DES-2017-0268 FERMILAB-PUB-17-445-A Density split statistics: joint model of counts and lensing in cells

1, 2, 3, 4, 5, 3 6 3 7 3, 4 2, 1 O. Friedrich, ∗ D. Gruen, † J. DeRose, D. Kirk, E. Krause, T. McClintock, E. S. Rykoff, S. Seitz, R. H. Wechsler,5, 3, 4 G. M. Bernstein,8 J. Blazek,9, 10 C. Chang,11 S. Hilbert,12, 13 B. Jain,8 A. Kovacs,14 O. Lahav,6 F. B. Abdalla,6, 15 S. Allam,16 J. Annis,16 K. Bechtol,17 A. Benoit-Lévy,18, 6, 19 E. Bertin,18, 19 D. Brooks,6 A. Carnero Rosell,20, 21 M. Carrasco Kind,22, 23 J. Carretero,14 C. E. Cunha,3 C. B. D’Andrea,8 L. N. da Costa,20, 21 C. Davis,3 S. Desai,24 H. T. Diehl,16 J. P. Dietrich,12, 13 A. Drlica-Wagner,16 T. F. Eifler,25, 26 P. Fosalba,27 J. Frieman,16, 11 J. García-Bellido,28 E. Gaztanaga,27 D. W. Gerdes,29, 30 T. Giannantonio,31, 32, 1 R. A. Gruendl,22, 23 J. Gschwend,20, 21 G. Gutierrez,16 K. Honscheid,9, 33 D. J. James,34 M. Jarvis,8 K. Kuehn,35 N. Kuropatkin,16 M. Lima,36, 20 M. March,8 J. L. Marshall,37 P. Melchior,38 F. Menanteau,22, 23 R. Miquel,39, 14 J. J. Mohr,13, 12, 2 B. Nord,16 A. A. Plazas,26 E. Sanchez,40 V. Scarpine,16 R. Schindler,4 M. Schubnell,30 I. Sevilla-Noarbe,40 E. Sheldon,41 M. Smith,42 M. Soares-Santos,16 F. Sobreira,43, 20 E. Suchyta,44 M. E. C. Swanson,23 G. Tarle,30 D. Thomas,45 M. A. Troxel,9, 33 V. Vikram,46 and J. Weller13, 2, 1 (DES Collaboration) 1Universitäts-Sternwarte, Fakultät für Physik, Ludwig-Maximilians Universität München, Scheinerstr. 1, 81679 München, Germany 2Max Institute for Extraterrestrial Physics, Giessenbachstrasse, 85748 Garching, Germany 3Kavli Institute for Particle Astrophysics & Cosmology, P. O. Box 2450, Stanford University, Stanford, CA 94305, USA 4SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA 5Department of Physics, Stanford University, 382 Via Pueblo Mall, Stanford, CA 94305, USA 6Department of Physics & Astronomy, University College London, Gower Street, London, WC1E 6BT, UK 7Department of Physics, University of Arizona, Tucson, AZ 85721, USA 8Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104, USA 9Center for Cosmology and Astro-Particle Physics, The Ohio State University, Columbus, OH 43210, USA 10Institute of Physics, Laboratory of Astrophysics, École Polytechnique Fédérale de Lausanne (EPFL), Observatoire de Sauverny, 1290 Versoix, Switzerland 11Kavli Institute for Cosmological Physics, University of Chicago, Chicago, IL 60637, USA 12Faculty of Physics, Ludwig-Maximilians-Universität, Scheinerstr. 1, 81679 Munich, Germany 13Excellence Cluster Universe, Boltzmannstr. 2, 85748 Garching, Germany 14Institut de Física d’Altes Energies (IFAE), The Barcelona Institute of Science and Technology, Campus UAB, 08193 Bellaterra (Barcelona) Spain 15Department of Physics and Electronics, Rhodes University, PO Box 94, Grahamstown, 6140, South Africa 16Fermi National Accelerator Laboratory, P. O. Box 500, Batavia, IL 60510, USA 17LSST, 933 North Cherry Avenue, Tucson, AZ 85721, USA 18CNRS, UMR 7095, Institut d’Astrophysique de Paris, F-75014, Paris, France 19Sorbonne Universités, UPMC Univ Paris 06, UMR 7095, Institut d’Astrophysique de Paris, F-75014, Paris, France 20Laboratório Interinstitucional de e-Astronomia - LIneA, Rua Gal. José Cristino 77, Rio de Janeiro, RJ - 20921-400, Brazil 21Observatório Nacional, Rua Gal. José Cristino 77, Rio de Janeiro, RJ - 20921-400, Brazil 22Department of Astronomy, University of Illinois, 1002 W. Green Street, Urbana, IL 61801, USA 23National Center for Supercomputing Applications, 1205 West Clark St., Urbana, IL 61801, USA 24Department of Physics, IIT Hyderabad, Kandi, Telangana 502285, India 25Department of Physics, California Institute of Technology, Pasadena, CA 91125, USA 26Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Dr., Pasadena, CA 91109, USA 27Institute of Space Sciences, IEEC-CSIC, Campus UAB, Carrer de Can Magrans, s/n, 08193 Barcelona, Spain 28Instituto de Fisica Teorica UAM/CSIC, Universidad Autonoma de Madrid, 28049 Madrid, Spain arXiv:1710.05162v4 [astro-ph.CO] 30 Jul 2021 29Department of Astronomy, University of Michigan, Ann Arbor, MI 48109, USA 30Department of Physics, University of Michigan, Ann Arbor, MI 48109, USA 31Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 32Kavli Institute for Cosmology, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 33Department of Physics, The Ohio State University, Columbus, OH 43210, USA 34Astronomy Department, University of Washington, Box 351580, Seattle, WA 98195, USA 35Australian Astronomical Observatory, North Ryde, NSW 2113, Australia 36Departamento de Física Matemática, Instituto de Física, Universidade de São Paulo, CP 66318, São Paulo, SP, 05314-970, Brazil 37George P. and Cynthia Woods Mitchell Institute for Fundamental Physics and Astronomy, and Department of Physics and Astronomy, Texas A&M University, College Station, TX 77843, USA 38Department of Astrophysical Sciences, Princeton University, Peyton Hall, Princeton, NJ 08544, USA 39Institució Catalana de Recerca i Estudis Avançats, E-08010 Barcelona, Spain 2

40Centro de Investigaciones Energéticas, Medioambientales y Tecnológicas (CIEMAT), Madrid, Spain 41Brookhaven National Laboratory, Bldg 510, Upton, NY 11973, USA 42School of Physics and Astronomy, University of Southampton, Southampton, SO17 1BJ, UK 43Instituto de Física Gleb Wataghin, Universidade Estadual de Campinas, 13083-859, Campinas, SP, Brazil 44Computer Science and Mathematics Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831 45Institute of Cosmology & Gravitation, University of Portsmouth, Portsmouth, PO1 3FX, UK 46Argonne National Laboratory, 9700 South Cass Avenue, Lemont, IL 60439, USA (Dated: August 3, 2021) We present density split statistics, a framework that studies lensing and counts-in-cells as a function of fore- ground galaxy density, thereby providing a large-scale measurement of both 2-point and 3-point statistics. Our method extends our earlier work on trough lensing and is summarized as follows: given a foreground (low red- shift) population of galaxies, we divide the sky into subareas of equal size but distinct galaxy density. We then measure lensing around uniformly spaced points separately in each of these subareas, as well as counts-in-cells statistics (CiC). The lensing signals trace the matter density contrast around regions of fixed galaxy density. Through the CiC measurements this can be related to the density profile around regions of fixed matter density. Together, these measurements constitute a powerful probe of cosmology, the skewness of the density field and the connection of galaxies and matter. In this paper we show how to model both the density split lensing signal and CiC from basic ingredients: a non-linear power spectrum, clustering hierarchy coefficients from perturbation theory and a parametric model for galaxy bias and shot-noise. Using N-body simulations, we demonstrate that this model is sufficiently accu- rate for a cosmological analysis on year 1 data from the Dark Energy Survey.

I. INTRODUCTION sarily mean that higher-order statistics are better than 2-point statistics in discriminating between particular choices of cos- The large-scale structure (LSS) observed today is thought mological parameters [3]. But they scale differently with pa- to originate from almost perfectly Gaussian density pertur- rameters such as Ωm, σ8, galaxy bias and galaxy stochastic- bations in the early Universe. This means that there was a ity than their 2-point counterparts. Hence, in a cosmologi- complete symmetry in the abundance and amplitude of under- cal analysis that varies a large number of parameters, probes dense and overdense regions in very early times. Gravitational that are sensitive to both 2-point and higher order statistics attraction then caused initial overdensities to collapse to small have the power to break degeneracies between these parame- but highly overdense structures such as galaxy clusters, while ters [11, 63, 76, 80]. initial underdensities expanded but stayed moderately under- Observations of higher-order statistical features of the den- dense and e.g. became voids. As a consequence the majority sity field include measurements of three point correlation of the volume in the late-time Universe is underdense, com- functions [68], shear peak statistics [49, 52, 53] and the clus- pensated by the presence of few highly overdense spots. Or, in ter mass function [58]. Also, a number of probes have been other words, a positive skewness in the distribution of density suggested (and in some cases measured in data) that study fluctuations emerges due to gravitational collapse. the correlation of 2-point statistics and background density. A variety of probes have been used to study the statisti- Chiang et al. [19] have measured this by means of the inte- cal properties of the late-time density field and to thereby un- grated bispectrum. Simpson et al. [70, 71, 72] have proposed derstand the physics of gravitational collapse as well as the a clipped power spectrum approach, where 2-point statistics processes responsible for the properties of the initial den- are measured on the sky after excluding high density regions. sity fluctuations. So far, the most extensive studies have They have shown that these measurements contain informa- been carried out on the 2-point statistics of density fluctu- tion complementary to the corresponding measurements on ations, i.e. on measuring the variance of density fluctu- the full sky. ations as a function of scale. This has e.g. been done A similar direction was investigated by Gruen et al. [39] through measurements of cosmic shear 2-point correlation who separately measured the lensing power spectrum in un- functions [e.g. 6, 8, 37, 42, 47, 50, 67, 79, 82, 85], galaxy derdense and overdense lines of sight. The framework pre- clustering [e.g. 1, 28, 34] and galaxy-galaxy lensing [e.g. sented in this paper is based on their concept of trough lens- 16, 17, 20, 44, 65, 84, 88] as well as combined measurements ing. We will call it density split lensing when only lensing thereof [e.g. 29, 57, 89]. measurements are involved and density split statistics when it While 2-point statistics are only sensitive to the overall is combined with counts-in-cells measurements. This method amplitude of density fluctuations, higher-order statistics also can be summarized as follows: we consider a foreground (low know about the skewness arising from the different behaviour redshift) population of galaxies and smooth their position field of underdense and overdense regions. This does not neces- with a circular top-hat aperture. This smoothed density field is then used to divide the sky into sub-areas of equal size but dis- tinct galaxy density. In this paper we consider in particular 5 sub-areas and call them quintiles of galaxy density. As a next ∗ Corresponding author: [email protected] step, we use a background (high redshift) population of galax- † Einstein Fellow ies to measure the tangential shear of these galaxies around a 3

set of uniformly spaced points within the area of each density our modeling of this signal (but postponing technical details quintile. The resulting lensing signals trace the matter density of this model to section IV). In section II C we provide fore- contrast around regions of fixed foreground galaxy density. casts on the cosmological information that can be obtained This data vector is then complemented by the histogram of with a measurement of density split statistics in year-1 data of counts-in-cells of the foreground galaxies to pin down their the Dark Energy Survey (DES Y1). bias and stochasticity. As we show in this paper, a cosmo- logical analysis based on this density split data vector has a number of desirable features: A. Measuring density split statistics it allows for an accurate analytic modeling with the help Density split lensing is a generalization of trough lensing •of cosmological perturbation theory and a non-linear power [39] and can be described in three steps: spectrum, it yields high signal-to-noise measurement, 1.) Splitting the sky into quantiles of different foreground • it avoids systematics common to cosmic shear such as ad- galaxy density •ditive shear biases or intrinsic alignments (as long as tracer sample and source sample do not overlap in redshift), Consider a sample of low-redshift galaxies that are tracing the line-of-sight density of matter with some redshift distribution it has a very intuitive interpretation. • nl(z). We will call these galaxies the foreground sample. For This paper is a companion paper of Gruen et al. [40], where an angular radius θT , which we will call the top-hat aper- we present our actual data analysis, including tests for sys- ture radius, we define NT (ˆn) to be the number of galaxies tematic effects as well as a description of how we estimate the found within a radius θT around the point on the sky specified covariance of our signal. This paper is presenting the mod- by the vector ˆn on the unit sphere. The field NT (ˆn) can be eling framework used in that analysis. Our section II gives a used to divide the sky into regions of different galaxy density. general overview of density split statistics: we describe our Gruen et al. [39] have done this by discretizing the sky with a 1 data vector, explain how it can be modelled and also present HEALPIX grid and sorting the pixels according to their value forecasts on cosmological parameter constraints, both for a of NT (ˆn). Then they considered the 20% of the pixels with ΛCDM model and an extended model that allows gravitational the lowest values of NT , calling them troughs. In the limit collapse to behave differently than within general relativity. In of a fine pixelization these pixels can be considered the most section III we describe the simulated data used in this work. underdense quintile of the sky area. This can be generalized Section IV explains details of the model presented in sec- to the second most underdense quintile, the third most under- tion II. There we also compare individual components of this dense quintile etc. or even to finer splits using more then just model directly to measurements in N-body simulations. In 5 quantiles. section V we show that our model for a data vector combining We stick to a division into 5 quantiles (quintiles) through- density split lensing and counts-in-cells statistics is accurate out this paper. The upper panel of figure 1 illustrates such a enough to recover the cosmology underlying our N-body sim- subdivision on a patch of a simulated sky (from the Buzzard ulations. Any possible deviation between our model and the flock, see section III and especially DeRose et al. [30] for de- simulations is shown to be well within statistical uncertainties tails). There we use a top-hat aperture radius θT = 200 and of year 1 data of the Dark Energy Survey (DES Y1). the tracer galaxies have the redshift distribution that is dis- In appendix A, we review a number of differential equa- played by the solid line in Figure 2. Figure 1 shows the most tions that govern gravitational collapse. In appendix B, we underdense quintile of the simulated patch in cyan, the most review the leading order perturbative calculation of the 3- overdense quintile in red and the three intermediate quintiles point statistics of the cosmic density field for a general ΛCDM in blue, green and orange. model. Appendix C qualitatively compares our model of the Note that the sum of the 5 lensing signals will vanish on cosmic density PDF to a second set of N-body simulations average (we subtracted 1/5 times the shear around random as a complement to the comparison carried out in the main points from each signal, but due to boundary effects their text. Appendix D derives properties of joint log-normal ran- sum will not vanish exactly). This means that roughly 4 dom fields and appendix E repeats the validation of our model of the 5 signals contain independent information. We have for an alternative shot-noise parametrization. not investigated, whether our choice of 5 quantiles is in any way optimal. Choosing 3 quantiles would leave us with 2 independent signals and would hence suffice to be sensitive II. DENSITY SPLIT STATISTICS: DATA VECTOR, to both the variance and skewness of the density field. 5 MODELING AND FORECASTS quantiles enable a sentitivity beyond the 3rd moment of the density field. And they also allow us to explicitly show, that This section provides an introduction to the program of den- the median universe is underdense (which we could not do sity split statistics. In section II A we describe how we ob- tain the density split lensing signal and how this signal can be further complemented with information on galaxy bias and stochasticity from counts-in-cells. In section II B we outline 1 See Górski et al. [38] for details on HEALPIX. 4

RA +80◦ +70◦ +60◦ +50◦ +40◦ +30◦ +20◦ +10◦ 30◦ −

40◦ − dec

50◦ − quintile 5 quintile 1

3.5 pert. theory 1.5 pert. theory 3.0 Gaussian Gaussian mean of 4 1.0 mean of 4 Buzzard 2.5 Buzzard + + DES Y1 errors DES Y1 errors 2

3 0.5 10 2.0 · 10 ) · t N 1.5 0.0 γ P(

1.0 0.5 − N 0.5 mean 1.0 − 0.0 20 40 60 80 101 102 N θ [arcmin]

FIG. 1. Top panel: splitting the lines of sight in one DES-Y1 like Buzzard simulation into 5 quantiles of galaxy density (color coding from cyan, most underdense, to red, most overdense). The map uses a 20 arcmin top-hat radius and REDMAGIC galaxies with a redshift range of 0.2 . z . 0.45. Bottom left: histogram of REDMAGIC galaxy counts in 20 arcmin radii (counts-in-cells). We show the mean histogram from 4 Buzzard realisations of DES-Y1 (black points), our model based on perturbation theory and cylindrical collapse (solid line) and a model that assumes the projected density contrast to be a Gaussian random field (dotted line). The color coding corresponds exactly to the density quantiles in the top panel. Bottom right: Lensing signals around random points split by the density quantile in which these points are located. We show the mean measurement from 4 Buzzard realisations (black points), our perturbation theory model (solid line) and a model that assumes projected density contrast and lensing convergence to be joint Gaussian random variables (dotted line). Color coding is the same as in the other panels. The asymmetry between the lensing signals around the most underdense and most overdense lines-of-sight indicates the skewness of the cosmic density PDF. 5

around the 20% most underdense pixels is lower than the am- 7 redshift distributions redMaGiC (lenses) plitude of the tangential shear around the 20% most overdense in Buzzard simulations source bin 1 pixels. This is reflecting the skewness of the cosmic density 6 source bin 2 PDF. Secondly, the signal around points in the third quintile is still significantly negative, which reflects the fact that the me- 5 dian universe is underdense. A more subtle feature is the fact that the underdense signals fall off less rapidly with increasing

) 4 scale than the overdense signals. This is because on large z

( scales the density field becomes Gaussian and hence recovers p 3 its initial symmetry between overdensities and underdensities.

2 3.) Measuring the average counts-in-cells in each density quintile to obtain additional information on galaxy bias and 1 stochasticity

0 If galaxy counts and the matter density field were perfectly correlated, then a split of the sky by galaxy density would be 0.0 0.5 1.0 1.5 identical to a split by matter density. Hence, in this limit the z density split lensing signals would be independent of the bias of the tracer galaxies. In a realistic scenario however, shot- FIG. 2. Redshift distributions of the tracer galaxy sample and the noise of the galaxies smears out our attempts to divide the sky source samples of our N-body realizations of DES-Y1. into areas of different matter density. Hence the density split lensing signals obtain a dependence on galaxy bias, but also on galaxy stochasticity. Increasing the linear bias of galaxy with 4 quantiles). In section IV we investigate different radii clustering will sharpen the tracers’ ability to distinguish be- of our top-hat aperture and find that θT = 200 is the smallest tween overdensities and underdensities. Thus, increasing this radius at which our model is reliable (given the redshift bias will increase the amplitudes of the signals. This means distribution we use in Gruen et al. [40]). that linear bias is to some degree degenerate with the ampli- tude of density fluctuations, σ8. But σ8 and bias influence the 2.) Tracing the mean dark matter density in each sky quintile third moments of the density field in different ways and their with gravitational lensing degeneracy is not complete. As a consequence, it is possi- ble to obtain constraints on cosmological parameters from the Now consider a second sample of galaxies at higher redshifts lensing signals alone (cf. section II C and the blue contour in than the foreground sample (the source sample, see e.g. the the left panel of figure 4). dashed and dotted redshift distributions in Figure 2). As But additional information on bias and stochasticity never- the light of these galaxies passes the large-scale structure of theless helps to tighten these constraints. In this paper we the foreground density distribution it undergoes gravitational decided to add that information in the form of normalized lensing effects such as gravitational shear [see e.g. 4]. The quantiles of the counts-in-cells (CiC) histogram of the tracer density split lensing signal around each quintile of the sky is galaxies: we measure the histogram of tracer counts within obtained by measuring the stacked radial profile of tangen- the same aperture that was used to identify our density quin- tial shear around random points located within that quintile. tiles. Then we identify the parts of this histogram that corre- These points are constrained to lie within the part of the sky spond to these quintiles (cf. lower left panel of Figure 1). For covered by a certain quintile of galaxy density but are other- each quintile q we then determine the mean galaxy count in wise random in their location. Because these random points that quintile, Nq, and normalize it by the overall mean galaxy ¯ ¯ are split according to the density quintile they are located in, count in our aperture, N. i.e. for each quintile we add Nq/N their stacked shear signals trace the average profile of density to our data vector. This indeed helps to tighten constraints on contrast around each quintile. cosmological parameters (cf. section II C and the green con- tour in the left panel of figure 4). In the lower right panel of Figure 1 we show the signals measured for each density quintile in our mock data. The points show the average measurement from 4 Buzzard realisations of DES year-1 data (using the highest redshift B. Modelling density split statistics source population shown in Figure 2) and the solid lines show predictions by the model presented in this paper. The error We now outline a general framework for modeling the data bars are derived from a set of log-normal realisations (using vector described above, leaving details of this framework to the FLASK tool by Xavier et al. [86]; in Gruen et al. [40] we section IV. Unless stated differently, we will assume a flat describe in detail how we configured FLASK to generate our ΛCDM universe throughout this paper. mock catalogs). Two main features of the density split lensing Let us start by introducing the quantities whose relations signals are apparent: first, the amplitude of the radial shear need to be modelled. First, we denote with δm,2D the line- 6

of-sight projection of the 3D density contrast according to the value κ<θ δm,T ? The tangential shear profile around that h | i redshift distribution nl(z) of our foreground galaxy sample, line-of-sight can then be inferred from the convergence profile i.e. as Z γt(θ) δm,T = κ<θ δm,T κθ δm,T δm,2D(nˆ) = dw ql(w) δm,3D(wnˆ, w) (II.1) h | i h | i − h | i cos θ 1 d = − κ<θ δm,T , (II.6) sin θ dθ where nˆ denotes a unit vector on the sky, w is co-moving dis- h | i κ θ tance and the projection kernel ql(w) is given in terms of nl(z) where θ is the average convergence at the radius . as The first of the above questions can be answered in the form δ N dz[w] of a conditional PDF of m,T given a certain value of T , i.e. ql(w) = nl(z[w]) . (II.2) p(δm,T NT ). Using Bayes’ theorem this can be written as dw | P (NT δm,T ) p(δm,T ) We furthermore define δm,T to be the average of δm,2D over p(δm,T NT ) = | , (II.7) | P (NT ) top-hat filters with aperture radius θT , i.e. where P (NT δm,T ) is the probability of finding a number of Z δ n | m,2D(ˆ0) galaxies NT given that the density contrast is δm,T and where δm,T (nˆ) = dΩ0 . (II.3) 2π(1 cos θT ) p(δm,T ) and P (NT ) are the total PDF of δm,T and the total 0 nˆ,nˆ <θT − | | probability of finding NT tracer galaxies. The average conver- gence profile around a circle with N galaxies is then given Here , denotes the angular distance between two points on T by the sky.|· ·| We identify regions of different density by means of our Z κ<θ NT = dδm,T κ<θ δm,T ,NT p(δm,T NT ) foreground galaxy sample. When smoothed with a top-hat h | i h | i | filter of radius θT , these galaxies are biased and possibly Z dδm,T κ<θ δm,T p(δm,T NT ) , (II.8) stochastic tracers of δm,T (nˆ). Hence our model also needs ≈ h | i | to include a description of how NT (nˆ), the number of tracer where in the second step we have assumed that the expected galaxies found within an angular radius θT around the line-of- convergence within θ only depends on the total matter density sight nˆ, relates to δm,T (nˆ). Finally, in order to describe the density split lensing sig- contrast within θT . nal, we need to consider the lensing convergence field for our We now divide the sky into different quintiles of tracer population of source galaxies. Given the source redshift dis- galaxy density. Let us denote with Q[0.0, 0.2] the 20% of the lines-of-sight on the sky that have the lowest value of NT . tribution ns(z), the convergence κ is given by the line-of-sight projection There will be a maximal value NT Nmax in this quin- tile and the stacked convergence profile≤ around these lines- Z of-sight is given by κ(nˆ) = dw Ws(w) δm,3D(wnˆ, w) , (II.4) κ<θ Q[0.0, 0.2] = h | i ! where Ws is the lensing efficiency, which is defined by 1 X P (N) κ<θ NT = N + α κ<θ NT = Nmax . 2 Z 0.2 h | i h | i 3Ω H ∞ w(w w) N

and further details). To model this function we use a cylindri- cal collapse approach based on the work of Bernardeau [9], which is also straightforward to generalize to other quantiles Bernardeau & Valageas [10] and Valageas [81]. Q[qmin, qmax]. The probabilities P (N) can be computed from the normal- (ii) Baseline model for κ<θ δm,T : ization of equation II.7. Hence, our model for the density split h | i lensing signal needs the following three ingredients: In our fiducial model we assume that κ<θ can be expressed as the sum of two random variables, (i) The PDF of matter density contrast, smoothed with a top- κ<θ = κ<θ,corr. + κ<θ,uncorr. , (II.15) hat filter of radius θT ,

where κ<θ,uncorr. is assumed to be completely uncorrelated to p(δm,T ) . (II.12) δm,T and hence doesn’t contribute to the density split lensing (ii) The expectation value of convergence inside a distance θ signal. As a result we have given the density contrast inside θ , T κ<θ δm,T κ<θ,corr. δm,T . (II.16) h | i ≡ h | i κ<θ δm,T . (II.13) We assume a joint log-normal PDF for the two ran- h | i dom variables δm,T and κ<θ,corr.. The expectation value (iii) The distribution of galaxy counts inside the top-hat radius κ<θ,corr. δm,T is then fixed by specifying the moments θT given the density contrast within that radius, h | i 2 3 δm,T , δm,T (II.17) P (NT δm,T ) . (II.14) h i h i | as well as Gruen et al. [39] assumed δm,T and κ<θ to have a joint Gaus- κ<θδm,T κ<θ,corr.δm,T (II.18) sian distribution. This allowed them to compute (i) and (ii) h i ≡ h i solely from the dark matter clustering power spectrum. To and compute (iii) they assumed a linear galaxy bias and Poisso- 2 2 nian shot-noise of the tracer galaxies. These assumptions al- κ<θδm,T κ<θ,corr.δm,T . (II.19) lowed a sufficient model for their measurements made on DES h i ≡ h i Science Verification data. But as can be seen from the dotted Second order moments are again computed from our non- lines in the lower panels of Figure 1, a Gaussian model for the linear power spectrum while third order moments are inferred density PDF is not sufficient within the much smaller uncer- from perturbation theory. (The introduction of κ<θ,uncorr. is tainty of DES-Y1. Also, in section IV we demonstrate that only necessary for consistency reasons: a joint log-normal the shot-noise of the tracer galaxies can not necessarily be as- PDF of δm,T and κ<θ characterized by the above moments would be inconsistent with the variance κ2 derived from sumed to be Poissonian. In this work we hence want to revise h <θi their model. our power spectrum.) For each of the model components (i) and (ii) we investi- Alternative model for κ<θ δm,T : gate two different modeling approaches - a baseline approach h | i and an approach with increased complexity used to check the As an alternative we compute κ<θ δm,T from the joint cu- h | i validity of the baseline model. In the following we briefly mulant generating function of the variables δm,T and κ<θ. outline each approach. The reader interested in details of each This function can also be modelled in a cylindrical collapse modeling ansatz is referred to section IV. Readers who are not approach. interested in this technical part of the paper should feel free to For model component (iii) we also investigate two different skip section IV. modeling approaches - one ansatz introducing 2 free param- eters and one ansatz introducing 3 free parameters. We find (i) Baseline model for p(δm,T ): that our simulated tracer catalogs are well described by the 2- In our fiducial model we assume δ to be a log-normal m,T parametric model. But - anticipating real galaxies to behave random field [41]. The PDF of such a variable can e.g. be more complicated than simulated ones - we do not consider ei- fixed by specifying the variance δ2 and skewness δ3 . m,T m,T ther of these models as our baseline model and instead apply We predict the variance of δ h fromi the the non-linearh i m,T both approaches to DES data in Gruen et al. [40]. We sum- matter power spectrum [cf. 39]. The latter is computed using marize both ansatzes here, but the interested reader is again halofit [73, 77] and an analytic transfer function [33]. We then referred to section IV for details of each ansatz. use leading order perturbation theory to compute a scaling

relation between the bispectrum and the power spectrum of (iii) Model 1 for P (NT δm,T ): the density field. Together with our power spectrum this fixes | In our fiducial model we introduce an auxilliary field δg,T the skewness of δm,T . such that our foreground galaxies are Poissonian tracers of that field, i.e. Alternative model for p(δm,T ): N As an alternative we compute the PDF p δ from its N¯(1 + δ ) T ( m,T ) g,T N¯(1+δg,T ) P (NT δg,T ) = e− . (II.20) cumulant generating function (see section IV for a definition | NT ! 8

δg,T can be thought of as a smooth (shot noise free) galaxy But in the parameter forecasts and likelihood contours shown density contrast. We then assume that δg,T and δm,T are joint in the following, we exclude bins with scales θT . This log-normal random variables with reduces all lensing signals to 17 log-spaced angular≤ bins with

2 2 2 3 3 3 δg,T = b δm,T , δg,T = b δm,T (II.21) θ < . h i h i h i h i 200 . 6000 (II.27) and The scales with θ < θT are excluded from our analysis for a 2 number of reasons: first, the signal-to-noise ratio of our lens- δg,T δm,T = br δm,T . (II.22) h i h i ing profiles drops quickly below θT (cf. Figure 1). Second, we The parameters b and r will be called galaxy bias and galaxy chose our fiducial value of θT = 200 because we still trust the stochasticity and are free parameters of the model. modeling of the density PDF described in section IV on this scale, and we do not want smaller angular scales to contribute Model 2 for P (NT δm,T ): to our fiducial data vector. Third, the approximation made in | As an alternative we assume P (NT δm,T ) to be a generaliza- equation II.8 might fail at scales smaller than our aperture. tion of the Poisson distribution, that| allows for The sum of all 5 lensing signals will be very close to zero

2 2 (though not exactly zero because of masking effects), so that N δm,T = NT δm,T + NT δm,T , (II.23) h T | i 6 h | i h | i they are not an independent set of observations. Hence, we i.e. for a shot-noise that is either enhanced or suppressed wrt. only include the 2 most overdense and the 2 most underdense the Poisson case. The enhancement of shot-noise is also al- quintiles in our analysis, i.e. the cyan, blue, orange and red lowed to be a function of δ of (approximately) the form lines in Figure 1. For the same reason, we also use the nor- m,T malized mean galaxy count only in four of the five quintiles of the counts-in-cells histogram to complement the lensing mea- 2 2 NT δm,T NT δm,T surement. The total number of data points in our data vector h | i − h | i α0 + α1δm,T . (II.24) NT δm,T ≈ is thus h | i

This model introduces an alternative bias parameter ˜b such Ndata = Nlens + NCiC that = [Nquant. 1] Nsource Nang + [Nquant. 1] − · · − NT δm,T = N¯[1 + ˜bδm,T ] . (II.25) = 4 2 17 + 4 h | i · · = 140 . (II.28)

For the model components (i) and (ii) and on the scales con- 2. Model parameters and forecast on constraining power sidered in this paper, the baseline and alternative modeling ap- proaches yield almost indistinguishable predictions (cf. sec- We now investigate what constraints on model parameters tion IV). For component (iii) the modeling ansatzes 1 and 2 can be expected from the above data vector when measured are not necessarily identical, because they introduce a differ- in DES-Y1 data. For this, we assume the optimistic case that ent number of degrees of freedom. Figure 1 as well as all component (iii) of our model is sufficiently described by two parameter contours shown in this paper are using the baseline parameters, i.e. by modeling approach 1 in the previous sec- model for components (i) and (ii) and ansatz 1 for compo- tion.2 This means that our model is determined by the follow- nent (iii). The predictions derived from different modelling ing 7 parameters: approaches are however almost indistinguishable (cf. figure 5). 1.) Ωm: present total matter density in units the critical density of the universe,

2.) σ8: amplitude of present day density fluctuations in spheres C. Data vector and forecasts on parameter constraints of 8Mpc/h radius as predicted by the linear power spectrum, 3.) Ω : present density of baryonic matter in units of the criti- 1. Binning and scales b cal density of the universe,

Throughout this paper, we assume that one sample of tracer 4.) ns: the spectral index of the primordial power spectrum, galaxies is used to identify density quintiles and that the lens- 5.) h100: the present day Hubble parameter in units of ing profiles around these quintiles are measured with two 100Mpc/(km/s), source redshift bins (cf. redshift distributions in Figure 2). To 6.) b: linear bias of the tracers w.r.t. matter density (cf. equa- identify the different density quintiles, we use a top-hat filter tion II.21), with fiducial smoothing radius of θT = 200. We measure the density split lensing signal in 24 log-spaced angular bins with 2 Gruen et al. [40] will also consider modeling approach 2. 50 < θ < 6000 . (II.26) 9

1.4 1.0 Forecast: constraints DES Y1 Forecast: constraints DES Y1 lensing only 1.3 on σ8 and Ωm 0.8 on Σ8 and hierarchical DES Y1 + Planck DES Y1 DES Y5 keeping ∆S3 0 lensing + CIC clustering coeff. S3 1.2 ≡ 0.6 DES Y5 + Planck 1.1

3 0.4 /S 8 3

σ 1.0 S 0.2 ∆ 0.9 0.0 0.8 0.2 − 0.7 0.4 0.6 − 0.1 0.2 0.3 0.4 0.5 0.35 0.40 0.45 0.50

Ωm Σ8 = σ8√Ωm

FIG. 3. Left panel: Forecast of 1σ and 2σ constraints on Ωm and σ8 achievable with density split lensing and counts-in-cells in DES Y1 data. The constraints are marginalized over Ωb, ns, h100, REDMAGIC galaxy bias b and galaxy-matter correlation coefficient r. For the parameters Ωb, ns, h100 we have assumed the same flat priors as used in the DES Y1 combined probes analysis presented in DES Collaboration [29]. hδ3i Right panel: ∆S3/S3 measures relative deviations from our fiducial perturbation theory prediction of the scaling coefficient S3 ≡ hδ2i2 (cf. equation II.30 and section IV for details). It can hence be thought of as the Bispectrum amplitude. We show 1σ constraints on this parameter achievable with density split lensing and counts-in-cells in DES data alone (solid lines) and using additional information on cosmology from Planck (dashed lines, no lensing). The sharp cut-off of the contours for low values of Σ8 is caused by the requirement that matter density and a shot-noise free galaxy density field must have a correlation coefficient r ≤ 1. All likelihoods are centered around our fiducial cosmology, i.e. the parameters describing the Buzzard simulations.

3.5 DES Y1 Forecast: constraints DES Y1 lensing only on galaxy bias 4 lensing only DES Y1 DES Y1 3.0 keeping ∆S3 0 lensing + CiC ≡ lensing + CiC

2.5 3 Forecast: constraints

) on gal. stochasticity ) b 2.0 r ( ( keeping ∆S 0

p 3 p 2 ≡ 1.5

1.0 1 0.5

0.0 0 0 1 2 3 0.2 0.4 0.6 0.8 1.0 b r

FIG. 4. Forecast of posterior constraints on galaxy bias b and galaxy-matter correlation coefficient r achievable with density split lensing and counts-in-cells in DES Y1 data. The constraints are marginalized over Ωm, σ8, Ωb, ns and h100, where for the last three parameters we have assumed the same flat priors as used in the DES Y1 combined probes analysis presented in DES Collaboration [29]. 10

parameter fiducial value prior in DES-Y1 constraints DES-Y1 constraints DES-Y1 constraints (in Buzzard) likelihood analysis without ∆S3/S3 with ∆S3/S3 without ∆S3/S3 (lensing only)

σ8 0.82 [0.2, 1.6] ±0.05 ±0.10 ±0.11 Ωm 0.286 [0.1, 0.9] ±0.04 ±0.06 ±0.06 Ωb 0.047 [0.01, 0.07] --- ns 0.96 [0.7, 1.3] -- h100 0.7 [0.55, 0.91] --- b 1.618 [0.8, 2.5] ±0.11 ±0.27 ±0.57 (lensing only:[0.0, 4.5]) r 0.956 [0.0, 1.0] ±0.10 ±0.11 ±0.18 ∆S3/S3 0.0 [−1.0, 2.0] - ±0.20 -

TABLE I. Model parameters of the forecast described in section II C 2. The second column shows our fiducial values (the values describing the Buzzard simulations). The third column shows the parameter priors used to cut our prediction for the posterior distribution of best-fit parameters. The priors on Ωb, ns and h100 are informative and chosen to be the same as used by the DES Collaboration [29]. The prior on r is needed for mathematical consistency. And the other priors only have to be introduced formally since we are approximating our analytic posterior by an MCMC. The 4th column shows the standard deviation of each parameter (as computed from the MCMC) after marginalization over all other parameters. The 5th column shows the same standard deviations but for the case where also ∆S3/S3 is introduced as a free parameter of our model. In column 6, we again fix ∆S3/S3 but assume that only the lensing part of the data vector is used. These forecasts can be compared to the actual errors we find in Gruen et al. [40, their tables 2 and 3] which are close despite marginalizing over systematic uncertainties. 11

7.) r: correlation coefficient between δm,T and δg,T (cf. equa- 8.) ∆S3/S3: a factor multiplying all third order statistics in tion II.22). our predictions. The notation for this parameter is based on the ratio We summarize these parameters and our fiducial values for 3 them in Table I. Throughout this paper, we assume the uni- δ S3 h i (II.30) verse to be flat. ≡ δ2 2 h i If πα and πβ are any two of the above parameters then let which connects third and second moments of the density con- πˆα,ML and πˆβ,ML be maximum likelihood estimates of these parameters based on a measurement of density split statistics. trast and hence characterizes the amplitude of the density bis- pectrum (see section IV A 2 for details). In our fiducial setup The covariance of πˆα,ML and πˆβ,ML can be estimated by we compute it from leading order perturbation theory and T ∆S3/S3 hence describes a relative deviation from that result. 1 ∂dth 1 ∂dth Cov[ˆπα,ML, πˆβ,ML]− Cd− , (II.29) ≈ ∂πα · · ∂πβ Within the ΛCDM model and at leading order in perturba- tion theory, the scaling between 2-point and 3-point statistics where dth is our model of the density split data vector and of the density field is almost independent of the cosmological Cd is the covariance matrix of a measurement of this signal parameters Ωm and σ8 [12]. Hence, a value of ∆S3/S3 = 0 in DES-Y1. We will in the following use an estimate of Cd would allow for deviations from the leading order result6 that from log-normal mocks and real DES Y1 shape noise (see cannot be compensated by changing Ωm or σ8. Such devia- section III B for a brief summary of our covariance estimation tions could be caused non-standard physics of dark matter and and Gruen et al. 40 for details). The parameter covariance dark energy that affect overdensities and underdensities differ- computed with equation II.29 can then be used to approximate ently (see e.g. Lue et al. [54], Multamäki et al. [61]; though the expected distribution of our best-fit parameter estimates f(R) modified gravity theories have been shown to largely with a multivariate Gaussian distribution. preserve the ΛCDM scaling, cf. Borisov & Jain [15], Jain Since the three parameters Ωb, ns and h100 are only poorly & Zhang [46]). Alternatively, they could indicate a break constrained by our data vector we are forced to assume prior down the the perturbative scaling relations due to highly non- knowledge on them. To do so, we cut the Gaussian poste- linear evolution of the density field or any small scale Bary- rior predicted from the parameter covariance with flat infor- onic physics that do not follow the scaling relations of pertur- mative priors. These priors are chosen to be the same used bation theory [cf. 12, 48, 80]. by the DES Collaboration [29]. For reasons of mathematical In the right panel of Figure 4 we show how density split consistency we furthermore have to demand that r [0, 1]. statistics including lensing and counts-in-cells can simultane- ∈ These hard cuts of our originally Gaussian approximation to ously constrain ∆S3/S3 and the parameter the posterior distribution of best-fit parameters make it diffi- p cult to marginalize over individual parameters. We hence nu- Σ8 = σ8 Ωm , (II.31) merically evaluate our analytic posterior with a Monte-Carlo Markov-Chain (MCMC). This chain is used in the following even after marginalization over the other model parameters. visualizations. Since we are using an MCMC to trace our an- We also project how these constraints will improve when alytic posterior, we have to formally define priors for all other moving to year 5 (Y5) data of DES or when adding cosmolog- model parameters. These are chosen to be flat and to extend ical information from the cosmic microwave background. For well beyond the single-parameter standard deviations of the the latter we estimated the parameter covariance in a Planck posterior. In the third column of table I, we summarize the chain3 and added this covariance as an additional Gaussian priors chosen for each model parameter. prior around our fiducial cosmology. The left panel of Figure 4 shows the 1σ and 2σ constraints With DES Y1 alone, we will be able to constrain the apli- achievable in the Ωm-σ8 plane. These contours are marginal- tude of third order statistics of the density field to about 20% ized over the other model parameters, using the priors men- accuracy (cf. last column of table I). Combining DES Y5 and tioned above. The blue contours assume that only the density Planck, this improves to about 5%. And this is even under- split lensing signal has been used while the green contours estimating the power of DES-Y5: to project our constraints allow for complementary information from the tracer counts- onto year-5 we simply divided our covariance by a factor of in-cells histogram. In the fourth column of table I we show the 4 in order to account for the larger area of the final DES sur- standard deviation of each parameter as found in our approx- vey. But this does not take into account the fact that DES Y5 imation to the posterior (and assuming the full data vector, will also be deeper than DES-Y1, which reduces shape noise including lensing and counts-in-cell). and opens up the possibility of analyzing a larger number of Density split statistics is complementary to an analysis redshift bins. based on 2-point statistics not just because it has a different dependence on the connection of galaxies and matter, but also since it is sensitive to higher order moments of the density field. We demonstrate this by introducing an additional degree 3 plikHM_TT_lowTEB in COM_CosmoParams_base-plikHM_R2.00.tar.gz of freedom in our model, described by an additional parame- from the Planck legacy archive https://pla.esac.esa.int/ ter: pla/, no lensing, cf. Planck Collaboration et al. [64] 12

III. SIMULATED DATA AND COVARIANCE MATRIX the source sample in the data. For in depth comparisons of these simulated samples to their data counterparts see DeRose In this work we use two sets of simulated data: the Buzzard et al. [30]. galaxy catalogs which are constructed from high-resolution For the density split analysis in this work, we use RED- N-body simulations [30, 55, 83] and simulated random fields MAGIC high density tracer galaxies (L > 0.5Lstar, ρ = 3 3 on the sky generated by the FLASK tool [86]. We briefly de- 10− Mpc− comoving density) selected at a REDMAGIC scribe these data sets in the following sections. photometric redshift estimate of 0.2 < z < 0.45. For the source redshift split lensing signals, we bin source galaxies by the expectation value of their p(z) as estimated with BPZ A. Buzzard mock galaxy catalogs [7, 43] from the Buzzard mock photometry. Bin limits are chosen such that the true mean redshifts of the bins match the Here we describe the key aspects of the Buzzard galaxy mean redshifts of the two highest redshift source samples de- catalogs for the purposes of this work and refer the reader to fined in Hoyle et al. [43]. more detailed descriptions elsewhere [30, 55, 83]. We use four independent realizations of a DES Y1-like survey in ver- sion Buzzard-v1.1. These catalogs were constructed from N- B. Simulated density and convergence fields from FLASK and Covariance matrix body simulations run using L-GADGET2 [74], a version of GADGET2 modified for memory efficiency. Second-order La- grangian perturbation theory initial conditions [27] were em- For testing the numerical implementation of the model de- ployed using 2LPTIC [27], and lightcones were output on scribed in the following section and for estimating a covari- the fly. Each galaxy catalog is built from a set of three nested ance matrix of the density split lensing and counts-in-cells sig- lightcones using progressively larger volume and lower res- nals in the Buzzard simulations and the DES data, we use log- olution at higher redshifts. The force resolution in each box normal realizations of matter density and convergence fields. is 20, 35 and 53 kpc/h with the boundaries between the light- We summarize their properties here, with details given in ap- cones falling at redshifts z = 0.34 and z = 0.9. pendix A of Gruen et al. [40]. The galaxy catalogs are constructed from the lightcones us- We generate these fields as all-sky HEALPIX maps of mat- ing the ADDGALS algorithm [83] which assigns galaxy lu- ter density and convergence using the FLASK software [86]. minosities and positions based on the relation between red- For the matter field, we choose the true redshift distribution shift, r-band absolute magnitude, and large-scale density, of REDMAGIC galaxies in the Buzzard simulations, selected p(δ Mr, z), found in a subhalo abundance matching (SHAM) as described in section III A. The matter field is sampled by a model| [25, 66, e.g.], in a high resolution N-body simulation. tracer population with REDMAGIC density, bias of b = 1.54, Spectral Energy Distributions (SEDs) are given to each sim- and Poissonian noise, from which lines of sight of different ulated galaxy by finding a SDSS DR7 galaxy [26] that has a density are identified by the same algorithm as in Buzzard or close match in Mr and distance to its fifth nearest neighbor in the data. For the convergence fields, we choose the esti- galaxy and assigning the SDSS galaxy’s SED to the simulated mated redshift distributions of DES source galaxies in the two galaxy. Galaxy sizes and ellipticities were assigned by draw- highest redshift bins of Hoyle et al. [43]. Log-normal param- ing from distributions fit to high resolution SuprimeCam i‘- eters of the density and convergence fields are set by the per- band data [59]. Observed magnitudes in griz were generated turbation theory formalism described in the following section. by redshifting the SEDs to the observer frame and integrating 960 of these realizations are used to estimate large-scale over the DES passbands. Photometric errors were added using structure and shot noise contributions to the covariance ma- the DES Y1 Multi Object Fitting (MOF) depth maps. trix of the signals modeled herein. The contribution of shape The effects of weak gravitational lensing are calculated us- noise is estimated by measuring the lensing signals in actual ing the multiple-plane raytracing algorithm CALCLENS [5]. DES Y1 data with 960 realizations of the METACALIBRATION The raytracing is done on a nside = 4096 HEALPIX [38] shape catalog [87] in which each galaxy ellipticity was rotated 0 grid yielding an effective angular resolution of 0.85 . At each by a random angle. lens plane, the inverse magnification matrix of the ray closest to every galaxy is interpolated to the galaxy’s 3D position and used to shear and magnify the galaxy. IV. MODELLING DETAILS AND COMPARISON TO With galaxy catalogs with magnitudes, sizes and lensing SIMULATIONS effects in hand, REDMAGIC and METACALIBRATION [45, 69, 87] samples of galaxies are selected from the simulations in In this section we present the approximations used to com- an effort to approximate the selections done in the data. In pute the ingredients (i), (ii) and (iii) of our model that are listed the case of REDMAGIC, the same algorithm which is used for at the end of section II B. We also test the model ingredients selection in the data is applied to the simulations, resulting in a (i) and (iii) directly against our N-body simulations. tracer galaxy catalog of equivalent volume density and at least Section IV A describes our model for the PDF of pro- comparable bias. For the METACALIBRATION sample, as we jected density contrast within the top-hat smoothing radius do not run full image simulations, we must make approximate θT , p(δm,T ). Section IV B describes how we model the con- cuts on signal to noise of the galaxies to create a facsimile of vergence profile κ<θ δm,T around apertures of fixed density h | i 13 contrast δm,T . And in section IV C we describe our modeling that correlations of δm,3D over a distance L vanish for all prac- of the probability P (NT δm,T ) of finding NT tracer galaxies tical purposes. Equation IV.5 employs a small angle approxi- | in an aperture with density contrast δm,T . mation and a Limber-like approximation [51]. This means it Section IV D summarizes our fiducial model and the ap- assumes that any n-point correlation function between density proximations used therein. constrast at different redshifts zi, i = 1...n, varies much more quickly with the redshift differences ∆zij than the projection kernel ql. Comparing IV.2 and IV.5 we see that the CGF of A. Projected density PDF δm,T can be computed in terms of the CGF of δcy.,R,L as Z ϕ (q (w)Ly, w) The computation of the PDF of the density field when ϕ(y) dw cy.,θw,L l . (IV.6) smoothed by top-hat filters has e.g. been adressed by ≈ L Bernardeau [9], Bernardeau & Valageas [10] and Valageas Hence, we have reduced the task of computing p(δ ) to the [81] (see also more recent developments in [13, 14, 23, 24, 80] m,T task of computing the connected moments of matter contrast which however do not yet enter our formalism). Bernardeau in a long 3D cylinder. & Valageas [10] demonstrated how to extend methods for the To proceed we consider two different ansatzes. The first is computation of the 3D density PDF to the weak lensing aper- to approximate p(δ ) by a log-normal distribution which is ture mass which is a projected quantity. In the following we m,T fixed by the first three connected moments of δ (nˆ). The show how to modify their formalism in order to compute the m,T second ansatz is to compute ϕ (y) as a whole in a way line-of-sight density PDF in angular circles of radius θ . To cy.,θw,L T similar to the one of Bernardeau [9] and Valageas [81] for the do so, we have to consider the cumulant generating function matter contrast in a 3-dimensional sphere. Using IV.6 we can (CGF) of the field δ (nˆ). The moment generating function m,T then attempt to solve the integral in IV.4 directly. We present (MGF) of δ (nˆ) is defined as m,T details of both approaches in the following subsections. k X δm,T (nˆ) ψ(y) = h i yk . (IV.1) k! k 1. Log-normal approximation for the PDF Due to the isotropy of the universe it does not depend on nˆ. Instead of computing the complete cumulant generating The CGF ϕ(y) is given in terms of the MGF ψ(y) as function of δm,T via equation IV.6 we start with an approach ϕ(y) = ln ψ(y) that only requires knowledge of the second and third cumu- 2 3 k lant, i.e. the moments δm,T c and δm,T c (implicitly we X δm,T (nˆ) c k h i h i h i y , (IV.2) also assume δm,T c 0 for the first cumulant). To do so, ≡ k! h i ≡ k we approximate δm,T as a shifted log-normal random variable where in the last line we have defined the connected moments [41, 86]. In this case the PDF of δm,T is given by k or cumulants δm,T (nˆ) c of δm,T (nˆ). The CGF of δm,T (nˆ)  2 2  h i [ln(δm,T /δ0 1)+σ /2] is related to its PDF via exp −2 − 2σ p(δm,T ) = (IV.7) Z √ πσ δ δ ϕ(y) yδm,T 2 ( m,T 0) e = dδm,T e p(δm,T ) , (IV.3) −

if δm,T > δ0 and p(δm,T ) = 0 otherwise. The expectation which is the Laplace transform of the PDF. Hence, if ϕ(y) is value of this PDF is zero, as is appropriate. The variance and a known analytic function, then p(δm,T ) can be computed by skewness of δm,T are given in terms of the parameters param- the inverse Laplace transform eters δ0 and σ by [41]

Z y  2  ∞ d iyδm,T +ϕ(iy) 2 2 σ p(δm,T ) = e− . (IV.4) δm,T c = δ0 e 1 2π h i − −∞ 3 1 δ3 δ2 2 δ2 3 . (IV.8) This integral in the complex plain is most efficiently evaluated m,T c = m,T c + 3 m,T c h i δ0 h i δ0 h i along the path of steepest descent (cf. [13], especially their appendix B). The cumulants of δm,T (nˆ) can be approximated The ansatz of a log-normal PDF has been shown to be con- as (cf. Bernardeau & Valageas [10]) sistent with early DES data by Clerkin et al. [22]. For the 3-dimensional density contrast δm,3D this can be reasonably D kE Z δcy.,θw,L(wnˆ, w)ql(w)L motivated from theory by observing that at leading order in k { } c δm,T (nˆ) c dw . perturbation theory the skewness of δ scales as h i ≈ L m,3D (IV.5) 3 2 2 δˆ c δˆ (IV.9) h m,3Di ∼ h m,3Dic Here ql is the line-of-sight density of tracer galaxies defined in ˆ2 3 equation II.2 and δcy.,R,L is the average of δm,3D over a cylin- with corrections of the order δm,3D c . This is exactly the der of length L and radius R, where L has to be chosen such scaling obeyed by the log-normalh distributioni and choosing 14 the parameter δ0 appropriately allows one to exactly repro- where δ∗ has to be determined by the implicit equation duce the scaling coefficients predicted by perturbation theory. At least for a power law power spectrum, the same kind of δ∗ 2 = yF 0[δ∗, τ] . (IV.15) scaling is observed also for 2-dimensional, projected versions σ3D,lin.(τ) of the density field. Hence, one of the ansatzes used in this It should be noted that equations IV.14 and IV.15 reproduce paper to compute the PDF of δm,T is to derive its variance exactly the tree-level results for the cumulant generating func- and skewness as described in appendix B and fix δ0 and σ by demanding that the PDF in IV.7 has the same second and third tion [cf. 9, 12, 14, 81] ! As described in Bernardeau et al. [12] moments. the coefficients δn S = c (IV.16) n h2 ni 1 δ c − 2. Tree level computation of ϕcy.,θw,L(y, w) in the cylindrical h i collapse model are given quite accurately by the lowest order of perturbation theory. Hence, using halofit [73, 77] and an analytic transfer Let us first consider the cumulant generating function function [33] to compute the non-linear matter power spec- 2 ϕ3D(y, τ) of the 3-dimensional density contrast δ3D(x, τ). To trum, we can compute the non-linear variance δ c,non.lin and compute ϕ3D at tree-level in perturbation theory it is sufficient then rescale the leading order CGF to its non-linearh i version to assume spherical symmetry around a particular point x [see via e.g. 81]. Doing so, we can directly compute δ3D(x, τ) as a δ2  δ2  function of the linear density contrast δ3D,lin.(x, τ) by means c,lin c,non.lin ϕnon.lin(y, τ) = h2 i ϕlead y h i2 , τ . of the spherical collapse model [35, 81], i.e. δ c,non.lin δ c,lin h i h i (IV.17) δ3D(x, τ) = F [δ3D,lin.(x, τ), τ] (IV.10) To perform the projection integral in equation IV.6 we need where F is determined by one of the differential equations to know the cumulant generating function of the density con- presented in appendix A. Hence, using the assumption that trast in a long cylinder of radius R and length L >> R, the linear density contrast has a Gaussian distribution with δcy.,R,L. Bernardeau [1994; see also 81, and the other ref- 2 erences above] have generalized equation IV.14 to the case of variance σ3D,lin.(τ) we can express the cumulant generating function as (cf. equation IV.3) matter density in a spherical aperture. Their results can easily be transferred to cylindrical apertures, yielding exp ϕ3D(y, τ) { } 2 Z δ∗ yδ3D ϕ (y, τ) yF [δ , τ] , = dδ3D e p(δ3D, τ) cy.,R,L cyl. ∗ σ2 τ ≈ − 2 ∗ ( ) R√1+Fcyl.[δ ],L,lin. Z 2 ! dδ3D,lin. δ3D,lin. (IV.18) = exp yF [δ , τ] , 2 q 3D,lin. 2 where σR,L,lin. is the variance of linear density contrast in a πσ2 τ − 2σ3D,lin.(τ) 2 3D,lin.( ) cylinder, Fcyl. is now determined by cylindrical collapse and (IV.11) δ∗ has to be determined by the implicit equation 2 where in the last step we simply performed a change of vari- 1 d δ∗ = yF 0[δ∗, τ] . (IV.19) ables from δ3D to δ3D,lin.. We now employ Laplace’s method, δ σ2 τ 2 d ∗ ∗ ( ) which states that a function f(x) which strongly peaks around R√1+Fcyl.[δ ],L,lin. x fulfils 0 Using equation IV.17 we can again rescale this leading order s Z 2π result for the generating function to its non-linear counterpart. dx ef(x) ef(x0) . (IV.12) The validity of equation IV.17 is ultimately limiting the ac- ≈ f (x0) | 00 | curacy of our model for the distribution of the density contrast This way we can approximate the last line of IV.11 as inside the aperture radius θT , p(δm,T ). In figure 5 we com- pare the PDF measured in the Buzzard simulations to both s 1 the log-normal model and the full CGF computation of the eϕ3D (y,τ) 2 PDF for aperture radii θT = 100, 200, 300. For both θT = 200 ≈ 1 yF 00[δ∗, τ] σ3D,lin.(τ) × | − | and 30 our model PDF’s and the Buzzard simulations agree ! 0 δ 2 within DES-Y1 as can be seen in the residuals yF δ , τ ∗ , exp [ ∗ ] 2 (IV.13) shown in the lower panels of each plot. Also, the difference × − 2σ3D,lin.(τ) between log-normal approximation and full CGF computation where δ∗ is the linear density contrast that maximizes the ex- is completely negligible. To investigate the agreement of Buz- ponent in IV.11 and 0 denotes derivation wrt. δ . In the quasi- zard and our models more quantitatively, we also compare the linear limit of σ2 0 this gives variance and skewness of each PDF. In figure 6 we show the 3D,lin. → ratio of these moments as found in Buzzard to our predictions. δ 2 ∗ For θT = 100, the density field in Buzzard seems to have a sig- ϕ3D(y, τ) yF [δ∗, τ] 2 , (IV.14) ≈ − 2σ3D,lin.(τ) nificantly higher skewness than predicted by our model. We 15

3.0 3.5

2.5 θT = 10 arcmin 3.0 θT = 20 arcmin

2.5 2.0 ) ) 2.0

m,T 1.5 m,T δ δ ( ( 1.5 p p 1.0 1.0

0.5 0.5

) 0.0 ) 0.0

m,T 1.2 m,T 1.2 δ δ ( ( . 1.1 . 1.1

Buzz 1.0 Buzz 1.0 / p / p

) 0.9 ) 0.9

m,T 0.8 m,T 0.8 δ δ ( (

p 2.0 p 0.4 0.2 0.0 0.2 0.4 0.6 0.8 0.4 0.2 0.0 0.2 0.4 0.6 − − − − δm,T δm,T

3.5 θT = 30 arcmin 3.0 1.8 Buzzard 2.5 ) Buzzard 2.0 m,T

δ Y1 sized patches (

p 1.5 1.6 cyl. collapse 1.0 log-normal 0.5 Gaussian

) 0.0

m,T 1.2 1.4 δ ( . 1.1

Buzz 1.0 / p

) 0.9

m,T 0.8 δ 1.2 ( p 0.4 0.2 0.0 0.2 0.4 0.6 − − δm,T

FIG. 5. The PDF of projected density contrast in Buzzard compared1.0 to several models for various smoothin scales (δm,T is the projected density contrast δm smoothed by our top-hat radius θT ). In the upper1.0 panel of each1. plot2 the black line1.4 shows a histogram1.6 of δm,T measured1.8 in 2.0 an all-sky map from Buzzard. The grey lines show histograms measured in 14 DES year1 shaped patches of that all-sky map. The blue lines show the PDF predicted by our tree-level computation of the cumulant generating function, the green lines show the PT-motivated log-normal model and the red lines show a Gaussian PDF with the same variance as the other two models. The bottom panels of each plot are showing the ratio of each PDF to the one measured in the Buzzard all-sky. For all aperture radii our halofit power spectrum is predicting a standard 0 0 deviation of δm,T that is & 2% higher than what we find in Buzzard (cf. figure 6). For θT = 20 and 30 this is the dominant source of mismatch. 16

1.5 projection of the CGF of matter density contrasts that are errorbars are std. dev. variance (n = 2) smoothed over concentrical cylindrical apertures with length 1.4 model from 14 Y1-sized patches i skewness (n = 3) L and radii R1,R2 , ϕcyl.,R1,R2,L(y, z, w): n ] 1.3 Z

m,T ϕcyl.,θT w,θw,L(ql(w)Ly, Ws(w)Lz, w) δ

[ ϕθ(y, z) dw . h 1.2 ≈ L / 1.1 (IV.22)

1.0 Here, ql(w) is again the line-of-sight density of lens galaxies simulation

i and Ws(w) is the lensing efficiency defined in eq. II.5. n

] 0.9 The joint PDF of κ<θ and δm,T is related to ϕθ via m,T

δ 0.8 [ h Z y z ∞ d d iys izr+ϕθ (iy,iz) p(δm,T = s, κ<θ = r) = 2 e− − . 10 20 30 (2π) −∞ θT [arcmin] (IV.23) The expectation value of κ<θ given a certain value of δm,T is FIG. 6. This figure shows the ratio of moments of δm,T measured in a then given by Buzzard all-sky density map to the moments predicted by our model as a function of θT . The errorbars represent the standard deviations κ<θ δm,T = s of the moments found in a set of 14 DES Y1-sized patches in the h | Zi 0 1 Buzzard map. At our fiducial radius θT = 20 we find a ∼ 2.4% = dr r p(δ = s, κ = r) p δ s m,T <θ deviation between the variance of δm,T measured in Buzzard and ( m,T = ) that predicted by our model. This would correspond to a ∼ 1.2% 1 Z ∞ dy dz Z e iys+ϕθ (iy,iz) r r e izr deviation in σ8. = 2 − d − p(δm,T = s) (2π) −∞ Z y z 1 ∞ d d iys+ϕθ (iy,iz) d = e− i δD(z) attribute this indeed to the failing of the Quasi-linear rescal- p(δm,T = s) 2π dz −∞ ing, eq. IV.17. For the other radii the skewness values agree to Z y 1 ∞ d iys+ϕm,T (iy) within 2 3%. A similar relative agreement is achieved for the = e− Gθ(iy) , − p(δm,T = s) 2π variance of the distributions. At our fiducial radius θT = 200 −∞ the variances of Buzzard and our model differ by about 2.4%. (IV.24) This corresponds to a disagreement in the fluctuation ampli- with tude σ8 of about 1.2%.

For comparison, we also show a Gaussian model for the d density PDF in figure 5, using the same variance as for the Gθ(y) = ϕθ(y, z) dz z=0 other PDF models. It can clearly be seen, that p(δm,T ) cannot k be well described by a Gaussian distribution for any of the X δm,T κ<θ c = h i yk . (IV.25) considered smoothing radii. k! k

Using equation IV.22 we can express Gθ(y) by the corre- B. Convergence profile around lines-of-sight of fix density sponding cylindrical, 3-dimensional quantity: contrast δm,T Z

Gθ(y) dw Ws(w) Gcyl.,θT w,θw,L(ql(w)Ly, w) . We now want to know the average lensing convergence κ<θ ≈ inside a radius θ around a line-of-sight with a given value of (IV.26) δm,T . By means of equation II.6 this can be turned into a pre- diction of the density split lensing signal. We start by looking We again pursue two ansatzes for computing Gθ(y): one in- volving a log-normal model for the joint cumulants of κ<θ at the joint moment generating function of κ<θ and δm,T , and δm,T and one involving the model of cylindrical collapse k l X δm,T κ<θ k l to compute a leading order perturbation theory prediction for ψθ(y, z) = h i y z , (IV.20) G y, w k! l! cyl.,θT w,θw,L( ). We are detailing these ansatzes in the k,l following subsections. and their joint cumulant generating function defined by ϕ y, z ψ y, z θ( ) = ln θ( ) 1. Log-normal model for the joint cumulants of κ<θ and δm,T k l X δm,T κ<θ c h i yk zl . (IV.21) ≡ k! l! From equation IV.25 one can see that only joint cumulants k,l of the form Using a Limber-like approximation similar to the one em- k ployed in eq. IV.6, one can write ϕθ(y, z) as a line-of-sight δ κ<θ c (IV.27) h m,T i 17 enter the convergence profile around a given density contrast. where the elements of the matrix are given by 11 = 2 2 C C Hence, in a spirit similar to section IV A 1 we only compute σR (δ∗), = σR (δ∗) and 12 = 21 is the linear co- 2 1,L 1 C 2,L 2 C C the leading order cumulants δm,T κ<θ c and δ κ<θ c as variance of density contrasts in concentric cylinders of radii h i h m,T i described in appendix B and use these moments to fix a joint R1,L(δ1∗) and R2,L(δ2∗). This time the critical linear density κ δ log-normal PDF for <θ and m,T . Note that this is indeed contrasts δ2∗ and δ2∗ are given by the implicit equations not assuming, that κ<θ is a log-normal random variable. It rather assumes that 1 ∂ X 1 δ∗δ∗ − = yF 0[δ∗] (IV.33) 2 ∂δ i j C ij 1 1∗ i,j κ<θ = κlog normal + κuncorr. , (IV.28) − 1 ∂ X 1 δ∗δ∗ − = zF 0[δ∗] . (IV.34) 2 ∂δ i j C ij 2 where κlog normal is log-normal and κuncorr. is an unspec- 2∗ i,j − ified random variable that is uncorrelated with δm,T . Only Note that these conditions force each δ to be a function of κlog normal will actually contribute to the density split lensing i∗ − both y and z, i.e. δ δ y, z . signal and we can model the expectation value κ<θ δm,T = i∗ = i∗( ) s by the following relation holding for two jointh log-normal| To predict the convergence profile around apertures of a variables:i given density contrast δm,T by means of equations IV.24 and IV.26 we are interested in computing the function κ δ s C s/δ V C  <θ m,T = (2 log(1 + 0) + ) h | i = exp − 1 ∂ κ0 2V − G(y) = ϕ(y, z) . (IV.35) (IV.29) ∂z z=0 with Using the conditions IV.33 and IV.34 one can see right away 2 ! δm,T c that V = log 1.0 + h 2 i δ0 G(y) = F [δ2∗(y, 0)] . (IV.36)   δm,T κ<θ c C = log 1.0 + h i δ κ Furthermore, for z = 0 equations IV.33 and IV.34 can be sim- 0 0 plified to 2 V δm,T κ<θ c e κ0 = h i (IV.30) 2 2 2 1 d δ∗ δm,T κ<θ c 2 δm,T κ<θ c δm,T c/δ0 1 h i − h i h i = yF 0[δ1∗] (IV.37) 2 dδ1∗ 11 and δ determined as described in section IV A 1. C 0 12 It should be noted that the log-normal parameter κ0 which δ2∗ = C δ1∗ . (IV.38) 11 we use to approximate the contribution of κ<θ to the lensing C signal is dependent on the the smoothing scale θ. This indi- This way we obtain a solution for G(y) at leading order in cates even further, that we do indeed not approximate κ as a perturbation theory. In appendix B we argue that for R2 R1 the cumulants δk δ approximately follow the scaling≥ log-normal field. R1 R2 c relation h i

δk δ δ δ δ2 k 1 . R1 R2 c R1 R2 c R1 c− (IV.39) 2. Tree level computation of Gcyl.,R1,R2,L(y, w) in the cylindrical h i ∼ h i h i collapse model This can be used to correct the tree-level approximation of G(y) for the non-linear evolution of the power spectrum. To For convenience we will shorten the notation of section do so, we first determine the proportionality factors of the re- IV A 2 by defining lation IV.39 at leading order by fitting a polynomial in y to the function G y and extracting the cumulants δk δ from ( ) R1 R2 c G(y) Gcyl.,R ,R ,L(y, w) h i ≡ 1 2 the polynomial coefficients. In practice, we do this with a ϕ(y, z) ϕcyl.,R1,R2,L(y, z, w) polynomial of degree 10, but already a polynomial of degree ≡ 4 F [δ] Fcyl.[δ, τ] 5 gives almost identical results . Then, we use the revised ≡ halofit of Takahashi et al. [77] to compute a late-time ver- σR σR,L,lin. ≡ p sion of the right-hand-side of IV.39. This, together with the Ri,L(δ) Ri 1 + F [δ] , i = 1, 2 . (IV.31) tree-level proportionality factors determined before, yields a ≡ non-linear approximation of the polynomial coefficients rep- The joint cumulant generating function of density contrast in resenting G(y). We use those to re-compute G(y) and then concentric cylinders is then [in complete analogy to equa- carry out the projection integral in equation IV.26. tion IV.18; see also 10, who present very similar calculations] given by

1 X 1 4 The coefficients of linear and quadratic order in y are always obtained from ϕ(y, z) yF [δ1∗] + zF [δ2∗] δi∗δj∗ − , (IV.32) ≈ − 2 C ij the exact perturbation theory computation of appendix B. i,j 18

Our rescaling of the coefficient corresponding to the cumu- We explore two ways to account for a possible stochastic- lant δ2 δ is in fact more complicated than described ity (see also Dekel & Lahav [31], who have discussed similar R1 R2 c here,h cf. appendixi B 5. But we find our prediction of the den- concepts). In our first approach we introduce a free param- sity split lensing signal to be insensitive to the details of the eter to our model - a Pearson correlation coefficient r = 1 6 rescaling procedure. between the random fields δg,T and δm,T . Within our log- normal framework we show that this automatically leads to a δm,T -dependence of the ratio in eq. IV.46. We explain the C. Shot-noise, stochasticity and Counts-in-Cells details of this in section IV C 1. In our second approach we employ a generalized Poisson We now want to model the conditional probability distribution for P (NT δm,T ) that allows for | P (NT δm,T ) of finding NT galaxies in an angular radius of | Var [NT δm,T ] θT , when the projected density contrast in that radius is δm,T . | = 1 . (IV.45) NT δm,T 6 This is the third ingredient of the framework described in sec- h | i tion II and completes our modeling of the density split lensing In this approach we introduce 2 parameters, α0 and α1, to our signal as well as the counts-in-cells histogram. model such that To analyze the relation of NT and δm,T in a systematic δ Var [NT δm,T ] way, let us introduce the auxiliary field g,T . We assume that | α0 + α1 δm,T . (IV.46) δ (nˆ) is a smooth field in the sky and that N is a Poisso- NT δm,T ≈ g,T T h | i nian tracer of this field. This means we will assume that The details of this are explained in section IV C 2. N Both of our approaches match our simulated data equally N¯(1 + δ ) g,T N¯(1+δg,T ) P (NT = N δg,T ) = e− , well (cf. Figure 8). This means that, for the galaxies in these | N! r (IV.40) realizations, the model based on the correlation coefficient is a sufficient description. It will thus be the fiducial model where N¯ NT . A consequence of this assumption is that the expectation≡ h valuei of N for fixed δ is given by in this paper, used in all figures unless otherwise noted. In T g,T Gruen et al. [40] we will nevertheless apply both this and the two-parametric model to account for the possibility that the NT δg,T = N¯(1 + δg,T ) (IV.41) h | i shot-noise of real galaxies behaves in a more complicated way and that the variance of NT for fixed δg,T fulfils than that of our simulated galaxies.

Var [NT δg,T ] | = 1 . (IV.42) 1. Shot-noise model 1: correlation r 6= 1 between galaxy density NT δg,T h | i and matter density To connect the galaxy field to the lensing convergence we however need to know the relation between NT and δm,T . As- In our fiducial model of P (NT δm,T ) we approximate the | suming a generic joint PDF of δm,T and δg,T we can write the joint distribution of both δm,T and δg,T with a joint log- expectation values of NT for fixed δm,T as normal distribution (cf. eq. IV.7 and Hilbert et al. [41] for properties of joint log-normal distributions). The joint PDF of Z two log-normal random variables is characterized by 5 param- NT δm,T = dδg,T p(δg,T δm,T ) NT δg,T h | i | h | i eters, e.g. by the variance and skewness of each variable and

= N¯(1 + δg,T δm,T ) . (IV.43) the covariance between the two variables. h | i In our case, we compute the variance and skewness of δm,T Also, it can be shown that the variance of NT for a fixed value as described in section IV A 1 and set the variance and skew- of δm,T is given by ness of δg,T to 2 2 2 ¯ 2 δ c = b δ c Var [NT δm,T ] = NT δm,T + N Var [δg,T δm,T ] . h g,T i h m,T i | h | i | 3 3 3 (IV.44) δg,T c = b δm,T c (IV.47) From equation IV.44 we can see that the distribution of h i h i galaxy bias b NT given δm,T can only be a Poisson distribution if where the is a free parameter. The covariance of δ δ Var [δg,T δm,T ] 0. This is the simplest model for the m,T and g,T is parametrized by their correlation coefficient | ≡ connection of NT and δm,T that we test in this work and in δg,T δm,T c Gruen et al. [40]. If Var [δg,T δm,T ] = 0 we will say that r = h i | 6 q 2 2 there is a stochasticity between the galaxy field and the mat- δ c δ c h g,T i h m,T i ter density field, and we cannot assume a Poisson distribution δ δ P N δ g,T m,T c for ( T m,T ). We note that “stochasticity” in this con- = h 2 i , (IV.48) | b δ c text could arise from a nonlinear biasing relationship between h m,T i δg,T and δm,T , including e.g. a dependence on higher powers i.e. of δm,T or effects from halo exclusion [2], or from physical 2 stochasticity in galaxy formation. δg,T δm,T c = rb δ c . (IV.49) h i h m,T i 19

The log-normal model for the joint PDF of δm,T and δg,T now Smoothing Scale b r allows us to compute the variance of galaxy counts as a func- [arcmin] tion of δm,T and more generally to compute P (NT δm,T ). We 10 1.644 ± 0.008 0.938 ± 0.001 present the necessary derivations in detail in appendix| D. 20 1.618 ± 0.008 0.956 ± 0.001 In our data analysis we consider b and r as free parameters. 30 1.603 ± 0.008 0.961 ± 0.001 The only restrictions we impose on them are TABLE II. Best-fit values galaxy bias and correlation coefficients of 0 < b , 0 r 1 . (IV.50) our simulated tracer galaxies within the model presented in section ≤ ≤ IV C 1. Error bars are estimated from a jackknife approach. To test how accurately this model describes the behaviour of our mock REDMAGIC catalogs based on the Buzzard N-body Smoothing Scale ˜b α0 α1 simulations we nevertheless want to determine what values of [arcmin] b and r are underlying our simulations. To this end, we gen- 10 1.54 ± 0.001 1.15 ± 0.001 0.22 ± 0.003 erate HEALPIX maps of δm,T with different top-hat aperture 20 1.54 ± 0.002 1.26 ± 0.002 0.29 ± 0.010 radii θT = 10, 20, 30 arcmin, based on particle count maps 30 1.54 ± 0.002 1.39 ± 0.003 0.45 ± 0.020 1 at resolution Nside = 8192 in slices of comoving 50h− Mpc thickness. We co-add these maps to reproduce a redshift range TABLE III. Best-fit values galaxy bias and shot-noise parameters of close to that of our fiducial analysis, z = 0.2100 ... 0.4453. our simulated tracer galaxies within the model presented in section We then select REDMAGIC galaxies with true redshift in this IV C 2. Error bars are again estimated from a jackknife approach. range and determine their counts around the same HEALPIX pixel centers and within the same aperture radii. The RED- MAGIC mock catalogs have a complex mask similar to that δg,T δm,T is consistent with a linear relation in δm,T . Inter- h | i of real DES data, which adds complication because the frac- estingly, the scale dependence of b and r we find in table II tion of masked area in each aperture must be equal in order to almost perfectly cancels to give a scale independent propor- meaningfully sort lines of sight by galaxy count. To this end, tionality coefficient we convert all counts to a masking fraction of 20 per-cent of area within the aperture radius using the stochastic method of rb 1.54 . (IV.55) ≈ Gruen et al. [40, their section 2.1]. This leaves us with sim- Next, we also measure the variance of galaxy counts N as ulated 2D maps of δm,T and NT within a DES-Y1 shaped T mask. a function of δm,T in our simulated maps and compare to the We can then measure the variances of these maps, prediction of the log-normal model. In Figure 8 we indeed find that Var(δm,T ) and Var(NT ), as well as their covariance Cov(NT , δm,T ). These fulfill the relations Var [NT δm,T ] | = 1 (IV.56) ¯ ¯ 2 2 NT δm,T 6 Var(NT ) = N + N b Var(δm,T ) . (IV.51) h | i and and that the δm,T -dependence is very well described by the log-normal model and the values of b and r we determined Cov(NT , δm,T ) = N¯ b r Var(δm,T ) (IV.52) before. Finally, in figure 9 we show the residuals between our baseline prediction of the counts-in-cells histogram with which then fixes b and r. The values determined in this way 5 θT = 20arcmin and the average of measurements in 4 Buz- are shown in table II. zard realizations of DES Y1 data (cf. figure 1). The residuals We now need to check whether these value for b and r to- are well contained within DES Y1 errorbars. gether with our assumption of a log-normal PDF for δm,T and δg,T describe the properties of our tracer galaxies well. Us- ing our simulated maps of δm,T and NT we can measure the 2. Shot-noise model 2: Parametric model for super-Poissonianity expectation value

NT δm,T Our model of shot-noise based on galaxy bias b and galaxy δg,T δm,T = h | i 1 (IV.53) h | i N¯ − matter correlation coefficient r describes our simulated tracer catalog well. But it contains the arbitrary assumption that both as a function of δm,T . Within the log-normal model (cf. ap- the variance and the skewness of δm,T and δg,T are related pendix D for the relevant formulae) this is very well approxi- through the bias parameter b (cf. equation IV.47). To account mated by for the possibility that real galaxies might behave in a more complicated way, we also consider a more flexible model of δg,T δm,T rb δm,T (IV.54) h | i ≈ the conditional distribution P (NT δm,T ). | which becomes exact for Gaussian random variables. In Fig- ure 7 we compare measurements of δg,T δm,T with the dif- h | i ferent smoothing radii θT = 10, 20, 30 arcmin to the predic- tion of the log-normal model. We find that in our simulations 5 This is the smoothing radius used in our data analysis [40]. 20

2 θT = 10 arcmin θT = 20 arcmin θT = 30 arcmin

1 g,T δ

0

1 − m,T

δ 0.1 · 54 . 0.0 1

− 0.1

g,T − δ 0.5 0.0 0.5 0.5 0.0 0.5 0.5 0.0 0.5 1.0 − − − δm,T δm,T δm,T

FIG. 7. Average galaxy overdensity δg,T as a function of matter overdensity, δm,T , for our simulated maps at different smoothing scales: 10 arcmin [left], 20 arcmin [middle] and 30 arcmin [right]. Solid lines show a linear bias model with the bias parameters obtained from maximizing the likelihood in equation IV.63 and the residual between the two are shown in the bottom panels. Note that the coefficient of linearity found with IV.63 (≈ 1.54) is almost identical to the value of the product b · r determined with equation IV.54. To indicate the range of δm,T that is relevant to our computation, we also show the density PDFs of figure 5 as shaded regions. The errorbars were derived from a jackknife approach.

0.20 θT = 20arcmin mean of 4 Buzzard + 2 0.15 DES Y1 error bars 2.0 10

· actual error bars )]

i 0.10 T N ( . N 1.5 h 0.05 / Buzz ] P T −

N 0.00 1.0 ) N ( Var[ 0.05 10 arcmin model 1: b, r − 0.5 ˜ theory 20 arcmin model 2: b, α0, α1 0.10 30 arcmin [P − 0.0 0.15 0.4 0.2 0.0 0.2 0.4 0.6 − 20 40 60 80 − − δm,T N

FIG. 8. Ratio of the variance in the galaxy count distribution to av- FIG. 9. For the fiducial smoothing radius of our data analysis pre- erage galaxy counts in our simulated maps as a function of matter sented in Gruen et al. [40], θT = 20arcmin, we show the residuals overdensity. Differently coloured solid lines show the result for each between our baseline prediction of the counts-in-cells histogram and smoothing scale. The dashed black lines show the predictions of the the average of measurements in 4 Buzzard realizations of DES Y1 2-parametric shot-noise model described in section IV C 2. The blue data. Blue error bars represent the uncertainties we expect for DES dotted lines show the corresponding prediction of the alternative, 1- Y1 while green error bars show the actual uncertainties of the mean parametric model described in section IV C 1. The horizontal solid measurement from our simulations. line shows the expectation if shot noise was purely Poissonian. The coloured regions show the 95% confidence limits derived from jack- knife resampling. 21

Gruen et al. [39] assumed that there is no stochasticity in the posterior distribution of α0 and α1 we infer with DES Y1 the relation of δm,T and δg,T and that galaxies trace the mat- like errors around the mean signal measured in Buzzard (after ter density with a linear bias and Poissonian shot noise. This marginalizing over our other model parameters, cf. appendix means they set E). Also, the contraints on α0 and α1 we derive in [40] on DES data are well contained within our prior distributions. P (NT = N δm,T = s) = Nevertheless, these priors must be considered mildly infor-  |  exp N ln[N¯(1 + ˜bs)] [N¯(1 + ˜bs)] ln Γ(N + 1) , mative. We expect that even stonger priors can be motivated − − theoretically. [2] find that for their most massive halos shot- (IV.57) noise is reduced wrt. Poisson expectation by a factor of 2, indicating that α . , while for halo masses comparable≈ to where ˜b is the galaxy bias, and where we now use a gener- 0 & 0 5 redmaGiC halo masses (cf. Clampitt et al. [21]) they find shot- alizable definition of the Poisson distribution based on the noise to be close to Poissonian. Also, there is evidence that the Gamma function Γ, for reasons that will appear later. The fraction of red galaxies that are satellites (resp. the fraction galaxy bias ˜b defined this way is not identical to the definition of satellite galaxies that are red) increases with environment in our fiducial model. We now rather have density (see e.g. [56, 62]). According to [2] this will cause an increase of galaxy stochasticity with environment density, δg,T ˜b δm,T . (IV.58) ≡ corresponding to α1 > 0.0. We intend to investigate impli- We test this in our simulated maps of NT and δm,T by fitting cations of models for halo occupation distributions (HOD) on a linear relation to the mean smoothed galaxy contrast as a our shot-noise parametrizations in future studies (see e.g. the function of dark matter contrast that was shown in Figure 7. work by [18, 32] on connecting HOD models and parametric We indeed find that this linear biasing model describes the models of galaxy bias and stochasticity). simulations very well and that ˜b br as expected from our To compare this parametric shot-noise model to our simu- arguments in section IV C 1. ≈ lations, we are nevertheless interested in the particular value The model used by Gruen et al. [39] however predicts that of α0 and α1 that describe these simulations. From the tuples of (NT , δm,T )i measured in our simulated maps, we can con- Var [NT δm,T ] strain bias and the α parameters with a likelihood L that is | 1 (IV.59) 0/1 NT δm,T ≡ simply the product of the probabilities of the individual tuples h | i from Equation IV.60, which is not what we find in Figure 8. To account for the X  i   i i  deviations we observe from pure Poissonian shot-noise, we ln L = Ni/α(δ ) ln (N¯i/α(δ ))(1 + b δ ) m,T m,T × m,T hence model the distribution of NT given δm,T as i  ¯ i  i Ni/α(δm,T ) (1 + b δm,T ) P (NT = N δm,T = s) = − × | N ×  i  i   ¯    ¯  ln Γ Ni/α(δm,T ) + 1 ln α(δm,T ) . N NT N NT − − exp ln (1 + bs) ln Γ + 1 (1 + bs) , (IV.63) α α − α − α Because the tuples have correlated counts and densities, this (IV.60) is not an exact expression for the likelihood of our measure- where the parameter α > 0 generalizes the distribution to one ments, but it should be sufficient to obtain reasonable best where groups of α galaxies appear with Poissonian noise and fit values for b, α0 and α1. We determine the uncertainties where the normalization coefficient is needed to ensure that of these best-fit values by finding the maximum of Equa- R N 1 tion IV.63 on jackknife resamplings of the simulations. The P (NT = N) dN = 1. We find to be very close to α− 1 N resulting parameter values are shown in Table III and dis- and identical to α− in the case where α is an integer value. To account for the observed increase of super-Poissonianity played in Figure 7 and Figure 8. We find that our simulated REDMAGIC galaxies are indeed well described as linearly bi- with density, we allow α to depend on δm,T , ased tracers of the density field with a small, but significant, α(δm,T ) = α0 + α1 δm,T . (IV.61) scale and density dependent super-Poissonian shot-noise. ×

This indeed leads to a δm,T -dependence of the variance of galaxy counts that is close to the relation mentioned in eq. D. Summary of fiducial model and approximations therein IV.46. In our analysis we treat α0 and α1 as free parameters within the ranges For each ingredient (i) to (iii) of the framework described in section II we have introduced at least two different mod- α0 [0.1, 3.0] ∈ eling ansatzes. We want to once more describe our baseline α1 [ 1.0, 4.0] . (IV.62) model built from these ansatzes (cf. also section II B). This is ∈ − the model we consider in section V and which we use in the In principle we could allow any value α > 0 but we choose 0 data analysis presented in Gruen et al. [40]. the boundary 0.1 < α0 because it is numerically difficult (and slow) to predict the CiC histogram for values close to α0 = (i) p(δm,T ): We find that the log-normal model (section 0.0. The other boundaries roughly enclose the 2-σ region of IV A 1) and our model based in cylindrical collapse (section 22

IV A 2) describe the PDF of projected density contrast equally Despite this long list of approximations, this baseline model well. The computations based on the log-normal model are describes our measurements in the Buzzard simulations well however significantly faster. Hence in our fiducial analysis within DES Y1 errorbars (cf. figure 1). As shown in the next we employ the log-normal model. section, the model is also accurate enough to recover the true cosmology of our simulation within DES Y1 uncertainties in a simulated likelihood analysis. In Gruen et al. [40] (and us- (ii) κ<θ δm,T : We also introduced a log-normal model h | i ing an extended set of simulations) we furthermore show that (section IV B 1) and a model based on cylindrical collapse the values of χ2 found between our fiducial model and indi- (section IV B 2) for the convergence profile around lines-of- vidual simulation measurements are consistent with the χ2- sight with fixed density contrast δm,T . Both models lead distribution expected from our number of data points, and that to almost identical predictions for the density split lensing the coverage (i.e. the fraction of times the true simulation cos- signal. Hence we again choose the log-normal model for our mology is within the confidence interval) matches expecta- fiducial analysis, because of the shorter computation time. tions.

(iii) P (NT δm,T ): We introduced two models for the distribu- | tion of tracer counts NT in lines-of-sight of matter density V. RECOVERING COSMOLOGY IN N-BODY SIMULATIONS δm,T . The first was based on linear galaxy bias b and galaxy- matter-correlation coefficient r (section IV C 1). The second was based on an alternative definition of galaxy bias and on In this section we want to test whether the modeling that two parameters α0 and α1 describing density dependent devi- was described in sections II and IV is sufficient to recover the ations from Poissonian shot-noise (section IV C 2). Both mod- cosmology underlying a density split data vector measured in els describe the behaviour of our simulated tracer galaxies in N-body simulations. The simulations we use are described Buzzard-v1.1 similarly well. But anticipating that real galax- in section III A. They are the same simulations against which ies might behave in a more complicated way, we will consider we tested the ingredients of our model in the previous sec- both ansatzes in our fiducial analysis. tion. A likelihood analysis based on a density split data vec- tor measured in these simulations in presented in section V A. In the following list, we are summarizing the approximations We only run a cosmological analysis on the mean data vec- that went into the derivation of our baseline model. tor measured on 4 DES-Y1 realisations. The goal of this is to show that any possible systematic deviations between our 1.) We assumed, that for fixed value of δm,T the convergence modeling of density split statistics and the behaviour of our within angular radius θ is not dependent on NT (cf. equation N-body simulations is smaller than the statistical uncertain- II.8). ties of DES-Y1. A more extensive validation of our likelihood 2.) All second order moments in our formalism are computed pipeline is presented in Gruen et al. [40]. with a halofit power spectrum [77] using an analytic approxi- mation for the transfer function [33]. A. Simulated likelihood analysis 3.) Equations IV.5, IV.6, IV.22 and IV.26 employ a small angle and a Limber-like approximation [following 10]. We now measure the data vector that was described in sec- 4.) We compute the cumulant generating function of density tion II C in 4 N-body realisations of DES-Y1. We always use contrast in long cylinders by means of the cylindrical collapse the mean of these 4 data vectors. In order to further reduce the approximation (cf. section IV A 2). noise of this measurement, we turn off the shape noise in our 5.) We assume that the tree-level result of the cumulant gen- simulated source catalogs, i.e. we measure our signal directly erating function can be corrected for the full non-linear evo- from the gravitational shear acting each galaxy. We then run lution of the density field by means of equations IV.16 and Monte-Carlo Markov Chains of our model around this data IV.17. vector. For this we assume a Gaussian likelihood function with the full covariance (i.e. including shape noise) that was 6.) We approximate the PDF of δm,T resulting from a cylindri- estimated by Gruen et al. [40] for a DES-Y1 data set. The cal collapse approximation by a log-normal PDF (cf. section goal of this analysis is to test whether the fiducial cosmol- IV A 1). ogy of the Buzzard simulations is well contained within the 7.) We employed approximations similar to 4.), 5.) and 6.) 1σ contraints derived from this likelihood analysis. A more extensive validation of our likelihood pipeline is presented in for the joint distribution of δm,T and κ<θ (cf. section IV B 2, equation IV.39 and section IV B 1). Gruen et al. [40]. In the top panel of Figure 10 we show the 1σ and 2σ con- 8.) We assume that galaxies are linearly biased tracers of the traints obtained from our simulated likelihood analysis in the density field. We consider two different models for shot- Ωm-σ8 plane after marginalizing over different sets of param- noise (resp. stochasticity), assuming that the full distribution eters. First, we only vary Ωm and σ8, setting other cosmo- P (NT δm,T ) is well described bei either two parameters (b, r) logical parameters to the inputs of the Buzzard simulations | or three parameters (˜b, α0, α1). and the parameters connecting galaxy count and matter den- 23

1.2 sity to the values we found from the Buzzard galaxy and den- chains around mean 7 params sity maps. Note that those values would be inaccessible in 1.1 of 4 Buzzard sims 4 params a real measurement. The corresponding constraints are very with Y1 covariance 2 params tight, but the fiducial values of our parameters are still well 1.0 contained in the 1σ contour. Then, we also marginalize over the galaxy bias b and the galaxy stochasticity r, demanding that 0 < r 1. The contours now widen, and the fiducial 0.9 ≤

8 values of Ωm and σ8 are still located well within the corre- σ sponding 1σ contour. Finally, we also marginalize over Ω , 0.8 b ns and h100, using the same informative priors that have been used in the DES-Y1 combined probes analysis [29]. The con- 0.7 tours widen further, but our model and our simulations still agree well within 1σ uncertainty. 0.6 In the bottom panel of Figure 10 we repeat this analysis, but now also vary the parameter ∆S3/S3 that was introduced 0.5 in section II C. This parameter allows for deviations of the 3- 0.1 0.2 0.3 0.4 0.5 point statistics of the density field from our fiducial model. Ωm Within our statistical uncertainties we find that the scaling be- 1.0 tween 3-point and 2-point statistics in our simulations is well 8 params described by our fiducial assumptions (∆S3/S3 = 0). 0.8 5 params We repeat this analysis with our alternative shot-noise 3 params model in appendix E. 0.6

0.4 VI. DISCUSSION & CONCLUSIONS 3 /S 3 0.2

S In this work we introduced density split statistics, a tech-

∆ nique to separately measure contributions to weak lensing and 0.0 counts-in-cells from regions of different foreground galaxy density. Based on the pioneering work of Bernardeau [9], 0.2 − Bernardeau & Valageas [10] and Valageas [81] (see also refer- 0.4 ences therein) on modeling the cosmic density PDF we were − able to model the density split lensing signal as well as the 0.6 counts-in-cells histogram from basic principles. With the help − 0.35 0.40 0.45 0.50 0.55 0.60 of this model, we then showed that density split statistics has two features that make it a potentially powerful cosmological Σ8 = σ8√Ωm probe:

FIG. 10. To test our model for possible systematic deviations from it is able to constrain the cosmological parameters Ωm and N-body simulations, we try to recover the Buzzard cosmology in a • σ8 even if the relation of galaxy density and matter density simulated likelihood analysis. Top panel: 1σ and 2σ contours in the is assumed to have 2 degrees of freedom: galaxy bias and Ω σ m- 8 plane from a likelihood computed around the mean of 4 shape galaxy-matter-correlation coefficient, noise free realisations of DES Y1 (but assuming the full covariance matrix for a single DES Y1). The green contours are marginalized it is able to constrain the amplitude of 3-point statistics of the • over Ωb, ns, h100, redMaGiC galaxy bias b and galaxy-matter cor- density field with almost no degeneracy to constraints on the relation coefficient r. For the parameters Ωb, ns, h100 we have as- amplitude of 2-point statistics. sumed the same flat priors as used in the DES Y1 combined probes analysis presented in DES Collaboration [29]. The red contours are In our fiducial model we predict 3-point statistics from cos- marginalized only over b and r and the blue contours only vary Ωm mological perturbation theory. Deviations from that fiducial and σ8. Even when going to this small parameter space, our model prediction may hint to non-standard physics, that affect over- agrees with Buzzard within 1σ errors of DES Y1. Bottom panel: dense and underdense parts of the matter field differently, or to Same contours but in the Σ8-∆S3/S3 plane and varying one addi- any non-linear dynamics or small scale physics that break the tional parameter, ∆S3/S3. scaling relations of ΛCDM perturbation theory. We showed that a DES-Y5 data set combined with data form the cosmic microwave background can measure the amplitude of 3-points statistics to a 1σ accuracy of . 5%. This is a conservative es- timate since our projections neglect the fact that DES-Y5 will be a deeper data set than DES-Y1. Also, we so far neglected the possibility of a combined analysis including density split statistics and measurements of 2-point correlation functions. 24

Using measurements in high-resolution N-body simula- logical Physics at the University of Chicago, the Center for tions we showed that our model of the density split lensing Cosmology and Astro-Particle Physics at the Ohio State Uni- signal and the counts-in-cells histogram is accurate to well versity, the Mitchell Institute for Fundamental Physics and within the statistical uncertainties of the DES-Y1 data set. Es- Astronomy at Texas A&M University, Financiadora de Estu- pecially, in a mock likelihood analysis we were able to recover dos e Projetos, Fundação Carlos Chagas Filho de Amparo à the input cosmology of our simulations to well within DES- Pesquisa do Estado do Rio de Janeiro, Conselho Nacional de Y1 parameter errors. Cosmological constraints from DES-Y1 Desenvolvimento Científico e Tecnológico and the Ministério data based on density split statistics are presented in Gruen da Ciência, Tecnologia e Inovação, the Deutsche Forschungs- et al. [40]. gemeinschaft and the Collaborating Institutions in the Dark Energy Survey. The DES data management system is sup- ported by the National Science Foundation under Grant Num- ACKNOWLEDGMENTS ber AST-1138766. The Collaborating Institutions are Argonne National Lab- OF was supported by SFB-Transregio 33 ‘The Dark Uni- oratory, the University of California at Santa Cruz, the Uni- verse’ by the Deutsche Forschungsgemeinschaft (DFG). Sup- versity of Cambridge, Centro de Investigaciones Enérgeticas, port for DG was provided by NASA through Einstein Post- Medioambientales y Tecnológicas-Madrid, the University of doctoral Fellowship grant number PF5-160138 awarded by Chicago, University College London, the DES-Brazil Consor- the Chandra X-ray Center, which is operated by the Smith- tium, the University of Edinburgh, the Eidgenössische Tech- sonian Astrophysical Observatory for NASA under contract nische Hochschule (ETH) Zürich, Fermi National Accelerator NAS8-03060. OF and SH acknowledge support by the DFG Laboratory, the University of Illinois at Urbana-Champaign, cluster of excellence ‘Origin and Structure of the Universe’ the Institut de Ciències de l’Espai (IEEC/CSIC), the Institut (www.universe-cluster.de). Part of our computa- de Física d’Altes Energies, Lawrence Berkeley National Lab- tions have been carried out on the computing facilities of the oratory, the Ludwig-Maximilians Universität München and Computational Center for Particle and Astrophysics (C2PAP). the associated Excellence Cluster Universe, the University This paper has gone through internal review by the DES of Michigan, the National Optical Astronomy Observatory, collaboration. We want to thank all the members of the DES the University of Nottingham, The Ohio State University, WL, LSS and Theory working groups that have contributed the University of Pennsylvania, the University of Portsmouth, with helpful comments and discussions. We also want to SLAC National Accelerator Laboratory, Stanford University, thank the anonymous journal referee for very helpful com- the University of Sussex, and Texas A&M University. ments. The DES participants from Spanish institutions are par- Funding for the DES Projects has been provided by the U.S. tially supported by MINECO under grants AYA2012-39559, Department of Energy, the U.S. National Science Foundation, ESP2013-48274, FPA2013-47986, and Centro de Excelencia the Ministry of Science and Education of Spain, the Science Severo Ochoa SEV-2012-0234. Research leading to these and Technology Facilities Council of the United Kingdom, the results has received funding from the European Research Higher Education Funding Council for England, the National Council under the European Union’s Seventh Framework Pro- Center for Supercomputing Applications at the University of gramme (FP7/2007-2013) including ERC grant agreements Illinois at Urbana-Champaign, the Kavli Institute of Cosmo- 240672, 291329, and 306478.

[1] Alam S., et al., 2017, MNRAS, 470, 2617 [16] Brainerd T. G., Blandford R. D., Smail I., 1996, ApJ, 466, 623 [2] Baldauf T., Seljak U., Smith R. E., Hamaus N., Desjacques V., [17] Brimioulle F., Seitz S., Lerchster M., Bender R., Snigula J., 2013, Phys. Rev. D, 88, 083507 2013, MNRAS, 432, 1046 [3] Barreira A., Bose S., Li B., Llinares C., 2017, JCAP, 2, 031 [18] Cacciato M., Lahav O., van den Bosch F. C., Hoekstra H., Dekel [4] Bartelmann M., Schneider P., 2001, Physics Reports, 340, 291 A., 2012, MNRAS, 426, 566 [5] Becker M. R., 2013, MNRAS, 435, 115 [19] Chiang C.-T., Wagner C., Sánchez A. G., Schmidt F., Komatsu [6] Becker M. R., et al., 2016, Phys. Rev. D, 94, 022002 E., 2015, Journal of Cosmology and Astroparticle Physics, 9, [7] Benítez N., 2000, ApJ, 536, 571 028 [8] Benjamin J., et al., 2007, MNRAS, 381, 702 [20] Clampitt J., et al., 2017a, MNRAS, 465, 4204 [9] Bernardeau F., 1994, A&A, 291, 697 [21] Clampitt J., et al., 2017b, MNRAS, 465, 4204 [10] Bernardeau F., Valageas P., 2000, A&A, 364, 1 [22] Clerkin L., et al., 2017, MNRAS, 466, 1444 [11] Bernardeau F., van Waerbeke L., Mellier Y., 1997, A&A, 322, [23] Codis S., Pichon C., Bernardeau F., Uhlemann C., Prunet S., 1 2016a, MNRAS, 460, 1549 [12] Bernardeau F., Colombi S., Gaztañaga E., Scoccimarro R., [24] Codis S., Bernardeau F., Pichon C., 2016b, MNRAS, 460, 1598 2002, Physics Reports, 367, 1 [25] Conroy C., Wechsler R. H., Kravtsov A. V., 2006, ApJ, 647, [13] Bernardeau F., Pichon C., Codis S., 2014, Phys. Rev. D, 90, 201 103519 [26] Cooper M. C., et al., 2011, ApJS, 193, 14 [14] Bernardeau F., Codis S., Pichon C., 2015, MNRAS, 449, L105 [27] Crocce M., Pueblas S., Scoccimarro R., 2006, MNRAS, 373, [15] Borisov A., Jain B., 2009, Phys. Rev. D, 79, 103506 369 25

[28] Crocce M., et al., 2016, Monthly Notices of the Royal Astro- Joachimi B., Verde L., 2016, Monthly Notices of the Royal As- nomical Society, 455, 4301 tronomical Society, 456, 278 [29] DES Collaboration 2017, preprint, (arXiv:1708.01530) [73] Smith R. E., et al., 2003, Monthly Notices of the Royal Astro- [30] DeRose J., Wechsler R., Rykoff E., et al., 2018, in prep. nomical Society, 341, 1311 [31] Dekel A., Lahav O., 1999, ApJ, 520, 24 [74] Springel V., 2005, MNRAS, 364, 1105 [32] Dvornik A., et al., 2018, MNRAS, 479, 1240 [75] Springel V., et al., 2005, Nature, 435, 629 [33] Eisenstein D. J., Hu W., 1998, ApJ, 496, 605 [76] Takada M., Jain B., 2002, MNRAS, 337, 875 [34] Elvin-Poole J., et al., 2017, preprint, (arXiv:1708.01536) [77] Takahashi R., Sato M., Nishimichi T., Taruya A., Oguri M., [35] Fosalba P., Gaztanaga E., 1998, MNRAS, 301, 503 2012, ApJ, 761, 152 [36] Friedrich O., Seitz S., Eifler T. F., Gruen D., 2016, MNRAS, [78] Takahashi R., Soma S., Takada M., Kayo I., 2014, Monthly 456, 2662 Notices of the Royal Astronomical Society, 444, 3473 [37] Fu L., et al., 2008, A&A, 479, 9 [79] Troxel M. A., et al., 2017, preprint, (arXiv:1708.01538) [38] Górski K. M., Hivon E., Banday A. J., Wandelt B. D., Hansen [80] Uhlemann C., et al., 2018, MNRAS, 473, 5098 F. K., Reinecke M., Bartelmann M., 2005, ApJ, 622, 759 [81] Valageas P., 2002, A&A, 382, 412 [39] Gruen D., Friedrich O., The DES Collaboration 2016, Monthly [82] Van Waerbeke L., et al., 2000, A&A, 358, 30 Notices of the Royal Astronomical Society, 455, 3367 [83] Wechsler R., DeRose J., Busha M., et al., 2018, in prep. [40] Gruen D., et al., 2018, Phys. Rev. D, 98, 023507 [84] Wilson G., Kaiser N., Luppino G. A., Cowie L. L., 2001, ApJ, [41] Hilbert S., Hartlap J., Schneider P., 2011, Astronomy & Astro- 555, 572 physics, 536, A85 [85] Wittman D. M., Tyson J. A., Kirkman D., Dell’Antonio I., [42] Hildebrandt H., et al., 2017, MNRAS, 465, 1454 Bernstein G., 2000, Nature, 405, 143 [43] Hoyle B., et al., 2018, MNRAS, 478, 592 [86] Xavier H. S., Abdalla F. B., Joachimi B., 2016, MNRAS, 459, [44] Hudson M. J., Gwyn S. D. J., Dahle H., Kaiser N., 1998, ApJ, 3693 503, 531 [87] Zuntz J., et al., 2017, preprint, (arXiv:1708.01533) [45] Huff E., Mandelbaum R., 2017, preprint, [88] van Uitert E., Hoekstra H., Velander M., Gilbank D. G., Glad- (arXiv:1702.02600) ders M. D., Yee H. K. C., 2011, A&A, 534, A14 [46] Jain B., Zhang P., 2008, Phys. Rev. D, 78, 063503 [89] van Uitert E., et al., 2018, MNRAS, 476, 4662 [47] Jee M. J., Tyson J. A., Hilbert S., Schneider M. D., Schmidt S., Wittman D., 2016, ApJ, 824, 77 [48] Jeong D., Komatsu E., 2009, ApJ, 703, 1230 Appendix A: Friedmann equations, linear growth, spherical [49] Kacprzak T., et al., 2016, MNRAS, 463, 3653 collapse and cylindrical collapse [50] Kilbinger M., et al., 2013, MNRAS, 430, 2200 [51] Limber D. N., 1953, ApJ, 117, 134 [52] Lin C.-A., Kilbinger M., 2015, A&A, 576, A24 Throughout this section we set G = 1 = c and we assume [53] Liu X., et al., 2015, MNRAS, 450, 2888 a flat ΛCDM universe. In proper co-moving time t the Fried- [54] Lue A., Scoccimarro R., Starkman G., 2004, Phys. Rev. D, 69, mann equations take the form 044005 [55] MacCrann N., et al., 2018, MNRAS, 2 8π H = (¯ρm +ρ ¯Λ) (A.1) [56] Mandelbaum R., Seljak U., Kauffmann G., Hirata C. M., 3 Brinkmann J., 2006, MNRAS, 368, 715 dH 2 4π [57] Mandelbaum R., Slosar A., Baldauf T., Seljak U., Hirata C. M., + H = (¯ρm 2¯ρΛ) , (A.2) Nakajima R., Reyes R., Smith R. E., 2013, MNRAS, 432, 1544 dt − 3 − [58] Mantz A. B., et al., 2016, MNRAS, 463, 3582 d where H = dt ln a. In conformal time, defined by dt = adτ, [59] Miyazaki S., et al., 2002, PASJ, 54, 833 this changes to [60] Mukhanov V., 2005, Physical Foundations of Cosmology. Cam- bridge University Press, doi:10.2277/0521563984 2 8π 2 [61] Multamäki T., Gaztañaga E., Manera M., 2003, MNRAS, 344, = a (¯ρm +ρ ¯Λ) (A.3) 761 H 3 [62] Peng Y.-j., Lilly S. J., Renzini A., Carollo M., 2012, ApJ, 757, d 4π 2 H = a (¯ρm 2¯ρΛ) , (A.4) 4 dτ − 3 − [63] Pires S., Leonard A., Starck J.-L., 2012, MNRAS, 423, 983 where = d ln a. We will from now put d ˙. [64] Planck Collaboration et al., 2015, preprint, H dτ dτ ≡ (arXiv:1502.01589) In the Newtonian approximation, i.e. on scales much [65] Prat J., et al., 2017, preprint, (arXiv:1708.01537) smaller that the curvature horizon of the universe, the evolu- [66] Reddick R. M., Wechsler R. H., Tinker J. L., Behroozi P. S., tion of a spherical, cylindrical or planar perturbation δ is given 2013, ApJ, 771, 30 by the equation [67] Schrabback T., et al., 2010, A&A, 516, A63 [68] Semboloni E., Schrabback T., van Waerbeke L., Vafaei S., Hart- d2δ dδ N + 1 1 dδ 2 lap J., Hilbert S., 2011, MNRAS, 410, 143 2 + 2H = 4πρ¯mδ(1 + δ) , [69] Sheldon E. S., Huff E. M., 2017, ApJ, 841, 24 dt dt − N 1 + δ dt [70] Simpson F., James J. B., Heavens A. F., Heymans C., 2011, (A.5) Physical Review Letters, 107, 271301 where N = 3 for a spherical perturbation, N = 2 for a cylin- [71] Simpson F., Heavens A. F., Heymans C., 2013, Phys. Rev. D, drical perturlation and N = 1 for a planar perturbation (see 88, 083510 Mukhanov [60] where this is demonstrated for N = 3 and [72] Simpson F., Harnois-Déraps J., Heymans C., Jimenez R., N = 1). 26

In conformal time this equation reads which is indeed independent of the particular shape of the per- turbation. ˙2 N + 1 δ 2 δ¨ + δ˙ = 4πρ¯ma δ(1 + δ) . (A.6) H − N 1 + δ To linear order in δ this becomes

2 δ¨ + δ˙ = 4πρ¯ma δ , (A.7) H

Appendix B: ΛCDM perturbation theory

In the following we are using the convention that the Fourier transform of a real space function f(x) is given by Z 3 d x ixk f˜(k) = f(x) e− . (B.1) (2π)3 Consider the matter density contrast δ and the divergence of the velocity field θ = ∇v. In the Newtonian approximation the Fourier space equations of motion of δ and θ are [cf. 12] ˜ Z ∂δ(k, τ) 3 3 + θ˜(k, τ) = d k1d k2 δD(k k12) α(k1, k2) δ˜(k1, τ) θ˜(k2, τ) ∂τ − − ˜ 0 2 Z ∂θ(k, τ) 3ΩmH0 3 3 + θ˜(k, τ) + δ˜(k, τ) = d k1d k2 δD(k k12) β(k1, k2) θ˜(k1, τ) θ˜(k2, τ) , (B.2) ∂τ H 2a − − where k12 = k1 + k2 and α and β are given by   1 k1 k2 k1 k2 α(k1, k2) = 1 + · + 2 k1k2 k2 k1   2 1 k1 k2 k1 k2 (k1 k2) β(k1, k2) = · + + 2· 2 . (B.3) 2 k1k2 k2 k1 k1k2 In the following we will abbreviate the integrals involving α and β as α[δ,˜ θ,˜ k] and β[θ,˜ θ,˜ k]. In perturbation theory we write δ˜ and θ˜ as

X∞ ∂ ln D+(τ) X∞ δ˜(k, τ) = δ (k, τ) and θ˜(k, τ) = θ (k, τ) , (B.4) n ∂τ n n=1 − n=1 where δn and θn are of order n in the linearly approximated fields δ1 and θ1 and D+(τ) is the linear growth factor. (We will ignore the decaying mode of linear growth here.) At linear order we have

D+(τ) δ1(k, τ) = θ1(k, τ) = δ1(k, τ0) D+(τ)δ1,1(k) , (B.5) D+(τ0) ≡ where we have assumed that D+(τ0) = 1 at present time τ0 and introduced the notation δ1,1(k) = δ1(k, τ0) whose purpose will become clear at the end of this section. To get δ˜ at second order we first note that

∂θ˜(k, τ) 1 ∂   + θ˜(k, τ) = aθ˜(k, τ) . (B.6) ∂τ H a ∂τ Hence, multiplying the first of equations B.2 with a and differentiating wrt. τ and then multiplying with 1/a we get " # " # ∂2δ˜(k, τ) ∂δ˜(k, τ) ∂θ˜(k, τ) ∂δ˜ ∂θ˜ h i + + + θ˜(k, τ) = α , θ,˜ k α δ,˜ , k α δ,˜ θ,˜ k . (B.7) ∂2τ H ∂τ ∂τ H − ∂τ − ∂τ − H

Now the second of equations B.2 can be used to eliminate θ˜ from the right-hand-side, giving

∂2δ˜(k, τ) ∂δ˜(k, τ) 3Ω0 H2 ∂δ˜ ∂θ˜ + m 0 δ˜(k, τ) = β[θ,˜ θ,˜ k] α[ , θ,˜ k] α[δ,˜ , k] α[δ,˜ θ,˜ k] . (B.8) ∂2τ H ∂τ − 2a − ∂τ − ∂τ − H 27

At second order in perturbation theory this equation becomes

2 0 2  2 2  2! ∂ δ2(k, τ) ∂δ2(k, τ) 3ΩmH0 ∂D+ ∂ D+ ∂D+ ∂D+ + δ2(k, τ) = β[δ1,1, δ1,1, k] + D + D + α[δ1,1, δ1,1, k] ∂2τ H ∂τ − 2a ∂τ ∂τ 2 H ∂τ ∂τ

 2 0 2  2! ∂D+ 3ΩmH0 2 ∂D+ = β[δ1,1, δ1,1, k] + D + α[δ1,1, δ1,1, k] ∂τ 2a ∂τ

0 2  2! 3ΩmH0 2 ∂D+ = α[δ1,1, δ1,1, k] D + 2 2a ∂τ  2 ∂D+ + (β[δ1,1, δ1,1, k] α[δ1,1, δ1,1, k]) . (B.9) − ∂τ This is solved by

δ2(k, τ) = D2,1(τ)δ2,1(k) + D2,2(τ)δ2,2(k) (B.10) where

2 D2,1(τ) D (τ) , δ2,1(k) = α[δ1,1, δ1,1, k] , δ2,2(k) = β[δ1,1, δ1,1, k] α[δ1,1, δ1,1, k] (B.11) ≡ + − and D2,2 is given by the differential equation

2 0 2  2 ∂ D2,2(τ) ∂D2,2(τ) 3ΩmH0 ∂D+ + D2,2(τ) = . (B.12) ∂2τ H ∂τ − 2a ∂τ

1. Second order of δ in Einstein-de Sitter universe

2 Let us define 1 µ D2,2/D . Then the general solution to B.9 is given by − ≡ + 2 δ2(k, τ) = D ([1 µ]β[δ1,1, δ1,1, k] + µα[δ1,1, δ1,1, k]) . (B.13) + − In an Einstein-de Sitter Universe where Ω0 = 1 and D a we have m ≡ 2 2 5 D2,2 = D , µ = (B.14) 7 + 7 and δ2 is hence given by   2 2 5 δ2(k, τ) = D β[δ1,1, δ1,1, k] + α[δ1,1, δ1,1, k] + 7 7 Z 3 3 = d k1d k2 δD(k k12) F2(k1, k2) δ1,1(k1) δ1,1(k2) − (B.15) with 5 2 F2(k1, k2) = α(k1, k2) + β(k1, k2) 7 7   2 5 1 k1 k2 k1 k2 2 (k1 k2) = + · + + 2· 2 . (B.16) 7 2 k1k2 k2 k1 7 k1k2

2. Second order of δ in ΛCDM universe

In a general ΛCDM universe the function F2 becomes time dependent. It is given by

  2 1 k1 k2 k1 k2 (k1 k2) F2(k1, k2, τ) = µ(τ) + · + + [1 µ(τ)] 2· 2 . (B.17) 2 k1k2 k2 k1 − k1k2 28

A useful property or this kernel is that

1 k2 k4 F2(k, k, τ) = µ(τ) + − (1 + 1) + [1 µ(τ)] = µ(τ) 1 + 1 µ(τ) = 0 . (B.18) − 2 k2 − k4 − −

Denoting the angle between k1 and k2 with φ one can also arive at the following form of F2(k1, k2, τ) which will be useful when computing the skewness of matter inside a long cylinder:     1 k1 k2 2 F2(k1, k2, τ) = 1 + cos φ + 1 + cos φ + [1 µ(τ)](cos φ 1) . (B.19) 2 k2 k1 − −

3. Bispectrum and 3-point function at leading order

The bispectrum B(k1, k2, k3, τ) of δ is defined by

δ˜(k1, τ)δ˜(k2, τ)δ˜(k3, τ) = δD(k1 + k2 + k3) B(k1, k2, k3, τ) . (B.20) h i At leading order in perturbation theory this can be calculated as Z 2 3 3 δ˜(k1, τ)δ˜(k2, τ)δ˜(k3, τ) 2nd. = D d q1d q2 δD(k3 q12) F2(q1, q2, τ) δ1,1(k1) δ1,1(k2)δ1,1(q1) δ1,1(q2) h i + − h i + cycl., (B.21) where 0cycl.0 indicates that the integral on the right-hand-side should appear for all possible permutations of k1, k3 and k3. Since we assume the linear density field to be a Gaussian random field, the expectation value on the left-hand-side factorizes as

δ1,1(k1)δ1,1(k2)δ1,1(q1)δ1,1(q2) = δ1,1(k1)δ1,1(k2) δ1,1(q1)δ1,1(q2) + δ1,1(k1)δ1,1(q1) δ1,1(k2)δ1,1(q2) h h ih i h ih i + δ1,1(k1)δ1,1(q2) δ1,1(k2)δ1,1(q1) h ih i = δD(k1 + k2)δD(q1 + q2)Plin,0(k1)Plin,0(q1) + δD(k1 + q1)δD(k2 + q2)Plin,0(k1)Plin,0(q2)

+δD(k1 + q2)δD(k2 + q1)Plin,0(k1)Plin,0(q1) (B.22) Because of equation B.18 the contribution of the first term to the bispectrum is zero. Using the symmetry 1 2 between the second and third term we hence get ↔

2 δ˜(k1, τ)δ˜(k2, τ)δ˜(k3, τ) 2nd. = 2D δD(k1 + k2 + k3) F2(k1, k2, τ)Plin,0(k1)Plin,0(k2) h i + + cycl.. (B.23)

4. Variance and skewness of long cylinder at leading order in perturbation theory

Consider a cylinder with radius R and length L. In Fourier space the tophat filter for this cylinder is given by

1 WR,L(k) = WL(k )WR(k ) (B.24) (2π)3 k ⊥ where we denote the component of k parallel to the cylinder axis with k and the components orthogonal to it are represented k by the two-dimensional vector k and WL and WR given by ⊥

sin(Lk /2) 2J1(Rk ) WL(k ) = k ,WR(k ) = ⊥ . (B.25) k Lk /2 ⊥ Rk k ⊥

Here k = k and Jν are the cylindrical bessel functions. At leading order or tree level in perturbation theory the variance of matter⊥ contrast| ⊥ within| the cylinder is then given by Z 2 2 2 2 δR,L tree(τ) = D+(τ) dk ,1dk ,2d k ,1d k ,2 WL(k ,1) WL(k ,2) WR(k ,1) WR(k ,2) δ1,1(k1)δ1,1(k2) h i k k ⊥ ⊥ k k ⊥ ⊥ h i Z 2 2 2 2 = D+(τ) dk d k WL(k ) WR(k ) Plin,0(k) , (B.26) k ⊥ k ⊥ 29 where Plin,0(k) is today’s linear power spectrum and in our Fourier convention the factors of 2π in Eqn. B.24 cancel with corresponding factors from the convolution theorem. For L R we can actually approximate WL by  2 2π WL(k ) δD(k ) (B.27) k ≈ L k such that in this limit we get 2 2 Z 2 (2π) D+ 2 δ tree(τ) dk k WR(k) Plin,0(k) . (B.28) h R,Li ≈ L The third moment at tree level is given by Z 3 2 2 2 2 δR,L tree(τ) = 3D+ dk ,1dk ,2dq d k ,1d k ,2d q WL(k ,1) WL(k ,2) WL(q ) WR(k ,1) WR(k ,2) WR(q ) h i k k k ⊥ ⊥ ⊥ k k k ⊥ ⊥ ⊥ δ1,1(k1)δ1,1(k2)δ2(q, τ) h Z i 4 2 2 2 3 3 = 3D+ dk ,1dk ,2dq d k ,1d k ,2d q d q1d q2 WL(k ,1) WL(k ,2) WL(q ) WR(k ,1) WR(k ,2) WR(q ) k k k ⊥ ⊥ ⊥ k k k ⊥ ⊥ ⊥

δD(q q1 q2)F2(q1, q2, τ) δ1,1(k1)δ1,1(k2)δ1,1(q1)δ1,1(q2) Z− − h i 4 2 2 3 3 = 3D+ dk ,1dk ,2d k ,1d k ,2d q1d q2 WL(k ,1) WL(k ,2) WR(k ,1) WR(k ,2) k k ⊥ ⊥ k k ⊥ ⊥

WR(q ,1 + q ,2) WL(q ,1 + q ,2)F2(q1, q2, τ) δ1,1(k1)δ1,1(k2)δ1,1(q1)δ1,1(q2) . (B.29) ⊥ ⊥ k k h i Since we assume the linear density field to be a Gaussian random field, the expectation value on the left-hand-side factorizes as

δ1,1(k1)δ1,1(k2)δ1,1(q1)δ1,1(q2) = δ1,1(k1)δ1,1(k2) δ1,1(q1)δ1,1(q2) + δ1,1(k1)δ1,1(q1) δ1,1(k2)δ1,1(q2) h h ih i h ih i + δ1,1(k1)δ1,1(q2) δ1,1(k2)δ1,1(q1) h ih i = δD(k1 + k2)δD(q1 + q2)Plin,0(k1)Plin,0(q1) + δD(k1 + q1)δD(k2 + q2)Plin,0(k1)Plin,0(q2)

+δD(k1 + q2)δD(k2 + q1)Plin,0(k1)Plin,0(q1) (B.30) Because of equation B.18 the contribution of the first term to the skewness is zero. Using the symmetry 1 2 between the second and third term we hence get ↔ Z 3 4 2 2 δR,L tree(τ) = 6D+ dq ,1dq ,2d q ,1d q ,2 WL(q ,1) WL(q ,2) WL(q ,1 + q ,2) WR(q1) WR(q2) WR(q1 + q2) h i i k k ⊥ ⊥ k k k k Plin,0(q1)Plin,0(q2)F2(q1, q2, τ) . (B.31) For L R we can use the approximation  2 (2π) 2 WL(q ,1) WL(q ,2) WL(q ,1 + q ,2) δD(q ,1, q ,2) . (B.32) k k k k ≈ L2 k k This gives

2 Z 3 4 (2π) 2 2 δ tree(τ) = 6D d q1d q2 WR(q1) WR(q2) WR(q1 + q2) Plin,0(q1)Plin,0(q2)F2(q1, q2, τ) , (B.33) h R,Li i + L2 where we will consider all vectors to be 2-dimensional from now on. Using equation B.19 to express F2 interms of q1, q2 and φ we can simplify this to 3 4 Z Z q  3 (2π) 6D+ 2 2 δ tree(τ) = dq1dq2 q1WR(q1) q2WR(q2) Plin,0(q1)Plin,0(q2) dφ WR q + q + 2q1q2 cos φ F2(q1, q2, φ, τ) . h R,Li i L2 1 2 (B.34) Using relations given in Bernardeau (1995) or Buchalter et al. (2000) one can simplify the integral over φ as Z q  2 2 dφ WR q1 + q2 + 2q1q2 cos φ F2(q1, q2, φ, τ) Z q      1 2 2 k1 k2 = dφ WR q1 + q2 + 2q1q2 cos φ 1 + cos φ + 1 + cos φ 2 k2 k1 30

Z q  2 2 2 +[1 µ(τ)] dφ WR q + q + 2q1q2 cos φ (cos φ 1) − 1 2 − ( ) π ∂WR(x) ∂WR(x) = 2πWR(q1)WR(q2) + WR(q1) Rq2 + WR(q2) Rq1 π[1 µ(τ)]WR(q1)WR(q2) 2 ∂x ∂x x=Rq2 x=Rq1 − − π ∂ = π[1 + µ(τ)]WR(q1)WR(q2) + WR(q1)WR(q2) . (B.35) 2 ∂ ln R { }

For the third moment of δR,L this gives 4 Z 3 4 (2π) 2 2 δ tree(τ) = [1 + µ(τ)]3D (τ) dq1dq2 q1q2 WR(q1) WR(q2) Plin,0(q1)Plin,0(q2) h R,Li + L2 4 Z 4 (2π) ∂ +3D (τ) dq1dq2 q1q2 WR(q1) WR(q2) WR(q1)WR(q2) Plin,0(q1)Plin,0(q2) + 2L2 ∂ ln R { } 2 (2π)2D2 (τ) Z  = 3[1 + µ(τ)] + dq q W (q)2 P (q) L 1 R lin,0 4 4 Z 3D+(τ) (2π) ∂ 2 2 + dq1dq2 q1q2WR(q1) WR(q2) Plin,0(q1)Plin,0(q2) 4 L2 ∂ ln R 2 (2π)2D2 (τ) Z  = 3[1 + µ(τ)] + dq q W (q)2 P (q) L 1 R lin,0  2 2 Z 2 3 ∂ (2π) D+(τ) 2 + dq1 q WR(q) Plin,0(q) 4 ∂ ln R L

2 2 3 ∂ 2 2 = 3[1 + µ(τ)] δ tree(τ) + δ tree(τ) . (B.36) h R,Li 4 ∂ ln R h R,Li Especially we have

3 2 δ tree(τ) 3 ∂ ln δ tree(τ) h R,Li h R,Li S3 2 2 = 3[1 + µ(τ)] + . (B.37) ≡ δ tree(τ) 2 ∂ ln R h R,Li n For an Einstein-de Sitter universe and a power law power spectrum P (k) k this gives S3 = 36/7 3/2 (n + 2). The leading ∼ − order prediction for S3 is surprisingly good, even in the mildly non-linear regime [see 12, and references therein]. Hence in order to predict the non-linear skewness, we simply employ the approximation

3 2 2 δR,L non lin.(τ) S3 δR,L non lin.(τ) , (B.38) h i − ≈ h i − where we compute the non-linear variance with the use of halofit as detailed in Takahashi et al. [78] which is a revised version of Smith et al. [73].

2 5. The moment hδRA,L δRB ,Litree

For predicting the density split lensing signal we are also interested in the moment δ2 δ , where R and R are RA,L RB ,L tree A B two different Radii. The above derivations can be generalized to give h i  1 ∂ ln Var(R ) ∂ ln Cov(R ,R ) δ2 δ τ R R ,R µ τ A A B RA,L RB ,L tree( ) = Var( A) Cov( A B) 2[1 + ( )] + + h i 2 ∂ ln RA ∂ ln RA   2 ∂ ln Cov(RA,RB) + Cov(RA,RB) [1 + µ(τ)] + (B.39) ∂ ln RB where we defined

2 Var(RA) = δ tree , Cov(RA,RB) = δR ,L δR ,L tree . (B.40) h RA,Li h A B i

To correct this expression for the non-linear evolution of the power spectrum, we compute Var(RA) and Cov(RA,RB) with our halofit power spectrum whenever they appear outside of the logarithmic derivatives. This is a generalization of the rescaling of Var(RA) by means of S3. 31

For RB RA this rescaling is dominated by first term on the right hand side of equation B.39. For RB RA it reduces to equation B.38. As a consequence, using the procedure described around IV.39 to correct for the non-linear≈ power spectrum evolution yields a predcition for the density split lensing signal that is almost identical to the procedure described here. Also, it can be considered accurate to the extend that equationB.38 is accurate. We nevertheless rescale the 3rd order moments in the more elaborate way described here.

n ng 6. The moment hδ δR ,Litree δg,T = [e 1] δg,0 , (D.3) RA,L B − where n and n have a joint Gaussian distribution and Using a diagrammatic representation of perturbation theory m g δ = bδ . The variances of n and n are given by (see e.g. [12]) one can see that the tree-level result for the g,0 m,0 m g n ( ) moment δR ,L δRB ,L c will consist of terms that scale as h A i 2 Varm σm = ln 1 + k n k δ2 Cov(RA,RB) Var(RA) − , 1 k n . (B.41) m,0 ∼ ≤ ≤ ( ) Var For RB RA each of these scalings reduces to σ2 = ln 1 + g n ≈ ∼ g 2 Var(RA) (cf. B.38 and the definition of Sn+1 in IV.16). δg,0 On the other hand, for RB RA the terms scaling as 2 n 1 = σm (D.4) Cov(RA,RB) Var(RA) − are the dominant contribu- ∼ tions (cf. the last section for the case n = 2). This is and their covariance is given by why we use IV.39 when rescaling moments with n > 2 in   Gcyl.,θ w,θw,L(ql(w)Ly, w) (see also IV.26). Covmg T ξmg = ln 1 + δm,0δg,0 ( ) Appendix C: Comparison with Millennium simulation Varm = ln 1 + r 2 . (D.5) δm,0 In Figure 11, we compare our model for the PDF of pro- Let us denote the correlation coefficient of the Gaussian field jected density contrast to another set of N-body simulations, by the Millennium Run [MR 75]. The MR has a smaller simu- 1 3 lation volume of only (500h− Mpc) co-moving, but a force ξmg 1 ρ = resolution of 5h kpc that is 4-10 times higher than that of 2 − σm the Buzzard simulations. The fiducial model and the log- n o Varm ln 1 + r 2 normal model describe the distribution of δm,T measured δm,0 from the MR well considering the large statistical uncertainty = n o . (D.6) Varm ln 1 + 2 on p(δm,T ) due to the limited simulated sky area. δm,0 Note that ρ will depend on scale even of b and r do not. Appendix D: Galaxy stochasticity Now we want to compute the conditional moments 2 δg,T δm,T and δ δm,T . First, h | i h g,T | i Consider the field of galaxy density contrast δg,T and the n n n +σ2(1 ρ2)/2 e g n = e g m g field of matter density contrast δ , where both fields are as- m h | i − m,T h | i 2 2 2 ρ(nm+σ /2) σ ρ /2 sumed to be smoothed over a fix circular aperture. The num- = e m − g 2 2 σ (ρ ρ )/2 ρnm ber of galaxies found inside such an aperture is assumed to be = e g − e . (D.7) a Poissonian random variable with first and second moments for a given value of δg,T are given by Second,

 2 2  2 2 ng σg (1 ρ ) 2 ng nm +σg (1 ρ ) Nˆ δg,T = N¯(1 + δg,T ) (D.1) Var (e nm) = e − 1 e h | i − h | i | −  2 2  2 2 σ (1 ρ ) σ (ρ ρ ) 2ρnm and = e g − 1 e g − e . (D.8) − 2 2 2 Nˆ δg,T = N¯(1 + δg,T ) + N¯ (1 + δg,T ) . (D.2) h | i Now what is Var(Nˆ δm,T )? 2 | Let Varm be the variance of δm,T and Varg = b Varm the Nˆ 2 δ = Nˆ 2 n variance of δg,T , where b is the galaxy bias. Then the galaxy g,T m h | i hZ | i stochasticity r is defined by Covmg = rbVarm, i.e. it is the = dδg,T p(δg,T nm) correlation coefficient of δg,T and δm,T . | × We will now assume both δg,T and δm,T to be joint log- 2 2 N¯[1 + δg,T ] + N¯ [1 + δg,T ] normal random variables, i.e. × Z ¯ ¯ 2 nm = N + N + dδg,T p(δg,T nm) δm,T = [e 1] δm,0 − | × 32

3.0 4 9 params θT =10 arcmin 2.5 8 params 3 5 params 2.0 2 params )

m,T 1.5 2 δ 1 p( 1.0 α 1 0.5

0.0 0 0.4 0.2 0.0 0.2 0.4 0.6 0.8 − − δm,T 3.5 1 − Millenium 0.5 1.0 1.5 2.0 2.5 3.0 3.0 θT =20 arcmin cyl. collapse α0 2.5 log-normal FIG. 12. 1σ and 2σ contours in the α0-α1 plane from a likelihood ) 2.0 Gaussian computed around the mean of 4 shape noise free realisations of DES

m,T Y1 (but assuming the full covariance matrix for a single DES Y1). δ 1.5 p( The blue contour only varies α0 and α1. The red contour marginal- izes over Ωm, σ8 and galaxy bias b. The green contour addition- 1.0 ally marginalizes over Ωb, ns, h100, assuming the priors used by [29]. And the black contour also allows variation of the parameter 0.5 ∆S3/S3. Dotted lines show the values of α0 and α1 that were found to describe our mock data best in section IV C 2. 0.0 0.4 0.2 0.0 0.2 0.4 0.6 − − δm,T 2 2 2  δg,T [N¯ + 2N¯ ] + N¯ δ × g,T 2 2 3.5 = N¯ + N¯ + [N¯ + 2N¯ ] δg,T nm + θT =30 arcmin 2 h | i 2 +N¯ Var (δg,T nm) + δg,T nm 3.0 | h | i 2.5 2 ) Var(Nˆ δm,T ) = N¯(1 + δg,T nm ) + N¯ Var (δg,T nm) 2.0 ⇒ | h | i | m,T 2 2 ng

δ ¯ ¯ = N(1 + δg,T nm ) + N δg,0Var (e nm)

p( h | i | 1.5 (D.9) 1.0 The probability P (NA δm,T ) can be computed in a similar 0.5 way, by numerically evaluating|

0.0 Z 0.4 0.2 0.0 0.2 0.4 0.6 P (NA δm,T ) = dδg,T p(δg,T δm,T ) P (NA δg,T ) , − − δ | | | m,T (D.10) where p(δg,T δm,T ) can be computed from basic relations for FIG. 11. The PDF of projected density contrast δm,T in the Millen- | nium Run (MR) compared to our model. In each plot, the black line joint log-normal random variables. 2 shows a histogram of δm,T measured from 64 patches of 4 × 4 deg made from the MR by projecting the 3D density contrast with a con- stant selection function ql between 0.19 . z . 0.43, i.e. with a Appendix E: Validation of alternative shot-noise model constant co-moving density between those redshifts. The blue lines display the PDF predicted by our PT-motivated log-normal model, In our data analysis [40] we investigate both shot-noise and the red lines show a Gaussian PDF with the same variance. The grey band is using the subsample covariance to estimate the error on parametrizations introduced in section IV C. We hence check the mean of all patches [36]. whether our alternative shot-noise parametrization, i.e. the one that uses three parameters to describe the relation between matter and galaxies (b, α0 and α1, cf. section IV C 2), recov- ers the true cosmology and shot-noise parameters of our mock data. 33

1.2 In figure 12 we show the posterior constraints derived for chains around mean 8 params the two shot-noise parameters α0 and α1, when marginalizing 1.1 of 4 Buzzard sims 5 params over different sets of model parameters. Our priors 0.1 < with Y1 covariance 2 params α0 < 3.0 and 1.0 < α1 < 4.0 are mildly informative. We − 1.0 however expect that even stronger priors can be motivated (cf. our discussion in section IV C 2) and will investigate this in 0.9 future work. Figure 13 shows that our alternative shot-noise

8 parametrization also recovers the correct Buzzard cosmology σ (cf. figure 10, which presents the same test for our baseline 0.8 model). 0.7

0.6

0.5 0.1 0.2 0.3 0.4 0.5

Ωm 1.0 9 params 0.8 6 params 3 params 0.6

0.4 3 /S 3 0.2 S ∆ 0.0

0.2 − 0.4 − 0.6 − 0.35 0.40 0.45 0.50 0.55 0.60

Σ8 = σ8√Ωm

FIG. 13. In analogy to figure 10, we test whether our alternative shot-noise parametrization can recover the Buzzard cosmology in a simulated likelihood analysis. Top panel: 1σ and 2σ contours in the Ωm-σ8 plane from a likelihood computed around the mean of 4 shape noise free realisations of DES Y1 (but assuming the full covariance matrix for a single DES Y1). The green contours are marginalized over Ωb, ns, h100, redMaGiC galaxy bias b as well as the shot-noise parameters α0 and α1. For the parameters Ωb, ns, h100 we have as- sumed the same flat priors as used in the DES Y1 combined probes analysis presented in Abbott et al. (in prep.). The red contours are marginalized only over bias and shot-noise parameters and the blue contours only vary Ωm and σ8. Even when going to this small pa- rameter space, our model agrees with Buzzard within 1σ errors of DES Y1. Bottom panel: Same contours but in the Σ8-∆S3/S3 plane and varying one additional parameter, ∆S3/S3. Dotted lines show the true Buzzard cosmology and our fiducial value of ∆S3/S3 = 0.