+ All Categories
Home > Documents > Planck Early Results: All sky temperature and dust optical depth from Planck and IRAS: Constraints...

Planck Early Results: All sky temperature and dust optical depth from Planck and IRAS: Constraints...

Date post: 25-Nov-2023
Category:
Upload: independent
View: 0 times
Download: 0 times
Share this document with a friend
17
Astronomy & Astrophysics manuscript no. early˙dark˙gas˙astroph c ESO 2011 January 12, 2011 Planck Early Results: All sky temperature and dust optical depth from Planck and IRAS : Constraints on the “dark gas” in our Galaxy Planck Collaboration: P. A. R. Ade 71 , N. Aghanim 47 , M. Arnaud 58 , M. Ashdown 56,77 , J. Aumont 47 , C. Baccigalupi 69 , A. Balbi 29 , A. J. Banday 75,6,63 , R. B. Barreiro 53 , J. G. Bartlett 3,54 , E. Battaner 79 , K. Benabed 48 , A. Benoˆ ıt 48 , J.-P. Bernard 75,6 ? , M. Bersanelli 27,42 , R. Bhatia 34 , J. J. Bock 54,7 , A. Bonaldi 38 , J. R. Bond 5 , J. Borrill 62,72 , F. R. Bouchet 48 , F. Boulanger 47 , M. Bucher 3 , C. Burigana 41 , P. Cabella 29 , J.-F. Cardoso 59,3,48 , A. Catalano 3,57 , L. Cay ´ on 20 , A. Challinor 78,56,9 , A. Chamballu 45 , L.-Y Chiang 50 , C. Chiang 19 , P. R. Christensen 66,30 , D. L. Clements 45 , S. Colombi 48 , F. Couchot 61 , A. Coulais 57 , B. P. Crill 54,67 , F. Cuttaia 41 , T. Dame 35 , L. Danese 69 , R. D. Davies 55 , R. J. Davis 55 , P. de Bernardis 26 , G. de Gasperis 29 , A. de Rosa 41 , G. de Zotti 38,69 , J. Delabrouille 3 , J.-M. Delouis 48 , F.-X. D´ esert 44 , C. Dickinson 55 , K. Dobashi 14 , S. Donzelli 42,51 , O. Dor´ e 54,7 , U. D ¨ orl 63 , M. Douspis 47 , X. Dupac 33 , G. Efstathiou 78 , T. A. Enßlin 63 , H. K. Eriksen 51 , F. Finelli 41 , O. Forni 75,6 , P. Fosalba 49 , M. Frailis 40 , E. Franceschi 41 , Y. Fukui 18 , S. Galeotta 40 , K. Ganga 3,46 , M. Giard 75,6 , G. Giardino 34 , Y. Giraud-H´ eraud 3 , J. Gonz´ alez-Nuevo 69 , K. M. G ´ orski 54,81 , S. Gratton 56,78 , A. Gregorio 28 , I. A. Grenier 58 , A. Gruppuso 41 , F. K. Hansen 51 , D. Harrison 78,56 , G. Helou 7 , S. Henrot-Versill´ e 61 , D. Herranz 53 , S. R. Hildebrandt 7,60,52 , E. Hivon 48 , M. Hobson 77 , W. A. Holmes 54 , W. Hovest 63 , R. J. Hoyland 52 , K. M. Huenberger 80 , A. H. Jae 45 , W. C. Jones 19 , M. Juvela 17 , A. Kawamura 18 , E. Keih¨ anen 17 , R. Keskitalo 54,17 , T. S. Kisner 62 , R. Kneissl 32,4 , L. Knox 22 , H. Kurki-Suonio 17,36 , G. Lagache 47 , J.-M. Lamarre 57 , A. Lasenby 77,56 , R. J. Laureijs 34 , C. R. Lawrence 54 , S. Leach 69 , R. Leonardi 33,34,23 , C. Leroy 47,75,6 , P. B. Lilje 51,8 , M. Linden-Vørnle 11 , M. L ´ opez-Caniego 53 , P. M. Lubin 23 , J. F. Mac´ ıas-P´ erez 60 , C. J. MacTavish 56 , B. Maei 55 , D. Maino 27,42 , N. Mandolesi 41 , R. Mann 70 , M. Maris 40 , P. Martin 5 , E. Mart´ ınez-Gonz´ alez 53 , S. Masi 26 , S. Matarrese 25 , F. Matthai 63 , P. Mazzotta 29 , P. McGehee 46 , P. R. Meinhold 23 , A. Melchiorri 26 , L. Mendes 33 , A. Mennella 27,40 , M.-A. Miville-Deschˆ enes 47,5 , A. Moneti 48 , L. Montier 75,6 , G. Morgante 41 , D. Mortlock 45 , D. Munshi 71,78 , A. Murphy 65 , P. Naselsky 66,30 , P. Natoli 29,2,41 , C. B. Netterfield 13 , H. U. Nørgaard-Nielsen 11 , F. Noviello 47 , D. Novikov 45 , I. Novikov 66 , I. J. O’Dwyer 54 , T. Onishi 15 , S. Osborne 74 , F. Pajot 47 , R. Paladini 73,7 , D. Paradis 75,6 , F. Pasian 40 , G. Patanchon 3 , O. Perdereau 61 , L. Perotto 60 , F. Perrotta 69 , F. Piacentini 26 , M. Piat 3 , S. Plaszczynski 61 , E. Pointecouteau 75,6 , G. Polenta 2,39 , N. Ponthieu 47 , T. Poutanen 36,17,1 , G. Pr´ ezeau 7,54 , S. Prunet 48 , J.-L. Puget 47 , W. T. Reach 76 , M. Reinecke 63 , C. Renault 60 , S. Ricciardi 41 , T. Riller 63 , I. Ristorcelli 75,6 , G. Rocha 54,7 , C. Rosset 3 , M. Rowan-Robinson 45 , J. A. Rubi ˜ no-Mart´ ın 52,31 , B. Rusholme 46 , M. Sandri 41 , D. Santos 60 , G. Savini 68 , D. Scott 16 , M. D. Seiert 54,7 , P. Shellard 9 , G. F. Smoot 21,62,3 , J.-L. Starck 58,10 , F. Stivoli 43 , V. Stolyarov 77 , R. Stompor 3 , R. Sudiwala 71 , J.-F. Sygnet 48 , J. A. Tauber 34 , L. Terenzi 41 , L. Toolatti 12 , M. Tomasi 27,42 , J.-P. Torre 47 , M. Tristram 61 , J. Tuovinen 64 , G. Umana 37 , L. Valenziano 41 , P. Vielva 53 , F. Villa 41 , N. Vittorio 29 , L. A. Wade 54 , B. D. Wandelt 48,24 , A. Wilkinson 55 , D. Yvon 10 , A. Zacchei 40 , and A. Zonca 23 (Aliations can be found after the references) Preprint online version: January 12, 2011 ABSTRACT An all sky map of the apparent temperature and optical depth of thermal dust emission is constructed using the Planck-HFI (350 μm to 2 mm) and IRAS (100 μm) data. The optical depth maps are correlated with tracers of the atomic (H i) and molecular gas traced by CO. The correlation with the column density of observed gas is linear in the lowest column density regions at high Galactic latitudes. At high N H , the correlation is consistent with that of the lowest N H , for a given choice of the CO-to-H 2 conversion factor. In the intermediate N H range, a departure from linearity is observed, with the dust optical depth in excess of the correlation. This excess emission is attributed to thermal emission by dust associated with a dark gas phase, undetected in the available H i and CO surveys. The 2D spatial distribution of the dark gas in the solar neighbourhood (|b II | > 10 ) is shown to extend around known molecular regions traced by CO. The average dust emissivity in the H i phase in the solar neighbourhood is found to be τ D /N tot H = 5.2 × 10 -26 cm 2 at 857 GHz. It follows roughly a power law distribution with a spectral index β = 1.8 all the way down to 3 mm, although the SED flattens slightly in the millimetre. Taking into account the spectral shape of the dust optical depth, the emissivity is consistent with previous values derived from FIRAS measurements at high latitudes within 10%. The threshold for the existence of the dark gas is found at N tot H = (8.0±0.58)×10 20 Hcm -2 (A V = 0.4 mag). Assuming the same high frequency emissivity for the dust in the atomic and the molecular phases leads to an average X CO = (2.54 ± 0.13) × 10 20 H 2 cm -2 /(K km s -1 ). The mass of dark gas is found to be 28% of the atomic gas and 118% of the CO emitting gas in the solar neighbourhood. The Galactic latitude distribution shows that its mass fraction is relatively constant down to a few degrees from the Galactic plane. A possible explanation for the dark gas lies in a dark molecular phase, where H 2 survives photodissociation but CO does not. The observed transition for the onset of this phase in the solar neighbourhood (A V = 0.4 mag) appears consistent with recent theoretical predictions. It is also possible that up to half of the dark gas could be in atomic form, due to optical depth eects in the H i measurements. Key words. ISM: general, dust, extinction, clouds – Galaxies: ISM – Infrared: ISM – Submillimeter: ISM 1. Introduction The matter that forms stars, that is left over after star formation, or that has never experienced star formation comprises the in- ? Corresponding author; email: [email protected]. terstellar medium (ISM). The life-cycle and the duration of the various observable phases remains largely unknown, because the nature of the diuse interstellar medium is dicult to discern, owing to its low temperatures and large angular scales. 1 arXiv:1101.2029v1 [astro-ph.GA] 11 Jan 2011
Transcript

Astronomy & Astrophysics manuscript no. early˙dark˙gas˙astroph c© ESO 2011January 12, 2011

Planck Early Results: All sky temperature and dust optical depthfrom Planck and IRAS: Constraints on the “dark gas” in our Galaxy

Planck Collaboration: P. A. R. Ade71, N. Aghanim47, M. Arnaud58, M. Ashdown56,77, J. Aumont47, C. Baccigalupi69, A. Balbi29,A. J. Banday75,6,63, R. B. Barreiro53, J. G. Bartlett3,54, E. Battaner79, K. Benabed48, A. Benoıt48, J.-P. Bernard75,6 ?, M. Bersanelli27,42, R. Bhatia34,

J. J. Bock54,7, A. Bonaldi38, J. R. Bond5, J. Borrill62,72, F. R. Bouchet48, F. Boulanger47, M. Bucher3, C. Burigana41, P. Cabella29,J.-F. Cardoso59,3,48, A. Catalano3,57, L. Cayon20, A. Challinor78,56,9, A. Chamballu45, L.-Y Chiang50, C. Chiang19, P. R. Christensen66,30,

D. L. Clements45, S. Colombi48, F. Couchot61, A. Coulais57, B. P. Crill54,67, F. Cuttaia41, T. Dame35, L. Danese69, R. D. Davies55, R. J. Davis55,P. de Bernardis26, G. de Gasperis29, A. de Rosa41, G. de Zotti38,69, J. Delabrouille3, J.-M. Delouis48, F.-X. Desert44, C. Dickinson55, K. Dobashi14,

S. Donzelli42,51, O. Dore54,7, U. Dorl63, M. Douspis47, X. Dupac33, G. Efstathiou78, T. A. Enßlin63, H. K. Eriksen51, F. Finelli41, O. Forni75,6,P. Fosalba49, M. Frailis40, E. Franceschi41, Y. Fukui18, S. Galeotta40, K. Ganga3,46, M. Giard75,6, G. Giardino34, Y. Giraud-Heraud3,

J. Gonzalez-Nuevo69, K. M. Gorski54,81, S. Gratton56,78, A. Gregorio28, I. A. Grenier58, A. Gruppuso41, F. K. Hansen51, D. Harrison78,56,G. Helou7, S. Henrot-Versille61, D. Herranz53, S. R. Hildebrandt7,60,52, E. Hivon48, M. Hobson77, W. A. Holmes54, W. Hovest63, R. J. Hoyland52,K. M. Huffenberger80, A. H. Jaffe45, W. C. Jones19, M. Juvela17, A. Kawamura18, E. Keihanen17, R. Keskitalo54,17, T. S. Kisner62, R. Kneissl32,4,

L. Knox22, H. Kurki-Suonio17,36, G. Lagache47, J.-M. Lamarre57, A. Lasenby77,56, R. J. Laureijs34, C. R. Lawrence54, S. Leach69,R. Leonardi33,34,23, C. Leroy47,75,6, P. B. Lilje51,8, M. Linden-Vørnle11, M. Lopez-Caniego53, P. M. Lubin23, J. F. Macıas-Perez60,

C. J. MacTavish56, B. Maffei55, D. Maino27,42, N. Mandolesi41, R. Mann70, M. Maris40, P. Martin5, E. Martınez-Gonzalez53, S. Masi26,S. Matarrese25, F. Matthai63, P. Mazzotta29, P. McGehee46, P. R. Meinhold23, A. Melchiorri26, L. Mendes33, A. Mennella27,40,

M.-A. Miville-Deschenes47,5, A. Moneti48, L. Montier75,6, G. Morgante41, D. Mortlock45, D. Munshi71,78, A. Murphy65, P. Naselsky66,30,P. Natoli29,2,41, C. B. Netterfield13, H. U. Nørgaard-Nielsen11, F. Noviello47, D. Novikov45, I. Novikov66, I. J. O’Dwyer54, T. Onishi15,

S. Osborne74, F. Pajot47, R. Paladini73,7, D. Paradis75,6, F. Pasian40, G. Patanchon3, O. Perdereau61, L. Perotto60, F. Perrotta69, F. Piacentini26,M. Piat3, S. Plaszczynski61, E. Pointecouteau75,6, G. Polenta2,39, N. Ponthieu47, T. Poutanen36,17,1, G. Prezeau7,54, S. Prunet48, J.-L. Puget47,W. T. Reach76, M. Reinecke63, C. Renault60, S. Ricciardi41, T. Riller63, I. Ristorcelli75,6, G. Rocha54,7, C. Rosset3, M. Rowan-Robinson45,

J. A. Rubino-Martın52,31, B. Rusholme46, M. Sandri41, D. Santos60, G. Savini68, D. Scott16, M. D. Seiffert54,7, P. Shellard9, G. F. Smoot21,62,3,J.-L. Starck58,10, F. Stivoli43, V. Stolyarov77, R. Stompor3, R. Sudiwala71, J.-F. Sygnet48, J. A. Tauber34, L. Terenzi41, L. Toffolatti12,

M. Tomasi27,42, J.-P. Torre47, M. Tristram61, J. Tuovinen64, G. Umana37, L. Valenziano41, P. Vielva53, F. Villa41, N. Vittorio29, L. A. Wade54,B. D. Wandelt48,24, A. Wilkinson55, D. Yvon10, A. Zacchei40, and A. Zonca23

(Affiliations can be found after the references)

Preprint online version: January 12, 2011

ABSTRACT

An all sky map of the apparent temperature and optical depth of thermal dust emission is constructed using the Planck-HFI (350 µm to 2 mm)and IRAS (100 µm) data. The optical depth maps are correlated with tracers of the atomic (H i) and molecular gas traced by CO. The correlationwith the column density of observed gas is linear in the lowest column density regions at high Galactic latitudes. At high NH, the correlation isconsistent with that of the lowest NH, for a given choice of the CO-to-H2 conversion factor. In the intermediate NH range, a departure from linearityis observed, with the dust optical depth in excess of the correlation. This excess emission is attributed to thermal emission by dust associated with adark gas phase, undetected in the available H i and CO surveys. The 2D spatial distribution of the dark gas in the solar neighbourhood (|bII| > 10◦)is shown to extend around known molecular regions traced by CO.The average dust emissivity in the H i phase in the solar neighbourhood is found to be τD/Ntot

H = 5.2 × 10−26 cm2 at 857 GHz. It follows roughly apower law distribution with a spectral index β = 1.8 all the way down to 3 mm, although the SED flattens slightly in the millimetre. Taking intoaccount the spectral shape of the dust optical depth, the emissivity is consistent with previous values derived from FIRAS measurements at highlatitudes within 10%. The threshold for the existence of the dark gas is found at Ntot

H = (8.0±0.58)×1020 Hcm−2 (AV = 0.4 mag). Assuming the samehigh frequency emissivity for the dust in the atomic and the molecular phases leads to an average XCO = (2.54 ± 0.13) × 1020 H2cm−2/(K km s−1).The mass of dark gas is found to be 28% of the atomic gas and 118% of the CO emitting gas in the solar neighbourhood. The Galactic latitudedistribution shows that its mass fraction is relatively constant down to a few degrees from the Galactic plane.A possible explanation for the dark gas lies in a dark molecular phase, where H2 survives photodissociation but CO does not. The observedtransition for the onset of this phase in the solar neighbourhood (AV = 0.4 mag) appears consistent with recent theoretical predictions. It is alsopossible that up to half of the dark gas could be in atomic form, due to optical depth effects in the H i measurements.

Key words. ISM: general, dust, extinction, clouds – Galaxies: ISM – Infrared: ISM – Submillimeter: ISM

1. Introduction

The matter that forms stars, that is left over after star formation,or that has never experienced star formation comprises the in-

? Corresponding author; email: [email protected].

terstellar medium (ISM). The life-cycle and the duration of thevarious observable phases remains largely unknown, because thenature of the diffuse interstellar medium is difficult to discern,owing to its low temperatures and large angular scales.

1

arX

iv:1

101.

2029

v1 [

astr

o-ph

.GA

] 1

1 Ja

n 20

11

Planck collaboration: Constraints on the dark gas in our galaxy

The distribution of diffuse interstellar gas, by which we meangas not in gravitationally–bound structures and not in the im-mediate vicinity of active star-formation regions, has primarilybeen assessed using the 21-cm hyperfine line of atomic hydro-gen. That line is easily excited by collisions and is opticallythin for gas with temperature TK > 50 K and velocity disper-sion δV > 10 km s−1 as long as the column density is less than9 × 1021 cm−2 (Kulkarni & Heiles 1988). Such conditions aretypical of the diffuse ISM pervaded by the interstellar radiationfield (ISRF), because photoelectric heating from grain surfaceskeeps the gas warm (T > 50 K), and observed velocity disper-sions (presumably due to turbulence) are typically > 10 km s−1.Based on the observed dust extinction per unit column density,N(HI)/AV = 1.9 × 1021 cm−2 mag−1 (Bohlin et al. 1978), theupper column density for optically thin 21-cm lines correspondsto visible extinctions AV < 4.7. Thus the 21-cm line is expectedto trace diffuse, warm atomic gas accutately throughout the dif-fuse ISM, except for lines of sight that are visibly opaque or areparticularly cold.

Molecular gas is typically traced by the 2.6-mm12CO(J=1→0) rotational line in emission, which, like the21-cm H i line, can be easily excited because it involves energylevels that can be obtained by collisions. The CO emission line,however, is commonly optically thick, due to its high radiativetransition rate. In the limit where the lines are optically thick, theprimary information determining the amount of molecular gasin the beam is the line width. If the material is gravitationallybound, then the virial mass is measured and CO can be usedas a tracer of molecular mass. It is common astronomicalpractice to consider the velocity-integrated CO line intensityas measuring the molecular column density, with the implicitassumption that the material is virialized and the mass of thevirialized structures is being measured. In the diffuse ISM,these conditions typically do not apply. On a physical scale of R(measured in parsecs), interstellar material is only virialized ifits column density N > 5.2 × 1021δV2R−1 cm−2 where δV is thevelocity dispersion (measured in km s−1). Thus the diffuse ISMis typically gravitationally unbound, invalidating the usage ofCO as a virial tracer of the molecular gas mass, except in verycompact regions or in regions that are visibly opaque. AlthoughCO can emit in gas with low density, the critical density requiredfor collisional equilibrium is of order 103 cm−3, which furthercomplicates the usage of CO as a tracer. This again is not typicalof the diffuse ISM.

To measure the amount and distribution of the molecularISM, as well as the cold atomic ISM, other tracers of the in-terstellar gas are required. At least three tracers have been usedin the past. These are UV absorption in Werner bands of H2,infrared emission from dust, and γ-ray emission from pion pro-ductiondue to cosmic-rays colliding with interstellar nucleons.The UV absorption is exceptionally sensitive to even very lowH2 column densities of 1017 cm−2. Using Copernicus (Savageet al. 1977) and FUSE data, atomic and molecular gas couldbe measured simultaneously on the sightlines to UV-bright starsand some galaxies. A survey at high Galactic latitudes withFUSE showed that the molecular fraction of the ISM, f(H2) ≡2N(H2)/[2N(H2) + N(HI)] < 10−3 for lines of sight with totalcolumn density less than 1020 cm−2, but there is a tremendousdispersion from 10−4 to 10−1 for higher-column density lines ofsight (Wakker 2006). Since UV-bright sources are preferentiallyfound towards the lowest-extinction sightlines, an accurate aver-age f(H2) is extremely difficult to determine from the stellar ab-sorption measurements. Along lines of sight toward AGNs be-hind diffuse interstellar clouds, Gillmon & Shull (2006) found

molecular hydrogen fractions of 1–30% indicating significantmolecular content even for low-density clouds.

The dust column density has been used as a total gas columndensity tracer, with the assumption that gas and dust are wellmixed. The possibility that dust traces the column density bet-ter than H i and CO was recognized soon after the first all-skyinfrared survey by IRAS , which for the first time revealed thedistribution of dust on angular scales down to 5′. Molecular gaswithout CO was inferred from comparing IRAS 100 µm surfacebrightness to surveys of the 21-cm and 2.6-mm lines of H i andCO on 9′ or degree scale de Vries et al. (1987); Heiles et al.(1988); Blitz et al. (1990). At 3′ scale using Arecibo, the cloudG236+39 was found to have significant infrared emission un-accounted for by 21-cm or 2.6-mm lines, with a large portionof the cloud being possibly H2 with CO emission below detec-tion threshold (Reach et al. 1994). Meyerdierks & Heithausen(1996) also detected IR emission surrounding the Polaris flare inexcess of what was expectated from the H i and CO emission,which they attributed to diffuse molecular gas. The all sky far-infrared observations by COBE -DIRBE (Hauser et al. 1998)made it possible to survey the molecular gas not traced by H ior CO at the 1◦ scale (Reach et al. 1998). This revealed numer-ous “infrared excess” clouds, many of which were confirmedas molecular after detection of faint CO with NANTEN (Onishiet al. 2001). Finally, there are also indications of more dust emis-sion than seen in nearby external galaxies such as the LargeMagellanic Cloud (Bernard et al. 2008; Roman-Duval et al.2010) and the Small Magellanic Cloud (Leroy et al. 2007). Thissuggests that large fractions of the gas masses of these galaxiesare not detected using standard gas tracers.

The γ-rays from the interstellar medium provide an indepen-dent tracer of the total nucleon density. As was the case with thedust column density, the γ-ray inferred nucleon column densityappears to show an extra component of the ISM not associatedwith detected 21-cm or 2.6-mm emission; this extra emissionwas referred to as ”dark gas” (e.g. Grenier et al. 2005; Abdo et al.2010), a term we will adopt in this paper to describe interstellarmaterial beyond what is traced by H i and CO emission. Grenieret al. (2005) inferred dark gas column densities of order 50% ofthe total column density toward locations with little or beyounddetection threshold CO emission, and general consistency be-tween infrared and γ-ray methods of detection. Recent observa-tions using FERMI have significantly advanced this method, al-lowing γ-ray emission to be traced even by the high-latitude dif-fuse ISM. In the Cepheus, Cassiopeia, and Polaris Flare clouds,the correlated excess of dust and γ rays yields dark gas massesthat range from 40 % to 60 % of the CO-bright molecular mass(Abdo et al. 2010).

Theoretical work predicts a dark molecular gas layer in re-gions where the balance between photodissociation and molec-ular formation allows H2 to form in significant quantity whilethe gas-phase C remains atomic or ionized (Wolfire et al. 2010;Glover et al. 2010). In this paper we describe new observa-tions made with Planck 1 (Planck Collaboration 2011a) that tracethe distribution of submillimeter emission at 350 µm and longerwavelengths. In combination with observations up to 100 µmwavelength by IRAS and COBE -DIRBE , we are uniquely able

1 Planck (http://www.esa.int/Planck) is a project of the EuropeanSpace Agency (ESA) with instruments provided by two scientific con-sortia funded by ESA member states (in particular the lead countries:France and Italy) with contributions from NASA (USA), and telescopereflectors provided in a collaboration between ESA and a scientific con-sortium led and funded by Denmark.

2

Planck collaboration: Constraints on the dark gas in our galaxy

Table 1. Characteristics of the data used in this study

Data λref νref θ σII σabs[ µm] [ GHz] [arcmin] [MJy/sr] [%]

IRAS 100.0 2998 4.30 0.06†‡ 13.6‡HFI 349.8 857 3.67 0.12[ 7%HFI 550.1 545 3.80 0.12[ 7%HFI 849.3 353 4.43 0.08[ <

∼ 2%HFI 1381.5 217 4.68 0.08[ <

∼ 2%HFI 2096.4 143 7.04 0.08[ <

∼ 2%HFI 2997.9 100 9.37 0.07[ <

∼ 2%

Data line λref θ σ σabs[ µm] [arcmin] [ Kkms−1] [%]

LAB HI 21 cm 36.0 1.70] 10.0DHT 12CO 2.6 mm 8.8 1.20 24.0Dame 12CO 2.6 mm 8.4 0.6 24.0NANTEN 12CO 2.6 mm 2.6 1.20 10.0† Assumed to be for the average IRAS coverage. σII computed by rescal-ing this value to actual coverage. ‡ From Miville-Deschenes & Lagache(2005). [ 1σ average value in one beam scaled from Planck HFI CoreTeam (2011b). We actually use internal variance maps for σII

] 1σ aver-age value. We actually use a map of the uncertainties (see Sect. 2.2.1).

to trace the distribution of interstellar dust with temperaturesdown to ∼ 10 K. The surface brightness sensitivity of Planck,in particular on angular scales of 5′ to 7◦, is unprecedented.Because we can measure the dust optical depth more accuratelyby including the Planck data, we can now reassess the relation-ship between dust and gas, and relate it to previous infrared andindependent UV and γ-ray results, and compare it to theoreticalexplanations to determine just how important the dark gas is forthe evolution of the interstellar medium.

2. Observations

2.1. Planck data

The Planck first mission results are presented in PlanckCollaboration (2011a) and the in-flight performances of the twofocal plane instruments HFI (High Frequency Instrument) andLFI (Low Frequency Instrument) are given in Planck HFI CoreTeam (2011a) and Mennella et al. (2011) respectively. The dataprocessing and calibration of the HFI and LFI data used hereis described in Planck HFI Core Team (2011b) and PlanckCollaboration (2011b) respectively.

Here we use only the HFI (DR2 release) data, the process-ing and calibration of which are described in Planck HFI CoreTeam (2011b). In this data the CMB component was identifiedand subtracted through a Needlet Internal Linear Combination(NILC) (Planck HFI Core Team 2011b).

We use the internal variance on intensity (σ2II) estimated dur-

ing the Planck data processing and provided with the Planck-HFI data, which we assume represents the white noise on theintensity. Note that this variance is inhomogeneous over the sky,owing to the Planck scanning strategy (Planck Collaboration2011a), with lower values in the Planck deep fields near theecliptic poles. We have checked that, within a small factor (< 2),the data variance above is consistent with “Jack-Knife” mapsobtained from differencing the two halves of the Planck rings.We also use the absolute uncertainties due to calibration uncer-tainties given in Planck HFI Core Team (2011b) for HFI and

summarized in Table 1. We note that, for a large scale analysissuch as carried out here, variances contribute to a small fractionof the final uncertainty resulting from combining data over largesky regions, so that most of the final uncertainty is due to abso-lute uncertainties.

2.2. Ancillary data

2.2.1. HI data

In order to trace the atomic medium, we use the LAB(Leiden/Argentine/Bonn) survey which contains the final datarelease of observations of the H i 21-cm emission line overthe entire sky (Kalberla et al. 2005). This survey merged theLeiden/Dwingeloo Survey (Hartmann & Burton 1997) of thesky north at δ > −30◦ with the IAR (Instituto Argentino deRadioastronomia) Survey (Arnal et al. 2000; Bajaja et al. 2005)of the Southern sky at δ < −25◦. The angular resolution and thevelocity resolution of the survey are ∼ 0.6◦ and ∼ 1.3 km s−1.The LSR velocity range −450 < VLSR < 400 km s−1 is fully cov-ered by the survey with 891 channels with a velocity separationof ∆Vch = 1.03 km s−1.

The data were corrected for stray radiation at the Institute forRadioastronomy of the University of Bonn. The rms brightness-temperature noise of the merged database is slightly lower in thesouthern sky than in the northern sky, ranging over 0.07-0.09 K.Residual uncertainties in the profile wings, due to defects in thecorrection for stray radiation, are for most of the data below alevel of 20 to 40 mK. We integrated the LAB data in the ve-locity range −400 < VLSR < 400 km s−1 to produce an all skymap of the H i integrated intensity (WHI), which was finally pro-jected into the HEALPix pixelisation scheme using the methoddescribed in Sect. 2.3.1.

We estimate the noise level of the WHI map as∆Trms∆Vch

√Nch where Nch(= 777) is the number of channels

used for the integration, and ∆Trms is the rms noise of the in-dividual spectra measured in the emission-free velocity rangemainly in −400 < VLSR < 350 km s−1. The resulting noise ofthe WHI map is mostly less than ∼ 2.5 Kkms−1 all over the skywith an average value of ∼ 1.7 Kkms−1, except for some limitedpositions showing somewhat larger noise (∼ 10 Kkms−1).

2.2.2. CO data

In order to trace the spatial distribution of the CO emission, weuse a combination of 3 large scale surveys in the 12CO(J=1→0)line.

In the Galactic plane, we use the Dame et al. (2001) surveyobtained with the CfA telescope in the north and the CfA-Chiletelescope in the south, referred to here as DHT (Dame, Hartmann& Thaddeus). These data have an angular resolution of 8.4′±0.2′and 8.8′±0.2′ respectively. The velocity coverage and the veloc-ity resolution for these data vary from region to region on the sky,depending on the individual observations composing the survey.The most heavily used spectrometer is the 500 kHz filter bankproviding a velocity coverage and resolution of 332 km s−1and1.3 km s−1, respectively. Another 250 kHz filter bank providingthe 166 km s−1coverage and 0.65 km s−1resolution was also fre-quently used . The rms noises of these data are suppressed downto 0.1–0.3 K (for details, see their Table 1). The data cubes havebeen transformed into the velocity-integrated intensity of the line(WCO) by integrating the velocity range where the CO emis-sion is significantly detected using the moment method proposedby Dame (2011). The noise level of the WCO map is typically

3

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 1. Map of the 12CO(J=1→0) integrated intensity used in this paper combining the Dame et al. (2001) and high latitude surveyand the NANTEN survey. The data shown cover 62.8% of the sky. The map is shown at a common resolution of all the sub-surveysof 8.8’. Many small clouds at high latitude are not visible in this rendering of the data.

∼ 1.2 Kkms−1, but it varies by a factor of a few depending on theintegration range used.

We also use the unpublished high latitude survey obtainedusing the CfA telescope (Dame et al. 2010, private communica-tion). This survey is still on-going and covers the northern skyup to latitudes as high as |bII| = 70◦ which greatly increasesthe overall sky coverage. The noise level of the CO spectra aresuppressed to ∼ 0.18 K for the 0.65 km s−1velocity resolution,and the total CO intensity was derived by integrating typically10–20 velocity channels, which results in a noise level of 0.4–0.6 Kkms−1.

Finally, we combined the above survey with the NANTEN12CO(J=1→0) survey obtained from Chile. This survey comple-ments some of the intermediate Galactic latitudes not covered bythe Dame et al. (2001) maps with an angular resolution of 2.6′.Most of the survey along the Galactic plane has a velocity cov-erage of ∼ 650 km s−1with a wide band spectrometer, but a partof the survey has a coverage of ∼ 100 km s−1with a narrow bandspectrometer. The noise level achieved was 0.4–0.5 K at a ve-locity resolution of 0.65 km s−1. The CO spectra were sampledwith the 2′ grid in the Galactic centre, and with the 4′ and 8′grid along the Galactic plane in the latitude range |b| < 5 ◦and|b| > 5 ◦, respectively. The integrated intensity maps were ob-tained by integrating over the whole velocity range, excludingregions of the spectra where no emission is observed. The re-sulting rms noise in the velocity-integrated intensity map variesdepending on the width of the emission. This survey along theGalactic plane is still not published in full, but parts of the surveyhave been analyzed (e.g. Fukui et al. (1999); Matsunaga et al.(2001); Mizuno & Fukui (2004)). A large amount of the sky at

intermediate Galactic latitude toward the nearby clouds is alsocovered with a higher velocity resolution of ∼ 0.1 km s−1 with anarrow band spectrometer with a . 100 km s−1 band (e.g. Onishiet al. (1999); Kawamura et al. (1999); Mizuno et al. (2001)).The velocity coverage, the grid spacing, and the noise level forthese data vary, depending on the characteristics of the individ-ual clouds observed, but the quality of the data is high enough totrace the total CO intensity of the individual clouds.

The three surveys were repixelised into the HEALPix pix-elisation scheme (Gorski et al. 2005) with the appropriate pixelsize to ensure Shannon sampling of the beam (Nside=2048 forthe NANTEN2 survey and Nside=1024 for the CfA surveys) us-ing the procedure described in Sect. 2.3.1.

Each survey was smoothed to a common resolution of8.8′through convolution with a Gaussian with kernel size ad-justed to go from the original resolution of each survey to agoal resolution of 8.8′, using the smoothing capabilities of theHEALPix software. We checked the consistency of the differ-ent surveys in the common region observed with NANTEN andCfA.

We found a reasonably good correlation between the two buta slope indicating that the NANTEN survey yields 24% largerintensities than the CfA values. The origin of this discrepancyis currently unknown. We should note that the absolute inten-sity scale in CO observations is not highly accurate as notedoften in the previous CO papers. Since the CfA survey coversmost of the regions used in this paper and has been widely usedfor calibrating the H2 mass calibration factor XCO, in particularby several generations of gamma ray satellites, we assumed theCfA photometry when merging the data, and therefore rescaled

4

Planck collaboration: Constraints on the dark gas in our galaxy

the NANTEN data down by 24% before merging. Note that thisan arbitrary choice. The implications on our results will be dis-cussed in Sec. 6.1.

The 3 surveys were then combined into a single map. In do-ing so, data from different surveys falling into the same pixelwere averaged using σ2 as a weight. The resulting combinedmap was then smoothed to the resolution appropriate to thisstudy. The resulting CO integrated intensity map is shown inFig. 1.

2.2.3. IR data

We use the IRIS (Improved Reprocessing of the IRAS Survey)IRAS 100 µm data (Miville-Deschenes & Lagache 2005, see) inorder to constrain the dust temperature. The data, provided inthe original format of individual tiles spread over the entire skywere combined into the HEALPix pixelisation using the methoddescribed in Sect. 2.3.1 at a HEALPix resolution (Nside = 2048corresponding to a pixel size of 1.7′). The IRAS coverage mapswere also processed in the same way. We assume the noise prop-erties given in Miville-Deschenes & Lagache (2005) and givenin Table 1. The noise level of 0.06 MJy sr−1 at 100 µm was as-sumed to represent the average data noise level and was appro-priately multiplied by the coverage map to lead to the pixel vari-ance of the data.

2.3. Additional Data processing

2.3.1. Common angular resolution and pixelisation

The individual maps are then combined into HEALPix using theintersection surface as a weight. This procedure was shown topreserve photometry accuracy.

The ancillary data described in Sect. 2.2 were brought to theHEALPix pixelisation, using a method where the surface of theintersection between each HEALPix pixel with each FITS pixelof the survey data is computed and used as a weight to regridthe data. The HEALPix resolution was chosen so as to matchthe Shannon sampling of the original data at resolution θ, witha HEALPix resolution set so that the pixel size is < θ/2.4. Theancillary data and the description of their processing will be pre-sented in Paradis & et. al. (2011).

All ancillary data were then smoothed to an appropriate res-olution by convolution with a Gaussian smoothing function withappropriate FWHM using the smoothing HEALPix function,and were brought to a pixel size matching the Shannon samplingof the final resolution.

2.3.2. Background levels

Computing the apparent temperature and optical depth of ther-mal dust over the whole sky requires accurate subtraction ofany offset (I0

ν ) in the intensity data, either of instrumental orastrophysical origin. Although both the IRIS and the Planck-HFI data used in this study have been carefully treated withrespect to residual offsets during calibration against the FIRASdata, the data still contains extended sources of emission un-related to the Galactic emission, such as the Cosmic InfraRedBackground (CIB) signal (Miville-Deschenes et al. 2002; PlanckCollaboration 2011d) or zodical light which could affect the de-termination of the dust temperature and optical depth at low sur-face brightness.

In order to estimate the above data offsets, we first computethe correlation between IR and H i emission in a reference re-

Table 2. Thermal Dust emissivity derived from the correlationwith HI emission in the reference region with |bII| > 20◦ andNHI

H < 1.2 × 1021 Hcm−2 ((Iν/NH)ref). Offsets derived from anempty region with NHI

H < 2.0 × 1019 Hcm−2, assuming the sameemissivity (Iν/NH)ref . The uncertainties are given at the 1σ level.The corresponding data are plotted in Fig. 2.

ν (Iν/NH)ref offset( GHz) [MJy/sr/1020 Hcm−2] [MJy/sr]

IRAS

2998 (6.95±0.94)×10−1 (7.36±0.03)×10−1

DIRBE :

2998 (6.60±0.01)×10−1 (7.72±0.22)×10−1

2141 1.16±0.01 1.41±0.161249 (8.85±0.04)×10−1 (8.46±0.90)×10−1

Planck-HFI :

857 (5.43±0.38)×10−1 (2.57±0.05)×10−1

545 (1.82±0.13)×10−1 (1.83±0.05)×10−1

353 (4.84±0.10)×10−2 (9.50±0.24)×10−2

217 (1.14±0.03)×10−2 (3.45±0.11)×10−2

143 (2.92±0.07)×10−3 (1.01±0.06)×10−2

100 (1.13±0.04)×10−3 (3.34±0.60)×10−3

Planck-LFI :

70.3 (8.74±2.66)×10−5 (-6.40±0.93) 10−3

44.1 (1.14±0.16)×10−4 (-7.02±0.45) 10−3

28.5 (2.11±0.14)×10−4 (-6.95±0.15) 10−3

WMAP :

93.7 (-1.05±0.57)×10−4 (-1.65±1.02)×10−3

61.2 (-1.14±0.27)×10−4 (-4.90±3.63)×10−4

41.1 (3.52±1.13)×10−5 (9.76±1.28)×10−4

32.9 (1.39±0.08)×10−4 (4.10±0.97)×10−4

23.1 (2.69±0.04)×10−4 (8.49±0.44)×10−4

gion such that |bII| > 20◦ and NHIH < 1.2 × 1021 Hcm−2. This was

done using the IDL regress routine and iterative removal of out-liers. The derived dust emissivities ((Iν/NH)ref) are given in Table2. The uncertainties given are those derived from the correla-tion using the data variance as the data uncertainty. The derivedemissivities are in agreement with the ensemble average of thevalues found for the local H i velocities in Planck Collaboration(2011i) (see their Table 2) for individual smaller regions at highGalactic latitude, within the uncertainties quoted in Table 2. Notethat these emissivities are used only to derive the offsets in thisstudy.

We then select all sky pixels with minimum H i column den-sity defined as NHI

H < 2.0×1019 Hcm−2 and compute the averageH i column density in this region to be Nhole

H = 1.75×1019 Hcm−2.The offsets are then computed assuming that the dust emissivityin this region is the same as in the reference region, ie,

I0ν = Ihole

ν − (Iν/NH)ref × NholeH (1)

where Iholeν is the average brightness in the hole region at fre-

quency ν.The offset values derived from the above procedure are given

in Table 2 and were subtracted from the maps used in the rest of

5

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 2. Upper panel: Thermal Dust emissivity (Iν/NH)ref fromTable 2. The dot curve showns a grey body at TD = 17.5 K andβ = 1.8 normalized at 857 GHz, for comparison. The variouscolours are for different instruments: IRAS (Yellow), DIRBE(light blue), Planck-HFI (red), WMAP (dark blue) and Planck-LFI (green). Lower panel: Offsets from Table 2. The error barsare plotted at ±3σ.

this analysis. The offset uncertainties also listed in Table 2 werederived from the emissivity uncertainties propagated to the offsetvalues through Eq. 1. When subtracting the above offsets fromthe IRAS and Planck intensity maps, the data variances werecombined with the offset uncertainties in order to reflect uncer-tainty on the offset determination. Note that, for consistency andfuture use, Table 2 also lists emissivities and offset values forFIR-mm datasets not used in this study. Note also that these off-sets for Planck data are not meant to replace the official valuesprovided with the data, since they suppress any large scale emis-sion not correlated with H i, whatever their origin.

3. Dust temperature and emissivity

3.1. Temperature determination

As shown in previous studies (e.g. Reach et al. 1995; Finkbeineret al. 1999; Paradis et al. 2009; Planck Collaboration 2011e,i),the dust emissivity spectrum in our Galaxy cannot be representedby a single dust emissivity index β over the full FIR-submmdomain. The data available indicate that β is usually steeper inthe FIR and flatter in the submm band, with a transition around500 µm. As dust temperature is best derived around the emissionpeak, we limit the range of frequencies used in the determina-

tion to the FIR, which limits the impact of a potential change ofβ with frequency.

In addition, the dust temperature derived will depend on theassumption made about β, since these two parameters are some-what degenerate in χ2 space. In order to minimize the above ef-fect, we derived dust temperature maps using a fixed value ofthe dust emissivity index β. The selected β value was derived byfitting each pixel of the maps with a modified black body of theform Iν ∝ νβBν(TD) in the above spectral range (method referredto as “free β”). This leads to a median value of TD = 17.7 K andβ = 1.8 in the region at |bII| > 10◦. Note that the β value is con-sistent with that derived from the combination of the FIRAS andPlanck-HFI data at low column density in Planck Collaboration(2011i). Inspection of the corresponding TD and β maps indeedshowed spurious values of both parameters, caused by their cor-relation and the presence of noise in the data, in particular in lowbrightness regions of the maps.

We then performed fits to the FIR emission using the fixedβ = 1.8 value derived above (method referred to as “fixed β”).In the determination of TD, we used the IRIS 100 µm map andthe two highest HFI frequencies at 857 and 545 GHz. Althoughthe median reduced χ2 is slightly higher than for the “free β”method, the temperature maps show many fewer spurious val-ues, in particular in low brightness regions. This results in asharper distribution of the temperature histogram. Since we lateruse the temperature maps to investigate the spectral distributionof the dust optical depth and the dust temperature is a source ofuncertainty, we adopt the “fixed β” method maps in the follow-ing. The corresponding temperature and uncertainty maps areshown in Fig. 3.

Temperature maps were derived at the common resolution ofthose three channels as well as at the resolution of lower inten-sity data. The model was used to compute emission in each pho-tometric channels of the instruments used here (IRAS , Planck-HFI ), taking into account the colour corrections using the ac-tual transmission profiles for each instrument and following theadopted flux convention. In the interest of computing efficiency,the predictions of a given model were tabulated for a large setof parameters (TD, β). For each map pixel, the χ2 was computedfor each entry of the table and the shape of the χ2 distributionaround the minimum value was used to derive the uncertainty onthe free parameters. This included the effect of the data varianceσ2

II and the absolute uncertainties.

3.2. Angular distribution of dust temperature

The all-sky map of the thermal dust temperature computed asdescribed in Sec. 3.1 for β = 1.8 is shown in Fig. 5. The elon-gated regions with missing values in the map correspond to theIRAS gaps, where the temperature cannot be determined fromthe Planck-HFI data alone. The distribution of the temperatureclearly reflects the large scale distribution of the radiation fieldintensity.

Along the Galactic plane, a large gradient can be seen fromthe outer Galactic regions, with TD ' 14 − 15 K to the in-ner Galactic regions around the Galactic center regions withTD ' 19 K. This asymmetry was already seen at lower angularresolution in the DIRBE (Sodroski et al. 1994) and the FIRAS(Reach et al. 1995) data. The asymmetry is probably due to thepresence of more massive stars in the inner Milky Way regions,in particular in the molecular ring. The presence of warmer dustin the inner Galaxy is actually clearly highlighted by the radialdistribution of the dust temperature derived from Galactic inver-sion of IR data (e.g. Sodroski et al. 1994; Paladini et al. 2007;

6

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 3. Upper panel: All sky map of the dust temperature in K. The temperature is derived from modeling the IRIS 100 µm andthe Planck-HFI emission at 857 and 545 GHz. Lower panel: All sky map of the dust temperature uncertainty in %. The maps areshown in Galactic coordinates with the Galactic centre at the centre of the image. Grey regions correspond to missing IRAS data.

Planck Collaboration 2011f). The origin of the large scale regionnear (lII,bII)=(340◦,−10◦) with TD ' 20 K is currently unclear,

but we note that it corresponds to a region of enhanced X-rayemission in the Rosat All-Sky Survey (RASS).

7

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 4. Details of the dust temperature (left column), dust optical depth at 857 GHz (central column) and dark gas column density(right column) for the Chamaeleon (first line), Aquila-Ophiuchus flare (second line), Polaris flare (third line) and Taurus (fourthline). The temperature and optical depth maps are shown in log scale with a colour scale ranging from 15 K to 20 K and 1 × 10−5

to 3 × 10−3 respectively. The dark gas column density derived from the optical depth at 857 GHz (see Sec. 4) and is shown in linearscale with a colour scale ranging from -3 to 7×1021 Hcm−2 . The contours show the 12CO(J=1→0) integrated intensity at 2, 10 and20 Kkms−1. The double line shows the limit of the CO surveys.

It may therefore correspond to warm dust associated with hotgas pervading the local bubble around the Sun, or a pocket of hotgas in Loop I. Similar large regions with enhanced dust temper-ature, such as near (lII,bII)=(340◦,−30◦) or (lII,bII)=(315◦,+30◦)may have a similar origin. Loop I (lII,bII)=(30◦,+45◦) is seen asa slightly warmer than average structure at TD ' 19 K. Runningparallel to it is the Aquila-Ophiuchus flare (lII,bII)=(30◦,+20◦)with apparent TD ' 14 K extending to latitudes as high as

60◦. The Cepheus and Polaris Flare (lII,bII)=(100–120◦,+10–+20◦) (see Planck Collaboration 2011i, for a detailed study) isalso clearly visible as a lower temperature arch extending up tobII=30◦ into the North Celestial Pole loop and containing a col-lection of even colder condensations (TD ' 12 − 13 K).

On small angular scales, which are accessible over the wholesky only with the combination of the IRAS and Planck-HFIdata at 5′, the map shows a variety of structures that can all

8

Planck collaboration: Constraints on the dark gas in our galaxy

be identified with local heating by known single stars or H iiregions for warmer spots and with molecular clouds for colderregions. Figure 4 illustrates the high resolution spatial distribu-tion of dust temperature and dust optical depth around someof these regions. Warmer regions include the tangent direc-tions to the spiral Galactic arms in Cygnus (lII,bII)=(80◦,0◦) andCarina (lII,bII)=(280◦,0◦), hosts to many OB associations, andmany H ii regions along the plane. At higher Galactic latitude,dust heated by individual hot stars such in the Ophiuchi region(lII,bII)=(340◦,+20◦) with individual stars σ − S co, ν − S co,ρ − Oph, ζ − Oph, in Orion (lII,bII)=(210◦,−20◦) with theTrapezium stars or in Perseus-Taurus (lII,bII)=(160◦,−20◦) withthe California Nebula (NGC1499) can clearly be identified. Notethe Spica HII region at (lII,bII)=(300◦,+50◦) where dust temper-atures are TD ' 20 K due to heating by UV photons from thenearby (80 pc) early-type, giant (B1III) star α Vir.

At intermediate and high latitudes, nearby molecular cloudsgenerally stand out as cold dust environments with TD '

13 K. The most noticeable ones are Taurus (lII,bII)=(160◦,−20◦)(see Planck Collaboration 2011j, for a detailed study), RCrA(lII,bII)=(0◦,−25◦), Chamaeleon (lII,bII)=(300◦,−20◦) and Orion(lII,bII)=(200◦,−20◦). Numerous cold small scale condensationscan readily be found when inspecting the temperature map,which mostly correspond to cold cores similar to those dis-covered at higher resolution in the Herschel data (e.g. Andreet al. 2010; Konyves et al. 2010; Molinari et al. 2010; Juvelaet al. 2010) and in the Planck Cold-Core catalog (see PlanckCollaboration 2011g,h).

Individual nearby Galaxies are also readily identified, in par-ticular the Large (lII,bII)=(279◦,−34◦) and the Small MagellanicCloud (lII,bII)=(301◦,−44◦) (see Planck Collaboration 2011c, fora detailed study), as well as M31 and M33.

Near the Galactic poles, the temperature determination be-comes noisy at the 5′resolution due to the low signal levels.

3.3. Optical depth determination

Maps of the thermal dust optical depth (τD(λ)) are derived using:

τD(λ) =Iν(λ)

Bν(TD), (2)

where Bν is the Planck function and Iν(λ) is the intensity mapat frequency ν. We used resolution–matched maps of TD andIν(λ) and derived τD(λ) maps at the various resolutions of thedata used here. The maps of the uncertainty on τD(λ) (∆τD) arecomputed as:

∆τD(ν) = τD

σ2II

I2ν

+

(δBν

δT(TD)

)2 ∆T2D

B2ν(TD)

1/2

. (3)

4. Dust/Gas correlation

We model the dust opacity (τM) as

τM(λ) =

(τD

NH

)ref

[NHI + 2XCOWCO], (4)

where(τDNH

)refis the reference dust emissivity measured in low

NH regions and XCO = NH2/WCO is the traditional H2/CO con-version factor. It is implicitly assumed that the dust opacity perunit gas column density is the same in the atomic and moleculargas. If this is not the case, this will directly impact our derived

XCO since only the product of XCO by the dust emissivity in theCO phase

(τDNH

)COcan be derived here. The fit to derive the free

parameters of the model is performed only in the portion of thesky covered by all surveys (infrared, H i, and CO) and where ei-ther (1) the extinction is less than a threshold ADG

V , or (2) the COis detected with WCO > 1 Kkms−1. Criterion (1) selects the low-column density regions that are entirely atomic and suffer verysmall H i optical depth effects, so that the dust in this region willbe associated with the H i emission at 21-cm. Criterion (2) se-lects regions where the CO is significantly detected and the dustis associated with both the H i and the 12CO emission lines. Wefit for the following three free parameters:

(τDNH

)ref, XCO and ADG

V .The threshold ADG

V measures the extinction (or equivalently thecolumn density) where the correlation between the dust opticaldepth and the Hi column density becomes non-linear.

The correlation between the optical depth for various photo-metric channels and the total gas column density (Ntot

H = NHI +2XCOWCO) is shown in Fig. 6. The correlations were computedin the region of the sky where the CO data is available (about63% of the sky) and at Galactic latitudes larger than bII> 10◦.The τD and WCO maps used were smoothed to the common res-olution of the H i map (0.6◦). For these plots, we used a fixedvalue of XCO = 2.3×1020 H2cm−2/(K km s−1). The colours showthe density of points in Ntot

H and τD bins. The dots show the NtotH

binned average correlation. The larger scatter of these points athigh Ntot

H comes from the limited number of points in the corre-sponding bins. The red line shows the τM model values derived

from the fit (slope=(τDNH

)ref) to the low Ntot

H part of the data.It can be seen that the correlation is linear at low Ntot

H valuesand then departs from linear at Ntot

H ' 8.0 × 1020 Hcm−2 (ADGV '

0.4 mag). Above NtotH ' 5 × 1021 Hcm−2 (AH2/CO

v ' 2.5 mag),where Ntot

H becomes dominated by the CO contribution, the dustoptical depth again is consistent with the observed correlation atlow Ntot

H for this given choice of the XCO value. Between thesetwo limits, the dust optical depth is in excess of the linear cor-relation. The same trend is observed in all photometric channelsshown, with a similar value for the threshold. It is also observedin the HFI bands at lower frequencies, but the increasing noiseat low Ntot

H prevents an accurate determination of the fit parame-ters.

The best fit parameters for(τDNH

)ref, XCO and ADG

V are givenin Table 3. They were derived separately for each frequency.The uncertainty was derived from the analysis of the fitted χ2

around the best value. The(τDNH

)refvalues decrease with increas-

ing wavelength, as expected for dust emission. The resulting dustoptical depth SED is shown in Fig. 7. The dust optical depthin low column density regions is compatible with β = 1.8 athigh frequencies. The best fit β value between the IRAS 100 µmand the HFI 857 GHz is actually found to be β = 1.75. TheSED then flattens slightly at intermediate frequencies with aslope of β = 1.57 around λ = 500 µm then steepens again toβ = 1.75 above 1 mm. The XCO values derived from the fitare constant within the error bars, which increase with wave-length. The average value, computed using a weight propor-tional to the inverse variance is given in Table 3 and is foundto be XCO = 2.54 ± 0.13 × 1020 H2cm−2/(K km s−1). Similarly,the 0.4 parameter does not significantly change over the wholefrequency range and the weighted average value is found to beADG

V = 0.4 ± 0.029 mag.The excess column density is defined using the difference be-

tween the best fit and the observed dust opacity per unit column

9

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 5. Maps of the dust optical depths on a log scale, in the IRAS 100 µm (first row left) and Planck-HFI bands at 857 (firstrow right), 545 (second row left), 353 (second row right), 217 (third row left), 143 (third row right) and 100 GHz (fourth row).All maps are shown in Galactic coordinates with the Galactic centre at the centre of the image. The missing data in all imagescorrespond to the IRAS gaps. The upper and lower bounds of the colour scale are set to τmin = 5 × 10−5 × (λ/100 µm)−1.8 andτmax = 10−2 × (λ/100 µm)−1.8 respectively.

density using,

NxH ≡ (τD − τM)/

(τD

NH

)ref

. (5)

The NxH map is used to derive the total excess mass (Mx

H) assum-ing a fiducial distance to the gas responsible for the excess.

We also computed the atomic and molecular total gas massesover the same region of the sky, assuming the same distance. In

10

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 6. Correlation plots between the dust optical depth at IRAS 100 µm (upper left), HFI 857 GHz (upper right), 545 GHz (lowerleft) and 353 GHz (lower right) and the total gas column density Nobs

H in the solar neighbourhood (|bII| > 10◦). The color scalerepresents the density of sky pixels on a log scale. The blue dots show a Nobs

H -binned average representation of the correlation. The

red line shows the best linear correlation derived at low NobsH values (τ =

(τDNH

)ref∗Nobs

H + cste). The vertical lines show the positionscorresponding to AV = 0.37 mag and AV = 2.5 mag. These figures are shown for a single XCO = 2.3 × 1020 H2cm−2/(K km s−1).

the region covered by the CO survey, the H i to CO mass ratio de-rived for XCO = 2.54× 1020 H2cm−2/(K km s−1) is MHI/MCO=4.Using the average

(τDNH

)refand XCO values above, the ratio of

the dark gas mass to the atomic gas mass (MxH/M

HIH ) and to the

molecular gas mass (MxH/M

COH ) are given in Table 3. On average,

at high Galactic latitudes, the dark gas masses are of the order of28%% ± 3%% of the atomic gas mass and ' 118%% ± 12%%of the molecular mass.

5. Dark-gas spatial distribution

The spatial distribution of the dark gas as derived from τD com-puted from the HFI 857 GHz channel is shown in Fig. 8. Itis shown in the region where the CO data are available andabove Galactic latitudes of |bII| > 5◦. Regions where WCO >1 Kkm/s have also been excluded. The map clearly shows thatthe dark gas is distributed mainly around the best known molec-ular clouds such as Taurus, the Cepheus and Polaris flares,Chamaeleon and Orion. The strongest excess region is in theAquila-Ophiuchus flare, which was already evident in Grenieret al. (2005).

Significant dark gas is also apparent at high latitudes, southof the Galactic plane in the anticenter and around known translu-

cent molecular clouds, such as MBM53 (lII= 90◦, bII= −30◦).As with all the molecular clouds, the spatial distribution of thedark gas closely follows that of the Gould-Belt (Perrot & Grenier2003) and indicates that most of the dark gas in the solar neigh-bourhood belongs to this dynamical structure.

6. Discussion

6.1. Dust emissivity in the atomic neutral gas

In the solar neighbourhood, Boulanger et al. (1996) measuredan emissivity value in the diffuse medium of 10−25 cm2/H at250 µm assuming a spectral index β = 2 which seemed con-sistent with their data. The optical depth of dust derived inour study in the low Ntot

H regions at |bII| > 10◦ is shown inFig. 7. The Figure also shows the reference value by Boulangeret al. (1996) which is in good agreement with the values de-rived here, interpolated at 250 µm (in fact 10% above when us-ing β = 1.8 and 6% above when using β = 1.75). Our studydoes not allow us to measure the emissivity in the moleculargas, since we are only sensitive to the product of this emissivitywith the XCO factor. However, we note that our derived averageXCO = 2.54 × 1020 H2cm−2/(K km s−1) is significantly higherthan previously derived values. Even if we account for the pos-

11

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 7. Dust optical depth derived from this study using theIRAS and Planck-HFI frequencies. The square symbol showsthe emissivity at 250 µm derived by Boulanger et al. (1996). Thedash and dash-dot lines show a power law emissivity with λ−1.8

and λ−1.75 respectively, normalized to the data at 100 µm. Theerror bars shown are ±1σ.

sible uncertainty in the calibration of the 12CO(J=1→0) emis-sion (24%) discussed in Sec. 2.2.2, increasing the CO emissionby the corresponding factor would only lower our XCO estimateto XCO = 2.2 H2cm−2/(K km s−1). In comparison, a value of(1.8 ± 0.3) × 1020 H2cm−2/(K km s−1) was found at |bII| > 5◦from the comparison of the H i, CO, and IRAS 100 µm maps(Dame et al. 2001). Similarly, values derived from γ-ray FERMIdata can be as low as XCO = 0.87 × 1020 H2cm−2/(K km s−1) inCepheus, Cassiopea and Polaris (Abdo et al. 2010). This couldbe evidence that the dust emissivity in the high-latitude molecu-lar material could be larger than in the atomic phase by a factor' 3. Such an increase in the dust emissivity in molecular regionshas been inferred in previous studies (e.g. Bernard et al. 1999;Stepnik et al. 2003) and was attributed to dust aggregation.

6.2. Dark Molecular gas

The nature of ‘dark molecular gas’ has recently been investi-gated theoretically by Wolfire et al. (2010), who specifically ad-dress the HI/H2 and C/C+ transition at the edges of molecularclouds. The nominal cloud modeled in their study is relativelylarge, with total column density 1.5 × 1022 cm−2, so the applica-bility of the results to the more translucent conditions of high-Galactic-latitude clouds is not guaranteed. The envelope of thecloud has an H i column density of 1.9×1021 cm−2 which is moretypical of the entire column density measured at high latitudes.Wolfire et al. (2010) define fDG as the fraction of molecular gasthat is dark, i.e. not detected by CO. In the nominal model, thechemical and photodissociation balance yields a total H2 col-umn density of 7.0 × 1021 cm−2, while the ‘dark’ H2 in the tran-sition region where CO is dissociated has a column density of1.9×1021 cm−2. The fraction of the total gas column density thatis molecular,

f (H2) =2N(H2)

2N(H2) + N(HI)(6)

is 93% in the nominal model, which suggestss that the line ofsight through such a cloud passes through material which is al-most entirely molecular. To compare the theoretical model to our

observational results, we must put them into the same units. Wedefine the dark gas fraction as the fraction of the total gas columndensity that is dark,

fDARK =2N(Hdark

2 )2N(H2) + N(HI)

= f (H2)fDG. (7)

For the nominal Wolfire et al. (2010) model, fDG =0.29 so wecan infer fDARK=0.27. The smaller clouds in Figure 11 of theirpaper have larger fDG, but f (H2) is also probably smaller (notgiven in the paper) so we cannot yet definitively match the modeland observations. These model calculations are in general agree-ment with our observational results, in that a significant fractionof the molecular gas can be in CO-dissociated ‘dark’ layers.

If we assume that all dark molecular gas in the solar neigh-bourhood is evenly distributed to the observed CO clouds, theaverage fDG measured is in the range fDG = 1.06 − 1.22. This ismore than three times larger than predicted by the Wolfire et al.(2010) mass fraction. This may indicate that molecular cloudsless massive than the ones assumed in the model actually have adark gas mass fraction higher by a factor of about three. Thiswould contradict their conclusion that the dark mass fractiondoes not depend on the total cloud mass.

The location of the H i-to-H2 transition measured here(ADG

V ' 0.4 mag) is comparable, although slightly higher thanthat predicted in the Wolfire et al. (2010) model (ADG

V '

0.2 mag). Again, this difference may indicate variations with thecloud size used, since UV shadowing by the cloud itself is ex-pected to be less efficient for smaller clouds, leading to a transi-tion deeper into the cloud.

6.3. Other possible origins

The observed departure from linearity between τD and the ob-servable gas column density could also in principle be caused byvariations of the dust/gas ratio (D/G). However, such variationswith amplitude of 30% in the solar neighbourhood and a sys-tematic trend for a higher D/G ratio in denser regions would bedifficult to explain over such a small volume and in the presenceof widespread enrichment by star formation. However, the factthat the dark gas is also seen in the γ-ray with comparable ampli-tudes is a strong indication that it originates from the gas phase.The dark gas column-densities inferred from the γ-ray observa-tions are also consistent with a standard D/G ratio (Grenier et al.2005).

The observed excess optical depth could also in principle bedue to variations of the dust emissivity in the FIR-Submm. Weexpect such variations to occur if dust is in the form of aggre-gates with higher emissivity (e.g. Stepnik et al. 2003) in the darkgas region. We note however that such modifications of the op-tical properties mainly affect the FIR-submm emissivity and arenot expected to modify significantly the absorption properties inthe Visible. Therefore, detecting a similar departure from linear-ity between large-scale extinction maps and the observable gaswould allow us to exclude this possibility.

Sky directions where no CO is detected at the sensitivity ofthe CO survey used (0.3-1.2 Kkms−1) may actually host signifi-cant CO emission, which could be responsible for the excess dustoptical depth observed. Evidences for such diffuse and weaklyemitting CO gas have been reported. For instance, in their studyof the large-scale molecular emission of the Taurus complex,Goldsmith et al. (2008) have found that half the mass of the com-plex is in regions of low column density NH < 2 × 1021 cm−2,seen below WCO ' 1 Kkms−1. However, Barriault et al. (2010)

12

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 8. Map of the excess column density derived from the 857 GHz data. The map is shown in Galactic coordinates with theGalactic centre at the centre of the image. The grey regions correspond to those where no IRAS data are available, regions withintense CO emission (WCO > 1 Kkms−1) and the Galactic plane (|bII| < 5◦).

Table 3. Derived parameters for the dark gas, computed in the region with available CO data and |bII| > 10◦.

Frequency τD/NobsH XCO ADG

V MXH/M

HIH MX

H/MCOH

[ GHz] [10−25cm2] [ H2cm−2/(K km s−1)] [mag] – –

2998 4.66±0.65 (2.60±0.18)×1020 (4.05±0.39)×10−1 (2.91±0.38)×10−1 1.27±0.16857 (5.25±0.49)×10−1 (2.52±0.29)×1020 (3.92±0.64)×10−1 (2.73±0.66)×10−1 1.22±0.30545 (2.53±0.29)×10−1 (2.52±0.36)×1020 (3.96±0.79)×10−1 (2.77±0.95)×10−1 1.24±0.43353 (1.18±0.17)×10−1 (2.31±0.47)×1020 (4.07±1.28)×10−1 (2.26±0.67)×10−1 1.10±0.33217 (6.03±1.19)×10−2 (2.58±0.67)×1020 (4.60±2.39)×10−1 (1.71±4.60)×10−1 (7.46±20.09)×10−1

143 (2.98±1.15)×10−2 (1.83±1.00)×1020 (4.62±4.81)×10−1 – –100 (2.08±0.48)×10−2 (4.05±1.93)×1020 (6.69±5.15)×10−1 – –

Average – (2.54±0.13)×1020 (4.03±0.29)×10−1 (2.78±0.28)×10−1 1.18±0.12

reported a poor spatial correlation between emission by dif-fuse CO and regions of FIR excess in two high Galactic lati-tude regions in the Polaris Flare and Ursa Major. The difficultyat finding the CO emission associated to dark gas is that theedges of molecular coulds tend to be highly structured spatially,which could explain why many attempts have been unsuccess-ful (see for instance Falgarone et al. 1991). In our case, it ispossible to obtain an upper limit to the contribution of weakCO emission below the survey detection threshold, by assum-ing that pixels with undetected CO emission actually emit withWCO = 0.5 Kkms−1. This is the detection limit of the survey weuse at |b| > 10◦ so this should be considered an upper limit to thecontribution of undetected diffuse CO emission. In that case, thedark gas mass is reduced by a factor lower than 20%. This indi-cates that, although diffuse weak CO emission could contribute

a fraction of the observed excess emission, it cannot produce thebulk of it.

Finally, we recognize that the optically thin approximationused here for the H i emission may not fully account for thewhole atomic gas present, even at high latitude. H i emission issubject to self absorption and NH can be underestimated fromapplying too high a spin temperature (Ts) while deriving columndensities. Ts is likely to vary from place to place depending onthe relative abundance of CNM clumps (with thermodynamicaltemperatures of 20-100 K) and WNM clouds (at several thou-sand K) in the telescope beam. The effective spin temperatureof 250-400 K to be applied to correct for this blending andtoretrieve the total column density from the H i spectra does notvary much in the Galaxy (Dickey et al. 2003, 2009). It indi-cates that most of the H i mass is in the warm phase and thatthe relative abundance of cold and warm H i is a robust frac-

13

Planck collaboration: Constraints on the dark gas in our galaxy

Fig. 9. Fractional mass of the dark gas with respect to the neutralgas mass as a function of the lower bIIvalue used in the analysis.The solid curve is computed under the assumption of opticallythin H i, the dashed curve is for NH i

H computed using Ts = 80 K.Error bars are 1σ.

tion across the Galaxy (outside of the inner molecular ring). Thecorrelation between the FERMI γ-ray maps and the H i columndensities derived for different spin temperatures also support anaverage (uniform) effective spin temperature > 250 K on andoff the plane (Ackermann et al. 2010). In order to test these ef-fects, we performed the analysis described in this paper using avery low choice for the H i spin temperature. We adopted a valueof Ts = 80 K when the observed H i peak temperature is below80 K and Ts = 1.1×Tpeak when above. Under this hypothesis, weobtained dark gas fractions which are about half of those givenin Table 3 under the optically thin approximation. We considerthis to indicate that significantly less than half of the detecteddark gas could be dense, cold atomic gas. We further note that,under the optically thin H i hypothesis, the dark gas fraction ap-pears very constant with Galactic latitude down to |bII| ' 3◦ (seeSec. 6.4), while it varies more strongly using Ts = 80 K. Thisdoes not support the interpretation that the bulk of the dust ex-cess results from underestimated H i column densities.

6.4. Dark-Gas variations with latitude

We investigate the distribution of the dark gas as a function ofGalactic latitude. This is important, since the dark gas templateproduced here for the solar neighbourhood is also used in direc-tions toward the plane for Galactic inversion purpose in PlanckCollaboration (2011f). We performed the calculations describedin Sec. 4 for various values of the Galactic latitude lower cutoff(bmin) in the range bmin < |bII| < 90◦ with bmin varying from 0◦to 10◦. For each value, we used the best fit parameters derivedfrom bmin = 10◦ and given in Table 3.

Figure 9 shows the evolution of the dark gas mass fractionwith respect to the atomic gas mass as a function of bmin. It canbe seen that the ratio changes only mildly (increases by a factor1.12 from bmin = 10◦ to bmin = 2◦) as we approach the Galacticplane. This indicates that a fairly constant fraction of the darkgas derived from the solar neighbourhood can be applied to therest of the Galaxy.

Figure 9 also shows the same quantity computed using theH i column density derived using Ts = 80 K. It can be seen that,

in that case, the dark gas fraction is predicted to decrease by afactor 2.12 from bmin = 10◦ to bmin = 2◦. This is caused bythe much larger inferred H i masses toward the plane under thishypothesis. We consider it unlikely that the dark gas fractionvaries by such a large factor from the solar neighbourhood to theGalactic plane, and consider it more likely that the correctionapplied to NH by using a spin temperature as low as Ts = 80 Kactually strongly overestimates the H i opacity, and therefore thefraction of the dark gas belonging to atomic gas.

7. Conclusions

We used the Planck-HFI and IRAS data to determine all skymaps of the thermal dust temperature and optical depth. Thetemperature map traces the spatial variations of the radiationfield intensity associated with star formation in the Galaxy. Thistype of map is very important for the detailed analysis of the dustproperties and their spatial variations.

We examined the correlation between the dust optical depthand gas column density as derived from H i and CO observations.These two quantities are linearly correlated below a thresholdcolumn density of Nobs

H < 8.0 × 1020 Hcm−2 corresponding toAV < 0.4 mag. Below this threshold, we observed dust emis-sivities following a power-law with β ' 1.8 below λ ' 500 µmand flattening at longer wavelengths. Absolute emissivity valuesderived in the FIR are consistent with previous estimates.

This linear correlation also holds at high column densities(Nobs

H > 5 × 1021 Hcm−2) corresponding to AV = 2.5 magwhere the total column density is dominated by the molecu-lar phase for a given choice of the XCO factor. Under the as-sumption that the dust emissivity is the same in both phases,this leads to an estimate of the average local CO to H2 factorof XCO = 2.54 × 1020 H2cm−2/(K km s−1). The optical depth inthe intermediate column density range shows an excess in allphotometric channels considered in this study. We interpret theexcess as dust emission associated with dark gas, probably in themolecular phase where H2 survives photodissociation, while theCO molecule does not.

In the solar neighbourhood, the derived mass of the dark gas,assuming the same dust emissivity as in the H i phase is found tocorrespond to ' 28%% of the atomic mass and ' 118%% of themolecular gas mass. The comparison of this value with the recentcalculations for dark molecular gas around clouds more mas-sive than the ones present in the solar neighbourhood indicates adark gas fraction about three times larger in the solar neighbour-hood. The threshold for the onset of the dark gas transition isfound to be ' 0.4 mand appears compatible to, although slightlylarger than, the thresholds predicted by this model. Finally, westress that the H i 21 cm line is unlikely to be optically thin andto measure all the atomic gas. Therefore, the dark gas detectedhere could well represent a mixture of dark molecular and darkatomic gas seen through its dust emission. For an average H ispin temperature of 80 K, the mixture is predicted to be 50%atomic and 50% molecular.

Acknowledgements. A description of the Planck Collaboration and a listof its members can be found at http://www.rssd.esa.int/index.php?project=PLANCK&page=Planck_Collaboration

ReferencesAbdo, A. A., Ackermann, M., Ajello, M., et al. 2010, ApJ, 710, 133Ackermann, M., Ajello, M., Baldini, L., et al. 2010, ApJAndre, P., Men’shchikov, A., Bontemps, S., et al. 2010, A&A, 518, L102+

14

Planck collaboration: Constraints on the dark gas in our galaxy

Arnal, E. M., Bajaja, E., Larrarte, J. J., Morras, R., & Poppel, W. G. L. 2000,A&AS, 142, 35

Bajaja, E., Arnal, E. M., Larrarte, J. J., et al. 2005, A&A, 440, 767Barriault, L., Joncas, G., Falgarone, E., et al. 2010, MNRAS, 406, 2713Bernard, J., Reach, W. T., Paradis, D., et al. 2008, AJ, 136, 919Bernard, J. P., Abergel, A., Ristorcelli, I., et al. 1999, A&A, 347, 640Blitz, L., Bazell, D., & Desert, F. X. 1990, ApJ, 352, L13Bohlin, R. C., Savage, B. D., & Drake, J. F. 1978, ApJ, 224, 132Boulanger, F., Abergel, A., Bernard, J., et al. 1996, A&A, 312, 256Dame, T. M. 2011, in preparationDame, T. M., Hartmann, D., & Thaddeus, P. 2001, ApJ, 547, 792de Vries, H. W., Thaddeus, P., & Heithausen, A. 1987, ApJ, 319, 723Dickey, J. M., McClure-Griffiths, N. M., Gaensler, B. M., & Green, A. J. 2003,

ApJ, 585, 801Dickey, J. M., Strasser, S., Gaensler, B. M., et al. 2009, ApJ, 693, 1250Falgarone, E., Phillips, T. G., & Walker, C. K. 1991, ApJ, 378, 186Finkbeiner, D. P., Davis, M., & Schlegel, D. J. 1999, ApJ, 524, 867Fukui, Y., Onishi, T., Abe, R., et al. 1999, PASJ, 51, 751Gillmon, K. & Shull, J. M. 2006, ApJ, 636, 908Glover, S. C. O., Federrath, C., Mac Low, M., & Klessen, R. S. 2010, MNRAS,

404, 2Goldsmith, P. F., Heyer, M., Narayanan, G., et al. 2008, ApJ, 680, 428Gorski, K. M., Hivon, E., Banday, A. J., et al. 2005, ApJ, 622, 759Grenier, I. A., Casandjian, J., & Terrier, R. 2005, Science, 307, 1292Hartmann, D. & Burton, W. B. 1997, Atlas of Galactic Neutral Hydrogen, ed.

Hartmann, D. & Burton, W. B.Hauser, M. G., Arendt, R. G., Kelsall, T., et al. 1998, ApJ, 508, 25Heiles, C., Reach, W. T., & Koo, B. 1988, ApJ, 332, 313Juvela, M., Ristorcelli, I., Montier, L. A., et al. 2010, A&A, 518, L93+Kalberla, P. M. W., Burton, W. B., Hartmann, D., et al. 2005, A&A, 440, 775Kawamura, A., Onishi, T., Mizuno, A., Ogawa, H., & Fukui, Y. 1999, PASJ, 51,

851Konyves, V., Andre, P., Men’shchikov, A., et al. 2010, A&A, 518, L106+Kulkarni, S. R. & Heiles, C. 1988, Neutral hydrogen and the diffuse interstellar

medium, ed. Kellermann, K. I. & Verschuur, G. L., 95–153Leroy, A., Bolatto, A., Stanimirovic, S., et al. 2007, ApJ, 658, 1027Matsunaga, K., Mizuno, N., Moriguchi, Y., et al. 2001, PASJ, 53, 1003Mennella et al. 2011, Planck early results 03: First assessment of the Low

Frequency Instrument in-flight performance (Submitted to A&A)Meyerdierks, H. & Heithausen, A. 1996, A&A, 313, 929Miville-Deschenes, M. & Lagache, G. 2005, ApJS, 157, 302Miville-Deschenes, M., Lagache, G., & Puget, J. 2002, A&A, 393, 749Mizuno, A. & Fukui, Y. 2004, in Astronomical Society of the Pacific Conference

Series, Vol. 317, Milky Way Surveys: The Structure and Evolution of ourGalaxy, ed. D. Clemens, R. Shah, & T. Brainerd, 59–+

Mizuno, A., Yamaguchi, R., Tachihara, K., et al. 2001, PASJ, 53, 1071Molinari, S., Swinyard, B., Bally, J., et al. 2010, A&A, 518, L100+Onishi, T., Kawamura, A., Abe, R., et al. 1999, PASJ, 51, 871Onishi, T., Yoshikawa, N., Yamamoto, H., et al. 2001, PASJ, 53, 1017Paladini, R., Montier, L., Giard, M., et al. 2007, A&A, 465, 839Paradis, D., Bernard, J., & Meny, C. 2009, A&A, 506, 745Paradis, D. & et. al. 2011, in preparationPerrot, C. A. & Grenier, I. A. 2003, A&A, 404, 519Planck Collaboration. 2011a, Planck early results 01: The Planck mission

(Submitted to A&A)Planck Collaboration. 2011b, Planck early results 08: The all-sky early Sunyaev-

Zeldovich cluster sample (Submitted to A&A)Planck Collaboration. 2011c, Planck early results 17: Origin of the submillimetre

excess dust emission in the Magellanic Clouds (Submitted to A&A)Planck Collaboration. 2011d, Planck early results 18: The power spectrum of

cosmic infrared background anisotropies (Submitted to A&A)Planck Collaboration. 2011e, Planck early results 19: All-sky temperature and

dust optical depth from Planck and IRAS — constraints on the “dark gas” inour Galaxy (Submitted to A&A)

Planck Collaboration. 2011f, Planck early results 21: Properties of the interstellarmedium in the Galactic plane (Submitted to A&A)

Planck Collaboration. 2011g, Planck early results 22: The submillimetre proper-ties of a sample of Galactic cold clumps (Submitted to A&A)

Planck Collaboration. 2011h, Planck early results 23: The Galactic cold corepopulation revealed by the first all-sky survey (Submitted to A&A)

Planck Collaboration. 2011i, Planck early results 24: Dust in the diffuse inter-stellar medium and the Galactic halo (Submitted to A&A)

Planck Collaboration. 2011j, Planck early results 25: Thermal dust in nearbymolecular clouds (Submitted to A&A)

Planck HFI Core Team. 2011a, Planck early results 04: First assessment of theHigh Frequency Instrument in-flight performance (Submitted to A&A)

Planck HFI Core Team. 2011b, Planck early results 06: The High FrequencyInstrument data processing (Submitted to A&A)

Reach, W. T., Dwek, E., Fixsen, D. J., et al. 1995, ApJ, 451, 188Reach, W. T., Koo, B., & Heiles, C. 1994, ApJ, 429, 672Reach, W. T., Wall, W. F., & Odegard, N. 1998, ApJ, 507, 507Roman-Duval, J., Israel, F. P., Bolatto, A., et al. 2010, A&A, 518, L74+Savage, B. D., Bohlin, R. C., Drake, J. F., & Budich, W. 1977, ApJ, 216, 291Sodroski, T. J., Bennett, C., Boggess, N., et al. 1994, ApJ, 428, 638Stepnik, B., Abergel, A., Bernard, J., et al. 2003, A&A, 398, 551Wakker, B. P. 2006, ApJS, 163, 282Wolfire, M. G., Hollenbach, D., & McKee, C. F. 2010, ApJ, 716, 1191

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

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

3 Astroparticule et Cosmologie, CNRS (UMR7164), UniversiteDenis Diderot Paris 7, Batiment Condorcet, 10 rue A. Domon etLeonie Duquet, Paris, France

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

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

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

7 California Institute of Technology, Pasadena, California, U.S.A.

8 Centre of Mathematics for Applications, University of Oslo,Blindern, Oslo, Norway

9 DAMTP, Centre for Mathematical Sciences, Wilberforce Road,Cambridge CB3 0WA, U.K.

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

11 DTU Space, National Space Institute, Juliane Mariesvej 30,Copenhagen, Denmark

12 Departamento de Fısica, Universidad de Oviedo, Avda. CalvoSotelo s/n, Oviedo, Spain

13 Department of Astronomy and Astrophysics, University of Toronto,50 Saint George Street, Toronto, Ontario, Canada

14 Department of Astronomy and Earth Sciences, Tokyo GakugeiUniversity, Koganei, Tokyo 184-8501, Japan

15 Department of Physical Science, Graduate School of Science,Osaka Prefecture University, 1-1 Gakuen-cho, Naka-ku, Sakai,Osaka 599-8531, Japan

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

15

Planck collaboration: Constraints on the dark gas in our galaxy

Canada

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

18 Department of Physics, Nagoya University, Chikusa-ku, Nagoya,464-8602, Japan

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

20 Department of Physics, Purdue University, 525 NorthwesternAvenue, West Lafayette, Indiana, U.S.A.

21 Department of Physics, University of California, Berkeley,California, U.S.A.

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

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

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

25 Dipartimento di Fisica G. Galilei, Universita degli Studi di Padova,via Marzolo 8, 35131 Padova, Italy

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

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

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

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

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

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

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

33 European Space Agency, ESAC, Planck Science Office, Caminobajo del Castillo, s/n, Urbanizacion Villafranca del Castillo,

Villanueva de la Canada, Madrid, Spain

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

35 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street,Cambridge, MA 02138, U.S.A.

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

37 INAF - Osservatorio Astrofisico di Catania, Via S. Sofia 78,Catania, Italy

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

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

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

41 INAF/IASF Bologna, Via Gobetti 101, Bologna, Italy

42 INAF/IASF Milano, Via E. Bassini 15, Milano, Italy

43 INRIA, Laboratoire de Recherche en Informatique, UniversiteParis-Sud 11, Batiment 490, 91405 Orsay Cedex, France

44 IPAG: Institut de Planetologie et d’Astrophysique de Grenoble,Universite Joseph Fourier, Grenoble 1 / CNRS-INSU, UMR 5274,Grenoble, F-38041, France

45 Imperial College London, Astrophysics group, Blackett Laboratory,Prince Consort Road, London, SW7 2AZ, U.K.

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

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

48 Institut d’Astrophysique de Paris, CNRS UMR7095, UniversitePierre & Marie Curie, 98 bis boulevard Arago, Paris, France

49 Institut de Ciencies de l’Espai, CSIC/IEEC, Facultat de Ciencies,Campus UAB, Torre C5 par-2, Bellaterra 08193, Spain

50 Institute of Astronomy and Astrophysics, Academia Sinica, Taipei,Taiwan

51 Institute of Theoretical Astrophysics, University of Oslo, Blindern,Oslo, Norway

52 Instituto de Astrofısica de Canarias, C/Vıa Lactea s/n, La Laguna,Tenerife, Spain

53 Instituto de Fısica de Cantabria (CSIC-Universidad de Cantabria),Avda. de los Castros s/n, Santander, Spain

54 Jet Propulsion Laboratory, California Institute of Technology, 4800Oak Grove Drive, Pasadena, California, U.S.A.

55 Jodrell Bank Centre for Astrophysics, Alan Turing Building, Schoolof Physics and Astronomy, The University of Manchester, Oxford

16

Planck collaboration: Constraints on the dark gas in our galaxy

Road, Manchester, M13 9PL, U.K.

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

57 LERMA, CNRS, Observatoire de Paris, 61 Avenue del’Observatoire, Paris, France

58 Laboratoire AIM, IRFU/Service d’Astrophysique - CEA/DSM -CNRS - Universite Paris Diderot, Bat. 709, CEA-Saclay, F-91191Gif-sur-Yvette Cedex, France

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

60 Laboratoire de Physique Subatomique et de Cosmologie, CNRS,Universite Joseph Fourier Grenoble I, 53 rue des Martyrs, Grenoble,

France

61 Laboratoire de l’Accelerateur Lineaire, Universite Paris-Sud 11,CNRS/IN2P3, Orsay, France

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

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

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

65 National University of Ireland, Department of ExperimentalPhysics, Maynooth, Co. Kildare, Ireland

66 Niels Bohr Institute, Blegdamsvej 17, Copenhagen, Denmark

67 Observational Cosmology, Mail Stop 367-17, California Institute ofTechnology, Pasadena, CA, 91125, U.S.A.

68 Optical Science Laboratory, University College London, GowerStreet, London, U.K.

69 SISSA, Astrophysics Sector, via Bonomea 265, 34136, Trieste, Italy

70 SUPA, Institute for Astronomy, University of Edinburgh, RoyalObservatory, Blackford Hill, Edinburgh EH9 3HJ, U.K.

71 School of Physics and Astronomy, Cardiff University, QueensBuildings, The Parade, Cardiff, CF24 3AA, U.K.

72 Space Sciences Laboratory, University of California, Berkeley,California, U.S.A.

73 Spitzer Science Center, 1200 E. California Blvd., Pasadena,California, U.S.A.

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

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

76 Universities Space Research Association, Stratospheric Observatoryfor Infrared Astronomy, MS 211-3, Moffett Field, CA 94035, U.S.A.

77 University of Cambridge, Cavendish Laboratory, Astrophysicsgroup, J J Thomson Avenue, Cambridge, U.K.

78 University of Cambridge, Institute of Astronomy, Madingley Road,Cambridge, U.K.

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

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

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

17


Recommended