<<

DES-2016-0211 -PUB-17-279-PPD Survey Year 1 Results: Cosmological Constraints from Cosmic Shear

1, 2, 1, 2 3 4, 5 6 7, 8 6, 9, 1, 10 M. A. Troxel, ⇤ N. MacCrann, J. Zuntz, T. F. Eifler, E. Krause, S. Dodelson, D. Gruen, † J. Blazek, O. Friedrich,11, 12 S. Samuroff,13 J. Prat,14 L. F. Secco,15 C. Davis,6 A. Ferte,´ 3 J. DeRose,16, 6 A. Alarcon,17 A. Amara,18 E. Baxter,15 M. R. Becker,16, 6 G. M. Bernstein,15 S. L. Bridle,13 R. Cawthon,8 C. Chang,8 A. Choi,1 J. De Vicente,19 A. Drlica-Wagner,7 J. Elvin-Poole,13 J. Frieman,7, 8 M. Gatti,14 W. G. Hartley,20, 18 K. Honscheid,1, 2 B. Hoyle,12 E. M. Huff,5 D. Huterer,21 B. Jain,15 M. Jarvis,15 T. Kacprzak,18 D. Kirk,20 N. Kokron,22, 23 C. Krawiec,15 O. Lahav,20 A. R. Liddle,3 J. Peacock,3 M. M. Rau,12 A. Refregier,18 R. P. Rollins,13 E. Rozo,24 E. S. Rykoff,6, 9 C. Sanchez,´ 14 I. Sevilla-Noarbe,19 E. Sheldon,25 A. Stebbins,7 T. N. Varga,11, 12 P. Vielzeuf,14 M. Wang,7 R. H. Wechsler,16, 6, 9 B. Yanny,7 T. M. C. Abbott,26 F. B. Abdalla,20, 27 S. Allam,7 J. Annis,7 K. Bechtol,28 A. Benoit-Levy,´ 29, 20, 30 E. Bertin,29, 30 D. Brooks,20 E. Buckley-Geer,7 D. L. Burke,6, 9 A. Carnero Rosell,23, 31 M. Carrasco Kind,32, 33 J. Carretero,14 F. J. Castander,17 M. Crocce,17 C. E. Cunha,6 C. B. D’Andrea,15 L. N. da Costa,23, 31 D. L. DePoy,34 S. Desai,35 H. T. Diehl,7 J. P. Dietrich,36, 37 P. Doel,20 E. Fernandez,14 B. Flaugher,7 P. Fosalba,17 J. Garc´ıa-Bellido,38 E. Gaztanaga,17 D. W. Gerdes,39, 21 T. Giannantonio,40, 41, 12 D. A. Goldstein,42, 43 R. A. Gruendl,32, 33 J. Gschwend,23, 31 G. Gutierrez,7 D. J. James,44 T. Jeltema,45 M. W. G. Johnson,33 M. D. Johnson,33 S. Kent,7, 8 K. Kuehn,46 S. Kuhlmann,47 N. Kuropatkin,7 T. S. Li,7 M. Lima,22, 23 H. Lin,7 M. A. G. Maia,23, 31 M. March,15 J. L. Marshall,34 P. Martini,1, 48 P. Melchior,49 F. Menanteau,32, 33 R. Miquel,50, 14 J. J. Mohr,36, 37, 11 E. Neilsen,7 R. C. Nichol,51 B. Nord,7 D. Petravick,33 A. A. Plazas,5 A. K. Romer,52 A. Roodman,6, 9 M. Sako,15 E. Sanchez,19 V. Scarpine,7 R. Schindler,9 M. Schubnell,21 M. Smith,53 R. C. Smith,26 M. Soares-Santos,7 F. Sobreira,54, 23 E. Suchyta,55 M. E. C. Swanson,33 G. Tarle,21 D. Thomas,51 D. L. Tucker,7 V. Vikram,47 A. R. Walker,26 J. Weller,36, 11, 12 and Y. Zhang7 (DES Collaboration) 1Center for and Astro-Particle Physics, The Ohio State University, Columbus, OH 43210, USA 2Department of Physics, The Ohio State University, Columbus, OH 43210, USA 3Institute for Astronomy, University of Edinburgh, Edinburgh EH9 3HJ, UK 4Department of Physics, California Institute of Technology, Pasadena, CA 91125, USA 5Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Dr., Pasadena, CA 91109, USA 6Kavli Institute for Particle Astrophysics & Cosmology, P. O. Box 2450, Stanford University, Stanford, CA 94305, USA 7Fermi National Accelerator Laboratory, P. O. Box 500, Batavia, IL 60510, USA 8Kavli Institute for Cosmological Physics, University of Chicago, Chicago, IL 60637, USA 9SLAC National Accelerator Laboratory, Menlo Park, CA 94025, USA 10Institute of Physics, Laboratory of Astrophysics, Ecole´ Polytechnique Fed´ erale´ de Lausanne (EPFL), Observatoire de Sauverny, 1290 Versoix, Switzerland 11Max Institute for Extraterrestrial Physics, Giessenbachstrasse, 85748 Garching, Germany 12Universitats-Sternwarte,¨ Fakultat¨ fur¨ Physik, Ludwig-Maximilians Universitat¨ Munchen,¨ Scheinerstr. 1, 81679 Munchen,¨ Germany 13Jodrell Bank Center for Astrophysics, School of Physics and Astronomy, University of Manchester, Oxford Road, Manchester, M13 9PL, UK 14Institut de F´ısica d’Altes Energies (IFAE), The Barcelona Institute of Science and Technology, Campus UAB, 08193 Bellaterra (Barcelona) Spain 15Department of Physics and Astronomy, University of Pennsylvania, Philadelphia, PA 19104, USA 16Department of Physics, Stanford University, 382 Via Pueblo Mall, Stanford, CA 94305, USA 17Institute of Space Sciences, IEEC-CSIC, Campus UAB, Carrer de Can Magrans, s/n, 08193 Barcelona, Spain 18Department of Physics, ETH Zurich, Wolfgang-Pauli-Strasse 16, CH-8093 Zurich, Switzerland 19Centro de Investigaciones Energeticas,´ Medioambientales y Tecnologicas´ (CIEMAT), Madrid, Spain 20Department of Physics & Astronomy, University College London, Gower Street, London, WC1E 6BT, UK 21Department of Physics, University of Michigan, Ann Arbor, MI 48109, USA 22Departamento de F´ısica Matematica,´ Instituto de F´ısica, Universidade de Sao˜ Paulo, CP 66318, Sao˜ Paulo, SP, 05314-970, Brazil 23Laboratorio´ Interinstitucional de e-Astronomia - LIneA, Rua Gal. Jose´ Cristino 77, Rio de Janeiro, RJ - 20921-400, Brazil 24Department of Physics, University of Arizona, Tucson, AZ 85721, USA 25Brookhaven National Laboratory, Bldg 510, Upton, NY 11973, USA 26Cerro Tololo Inter-American Observatory, National Optical Astronomy Observatory, Casilla 603, La Serena, Chile 27Department of Physics and Electronics, Rhodes University, PO Box 94, Grahamstown, 6140, South Africa 28LSST, 933 North Cherry Avenue, Tucson, AZ 85721, USA 29CNRS, UMR 7095, Institut d’Astrophysique de Paris, F-75014, Paris, France 30Sorbonne Universites,´ UPMC Univ Paris 06, UMR 7095, Institut d’Astrophysique de Paris, F-75014, Paris, France 31Observatorio´ Nacional, Rua Gal. Jose´ Cristino 77, Rio de Janeiro, RJ - 20921-400, Brazil 32Department of Astronomy, University of Illinois, 1002 W. Green Street, Urbana, IL 61801, USA 2

33National Center for Supercomputing Applications, 1205 West Clark St., Urbana, IL 61801, USA 34George 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 35Department of Physics, IIT Hyderabad, Kandi, Telangana 502285, India 36Excellence Cluster , Boltzmannstr. 2, 85748 Garching, Germany 37Faculty of Physics, Ludwig-Maximilians-Universitat,¨ Scheinerstr. 1, 81679 Munich, Germany 38Instituto de Fisica Teorica UAM/CSIC, Universidad Autonoma de Madrid, 28049 Madrid, Spain 39Department of Astronomy, University of Michigan, Ann Arbor, MI 48109, USA 40Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 41Kavli Institute for Cosmology, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK 42Department of Astronomy, University of California, Berkeley, 501 Campbell Hall, Berkeley, CA 94720, USA 43Lawrence Berkeley National Laboratory, 1 Cyclotron Road, Berkeley, CA 94720, USA 44Astronomy Department, University of Washington, Box 351580, Seattle, WA 98195, USA 45Santa Cruz Institute for Particle Physics, Santa Cruz, CA 95064, USA 46Australian Astronomical Observatory, North Ryde, NSW 2113, Australia 47Argonne National Laboratory, 9700 South Cass Avenue, Lemont, IL 60439, USA 48Department of Astronomy, The Ohio State University, Columbus, OH 43210, USA 49Department of Astrophysical Sciences, Princeton University, Peyton Hall, Princeton, NJ 08544, USA 50Institucio´ Catalana de Recerca i Estudis Avanc¸ats, E-08010 Barcelona, Spain 51Institute of Cosmology & Gravitation, University of Portsmouth, Portsmouth, PO1 3FX, UK 52Department of Physics and Astronomy, Pevensey Building, University of Sussex, Brighton, BN1 9QH, UK 53School of Physics and Astronomy, University of Southampton, Southampton, SO17 1BJ, UK 54Instituto de F´ısica Gleb Wataghin, Universidade Estadual de Campinas, 13083-859, Campinas, SP, Brazil 55Computer Science and Mathematics Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831 (Dated: August 4, 2017) We use 26 million from the (DES) Year 1 shape catalogs over 1321 deg2 of the sky to produce the most significant measurement of cosmic shear in a survey to date. We constrain cosmological parameters in both the flat ⇤CDM and wCDM models, while also varying the neutrino mass density. These results are shown to be robust using two independent shape catalogs, two independent photo- z calibration methods, and two independent analysis pipelines in a blind analysis. We find a 3% fractional 0.5 +0.024 uncertainty on 8(⌦m/0.3) =0.789 0.026 at 68% CL, which is a factor of three improvement over the fractional constraining power of our DES Science Verification results and a factor 1.5 tighter than the previous 0.5 state-of-the-art cosmic shear results. In wCDM, we find a 5% fractional uncertainty on 8(⌦m/0.3) = +0.036 +0.26 0.789 0.038 and a dark energy equation-of-state w = 0.82 0.48. Though we find results that are consistent with previous cosmic shear constraints in 8 – ⌦m, we nevertheless see no evidence for disagreement of our weak lensing data with data from the CMB. Finally, we find no evidence preferring a wCDM model allowing w = 1. We expect further significant improvements with subsequent years of DES data, which will more than triple6 the sky coverage of our shape catalogs and double the effective integrated exposure time per galaxy. Keywords: gravitational lensing: weak; ; dark energy; methods: data analysis; cosmology: observations; cosmo- logical parameters

I. INTRODUCTION cal study, however, there remain several fundamental myster- ies that enter the model as purely phenomenological parame- The study of cosmology over the previous few decades has ters. These include our lack of understanding of the value of been very successful at building a minimal model that is based the cosmological constant or of any motivation for a different on predictions from General Relativity at cosmological scales, driver of cosmic acceleration. Further, although there are can- and validated through a wide range of increasingly sophisti- didates within particle physics models for dark matter, there cated experimental probes. In this model, ⇤CDM, the gravi- has been no detection of such a new particle. As our ability tational dynamics of matter on large scales are dominated by a to constrain cosmological models at lower continues cold dark matter component that only interacts gravitationally to increase, we can also begin to explore subtle discrepancies (CDM) [1], while the accelerated expansion of the Universe between low- and high-redshift observations. This could indi- is driven by a cosmological constant ⇤. These components cate that ⇤CDM, which has explained measurements like the make up about 25% and 70% of the Universe, respectively, power of the cosmic microwave background (CMB) while the remainder is composed of , radiation, and so well, may not be sufficient to connect observations across neutrinos. Despite the overall success of modern cosmologi- cosmic times as the Universe undergoes significant evolution and becomes strongly inhomogeneous. It is with these fundamental mysteries and potential for new physics in mind that several major observing programs ⇤ Corresponding author: [email protected] have been undertaken to measure cosmic shear, a probe that † Einstein Fellow is sensitive to both the expansion of the Universe and the 3 growth and evolution of structure across vast volumes of space duced their first results. More recent observations of cosmic [2–4]. Cosmic shear measures the correlated distortion of shear include analyses of data from the Sloan Digital Sky Sur- the envelopes of light bundles that are emitted from distant vey (SDSS), Canada-France-Hawaii-Telescope Legacy Sur- sources (i.e., galaxies) due to gravitational lensing by large- vey (CFHTLS), the Deep Lens Survey (DLS), the Kilo- scale structure in the Universe. It is sensitive to both the Degree Survey (KiDS)1, and the Dark Energy Survey (DES)2. growth rate and evolution of matter clustering as well as the A portion of the SDSS Stripe 82 (S82) region was analyzed in relative distances between objects, and thus, the expansion [22, 23]. Cosmological constraints from cosmic shear using history of the Universe. With the significant improvements DLS data taken with the Mosaic Imager on the Blanco tele- in constraining power from the work described herein, cosmic scope between 2000 and 2003 were shown in [24, 25]. Sev- shear has now become a leading probe of the nature of dark eral cosmological constraints have resulted from the CFHT energy, dark matter, and astrophysical models of structure for- Lensing Survey (CFHTLenS) [26–30], which was recently re- mation. analyzed in [31]. Cosmic shear from 139 deg2 of the DES Cosmic shear directly measures inhomogeneities along the Science Verification (SV) data was used to place the first con- line-of-sight to an observed galaxy, typically labelled a source straint on cosmology with DES [32] using both real- and (of the light bundle), with a weighted kernel that depends on harmonic-space measurements. Cosmic shear has also been the ratio of distances between the lensing mass, the observed recently measured using KiDS data [33, 34], which was used galaxy, and the observer. The distortion of light bundles due in [34–38] to place tomographic constraints on cosmology. to cosmic shear can be expressed in terms of the convergence, There are currently three ongoing Stage III surveys de- 2 rs signed for measuring cosmic shear: DES using the Blanco  =(3/2)(H0/c) ⌦m 0 dr (r) r(rs r)/(a(r)rs), where r is the comoving distance to an element of lensing mass, and telescope, the Hyper-Suprime Cam (HSC)3 survey using the R this is integrated from the observer to the source at rs; the Subaru telescope, and KiDS using the VLT Survey Tele- matter density fluctuation is , the fractional matter density scope (VST). Preparations are also underway for four Stage parameter ⌦m, a is the scale factor, c is the , and IV weak lensing surveys to operate over the next decade: 4 5 H0 the Hubble parameter. Cosmic shear probes the weak- the Large Synoptic Survey Telescope (LSST) , Euclid , the lensing regime,  1, meaning that distortions are small and Square Kilometer Array (SKA)6, and the Wide linearly related to the⌧ potential. In this regime the convergence Survey Telescope (WFIRST)7. These surveys have signifi- () and shear () are simply related (see e.g., [5]). However, cantly different observing strategies, and in most cases, com- the associated density fluctuations that are probed are not nec- plementary overlap in observing fields to allow for joint cal- essarily in the linear regime. For example, we measure shear ibration and measurements. Two, Euclid and WFIRST, will correlations down to angular separations of about 3 arcmin use space-based telescopes to remove the obstacle of dealing for source bins of mean redshift z =0.38 to 0.96: this corre- with distortions in observed shapes due to the Earth’s atmo- 1 sponds to comoving separations of 0.9 2 h Mpc, on which sphere. scales the has strong nonlinear correc- While cosmic shear can now be said to be one of the most tions. The lensing kernel peaks about halfway to the source powerful probes of cosmology and the nature of dark energy galaxy, so we are mainly sensitive to fluctuations on scales a and dark matter, it is also a challenging measurement to make. further factor of two smaller. The weak distortion in the shapes of objects that we mea- The amplitude of the lensing signal is primarily sensitive sure is of the order of 1%. In the presence of noise and the intrinsic scatter in galaxy shapes of individual objects, grav- to the normalization of the matter fluctuations, 8, and to the itational shear must be statistically measured typically over matter density ⌦m. The combination most tightly constrained ↵ many millions of galaxies, each of which must have a robust is S8 8(⌦m/0.3) [6], where empirically ↵ 0.5 for cosmic⌘ shear. The degeneracy is not exact, so that contours⇡ in shape measurement constrained to high accuracy, to precisely reconstruct the cosmic shear signal. Several technical ad- the 8 ⌦m plane take the shape of a banana, but our most sensitive measurement is perpendicular to the banana and is vances in the robust measurement and interpretation of galaxy shapes are discussed in [36], and we exploit these in the cur- captured by the value of S8. This measurement via cosmic shear of inhomogeneities at z<1 is often compared to infor- rent analysis. In order to interpret the measured cosmic shear mation on the amplitude of primordial fluctuations at recom- signal, one also must have robust estimates of the distribution bination, which is encoded in the temperature fluctuations of of galaxies in redshift, which is a challenging and evolving the CMB. The CMB constraint on fluctuations can be evolved field of study. This is discussed further in [39–42]. Finally, forward in time, assuming a model such as ⇤CDM, to predict one must also be able to interpret the measured signal in the the lensing signal at z<1 and allow a direct comparison with the measurements of structure made using cosmic shear. A closely related and growing field of study is - ing of the CMB (e.g., [7–9]); this probes z 2, and 1 http://kids.strw.leidenuniv.nl  2 http://www.darkenergysurvey.org has demonstrated promising results with constraining power 3 comparable to previous cosmic shear results. http://hsc.mtk.nao.ac.jp/ssp/ 4 http://www.lsst.org The statistical power of cosmic shear has increased rapidly 5 http://sci.esa.int/euclid over the last several years relative to the first detections in the 6 http://skatelescope.org 2000s [10–21] as the current generation of surveys have pro- 7 http://wfirst.gsfc.nasa.gov 4 presence of the correlated intrinsic shapes of galaxies (‘intrin- The paper is organized as follows. We discuss in Sec. II sic alignment’) [43, 44], as well as other astrophysical effects the DES data, and our shape and (photo- that impact the measured cosmic shear signal, such as poorly z) catalogs. Simulations and mock catalogs are discussed in understood baryonic physics (e.g., [45–47]). Sec. III. We present our measurements of cosmic shear in In this work, we present the first cosmological constraints Sec. IV and covariance validation in Sec V. We discuss our from cosmic shear in the main DES wide-field survey, using blinding strategy in Sec. VI, and describe our analysis choices data taken during its first year of observations. Preliminary in Sec. VII. Cosmological parameter constraints are shown in constraints from DES were the result of an analysis of cos- Sec. VIII and further robustness tests in Sec. IX. Finally, we mic shear with data from the DES Science Verification (SV) conclude in Sec. X. observing period in [32, 48]. This was followed by parame- ter constraints from weak lensing peak statistics [49] and the combination of galaxy-galaxy lensing and galaxy clustering II. DARK ENERGY SURVEY YEAR 1 DATA [50]. We combine galaxy-galaxy lensing and - ing with our cosmic shear results in a joint DES Y1 analysis The Dark Energy Survey (DES) is a five year observ- [51], which shares most components of the analysis pipeline ing program using the 570 megapixel DECam [60] on the used in this work. We present further information support- Blanco telescope at the Cerro Tololo Inter-American Obser- ing the DES Y1 cosmological analyses in several concurrent vatory (CTIO). The nominal DES wide-field survey images papers: 5000 square degrees of the southern sky to 24th i-band lim- iting magnitude in the grizY bands spanning 0.40-1.06 µm. The construction and validation of the ‘Gold’ catalog of • The survey tiling strategy ultimately consists of ten overlap- objects in DES Y1 is described in [52]. ping 90 second exposures in each of griz and 45s exposures The DES Y1 REDMAGIC galaxy sample and clustering in Y over the full wide-field area. • systematics that enter our clustering photo-z constraints The DES Year 1 (Y1) shape catalogs used for this analysis are described in [53]. are based on observations taken between Aug. 31 2013 and Shape measurement and calibration techniques and val- Feb. 9 2014 during the first full season of DES operations. • idation of the two shape catalogs, METACALIBRATION DES Y1 wide-field observations were targeted to a large re- and IM3SHAPE, are described in [54]. gion overlapping the (SPT) survey foot- print extending between approximately 60 << 40, Further exploration of the IM3SHAPE image simula- • tions is discussed in [55]. and a much smaller area overlapping the ‘Stripe 82’ region of the (SDSS), which is not included Additional null tests of the reconstructed shear or con- • in this analysis. The observed area was limited in the DES vergence fields are discussed in [56]. Y1 period to reach a sufficient number of overlapping expo- Construction and validation of the redshift distributions sures across the observed footprint. In practice, this resulted • are discussed in [39]. in a total area of about 1514 deg2 with a mean depth of three Constraints on the accuracy of the REDMAGIC red- exposures, after masking potentially bad regions not used for • shifts that enter our clustering photo-z constraints are weak lensing [52]. shown in [40]. The DES Y1 data incorporate a variety of improvements The accuracy of the clustering cross-correlation meth- over the DES Science Verification (SV) data used in prelim- • ods to constrain the source photometric redshifts is de- inary DES weak lensing analyses, including updates to the tailed in [41]. telescope and systems components and to data processing. These are discussed in detail in [52], which describes the pro- The final construction of the clustering cross- • duction and validation of a ‘Gold’ catalog of 137 million ob- correlation constraints on the source photometric jects prior to the ‘bad region’ masking referred to above, and redshifts using the DES Y1 REDMAGIC galaxy sample in [54], where the shape catalog production and validation is is described in [42]. described. Galaxy-galaxy lensing measurements and further shape • catalog tests and validation are shown in [57]. The general methodology, likelihood analysis, and co- A. Shape Catalogs • variance matrix used in the cosmological analyses shown in this work and [51] are described and validated We test the robustness of our results with two indepen- in [58]. dent shape measurement pipelines, METACALIBRATION and Finally, this methodology is independently validated us- IM3SHAPE, which are fully described and characterized in the • ing complex simulations in [59]. accompanying catalog paper [54]. The pipelines use differ- ent subsets of the DES Y1 data, different measurement tech- We encourage the reader to refer to these papers for further ex- niques, and different calibration strategies. Each was devel- tensive information about the DES Y1 data production, test- oped without direct comparison to the other at the two-point ing, and analysis framework that is not repeated in detail in level – blinded measurements of ⇠ were compared only once the current work. the two catalogs were finalized.± Unlike in the DES Science 5

Verification (SV) analysis [61], no effort was made to modify m =0.012 0.013. This error budget is dominated by our es- them to ensure they agreed, beyond applying the same suite timate of the± unaccounted effects of contaminating light from of null tests to both catalogs in [54]. neighboring objects on the shear estimation. The median measured seeing (FWHM of stars selected for With METACALIBRATION, corrections are calculated for point-spread function (PSF) modeling) in the riz bands is 0.96 both the response of the shape estimator to a shear and the arcsec for the DES Y1 shape catalog, which is an improve- response of object selections to a shear. The METACALI- ment over the DES SV seeing. This value is after routine BRATION procedure produces a noisy estimate of the shear nightly rejection of exposures [52] and the blacklisting of a response R for each galaxy, which is then averaged to pro- small number of exposures during the PSF model building duce R . The induced selection bias is calculated only in h i process [61] due to imaging and processing anomalies. At the mean RS . These quantities are in general 2x2 matri- h i the catalog level, objects are removed based on a set of crite- ces of the ellipticity components. The explicit calculation of ria unique to each shape catalog, but which generally satisfy these corrections using the four sheared catalogs is described a lower S/N cut and a rigorous size cut relative to the PSF in Secs. 4.1 & 7.4 of [54]. The application of these correc- size. More details on these selections are described in [54], tions depends on the details of the shear statistic that is being which discusses further the impact of data quality and various calibrated; some examples are derived in [63]. selections on the final catalog number density. In this work we adopt a number of approximations that sim- For both catalogs, we use only objects that pass the default plify this process. First, we assume that the shear response is recommended selection FLAGS SELECT (see selection criteria independent of environment, and thus not dependent on the defined in [54]). We additionally limit objects to have photo-z separation of galaxies. Under this assumption, the correc- point estimates within the redshift range 0.2 – 1.3 (cf. sec- tion to the shear two-point function is simply the square of tion II B) and to fall within the large, contiguous southern por- the mean response (see section 3.2 in [63]). We further make tion of the footprint (cutting dec < 35) that overlaps with the assumption that the correction is independent of the rela- the SPT survey. Finally, we limit our study to objects that tive orientation of galaxies, so that the mean response can be are contained within the REDMAGIC mask described in [53], calculated without the shape rotations that are applied when which additionally removes a few tens of deg2 from the orig- measuring the shear two-point function. We find that the mean inal shape catalog footprint, bringing its final effective area to response matrices are consistent with being diagonal, which 1321 deg2. This final mask, while not strictly necessary for further simplifies the calibration procedure. While these as- cosmic shear, is applied to make this work consistent with the sumptions appear to be valid for the current analysis, fully joint cosmological constraints combining weak lensing and testing the propagation of the full rotated selection response galaxy clustering in [51], where the same footprint is assumed through the shear two-point function is left to a future work. in our covariance calculation. This has the added benefit of There should be no additive correction of the response neces- reducing depth variation across the field, and thus spatial vari- sary for METACALIBRATION, due to the symmetric reconvo- ations in our redshift distribution. lution function used during the metacalibration process [54], though we discuss the impact of residual mean shears in Ap- pendix A. The use of the response corrections is discussed 1. METACALIBRATION further in Sec. IV. The METACALIBRATION catalog yields a total of 35 mil- lion objects, 26 million of which are used in the selection for METACALIBRATION is a method to calibrate a shear statis- the current analysis. The final effective number density of the tic, such as a mean shear estimate or shear two-point func- selection used in this analysis is 5.07 galaxies arcmin 2. The tion, from available imaging data, without requiring signifi- raw mean number density of objects are shown in Fig. 1 as cant prior information about galaxy properties or calibration a function of position on the sky, drawn by SKYMAPPER8 in from image simulations [62, 63]. METACALIBRATION has 9 HEALPIX [64] cells of Nside 1024. Overlaid are the bounds been tested with complex image simulations and shown to be of the nominal five year DES survey footprint. The region accurate at the part per thousand level [63]. The implemen- shown is truncated to magnify the region of interest for this tation used in DES is described in detail in [54], where the analysis, but we refer the reader to [54], where the full shape ellipticity is measured using a single Gaussian model that is catalog footprint is shown, including the ‘Stripe 82’ region. fit to the galaxy image in the riz bands. The galaxy image is then artificially sheared and the ellipticity remeasured to con- struct the shear response matrix via numerical derivatives of 2. IM3SHAPE the ellipticity. We do this by deconvolving the PSF, apply- ing a shear, and then reconvolving by a symmetrized version IM3SHAPE measures galaxy shapes by fitting bulge and of the PSF. This results in one unsheared and four sheared disc models to each object with a Levenberg-Marquardt al- versions of the shape catalog (one for each direction ( ) and gorithm and then taking the best fitting of the two models. component of shear), all of which include flux measurements± for photo-z measurement. Some limitations in the applica- tion of METACALIBRATION to DES Y1 data are discussed in [54], which leads us to assign a non-zero mean for the Gaus- 8 https://github.com/pmelchior/skymapper sian prior assumed in this analysis on the shear calibration of 9 http://healpix.sourceforge.net 6

2 FIG. 1. The footprint of the DES Y1 METACALIBRATION catalog selection used in this work, covering 1321 deg . The joint REDMAGIC mask described in Sec. II A is not included. The raw mean number density of objects drawn by SKYMAPPER in HEALPIX cells of Nside 1024 is shown, which is uncorrected for the coverage fraction at subpixel scales. Overlaid are the bounds of the nominal five year DES survey footprint. The full shape catalog footprint, which includes the ‘Stripe 82’ region, is shown in [54].

This measurement process is unchanged from that used in the 2 DES SV catalog [61]. The code is described in detail in [65]. TABLE I. Effective number density neff (gal arcmin ) and shape Noise, model, and selection biases on the galaxy shapes are noise e (per component) estimates for each tomographic redshift bin of the METACALIBRATION catalog. We reference ne↵ as defined calibrated using a suite of simulations designed to closely re- in both [67] (C13) and [26] (H12). We include H12 for comparison flect real data, which are described in [54, 55]. to previous results, because many have been characterized with this The calibration of IM3SHAPE produces a multiplicative m definition. These quantities are discussed further in [54]. and additive ci bias corrections per object, where i is the el- lipticity component, such that the observed ellipticity (eo) is Bin Extent n e t o t e↵ related to the true ellipticity (e ) as ei =(1+m)ei + ci. The C13 H12 C13 H12 multiplicative bias is assumed to be the same for both shear Full 0.20 – 1.30 5.07 5.40 0.27 0.28 components. Based on the work described in [54], we assign 1 0.20 – 0.43 1.45 1.50 0.26 0.26 a Gaussian prior on the shear calibration of m =0 0.025, 2 0.43 – 0.63 1.43 1.52 0.29 0.30 ± which is wider than that obtained with METACALIBRATION 3 0.63 – 0.90 1.47 1.59 0.26 0.27 due primarily to our estimates of uncertainties related to the 4 0.90 – 1.30 0.70 0.79 0.27 0.28 accuracy of reproducing the real survey in our image simula- tions. The bias corrections m and ci are applied in the same way as previous cosmic shear studies (e.g., [32]), and thus not discussed in detail here. We propagate the impact of the shear for doing so, and for assigning uncertainties to ni(z), are de- calibration m for IM3SHAPE through the two-point estimator scribed fully in [39] and the companion papers [41, 42]. In to produce a two-point correction as a function of scale. This this analysis, galaxies in the shape catalogs are assigned to the is described further in Sec. IV. four redshift bins listed in Table I by the mean of the photo-z IM3SHAPE was applied only to r band images, yielding a posterior p(z) estimated from DES griz flux measurements. smaller catalog of 22 million objects, which is reduced to 18 The redshift distribution of each bin is constructed by stack- million with the selection for the current analysis. The final ing a random sample from the p(z) of each galaxy, weighted effective number density of the selection used in this analysis according to WiSi, which is defined in Sec. IV. The photo-z 2 i is 2.93 galaxies arcmin . posteriors used for bin assignment and n (z) estimation in the fiducial analysis are derived using the Bayesian photometric redshift (BPZ) methodology [66]. Details are given in Sec. 3.1 of [39]. The estimated redshift distributions for META- B. Photometric Redshift Estimates CALIBRATION are shown in Fig. 2. Our adopted model for the redshift distribution assumes A tomographic cosmic shear measurement requires an as- that the true redshift distribution in each bin is related to our signment of each source galaxy to a redshift bin i, and its inter- measured distribution such that: pretation requires an accurate estimation of the redshift distri- bution of galaxies in each redshift bin, ni(z). The procedures ni(z)=ni (z zi), (1) PZ 7

BPZ TABLE II. Summary of cosmological, systematic, and astrophysical parameters used in the fiducial analysis. In the case of flat priors, the 5.0 prior is identical to the listed range. Gaussian priors are indicated by their mean and 1 width listed in the prior column. In the case of w, it is fixed to 1 for ⇤CDM and varies over the range given for n(z) i 2.5 wCDM. For m , the values listed are for METACALIBRATION, which are inflated from the original 1.3% constraint to preserve the overall m uncertainty when combining tomographic pairs in the likelihood analysis. 0.0 Resampled 5.0 COSMOS Parameter Range Prior Cosmological 9 As 10 0.5 . . . 5.0 Flat ⇥

n(z) ⌦m 0.1 . . . 0.9 Flat 2.5 ⌦b 0.03 . . . 0.07 Flat 2 ⌦⌫ h 0.0006 . . . 0.01 Flat 1 1 H0 (km s Mpc ) 55 . . . 90 Flat 0.0 ns 0.87 . . . 1.07 Flat w 2.0 ... 0.333 Fixed/Flat 0.0 0.5 1.0 1.5 2.0 z ⌦k 0.0 Fixed ⌧ 0.08 Fixed Systematic (m1 – m4) 102 10 . . . 10 1.2 2.3 1 2⇥ ± FIG. 2. The measured BPZ and resampled COSMOS redshift dis- z 10 10 . . . 10 0.1 1.6 2 ⇥ 2 ± tributions for the METACALIBRATION shape catalog, binned by the z 10 10 . . . 10 1.9 1.3 3 ⇥ 2 ± means of the photo-z posteriors into the four tomographic ranges in z 10 10 . . . 10 0.9 1.1 4 ⇥ 2 ± Table I and marked by the color shading. The normalization of each z 10 10 . . . 10 1.8 2.2 ⇥ ± bin reflects their relative neff. The BPZ distributions are corrected Astrophysical i by the mean of the redshift bias priors z . The contribution of each A 5.0 . . . 5.0 Flat galaxy is weighted by WiSi, as defined in Sec. IV. The IM3SHAPE ⌘ 5.0 . . . 5.0 Flat redshift distributions are similar to those shown for METACALIBRA- z0 0.62 Fixed TION. The second bin is clearly most different between the resam- pled COSMOS estimate and BPZ – we explore this further in Sec. IX C and show that it does not significantly impact the inferred cos- mological parameters. on the systematic parameters zi at the level of 0.02 [39]. ± For METACALIBRATION the zi are listed in Table II and are where z is the difference in the mean redshift of the true consistent with the original BPZ estimate. The agreement be- and measured n(z). This is a sufficient description of the tween these validation methods provides further justification photo-z uncertainty for the current cosmic shear analysis, as of our reliance on the accuracy of the COSMOS2015 30-band we demonstrate in Sec. IX C. Deviations in the shape of the photo-zs. The priors for alternate combinations of shear and n(z) are subdominant to the impact of the mean z, for reason- photo-z pipelines are given in [39] and require statistically able variance in the shape at the level of precision necessary significant shifts to the redshift distributions in some cases. i for the DES Y1 analysis. We derive constraints on z for One notable complication is the correction of photo-z the estimated redshift distributions by comparison of the mean induced selection biases in METACALIBRATION. For the redshift in each bin to that from two independent methods: METACALIBRATION catalog, six versions of the BPZ cata- log are produced. In order to calculate selection biases on 1. the mean, high-quality photo-z of a sample of galaxies the shear measurement, BPZ redshifts are estimated from from the COSMOS2015 catalog [68], matched to re- the griz photometry produced during shape fitting in the semble the source galaxies in griz flux and pre-seeing four sheared METACALIBRATION catalogs, as well as the un- size [39] sheared catalog—redshifts and bin assignments may change 2. in the lowest three redshift bins, the clustering of source when fluxes are measured from a sheared image. We use galaxies with REDMAGIC galaxies at 0.15 0.9. For the IM3SHAPE catalog, BPZ redshifts estimated from The ADDGALS [69, 70] algorithm is used to add galaxies MOF photometry are used both for binning and for recon- to the simulations by assigning r-band absolute magnitudes structing the redshift distribution, since IM3SHAPE does not to particles in the simulations based on large-scale density. actively calibrate for such biases via the data. It has been The particles to which galaxies are assigned are not necessar- confirmed for IM3SHAPE that there is no apparent residual ily in resolved dark matter halos, but all resolved central dark redshift-dependent bias m in its image simulation, for the matter halos have galaxies assigned to them. Galaxy spectral same redshift bins used in this work and using the associated energy distributions (SEDs) are then assigned to the galaxies COSMOS redshifts of each input object, as discussed in [54]. from a training set derived from SDSS DR7 [73], and DES This test is performed by constructing the calibration from one griz magnitudes are generated by convolving the SEDs with half of the simulation and testing the residual in the other, split the DES pass bands. Galaxy ellipticities and shapes are as- both randomly and by input COSMOS objects. signed by drawing from distributions fit to SuprimeCam i0 - Estimation of the redshift distribution of the lensing source band data degraded to the mean predicted DES PSF. Galaxy galaxies is one of the most difficult components of a broad- magnitudes, ellipticities, and shapes are then lensed using the band cosmic shear survey like DES. Along with the use of multiple-plane ray-tracing algorithm CALCLENS [74]. To two independent methods to constrain zi, we present in mimic DES depth fluctuations, we apply photometric errors Sec. IX C several tests of the robustness of our cosmological using the DES Y1 MAG AUTO depth maps according to an- results to the methods and assumptions of our ni(z) estimates. gular position within the footprint and the true apparent mag- One such test is to replace the shifted BPZ estimator in Eq. (1) nitude of the galaxy. with the resampled COSMOS ni(z) to confirm robustness to Flux and size cuts are applied to the simulated galaxies to the shape of ni(z). approximate the signal-to-noise distribution of a weak lens- ing sample roughly mimicking the selection of the META- CALIBRATION shape catalog. To bring the shape noise in the simulated sample to that of the METACALIBRATION cat- III. SIMULATIONS AND MOCK CATALOGS alog, we then apply an additional redshift dependent flux cut to the mock catalogs. After final cuts and unblinding of the In this analysis, we have employed both a limited number of METACALIBRATION catalog, the effective number density of mock shear catalogs produced from dark-matter-only N-body the mocks is about 7 per cent larger than in the data. simulations, described in Secs. III A & III B, and a large num- ber of lognormal mock shear catalogs, described in Sec. III C. The full N-body mock catalogs have been used to validate B. MICE-GC Mock Catalogs our analysis pipeline [59] and our covariance estimation [58]. There is not a sufficient number of N-body mock catalogs to The MICE Grand Challenge (MICE-GC) simulation is a produce an independent covariance matrix for our data vector large N-body simulation which evolved 40963 particles in to compare to the halo model covariance described in Sec. V, 1 3 a volume of (3072 h Mpc) using the GADGET-2 code so we also employ a large suite of lognormal mock catalogs 10 1 [71]. This results in a particle mass of 2.93 10 h M . to test certain pieces of the full halo model covariance cal- ⇥ The initial conditions were generated at zi = 100 using the culation (also discussed in [58]) and to construct covariance Zel’dovich approximation and a linear power spectrum gen- matrices for various null tests. erated with CAMB10. On-the-fly light-cone outputs of dark- matter particles up to z =1.4 were produced without repe- tition in one octant. A set of 256 all-sky maps with angular A. Buzzard Mock Catalogs HEALPIX resolution Nside = 8192 of the projected mass density field in narrow redshift shells were measured. The The Buzzard simulations, built from dark-matter-only N- process used to compute weak lensing maps from HEALPIX body simulations, are a suite of 18 mock realizations of the mass maps in z-slices was first discussed in [75]. These were DES Y1 survey. The most important aspects of these simula- used to derive the convergence field  in the Born approxima- tions are summarized below, but we refer the reader to more tion by integration along the line-of-sight. The convergence detailed descriptions in [59, 69, 70]. The 18 mock catalogs was transformed to harmonic space, converted to an E-mode are composed of 3 sets of 6 catalogs each, where each set is shear map, and transformed back to angular space to obtain built from a combination of three separate N-body simula- the Stokes (1,2) shear fields, following [76]. In this way 1 3D lensing maps of convergence and shear were produced. tions. These have box lengths of 1.05, 2.6, and 4.0 h Gpc, and 14003, 20483, and 20483 particles, giving mass resolu- 10 11 11 1 tions of 2.7 10 , 1.3 10 , and 4.8 10 h M , re- ⇥ ⇥ ⇥ 10 spectively. The simulations were run using the L-GADGET2 http://camb.info 9

Halos in the light-cone were identified using a Friends-of- effective shape noise of either shape catalog after reweighting Friends algorithm. A combination of Halo Occupation Distri- the objects to match redshift distributions of subsets of the bution (HOD) and SubHalo Abundance Matching (SHAM) catalogs. For further details of how this was implemented, see techniques were then implemented to populate halos with Appendix B. galaxies, assigning positions, velocities, and col- ors to reproduce the function, (g r) color distri- bution, and clustering as a function of color and luminosity in IV. COSMIC SHEAR MEASUREMENT SDSS [77, 78]. Spectral energy distributions (SEDs) are then assigned to the galaxies resampling from the COSMOS cata- We present in this section the measurements of the real- griz log of [79]. Finally, DES magnitudes are generated by space two-point correlation function ⇠ from the METACAL- ± convolving the SEDs with the DES pass bands. The catalogs IBRATION and IM3SHAPE catalogs. These results are derived are available at cosmohub.pic.es and a detailed descrip- from measurements in a contiguous area 1321 deg2 on the sky, tion is given in [76, 80–82]. which has been split into four tomographic bins as described in Sec. II B. These measurements are the highest signal-to- noise measurements of cosmic shear in a galaxy survey to C. Lognormal Mock Catalogs date, with total detection significance (S/N)13 of 26.8 for the fiducial METACALIBRATION measurement using all an- In order to generate large numbers of realizations of mock gular scales and redshift bin pairs. shear data, we take advantage of the fact that a lognormal Cosmic shear is a quantity with two components, based on shear field can be produced quickly and with reasonable lev- two 2nd order angular derivatives of the lensing potential, : els of non-Gaussianity. The potential use of lognormal ran- 1 =( 11 22)/2; 2 = 12 (where the angular deflec- dom fields in cosmological analyses was first outlined in [83], tion is ). For points along the 1 axis, these compo- and the lognormal distribution of shear fields has shown good nents giver a simple definition of the tangential- and cross- agreement with N-body simulations and real data up to non- shear components: t = 1; = 2. There are thus linear scales [84–86]. The production of such mock cata- three 2-point functions to consider, ⇥ but in practice the cross- logs has a significantly smaller computational expense than correlation t vanishes, leaving the two standard quanti- a full N-body simulation and ray-tracing. Thus, lognormal ties that areh the focus⇥i of most weak-lensing studies [88]: mock simulations provide a compromise between accuracy and computational cost that allows us to quantify how the ⇠ = tt . (2) non-Gaussianity of cosmic fields and incomplete sky cover- ± h i±h ⇥ ⇥i age propagate into the covariance of cosmic shear. We estimate ⇠ for redshift bin pair i, j as We use the publicly available code FLASK11 [87], which ± generates consistent density and convergence fields, to pro- i ~ j ~ i ~ j ~ ab WaWb eˆa,t(✓)ˆeb,t(✓) eˆa, (✓)ˆeb, (✓) duce 150 mock full-sky shear maps that reproduce a set of in- ⇠ˆij(✓)= ± ⇥ ⇥ , 12 h i put power spectra that fit a fiducial cosmology and the actual ± P ab WaWbSaSb redshift distribution of sources in our data. These maps are (3) P produced on a HEALPIX grid with resolution set by an Nside where eˆa,t is the tangential component of the corrected ellip- parameter of 4096. In this resolution, the typical pixel area is ticity of galaxies a along the direction towards galaxy b and around 0.73 arcmin2. The full sky mocks are then divided into eˆa, is the cross component, W is a per-object weight, and ⇥ eight non-overlapping DES Y1 footprints per full-sky simula- S is either a multiplicative bias correction (IM3SHAPE) or a tion. To a good approximation, the footprints belonging to shear response correction (METACALIBRATION). The sums the same full-sky are uncorrelated for sufficiently high multi- are each computed for a subset of galaxy pairs a, b within each poles. This produces a total of 1,200 mock shear maps. For angular separation ✓ for each ✓ = ✓~b ✓~a . These angu- each mock realization, we simulate four shear fields corre- lar bins are chosen to be logarithmic| with a total| of 20 bins sponding to the redshift distributions of the four redshift bins between 2.5 and 250 arcmin, though only a subset of these shown in Fig. 2. angular bins are used in parameter estimation, as discussed in To capture the expected noise properties of the shear fields, Sec. VII A. All two-point calculations are done with the pub- we then add appropriate shape-noise by sampling each pixel lic code TREECORR14 [89]. The estimator for ⇠ is in prac- ± of the map to match the measured ne↵ and e shown in Table tice calculated quite differently for the two shape catalogs, be- I for each tomographic shear bin. Covariance validation was cause they each estimate the ellipticity of an object and any done using METACALIBRATION parameters, while the mock shear calibrations via fundamentally different processes. catalogs were remade for each null test in Sec. B to match the

1 13 S/N = ⇠dataC ⇠model 11 Signal-to-noise is derived here as in [48] as 1 , http://www.astro.iag.usp.br/ flask/ p⇠modelC ⇠model 12 ⇠ The cosmology used for the FLASK simulations is described by a flat where C is the covariance described in Sec. V and ⇠model is the best-fit ⇤CDM model with: ⌦m =0.295, ⌦b =0.0468, As =2.260574 model obtained from the analysis in Sec. VIII. 9 ⇥ 14 10 , h =0.6881, ns =0.9676. https://github.com/rmjarvis/TreeCorr 10

For METACALIBRATION, the kth component of the unro- 4 metacalibration im3shape 10 tated ellipticity is given by eˆk = ek ek , where e is the residual mean shear in a given tomographich i bin.h Thei

METACALIBRATION catalog does not use a galaxy weight + 6 10 (W =1), and the shear response correction (S) is given by S R = R + RS. In general R is a 2x2 matrix, where ⌘ 10 8 Rii = R,ii + RS,ii is the sum of the iith element of the measured shear response and shear selection bias correction 5 10

matrix for METACALIBRATION. We simply use the average of the components of R, where R =(R11 + R22)/2. For IM3SHAPE, eˆk = ek ck ek ck , where c is the addi- h i 7 tive shear correction and e c is the residual mean shear 10 h i 101 102 101 102 for a tomographic bin. The IM3SHAPE catalog uses an em- (arcmin) (arcmin) pirically derived weight (W = w), and a multiplicative shear correction S =1+m, where m is defined irrespective of the ellipticity component. For more details about the calculation of c, w, m, and R, see [54]. FIG. 3. The measured non-tomographic shear correlation function ⇠ for the DES Y1 shape catalogs. ± The redshift distribution of each tomographic bin for the METACALIBRATION measurements is shown in Fig. 2. The redshift boundaries, effective number density, and per compo- V. COVARIANCE MATRIX nent e of each tomographic bin for METACALIBRATION are given in Table I. Due to the inherent weighting of each object The calculation of the covariance matrix of ⇠ and tests to in the estimator in Eq. (3), the objects contributing to the n(z) ± for a tomographic bin have been weighted by the factor W S . validate its quality can be found in [58]. A large part of our i i covariance is caused by the shape-noise and Gaussian compo- We show the measured two-point correlation function ⇠ nents of the covariance, i.e., covariance terms that involve at for each shape catalog in Figs. 3 – 5. Scales not used to± most two-point statistics of the cosmic shear fields. To guar- constrain cosmological parameters are shaded in Fig. 4 & 5. antee that our covariance model captures these error contribu- This is the first measurement to correct, through the metacal- tions correctly, the Gaussian parts of the model are compared ibration process, the shear selection effects RS, e.g., due to to a sample covariance from 1200 Gaussian random realiza- photo-z binning in the data. This effect can be only roughly tions of the shear fields in our tomographic bins. The un- approximated in traditional image simulation calibrations by certainties on cosmological parameters projected from each assigning redshifts based on the original redshift measurement of these covariances agree very well [58]. The non-Gaussian of the input objects, which is not the same as the redshift mea- parts of our covariance, i.e., the parts involving higher order surement used in the data and not even necessarily correlated correlations of the shear field, are modeled in a halo model with magnitude or color in a natural way in the simulation15. framework [90]. To measure the influence of realistic survey The measured selection effects R vary from 0.6% to 2.5% geometry on the covariance matrices, covariance matrices de- S of the shear response correction Rg in the four tomographic termined in three different ways are compared: 1) the full halo bins, with an of up to 0.014, which is a model covariance, 2) a sample covariance from 1200 lognor- significant level of bias if it were uncorrected – larger than the mal realizations (see Sec. III C) of the convergence field in Gaussian prior width on the multiplicative bias of 0.013 for our tomographic bins that assumes a circular survey footprint, the METACALIBRATION catalog. We can estimate its effect and 3) a sample covariance from 1200 lognormal realizations by comparing to the selection bias correction with no tomo- using our actual DES Y1 footprint. The parameter uncertain- graphic binning, which is 0.011. Thus, the inclusion of the ties derived from each of these matrices agree very well [58], selection bias correction calculated from the four runs of BPZ indicating that masking and footprint shape have a negligible on the sheared photometry from METACALIBRATION is likely impact in the DES Y1 analysis. a significant contribution to the corrected selection bias, and We show the full halo model correlation matrix for ⇠ as the the additional computational resources and complexity intro- lower triangle in Fig. 6. The upper triangle is the difference± duced are warranted. of the full halo model correlation matrix and the correlation matrix resulting from the 1200 lognormal realizations masked by the DES Y1 footprint. Following the suggestion of an iter- ative approach to dealing with the cosmological dependence of covariance matrices proposed by [91], an initial covariance matrix was calculated using an arbitrary cosmology, but the fi- nal covariance matrix used in this work was recalculated with 15 We preserve the original COSMOS magnitudes of objects in the simula- the best-fit cosmology of the initial fiducial result from [51]. tions used to calibrate IM3SHAPE, so the assigned redshifts do correspond We found no significant change in our inferred cosmology due to the flux and morphology of the simulated image. to this covariance change. 11

2 1,4 1,3 1,2 1,1 4 10 / + 0

2,4 2,3 2,2 4

2.5 10 / +

0.0

3,4 3,3 4,4 5 4 4

10 2.5 10 / / +

0.0 0 5 4,4 3,3 3,4 4 4

10 2.5 10 / / +

0 0.0 2,2 2,3 2,4 4

2.5 10 /

0.0 1,1 1,2 1,3 1,4 2 4 10 / 0

101 102 101 102 101 102 101 102 (arcmin) (arcmin) (arcmin) (arcmin)

FIG. 4. The measured shear correlation function ⇠+ (top triangle) and ⇠ (bottom triangle) for the DES Y1 METACALIBRATION catalog. Results are scaled by the angular separation (✓) to emphasize features and differences relative to the best-fit model. The correlation functions are measured in four tomographic bins spanning the redshift ranges listed in Table I, with labels for each bin combination in the upper left corner of each panel. The assignment of galaxies to tomographic bins is discussed in Sec. II B. Scales which are not used in the fiducial analysis are shaded (see Sec. VII A). The best-fit ⇤CDM theory line from the full tomographic analysis is also plotted as the solid line. We find a 2 of 268 for 211 degrees of freedom in the non-shaded regions, which is discussed in Sec. VIII A.

We also test the level of shape noise in the covariance ma- blinding 16 was preserved until the catalogs and primary DES trix by comparing halo model covariance predictions for ⇠ Y1 cosmological analyses and papers (this work and [51]) on small scales (2.5 <✓<10 arcmin), where shape noise completed a first round of the DES internal review process. dominates, to jackknife estimates for both shape catalogs from All calculations were then repeated with the unblinded cata- the data. We find very good agreement for METACALIBRA- logs for the final version of this paper. TION, but there is an indication that in two tomographic bins, In addition to this catalog-level blinding, no comparison the shape noise of IM3SHAPE may be underestimated by up to theory at the two-point level (⇠ ) or of cosmological con- to 20%. We believe this is due to an unresolved issue with the tours was made, nor were central± values of any cosmological empirically derived weights as a function of redshift. Since we use IM3SHAPE only to validate that our shape measurement and calibration is robust, this would only result in a slight in- 16 During the internal review process for [54], it was discovered that separate, flation of the significance of this test in Sec. IX B. but equivalent, oversights in the shear calibration of the two catalogs led to a substantial fraction (e.g., the linear part in e) of the blinding factor being calibrated. This was undiscovered until the catalogs were finalized, and thus had no impact on catalog-level choices. It is valid to question VI. BLINDING whether this invalidated our blinding strategy at the parameter estimation level. It did not, for two reasons: 1) only a few people in the collaboration were aware of the potential issue until after we unblinded the cosmolog- For the DES Y1 analysis, we have maintained a catalog- ical parameters, minimizing any impact, and 2) The secondary blinding level blinding scheme similar to the DES SV analyses, but enforced at the two-point and parameter level ensured that even had we rescaling ⌘ = 2 arctanh e by a factor between 0.9 and 1.1 become aware of this oversight much sooner, it could not have led to ex- (see [92]| for| a review of blinding| | in general). This catalog perimenter bias in our analyses. 12

2 1,4 1,3 1,2 1,1 4 10 / + 0

2,4 2,3 2,2 4

2.5 10 / +

0.0

3,4 3,3 4,4 5 4 4

10 2.5 10 / / +

0.0 0 5 4,4 3,3 3,4 4 4

10 2.5 10 / / +

0 0.0 2,2 2,3 2,4 4

2.5 10 /

0.0 1,1 1,2 1,3 1,4 2 4 10 / 0

101 102 101 102 101 102 101 102 (arcmin) (arcmin) (arcmin) (arcmin)

FIG. 5. The measured shear correlation function ⇠+ (top triangle) and ⇠ (bottom triangle) for the DES Y1 IM3SHAPE catalog (see caption of Fig. 4). The uncertainty on ⇠ is clearly larger for IM3SHAPE compared to METACALIBRATION in Fig. 4 due to the lower number density of ± objects. We find a 2 of 290 for 211 degrees of freedom in the non-shaded regions, which is discussed in Sec. VIII A. inferences revealed, until after the shape catalogs and priors VII. MODELING CHOICES were finalized. A qualitative comparison of results from the two shape catalogs with axes and values suppressed was per- The measured ⇠ for tomographic bins i and j can be re- formed to confirm that they produced consistent results af- lated to the angular± convergence power spectrum in a flat uni- ter their development was complete and before finalizing the verse through the integral shear priors. The results of this test were acceptable, and no modification to the shape catalogs or priors was necessary. All ˆij 1 ij ⇠ (✓)= d``J0/4(✓`)P (`), (4) measurement, processing, and plotting routines were tested ± 2⇡ either on measurements of the mock catalogs or on synthetic Z data vectors before use on the DES data. Only after analysis where Jn is the nth order Bessel function of the first kind. plans were finalized was any comparison to theory allowed P is then related to the matter power spectrum P with the or inferred cosmological parameter values revealed. Several harmonic-space version [93, 94] of the Limber approximation negligible updates to the precise values of the shear and photo- [95] z priors, and some bugs in the IM3SHAPE catalog selection related to blacklisted images, were approved after unblind- H qi()qj() ` +1/2 P ij(`)= d P , , (5) ing the catalogs but before unblinding the parameter values.  2 NL 0 ✓ ◆ These changes did not have an impact on the final results – Z changes due to updates to the IM3SHAPE selection, for exam- where is radial comoving distance, H is the distance to the ple, were entirely negligible at the two-point function level. horizon, and q() is the lensing efficiency function

2 H i 3 H0 i dz 0 q ()= ⌦ d0n (0) , 2 m c a() d ✓ ◆ Z 0 0 (6) 13

1.00 LIKE [96] and COSMOSIS17 [97]. These pipelines were val- idated against each other in [58] and through an analysis by

0.75

j [59] on simulations. To calculate the matter power spec- s ,z

i trum PNL(k, z),COSMOLIKE uses CLASS [98], while COS- s 0.50

, z MOSIS uses CAMB [99, 100]. We sample the parameter

ULTI EST EMCEE

space using both M N [101–103] and [104] – 0.25 MULTINEST in particular produces estimates for the Bayesian evidence that we use for model and data comparison. All 0.00 constraints shown are produced with MULTINEST results, pri- marily due to the speed of convergence and availability of the 18

0.25

j Bayesian evidence estimate.

s ,z

i We perform likelihood evaluation using the cosmic shear s 0.50 , z measurements described in Sec. IV, the redshift distributions +

described in Sec. II B, and the covariance matrix described in 0.75 Sec. V. Cosmological, astrophysical, and systematic parame- ters are constrained for both the ⇤CDM model and the wCDM 1.00 model, where the equation of state of dark energy is described + , zs ,zs , zs ,zs i j i j by a single parameter w. We leave exploration of models with non-zero curvature to future work. A varying neutrino mass density is included in both models, which we believe is FIG. 6. The cosmic shear correlation matrix. The fiducial halo model strongly motivated, but is one reason we must recompute the correlation matrix is shown in the lower triangle, while the difference likelihood of some external data (see Sec. VIII C) to compare of this with the correlation matrix derived from 1200 FLASK log- directly to our own. The parameters varied in the fiducial anal- normal simulations with the DES Y1 mask applied is shown in the ysis are listed in Table II, along with their range and any pri- upper triangle. This shows primarily the noise in the FLASK co- ors applied. For the ⇤CDM model, w is fixed to 1, while in variance, and any differences between the two derived covariance wCDM it is allowed to vary in the given range. For those cos- matrices were shown to be entirely negligible in [58]. Elements are ordered to the right (upward) by increasing redshift bin pair index ij mological parameters we expect to constrain well with cosmic with i 1/3 that do not produce acceler- ation and impose a limit ofw> 2, which is a broad enough range to allow our 1 contour to close assuming we had found where ⌦ is the matter density parameter, H is the Hubble m 0 a posterior centered at w = 1. Those parameters that are not constant, c is the speed of light, a is the scale factor, and ni() constrained well by cosmic shear have informative priors that is the effective number density of galaxies as a function of co- widely bracket allowed values from external experiments. In moving radial distance normalized such that dni()=1. particular, for ⌦ h2 we take a lower limit obtained from os- The appropriateness of the Limber and flat-sky approxima- ⌫ cillation experiments [105], and an upper limit of 0.01 that is tions in these relationships is tested in [58] forR the DES Y1 roughly five times the 95% confidence limit (CL) of the typi- statistical precision. 2 cal limiting value found by external data, ⌦⌫ h 0.002 (see Our data vector D (i.e., Figs. 4 & 5) contains 227 data e.g. Planck Collaboration et al. [106])19. ⇡ points after the cuts described in Sec. VII A. We sample the Though we sample over the normalization of the matter likelihood, which is assumed to be Gaussian in the multi- power spectrum As, we present results in terms of the com- dimensional parameter space: monly used parameter S (⌦ /0.3)↵. Choosing ↵ = 8 ⌘ 8 m 0.5 largely decorrelates S8 and ⌦m in constraints from cos- 1 1 ln (p)= (Di Ti(p)) C ij (Dj Tj(p)) (7) mic shear experiments. The amplitude of the shear correla- L 2 2 ij tion function is roughly S8 . We will refer to both the 68% X confidence limit, which is/ the area around the peak of the pos- where p is the full set of parameters and Ti(p) are the theo- retical predictions for ⇠ as given above. The likelihood also depends upon the covariance± matrix C from Sec. V, which 17 https://bitbucket.org/joezuntz/cosmosis/wiki/Home describes how the measurement in each angular and redshift 18 We find a minor difference in results in poorly constrained directions of bin is correlated with every other measurement. The covari- parameter space, however, with EMCEE giving slightly broader results for ance matrix should also depend on the model parameters p, the longest chain we compared. In general, these do not matter for the but we assume a fiducial set of parameters and use a fixed interpretation of our results (e.g., the edges of the ⌦b prior range), and in covariance. This has been shown to not impact the inferred the primary constraint direction S8, this amounts to a change of 0.5% or less in the fractional constraint. cosmology (see Sec. V). The posterior is then the product of 19 The Planck limit is derived assuming the validity of ⇤CDM. The conserva- the likelihood with the priors, (p), as given in Table II. tive inflation of this limit in our prior leaves us confident in using the prior P Results are derived via two analysis pipelines: COSMO- for tests of ⇤CDM. 14 terior within which 68% of the probability lies, as well as the to COSMIC EMU [115, 116], a power spectrum emulator, figure of merit (FoM), which is useful for comparing the rel- with the impact of baryonic effects from OWLS AGN, which ative constraining power of results in 2D. In two parameter is comparable or larger at all k. dimensions, the FoM is defined for parameters p1 and p2 as [107, 108]:

1 B. Intrinsic Alignment Modeling FoMp1 p2 = , (8) det Cov(p1,p2) In addition to coherent shape distortions induced by lens- p which is a generalization of the Dark Energy Task Force ing, galaxies can exhibit physical shape correlations due to (DETF) recommendation for the dark energy FoM [2]. This their formation and evolution in the same large-scale gravi- kind of statistic is naturally motivated by the form taken by a tational environment. Along with baryonic effects, ‘intrinsic change in relative entropy driven by a gain in information. alignment’ (IA) constitutes the most significant astrophysical systematic to cosmic shear. IA includes both an ‘Intrinsic- Intrinsic’ (II) term due to physically nearby galaxy pairs [117– A. Matter Power Spectrum Modeling and Baryonic Effects 120] and a ‘Gravitational-Intrinsic’ (GI) term from the corre- lation of galaxies that are aligned with those that are lensed Approximations in the nonlinear clustering of matter on by the same structure [121]. The total measured cosmic shear small scales, including the impact of baryonic effects, is a key signal is the sum of the pure lensing contribution and the two modeling choice for the cosmic shear signal. The discussion IA terms: in [32], to which we refer the reader, remains applicable to the current analysis, though some updates to scale selection ij ij ij ij ij Pobs(`)=PGG(`)+PGI(`)+PIG(`)+PII (`), (10) are necessary and are discussed further in Sec. IX D. We also explore the impact of these modeling choices in [58]. where ‘GG’ refers to the cosmic shear signal. Neglecting IA To model the nonlinear matter power spectrum, we use can lead to significantly biased cosmological constraints [117, HALOFIT [109] with updates from [110]. The impact of 122–124]. See the reviews [43, 44] for further information on neutrino mass on the matter power spectrum is implemented these effects. in HALOFIT from [111], which introduces some additional For our fiducial analysis, we treat IA in the ‘tidal alignment’ uncertainty. The fiducial analysis removes scales that could paradigm, which assumes that intrinsic galaxy shapes are lin- be significantly biased by baryonic effects. For scale selec- early related to the tidal field [119]. While a complete under- tion, these effects are modeled as a rescaling of the nonlinear standing of alignment mechanism(s) remains a topic of active matter power spectrum study, tidal alignment has been shown to accurately describe red/elliptical galaxy alignment and is expected to dominate PDM+ PNL(k, z) PNL(k, z), (9) the IA signal on linear scales [122, 125, 126]. We also perform ! P DM an analysis that includes the potential impact of alignments where ‘DM’ refers to the power spectrum from the OWLS from angular momentum correlations (e.g. tidal torque the- (OverWhelmingly Large Simulations project) dark-matter- ory [127]) which are expected to contribute for spiral galax- only simulation, while ‘DM+Baryon’ refers to the power ies, although the amplitude is likely smaller than tidal align- spectrum from the OWLS AGN simulation [45, 112]. OWLS ment of ellipticals [128]. This work is thus the first to include is a suite of hydrodynamic simulations with different sub- both tidal alignment and tidal torque-type alignments, as well grid prescriptions for baryonic effects. We use this partic- as their cross-correlation [129]. This more general, ‘mixed’ ular OWLS simulation for two reasons. First, it is the one model is completely analogous to a perturbative expansion of which deviates most from the dark matter-only case in the galaxy bias beyond linear order and is thus expected to capture relevant scales of the matter power spectrum; given we are the relevant alignment effects at next-to-leading order, even if cutting scales based on the size of this deviation, this is a con- they are not due to classical ‘tidal torquing.’ servative choice. Secondly, McCarthy et al. [113] find that of The amplitude (A) of the ‘nonlinear alignment’ (NLA) the OWLS, the AGN simulation best matches observations of model [130] and its redshift evolution (⌘) are allowed to vary galaxy groups in the X-ray and optical, so arguably it is the in our fiducial analysis, such that the amplitude is described by most realistic. A A[(1 + z)/(1 + z )]⌘, where the pivot redshift is chosen ⌘ 0 We remove any scales from the ⇠ data vector that would to be approximately the mean redshift of sources z0 =0.62. have a fractional contribution from baryonic± effects exceeding This is an improvement over the fiducial analyses of previous 2% at any physical scale. This removes a significant number cosmic shear studies, which fixed this power-law dependence of data points, particularly in ⇠ , on small scales. In general (or neglected it entirely), and this fiducial model is the one we find that our cuts in scale to remove parts of the cosmic employed in our combined probes analysis [51]. The ampli- shear data vector contaminated by potential baryonic effects tude of the terms are then scaled as PGI(`) A (GI IG) 2 / $ are sufficient to alleviate any potential bias due to uncertain- and PII(`) A , following the standard tidal alignment sign ties in modeling nonlinear matter clustering. This can be seen convention./ In Sec. IX E we vary the IA model to demonstrate in [114], which compares inaccuracies in HALOFIT relative robustness to the specific modeling choice. 15

C. Modeling Shear Systematics 1.00 DES Y1 The shear multiplicative bias is modeled as [131, 132] 0.95 KiDS-450 Planck ij i j ij ⇠ =(1+m )(1 + m )⇠true, (11) 0.90 where mi are free to independently vary in each tomographic bin. We do not explicitly marginalize over the potential im- 0.85 pact of additive systematics. We use a Gaussian prior on mi of 0.012 0.023 for METACALIBRATION, given in Ta- 0.80

± 8 ble II, which is rescaled from the non-tomographic prior m =0.012 0.013 due to potential correlations between 0.75 tomographic± bins as discussed in Appendix D of [54]. The equivalent IM3SHAPE prior on mi is 0.0 0.035. Both are 0.70 allowed to vary independently in each tomographic± bin. The only potential source of additive systematics we have 0.65 identified in [54] is related to incorrect modeling of the PSF. We can model the impact of the PSF model errors in cosmic 0.60 shear and this is described in detail in Appendix A along with a discussion of the residual mean shear in each tomographic 0.55 bin, which is not fully described by PSF model errors. We find 0.2 0.3 0.4 0.5 that after correcting the signal for the mean shear, the effect CDM 0.90 m of PSF modeling errors is negligible.

D. Modeling Photo-z Systematics 0.85 5 . 0 3)

The photo-z bias is modeled as an additive shift of the n(z) . 0 / 0.80 i i i m n = nPZ(z z ), (12) ( 8 where zi are free to independently vary in each tomographic ⌘ bin. As discussed in Sec. II B, this is a sufficient approx- 8 0.75 imation for the DES Y1 cosmic shear analysis, and this is S further validated in Sec. IX C. The Gaussian priors on zi for the METACALIBRATION measurements are listed in Ta- 0.70 ble II. We separately calibrate priors for the IM3SHAPE mea- surements, which have Gaussian priors of zi =(0.004 0.015; 0.024 0.013; 0.003 0.011; 0.057 0.022)± [39, 42]. When± using the resampled± COSMOS n±i(z), the 0.65 same width for the prior on zi is used, but it is centered 0.2 0.3 0.4 0.5 at zero. All zi are allowed to vary independently in each to- m mographic bin. As in the case of shear calibration, the width of these priors accounts for correlations between tomographic bins as described in Appendix A of [39]. FIG. 7. Fiducial constraints on the clustering amplitude 8 and S8 with the matter density ⌦m in ⇤CDM. The fiducial DES Y1 cosmic shear constraints are shown by the gray filled contours, with Planck VIII. COSMOLOGICAL PARAMETER CONSTRAINTS CMB constraints given by the filled green contours, and cosmic shear constraints from KiDS-450 by unfilled blue contours. In all cases, Given the size and quality of the DES Y1 shape catalogs, 68% and 95% confidence levels are shown. we are able to make a highly significant statement about the robustness of the standard ⇤CDM cosmological model. Our measurements of cosmic shear probe the evolution of nonlin- before light left the galaxies we now observe in DES. Com- ear fluctuations in the underlying matter field and expansion paring the prediction of these very different probes at the same of space across a very large volume around z 0.6. By com- redshift via the parameter S8 allows us to test whether these parison, equally constraining measurements⇡ of the CMB at results are consistent within the ⇤CDM model to high preci- z = 1100 use information from linear perturbations in the ra- sion. diation field to constrain the same model eight billion years Using the fiducial modeling choices described in the pre- 16

2 space used in this work (see Table II), including varying ⌦⌫ h , DES Y1 which is discussed further in Sec. VIII C. We show the impact 0.90 Fixed h2 2 of fixing ⌦⌫ h in our fiducial ⇤CDM analysis in Fig. 8. i i Fixed m , z The 1D peak value of S8 and ⌦m are listed in Table III Fixed mi, zi, A, along with 68% CL about the peak for both our fiducial 0.85 ⇤CDM results and a large variety of consistency checks and constraints from external data. The FoM in Eq. (8) for the S8 – ⌦m plane is also listed. These constraints on S8 are visu- 0.80 ally summarized in Fig. 9. In both, we distinguish variations 8

S on the fiducial setup that are not necessarily expected to give consistent results (e.g., by neglecting astrophysical systemat- ics) by an asterisk. We find a 3% fractional uncertainty on 0.75 +0.024 S8 =0.789 0.026 at 68% CL, which is a significant improve- ment over the previous best constraint from cosmic shear from KiDS-450 [34] of about 5% (see Table III), and a factor of 0.70 three improvement over the constraining power of our Sci- ence Verification results. We see similar improvements in the constraint on ⌦m, which is more representative of the gain 0.65 in the direction of degeneracy. We expect further significant 0.2 0.3 0.4 0.5 improvements with subsequent years of DES data, which will m more than triple the sky coverage of our shape catalogs and double the effective integrated exposure time per galaxy. We have plotted the best-fit ⇤CDM prediction in Figs. 4 & 2 FIG. 8. A comparison of the fiducial constraints in ⇤CDM (filled 5 for both shape catalogs. We find a total for the fiducial 2 gray contours) to constraints where we: 1) fix ⌦⌫ h (orange con- METACALIBRATION measurement of 268 with 211 degrees tours), 2) fix all photo-z and shear systematic parameters (green), of freedom20 (227 data points and 16 free parameters) for the and 3) fix all systematic parameters and intrinsic alignment (IA) pa- ⇤CDM best-fit model.21 The probability p of getting a higher rameters (green). We find no visually significant bias correction or 2 value can be derived assuming our data vector is drawn decrease in constraining power including systematics parameters, but 2 from a multi-variate Gaussian likelihood around the best-fit varying ⌦⌫ h and IA parameters both shift and enlarge the resulting theory vector and that our covariance matrix is precisely and contours. Both 68% and 95% confidence levels are shown. 2 fully characterized. This leads to p =0.5 10 , which is formally unlikely. We have verified that no⇥ single tomo- vious section, we use cosmic shear from the first year of the graphic measurement (out of 20) can be identified as a signif- Dark Energy Survey to constrain both the ⇤CDM and wCDM icant outlier in this sense. However, we have found that some models with varying neutrino mass to produce the tightest cos- of our robustness checks, (i.e., removing the highest redshift bin from the analysis or removing the largest scales) do sig- mological constraints from cosmic shear to date. In [51], our 2 nificantly improve this statistic, increasing p to 5.0 10 cosmic shear results are further combined with galaxy-galaxy 2 ⇥ and 5.7 10 , respectively, while not significantly altering lensing and galaxy clustering to significantly improve these ⇥ constraints. When comparing with external data, it is impor- the inferred cosmology. However, the formal p value assumes 2 the covariance matrix to be known exactly, and this cannot be tant to note that we vary ⌦⌫ h in our fiducial analysis, and thus all results we compare to, and so the central values and completely realistic. If the true covariance values were sys- tematically 10% larger than our estimates, then the p value for uncertainties of parameters may differ from those previously 2 published for these data. the full would rise to an acceptable 0.075. Although the covariance matrix has been validated in several ways in [58] and we have confirmed the level of shape noise in Sec. V, it is A. Fiducial ⇤CDM Results difficult to rule out shifts of this order – even the uncertainty in the correct true cosmology can have an effect of the right mag- nitude on the covariance matrix. Alternatively, it may be that We marginalize over a total of 6 cosmological parameters we are witnessing small failures of the Gaussian likelihood as- ⇤ in the fiducial CDM model, including a free neutrino mass sumption, as recently proposed by [133] relating to large-scale density, and 10 systematic or astrophysical parameters. These are listed in Table II. Our fiducial ⇤CDM constraints in the 8 – ⌦m and S8 – ⌦m planes are shown in Fig. 7. The DES Y1 cosmic shear constraints are shown by the gray filled contours, 20 It is worth noting that half of these parameters are tightly constrained by while the previous best real-space cosmic shear constraints priors, such that fixing them does not significantly alter the final constraint, and thus the number of dof may be overestimated. from the KiDS survey are shown in blue, and Planck con- 21 The 2 for IM3SHAPE is slightly larger – 290 for the same dof. Given straints from the CMB in filled green, for comparison. Both the likely underestimation of shape noise in the IM3SHAPE covariance (see 68% and 95% confidence levels are shown. For consistency, Sec. V), this is an equivalently good fit to that found for METACALIBRA- previous constraints have been reanalyzed in the parameter TION. 17

TABLE III. Summary of constraints on the 1D peak value of S8 and ⌦m in ⇤CDM. The FoM in Eq. (8) for the S8 ⌦m plane is also shown. In the wCDM model, we show the 1D peak value of S8, ⌦m, and w, along with the FoM for the S8 w plane. 68% CL are included for each parameter, which are not symmetric about the peak in most cases. We distinguish variations on the fiducial model that are not required to give consistent results (e.g., by neglecting astrophysical systematics) by an asterisk. The constraints on S8 are also visually summarized in Fig. 9. ⇤CDM wCDM Model S8 ⌦m FoMS ⌦ S8 ⌦m w0 FoMS w 8 m 8 0 +0.024 +0.068 +0.036 +0.074 +0.26 Fiducial 0.789 0.026 0.311 0.061 510 0.789 0.038 0.317 0.054 0.82 0.48 100 +0.032 +0.038 +0.037 +0.079 +0.35 *Fixed neutrino mass density 0.792 0.021 0.304 0.062 750 0.797 0.037 0.290 0.051 0.92 0.40 96 +0 .020 +0 .060 +0 .028 +0 .078 +0 .33 *No photo-z or shear systematics 0.789 0.027 0.320 0.068 664 0.801 0.050 0.312 0.055 0.79 0.48 111 +0 .017 +0 .060 +0 .042 +0 .067 +0.44 *Only cosmological parameters 0.766 0.026 0.250 0.040 896 0.737 0.027 0.241 0.039 1.28 0.35 138 Shape Measurement

+0.029 +0.066 +0.034 +0.066 +0.30 *No shear systematics 0.787 0.017 0.305 0.060 598 0.789 0.039 0.330 0.069 0.89 0.49 97 +0 .039 +0 .117 +0 .040 +0 .074 +0 .44 IM3SHAPE 0.812 0.048 0.293 0.052 250 0.794 0.060 0.307 0.058 1.37 0.38 53 Photometric Redshifts

+0.021 +0.060 +0.030 +0.068 +0.28 *No photo-z systematics 0.785 0.026 0.302 0.064 639 0.789 0.044 0.321 0.063 0.79 0.50 104 +0 .025 +0 .064 +0 .035 +0 .079 +0 .37 Resampled COSMOS photo-zs 0.778 0.028 0.286 0.053 604 0.786 0.043 0.299 0.053 0.88 0.34 96 +0 .033 +0 .063 +0 .043 +0 .082 +0 .16 Removing highest z-bin 0.780 0.041 0.286 0.047 354 0.789 0.048 0.298 0.045 0.59 0.56 63 Data Vector Choices

+0.015 +0.039 +0.027 +0.073 +0.28 *Extended angular scales 0.770 0.023 0.320 0.069 887 0.772 0.038 0.315 0.052 0.83 0.47 124 +0 .040 +0 .069 +0 .061 +0 .080 +0 .52 Large angular scales 0.813 0.053 0.374 0.091 260 0.785 0.058 0.349 0.068 1.29 0.40 46 +0 .027 +0 .072 +0 .044 +0 .116 +0 .33 Small angular scales 0.778 0.035 0.325 0.081 349 0.788 0.044 0.310 0.054 0.75 0.32 74 Intrinsic Alignment Modeling

+0.025 +0.067 +0.029 +0.039 +0.31 *No IA modeling 0.760 0.021 0.255 0.032 910 0.753 0.038 0.281 0.061 1.07 0.44 130 +0 .021 +0 .054 +0 .030 +0 .077 +0 .34 *NLA 0.790 0.031 0.338 0.082 622 0.803 0.050 0.343 0.066 0.80 0.43 99 +0 .031 +0 .055 +0 .038 +0 .065 +0 .29 NLA w/ free amp. per z-bin 0.794 0.033 0.306 0.059 506 0.791 0.048 0.319 0.059 0.87 0.61 70 +0 .028 +0 .057 +0 .044 +0 .059 +0.34 Mixed Alignment Model 0.766 0.030 0.298 0.050 489 0.727 0.034 0.282 0.051 1.39 0.43 86 Baryonic Effects

+0.032 +0.090 *Baryonic P (k) model 0.796 0.027 0.287 0.047 535 — —- —- —- Other Lensing Data

+0.062 +0.057 +0.068 +0.068 +0.37 DES SV 0.769 0.072 0.268 0.049 250 0.758 0.109 0.264 0.040 1.20 0.60 27 +0 .033 +0 .091 +0 .044 +0 .061 +0 .24 KiDS-450 0.757 0.038 0.265 0.045 405 0.759 0.042 0.326 0.078 0.64 0.38 77 +0 .024 +0 .037 +0 .035 +0 .064 +0 .34 Planck TT + lowP 0.840 0.028 0.334 0.020 1330 0.798 0.035 0.221 0.025 1.50 0.18 152 +0 .013 +0 .007 +0 .013 +0 .008 +0.053 Planck (+lensing) + BAO + JLA 0.815 0.015 0.306 0.007 9954 0.814 0.016 0.304 0.011 1.025 0.047 1337 measurements of cosmic shear; this must be explored further Science Verification results. We find a dark energy equation- +0.26 in future cosmic shear constraints. But whatever the explana- of-state w = 0.82 0.48 using cosmic shear alone. tion of the formally high 2, there seems no reason to believe We find an equally good fit to the wCDM model as for that it causes a significant bias in our cosmological results. ⇤CDM, with best-fit 2 of 268 for the 227 data points in the non-shaded region for the METACALIBRATION measurement. This is identical (with rounding) to the 2 for ⇤CDM. We can further compare the relative Bayesian evidence for each model B. Fiducial wCDM Results via the Bayes factor. The Bayesian evidence, or probability of observing a dataset D given a model M with parameters p, is We marginalize over a total of 7 cosmological parame- ters in the fiducial wCDM model, including a free neutrino P (D M)= dN pP (D p,M)P (p M). (13) | | | mass density, and 10 systematic or astrophysical parameters. Z w These are listed in Table II. Our fiducial CDM constraints The Bayes factor comparing the evidence for the wCDM and are shown in Fig. 10. We find a 5% constraint on S8, of +0.036 ⇤CDM models is then 0.789 0.038 at 68% CL, which is a significant improvement over the previous constraints from cosmic shear and more than P (D wCDM) K = | . (14) a factor of 2 improvement over the constraining power of our P (D ⇤CDM) | 18

0.5 FIG. 9. Summary of constraints on the 1D peak value of S8 8(⌦m/0.3) in ⇤CDM. The 68% CL are shown as horizontal bars. We distinguish variations on the fiducial setup that are not necessarily⌘ required to give consistent results (e.g., by neglecting astrophysical systematics) by an asterisk. The numerical parameter values are listed in Table III. Especially for external data, it is important to remember 2 that we vary ⌦⌫ h in our fiducial analysis, and thus all results we compare to, and so the central values and uncertainties of parameters may not follow intuition from previous results.

The interpretation of the Bayes factor can be characterized in 450 [34], we use the original measurements of ⇠ij, ni(z), multiple ways [134, 135]. We find log(K)= 0.6, which and covariance matrix from these works, but enforce± our fidu- indicates no preference for a model which allowsw = 1. cial model choices for intrinsic alignment and baryon scale 6 cuts, in addition to the recommended scale cuts in the origi- nal analysis. This changes the precise appearance of contours C. Comparison to External Measurements and quoted parameter values relative to the original works. We also recompute the posterior of other external data sets In order to place our constraints in the context of both other of complementary probes in our parameter spaces, including cosmic shear constraints and constraints from complementary varying neutrino mass. These include: probes of the Universe, we recompute the posterior of external results in our fiducial parameter spaces for both ⇤CDM and Constraints on the angular diameter distance from wCDM with a varying neutrino mass density. For recent real- • baryon acoustic oscillation (BAO) measurements – the space cosmic shear data, including DES SV [32] and KiDS- 6dF Galaxy Survey [136], the SDSS Data Release 7 19

1.00 Main Galaxy Sample [137], and the BOSS Data Re- DES Y1 lease 12 [138]. The BAO distances are measured rela- 0.95 KiDS-450 tive to a true physical BAO scale rd, which leads to a Planck factor that depends on the cosmological model, which 0.90 must be calculated at each likelihood step (see [138]).

0.85 measurements – the Joint • 0.80 Lightcurve Analysis of Type Ia Supernovae (SNe) from 8 Betoule et al. [139]. 0.75 Cosmic microwave background (CMB) temperature 0.70 • and polarization measurements – Planck [140] (‘Planck TT + lowP’), using TT (` = 30 – 2508) and 0.65 TT+TE+EE+BB (` =2– 29), and additionally includ- 0.60 ing lensing when combined with other external data. 0.55 We show a comparison of our fiducial ⇤CDM results to w0CDM.2 0.3 0.4 0.5 KiDS-450 and Planck in Fig. 7, excluding CMB lensing to 0.90 attempt to emphasize any differences between our low-z mea- surements and Planck CMB predictions from high-z. We find a constraint from cosmic shear that lies between the previ-

5 0.85 .

0 ous cosmic shear results and constraints from Planck. There 3)

. has been significant discussion in the literature regarding con- 0

/ sistency of cosmic shear constraints with the CMB. Though

m 0.80 we find results that are consistent with previous cosmic shear ( 8 constraints, we nevertheless see no significant evidence for

⌘ 0.75 disagreement with CMB data. It is worth noting that the in- 8 clusion of Planck lensing, not included in Fig. 7, lowers the S Planck estimate of S8 and ⌦m (Table 4 of [140]), thus fur- 0.70 ther reducing any apparent differences with our results in this plane. We leave a detailed discussion and interpretation of consistency of our data with external probes to Dark Energy 0.65 Survey et al. [51], where significantly tighter cosmological 0.4 0.2 0.3 0.4 0.5 constraints are presented when combining galaxy clustering 0.6 and galaxy-galaxy lensing with our cosmic shear results. We also compare our fiducial results in wCDM to results from 0.8 KiDS-450 and Planck in Fig. 10, where our conclusion about consistency is unchanged. Finally, we show in Fig. 9 and Ta- 1.0 ble III a summary of comparisons in ⇤CDM and wCDM of our fiducial constraints with DES SV, KiDS-450, Planck, and w 1.2 Planck + BAO + JLA. We find our constraints in S8 and ⌦m agree well with the combination Planck + BAO + JLA. We di- 1.4 rectly compare these external data sets to our fiducial ⇤CDM 1.6 constraints in Fig. 11. 1.8 IX. ROBUSTNESS TO MODELING AND DATA CHOICES 2.0 0.2 0.3 0.4 0.5 Validating the robustness of the cosmic shear signal to var- m ious potential residual systematic effects is technically chal- lenging. The two primary measurement uncertainties in the cosmic shear signal, multiplicative shear and photometric red-

FIG. 10. Fiducial constraints on the clustering amplitude 8, S8, shift biases, are each difficult to constrain directly. A resid- and w with the matter density ⌦m in wCDM. The fiducial DES Y1 ual bias due to multiplicative shear is typically constrained cosmic shear constraints are shown by the gray filled contours, with in two ways: measuring residual bias in shape measurement Planck CMB constraints given by the filled green contours, and cos- pipelines on simulated data, which relies critically on the fi- mic shear constraints from KiDS-450 by unfilled blue contours. In delity with which the image simulation matches the real data; all cases, 68% and 95% confidence levels are shown. A dotted line or comparing results between two independent shape mea- at w = 1 indicates the ⇤CDM value. surement pipelines. Results from both these methods are 20

DES Y1 TABLE IV. A comparison of the priors and posteriors of non- 0.95 Planck cosmological parameters in the fiducial analysis. BAO + JLA Planck + BAO + JLA Parameter Prior Posterior 0.90 Systematic 1 2 +2.3 +2.0 m 10 1.2 2.3 1.5 2.1 2 ⇥ 2 +2 .3 +2 .0 m 10 1.2 2.3 1.4 1.9 3 ⇥ 2 +2 .3 +1 .8 m 10 1.2 2.3 0.5 2.1 0.85 4 ⇥ 2 +2 .3 +2 .0 8 m 10 1.2 2.3 1.3 1.8 S 1⇥ 2 +1 .6 +1 .0 z 10 0.1 1.6 0.7 1.6 2 ⇥ 2 +1.3 +1.1 z 10 1.9 1.3 2.0 1.0 3 ⇥ 2 +1.1 +0.9 0.80 z 10 0.9 1.1 0.9 1.0 4 ⇥ 2 +2.2 +2.2 z 10 1.8 2.2 2.2 1.5 ⇥ Astrophysical +5.0 +0.5 A 0.0 5.0 1.3 0.6 0.75 +5 .0 +1 .0 ⌘ 0.0 5.0 3.7 2.3

0.70 0.2 0.3 0.4 0.5 m data selection and model design that can have a significant im- pact on any conclusions drawn from the analysis. Our choices are informed by an extensive battery of null tests, which were FIG. 11. A comparison of the DES Y1 constraints on S8 and the mat- performed while blind to the consequences on cosmological ter density ⌦m in ⇤CDM with external data sets excluding cosmic parameters. We also performed many robustness tests blinded, shear. The fiducial DES Y1 cosmic shear constraints are shown by where relative deviations in constraints are examined without the gray filled contours, with Planck CMB constraints given by the knowledge of the absolute value of any parameter. Some of green contours, the combination of BAO and JLA (SNe) constraints these variations and tests are summarized in Fig. 9 and Table in blue, and the combination Planck + BAO + JLA in orange. In all III, and we reproduce many of them here unblinded. cases, 68% and 95% confidence levels are shown. discussed in [54] and Sec. IX B below. Biases in photo- metric redshift measurement have typically been constrained A. Impact of Nuisance Parameter Marginalization by comparison to reference spectroscopic fields. However, spectroscopic samples are often incomplete and selected with magnitude or color criteria that cannot be reproduced in our While we marginalize over a large number of non- griz photometry, which leads to uncharacterized selection bi- cosmological parameters in the fiducial analysis, nearly twice ases. We instead constrain the photo-z distribution by com- as many as cosmological parameters, we find that our con- parison to a sample with accurate photo-z’s determined from straining power is not strongly impacted by marginalizing deep 30-band photometry. We combine with further con- over shear and photo-z systematic parameters in the S8 – straints from the angular cross-correlation of source galaxies ⌦m plane. In terms of S8 alone, marginalizing over shear with a subsample selected to have the most accurate photo-zs, and photo-z systematic parameters degrades our constraint i.e. bright red-sequence galaxies, as discussed in [39] and Sec. by about 10%. Our constraints are significantly degraded IX C below. by marginalizing over a model for intrinsic alignment (IA), The DES Y1 shape catalogs have undergone extensive null which is one reason that combining cosmic shear with the testing, both during catalog development and at the level other probes in [51] is so powerful. We also see a significant of specific probes or measurements, including those demon- bias when ignoring intrinsic alignment. This is demonstrated strated below for cosmic shear. Primary catalog-level tests in Fig. 8, where we compare the fiducial analysis (DES Y1 - are discussed in detail in [54], which includes tomographic filled gray contours) in ⇤CDM to the cases where we sample constraints on the shear B-mode signal, while other tests of only over cosmological parameters (blue contours) and both the shape and photo-z catalogs have been carried out in [39– cosmological and astrophysical parameters (green contours). 42, 53, 57]. We find no significant B-mode signal in [54], but Further discussion of the impact of intrinsic alignment mod- identify both a significant PSF model residual that we are able eling can be found in Sec. IX E. We compare the priors and to model and an unidentified source of mean shear, which we posteriors for the non-cosmological parameters in Table IV correct. These effects and their impact on the cosmic shear and Appendix C, and find that we do not gain a significant signal are discussed in Appendix A. Further successful null amount of information in constraining shear and photo-z bias tests of the cosmic shear signal are discussed in Appendix B. parameters in the fiducial analysis, while providing significant In designing a fiducial analysis, many choices are made in constraints on the intrinsic alignment model. 21

B. Shear Pipeline Comparison DES Y1 (METACALIBRATION) 0.90 In the DES SV shape catalog paper [61], we made explicit IM3SHAPE comparisons between the two shape measurement methods DES SV NGMIX & IM3SHAPE in simulations and at the two-point es- 0.85 timator level. This was informed by simultaneous measure- ments on simulated data by both shape measurement methods, which gave us an estimate of the relative selection bias, and resulted in choices that made the catalogs more similar. Resid- 0.80 8 ual differences ultimately provided the basis for the final prior S for m. This is even more complicated to do with METACALI- BRATION and IM3SHAPE due to the very different ways each 0.75 are calibrated. Instead, [54] perform detailed independent, ab initio estimations of uncertainty in m for each pipeline. This paper also demonstrates that there is no significant B-mode 0.70 signal in the data. To confirm that the two shear measurements from META- CALIBRATION and IM3SHAPE agree, we have relied on a 0.65 quantitative comparison of their agreement only at the level 0.2 0.3 0.4 0.5 of cosmological parameter constraints, where the differing se- lection of objects in each catalog is naturally accounted for. m This comparison was performed only once the two shape cat- alogs were finalized based on results of tests in [54], and is shown in Fig. 12 for ⇤CDM. The resulting contours in the FIG. 12. A comparison of ⇤CDM constraints in the S8 – ⌦m S8 – ⌦m plane are entirely consistent, though the mean of plane from the two shape measurement pipelines, METACALIBRA- TION (gray filled contours) and IM3SHAPE (blue contours). This is the IM3SHAPE constraint in S8 is shifted to slightly higher values. This is more significant in [51] where the contours a strong test of robustness to assumptions and differences in mea- surement and calibration methodology. Each pipeline utilizes very shrink significantly. The weaker constraint for IM3SHAPE is different and independent methods of shape measurement and shear due primarily to using only the r band for shape measurement, calibration. We also compare the DES SV results in green. Both relative to riz bands for METACALIBRATION, and additional 68% and 95% confidence levels are shown. necessary catalog selections to remove objects that cannot be calibrated accurately due to limitations of galaxy morphology in the input COSMOS catalog. These contribute to a signif- redshift bin, because the REDMAGIC sample ends at redshift icantly smaller ne↵ for IM3SHAPE. This agreement, reached z =0.9. Thus, we also tested removing the fourth bin from through two very different and independent shape measure- our analysis and confirmed in Fig. 13 that it does not produce ment and calibration strategies, is thus a very strong test of a significant shift in the inferred cosmology. Both tests were robustness to shape measurement errors in the final cosmo- performed before unblinding. logical constraints.

D. Scale Selection C. Photometric Redshift Comparison We remove any scales from the ⇠ data vector that would As discussed in Sec. II B, we rely on a combination of 1) have a fractional contribution from baryonic± interactions ex- comparisons to redshift distributions of resampled COSMOS ceeding 2%. We use results from the OWLS ‘AGN’ simula- objects and 2) constraints due to clustering cross-correlations tion, discussed in Sec. VII A, to determine this limit. This between source galaxies and REDMAGIC galaxies with very removes a significant number of data points on small scales, good photometric redshifts. We parameterize corrections to particularly in ⇠ where the impact of baryonic interactions is the n(z) as a shift in the mean redshift of the distribution larger. Similarly, we correct a significant bias due to residual of galaxies. As an independent test of whether shifting the mean shear in the signal that is partly due to PSF modeling mean of the redshift distribution captures the full effect of errors. This impacts the signal primarily on the largest scales. photo-z bias uncertainty, we show cosmological constraints To verify that our scale cuts are robust, we repeat the fiducial directly using the resampled COSMOS ni(z) redshift distri- analysis by splitting the angular range in two, separately con- butions measured from COSMOS (see Fig. 2). The resulting straining cosmology with the smaller and larger scales. This constraints are shown in Fig. 13, illustrating that differences split is at ✓ = 200 for ⇠+ and ✓ = 1500 for ⇠ . The re- in the shape of the redshift distribution are sub-dominant for sults of this test are shown in Fig. 14, with the fiducial analy- cosmic shear when matching the mean at the current statis- sis shown as the filled gray contours, along with results from tical precision (see also [141]). The independent constraint the smaller (green contours) and larger (blue contours) scale from clustering cross-correlations is unavailable in the fourth selection. We find consistent results in all three cases. To 22

DES Y1 (BPZ) DES Y1 0.90 Resampled COSMOS 0.90 Extended angular scales BPZ w/o highest z-bin Large angular scales Small angular scales 0.85 0.85

0.80 0.80 8 8 S S

0.75 0.75

0.70 0.70

0.65 0.65 0.2 0.3 0.4 0.5 0.2 0.3 0.4 0.5 m m

FIG. 13. A comparison of ⇤CDM constraints in the S8 – ⌦m plane FIG. 14. A comparison of ⇤CDM constraints in the S8 – ⌦m plane for different photo-z choices. The gray filled contours show the fidu- for different data vector choices. The filled gray contours show the cial analysis using the BPZ redshift distribution shown in Fig. 2 with fiducial analysis, while the orange contours small scales outside the the offset priors listed in Table II. The green contours are the result fiducial selection, the blue contours use fiducial angular scales for ⇠+ of the same analysis, but removing the fourth tomographic bin. The (⇠ ) with ✓>200 (✓>1500), and the green contours use fiducial blue contours use the COSMOS n(z) from which part of the BPZ scales for ⇠+ (⇠ ) with ✓<200 (✓<1500). The orange contours prior is derived. By design, the mean redshift of each tomographic should not necessarily agree with the fiducial case due to the impact bin agrees between the resampled COSMOS n(z) and the BPZ red- of baryonic effects, while we find consistent results using subsets shift distribution used in the fiducial analysis, but the shapes of the of our fiducial angular scale range. Both 68% and 95% confidence n(z) are significantly different in some tomographic bins, providing levels are shown. a test of whether parameterizing the photo-z bias as only a shift in the mean redshift is sufficient. Both 68% and 95% confidence levels are shown. model over which we marginalize. These include: 1) fixing the power-law redshift scaling of the fiducial model to have ⌘ =0, leaving a single-parameter (A) model; 2) removing demonstrate the potential degradation in constraining power the power-law dependence of redshift evolution to marginal- due to our baryon cuts on small scales, we also use the full ize over four free amplitudes in each redshift bin (Ai); 3) data vector from 2.50 <✓<2500. This is shown as the orange allowing for both tidal alignment and tidal torquing (linear contours in Fig. 14, but is likely biased due to inaccuracies and quadratic) alignment amplitudes (‘mixed’ model, [129]). in the modeling of baryonic effects. Finally, we verify (see Note that the mixed model includes IA B-mode contributions, Fig. 9 and Table III) that replacing our power spectrum cal- which are incorporated through P PE PB in Eq. 4. This culation with that from [114] and marginalizing additionally model also has mild dependence on! the source± galaxy bias, over the included two-parameter baryon feedback model in which we assume to be 1. Fig. 15 shows constraints in ⇤CDM ⇤CDM does not significantly change our inferred cosmology. for the fiducial model (NLA + z-power law – gray contours), Since the COSMOSIS interface to the model in [114] does not compared to the single-parameter NLA model (green con- yet incorporate more recent changes that account more accu- tours), the NLA model with a free amplitude in each tomo- rately for massive neutrinos, the neutrino mass density cannot graphic bin (orange contours), and the mixed alignment model be properly marginalized over and this result should be com- (blue contours). There is no significant difference in inferred pared to the fixed neutrino mass density constraint. cosmology between the first three models, which are well- tested and have been implemented in the literature before. The mixed alignment model, which includes alignment due to non- E. Intrinsic Alignment Modeling linear effects in the tidal field, including tidal torquing, does cause a non-negligible shift in inferred parameters. In addition to the fiducial intrinsic alignment (IA) model, We caution against concluding that the fiducial results pre- we also consider several variants to test the robustness of sented here are biased due to the shift in cosmology observed our results with respect to the choice of intrinsic alignment in Fig. 15 when using the mixed alignment model, however, 23

because we have seen similar trends to lower S8 and ⌦m in CDM DES Y1 (NLA + z-power law) less constraining data sets when marginalizing over too flex- 0.90 NLA ible an intrinsic alignment model. For example, the DES SV NLA (free amp. per z-bin) (and to a lesser degree IM3SHAPE) result in Fig. 12 (see also Mixed Alignment Model IA discussion in [32]), shows a similar trend toward this area 0.85 of parameter space with even the fiducial IA model in this work, which disappears with our more constraining DES Y1 data. We also see a similar trend toward this region when re- moving the highest redshift bin in Sec. IX C, which degrades 0.80 our ability to constrain the redshift evolution of the fiducial 8

S IA model. We further see much less significant an impact on cosmology in the full combined clustering and weak lensing 0.75 analysis when injecting a tidal torque signal of greater ampli- tude than we find here into a pure lensing signal [58]. It is worth noting that there is no significant difference in 2 or Bayesian evidence whether we include or not the tidal torque 0.70 contribution of the mixed alignment model. We thus conclude that while this is an interesting result, it requires further ex- ploration that we defer to a future work. Nevertheless, this result highlights the importance of considering the impact of 0.65 0.2 0.3 0.4 0.5 IA models beyond the tidal alignment (linear) paradigm in fu- wCDM ture cosmic shear studies, and it may indicate a real bias in 0.90 m cosmic shear at our statistical precision when using the fidu- cial tidal alignment model. A more conclusive answer for this question will require more constraining data, which we are 0.85 already working toward with DES Y3+ results, or better ex- ternal priors on the amplitude of the tidal torquing component (and orientation). 0.80 Given the constraining power of the DES Y1 analysis, it is 8

S clear that we can learn not just about cosmology, but also in- teresting astrophysical effects like IA. In Fig. 16 we compare the recovered value of A, the amplitude of the tidal alignment 0.75 (TA) model as a function of redshift in the four models consid- ered in this analysis. For the mixed alignment model, we also show the constraint on A2, the amplitude of the tidal torquing 0.70 (TT) component of the model. Note that subscripts are used with the amplitudes in the mixed alignment case and that A1 corresponds to the fiducial A parameter. We find good agree- ment in the TA amplitude between all four models, including 0.65 the mixed alignment case, where the contributions from TT 0.2 0.3 0.4 0.5 terms appear to primarily shift the inferred cosmological pa- m rameters rather than the TA amplitude. For the fiducial IA model and the mixed alignment model, which have a smooth functional form with redshift, we derive the amplitude at the FIG. 15. A comparison of the impact of different intrinsic align- mean redshifts of each redshift bin and report this value and its ment (IA) models on ⇤CDM and wCDM constraints in the S8 – ⌦m uncertainty. This analysis provides a significant improvement plane. The fiducial model (NLA + z-power law – gray contours), is in IA constraining power compared to previous analyses, with compared to the single-parameter NLA model (green contours), the detection of nonzero A =1.3 at the 92% CL when allowing a NLA model with a free amplitude in each tomographic bin (orange power-law redshift scaling, which is comparable to that when contours), and the mixed alignment model (blue contours). There assuming a fixed ⌘. The fiducial power-law ⌘ =3.7 is con- is no significant difference in inferred cosmology between the first three models, which are well-tested and have been implemented in strained to be positive at the 87% CL. In the mixed model, the literature before. The mixed alignment model, which includes A1 =1.1 is still constrained to be non-zero at 82% CL with alignment due to tidal torquing (or other nonlinear contributions), ⌘1 =2.4 constrained to be positive at the 80% CL. The tidal does cause a non-negligible shift in inferred parameters, which is torque amplitude A2 = 0.8 is nonzero at the 84% CL, with discussed further in Sec. IX E. Both 68% and 95% confidence levels a negative amplitude, and power-law ⌘2 = 1.5, which is are shown. consistent with zero. As discussed in [129], the sign conven- tion for A1 and A2 is such that positive values correspond to galaxy alignment towards overdense regions and thus a nega- 24

5 X. CONCLUSIONS DES Y1 (NLA + z-power law) NLA We have used 26 million galaxies from Dark Energy Sur- 4 NLA (free amp. per z-bin) vey (DES Y1) shape catalogs over 1321 deg2 of the southern Mixed Alignment Model (A1) sky to produce the most significant measurement of cosmic 3 Mixed Alignment Model (A2) shear in a galaxy survey to date. We constrain cosmologi- cal parameters in both the ⇤CDM and wCDM models, while i

A also varying the neutrino mass density. We find a 3% frac- 2 +0.024 tional uncertainty on S8 =0.789 0.026 at 68% CL, which is a factor of 3 improvement over the constraining power of 1 our SV results [61] and a factor 1.5 tighter than the previous

Effective state-of-the-art cosmic shear survey, KiDS [34]. In wCDM, +0.036 we find a 5% fractional uncertainty on S8 =0.789 0.038 and 0 +0.26 w = 0.82 0.48. We find no evidence preferring the addi- tion ofw = 1 using cosmic shear alone, and no constraint 6 1 beyond our prior on the neutrino mass density. Our constraints from cosmic shear lie between the previ- ous cosmic shear results from KiDS-450 and CMB data from 2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 Planck. Though we find results that are consistent with previ- z ous cosmic shear constraints in S8 – ⌦m, preferring a slightly lower value of S8 than Planck, we nevertheless see no evi- dence for disagreement of our weak lensing data with data from the CMB. Significantly tighter cosmological constraints FIG. 16. The constraint on the intrinsic alignment amplitude A as a when galaxy clustering and galaxy-galaxy lensing is added to function of redshift in the four models considered. For all models, our fiducial cosmic shear measurements are discussed in Dark we show A1, the amplitude of the tidal alignment (TA) model, while Energy Survey et al. [51], and we expect further significant for the mixed alignment model, we also show the constraint on A2, the amplitude of the tidal torquing (TT) component of the model. improvements with subsequent years of DES data, which will We find good agreement between the redshift evolution of the tidal more than triple the sky coverage of our shape catalogs and alignment amplitude in the four models. double the effective integrated exposure time per galaxy. We have detailed a suite of rigorous null tests at the data- and catalog-level [39–42, 52–54, 57] and robustness and vali- dation checks of our measurements in this work and [51] that tive GI term. provide us with confidence in the accuracy of these results. The measured fiducial IA amplitude is in reasonable agree- We employ two independent and very different shape mea- ment with our prediction of A 0.5 at z =0.62, obtained surement and calibration methods to measure cosmic shear, ⇡ piv from extrapolating IA amplitude scalings calibrated on galax- METACALIBRATION and IM3SHAPE, which agree very well ies that are significantly more luminous than our lensing sam- with each other despite being developed without attempting ple [58]. Our analysis thus provides significant improvement to study or enforce agreement at the two-point level. We in constraining the IA signal in weak lensing measurements. further employ two independent methods of calibrating our Moreover, it is the first indication of nonlinear alignment photometric redshift distributions, including clustering cross- mechanisms, such as tidal torquing, in a general weak lens- correlations with a photometric sample of galaxies with pre- ing sample. Previous weak lensing studies (e.g., [32, 34]) did cise and accurate photo-zs. We account for the intrinsic align- not account for the potential presence of these higher-order ment of galaxies, finding evidence for tidal alignment in our effects, while spectroscopic alignment studies on blue/spiral fiducial analysis and for tidal torquing (quadratic) alignment galaxies have placed comparatively weak constraints on these in an extended analysis. We employ two independently devel- contributions (e.g., [128]). Recent hydrodynamic simulations oped parameter inference pipelines, COSMOLIKE and COS- have also examined the expected alignment of both disk and MOSIS, that have been validated against one another, and we elliptical galaxies (e.g., [142–144]). These simulations con- validate components of our covariance matrix using a limited sistently find an overall alignment towards overdense regions, number of N-body mock catalogs with ray-traced shear, a dominated by elliptical galaxies, in agreement with the sign of large number of lognormal simulations, and jackknife mea- our measured A. However, they disagree on the IA behavior surements in our data to validate shape-noise contributions of spiral galaxies (as well as other kinematic properties), with to the covariance. Finally, we present a series of robust- [142] finding tangential alignment of the major axis with over- ness checks to variations on the fiducial analysis to demon- densities, consistent with our tentative measurement of A2, strate that our analysis choices are secure and well moti- while others find radial alignment (see [143] for a compari- vated. The statistical power of our weak lensing data set, par- son). Improved observational data and hydrodynamic simula- ticularly when combined with galaxy clustering and galaxy- tions, along with advances in analytic modeling, will clarify galaxy lensing measurements in [51], now constrains low- this issue. redshift clustering as strongly as do CMB measurements. 25

ACKNOWLEDGEMENTS de F´ısica d’Altes Energies, Lawrence Berkeley National Lab- oratory, the Ludwig-Maximilians Universitat¨ Munchen¨ and the associated Excellence Cluster Universe, the University We thank C. M. Hirata for useful conversations in the of Michigan, the National Optical Astronomy Observatory, course of preparing this work. TE is supported by NASA the University of Nottingham, The Ohio State University, ROSES ATP 16-ATP16-0084 grant and by NASA ROSES the University of Pennsylvania, the University of Portsmouth, 16-ADAP16-0116; TE’s research was carried out at the Jet SLAC National Accelerator Laboratory, Stanford University, Propulsion Laboratory, California Institute of Technology, un- the University of Sussex, Texas A&M University, and the der a contract with the National Aeronautics and Space Ad- OzDES Membership Consortium. ministration. EK was supported by a Kavli Fellowship at Based in part on observations at Cerro Tololo Inter- Stanford University. OF was supported by SFB-Transregio American Observatory, National Optical Astronomy Observa- 33 ‘The Dark Universe’ of the Deutsche Forschungsgemein- tory, which is operated by the Association of Universities for schaft (DFG) and by the DFG Cluster of Excellence ‘Origin Research in Astronomy (AURA) under a cooperative agree- and Structure of the Universe’. Support for DG was provided ment with the National Science Foundation. by NASA through Einstein Postdoctoral Fellowship grant The DES data management system is supported by the number PF5-160138 awarded by the Chandra X-ray Center, National Science Foundation under Grant Numbers AST- which is operated by the Smithsonian Astrophysical Observa- 1138766 and AST-1536171. The DES participants from tory for NASA under contract NAS8-03060. MJ, BJ, and GB Spanish institutions are partially supported by MINECO are partially supported by the US Department of Energy grant under grants AYA2015-71825, ESP2015-88861, FPA2015- DE-SC0007901 and funds from the University of Pennsylva- 68048, SEV-2012-0234, SEV-2016-0597, and MDM-2015- nia. SS acknowledges receipt of a Doctoral Training Grant 0509, some of which include ERDF funds from the Euro- from the UK Science and Technology Facilities Council. pean Union. IFAE is partially funded by the CERCA pro- Funding for the DES Projects has been provided by the U.S. gram of the Generalitat de Catalunya. Research leading to Department of Energy, the U.S. National Science Foundation, these results has received funding from the European Re- the Ministry of Science and Education of Spain, the Science search Council under the European Union’s Seventh Frame- and Technology Facilities Council of the United Kingdom, the work Program (FP7/2007-2013) including ERC grant agree- Higher Education Funding Council for England, the National ments 240672, 291329, and 306478. We acknowledge sup- Center for Supercomputing Applications at the University of port from the Australian Research Council Centre of Excel- Illinois at Urbana-Champaign, the Kavli Institute of Cosmo- lence for All-sky Astrophysics (CAASTRO), through project logical Physics at the University of Chicago, the Center for number CE110001020. Cosmology and Astro-Particle Physics at the Ohio State Uni- This manuscript has been authored by Fermi Research Al- versity, the Mitchell Institute for Fundamental Physics and liance, LLC under Contract No. DE-AC02-07CH11359 with Astronomy at Texas A&M University, Financiadora de Estu- the U.S. Department of Energy, Office of Science, Office of dos e Projetos, Fundac¸ao˜ Carlos Chagas Filho de Amparo a` High Energy Physics. The United States Government retains Pesquisa do Estado do Rio de Janeiro, Conselho Nacional de and the publisher, by accepting the article for publication, ac- Desenvolvimento Cient´ıfico e Tecnologico´ and the Ministerio´ knowledges that the United States Government retains a non- da Ciencia,ˆ Tecnologia e Inovac¸ao,˜ the Deutsche Forschungs- exclusive, paid-up, irrevocable, world-wide license to publish gemeinschaft and the Collaborating Institutions in the Dark or reproduce the published form of this manuscript, or allow Energy Survey. others to do so, for United States Government purposes. The Collaborating Institutions are Argonne National Lab- Plots in this work were produced partly with MATPLOTLIB oratory, the University of California at Santa Cruz, the Uni- [145], and has made use of NASA’s Astrophysics Data Sys- versity of Cambridge, Centro de Investigaciones Energeticas,´ tem Bibliographic Services. This research used resources of Medioambientales y Tecnologicas-Madrid,´ the University of the National Energy Research Scientific Computing Center, a Chicago, University College London, the DES-Brazil Consor- DOE Office of Science User Facility supported by the Office tium, the University of Edinburgh, the Eidgenossische¨ Tech- of Science of the U.S. Department of Energy under Contract nische Hochschule (ETH) Zurich,¨ Fermi National Accelerator No. DE-AC02-05CH11231. This work also used resources Laboratory, the University of Illinois at Urbana-Champaign, at the Ohio Supercomputing Center [146] and computing re- the Institut de Ciencies` de l’Espai (IEEC/CSIC), the Institut sources at SLAC National Accelerator Laboratory.

[1] G. R. Blumenthal, S. M. Faber, J. R. Primack, and M. J. Rees, [4] D. H. Weinberg, M. J. Mortonson, D. J. Eisenstein, C. Hirata, Nature 311, 517 (1984). A. G. Riess, and E. Rozo, Phys. Rep. 530, 87 (2013). [2] A. Albrecht, G. Bernstein, R. Cahn, W. L. Freedman, J. He- [5] M. Bartelmann and P. Schneider, Phys. Rep. 340, 291 (2001). witt, W. Hu, J. Huth, M. Kamionkowski, E. W. Kolb, L. Knox, [6] B. Jain and U. Seljak, Astrophys. J. 484, 560 (1997). J. C. Mather, S. Staggs, and N. B. Suntzeff, ArXiv Astro- [7] R. A. Battye and A. Moss, Phys. Rev. Lett. 112, 051303 physics e-prints (2006). (2014). [3] J. Peacock and P. Schneider, The Messenger 125, 48 (2006). [8] B. Leistedt, H. V. Peiris, and L. Verde, Phys. Rev. Lett. 113, 26

041301 (2014). [44] B. Joachimi, M. Cacciato, T. D. Kitching, A. Leonard, [9] F. Beutler et al., Mon. Not. R. Astron. Soc. 444, 3501 (2014). R. Mandelbaum, B. M. Schafer,¨ C. Sifon,´ H. Hoekstra, [10] D. J. Bacon, A. R. Refregier, and R. S. Ellis, Mon. Not. R. A. Kiessling, D. Kirk, and A. Rassat, Sp. Sci. Rev. 193,1 Astron. Soc. 318, 625 (2000). (2015), arXiv:1504.05456. [11] N. Kaiser, G. Wilson, and G. A. Luppino, arXiv:astro- [45] M. P. van Daalen, J. Schaye, C. M. Booth, and C. Dalla Vec- ph/0003338 (2000). chia, Mon. Not. R. Astron. Soc. 415, 3649 (2011). [12] D. M. Wittman, J. A. Tyson, D. Kirkman, I. Dell’Antonio, and [46] E. Semboloni, H. Hoekstra, J. Schaye, M. P. van Daalen, and G. Bernstein, Nature 405, 143 (2000). I. G. McCarthy, Mon. Not. R. Astron. Soc. 417, 2020 (2011). [13] L. van Waerbeke et al., Astron. Astrophys. 358, 30 (2000). [47] J. Harnois-Deraps,´ L. van Waerbeke, M. Viola, and C. Hey- [14] H. Hoekstra, H. K. C. Yee, M. D. Gladders, L. F. Barrientos, mans, Mon. Not. R. Astron. Soc. 450, 1212 (2015). P. B. Hall, and L. Infante, Astrophys. J. 572, 55 (2002). [48] M. R. Becker et al. (DES Collaboration), Phys. Rev. D 94, [15] L. Van Waerbeke, Y. Mellier, and H. Hoekstra, Astron. Astro- 022002 (2016). phys. 429, 75 (2005). [49] T. Kacprzak et al. (DES Collaboration), Mon. Not. R. Astron. [16] M. Jarvis, B. Jain, G. Bernstein, and D. Dolney, Astrophys. J. Soc. 463, 3653 (2016). 644, 71 (2006). [50] J. Kwan et al. (DES Collaboration), Mon. Not. R. Astron. Soc. [17] E. Semboloni, Y. Mellier, L. van Waerbeke, H. Hoekstra, 464, 4045 (2017). I. Tereno, K. Benabed, S. D. J. Gwyn, L. Fu, M. J. Hudson, [51] DES Collaboration et al., to be submitted to Phys. Rev. D R. Maoli, and L. C. Parker, Astron. Astrophys. 452, 51 (2006). (2017). [18] J. Benjamin, C. Heymans, E. Semboloni, L. van Waerbeke, [52] A. Drlica-Wagner et al. (DES Collaboration), submitted to As- H. Hoekstra, T. Erben, M. D. Gladders, M. Hetterscheidt, trophys. J. Suppl. Ser. (2017). Y. Mellier, and H. K. C. Yee, MNRAS 381, 702 (2007), astro- [53] J. Elvin-Poole et al. (DES Collaboration), to be submitted to ph/0703570. Phys. Rev. D (2017). [19] R. Massey et al., Astrophys. J. 172, 239 (2007). [54] J. Zuntz et al. (DES Collaboration), submitted to Mon. Not. R. [20] M. Hetterscheidt, P. Simon, M. Schirmer, H. Hildebrandt, Astron. Soc. (2017). T. Schrabback, T. Erben, and P. Schneider, Astron. Astrophys. [55] S. Samuroff et al. (DES Collaboration), to be submitted to 468, 859 (2007). Mon. Not. R. Astron. Soc. (2017). [21] T. Schrabback et al., Astron. Astrophys. 516, A63 (2010). [56] C. Chang et al. (DES Collaboration), submitted to Mon. Not. [22] H. Lin, S. Dodelson, H.-J. Seo, M. Soares-Santos, J. Annis, R. Astron. Soc. (2017). J. Hao, D. Johnston, J. M. Kubo, R. R. R. Reis, and M. Simet, [57] J. Prat et al. (DES Collaboration), to be submitted to Phys. Astrophys. J. 761, 15 (2012). Rev. D (2017). [23] E. M. Huff, T. Eifler, C. M. Hirata, R. Mandelbaum, [58] E. Krause et al. (DES Collaboration), submitted to Phys. Rev. D. Schlegel, and U. Seljak, Mon. Not. R. Astron. Soc. 440, D (2017), arXiv:1706.09359 [astro-ph.CO]. 1322 (2014). [59] N. MacCrann et al. (DES Collaboration), in prep. (2017). [24] M. J. Jee, J. A. Tyson, M. D. Schneider, D. Wittman, [60] B. Flaugher et al. (DES Collaboration), Astron. J. 150, 150 S. Schmidt, and S. Hilbert, Astrophys. J. 765, 74 (2013). (2015). [25] M. J. Jee, J. A. Tyson, S. Hilbert, M. D. Schneider, S. Schmidt, [61] M. Jarvis et al. (DES Collaboration), Mon. Not. R. Astron. and D. Wittman, Astrophys. J. 824, 77 (2016). Soc. 460, 2245 (2016). [26] C. Heymans et al., Mon. Not. R. Astron. Soc. 427, 146 (2012). [62] E. Huff and R. Mandelbaum, ArXiv e-prints (2017), [27] M. Kilbinger et al., Mon. Not. R. Astron. Soc. 430, 2200 arXiv:1702.02600. (2013). [63] E. S. Sheldon and E. M. Huff, Astrophys. J. 841, 24 (2017). [28] C. Heymans et al., Mon. Not. R. Astron. Soc. 432, 2433 [64] K. M. Gorski,´ E. Hivon, A. J. Banday, B. D. Wandelt, F. K. (2013). Hansen, M. Reinecke, and M. Bartelmann, Astrophys. J. 622, [29] J. Benjamin et al., Mon. Not. R. Astron. Soc. 431, 1547 759 (2005). (2013). [65] J. Zuntz, T. Kacprzak, L. Voigt, M. Hirsch, B. Rowe, and [30] T. D. Kitching et al., Mon. Not. R. Astron. Soc. 442, 1326 S. Bridle, Mon. Not. R. Astron. Soc. 434, 1604 (2013). (2014). [66] N. Ben´ıtez, Astrophys. J. 536, 571 (2000). [31] S. Joudaki, C. Blake, C. Heymans, A. Choi, J. Harnois- [67] C. Chang, M. Jarvis, B. Jain, S. M. Kahn, D. Kirkby, A. Con- Deraps, H. Hildebrandt, B. Joachimi, A. Johnson, A. Mead, nolly, S. Krughoff, E.-H. Peng, and J. R. Peterson, Mon. Not. D. Parkinson, M. Viola, and L. van Waerbeke, Mon. Not. R. R. Astron. Soc. 434, 2121 (2013). Astron. Soc. 465, 2033 (2017). [68] C. Laigle, H. J. McCracken, O. Ilbert, B. C. Hsieh, I. David- [32] DES Collaboration et al., Phys. Rev. D 94, 022001 (2016). zon, P. Capak, and others, Astrophys. J.s 224, 24 (2016). [33] K. Kuijken et al., Mon. Not. R. Astron. Soc. 454, 3500 (2015). [69] J. DeRose, R. Wechsler, E. Rykoff, et al., in prep. (2017). [34] H. Hildebrandt et al., Mon. Not. R. Astron. Soc. 465, 1454 [70] R. Wechsler, J. DeRose, M. Busha, et al., in prep. (2017). (2017). [71] V. Springel, Mon. Not. R. Astron. Soc. 364, 1105 (2005). [35] S. Joudaki et al., ArXiv e-prints (2016), arXiv:1610.04606. [72] M. Crocce, S. Pueblas, and R. Scoccimarro, Mon. Not. R. [36] F. Kohlinger¨ et al., ArXiv e-prints (2017), arXiv:1706.02892. Astron. Soc. 373, 369 (2006). [37] E. van Uitert et al., ArXiv e-prints (2017), arXiv:1706.05004. [73] M. C. Cooper, J. A. Aird, A. L. Coil, M. Davis, S. M. Faber, [38] S. Joudaki et al., ArXiv e-prints (2017), arXiv:1707.06627. S. Juneau, J. M. Lotz, K. Nandra, J. A. Newman, C. N. A. [39] B. Hoyle et al. (DES Collaboration), to be submitted to Mon. Willmer, and R. Yan, Astrophys. J.s 193, 14 (2011). Not. R. Astron. Soc. (2017). [74] M. R. Becker, Mon. Not. R. Astron. Soc. 435, 115 (2013). [40] R. Cawthon et al. (DES Collaboration), in prep. (2017). [75] P. Fosalba, E. Gaztanaga,˜ F. J. Castander, and M. Manera, [41] M. Gatti et al. (DES Collaboration), in prep. (2017). Mon. Not. R. Astron. Soc. 391, 435 (2008). [42] C. Davis et al. (DES Collaboration), in prep. (2017). [76] P. Fosalba, E. Gaztanaga,˜ F. J. Castander, and M. Crocce, [43] M. Troxel and M. Ishak, Physics Reports 558, 1 (2015). Mon. Not. R. Astron. Soc. 447, 1319 (2015). 27

[77] M. R. Blanton et al., Astrophys. J. 592, 819 (2003). 1536 (2010). [78] I. Zehavi et al., Astrophys. J. 736, 59 (2011). [113] I. G. McCarthy, J. Schaye, T. J. Ponman, R. G. Bower, C. M. [79] O. Ilbert et al., Astrophys. J. 690, 1236 (2009). Booth, C. Dalla Vecchia, R. A. Crain, V. Springel, T. Theuns, [80] P. Fosalba, M. Crocce, E. Gaztanaga,˜ and F. J. Castander, and R. P. C. Wiersma, MNRAS 406, 822 (2010). Mon. Not. R. Astron. Soc. 448, 2987 (2015). [114] A. J. Mead, J. A. Peacock, C. Heymans, S. Joudaki, and A. F. [81] M. Crocce, F. J. Castander, E. Gaztanaga,˜ P. Fosalba, and Heavens, Mon. Not. R. Astron. Soc. 454, 1958 (2015). J. Carretero, Mon. Not. R. Astron. Soc. 453, 1513 (2015). [115] K. Heitmann, M. White, C. Wagner, S. Habib, and D. Higdon, [82] J. Carretero, F. J. Castander, E. Gaztanaga,˜ M. Crocce, and The Astrophysical Journal 715, 104 (2010). P. Fosalba, Mon. Not. R. Astron. Soc. 447, 646 (2015). [116] K. Heitmann, E. Lawrence, J. Kwan, S. Habib, and D. Hig- [83] P. Coles and B. Jones, ”MNRAS” 248, 1 (1991). don, The Astrophysical Journal 780, 111 (2014). [84] I. Kayo, A. Taruya, and Y. Suto, Astrophys. J. 561, 22 (2001). [117] A. Heavens, A. Refregier, and C. Heymans, Mon. Not. R. [85] O. Lahav and Y. Suto, Living Rev. Rel. 7, 8 (2004). Astron. Soc. 319, 649 (2000). [86] S. Hilbert, J. Hartlap, and P. Schneider, Astron. Astrophys. [118] R. A. C. Croft and C. A. Metzler, Astrophys. J. 545, 561 536, A85 (2011). (2000). [87] H. S. Xavier, F. B. Abdalla, and B. Joachimi, Mon. Not. Roy. [119] P. Catelan, M. Kamionkowski, and R. D. Blandford, Mon. Astron. Soc. 459, 3693 (2016). Not. R. Astron. Soc. 320, L7 (2001). [88] P. Schneider, L. van Waerbeke, and Y. Mellier, A&A 389, 729 [120] R. G. Crittenden, P. Natarajan, U.-L. Pen, and T. Theuns, As- (2002), astro-ph/0112441. trophys. J. 559, 552 (2001). [89] M. Jarvis, G. Bernstein, and B. Jain, Mon. Not. R. Astron. [121] C. M. Hirata and U. Seljak, Phys. Rev. D 70, 063526 (2004). Soc. 352, 338 (2004). [122] B. Joachimi, R. Mandelbaum, F. B. Abdalla, and S. L. Bridle, [90] E. Krause and T. Eifler, Mon. Not. R. Astron. Soc. 470, 2100 Astron. Astrophys. 527, A26 (2011). (2017). [123] D. Kirk, A. Rassat, O. Host, and S. Bridle, Mon. Not. R. [91] T. Eifler, P. Schneider, and J. Hartlap, Astron. Astrophys. 502, Astron. Soc. 424, 1647 (2012). 721 (2009). [124] E. Krause, T. Eifler, and J. Blazek, Mon. Not. R. Astron. Soc. [92] J. R. Klein and A. Roodman, Ann. Rev. Nucl. Part. Sci. 55, 456, 207 (2016). 141 (2005). [125] J. Blazek, M. McQuinn, and U. Seljak, J. Cosmology As- [93] N. Kaiser, Astrophys. J. 388, 272 (1992). tropart. Phys. 5, 010 (2011). [94] N. Kaiser, Astrophys. J. 498, 26 (1998). [126] J. Blazek, Z. Vlah, and U. Seljak, J. Cosmology Astropart. [95] D. N. Limber, Astrophys. J. 117, 134 (1953). Phys. 8, 015 (2015). [96] T. Eifler, E. Krause, P. Schneider, and K. Honscheid, Mon. [127] J. Lee and U.-L. Pen, Astrophys. J. Lett. 532, L5 (2000). Not. R. Astron. Soc. 440, 1379 (2014). [128] R. Mandelbaum et al., Mon. Not. R. Astron. Soc. 410, 844 [97] J. Zuntz, M. Paterno, E. Jennings, D. Rudd, A. Manzotti, (2011). S. Dodelson, S. Bridle, S. Sehrish, and J. Kowalkowski, As- [129] J. Blazek, M. MacCrann, M. A. Troxel, and X. Fang, in prep. tronomy and Computing 12, 45 (2015). (2017). [98] D. Blas, J. Lesgourgues, and T. Tram, J. Cosmology As- [130] S. Bridle and L. King, New Journal of Physics 9, 444 (2007). tropart. Phys. 7, 034 (2011). [131] C. Heymans et al., Mon. Not. R. Astron. Soc. 368, 1323 [99] A. Lewis, A. Challinor, and A. Lasenby, Astrophys. J. 538, (2006). 473 (2000). [132] D. Huterer, M. Takada, G. Bernstein, and B. Jain, Mon. Not. [100] C. Howlett, A. Lewis, A. Hall, and A. Challinor, J. Cosmol- R. Astron. Soc. 366, 101 (2006). ogy Astropart. Phys. 4, 027 (2012). [133] E. Sellentin and A. F. Heavens, ArXiv e-prints (2017), [101] F. Feroz, M. P. Hobson, and M. Bridges, Mon. Not. R. Astron. arXiv:1707.04488. Soc. 398, 1601 (2009). [134] H. Jeffreys, Theory of Probability, 3rd ed. (Oxford, Oxford, [102] F. Feroz and M. P. Hobson, Mon. Not. R. Astron. Soc. 384, England, 1961). 449 (2008). [135] R. E. Kass and A. E. Raftery, Journal of the American Statis- [103] F. Feroz, M. P. Hobson, E. Cameron, and A. N. Pettitt, ArXiv tical Association 90, 773 (1995). e-prints (2013), arXiv:1306.2144. [136] F. Beutler, C. Blake, M. Colless, D. H. Jones, L. Staveley- [104] D. Foreman-Mackey, D. W. Hogg, D. Lang, and J. Goodman, Smith, L. Campbell, Q. Parker, W. Saunders, and F. Watson, Publ. Astron. Soc. Pac. 125, 306 (2013). Mon. Not. Roy. Astron. Soc. 416, 3017 (2011). [105] C. Patrignani et al. (Particle Data Group), Chin. Phys. C40, [137] A. J. Ross, L. Samushia, C. Howlett, W. J. Percival, A. Burden, 100001 (2016). and M. Manera, Mon. Not. Roy. Astron. Soc. 449, 835 (2015). [106] Planck Collaboration, P. A. R. Ade, et al., ArXiv e-prints [138] S. Alam et al., submitted to Mon. Not. Roy. Astron. Soc. (2015), arXiv:1502.01589. (2016). [107] D. Huterer and M. S. Turner, Phys. Rev. D 64, 123527 (2001). [139] M. Betoule et al., Astron. Astrophys. 568, A22 (2014). [108] Y. Wang, Phys. Rev. D 77, 123525 (2008). [140] P. A. R. Ade et al., Astron. Astrophys. 594, A13 (2016). [109] R. E. Smith, J. A. Peacock, A. Jenkins, S. D. M. White, C. S. [141] C. Bonnett et al. (DES Collaboration), Phys. Rev. D 94, Frenk, F. R. Pearce, P. A. Thomas, G. Efstathiou, and H. M. P. 042005 (2016). Couchman, Mon. Not. R. Astron. Soc. 341, 1311 (2003). [142] N. Chisari, C. Laigle, S. Codis, Y. Dubois, J. Devriendt, [110] R. Takahashi, M. Sato, T. Nishimichi, A. Taruya, and L. Miller, K. Benabed, A. Slyz, R. Gavazzi, and C. Pichon, M. Oguri, Astrophys. J. 761, 152 (2012). Mon. Not. R. Astron. Soc. 461, 2702 (2016). [111] S. Bird, M. Viel, and M. G. Haehnelt, Mon. Not. R. Astron. [143] A. Tenneti, R. Mandelbaum, and T. Di Matteo, Mon. Not. R. Soc. 420, 2551 (2012). Astron. Soc. 462, 2668 (2016). [112] J. Schaye, C. Dalla Vecchia, C. M. Booth, R. P. C. Wiersma, [144] S. Hilbert, D. Xu, P. Schneider, V. Springel, M. Vogelsberger, T. Theuns, M. R. Haas, S. Bertone, A. R. Duffy, I. G. Mc- and L. Hernquist, Mon. Not. R. Astron. Soc. 468, 790 (2017). Carthy, and F. van de Voort, Mon. Not. R. Astron. Soc. 402, [145] J. D. Hunter, Computing In Science & Engineering 9, 90 28

(2007). [146] O. S. Center, “Ohio supercomputer center,” http://osc. TABLE V. Constraints on ‘ ’, the proportionality constant when as- edu/ark:/19495/f5s1ph73 (1987). suming a linear relationship between inferred shear and PSF model [147] S. Paulin-Henriksson, A. Amara, L. Voigt, A. Refregier, and ellipticity residual (see Eqn. A2). Errors quoted are 68% confidence intervals. S. L. Bridle, Astron. Astrophys. 484, 67 (2008). [148] B. Rowe, Mon. Not. R. Astron. Soc. 404, 350 (2010). Redshift bin METACALIBRATION IM3SHAPE 1 0.72 0.26 1.65 0.54 ± ± Appendix A: Residual PSF Model Bias and Mean Shear 2 0.99 0.32 2.45 0.64 3 0.72 ± 0.32 1.4 ±0.60 4 1.31 ± 0.43 0.92± 0.78 1. Residual PSF Model Errors ± ±

A robust treatment of the PSF is crucial for unbiased cos- mic shear measurements. Imperfect modeling or deconvolu- we use the following cross-correlations tion of the PSF can produce coherent additive and multiplica- tive shear biases, both of which contaminate the cosmic shear eobsep = ↵ epep + qep (A3) signal [147]. In [54], we identified spatially correlated ellip- h i h i obs p ticity errors in the PSF modeling. We model the impact of PSF e q = ↵ e q + qq . (A4) ⌦ ↵ h i h i model ellipticity errors on the inferred shear using the linear ⌦ ↵ relation Note that in the above, the angle-brackets denote correlations of spin-2 quantities; we use the ⇠+(✓) statistic for all of these. sys p ei = (ei ei⇤)=qi (A1) Eqs. (A3) and (A4) follow from Eq. (A2), and provide a means to find ↵ and , which are taken to be free parameters p where i denotes the shear component, ✏ is the PSF model el- (for each redshift bin). The correlations epep , qep , epq , i h i h i h i lipticity, ✏⇤ is the true PSF ellipticity and therefore qi is the ith and qq are measured from the star catalog described in [54]. i h i component of the PSF model ellipticity residual. This relation Figure 17 shows these two measured correlations for the is exact in the case of an unweighted quadrupole ellipticity METACALIBRATION catalog, for each redshift bin. Uncer- estimator, if both the galaxy and PSF profiles are Gaussian. tainties are estimated using the lognormal mock shear cata- While neither of these conditions are satisfied in our case, we logs described in Sec. III C. For both catalogs, we find the use this model as a first-order approximation. If, as well as ↵ values to be consistent with zero. Given this, and the fact PSF modeling errors, there are also errors in the deconvolu- we have no a priori reason to expect a non-zero ↵, we assume tion of the PSF model from the galaxy image, one might also ↵ =0from now on. As expected we find 1 for all red- expect a systematic bias that is proportional to the PSF model shift bins; constraints are shown in Table V.⇠ Solid lines show ellipticity (sometimes this term is referred to as PSF leakage), the model predictions with the best-fit . Note that these are such that the model for the shear bias becomes actually constraints on the mean for each redshift bin - the value of for a specific galaxy will depend on the specific sys p i ei = ↵iei + qi. (A2) galaxy and PSF properties. Given these estimates of , we estimate the impact of PSF Note that we have no reason to expect non-zero ↵i from either model ellipticity errors on our cosmological parameter infer- shape measurement method; METACALIBRATION uses a cir- ence as follows. The expected systematic contamination of cularized PSF, and IM3SHAPE uses calibration simulations to ij ⇠ , where ij denotes the redshift bin pair, is remove any correlation with the PSF ellipticity. On the other + hand we expect all shear estimation algorithms to have a non- ⇠ij,sys = ij qq q 2 q 2 (A5) zero ; even a ‘perfect’ shear estimator has to assume a PSF + h ih 1i h 2i model, and errors in that PSF model will propagate to errors in ⌦ ↵ ⇣ ⌘ the shear estimation (Paulin-Henriksson et al. [147] estimate where the second and third terms on the RHS arise because we i for an un-weighted moments shear estimator). In [54], we are subtracting the mean ellipticity from each tomographic bin measure a significant correlation between the estimated shear to correct for scale-independent additive biases (see Sec. A 2). and the PSF model ellipticity. This could be evidence for non- We expect that on large scales (where additive biases are most zero ↵, but could also arise from correlations between the significant), is uncorrelated between galaxies, and there- PSF model ellipticity and the PSF ellipticity residuals even fore make the assumption that ij = i j . Using for ↵ =0. We demonstrate below that the latter is the most the measured qq (also known as the first ‘⇢-statistic’, ⇢1(✓) likely explanation. [148]), and theh best-fiti values⌦ from↵ Table⌦ V,↵⌦ we produce↵ a While we have an estimate of epi at each galaxy position, contaminated prediction of our data vector, which we then an- we can only estimate qi at the position of stars (where we can alyze using our parameter estimation framework to check for evaluate the PSF model and directly measure the star’s pro- biases in cosmological parameters that this level of contami- file). Therefore, we use cross-correlations between the galaxy nation would induce. We thus verify that the level of impact and star samples in order to simultaneously estimate ↵ and on cosmological parameters is entirely negligible, following (we assume from now on that ↵1 = ↵2,1 = 2). To do this, the subtraction of the mean shear discussed below. 29

5 10 1.5 ⇥ TABLE VI. Values for the mean shear of the METACALIBRATION bin 1 and IM3SHAPE shape catalogs for each redshift bin. bin 2 1.0 bin 3 ETACALIBRATION IM SHAPE bin 4 M 3 Redshift bin ei 104 ei 104 ei 104 ei 104 0.5 h 1i⇥ h 2i⇥ h 1i⇥ h 2i⇥ 1 0.84.3 2.12.7 ) 2 1.64.51.71.6 ✓ (

p 0.0 3 3.8 2.04.50.2 e ⇠ 4 8.9 1.75.96.6 0.5

1.0 straints. The contribution to ⇠+ of some additive shear bias cj = ej that is constant within a tomographic bin is 1.5 i h i i 100 101 102 ⇠ij = ⇠ij +( ei ej + ei ej ). (A6) ✓ (arcmin) +,pred +,true h 1ih 1i h 2ih 2i

6 Artificially introducing this bias due to mean shear to an un- 10 ⇥ biased simulated data vector results in non-negligible biases bin 1 4 to cosmological parameters, and so we subtract the impact of bin 2 the mean shear from the ellipticity before measuring ⇠ , as bin 3 ± bin 4 described in Sec. IV. This mean shear is different for both 2 shape catalogs, and its origin remains unclear. To verify that

) this is a sufficient correction for any large-scale problems, we ✓ (

q also test that our inferred cosmological results are unchanged e ⇠ 0 when we add a constant factor to the covariance that is equal ⇥ ij ✓ to the impact of the mean shear on each ⇠+ . This factor is 2 much larger in many blocks than the impact of cosmic vari- ance, and thus removes most of the large-scale information from the signal for this test. 4 100 101 102 Appendix B: Testing for residual systematics in the cosmic shear ✓ (arcmin) signal

Even after targeted tests for systematic bias in shape and photo-z measurements, there can persist biases that have not FIG. 17. The measured correlation functions ⇠ep (Eq. (A3)) and been identified that may impact the cosmic shear signal. In ⇠eq (Eq. (A4)). The resulting model from the best-fit ↵ and from Eqs. (A3) & (A4) for ↵ and is shown as a solid line for each the DES SV cosmic shear analysis [48], we tested for this tomographic bin. The best-fit ↵ is consistent with zero, while the by splitting the shear catalog into halves along a large num- values of METACALIBRATION used are shown in Table V. ber of either catalog or survey properties that could correlate with shape or photo-z systematics. These tests assume that the quantity being used to split the shear catalog either should not be correlated with gravitational shear or is simply correlated 2. Mean Residual Shear with redshift, and thus any non-cosmological signal induced by the selection can be corrected to first order by reweighting While we conclude that the propagation of PSF model er- the redshift distribution. This allows us to identify potential rors into the shape measurement does not produce a signifi- residual systematics that may either impact our cosmological cant bias on cosmological constraints with cosmic shear, we constraints or indicate limitations in analyses that are sensitive are left with an unidentified mean shear that is too large in to subsets of the shape or photo-z catalogs. some tomographic bins to be consistent with shape noise and cosmic variance. The values of the mean shear are listed in Ta- j 1. Methodology ble VI. Three of the eight ei for METACALIBRATION have a magnitude greater than 2.5h i times the predicted shear vari- ance calculated using the lognormal mock catalogs described The method of implementing these tests is similar to that in Sec. III C. We cannot rule this out as being sourced by some described in Section VI.C of [48]. For each shape catalog, unidentified additive bias or failure in our PSF bias model, so we select ten quantities that are most likely to be correlated we test what impact this could have on our cosmological con- with residual shear systematics. These are: signal-to-noise 30

METACALIBRATION

3.0 S/N PSF e1 PSF e2 T E(B V ) ⇠+ 2 2.5 ⇠ ND 1 2.0 1.5 1.0 0.5 0.0 3.0 Sky Brightness PSF FWHM Airmass r-band Lim. Mag. Color (r i) 2 2.5 ND 1 2.0 1.5 1.0 0.5 0.0

CumulativeTomo. binTomo. 1 binTomo. 2 binTomo. 3 binCumulative 4 Tomo. binTomo. 1 binTomo. 2 binTomo. 3 binCumulative 4 Tomo. binTomo. 1 binTomo. 2 binTomo. 3 binCumulative 4 Tomo. binTomo. 1 binTomo. 2 binTomo. 3 binCumulative 4 Tomo. binTomo. 1 binTomo. 2 binTomo. 3 bin 4

IM3SHAPE

3.0 S/N PSF e1 PSF e2 Rgp/Rpp E(B V ) 2 2.5 ND 1 2.0 1.5 1.0 0.5 0.0 3.0 Sky Brightness PSF FWHM Airmass r-band Lim. Mag. Color (r i) 2 2.5 ND 1 2.0 1.5 1.0 0.5 0.0

CumulativeTomo. binTomo. 1 binTomo. 2 binTomo. 3 binCumulative 4 Tomo. binTomo. 1 binTomo. 2 binTomo. 3 binCumulative 4 Tomo. binTomo. 1 binTomo. 2 binTomo. 3 binCumulative 4 Tomo. binTomo. 1 binTomo. 2 binTomo. 3 binCumulative 4 Tomo. binTomo. 1 binTomo. 2 binTomo. 3 bin 4

FIG. 18. The significance of differences in the cosmic shear signal, ⇠ , between subsets of the shape catalogs split by ten quantities that are ± most likely to be correlated with residual shear systematics. These are: signal-to-noise (S/N), PSF ellipticity (PSF e1, e2), galaxy size (T -METACALIBRATION, Rgp/Rpp - IM3SHAPE), r i color, dust (E(B V )), sky brightness, PSF FWHM, airmass, and r-band limiting magnitude. We report the normalized 2 of⇠ with the null hypothesis in each case for both ⇠ . The most interesting differences ± ± occur in the highest redshift bin, and when taking selection biases in the photo-z catalog into account, are not concerning as a contaminant to the cosmic shear signal. In all cases, the cumulative impact on the data vector combining all four auto-correlations is consistent with there being no significant residual systematic effects in the fiducial cosmic shear analysis.

(S/N), PSF ellipticity (PSF e1, e2), galaxy size (T -META- catalogs with improved DES Y1 data. CALIBRATION, Rgp/Rpp - IM3SHAPE), r i color, dust ex- For each quantity and shape catalog, we split the objects tinction (E(B V )), sky brightness, PSF FWHM, airmass, used in the cosmic shear measurements into two halves, each and r-band limiting magnitude. The first five are intrinsic with the same effective shape weight. We then correct any properties of each galaxy image measured by our shape and resulting differences in the redshift distribution of each half PSF measurement pipelines, while the last five are taken to be of the catalogs by constructing weights for each galaxy that the mean value of each property across exposures at a given match each redshift distribution to that of the full catalog. For position on the sky in HEALPIX cells of Nside 4096. We an example of the impact of the redshift reweighting proce- exclude several properties tested in the DES SV analysis that dure, see Fig. 8 of [48]. We also recalculate RS for META- are highly degenerate with the remaining quantities, and also CALIBRATION in each half, which should correct for any do not include surface brightness, since we have identified no shear selection effects. Finally, we measure ⇠ with the same need to make an explicit surface brightness cut in the shape binning as used in the fiducial measurement for± each half us- 31 ing a weight that is the product of this redshift weight with any shape weight. Unlike in [48], we use a single zmc drawn from the photo-z pdf of each galaxy, rather than construct the redshift weights using the full pdfs. For Gaussian pdfs, this 0.2 0.3 0.4 0.5 0.6 1 2 3 4 3.5 4.0 4.5 5.0 5.5 6.0 60 65 70 75 80 85 9 2 procedure reproduces the correct peak of the original redshift m As 10 b 10 H0 distribution, but the reweighted distribution remains skewed. ⇥ ⇥ The effect of skewness in the reweighted redshift distribution should be subdominant to selection effects in the photo-z bias correction, however, and we ignore it. 0.90 0.93 0.96 0.99 1.02 2 3 4 5 6 7 8 2 1 0 1 2 3 4 2 0 2 4 We construct a covariance for these tests using 250 of the 2 3 ns h 10 AIA IA simulated shear maps described in Sec. III C. We build a mock ⇥ catalog by sampling from the simulated shear maps an appro- priate number density of objects and shape noise based on a mapping of the weights by position on the sky for each cata- log subset to each simulated map. This captures the impact of 6 4 2 0 2 4 6 6 4 2 0 2 4 6 6 4 2 0 2 4 6 6 4 2 0 2 4 6 m 102 m 102 m 102 m 102 added effective shape noise in the redshift reweighting proce- 1 ⇥ 2 ⇥ 3 ⇥ 4 ⇥ dure, but does not capture any correlation of the selection or weights with redshift in each tomographic bin. We then cal- culate ⇠ for each quantity and its covariance. This requires approximately± 200,000 measurements of ⇠ in total. ± 6 4 2 0 2 4 6 6 4 2 0 2 4 6 6 4 2 0 2 4 6 6 4 2 0 2 4 6 z 102 z 102 z 102 z 102 A failure of these tests may not indicate a bias in any cos- 1 ⇥ 2 ⇥ 3 ⇥ 4 ⇥ mological parameter constraints, but rather a failure due to se- lection effects not accounted for in catalog creation that does FIG. 19. The full 1D posteriors on all 16 parameters in our ⇤CDM not impact the full sample. In the case of METACALIBRA- model space. Shaded regions show the approximate 68% confidence TION, we can explicitly correct for selection biases caused intervals. The smooth curves show the Gaussian priors on systematic by splitting the catalog by any quantity measured during the parameters. metacalibration procedure. However, we have measured, us- ing the resampled COSMOS catalog, significant selection bi- ases in the photo-z distribution when splitting the catalog by, of subsamples. for example, S/N and size. Finally, it is worth noting that a non-null detection in a test does not necessarily indicate a sys- tematic that translates into a significant bias in cosmological Appendix C: Full 1D Parameter Constraints parameter constraints, even though we may have the statistical power to detect it in this test, since we are taking the difference We show 1D posteriors for the full ⇤CDM parameter space of two signals that share cosmic variance. in Fig. 19, including cosmological, astrophysical, and sys- tematic parameters. We find no significant constraints beyond 2 the prior imposed on the parameters ⌦b, H0, ns, and ⌦⌫ h . 2. Results The priors for all 16 parameters are listed in Table II and a quantitative comparison of priors and posteriors of the non- 2 cosmological parameters in Table IV. We report the resulting significance ( /(ND 1)) for each quantity in Fig. 18 for each tomographic auto-correlation (Bin. 1-4) and their combination (Cumulative), where ND is the length of the data vector. We show only the impact on auto-correlations, since any correlated systematic should be strongest there. The mean 2/(N 1) falls close to 1, as D generally expected: 0.91 for METACALIBRATION and 1.1 for IM3SHAPE. There is no significant indication of bias in ei- ther shape catalog for the ‘cumulative’ points that combine the impact of all four tomographic bins. However, there does seem to be a higher significance for the ⇠ in the highest tomographic bin in splits along several quantities± for both cat- alogs. For METACALIBRATION, the highest significance de- tections of ⇠ all occur with quantities that require signifi- cant redshift reweighting,± and we have confirmed that a suf- ficient component of this 2 is due to photo-z selection bias in splitting the samples, such that the apparent non-null result is not significant. We do not explicitly correct this in the fig- ures, because there is a large uncertainty on the relative bias