimilar-toabsent1.5𝜎\sim 1.5\sigma∼ 1.5 italic_σ range, suggesting that much of the variation at fixed cvirsubscript𝑐virc_{\rm vir}italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT comes from orientation effects.

Figure 8 plots the probability distribution (black curve) of 𝒞vir/cvirsubscript𝒞virsubscript𝑐vir\mathcal{C}_{\rm vir}/c_{\rm vir}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT of all halos. The curve is nicely described by a Gaussian (dashed magenta) with mean μ=1.45𝜇1.45\mu=1.45italic_μ = 1.45 and standard deviation σ=0.22𝜎0.22\sigma=0.22italic_σ = 0.22. From this, we can infer that the parameters have a mean relation of Cvir1.5cvirsimilar-to-or-equalssubscript𝐶vir1.5subscript𝑐virC_{\rm vir}\simeq 1.5\,c_{\rm vir}italic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ≃ 1.5 italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT at this fixed redshift; a larger sample of halos, sampled in cosmological volumes, and studied at a wider range of redshifts, will be needed to explore these relationships more completely.

Given the relationship 𝒞virccvir0.87proportional-tosubscript𝒞virsuperscriptsubscript𝑐cvir0.87\mathcal{C_{\rm vir}}\propto c_{\rm cvir}^{0.87}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ∝ italic_c start_POSTSUBSCRIPT roman_cvir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0.87 end_POSTSUPERSCRIPT shown in Figure 7, we have also explored an alternative to Figure 8 that looks at the distribution of ratio 𝒞vir/cvir0.87subscript𝒞virsuperscriptsubscript𝑐vir0.87\mathcal{C_{\rm vir}}/c_{\rm vir}^{0.87}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0.87 end_POSTSUPERSCRIPT. We find a similarly-shaped distribution, also well-described by a Gaussian, this with mean μ=2.14𝜇2.14\mu=2.14italic_μ = 2.14 and standard deviation σ=0.28𝜎0.28\sigma=0.28italic_σ = 0.28. In this case the scatter relative to the mean is σ/μ0.13similar-to-or-equals𝜎𝜇0.13\sigma/\mu\simeq 0.13italic_σ / italic_μ ≃ 0.13.

Refer to caption
Figure 8: The probability distribution of the ratio between the projected concentration and the halo concentration 𝒞vir/cvirsubscript𝒞virsubscript𝑐vir\mathcal{C}_{\rm vir}/c_{\rm vir}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT (black line), along with the best fit Gaussian fit (dashed magenta line).
Refer to caption
Figure 9: Comparison of lensing outcomes for Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits and assumed spherical NFW fits for an example Mvir109Msimilar-to-or-equalssubscript𝑀virsuperscript109subscript𝑀direct-productM_{\rm vir}\simeq 10^{9}\ M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ≃ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT dark matter halo perturber. Left: Surface density profiles for the example halo. The thick, solid black curve is the implied surface density profile for the best-fit ρNFWsubscript𝜌NFW\rho_{\rm NFW}italic_ρ start_POSTSUBSCRIPT roman_NFW end_POSTSUBSCRIPT for this halo. The dashed color curves show projected density profiles along the three primary density axes. The respective Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits are solid lines plotted in the same color as the dashed curves. Right: the relative change in magnification for an image relative to a smooth lens derived from the same analytic profiles on the left. We assume a source at redshift zs=1.3subscript𝑧𝑠1.3z_{s}=1.3italic_z start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 1.3 and lens at z=0.2subscript𝑧0.2z_{\ell}=0.2italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT = 0.2. The horizontal axis shows the angular separation between the perturber halo and the image in arcseconds. The Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profiles (colored curves) produce significant differences in relative magnification compared to the predicted NFW signal.

4 Implications for gravitational lensing

Our primary motivation for characterizing the surface density profiles of dark matter halos is to provide a useful and accurate tool for gravitational lensing applications. As shown in sections A.4, A.5, and A.6, the ΣβsubscriptΣ𝛽\Sigma_{\beta}roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT profile given by Equation 13 has an analytic deflection potential, deflection angle, and lensing shear. One example use case is in the search for substructure within multiply-imaged quasar strong-lens systems. For example, the magnification ratios of images compared to those of a smoothly-parameterized mass distributions from a macro model provides a way to discover and characterize invisible dark matter halo perturbers (Mao & Schneider, 1998; Metcalf & Madau, 2001; Dalal & Kochanek, 2002; Nierenberg et al., 2014; Nierenberg et al., 2017; Gilman et al., 2020b, a; Hsueh et al., 2020).

Figure 9 shows an example of how the use of the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profile gives different magnification predictions than an NFW profile fit for the same halo. For this calculation we use the open-source gravitational lensing software LENSTRONOMY444https://github.com/sibirrer/lenstronomy, which performs the main lensing computations such as ray-tracing and magnifications (Birrer & Amara, 2018; Birrer et al., 2021), combined with the open-source pyHalo555https://github.com/dangilman/pyHalo, which generates realizations for subhalos and LOS halos within a lens system (Gilman et al., 2021). The left side of Figure 9 shows the projected surface density profile for an example M109Msimilar-to-or-equals𝑀superscript109subscript𝑀direct-productM\simeq 10^{9}\ M_{\odot}italic_M ≃ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT halo at z=0.2𝑧0.2z=0.2italic_z = 0.2 along the three primary projections (colored lines). The lines are truncated at the convergence radius for this simulation. Also shown is the implied surface-density profile for the best-fit (spherical) NFW profile for the same halo (solid black). The three solid colored lines show the best-fit Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT profile fits for each projection. Note that he horizontal axis on the bottom shows projected radius in physical units (kpc) and on the top we show the same quantity in angular size for a z=0.2𝑧0.2z=0.2italic_z = 0.2 lens in seconds of arc.

The right panel depicts the relative magnification cross-sections compared to a smooth model for the three-dimensional NFW fit (thick, solid black line) and the three best-fit Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT models for each orientation. The colored lines on the right panel were calculated using the analytic fits shown in the same color on the left. We assume a Gaussian source light distribution with a FWHM of 35 parsecs (comparable to the size of the nuclear narrow-line region of a quasar; e.g., Nierenberg et al. 2014, Müller-Sánchez et al. 2011) for a source redshift zs=1.3subscript𝑧s1.3z_{\rm s}=1.3italic_z start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT = 1.3 and lens redshift z=0.2subscript𝑧0.2z_{\ell}=0.2italic_z start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT = 0.2. This figure is analogous to Figure 2 in Gilman et al. (2020b). The horizontal axis shows the perturber position relative to the image’s position in arcseconds. The resulting change in magnification compared to a smooth-lens model is shown on the vertical axis.

We see that when a halo is projected along the densest and least-densest axis, relative magnification changes by a factor of 2similar-toabsent2\sim 2∼ 2. The best-fit NFW density profile (black solid curve) fails to match these two extremes. This level of difference could bias the inferred parameters of the perturbing halo systematically at the 50%similar-toabsentpercent50\sim 50\%∼ 50 % level. For example, if we tried to interpret the major or minor axis orientations’ magnification signals with NFW profiles of the same concentration, we would infer halos that are more or less massive than they actually are by factors of order unity.

5 Summary

We have studied the surface density structure dark matter halos using a suite of dark matter only zoom simulations that contain halos of mass Mvir=10711Msubscript𝑀virsuperscript10711subscript𝑀direct-productM_{\rm vir}=10^{7-11}\ M_{\odot}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 7 - 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. We find that their cylindrically-averaged surface density profiles along any projection can be modeled as a function of projected radius R𝑅Ritalic_R using an easy-to-use, analytic profile, Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT, defined in Equation 13 with β=0.3𝛽0.3\beta=0.3italic_β = 0.3. The function has two free parameters: a scale radius R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT and a normalization Σ1subscriptΣ1\Sigma_{-1}roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT. A primary motivation is to provide a profile that is more accurate than a projected spherical NFW profile and that has analytic lensing properties (A.4, A.5, and A.6) that can be used in strong lensing studies for line-of-sight perturbing halos.

The common approach to modeling small-halo perturbers is to assume spherical NFW profiles with properties drawn from mass-concentration relation predictions (e.g. Gilman et al., 2020a). As shown in Figures 2 and 3, the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT approach allows for significant improvement, with the same number of free parameters, especially at small projected radii, where the densities are highest and of most relevance for lensing perturbations. An important by-product introduced here is the projected concentration, defined by the size of the halo and the projected scale radius, 𝒞vir:=rvir/R1assignsubscript𝒞virsubscript𝑟virsubscript𝑅1\mathcal{C}_{\rm vir}:=r_{\rm vir}/R_{-1}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT := italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT. For any individual halo, the projected concentration can be higher or lower depending on the projected orientation. By fitting halos along an ensemble of orientations, the projected concentration provides a way to statistically account for asphericity while still utilizing a two-parameter foundation. The the projected concentration correlates with the standard three-dimensional concentration in a way that allows ease of use (Figures 7 and 8).

In order to illustrate implications for gravitational lensing, we set up a mock-lensing analysis to explore the implied perturber magnification with the Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT model compared to the standard NFW approach (Figure 9). In this exercise, we examined the projected density structure of an example halo along three primary density axes and showed that the implied magnification perturbation as captured by Σ0.3subscriptΣ0.3\Sigma_{0.3}roman_Σ start_POSTSUBSCRIPT 0.3 end_POSTSUBSCRIPT fits for different orientations give as much as 50%similar-toabsentpercent50\sim 50\%∼ 50 % relative differences compared to the same halo’s NFW fit at fixed mass. Flux ratio studies of lensing by halos draw populations of halos with halo to halo scatter in the mass-concentration relation. To studying the impact of the orientation observed in this work on a full dark matter inference would require full simulations of gravitational lenses. However we can see that the effect should in principle be similar to increasing the “effective scatter” in the lensing mass-concentration relation (Gilman et al., 2020b, a). We leave a detailed analysis of this to a future work.

One weakness of this analysis is that we have relied on zoom simulations, and not a fair cosmological sample. A next step in this analysis will be to use cosmological samples of simulated halos to provide statistically meaningful predictions for effective concentration distributions as a function of halo mass. We expect that the relative relationships we have found between the three-dimensional concentration and effective (projected) concentrations for our halos will be robust. From the analysis done here, the average mapping obeys 𝒞vir=2.3cvir0.87subscript𝒞vir2.3superscriptsubscript𝑐vir0.87\mathcal{C_{\rm vir}}=2.3\,c_{\rm vir}^{0.87}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT = 2.3 italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0.87 end_POSTSUPERSCRIPT (Figure 7) with a Gaussian scatter of σ0.2similar-to-or-equals𝜎0.2\sigma\simeq 0.2italic_σ ≃ 0.2 at fixed cvirsubscript𝑐virc_{\rm vir}italic_c start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT for any individual halo. Future work will allow us to test these expectations in a fair cosmological context.

Acknowledgements

We thank the anonymous referee for their helpful comments in improving the early version of this article. Part of this research was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration. JSB was supported by the National Science Foundation (NSF) grant AST-1910965 and NASA grant 80NSSC22K0827. AL was supported by NASA grant 80NSSSC20K1469. MBK acknowledges support from NSF CAREER award AST-1752913, NSF grants AST-1910346 and AST-2108962, NASA grant 80NSSC22K0827, and HST-AR-15809, HST-GO-15658, HST-GO-15901, HST-GO-15902, HST-AR-16159, HST-GO-16226, HST-GO-16686, HST-AR-170248, and HST-AR-17043. LAM acknowledges partial support by the NASA Astrophysics Theory Program investigation 17-ATP17-120. The analysis in this manuscript made extensive use of the python packages NumPy (van der Walt et al., 2011), SciPy (Oliphant, 2007), Matplotlib (Hunter, 2007), and COLOSSUS (Diemer, 2018); We are thankful to the developers of these tools. This research has made all intensive use of NASA’s Astrophysics Data System (https://ui.adsabs.harvard.edu/) and the arXiv eprint service (http://arxiv.org).

Data Availability

The data supporting the plots within this article are available on reasonable request to the corresponding author. A public version of the GIZMO code is available at http://www.tapir.caltech.edu/~phopkins/Site/GIZMO.html. Additional data from the FIRE project, including simulation snapshots, initial conditions, and derived data products, are available at https://fire.northwestern.edu/data/.

References

  • Allgood et al. (2006) Allgood B., Flores R. A., Primack J. R., Kravtsov A. V., Wechsler R. H., Faltenbacher A., Bullock J. S., 2006, MNRAS, 367, 1781
  • Benitez-Llambay & Frenk (2020) Benitez-Llambay A., Frenk C., 2020, MNRAS, 498, 4887
  • Birrer & Amara (2018) Birrer S., Amara A., 2018, Physics of the Dark Universe, 22, 189
  • Birrer et al. (2021) Birrer S., et al., 2021, The Journal of Open Source Software, 6, 3283
  • Bode et al. (2001) Bode P., Ostriker J. P., Turok N., 2001, ApJ, 556, 93
  • Bond et al. (1996) Bond J. R., Kofman L., Pogosyan D., 1996, Nature, 380, 603
  • Boyarsky et al. (2014) Boyarsky A., Ruchayskiy O., Iakubovskyi D., Franse J., 2014, Phys. Rev. Lett., 113, 251301
  • Bryan & Norman (1998) Bryan G. L., Norman M. L., 1998, ApJ, 495, 80
  • Bulbul et al. (2014) Bulbul E., Markevitch M., Foster A., Smith R. K., Loewenstein M., Randall S. W., 2014, ApJ, 789, 13
  • Bullock & Boylan-Kolchin (2017) Bullock J. S., Boylan-Kolchin M., 2017, ARA&A, 55, 343
  • Bullock et al. (2001) Bullock J. S., Kolatt T. S., Sigad Y., Somerville R. S., Kravtsov A. V., Klypin A. A., Primack J. R., Dekel A., 2001, MNRAS, 321, 559
  • Cardone (2004) Cardone V. F., 2004, A&A, 415, 839
  • Child et al. (2018) Child H. L., Habib S., Heitmann K., Frontiere N., Finkel H., Pope A., Morozov V., 2018, ApJ, 859, 55
  • Cole & Lacey (1996) Cole S., Lacey C., 1996, MNRAS, 281, 716
  • Colín et al. (2000) Colín P., Avila-Reese V., Valenzuela O., 2000, ApJ, 542, 622
  • Dalal & Kochanek (2002) Dalal N., Kochanek C. S., 2002, ApJ, 572, 25
  • Davis et al. (1985) Davis M., Efstathiou G., Frenk C. S., White S. D. M., 1985, ApJ, 292, 371
  • Despali et al. (2018) Despali G., Vegetti S., White S. D. M., Giocoli C., van den Bosch F. C., 2018, MNRAS, 475, 5424
  • Dhar (2021) Dhar B. K., 2021, MNRAS, 504, 4583
  • Dhar & Williams (2010) Dhar B. K., Williams L. L. R., 2010, MNRAS, 405, 340
  • Diemand et al. (2007) Diemand J., Kuhlen M., Madau P., 2007, ApJ, 667, 859
  • Diemer (2018) Diemer B., 2018, ApJS, 239, 35
  • Einasto (1965) Einasto J., 1965, Trudy Astrofizicheskogo Instituta Alma-Ata, 5, 87
  • Elbert et al. (2015) Elbert O. D., Bullock J. S., Garrison-Kimmel S., Rocha M., Oñorbe J., Peter A. H. G., 2015, MNRAS, 453, 29
  • Feng et al. (2009) Feng J. L., Kaplinghat M., Tu H., Yu H.-B., 2009, J. Cosmology Astropart. Phys, 2009, 004
  • Feroz & Hobson (2012) Feroz F., Hobson M. P., 2012, MNRAS, 420, 596
  • Fitts et al. (2017) Fitts A., et al., 2017, MNRAS, 471, 3547
  • Frenk & White (2012) Frenk C. S., White S. D. M., 2012, Annalen der Physik, 524, 507
  • Frenk et al. (1988) Frenk C. S., White S. D. M., Davis M., Efstathiou G., 1988, ApJ, 327, 507
  • Gao et al. (2008) Gao L., Navarro J. F., Cole S., Frenk C. S., White S. D. M., Springel V., Jenkins A., Neto A. F., 2008, MNRAS, 387, 536
  • Geller & Huchra (1989) Geller M. J., Huchra J. P., 1989, Science, 246, 897
  • Gilman et al. (2018) Gilman D., Birrer S., Treu T., Keeton C. R., Nierenberg A., 2018, MNRAS, 481, 819
  • Gilman et al. (2019) Gilman D., Birrer S., Treu T., Nierenberg A., Benson A., 2019, MNRAS, 487, 5721
  • Gilman et al. (2020a) Gilman D., Birrer S., Nierenberg A., Treu T., Du X., Benson A., 2020a, MNRAS, 491, 6077
  • Gilman et al. (2020b) Gilman D., Du X., Benson A., Birrer S., Nierenberg A., Treu T., 2020b, MNRAS, 492, L12
  • Gilman et al. (2021) Gilman D., Bovy J., Treu T., Nierenberg A., Birrer S., Benson A., Sameie O., 2021, MNRAS, 507, 2432
  • Gilman et al. (2022) Gilman D., Benson A., Bovy J., Birrer S., Treu T., Nierenberg A., 2022, MNRAS, 512, 3163
  • Gradshteyn & Ryzhik (1980) Gradshteyn I. S., Ryzhik I. M., 1980, Table of integrals, series and products
  • Green et al. (2005) Green A. M., Hofmann S., Schwarz D. J., 2005, J. Cosmology Astropart. Phys, 2005, 003
  • Hahn & Abel (2011) Hahn O., Abel T., 2011, MNRAS, 415, 2101
  • He et al. (2022a) He Q., et al., 2022a, MNRAS, 511, 3046
  • He et al. (2022b) He Q., et al., 2022b, MNRAS, 512, 5862
  • Hezaveh et al. (2016) Hezaveh Y. D., et al., 2016, ApJ, 823, 37
  • Hopkins (2015) Hopkins P. F., 2015, MNRAS, 450, 53
  • Hopkins et al. (2018) Hopkins P. F., et al., 2018, MNRAS, 480, 800
  • Hsueh et al. (2020) Hsueh J. W., Enzi W., Vegetti S., Auger M. W., Fassnacht C. D., Despali G., Koopmans L. V. E., McKean J. P., 2020, MNRAS, 492, 3047
  • Hunter (2007) Hunter J. D., 2007, Computing in Science and Engineering, 9, 90
  • Klypin et al. (2016) Klypin A., Yepes G., Gottlöber S., Prada F., Heß S., 2016, MNRAS, 457, 4340
  • Knollmann & Knebe (2009) Knollmann S. R., Knebe A., 2009, ApJS, 182, 608
  • Koopmans (2005) Koopmans L. V. E., 2005, MNRAS, 363, 1136
  • Lazar et al. (2020) Lazar A., et al., 2020, MNRAS, 497, 2393
  • Lazar et al. (2021) Lazar A., Bullock J. S., Boylan-Kolchin M., Feldmann R., Çatmabacak O., Moustakas L., 2021, MNRAS, 502, 6064
  • Li et al. (2017) Li R., Frenk C. S., Cole S., Wang Q., Gao L., 2017, MNRAS, 468, 1426
  • Lovell et al. (2012) Lovell M. R., et al., 2012, MNRAS, 420, 2318
  • Lovell et al. (2014) Lovell M. R., Frenk C. S., Eke V. R., Jenkins A., Gao L., Theuns T., 2014, MNRAS, 439, 300
  • Mao & Schneider (1998) Mao S., Schneider P., 1998, MNRAS, 295, 587
  • Metcalf (2005) Metcalf R. B., 2005, ApJ, 622, 72
  • Metcalf & Madau (2001) Metcalf R. B., Madau P., 2001, ApJ, 563, 9
  • Minor et al. (2017) Minor Q. E., Kaplinghat M., Li N., 2017, ApJ, 845, 118
  • Minor et al. (2021) Minor Q., Kaplinghat M., Chan T. H., Simon E., 2021, MNRAS, 507, 1202
  • Moustakas & Metcalf (2003) Moustakas L. A., Metcalf R. B., 2003, MNRAS, 339, 607
  • Müller-Sánchez et al. (2011) Müller-Sánchez F., Prieto M. A., Hicks E. K. S., Vives-Arias H., Davies R. I., Malkan M., Tacconi L. J., Genzel R., 2011, ApJ, 739, 69
  • Navarro et al. (1997) Navarro J. F., Frenk C. S., White S. D. M., 1997, ApJ, 490, 493
  • Navarro et al. (2004) Navarro J. F., et al., 2004, MNRAS, 349, 1039
  • Navarro et al. (2010) Navarro J. F., et al., 2010, MNRAS, 402, 21
  • Nierenberg et al. (2014) Nierenberg A. M., Treu T., Wright S. A., Fassnacht C. D., Auger M. W., 2014, MNRAS, 442, 2434
  • Nierenberg et al. (2017) Nierenberg A. M., et al., 2017, MNRAS, 471, 2224
  • Nierenberg et al. (2020) Nierenberg A. M., et al., 2020, MNRAS, 492, 5314
  • Oñorbe et al. (2014) Oñorbe J., Garrison-Kimmel S., Maller A. H., Bullock J. S., Rocha M., Hahn O., 2014, MNRAS, 437, 1894
  • O’Riordan et al. (2021) O’Riordan C. M., Warren S. J., Mortlock D. J., 2021, MNRAS, 501, 3687
  • Oliphant (2007) Oliphant T. E., 2007, Computing in Science and Engineering, 9, 10
  • Planck Collaboration et al. (2016) Planck Collaboration et al., 2016, A&A, 594, A13
  • Power et al. (2003) Power C., Navarro J. F., Jenkins A., Frenk C. S., White S. D. M., Springel V., Stadel J., Quinn T., 2003, MNRAS, 338, 14
  • Prada et al. (2012) Prada F., Klypin A. A., Cuesta A. J., Betancort-Rijo J. E., Primack J., 2012, MNRAS, 423, 3018
  • Press & Schechter (1974) Press W. H., Schechter P., 1974, ApJ, 187, 425
  • Retana-Montenegro et al. (2012) Retana-Montenegro E., van Hese E., Gentile G., Baes M., Frutos-Alfaro F., 2012, A&A, 540, A70
  • Rocha et al. (2013) Rocha M., Peter A. H. G., Bullock J. S., Kaplinghat M., Garrison-Kimmel S., Oñorbe J., Moustakas L. A., 2013, MNRAS, 430, 81
  • Sánchez et al. (2006) Sánchez A. G., Baugh C. M., Percival W. J., Peacock J. A., Padilla N. D., Cole S., Frenk C. S., Norberg P., 2006, MNRAS, 366, 189
  • Sawala et al. (2016) Sawala T., et al., 2016, MNRAS, 456, 85
  • Schneider et al. (1992) Schneider P., Ehlers J., Falco E. E., 1992, Gravitational Lenses, doi:10.1007/978-3-662-03758-4.
  • Schneider et al. (2012) Schneider A., Smith R. E., Macciò A. V., Moore B., 2012, MNRAS, 424, 684
  • Sersic (1968) Sersic J. L., 1968, Atlas de Galaxias Australes
  • Springel et al. (2008) Springel V., et al., 2008, MNRAS, 391, 1685
  • Tegmark et al. (2004) Tegmark M., et al., 2004, ApJ, 606, 702
  • Tessore & Metcalf (2015) Tessore N., Metcalf R. B., 2015, A&A, 580, A79
  • Tulin & Yu (2018) Tulin S., Yu H.-B., 2018, Phys. Rep., 730, 1
  • Vegetti & Koopmans (2009a) Vegetti S., Koopmans L. V. E., 2009a, MNRAS, 392, 945
  • Vegetti & Koopmans (2009b) Vegetti S., Koopmans L. V. E., 2009b, MNRAS, 400, 1583
  • Vegetti et al. (2010) Vegetti S., Koopmans L. V. E., Bolton A., Treu T., Gavazzi R., 2010, MNRAS, 408, 1969
  • Vegetti et al. (2012) Vegetti S., Lagattuta D. J., McKean J. P., Auger M. W., Fassnacht C. D., Koopmans L. V. E., 2012, Nature, 481, 341
  • Wang et al. (2020) Wang J., Bose S., Frenk C. S., Gao L., Jenkins A., Springel V., White S. D. M., 2020, Nature, 585, 39
  • Weinberg et al. (2013) Weinberg D. H., Mortonson M. J., Eisenstein D. J., Hirata C., Riess A. G., Rozo E., 2013, Phys. Rep., 530, 87
  • Wright & Brainerd (2000) Wright C. O., Brainerd T. G., 2000, ApJ, 534, 34
  • van der Walt et al. (2011) van der Walt S., Colbert S. C., Varoquaux G., 2011, Computing in Science and Engineering, 13, 22

Appendix A Lensing profile Properties

A.1 Projected mass distribution

This paper discusses a surface density profile Σβ(R)subscriptΣ𝛽𝑅\Sigma_{\beta}(R)roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R ) for dark matter halos described by Equation 13. The projected cumulative mass of the dark matter, \mathcal{M}caligraphic_M, with a projected radius, R𝑅Ritalic_R, for such a functional form is given by

(R)𝑅\displaystyle\mathcal{M}(R)caligraphic_M ( italic_R ) =02πdϕ0RdRRΣβ(R)absentsuperscriptsubscript02𝜋differential-ditalic-ϕsuperscriptsubscript0𝑅differential-dsuperscript𝑅superscript𝑅subscriptΣ𝛽superscript𝑅\displaystyle=\int_{0}^{2\pi}\mathrm{d}{\phi}\int_{0}^{R}\mathrm{d}R^{\prime}% \,R^{\prime}\,\Sigma_{\beta}(R^{\prime})= ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT roman_d italic_ϕ ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_R end_POSTSUPERSCRIPT roman_d italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (15)
=2π0RdRRΣβ(R),absent2𝜋superscriptsubscript0𝑅differential-dsuperscript𝑅superscript𝑅subscriptΣ𝛽superscript𝑅\displaystyle=2\pi\int_{0}^{R}\mathrm{d}R^{\prime}\,R^{\prime}\,\Sigma_{\beta}% (R^{\prime})\,,= 2 italic_π ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_R end_POSTSUPERSCRIPT roman_d italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) , (16)

where the last line assumes spherical symmetry. The projected mass profile for the enclosed mass is

(R)=2πΣ1R12βexp(1+2lnββ)γ[2β,1β(RR1)β],𝑅2𝜋subscriptΣ1superscriptsubscript𝑅12𝛽12𝛽𝛽𝛾2𝛽1𝛽superscript𝑅subscript𝑅1𝛽\displaystyle\mathcal{M}\left(R\right)=\frac{2\pi\Sigma_{-1}R_{-1}^{2}}{\beta}% \exp\left({\frac{1+2\ln\beta}{\beta}}\right)\,\gamma\left[\frac{2}{\beta},\,% \frac{1}{\beta}\left(\frac{R}{R_{-1}}\right)^{\beta}\right]\,,caligraphic_M ( italic_R ) = divide start_ARG 2 italic_π roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_β end_ARG roman_exp ( divide start_ARG 1 + 2 roman_ln italic_β end_ARG start_ARG italic_β end_ARG ) italic_γ [ divide start_ARG 2 end_ARG start_ARG italic_β end_ARG , divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ( divide start_ARG italic_R end_ARG start_ARG italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT ] , (17)

where γ(a,x)𝛾𝑎𝑥\gamma(a,x)italic_γ ( italic_a , italic_x ) is the lower incomplete gamma function.

A.2 Projected mass normalization

To determine the normalization relation, assuming circular symmetry, we integrate R𝑅Ritalic_R out to a virial radius, rΔsubscript𝑟Δr_{\Delta}italic_r start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT:

effsubscripteff\displaystyle\mathcal{M}_{\rm eff}caligraphic_M start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT =2π0rΔdRRΣβ(R)absent2𝜋superscriptsubscript0subscript𝑟Δdifferential-dsuperscript𝑅superscript𝑅subscriptΣ𝛽superscript𝑅\displaystyle=2\pi\int_{0}^{r_{\Delta}}\mathrm{d}R^{\prime}\,R^{\prime}\,% \Sigma_{\beta}(R^{\prime})= 2 italic_π ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_d italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (18)
effsubscripteff\displaystyle\mathcal{M}_{\rm eff}caligraphic_M start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT =2πΣ10rΔdRRexp{1β[(RR1)β1]}.absent2𝜋subscriptΣ1superscriptsubscript0subscript𝑟Δdifferential-dsuperscript𝑅superscript𝑅1𝛽delimited-[]superscriptsuperscript𝑅subscript𝑅1𝛽1\displaystyle=2\pi\,\Sigma_{-1}\int_{0}^{r_{\Delta}}\mathrm{d}R^{\prime}\,R^{% \prime}\,\exp\left\{-\frac{1}{\beta}\left[\left(\frac{R^{\prime}}{R_{-1}}% \right)^{\beta}-1\right]\right\}\,.= 2 italic_π roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_d italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_exp { - divide start_ARG 1 end_ARG start_ARG italic_β end_ARG [ ( divide start_ARG italic_R start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_ARG start_ARG italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT - 1 ] } . (19)

Note that we have chosen rΔsubscript𝑟Δr_{\Delta}italic_r start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT, and then later 𝒞Δsubscript𝒞Δ\mathcal{C}_{\Delta}caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT, to explicitly show that this works for any over-density definition used. We have also defined (rΔ)effsubscript𝑟Δsubscripteff\mathcal{M}(r_{\Delta})\equiv\mathcal{M}_{\rm eff}caligraphic_M ( italic_r start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ) ≡ caligraphic_M start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT, since integrating out to the virial radius does not always ensure we recover the virial mass of the halo, i.e. (rΔ)Mvirsubscript𝑟Δsubscript𝑀vir\mathcal{M}(r_{\Delta})\neq M_{\rm vir}caligraphic_M ( italic_r start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ) ≠ italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, as the depth of the projection is set by a cylinder depth defined in the main text,

. Moving forward, we can make the substitution x=R/R1𝑥𝑅subscript𝑅1x=R/R_{-1}italic_x = italic_R / italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, which leads us to the relation:

effsubscripteff\displaystyle\mathcal{M}_{\rm eff}caligraphic_M start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT =2πΣ1rΔ2(𝒞Δ|β),absent2𝜋subscriptΣ1superscriptsubscript𝑟Δ2conditionalsubscript𝒞Δ𝛽\displaystyle=2\pi\Sigma_{-1}r_{\Delta}^{2}\,\mathcal{I}(\mathcal{C}_{\Delta}|% \beta)\,,= 2 italic_π roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT italic_r start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT caligraphic_I ( caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT | italic_β ) , (20)

where

(𝒞Δ|β)conditionalsubscript𝒞Δ𝛽\displaystyle\mathcal{I}(\mathcal{C}_{\Delta}|\beta)caligraphic_I ( caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT | italic_β ) =1𝒞Δ20𝒞Δdxxexp{1β[xβ1]},absent1superscriptsubscript𝒞Δ2superscriptsubscript0subscript𝒞Δdifferential-d𝑥𝑥1𝛽delimited-[]superscript𝑥𝛽1\displaystyle=\frac{1}{\mathcal{C}_{\Delta}^{2}}\int_{0}^{\mathcal{C}_{\Delta}% }\mathrm{d}x\,x\exp\left\{-\frac{1}{\beta}\left[x^{\beta}-1\right]\right\}\,,= divide start_ARG 1 end_ARG start_ARG caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT roman_d italic_x italic_x roman_exp { - divide start_ARG 1 end_ARG start_ARG italic_β end_ARG [ italic_x start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT - 1 ] } , (21)

and the projected concentration is 𝒞Δsubscript𝒞Δ\mathcal{C}_{\Delta}caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT.

Refer to caption
Figure 10: Fit to the integral function for β=0.3𝛽0.3\beta=0.3italic_β = 0.3 as a function of the projected concentration. The top panel plots Equation 21 as the black line while the dashed cyan curve is the resulting fit using Equation 22. The bottom panel plots the fit quality. The fit is accurate to 2% and better for a large range of projected concentrations.

The integral \mathcal{I}caligraphic_I cannot be solved analytically, but can be numerically integrated for a given value of 𝒞Δsubscript𝒞Δ\mathcal{C}_{\Delta}caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT and β𝛽\betaitalic_β. We provide a generalized fitting fitting function:

=c0exp[n(𝒞Δ+c1c2)1n].subscript𝑐0𝑛superscriptsubscript𝒞Δsubscript𝑐1subscript𝑐21𝑛\displaystyle\mathcal{I}=c_{0}\exp\left[-n\left(\frac{\mathcal{C}_{\Delta}+c_{% 1}}{c_{2}}\right)^{\frac{1}{n}}\right]\,.caligraphic_I = italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_exp [ - italic_n ( divide start_ARG caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG italic_n end_ARG end_POSTSUPERSCRIPT ] . (22)

Once fitted with β=0.3𝛽0.3\beta=0.3italic_β = 0.3, the best-fit parameters c0=5756subscript𝑐05756c_{0}=5756italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 5756, c1=0.628subscript𝑐10.628c_{1}=0.628italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0.628, c2=0.438subscript𝑐20.438c_{2}=0.438italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0.438, n=7.41𝑛7.41n=7.41italic_n = 7.41. The results are presented in Figure 10, where the black line is the above integral with a fixed β=0.3𝛽0.3\beta=0.3italic_β = 0.3 as a function of the projected concentration, while the red-dashed line is the resulting fit. The bottom panel plots the residuals between the integral and the analytical fit. The best fit parameters is accurate to similar-to\sim 2% and better for ranges of 𝒞Δ[1,100]subscript𝒞Δ1100\mathcal{C}_{\Delta}\in[1,100]caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ∈ [ 1 , 100 ]. While not shown here, Equation 22 also fits accurately for shapes with β[0.01,0.40]𝛽0.010.40\beta\in[0.01,0.40]italic_β ∈ [ 0.01 , 0.40 ], which can be applicable for halos with differing shapes. With above results, we arrive with the final form of the profile normalization as a function of mass and projected concentration:

Σ1subscriptΣ1\displaystyle\Sigma_{-1}roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT =eff2πrΔ2(𝒞Δ;β).absentsubscripteff2𝜋superscriptsubscript𝑟Δ2subscript𝒞Δ𝛽\displaystyle=\frac{\mathcal{M}_{\rm eff}}{2\pi\,r_{\Delta}^{2}\,\mathcal{I}(% \mathcal{C}_{\Delta};\beta)}\,.= divide start_ARG caligraphic_M start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_π italic_r start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT caligraphic_I ( caligraphic_C start_POSTSUBSCRIPT roman_Δ end_POSTSUBSCRIPT ; italic_β ) end_ARG . (23)

Note that this normalization can be connected to the virial definition of the halo by imposing an established effsubscripteff\mathcal{M}_{\rm eff}caligraphic_M start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPTMvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT relation.

A.3 Unit-convergence radius

The surface mass-density of a dark matter halo with a spherical ΣβsubscriptΣ𝛽\Sigma_{\beta}roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT profile (Equation 13) expressed in units of the critical surface density yields the dimensionless convergence profile:

κβ(R)=Σβ(R)/Σc=κ1exp{1β[(RR1)β1]}subscript𝜅𝛽𝑅subscriptΣ𝛽𝑅subscriptΣcsubscript𝜅11𝛽delimited-[]superscript𝑅subscript𝑅1𝛽1\displaystyle\kappa_{\beta}(R)=\Sigma_{\beta}(R)/\Sigma_{\rm c}=\kappa_{-1}% \exp\left\{-\frac{1}{\beta}\left[\left(\frac{R}{R_{-1}}\right)^{\beta}-1\right% ]\right\}italic_κ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R ) = roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R ) / roman_Σ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT = italic_κ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT roman_exp { - divide start_ARG 1 end_ARG start_ARG italic_β end_ARG [ ( divide start_ARG italic_R end_ARG start_ARG italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT - 1 ] } (24)

with κ1:=Σ1/Σcassignsubscript𝜅1subscriptΣ1subscriptΣc\kappa_{-1}:=\Sigma_{-1}/\Sigma_{\rm c}italic_κ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT := roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT / roman_Σ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT such that Σ1:=Σ(R1)assignsubscriptΣ1Σsubscript𝑅1\Sigma_{-1}:=\Sigma(R_{-1})roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT := roman_Σ ( italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ) and ΣcsubscriptΣc\Sigma_{\rm c}roman_Σ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is the critical surface density for lensing,

Σc:=c2Ds4πDDs,assignsubscriptΣcsuperscript𝑐2subscript𝐷𝑠4𝜋subscript𝐷subscript𝐷𝑠\displaystyle\Sigma_{\rm c}:=\frac{c^{2}D_{s}}{4\pi D_{\ell}D_{\ell s}}\,,roman_Σ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT := divide start_ARG italic_c start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_D start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT roman_ℓ italic_s end_POSTSUBSCRIPT end_ARG , (25)

such that Ds,Dsubscript𝐷𝑠subscript𝐷D_{s},\,D_{\ell}italic_D start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , italic_D start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT and Dssubscript𝐷𝑠D_{\ell s}italic_D start_POSTSUBSCRIPT roman_ℓ italic_s end_POSTSUBSCRIPT are the angular diameter distances between the observer from the source, the observer from the lens, and the lens from the source. The unit-convergence radius, Rκsubscript𝑅𝜅R_{\kappa}italic_R start_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT, at which the convergence κβ(Rκ)=1subscript𝜅𝛽subscript𝑅𝜅1\kappa_{\beta}(R_{\kappa})=1italic_κ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT ) = 1, is simply

Rκ=R1[βlog(κ1)+1]1/β.subscript𝑅𝜅subscript𝑅1superscriptdelimited-[]𝛽subscript𝜅111𝛽\displaystyle R_{\kappa}=R_{-1}\left[\beta\log\left(\kappa_{-1}\right)+1\right% ]^{1/\beta}\,.italic_R start_POSTSUBSCRIPT italic_κ end_POSTSUBSCRIPT = italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT [ italic_β roman_log ( italic_κ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ) + 1 ] start_POSTSUPERSCRIPT 1 / italic_β end_POSTSUPERSCRIPT . (26)

A.4 Deflection potential

A crucial component of modern lensing software is to analytically implement the lensing potential (and deflection angle) for structure in projection. Our projected mass profile is conveniently in a form similar to that of a two-dimensional Sersic (1968) light profile. Moreover, the lens properties for a Sersic (1968) profile are already derived in Cardone (2004), making it a straightforward procedure to map our parameters to those results. For completeness, we re-derive the lensing properties in a similar manner to Cardone (2004).

With the ΣβsubscriptΣ𝛽\Sigma_{\beta}roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT profile (Equation 13), it is convenient to introduce a dimensionless variable y:=(R/R1)βassign𝑦superscript𝑅subscript𝑅1𝛽y:=\left(R/R_{-1}\right)^{\beta}italic_y := ( italic_R / italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT. The two-dimensional lens potential, ψ(R)𝜓𝑅\psi(R)italic_ψ ( italic_R ), can be determined by solving the two-dimensional Poisson equation (e.g. Schneider et al. 1992):

2ψ(R)=2κβ(R),superscript2𝜓𝑅2subscript𝜅𝛽𝑅\displaystyle\nabla^{2}\psi(R)=2\,\kappa_{\beta}(R)\,,∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ ( italic_R ) = 2 italic_κ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R ) , (27)

where κβ(R)subscript𝜅𝛽𝑅\kappa_{\beta}(R)italic_κ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( italic_R ) is the convergence profile for our model defined previously in Equation 24. Expanding out the two-dimensional Laplacian on the left-hand side to polar coordinates, the following change of variables transforms the differential equation as

y2(11β)d2ψdy2+y(12β)dψdy=2R12κ1β2exp[1β(y1)].superscript𝑦211𝛽superscriptd2𝜓dsuperscript𝑦2superscript𝑦12𝛽d𝜓d𝑦2superscriptsubscript𝑅12subscript𝜅1superscript𝛽21𝛽𝑦1\displaystyle y^{2\left(1-\frac{1}{\beta}\right)}\frac{\mathrm{d}^{2}\psi}{% \mathrm{d}y^{2}}+y^{\left(1-\frac{2}{\beta}\right)}\frac{\mathrm{d}\psi}{% \mathrm{d}y}=\frac{2\,R_{-1}^{2}\,\kappa_{-1}}{\beta^{2}}\exp\left[-\frac{1}{% \beta}(y-1)\right]\,.italic_y start_POSTSUPERSCRIPT 2 ( 1 - divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ) end_POSTSUPERSCRIPT divide start_ARG roman_d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ end_ARG start_ARG roman_d italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_y start_POSTSUPERSCRIPT ( 1 - divide start_ARG 2 end_ARG start_ARG italic_β end_ARG ) end_POSTSUPERSCRIPT divide start_ARG roman_d italic_ψ end_ARG start_ARG roman_d italic_y end_ARG = divide start_ARG 2 italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_β start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG roman_exp [ - divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ( italic_y - 1 ) ] . (28)

The above equation has the solution

ψ(y)=ψ1y22βF2(2β,2β;1+2β,1+2β;yβ),𝜓𝑦subscript𝜓1subscriptsuperscript𝑦2𝛽2subscript𝐹22𝛽2𝛽12𝛽12𝛽𝑦𝛽\displaystyle\psi(y)=\psi_{-1}y^{\frac{2}{\beta}}\,_{2}F_{2}\left(\frac{2}{% \beta},\frac{2}{\beta};1+\frac{2}{\beta},1+\frac{2}{\beta};\frac{y}{\beta}% \right)\,,italic_ψ ( italic_y ) = italic_ψ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT italic_y start_POSTSUPERSCRIPT divide start_ARG 2 end_ARG start_ARG italic_β end_ARG end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_F start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( divide start_ARG 2 end_ARG start_ARG italic_β end_ARG , divide start_ARG 2 end_ARG start_ARG italic_β end_ARG ; 1 + divide start_ARG 2 end_ARG start_ARG italic_β end_ARG , 1 + divide start_ARG 2 end_ARG start_ARG italic_β end_ARG ; divide start_ARG italic_y end_ARG start_ARG italic_β end_ARG ) , (29)

where we have used the conditions that ψ0𝜓0\psi\rightarrow 0italic_ψ → 0 as y0𝑦0y\rightarrow 0italic_y → 0 or \infty, Fqp(a1,,ap;b1,,bp;z)subscriptsubscript𝐹𝑞𝑝subscript𝑎1subscript𝑎𝑝subscript𝑏1subscript𝑏𝑝𝑧{}_{p}F_{q}\left(a_{1},\dots,a_{p};b_{1},\dots,b_{p};z\right)start_FLOATSUBSCRIPT italic_p end_FLOATSUBSCRIPT italic_F start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_a start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ; italic_b start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_b start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ; italic_z ) is the generalized hypergeometric function (Gradshteyn & Ryzhik, 1980), and

ψ1subscript𝜓1\displaystyle\psi_{-1}italic_ψ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT :=R12κ12e1/βassignabsentsuperscriptsubscript𝑅12subscript𝜅12superscript𝑒1𝛽\displaystyle:=\frac{R_{-1}^{2}\kappa_{-1}}{2\,e^{1/\beta}}\,:= divide start_ARG italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_κ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_e start_POSTSUPERSCRIPT 1 / italic_β end_POSTSUPERSCRIPT end_ARG (30)

is the value of the potential for y=1𝑦1y=1italic_y = 1.

A.5 Scaled deflection angle

Our lens profile model imposes circular symmetry. Therefore, the deflection angle is purely radial and will have a magnitude α𝛼\alphaitalic_α given by dψ/dRd𝜓d𝑅\mathrm{d}\psi/\mathrm{d}Rroman_d italic_ψ / roman_d italic_R. The complete differentiation of the potential with respect to y𝑦yitalic_y gives

α(y)𝛼𝑦\displaystyle\alpha(y)italic_α ( italic_y ) =2α1y1β[Γ(1+2β)2βΓ[2β,yβ]],absent2subscript𝛼1superscript𝑦1𝛽delimited-[]Γ12𝛽2𝛽Γ2𝛽𝑦𝛽\displaystyle=2\,\alpha_{-1}\,y^{-\frac{1}{\beta}}\,\left[\Gamma\left(1+\frac{% 2}{\beta}\right)-\frac{2}{\beta}\,\Gamma\left[\frac{2}{\beta},-\frac{y}{\beta}% \right]\right]\,,= 2 italic_α start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT italic_y start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG italic_β end_ARG end_POSTSUPERSCRIPT [ roman_Γ ( 1 + divide start_ARG 2 end_ARG start_ARG italic_β end_ARG ) - divide start_ARG 2 end_ARG start_ARG italic_β end_ARG roman_Γ [ divide start_ARG 2 end_ARG start_ARG italic_β end_ARG , - divide start_ARG italic_y end_ARG start_ARG italic_β end_ARG ] ] , (31)

where Γ(x,a)Γ𝑥𝑎\Gamma(x,a)roman_Γ ( italic_x , italic_a ) is the incomplete Gamma function, Γ(a)Γ𝑎\Gamma(a)roman_Γ ( italic_a ) is the standard Gamma function, and

α1subscript𝛼1\displaystyle\alpha_{-1}italic_α start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT :=(1β)2βe1/ββR1κ1.assignabsentsuperscript1𝛽2𝛽superscript𝑒1𝛽𝛽subscript𝑅1subscript𝜅1\displaystyle:=\frac{\left(-\frac{1}{\beta}\right)^{-\frac{2}{\beta}}}{e^{1/% \beta}\,\beta}\,R_{-1}\,\kappa_{-1}\,.:= divide start_ARG ( - divide start_ARG 1 end_ARG start_ARG italic_β end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 2 end_ARG start_ARG italic_β end_ARG end_POSTSUPERSCRIPT end_ARG start_ARG italic_e start_POSTSUPERSCRIPT 1 / italic_β end_POSTSUPERSCRIPT italic_β end_ARG italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT . (32)

A.6 Lensing shear

The radial dependence of a spherically symmetric system can be written as

γ(y)=Σ¯(y)Σ(y)Σc,subscript𝛾𝑦¯Σ𝑦Σ𝑦subscriptΣc\displaystyle\gamma_{\ell}(y)=\frac{\overline{\Sigma}(y)-\Sigma(y)}{\Sigma_{% \rm c}}\,,italic_γ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ( italic_y ) = divide start_ARG over¯ start_ARG roman_Σ end_ARG ( italic_y ) - roman_Σ ( italic_y ) end_ARG start_ARG roman_Σ start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT end_ARG , (33)

where

Σ¯(y)=2y20ydyyΣ(y)¯Σ𝑦2superscript𝑦2superscriptsubscript0𝑦differential-dsuperscript𝑦superscript𝑦Σsuperscript𝑦\displaystyle\overline{\Sigma}(y)=\frac{2}{y^{2}}\int_{0}^{y}\mathrm{d}y^{% \prime}\,y^{\prime}\,\Sigma(y^{\prime})over¯ start_ARG roman_Σ end_ARG ( italic_y ) = divide start_ARG 2 end_ARG start_ARG italic_y start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_y end_POSTSUPERSCRIPT roman_d italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT roman_Σ ( italic_y start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) (34)

is the mean surface mass density inside the dimensionless radius y𝑦yitalic_y. For our lensing model, if we define y~:=yβ/βassign~𝑦superscript𝑦𝛽𝛽\tilde{y}:=y^{\beta}/\betaover~ start_ARG italic_y end_ARG := italic_y start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT / italic_β, the mean surface mass density is then

Σ¯β(y~)=2Σ1exp(1βlnββ)1y~2/βγ[2β,y~],subscript¯Σ𝛽~𝑦2subscriptΣ11𝛽𝛽𝛽1superscript~𝑦2𝛽𝛾2𝛽~𝑦\displaystyle\overline{\Sigma}_{\beta}(\tilde{y})=2\,\Sigma_{-1}\exp\left(% \frac{1-\beta\ln\beta}{\beta}\right)\frac{1}{\tilde{y}^{2/\beta}}\,\gamma\left% [\frac{2}{\beta},\,\tilde{y}\right]\,,over¯ start_ARG roman_Σ end_ARG start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ( over~ start_ARG italic_y end_ARG ) = 2 roman_Σ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT roman_exp ( divide start_ARG 1 - italic_β roman_ln italic_β end_ARG start_ARG italic_β end_ARG ) divide start_ARG 1 end_ARG start_ARG over~ start_ARG italic_y end_ARG start_POSTSUPERSCRIPT 2 / italic_β end_POSTSUPERSCRIPT end_ARG italic_γ [ divide start_ARG 2 end_ARG start_ARG italic_β end_ARG , over~ start_ARG italic_y end_ARG ] , (35)

which gives us the shearing quantity in its compact form:

γ(y)=κ1{exp(1βlnββ)2y~2/βγ[2β,y~]exp(y~+1)}.subscript𝛾𝑦subscript𝜅11𝛽𝛽𝛽2superscript~𝑦2𝛽𝛾2𝛽~𝑦~𝑦1\displaystyle\gamma_{\ell}(y)=\kappa_{-1}\left\{\exp\left(\frac{1-\beta\ln% \beta}{\beta}\right)\frac{2}{\tilde{y}^{2/\beta}}\,\gamma\left[\frac{2}{\beta}% ,\,\tilde{y}\right]-\exp\left(-\tilde{y}+1\right)\right\}\,.italic_γ start_POSTSUBSCRIPT roman_ℓ end_POSTSUBSCRIPT ( italic_y ) = italic_κ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT { roman_exp ( divide start_ARG 1 - italic_β roman_ln italic_β end_ARG start_ARG italic_β end_ARG ) divide start_ARG 2 end_ARG start_ARG over~ start_ARG italic_y end_ARG start_POSTSUPERSCRIPT 2 / italic_β end_POSTSUPERSCRIPT end_ARG italic_γ [ divide start_ARG 2 end_ARG start_ARG italic_β end_ARG , over~ start_ARG italic_y end_ARG ] - roman_exp ( - over~ start_ARG italic_y end_ARG + 1 ) } . (36)

Appendix B The impact of projection depth

A major component within the analysis done in this paper is how the mass is collected within a given a dark matter halo along the projection. Here, we treat the line-of-sight projection as a face-on-viewed cylinder, with a radius and length, \mathcal{R}caligraphic_R and \mathcal{L}caligraphic_L. The total mass within this volume is denoted at the effective mass, effsubscripteff\mathcal{M}_{\rm eff}caligraphic_M start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT, which is assumed to relate to Mvirsubscript𝑀virM_{\rm vir}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT in some way. In the main text, we chose the radius to be the size of the halo, =rvirsubscript𝑟vir\mathcal{R}=r_{\rm vir}caligraphic_R = italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT, while we chose the cylinder length to be proportional to the halo virial radius =2ξrvir2𝜉subscript𝑟vir\mathcal{L}=2\xi\,r_{\rm vir}caligraphic_L = 2 italic_ξ italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. The chosen value ξ𝜉\xiitalic_ξ plays a small part in inferring the effective mass. Indeed, as Figure 11 shows that the effective mass, based on our chosen cylinder volume, and the halo virial mass are roughly one-to-one, i.e., effMvirsimilar-to-or-equalssubscripteffsubscript𝑀vir\mathcal{M}_{\rm eff}\simeq M_{\rm vir}caligraphic_M start_POSTSUBSCRIPT roman_eff end_POSTSUBSCRIPT ≃ italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT for most chosen values of ξ𝜉\xiitalic_ξ. Notably, as ξ𝜉\xiitalic_ξ increases, the scatter (which has reduced Poisson noise) for the lower-mass halos increases.

While the effective mass is consistent for different values of ξ𝜉\xiitalic_ξ, that is only one part of the complete story, where fortunately, the the inferred projected concentration is not impacted that much either. Figure 12 shows very little variance in the normalization of the mass function for the corresponding values of ξ𝜉\xiitalic_ξ previously shown in Figure 11. These projected concentration functions are compiled using our sample of zoom simulations, which has varying initial conditions, under-dense regions, as compared to fully realized cosmological boxes, and a relation riddled Poisson noise; these results are to not to be taken as legitimate statistical relations and are plotted for demonstration purposes. The faded-solid lines depict the bin values of 𝒞virsubscript𝒞vir\mathcal{C}_{\rm vir}caligraphic_C start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT while the solid lines are power-law fits to these curves, as we want to emphasize the normalization values. As we increase to larger values of ξ𝜉\xiitalic_ξ, we see the normalizations remains relatively the same with slight adjustments of the slopes.

From both figures, it was decided to use a value of ξ=1.5𝜉1.5\xi=1.5italic_ξ = 1.5 as this is where the projected concentration and mass relation converges while also minimizing the scatter down at the low-mass range for the effective mass and halo mass relation (which otherwise large for ξ=2𝜉2\xi=2italic_ξ = 2).

Refer to caption
Refer to caption
Figure 11: The virial mass plotted against the effective (projected) mass in multiple depths of projections for our chosen cylinder geometry: =2ξrvir2𝜉subscript𝑟vir\mathcal{L}=2\xi\,r_{\rm vir}caligraphic_L = 2 italic_ξ italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT. The color codes map to the value of ξ𝜉\xiitalic_ξ as shown. Note that the cylinder radius is fixed to be the virial radius. From the top plot, we see that the two masses are roughly be same. The cyan curve is a one-to-one relationship. The bottom plot shows residuals. The shaded pink band indicates the region where the convergence radius of each halo is larger than R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT.
Refer to caption
Figure 12: Same ξ𝜉\xiitalic_ξ values from the previous figure, now plotting the implied projected concentration-mass relations. No changes of the normalization are present.

Appendix C Logarithmic slope of surface density profile

Here, we demonstrate that the behavior of the logarithmic slope for the projected density profile is captured by a radially-dependent power law profile that is simple in form:

dlogΣβdlogR=(RR1)β.dsubscriptΣ𝛽d𝑅superscript𝑅subscript𝑅1𝛽\displaystyle\frac{\mathrm{d}\log{\Sigma_{\beta}}}{\mathrm{d}\log{R}}=-\Bigg{(% }\frac{R}{R_{-1}}\Bigg{)}^{\beta}\,.divide start_ARG roman_d roman_log roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT end_ARG start_ARG roman_d roman_log italic_R end_ARG = - ( divide start_ARG italic_R end_ARG start_ARG italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_β end_POSTSUPERSCRIPT . (37)

Here, R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT is the radius where the logarithmic slope of the surface density is equal to 11-1- 1 and β𝛽\betaitalic_β is the shape parameters of that effectively tailors itself to any dark matter halo.

Refer to caption
Figure 13: The logarithmic slope of the surface-mass density profiles, Σ(R)Σ𝑅\Sigma(R)roman_Σ ( italic_R ), along three density axis. Presented are several example halos for mass bins of Mvir1011similar-to-or-equalssubscript𝑀virsuperscript1011M_{\rm vir}\simeq 10^{11}italic_M start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT ≃ 10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT (right-most panel), 1010superscript101010^{10}10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT (center panel), and 109Msuperscript109subscript𝑀direct-product10^{9}\ M_{\odot}10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT (left-most panel) for demonstration purposes. The three density axes for a single halo are is plotted within a single panel. A power-law radial dependence of the slope seems to fit the different density axes of our simulations very well. Our predictive model, ΣβsubscriptΣ𝛽\Sigma_{\beta}roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT, introduced in Equations 12 and 13, are shown as thin-solid lines for the best-fit parameters to ΣβsubscriptΣ𝛽\Sigma_{\beta}roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT with fixed β=0.3𝛽0.3\beta=0.3italic_β = 0.3.

This is shown in Figure 13, where we selected three representative halos within three of the higher-mass bins: 109, 1010,superscript109superscript101010^{9},\ 10^{10},10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT , and 1011Msuperscript1011subscript𝑀direct-product10^{11}\ M_{\odot}10 start_POSTSUPERSCRIPT 11 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT in the left-most, center, and right-most panel, respectively. The logarithmic slopes in each mass bin follows a similar shape out to rvirsubscript𝑟virr_{\rm vir}italic_r start_POSTSUBSCRIPT roman_vir end_POSTSUBSCRIPT along each density-axes projection, minus the sporadic-like behavior seen in the outer region. It is until the profiles approaches the inner-region where we begin to see differences between all three profiles. A clear deviation, found for all of these masses, between each density axes is seen around the region at which the log-slope is equal to 11-1- 1, which is marked by the dotted-gray horizontal line. Another important feature here is how each profile does not converge to some finite value. This would imply that the actual two-dimensional profile does not centrally converge to some power law.

The thinner solid lines in each panel of Figure 13 depicts Equation 12. The lines are determined from fitting the integrated profile of ΣβsubscriptΣ𝛽\Sigma_{\beta}roman_Σ start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT to the local surface density profiles with a fixed shape β=0.3𝛽0.3\beta=0.3italic_β = 0.3. The power-law profile is shown to reproduce the radial dependence fairly well within the three different density axes. Not that by eliminating the dependence of the shape parameter, β𝛽\betaitalic_β, for a given halo, a given orientation is parametrized by the projected scale radius, R1subscript𝑅1R_{-1}italic_R start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT, which in-turn captures the variation and scatter expected from halo-to-halo.