2013 results. xviii. gravitational lensing-infrared ... · pdf fileastronomy &...

25
Astronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation Planck Collaboration: P. A. R. Ade 84 , N. Aghanim 58 , C. Armitage-Caplan 89 , M. Arnaud 71 , M. Ashdown 68,6 , F. Atrio-Barandela 19 , J. Aumont 58 , C. Baccigalupi 83 , A. J. Banday 92,10 , R. B. Barreiro 65 , J. G. Bartlett 1,66 , S. Basak 1 , E. Battaner 93 , K. Benabed 59,91 , A. Benoˆ ıt 56 , A. Benoit-L´ evy 26,59,91 , J.-P. Bernard 10 , M. Bersanelli 35,49 , M. Bethermin 71 , P. Bielewicz 92,10,83 , J. Bobin 71 , J. J. Bock 66,11 , A. Bonaldi 67 , J. R. Bond 9 , J. Borrill 14,86 , F. R. Bouchet 59,91 , F. Boulanger 58 , M. Bridges 68,6,62 , M. Bucher 1 , C. Burigana 48,33 , R. C. Butler 48 , J.-F. Cardoso 72,1,59 , A. Catalano 73,70 , A. Challinor 62,68,12 , A. Chamballu 71,16,58 , L.-Y Chiang 61 , H. C. Chiang 28,7 , P. R. Christensen 79,38 , S. Church 88 , D. L. Clements 54 , S. Colombi 59,91 , L. P. L. Colombo 25,66 , F. Couchot 69 , A. Coulais 70 , B. P. Crill 66,80 , A. Curto 6,65 , F. Cuttaia 48 , L. Danese 83 , R. D. Davies 67 , P. de Bernardis 34 , A. de Rosa 48 , G. de Zotti 45,83 , J. Delabrouille 1 , J.-M. Delouis 59,91 , F.-X. D´ esert 52 , J. M. Diego 65 , H. Dole 58,57 , S. Donzelli 49 , O. Dor´ e 66,11? , M. Douspis 58 , X. Dupac 41 , G. Efstathiou 62 , T. A. Enßlin 76 , H. K. Eriksen 63 , F. Finelli 48,50 , O. Forni 92,10 , M. Frailis 47 , E. Franceschi 48 , S. Galeotta 47 , K. Ganga 1 , M. Giard 92,10 , G. Giardino 42 , Y. Giraud-H´ eraud 1 , J. Gonz´ alez-Nuevo 65,83 , K. M. G ´ orski 66,95 , S. Gratton 68,62 , A. Gregorio 36,47 , A. Gruppuso 48 , J. E. Gudmundsson 28 , F. K. Hansen 63 , D. Hanson 77,66,9 , D. Harrison 62,68 , S. Henrot-Versill´ e 69 , C. Hern´ andez-Monteagudo 13,76 , D. Herranz 65 , S. R. Hildebrandt 11 , E. Hivon 59,91 , M. Hobson 6 , W. A. Holmes 66 , A. Hornstrup 17 , W. Hovest 76 , K. M. Huenberger 94 , T. R. Jae 92,10 , A. H. Jae 54 , W. C. Jones 28 , M. Juvela 27 , E. Keih¨ anen 27 , R. Keskitalo 23,14 , T. S. Kisner 75 , R. Kneissl 40,8 , J. Knoche 76 , L. Knox 29 , M. Kunz 18,58,3 , H. Kurki-Suonio 27,44 , F. Lacasa 58 , G. Lagache 58 , A. L¨ ahteenm¨ aki 2,44 , J.-M. Lamarre 70 , A. Lasenby 6,68 , R. J. Laureijs 42 , C. R. Lawrence 66 , R. Leonardi 41 , J. Le ´ on-Tavares 43,2 , J. Lesgourgues 90,82 , M. Liguori 32 , P. B. Lilje 63 , M. Linden-Vørnle 17 , M. L ´ opez-Caniego 65 , P. M. Lubin 30 , J. F. Mac´ ıas-P´ erez 73 , B. Maei 67 , D. Maino 35,49 , N. Mandolesi 48,5,33 , M. Maris 47 , D. J. Marshall 71 , P. G. Martin 9 , E. Mart´ ınez-Gonz´ alez 65 , S. Masi 34 , S. Matarrese 32 , F. Matthai 76 , P. Mazzotta 37 , A. Melchiorri 34,51 , L. Mendes 41 , A. Mennella 35,49 , M. Migliaccio 62,68 , S. Mitra 53,66 , M.-A. Miville-Deschˆ enes 58,9 , A. Moneti 59 , L. Montier 92,10 , G. Morgante 48 , D. Mortlock 54 , D. Munshi 84 , P. Naselsky 79,38 , F. Nati 34 , P. Natoli 33,4,48 , C. B. Netterfield 21 , H. U. Nørgaard-Nielsen 17 , F. Noviello 67 , D. Novikov 54 , I. Novikov 79 , S. Osborne 88 , C. A. Oxborrow 17 , F. Paci 83 , L. Pagano 34,51 , F. Pajot 58 , D. Paoletti 48,50 , G. Patanchon 1 , O. Perdereau 69 , L. Perotto 73 , F. Perrotta 83 , F. Piacentini 34 , M. Piat 1 , E. Pierpaoli 25 , D. Pietrobon 66 , S. Plaszczynski 69 , E. Pointecouteau 92,10 , G. Polenta 4,46 , N. Ponthieu 58,52 , L. Popa 60 , T. Poutanen 44,27,2 , G. W. Pratt 71 , G. Pr´ ezeau 11,66 , S. Prunet 59,91 , J.-L. Puget 58 , J. P. Rachen 22,76 , R. Rebolo 64,15,39 , M. Reinecke 76 , M. Remazeilles 58,1 , C. Renault 73 , S. Ricciardi 48 , T. Riller 76 , I. Ristorcelli 92,10 , G. Rocha 66,11 , C. Rosset 1 , G. Roudier 1,70,66 , M. Rowan-Robinson 54 , B. Rusholme 55 , M. Sandri 48 , D. Santos 73 , G. Savini 81 , D. Scott 24 , M. D. Seiert 66,11 , P. Serra 58 , E. P. S. Shellard 12 , L. D. Spencer 84 , J.-L. Starck 71 , V. Stolyarov 6,68,87 , R. Stompor 1 , R. Sudiwala 84 , R. Sunyaev 76,85 , F. Sureau 71 , D. Sutton 62,68 , A.-S. Suur-Uski 27,44 , J.-F. Sygnet 59 , J. A. Tauber 42 , D. Tavagnacco 47,36 , L. Terenzi 48 , L. Toolatti 20,65 , M. Tomasi 49 , M. Tristram 69 , M. Tucci 18,69 , J. Tuovinen 78 , L. Valenziano 48 , J. Valiviita 44,27,63 , B. Van Tent 74 , P. Vielva 65 , F. Villa 48 , N. Vittorio 37 , L. A. Wade 66 , B. D. Wandelt 59,91,31 , S. D. M. White 76 , D. Yvon 16 , A. Zacchei 47 , and A. Zonca 30 (Aliations can be found after the references) ABSTRACT The multi-frequency capability of the Planck satellite provides information both on the integrated history of star formation (via the cosmic infrared background, or CIB) and on the distribution of dark matter (via the lensing eect on the cosmic microwave background, or CMB). The conjunction of these two unique probes allows us to measure directly the connection between dark and luminous matter in the high redshift (1 z 3) Universe. We use a three-point statistic optimized to detect the correlation between these two tracers. Following a thorough discussion of possible contaminants and a suite of consistency tests, using lens reconstructions at 100, 143 and 217 GHz and CIB measurements at 100–857 GHz, we report the first detection of the correlation between the CIB and CMB lensing. The well matched redshift distribution of these two signals leads to a detection significance with a peak value of 42 σ at 545 GHz and a correlation as high as 80 % across these two tracers. Our full set of multi- frequency measurements (both CIB auto- and CIB-lensing cross-spectra) are consistent with a simple halo-based model, with a characteristic mass scale for the halos hosting CIB sources of log 10 (M/M ) = 10.5 ± 0.6. Leveraging the frequency dependence of our signal, we isolate the high redshift contribution to the CIB, and constrain the star formation rate (SFR) density at z 1. We measure directly the SFR density with around 2 σ significance for three redshift bins between z = 1 and 7, thus opening a new window into the study of the formation of stars at early times. Key words. Gravitational lensing – Galaxies: star formation – cosmic background radiation – dark matter – large-scale structure of Universe 1. Introduction This paper, one of a set associated with the 2013 release of data from the Planck 1 mission (Planck Collaboration I 2013), ? Corresponding author: Olivier Dor´ e <olivier.p.dore@jpl. nasa.gov> 1 Planck (http://www.esa.int/Planck) is a project of the European Space Agency (ESA) with instruments provided by two sci- entific consortia funded by ESA member states (in particular the lead presents a first detection of a strong correlation between the in- frared background anisotropies and a lensing-derived projected mass map. The broad frequency coverage of the Planck satel- lite provides two important probes of the high redshift Universe. In the central frequency bands of Planck (70, 100, 143, and 217 GHz), cosmic microwave background (CMB) fluctuations countries France and Italy), with contributions from NASA (USA) and telescope reflectors provided by a collaboration between ESA and a sci- entific consortium led and funded by Denmark. 1 arXiv:1303.5078v1 [astro-ph.CO] 20 Mar 2013

Upload: buicong

Post on 26-Mar-2018

215 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Astronomy & Astrophysics manuscript no. paper˙cib˙lensing c© ESO 2013March 22, 2013

Planck 2013 results. XVIII.Gravitational lensing-infrared background correlation

Planck Collaboration: P. A. R. Ade84, N. Aghanim58, C. Armitage-Caplan89, M. Arnaud71, M. Ashdown68,6, F. Atrio-Barandela19, J. Aumont58,C. Baccigalupi83, A. J. Banday92,10, R. B. Barreiro65, J. G. Bartlett1,66, S. Basak1, E. Battaner93, K. Benabed59,91, A. Benoıt56,

A. Benoit-Levy26,59,91, J.-P. Bernard10, M. Bersanelli35,49, M. Bethermin71, P. Bielewicz92,10,83, J. Bobin71, J. J. Bock66,11, A. Bonaldi67,J. R. Bond9, J. Borrill14,86, F. R. Bouchet59,91, F. Boulanger58, M. Bridges68,6,62, M. Bucher1, C. Burigana48,33, R. C. Butler48, J.-F. Cardoso72,1,59,A. Catalano73,70, A. Challinor62,68,12, A. Chamballu71,16,58, L.-Y Chiang61, H. C. Chiang28,7, P. R. Christensen79,38, S. Church88, D. L. Clements54,

S. Colombi59,91, L. P. L. Colombo25,66, F. Couchot69, A. Coulais70, B. P. Crill66,80, A. Curto6,65, F. Cuttaia48, L. Danese83, R. D. Davies67, P. deBernardis34, A. de Rosa48, G. de Zotti45,83, J. Delabrouille1, J.-M. Delouis59,91, F.-X. Desert52, J. M. Diego65, H. Dole58,57, S. Donzelli49,

O. Dore66,11?, M. Douspis58, X. Dupac41, G. Efstathiou62, T. A. Enßlin76, H. K. Eriksen63, F. Finelli48,50, O. Forni92,10, M. Frailis47,E. Franceschi48, S. Galeotta47, K. Ganga1, M. Giard92,10, G. Giardino42, Y. Giraud-Heraud1, J. Gonzalez-Nuevo65,83, K. M. Gorski66,95,

S. Gratton68,62, A. Gregorio36,47, A. Gruppuso48, J. E. Gudmundsson28, F. K. Hansen63, D. Hanson77,66,9, D. Harrison62,68, S. Henrot-Versille69,C. Hernandez-Monteagudo13,76, D. Herranz65, S. R. Hildebrandt11, E. Hivon59,91, M. Hobson6, W. A. Holmes66, A. Hornstrup17, W. Hovest76,K. M. Huffenberger94, T. R. Jaffe92,10, A. H. Jaffe54, W. C. Jones28, M. Juvela27, E. Keihanen27, R. Keskitalo23,14, T. S. Kisner75, R. Kneissl40,8,J. Knoche76, L. Knox29, M. Kunz18,58,3, H. Kurki-Suonio27,44, F. Lacasa58, G. Lagache58, A. Lahteenmaki2,44, J.-M. Lamarre70, A. Lasenby6,68,

R. J. Laureijs42, C. R. Lawrence66, R. Leonardi41, J. Leon-Tavares43,2, J. Lesgourgues90,82, M. Liguori32, P. B. Lilje63, M. Linden-Vørnle17,M. Lopez-Caniego65, P. M. Lubin30, J. F. Macıas-Perez73, B. Maffei67, D. Maino35,49, N. Mandolesi48,5,33, M. Maris47, D. J. Marshall71,

P. G. Martin9, E. Martınez-Gonzalez65, S. Masi34, S. Matarrese32, F. Matthai76, P. Mazzotta37, A. Melchiorri34,51, L. Mendes41, A. Mennella35,49,M. Migliaccio62,68, S. Mitra53,66, M.-A. Miville-Deschenes58,9, A. Moneti59, L. Montier92,10, G. Morgante48, D. Mortlock54, D. Munshi84,

P. Naselsky79,38, F. Nati34, P. Natoli33,4,48, C. B. Netterfield21, H. U. Nørgaard-Nielsen17, F. Noviello67, D. Novikov54, I. Novikov79, S. Osborne88,C. A. Oxborrow17, F. Paci83, L. Pagano34,51, F. Pajot58, D. Paoletti48,50, G. Patanchon1, O. Perdereau69, L. Perotto73, F. Perrotta83, F. Piacentini34,M. Piat1, E. Pierpaoli25, D. Pietrobon66, S. Plaszczynski69, E. Pointecouteau92,10, G. Polenta4,46, N. Ponthieu58,52, L. Popa60, T. Poutanen44,27,2,G. W. Pratt71, G. Prezeau11,66, S. Prunet59,91, J.-L. Puget58, J. P. Rachen22,76, R. Rebolo64,15,39, M. Reinecke76, M. Remazeilles58,1, C. Renault73,S. Ricciardi48, T. Riller76, I. Ristorcelli92,10, G. Rocha66,11, C. Rosset1, G. Roudier1,70,66, M. Rowan-Robinson54, B. Rusholme55, M. Sandri48,

D. Santos73, G. Savini81, D. Scott24, M. D. Seiffert66,11, P. Serra58, E. P. S. Shellard12, L. D. Spencer84, J.-L. Starck71, V. Stolyarov6,68,87,R. Stompor1, R. Sudiwala84, R. Sunyaev76,85, F. Sureau71, D. Sutton62,68, A.-S. Suur-Uski27,44, J.-F. Sygnet59, J. A. Tauber42, D. Tavagnacco47,36,

L. Terenzi48, L. Toffolatti20,65, M. Tomasi49, M. Tristram69, M. Tucci18,69, J. Tuovinen78, L. Valenziano48, J. Valiviita44,27,63, B. Van Tent74,P. Vielva65, F. Villa48, N. Vittorio37, L. A. Wade66, B. D. Wandelt59,91,31, S. D. M. White76, D. Yvon16, A. Zacchei47, and A. Zonca30

(Affiliations can be found after the references)

ABSTRACT

The multi-frequency capability of the Planck satellite provides information both on the integrated history of star formation (via the cosmic infraredbackground, or CIB) and on the distribution of dark matter (via the lensing effect on the cosmic microwave background, or CMB). The conjunctionof these two unique probes allows us to measure directly the connection between dark and luminous matter in the high redshift (1 ≤ z ≤ 3)Universe. We use a three-point statistic optimized to detect the correlation between these two tracers. Following a thorough discussion of possiblecontaminants and a suite of consistency tests, using lens reconstructions at 100, 143 and 217 GHz and CIB measurements at 100–857 GHz, wereport the first detection of the correlation between the CIB and CMB lensing. The well matched redshift distribution of these two signals leadsto a detection significance with a peak value of 42σ at 545 GHz and a correlation as high as 80 % across these two tracers. Our full set of multi-frequency measurements (both CIB auto- and CIB-lensing cross-spectra) are consistent with a simple halo-based model, with a characteristic massscale for the halos hosting CIB sources of log10 (M/M) = 10.5 ± 0.6. Leveraging the frequency dependence of our signal, we isolate the highredshift contribution to the CIB, and constrain the star formation rate (SFR) density at z ≥ 1. We measure directly the SFR density with around2σ significance for three redshift bins between z = 1 and 7, thus opening a new window into the study of the formation of stars at early times.

Key words. Gravitational lensing – Galaxies: star formation – cosmic background radiation – dark matter – large-scale structure of Universe

1. Introduction

This paper, one of a set associated with the 2013 release ofdata from the Planck1 mission (Planck Collaboration I 2013),

? Corresponding author: Olivier Dore <[email protected]>

1 Planck (http://www.esa.int/Planck) is a project of theEuropean Space Agency (ESA) with instruments provided by two sci-entific consortia funded by ESA member states (in particular the lead

presents a first detection of a strong correlation between the in-frared background anisotropies and a lensing-derived projectedmass map. The broad frequency coverage of the Planck satel-lite provides two important probes of the high redshift Universe.In the central frequency bands of Planck (70, 100, 143, and217 GHz), cosmic microwave background (CMB) fluctuations

countries France and Italy), with contributions from NASA (USA) andtelescope reflectors provided by a collaboration between ESA and a sci-entific consortium led and funded by Denmark.

1

arX

iv:1

303.

5078

v1 [

astr

o-ph

.CO

] 2

0 M

ar 2

013

Page 2: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

dominate over most of the sky. Gravitational lensing by large-scale structure produces small shear and magnification effects inthe observed fluctuations, which can be exploited to reconstructan integrated measure of the gravitational potential along the lineof sight Okamoto & Hu (2003). This “CMB lensing potential”is sourced primarily by dark matter halos located at 1 . z . 3,halfway between ourselves and the last scattering surface (seeBlandford & Jaroszynski 1981; Blanchard & Schneider 1987, orLewis & Challinor 2006 for a review). In the upper frequencybands (353, 545, and 857 GHz), the dominant extragalactic sig-nal is not the CMB, but the cosmic infrared background (CIB),composed of redshifted thermal radiation from UV-heated dust,enshrouding young stars. The CIB contains much of the energyfrom processes involved in structure formation. According tocurrent models, the dusty star-forming galaxies (DSFGs), whichform the CIB have a redshift distribution peaked between z ∼ 1and z ∼ 2, and tend to live in 1011–1013M dark matter halos(see, e.g., Bethermin et al. 2012, and references therein).

As first pointed out by Song et al. (2003), the halo mass andredshift dependence of the CMB lensing potential and the CIBfluctuations are well matched, and as such a significant correla-tion between the two is expected. This point is illustrated quan-titatively in Fig. 1, where we plot estimates for the redshift- andmass- kernels of the two tracers. In this paper we report on thefirst detection of this correlation.

Measurements of both CMB lensing and CIB fluctuationsare currently undergoing a period of rapid development. Whilethe CIB mean was first detected using the FIRAS and DIRBEinstruments aboard COBE (Puget et al. 1996; Fixsen et al. 1998;Hauser et al. 1998), CIB fluctuations were later detected bythe Spitzer Space Telescope (Lagache et al. 2007) and by theBLAST balloon experiment (Viero et al. 2009) and the HerschelSpace Observatory (Amblard et al. 2011; Viero et al. 2012),as well as the new generation of CMB experiments, includ-ing Planck, which have extended these measurements to longerwavelengths (Hall et al. 2010; Dunkley et al. 2011; PlanckCollaboration XVIII 2011; Reichardt et al. 2012). The Planckearly results paper: Planck Collaboration XVIII (2011) (hence-forth referred to as PER) presented measurements of the angu-lar power spectra of CIB anisotropies from arc-minute to degreescales at 217, 353, 545, and 857 GHz, establishing Planck as apotent probe of the clustering of the CIB, both in the linear andnon-linear regimes. A substantial extension of PER is presentedin a companion paper to this work (Planck Collaboration 2013,henceforth referred to as PIR).

The CMB lensing potential, on the other hand, which wasfirst detected statistically through cross-correlation with galaxysurveys (Smith et al. 2007; Hirata et al. 2008, and more recentlyBleem et al. 2012; Sherwin et al. 2012), has now been observeddirectly in CMB maps by the Atacama Cosmology Telescopeand the South Pole Telescope (Das et al. 2011; van Engelen et al.2012).

Planck’s frequency coverage, sensitivity and survey area, al-low high signal-to-noise measurements of both the CIB and theCMB lensing potential. Accompanying the release of this pa-per, Planck Collaboration XVII (2013) reports the first measure-ment and characterisation of the CMB lensing potential with thePlanck data, which has several times more statistical power thanprevious measurements, over a large fraction (approximately70% of the sky). We will use this measurement of the lensingpotential in cross-correlation with measurements of the CIB inthe PlanckHFI bands to make the first detection of the lensing-infrared background correlation. In addition to our measure-ment, we discuss the implications for models of the CIB fluc-

Fig. 1. Redshift- and mass- integrand for the CIB and CMB lens-ing potential power spectra at ` = 500, calculated using theCIB halo model of Planck Collaboration XVIII (2011), evalu-ated at 217 GHz. The good match between the redshift and halomass distributions leads to an expected correlation up to 80 %.The sharper features in the CIB kernel are artefacts from theBethermin et al. (2012) model. We note that the low mass, highz behavior of our calculation is limited by the accuracy of themass function we use (Tinker & Wetzel 2010). All of our massintegrals use Mmin = 105 M.

tuations. The outline of this paper is as follows. In Sect. 2 wedescribe the data we will use, followed by a description of ourpipeline for correlating the CIB and lensing signals in Sect. 3.Our main result is presented in Sect. 4, with a description of ourerror budget, consistency tests and an array of systematic tests inSect. 5. We discuss the implications of the measured correlationfor CIB modelling in Sect. 6.

2. Data sets

2.1. Planck maps

Planck (Tauber et al. 2010; Planck Collaboration I 2011) is thethird generation space mission to measure the anisotropy of theCMB. It observes the sky with high sensitivity in nine frequencybands covering 30–857 GHz at an angular resolution from 31′ to5′. The Low Frequency Instrument (LFI; Mandolesi et al. 2010;Bersanelli et al. 2010; Mennella et al. 2011) covers the 30, 44,and 70 GHz bands with radiometers that incorporate amplifierscooled to 20 K. The High Frequency Instrument (HFI; Lamarreet al. 2010; Planck HFI Core Team 2011a) covers the 100, 143,217, 353, 545, and 857 GHz bands with bolometers cooled to0.1 K. Polarization is measured in all but the highest two bands

2

Page 3: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

(Leahy et al. 2010; Rosset et al. 2010). A combination of radia-tive cooling and three mechanical coolers produces the temper-atures needed for the detectors and optics (Planck CollaborationII 2011). Two data processing centres (DPCs) check and cali-brate the data and make maps of the sky (Planck HFI Core Team2011b; Zacchei et al. 2011). Planck’s sensitivity, angular reso-lution, and frequency coverage make it a powerful instrumentfor Galactic and extragalactic astrophysics as well as cosmol-ogy. Early astrophysics results are given in Planck CollaborationVIII–XXVI 2011, based on data taken between 13 August 2009and 7 June 2010. Intermediate astrophysics results are now be-ing presented in a series of papers based on data taken between13 August 2009 and 27 November 2010. This paper uses datacorresponding to the second Planck data release, with data ac-quired in the period up to 27 November 2010 and undergoingimproved processing.

We use the Planck HFI temperature maps at all six frequen-cies, i.e., 100, 143, 217, 353, 545, and 857 GHz. The maps ateach frequency were created using almost three full-sky sur-veys. Here we give an overview of the HFI map-making pro-cess with additional details given in Planck HFI Core Team(2011b); Planck Collaboration VI (2013). The data are organizedas time-ordered information, hereafter TOI. The attitude of thesatellite as a function of time is provided by two star trackerson the spacecraft. The pointing for each bolometer is computedby combining the attitude with the location of the bolometer inthe focal plane, as determined by planet observations. The rawbolometer TOI for each channel is first processed to producecleaned timelines and to set flags that mark bad data (for ex-ample data immediately following a cosmic ray strike on the de-tector). This TOI processing includes: (1) signal demodulationand filtering; (2) deglitching, which flags the strong part of anyglitch and subtracts the tails; (3) conversion from instrumentalunits (volts) to physical units (watts of absorbed power, after acorrection for the weak non-linearity of the response); (4) de-correlation of thermal stage fluctuations; (5) removal of the sys-tematic effects induced by 4 K cooler mechanical vibrations; and(6) deconvolution of the bolometer time response. Focal planereconstruction and beam shape estimation is made using obser-vations of Mars. The simplest description of the beams, an el-liptical Gaussian, leads to full-width at half-maximum (FWHM)values of 9.65, 7.25, 4.99, 4.82, 4.68 and 4.33 ′as given in Table4 of Planck Collaboration VI (2013). As explained in this paper,the inter-calibration accuracy between channels is better than theabsolute calibration. This leads us to adopt conservative abso-lute calibration uncertainties of 0.64, 0.53, 0.69, 2.53, 10., 10. %at 100, 143 217, 353, 545 and 857 GHz respectively. We con-vert between emissivities given in MJy sr−1(with the photomet-ric convention νIν = constant) and temperatures in µK, usingthe measured bandpass filters (see PER and PIR for details).

For the sake of consistency testing (presented in particular inSect. 5), we will sometimes use temperature maps where only afraction of the TOI is used to generate the sky map. In particular,throughout this paper we use the terminology “half-ring” (HR)maps to refer to maps made using the first and second half of thestable pointing period, “survey” for individual full-sky surveymaps (note that the third survey is incomplete for the particu-lar data release used in the intermediate papers), and “detset”for maps made using two independent sets of detectors per fre-quency (for details see Planck HFI Core Team 2011b).

We create three masks to exclude regions with brightGalactic and extragalactic foreground emission. The first maskaccounts for diffuse Galactic emission as observed in the Planckdata. To allow us to test for the effects of residual Galactic

Fig. 2. Combined Galactic, point-source and H i mask with skyfractions 16, 30 and 43 %.

emission on our results we create three different versions ofthis mask, each with a different masked area, such that 20, 40or 60 % of the sky is unmasked. Each version of this maskis created directly from the Planck 353 GHz map, from whichwe remove the CMB using the 143 GHz channel as a CMBtemplate before smoothing by a Gaussian with FWHM of 5.The map is then thresholded such that the mask has the re-quired sky fraction. Although the Galactic emission is strongerat 857 GHz, we expect the 353 GHz mask to better trace dustemission at the lower frequencies we use. The mask therefore ac-counts for Galactic dust and Galactic CO emission as explainedin Planck Collaboration XII (2013). We will not worry aboutsynchrotron emission, which is important at low frequencies,since its contribution at 100 GHz and at high Galactic latitudesis small, and, as with the dust component, will be uncorrelatedwith the lensing potential. The second mask covers bright pointsources. This mask is created using algorithms tailored to de-tect point sources in the Planck data and is optimized for eachfrequency, as detailed in Planck Collaboration VII (2011) andPlanck Collaboration (2011). The third mask is designed to re-move extended high-latitude Galactic dust emission (“cirrus”),as traced by external H i data, as we will describe in Sect. 2.2.1.While the first two masks are described in Planck CollaborationXII (2013), the latter is specific to our cross-correlation analy-sis, as it provides a method to reduce the large-scale noise in ourmeasurement, and the 3-point nature of our estimate ensures thatit will not introduce a bias (although we test for this in Sect. 5).Ultimately, when we combine the three masks we obtain an ef-fective sky fraction of 16, 30 and 43 % for the 20, 40 and 60 %Galactic masks, respectively.

2.2. External data sets

2.2.1. H i maps

We use measurements of 21-cm emission from Galactic neutralhydrogen (H i) as a cirrus monitor. Outside of our Galactic andpoint source masks we use the H i data to construct a templateof the dust emission in regions where the H i column densityis low (less than NHI ≤ 2 × 1020 cm−2), and we mask regionswhere it is high, since in these regions the H i and dust emis-sion are not well correlated (Boulanger et al. 1996; Boulanger &Perault 1988, PER). The masking procedure that we use is de-scribed in detail in Planck Collaboration XXIV (2011). It con-sists of subtracting the H i dust template from the Planck tem-perature map at 857 GHz and calculating the skewness of theresiduals in 5 deg2 regions. If the skewness is larger than a givenvalue then the region is masked. This is an improvement over

3

Page 4: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

the usual cut-off in H i column density. We use the latest releasefrom the Leiden/Argentina/Bonn (LAB) survey (Kalberla et al.2005), which consists of the Leiden/Dwingeloo Survey (LDS)(Hartmann & Burton 1997) north of −30 declination, com-bined with the Instituto Argentino de Radioastronomia Survey(Arnal et al. 2000; Bajaja et al. 2005) south of −25 declina-tion. The angular resolution of the combined map is approx-imately 0.6 FWHM. The LAB Survey is the most sensitiveMilky Way H i survey to date, with the greatest coverage bothspatially and kinematically. We make use of projections of theLAB maps onto Nside = 512 HEALPix maps performed by Land& Slosar (2007) and made available through the LAMBDA web-site1. The local standard of rest velocity coverage spans the in-terval −450 km s−1 to +400 km s−1, at a resolution of 1.3 km s−1,with an rms brightness-temperature noise of 0.07–0.09 K, andwith additional errors due to defects in the correction for strayradiation that are less than 20–40 mK for most of the data.

2.2.2. IRIS/IRAS maps

As a consistency test we will use an additional tracer of the CIBthat derives from re-processed IRAS maps at 60 and 100 µm.This new generation of IRAS maps, known as IRIS, benefits fromimproved zodiacal light subtraction, a calibration and zero levelcompatible with DIRBE, and an improved de-striping procedure(Miville-Deschenes & Lagache 2005). IRAS made two full-skymaps (HCON-1 and HCON-2), as well as a final map that cov-ers 75 % of the sky (HCON-3). The three maps had identicalprocessing that included deglitching, checking of the zero-levelstability, visual examination for glitches and artifacts, and zodia-cal light removal. The three HCONs were then co-added, takinginto account the inhomogeneous sky coverage maps, to gener-ate the average map (HCON-0). Note that the Finkbeiner et al.(1999) maps are also constructed from the IRAS 100 µm data,and as such we will not investigate their cross-correlation prop-erties since the IRIS map contains the same information. Forsimplicity we will assume that the effective IRIS beam is uni-form across the sky and described by a Gaussian with FHWM of4.3′.

3. Cross-correlation formalism and implementation

We now describe our statistical formalism and its implementa-tion, with additional technical details given in the appendices.Our analysis consists of cross-correlating a full-sky reconstruc-tion of the CMB lensing potential with a temperature map.

3.1. Reconstructing the CMB lensing potential

The CMB is lensed by the gravitational potential of all matteralong the photon trajectory from the last scattering surface to us.The lensed CMB is a remapping of the unlensed CMB with thelensed temperature equal to Θ(n ) = Θ(n + ∇φ), where Θ(n )is the unlensed CMB temperature and φ is the lensing poten-tial. We use the methodology described in Planck CollaborationXVII (2013) to obtain estimates φLM of the lensing potentialin harmonic space, using the standard Okamoto & Hu (2003)quadratic estimator.

Complete details on the lens reconstruction procedure, whichwe use are given in Planck Collaboration XVII (2013), althoughwe review it briefly in point form here. Our estimates of φ areobtained by the following set of steps:

1 http://lambda.gsfc.nasa.gov/product/foreground/

1. Inverse variance filter the CMB map.2. Use the filtered CMB map as the input to a quadratic lensing

estimator, which is designed to extract the off-diagonal con-tributions to the CMB covariance matrix induced by lensing.

3. Subtract a “mean-field bias”, which corrects for known non-lensing contributions to the covariance matrix, including in-strumental noise inhomogeneity, beam asymmetry, and theGalaxy+point source mask.

The output from this pipeline is an estimate of the lensing po-tential in harmonic space φLM and an associated noise powerspectrum Nφφ

L , which we use to weight our cross-correlation es-timates. We also produce a set of simulated lens reconstruction,which we use to establish our statistical error bars.

Our nominal lens reconstructions use the 143 GHz channel,however there is almost equivalent power to measure lensing us-ing the 217 GHz channel. Combining both channels would re-duce the noise power spectrum of our lens reconstruction by ap-proximately 25 %, compared with using either individually (theimprovement is significantly less than 50 % because a signifi-cant portion of the lens reconstruction noise is due to the finitenumber of CMB modes, which we are able to observe, and iscorrelated between the two channels). We choose to focus on143 GHz here because it is significantly less susceptible to CIBcontamination. We will use lens reconstructions based on the100 and 217 GHz data for consistency tests.

3.2. Decreasing the foreground noise

An important source of noise (but, as we will explain below,not bias) in our cross-correlation measurement is Galactic fore-ground emission. Dust emission is the dominant Galactic com-ponent at HFI frequencies above 217 GHz (see Sect. 5.1 for aquantitative discussion). In order to reduce the Galactic dustemission we create a dust template and subtract it from the tem-perature maps described in Sect. 2.1. At 100 and 143 GHz theCMB signal is significantly brighter than the dust emission out-side the Galactic mask. We therefore do not create and subtracta dust template at these frequencies. Note that while we coulduse other frequency maps to trace the CMB and remove it, toquantify the non-negligible amount of CIB that would be re-moved this way is not easy given the uncertainties in the cross-frequency CIB correlation structure.

We rely on the well documented (but complex) correlationbetween Galactic H i and dust (e.g., Boulanger & Perault 1988;Boulanger et al. 1996; Lagache et al. 1998, PER) to reduce thecontamination by subtracting the H i-correlated dust component.As was performed in PER, we split the H i map into two ve-locity components: a low-velocity local component (LC) typicalof high-latitude H i emission, and a component of intermediate-velocity clouds (IVC). We found that the inclusion of a high-velocity component makes a negligible difference to the dust-cleaned map. Unlike the dedicated high-resolution H i observa-tions used in PER and PIR that only partially cover the sky, herewe use the full-sky, low resolution LAB survey introduced inSect. 2.2.1 as our H i tracer. Although the resolution of this sur-vey is lower than the Planck resolution, it allows us to performdust cleaning on large scales, where our cross-correlation mea-surement has high signal-to-noise ratio. The emissivity of thedust varies across the sky, and so the correlation between thedust and H i emission is expected to vary. To account for thiswe divide the sky into regions where we assume that the dust-H i correlation is constant. For the sake of convenience, we useregions of approximate size 13 (52) deg2 defined by the Healpix

4

Page 5: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

pixels at resolution Nside = 16 (8) that are outside the Galacticmask. We test that our conclusions do not depend on this resolu-tion.

The details of our procedure is as follows. We subtractthe 143 GHz Planck temperature map from each of the 217–857 GHz temperature maps to remove the CMB signal (thisCMB subtraction is only done for the purposes of creating thedust template). We upgrade each of the Nside = 512 LAB mapscompiled in Land & Slosar (2007) to the Planck map resolu-tion of Nside = 2048. Within each region we then simultane-ously fit for the amplitude of each H i velocity component inthe CMB-subtracted maps, and use the two coefficients per re-gion to assemble a full-sky (minus the mask) dust template foreach of the 217–857 GHz channels. We smooth each templatewith a Gaussian of FWHM 10′ to remove the discontinuity atthe patch boundaries, and then subtract the template from theoriginal (CMB-unsubtracted) Planck maps.

We note that the accuracy of this procedure would be diffi-cult to evaluate for all possible uses of the map, i.e., whether itmight constitute a robust component separation method remainsto be demonstrated. However, in the case of our cross-correlationanalysis the dust-removal requirements are less severe, since thedust emission only contributes to our measurement as noise. Wewill describe later in Sect. 5.2 the effect on the cross-spectrumof removing this emission, and will place limits on the residualGalactic contamination in Sect. 5.3.5.

3.3. Measuring cross-correlations

To estimate the cross-correlation between the lensing potentialand a tracer t, we calculate

Ctφ`

=1

2` + 1

∑m

t`mφ∗`m. (1)

As the CIB has an approximately `−1 dependence and the lensingpotential has an `−2 dependence, we multiply the cross-spectraby `3, and then bin it in 15 linearly spaced bins between 100 <` < 2000. As we will discuss in Sect. 5, modes with ` < 100are not considered, due to possible lens mean-field systematiceffects, and modes with ` > 2000 are removed due to possibleextragalactic foreground contamination. We have tested that ourresults are robust to an increase or decrease in the number of`-bins.

We expect the error bars to be correlated across bins to someextent, due to pseudo-C` mixing induced by the mask, and be-tween frequencies, because the lens reconstruction noise is com-mon. In addition, any foregrounds that are present in multiplechannels will introduce correlated noise. The foreground maskwill also induce a coupling between different modes of the un-masked map. This extra coupling can be calculated explicitlyusing the mixing matrix formalism introduced in Hivon et al.(2002). Using this formalism and our best-fit models we haveevaluated the correlation between different bins of the cross-correlation signal for our nominal binning scheme. We find thatthe mask-induced correlation is less than 2 % across all bins atall frequencies. We will thus neglect it in our analysis. For thisreason, and based on the results we obtain from simulations,we do not attempt to “deconvolve” the mask from the cross-spectrum (see e.g., Hivon et al. 2002) and instead correct for thepower lost through masking the maps by a single sky fraction,fsky, ignoring the mode coupling.

As will be discussed later in Sect. 6.1, when we fit modelsto the cross-spectrum we will assume that the noise correlationbetween bins can be neglected and that the band-powers are flat.

3.4. Simulating cross-correlations

In order to validate our measurement pipeline and to confirm thatour estimate of the cross-spectrum is unbiased we create simu-lated maps of the lensed CMB and CIB that have the expectedstatistical properties.

Using the Planck only favored ΛCDM cosmology as de-scribed in Planck Collaboration XVI (2013) we generate a the-oretical prediction of the lensing potential spectrum using CAMB(Lewis et al. 2000), from which we generate 300 maps of φ thatare used to lens 300 CMB realizations using the approach de-scribed in Planck Collaboration XVII (2013). We then use thePER best-fit CIB model to generate CIB auto- and CIB-φ cross-spectra, from which we create CIB realizations that are correctlycorrelated with φ in each HFI band. The PER model that weuse describes the CIB clustering at HFI frequencies using a haloapproach, and simultaneously reproduces known number countand luminosity function measurements. At each frequency weadd a lensed CMB realization to each of the CIB realizationsand then smooth the maps using a symmetric beam with thesame FWHM as the beam described in Sect. 2.1. Once this setof realizations has been generated we apply the reconstructionprocedure described above to produce an estimate of the lens-ing potential map, and then calculate the cross-power spectrumusing our measurement pipeline.

These simulations will miss some complexities inherent inthe Planck mission. They do not take into account inhomoge-neous and correlated noise, and we do not simulate asymmetricbeam effects. In addition, we do not simulate any foregroundcomponents, and we instead take a different approach to deter-mine their contribution. While simplistic, we believe that oursimulations are good enough for the purposes of this particularmeasurement. In Sect. 5 we will discuss possible limitations, aswell as how we test for systematic effects that are not includedin the simulations.

We use the simulated maps to check that our pipeline cor-rectly recovers the cross-spectrum that was used to generate thesimulations. For the `-bins used in our analysis, we find that therecovered spectrum is unbiased (to within the precision achiev-able with 300 simulations), and with a noise level consistent withexpectations. The noise in the recovered spectrum is discussedin Sect. 5.1.

4. A strong signal using Planck HFI data

We now describe the result of applying our pipeline to our nom-inal data set, i.e., the lens reconstruction at 143 GHz and theforeground reduced Planck HFI temperature maps with a 40 %Galactic mask, which when combined with the point sourcemask and H i mask leaves 30.4 % of the sky unmasked. The re-sults are presented in Fig. 3. The error bars correspond to thenaive scatter measured within each bin. The thin black line corre-sponds to the expected CIB-lensing correlation predicted usingthe PER CIB model (the HOD parameters of the PER 217 GHzbest-fit model were used at 100 and 143 GHz since no CIB clus-tering measurement at these frequencies is available). As can beseen from these plots, the noise is strongly correlated across fre-quencies, especially at the lowest frequencies where the CMBdominates the error budget. A detailed analysis of the uncertain-ties and potential systematic errors attached to this result is pre-sented in Sect. 5.

As clearly visible in Fig. 3, a strong signal is detected. Toset a reference point and naively quantify its statistical signif-icance when taken at face value, we define a detection signifi-

5

Page 6: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

100 500 1000 1500 2000`

1.5

1.0

0.5

0.0

0.5

1.0

1.5`3C

`[µ

K.s

r]ν=100 GHz (x 100)

100 500 1000 1500 2000`

0.2

0.0

0.2

0.4

0.6

0.8

1.0 ν=143 GHz (x 100)

100 500 1000 1500 2000`

0.2

0.0

0.2

0.4

0.6

0.8

1.0

1.2 ν=217 GHz (x 100)

100 500 1000 1500 2000`

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

`3C

`[µ

K.s

r]

ν=353 GHz (x 100)

100 500 1000 1500 2000`

0

1

2

3

4 ν=545 GHz (x 10)

100 500 1000 1500 2000`

0.0

0.5

1.0

1.5

2.0 ν=857 GHz (x 0.1)

Fig. 3. Angular cross-spectra between the reconstructed lensing map and the temperature map at the six HFI frequencies. The errorbars correspond to the scatter within each band. The solid line is the expected result based on the PER model and is not a fit tothese data (see Fig. 16 for an adjusted model), although it is already a satisfying model. In each panel we also show the correlationbetween the lens reconstruction at 143 GHz and the 143 GHz temperature map in grey. This is a simple illustration of the frequencyscaling of our measured signal and also the strength of our signal as compared to possible intra-frequency systematic errors.

cance as follows. We count the number of standard deviations asthe quadrature sum of the significance in the different multipolebins:

sν =

√√√15∑i=1

CTφi

∆CTφi

2

. (2)

For our nominal parameters this gives us 3.6σ, 4.3σ, 8.3σ,31σ, 42σ, and 32σ, at, respectively, 100, 143, 217, 343, 545and 857 GHz. Note that these numbers include an additional20 % contribution to the statistical error to account for mode cor-relations (which we discuss in Sect. 5.1), but do not include sys-tematic errors or our point source correction. As a comparison, ineach panel we plot the correlation between the lens reconstruc-tion at 143 GHz and the 143 GHz map in grey. This shows thefrequency scaling of our measured signal and also the strengthof the signal, as compared to possible intra-frequency systematiceffects. This will be studied in depth in Sect. 5.

This first pass on our raw data demonstrates a strong detec-tion that is in good agreement with the expected CIB-lensingsignal. To get a better intuition for this detection, we show inFig. 4 the real-space correlation between the observed tempera-

ture and the lens deflection angles. This figure allows us to vi-sualize the correlation between the CIB and the CMB lensingdeflection angles for the first time. These images were generatedusing the following stacking technique. We first mask the 545and 857 GHz temperature maps with our combined mask thatincludes the 20 % Galaxy mask, and identify 20,000 local max-ima and minima in these maps. We also select 20,000 randomlocations outside the masked region to use in a null test. We thenband pass filter the lens map between ` = 400–600 to removescales larger than our stacked map as well as small-scale noise.We stack a 1 deg2 region around each point in both the filteredtemperature map and lensing potential map, to generate stackedCIB and stacked lensing potential images. We take the gradientof the stacked lensing potential to calculate the deflection angles,which we display in Fig. 4 as arrows. The result of the stack-ing over the maxima, minima and random points is displayedfrom left to right in Fig. 4. The strong correlation seen alreadyin the cross-power spectrum is clearly visible in both the 545 and857 GHz extrema, while the stacking on random locations leadsto a lensing signal consistent with noise. From simulations, weexpect a small off-set (' 1′′) in the deflection field. This offset

6

Page 7: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

Fig. 4. Temperature maps of size 1 deg2 at 545 and 857 GHz stacked on the 20,000 brightest peaks (left column), troughs (centrecolumn) and random map locations (right column). The stacked (averaged) temperature maps is in K. The arrows indicate thelensing deflection angle deduced from the gradient of the band-pass filtered lensing potential map stacked on the same peaks. Thelongest arrow corresponds to a deflection of 6.3′′, which is only a fraction of the total deflection angle because of our filtering. Thisstacking allows us to visualize in real space the lensing of the CMB by the galaxies that generate the CIB. A small and expectedoffset ('1′) was corrected by hand when displaying the deflection field.

was corrected for in this plot. We have verified in simulationsthat this is due to noise in the stacked lensing potential map thatshifts the peak. As expected, we see that the temperature max-ima of the CIB, which contain a larger than average number ofgalaxies, deflect light inward, i.e., they correspond to gravita-tional potential wells, while temperature minima trace regionswith fewer galaxies and deflect light outward, i.e., they corre-spond to gravitational potential hills.

5. Statistical and systematic error budget

The first pass of our pipeline suggests a strong correlation ofthe CIB with the CMB lensing potential. We now turn to in-vestigate the strength and the origin of this signal. We will firstdiscuss the different contributions to the statistical error budgetin Sect. 5.1, and then possible systematic effects in Sect. 5.2.Although the most straightforward interpretation of the signal isthat it arises from dusty star-forming galaxies tracing the large-scale mass distribution, in Sect. 5.3 we consider other potentialastrophysical origins for the observed correlation.

5.1. Statistical error budget

In this section we discuss any noise contribution that does notlead to a bias in our measurement. The prescription adoptedthroughout this paper is to obtain the error estimates from thenaive Gaussian analytical error bars calculated using the mea-sured auto-spectra of the CIB and lensing potential. We find thatthese errors are approximately equal to 1.2 times the naive scat-ter within an `-bin, and we will sometimes use this prescriptionwhere appropriate for convenience (as will be stated in the text).This is justified in Appendix A where we consider six differentmethods of quantifying the statistical errors using both simula-tions and data. The Gaussian analytical errors, ∆CTφ

`, are calcu-

lated using the naive prescription

fsky (2` + 1) ∆`(∆CTφ

`

)2= CTT

` Cφφ`

+(CTφ`

)2, (3)

where as before fsky is the fraction of the sky that is unmasked,∆` = 126 for our 15 linear bins between ` = 100 and ` = 2000,CTT` and Cφφ

`are the spectra measured using the data, and CTφ

`is the model cross spectrum. This last term provides a negligi-

7

Page 8: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

10-5

10-4

10-3

10-2

10-1

100

∆(`

3C

`)

[µK.s

r] ν=100 GHz (x 10) ν=143 GHz (x 100) ν=217 GHz (x 100)

100 500 1000 1500`

10-5

10-4

10-3

10-2

10-1

100

∆(`

3C

`)

[µK.s

r] ν=353 GHz (x 100)

100 500 1000 1500`

ν=545 GHz (x 10)

100 500 1000 1500 2000`

ν=857 GHz (x 0.1)

Fig. 5. Naive analytical estimates of the contribution to the CTφ`

variance as a function of multipole and frequency as given inEq. 4. We assume the same bin sizes as in Fig. 3. The differentlines are the contribution to the analytical error from the signalonly: Cφφ

`CCIB`

+(CCIB φ`

)2(green), noise only: Cφφ,N

`CCIB,N`

(blue),and the mixed signal and noise terms: Cφφ

`CCIB,N`

(yellow) andCφφ,N`

CCIB`

(orange). The total contribution is the solid black line,

and the theory spectrum,(CCIBφ`

)2, is the dashed line.

ble contribution due to the large noise bias on Cφφ`

, as we nowdescribe.

The statistical error has two sources: instrumental and astro-physical. The measured auto-spectra in Eq. 3 contain a signaland noise contribution: CXX

` = CXX` + CXX,N

`. It is informative to

split the right hand side of Eq. 3 into four pieces:

rhs =

[Cφφ`

CCIB` +

(CCIBφ`

)2]

+ Cφφ,N`

CCIB,N`

+ Cφφ`

CCIB,N`

+ Cφφ,N`

CCIB` .

(4)

Here the first term is a signal-only piece, the second is a noise-only piece, and the remaining two terms are mixed signal andnoise pieces. To discuss the relative importance of these terms,we will use for the signal terms the model spectra, and for thenoise terms we subtract the model spectra from the measuredspectra: CXX,N

`= CXX

` − CXX` . With this definition, the noise will

include the instrumental contribution, as well as other astrophys-ical signals including the CMB, which we do not remove fromour data for reasons previously mentioned. We show the differ-ent terms in Fig. 5. Up to 353 GHz the measured temperaturespectrum, CTT

` , is dominated by the CMB at low ` and the instru-mental noise at high `. At higher frequencies Galactic emissiondominates at low ` and the CIB at high `. For all frequencies upto 353 GHz the dominant contribution to the errors in our signalcomes from the noise-only term (in blue), which is proportionalto the temperature noise spectrum. At 353 GHz and above themixed signal-noise term Cφφ,N

`CCIB`

(orange) becomes importantand is the largest contribution at 545 and 857 GHz at high `.

5.2. Instrumental and observational systematic effects

A number of systematic errors affect the Planck HFI analysisand we briefly discuss some of them here. A more complete dis-cussion can be found in Planck HFI Core Team (2011b). We

will illustrate how the very nature of our measurement, a 3-pointfunction, makes it particularly robust to many systematic effects,and we will check for their signatures using null tests. For exam-ple, there is no noise bias in the 3-point measurement, and manyeffects that can lead to biases in the auto-spectrum of φ do notaffect us.

5.2.1. Potential sources of systematic error

We begin by describing our knowledge of known systematic ef-fects, before discussing others that could bias our result. To ac-count for an error in the calibration of the temperature maps, wesimply add in quadrature a calibration uncertainty to our errorbars. In Sect. 5.2.2 we use null tests to check that these errorsare consistent with the data. In addition we use the null tests tosearch for evidence that the calibration has changed between sur-veys, for example due to gain drifts. We account for beam errorsin a conservative manner by using a constant error at each fre-quency equal to the maximum error in the beam multipoles, B`,at any ` (see PIR for details). The B` uncertainty is larger at high-` with, for example, values at ` = 1500 of 79.5 %, 8.2 %, 0.53 %,0.95 %, 0.31 %, and 0.70 %, at 100–857 GHz, respectively. Thecalibration error is therefore larger than the beam error at all` between 217 and 857 GHz but smaller at high ` in the 100and 143 GHz channels. We add the beam error in quadrature inan `- and frequency-dependent manner. As discussed in PlanckCollaboration XVII (2013), uncertainties in the beam transfer(as well as the fiducial CMB power spectrum CT

` T ) propagatedirectly to a normalization uncertainty in the lens reconstruc-tion. Based on the beam eigenmodes of Planck CollaborationVII (2013), it is estimated in Planck Collaboration XVII (2013)that beam uncertainty leads to an effective normalization uncer-tainty of approximately 0.2% and 143 and 217 GHz, and 0.8% at100 GHz. To be conservative, on top of the calibration and beamerror we will add in quadrature a 2 % uncertainty on the overalllens normalization.

CMB lens reconstruction recovers modes of the lensingpotential through their anisotropic distorting effect on small-scale hot and cold spots in the CMB. The quadratic estima-tor, which we use to reconstruct the lensing potential is opti-mized to measure the lensing induced statistical anisotropy inCMB maps. However, other sources of statistical anisotropy,such as the sky mask, inhomogeneous noise, and beam asymme-tries, produce signals, which can potentially overlap with lens-ing. These introduce a “mean-field” bias, which we estimate us-ing Monte Carlo simulations and subtract from our lensing esti-mates. Innaccuracies in the simulation procedure will lead to er-rors in this correction, particularly if the correction is large. Themean-field introduced by the application of a Galaxy and point-source mask, for example, which can be several orders abovemagnitude larger than the lensing signal at ` < 100. This isdiscussed further in Appendix B of Planck Collaboration XVII(2013). The mask mean-field is a particular concern for us be-cause it has the same phases as the harmonic transform of themask. If our masked CIB maps have a non-zero monopole, forexample, it will correlate strongly with any error in the maskmean-field correction. For this reason we do not use any databelow ` = 100 in our analysis.

To summarize, we do not expect these known systematic ef-fects to be present at a significant level. Nevertheless, we stillperform a set of consistency tests that would be sensitive to themor other unknown effects.

8

Page 9: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

5.2.2. Null tests

The Planck scanning strategy, its multiple frequency bands andits numerous detectors per frequency, offer many opportunitiesto test the consistency of our signal (see Sect. 2.1). We focus onsuch tests in this section. Our aim is to reveal any systematic ef-fects that could lead to a spurious correlation. For all of the testspresented, we will quote a χ2 value as well as the number of de-grees of freedom (Ndof) as a measure of the consistency with theexpected null result. Throughout this section, black error barsin plots will correspond to the measured scatter within an ` binmultiplied by 1.2, as was justified in Sect. 5.1 and Appendix A,and will also include a CIB calibration error and a beam error,while the coloured boxes correspond to the analytical errors ofthe corresponding signal (i.e., not the difference correspondingto the null test). Plotting these two error bars illustrates how im-portant any deviation could be to our signal. Throughout thissection, we will illustrate our findings with the 545 GHz corre-lation, since it is our prime band for this measurement, but ourconclusions hold at other frequencies.

The first test we conduct is to take the temperature differ-ence between the two half-ring (HR) maps to cancel any sig-nal contribution, and therefore investigate the consistency of ourmeasurements with our statistical errors on all time scales. Wenull the temperature maps and correlate with our nominal lens-ing map. The results are shown in the left panel of Fig. 6. We seea significant deviation from null only when considering surveydifferences. This particular failure can probably be explained byapparent gain drifts due to nonlinearity in the analog-digital con-version (Planck Collaboration VI 2013; Planck CollaborationVIII 2013), not yet corrected at this frequency. Note howeverthat the predicted variation is about 1% while the deviation fromnull would call for a variation of 1.5-2%. But in any case, itsamplitude is too small to significantly affect our measurement.

We see a significant deviation from null only when con-sidering survey differences. This particular failure can proba-bly be explained by apparent gain drifts due to nonlinearity inthe analog-digital conversion (Planck Collaboration VI 2013;Planck Collaboration VIII 2013), not yet corrected at these fre-quency. But in any case, its amplitude is too small to significantlyaffect our measurment.

The second test uses multiple detectors at a given frequencythat occupy different parts of the focal plane. These detector setsare used to construct the “detset” maps that were described inSect. 2.1. The two “detset” maps are subtracted and then corre-lated with our nominal lens reconstruction. This test is particu-larly sensitive to long term noise properties or gain variations, aswe do not expect these to be correlated from detector to detector.Since this detector division breaks the focal plane symmetry, itis also a good check for beam asymmetry effects. Here again, wedo not find any significant deviation, as illustrated in the middlepanel of Fig. 6.

The third test we conduct makes use of the redundant skycoverage, using multiple surveys to cancel the signal. As above,we null the temperature signal and correlate with the nominallens reconstruction. This test is particularly sensitive to any longterm, i.e., month timescale drifts that could affect our measure-ment. It is also a good test for any beam asymmetry effects, asindividual pixels are observed with a different set of orientationsin each survey. Since only the first two surveys are completefor this particular data release, we only use the two full surveymaps to avoid complications with the partially completed thirdsurvey. Here again, we do not find any significant deviation, asillustrated in the right panel of Fig. 6.

100 500 1000 1500`

1.5

1.0

0.5

0.0

0.5

1.0

1.5

∆( `3

CTφ

`

)[µ

K.s

r]

ν=545 GHz (x10)

χ2 (dof) = 9.2 (15)

100 500 1000 1500`

ν=545 GHz (x10)

χ2 (dof) = 27.8 (15)

100 500 1000 1500 2000`

ν=545 GHz (x10)

χ2 (dof) = 40.5 (15)

Fig. 6. Null tests at 545 GHz. Left: difference spectra obtained bynulling the signal in the HR temperature map before correlatingit with our nominal φ reconstruction. Middle: temperature signalnulled using different detectors at 545 GHz. Right: temperaturesignal nulled using the first and second survey maps. The blackerror bars correspond to the scatter measured within an `-bin,while the coloured bands correspond to the analytical estimate.Except for the survey null test (see text for details), these testsare passed satisfactorily except, as illustrated by the quoted χ2

and Ndof , thus strengthening confidence in our signal.

100 500 1000 1500`

1.0

0.5

0.0

0.5

1.0

∆( `3

CTφ

`

)[µ

K.s

r]

ν=545 GHz (x10)

χ2 (dof) = 25.5 (15)

100 500 1000 1500`

ν=545 GHz (x10)

χ2 (dof) = 13.5 (15)

100 500 1000 1500 2000`

ν=545 GHz (x10)

χ2 (dof) = 11.8 (15)

Fig. 7. Left: difference between the cross-spectra measured us-ing the 20 % Galactic mask (20 % is the unmasked sky frac-tion) from that measured with our default 40 % Galactic mask.Middle: spectra obtained when differencing the 60 % and 40 %Galaxy mask measurements. For both left and middle panels andall Galactic masks, the same point source and H imasks are used,which removes an additional fraction of the sky. Right: differ-ence between the cross-spectra calculated with the H i cleanedtemperature maps from those with no H i cleaning. This cross-spectrum is thus the correlation between the H i template and theφ reconstruction. The error bars are calculated in the same wayas in Fig. 6. Again, the null tests are passed with an acceptableχ2.

To conclude, this first set of stringent consistency tests haveshown that there is no obvious contamination of our measure-ments due to instrumental effects. In addition, the reasonableχ2/Ndof obtained gives us confidence in our statistical noise eval-uation. Although we measure the noise directly from the data,this success was not guaranteed.

5.3. Astrophysical contamination

We now turn to possible astrophysical biases to our measure-ment. We will discuss successively known astrophysical contam-inants that can either come from Galactic or extragalactic origin.Once again, besides our knowledge of these signals, we will relyheavily on consistency tests made possible by having multiplefull sky frequency maps.

9

Page 10: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

100 500 1000 1500`

1.5

1.0

0.5

0.0

0.5

1.0

1.5

∆( `3

CTφ

`

)[µ

K.s

r]

ν=545 GHz (x10)

χ2 (dof) = 12.1 (15)

100 500 1000 1500`

ν=545 GHz (x10)

χ2 (dof) = 15.6 (15)

100 500 1000 1500 2000`

ν=545 GHz (x10)

χ2 (dof) = 2.6 (15)

Fig. 8. Left: difference between cross-spectra calculated usingthe lens reconstruction at 100 GHz with the nominal 143 GHzreconstruction. We see an overall shift, which leads to a highreduced χ2. This shift can be explained by the expected overallnormalization uncertainties of the 100 GHz and 143 GHz recon-structions. While this uncertainty is not included in the χ2 orthe solid bars, it is included later in our analysis in Sect. 6.1.Middle: same as the left panel, but the 217 GHz reconstructionis used instead of the 100 GHz reconstruction. Right: differencebetween cross-spectra when we consider the 143 GHz lens re-construction calculated with a less restrictive Galaxy mask (thatexcludes only 20 % of the sky) and the nominal reconstructionmask that excludes 40 % of the sky.

5.3.1. Galactic foregrounds

Galactic foregrounds have two possible effects on our measure-ment. The first is the introduction of an extra source of noise. Thesecond is that contamination of the lensing reconstruction by anyGalactic signal, e.g., synchrotron, free-free or dust, which couldthen correlate with foreground emission present in the tempera-ture maps, remains a source of bias that has to be investigated.We will show that this bias is small. To do so, we take three ap-proaches. We first investigate various Galactic masks, then per-form the lensing reconstruction at various frequencies, and fi-nally investigate the effect of a dust-cleaning procedure.

First, we consider two additional masks, either more aggres-sive or more conservative than our fiducial one. Both were intro-duced in Sect. 2.1. The first one leaves approximately 20 % ofthe sky unmasked, while the second one leaves approximately60 % of the sky. Given the strong dependence of Galactic fore-grounds on Galactic latitude, any Galactic contamination shouldvary strongly when we switch between masks. Comparing themeasurements using these masks with our fiducial 40 % mask inthe left and centre panels of Fig. 7, we do not see any substantialdeviation from our fiducial measurements. This excludes strongGalactic contamination of our results.

Second, we perform the lens reconstruction at 100 and217 GHz, different from the fiducial frequency of 143 GHz, andcompare their correlation with the temperature maps. Given thestrong dependence of the Galactic emission with frequency, T ∝ν−3 for synchrotron and T ∝ ν2 for dust in this frequency range,any contamination of our signal would have a strong frequencydependence. The comparison with the 100 GHz (217 GHz) re-construction is presented in the left (centre) panel of Fig. 8. Theright panel shows the difference of the cross-spectra calculatedusing the 143 GHz reconstruction with a more aggressive Galaxymask (20 % instead of 40%), to reduce possible Galactic contam-inants in the reconstruction, and the nominal reconstruction. Thethree differences are consistent with null as demonstrated by thequoted χ2 and Ndof .

Third, we investigate more specifically how cirrus, the domi-nant Galactic contaminant for our higher frequency channels, af-fects our measurements. We rely on the dust cleaning procedure

detailed in Sect. 3.2 that aims to remove the H i-correlated dustcomponent. This procedure leads to a decrease in the variancemeasured outside the mask of 22, 65, 73 and 73 % in the 217,353, 545 and 857 GHz maps, respectively. This frequency de-pendence is expected given the dust scaling. However, in Fig. 7,where we show the differences between the cleaned and non-cleaned cross-spectra, we observe that the large scale H i clean-ing, even though it makes a substantial impact on the powerwithin our map, only makes a small change at low-` in the cross-spectrum, as well as reducing the noise at all multipoles. If wequantify the effect of our “local” H i cleaning on the detectionsignificance level computed using only statistical errors, we findthat the significance is increased by 4, 4, 28, and 36 % at 217,353, 545 and 857 GHz, respectively. Also, not surprisingly, weobserve that for frequencies up to 353 GHz where the statisticalerrors are dominated by the CMB, the H i cleaning has almost noeffect on the cross-spectra. From the three studies in this sectionwe conclude that there are no obvious signs of Galactic fore-ground contamination in our cross-correlation.

5.3.2. Point source contamination

We now discuss another well-known potential source of contam-ination, namely the contribution of unresolved point sources vis-ible either through their radio or dust emission. Our concern isthat a correlation between a spurious lens reconstruction causedby unresolved point sources can correlate with sources in thetemperature map. Although in Sect. 5.3.1 our null test using lensreconstructions at different frequencies suggests that unresolvedpoint sources are not an obvious contaminant, we will now per-form a more detailed test designed specifically to search for pointsource contamination. Following Smith et al. (2007); Osborneet al. (2013), we will construct a point source estimator designedto be more sensitive than the lensing estimator to point sourcecontamination. Our focus here will be on possible contamina-tion from the point source shot-noise bispectrum. In Sect. 5.3.5we will discuss contamination from a scale dependent bispec-trum.

Our (unnormalized) quadratic estimator, which is designedto detect point source contributions is given by

(Θ143(n))2LM ≡

∑LM

Y∗LM(n) (Θ143(n))2, (5)

where Θ is the inverse-variance filtered sky map. This estima-tor is simply the square of the inverse-variance filtered sky map,which is a more sensitive probe of point sources than the stan-dard lensing estimator.

In Fig. 9 we plot the cross-spectrum of (Θ143(n))2LM mea-

sured at 143 GHz and ΘνLM for the full set of HFI channels. This

cross-spectrum is probing the same point source contributionsthat could bias our estimates of CTφ

`, however with a greater

signal-to-noise ratio.There is one complication here, which is that just as lens

reconstruction may be biased by point source contributions, thepoint source estimator is correspondingly biased by lensing. Thebias to the plotted cross-spectra is given by

〈(Θ143(n))2LMΘν∗

LM〉φ =Cνφ

L

CννL

∑`1m1

∑`2m2

G−Mm1m2L`1`2

Ctot`1

Ctot`2

×((−1)m1 Im2−M−m1

`2L`1C`1 + (−1)m2 Im1−M−m2

`1L`2C`2

), (6)

10

Page 11: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

where Gm1m2m3`1`2`3

=∫

dn Y`1m1 (n)Y`2m2 (n)Y`3m3 (n), C` is the un-lensed CMB spectrum, Ctot

` is the spectrum of Θ143(n), Im1m2m3`1`2`3

is as defined in Okamoto & Hu (2003), and we have used thefact that for our inverse-variance filtering, Θ`m ≈ Θ`m/Ctot

` . Wehave calculated this contribution using our measured CTφ

L andsubtracted it from the data points of Fig. 9.

We can consider the effect of shot noise on this cross-spectrum. With the shot-noise bispectrum defined by

〈Θ`1m1Θ`2m2Θν`3m3〉S 3 = Gm1m2m3

`1`2`3

⟨S 3

⟩(7)

the bias to the plotted cross-spectrum is given as

〈(Θ143(n))2LMΘν

LM〉S 3 =⟨S 3

⟩ (−1)M

CννL

∑`1m1

∑`2m2

G−Mm1m2L`1`2

GMm1m2L`1`2

Ctot`1

Ctot`2

.

(8)This bias is plotted for best-fit values of 〈S 3〉 as the black lines inFig. 9. To minimize systematic effects that might bias the

⟨S 3

⟩estimator, we have estimated S 3 from the spectra of Fig. 9 be-tween multipoles between ` = 500 and 2000. The fitted

⟨S 3

⟩amplitudes are given in Table 1.

These amplitudes match our expectations, for examplesee Planck Collaboration XIII (2011). We observe a decrease inthe amplitude of the point source contribution going from 100 to217 GHz, which corresponds to a dominant contribution from ra-dio point sources. We do not see any evidence of a dusty galaxycontribution to the shot-noise bias. These conclusions have beenverified using less restrictive point source masks that cover fewersources.

With estimates of S 3 in hand, we estimate a correspondingbias to CTφ

`, given by

⟨φLMΘν∗

LM

⟩S 3

= (−1)M

⟨S 3

⟩Cνν

L

∑`1m1

∑`2m2

Wm1m2 M`1`2L

Ctot`1

Ctot`2

Gm1m2−M`1`2L (9)

where Wm1m2 M`1`2L = (−1)m2 (I−m2 Mm1

`2L`1C`1 + Im1 M−m2

`1L`2C`2 )/2RφφL with

RφφL defined in Planck Collaboration XVII (2013). We show this

contribution later in Fig. 11 as the dotted line. While non-zero,we see that the point source shot noise contribution is alwaysnegligible in the ` range we consider, except at lower frequencieswhere the radio point sources are important (but still not strongenough to lead to any clear signal in the cross-spectra).

5.3.3. SZ contamination

A fraction of CMB photons travelling from the surface of recom-bination are scattered by hot electrons in galaxy clusters. In the

Table 1. Point source estimator. The measured quantity⟨S 3

⟩, as

defined in Eq. 9 is given as a function of frequency.

Frequency⟨S 3

⟩(No. of σ)

[ GHz] [×109µK3]

100 . . . . . . . 11.7 ± 5.8 (2.0)143 . . . . . . . 4.3 ± 1.8 (2.3)217 . . . . . . . 3.7 ± 1.6 (2.2)353 . . . . . . . 6.1 ± 3.9 (1.6)545 . . . . . . . −79 ± 39 (−2.0)857 . . . . . . . (−1.9 ± 2.1) × 103 (−0.9)

100 143 217 353 545 857ν [GHz]

10-4

10-3

10-2

10-1

100

101

CTφ

`[µ

K.s

r]

`=1200-1450

100 143 217 353 545 857ν [GHz]

10-4

10-3

10-2

10-1

100

101

CTφ

`[µ

K.s

r]

`=300-450

CIB (G)CIB (F)|SZ|

CIB (G) + SZ

Fig. 10. Frequency spectrum of our cross-spectra averagedwithin `-bins (black points with associated error bars). The lightshaded regions correspond to the HFI frequency bands. The solidblack curve corresponds to the best-fit CIB assuming a Gispertet al. (2000) spectrum, while the dot-dashed line assumes aFixsen et al. (1998) spectrum. The dashed black line correspondsto the best-fit model when allowing for an SZ component in ad-dition to the Gispert et al. (2000) CIB shape. The blue dots cor-respond to the associated absolute value of the best-fit SZ com-ponent. We conclude from this plot that the SZ effect is not animportant contaminant.

most massive clusters approximately 1 % of CMB photons pass-ing through them get scattered. On average, their energy will beincreased, which leads to a measurable spectral distortion. Thisis the so-called thermal Sunyaev-Zeldovich (SZ) effect (Sunyaev& Zeldovich 1970). At the location of a galaxy cluster the CMBappears colder at frequencies below about 220 GHz and hotterat higher frequencies, with a temperature change proportionalto the cluster optical depth to Compton scattering and to theelectron temperature. Since hot electrons in clusters also tracethe large scale matter potential that is traced by CMB lensing,we expect an SZ-induced contamination in our measurement atsome level. We will show below that the level of contaminationis negligible. In these calculations we ignore the small relativis-tic corrections to the thermal SZ spectrum (e.g., Nozawa et al.2000). We also ignore the kinetic SZ signal coming from the bulkmotion of hot electrons in clusters, since it is subdominant tothe thermal signal (Sunyaev & Zeldovich 1980; Reichardt et al.2012; Hand et al. 2012).

The frequency dependence of the SZ signal in our map de-pends on the detector bandpasses and is

f (ν) =

∫dν h(ν) g(ν)∫

dν h(ν), (10)

where h(ν) is the detector bandpass and g(ν) is the SZ fre-quency dependence, which in the non-relativistic limit is g(ν) =x (ex + 1)/(ex − 1) − 4, with x = hν/kBTCMB. The effect of thebandpass only makes a large difference at 217 GHz near the nullof the SZ signal. The thermal SZ affects our measurement in twoways. First, since the SZ emission in our maps is not a Gaussianrandom field (e.g., Wilson et al. 2012) it introduces a spurioussignal into our lens reconstruction that will correlate with theSZ signal in our CIB map. As shown in Osborne et al. (2013),

11

Page 12: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

100 500 1000 1500 2000`

40

20

0

20

40

60

80

∑ M

( ( Θ14

3) 2) L

MΘ∗ LM

[x10

12µK

3]

ν=100 GHz

<S3 >

10−9µK=(11.7 ± 5.85)

100 500 1000 1500 2000`

15

10

5

0

5

10

15

20 ν=143 GHz

<S3 >

10−9µK=(4.31 ± 1.84)

100 500 1000 1500 2000`

5

0

5

10

ν=217 GHz

<S3 >

10−9µK=(3.69 ± 1.65)

100 500 1000 1500 2000`

20

10

0

10

20

∑ M

( ( Θ14

3) 2) L

MΘ∗ LM

[x10

12µK

3]

ν=353 GHz

<S3 >

10−9µK=(6.09 ± 3.87)

100 500 1000 1500 2000`

200

150

100

50

0

50

100

150 ν=545 GHz

<S3 >

10−9µK=(-79.0 ± 39.1)

100 500 1000 1500 2000`

8000

6000

4000

2000

0

2000

4000 ν=857 GHz

<S3 >

10−9µK=(-1910 ± 2090)

Fig. 9. Results from the point source contamination estimator of Eq. (5). The best-fit cross-spectra associated with shot noise areplotted in black. We do not show the best-fit at 545 and 857 GHz since the signal-to-noise ratio is low. The grey line is a predictionfor the bias from the CMB lensing - infrared correlation, and has been subtracted from the spectra (plotted as black points). Wesee that with the subtraction of the bias from CMB lensing, the measured bispectrum-related spectrum is generally consistent eitherwith zero, or with the shape expected for shot noise.

this is well approximated by a Poisson noise term and is thus al-ready addressed by our treatment of point sources in Sect. 5.3.2.The spurious lensing signal will also correlate with other com-ponents in our map such as the CIB. However, we ignore theseterms since they will be smaller than those that correlate directlywith the SZ emission. Additionally, a contribution comes fromSZ emission in our CIB map that correlates with the lensing po-tential itself. The latter is the dominant term and we discuss it inthis section.

To measure a contribution from the SZ-lensing correlationwe attempt to separate the SZ and CIB emission based on theirdiffering spectral shapes. We consider all frequencies from 100to 857 GHz, but we will illustrate this procedure by consideringonly two ` bands: ` = 300–450; and 1200–1450. The first is wellinside the linear regime, while the second receives a more im-portant non-linear contribution. However, we have checked thatif we consider different `-bins we obtain similar conclusions. Wemodel the signal within each ` band as s`(ν) = a1,`c(ν)+a2,` f (ν),where c(ν) and f (ν) are, respectively, the CIB frequency depen-dence (as proposed in Fixsen et al. 1998 or Gispert et al. 2000)and the SZ frequency dependence obtained from Eq. 10. For

each ` band, we will solve for a1,` and a2,` minimizing the as-sociated χ2 while forcing both amplitudes to be positive. As anapproximation to the error in each multipole band we calculatethe scatter of the signal within the band and multiply it by 1.2,as discussed in Sect. 5.1.

In Fig. 10 we show the measured frequency spectrum withineach ` band, along with the best-fit SZ-lensing and CIB-lensingspectra. For the CIB-only fit with the Gispert et al. (2000) fre-quency dependence we find a relatively poor fit in the lowest`-bin, χ2 (dof) = 15.5 (5), but an improved fit in the higher `-bin, χ2 (dof) = 4.15 (5). Including the SZ component gives ∆χ2

= 0.52 and 1.34 in the low and high ` bins for one extra degreeof freedom. When we use the Fixsen et al. (1998) frequency de-pendence we find an improved fit, with χ2 (dof) = 2.25 (5) and5.49 (5) in the low and high-`-bins, respectively. Overall, the im-provement in the χ2/dof when including the SZ component doesnot justify inclusion of the SZ component in the model, withthe poor fit driven by the lowest frequency bands where the CIBscaling is rather unconstrained. In fact, our measurements mightconstitute the first constraints to date on this scaling. From theseresults we conclude that including the SZ-lensing correlation in

12

Page 13: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

our data does not improve the fit in the ` range of interest to usand thus we do not consider it necessary to correct for.

As an extra validation of this result, we now verify its consis-tency with current models of the CIB and SZ emission. For thispurpose, we use the calculation of the correlation from Osborneet al. (2013), based on Babich & Pierpaoli (2008), which modelsthe SZ emission as a statistically isotropic signal modulated bya biased density contrast, where the bias depends on the clus-ter mass and redshift. To obtain an estimate of the contributionto the cross-spectrum at 217–857 GHz we assume that the mea-sured cross-spectrum at 143 GHz is entirely due to thermal SZemission (note that we do this to find what we believe to be anupper limit on the SZ contribution at 217–857 GHz; for the rea-sons stated above we do not expect the 143 GHz correlation tobe due to SZ). Since the SZ signal at 143 GHz gives a decrementin the CMB, and the CIB emission gives an enhanced signal, itis possible that this approach could still underestimate the SZsignal. We find that in order to fit the cross-spectrum at 143 GHzusing only the SZ-lensing correlation requires an amplitude of(2.4 ± 1.6) times our calculated SZ-lensing cross-spectrum. InFig. 11 the dashed line shows the magnitude of this SZ signalscaled to each frequency using Eq. 10. The small contribution itmakes at 217–857 GHz further suggests that we can neglect thiscomponent. At 217 GHz the signal is negative, while at higherfrequencies it is positive.

5.3.4. ISW contamination

The Integrated Sachs-Wolfe (ISW) effect describes the redshift-ing (blueshifting) of photons travelling through gravitational po-tential wells (hills) that decay as the photons travel throughthem (Sachs & Wolfe 1967). The induced modulation of theCMB mean by the gravitational potential generates CMB fluc-tuations that correlate with the lensing potential, which alsotraces out the gravitational potential perturbations (Seljak &Zaldarriaga 1999; Goldberg & Spergel 1999; Lewis et al. 2011).Note that because the mean of the CIB is relatively much smallerthan its fluctuation, the ISW induced CIB fluctuations make anegligible change to total CIB anisotropy. The CMB ISW in-duced signal has the same frequency dependence as the CMBand so is only a significant contaminant for us at low frequen-cies. We evaluate this signal using a theoretical calculation per-formed in CAMB (Lewis et al. 2000). The results are shown as thesolid line in Fig. 11. It is a negligible contribution at all frequen-cies, except in the lowest `-bin of the lowest frequencies, wherethe measured cross-spectrum is consistent with zero.

5.3.5. CIB bispectrum

Having calculated the bias from the point source shot noise inSect. 5.3.2, we now discuss a more complicated form of theunresolved point source 3-point function that could be presentin our data, namely the clustering contribution. While unknown(although the first detection was recently reported in Crawfordet al. (2013)), the CIB bispectrum is potentially a direct contam-inant to our measurement. Because of the non-linear clusteringof DSFGs (PER), it has to exist. But because of the very largeredshift kernel that characterizes the CIB, this non-Gaussian ef-fect will be washed out, reducing its importance. Nevertheless,we ought to study it carefully.

If important, this effect would show up as a departure of thedata from the best-fit curve in Fig. 9, since the best-fit modelthat we used assumes only a Poissonian shot-noise contribution.

500 1000 1500 2000`

4000

3000

2000

1000

0

1000

2000

`3C

`[µ

K.s

r]

fsky =0.09

fsky =0.16

fsky =0.30

Fig. 12. Cross-spectrum of the 545 GHz lens reconstructioncorrelated with the 545 GHz temperature map with differentGalactic masks. The legend gives the visible sky fractions. Thesolid line represents the analytic unclustered shot-noise contri-bution fit to the fsky = 0.09 points above ` = 1300.

We do not see any significant deviation in Fig. 9. Still, in or-der to isolate this effect we create cross-spectra with increasedsensitivity to the clustered point source signal. We do this bycalculating the lens reconstruction at 100 GHz and 545 GHz,where, respectively, the radio and dusty point source contribu-tion is stronger. The 545 GHz map has a much larger Galacticdust signal than our nominal 143 GHz map. However, unlike inour fiducial estimates, here we do not attempt to project out dustcontamination from the map used to perform our lens recon-struction as this would also remove some of the CIB signal inthe bispectrum. As was found in Sect. 5.3.1 the cross-correlationbetween the 100 GHz reconstruction and the 100 GHz temper-ature map does not show any large difference with the cross-spectrum obtained using the 143 GHz signal. We are thus notsensitive enough to detect a bias from the clustering of radiosources using this method. However, we do detect a strong cross-correlation between the raw 545 GHz lens reconstruction andthe 545 GHz temperature map. This cross-spectrum is shownin Fig. 12 for three different Galaxy masks. The line shows thepoint source shot-noise template derived in Sect. 5.3.2, fit to thecross-spectrum with the 10 % Galaxy mask at ` above 1300. Ifthe signal were entirely due to extragalactic point sources, thenthe signal would be independent of masking, and we do see aconvergence of the signal at high ` as the size of the Galacticmask is increased. At low `, however, there is a large Galacticcontribution and the convergence with the reduced mask sizeis less clear. We thus conclude that a strong contribution fromGalactic dust is present in this measurement at all `.

We do not attempt to calculate accurately the shape of theclustering contribution to the CIB bispectrum here, since it isbeyond the scope of this work, even though a simple prescrip-tion for it has recently been proposed in Lacasa et al. (2012).To separate the Galactic from non-Galactic contributions in ourbispectrum measurement is difficult, even if a strong Galacticsignal is clearly present, given the strong dependence of the sig-nal on variations of the Galactic mask in Fig. 12. However, thecombination of dust cleaning that we perform in our nominalpipeline, coupled with the fact that our nominal pipeline usesthe 143 GHz map for lens reconstruction, means that we do not

13

Page 14: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

100 500 1000 1500 2000`

10-8

10-7

10-6

10-5

10-4

10-3

10-2

10-1`3C

`[µ

K.s

r]ν=100 GHz

100 500 1000 1500 2000`

10-8

10-7

10-6

10-5

10-4

10-3

10-2 ν=143 GHz

100 500 1000 1500 2000`

10-8

10-7

10-6

10-5

10-4

10-3

10-2

10-1

ν=217 GHz

100 500 1000 1500 2000`

10-8

10-7

10-6

10-5

10-4

10-3

10-2

10-1

`3C

`[µ

K.s

r]

ν=353 GHz

100 500 1000 1500 2000`

10-8

10-7

10-6

10-5

10-4

10-3

10-2

10-1

100 ν=545 GHz

100 500 1000 1500 2000`

10-810-710-610-510-410-310-210-1100101102103

ν=857 GHz

Fig. 11. Foreground components at each frequency. The data points and error bars show our results. The dashed line is an estimatedupper limit on the magnitude of the SZ contamination derived in Sect. 5.3.3. We show the absolute value of this contribution, whichis negative at frequencies less than 217 GHz. The dot-dashed line is the extragalactic point source contribution, with an amplitudemeasured from our data as derived in Sect. 5.3.2. Again we show the absolute value, with the signal being negative below ` ∼ 1200.The oscillating solid line corresponds to the calculated ISW contamination.

observe any dependence with masking in our measurement, asseen in Fig. 7. Because of this, the CIB bispectrum is unlikely tobe a large contribution to our measurement. Furthermore, evenif we were to assume that all of the signal seen in Fig. 12 wasextragalactic in nature, using the Gispert et al. (2000) frequencyscaling for the CIB (also appropriate for the Galactic dust in fact,Planck Collaboration XXIV 2011), the roughly −1700 µK.sr ob-served at ` = 400 for the 40 % Galactic mask would only leadto a −0.02 µK.sr signal in Fig. 11, which is an order of magni-tude smaller than our measured signal. To conclude, althoughour analysis does not lead to a clean measurement of the CIBbispectrum, we can safely assume that it is not a contaminant toour measurement.

5.4. Final statistical and systematic error budget

Throughout the suite of tests for instrumental and observationalsystematic errors presented in Sect. 5.2, as well as the suiteof tests for possible astrophysical contaminants presented inSect. 5.3, we have established the robustness of our measure-ment. The fact that our consistency tests do not lead to any sig-

nificant deviation gives us confidence in our error budget. Asdescribed in Sect. 5.2 we add to them an overall calibration un-certainty, beam uncertainty, and lens normalization uncertainty,consistent with the Planck data processing paper (Planck HFICore Team 2011b). We gather the measured band-powers inTable 2, along with our statistical and systematic errors. Theseband-powers have been corrected for the point source compo-nent measured in Sect. 5.3.2, whose amplitude is also given inTable 2.

Once all systematic effects are factored in, we claim a de-tection significance of 3.6 (3.5), 4.3 (4.2), 8.3 (7.9), 31 (24), 42(19), and 32 (16) σ statistical (statistical and systematic) at 100,143, 217, 353, 545 and 857 GHz, respectively.

6. Interpretation and discussion

The correlation we have investigated leads to a very strong sig-nal at most HFI frequencies. After a thorough examination ofpossible instrumental and astrophysical origins, we interpret itas originating from the spatial correlation between the sourcesof the CIB and the matter responsible for the gravitational de-flection of CMB photons. In this section, we build on this result

14

Page 15: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

Table 2. Cross-spectrum detection band-powers. All values are in units of µK.sr. The extragalactic foreground contribution, Cfore`

has been removed from C`. Both statistical and systematic errors are given (see Sect. 5 for details).

`mean 163 290 417 543 670 797 923 1050 1177 1303 1430 1557 1683 1810 1937

100 GHz `3C` × 1000 1.32 5.44 0.26 0.54 −0.46 −0.37 −0.64 −0.69 −1.72 0.59 −2.63 2.96 2.61 −1.56 10.01∆

(`3Cstat

`

)× 1000 2.48 2.15 1.30 1.33 1.17 1.27 1.17 1.12 1.29 1.51 1.97 2.66 3.78 5.69 8.93

∆(`3Csys

`

)× 10000 0.28 1.14 0.05 0.12 0.09 0.07 0.13 0.14 0.37 0.10 0.62 0.50 0.34 0.65 1.63

`3Cfore` × 10000 -0.00 −0.01 −0.03 −0.08 −0.16 −0.25 −0.28 −0.14 0.33 1.31 3.05 5.79 9.78 15.29 22.48

143 GHz `3C` × 1000 2.05 6.28 1.11 1.37 0.09 0.05 −0.60 0.18 −1.16 −0.39 2.21 1.13 −0.65 −0.39 2.05∆

(`3Cstat

`

)× 1000 2.49 2.16 1.30 1.32 1.15 1.23 1.07 0.94 0.98 0.93 0.99 0.99 1.05 1.22 1.48

∆(`3Csys

`

)× 10000 0.42 1.30 0.23 0.28 0.02 0.01 0.12 0.04 0.24 0.09 0.43 0.19 0.21 0.20 0.25

`3Cfore` × 10000 -0.00 -0.00 −0.01 −0.03 −0.06 −0.09 −0.10 −0.05 0.12 0.48 1.13 2.14 3.61 5.65 8.31

217 GHz `3C` × 1000 5.08 8.99 3.19 3.75 2.15 1.91 1.15 1.47 0.66 2.11 3.15 2.04 −0.15 1.84 3.16∆

(`3Cstat

`

)× 1000 2.49 2.17 1.31 1.34 1.16 1.24 1.07 0.93 0.97 0.90 0.95 0.90 0.89 0.94 1.02

∆(`3Csys

`

)× 10000 1.07 1.90 0.68 0.79 0.46 0.41 0.24 0.31 0.14 0.44 0.65 0.39 0.10 0.29 0.52

`3Cfore` × 10000 -0.00 -0.00 −0.01 −0.03 −0.05 −0.08 −0.09 −0.04 0.10 0.41 0.96 1.83 3.09 4.83 7.10

353 GHz `3C` × 1000 · · · · · · 21.47 21.79 16.56 16.08 14.83 12.76 11.76 15.60 14.98 10.44 11.33 10.67 12.76∆

(`3Cstat

`

)× 1000 · · · · · · 1.88 1.87 1.75 1.81 1.79 1.81 1.96 2.13 2.34 2.55 2.78 3.07 3.44

∆(`3Csys

`

)× 1000 · · · · · · 0.69 0.70 0.53 0.52 0.48 0.41 0.38 0.50 0.48 0.33 0.35 0.32 0.37

`3Cfore` × 10000 · · · · · · −0.01 −0.04 −0.09 −0.13 −0.15 −0.07 0.17 0.68 1.59 3.02 5.10 7.97 11.72

545 GHz `3C` × 100 · · · · · · 28.92 29.05 23.38 20.12 21.37 18.32 15.38 19.36 18.78 12.94 12.67 11.70 7.13∆

(`3Cstat

`

)× 100 · · · · · · 2.09 1.97 1.91 1.87 1.94 2.00 2.12 2.31 2.44 2.60 2.73 2.87 3.04

∆(`3Csys

`

)× 100 · · · · · · 2.95 2.96 2.38 2.05 2.18 1.87 1.57 1.98 1.94 1.36 1.36 1.30 0.88

`3Cfore` × 1 · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · ·

857 GHz `3C` × 1 · · · · · · 12.34 14.32 11.08 8.73 9.00 8.19 7.37 9.85 9.35 4.42 5.75 5.99 4.09∆

(`3Cstat

`

)× 1 · · · · · · 1.40 1.27 1.21 1.14 1.14 1.14 1.18 1.25 1.30 1.36 1.39 1.43 1.48

∆(`3Csys

`

)× 1 · · · · · · 1.26 1.46 1.13 0.89 0.92 0.83 0.75 1.01 0.96 0.46 0.60 0.64 0.45

`3Cfore` × 1 · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · · ·

and interpret the measurement using both angular and frequencyinformation.

Before doing so, we highlight the spectral information con-tained in the signal. It was shown in PER that both the fre-quency spectrum of the CIB mean and fluctuation rms are wellapproximated by the two modified blackbody spectra proposedby Fixsen et al. (1998) and Gispert et al. (2000), with a slightpreference for the latter. We expect our measurements to followthe same spectral energy distribution (SED). Following the pro-cedure outlined in Sect. 5.3.3, we plot in Fig. 10 the best-fit CIBcomponent with either a Fixsen et al. (1998) – dot-dashed blackline – or Gispert et al. (2000) – solid black line – SED. We doso for two `-bins. We can see that, indeed, for a given `-bin,our measurements qualitatively follow the expected CIB spec-trum. Unlike PER, we do not find a preference for the Gispertet al. (2000) shape in our low `-bin and only a slight one inthe high ` bin. It is worth emphasizing that by carrying out across-correlation measurement we can obtain constraints at thelowest frequency, which is usually heavily contaminated by theCMB (see PIR for discussion). This is particularly interesting,because these measurements are simultaneously the most sensi-tive to high-z star formation processes and the most discrepantwith either of the SEDs, i.e., they are both systematically low byabout 0.5σ. The models presented in this section will allow usto use both the spectral dependence and the relative amplitudesof the `-bins that was lost in Fig. 10. We now describe the gen-

eral methodology we will use, before describing our models indetail.

6.1. Model comparison methodology

For the purpose of model fitting, we will utilize both the CIB-lensing cross-spectra measured in this paper and the CIB auto-spectra obtained from PIR. We use the CIB-lensing cross-spectrafor two purposes: to improve constraints on the model parame-ters; and to provide a consistency test of models fit to the CIBauto- and frequency cross-spectra alone. As will be seen in PIR,the cross-spectra of the CIB at different frequencies providepowerful constraints on the CIB emissivity.

We use the log-likelihood,

lnL(p) = −12

∑`

∑i j

[Ci` − Ci

`(p)] (

N i j`

)−1 [C j`− C j

`(p)

]†,

(11)where C and C are the data and theory spectra with parameters p,the i and j indices denote the type of spectra (e.g., 100 GHz× φor 100 GHz× 100 GHz), and N is the covariance matrix that in-cludes both statistical and systematic errors. We make the ap-proximation that the covariance matrix is diagonal, i.e., we treatthe errors for different bins of each auto- and cross-spectrumas being uncorrelated. The small (' 2 %) mask-induced mode-coupling between neighbouring bins supports this approxima-tion. However, calibration and beam errors (which are corre-

15

Page 16: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

lated between the auto- and cross-spectra at a given frequency),as well as the lens normalization error (which is also corre-lated across spectra) are not accounted for in this approxima-tion. In addition, the lens reconstruction has some sensitivity toall modes of the temperature maps, and so different φ modesare correlated to some degree. We also neglect the fact that thecontribution to the error from the CIB signal itself (the orangeline in Fig. 5) is also substantially correlated from frequency tofrequency. However, our evaluation using simulations suggeststhat these effects are too small to significantly affect our pro-cedure. We thus resort to simply adding the beam, calibrationand normalization uncertainties in quadrature to the statisticalerrors. The posterior probability distributions of model param-eters are determined using now standard Markov Chain MonteCarlo techniques (e.g., Knox et al. 2001; Lewis & Bridle 2002).

6.2. Two modelling approaches

The strength of the correlation signal should come as no sur-prise, given our current knowledge of CMB lensing and the CIB.The PER model predicts a high correlation between the CIB andthe lensing potential. As clearly illustrated in Fig. 1, the broadoverlap of the redshift distributions of the CIB with the lensingkernel peaking at z ≈ 2–3 leads to a correlation of 60–80 %. Thisis comparable to our measurements at all of the HFI frequencies,as illustrated in Fig. 13,

In models of the cross-correlation, the underlying propertieswe can probe come from a combination of the lensing potentialand the characteristics of the DSFGs, in particular their emis-sivity and clustering properties. Mostly driven by linear physics,the former is well understood theoretically, as confirmed by re-cent observations (Smith et al. 2008; Hirata et al. 2008; Das et al.2011; van Engelen et al. 2012). Assuming the currently favoredΛCDM cosmology, we can consider it to be known to better than10 % in the multipole range of interest to us, an uncertainty dom-inated by the uncertainty in the normalization of the primordialpower spectrum. Given that this is much smaller than the theoret-ical uncertainties related to DSFGs, we will fix the cosmology tothe currently favoured Planck alone flat ΛCDM model in PlanckCollaboration XVI (2013), and will focus our analysis on themodelling of the DSFGs.

At a given redshift we model the fluctuations in the meanCIB emission, j, as being proportional to the fluctuations in thenumber of galaxies, ng (Haiman & Knox 2000),

δ j ∝ jδng

ng. (12)

With this hypothesis, the goal of the CIB modelling becomestwofold: first, to better understand the statistical properties ofthe dusty galaxy number density, δng; and second, to reconstructthe mean emissivity of the CIB as a function of redshift.

In this paper we will use two different models of the CIBemission. The first model (described in Sect. 6.2.1 and inspiredby Hall et al. 2010) uses a single bias parameter at all frequen-cies with the mean CIB emissivity modelled as a two parameterGaussian. This model is not designed to be physically realistic,and furthermore we will marginalize over an arbitrary amplitudein this case. Nevertheless, we present this simple model to showthat our measurements are quite consistent with broad expecta-tions of the CIB. The second model, described in Sect. 6.2.2, isa natural extension of the Halo Occupation Density (HOD) ap-proach used in PER (see also Penin et al. 2011, and referencestherein). But unlike the results obtained in PER we now use a

100 500 1000 1500 2000`

0.4

0.6

0.8

1.0

1.2

CTφ

`/(C

TT

`Cφφ

`)1/

2

100 GHz143 GHz

217 GHz353 GHz

545 GHz857 GHz

Fig. 13. Cross-correlation coefficients calculated from the modelφ spectrum and best-fit halo model at each frequency. The CIBis a spectacular tracer of CMB lensing, and vice-versa. The datapoints represent the measured cross-correlation divided by thebest-fit auto power spectra models at 545 GHz.

0

1

2

3

4

5

σz

ν=217 GHz ν=353 GHz

0 1 2 3 4 5zc

0

1

2

3

4

5

σz

ν=545 GHz

0 1 2 3 4 5zc

ν=857 GHz

Fig. 14. Marginalized 2-D distributions of zc andσz for the linearbias model, fit to all frequencies simultaneously. The orange dotsindicate the parameter values at the minimum χ2.

single HOD to describe the spectra at all frequencies. This ispossible by allowing for deviations from the Bethermin et al.(2011) model (hereafter B11) that was used to fix the emissivity.

6.2.1. Linear bias model

As a first pass at interpreting our measurement, we will considera redshift independent linear bias model with a simple paramet-ric SED. This model was found to provide a reasonable fit to theauto-spectra in the linear regime in PER. Throughout this paperwe use the Limber approximation, and in this section, since we

16

Page 17: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

are using a linear model, we write the relevant angular spectra as

CXY` =

∫ χ∗

0dχ WX(χ)WY (χ) Pδδ(k = `/χ, χ), (13)

where X and Y are either the CIB at frequency ν or the lensingpotential φ, the integral is over χ, the comoving distance alongthe line of sight, χ∗ is the comoving distance to the last scatteringsurface, Pδδ(k, χ) is the matter power spectrum at distance χ, andthe WX functions are the redshift weights for each of the signalsX:

Wν(χ) = ba jν(χ)χ

;

Wφ(χ) = −3`2 ΩmH2

a

(χ∗ − χ

χ∗χ

).

(14)

Here b is the DSFG bias that we assume to be redshift indepen-dent, a is the scale factor, jν(χ) is the mean CIB emissivity atfrequency ν, as defined in PER, Ωm is the matter density todayin critical density units and H0 is the Hubble parameter today.We use the Hall et al. (2010) model for the CIB kernel, whichtreats the CIB emissivity as a Gaussian in redshift:

jν(χ) ∝ a χ2 exp[−(z − zc)2/2σ2

z

]fν(1+z) (15)

where we use a modified blackbody frequency dependence

fν(1+z) ∝ νβBν(Td). (16)

We fix the dust temperature to Td = 34 K, the spectral in-dex to β = 2 (Hall et al. 2010), and assume a constant biasb. We include a single normalization parameter for j, whichwe marginalize over. Since the normalization and bias param-eters are degenerate in Eq. 13, if we were to only use the mea-sured auto- and cross-spectra this approach would be equivalentto marginalizing over a frequency independent bias parameter.However, we will further constrain our model using the FIRASdata, which breaks this degeneracy. We constrain the zc and σzparameters at each frequency, giving us a total of 13 free param-eters.

For 217–857 GHz, we use the FIRAS measurements of theCIB mean intensity from Lagache et al. (2000) as an additionalconstraint to our model. The mean intensity is simply

Iν =

∫ χ∗

0dχ a jν(χ) . (17)

Using this equation and the measured FIRAS mean and uncer-tainty we calculate a χ2 value and add this to the χ2 in Eq. 11.Since there are no FIRAS constraints at 100 and 143 GHz, aswell as no CIB auto-spectra measurements, and noisier cross-spectra measurements at these frequencies, our constraints forthe 100 and 143 GHz redshift parameters are weaker than for theother parameters.

The linear bias model considers only linear clustering, andso when fitting the auto-spectra we restrict ourselves to ` < 500,where non-linear contributions are negligible. Because we do notconsider the high-` modes, we also neglect the shot-noise con-tribution to the auto-spectra. The best-fit model is shown as thecoloured dashed lines in Fig. 15, with χ2 values of 13.4, 16.8,25.2, 21.8, 9.1, and 9.4 if we break up the χ2 contribution perfrequency from 100 to 857 GHz, leading to an overall χ2 of 95.7for Ndof = 59. We see that the model captures some features ofthe data, but we also have evidence it is significantly missing

100 500 1000 1500 2000

`

0.5

0.0

0.5

1.0

1.5

2.0

`3C

`[µ

K.s

r]

ν=100 GHz (x 100)χ2 =14.08 (15)

100 500 1000 1500 2000

`

0.2

0.0

0.2

0.4

0.6

0.8 ν=143 GHz (x 100)χ2 =16.81 (15)

100 500 1000 1500 2000

`

0.0

0.2

0.4

0.6

0.8

1.0ν=217 GHz (x 100)χ2 =20.22 (15)

100 500 1000 1500 2000

`

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

`3C

`[µ

K.s

r]

ν=353 GHz (x 100)χ2 =7.28 (15)

100 500 1000 1500 2000

`

0

1

2

3

4ν=545 GHz (x 10)χ2 =5.58 (15)

100 500 1000 1500 2000

`

0.0

0.5

1.0

1.5

2.0ν=857 GHz (x 0.1)χ2 =7.98 (15)

Fig. 15. Measured cross-spectra with the best-fit j reconstructionmodel fit to both the CIB auto- and CIB-lensing cross-spectra(solid coloured), and the best-fit linear bias model (dashedcoloured). The χ2 values quoted in each panel are the contribu-tion to the global χ2 from the data in the panel for the halo model,and loosely indicate the goodness of fit (see text for details).The one and two-halo contributions are shown as the dashed andsolid black lines, respectively. A light dashed black horizontalline is indicating the zero level.

200 500 1000 2000

`

0.1

0.5

2.0

`(`+

1)C

TT

`/(

2π)

[µK

2]

ν=217 GHz (x 10−2 )χ2 =4.94 (9)

200 500 1000 2000

`

0.05

0.20

0.50ν=353 GHz (x 10−4 )χ2 =14.53 (9)

200 500 1000 2000

`

0.01

0.04

0.10

`(`+

1)C

TT

`/(

2π)

[µK

2]

ν=545 GHz (x 10−7 )χ2 =1.87 (9)

200 500 1000 2000

`

0.03

0.10

0.30ν=857 GHz (x 10−10 )χ2 =8.81 (9)

Fig. 16. PIR auto-spectra with the best-fit mean emissivity j re-construction model fit for the CIB auto and CIB-lensing cross-spectra (solid coloured). The χ2 values are defined as in Fig. 15.The one and two-halo contributions are shown as the dashed andsolid black lines, respectively, while shot noise is the dot-dashedblack line.

some as well. This is perhaps not surprising given the simplic-ity of the model. The two-dimensional marginal distributions ofthe zc, σz parameters are shown in Fig. 14. Although we allowedfor these parameters to be frequency dependent we note that thepoint zc = 1 and σz = 2.2 is in a region of high probability at allfrequencies, and gives a redshift distribution for the emissivitydensity roughly consistent with our expectations, rising towardz = 1 due to the χ2 term and then only slowly falling off towardhigher redshifts.

17

Page 18: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

Althought it is useful to see the extent to which such a simplemodel can explain our data, we now turn to make a stronger con-nection between the properties of the infrared light and the dis-tribution of the underlying dark matter applicable into the non-linear regime.

6.2.2. An extended halo model based analysis

In this section we use a description of the CIB motivated by thehalo model, which has been used successfully to describe thetransition between the linear and non-linear clustering regimesfor optical galaxies. We use the halo model to attempt to recon-struct the CIB emissivity as a function of redshift. This is an ex-tension of the approach taken in PER, where the modelled CIBemissivity at high redshift was treated as a single bin with theamplitude constrained by the data. The goal of this approach isto isolate the high-redshift contribution to the CIB, which is dif-ficult to probe using observations of individual galaxies, due totheir low brightness. The power of such an approach is furtherdemonstrated in PIR.

We replace the linear bias used in Sect. 6.2.1 with a halomodel and an HOD prescription that assigns galaxies to hostdark matter halos (see PER for references and definitions). Itallows a consistent description of the linear and non-linear partof the galaxy power spectrum and its redshift evolution. Becauseit is built on the clustering of dark matter halos, the halo modelallows us to describe the clustering of DSFGs and the gravita-tional lensing caused by the halos in a consistent way. However,it is important to realise that the HOD prescription was devel-oped to describe stellar mass within dark matter halos – an ap-plication for which it has been thoroughly tested – while herewe are applying it to star formation within halos. The accuracyof this approach needs to be further quantified. However, it pro-vides a good phenomenological description of our data as well asother CIB measurements, but also of other astrophysical probesof the relation between dark matter and light (e.g., Leauthaudet al. (2012); Hikage et al. (2012)).

Unlike the model presented in PER we use a single HODto describe our data at all frequencies. This is made possible byallowing for a deviation from the B11 emissivity model. Notehowever that we will still consider the CIB emissivity to dependonly on redshift and not on the galaxy host halo mass, a simpli-fication highlighted in Shang et al. (2012) that will be relaxedin the PIR model. The emissivity of the CIB is inhomogeneous,due to spatial variations in the number density of galaxies:

δ jνjν

(n , z) =δng

ng(n , z) ≡ δg(n , z). (18)

Here j(n , z) is the CIB emissivity at redshift z with mean j(z),ng(n , z) is the number density of DSFGs with mean ng(z),and δg(n , z) is the DSFG overdensity, with power spectrum⟨δg(k, z) δg(k′, z)∗

⟩= (2π)3 δ(k − k′) Pgg(k, z). We calculate this

power spectrum, including the constituent 1 and 2-halo terms,using the procedure described in appendix C of PER, with theconstraint αsat = 1, a theoretically favoured value (Tinker &Wetzel 2010). We remove the relationship between Msat, a char-acteristic satellite mass scale, and Mmin, the halo mass at whicha halo has a 50 % probability of containing a central galaxy thatwas enforced in PER (i.e., Msat = 3.3Mmin), and allow both Msatand Mmin to vary independently.

At redshift z < 1 we fix the emissivity to the B11 value, but athigher redshift we assume that the emissivity is constant withinz-bins and solve for the amplitude of the bins. Two factors affect

9 10 11 12 13log10

(Mmin/Msun

)9

10

11

12

13

log 1

0( M sat/M

sun

)Fig. 17. Marginalized 2-D distribution of log10 (Mmin/M) andlog10 (Msat/M) for our overall HOD model when the CIB-lensing cross-spectra are combined with the CIB auto-spectraand FIRAS measurements. The orange dot shows the best-fitvalue. The contours correspond to 68%, 95% and 99.7% con-fidence intervals.

the number of bins that we choose. The auto-spectra have a j2dependence, and so if the true value of j has a strong z depen-dence within a bin then the best-fit emissivity in the bin will bedifficult to interpret. The best-fit bin values could be significantlydifferent from those that would be obtained by binning the trueemissivity. However, as more bins are used and the number ofparameters increases, it becomes more difficult to determine thebest-fit parameters and the parameters will be highly correlated.After investigation using simulations, we found that three binswas a good compromise, given the expected slow redshift evo-lution. The bins are defined by: 1 < z ≤ 1.5; 1.5 < z ≤ 3; and3 < z ≤ 7. As in Sect. 6.2.1 we use the FIRAS results at 217–857 GHz to add an integral constraint on the emissivity. The CIBauto and lensing cross-spectra are (Song et al. 2003):

Cνν′

` =

∫dχ Wν(χ)Wν′ (χ) Pgg(k = `/χ, χ);

Cνφ`

=

∫dχ Wν(χ)Wφ(χ) Pδg(k = `/χ, χ) .

(19)

Since we fix j at z < 1, the model spectra consist of a low red-shift part that is independent of the emissivity parameters, anda contribution from z > 1 that is proportional to jν jν′ for theauto-spectra and jν for the lensing cross-spectra.

Overall, the halo-based model contains two halo param-eters that describe the galaxy clustering and are independentof frequency, and three j amplitudes at each frequency, giv-ing a total of 20 parameters for the six frequencies of interestto us. The auto and cross-spectra have a total of 120 `-bins,with four additional FIRAS data points. Solving for the likeli-hood described in Sect. 6.1, gives the best-fit models shown inFigs. 15 and 16 as solid lines. The χ2 values in each panel arethe contribution to the total χ2 from the data within the panel.

18

Page 19: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

The combined reduced χ2 is 102.1 for Ndof = 104, indicat-ing a good fit. The constraints we find on Msat and Mmin areshown in Fig. 17. We force Msat ≥ Mmin in the MCMC fit-ting procedure, with the dashed line in Fig. 17 showing equal-ity. The red cross corresponds to the parameter values that givethe minimum χ2 in the fit, and are log10 (Mmin/M) = 12.18and log10 (Msat/M) = 12.76, which gives Msat/Mmin = 3.80.The mean parameter values are log10 (Mmin/M) = 10.53± 0.62and log10 (Msat/M) = 10.80 ± 0.74. The best-fit value of Mminis consistent with those derived in PER at multiple frequencies,even though we now set αsat = 1 and reconstruct the mean emis-sivity as a function of redshift. The associated mean emissivityparameters are given in Table 3 and displayed in Fig. 18, wherewe also plot the B11 model for reference. As can be seen inFig. 18, we remain consistent with the B11 model in most red-shift bins.

6.3. Interpreting the reconstructed emissivities

We now illustrate one interesting consequence of this measure-ment and show how using the constrained emissivities, jν(z), wecan estimate the star formation rate (SFR) density at differentredshifts. Following Penin et al. (2011), we begin by writing theemissivity as an integral over the galaxy flux densities:

jν(z) =

(a

dχdz

)−1 ∫S ν

d2NdS ν dz

dS ν . (20)

Here S ν is the flux density, and d2N/dS ν dz is the number ofgalaxies per flux element and redshift interval. The galaxies con-tributing to the CIB can be divided into various populations (la-belled as p) based on the galaxy SED, e.g., according to galaxytype or dust temperature:

jν(z) =

(a

dχdz

)−1 ∑p

∫S ν

d2Np

dS ν dzdS ν . (21)

If we define sν as the flux density of an LIR = L source with theSED of a given population, i.e., S ν = sνLIR (with LIR in units ofL), then we can write Eq. 21 as (Penin et al. 2011):

jν(z) =

(a

dχdz

)−1 dVdz

∑p

∫LIR

d2Np

dLIR dVdLIR. (22)

The contribution to the infrared luminosity density from a givenpopulation is

ρIR,p =

∫LIR

d2Np

dLIR dVdLIR. (23)

We assume a simple conversion between LIR and the star forma-tion rate density, ρSFR, using the Kennicutt constant K (Kennicutt1998). Since by definition ρSFR = K

∑p ρIR,p, we can rewrite

Eq. 22 as:

jν(z) = (1 + z) χ2 ρSFR

K

(∑p sν ρIR,p∑

p ρIR,p

), (24)

where the final term in brackets is the effective SED of infraredgalaxies, which we write as sν,eff . We derive these SEDs follow-ing the evolution model of Bethermin et al. (2012) using Magdiset al. (2012) templates. The construction of these effective SEDs

0 1 2 3 4 5 6 7z

0

5

10

15

20

25

30

j[J

y/M

pc/

sr]

ν=100 GHz

0 1 2 3 4 5 6 7z

0

10

20

30

40

50

60ν=143 GHz

0 1 2 3 4 5 6 7z

0

10

20

30

40

50

60

70ν=217 GHz

0 1 2 3 4 5 6 7z

0

50

100

150

200

250

j[J

y/M

pc/

sr]

ν=353 GHz

0 1 2 3 4 5 6 7z

0

200

400

600

800

1000ν=545 GHz

0 1 2 3 4 5 6 7z

0

200

400

600

800

1000

1200

1400

1600ν=857 GHz

Fig. 18. Reconstructed mean emissivity, j, for each frequency asa function of redshift. The solid line at low z and the dashed lineat higher z correspond to the B11 model. The B11 emissivitymodel at z > 1 is not used, and is shown only for reference.The black error bars correspond to the 68% C.L. while the colorshading display the full posterior distribution.

0 50 100 150 200 250j [Jy/Mpc/sr]

0.0

0.2

0.4

0.6

0.8

1.0

1.2

Like

lihood

ν=353 GHz, z=3−7

Fig. 19. Marginalized 1-D distribution of the emissivity in thehigh redshift bin at 353 GHz with (black line) or without (blueline) including the CIB-lensing correlation. Its inclusion helps toconstrain the emissivity at high redshift, transforming an upperlimit into a detection.

will be explained in detail in future work. Finally, we obtain theconversion factor between mean emissivity and SFR density,

ρSFR(z) =K

(1 + z) χ2(z) S ν,eff(z)jν(z) . (25)

Using Eq. 25 we find the coefficients for each of the redshift binsand frequencies used in Table 3.

6.4. Discussion and outlook

In the previous section we described a model that simultaneouslyfits the CIB auto-spectra and the CIB-lensing cross-spectra, at allfrequencies and with a single HOD prescription. Given that we

19

Page 20: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

Table 3. Reconstructed emissivity as a function of redshift and associated star formation rate. At each frequency and for each ofthe three redshift bins the first quantity corresponds to the mean emissivity in the corresponding redshift bin, j(z), in Jy Mpc−1 sr−1,while the second corresponds to the SFR density, ρSFR, in MMpc−3 yr−1.

1 < z ≤ 1.5 1.5 < z ≤ 3 3 < z ≤ 7

j(z) ρSFR j(z) ρSFR j(z) ρSFR

100 GHz . . . . 7.16±5.77 1.96±1.58 3.53±3.05 0.655±0.564 5.49±4.78 0.271±0.236143 GHz . . . . 12.7±9.60 1.37±0.964 6.82±5.46 0.438±0.351 10.5±9.05 0.178±0.153217 GHz . . . . 11.9±6.33 0.310±0.165 17.3±7.23 0.282±0.118 36.6±13.8 0.182±0.068353 GHz . . . . 116±17.1 0.671±0.099 75.5±27.5 0.286±0.104 164±47.3 0.320±0.092545 GHz . . . . 185±106 0.320±0.183 224±148 0.317±0.210 417±251 0.659±0.396857 GHz . . . . 193±139 0.144±0.104 354±212 0.317±0.190 609±359 1.37±0.809

500 1000 1500 2000`

10

0

10

20

30

40

50

`3C

`[M

Jy/s

r]

Fig. 20. Correlation between the lensing potential and the IRISmap at 100 µm using our nominal lens reconstruction. We clearlysee a correlation and estimate the significance to be 9σ, ignoringpossible systematic effects. The solid line represents a simplereasonnable prediction for this signal.

use an emissivity function that is close to the B11 emissivities(to within our uncertainties), we expect predictions of the galaxynumber counts derived from our best-fit emissivity to agree withcurrent estimates (Bethermin et al. 2012). The fact that our mea-surement is consistent with previous models of the CIB lendssupport to our current understanding of its origin. For exam-ple, the characteristic mass scale at which halos host galaxies,Mmin, is consistent with values derived in PER, and is consistentwith, but slightly higher than, the value derived more recentlyin Viero et al. (2012), log10 (Mmin/M) = 9.9 ± 0.5 (although adirect comparison could be misleading given the different modelassumptions). In particular, it is clear that our model has limi-tations, some of which have been partially addressed in recentwork (Shang et al. 2012; Bethermin et al. 2012; De Bernardis &Cooray 2012; Bethermin et al. 2012; Viero et al. 2012; Addisonet al. 2012) and are points of focus in PIR, amongst them themass independence of the emissivity. Another question worthfurther investigation is the dependence of our results on the bin-ning scheme chosen for the emissivity, which will be addressedin a future paper.

Given the consistency of our model with the PER results,the information added by our cross-spectrum measurement isworth quantifying. As an example, we show in Fig. 19 the high-est redshift emissivity bin in the 353 GHz band. Adding the CIB-

lensing cross-spectrum information tightens the constraint on thehigh-redshift part of the emissivity. This statement also holds forthe other frequencies and stems from the fact that the CMB lens-ing kernel peaks at high redshift, making the cross-correlationmore sensitive to the high-redshift CIB signal than the CIB auto-spectrum, as is illustrated in Fig. 1. Although this gain does nottranslate into a substantial improvement in Mmin, it leads to inter-esting constraints on the SFR density, as can be seen in Table 3.

The results at frequencies above 217 GHz each lead toaround 2σ evidence for a non-zero SFR density for 1.5 < z < 3and for 3 < z < 7. The values inferred are consistent with otherprobes of the SFR in these redshift ranges, as compiled for ex-ample in Fig. 1 of Hopkins & Beacom (2006). Assuming thateach frequency is independent, we obtain SFR densities for thethree redshift bins of 0.423 ± 0.123, 0.292 ± 0.138 and 0.226± 0.100 MMpc−3 yr−1, respectively where the errors are 68%C.L.. We note that the j distributions are rather non-Gaussian sothat the 95% C.L. become 0.228, 0.246 and 0.191 respectively.This roughly 2σ detection per bin compares very favourablywith other published measurements. These constraints clearlyillustrate how this particular correlation can be used to betterisolate the high redshift component of the CIB and improve ourconstraints on the star formation rate at high redshift. Such con-strains will improve with future measurements, in particular ifwe can increase the signal-to-noise ratio in our lower frequencychannels, where the high redshift contribution is the greatest.This will likely require an accurate removal of the CMB, ourdominant source of noise at low frequencies. A more thoroughdiscussion of this possibility will be given in PIR.

To fully utilize the richness of the correlation will requiremore studies. Future work could involve using more sophisti-cated halo models specifically designed to model star formationwithin halos, as well as relaxing some of the assumptions madehere, such as the mass independent luminosity function. In addi-tion the use of map-based methods that enable estimates of thegalaxy host halo mass by stacking the lensing potential maps isworth pursuing, as is the extension to other data-sets. For illus-tration purposes, we show in Fig. 20 raw measurements of thecorrelation between our lensing potential map and the IRIS mapat 100 µm that was introduced in Sect. 2.2.2. We use our nominalmask and lens reconstruction, with no H i cleaning performed onthe IRIS map. We clearly see a strong correlation, whose signif-icance we estimate to be 9σ, ignoring any possible systematiceffects. To guide the eye we have added a prediction (not a fit)based on the HOD model presented in Penin et al. (2011). Thefull analysis of this signal is beyond the scope of this paper, butit illustrates possible future uses of the lensing potential map. Inthis case the IRIS wavelength range will help us to isolate thelow-redshift contribution to the CIB.

20

Page 21: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

To conclude, we have presented the first measurement ofthe correlation between lensing of the CMB and the CIB.Planck’s unprecedented full-sky, multi-frequency, deep surveyenables us to make an internal measurement of this correlation.Measurements with high statistical significance are obtained,even after accounting for possible systematic errors. The highdegree of correlation that is measured (around 80 %) allows forunprecedented visualization of lensing of the CMB and holdsgreat promise for new CIB and CMB focused science. CMBlensing appears promising as a probe of the origin of the CIB,while the CIB is now established as an ideal tracer of CMB lens-ing.

Acknowledgements. Based on observations obtained with Planck (http://www.esa.int/Planck), an ESA science mission with instruments and contri-butions directly funded by ESA Member States, NASA, and Canada. The devel-opment of Planck has been supported by: ESA; CNES and CNRS/INSU-IN2P3-INP (France); ASI, CNR, and INAF (Italy); NASA and DoE (USA); STFC andUKSA (UK); CSIC, MICINN and JA (Spain); Tekes, AoF and CSC (Finland);DLR and MPG (Germany); CSA (Canada); DTU Space (Denmark); SER/SSO(Switzerland); RCN (Norway); SFI (Ireland); FCT/MCTES (Portugal); andPRACE (EU). A description of the Planck Collaboration and a list of itsmembers, including the technical or scientific activities in which they havebeen involved, can be found at http://www.sciops.esa.int/index.php?project=planck&page=Planck_Collaboration. We acknowledge the useof the HEALPix package, and the LAMBDA archive (http://lambda.gsfc.nasa.gov).

ReferencesAddison, G. E., Dunkley, J., & Bond, J. R., Constraining thermal dust emis-

sion in distant galaxies with number counts and angular power spectra. 2012,ArXiv e-prints, arXiv:1210.6697

Amblard, A., Cooray, A., Serra, P., et al., Submillimetre galaxies reside in darkmatter haloes with masses greater than 3×1011 solar masses. 2011, Nature,470, 510, arXiv:1101.1080

Arnal, E. M., Bajaja, E., Larrarte, J. J., Morras, R., & Poppel, W. G. L., A highsensitivity Hi survey of the sky at delta ≤ = -25o. 2000, A&AS, 142, 35

Babich, D. & Pierpaoli, E., Point Source Contamination in CMB Non-Gaussianity Analyses. 2008, Phys.Rev., D77, 123011, arXiv:0803.1161

Bajaja, E., Arnal, E. M., Larrarte, J. J., et al., A high sensitivity HI survey ofthe sky at δ ≤ -25 . Final data release. 2005, A&A, 440, 767, arXiv:astro-ph/0504136

Bersanelli, M., Mandolesi, N., Butler, R. C., et al., Planck pre-launch status:Design and description of the Low Frequency Instrument. 2010, A&A, 520,A4+, arXiv:1001.3321

Bethermin, M., Daddi, E., Magdis, G., et al., A unified empirical model for in-frared galaxy counts based on observed physical evolution of distant galaxies.2012, arXiv:1208.6512

Bethermin, M., Dole, H., Lagache, G., Le Borgne, D., & Penin, A., Modeling theevolution of infrared galaxies: a parametric backward evolution model. 2011,A&A, 529, A4, arXiv:1010.1150

Bethermin, M., Dore, O., & Lagache, G., Where stars form and live at highredshift: clues from the infrared. 2012, A&A, 537, L5, arXiv:1201.0546

Bethermin, M., Floc’h, E. L., Ilbert, O., et al., HerMES: deep number counts at250, 350, and 500 microns in the COSMOS and GOODS-N fields and thebuild-up of the cosmic infrared background. 2012, arXiv:1203.1925

Blanchard, A. & Schneider, J., Gravitational lensing effect on the fluctuations ofthe cosmic background radiation. 1987, A&A, 184, 1

Blandford, R. D. & Jaroszynski, M., Gravitational distortion of the images ofdistant radio sources in an inhomogeneous universe. 1981, ApJ, 246, 1

Bleem, L. E., van Engelen, A., Holder, G. P., et al., A Measurement of theCorrelation of Galaxy Surveys with CMB Lensing Convergence Maps fromthe South Pole Telescope. 2012, ApJ, 753, L9, arXiv:1203.4808

Boulanger, F., Abergel, A., Bernard, J. P., et al., The dust/gas correlation at highGalactic latitude. 1996, A&A, 312, 256

Boulanger, F. & Perault, M., Diffuse infrared emission from the galaxy. I - Solarneighborhood. 1988, ApJ, 330, 964

Crawford, T., Schaffer, K., Bhattacharya, S., et al., A measurement of thesecondary-CMB and millimeter-wave-foreground bispectrum using 800square degrees of South Pole Telescope data. 2013, arXiv:1303.3535

Das, S., Sherwin, B. D., Aguirre, P., et al., Detection of the Power Spectrumof Cosmic Microwave Background Lensing by the Atacama CosmologyTelescope. 2011, Physical Review Letters, 107, 021301, arXiv:1103.2124

De Bernardis, F. & Cooray, A., A Conditional Luminosity Function Modelof the Cosmic Far-Infrared Background Anisotropy Power Spectrum. 2012,arXiv:1206.1324

Dunkley, J., Hlozek, R., Sievers, J., et al., The Atacama Cosmology Telescope:Cosmological Parameters from the 2008 Power Spectra. 2011, Astrophys.J.,739, 52, arXiv:1009.0866

Finkbeiner, D. P., Davis, M., & Schlegel, D. J., Extrapolation of GalacticDust Emission at 100 Microns to Cosmic Microwave Background RadiationFrequencies Using FIRAS. 1999, ApJ, 524, 867, arXiv:astro-ph/9905128

Fixsen, D. J., Dwek, E., Mather, J. C., Bennett, C. L., & Shafer, R. A., TheSpectrum of the Extragalactic Far-Infrared Background from the COBEFIRAS Observations. 1998, ApJ, 508, 123, arXiv:astro-ph/9803021

Gispert, R., Lagache, G., & Puget, J. L., Implications of the cosmic in-frared background for light production and the star formation history in theUniverse. 2000, A&A, 360, 1

Goldberg, D. M. & Spergel, D. N., Microwave background bispectrum. 2.A probe of the low redshift universe. 1999, Phys.Rev., D59, 103002,arXiv:astro-ph/9811251

Haiman, Z. & Knox, L., Correlations in the far infrared background. 2000,Astrophys.J., 530, 124, arXiv:astro-ph/9906399

Hall, N., Knox, L., Reichardt, C., et al., Angular Power Spectra of the MillimeterWavelength Background Light from Dusty Star-forming Galaxies with theSouth Pole Telescope. 2010, Astrophys.J., 718, 632, arXiv:0912.4315

Hall, N. R., Keisler, R., Knox, L., et al., Angular Power Spectra of theMillimeter-wavelength Background Light from Dusty Star-forming Galaxieswith the South Pole Telescope. 2010, ApJ, 718, 632, arXiv:0912.4315

Hand, N., Addison, G. E., Aubourg, E., et al., Evidence of GalaxyCluster Motions with the Kinematic Sunyaev-Zel’dovich Effect. 2012,Phys.Rev.Lett., 109, 041101, arXiv:1203.4219

Hanson, D., Challinor, A., Efstathiou, G., & Bielewicz, P., CMB tem-perature lensing power reconstruction. 2011, Phys.Rev., D83, 043005,arXiv:1008.4403

Hartmann, D. & Burton, W. B. 1997, Atlas of Galactic Neutral HydrogenHauser, M. G., Arendt, R. G., Kelsall, T., et al., The COBE Diffuse Infrared

Background Experiment Search for the Cosmic Infrared Background. I.Limits and Detections. 1998, Astrophys.J., 508, 25

Hikage, C., Mandelbaum, R., Takada, M., & Spergel, D. N., Where are theLuminous Red Galaxies (LRGs)? Using correlation measurements and lens-ing to relate LRGs to dark matter halos. 2012, arXiv:1211.1009

Hirata, C. M., Ho, S., Padmanabhan, N., Seljak, U., & Bahcall, N. A., Correlationof CMB with large-scale structure: II. Weak lensing. 2008, Phys. Rev., D78,043520, arXiv:0801.0644

Hivon, E., Gorski, K., Netterfield, C., et al., Master of the cosmic microwavebackground anisotropy power spectrum: a fast method for statistical anal-ysis of large and complex cosmic microwave background data sets. 2002,Astrophys.J., 567, 2, arXiv:astro-ph/0105302

Hopkins, A. M. & Beacom, J. F., On the normalisation of the cosmic star forma-tion history. 2006, Astrophys.J., 651, 142, arXiv:astro-ph/0601463

Kalberla, P. M., Burton, W., Hartmann, D., et al., The Leiden/Argentine/Bonn(LAB) survey of Galactic HI: Final data release of the combinedLDS and IAR surveys with improved stray-radiation corrections. 2005,Astron.Astrophys., 440, 775, arXiv:astro-ph/0504140

Kennicutt, R. C., Star Formation in Galaxies Along the Hubble Sequence. 1998,ARAA, 36, 189

Knox, L., Christensen, N., & Skordis, C., The age of the universe and the cosmo-logical constant determined from cosmic microwave background anisotropymeasurements. 2001, Astrophys.J., 563, L95, arXiv:astro-ph/0109232

Lacasa, F., Aghanim, N., Kunz, M., & Frommert, M., Characterization of thenon-Gaussianity of radio and IR point sources at CMB frequencies. 2012,MNRAS, 421, 1982, arXiv:1107.2251

Lagache, G., Abergel, A., Boulanger, F., & Puget, J. L., The interstellar cold dustobserved by COBE. 1998, A&A, 333, 709

Lagache, G., Bavouzet, N., Fernandez-Conde, N., et al., Correlated Anisotropiesin the Cosmic Far-Infrared Background Detected by the Multiband ImagingPhotometer for Spitzer: Constraint on the Bias. 2007, ApJ, 665, L89,arXiv:0707.2443

Lagache, G., Haffner, L. M., Reynolds, R. J., & Tufte, S. L., Evidence for dustemission in the Warm Ionised Medium sing WHAM data. 2000, A&A, 354,247

Lamarre, J., Puget, J., Ade, P. A. R., et al., Planck pre-launch status: The HFIinstrument, from specification to actual performance. 2010, A&A, 520, A9+

Land, K. & Slosar, A., Correlation between galactic HI and the cosmic mi-crowave background. 2007, Phys. Rev. D, 76, 087301, arXiv:0706.1703

Leahy, J. P., Bersanelli, M., D’Arcangelo, O., et al., Planck pre-launch status:Expected LFI polarisation capability. 2010, A&A, 520, A8+

Leauthaud, A., George, M. R., Behroozi, P. S., et al., The integrated stellar con-tent of dark matter halos. 2012, Astrophys.J., 746, 95, arXiv:1109.0010

21

Page 22: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

Lewis, A. & Bridle, S., Cosmological parameters from CMB and other data:A Monte Carlo approach. 2002, Phys.Rev., D66, 103511, arXiv:astro-ph/0205436

Lewis, A. & Challinor, A., Weak gravitational lensing of the cmb. 2006,Phys.Rept., 429, 1, arXiv:astro-ph/0601594

Lewis, A., Challinor, A., & Hanson, D., The shape of the CMB lensing bispec-trum. 2011, JCAP, 1103, 018, arXiv:1101.2234

Lewis, A., Challinor, A., & Lasenby, A., Efficient computation of CMBanisotropies in closed FRW models. 2000, Astrophys.J., 538, 473,arXiv:astro-ph/9911177

Magdis, G. E., Daddi, E., Bethermin, M., et al., The Evolving InterstellarMedium of Star Forming Galaxies Since z=2 as Probed by Their InfraredSpectral Energy Distributions. 2012, Astrophys.J., 760, 6, arXiv:1210.1035

Mandolesi, N., Bersanelli, M., Butler, R. C., et al., Planck pre-launch status: ThePlanck-LFI programme. 2010, A&A, 520, A3+, arXiv:1001.2657

Mennella, A., Butler, R. C., Curto, A., et al., Planck early results. III. First assess-ment of the Low Frequency Instrument in-flight performance. 2011, A&A,536, A3

Miville-Deschenes, M.-A. & Lagache, G., IRIS: A New Generation of IRASMaps. 2005, ApJS, 157, 302, arXiv:astro-ph/0412216

Nozawa, S., Itoh, N., Kawana, Y., & Kohyama, Y., Relativistic Correctionsto the Sunyaev-Zeldovich Effect for Clusters of Galaxies. IV. AnalyticFitting Formula for the Numerical Results. 2000, ApJ, 536, 31, arXiv:astro-ph/9912008

Okamoto, T. & Hu, W., CMB Lensing Reconstruction on the Full Sky. 2003,Phys. Rev., D67, 083002, arXiv:astro-ph/0301031

Osborne, S., Hanson, D., & Dore, O., Unresolved Foreground Contamination ofCMB Lens Reconstruction. 2013, In prep.

Penin, A., Dore, O., Lagache, G., & Bethermin, M., Modeling galaxies clusteringin the Cosmic Infrared Background. 2011, A&A, submitted

Planck Collaboration. 2011, The Explanatory Supplement to the Planck EarlyRelease Compact Source Catalogue (ESA)

Planck Collaboration, Planck intermediate results. The multi-frequency andmulti-scale CIB anisotropy. 2013, A&A in preparation

Planck Collaboration I, Planck early results. I. The Planck mission. 2011, A&A,536, A1

Planck Collaboration I, Planck 2013 results: Overview of Planck Products andScientific Results (p01). 2013, In preparation

Planck Collaboration II, Planck early results. II. The thermal performance ofPlanck. 2011, A&A, 536, A2

Planck Collaboration VI, Planck 2013 results: The High Frequency InstrumentData Processing (p03). 2013, In preparation

Planck Collaboration VII, Planck early results. VII. The Early Release CompactSource Catalogue. 2011, A&A, 536, A7

Planck Collaboration VII, Planck 2013 results: HFI Time transfer functions andbeams (p03c). 2013, In preparation

Planck Collaboration VIII, Planck 2013 results: HFI Photometric calibration andMap-making (p03f). 2013, In preparation

Planck Collaboration XII, Planck 2013 results: Separation of the Planck skymaps into physical components (p06). 2013, In preparation

Planck Collaboration XIII, Planck early results. XIII. Statistical properties ofextragalactic radio sources in the Planck Early Release Compact SourceCatalogue. 2011, A&A, 536, A13

Planck Collaboration XVI, Planck 2013 results: Cosmological parameters fromPlanck combined with astrophysical data sets (p11). 2013, In preparation

Planck Collaboration XVII, Planck 2013 results: Gravitational lensing by large-scale structures observed by Planck (p12). 2013, In preparation

Planck Collaboration XVIII, Planck early results. XVIII. The power spectrum ofcosmic infrared background anisotropies. 2011, A&A, 536, A18

Planck Collaboration XXIV, Planck early results. XXIV. Dust in the diffuse in-terstellar medium and the Galactic halo. 2011, A&A, 536, A24

Planck HFI Core Team, Planck early results, IV. First assessment of the HighFrequency Instrument in-flight performance. 2011a, A&A, 536, A4

Planck HFI Core Team, Planck early results. VI. The High Frequency Instrumentdata processing. 2011b, A&A, 536, A6

Puget, J. L., Abergel, A., Bernard, J. P., et al., Tentative detection of a cosmicfar-infrared background with COBE. 1996, A&A, 308, L5

Reichardt, C., Shaw, L., Zahn, O., et al., A measurement of secondary cosmicmicrowave background anisotropies with two years of South Pole Telescopeobservations. 2012, Astrophys.J., 755, 70, arXiv:1111.0932

Rosset, C., Tristram, M., Ponthieu, N., et al., Planck pre-launch status: HighFrequency Instrument polarization calibration. 2010, A&A, 520, A13+,arXiv:1004.2595

Sachs, R. & Wolfe, A., Perturbations of a cosmological model and angular vari-ations of the microwave background. 1967, Astrophys.J., 147, 73

Seljak, U. & Zaldarriaga, M., Direct signature of evolving gravitational po-tential from cosmic microwave background. 1999, Phys.Rev., D60, 043504,arXiv:astro-ph/9811123

Shang, C., Haiman, Z., Knox, L., & Oh, S. P., Improved models for cosmic in-frared background anisotropies: new constraints on the infrared galaxy popu-lation. 2012, MNRAS, 421, 2832, arXiv:1109.1522

Sherwin, B. D., Das, S., Hajian, A., et al., The Atacama Cosmology Telescope:Cross-correlation of cosmic microwave background lensing and quasars.2012, Phys. Rev. D, 86, 083006, arXiv:1207.4543

Smith, K. M., Zahn, O., & Dore, O., Detection of Gravitational Lensingin the Cosmic Microwave Background. 2007, Phys. Rev., D76, 043510,arXiv:0705.3980

Smith, K. M., Zahn, O., & Dore, O., Detection of gravitational lensingin the cosmic microwave background. 2007, Phys. Rev. D, 76, 043510,arXiv:0705.3980

Smith, K. M. et al., CMBPol Mission Concept Study: Gravitational Lensing.2008, arXiv:0811.3916

Song, Y.-S., Cooray, A., Knox, L., & Zaldarriaga, M., The Far-infraredbackground correlation with CMB lensing. 2003, Astrophys.J., 590, 664,arXiv:astro-ph/0209001

Sunyaev, R. A. & Zeldovich, I. B., The velocity of clusters of galaxies relativeto the microwave background - The possibility of its measurement. 1980,MNRAS, 190, 413

Sunyaev, R. A. & Zeldovich, Y. B., Small-Scale Fluctuations of Relic Radiation.1970, Ap&SS, 7, 3

Tauber, J. A., Mandolesi, N., Puget, J., et al., Planck pre-launch status: ThePlanck mission. 2010, A&A, 520, A1+

Tinker, J. L. & Wetzel, A. R., What does Clustering Tell us About the Buildupof the Red Sequence? 2010, ApJ, 719, 88, arXiv:0909.1325

van Engelen, A., Keisler, R., Zahn, O., et al., A Measurement of GravitationalLensing of the Microwave Background Using South Pole Telescope Data.2012, ApJ, 756, 142, arXiv:1202.0546

Viero, M., Wang, L., Zemcov, M., et al., HerMES: Cosmic Infrared BackgroundAnisotropies and the Clustering of Dusty Star-Forming Galaxies. 2012,arXiv:1208.5049

Viero, M. P., Ade, P. A. R., Bock, J. J., et al., BLAST: Correlations in the CosmicFar-Infrared Background at 250, 350, and 500 µm Reveal Clustering of Star-forming Galaxies. 2009, ApJ, 707, 1766, arXiv:0904.1200

Wilson, M. J., Sherwin, B. D., Hill, J. C., et al., Atacama Cosmology Telescope:A measurement of the thermal Sunyaev-Zel’dovich effect using the skew-ness of the CMB temperature distribution. 2012, Phys. Rev. D, 86, 122005,arXiv:1203.6633

Zacchei, A., Maino, D., Baccigalupi, C., et al., Planck early results. V. The LowFrequency Instrument data processing. 2011, A&A, 536, A5

Appendix A: Statistical Errors

In this section we compare six different methods to estimate ourstatistical errors. This comparison is used to validate our claim(presented in the main text) that we can obtain our errors fromthe naive analytical errors calculated from the data, i.e., method1 below. The six different methods we compare are:

1. The naive, Gaussian, analytical errors estimated from thedata through the measured total power of the T and φ fields,respectively CTT

` and Cφφ`

and the model cross-spectrum CTφ`

.2. As above but instead of the data maps we use one of our sim-

ulations of the CIB and CMB temperature maps described inSect. 3.4. The lens reconstruction is obtained from the sim-ulated maps using the same procedure that we use for thedata.

3. The scatter directly within individual `-bins in the data-determined cross-spectrum.

4. As above but the scatter is measured in each `-bin for eachsimulation realization, and the errors are averaged over 100realizations.

5. The scatter of the bins is calculated using the cross-spectrameasured from our simulated maps. This is a direct mea-surement of the statistical error we require (to the extent thatour simulations are realistic), and will differ from the scat-ter within the `-bins, for example due to noise correlationsbetween different multipoles of the cross-spectrum.

6. The error in the cross-spectrum of the reconstructed lensingpotential in simulated maps with the measured temperature

22

Page 23: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

0.7

0.8

0.9

1.0

1.1

1.2

1.3

Err

or

rati

o

ν=100 GHz ν=143 GHz ν=217 GHz

100 500 1000 1500`

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

1.2

Err

or

rati

o

ν=353 GHz

100 500 1000 1500`

ν=545 GHz

100 500 1000 1500 2000`

ν=857 GHz

Fig. A.1. Ratio of various error estimation procedures to the errors obtained with the data-based analytical estimate. At each fre-quency the numerator is given by: (i) the scatter within an `-bin in simulations (solid black line); (ii) the scatter within an `-binin the data (solid dashed black line); (iii) the scatter of bins across simulated realizations (solid coloured line); (iv) the analyticalerrors calculated from the simulations (dashed coloured line); (v) the scatter across realizations for the cross-correlation betweenthe simulated temperature map and the lensing potential reconstructed from the data (coloured dot-dashed line). The grey envelopeis the precision of the simulated errors expected from 100 simulations (shown as a spread around unity).

maps. This will only give part of the contribution to the errorsince the temperature maps are fixed, but it is still a usefulcross-check.

In Fig. A.1 we show a comparison of the errors found fromour six measurement methods. The precision achievable with100 simulations is indicated by the grey envelope. We showthe errors in each ` bin from the different methods divided bythe data-based analytical estimate. To discuss the implicationsof these results we shall focus on the 100 GHz panel first. Thescatter measured within an ` bin is fairly consistent in the sim-ulations (method 4, black solid line) and in the data (method3, black dashed line) giving us confidence in our simple simula-tion procedure. This rules out important systematic contributionsand shows that our signal is mostly Gaussian, as expected. Notethat the consistency with the simulations is not surprising, sinceat low frequencies we are dominated by CMB and instrumen-tal noise, which are well understood. In addition, the fact thatthe analytical errors calculated on the simulations (method 2,coloured dashed line) are mostly within the grey shaded region,and are therefore close to the analytical errors calculated fromthe data (method 1), gives us further confidence in the simula-tions. To the extent that the simulations are accurate, the scatterof the `-bins in simulations (solid coloured line) is the error thatwe require. The fact that it is essentially all within the shadedenvelope means that this method gives errors that are consistentwith the analytical errors measured using the data, justifying ournominal choice for calculating the errors at low frequencies.

However, comparing the black lines with the coloured linesclearly indicates that in the data and simulations obtaining theerror bars by measuring the scatter within the `-bins leads toan underestimation of the errors by approximately 20 %. Given

the fact that this difference is observed in both the data and thesimulations, we exclude any instrumental systematic effect as itscause and explain it as being due to noise correlations withinthe `-bins. Such a correlation is expected, since most of the lensreconstruction signal in the `-range of interest to us comes frommodes in the CMB map within a relatively narrow range at ` '1500, and so multipoles in the lens reconstruction are correlated.We have also checked that the mask induced `-bin coupling isnegligible, given the bin width we have chosen, and is alwayssmaller than 2 %.

All of these conclusions remain valid up to 353 GHz.However, at 545 and 857 GHz, we see by looking at the er-rors measured using simulations (solid black for method 4, solidcoloured for method 5, and dashed coloured for method 2) thatthe errors deduced from the analytical estimates measured fromthe data are substantially higher than those we measure in thesimulations. This is easily explained through the fact that we areomitting any foreground emission in our simulations. The rel-ative contribution of Galactic foreground emission is higher atlow `, which is expected because the Galactic cirrus emissionhas a steep power spectrum. Overall, the amplitude of this con-tribution is also consistent with what is seen in Fig. 5.

Since the scatter within the `-bins measured in simulations(black dashed line) is about 20 % lower than the data-based ana-lytical estimates at all frequencies, we use the data-based analyti-cal estimates as the basis for our statistical errors. We could alter-natively scale the scatter-determined errors by 20 % and obtainconsistent results. This approach accounts for the foregroundemission seen at 545 GHz and 857 GHz, but will in practice ne-glect the contribution to the errors from the non-Gaussian partof the foregrounds. However, we show in Sect. 5.3.5 that thiscontribution is small enough that we can ignore it.

23

Page 24: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

The remaining method to discuss is obtained from the cross-spectrum of the reconstructed lensing potential in simulatedmaps with the data-measured temperature maps (method 6,coloured dashed-dotted line in Fig. A.1). At 545 and 857 GHzthe CIB signal is dominant over a large ` range, and so the er-ror obtained from this method is equal to the “signal” terms inEq. 4, which are the orange and green lines in Fig. 5. Thesetwo lines make up a significant fraction of the total error andprovide a reasonable approximation to the true error at high-`.However, at low-` where Galactic emission is important, and at100–353 GHz where the CMB and instrumental noise are thelargest components, the orange and green curves do not accu-rately describe the total error. We can see from Fig. A.1 that theerrors obtained using this method are close to the errors mea-sured using the other techniques. However, this method will un-derestimate the true errors, since the scatter in the CMB andnoise components is neglected.

Note that the results presented in Fig. A.1 are all computedusing the 40 % Galaxy mask, but we have checked that they holdwhen using the 20 % and 60 % Galaxy masks (which are dis-cussed in detail in Sect. 5.2) and that the results show the appro-priate fsky scaling.

1 APC, AstroParticule et Cosmologie, Universite Paris Diderot,CNRS/IN2P3, CEA/lrfu, Observatoire de Paris, Sorbonne ParisCite, 10, rue Alice Domon et Leonie Duquet, 75205 Paris Cedex13, France

2 Aalto University Metsahovi Radio Observatory, Metsahovintie 114,FIN-02540 Kylmala, Finland

3 African Institute for Mathematical Sciences, 6-8 Melrose Road,Muizenberg, Cape Town, South Africa

4 Agenzia Spaziale Italiana Science Data Center, c/o ESRIN, viaGalileo Galilei, Frascati, Italy

5 Agenzia Spaziale Italiana, Viale Liegi 26, Roma, Italy6 Astrophysics Group, Cavendish Laboratory, University of

Cambridge, J J Thomson Avenue, Cambridge CB3 0HE, U.K.7 Astrophysics & Cosmology Research Unit, School of Mathematics,

Statistics & Computer Science, University of KwaZulu-Natal,Westville Campus, Private Bag X54001, Durban 4000, SouthAfrica

8 Atacama Large Millimeter/submillimeter Array, ALMA SantiagoCentral Offices, Alonso de Cordova 3107, Vitacura, Casilla 7630355, Santiago, Chile

9 CITA, University of Toronto, 60 St. George St., Toronto, ON M5S3H8, Canada

10 CNRS, IRAP, 9 Av. colonel Roche, BP 44346, F-31028 Toulousecedex 4, France

11 California Institute of Technology, Pasadena, California, U.S.A.12 Centre for Theoretical Cosmology, DAMTP, University of

Cambridge, Wilberforce Road, Cambridge CB3 0WA U.K.13 Centro de Estudios de Fısica del Cosmos de Aragon (CEFCA),

Plaza San Juan, 1, planta 2, E-44001, Teruel, Spain14 Computational Cosmology Center, Lawrence Berkeley National

Laboratory, Berkeley, California, U.S.A.15 Consejo Superior de Investigaciones Cientıficas (CSIC), Madrid,

Spain16 DSM/Irfu/SPP, CEA-Saclay, F-91191 Gif-sur-Yvette Cedex,

France17 DTU Space, National Space Institute, Technical University of

Denmark, Elektrovej 327, DK-2800 Kgs. Lyngby, Denmark18 Departement de Physique Theorique, Universite de Geneve, 24,

Quai E. Ansermet,1211 Geneve 4, Switzerland19 Departamento de Fısica Fundamental, Facultad de Ciencias,

Universidad de Salamanca, 37008 Salamanca, Spain20 Departamento de Fısica, Universidad de Oviedo, Avda. Calvo

Sotelo s/n, Oviedo, Spain21 Department of Astronomy and Astrophysics, University of

Toronto, 50 Saint George Street, Toronto, Ontario, Canada

22 Department of Astrophysics/IMAPP, Radboud UniversityNijmegen, P.O. Box 9010, 6500 GL Nijmegen, The Netherlands

23 Department of Electrical Engineering and Computer Sciences,University of California, Berkeley, California, U.S.A.

24 Department of Physics & Astronomy, University of BritishColumbia, 6224 Agricultural Road, Vancouver, British Columbia,Canada

25 Department of Physics and Astronomy, Dana and David DornsifeCollege of Letter, Arts and Sciences, University of SouthernCalifornia, Los Angeles, CA 90089, U.S.A.

26 Department of Physics and Astronomy, University CollegeLondon, London WC1E 6BT, U.K.

27 Department of Physics, Gustaf Hallstromin katu 2a, University ofHelsinki, Helsinki, Finland

28 Department of Physics, Princeton University, Princeton, NewJersey, U.S.A.

29 Department of Physics, University of California, One ShieldsAvenue, Davis, California, U.S.A.

30 Department of Physics, University of California, Santa Barbara,California, U.S.A.

31 Department of Physics, University of Illinois atUrbana-Champaign, 1110 West Green Street, Urbana, Illinois,U.S.A.

32 Dipartimento di Fisica e Astronomia G. Galilei, Universita degliStudi di Padova, via Marzolo 8, 35131 Padova, Italy

33 Dipartimento di Fisica e Scienze della Terra, Universita di Ferrara,Via Saragat 1, 44122 Ferrara, Italy

34 Dipartimento di Fisica, Universita La Sapienza, P. le A. Moro 2,Roma, Italy

35 Dipartimento di Fisica, Universita degli Studi di Milano, ViaCeloria, 16, Milano, Italy

36 Dipartimento di Fisica, Universita degli Studi di Trieste, via A.Valerio 2, Trieste, Italy

37 Dipartimento di Fisica, Universita di Roma Tor Vergata, Via dellaRicerca Scientifica, 1, Roma, Italy

38 Discovery Center, Niels Bohr Institute, Blegdamsvej 17,Copenhagen, Denmark

39 Dpto. Astrofısica, Universidad de La Laguna (ULL), E-38206 LaLaguna, Tenerife, Spain

40 European Southern Observatory, ESO Vitacura, Alonso de Cordova3107, Vitacura, Casilla 19001, Santiago, Chile

41 European Space Agency, ESAC, Planck Science Office, Caminobajo del Castillo, s/n, Urbanizacion Villafranca del Castillo,Villanueva de la Canada, Madrid, Spain

42 European Space Agency, ESTEC, Keplerlaan 1, 2201 AZNoordwijk, The Netherlands

43 Finnish Centre for Astronomy with ESO (FINCA), University ofTurku, Vaisalantie 20, FIN-21500, Piikkio, Finland

44 Helsinki Institute of Physics, Gustaf Hallstromin katu 2, Universityof Helsinki, Helsinki, Finland

45 INAF - Osservatorio Astronomico di Padova, Vicolodell’Osservatorio 5, Padova, Italy

46 INAF - Osservatorio Astronomico di Roma, via di Frascati 33,Monte Porzio Catone, Italy

47 INAF - Osservatorio Astronomico di Trieste, Via G.B. Tiepolo 11,Trieste, Italy

48 INAF/IASF Bologna, Via Gobetti 101, Bologna, Italy49 INAF/IASF Milano, Via E. Bassini 15, Milano, Italy50 INFN, Sezione di Bologna, Via Irnerio 46, I-40126, Bologna, Italy51 INFN, Sezione di Roma 1, Universita di Roma Sapienza, Piazzale

Aldo Moro 2, 00185, Roma, Italy52 IPAG: Institut de Planetologie et d’Astrophysique de Grenoble,

Universite Joseph Fourier, Grenoble 1 / CNRS-INSU, UMR 5274,Grenoble, F-38041, France

53 IUCAA, Post Bag 4, Ganeshkhind, Pune University Campus, Pune411 007, India

54 Imperial College London, Astrophysics group, BlackettLaboratory, Prince Consort Road, London, SW7 2AZ, U.K.

55 Infrared Processing and Analysis Center, California Institute ofTechnology, Pasadena, CA 91125, U.S.A.

24

Page 25: 2013 results. XVIII. Gravitational lensing-infrared ... · PDF fileAstronomy & Astrophysics manuscript no. paper˙cib˙lensing c ESO 2013 March 22, 2013 Planck 2013 results. XVIII

Planck Collaboration: Planck 2013 results. XVIII. Gravitational lensing-infrared background correlation

56 Institut Neel, CNRS, Universite Joseph Fourier Grenoble I, 25 ruedes Martyrs, Grenoble, France

57 Institut Universitaire de France, 103, bd Saint-Michel, 75005,Paris, France

58 Institut d’Astrophysique Spatiale, CNRS (UMR8617) UniversiteParis-Sud 11, Batiment 121, Orsay, France

59 Institut d’Astrophysique de Paris, CNRS (UMR7095), 98 bisBoulevard Arago, F-75014, Paris, France

60 Institute for Space Sciences, Bucharest-Magurale, Romania61 Institute of Astronomy and Astrophysics, Academia Sinica, Taipei,

Taiwan62 Institute of Astronomy, University of Cambridge, Madingley Road,

Cambridge CB3 0HA, U.K.63 Institute of Theoretical Astrophysics, University of Oslo, Blindern,

Oslo, Norway64 Instituto de Astrofısica de Canarias, C/Vıa Lactea s/n, La Laguna,

Tenerife, Spain65 Instituto de Fısica de Cantabria (CSIC-Universidad de Cantabria),

Avda. de los Castros s/n, Santander, Spain66 Jet Propulsion Laboratory, California Institute of Technology, 4800

Oak Grove Drive, Pasadena, California, U.S.A.67 Jodrell Bank Centre for Astrophysics, Alan Turing Building,

School of Physics and Astronomy, The University of Manchester,Oxford Road, Manchester, M13 9PL, U.K.

68 Kavli Institute for Cosmology Cambridge, Madingley Road,Cambridge, CB3 0HA, U.K.

69 LAL, Universite Paris-Sud, CNRS/IN2P3, Orsay, France70 LERMA, CNRS, Observatoire de Paris, 61 Avenue de

l’Observatoire, Paris, France71 Laboratoire AIM, IRFU/Service d’Astrophysique - CEA/DSM -

CNRS - Universite Paris Diderot, Bat. 709, CEA-Saclay, F-91191Gif-sur-Yvette Cedex, France

72 Laboratoire Traitement et Communication de l’Information, CNRS(UMR 5141) and Telecom ParisTech, 46 rue Barrault F-75634Paris Cedex 13, France

73 Laboratoire de Physique Subatomique et de Cosmologie,Universite Joseph Fourier Grenoble I, CNRS/IN2P3, InstitutNational Polytechnique de Grenoble, 53 rue des Martyrs, 38026Grenoble cedex, France

74 Laboratoire de Physique Theorique, Universite Paris-Sud 11 &CNRS, Batiment 210, 91405 Orsay, France

75 Lawrence Berkeley National Laboratory, Berkeley, California,U.S.A.

76 Max-Planck-Institut fur Astrophysik, Karl-Schwarzschild-Str. 1,85741 Garching, Germany

77 McGill Physics, Ernest Rutherford Physics Building, McGillUniversity, 3600 rue University, Montreal, QC, H3A 2T8, Canada

78 MilliLab, VTT Technical Research Centre of Finland, Tietotie 3,Espoo, Finland

79 Niels Bohr Institute, Blegdamsvej 17, Copenhagen, Denmark80 Observational Cosmology, Mail Stop 367-17, California Institute

of Technology, Pasadena, CA, 91125, U.S.A.81 Optical Science Laboratory, University College London, Gower

Street, London, U.K.82 SB-ITP-LPPC, EPFL, CH-1015, Lausanne, Switzerland83 SISSA, Astrophysics Sector, via Bonomea 265, 34136, Trieste,

Italy84 School of Physics and Astronomy, Cardiff University, Queens

Buildings, The Parade, Cardiff, CF24 3AA, U.K.85 Space Research Institute (IKI), Russian Academy of Sciences,

Profsoyuznaya Str, 84/32, Moscow, 117997, Russia86 Space Sciences Laboratory, University of California, Berkeley,

California, U.S.A.87 Special Astrophysical Observatory, Russian Academy of Sciences,

Nizhnij Arkhyz, Zelenchukskiy region, Karachai-CherkessianRepublic, 369167, Russia

88 Stanford University, Dept of Physics, Varian Physics Bldg, 382 ViaPueblo Mall, Stanford, California, U.S.A.

89 Sub-Department of Astrophysics, University of Oxford, KebleRoad, Oxford OX1 3RH, U.K.

90 Theory Division, PH-TH, CERN, CH-1211, Geneva 23,Switzerland

91 UPMC Univ Paris 06, UMR7095, 98 bis Boulevard Arago,F-75014, Paris, France

92 Universite de Toulouse, UPS-OMP, IRAP, F-31028 Toulouse cedex4, France

93 University of Granada, Departamento de Fısica Teorica y delCosmos, Facultad de Ciencias, Granada, Spain

94 University of Miami, Knight Physics Building, 1320 Campo SanoDr., Coral Gables, Florida, U.S.A.

95 Warsaw University Observatory, Aleje Ujazdowskie 4, 00-478Warszawa, Poland

25