arxiv:0912.4029v2 [astro-ph.he] 21 jan 2010

17
arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010 DRAFT VERSION OCTOBER 25, 2018 Preprint typeset using L A T E X style emulateapj v. 04/20/08 PKS 1502+106: A NEW AND DISTANT GAMMA-RAY BLAZAR IN OUTBURST DISCOVERED BY THE FERMI LARGE AREA TELESCOPE A. A. ABDO 2,3 , M. ACKERMANN 4 , M. AJELLO 4 , W. B. ATWOOD 5 , M. AXELSSON 6,7 , L. BALDINI 8 , J. BALLET 9 , G. BARBIELLINI 10,11 , D. BASTIERI 12,13 , B. M. BAUGHMAN 14 , K. BECHTOL 4 , R. BELLAZZINI 8 , B. BERENJI 4 , E. D. BLOOM 4 , G. BOGAERT 15 , E. BONAMENTE 16,17 , A. W. BORGLAND 4 , J. BREGEON 8 , A. BREZ 8 , M. BRIGIDA 18,19 , P. BRUEL 15 , T. H. BURNETT 20 , G. A. CALIANDRO 21 , R. A. CAMERON 4 , P. A. CARAVEO 22 , J. M. CASANDJIAN 9 , E. CAVAZZUTI 23 , C. CECCHI 16,17 , ¨ O. C ¸ ELIK 24,25,26 , A. CHEKHTMAN 2,27 , C. C. CHEUNG 2,3 , J. CHIANG 4 , S. CIPRINI 17,1 , R. CLAUS 4 , J. COHEN-TANUGI 28 , J. CONRAD 29,7,30 , S. CUTINI 23 , C. D. DERMER 2 , A. DE ANGELIS 31 , F. DE PALMA 18,19 , S. W. DIGEL 4 , E. DO COUTO E SILVA 4 , P. S. DRELL 4 , R. DUBOIS 4 , D. DUMORA 32,33 , C. FARNIER 28 , C. FAVUZZI 18,19 , S. J. FEGAN 15 , E. C. FERRARA 24 , W. B. FOCKE 4 , M. FRAILIS 31 , L. FUHRMANN 34 , Y. FUKAZAWA 35 , S. FUNK 4 , P. FUSCO 18,19 , F. GARGANO 19 , D. GASPARRINI 23 , N. GEHRELS 24,36,37 , S. GERMANI 16,17 , B. GIEBELS 15 , N. GIGLIETTO 18,19 , F. GIORDANO 18,19 , M. GIROLETTI 38 , T. GLANZMAN 4 , G. GODFREY 4 , I. A. GRENIER 9 , M.-H. GRONDIN 32,33 , J. E. GROVE 2 , L. GUILLEMOT 34 , S. GUIRIEC 39 , Y. HANABATA 35 , A. K. HARDING 24 , M. HAYASHIDA 4 , E. HAYS 24 , R. E. HUGHES 14 , G. J ´ OHANNESSON 4 , A. S. J OHNSON 4 , R. P. J OHNSON 5 , W. N. J OHNSON 2 , M. KADLER 40,25,41,42 , T. KAMAE 4 , H. KATAGIRI 35 , J. KATAOKA 43 , M. KERR 20 , J. KN ¨ ODLSEDER 44 , M. L. KOCIAN 4 , F. KUEHN 14 , M. KUSS 8 , J. LANDE 4 , L. LATRONICO 8 , M. LEMOINE-GOUMARD 32,33 , F. LONGO 10,11 , F. LOPARCO 18,19 , B. LOTT 32,33 , M. N. LOVELLETTE 2 , P. LUBRANO 16,17 , G. M. MADEJSKI 4 , A. MAKEEV 2,27 , M. MARELLI 22 , E. MASSARO 45 , W. MAX-MOERBECK 46 , M. N. MAZZIOTTA 19 , W. MCCONVILLE 24,37 , J. E. MCENERY 24,37 , C. MEURER 29,7 , P. F. MICHELSON 4 , W. MITTHUMSIRI 4 , T. MIZUNO 35 , A. A. MOISEEV 25,37 , C. MONTE 18,19 , M. E. MONZANI 4 , A. MORSELLI 47 , I. V. MOSKALENKO 4 , S. MURGIA 4 , P. L. NOLAN 4 , J. P. NORRIS 48 , E. NUSS 28 , T. OHSUGI 35 , N. OMODEI 8 , E. ORLANDO 49 , J. F. ORMES 48 , M. OZAKI 50 , D. PANEQUE 4 , J. H. PANETTA 4 , D. PARENT 32,33 , V. PAVLIDOU 46 , T. J. PEARSON 46 , V. PELASSA 28 , M. PEPE 16,17 , M. PESCE-ROLLINS 8 , F. PIRON 28 , T. A. PORTER 5 , S. RAIN ` O 18,19 , R. RANDO 12,13 , M. RAZZANO 8 , S. RAZZAQUE 2,3 , A. READHEAD 46 , A. REIMER 51,4 , O. REIMER 51,4 , T. REPOSEUR 32,33 , J. L. RICHARDS 46 , S. RITZ 5,5 , L. S. ROCHESTER 4 , A. Y. RODRIGUEZ 21 , R. W. ROMANI 4 , M. ROTH 20 , F. RYDE 52,7 , H. F.-W. SADROZINSKI 5 , D. SANCHEZ 15 , A. SANDER 14 , P. M. SAZ PARKINSON 5 , J. D. SCARGLE 53 , C. SGR ` O 8 , M. S. SHAW 4 , E. J. SISKIND 54 , D. A. SMITH 32,33 , P. D. SMITH 14 , G. SPANDRE 8 , P. SPINELLI 18,19 , M. STEVENSON 46 , M. S. STRICKMAN 2 , D. J. SUSON 55 , H. TAJIMA 4 , H. TAKAHASHI 35 , T. TANAKA 4 , J. B. THAYER 4 , J. G. THAYER 4 , D. J. THOMPSON 24 , L. TIBALDO 12,13,9 , O. TIBOLLA 56 , D. F. TORRES 57,21 , G. TOSTI 16,17 , A. TRAMACERE 4,58 , P. UBERTINI 59 , Y. UCHIYAMA 4 , T. L. USHER 4 , V. VASILEIOU 25,26 , N. VILCHEZ 44 , V. VITALE 47,60 , A. P. WAITE 4 , P. WANG 4 , B. L. WINER 14 , K. S. WOOD 2 , H. YASUDA 35 , T. YLINEN 52,61,7 , J. A. ZENSUS 34 , M. ZIEGLER 5 , (THE FERMI LAT COLLABORATION), AND E. ANGELAKIS 34 , T. HOVATTA 62 , E. HOVERSTEN 36 , Y. I KEJIRI 35 , K. S. KAWABATA 63 , Y. Y. KOVALEV 64,34 ,YU. A. KOVALEV 64 , T. P. KRICHBAUM 34 , M. L. LISTER 65 , A. L ¨ AHTEENM ¨ AKI 62 , N. MARCHILI 34 , P. OGLE 46 , C. PAGANI 36 , A. B. PUSHKAREV 66,34,67 , K. SAKIMOTO 35 , M. SASADA 35 , M. TORNIKOSKI 62 , M. UEMURA 63 , M. YAMANAKA 35 , T. YAMASHITA 63 Draft version October 25, 2018 ABSTRACT The Large Area Telescope (LAT) on board the Fermi Gamma-ray Space Telescope discovered a rapid (5 days duration), high-energy (E> 100 MeV) gamma-ray outburst from a source identified with the blazar PKS 1502+106 (OR 103, S3 1502+10, z=1.839) starting on August 05, 2008 (23 UTC, MJD 54683.95), and followed by bright and variable flux over the next few months. Results on the gamma-ray localization and identification, as well as spectral and temporal behavior during the first months of the Fermi all-sky survey are reported here in conjunction with a multi-waveband characterization as a result of one of the first Fermi multi- frequency campaigns. The campaign included a Swift ToO (followed up by 16-day observations on August 07- 22, MJD 54685-54700),VLBA (within the MOJAVE program), Owens Valley (OVRO) 40m, Effelsberg-100m, Mets¨ ahovi-14m, RATAN-600 and Kanata-Hiroshima radio/optical observations. Results from the analysis of archival observations by INTEGRAL, XMM-Newton and Spitzer space telescopes are reported for a more complete picture of this new gamma-ray blazar. PKS 1502+106 is a sub-GeV peaked, powerful flat spectrum radio quasar (luminosity at E> 100 MeV, L γ , is about 1.1 × 10 49 erg s 1 , and black hole mass likely close to 10 9 M ), exhibiting marked gamma-ray bolometric dominance, in particular during the asymmetric outburst (L γ /L opt 100, and 5-day averaged flux F E>100 MeV =2.91 ± 1.4 × 10 6 ph cm 2 s 1 ), which was characterized by a factor greater than 3 of flux increase in less than 12 hours. The outburst was observed simultaneously from optical to X-ray bands (F 0.310 keV =2.18 +0.15 0.12 × 10 12 erg cm 2 s 1 , and hard photon index 1.5, similar to past values) with a flux increase of less than one order of magnitude with respect to past observations, and was likely controlled by Comptonization of external-jet photons produced in the broad line region (BLR) in the gamma-ray band. No evidence of a possible blue bump signature was observed in the optical-UV continuum spectrum, while some hints for a possible 4-day time-lag with respect to the gamma-ray flare were found. Nonetheless, the properties of PKS 1502+106 and the strict optical/UV, X- and gamma-ray cross correlations suggest the contribution of the synchrotron self Compton (SSC), in-jet, process should dominate from radio to X-rays. This mechanism may also be responsible for the consistent gamma-ray variability observed by the LAT on longer timescales, after the ignition of activity at these energies provided by the BLR-dissipated outburst. Modulations and subsequent minor, rapid flare events were detected, with a

Upload: others

Post on 03-Dec-2021

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

arX

iv:0

912.

4029

v2 [

astr

o-ph

.HE

] 21

Jan

201

0DRAFT VERSIONOCTOBER25, 2018Preprint typeset using LATEX style emulateapj v. 04/20/08

PKS 1502+106: A NEW AND DISTANT GAMMA-RAY BLAZAR IN OUTBURSTDISCOVERED BY THE FERMI LARGE AREA TELESCOPE

A. A. A BDO2,3, M. ACKERMANN4, M. AJELLO4, W. B. ATWOOD5, M. AXELSSON6,7, L. BALDINI 8, J. BALLET 9, G. BARBIELLINI 10,11,D. BASTIERI12,13, B. M. BAUGHMAN 14, K. BECHTOL4, R. BELLAZZINI 8, B. BERENJI4, E. D. BLOOM4, G. BOGAERT15,E. BONAMENTE16,17, A. W. BORGLAND4, J. BREGEON8, A. BREZ8, M. BRIGIDA18,19, P. BRUEL15, T. H. BURNETT20,

G. A. CALIANDRO 21, R. A. CAMERON4, P. A. CARAVEO22, J. M. CASANDJIAN9, E. CAVAZZUTI 23, C. CECCHI16,17, O. CELIK 24,25,26,A. CHEKHTMAN 2,27, C. C. CHEUNG2,3, J. CHIANG4, S. CIPRINI17,1, R. CLAUS4, J. COHEN-TANUGI28, J. CONRAD29,7,30, S. CUTINI 23,

C. D. DERMER2, A. DE ANGELIS31, F. DE PALMA 18,19, S. W. DIGEL4, E. DO COUTO E SILVA 4, P. S. DRELL4, R. DUBOIS4,D. DUMORA32,33, C. FARNIER28, C. FAVUZZI 18,19, S. J. FEGAN15, E. C. FERRARA24, W. B. FOCKE4, M. FRAILIS31, L. FUHRMANN34,Y. FUKAZAWA 35, S. FUNK4, P. FUSCO18,19, F. GARGANO19, D. GASPARRINI23, N. GEHRELS24,36,37, S. GERMANI16,17, B. GIEBELS15,

N. GIGLIETTO18,19, F. GIORDANO18,19, M. GIROLETTI38, T. GLANZMAN 4, G. GODFREY4, I. A. GRENIER9, M.-H. GRONDIN32,33,J. E. GROVE2, L. GUILLEMOT 34, S. GUIRIEC39, Y. HANABATA 35, A. K. HARDING24, M. HAYASHIDA 4, E. HAYS24, R. E. HUGHES14,

G. JOHANNESSON4, A. S. JOHNSON4, R. P. JOHNSON5, W. N. JOHNSON2, M. KADLER40,25,41,42, T. KAMAE 4, H. KATAGIRI 35,J. KATAOKA 43, M. KERR20, J. KNODLSEDER44, M. L. KOCIAN4, F. KUEHN14, M. KUSS8, J. LANDE4, L. LATRONICO8,M. L EMOINE-GOUMARD32,33, F. LONGO10,11, F. LOPARCO18,19, B. LOTT32,33, M. N. LOVELLETTE2, P. LUBRANO16,17,G. M. MADEJSKI4, A. MAKEEV2,27, M. MARELLI 22, E. MASSARO45, W. MAX -MOERBECK46, M. N. MAZZIOTTA 19,W. MCCONVILLE 24,37, J. E. MCENERY24,37, C. MEURER29,7, P. F. MICHELSON4, W. MITTHUMSIRI4, T. MIZUNO35,

A. A. M OISEEV25,37, C. MONTE18,19, M. E. MONZANI4, A. MORSELLI47, I. V. M OSKALENKO4, S. MURGIA4, P. L. NOLAN4,J. P. NORRIS48, E. NUSS28, T. OHSUGI35, N. OMODEI8, E. ORLANDO49, J. F. ORMES48, M. OZAKI 50, D. PANEQUE4, J. H. PANETTA4,D. PARENT32,33, V. PAVLIDOU 46, T. J. PEARSON46, V. PELASSA28, M. PEPE16,17, M. PESCE-ROLLINS8, F. PIRON28, T. A. PORTER5,

S. RAIN O18,19, R. RANDO12,13, M. RAZZANO8, S. RAZZAQUE2,3, A. READHEAD46, A. REIMER51,4, O. REIMER51,4, T. REPOSEUR32,33,J. L. RICHARDS46, S. RITZ5,5, L. S. ROCHESTER4, A. Y. RODRIGUEZ21, R. W. ROMANI 4, M. ROTH20, F. RYDE52,7,

H. F.-W. SADROZINSKI5, D. SANCHEZ15, A. SANDER14, P. M. SAZ PARKINSON5, J. D. SCARGLE53, C. SGRO8, M. S. SHAW4,E. J. SISKIND54, D. A. SMITH 32,33, P. D. SMITH 14, G. SPANDRE8, P. SPINELLI18,19, M. STEVENSON46, M. S. STRICKMAN 2,

D. J. SUSON55, H. TAJIMA 4, H. TAKAHASHI 35, T. TANAKA 4, J. B. THAYER4, J. G. THAYER4, D. J. THOMPSON24, L. TIBALDO 12,13,9,O. TIBOLLA 56, D. F. TORRES57,21, G. TOSTI16,17, A. TRAMACERE4,58, P. UBERTINI59, Y. UCHIYAMA 4, T. L. USHER4,

V. VASILEIOU25,26, N. VILCHEZ44, V. V ITALE 47,60, A. P. WAITE4, P. WANG4, B. L. WINER14, K. S. WOOD2, H. YASUDA35,T. YLINEN52,61,7, J. A. ZENSUS34, M. ZIEGLER5,

(THE FERMI LAT COLLABORATION),AND

E. ANGELAKIS34, T. HOVATTA 62, E. HOVERSTEN36, Y. IKEJIRI35, K. S. KAWABATA 63, Y. Y. KOVALEV 64,34, YU. A. KOVALEV 64,T. P. KRICHBAUM34, M. L. L ISTER65, A. L AHTEENMAKI 62, N. MARCHILI 34, P. OGLE46, C. PAGANI 36, A. B. PUSHKAREV66,34,67,

K. SAKIMOTO 35, M. SASADA35, M. TORNIKOSKI62, M. UEMURA63, M. YAMANAKA 35, T. YAMASHITA 63

Draft version October 25, 2018

ABSTRACTThe Large Area Telescope (LAT) on board theFermi Gamma-ray Space Telescope discovered a rapid (∼

5 days duration), high-energy (E > 100 MeV) gamma-ray outburst from a source identified with the blazarPKS 1502+106 (OR 103, S3 1502+10, z=1.839) starting on August 05, 2008 (∼ 23 UTC, MJD 54683.95),and followed by bright and variable flux over the next few months. Results on the gamma-ray localization andidentification, as well as spectral and temporal behavior during the first months of theFermiall-sky survey arereported here in conjunction with a multi-waveband characterization as a result of one of the firstFermi multi-frequency campaigns. The campaign included aSwiftToO (followed up by 16-day observations on August 07-22, MJD 54685-54700), VLBA (within the MOJAVE program), Owens Valley (OVRO) 40m, Effelsberg-100m,Metsahovi-14m, RATAN-600 and Kanata-Hiroshima radio/optical observations. Results from the analysis ofarchival observations by INTEGRAL, XMM-Newtonand Spitzerspace telescopes are reported for a morecomplete picture of this new gamma-ray blazar. PKS 1502+106is a sub-GeV peaked, powerful flat spectrumradio quasar (luminosity atE > 100 MeV, Lγ , is about1.1× 1049 erg s−1, and black hole mass likely close to109 M⊙), exhibiting marked gamma-ray bolometric dominance, in particular during the asymmetric outburst(Lγ/Lopt ∼ 100, and 5-day averaged flux FE>100 MeV = 2.91 ± 1.4 × 10−6 ph cm−2 s−1), which wascharacterized by a factor greater than 3 of flux increase in less than 12 hours. The outburst was observedsimultaneously from optical to X-ray bands (F0.3−10 keV = 2.18+0.15

−0.12× 10−12 erg cm−2 s−1, and hard photonindex∼ 1.5, similar to past values) with a flux increase of less than one order of magnitude with respect topast observations, and was likely controlled by Comptonization of external-jet photons produced in the broadline region (BLR) in the gamma-ray band. No evidence of a possible blue bump signature was observedin the optical-UV continuum spectrum, while some hints for apossible 4-day time-lag with respect to thegamma-ray flare were found. Nonetheless, the properties of PKS 1502+106 and the strict optical/UV, X- andgamma-ray cross correlations suggest the contribution of the synchrotron self Compton (SSC), in-jet, processshould dominate from radio to X-rays. This mechanism may also be responsible for the consistent gamma-rayvariability observed by the LAT on longer timescales, afterthe ignition of activity at these energies providedby the BLR-dissipated outburst. Modulations and subsequent minor, rapid flare events were detected, with a

Page 2: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

2 Abdo et al.

general fluctuation mode between pink-noise and a random-walk. The averaged gamma-ray spectrum showeda deviation from a simple power-law, and can be described by alog-parabola curved model peaking around 0.4-0.5 GeV. The maximum energy of photons detected from the source in the first four months of LAT observationswas15.8GeV, with no significant consequences on extragalactic background light predictions. A possible radiocounterpart of the gamma-ray outburst can be assumed only ifa delay of more than 3 months is consideredon the basis of opacity effects at cm and longer wavelengths.The rotation of the electric vector position angleobserved by VLBA from 2007 to 2008 could represent a slow fieldordering and alignment with respect to thejet axis, likely a precursor feature of the ejection of a superluminal radio knot and the high-energy outburst.This observing campaign provides more insight into the connection between MeV-GeV flares and the moving,polarized structures observed by the VLBI.Subject headings:gamma-rays: observations – quasars: individual: PKS 1502+106 – quasars: general – galax-

ies: active – galaxies: jets – X-rays: galaxies

1 Corresponding author: S. Ciprini, [email protected] Space Science Division, Naval Research Laboratory, Washington, DC

20375, USA3 National Research Council Research Associate, National Academy of

Sciences, Washington, DC 20001, USA4 W. W. Hansen Experimental Physics Laboratory, Kavli Institute for Parti-

cle Astrophysics and Cosmology, Department of Physics and SLAC NationalAccelerator Laboratory, Stanford University, Stanford, CA 94305, USA

5 Santa Cruz Institute for Particle Physics, Department of Physics and De-partment of Astronomy and Astrophysics, University of California at SantaCruz, Santa Cruz, CA 95064, USA

6 Department of Astronomy, Stockholm University, SE-106 91 Stockholm,Sweden

7 The Oskar Klein Centre for Cosmoparticle Physics, AlbaNova, SE-10691 Stockholm, Sweden

8 Istituto Nazionale di Fisica Nucleare, Sezione di Pisa, I-56127 Pisa, Italy9 Laboratoire AIM, CEA-IRFU/CNRS/Universite Paris Diderot, Service

d’Astrophysique, CEA Saclay, 91191 Gif sur Yvette, France10 Istituto Nazionale di Fisica Nucleare, Sezione di Trieste,I-34127 Tri-

este, Italy11 Dipartimento di Fisica, Universita di Trieste, I-34127 Trieste, Italy12 Istituto Nazionale di Fisica Nucleare, Sezione di Padova, I-35131

Padova, Italy13 Dipartimento di Fisica “G. Galilei”, Universita di Padova, I-35131

Padova, Italy14 Department of Physics, Center for Cosmology and Astro-Particle

Physics, The Ohio State University, Columbus, OH 43210, USA15 Laboratoire Leprince-Ringuet,Ecole polytechnique, CNRS/IN2P3,

Palaiseau, France16 Istituto Nazionale di Fisica Nucleare, Sezione di Perugia,I-06123 Peru-

gia, Italy17 Dipartimento di Fisica, Universita degli Studi di Perugia, I-06123 Peru-

gia, Italy18 Dipartimento di Fisica “M. Merlin” dell’Universita e del Politecnico di

Bari, I-70126 Bari, Italy19 Istituto Nazionale di Fisica Nucleare, Sezione di Bari, 70126 Bari, Italy20 Department of Physics, University of Washington, Seattle,WA 98195-

1560, USA21 Institut de Ciencies de l’Espai (IEEC-CSIC), Campus UAB, 08193

Barcelona, Spain22 INAF-Istituto di Astrofisica Spaziale e Fisica Cosmica, I-20133 Milano,

Italy23 Agenzia Spaziale Italiana (ASI) Science Data Center, I-00044 Frascati

(Roma), Italy24 NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA25 Center for Research and Exploration in Space Science and Technology

(CRESST) and NASA Goddard Space Flight Center, Greenbelt, MD 20771,USA

26 Department of Physics and Center for Space Sciences and Technology,University of Maryland Baltimore County, Baltimore, MD 21250, USA

27 George Mason University, Fairfax, VA 22030, USA28 Laboratoire de Physique Theorique et Astroparticules, Universite Mont-

pellier 2, CNRS/IN2P3, Montpellier, France29 Department of Physics, Stockholm University, AlbaNova, SE-106 91

Stockholm, Sweden30 Royal Swedish Academy of Sciences Research Fellow, funded by a

grant from the K. A. Wallenberg Foundation31 Dipartimento di Fisica, Universita di Udine and Istituto Nazionale di

Fisica Nucleare, Sezione di Trieste, Gruppo Collegato di Udine, I-33100Udine, Italy

32 Universite de Bordeaux, Centre d’Etudes Nucleaires Bordeaux Gradig-nan, UMR 5797, Gradignan, 33175, France

33 CNRS/IN2P3, Centre d’Etudes Nucleaires Bordeaux Gradignan, UMR5797, Gradignan, 33175, France

34 Max-Planck-Institut fur Radioastronomie, Auf dem Hugel69, 53121Bonn, Germany

35 Department of Physical Sciences, Hiroshima University, Higashi-Hiroshima, Hiroshima 739-8526, Japan

36 Department of Astronomy and Astrophysics, Pennsylvania State Uni-versity, University Park, PA 16802, USA

37 Department of Physics and Department of Astronomy, University ofMaryland, College Park, MD 20742, USA

38 INAF Istituto di Radioastronomia, 40129 Bologna, Italy39 Center for Space Plasma and Aeronomic Research (CSPAR), University

of Alabama in Huntsville, Huntsville, AL 35899, USA40 Dr. Remeis-Sternwarte Bamberg, Sternwartstrasse 7, D-96049 Bam-

berg, Germany41 Erlangen Centre for Astroparticle Physics, D-91058 Erlangen, Germany42 Universities Space Research Association (USRA), Columbia, MD

21044, USA43 Waseda University, 1-104 Totsukamachi, Shinjuku-ku, Tokyo, 169-

8050, Japan44 Centre d’Etude Spatiale des Rayonnements, CNRS/UPS, BP 44346, F-

30128 Toulouse Cedex 4, France45 Universita di Roma “La Sapienza”, I-00185 Roma, Italy46 Cahill Center for Astronomy and Astrophysics, California Institute of

Technology, Pasadena, CA 91125, USA47 Istituto Nazionale di Fisica Nucleare, Sezione di Roma “TorVergata”,

I-00133 Roma, Italy48 Department of Physics and Astronomy, University of Denver,Denver,

CO 80208, USA49 Max-Planck Institut fur extraterrestrische Physik, 85748 Garching, Ger-

many50 Institute of Space and Astronautical Science, JAXA, 3-1-1 Yoshinodai,

Sagamihara, Kanagawa 229-8510, Japan51 Institut fur Astro- und Teilchenphysik and Institut fur Theoretische

Physik, Leopold-Franzens-Universitat Innsbruck, A-6020 Innsbruck, Austria52 Department of Physics, Royal Institute of Technology (KTH), Al-

baNova, SE-106 91 Stockholm, Sweden53 Space Sciences Division, NASA Ames Research Center, Moffett Field,

CA 94035-1000, USA54 NYCB Real-Time Computing Inc., Lattingtown, NY 11560-1025, USA55 Department of Chemistry and Physics, Purdue University Calumet,

Hammond, IN 46323-2094, USA56 Max-Planck-Institut fur Kernphysik, D-69029 Heidelberg, Germany57 Institucio Catalana de Recerca i Estudis Avancats (ICREA), Barcelona,

Spain58 Consorzio Interuniversitario per la Fisica Spaziale (CIFS), I-10133

Torino, Italy59 INAF-Istituto di Astrofisica Spaziale e Fisica Cosmica, I-00133 Roma,

Italy60 Dipartimento di Fisica, Universita di Roma “Tor Vergata”,I-00133

Roma, Italy61 School of Pure and Applied Natural Sciences, University of Kalmar,

SE-391 82 Kalmar, Sweden

Page 3: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

PKS 1502+106: a new and distant gamma-ray blazar in outburstdiscovered by the Fermi LAT 3

1. INTRODUCTION

The Large Area Telescope (LAT), on board theFermiGamma-ray Space Telescope (formerly GLAST; Ritz 2007),was successfully launched by NASA on 2008, June 11, fromCape Canaveral, Florida, on a Delta II Heavy launch vehi-cle. While still in the commissioning and checkout phase,it discovered and monitored bright, flaring gamma-ray emis-sion above 100 MeV from a source identified with the blazarPKS 1502+106 (historically also known as OR 103 and S31502+10). The large field of view, effective area and sensitiv-ity and the nominal survey observational mode makeFermi-LAT an unprecedented all-sky monitor ofγ-ray flares andsource variability (see, e.g. McEnery 2006; Thompson 2006;Lott et al. 2007; Atwood et al. 2009).

At the beginning of August 2008, PKS 1502+106 was thesecond brightest extragalactic source in theγ-ray sky, ex-hibiting a sudden high-energy outburst announced in ATel#1650. This outburst successfully triggered the first (un-planned)Fermi multi-frequency campaign. Major renewedgamma-ray activity observed byFermi in January 2009 wasannounced via ATel #1905.

PKS 1502+106 is a luminous, quasar-like (optically broad-line and flat radio spectrum) AGN discovered during the 178MHz pencil beam survey from the Mullard Radio AstronomyObservatory, Cambridge, UK, (appearing in a list not includedin the 4C catalog; Crowther & Clarke 1966; Williams et al.1967), and was re-observed and characterized as an extra-galactic source by both the Australian National Radio Astron-omy Observatory of Parkes, NSW, Australia, (Day et al. 1966,id.: PKS 1502+106), and the Ohio State University (“BigEar”) Radio Observatory, Delaware, OH, USA, (Fitch et al.1969, id.: OR 103). The source exhibited substantial radioflux variations (factor> 2), a high degree of linear polariza-tion, a core-dominated, one-sided and curved radio jet witha variable, a complex morphology at VLBI scales (An et al.2004; Lister et al. 2009a), and a compact large scale struc-ture. 11 VLBA observations at 15.4 GHz performed betweenAug. 1997 and Aug. 2007 showed a FWHM major beamaxis in the range 1.02-1.57 mas, a minor axis beam axis of0.5 mas, and a total flux density in the range 0.88-1.93 Jy(Lister et al. 2009a), and apparent jet speed of(14.8 ± 1.2)c(Lister et al. 2009b). The 22 and 37 GHz flux history showsseveral long-term flares (> 1 Jy variations, i.e.∼ 60% of thetotal flux range span, on typical timescales of a year, and peakfluxes well above 2 Jy), with at least five flares and an aver-age trend that was slightly increasing from 1988 to mid-2004(Terasranta et al. 2005). WMAP fluxes at similar frequencies(K, Ka, Q bands) are in agreement with these flux ranges(Lopez-Caniego et al. 2007). The Doppler factor estimatedfrom the observed 37 GHz variability and brightness temper-ature (Hovatta et al. 2009) agrees with the jet speed (14.6c)cited above, a Doppler factorDvar = 12 and viewing angleθvar = 4.7.

62 Metsahovi Radio Observatory, Helsinki University of Technology TKK,FIN-02540 Kylmala, Finland

63 Hiroshima Astrophysical Science Center, Hiroshima University,Higashi-Hiroshima, Hiroshima 739-8526, Japan

64 Astro Space Center of the Lebedev Physical Institute, 117810 Moscow,Russia

65 Department of Physics, Purdue University, West Lafayette,IN 47907,USA

66 Crimean Astrophysical Observatory, 98409 Nauchny, Crimea, Ukraine67 Pulkovo Observatory, 196140 St. Petersburg, Russia* Correspondence: [email protected]

This radio blazar was identified in the optical band byBlake (1970) with a position refinement by Argue & Sullivan(1980), while an initial spectroscopic inspection was per-formed by Burbidge & Strittmatter (1972). Variations> 2.5mag were observed in its optical flux history (Palomar-Questand Catalina Sky Surveys, ATel #1661), together with a vari-able and relatively high degree (up to20%) of linear polar-ization, pointing out a dominant synchrotron emission withno observed dilution by thermal components. The redshift ofPKS 1502+106, as confirmed by the good S/N spectrum of theSloan Digital Sky Survey (z = 1.8385± 0.0024 at high con-fidence), is in agreement with the valuez = 1.839 estimatedpreviously by Smith et al. (1977). A less remote value (z =0.56) is reported in other works (Burbidge & Strittmatter1972; Wright et al. 1979; Wilkes et al. 1983), although thepossible multiple MgII absorption system (pointed out by afeature shortward of the 4388A emission line) would be veryunusual for a low redshift object.

Serendipitous X-ray data of PKS 1502+106 are availablebecause the source lies about7′ NE of the bright Seyfert type-1 galaxy Mkn 841, although only one multifrequency workdedicated to this blazar (George et al. 1994) has appeared pre-viously. Early X-ray observations (ROSAT, ASCA) showedlow-amplitude variations on short timescales (factor> 2 ontimescales of a year), a flat 0.1-10 keV photon indexΓX

between 1.4 and 1.9, and an intrinsic X-ray luminosity ofL2−10keV = 1.2 × 1046 erg s−1, and a 2-10 keV flux in therange4.9 − 6.54 × 10−13 erg cm−2 s−1(George et al. 1994;Akiyama et al. 2003; Watanabe et al. 2004). PKS 1502+106was speculated to be a possibleγ-ray source before theLAT discovery because of the superluminal motions of jetcomponents (up to187 ± 15µas/year Lister & Homan 2005;An et al. 2004), and the multiwaveband spectral indexesαrx

andαox (consistent with other FSRQs detected by EGRET,George et al. 1994). Only modest intrinsic X-ray absorptionwas suggested by this work, and the optical and near-IR red-dening claimed in Watanabe et al. (2004) is probably due tothe synchrotron jet dominance at these low frequencies ratherthan by absorption from inner nuclear light.

A relation involving the misalignment between the pc-and kpc-scale radio structure (position angle) and theγ-ray emission was postulated as well (Cooper et al. 2007).However, only a cumulative2σ upper limit by EGRETof 7 × 10−8 ph cm−2 s−1 was reported (Phase/Cycle I,combined Viewing Periods: 24.0 to 25.0, i.e. April 02-23,1992; Fichtel et al. 1994), and the source was likewise unde-tected in the following EGRET cycles (Hartman et al. 1999;Casandjian & Grenier 2008).

In the following we use aΛCDM (concordance) cos-mology with values given within 1σ of the WMAP results(Komatsu et al. 2009), namelyh = 0.71, Ωm = 0.27 andΩΛ = 0.73, and a Hubble constant valueH0 = 100h km s−1

Mpc−1.In Section 2, first results on theγ-ray identification, the ob-

served MeV-GeV outburst and the subsequent four monthsof monitoring by theFermi-LAT are described. In Sec-tion 3, multifrequency results obtained through simultaneousoptical-UV-X-ray observations bySwift (thanks to a 16-daylong monitoring following a triggered Target of Opportunity,ToO), and by radio-optical observatories (the 40m dish tele-scope of the Owens Valley Radio Observatory, the Effelsberg100m dish radio telescope, the ring radio telescope RATAN-600, the VLBA within the MOJAVE program, and the Kanata

Page 4: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

4 Abdo et al.

FIG. 1.— Left panel: LAT count map cumulated on a nine-month (Aug.2008 - Apr. 2009) baseline, weighted and smoothed by the point spreadfunction (PSF) such that higher energy photons are mapped tohigher in-tensities. The map is in arbitrary units in the energy range 0.1-100 GeV andin a 2.5 × 2.5 region centered on PKS 1502+106. The qualitative circlesizes of the PSF at 200MeV, 2GeV and 20GeV are outlined for reference.Right panel: LAT source localization with 95% and 68% uncertainty radii(red circles) superimposed on an arcmin-scale (R-band) optical image show-ing also the X-ray counterpart error box by theSwift-XRT observations andthe radio position and intensity contours by VLA of PKS 1502+106. Thebest LAT source position, calculated on the same nine-monthperiod, withthepointlike tool is RA: 226.10179 , Dec: +10.4927, ∆ = 0.0027,with 68% and 95% LAT error circles of0.0077 and0.0124 respectively.

telescope of the Higashi-Hiroshima Observatory) are summa-rized. In addition, past and unpublished observations by theXMM-Newton and Spitzer space telescopes are analyzed andpresented in Section 4 for a more complete picture. Finally,in Section 5 and 6, discussion and concluding remarks are re-ported.

2. GAMMA-RAY OBSERVATIONS AND RESULTS BYFERMI-LAT

2.1. LAT observations

The LAT instrument is a pair tracker-converter telescopecomprising a modular array of 16 towers—each with atracker based on silicon micro-strip detector technology—and a calorimeter based on a hodoscopic array of 96 CsI(Tl)crystals, surrounded by an Anti-Coincidence Detector capa-ble of measuring the directions and energies of cosmicγ-ray photons with energies from 20 MeV to> 300 GeV (fordetails, see, e.g. Bellazzini et al. 2002; Michelson 2007;Atwood et al. 2007, 2009; Abdo et al. 2009h).

The reduction and analysis of LAT data was performed us-ing the Science Tools v.9.8, based in particular on a unbinnedmaximum-likelihood estimator of the spectral model param-eters (gtlike tool). Events were selected using the Instru-ment Response Functions (IRFs) P6V1 DIFFUSE. This se-lection provides the cleanest set of events (in terms of direc-tion, energy reconstruction and background rejection) at thecost of reduced effective area at low energies, and takes intoaccount the differences between front- and back-convertingevents. To minimize contamination by Earth albedoγ-rayevents that have reconstructed directions with angles withre-spect to the local zenith> 105 have been excluded. For thisobject with high Galactic latitude, events are extracted withina 10 acceptance cone centered at the PKS 1502+106 radioposition. This cone, substantially larger than the 68% contain-ment angle of the PSF at the lowest energies, provides suffi-cient events to accurately constrain the diffuse emission com-ponents. Thegtlike model includes the PKS 1502+106point source component, two other point sources from the 3month catalog (both faint and low-confidence sources withTS ≃ 0.9% of the TS value of PKS 1502+106 for the sameperiod), a component for the Galactic diffuse emission (GAL-PROP code, see, e.g. Moskalenko et al. 2003, and references

therein), and an isotropic component including the extragalac-tic diffuse emission and the residual background from cosmicrays.

TheFermi-LAT data of PKS 1502+106 presented here wereobtained during the first four months of the LAT survey (Aug.-Dec. 2008). In this period PKS 1502+106 was one of themost persistently bright, variable sources in the high-energysky and almost certainly the source with the highest luminos-ity. The background contribution within a few degrees wasonly a small fraction of the source count rate, with no nearbysource confusion. The time interval was sufficient for a finedetermination of the average spectrum, for a first look at themid-timescale variability and detection of posterior flares, fora refined localization, and a first cross comparison with theother multifrequency monitoring data. The firstγ-ray detec-tion of PKS 1502+106 by the LAT occurred in July 2008,when it was confirmed by the high-level Automatic ScienceProcessing pipeline monitoring (ASP; Chiang et al. 2006,2007), based on a wavelet-based (pgwave) quick-look de-tection tool (e.g. Damiani et al. 1997; Marcucci et al. 2004;Ciprini et al. 2007a) and a maximum likelihood analysis, andby the LAT Source Catalog algorithm (mr filter) based onwavelet analysis in the Poisson regime (Starck & Pierre 1998)and the peak-finding toolsExtractor (Bertin & Arnouts1996). The rapid and markedly time-asymmetricγ-ray out-burst announced in ATel #1650 and triggering an unplannedTarget of Opportunity (ToO) multifrequency campaign wasseen from Aug. 05 (about 23 UTC) until about Aug. 11, 2008(∼ 5 days duration), with a fast rise, slower decay, and anapproximately two-day sustained peak flux.

Some caveats related to the preflight instrument responsefunctions (P6V1), which overestimated the acceptance at lowenergies, are briefly described in Abdo et al. (2009b).

2.2. Gamma-ray source localization, association andidentification

The LAT PSF and sensitivity provides an unprecedentedangular resolution in gamma-rays (68% containment ra-dius better than∼ 1 at 1 GeV, Atwood et al. 2007;Cecchi et al. 2007; Atwood et al. 2009; Abdo et al. 2009c,h),making the association and identification processes less dif-ficult than in previous experiments. In the case of this verybright γ-ray source, we obtained—beyond the good spatial“association”—a firm “identification” with PKS 1502+106.The 3-month bright source list results (Abdo et al. 2009c, id:0FGL J1504.4+1030), provided a good initial localization:RA: 226.12, Dec: +10.51; r95 = 0.05 and

√TS = 88.2,

(r95 being the radius of 95% confidence region, TS the like-lihood test statistic from the 200 MeV to 100 GeV analy-sis). Application of thepointlike tool (Burnett 2007;Abdo et al. 2009c) on a much longer (nine-month, Aug. 2008- Apr. 2009) LAT dataset with very high statistics, providedan excellent localization (outlined in Fig.1): RA: 226.10179,Dec: +10.4927,∆ = 0.0027, with 68% and 95% LAT errorcircles of0.0077 and0.0124 respectively (statistical only).Studies of bright source localizations indicate a systematic un-certainty in the localization of< 30”, that can be taken as anestimate of the systematics with this tool (more details on theproduction ofpointlike density maps and localization aredescribed in Camilo et al. 2009). The relevant improvementin the localization carried out on a nine-month baseline is dueto the high variability that occurred with this source. Theselocalization values are in agreement with the VLBI radio and

Page 5: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

PKS 1502+106: a new and distant gamma-ray blazar in outburstdiscovered by the Fermi LAT 5

FIG. 2.— Main panel:Likelihood flux (E>100MeV) light curve obtained in daily bins from Aug. 02 to Dec. 15, 2008. The outburst state and the subsequentpost-flare (a lower and intermediate level brightness, far from the faintest state observed) period with simultaneous monitoring bySwift are represented by thetwo horizontal lines.Left inset plot:A zoom on the corresponding flux light curve around the outburst period obtained using finer, 12-hour bins (lower statistics).Right inset plot:The gamma-ray (E>100MeV) photon index values for the same period using daily bins as in the main panel light curve.

optical positions of PKS 1502+106, the VLA contours andtheSwiftXRT error box (Fig. 1). PKS 1502+106 is the onlybright VLA radio source (calibrator source list) located withinthe LAT 95% confidence circle. The Seyfert galaxy Mkn 841(observed with a hard X-ray cutoff, see Sect.4) is positionedwell outside of these localization circles, as are other HB89-catalog quasars in this region.

Beyond the excellent spatial association, the most secureand distinctive signature for firm identification of this newgamma-ray source found byFermi is the observed correla-tion between theγ-ray, X-ray and optical-UV variability (seesections 3 and 4). This object was also a member of thepre-launch CGRaBS (Healey et al. 2008, object id: CGRaBSJ1504+1029) and Roma-BZCAT (Massaro et al. 2009, objectid: BZQ J1504+1029 ) catalogs listing candidate gamma-rayblazars. Finally, a method based on a “figure of merit” (de-scribed in Sowards-Emmerd et al. 2003, 2005) for this LATsource position provides a very high likelihood of identifica-tion with PKS 1502+106.

2.3. Gamma-ray temporal behavior

The typically brightγ-ray flux and the enduring activityshown by PKS 1502+106 inγ-rays, allowed a firm detec-tion of the source by the LAT on a daily basis. Fig. 2shows the light curve (daily bins,E > 100MeV) extractedwith the gtlike tool over the first four and half monthsof LAT all-sky survey. A fast-rising, markedly asymmet-ric and bright outburst was found, with a factor> 3 of in-crease in flux in less than 12 hours. The integrated flux atE > 100 MeV averaged in the 12h bin of the peak emission(between Aug. 05 and Aug. 06, 2008, i.e. DOY 218-219)wasFE>100 MeV = (3.7± 0.7)× 10−6 ph cm−2 s−1 (statis-tical only), as measured when the LAT instrument was still incommissioning and checkout phase (all-sky nominal mode).The emission from PKS 1502+106 then faded more slowly inthe following days, and the entire outburst interval spannedAug. 05 around 23 UTC until to Aug. 11 around 00 UTC,2008 (i.e. DOY, 218.95-224.0,∼ 5 days duration Fig. 2).The peak flux appeared elevated for less than two days, rival-ing the brighest apparent flux of other extragalactic sources atthat time (Section 2.1). The finer, 12-hour bin light curve (∼ 8Fermi orbits, ensuring still similar exposures) reported in theright inset panel of Fig. 2, shows the elevated flux held forat least one 1.5 days, while the slower fading decay exhibits

FIG. 3.— Time series analysis of the LAT light curve presented inFig.2:the periodogram, first order structure function (upper inset) and power spec-trum (lower inset). These functions show a variability witha power spectrumconsistent with1/f1.3 fluctuations. This indicates a variability mode placedbetween flickering and shot noise. The horizontal dashed line represents the0.01 false alarm probability threshold (99% significance that the signal de-tection is not wrong).

a high flux (FE>100 MeV > 2 ×10−6 ph cm−2 s−1) plateau,or a secondary superposed flare, that extended for about 2.5days. During this outburst a “harder when brighter” spectraltrend is suggested, despite the photon index error dispersion(see the daily photon indices reported on the right inset panelof Fig. 2). Thegtlike performances and the current IRFsused did not allow to go below a daily binning in the extrap-olation of the photon index values, and this was possible inan acceptable way with respect to amplitude of the statisticalerror only for the high-flux and high photon count statisticsavailable around the outburst epoch.

A consistent level of variability, with a couple of minor butrelevant rapid flares, occurred after the major outburst withfluctuations on timescales of weeks. Renewed activity and in-creased average brightness from the end of Nov. 2008 wereobserved. Two rapid flares approached a maximum peak fluxof FE>100 MeV ∼ 2 ×10−6 ph cm−2 s−1(daily bin estima-tions), on Sept. 06 (DOY 250), where a simultaneous cross-correlated optical flare was observed as well, and on Oct. 30(DOY 304). Visual inspection of the light curve reported inFig.2 suggests a period of higher activity beginning in midNov. 2008 (after about DOY 320), and, in general, a series of

Page 6: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

6 Abdo et al.

modulations occurring on about a one-month timescale, withfaster fluctuations and rapid flare episodes superposed. Thepower spectral density (PSD) shows a power-law dependence,P (f) ∝ 1/f1.3. Similar time scale dependence is exhibitedby the first order structure function (Fig. 3) and by the au-tocorrelation function. More detailed variability analysis forthis and other blazars using a longer dataset will be presentedin Abdo et al. (2009d).

2.4. Gamma-ray spectra

We have analyzed the time-averaged spectra of PKS1502+106 for three intervals: the high-state of the outburst(DOY, 218.95-224.0, i.e. Aug. 05 - Aug. 11, 2008, about5 days); the post-flare period characterized by an intermedi-ate brightness level and during which simultaneousSwiftob-servations were performed (DOY 224.0-235.42, i.e.Aug. 11-Aug. 22, 2008, about 11.4 days); the longer and heteroge-neous period that includes the outburst and the following 4months ofFermi-LAT monitoring (Aug.02 to Dec.07, 2008,∼ 126 days, where the source displayed different stages ofactivity and significant variability). Events below 200 MeVwere excluded from these analyses because of calibration un-certainties at those energies. An isotropic background modelwas used as PKS 1502+106 was very bright relative to otherpoint sources during the period stated above, because is lo-cated at high Galactic latitude, and because checks with morecomplex models provided no significant difference. Further-more, no appreciable differences were observed using differ-ent acceptance cone radii for the event extraction.

The spectra for the post-flare and cumulative 4-monthdatasets can be consistent with a log-parabola (LP) model,dN/dE ∝ E−(α+β log(E)) (see, e.g. Landau et al. 1986;Inoue & Takahara 1996; Fossati et al. 2000; Massaro et al.2004; Perlman et al. 2005). The likelihood ratio test(Mattox et al. 1996) rejects the hypothesis that the spectrumis a power law (null hypothesis) against the one that the spec-trum is curved as a log-parabola model. This model is pre-ferred over a simple power-law model at the 11.4 sigma sig-nificance level. Broken power-law (BPL) fits show a similarimprovement over single power-law models, but we find noevidence to prefer BPL over the log-parabola representation.For the full time interval characterized by very high statistics,the logarithm of the likelihood increases significantly whenallowing β to vary, and an increase of the value for the BPLwith respect to simple power-law model of the same order ofthe increase for the LP vs. power-law test is not observed.

TABLE 1SUMMARY OF THE UNBINNED LIKELIHOOD SPECTRAL FIT

ABOVE 100 MEV

Interval [MJD (DOY)] Best-fit Model and Parameters

All observations Log-parabola54682.680 (217.680) α = 1.94± 0.0554775.580 (310.580) β = 0.10 ± 0.02

FE>100 MeV = 6.90± 0.34× 10−7 [ph cm−2 s−1]Outburst/high state Power Law54683.955 (218.955) Γ = −2.06 ± 0.01754688.985 (223.985) FE>100 MeV = 29.1± 1.4× 10−7 [ph cm−2 s−1]Post-flare/lower state Log-parabola54689.063 (224.063) α = 1.87± 0.2054775.580 (310.580) β = 0.18 ± 0.08

FE>100 MeV = 5.32± 1.03× 10−7 [ph cm−2 s−1]

The LP description introduces the advantage of only one ex-tra parameter (while BPL model adds two parameters) withrespect to the simple power-law model, it allows modelingof mild spectral curvatures with no abrupt cutoffs, and pro-vides a better phenomenological physical interpretation.Onthe other hand there can be still caveats when usinggtlike

with a broken power-law model, in particular in determiningthe break energy when statistics do not allow a high numberof energy bins. It is also plausible that an energy spectrumaveraged over a long period of time, and containing differentactivity stages with time varying hardness, may exhibit an ap-parent curvature. Finally this does not exclude BPL model ifthe spectrum is extracted in different time intervals.

The average spectrum during the outburst state is consistentwith a simple power-law model,dN/dE ∝ EΓ. The outburststate shows a rather hard spectrum, suggesting a maximumpeak in the MeV energy bands (in theνFν representation),in agreement with the LP peaks found for the spectra citedabove. The extrapolated and averaged fluxes integrated above100 MeV, and the spectral fit parameters for all three periodsare shown in Table 1.

3. SIMULTANEOUS MULTIFREQUENCY OBSERVATIONS

Because of the reasonably uniform exposure and high sen-sitivity of the LAT, and the broad-band radio-to-gamma-rayemission of this kind of AGN, simultaneous multifrequencydata are very important to the investigation of the physicalproperties of supermassive black holes and relativistic jets,beyond the benefit of a firm source identification (Section2.2). With this in mind, several campaigns on a few selectedobjects, or ToO list of candidates for flaring sources, wereprepared pre-launch by theFermi collaboration (Tosti 2007;Thompson 2007). PKS 1502+106 was a previously unknownγ-ray source, with no pre-planned multifrequency campaign.But following the LAT outburst (reported in ATel #1650), aToO campaign was initiated on Aug. 07, 2008. This wasthe first Fermi multifrequency campaign that had not beenplanned pre-launch, and saw triggers for ToO pointings by IN-TEGRAL andSwift, long-term radio flux and structure moni-toring, as well as optical observations by ground based facili-ties.

The fast response ToO pointing by INTEGRAL provided200ks of observations during the period Aug. 09, 01:53 UT -Aug. 11, 15:12 UT, 2008 (revolution 711). However, PKS1502+106 was not detected (preliminarily) by the imagerIBIS on board INTEGRAL. Extrapolating the X-ray flux ob-served by Swift, the hard-X-ray flux had likely already fadedto slightly below the IBIS detection threshold in this epoch(more details will appear in Pian et al. 2009).

3.1. Simultaneous X-ray and UV-optical observations andresults by Swift

The Swift satellite (Gehrels et al. 2004) performed a ToOmonitoring campaign of PKS 1502+106 with daily snapshotsfrom Aug. 07 to Aug. 22, 2008. This quite long-term ob-serving campaign bySwift allowed extended daily snapshotsfor about 16 days, using the three instruments onboard: theX-ray telescope (XRT) for the 0.2-10 keV energy band, theUltraviolet/Optical Telescope (UVOT) for multiband photom-etry, and the Burst Alert Telescope (BAT) for the 15-150 keVhard X-ray band. BAT data were not used because of sourceconfusion problems with Mkn 841 which is about a factor of10 brighter than PKS 1502+106 in the hard X-ray band. The16 days of observations by Swift allow for cross-correlation

Page 7: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

PKS 1502+106: a new and distant gamma-ray blazar in outburstdiscovered by the Fermi LAT 7

FIG. 4.— Simultaneous gamma-ray and multifrequency light curves ob-tained during the multiwavelength campaign of August 2008 triggered by thehigh energy outburst discovered byFermi-LAT. The flux above 100 MeV,the X-ray flux (0.3-10keV) bySwift-XRT, the six-band fluxes monitoredby Swift-UVOT, the Kanata-TRISPEC differential relative magnitude lightcurves (optical∆V and near-IR∆J bands) and corresponding measures ofthe linear polarization degree, and the 15 GHz radio light curve from OVRO40-m are reported.

studies between theγ-ray, X-ray and UV-optical bands duringboth the active flaring stage and the fading post-flare stage ofPKS 1502+106.

The XRT was set in photon counting mode, and the datawere processed by the xrtpipeline with the use of standardsoftware (HEADAS software, v6.4) and standard filtering andscreening criteria. The XRT events in the 0.3–10 keV energyband were extracted using theXRTGRBLC FTOOL from cir-cular regions centered on the source position with variableradii depending on the source intensity and applying correc-tion for vignetting, Point-Spread Function corrections and hotpixels and columns with the use of exposure maps. The XRTX-ray flux light curve is shown in the second panel of Fig.4. The 0.3-10keV count rate of PKS 1502+106 measured byXRT was at a level 0.05 counts/sec (from our data), up from alevel of 0.02 counts/sec (archival past observations). The0.3-10 keV XRT 16-day long light curve obtained in August 2008(Fig.4) shows an initial count rate of 0.05 counts/sec, and agradual decay down to the level of about 0.02 counts/sec.

10−

30.

015×

10−

42×

10−

35×

10−

30.

02

Cou

nts

s−1

keV

−1

Outburst

Post−flare

10.5 2 5

11.

5

Rat

io

Energy [keV]

FIG. 5.— Swift-XRT combined 0.3-10 KeV spectra of PKS 1502+106 ex-tracted for the high state (MJD: 54685-54689) and the subsequent low state(MJD 54690-54701), mapping the X-ray behavior simultaneous to the LATflare and to the post-flare, relaxing activity and brightness.

TABLE 2ANALYSIS SUMMARY OF THE SIMULTANEOUS DATA OBTAINED BY THE

XRT INSTRUMENT ON BOARDSwift.

Obs. id. (date) Best-fit Model and Parameters

All observations Power Law(MJD 54685-54701) ΓX = 1.53+0.06

−0.07

texp:52910s χ2r=1.05/80

F0.3−10 keV = 1.79+0.08−0.11 × 10−12 erg cm−2 s−1

Outburst/high state Power Law(MJD 54685-54689) ΓX = 1.54± 0.08texp:27680s χ2

r=1.11/52F0.3−10 keV = 2.18+0.15

−0.12 × 10−12 erg cm−2 s−1

Post-flare/lower state Power Law(MJD 54690-54701) ΓX = 1.45+0.12

−0.11

texp:25230s χ2=0.76/32F0.3−10 keV = 1.39+0.14

−0.12 × 10−12 erg cm−2 s−1

The X-ray spectrum of each observation segment was fit-ted with an absorbed power law. Because of the low num-ber of events from the source, event were not grouped andC-statistics was used, fixing the column densityNH to theGalactic valueNHI = 2.19 × 1020 cm−2 in that direction(in agreement with values used, for example, in George et al.1994; Akiyama et al. 2003), and usingz = 1.839. The erroron the photon index and the flux (0.3-10keV) is large due tothe low statistics. The background photons were selected ina circular region close to the source. No significant photonindex variation was observed between the high and the lowstate, while the count rate and flux did vary.

The Swift Ultraviolet/Optical Telescope (UVOT) photom-etry was done using the publicly available UVOT FTOOLSdata reduction suite, and is in the UVOT photometric systemdescribed in Poole et al. (2008). The photometric data pointswere corrected for Galactic extinction using the dust maps ofSchlegel et al. (1998) and the Milky Way extinction curve ofPei (1992). These simultaneous multi-band optical and UVdata show an increase of about 2 magnitudes in all filters whencompared with the past-years archival values (i.e., from about19 to 17 in B band). The flux light curves in the six UVOTbands are shown in the third and fourth panel of Fig. 4. Thesefluxes appears to be well correlated. A slight rise in flux of 3days is observed in all the UVOT bands, followed by a fading

Page 8: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

8 Abdo et al.

similar to the flux decrease seen inγ-rays and X-ray bands. Ifthe time of the observed UV and optical maximum is relatedto the flare activity at higher energies, this would imply aninteresting time lag of about 4 days.

3.2. Simultaneous near-infrared and optical monitoring

PKS 1502+106 was also monitored in the optical V andnear-infrared J bands with some photometric and polarimetricsnapshots by the TRISPEC istrument attached to the 1.5-m“KANATA” telescope at the Higashi-Hiroshima Observatory,Japan (Watanabe et al. 2005; Uemura et al. 2008), within atwofold program of optical follow up for LAT flaring sourcesand regular monitoring of about 20 blazars. Imaging relativephotometry was performed using some comparison stars inthe same field, but due to the absence of an accurate calibra-tion for this field we prefer to report only the relative magni-tude difference∆mag with respect to the minimum level (fifthpanel Fig. 4). The optical and near-infrared band imagingphotometry is performed simultaneously in TRISPEC with aunit of polarimetric sequence (consisting of successive expo-sures at four position angles of the half-wave plate, where aset of linear polarization parameters, Q/I, U/I, are derived).

These flux observations, performed for a longer intervalwith respect to theFermi-Swiftcampaign (i.e. until Sept. 22,2008), show a high correlation between the V- and J-bandlight curves and show an optical decay phase comparable tothat observed in the UVOT photometric observations (fifthpanel of Fig.4). Remarkably, a strong correlation with theLAT gamma-ray light curve is found, including the first (Sept.04-07, 2008) of the possible minor flares occurring after theinitial large outburst. The observation of a flare in the opti-cal (V) and near-IR (J) bands, simultaneous with a secondγ-ray flare have the twofold advantage of providing a validationof such a minor LAT flare as a real feature displayed by thisblazar, and, even more crucially, in confirming the firm identi-fication of the new gamma-ray source seen byFermi with theblazar PKS 1502+106.

Comparing the Kanata-TRISPEC V-band and J-band col-ors, theV − J color index varies between 2.05 and 1.69(during the Sept.04-07 minor, rapid flare cited above). Onthe other hand, the degreeP of linear optical (inV, J bands)polarization observed (sixth panel of Fig.4) , remains ratherscattered by error dispersion irrespective the flux level, evenduring the minor flare mentioned (the maximum degreesrecorded during the monitoring wereP (V )max = 15 ± 3%andP (J)max = 13± 4%).

3.3. Simultaneous radio flux-structure data by single-dishand VLBI observations

As part of an ongoing blazar monitoring program, theOwens Valley Radio Observatory (OVRO) 40-m radio tele-scope has observed PKS 1502+106 at 15 GHz approximatelyevery two days since mid-2007. Flux densities for the peri-ods from July 26 to September 3, 2008 (MJD 54673–54711)and October 23 to December 9, 2008 (MJD 54762–54809)are shown in Figure 6. Flux densities were measured using az-imuth double switching as described in Readhead et al. (1989)after peaking up on-source. The relative flux density uncer-tainty for this source is dominated by a conservative 1.6%systematic error with a typical thermal error contributionof5 mJy. Absolute flux density is calibrated to about 5% usingthe Baars al. (1977) model for 3C 286. This absolute uncer-tainty is not included in the plotted errors.

The measured flux densities in the MJD 54673–54711(DoY 208-246) time period fit a 1.69 Jy constant-flux modelwith χ2/(N − 1) = 0.70 (N = 15 data points). This in-dicates no statistically significant variability in this time pe-riod. The beginning of a bright radio flare is apparent in theMJD 54762–54809 (DoY 297-344) time period with an in-crease of at least 30% over the earlier mean flux density.

A less intensive monitoring at 37 GHz was carried out withthe 13.7m radio telescope at Metsahovi Radio Observatory,Helsinki University of Technology, Finland. The flux den-sity scale is set by observations of DR 21, and sources 3C 84and 3C 274 are used as secondary calibrators. A detailed de-scription on the data reduction and analysis can be found inTerasranta et al. (1998).

The PKS 1502+106 flare was also followed-up by the Ef-felsberg 100-m radio telescope with four multi-frequency ra-dio spectra obtained on August 23, September 16, October 18,and December 6, 2008 (within the F-GAMMA project, seeFuhrmann et al. 2007). Each radio spectrum consists of si-multaneous measurements at various frequencies between 2.6and 42 GHz. The observations were performed using cross-scans in azimuth/elevation with the number of sub-scansmatching the source’s brightness at the given frequencies.Thedata reduction was done using standard procedures describedin Fuhrmann et al. (2008); Angelakis et al. (2008).

Other radio observations are available at six frequencies be-tween 1 and 21.7 GHz thanks to the 600-m ring transit radiotelescope of the Russian Academy of Sciences RATAN-600(Korolkov & Parijskij 1979), which observed the source onSeptember 10, 26, and October 2, 2008. A weighted averageof these three observations, is presented in Fig 7. PreviousRATAN-600 data which cover the period between 1997 andMarch 2008 are also shown for comparison. The observingmethods, the data processing, and the amplitude calibration isdescribed in Kovalev et al. (1999).

All single-dish spectra obtained with the Effelsberg 100-mand RATAN-600 radio telescopes are presented in Figure 7.Here, no indication of a flare or strong difference/variabilitybetween August and September can be noted. However, theDecember 2008 spectrum shows the beginning of a bright ra-dio flare with a clear spectral steepening towards higher fre-quencies (ν > 10 GHz). This is in good agreement with thestrong flux density increase seen in the 15 GHz light curveduring November/December (Fig.7).

Detailed radio images at sub-milliarcsecond scale of the

FIG. 6.— Long term radio flux light curve at 15 GHz obtained by the OwensValley Radio Observatory (OVRO) 40m dish radio telescope (filled circles),showing the rising part of a radio outburst started in late November 2008,i.e. almost 4 months after the gamma-ray outburst detected by Fermi. Thefill diamonds represent the flux measurements performed by the Metsahovi14-m radio telescope at 37 GHz (right y-axis scale), confirming the start of aradio outburst at a higher frequency. The scaled LAT daily light curve on thesame period is reported for comparison (small light grey bars).

Page 9: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

PKS 1502+106: a new and distant gamma-ray blazar in outburstdiscovered by the Fermi LAT 9

FIG. 7.— Variable broad-band radio spectra observed with the Effelsberg100-m and RATAN-600 radio telescopes simultaneous to the LAT data. His-torical RATAN-600 data (grey open triangles) and archival data from the lit-erature until March 2008 (grey open circles) are shown in thebackground forcomparison.

PKS 1502+106 superluminal jet were obtained during threeepochs in 2008 withFermi already in orbit: on June 25, Au-gust 06 (during the maximum peak of theγ-ray outburst), andNovember 19. These observations were performed as part ofthe MOJAVE monitoring program conducted with the VeryLong Baseline Array (VLBA) atλ = 2 cm (Lister & Homan2005; Lister et al. 2009a) and provided useful high resolutiontotal intensity and linear polarization images. These VLBAimages close to theγ-ray flare are reported and compared tothe map obtained one year earlier in Figure 8. The highestintegrated flux density value since the beginning of the 2 cmVLBA monitoring in 1997 (Kellermann et al. 2004) was mea-sured on November 19, 2008 (DOY 324) asF15 GHz = 2.0 Jy,with a peak intensity of 1.6 Jy beam−1. These values are sig-nificantly higher than the typical level of 1.3 Jy reported inthe program (Kovalev et al. 2005; Lister et al. 2009a,b) andindicate a radio flare happening in the source VLBI core.The core flux density and brightness temperature raised tohigher values as well which means that the flare happens inthe VLBI core, as expected. These finding are in good agree-ment with the single-dish results presented above (Figure 6and Figure 7). The second relevant feature is the directionof the electric vector position angle (EVPA) in the core re-gion, which rotated between the 2007 and 2008 epochs by90 degrees most probably indicating an opacity change — aprecursor of an outburst in the VLBI core.

In summary, our single-dish and VLBI radio monitoringof PKS 1502+106 simultaneous to theFermi LAT observa-tions has revealed (i) no significant radio 15 GHz variabil-ity during the strong LATγ-ray flare seen in August 2008,and (ii) a strong radio flare which becomes clearly visible atν > 10 GHz during October/November 2008 (Figure 6) witha rise phase lasting for at least 20 days. If this flaring behav-ior is associated with the brightγ-ray flare of August 2008,a delay of more than three months (∼ 98 days if the lagbetween the starting days of theγ-ray outburst, DoY 218.2,and the 15 GHz outburst, DoY 316.7, is considered) could beexplained by opacity effects in the core region of the source(e.g., Aller et al. 1999). Although we cannot exclude the pos-sibility of radio activity at 15 GHz during the OVRO 40-m ob-servations outage (September 3 - Otober 23 2008) that mightalso be associated with the gamma-ray flare of August 2008,

FIG. 8.— Total intensity and linear polarization images observed by theVLBA at 15 GHz as part of the largeFermi-supporting MOJAVE program.Naturally-weighted total intensity images are shown by black contours. Thecontours are in successive powers of two times the base contour level of1.0 mJy beam−1. Electric polarization vector directions are indicated ontheleft hand side by blue sticks, with their length being proportional to the polar-ized intensity. Linear fractional polarization is shown onthe right hand sideoverlaid according to the color wedge.

OVRO data before and after the outage as well as Septemberand October Effelsberg 100-m data are consistent with verylittle change in the 15 GHz flux over this time period. Thethree flux measures at 37 GHz obtained at the Metsahovi ra-dio observatory during the OVRO outage, confirms this trendwith very little variability during this period. However, theobserved radio flare could also be associated with, e.g., themore recent, prominent variability seen in the LATγ-ray dataduring November/December (Figure 2, DoY∼ 320–333). Amore detailed analysis of such possible correlations and thesource’s overall radio/γ-ray behavior seen with LAT and si-multaneous radio observations over a longer period of timewill be the subject of a subsequent work.

4. ARCHIVAL MULTIWAVEBAND DATA

Page 10: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

10 Abdo et al.

A full multiwavelength analysis dedicated on PKS1502+106 is available only from George et al. (1994), whereold archival and broadband radio to X-ray data obtained in1993-94 were presented. Data and analysis on PKS 1502+106reported in other papers are mostly limited to the radio regime.In order to compare our multifrequency findings with the past,and to form a more complete characterization of this blazar,we briefly present here, for the first time, results from un-published past observations by INTEGRAL, XMM-Newton,Swift, and Spitzer space telescopes performed in 2001, 2005and 2006.

4.1. INTEGRAL observations in 2006

The sky region containing PKS 1502+106 was observed in2006 by IBIS (83ksec, MJDs 53760.4 to 53762.4, Jan.25-27,2006), and a new softγ-ray source (IGR J15039+1022) wasdetected with a flux density of 1.6 mCrab in the 18-60 keVenergy range (corresponding to1.2 × 10−11 erg cm−2 s−1,see ATel #1652). This IBIS source was identified with Mkn841, a Seyfert galaxy known to display a well detected highenergy cut-off around 100 keV (Petrucci et al. 2002), makingit unlikely to emit in theγ-ray domain. The angular distanceof PKS 1502+106 from IGR J15039+1022,∼ 11′, points toa clear non-detection during this Jan. 2006 INTEGRAL ob-servation, while a 2σ upper limit for PKS 1502+106 of 0.7mCrab in the range 18-60 keV (0.52× 10−11 erg cm−2 s−1)is inferred.

4.2. XMM-Newton and Swift

Four serendipitous, unpublished XMM-Newton X-ray ob-servations of PKS 1502+106 by the EPIC (MOS detectoronly) camera are available as the source was in the frame ofthe target Seyfert galaxy Mkn 841. PKS 1502+106 was al-ways on the border of the MOS chips and out of the PN frame,and therefore subject to low X-ray statistics, regardless of itsintrinsic brightness. The four X-ray EPIC-MOS observations(three in 2001 and one performed in Jul.17, 2005, see Table 3for an analysis summary) do not show variations in the 0.2-10keV photon index, while the 0.2-10 keV flux intensity var-ied by a factor of a few (in the range3.5 − 6.8 × 10−13 ergcm−2 s−1in 2001, comparable to the lower states observed byASCA, and a mildly active state with1.1 × 10−12 erg cm−2

s−1in July 2005, Table 3).PKS 1502+106 was also observed twice in the past with

theSwift-XRT as a fill-in target (TargetID: 36388), showing afainter X-ray flux (0.02 counts/sec) than the flux recorded inthe August 2008 campaign observations (0.05 counts/sec).

4.3. Spitzer observations and the multifrequency behavior onJul-Aug.2005

In the past, only upper limits in the far-/near-infraredbands by IRAS were available for PKS 1502+106(Neugebauer et al. 1986, and Figures 9 and 11). PKS1502+106 was observed serendipitously in the mid-infraredband for the first and only time by Spitzer on August 13, 2005,09:10-09:18 UT (PID 117, AOR Request Key 5011456). TheInfrared Spectrograph (IRS, Houck et al. 2004), low resolu-tion (R=60-130) module, recorded the mid-IR spectrum from5-14 µm (shown in Fig.9 with the optical SDSS spectrumand near-IR photometric data point). High resolution IRSmodule spectra were also taken, but could not be used sincethere were no accompanying background observations. TheShort-Low (SL) coadded 2D spectra were reduced using

FIG. 9.— The unique observed mid-IR spectrum in the range 5-14µmobtained by the Spitzer Infrared Spectrograph (IRS) low resolution (R =60 − 130) module, in August 13, 2005. The position of redshifted Brackettemission lines and PAH line are indicated, even if they are not detected inthe IRS spectrum that is consistent with a simple power law model. In addi-tion the optical spectrum by the SDSS on April 23, 2006, and high precisionJHKS photometric flux measurements of the Kitt Peak National Observatory(KPNO) 2.1m telescope of March 29, 2002 are also reported.

the standard Spitzer IRS pipeline (ver. S17.2). Backgroundwas subtracted using the two nod positions along the slit.The spectra were extracted and flux-calibrated with SPICEver. 2.1.2, in a standard, expanding point-source aperture.The two spectral orders match well at 7µm, indicating awell-pointed observation. The mid-IR continuum of PKS1502+106 rises to the near infrared and appears to be ratherfeatureless, consistent with pure synchrotron emission (apower lawFν ∝ ν−0.9 in the 5-14µm range, Fig.9). Thewave-like deviation of the data can be simply explained aswavelength-dependent spectrograph slit losses, while thered-shifted Brackett emission line series (like the 3.3µmPAH feature) falling in this wavelength range are indicated,but they are not well detected in the spectrum. This IRSspectrum is similar to other blazars which have been observedby Spitzer, including BL Lac and 3C 454.3 (Leipski et al.2008; Ogle et al. et al. 2009), while the near IR (J,H,K) fluxdata from the Kitt Peak National Observatory (KPNO) 2.1mtelescope Watanabe et al. (2004) reported in the same figureindicate a lower flux state and steeper H-KS (1.65-2.15µm),

TABLE 3ANALYSIS SUMMARY OF THE EPIC-MOSINSTRUMENT OBSERVATIONS

(JAN . 2001AND JUL . 2005)ON BOARD OFXMM-N EWTON.

Obs. id. (date) Best-fit Model and Parameters

ObsID 0070740101 Power Law(Jan 13, 2001, 09:20 UTC) ΓX= 1.6± 0.2

χ2r=1.69/11

F0.2−10 keV = 3.5× 10−13 erg cm−2 s−1

ObsID 0070740301 Power Law(Jan 14, 2001, 00:30 UTC) ΓX= 1.7± 0.2

χ2r=1.27/17

F0.2−10 keV = 6.8× 10−13 erg cm−2 s−1

ObsID 0112910201 Power Law(Jan 13, 2001, 04:58 UTC) ΓX= 1.6± 0.2

χ2r=1.05/8

F0.2−10 keV = 3.6× 10−13 erg cm−2 s−1

ObsID 0205340401 Power Law(Jul 17, 2005, 06:32 UTC) ΓX= 1.69± 0.08

χ2r=0.99/58

F0.2−10 keV = 11.0× 10−13 erg cm−2 s−1

Page 11: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

PKS 1502+106: a new and distant gamma-ray blazar in outburstdiscovered by the Fermi LAT 11

spectral index(

α(H−KS) = 1.66)

.The July-August 2005 SED assembled with these XMM-

Newton (July 17, 2005) and Spitzer (August 13, 2005) ob-servations (joined with a couple of radio-optical observationsin these two months by the Metsahovi, RATAN and Catalinaobservatories) is consistent with a low or mildly active stagethat can be explained by a simple SSC model (inset plot inFig.11). Based on these data there are therefore no hints fora deviation from a SSC scenario in this flat spectrum radioquasar. In any case, this SED cannot provide significant con-straints, because it contains observations obtained over abouttwo months and does not include anyγ-ray data, aside fromthe older (1992) EGRET upper limit (calculated with the pre-scriptions in Thompson et al. 1996).

XMM-Newton observations pointed out an (0.2-10 keV)photon index that shown almost no variations, with val-ues (ΓX ∼ 1.7) in agreement with the previous X-ray ob-servations performed by ROSAT and ASCA (George et al.1994; Watanabe et al. 2004). The integrated X-ray fluxF0.2−10 keV = 1.0×10−12 suggests a mildly active state dur-ing this observation, comparable to theSwift XRT spectrumobserved in the days soon after the LAT outburst (red filledsymbols, data points on Fig.11). More complicated and pos-sibly more accurate emission models, beyond the SSC, canbe therefore investigated for the first time only thanks to theFermimultifrequency campaign of August 2008, whose SEDis described in the next section. TheSwift XRT simultane-ous spectrum of Aug.2008 has a slightly harder spectrum withrespect to these archival XMM-Newton observations whilethe XRT flux was about one order of magnitude greater thanthe flux observed in the past by XMM-Newton, ASCA andROSAT.

5. DISCUSSION

5.1. Gamma-ray outburst and longer-term variability

During the first several months of the LAT survey, PKS1502+106 was one of the brightest, as well as the most lu-minous, blazar in the MeV–GeV band. The threefold flux in-crease in. 12 hours between Aug. 05 and Aug. 06, 2008(DoY 218-219, Fig. 2) constrains the rest-frame size (R′) ofthe flaring region:R′ ≤ c∆tD/(1 + z) ≃ 6.8 × 1015 cm(whereD = 1/(Γ(1− β cos θ)) is the macroscopic Dopplerfactor,Γ the bulk Lorentz factor andθ ≃ 1/Γ the angle ofsight). The valueD ≃ Γ ≃ βapp = 14.8 is assumed fromMOJAVE VLBA measurements (Lister et al. 2009b), whereβapp is the kinematic apparent jet speed in units of c. In thishigh speed regime an upper limit on the viewing angle can bealso estimated:θ < 2 arctan(1/βapp) = 7.7. This is con-sistent with an independent estimation of the Doppler factorbased on the 37 GHz flux variability, made in Hovatta et al.(2009): a factorDvar = 12 andβapp var = 14.6, with bright-ness temperatureTbr = 8.7 × 1013 K, and angle of sightθ = 4.7.

Assuming the “concordance” cosmology (Sect. 1 andz =1.839), the luminosity distance of PKS 1502+106 isdL =14.2 Gpc, and the inferred, apparent and isotropic, monochro-matic luminosity atE0 = 100 MeV of PKS 1502+106 duringthe outburst phase isLE>100 ≃ 4πd2L · (Γ− 1)E0FE>E0

≃1.1× 1049 erg s−1,, where the average fluxFE>E0

= 2.91×10−6 ph cm−2 s−1 in the outburst interval (DoY 2008: 218.95- 224.0) is used. The bolometric luminosity is expected to beeven higher than this value, since the measured LAT spectrumappears to be beyond the peak of the high energy component,

and therefore this LAT blazar has probably one of the highestLγ/∆t ratios (2.5 × 1043 erg s−2) known in the MeV-GeVregime.

Relativistic motion provides a solution for the problem ofintrinsic excess absorption by pair-production in powerfulγ-ray sources like PKS 1502+106 (see, e.g. Mattox et al.1993; Madejski et al. 1996) which have a significantLBLR.Adopting the flux tripling time scale of the outburst rise (i.e.∆t = 12 hours), and the outburst state averaged X-ray flux(F0.3−10 keV = 2.18+0.15

−0.12 × 10−12, Table 2) at the observedphoton frequencyνX = 1018 Hz, the minimum DopplerfactorD required for the photon-photon annihilation opticaldepth to beτγγ ≤ 1 can be estimated. Using the derivedrelation1 = τγγ ≃ σTd

2LFX/

(

3∆t c2EXD4)

and takingthe region sizeR = c∆tD/(1 + z), the source-frame photonenergyE′

X = (1 + z)hνX/D and the intrinsic X-ray lumi-nosityL′

X = 4πd2LD−4FX we obtainD & 7.7 (omitting therequirement of co-spatiality of the X-ray andγ-ray emissionregions relaxes this limit). This is in agreement with the val-ues found from radio flux-structure observations and with theSED modeling parameters found.

The asymmetry of the August 5–6γ-ray outburst can sug-gest a more complex emission geometry than a simple one-zone model. The temporal structure —∼ 0.5 day rise, fol-lowed by a∼ 4.5 day decay where a∼ 2.5 day intermediate-level plateau is likely observed — implies particle accelera-tion and cooling times that are greater than the light cross-ing time, i.e.,tinj, tcool > R/c, where these quantities areevaluated in the jet comoving frame. A synchrotron self-Compton (SSC) emitting blob in the jet should be relativelyconfined (< 0.01 pc), although relativistic beaming wouldpermit the region to be as much as an order of magnitudelarger. The hinted intermediate plateau could mark a twofoldactive region, and two SSC emitting components. Descrip-tions making use of a multi-zone SSC or multi-emissioncomponent SSC (second order, superquadratic components)are reported, for example, in Sokolov & Marscher (2005);Georganopoulos et al. (2006); Graff et al. (2008). On theother hand, if the injection of relativistic electrons is impulsiveand repeated, some single-zone SSC models already predictplateaus during an outburst (see e.g., Chiaberge & Ghisellini1999; Bottcher & Chiang 2002; Sokolov et al. 2004). Atlower frequencies (IR-optical-UV), where cooling times arelonger, the electron distributions corresponding to differentinjections can build up, and the memory of the individual in-jection phases can be lost, providing a smoother decay with-out intermediate plateaus (as shown in X-ray, optical-UV lightcurves bySwift, Fig. 4). Apparent delays (like the 4-day laghinted by UVOT data) can also be explained within this sce-nario.

On the other hand this asymmetric (fast rise, slower decay)shape of theγ-ray outburst can be also an evidence for a dom-inant contribution by Comptonization of photons producedoutside the jet (Sect. 5.3 and 5.4) during this event, as pre-dicted for example in Sikora et al. (2001). Gamma-ray flaresproduced by short-lasting energetic electron injections and atlarger jet opening angles are predicted to be more asymmet-ric showing much faster increase than decay, the latter deter-mined by the light travel time effects.

The “harder when brighter” trend of the gamma-ray pho-ton index during the outburst (right inset panel of Fig. 2),hints a narrow hysteresis evolution of the spectral indexagainst the flux, a signature produced by non-thermal cool-

Page 12: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

12 Abdo et al.

ing and high to low energy propagation of the electron in-jection rate (Kirk et al. 1998; Georganopoulos & Marscher1998; Bottcher & Chiang 2002). The photon index extractedwith a power-law model over the daily bins, was found quitescattered irrespective of the flux level in the remain part ofthelight curve following the outburst.

The outburst of Aug. 2008 appeared to have ignited an en-during relaxing state ofγ-ray brightness and activity, dur-ing the following four months (Sections 2.3 and 5.1). The1/fa (with a ∼ 1.3) PSD points out a general fluctua-tion mode placed between a pink-noise (flickering) and arandom-walk, (Brownian motion or brown noise), staggeredby some rapid flares, similar to the long-term variability ofblazars observed in radio and optical wavebands (for example,Hughes et al. 1992; Ciaramella et al. 2004; Terasranta et al.2005; Ciprini et al. 2007b; Hovatta et al. 2007). In con-trast, this variability mode is rather different than the fullBrownian regime shown by the short-term (intra-hour reso-lution) light curve of the large TeV outburst of PKS 2155-304(Aharonian et al. 2007). This PSD indicates the occurrencefrequency of a specific variation is inversely proportionaltoits strength, as found in processes driven by stochastic relax-ation, and rapid flares/outbursts (phenomenologically calledintermittence), are not occasional events produced by physicalprocesses of different nature with respect to the mechanismresponsible for the long-term flickering, but can be consideredas statistical tails of the same dynamic process, possibly con-nected to disk or jet instabilities, to viscosity and magneticturbulence, or to inhomogeneities and shocks (for exampleBegelman et al. 1984). Even in the case of quite nonhomoge-neous structures, the jet is seen under a small viewing angle(θ < 7.7 here), therefore flaring events and variability trendsare the result of emission components originating from differ-ent regions, excluding, in most cases, causality.

5.2. Gamma-ray and radio connection

The absence of significant radio 15 GHz and 37 GHz vari-ability before, during and immediately after the LAT outburst(Fig. 6), and strictly correlated to the near-IR toγ-ray out-burst, has consequences for opacity at lower radio frequen-cies. However, the beginning of the strong radio flare seen inthe Oct.-Dec. 2008 data of Fig. 6 and Fig. 7, could be as-sociated with the period of increasedγ-ray activity seen after∼DoY 330 (Nov. 25, 2008, Fig. 2). If the beginning of the ra-dio outburst in Fig. 6 is associated with the LAT outburst, thiswould imply a delay of more than about 98 days (Sect. 3.3),meaning that the radio emission region becomes transparentat the cm and longer wavelengths significantly after the out-burst seen at wavelengths where the source is truly opticallythin (mm, IR, optical bands).

The rotation direction of the electric vector position angle(EVPA) in the milliarcsecond-scale core of PKS 1502+106between 2007 and 2008 was a possible precursor signatureof an outburst that occurred in the core (Fig.8). The elec-tric polarization vectors appear well aligned to jet axis intheJune 25, 2008 map and even more in the Aug. 06, 2008 map.This latter map represents an unprecedented example of radio-structure snapshot truly simultaneous to the peak of a MeV-GeV outburst (Fig.8). About 3 months after the outburst, onNov. 11, 2008, the alignment is again decreasing. These in-teresting findings are in agreement with the scenario that as-sumes very brightγ-ray flares occurring after the ejections ofsuperluminal radio knots, with accompanying increases in po-larized radio flux, and a field ordering and alignment with re-

spect to the jet axis. A correspondence between variations inpolarization direction and intensity in different bands atpar-sec scales can help localize the primary site of the high energyemission. We found that the prominentγ-ray outburst andthe possible 3-month delayed radio-outburst are likely pro-duced in the core of the parsec-scale jet of PKS 1502+106.More details on connections between VLBI radio structuresand high energy emission, and their interpretation are reportedrecently in Jorstad et al. (2007); d’Arcangelo et al. (2007);Marscher et al. (2008); Kovalev et al. (2009). To test if thisis indeed the case for the discussed gamma-ray flare, VLBAmonitoring of 1502+106 is continued at 15 and 43 GHz.

5.3. Gamma-ray, X-ray and UV-optical cross correlations

During the EGRET era a similar degree of simultaneousX-ray andγ-ray monitoring was achieved, for instance, forthe blazar 3C 279 (Wehrle et al. 1998; Hartman et al. 2001),a typicalγ-ray powerful FSRQ resembling PKS 1502+106.The correlation found in PKS 1502+106 between theγ-rayflux and the X-ray and UV-optical emission during the 16-days ofSwift follow up, is evidenced by the multi-panel andmultifrequency light curves in Fig. 4. The apparent linear cor-relation between the emission in the LAT band and the X-rays (Fig. 10, left panel) suggests that the observed X-ray andMeV-GeV photons may be part of a single SSC component(continuous line, labeled “SSC2”, on Fig.11). This could alsoexplain the nearly constant X-ray (0.3-10 keV) spectral in-dex observed during the outburst with respect to the post-flarephase.

The γ-ray peak power observed during the outburst wasνFν ∼ 1049 erg s−1 , and decreased to less than about2×1048

erg s−1 in the post-flare period, meaning a difference of atleast a factor of∼ 6 (Fig. 11). The difference between thesetwo emission states in X-ray output was instead a factor of∼ 2.5, and a factor of∼ 3.5 in the optical-UV band (frequen-cies above the synchrotron peak). These sub-quadratic dif-ferences between the outburst and the subsequent state couldbe still described by SSC descriptions. The strict correlationbetween the X-ray flare and theγ-ray outburst supports, atleast, the dominance in the SED (Fig.11) of the SSC, in-jet,emission from radio to X-ray bands in agreement with resultsdetailed in Sikora et al. (2001). A simple single-zone SSCmodel has nevertheless problems explaining and reproducingthe largeγ-ray dominance observed during the outburst, andcould require very sub-equipartition magnetic fields.

Fig. 4 and the right panel of Fig. 10 show the flux mon-itored by UVOT (U-band reported there, but similar resultsare found in the other 5 filters), indicating the same strongcorrelation with respect toγ-rays, but with a possible timelag of about 4 days (also hinted at by the peak on the dis-crete cross-correlation function, DCCF, inset plot). Thispos-sible time lag can be reasonable only by assuming that theoptical-UV brightness during the start and rise of the gamma-ray outburst (i.e. during DoY 216-219) was comparable orlower than the flux observed during the first UVOT observa-tion (performed on DoY 220.8). Relative delays between thesynchrotron (ourSwiftUVOT data) and the inverse Compton(our Swift XRT andFermi LAT data) counterparts, are dom-inated by energy stratification and geometry of the emittingregion. In SSC-dominated zones the synchrotron emissionis co-located with the inverse-Compton production site, andsuch 4-day optical-UV toγ-ray lag would depend on light-travel time effects in the the emitting region (characterizedby sizeR and viewing angleθ), and by the particle cooling

Page 13: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

PKS 1502+106: a new and distant gamma-ray blazar in outburstdiscovered by the Fermi LAT 13

FIG. 10.—Left panel:the gamma-ray flux measured byFermiLAT versusthe X-ray flux measured bySwift-XRT. The cross-correlation, without lagsis well displayed.Right panel:the LAT gamma-ray flux versus the UV fluxmeasured bySwift-UVOT in the representative U-band, and the discrete crosscorrelation function diagram (DCCF, inset plot) between the gamma-ray fluxand the U-band flux. Here a clear correlation with a 4 day delayof the UVemission flare with respect to the gamma-ray flare is suggested.

time and decay time of the synchrotron and inverse Comp-ton (IC) emissions. When the region is not so compact, bothdecay times can be comparable to the apparent light cross-ing timeR/c, and significant shifts between light curves atthese different energy bands are expected. Synchrotron flares(our optical-UV data) can be delayed with respect to the ICγ-ray flares (e.g. Chiaberge & Ghisellini 1999; Sokolov et al.2004). Rather long time lags such as this could also be ex-plained by a prolonged disturbance traveling down the jet.The disturbance, being radially inhomogeneous in both den-sity and velocity, could induce shocks and collisions leadingto the formation of two adjacent emission zones with similarproperties (multi-zone SSC), thus explaining the flare shapeasymmetry and the intermediate-level plateau during the de-cay (Section 5.1).

Unlike SSC emission, outbursts dominated by inverseCompton external radiation involve a constant field of seedphotons, and are not delayed by light-travel time of pho-tons. The resulting frequency stratification behind the shockfront could extinguish an outburst first at the highest ener-gies, then progressing to lower frequencies as time advances(Sokolov & Marscher 2005), and energy-dependent radiativelosses induce delays in the declining part of the emission pro-duced by the lower energy particles. The variability patternsshown by our data could rule out radiative cooling as the onlymechanism for causing the delay, since the rising part of theoptical emission follows the rising part of theγ-rays. In ad-dition, the decay time scale in the UVOT data seems com-parable to that seen in the LAT, and not longer as requiredby energy-dependent dominated cooling. A reasonable mix-ture of SSC and an extra-contribution by Comptonization ofthe BLR photon field (external jet origin) by the same popula-tion of energetic electrons (ERC, e.g. Sikora et al. 1994, 2001;Dermer & Schlickeiser 2002; Ghisellini & Tavecchio 2009a)could be more adequate to explain also the SED, and suchSSC+ERC scenario is still compatible with the X-ray toγ-raycorrelation found in PKS 1502+106.

As seen in Section 3.2, the first of the fast, minorγ-rayflares occurring in Sept. 2008 (DoY 248-251) after the bigoutburst, was well observed by the Kanata-TRISPEC tele-scope in bothJ andV bands, with no significant time lag.The lower intensity and duration of the flare, the limited tem-poral resolution of the data, and a possible dominance of theSSC process during this episode can explain this difference

with respect to August’s big outburst. The match between theMeV-GeV and optical-near-IR flare, was also crucial for thefirm identification with PKS 1502+106 (Sect. 2.2), and was areasonable confirmation of the source-intrinsic nature of thisvariation seen in the LAT light curve (Fig. 2 and Fig. 4). TheseV andJ flux measurements, obtained after the conclusion ofthe outburst phase but on a period longer than theSwiftmoni-toring (Fig. 4), appear well correlated, though theV −J colorwas less pronounced than the multi-band near-IR colors re-ported in Watanabe et al. (2004) and Fig. 9, while the degreeof optical polarization remained almost constant.

5.4. Spectral energy distribution

Variability is a powerful diagnostic to investigate blazarphysics, but represents also a supplementary problem for theanalysis of broad band spectral energy distributions (SED),where model constraints are provided by simultaneous andwell time-resolved multifrequency data. In the previous sec-tion we indicated that dominant synchrotron and SSC (in-jet) mechanisms can explain the radio-to-X-ray emission andcorrelations, whereas the origin of the high-power MeV-GeV bolometric emission could be better constrained by ex-ternal Comptonization of the radiation from the broad lineregion (BLR), as invoked by (Ghisellini et al. 2009b), andfound already in similar FSRQs of the EGRET era (e.g.,Sokolov & Marscher 2005; Sikora et al. 2008, 2009). Thefeatureless mid-IR continuum observed by Spitzer-IRS in2005 (Fig. 9), supports the hypothesis of a prominent syn-chrotron emission by the jet, controlling the lower energycomponent. It could also reflect the lack of detectable am-bient dust radiation, thereby supporting the idea that the ERCdissipation occurred within the BLR for this LAT outburst, inagreement with prescriptions of (Sikora et al. 2002).

PKS 1502+106 might be considered a blazar peaked at theborder of the MeV and GeV bands (a peak around 0.4-0.5GeV is suggested by the curved model fit of the 4-month spec-trum, Section 2.4). In other words, PKS 1502+106 is likelyat the border of the family of BLR-dissipated FSRQs andcircum-nuclear ambient/torus dust-dissipated FSRQs, with animportant SSC power output from radio-to-X-ray bands (Sec-tion 5.3), as depicted by the multiband correlations, the ab-sence of hints for a bulk Compton feature produced by cold,adiabatically cooling electrons (as observed in PKS 1510-089,Kataoka et al. 2008; Abdo et al. 2009f), the lack of evidencefor a blue bump.

The medium or high black hole mass of this blazar(likely in the range0.5 − 1 × 109 M⊙, as calculated byD’Elia et al. 2003; Liu et al. 2006, respectively) could con-note a BLR radiation field and inverse Compton dissipationmoderately stronger than the magnetic energy density and theSSC luminosity in the gamma-rays, as reported in Fig. 11.The estimated accretion rate is 2 M⊙ yr−1 (D’Elia et al.2003). In this case we haveLERC/Lsyn = U ′

BLR/UB ≃LBLRD2/(4πcR2

BLRUB) whereU ′BLR andUB are the elec-

tromagnetic energy density in the BLR and the magneticfield energy density in the jet blob, respectively. With a109

M⊙ mass the Schwarzschild radius of the supermassive blackhole is RS = 2GM/c2 ≃ 3 × 1014 cm. The bolomet-ric luminosity of the BLR in PKS 1502+106, evinced by theMgII emission line profile, isLBLR = 3.7 × 1045 erg s−1

(Liu et al. 2006), implying that the BLR is located at a radiusRBLR = 1017

L/1045 ≃ 2× 1017 cm. We can expect rela-tively smaller magnetic fields in the dissipation region andan

Page 14: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

14 Abdo et al.

FIG. 11.—Major panel: Overall radio-to-gamma-ray spectral energy distribution(SED) of PKS 1502+106 assembled with data from the 2-weekFermi LATmultifrequency campaign of August 2008. The two representative, time-averaged, states for this campaign (time intervals outlined in Fig.2), i.e. the high state(Aug. 05-10, 2008, DoY 2008: 218.95 - 224.0; blue filled circle symbol), and the post-flare (intermediate brightness) state (Aug. 11-22, 2008; DoY 2008: 224.0 -235.42; red filled square symbol), are represented along with their SSC and ERC model attempts. Archival non-simultaneous data (including the Jul.-Aug. 2005data, and the whole 4-month cumulated LAT spectrum, open orange circle symbols and strip) are reported in the backgroundfor comparison. The high Comptondominance and gamma-ray bolometric luminosity reached during the outburst is evident. The SED of the outburst state is reported with a superposed one-zone“SSC+ERC” model fit attempt (where the SSC dominates the radio-to-X-ray SED and the ERC produces the gamma-ray component, blue dashed line), whilethe SED of the post-flare state is reported with two possible models superposed: the same SSC+ERC modeling (SSC radio-to-X-ray and ERC for gamma-rayband, red dotted line) and also a pure one-zone SCC attempt (labeled with “SSC2”, a second type of stand-alone SSC model for the entire radio-to-gamma-raySED, continuous red line).Inset panel:The July-August 2005 non-simultaneous SED (filled orange circles) assembled with the XMM-Newton EPIC-MOS datafrom Jul. 17, 2005, the Spitzer IRS observation from Aug. 13,2005, radio flux data from Metsahovi and RATAN radio observatories, and optical data from theCatalina Sky Survey. Archival non-simultaneous data (opencircles) are reported in background. This two-month averaged SED is consistent with a low or mildlyactive stage (suggested by the X-ray flux above1× 10−12 erg cm−2 s−1) and can still be explained by a pure one-zone SSC model.

ERC, external-jet process that is the dominant cooling mech-anism when the source is in a highγ-ray state, as suggested inSection 5.1. In fact a SSC-only model would require a mag-netic field in very sub-equipartition conditions, even if onlythe BLR fraction lying within the varying region’s beamingcone could contribute to gamma-rays. The simultaneous SEDcorresponding to the outburst state reported in Fig. 11 cor-roborates the possible needs for an ERC contribution, show-ing a large (Lγ/Lopt ∼ 100) gamma-ray dominance over thesynchrotron component. The SSC+ERC model can be con-sidered in some way in agreement with the redshift value ofthis blazar, if we assume the FSRQs→ BL Lacs cosmolog-ical evolutionary scenario with dimming jet power, possiblyrelated to star formation rate or the the far-IR/submillimeterluminosity density (Dermer 2007).

In Fig. 11 we report the data of the two SED states (the“outburst state” of Aug. 05-10, 2008 (blue points) and the“post-flare” state (red points), characterized by a intermediateluminosity, of Aug. 11-22, 2008 (both intervals are outlinedin Fig. 2), in conjunction with the unpublished archival andliterature data collected for comparison. Above the two si-multaneous SEDs of Aug. 2008, SSC and ERC “strawman”models are reported for each state (SSC+ERC modeling, bluelines for the outburst period, and SSC+ERC plus a SSC-stand-alone modeling, red lines, both for the post-flare period).

The pure SSC stand-alone model for the post-flare state (in-dicated in Fig. 11 as a version 2 of the first order SSC model-ing with the label “SSC2”, continuous red line, spanning fromradio to gamma-ray bands) implements the temporal evolu-tion of the synchrotron and SSC spectral components in a sin-gle flaring blob within the jet, where a population of acceler-ated electrons having a power-law with exponential cutoff dis-tribution (dN/dE ∝ E−pe−(E/Emax)), is instantly injected(for more details on the analytical and numerical model see,for example, Ciprini 2008). This form fordN/dE is plausi-ble in presence of time-dependent acceleration or radiative-loss limits (Webb et al. 1984; Drury 1991). The post-flare(lower) state is reproduced by setting the electron energy in-dex p = 1.77, minimal and maximal electron Lorentz fac-tors γmin = 200, γmax = 8 × 104, compactness injectedℓinj = LinjσT /(Rmec

3) = 10−3, radius of the emittingregion in the comoving frameR = 6.5 × 1017 cm, a bulkDoppler factorD = 8 and magnetic field intensityB = 0.024G. If the X-ray spectrum is produced by electrons that cool ontimescales longer than the light crossing time,R/c, the X-rayspectral index would beαX = (p− 1)/2 ≃ 0.4, a value quitesimilar to the averaged value (αX = 0.45±0.03) measured bySwiftXRT for the post-flare interval (bottom segment of Table2). The rather large size of the emitting region (R ≃ 0.2pc)

Page 15: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

PKS 1502+106: a new and distant gamma-ray blazar in outburstdiscovered by the Fermi LAT 15

and the very sub-equipartition magnetic field and reducedDsuggested by this “SSC2” (stand-alone and first order radio-to-gamma-ray SSC) model attempt indicate that even for thislower brightness state, an ERC contribution can be reason-able. Furthermore, the value of the optical-UV spectral in-dex for this post-flare state (αUV OT = 1.9 ± 0.3) is softerthan the averaged index (3/2) expected (Chiang & Bottcher2002), when a totally SSC-dominated loss is considered. Thehuge Compton dominance of the outburst state might precludea first order SSC-stand-alone model attempt for such SEDstate. On the other hand for high-energyγ-ray loud blazarslike PKS 1502+106 superquadratic variations produced byhigher order scatters in SSC are predicted by other models(Georganopoulos et al. 2006; Perlman et al. 2008) and Sec-tion 5.1, in fact in luminous blazars, the optical depth toCompton scattering (related to particle density) increases andsecond and higher-order scatters become more important. Inmore Compton-dominated objects these scattering reactionscan dominate the energy output from the SSC process, andcan produce superquadratic behaviors during big flares.

As mentioned above a “SSC plus ERC” model descrip-tion, joining a SSC radio-to-X-ray component and a ERCgamma-ray component, to both the post-flare and outburststates (Fig.11, dashed blue and red dotted lines) is plausibleas well for these SEDs. The description of an initial versionof this composite modeling, implementing the prescriptionsof Sikora et al. (1994); Dermer & Schlickeiser (2002), can befound, for example, in Tramacere & Tosti (2003). The lowstate is described with the following “SSC+ERC” parametersB = 0.5 G, blob sizeR = 7.9 × 1016 cm,D = 20, a log-parabola electron injection function betweenγmin = 100 andγmax = 3 × 104, Ldisk = 1.1 × 1046, τBLR = 0.1, Tdisk =1.5× 105 K and distance from the disk of1018 cm. The out-burst state is described with the following “SSC+ERC” pa-rametersB = 0.5 G, blob sizeR = 6.3×1016 cm,D = 24, alog-parabola electron injection function betweenγmin = 100andγmax = 3 × 104, Ldisk = 1.1 × 1046, τBLR = 0.1,Tdisk = 1.5× 105 K and distance from the disk of7× 1017

cm.In Fig. 11, the LAT averaged spectrum (E>200 MeV) of

the entire Aug.-Dec. 2008 period considered (characterizedby high statistics and consistent with an intrinsically curvedshape described by a log-parabola model) is reported as well(the strip in the same orange color used for non-simultaneousdata). Log-parabola curvature (Section 2.4) at high energycanbe produced by several plausible models. For example by ra-diative particle cooling and stochastic acceleration processesdriven by magnetic turbulence (rather than systematic particleacceleration) acting near the shock front (Stawarz & Petrosian2008). If the energy where losses balance the accelerationrate if the acceleration time decreases more slowly than theloss time. The probability of energetic gain is lower whenparticle energy increases, because particles are confined by amagnetic field with a confinement efficiency decreasing foran increasing gyration radius. The integral energy distribu-tion of the accelerated particles results in a log-parabolic law:N(γ) = N0(γ/γ0)

−p−1+r log(γ/γ0), wherer is a curvatureterm (Landau et al. 1986; Fossati et al. 2000; Massaro et al.2004; Perlman et al. 2005; Tramacere et al. 2007). Non-power law radiation spectra could also originate in a non-linear regime, such as in the presence of shock modifica-tions, precursors, and when diffusion coefficients vary withparticle momentum (Amato et al. 2008). Another explanation

takes into account episodic particle acceleration, that isappli-cable to high-energy flares with intrinsic spectral curvature(Perlman et al. 2005; Perlman & Wilson 2005). The fillingfactor of the regions within which particles are acceleratedis a function of both position and energy. If the light-crossingtime of the emission region or the integration time of obser-vations is greater than the characteristic particle accelerationtime, we effectively observe an electron distribution which isthe product of a power law multiplied by a logarithmic termproducing a spectral curvature.

Another contribution to such deviations might also be theaveraging over long periods of time, hence combining vari-ability effects, resulting in a cumulation of photons fromdifferent activity “flavors” (different source brightnessandspectral hardness) being included in the evaluation of thespectrum. In the case of PKS 1502+106, the same spectralcurved shape was also preferred for the much shorter, post-flare, interval, while the whole period includes many photonsand covers a wider energy range (photons observed from thesource have energies up to 15.8 GeV). PKS 1502+106, and 3C454.3 (Abdo et al. 2009a) for the first time show this departurefrom a simple power law, while EGRET observed only simplepower laws, probably due to its lower high-energy observa-tional limit. The maximum peak observed in the LP modelsof PKS 1502+106 (in theνFν representation) are around en-ergies of about 390 MeV for the whole period and 480 MeVfor the post-flare period. These peaks are consistent with asub-GeV FSRQ blazar class.

From Sect. 5.1 we remember that the inferred, apparentand isotropic, monochromatic luminosity atE0 = 100 MeVduring the outburst phase (DoY 2008: 218.95 - 224.0) isLE>100 ≃ 1.1 × 1049 erg s−1, and the bolometric lumi-nosity is expected to be even higher than this value, mak-ing PKS 1502+106 one of the most powerful and luminoushigh-energy blazars observed during the first year theFermiLAT all-sky survey. On the other hand The highest energy ofphotons detected from the source during the first four monthsof survey (15.8 GeV, assuming a strict PSF size criterium),has marginal consequences for extragalactic background light(EBL) predictions. The optical depth forγγ → e+e− pairproduction of 16 GeV photons propagating through the EBLfrom a redshiftz = 1.839 source to Earth approaches unityfor rather high-density EBL models (e.g, Stecker et al. 2006,τγγ(z = 1.839, E = 16 GeV) ∼ 1.0 − 1.3), while most lowdensity EBL models predict rather small interaction probabil-ities at such energies. An in-depth exploration of this findingwill be presented elsewhere (Abdo et al. 2009g).

6. FINAL REMARKS

This was the first time that PKS 1502+106, a distant radio-and X-ray-selected FSRQ, has been announced to have ob-servable high-energy gamma-ray emission above 100 MeV.FermiLAT as basically a sort of all-space, -time, and -energymonitor allowed an excellent spatial localization of this newγ-ray source (Section 2.2 and Fig.1, with a firm identifica-tion thanks to the optical-γ-ray match of the outburst and, re-markably, also of a second flare). It allowed detailed analy-sis of the energy spectrum (Section 2.4 and Fig. 11, demon-strating the possibility for an intrinsic spectral curvature inγ-rays), and was able to provide regular, daily monitoring fluxlight curve (Section 2.3 2). This has made possible the dis-covery of consistentγ-ray brightness and activity ignited bythe big outburst over the following 4 months (pointing out a1/f1.3 variability behavior), the discovery of further minor

Page 16: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

16 Abdo et al.

and rapid flares, and the disclosure of the outburst’s tempo-ral shape. This type of “PSD-SED” monitoring performed bythe LAT yielded advantage in developing the first unplannedFermi multifrequency campaign, with a strategic 16 days ofsimultaneousFermi-Swiftmonitoring.

PKS 1502+106 is a powerful gamma-ray (∼ 1049 erg s−1

atE > 100 MeV) FSRQ that showed, especially during thefast-rising outburst, a dominant MeV-GeV bolometric emis-sion similar to other FSRQs of the EGRET era. Dissipationprobably occurred within the BLR, and, assuming for a blackhole mass of∼ 109 M⊙, theγ-ray emission was likely domi-nated by the ERC process. The SSC, in-jet, emission appearsto dominate the observed SED from radio to X-rays bands.PKS 1502+106 might be considered an example of a sub-GeVpeaked blazar, placed at the border of the BLR dissipated andthe dusty torus/ambient-radiation dissipated FSRQs classes.The level of correlation found among theγ-ray, X-ray andoptical-UV outburst and post-flare relaxing phases, supportthis idea. Opacity effects at cm and longer radio wavelengths,possible links between field ordering, jet-axis alignment,su-perluminal radio knots, and MeV-GeV outburst, are also de-picted by our results.

In conclusion, theFermi LAT performance in blazar sci-ence (as a stand alone observatory, or leading instrument formultifrequency campaigns), and the synergy betweenFermiand Swift in particular, is evidenced by this work on PKS1502+106. By itself, this blazar is emerging as a major, lu-minous, energetic andγ-ray variable source, with promisingdiagnostic and discovery potential in emission modeling, inspectral and temporal variability studies, and in understand-ing the radio-gamma-ray connection.

7. ACKNOWLEDGMENTS

This research is based on observations obtained with theFermi Gamma-ray Space Telescope. TheFermi LAT Col-laboration acknowledges generous ongoing support from anumber of agencies and institutes that have supported boththe development and the operation of the LAT as well as sci-entific data analysis. These include the National Aeronau-tics and Space Administration and the Department of Energyin the United States, the Commissariat a l’Energie Atomiqueand the Centre National de la Recherche Scientifique / Insti-tut National de Physique Nucleaire et de Physique des Par-ticules in France, the Agenzia Spaziale Italiana (ASI) andthe Istituto Nazionale di Fisica Nucleare (INFN) in Italy, theMinistry of Education, Culture, Sports, Science and Technol-ogy (MEXT), High Energy Accelerator Research Organiza-tion (KEK) and Japan Aerospace Exploration Agency (JAXA)in Japan, and the K. A. Wallenberg Foundation, the SwedishResearch Council and the Swedish National Space Board in

Sweden.Additional support for science analysis during the oper-

ations phase is gratefully acknowledged from the IstitutoNazionale di Astrofisica (INAF) in Italy and the Centre Na-tional d’Etudes Spatiales in France.

SC acknowledges funding by grant ASI-INAF n.I/047/8/0related to Fermi on-orbit activities.

This research has made use of the NASA/IPAC NEDdatabase (JPL CalTech and NASA, USA), the HEASARCdatabase (LHEA NASA/GSFC and SAO, USA), the Smith-sonian/NASA’s ADS bibliographic databases, and the SIM-BAD database (CDS, Strasbourg, France). This work includesobservations obtained with the NASASwiftgamma-ray burstExplorer. This work includes observations obtained with theSpitzer Space Telescope (operated by the Jet Propulsion Lab-oratory, California Institute of Technology under a contractwith NASA). This work includes observations obtained withXMM-Newton, an ESA science mission with instruments andcontributions directly funded by ESA Member States andNASA. This work has made use of observations obtained withOwens Valley Radio Observatory. The monitoring program atthe OVRO is supported by NASA award NNX08AW31G, andNSF award# AST-0808050. This research has made use ofobservations from the MOJAVE database that is maintainedby the MOJAVE team. The MOJAVE project is supportedunder National Science Foundation grant 0807860-AST andNASA-Fermi grant NNX08AV67G. The National Radio As-tronomy Observatory (NRAO VLBA) is a facility of the Na-tional Science Foundation operated under cooperative agree-ment by Associated Universities, Inc. This research has madeuse of observations obtained with the 100-m telescope of theMPIfR (Max-Planck-Institut fur Radioastronomie) at Effels-berg, Germany. This research has made use of observationsfrom the RATAN–600 that is partly supported by the RussianFoundation for Basic Research (projects 01-02-16812, 05-02-17377, 08-02-00545). This work has made use of observa-tions obtained with the 14m Metsahovi Radio Observatory, aseparate research institute of the Helsinki University of Tech-nology. The Metsahovi team acknowledges the support fromthe Academy of Finland. This work has made use of observa-tions obtained with the TRISPEC instrument on the Kanatatelescope that is operated by Hiroshima University, Japan.YYK is a Research Fellow of the Alexander von HumboldtFoundation.

The LAT team and multifrequency collaboration extendthanks to the anonymous referee who made very usefulcomments.

Facilities: FermiLAT.

REFERENCES

Abdo, A. A., et al., 2009a, ApJ, 699, 817Abdo, A. A., et al., 2009b, ApJ, 700, 597Abdo, A. A., et al., 2009c, ApJS, 183, 46Abdo, A. A., et al., 2009d, ApJ, submitted, (AGN variability)Abdo, A. A., et al., 2009e, ApJ, 697, 934Abdo, A. A., et al., 2009f, ApJ, submitted, (PKS 1510-089)Abdo, A. A., et al., 2009g, ApJ, in prep. (EBL study)Abdo, A. A., et al., 2009h, Atropart. Physics, 32, 193Aharonian, F., et al. 2007, ApJ, 664, L71Akiyama, M., Ueda, Y., Ohta, K., Takahashi, T., & Yamada, T. 2003, ApJS,

148, 275Aller, M. F., Aller, H. D., Hughes, P. A., & Latimer, G. E. 1999, ApJ, 512,

601

Amato, E., Blasi, P., & Gabici, S. 2008, MNRAS, 385, 1946An, T., Hong, X. Y., Venturi, T., Jiang, D. R., & Wang, W. H. 2004, A&A,

421, 839Angelakis, E., Fuhrmann, L., Marchili, N., Krichbaum, T. P., & Zensus, J. A.

2008, Memorie SAIt, 79, 1042Argue, A. N., & Sullivan, C. 1980, The Observatory, 100, 152Atwood, W. B., et al., 2009, ApJ, 697, 1071Atwood, W. B., Bagagli, R., Baldini, L., et al. 2007, Astropart. Phys., 28,

422Baars, J. W. M., Genzel, R. Pauliny-Toth, I. I. K., Witzel, A., 1977, A&A,

61, 99Begelman, M. C., Blandford, R. D., & Rees, M. J. 1984, Reviewsof Modern

Physics, 56, 255

Page 17: arXiv:0912.4029v2 [astro-ph.HE] 21 Jan 2010

PKS 1502+106: a new and distant gamma-ray blazar in outburstdiscovered by the Fermi LAT 17

Bellazzini, R., et al. 2002, Nucl. Phys. B Proc. Sup. 113, 303Bertin, E., & Arnouts, S. 1996, A&AS, 117, 393Blake, G. M. 1970, Astrophys. Lett., 6, 201Burbidge, E. M., & Strittmatter, P. A. 1972, ApJ, 174, L57Burnett, T. H. 2007, AIP Conf. Proc., 921, 530Bottcher, M., & Chiang, J. 2002, ApJ, 581, 127Camilo, F., Ray, P. S., Ransom, S. M. et al. 2009, ApJ, 705, 1Casandjian, J.-M., & Grenier, I. A. 2008, A&A, 489, 849Cecchi, C., Germani, S., Pepe, M., et al. 2007, AIP Conf. Proc., 921, 540Chiaberge, M., & Ghisellini, G. 1999, MNRAS, 306, 551Chiang, J., Carson, J., & Focke, W. 2007, AIP Conf. Proc., 921, 544Chiang, J., Digel, S., Silva, E. D. C. E., & Reimer, O. 2006, Bull. American

Astron. Soc., 38, 382Chiang, J., Bottcher, M. 2002, ApJ, 564, 92Ciaramella, A., Bongardo, C., Aller, H.D. et al. 2004, A&A, 419, 485Ciprini, S. 2008, in proc. of Blazar Variability across the Electromagnetic

Spectrum, PoS(BLAZARS2008), n.073Ciprini, S., Tosti, G., Marcucci, F. et al. 2007a, AIP Conf. Proc., 921, 546Ciprini, S., Takalo, L. O., Tosti, G., et al. 2007b, A&A, 467,465Cooper, N. J., Lister, M. L., & Kochanczyk, M. D. 2007, ApJS, 171, 376Crowther, J. H., & Clarke, R. W. 1966, MNRAS, 132, 405Damiani, F., Maggio, A., Micela, G., & Sciortino, S. 1997, ApJ, 483, 350d’Arcangelo, et al. 2007, ApJ, 659, L107Day, G. A., Shimmins, A. J., Ekers, R. D., & Cole, D. J. 1966, Australian

Journ. Phys., 19, 35D’Elia, V., Padovani, P., & Landt, H. 2003, MNRAS, 339, 1081Dermer, C. D., & Schlickeiser, R. 2002, ApJ, 575, 667Dermer, C. D. 2007, ApJ, 659, 958Drury, L. O. 1991, MNRAS, 251, 340Fichtel, C. E., Bertsch, D. L., Chiang, J., et al. 1994, ApJS,94, 551Fitch, L. T., Dixon, R. S., & Kraus, J. D. 1969, AJ, 74, 612Foschini L., Ghisellini G., Raiteri C.M., et al., 2006, A&A453, 829Fossati, G., Celotti, A., Chiaberge, M., et al. 2000, ApJ, 541, 166Franceschini, A., Rodighiero, G., & Vaccari, M. 2008, A&A, 487, 837Fuhrmann, L., Zensus, J. A., Krichbaum, T. P., Angelakis, E., & Readhead,

A. C. S. 2007, AIP Conf. Proc., 921, 249Fuhrmann, L., Krichbaum, T. P., Witzel, A. et al. 2008, A&A, 490, 1019Gehrels, N., et al. 2004, ApJ, 611, 1005George, I. M., Nandra, K., Turner, T. J., & Celotti, A. 1994, ApJ, 436, L59Georganopoulos, M., Perlman, E. S., Kazanas, D., & Wingert,B. 2006, AIP

Conf. Proc., 350, 178Georganopoulos, M., & Marscher, A. P. 1998, ApJ, 506, L11Ghisellini, G., & Tavecchio, F. 2009a, MNRAS, 397, 985Ghisellini, G., Tavecchio, F., & Ghirlanda, G. 2009b, accepted,arXiv:0906.2195

Graff, P. B., Georganopoulos, M., Perlman, E. S., & Kazanas,D. 2008, ApJ,689, 68

Hartman, R. C., Bottcher, M., Aldering, G., et al. 2001, ApJ, 553, 683Hartman, R. C., Bertsch, D. L., Bloom, S. D., et al. 1999, ApJS, 123, 79Healey, S. E., Romani, R. W., Cotter, G., Michelson, P. F., Schlafly, E. F.,

Readhead, A. C. S., Giommi, P., Chaty, S., Grenier, I. A.; & Weintraub, L.C. 2008, ApJS, 175, 97

Houck, J. R., et al. 2004, ApJS, 154, 18Hovatta, T., Valtaoja, E., Tornikoski, M., Lahteenmaki,A. 2009, A&A, 494,

527Hovatta, T., Tornikoski, M., Lainela, M., Lehto, H. J., Valtaoja, E.,

Torniainen, I., Aller, M. F., & Aller, H. D. 2007, A&A, 469, 899Hughes, P. A., Aller, H. D., & Aller, M. F. 1992, ApJ, 396, 469Inoue, S., & Takahara, F. 1996, ApJ, 463, 555Jorstad, S. et al. 2007, AJ, 134,Kataoka, J., Madejski, G., Sikora, M., et al. 2008, ApJ, 672,787Kellermann K. I., Lister M. L., Homan D. C., et al., 2004, ApJ,609, 539Kirk, J. G., Rieger, F. M., Mastichiadis, A. 1998, A&A, 333, 452Komatsu, E., Dunkley, J., Nolta, M. R., et al. 2009, ApJS, 180, 330Kovalev, Y. Y., Aller, H. D., Aller, M. F., et al. 2009, ApJ, 696, L17Kovalev, Y. Y., Kellermann, K. I., Lister, M. L., et al. 2005,AJ, 130, 2473Kovalev, Y. Y., Nizhelsky, N. A., Kovalev, Yu. A., Berlin, A.B., Zhekanis,

G. V., Mingaliev, M. G., & Bogdantsov, A. V. 1999, A&AS, 139, 545Korolkov, D. V., & Parijskij, Yu. N. 1979, Sky Telesc., 57, 324Landau, R., et al. 1986, ApJ, 308, 78Leipski et al. 2009, ApJ, 701, 891Lister M. L., Aller, H. D., Aller, M. F., 2009, AJ, 137, 3718Lister, M. L., Homan, D. C., Kadler, M., Kellermann, K. I., Kovalev, Y. Y.,

Ros, E., Savolainen, T., & Zensus, J. A. 2009, ApJ, 696, L22Lister M. L., & Homan D. C., 2005, AJ, 130, 1389Liu, Y., Jiang, D. R., & Gu, M. F. 2006, ApJ, 637, 669

Lopez-Caniego, M., Gonzalez-Nuevo, J., Herranz, D., et al. 2007, ApJS,170, 108

Lott, B., Carson, J., Ciprini, S., et al. 2007, AIP Conf. Proc., 921, 347Madejski, G., Takahashi, T., Tashiro, M., et al. 1996, ApJ, 459, 156Marcucci, F., Cecchi, C., & Tosti, G., 2004, Frascati Physics Ser. XXXVII,

285Marscher A. P.; Jorstad, S. G.; d’Arcangelo, F. D., et al. 2008, Nature, 452,

966Massaro, E., Giommi, P., Leto, C., et al. 2009, A&A, 495, 691Massaro, E., Perri, M., Giommi, P., & Nesci, R. 2004, A&A, 413, 489Mattox, J. R., Bertsch, D. L., Chiang, J., et al. 1996, ApJ, 461, 396Mattox, J. R., et al. 1993, ApJ, 410, 609McEnery, J. 2006, ASP Conf. Proc. 350, 229Michelson, P. F. 2007, AIP Conf. Proc., 921, 8Moskalenko, I. V., Jones, F. C., Mashnik, S. G., Ptuskin, V. S., & Strong,

A. W. 2003, Int. Cosmic Ray Conf., 4, 1925Neugebauer, G., Miley, G. K., Soifer, B. T., & Clegg, P. E. 1986, ApJ, 308,

815Nolan, P. L., Tompkins, W. F., Grenier, I. A., & Michelson, P.F. 2003, ApJ,

597, 615Ogle, P., et al. 2009, in prep.Pei, Y. C. 1992, ApJ, 395, 130Perlman, E., Addison, B., Georganopoulos, M., Wingert, B.,& Graff, P.

2008, in proc. of Blazar Variability across the Electromagnetic Spectrum,PoS(BLAZARS2008), n.009

Perlman, E. S., Madejski, G., Georganopoulos, M. et al. 2005, ApJ, 625, 727Perlman, E. S., & Wilson, A. S. 2005, ApJ, 627, 140Petrucci, P. O., Henri, G., Maraschi, L., et al. 2002, A&A, 388, L5Pian, E., et al., 2009, in prep.Poole, T. S., Breeveld, A. A., Page, M. J., et al. 2008, MNRAS,383, 627Readhead, A. C. S., Lawrence, C. R., Myers, S. T., Sargent, W.L. W.,

Hardebeck, H. E., & Moffet, A. T. 1989, ApJ, 346, 566Ritz, S. 2007, AIP Conf. Proc., 921, 3Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJS, 500, 525Sikora, M., Stawarz, Ł., Moderski, R., Nalewajko, K., & Madejski, G. M.

2009, ApJ, 704, 38Sikora, M., Moderski, R., & Madejski, G. M. 2008, ApJ, 675, 71Sikora, M., Błazejowski, M., Moderski, R., & Madejski, G. M. 2002, ApJ,

577, 78Sikora, M., Błazejowski, M., Begelman, M. C., & Moderski, R. 2001, ApJ,

554, 1Sikora, M., Begelman, M. C., & Rees, M. J. 1994, ApJ, 421, 153Smith, H. E., Burbidge, E. M., Baldwin, J. A., Tohline, J. E.,Wampler, E. J.,

Hazard, C., & Murdoch, H. S. 1977, ApJ, 215, 427Sokolov, A., & Marscher, A. P. 2005, ApJ, 629, 52Sokolov, A., Marscher, A. P., & McHardy, I. M. 2004, ApJ, 613,725Sowards-Emmerd, D., Romani, R. W., Michelson, P. F., et al. 2005, ApJ,

626, 95Sowards-Emmerd, D., Romani, R. W., & Michelson, P. F. 2003, ApJ, 590,

109Starck, J.-L., & Pierre, M., A&AS, 128, 397Stawarz, Ł., & Petrosian, V. 2008, ApJ, 681, 1725Stecker, F. W., Malkan, M. A., & Scully, S. T. 2006, ApJ, 648, 774Struder L., Briel U., Dennerl K., et al., 2001, A&A, 365, L18Thompson, D. J. 2007, AIP Conf. Proc., 921, 86Thompson, D.J., 2006, ASP Conf. Ser., 350, 113Thompson, D. J., Bertsch, D. L., Dingus, B. L. et al. 1996, ApJS, 107, 227Terasranta, H., Wiren, S., Koivisto, P., Saarinen, V., & Hovatta, T. 2005,

A&A, 440, 409Terasranta, H., Tornikoski, M., Mujunen, A., et al. 1998, A&AS, 132, 305Tosti, G. 2007, AIP Conf. Proc., 921, 255Tramacere, A., Massaro, F., & Cavaliere, A. 2007, A&A, 466, 521Tramacere, A., & Tosti, G. 2003, New Astron. Rev., 47, 697Uemura, M., Sasada, M., Arai, A., et al. 2008, in proc. of Blazar Variability

across the Electromagnetic Spectrum, PoS(BLAZARS2008), n.070Watanabe, M., Nakaya, H., Yamamuro, T., et al. 2005, PASP, 117, 870Watanabe, C., Ohta. K., Akiyama, M., & Ueda, Y., 2004, ApJ, 610, 128Webb, G. M., Drury, L. O., & Biermann, P. 1984, A&A, 137, 185Wehrle, A., et al. 1998, ApJ, 497, 178Wilkes, B. J., Wright, A. E., Jauncey, D. L., & Peterson, B. A.1983, Proc. of

Astronomical Society of Australia, 5, 2Williams, P. J. S., Kenderdine, S., & Baldwin, J. E. 1967, MmRAS, 70, 53Wright, A. E., Peterson, B. A., Jauncey, D. L., & Condon, J. J.1979, ApJ,

229, 73