arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

22
arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017 Draft version February 7, 2017 Preprint typeset using L A T E X style emulateapj v. 5/2/11 SPACE TELESCOPE AND OPTICAL REVERBERATION MAPPING PROJECT. V. OPTICAL SPECTROSCOPIC CAMPAIGN AND EMISSION-LINE ANALYSIS FOR NGC 5548 L. Pei 1,2 , M. M. Fausnaugh 3 , A. J. Barth 1 , B. M. Peterson 3,4,5 , M. C. Bentz 6 , G. De Rosa 5 , K. D. Denney 3,4,7 , M. R. Goad 8 , C. S. Kochanek 3,4 , K. T. Korista 9 , G. A. Kriss 5,10 , R. W. Pogge 3,4 , V. N. Bennert 11 , M. Brotherton 12 , K. I. Clubb 13 , E. Dalla Bont` a 14,15 , A. V. Filippenko 13 , J. E. Greene 16 , C. J. Grier 3,17,18 , M. Vestergaard 19,20 , W. Zheng 13 , Scott M. Adams 3,21 , Thomas G. Beatty 3,17,22 , A. Bigley 13 , Jacob E. Brown 23 , Jonathan S. Brown 3 , G. Canalizo 24 , J. M. Comerford 25 , Carl T. Coker 3 , E. M. Corsini 14,15 , S. Croft 13 , K. V. Croxall 3,4 , A. J. Deason 26 , Michael Eracleous 17,18,27,28 , O. D. Fox 13 , E. L. Gates 29 , C. B. Henderson 3,30,31 , E. Holmbeck 32 , T. W.-S. Holoien 3,4 , J. J. Jensen 19 , C. A. Johnson 33 , P. L. Kelly 34,35,36 , S. Kim 3,4 , A. King 37 , M. W. Lau 26 , Miao Li 38 , Cassandra Lochhaas 3 , Zhiyuan Ma 23 , E. R. Manne-Nicholas 6 , J. C. Mauerhan 13 , M. A. Malkan 32 , R. McGurk 26,39 , L. Morelli 14,15 , Ana Mosquera 3,40 , Dale Mudd 3 , F. Muller Sanchez 25 , M. L. Nguyen 12 , P. Ochner 14,15 , B. Ou-Yang 6 , A. Pancoast 41,42,43 , Matthew T. Penny 3,44 , A. Pizzella 14,15 , Radoslaw Poleski 3 , Jessie Runnoe 17,18,45 , B. Scott 24 , Jaderson S. Schimoia 3,46 , B. J. Shappee 47,48 , I. Shivvers 13 , Gregory V. Simonian 3 , A. Siviero 14 , Garrett Somers 3,49 , Daniel J. Stevens 3 , M. A. Strauss 16 , Jamie Tayar 3 , N. Tejos 50,51 , T. Treu 32,41,52 , J. Van Saders 47 , L. Vican 32 , S. Villanueva Jr. 3 , H. Yuk 13 , N. L. Zakamska 10 , W. Zhu 3 , M. D. Anderson 6 , P. Ar´ evalo 53 , C. Bazhaw 6 , S. Bisogni 3,54 , G. A. Borman 55 , M. C. Bottorff 56 , W. N. Brandt 17,18,57 , A. A. Breeveld 58 , E. M. Cackett 59 , M. T. Carini 60 , D. M. Crenshaw 6 , A. De Lorenzo-C´ aceres 61 , M. Dietrich 62,63 , R. Edelson 64 , N. V. Efimova 65 , J. Ely 5 , P. A. Evans 8 , G. J. Ferland 66 , K. Flatland 67 , N. Gehrels 68 , S. Geier 69,70,71 , J. M. Gelbord 72,73 , D. Grupe 74 , A. Gupta 3 , P. B. Hall 75 , S. Hicks 60 , D. Horenstein 6 , Keith Horne 61 , T. Hutchison 56 , M. Im 76 , M. D. Joner 77 , J. Jones 6 , J. Kaastra 78,79,80 , S. Kaspi 81,82 , B. C. Kelly 41 , J. A. Kennea 17 , M. Kim 83 , S. C. Kim 83 , S. A. Klimanov 66 , J. C. Lee 83 , D. C. Leonard 67 , P. Lira 84 , F. MacInnis 56 , S. Mathur 3,4 , I. M. M c Hardy 85 , C. Montouri 86 , R. Musso 56 , S. V. Nazarov 55 , H. Netzer 81 , R. P. Norris 6 , J. A. Nousek 17 , D. N. Okhmat 55 , I. Papadakis 87,88 , J. R. Parks 6 , J.-U. Pott 39 , S. E. Rafter 82,89 , H.-W. Rix 39 , D. A. Saylor 6 , K. Schn¨ ulle 39 , S. G. Sergeev 55 , M. Siegel 90 , A. Skielboe 19 , M. Spencer 77 , D. Starkey 61 , H.-I. Sung 83 , K. G. Teems 6 , C. S. Turner 6 , P. Uttley 91 , C. Villforth 92 , Y. Weiss 82 , J.-H. Woo 76 , H. Yan 23 , S. Young 64 , and Y. Zu 93,4 Draft version February 7, 2017 ABSTRACT We present the results of an optical spectroscopic monitoring program targeting NGC 5548 as part of a larger multi-wavelength reverberation mapping campaign. The campaign spanned six months and achieved an almost daily cadence with observations from five ground-based telescopes. The Hβ and He II λ4686 broad emission-line light curves lag that of the 5100 ˚ A optical continuum by 4.17 +0.36 0.36 days and 0.79 +0.35 0.34 days, respectively. The Hβ lag relative to the 1158 ˚ A ultraviolet continuum light curve measured by the Hubble Space Telescope is roughly 50% longer than that measured against the optical continuum, and the lag difference is consistent with the observed lag between the optical and ultraviolet continua. This suggests that the characteristic radius of the broad-line region is 50% larger than the value inferred from optical data alone. We also measured velocity-resolved emission-line lags for Hβ and found a complex velocity-lag structure with shorter lags in the line wings, indicative of a broad-line region dominated by Keplerian motion. The responses of both the Hβ and He II emission lines to the driving continuum changed significantly halfway through the campaign, a phenomenon also observed for C IV, Lyα, He II(+O III]), and Si IV(+O IV]) during the same monitoring period. Finally, given the optical luminosity of NGC 5548 during our campaign, the measured Hβ lag is a factor of five shorter than the expected value implied by the R BLR L AGN relation based on the past behavior of NGC 5548. 1 Department of Physics and Astronomy, 4129 Frederick Reines Hall, University of California, Irvine, CA 92697, USA 2 Department of Astronomy, University of Illinois at Urbana- Champaign, Urbana, IL 61801, USA 3 Department of Astronomy, The Ohio State University, 140 W 18th Ave, Columbus, OH 43210, USA 4 Center for Cosmology and AstroParticle Physics, The Ohio State University, 191 West Woodruff Ave, Columbus, OH 43210, USA 5 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA 6 Department of Physics and Astronomy, Georgia State University, 25 Park Place, Suite 605, Atlanta, GA 30303, USA 7 NSF Postdoctoral Research Fellow 8 Department of Physics and Astronomy, University of Leicester, Leicester, LE1 7RH, UK 9 Department of Physics, Western Michigan University, 1120 Everett Tower, Kalamazoo, MI 49008, USA 10 Department of Physics and Astronomy, The Johns Hopkins University, Baltimore, MD 21218, USA 11 Physics Department, California Polytechnic State Univer- sity, San Luis Obispo, CA 93407, USA 12 Department of Physics and Astronomy, University of Wyoming, 1000 E. University Ave. Laramie, WY 82071, USA 13 Department of Astronomy, University of California, Berke- ley, CA 94720-3411, USA 14 Dipartimento di Fisica e Astronomia “G. Galilei,” Univer- sit` a di Padova, Vicolo dell’Osservatorio 3, I-35122 Padova, Italy 15 INAF-Osservatorio Astronomico di Padova, Vicolo dell’Osservatorio 5 I-35122, Padova, Italy 16 Department of Astrophysical Sciences, Princeton Univer- sity, Princeton, NJ 08544, USA 17 Department of Astronomy and Astrophysics, Eberly Col-

Transcript of arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

Page 1: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

arX

iv:1

702.

0117

7v1

[as

tro-

ph.G

A]

3 F

eb 2

017

Draft version February 7, 2017Preprint typeset using LATEX style emulateapj v. 5/2/11

SPACE TELESCOPE AND OPTICAL REVERBERATION MAPPING PROJECT. V. OPTICALSPECTROSCOPIC CAMPAIGN AND EMISSION-LINE ANALYSIS FOR NGC 5548

L. Pei1,2, M. M. Fausnaugh3, A. J. Barth1, B. M. Peterson3,4,5, M. C. Bentz6, G. De Rosa5, K. D. Denney3,4,7,M. R. Goad8, C. S. Kochanek3,4, K. T. Korista9, G. A. Kriss5,10, R. W. Pogge3,4, V. N. Bennert11,

M. Brotherton12, K. I. Clubb13, E. Dalla Bonta14,15, A. V. Filippenko13, J. E. Greene16, C. J. Grier3,17,18,M. Vestergaard19,20, W. Zheng13, Scott M. Adams3,21, Thomas G. Beatty3,17,22, A. Bigley13, Jacob E. Brown23,

Jonathan S. Brown3, G. Canalizo24, J. M. Comerford25, Carl T. Coker3, E. M. Corsini14,15, S. Croft13,K. V. Croxall3,4, A. J. Deason26, Michael Eracleous17,18,27,28, O. D. Fox13, E. L. Gates29, C. B. Henderson3,30,31,

E. Holmbeck32, T. W.-S. Holoien3,4, J. J. Jensen19, C. A. Johnson33, P. L. Kelly34,35,36, S. Kim3,4, A. King37,M. W. Lau26, Miao Li38, Cassandra Lochhaas3, Zhiyuan Ma23, E. R. Manne-Nicholas6, J. C. Mauerhan13,M. A. Malkan32, R. McGurk26,39, L. Morelli14,15, Ana Mosquera3,40, Dale Mudd3, F. Muller Sanchez25,M. L. Nguyen12, P. Ochner14,15, B. Ou-Yang6, A. Pancoast41,42,43, Matthew T. Penny3,44, A. Pizzella14,15,

Rados law Poleski3, Jessie Runnoe17,18,45, B. Scott24, Jaderson S. Schimoia3,46, B. J. Shappee47,48, I. Shivvers13,Gregory V. Simonian3, A. Siviero14, Garrett Somers3,49, Daniel J. Stevens3, M. A. Strauss16, Jamie Tayar3,

N. Tejos50,51, T. Treu32,41,52, J. Van Saders47, L. Vican32, S. Villanueva Jr.3, H. Yuk13, N. L. Zakamska10,W. Zhu3, M. D. Anderson6, P. Arevalo53, C. Bazhaw6, S. Bisogni3,54, G. A. Borman55, M. C. Bottorff56,

W. N. Brandt17,18,57, A. A. Breeveld58, E. M. Cackett59, M. T. Carini60, D. M. Crenshaw6,A. De Lorenzo-Caceres61, M. Dietrich62,63, R. Edelson64, N. V. Efimova65, J. Ely5, P. A. Evans8, G. J. Ferland66,K. Flatland67, N. Gehrels68, S. Geier69,70,71, J. M. Gelbord72,73, D. Grupe74, A. Gupta3, P. B. Hall75, S. Hicks60,

D. Horenstein6, Keith Horne61, T. Hutchison56, M. Im76, M. D. Joner77, J. Jones6, J. Kaastra78,79,80, S. Kaspi81,82,B. C. Kelly41, J. A. Kennea17, M. Kim83, S. C. Kim83, S. A. Klimanov66, J. C. Lee83, D. C. Leonard67, P. Lira84,

F. MacInnis56, S. Mathur3,4, I. M. McHardy85, C. Montouri86, R. Musso56, S. V. Nazarov55, H. Netzer81,R. P. Norris6, J. A. Nousek17, D. N. Okhmat55, I. Papadakis87,88, J. R. Parks6, J.-U. Pott39, S. E. Rafter82,89,

H.-W. Rix39, D. A. Saylor6, K. Schnulle39, S. G. Sergeev55, M. Siegel90, A. Skielboe19, M. Spencer77,D. Starkey61, H.-I. Sung83, K. G. Teems6, C. S. Turner6, P. Uttley91, C. Villforth92, Y. Weiss82, J.-H. Woo76,

H. Yan23, S. Young64, and Y. Zu93,4

Draft version February 7, 2017

ABSTRACT

We present the results of an optical spectroscopic monitoring program targeting NGC 5548 as partof a larger multi-wavelength reverberation mapping campaign. The campaign spanned six monthsand achieved an almost daily cadence with observations from five ground-based telescopes. The Hβand He II λ4686 broad emission-line light curves lag that of the 5100 A optical continuum by 4.17+0.36

−0.36

days and 0.79+0.35−0.34 days, respectively. The Hβ lag relative to the 1158 A ultraviolet continuum light

curve measured by the Hubble Space Telescope is roughly ∼50% longer than that measured againstthe optical continuum, and the lag difference is consistent with the observed lag between the opticaland ultraviolet continua. This suggests that the characteristic radius of the broad-line region is ∼50%larger than the value inferred from optical data alone. We also measured velocity-resolved emission-linelags for Hβ and found a complex velocity-lag structure with shorter lags in the line wings, indicative ofa broad-line region dominated by Keplerian motion. The responses of both the Hβ and He II emissionlines to the driving continuum changed significantly halfway through the campaign, a phenomenonalso observed for C IV, Lyα, He II(+O III]), and Si IV(+O IV]) during the same monitoring period.Finally, given the optical luminosity of NGC 5548 during our campaign, the measured Hβ lag is afactor of five shorter than the expected value implied by the RBLR −LAGN relation based on the pastbehavior of NGC 5548.

1 Department of Physics and Astronomy, 4129 FrederickReines Hall, University of California, Irvine, CA 92697, USA

2 Department of Astronomy, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA

3 Department of Astronomy, The Ohio State University, 140W 18th Ave, Columbus, OH 43210, USA

4 Center for Cosmology and AstroParticle Physics, The OhioState University, 191 West Woodruff Ave, Columbus, OH 43210,USA

5 Space Telescope Science Institute, 3700 San Martin Drive,Baltimore, MD 21218, USA

6 Department of Physics and Astronomy, Georgia StateUniversity, 25 Park Place, Suite 605, Atlanta, GA 30303, USA

7 NSF Postdoctoral Research Fellow8 Department of Physics and Astronomy, University of

Leicester, Leicester, LE1 7RH, UK9 Department of Physics, Western Michigan University, 1120

Everett Tower, Kalamazoo, MI 49008, USA10 Department of Physics and Astronomy, The Johns Hopkins

University, Baltimore, MD 21218, USA11 Physics Department, California Polytechnic State Univer-

sity, San Luis Obispo, CA 93407, USA12 Department of Physics and Astronomy, University of

Wyoming, 1000 E. University Ave. Laramie, WY 82071, USA13 Department of Astronomy, University of California, Berke-

ley, CA 94720-3411, USA14 Dipartimento di Fisica e Astronomia “G. Galilei,” Univer-

sita di Padova, Vicolo dell’Osservatorio 3, I-35122 Padova, Italy15 INAF-Osservatorio Astronomico di Padova, Vicolo

dell’Osservatorio 5 I-35122, Padova, Italy16 Department of Astrophysical Sciences, Princeton Univer-

sity, Princeton, NJ 08544, USA17 Department of Astronomy and Astrophysics, Eberly Col-

Page 2: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

2

1. INTRODUCTION

Broad emission lines are among the most striking fea-tures of quasars and active galactic nuclei (AGNs). TheseDoppler-broadened lines are emitted by gas occupyingthe broad-line region (BLR), which is located within sev-eral light-days to light-months of the central supermas-sive black hole (SMBH; e.g., Antonucci & Cohen 1983;Clavel et al. 1991; Peterson et al. 1998, 2004; Bentz et al.

lege of Science, The Pennsylvania State University, 525 DaveyLaboratory, University Park, PA 16802, USA

18 Institute for Gravitation and the Cosmos, The Pennsylva-nia State University, University Park, PA 16802, USA

19 Dark Cosmology Centre, Niels Bohr Institute, Universityof Copenhagen, Juliane Maries Vej 30, DK-2100 CopenhagenØ, Denmark

20 Steward Observatory, University of Arizona, 933 NorthCherry Avenue, Tucson, AZ 85721, USA

21 Cahill Center for Astrophysics, California Institute ofTechnology, Pasadena, CA 91125, USA

22 Center for Exoplanets and Habitable Worlds, The Penn-sylvania State University, University Park, PA 16802, USA

23 Department of Physics and Astronomy, University ofMissouri, Columbia, MO 65211, USA

24 Department of Astronomy, University of California, River-side, CA 92521, USA

25 Department of Astrophysical and Planetary Sciences,University of Colorado, Boulder, CO 80309, USA

26 Department of Astronomy and Astrophysics, Universityof California Santa Cruz, 1156 High Street, Santa Cruz, CA95064, USA

27 Center for Relativistic Astrophysics, Georgia Institute ofTechnology, Atlanta, GA 30332, USA

28 Department of Astronomy, University of Washington, Box351580, Seattle, WA 98195, USA

29 Lick Observatory, P.O. Box 85, Mt. Hamilton, CA 95140,USA

30 Jet Propulsion Laboratory, California Institute of Technol-ogy, 4800 Oak Grove Drive, Pasadena, CA 91109, USA

31 NASA Postdoctoral Program Fellow32 Department of Physics and Astronomy, University of

California, Los Angeles, CA 90095, USA33 Santa Cruz Institute for Particle Physics and Department

of Physics, University of California, Santa Cruz, CA 95064,USA

34 Department of Physics, Stanford University, 382 ViaPueblo Mall, Stanford, CA 94305, USA

35 Kavli Institute for Particle Astrophysics and Cosmology,Stanford University, Stanford, CA 94305, USA

36 SLAC National Accelerator Laboratory, 2575 Sand HillRoad, Menlo Park, CA 94025, USA

37 School of Physics, University of Melbourne, Parkville, VIC3010, Australia

38 Department of Astronomy, Columbia University, 550W120th Street, New York, NY 10027, USA

39 Max Planck Institut fur Astronomie, Konigstuhl 17,D–69117 Heidelberg, Germany

40 Physics Department, United States Naval Academy, An-napolis, MD 21403, USA

41 Department of Physics, University of California, SantaBarbara, CA 93106, USA

42 Harvard-Smithsonian Center for Astrophysics, 60 GardenStreet, Cambridge, MA 02138, USA

43 Einstein Fellow44 Sagan Fellow45 Department of Astronomy, University of Michigan, 500

Church Street, Ann Arbor, MI 48109, USA46 Instituto de Fısica, Universidade Federal do Rio do Sul,

Campus do Vale, Porto Alegre, Brazil47 Carnegie Observatories, 813 Santa Barbara Street,

Pasadena, CA 91101, USA48 Carnegie-Princeton Fellow, Hubble Fellow49 Department of Physics and Astronomy, Vanderbilt Univer-

sity, 6301 Stevenson Circle, Nashville, TN 37235, USA50 Millennium Institute of Astrophysics, Santiago, Chile51 Instituto de Astrofısica, Pontificia Universidad Catolica de

Chile, Vicuna Mackenna 4860, Santiago, Chile

2009b; Grier et al. 2013). The geometry and kinemat-ics of the BLR play a significant role in AGN researchbecause these properties can be used to infer the massof the central black hole (e.g., Gaskell & Sparke 1986;Clavel et al. 1991; Kaspi et al. 2000; Denney et al. 2006,2010; Pancoast et al. 2014). Additionally, it is possiblethat infalling BLR gas may fuel SMBH accretion (e.g.,Peterson 2006; Gaskell & Goosmann 2016) and outflow-

52 Packard Fellow53 Instituto de Fısica y Astronomıa, Facultad de Ciencias,

Universidad de Valparaıso, Gran Bretana N 1111, Playa Ancha,Valparaıso, Chile

54 Osservatorio Astrofisico di Arcetri, largo E. Fermi 5, 50125,Firenze, Italy

55 Crimean Astrophysical Observatory, P/O Nauchny, Crimea298409, Russia

56 Fountainwood Observatory, Department of Physics FJS149, Southwestern University, 1011 E. University Ave., George-town, TX 78626, USA

57 Department of Physics, 104 Davey Laboratory, ThePennsylvania State University, University Park, PA 16802, USA

58 Mullard Space Science Laboratory, University CollegeLondon, Holmbury St. Mary, Dorking, Surrey RH5 6NT, UK

59 Department of Physics and Astronomy, Wayne StateUniversity, 666 W. Hancock St, Detroit, MI 48201, USA

60 Department of Physics and Astronomy, Western KentuckyUniversity, 1906 College Heights Blvd #11077, Bowling Green,KY 42101, USA

61 SUPA Physics and Astronomy, University of St. Andrews,Fife, KY16 9SS Scotland, UK

62 Department of Physics and Astronomy, Ohio University,Athens, OH 45701, USA

63 Department of Earth, Environment and Physics, WorcesterState University, Worcester, MA 01602, USA

64 Department of Astronomy, University of Maryland, CollegePark, MD 20742, USA

65 Pulkovo Observatory, 196140 St. Petersburg, Russia66 Department of Physics and Astronomy, The University of

Kentucky, Lexington, KY 40506, USA67 Department of Astronomy, San Diego State University, San

Diego, CA 92182, USA68 Astrophysics Science Division, NASA Goddard Space

Flight Center, Mail Code 661, Greenbelt, MD 20771, USA69 Instituto de Astrofısica de Canarias, 38200 La Laguna,

Tenerife, Spain70 Departamento de Astrofısica, Universidad de La Laguna,

E-38206 La Laguna, Tenerife, Spain71 Gran Telescopio Canarias (GRANTECAN), 38205 San

Cristobal de La Laguna, Tenerife, Spain72 Spectral Sciences Inc., 4 Fourth Ave., Burlington, MA

01803, USA73 Eureka Scientific Inc., 2452 Delmer St. Suite 100, Oakland,

CA 94602, USA74 Space Science Center, Morehead State University, 235

Martindale Dr., Morehead, KY 40351, USA75 Department of Physics and Astronomy, York University,

Toronto, ON M3J 1P3, Canada76 Astronomy Program, Department of Physics & Astronomy,

Seoul National University, Seoul, Republic of Korea77 Department of Physics and Astronomy, N283 ESC,

Brigham Young University, Provo, UT 84602, USA78 SRON Netherlands Institute for Space Research, Sorbon-

nelaan 2, 3584 CA Utrecht, The Netherlands79 Department of Physics and Astronomy, Univeristeit

Utrecht, P.O. Box 80000, 3508 Utrecht, The Netherlands80 Leiden Observatory, Leiden University, PO Box 9513, 2300

RA Leiden, The Netherlands81 School of Physics and Astronomy, Raymond and Beverly

Sackler Faculty of Exact Sciences, Tel Aviv University, Tel Aviv69978, Israel

82 Physics Department, Technion, Haifa 32000, Israel83 Korea Astronomy and Space Science Institute, Republic of

Korea84 Departamento de Astronomia, Universidad de Chile,

Camino del Observatorio 1515, Santiago, Chile85 University of Southampton, Highfield, Southampton, SO17

Page 3: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

3

ing gas may be part of disk winds that carry away an-gular momentum from the disk and provide energy andmomentum feedback to the host galaxy (e.g., Emmeringet al. 1992; Murray & Chiang 1997; Kollatschny 2003;Leighly & Moore 2004). Understanding the dynamicalstate and physical conditions of gas in the BLR is of keyimportance in completing our understanding of the AGNphenomenon.

Owing to its small angular size, the BLR is currentlyimpossible to resolve spatially even for the closest AGNs.An alternative method to study this region is to resolve itin the time domain using reverberation mapping (RM),a technique that leverages the variable nature of quasarsand Seyferts (Blandford & McKee 1982; Peterson 1993,2014). AGNs exhibit stochastic flux variations, possiblybecause of inhomogeneous accretion and thermal fluctu-ations in the accretion disk (Czerny et al. 1999; Collier& Peterson 2001; Czerny et al. 2003; Kelly et al. 2009;Koz lowski et al. 2010; MacLeod et al. 2010). Photonsfrom the central engine ionize the BLR gas, which thenechoes continuum-flux variations with a light-travel timelag, τ . The emission-line flux L(vr, t) at time t and line-of-sight velocity vr is related to the ionizing continuumby

L(vr, t) =

∫ ∞

0

Ψ(vr, τ)C(t − τ)dτ, (1)

where C(t − τ) is the continuum emission at an earliertime t−τ , and Ψ(v, τ) is the transfer function that mapsthe continuum light curve to the time-variable line profile(Blandford & McKee 1982).

The transfer function—also known as the velocity-delay map—encodes important information about theBLR’s geometry and kinematics. There has been tremen-dous effort by many groups to recover velocity-delaymaps (Rosenblatt & Malkan 1990; Horne et al. 1991;Krolik et al. 1991; Ulrich & Horne 1996; Bentz et al.2010a; Pancoast et al. 2011; Li et al. 2013; Grier et al.2013; Pancoast et al. 2014; Skielboe et al. 2015) andvelocity-resolved line lags (e.g., Kollatschny 2003; Bentzet al. 2009b; Denney et al. 2010; Barth et al. 2011; Duet al. 2016a). In order to obtain Ψ(v, τ), RM cam-paigns must have a combination of high cadence, longduration, high photometric precision, and high signal-to-noise ratios (SNR), which is often not achievable byground-based programs. More typically, RM campaigns

1BJ, UK86 DiSAT, Universita dell’Insubria, via Valleggio 11, 22100,

Como, Italy87 Department of Physics and Institute of Theoretical and

Computational Physics, University of Crete, GR-71003 Herak-lion, Greece

88 IESL, Foundation for Research and Technology, GR-71110Heraklion, Greece

89 Department of Physics, Faculty of Natural Sciences,University of Haifa, Haifa 31905, Israel

90 Las Cumbres Observatory Global Telescope Network, 6740Cortona Drive, Suite 102, Goleta, CA 93117, USA

91 Astronomical Institute ‘Anton Pannekoek,’ University ofAmsterdam, Postbus 94249, NL-1090 GE Amsterdam, TheNetherlands

92 University of Bath, Department of Physics, ClavertonDown, BA2 7AY, Bath, UK

93 Department of Physics, Carnegie Mellon University, 5000Forbes Avenue, Pittsburgh, PA 15213, USA

are able to only measure the mean emission-line lag τ ,which represents the response-weighted mean light-traveltime from the ionizing continuum to the BLR.

Assuming that the broad-line width is a result of thevirialized motion of gas within the black hole’s potentialwell, the emission-line lag and gas velocity dispersion in-ferred from the line width (∆V ) can be used to infer theblack hole (BH) mass using

MBH = fcτ∆V 2

G. (2)

Here, cτ = RBLR is the characteristic radius of the BLR,and f is a dimensionless calibration factor of order unitythat accounts for the unknown BLR geometry and kine-matics. Ground-based RM campaigns have produced BHmass measurements for ∼60 local AGNs to date (seeBentz & Katz 2015 for references and a recent compi-lation). RM is also starting to be used for objects at cos-mological distances (Kaspi et al. 2007; King et al. 2015;Shen et al. 2016) with the aims of studying the UV con-tinuum and emission lines and calibrating BH masses athigh redshifts.

The ionizing continuum is emitted at wavelengths <912 A and is generally unobservable due to the Lymanlimit of the host galaxy. Given this limitation, the far-ultraviolet (UV) continuum at λ ≈ 1100–1500 A shouldbe used to derive emission-line lags because it is close inwavelength to the ionizing continuum and should there-fore serve as an accurate proxy. However, wavelengthsshorter than ∼3200 A are inaccessible from the ground,so the rest-frame optical continuum is often used as aproxy for the ionizing source in low-redshift AGNs. Al-though the far-UV and optical continua have been shownto vary almost simultaneously in some cases (e.g., Clavelet al. 1991; Reichert et al. 1994; Korista et al. 1995; Wan-ders et al. 1997), more recent high-cadence studies havefound that the optical continuum can lag the UV contin-uum by up to a few days (Collier et al. 1998; Sergeev et al.2005; McHardy et al. 2014; Shappee et al. 2014; Edelsonet al. 2015; Fausnaugh et al. 2016). This can significantlyaffect the measured broad-line lag if the BLR has a char-acteristic radius on the order of light days. The variableoptical continuum has also been shown to have smootherfeatures and smaller amplitudes than its UV counterpart(e.g., Peterson et al. 1991; Dietrich et al. 1993; Stirpeet al. 1994; Santos-Lleo et al. 1997; Dietrich et al. 1998;Shappee et al. 2014; Fausnaugh et al. 2016). These differ-ences between the UV and optical continua suggest thatthe optical continuum is not fully interchangeable withthe ionizing source for determining reverberation lags.

Furthermore, a long-standing assumption in RM isthat the source of the ionizing photons in a typicalSeyfert galaxy is physically much smaller than the BLR(about a factor of 100; e.g., Peterson 1993; Peterson &Horne 2004). This assumption implies that the disksize can be neglected when determining RBLR from RMdata. However, Fausnaugh et al. (2016) have shown thatthe optically emitting portion of the accretion disk hasa lag similar to that of the inner portion of the BLR.If we assume a model in which the measured lags arepurely dependent on the radial distance from the ioniz-ing source, then the emission-line lags measured using theoptical continuum may significantly underestimate the

Page 4: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

4

BLR characteristic radius. Since most RM campaignsuse only optical data, it is imperative that we under-stand the systematic effects of using the optical ratherthan the UV continuum in RM studies and the relevantimplications for BH mass estimates.

To this end, we present the results of a six-monthground-based RM program monitoring the galaxy NGC5548 (redshift z = 0.0172). This paper is the fifth in a se-ries describing results from the AGN Space Telescope andOptical Reverberation Mapping (AGN STORM) cam-paign, the most intensive multi-wavelength AGN mon-itoring program to date. The campaign is centeredaround 171 epochs of daily cadence observations usingthe Cosmic Origins Spectrograph on the Hubble SpaceTelescope (HST ). Concurrent with the HST programwere four months of Swift observations and six monthsof ground-based photometric and spectroscopic observa-tions. First results of the HST , Swift , and ground-basedphotometry programs were presented by De Rosa et al.(2015), Edelson et al. (2015), and Fausnaugh et al. (2016)(Papers I–III, respectively). Goad et al. (2016) (PaperIV) explore the anomalous behavior of the UV contin-uum and broad emission-line light curves observed dur-ing a portion of this campaign. This paper focuses onthe ground-based spectroscopic data and emission-lineanalysis.

NGC 5548 is one of the best-studied Seyfert galaxiesand has been the subject of many past RM programs.Most notably, it was the target of a 13-year campaigncarried out by the AGN Watch consortium (Petersonet al. 2002, and references therein), which was initiallydesigned to support UV monitoring of NGC 5548 car-ried out by the International Ultraviolet Explorer (IUE;Clavel et al. 1991). Individual years of this campaignachieved median sampling cadences of 1–3 days for spec-troscopic observations. Subsequently, NGC 5548 wasmonitored in programs described by Bentz et al. (2007),Denney et al. (2009), Bentz et al. (2009b), and De Rosa etal. (AGN12, results in preparation) with campaign dura-tions of 40 days, 135 days, 64 days, and 120 days (respec-tively), and each with a median sampling cadence of ∼ 1day. A more recent RM program described by Lu et al.(2016) monitored this AGN for 180 days with a medianspectroscopic sampling of ∼ 3 days. The 2014 AGNSTORM campaign’s combination of daily cadence, six-month duration, and multi-wavelength coverage makes itthe most intensive RM campaign ever conducted.

There are two primary goals of the present work. Thefirst is to compare the Hβ emission-line lag measuredagainst simultaneously observed far-UV and optical con-tinua in order to understand the effects of substitutingthe optical continuum for the ionizing continuum in re-verberation measurements. The second goal is to exam-ine in detail the responses of the optical emission lines tocontinuum variations and compare them to those of theUV lines, which will provide a more complete picture ofthe structure and kinematics of the BLR than previousstudies that used only optical data.

We describe the spectroscopic observations and reduc-tions in Section 2. Section 3 details our procedures forflux and light-curve measurements. In Section 4, wepresent our analysis of emission-line lags, line responses,line profiles, and BH mass measurements. We discussthe implications of our results and compare our mea-

Fig. 1.— Mean spectrum of NGC 5548 from the Asiago dataset,which includes 21 epochs of spectra with spectral resolution of 1.0A pixel−1 and has a median SNR of 160. Labeled are the He IIλ4686 Hβ λ4861, and [O III] λλ4959, 5007 emission lines.

surements to those from previous campaigns in Section5. Section 6 summarizes our findings. We quote wave-lengths in the rest frame of NGC 5548 unless otherwisestated.

2. OBSERVATIONS AND DATA REDUCTION

Spectroscopic data were obtained from five telescopes:the McGraw-Hill 1.3-m telescope at the MDM Observa-tory, the Shane 3-m telescope at the Lick Observatory,the 1.22-m Galileo telescope at the Asiago AstrophysicalObservatory, the 3.5-m telescope at Apache Point Obser-vatory (APO), and the 2.3-m telescope at the WyomingInfrared Observatory (WIRO). Observations at MDMwere carried out with a slit width of 5′′ oriented in thenorth-south direction, and spectra at the other telescopeswere taken with a 5′′-wide slit oriented at the parallac-tic angle (Filippenko 1982). The optical spectroscopicmonitoring began on 2014 January 4 (UT dates are usedthroughout this paper) and continued through 2014 July6 with approximately daily cadence.

Table 1 lists the properties of the telescopes and in-struments used to obtain spectroscopic data, and Fig-ure 1 shows the mean spectrum constructed using datafrom Asiago, which obtained the only spectra that coverthe full optical wavelength range. MDM contributedthe largest number of spectra with 143 epochs. The 35epochs of Lick spectra were obtained by several groupsof observers who used slightly different setups and cali-brations. The Kast spectrograph (Miller & Stone 1993)at Lick Observatory has red-side and blue-side cameras,but since the red-side setup was very different for eachgroup, we present only the blue-side data here. Asiago,APO, and WIRO contributed 21, 13, and 6 epochs ofspectra, respectively. Our analysis focuses primarily onthe MDM dataset for homogeneity.

Data-reduction procedures included bias subtraction,flat fielding, and cosmic ray removal using the L.A. Cos-mic routine (van Dokkum 2001). The one-dimensionalspectra were extracted from a 15′′-wide region centeredon the AGN and with consistent background sky aper-tures for all observations. We used optimally-weightedextractions for the stellar spectra (Horne 1986) but un-weighted extractions for the AGN spectra. This is be-

Page 5: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

5

TABLE 1Instrument characteristics and data-reduction parameters for all telescopes

Telescope Instrument Number of Median Wavelength Wavelength Pixel Median [O III]Epochs Seeing Dispersion Coverage Scale SNR Fvar

(′′) (A pixel−1) (A) (′′ pixel−1) (%)

MDM Boller & Chivens CCD Spectrograph 143 1.7 1.25 4225−5775 0.75 118 0.62Lick Kast Double Spectrograph 35 1.5 1.02 3460−5500 0.43 194 0.32Asiago Boller & Chivens CCD Spectrograph 21 4.0 1.00 3250−7920 1.00 160 0.27APO Dual Imaging Spectrograph 13 1.4 1.00 4180−5400 0.41 160 0.28WIRO WIRO Long Slit Spectrograph 6 2.1 0.74 5599−4399 0.52 217 0.47

Note. — The wavelength coverage for Lick refers to only the Kast blue-side camera. The SNR value refers to the median SNR perpixel over the rest wavelength range 5070−5130 A. The [O III] Fvar is the amount of residual variations in the [O III] light curve afterspectral scaling and gives an indication of the flux scaling accuracy.

cause the optimal extraction method requires the spatialprofile of the target to be a smooth function of wave-length, and tends to truncate the peaks of strong emis-sion lines such as [O III] that have different spatial ex-tents from the surrounding continuum.

The data were wavelength calibrated using night skylines and flux calibrated using standard stars. Our mostfrequently used flux standard stars were Feige 34, BD332642, and HZ 44. For nights when multiple expo-sures were taken, we aligned the flux-calibrated one-dimensional spectra by applying small wavelength shiftsto each spectrum before combining them. We do not ex-pect significant differential atmospheric refraction (Fil-ippenko 1982) because of the large slit width use for ourobservations.

For the MDM data, the first 133 epochs were flux cal-ibrated using Feige 34, while the last 10 epochs, takenfrom 2014 June 20 to 2014 June 30, were flux calibratedwith BD 332642. This caused spurious changes in theshape of some emission-line features, so we use only thefirst 133 MDM epochs for our present analysis.

2.1. Spectral Flux Calibrations

To place the instrumental fluxes on an absolute fluxscale, we measured the narrow [O III] λ5007 line fluxfrom spectra taken under photometric conditions andscaled all other nightly spectra to have the same [O III]flux. There were 21 epochs identified as having been ob-served under photometric conditions by the MDM ob-servers. We determined the flux of the [O III] line(λobserved = 5093 A) by first subtracting a linear fitto continuum windows on either side of the line, thenintegrating over a fixed wavelength range. We usedthe rest-frame wavelength ranges 4976.5−4948.0 A and5027.7−5031.6 A to fit the continuum and integratedover the range 4980.5−5026.7 A for the line flux. The 2σoutliers from this set of [O III] flux measurements werediscarded, the mean was recomputed, and this processwas repeated until there were no more 2σ outliers, whichresulted in a total of 16 final photometric spectra. Themean spectrum of these 16 epochs has an [O III] λ5007line flux of (5.01 ± 0.11) × 10−13 erg s−1 cm−2, whichrepresents our best estimate of the true [O III] flux forNGC 5548 during this campaign and is not expected tovary over a six-month period. For comparison, Petersonet al. (2013) found the [O III] flux in NGC 5548 to be(4.77 ± 0.14) × 10−13 erg s−1 cm−2 in their 2012 moni-toring campaign, and the difference is within the rangeof total [O III] variability observed for NGC 5548 over

the course of 21 years (see Peterson et al. 2013).In addition to the intrinsic variability of the AGN,

many other factors contribute to nightly variations inthe spectra. These include changes in transparency dueto clouds, changes in seeing conditions, inconsistent in-strument focus, and miscentering of the AGN in the slitduring observations. We used the flux scaling methoddescribed by van Groningen & Wanders (1992) to alignthe nightly spectra and place them on a consistent fluxscale. For each spectrum in the dataset, the algorithmlooks for a combination of wavelength shift, multiplica-tive scale factor, and Gaussian kernel convolution thatminimizes the residual between each individual spectrumand a reference spectrum over a region containing thenarrow [O III] line.

We constructed a separate reference spectrum for eachtelescope by averaging the highest SNR spectra in eachdataset, then broadened the reference spectrum so thatthe [O III] line width matches the broadest [O III] linewidth in the dataset. This extra broadening of the refer-ence spectrum helps to reduce the [O III] residuals fromspectral scaling (Fausnaugh 2016). We then scaled eachspectrum to have the same [O III] flux as the photomet-rically calibrated mean MDM spectrum. This brings allspectra to a common flux scale after spectral scaling.

To assess the accuracy of spectral scaling, we estimatedthe intrinsic fractional variability of the residual [O III]λ5007 light curve after correcting for random measure-ment errors,

Fvar =

σ2 − 〈δ2〉

〈f〉, (3)

where σ2 is the [O III] flux variance, 〈δ2〉 is the mean-square value of the measurement uncertainties deter-mined from the nightly error spectra produced by thedata reduction pipeline, and 〈f〉 is the unweighted meanflux. The Fvar for the [O III] λ5007 light curve gives agood estimate of the residual flux-scaling errors (Barth &Bentz 2016), and the value for each telescope is listed inthe last column of Table 1. We found Fvar to be between0.27% and 0.62% for all telescopes, which means thereis an additional scatter of less than 1% in the [O III]light curve above the measurement errors. These Fvar

values are consistent with or better than the best val-ues typically obtained in ground-based campaigns. Forexample, Barth et al. (2015) found Fvar values rangingfrom 0.5% to 3.3% for individual AGNs in the 2011 LickAGN Monitoring Project.

Page 6: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

6

Fig. 2.— The mean and excess rms (Eq. 4) spectra from theMDM dataset are shown in black, and the rms spectrum with theAGN and stellar continuum removed is shown in red (see Section3.1).

Figure 2 shows (in black) the mean and root-mean-square (rms) residual spectra for the MDM dataset. Therms spectrum indicates the degree of variability at eachwavelength over the course of the campaign. Both thebroad Hβ and He II λ4686 emission lines exhibit strongvariations, and the Hβ rms profile appears to have multi-ple peaks. Traditionally, the rms spectrum is constructedsuch that the value at each wavelength is taken to be thestandard deviation of fluxes from all epochs, but this doesnot take into account Poisson or detector noise, whichmay bias the rms profile by a small amount (Barth et al.2015). Park et al. (2012b) suggest using the SNR foreach spectrum as the weight for that spectrum in calcu-lating the rms, or using a maximum-likelihood method toobtain the rms. We adopt a simpler approach that usesthe excess variance as a way to exclude variations thatare not intrinsic to the AGN. This “excess rms” value ateach wavelength is defined as

e-rmsλ =

1

N − 1

N∑

i=1

[(Fλ,i − 〈Fλ〉)2 − δ2λ,i], (4)

where N is the total number of spectra in the dataset,〈Fλ〉 is the mean flux at each wavelength, and Fλ,i andδλ,i are the wavelength-specific fluxes and associatedmeasurement uncertainties from individual epochs, re-spectively. This method estimates the degree of vari-ability above what is expected given the measurementuncertainties and pixel-to-pixel noise.

3. SPECTROSCOPIC FLUX MEASUREMENTS

The 5100 A continuum flux density was determined byaveraging the flux over the rest-frame wavelength range

5070−5130A. The Hβ line fluxes were measured fromthe scaled spectra using the same method as for [O III]λ5007, where we subtracted a linear fit to the surround-ing continuum (wavelength windows 4483.0−4542.0 Aand 5033.5−5092.5 A) and integrated across the line pro-file (4748.4−4945.1 A). The uncertainty in each measure-ment is a combination of Poisson noise and residuals fromspectral scaling. We computed the spectral scaling un-certainty by multiplying each flux measurement by the[O III] Fvar value for that dataset, then adding this valuein quadrature to the Poisson noise to obtain the finalflux uncertainty for each measurement. There is an ad-ditional source of spectral scaling uncertainty from slightdifferences in the overall spectral shape from night tonight. This effect is likely small for Hβ because it is veryclose to the [O III] λ5007 line that anchors the spectralscaling.

Spectrophotometric calibrations of the reference spec-tra, as described in the previous section, converted allinstrumental fluxes to absolute fluxes, which means thatmeasurements from all telescopes should now be on thesame flux scale. However, light curves from differentobserving sites may be offset from each other owing toaperture effects (Peterson et al. 1995, 1999). While ourobservations were standardized to have the same 5′′ ×15′′ aperture size, significant differences in image qualitybetween observing sites could still cause flux offsets.

To intercalibrate the Hβ light curves, we used datapoints from each non-MDM telescope (FHβ,t) thatare nearly contemporaneous with MDM observations(FHβ,MDM) and performed a least-squares fit to the equa-tion

FHβ,MDM = φFHβ,t (5)

to find the scale factor φ that puts each line light curveon the same flux scale as the MDM data. For the con-tinuum intercalibration, we also include an additive shiftG to account for the differences in the host-galaxy fluxadmitted by different apertures:

F5100,MDM = φF5100,t + G. (6)

The scale factors for the Lick, Asiago, APO, and WIROlight curves are φ = [0.961, 0.963, 1.037, 0.918], and theshift constants are G =[−0.155, −0.640, −0.041, 0.024]

in units of 10−15 erg s−1 cm−2 A−1

. The combined con-tinuum and Hβ light curves are shown in Figure 3 (THJD= HJD − 2,450,000), and the 5100 A continuum and Hβfluxes are listed in Table 2.

We attempted to measure the He II λ4686 flux fromthe nightly spectra. However, this line is very weak andalso heavily blended with the broad Hβ, as shown inFigure 2. Thus, we were unable to obtain a He II lightcurve using the linear interpolation method to removethe continuum.

3.1. Spectral Decomposition

To more accurately remove the continuum underly-ing the emission lines and to deblend the broad emis-sion features from each other, we employed the spec-tral decomposition algorithm described by Barth et al.(2015). The components fitted in this procedure include

Page 7: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

7

Fig. 3.— Continuum (10−15 erg s−1 cm−2 A−1

) and Hβ (10−15 erg s−1 cm−2) light curves (THJD = HJD − 2,450,000). The Lick,APO, Asiago, and WIRO light curves were scaled and shifted to match the MDM light curve, which has the longest temporal coverageand highest sampling cadence. The plotted uncertainties include Poisson noise and the normalized excess variance of the [O III] light curve(Section 2.1.)

TABLE 2Flux measurements for continuum and emission lines.

HJD−2,450,000 Telescope F5100 FHβ FHβ,SD FHe II,SD

6663.00 MDM 10.766 ± 0.075 726.012 ± 4.985 710.187 ± 4.897 21.720 ± 2.6066663.65 Asiago 10.921 ± 0.040 741.771 ± 3.586 — —6664.03 MDM 11.154 ± 0.075 732.511 ± 5.156 715.057 ± 5.061 28.154 ± 3.2226665.02 MDM 10.788 ± 0.075 724.537 ± 4.946 709.473 ± 4.860 25.135 ± 2.4516667.02 MDM 10.872 ± 0.076 735.001 ± 5.393 711.347 ± 5.288 37.008 ± 3.964

Note. — The 5100 A continuum flux density (10−15 erg s−1 cm−2 A−1

) includes contributionsfrom both the AGN and the host galaxy. The Hβ and Hβ SD fluxes were obtained using a linearcontinuum model and the spectral decomposition method, respectively. The He II flux is based on thespectral decomposition model. All emission-line fluxes are in units of 10−15 erg s−1 cm−2 and includecontributions from both broad- and narrow-line components. The full table is available in the onlineversion.

narrow [O III], broad and narrow Hβ, broad and nar-row He II, Fe II emission blends, the stellar continuum,and the AGN continuum. The host-galaxy starlight wasmodeled with an 11 Gyr, solar metallicity, single-burstspectrum from Bruzual & Charlot (2003). For the Fe II

model component, we tested three different templatesfrom Boroson & Green (1992), Veron-Cetty et al. (2004),and Kovacevic et al. (2010). The Fe II templates werebroadened by convolution with a Gaussian kernel in ve-locity. The free fit parameters for Fe II include the ve-locity shift relative to broad Hβ, the broadening ker-nel width, and the flux normalization of the broadenedtemplate spectrum. The Boroson & Green (1992) andVeron-Cetty et al. (2004) templates are monolithic andrequire only one flux normalization parameter, whereasthe Kovacevic et al. (2010) template has five componentsthat can vary independently in flux. The Kovacevic et al.(2010) template achieves the best fit to the nightly spec-

tra, presumably a result of the larger number of free fitparameters due to the multi-component Kovacevic et al.(2010) template.

We made several modifications to the spectral fittingprocedures used by Barth et al. (2015). First, becauseof the complex line profiles, we used sixth-order Gauss-Hermite functions (van der Marel & Franx 1993) to fitthe broad and narrow Hβ and narrow [O III] lines insteadof fourth-order functions. Second, there is significant de-generacy between the weak Fe II blend and the contin-uum flux in the nightly fits. Since the Fe II fit is poorlyconstrained and sometimes varied drastically from nightto night, the continuum model flux also varied signifi-cantly as a result, which in turn introduced noise to thebroad Hβ fit component. To address this issue, we con-strained the Fe II flux to lie within 10% of the value fromthe fit to the mean spectrum (Barth et al. 2013). We alsofixed the Fe II redshift to that of the mean spectrum and

Page 8: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

8

Fig. 4.— Top: Spectral decomposition components of the meanMDM spectrum. The red spectrum is the sum of all model compo-nents and traces the data (black spectrum) well over most of thespectral range. Bottom: Residuals from the full model fit.

constrained the Fe II broadening kernel to be within 5%of its value from the mean spectrum fit. The He I λ4922and λ5016 lines are very weak and are heavily blendedwith broad Hβ, making it impossible to constrain theirfit parameters. We therefore do not fit for these compo-nents in our model.

The broad He II λ4686 component has very low am-plitude compared to the other fit components and it isblended with the blue wing of broad Hβ. It is also highlyvariable, as demonstrated by the broad bump in the rmsspectrum. This made it difficult to fit the He II broad-line profile accurately, and the width varied significantlyfrom night to night when fitted as a free parameter. Sincethe He II λ1640 and λ4868 lines are expected to form un-der the same physical conditions and should thus havesimilar widths, we used fits to the λ1640 line in concur-rent HST spectra to constrain the λ4686 line width.

The He II λ1640 line was modeled with five Gaussiancomponents (De Rosa et al., in prep.), and we took thethree broadest components to represent the broad He II

λ1640 line profile. For each MDM spectrum, the He II

λ4686 broad-line full width at half-maximum intensity(FWHM) was allowed to vary within 3 A of the He II

λ1640 FWHM measured from the closest HST epoch.The first 23 epochs from the MDM campaign do nothave corresponding HST spectra, so for each of these“pre-HST” epochs, we found the three epochs from laterin the campaign with the closest matching 5100 A con-tinuum flux density. We then used the weighted mean ofthe broad He II λ1640 widths from these three nights asthe width constraint for the pre-HST epoch, where theweight were determined by how closely the 5100 A fluxesof the later epochs matched that of the pre-HST epoch.The He II λ1640 line width was highly variable duringthe HST campaign, and the model FWHM widths usedto constrain the spectral decomposition have a mean of48 A, with a minimum of 28 A and maximum of 59 A.

We applied spectral decomposition to the data from alltelescopes, but since the MDM dataset is the largest and

has the highest data quality and consistency, we use thisdataset for all subsequent analysis. Figure 4 shows thefit components for the mean MDM spectrum, where theblack spectrum is the data and the red spectrum is thesum of all the model components. The model does notfit the detailed structure of the broad Hβ line well, es-pecially in the line core. To prevent this from impactingour measured Hβ fluxes, we subtracted all the other well-modeled fit components except the broad and narrowHβ components from the full spectrum, then obtainedthe Hβ line flux by integrating over the same wavelengthrange used to measure the flux without spectral decom-position. The He II λ4686 flux was taken to be the totalflux in the broad- and narrow-line models for each night.The narrow Hβ and He II λ4686 line fluxes from fits tothe mean spectrum are 48.4 × 10−15 erg s−1 cm−2 and8.5× 10−15 erg s−1 cm−2 (respectively), with uncertain-ties of ∼ 2% from the overall photometric scale of thedata. The ratio of the narrow Hβ flux to the [O III] λ5007flux is FHβ/F[O III] = 0.099 ± 0.002, which is in goodagreement with the value of FHβ/F[O III] = 0.110±0.010found by Peterson et al. (2004).

The red spectrum in Figure 2 shows the rms of theMDM spectra after subtracting the AGN continuum andstellar continuum models from individual spectra so thatonly the emission-line components remain. This rmsspectrum is expected to be a more accurate represen-tation of the emission-line variability than the rms of thefull spectra (Barth et al. 2015). We show the rms hereand not the excess rms defined by Equation 4 because,for parts of the spectra dominated by continuum emis-sion, the continuum-subtracted flux could be lower thanthe total flux uncertainties and the e-rms would be un-defined. Thus, we use the excess rms only for the fullspectrum and not for individual fit components.

Panels (a) and (c) of Figure 5 show the Hβ mean andrms fluxes as a function of line-of-sight velocity (vr) afterspectral decomposition, and panel (b) shows the differ-ence between the T1 and T2 mean fluxes. The rms fluxhas a statistical uncertainty of ∼ 12% and the [O III]residuals are much lower compared to the case with nospectral decomposition (Fig. 2). The rms profile still hasjagged features, which likely reflect real variability acrossthe broad emission line.

Figure 6 shows the 1158 A UV continuum light curvefrom Paper I, the MDM optical 5100 A continuum lightcurve, the V -band photometric light curve from PaperIII, and the MDM Hβ and He II λ4686 emission-line lightcurves. The He II light curve reaches a flat-bottomedminimum near THJD = 6720. This is because the He II

flux includes contributions from the broad- and narrow-line components, so when the broad-line flux is near zero,the total He II line flux stays at a minimum value equal tothe narrow-line flux. Light-curve statistics that quantifythe variability of NGC 5548 during the monitoring periodare given in Table 3. Fvar is as defined in Equation 3 andRmax is the ratio between the maximum and minimumfluxes.

3.2. Host-Galaxy Flux Removal

We measured the host-galaxy contribution to the con-tinuum using an “AGN-free” image of NGC 5548 gen-erated by Bentz et al. (2013) after performing two-

Page 9: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

9

TABLE 3Statistics for HST and MDM Light Curves

Emission Component Epochs Mean Flux rms Flux Fvar Rmax

Fλ (1158 A) 171 43.48 ± 0.86 11.14 ± 1.21 0.255 4.07 ± 0.18Fλ (5100 A) 133 11.96 ± 0.07 0.80 ± 0.09 0.066 1.33 ± 0.01Hβ 133 738.49 ± 2.40 28.29 ± 3.38 0.038 1.22 ± 0.01He II λ4686 133 78.71 ± 2.95 35.14 ± 4.17 0.444 7.48 ± 0.91

Fλ (1158 A, T1) 51 35.85 ± 1.79 12.61 ± 2.54 0.351 3.31 ± 0.15Fλ (1158 A, T2) 120 46.72 ± 0.80 8.66 ± 1.13 0.184 2.36 ± 0.08Fλ (5100 A, T1) 67 11.31 ± 0.08 0.67 ± 0.12 0.059 1.27 ± 0.01Fλ (5100 A, T2) 67 12.51 ± 0.04 0.37 ± 0.06 0.029 1.13 ± 0.01Hβ (T1) 67 725.10 ± 3.80 30.47 ± 5.36 0.041 1.20 ± 0.01Hβ (T2) 67 750.14 ± 2.38 20.11 ± 3.33 0.026 1.13 ± 0.01He II λ4686 (T1) 67 65.84 ± 4.16 33.61 ± 5.89 0.507 5.95 ± 0.73He II λ4686 (T2) 67 89.90 ± 3.75 32.71 ± 5.34 0.362 4.07 ± 0.37

Note. — Continuum flux densities are in units of 10−15 erg s−1 cm−2 A−1

andemission-line fluxes are in units of 10−15 erg s−1 cm−2. T1 and T2 denote the first andsecond halves of the campaign (respectively) divided at THJD = 6747.

dimensional surface brightness decomposition on HSTimages of the galaxy. We found that the amount ofstarlight expected through a 5′′× 15′′ aperture with aslit position angle of 0◦ is F5100,gal = (4.52 ± 0.45) ×

10−15 erg s−1 cm−2 A−1

. Subtracting this from the meancontinuum flux density of F5100 = (11.96 ± 0.07) ×

10−15 erg s−1 cm−2 A−1

gives a mean AGN flux of

F5100,AGN = (7.44 ± 0.50) × 10−15 erg s−1 cm−2 A−1

,which is consistent with the value of F5100,AGN =

(7.82 ± 0.02)× 10−15 erg s−1 cm−2 A−1

measured fromthe power-law component of the spectral decompositionfor the mean MDM spectrum.

3.3. Anomalous Emission-Line Light-Curve Behavior

RM analyses typically assume that the emission-linelight curve responds linearly to continuum variations andis a lagged, scaled, and smoothed version of the contin-uum light curve. However, this does not appear to be thecase for a portion of our campaign. As described in PaperI and Paper IV, there are significant differences betweenthe UV continuum and emission-line light curves afterTHJD = 6780. The continuum flux increased while theemission-line fluxes either decreased or remained roughlyconstant in a suppressed state for the remainder of thecampaign.

The Hβ emission line shows similar behavior to thatof Lyα and C IV, in that there is a marked differencein the line response between the first and second halvesof the campaign. This “decorrelation” phenomenon isillustrated in Figure 7(a), where the Hβ and UV contin-uum light curves trace each other well in the first half ofthe campaign, but the continuum flux continues to trendupward beyond THJD = 6740, while the Hβ flux beginsto fall. For the remainder of the campaign, the Hβ fluxremains in a suppressed state and the light curve doesnot follow the continuum light curve well in that promi-nent features in the continuum light curves (e.g. THJD≈ 6770, 6785) are not present in the emission-line lightcurve. The He II λ4686 light curve behaves similarlyand decorrelates from the continuum at around THJD= 6760 (Figure 7b). This phenomenon is also appar-ent when comparing the Hβ light curve to the 5100 A

continuum, as shown in Figure 7(c). Though there aresmall differences between the light curves in T1 (suchas around THJD 6685), the overall correlation in T1 issignificantly better than in T2.

Owing to this change in the emission-line response, wefollowed the procedures presented in Paper I for deter-mining the UV emission-line lags, and divided the 5100 Acontinuum and optical emission-line light curves into twosubsets, labeled T1 and T2, to examine the lag of eachsegment separately. The subsets are divided at THJD =6747 and each has 67 epochs. The 1158 A and V -bandcontinuum light curves were also separated into two seg-ments at THJD = 6747. Note that this is a differentdividing epoch from the one used in Paper I. The char-acteristics of the half-campaign light curves are given inthe bottom portion of Table 3. Figure 5 shows the MDMmean and rms spectra for T1 and T2 in gray and orange,respectively. While the three mean spectra look almostidentical, the T1 and T2 rms spectra are significantlydifferent, which indicates changes in the amount of vari-ability between the two campaign halves.

4. DATA ANALYSIS

In the following sections, we examine properties of theBLR by measuring the emission-line responses to contin-uum variations. We also discuss the anomalous behav-ior of the emission-line light curves observed during thiscampaign and BH mass measurements using this dataset.

4.1. Emission-Line Lags

We measured the Hβ and He II λ4686 lags relativeto both the 5100 A continuum and the 1158 A contin-uum. All light curves were detrended by subtracting alinear least-squares fit to the data to remove long-termtrends that may bias lag calculations (Welsh 1999). Inthis case, we found very weak trends for all the lightcurves, and detrending has a very small effect (∼ 0.01days) on the measured lags. We computed the cross-correlation coefficient r for lags between −20 and 40 daysin increments of 0.25 days using the interpolated cross-correlation function (ICCF; White & Peterson 1994).Two lag estimates were made for each light-curve pair—the value corresponding to rmax (τpeak) and the centroidof all values with r > 0.8rmax (τcen). Estimates for the

Page 10: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

10

Fig. 5.— The MDM mean spectrum (a), difference between theT1 (THJD < 6747) and T2 (THJD > 6747) mean spectra (b),and the rms spectra (c−e) for Hβ after subtracting all other fitcomponents from spectral decomposition. The colors are for thefull campaign (black), T1 (gray), and T2 (orange, see Section 3.3).Zero velocity is determined by the peak of the narrow Hβ line in themean spectrum, and the gray bands indicate regions contaminatedby [O III] residuals. The rms spectra have statistical uncertaintiesof ∼ 12%.

final τpeak and τcen values and their uncertainties wereobtained using Monte Carlo bootstrapping analysis (Pe-terson et al. 2004), where many realizations of the con-tinuum and emission-line light curves were created byrandomly choosing n data points with replacement fromthe observed light curves, where n is the total number ofpoints in the dataset. If a data point is picked m times,then the uncertainty on that point is decreased by a fac-tor of m1/2. Each value is then varied by a random Gaus-sian deviate scaled by the measured flux uncertainty. Weconstructed 103 realizations of each light curve and com-puted the cross-correlation function (CCF) for each pairof line and continuum light-curve realizations to create adistribution of τpeak and τcen values. The median valuefrom each distribution and the central 68% interval are

then taken to be the final lag and its uncertainty.Table 4 lists the ICCF lags for Hβ and He II λ4686

measured against the 1158 A, 5100 A, and V -band con-tinua. The lag between the 5100 A and 1158 A con-tinua is also given. For comparison with Paper I, whichpresents the UV emission-line lags against the 1367 Acontinuum, we also include Hβ and He II lags measuredagainst this continuum. Distributions of τcen values fromthe Monte Carlo bootstrap analysis using the 1158 A con-tinuum are shown in the top panels of Figure 8.

We also computed Hβ and He II λ4686 lags for timeperiods T1 and T2, and found the T1 lags to be consis-tently longer by about 2σ. For Hβ, the T1 and T2 lagsbracket the full-campaign lag, while for He II λ4686, thefull-campaign lag is shorter than both T1 and T2 lags.For comparison, Paper I found the Lyα, Si IV, C IV, andHe II λ1640 lags to be longer for T2 than for T1, whichis the opposite of what we find for Hβ.

To illustrate the effects of spectral decomposition, wealso include in Table 4 the ICCF lags for the Hβ lightcurve where the fluxes were measured using the straight-line continuum-subtraction method and without spectraldecomposition. We calculated lags for both the MDM-only and the multi-site Hβ light curves, which were trun-cated at THJD = 6828.75 to exclude the last 10 epochsof MDM data (see Section 2). The lags measured withand without using spectral decomposition are consistentto within 1σ.

In addition to the ICCF method, we computed theemission-line lags using the JAVELIN suite of Pythoncodes (Zu et al. 2011). We used JAVELIN to linearly de-trend the light curves and model the AGN continuumvariability as a damped random walk process (DRW;Kelly et al. 2009; Zu et al. 2013). JAVELIN explic-itly models the emission-line light curves as smoothed,scaled, and lagged versions of the continuum light curve.Since the decorrelation of the line and continuum lightcurves during the latter half of the campaign clearly vi-olates these assumptions, it is of interest to examine theconsequences for the JAVELIN models. For these mod-els, we simultaneously fit the Hβ and He II light curvesusing either the 1158 A, 5100 A, or V -band continuumlight curve. The AGN STORM light curves are too shortto accurately determine the DRW damping timescale, sothis value was fixed to τDRW ≈ 164 days as derived by Zuet al. (2011) from fits to the 13-year light curve of NGC5548 (Peterson et al. 2002). The precise value of τDRW

is not critical for the algorithm to work provided that itis approximately correct.

The top three panels of Figure 9 show the results ofusing JAVELIN directly on the observed data. Despitethe long DRW time scale, the light-curve models showrapid fluctuations. The algorithm tries to match the sup-pressed flux in the line light curve to the continuum lightcurve, and because of the small uncertainties, it stronglyprefers lag times for which the continuum and line lightcurves have minimal temporal overlap. This caused theposterior lag distribution to have multiple narrow peakscorresponding to lags that best desynchronize the lightcurves.

We can attempt to compensate for this problem by in-creasing the uncertainties to encompass the amplitudeof the decorrelation. This requires scaling up the full-

Page 11: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

11

Fig. 6.— Left: Light curves for the UV 1158 A continuum, optical 5100 A continuum, V -band continuum, Hβ, and He II λ4686. The

continuum light curves are in units of 10−15 erg s−1 cm−2 A−1

and the line light curves are in units of 10−15 erg s−1 cm−2. The Hβand He II fluxes include contributions from both broad and narrow line components. The solid green vertical line indicates where theemission-line light curves were truncated for the lag analysis (see text) and the dashed black vertical line shows the division between theT1 and T2 periods. Right: Cross-correlation functions for each light curve measured against the 1158 A continuum. The top right panelshows the auto-correlation of the 1158 A light curve. The black, gray, and orange solid lines represent the CCFs for the full campaign, T1,and T2 (respectively), and the dotted vertical lines denote τcen for the full campaign.

campaign light-curve uncertainties by factors of 5 and 3for the continuum and line light curves, respectively. Thelower panels in Figure 9 show the JAVELIN results fromfitting to the light curves with scaled errors, where thebroader uncertainties allow the algorithm to constructsmooth light-curve models. Since there is more statis-tical weight from fitting both lines simultaneously, themodels track the line light curves best and show a smoothsystematic offset for the continuum where the line andcontinuum light curves are decorrelated. When comput-ing the half-campaign lags, JAVELIN favors a smaller lineflux scale factor for T2 to account for the suppressedline fluxes, which begins near the epoch separating T1and T2. We therefore did not need to scale the flux er-rors by as much as for the full campaign to account for

the decorrelation, and used an error scaling factor of 3for both continuum and line light curves. The result-ing JAVELIN lags, shown in Table 4, are consistent withthose measured using the ICCF method. Similar to theICCF lags, the JAVELIN lags for the full-campaign lightcurves are also shorter than those for T1 and T2. How-ever, since the JAVELIN assumptions of the relationshipbetween the line and continuum light curves are not validfor this campaign and the flux errors were scaled for thesole purpose of producing convergent solutions, we donot use the JAVELIN lags for subsequent analysis.

We examined the velocity-resolved emission-line re-sponse by dividing the Hβ line profile into bins with ve-locity width of 500 km s−1 and set zero velocity using thepeak of the narrow Hβ component in the mean spectrum.

Page 12: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

12

TABLE 4Rest-Frame Emission-Line Lags

Light Curves τpeak τcen τcen,T1 τcen,T2 τJAVELIN τJAVELIN,T1 τJAVELIN,T2

Hβ vs. Fλ(1158 A) 6.14+0.74−0.98 6.23+0.39

−0.44 7.62+0.49−0.49 5.99+0.71

−0.75 6.56+0.48−0.49 6.91+0.64

−0.63 7.42+0.97−1.07

Hβ vs. Fλ(1367 A) 5.90+0.25−0.74 5.89+0.37

−0.37 7.24+0.49−0.48 5.99+0.76

−0.82 6.12+0.46−0.47 6.52+0.60

−0.57 7.11+1.03−1.06

Hβ vs. Fλ(5100 A) 4.42+0.98−0.25 4.17+0.36

−0.36 4.99+0.40−0.47 3.10+0.77

−0.80 3.84+0.57−0.59 5.15+0.68

−0.69 4.78+1.13−1.17

Hβ vs. V band 3.93+0.98−0.98 3.79+0.37

−0.34 3.82+0.57−0.47 4.13+0.55

−0.58 3.54+0.45−0.46 4.89+0.66

−0.71 4.05+0.93−0.78

He II vs. Fλ(1158 A) 2.46+0.49−0.25 2.69+0.24

−0.25 3.71+0.39−0.38 3.19+0.36

−0.35 2.65+0.27−0.27 3.27+0.35

−0.35 2.99+0.25−0.26

He II vs. Fλ(1367 A) 2.21+0.25−0.25 2.45+0.25

−0.24 3.43+0.36−0.43 3.16+0.29

−0.33 2.41+0.25−0.26 3.04+0.35

−0.36 2.79+0.25−0.25

He II vs. Fλ(5100 A) 0.49+0.25−0.25 0.79+0.35

−0.34 1.21+0.28−0.36 0.85+0.36

−0.36 0.16+0.37−0.37 1.13+0.51

−0.48 0.85+0.38−0.35

He II vs. V band 0.49+0.25−0.49 0.50+0.34

−0.26 0.40+0.43−0.39 1.46+0.34

−0.27 0.44+0.23−0.23 0.63+0.37

−0.36 0.91+0.23−0.21

Fλ(5100 A) vs. Fλ(1158 A) 1.97+0.25−0.49 2.23+0.31

−0.26 2.55+0.28−0.33 2.77+0.41

−0.45

Hβ full, MDM vs. Fλ(1158 A) 5.90+0.74−0.49 6.26+0.38

−0.37

Hβ full, all sites vs. Fλ(1158 A) 5.65+0.74−0.25 6.49+0.37

−0.37

Note. — Rest-frame Hβ and He II λ4686 lags (days) for the full campaign and for the T1 (THJD < 6747) and T2(THJD > 6747) subsets, measured using both ICCF and JAVELIN. The last two lines show the Hβ lags measured usingthe light curve derived without spectral decomposition up to THJD = 6828.75.

Fig. 7.— 1158 A and 5100 A continuum light curves (black)compared with scaled and shifted emission-line light curves (blue).The vertical line at THJD = 6747 indicates the epoch separatingthe T1 and T2 segments. In each of the panels, the emission-linelight curve closely tracks the continuum light curve in T1, butappears to correlate less closely with the continuum variations inT2.

We constructed light curves for each velocity bin sepa-rately, and Figure 11 shows the light curves for 1500 kms−1 velocity bins across the Hβ line profile (black), withthe velocity at the center of each bin shown in the top leftof each panel. There were six epochs of spectra† that pro-duced outlying Hβ fluxes and significantly higher-than-average flux uncertainties for individual velocity bins, sowe removed them from the velocity-resolved light curvesin order to improve the lag measurements.

We determined the ICCF lag for each of these binned

†THJD = [6689.0, 6699.0, 6757.8, 6758.9, 6796.8, 6797.8]

Fig. 8.— τcen (top) and JAVELIN (bottom) lag probability distri-butions for Hβ (left) and He II (right) measured against the 1158 Acontinuum. Black solid lines are for the full campaign, gray dashedlines are for T1, and orange dot-dash lines are for T2.

light curves with respect to both the UV and optical con-tinua, and show these lags as a function of line-of-sightvelocity in the top left panel of Figure 10. The secondleft panel shows the velocity-resolved Hβ–UV lag for T1and T2, and the maximum cross-correlation coefficients(rmax) are shown in the right-side panels. The bottompanel of Figure 10 shows the MDM full-campaign meanspectrum for reference, and Table 5 lists the velocity-resolved lags. Emission-line variations have lower ampli-tudes during T2 compared with T1, which led to poorly-constrained ICCF lags with large uncertainties for ve-locity bins with low line flux. We therefore reduced theupper limit of the CCF lag range from 40 days to 20 dayswhen computing the T2 lags.

The velocity-resolved Hβ–UV lag for the full campaignis shortest (τcen ≈ 2 days) in the line wings where vr ≈±7000 km s−1. The lag increases as vr approaches zero

Page 13: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

13

Fig. 9.— JAVELIN light curves from simultaneously modeling theHβ and He II λ4686 emission lines with the UV 1158 A continuum.The data points are measured from observations, the black solidlines are the weighted means of the model light curves consistentwith the data, and the thickness of the shaded regions indicatesthe 1σ spread of those light curves. The top three panels show theobserved data and JAVELIN model light curves without any errorscaling, and the bottom three panels show the data and light-curvemodels with scaled uncertainties.

from both sides of the lag profile, reaches local maximaof τcen ≈ 10 days at about vr ≈ ±3000 km s−1, and thensteadily decreases until it reaches a local minimum ofτcen ≈ 4 days near the line profile center. The lag profilemeasured against the optical continuum has a similarshape, but with all lags ∼ 2–3 days shorter, as we wouldexpect from the ∼ 2-day lag between the two continua(Table 4). A similar double-peaked lag profile is alsoobserved for Lyα (see Paper I). The T1 lag profile closelyresembles that of the full campaign, but the T2 lag profileshows a slightly different structure. The bins where theT1 and T2 lags are most discrepant are also where theT2 light curves are least correlated with the continuum(rmax < 0.4). However, even excluding these outliers,there are still discernible differences between the T1 andT2 lag profiles. Additionally, the T2 rmax values arelower than those of T1 in every velocity bin, which clearlydemonstrates that the line and continuum light curvesare less correlated in T2 than in T1.

The shape of the velocity-resolved lag profile can pro-vide qualitative information about the kinematics of the

line-emitting gas (e.g., Kollatschny 2003; Bentz et al.2009b; Denney et al. 2010; Barth et al. 2011; Du et al.2016a). In simple models of the BLR (Ulrich et al. 1984;Gaskell 1988; Welsh & Horne 1991; Horne et al. 2004;Goad et al. 2012; Gaskell & Goosmann 2013; Grier et al.2013), pure infall motion would lead to longer lags onthe blue side of the line profile, and for outflow, the mostredshifted gas would have the longest lag. For gas inKeplerian orbits, the shortest lags would be in the linewings, since gas with higher vr is closer to the centralblack hole. Gas with very low vr could have a widerange of lags, and a spherical or flat disk distribution ofBLR clouds in Keplerian motion could lead to a double-peaked velocity-resolved lag profile if the ionizing sourceis emitting anisotropically (Welsh & Horne 1991; Goad& Wanders 1996; Horne et al. 2004).

Previous studies of the UV and optical lines in NGC5548 have inferred either Keplerian orbits (Horne et al.1991; Wanders et al. 1995; Denney et al. 2009; Bentzet al. 2010b) or infalling motion (Crenshaw & Blackwell1990; Done & Krolik 1996; Welsh et al. 2007; Pancoastet al. 2014; Gaskell & Goosmann 2016) for the BLR gas.From our data, the shape of the Hβ velocity-resolvedlag profile suggests a BLR dominated by Keplerian mo-tion. The discrepancy between the T1 and T2 lag profilesmay suggest a change in the distribution or dynamics ofthe BLR gas, though such changes typically occur ontimescales much longer than our campaign. More de-tailed interpretation requires comparison with transferfunctions generated for various dynamical models of theBLR, which will be the subject of future work in thisseries (Pancoast et al., in prep.).

Figure 11 also shows modified versions of the 1158 Acontinuum light curve in red. These light curves wereshifted in time by the average T1 lag of the bins incorpo-rated in each Hβ light curve, which are shown in the bot-tom right of each panel. The fluxes were roughly scaledand shifted to match the first half of the Hβ light curves.Comparing the line and continuum light curves, it is evi-dent that the Hβ response in T2 is heavily dependent onthe line-of-sight velocity.

4.2. Anomalous Emission-Line Response to Continuum

Our data show that previous assumptions about the re-lationship between the continuum and emission-line lightcurves—namely that the emission-line light curves aresmoothed, scaled, and time-shifted versions of the con-tinuum light curve—are not always valid, as we observedall the UV and optical emission-line light curves decor-relating from the UV continuum about halfway throughthe monitoring period. Paper IV examined this effectfor the UV emission lines by measuring changes in theemission-line equivalent width (EW),

EW =Fline

Fcont, (7)

and the responsivity (ηeff), which is the power-law indexthat relates the driving continuum flux to the respondingemission-line fluxes,

log Fline = A + ηeff [log Fcont]. (8)

In the case of no line response, ηeff = 0, and if the line re-sponds linearly to continuum variations (i.e. the transfer

Page 14: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

14

Fig. 10.— Top left: ICCF Hβ lags (τcen) for 500-km s−1 bins measured against the 1158 A and 5100 A continua. Middle left: Lagsfor T1 and T2 measured against the 1158 A continuum. Bottom left: MDM mean spectrum for the full campaign. Right: Maximumcross-correlation coefficients (rmax) for individual velocity bins.

function is a δ function), then ηeff = 1. The emission-lineEW and the continuum flux are related via the Baldwinrelation (Baldwin 1977), which is described by

log EWline = B + β[log Fcont], (9)

where the choice for Fcont is assumed to be a reasonableproxy for the ionizing continuum. Thus, β is also knownas the slope of the Baldwin relation.

Following the same procedures as in Paper IV, we com-pute the responsivity ηeff and EW for the portion of theHβ light curve that correlates with the UV continuum,then examine how these values change in different seg-ments of the light curves. The values Fcont and Fline

refer to the continuum and emission-line fluxes after re-moving non-variable components such as host galaxy andnarrow-line flux contributions, and after correcting forthe mean time delay between the continuum and line

light curves (Pogge & Peterson 1992; Gilbert & Peter-son 2003; Goad et al. 2004). There is very little hostgalaxy flux in the 1158 A continuum, which is dominatedby the variable AGN. For the line fluxes, we took Hβfluxes measured after linear continuum removal (withoutspectral decomposition) and subtracted a constant nar-row Hβ flux measured from the MDM mean spectrum fitto remove the non-variable line component. To correctfor the emission-line time delay, we shifted the Hβ lightcurve by 8 days, which corresponds to the lag for theportion of the line light curve closely correlated with theUV continuum. Figure 12(a) shows the 1158 A contin-uum light curve (black) with the time-shifted and flux-scaled Hβ light curve, which has been truncated at thebeginning to match the first epoch of continuum obser-vations. To show the Hβ light curve’s general behaviortoward the end of the campaign, we have shown here the

Page 15: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

15

TABLE 5Rest-Frame Hβ Velocity-Resolved Lags

Wavelength vr τFull rmax,Full τT1 rmax,T1 τT2 rmax,T2

(A) (km s−1) (days) (days) (days)

4743.82−4751.92 7000 1.74+0.59−0.60 0.61 2.05+1.04

−0.94 0.79 2.13+0.59−0.66 0.42

4751.92−4760.03 6500 2.34+0.49−0.49 0.72 2.57+0.60

−0.71 0.88 3.08+0.98−0.77 0.52

4760.03−4768.13 6000 2.94+0.45−0.48 0.73 3.41+0.63

−0.70 0.88 3.48+0.71−0.66 0.62

4768.13−4776.23 5500 3.18+0.38−0.46 0.75 3.25+0.57

−0.54 0.83 3.46+0.72−0.62 0.62

4776.23−4784.33 5000 3.92+0.50−0.47 0.73 3.12+0.57

−0.53 0.85 4.95+1.04−0.94 0.56

4784.33−4792.43 4500 4.69+0.56−0.49 0.73 3.94+0.63

−0.49 0.85 5.88+0.63−0.70 0.65

4792.43−4800.53 4000 6.19+0.63−0.68 0.63 7.00+0.74

−0.85 0.83 6.26+1.34−1.29 0.53

4800.53−4808.63 3500 7.26+0.62−0.61 0.68 8.45+0.94

−0.87 0.79 6.93+2.34−1.57 0.51

4808.63−4816.73 3000 8.69+0.64−0.67 0.63 10.90+0.89

−0.89 0.81 6.40+2.71−1.48 0.42

4816.73−4824.84 2500 8.70+1.29−1.13 0.53 10.54+1.55

−0.96 0.77 4.92+1.65−0.86 0.42

4824.84−4832.94 2000 7.63+0.87−1.01 0.58 10.46+1.24

−1.07 0.77 4.31+0.62−0.50 0.66

4832.94−4841.05 1500 5.91+0.58−0.53 0.68 8.71+0.78

−0.87 0.80 5.36+0.62−0.74 0.70

4841.05−4849.15 1000 4.65+0.48−0.45 0.71 5.79+0.63

−0.53 0.89 5.36+0.55−0.60 0.74

4849.15−4857.25 500 5.27+0.38−0.49 0.74 5.80+0.61

−0.65 0.90 6.11+0.51−0.67 0.76

4857.25−4865.35 0 6.01+0.37−0.49 0.79 6.50+0.52

−0.60 0.90 6.59+0.57−0.76 0.71

4865.35−4873.45 500 6.40+0.47−0.39 0.75 8.06+0.67

−0.70 0.84 6.43+0.64−0.71 0.67

4873.45−4881.56 1000 6.99+0.38−0.46 0.77 8.86+0.55

−0.39 0.90 6.87+0.53−0.59 0.74

4881.56−4889.66 1500 7.77+0.44−0.37 0.80 9.80+0.41

−0.46 0.92 7.05+0.57−0.52 0.73

4889.66−4897.76 2000 8.59+0.40−0.39 0.81 10.65+0.53

−0.55 0.90 6.38+0.75−0.68 0.65

4897.76−4905.86 2500 10.18+0.66−0.67 0.74 11.13+0.67

−0.68 0.87 6.47+2.01−2.38 0.27

4905.86−4913.96 3000 10.19+0.98−0.92 0.67 11.18+0.71

−0.61 0.86 2.58+1.21−0.98 0.21

4913.96−4922.06 3500 8.23+1.03−0.96 0.66 9.82+0.86

−0.96 0.81 3.78+2.74−1.46 0.27

4922.06−4930.16 4000 7.25+0.87−0.91 0.65 8.80+0.71

−0.70 0.87 4.06+1.57−0.97 0.50

4930.16−4938.27 4500 6.01+0.55−0.59 0.62 7.01+0.81

−0.72 0.88 5.52+0.64−0.70 0.62

4938.27−4946.38 5000 5.04+0.59−0.53 0.47 6.63+0.61

−0.56 0.88 4.85+0.69−0.68 0.58

4946.38−4954.48 5500 4.00+0.56−0.57 0.40 5.22+0.84

−0.83 0.80 4.55+0.72−0.64 0.59

4954.48−4962.58 6000 3.41+0.63−0.60 0.43 3.16+1.05

−1.30 0.83 4.89+0.65−0.70 0.62

4962.58−4970.68 6500 3.20+0.61−0.53 0.53 3.46+0.74

−0.72 0.85 4.30+0.87−0.96 0.53

Note. — ICCF Hβ lags (τcen) for the full campaign and for the T1 (THJD < 6747) and T2 (THJD> 6747) segments. The first column gives the rest-frame wavelength range for each bin and the secondcolumn gives the line-of-sight velocity at the center of each bin. The rmax values give the maxima ofthe cross-correlation functions between the light curve for each velocity bin and the UV continuum.

full 143 epochs of Hβ flux measurements instead of the133-epoch light curve we used in the Hβ lag analysis.However, since the last 10 epochs of spectra suffer frominconsistent spectral flux calibration (see Section 2), wedo not use these points in calculating ηeff or β.

We divided the Hβ light curve into five segments. Thefirst segment corresponds to the period when the linelight curve closely follows the continuum light curve (bluepoints); the second and third segments correspond to pe-riods when the line light curve decouples from the con-tinuum (cyan points) and remains in a state of depressedflux (red points); the last two segments correspond to theline light curve recovering from the depressed state (ma-genta points) and correlating once again with the contin-uum light curve (green points). The epochs that dividethese segments are THJD = [6743, 6772, 6812, 6827].

Figure 12(d) and 12(e) show the Hβ broad-line flux andEW as a function of the 1158 A continuum flux densitydetermined from the HST epoch closest to each MDMepoch. The red lines represent linear least-squares fits toEquations 8 and 9 using only the blue points. We foundthat the cyan, red, and magenta points—correspondingto when the light curves are not well-correlated—lie wellbelow the best fits of Equations 8 and 9 to the blue

points. Furthermore, the epochs during the anomaly (redpoints) are characterized by an ηeff value consistent withzero (ηeff = 0.02± 0.03, black dotted line in Figure 12d),which shows that the emission-line strength remainedconstant independent of continuum strength. Similar re-sults were found for the C IV, Lyα, He II(+O III]) andSi IV(+O IV]) emission lines in Paper IV. Table 6 sum-marizes the ηeff and β values for all broad emission linesmeasured during this campaign. Note that the valuesfrom Paper IV were computed using epochs both beforeand after the anomaly (blue and green points), whereasour values were calculated using only epochs before theanomaly (blue points).

We measured the amplitude of the anomaly by compar-ing the observed Hβ light curve to a simulated light curvethat represents what the line response would have beenwithout the anomaly. The simulated line fluxes (Fsim)were calculated using the observed continuum fluxes andEquation 8, where ηeff and A are chosen to be the best-fit values for the period when the light curves are well-correlated (blue points). This simulated light curve isshown in Figure 12(b) with gray dots, and comparing itwith the observed data (colored points) shows that theHβ line had lower-than-expected variability amplitude

Page 16: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

16

Fig. 11.— Velocity-resolved Hβ light curves for 1500-km s−1 bins(black) with the central velocity for each bin shown at the top leftof each panel. HST 1158 A light curves that have been scaled andshifted to match the first half of the Hβ light curves are shown inred, and the dashed line indicates the epoch that separates T1 andT2 for this analysis. The Hβ response to the continuum is distinctlyvelocity-dependent during the second half of the campaign.

and mean flux during T2. If we define the fractional fluxloss as Flost = (Fsim −Fobs)/Fsim (Figure 12c), then thisimplies an Hβ flux deficit of ∼ 6% during the anomaly(red points), which is close to the deficit for Lyα butmuch smaller than that of the other UV emission lines,as shown in Table 6.

Table 7 summarizes the Hβ responsivities and AGNoptical continuum flux densities measured by Goad et al.(2004, GKK04) for every year of the 13-year moni-toring campaign carried out by the AGN Watch con-

Fig. 12.— Panel (a) shows 1158 A continuum light curve withthe time-shifted Hβ light curve (see text for color scheme). Panels(b) and (c) show the reconstructed Hβ light curve (gray dots) andthe percent of flux lost during the anomaly, respectively. Panels(d) and (e) show the Hβ broad-line flux (10−15 erg s−1 cm−2)and EW (A) as a function of the 1158 A continuum flux density

(10−15 erg s−1 cm−2 A−1

). The solid red lines are linear least-squares fits to the blue points, which are from the period when theline and continuum light curves are well-correlated. The slopes ofthese fits are shown in each panel. The dotted black line in panel(d) is the least-squares fit to the red points (anomaly) and has aslope of ηeff = 0.02 ± 0.03.

TABLE 6Responsivity, Slope of the Baldwin

Relation, and Flux Deficit During theAnomaly for All Emission Lines

Line ID ηeff β Flost

Lyα 0.30 ± 0.01 −0.73 ± 0.02 9%Si IV+O IV] 0.45 ± 0.01 −0.58 ± 0.03 23%

C IV 0.25 ± 0.01 −0.75 ± 0.01 18%He II+O III] 0.58 ± 0.04 −0.48 ± 0.04 21%

Hβ 0.15 ± 0.01 −0.85 ± 0.02 6%

Note. — All values were measured using the1158 A continuum fluxes from this campaign. Thefirst four rows show values from Goad et al. (2016).

Page 17: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

17

TABLE 7Hβ Responsivity and AGN Optical

Continuum Flux Density for NGC 5548Over 25 Years

Year ηeff Mean Fcont rms Fcont

1989 0.56 ± 0.04 6.54 1.271990 0.84 ± 0.03 3.79 0.911991 0.95 ± 0.09 6.06 0.921992 0.94 ± 0.05 3.34 1.171993 0.43 ± 0.04 5.69 0.871994 0.74 ± 0.04 6.40 1.111995 0.68 ± 0.04 8.71 1.011996 0.54 ± 0.03 7.07 1.521997 0.80 ± 0.07 4.73 0.891998 0.51 ± 0.02 10.05 1.441999 0.41 ± 0.04 8.48 1.822000 0.65 ± 0.06 3.59 1.202001 1.00 ± 0.12 3.65 0.862014 0.59 ± 0.01 7.44 0.50

Note. — Flux densities are in units of10−15 erg s−1 cm−2 A

−1. The first 13 rows

show values from Goad et al. (2004)‡, and thelast row shows values from this campaign. Allηeff values listed were measured with respect tothe 5100 A continuum.

sortium (Peterson et al. 2002)‡, along with the val-ues of ηeff = 0.59 ± 0.01 and 〈F5100,AGN〉 = (7.44 ±

0.50) × 10−15 erg s−1 cm−2 A−1

calculated from thisdataset. Both the responsivity and optical continuumflux density for this campaign are close to those mea-sured from the 1998 data. Using the mean Hβ flux with-out narrow-line contributions of 〈FHβ〉 = (690 ± 2.40) ×10−15 erg s−1 cm−2, we find a mean EW for this cam-paign of 92.75± 2.45 A, which is lower than values mea-sured by Goad & Korista (2014) for previous campaigns.

The Hβ light curve appears to decorrelate from the UVcontinuum light curve at a somewhat earlier time (THJD≈ 6742) than C IV (THJD ≈ 6765 as found in PaperIV). If we assume that the Hβ light curve decorrelatesand re-correlates with the continuum light curve at thesame times as C IV and use the same dividing epochsas those used in Paper IV (THJD = [6766, 6777, 6814,6830]), then ηeff = 0.13 ± 0.01 and β = −0.88 ± 0.01 forthe blue points.

While He II λ4686 also shows anomalous behavior dur-ing the T2 period (Fig. 7), its broad line component isvery weak and the fitted profile is poorly constrained inthe spectral decomposition process. The He II light curveis thus noisier than that of Hβ, and the Fline and EWvalues are poorly correlated with Fcont. We therefore donot perform a detailed analysis of the responsivity forHe II.

4.3. Line Width and MBH Estimate

The black hole mass in NGC 5548 has been estimatedfor several previous RM campaigns, including the AGNWatch consortium (Peterson et al. 2002), Bentz et al.(2007), Denney et al. (2010), and Bentz et al. (2009b).The AGN Watch group determined the BH mass usingdata from each of the program’s monitoring years, andsubsequent campaigns each produced an independent BH

‡The continuum flux densities from Peterson et al. (2002) havebeen updated by Bentz et al. (2013) and Kilerci Eser et al. (2015).

mass value. Here we compute the BH mass using datafrom this campaign and compare it with previous results.

We measured the line widths from the Hβ mean andrms spectra as shown in Figure 5. The narrow Hβ lineessentially disappears in the rms spectrum but is stillpresent in the mean spectrum. We removed this emis-sion component by subtracting a Gaussian fit to the Hβnarrow line in the mean MDM spectrum, then linearlyinterpolated over the narrow Hβ residuals and the [O III]λ4959 and λ5007 residuals.

Two emission-line width values are typically measuredin RM: the FWHM and the line dispersion, defined by

σ2line =

(

c

λ0

)2(∑

λ2iSi

Si− λ2

0

)

, (10)

where Si is the flux density at wavelength bin λi and λ0

is the flux-weighted centroid wavelength of the line pro-file. We measured σline and FWHM for the mean andrms spectra using the line profile within the rest-framewavelength range 4669.8−5063.0 A. We treated the meanprofile as double-peaked and followed the procedures de-scribed by Peterson et al. (2004) to measure its FWHM.From each of the two peaks at 4827.1 A and 4886.1 A,we traced the line profile outward until the flux reached0.5Fmax, then traced the line profile inward from the con-tinuum until the flux again reached 0.5Fmax. The twowavelengths at 0.5Fmax—which generally agree well fora smooth line profile—are then averaged to obtain thewavelength at half maximum on each side of the profile.The rms profile is more complicated because it has morethan two peaks and the troughs between them can reachwell below half of the peak fluxes. We therefore identifieda single maximum flux Fmax and traced the profile fromthe continuum toward the center on both sides until theflux reached 0.5Fmax. The separation between the twowavelengths at 0.5Fmax is taken to be the FWHM of therms spectrum.

The Hβ line widths and their uncertainties were de-termined using Monte Carlo bootstrap analysis. With ntotal spectra, we randomly selected n spectra from thedataset with replacement, constructed mean and rms lineprofiles, and measured the line dispersion and FWHM.The median and standard deviation of 104 bootstrap re-alizations are used for the σline and FWHM and theirestimated uncertainties.

There are additional systematic uncertainties in theline widths from using different Fe II templates in thespectral decomposition. We repeated the bootstrap anal-ysis after performing spectral decompositions using eachof the Boroson & Green (1992), Veron-Cetty et al. (2004)and Kovacevic et al. (2010) Fe II templates, then took thestandard deviation of the Hβ widths from using the dif-ferent templates as the systematic uncertainty for the linewidth. This systematic error dominates the error budgetfor all σline measurements and the FWHM of the meanspectrum, but is comparable to the uncertainty frombootstrapping analysis for the FWHM of the rms spec-trum. We added this systematic uncertainty in quadra-ture to the statistical uncertainty from the Kovacevicet al. (2010) line width to obtain the final Hβ line widthuncertainty.

The line widths are also affected by instrumental

Page 18: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

18

TABLE 8Hβ Rest-Frame Line Widths

Segment Spectrum σline FWHM(km s−1) (km s−1)

Full rms 4278 ± 671 10161 ± 587T1 rms 4155 ± 513 10861 ± 739T2 rms 4856 ± 731 9103 ± 1279

Full Mean 3691 ± 162 9496 ± 418T1 Mean 3983 ± 150 9612 ± 427T2 Mean 3939 ± 177 9380 ± 158

broadening due to the use of a wide slit. The observedline width is the quadratic sum of the intrinsic and in-strumental line widths (σ2

observed = σ2intrinsic+σ2

instrumentalor similarly for FWHM). To calculate the instrumentalbroadening for a 5′′ slit, we followed the methods de-scribed by Bentz et al. (2009b) and compare the [O III]λ5007 line width measured from our observations tothe width measured from a higher-resolution observationtaken using a narrow (∼ 2′′) slit, which represents the in-trinsic line width. The FWHM of the [O III] λ5007 modelfrom the MDM rest-frame mean spectrum is 9.79 A (572km s−1), while Whittle (1992) found a FWHM of 410 kms−1 using a 2′′ slit. This implies an instrumental broad-ening of FWHMinstrument = 399 km s−1, correspondingto σinstrument = 170 km s−1 for a Gaussian model of the[O III] line profile. We subtract this instrumental widthin quadrature from the Hβ line width measurements toobtain the intrinsic Hβ line widths, which are listed inTable 8. The large uncertainties on the FWHM measure-ments from the rms spectrum are a result of the jaggedshape of the rms profile.

We use the Hβ line dispersion measured from the rmsspectrum as the velocity dispersion ∆V , as is commonpractice in RM (Peterson et al. 2004, but also see Collinet al. 2006 and Mejıa-Restrepo et al. 2016 for comparisonof FWHM and sigma as indicators of BLR virial velocityfor BH mass estimation), and calculate the virial prod-uct, defined as VP = cτ∆V 2/G. The f factor in Equa-tion 2, which incorporates the geometry and kinemat-ics of the BLR, is generally unknown for any individualAGN, so a single value 〈f〉 is often used to represent theaverage normalization for all AGNs. This value is usu-ally taken to be the scale factor that puts the sample ofRM virial products onto the same MBH − σ⋆ relation asnearby inactive galaxies (see Kormendy & Ho 2013 for adiscussion of the uncertainties in the MBH−σ⋆ relation),and can vary depending on the sample of AGNs used inthe fit as well as the fitting method (Onken et al. 2004;Watson et al. 2007; Woo et al. 2010; Graham et al. 2011;Park et al. 2012a; Grier et al. 2013; Ho & Kim 2015). Weadopt a value of 〈f〉 = 4.47 as calculated by Woo et al.(2015). Since 〈f〉 is calibrated using Hβ lags measuredagainst the optical continuum, we used the Hβ–5100 Alag to calculate the VP and BH mass.

Table 9 lists the virial products and the inferred BHmasses for the full, T1, and T2 segments. Our MBH

uncertainties do not include uncertainties in 〈f〉 (Wooet al. 2015) or scatter in the distribution of f for differentAGNs (e.g. Ho & Kim 2015). The T1 and T2 MBH esti-mates are consistent even though the two lags are differ-ent by more than 1σ. Compared to recent measurements,

the full-campaign virial product (1.49+0.49−0.49 × 107 M⊙) is

entirely consistent with the values of 1.50+0.37−0.51× 107 M⊙

and 1.38+0.51−0.41×107 M⊙ obtained by Bentz et al. (2009b)

and Lu et al. (2016), respectively. Our BH mass mea-surement (6.66+2.17

−2.17 × 107 M⊙) is also consistent to 1σ

with the dynamical mass of 3.24+2.26−0.90 × 107 M⊙ as de-

termined by Pancoast et al. (2014, P14). P14 use theV -band continuum as the ionizing source but do not usethe f -factor, which compensates for the difference in lagbetween using the UV and optical continua (see nextsection). This will lead to an underestimate of RBLR

and hence a proportional underestimate of the BH mass(Eq. 2). If we scale the P14 mass by the ratio betweenthe Hβ–1158 A and the Hβ–V -band lags from Table 4(τHβ−UV/τHβ−V = 1.64± 0.12), then the two MBH mea-surements become much more congruent.

5. DISCUSSION

5.1. Implications of UV and Optical Hβ Lags

Ground-based RM campaigns have traditionally usedthe optical continuum light curve—by necessity—to de-termine emission-line lags, even though the far-UV con-tinuum is a better proxy for the ionizing source. Lags rel-ative to the UV continuum (τHβ−UV) thus should yieldmore accurate estimates of the BLR characteristic ra-dius than lags measured relative to the optical contin-uum. Our Hβ–UV lag (τHβ−UV = 6.23+0.39

−0.44 days) is ∼ 2

days longer than the Hβ–optical lag (τHβ−opt = 4.17+0.36−0.36

days). Given that past measurements of the Hβ-opticallag for this object range from ∼ 4 to ∼ 25 days (Bentzet al. 2013, and references therein), and assuming thatτHβ−opt is always ∼ 2 days longer than τHβ−UV, then theBLR characteristic radius estimated from previous cam-paigns using only optical data is biased low by 10−50%.The difference in these Hβ lags is also consistent withthe optical-to-UV continuum lag of 2.24+0.24

−0.24 days foundin Paper III.

Since virial estimates of MBH scale with RBLR, it mayseem that this difference between τHβ−UV and τHβ−opt

will change the BH mass estimate for NGC 5548 andother reverberation-mapped AGNs. However, the virialproduct—not the BH mass—is the quantity that is di-rectly affected, and the normalization factor f is stillneeded to scale the virial product to a calibrated BHmass (Eq. 2). If the ratio of τHβ−UV/τHβ−opt is the samefor all AGNs, then all RM virial products would be scaledup by a constant value, so simply changing the value off would remove this bias and leave the RM black holemasses unchanged. Even if the lag ratio is not constantfor all AGNs, its effect on the BH mass scale is still smallbecause the largest source of MBH uncertainty in RM isthe calibration uncertainty for f at ∼ 0.12 dex (Woo et al.2015). Furthermore, for NGC 5548, the Hβ lag measuredduring this campaign is relatively short compared to itshistorical values (see Section 5.3). If the lag had beenlonger, the ratio of τHβ−UV/τHβ−opt would be closer tounity and the bias in the BLR characteristic size and BHmass estimate would be much smaller.

While the discrepancy between τHβ−UV and τHβ−opt

may not change MBH measurements that use the f -factor, dynamical models that directly infer MBH and

Page 19: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

19

TABLE 9Hβ Line Measurements, MBH Estimates, and Continuum Flux Densities

Segment σline τHβ−opt Virial Product MBH F5100,total F5100,AGN

(km s−1) (days) (107 M⊙) (107 M⊙) (10−15 erg s−1 cm−2 A−1

)

Full 4278 ± 671 4.17+0.36−0.36 1.49+0.49

−0.49 6.66+2.17−2.17 11.96 ± 0.07 7.44 ± 0.50

T1 4155 ± 513 4.99+0.40−0.47 1.68+0.44

−0.45 7.53+1.96−1.99 11.31 ± 0.08 6.79 ± 0.46

T2 4856 ± 731 3.10+0.77−0.80 1.43+0.55

−0.56 6.38+2.49−2.53 12.51 ± 0.04 7.99 ± 0.45

Note. — τHβ−opt is the rest-frame ICCF τcen value measured against the 5100 A continuum.

BLR characteristics (e.g., Pancoast et al. 2011; Li et al.2013) using only optical continuum data are affectedsince they do not depend on this virial normalization.The BLR characteristic size for each AGN inferred usingoptical data alone would thus be biased by a factor thatdepends on the value of τHβ−UV/τHβ−opt for that par-ticular object. Changes in the inferred RBLR could alsoimpact single-epoch MBH estimates, which rely on theempirical relation between the BLR characteristic radiusand the AGN continuum luminosity (RBLR ∝ Lα

AGN withα ≈ 1/2, e.g., Laor 1998; Wandel et al. 1999; Kaspi et al.2000; McLure & Jarvis 2002; Kaspi et al. 2005; Vester-gaard & Peterson 2006; Bentz et al. 2006, 2009a, 2013).If τHβ−UV/τHβ−opt correlates with AGN luminosity, thenthe expected value of α would change, and if this ratio isdifferent for all AGNs but is uncorrelated with any otherAGN properties, then this would introduce additionalscatter to the scaling relation.

We can examine the expected scaling ofτHβ−UV/τHβ−opt with respect to MBH and LAGN

by using simple disk and BLR ionization models.Assuming τHβ−UV is the sum of τHβ−opt and theoptical–UV inter-band continuum lag τopt−UV, then

τHβ−opt

τHβ−UV= 1 −

τopt−UV

τHβ−UV. (11)

For a standard thin disk, the characteristic size scale ofthe disk region emitting at wavelength λ scales as

Rλ ∝ M2/3BH l1/3λ4/3, (12)

where l = L/LEdd. For a simple photoionization equilib-rium model, the BLR characteristic radius scales as

RBLR ∝ M1/2BH l1/2. (13)

If the inter-band continuum lags are due to light-traveltime across the accretion disk, then

τopt−UV

τHβ−UV∝ M

1/6BH l−1/6λ4/3. (14)

The lag ratio is thus weakly dependent on both MBH andaccretion rate, where even a factor of 103 increase in theblack hole mass or accretion rate will change the ratio byonly a factor of three. Empirically, Bentz et al. (2013)found a low scatter of 0.13 dex around the RBLR−LAGN

relation for 41 AGNs over four orders of magnitude inluminosity, which suggests this effect is indeed small formost AGNs. However, it is important to directly exam-ine the potential consequences of this effect by obtainingmore simultaneous observations of the UV/optical con-tinua and the broad emission lines for AGNs over a wide

range of luminosities.

5.2. Anomalous Emission-Line Light Curve Behavior

Despite apparent differences in the C IV and Hβ decor-relation start times and flux deficits during T2, it is likelythat the cause of the anomalous light-curve behavior isthe same for the UV and optical emission lines. The lineresponse during the anomaly is also heavily dependenton the line-of-sight velocity for Hβ (Fig. 11) and the UVemission lines (Goad et al, in prep.). Paper IV suggeststwo scenarios that could produce the anomaly: (1) a tem-porary obscuration of the ionizing source from parts ofthe BLR by a moving veil of gas between the accretiondisk and BLR, or (2) a temporary change in spectral en-ergy distribution of the ionizing source. Future paperswill investigate in detail the timing and magnitude ofthe anomaly for all UV and optical lines using the fullmulti-wavelength dataset from this campaign.

The bottom panel of Figure 7 shows that the Hβlight curve decorrelation was also clearly detected us-ing only the optical data. If our campaign had lastedfor only the duration of T2, we still would have beenable to measure the Hβ lags with fairly high precision(τHβ−UV = 5.99+0.71

−0.75 days and τHβ−opt = 3.10+0.77−0.80

days), but the lag signal would be contaminated by otherunknown factors and would lead to a somewhat biasedestimate of the BLR characteristic radius. Depending onhow common this decorrelation behavior is for Hβ, thiseffect could contribute to additional scatter in the single-object RBLR − LAGN relations for NGC 5548 and otherAGNs, which can account for about half of the observedscatter in the global RBLR − LAGN relation for the en-tire sample of reverberation-mapped AGNs (Kilerci Eseret al. 2015).

The high cadence and long duration of this campaignhave both been crucial in detecting this decorrelationphenomenon. Horne et al. (2004) found that a campaignduration of at least three times the maximum BLR light-crossing time is needed to recover high-fidelity velocity-delay maps from reverberation mapping data. Giventhe Hβ–UV lag of ∼ 6 days for NGC 5548 during ourmonitoring period, the BLR characteristic light-crossingtime is ∼ 12 days and the minimum campaign lengthneeded to recover velocity-delay maps would correspondto ∼ 40 days, which would not have allowed us to see thisdecorrelation. In order to detect and characterize theseanomalous behaviors in the emission-line light curves,RM campaigns must be much longer than the minimumrequirement for obtaining velocity-delay maps.

Finally, while there are no previously published resultsdocumenting similar emission-line light curve behavior,it is possible that this decorrelation phenomenon was

Page 20: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

20

Fig. 13.— NGC 5548 optical continuum luminosity and Hβ–optical lags from all monitoring campaigns to date. The solid lineis the linear least-squares fit to the Kilerci Eser et al. (2015) andDenney et al. (2009) points.

indeed observed in other AGNs in previous RM cam-paigns, but were not recognized as such because thecampaign had relatively low cadence and/or short du-ration. RM programs designed to study large numbersof sources with lower cadence would also not be ableto detect these decorrelations. This further highlightsthe importance of high-cadence, long-duration and high-SNR multi-wavelength reverberation datasets in order todetermine the prevalence of this phenomenon.

5.3. Comparison to Previous Campaigns: theRBLR − LAGN Relation

We compare the NGC 5548 Hβ–optical lags and op-tical continuum luminosity (L5100) from this campaignto those from previous RM campaigns targeting this ob-ject, as compiled by Kilerci Eser et al. (2015). The total5100 A flux and the AGN continuum flux densities forthe full, T1, and T2 segments are listed in columns 6 and7 of Table 9, respectively. We applied a Galactic extinc-tion correction of E(B−V ) = 0.017 mag (Schlegel et al.1998; Schlafly & Finkbeiner 2011) and used a luminositydistance of 75 Mpc in converting fluxes to luminosities.Figure 13 shows the RBLR−LAGN relation for NGC 5548,where the uncertainties are from absolute photometriccalibration using [O III] λ5007 and do not include theluminosity distance uncertainty. The lags from this cam-paign have smaller uncertainties due to the high cadenceand long duration of the ground-based monitoring. Den-ney et al. (2010) monitored NGC 5548 as part of a multi-object RM campaign and found τHβ−opt = 12.40+2.74

−3.85days. However, the light curves were dominated by alarge, long-term trend, and after detrending the lightcurves with a 3rd order polynomial, the Hβ lag becomes5.07+2.46

−2.37 days (open circle in Fig. 13).The surprising result is that, given the AGN luminosity

during this campaign, the Hβ lags are nearly five timesshorter than expected based on past measurements.NGC 5548 was also monitored in 2012 (AGN12, De Rosaet al., in prep.) and had an Hβ lag of 4.1± 0.9 days anda luminosity of log(λL5100) = 43.22 ± 0.1, whereas Luet al. (2016) found the Hβ lag to be 7.2+1.33

−0.35 days and

the AGN luminosity to be log(λL5100) = 43.21±0.1 dur-ing their 2015 campaign, as shown in Figure 13. Both ofthese lags are also shorter than expected based on pastresults, though the point from Lu et al. (2016) suggeststhat the AGN may be returning to its previously mea-sured RBLR − LAGN relation.

In numerical simulations of the emission-line responseto continuum variations for a model BLR with fixed ra-dial extent, Goad & Korista (2014) found that if thecharacteristic continuum variability timescale (τchar) issmaller than the BLR light-crossing time, then there ex-ists a strong correlation between τchar, the line responsiv-ity (ηeff), and measured lag, such that both ηeff and thelag decrease as τchar decreases (their Figure 9). This isbecause short timescale continuum variations only probethe inner parts of the BLR, so the measured lag and re-sponsivity would be biased low. If we use the FWHMof the continuum light curve auto-correlation functionas a crude proxy for the characteristic variability timescale, then τchar ∼ 10 days, which is significantly shorterthan the values measured for this source for the 1989IUE campaign and the 1993 HST campaign, and couldexplain the shorter than expected Hβ lags.

Regardless of the physical causes of the short lags,these results suggest that the RBLR − LAGN relationis more complex than previously realized. Bentz et al.(2013) found a tight correlation between RBLR and LAGN

for a sample of ∼ 40 reverberation mapped AGNs, butrecent studies by Du et al. (2015) and Du et al. (2016b)have shown that many AGNs with very high accretionrates tend to have considerably shorter Hβ lags com-pared to low-accretion AGNs with similar luminosities.Now, we have shown that even for a single AGN withlow accretion rate (λEdd = 0.021 for NGC 5548; Vasude-van et al. 2010), the RBLR − LAGN relation does not al-ways follow a tight power law and that more complexphysical processes may contribute significantly to thescatter. In order to further investigate the single-objectRBLR − LAGN relation, repeated monitoring campaignsfor individual AGNs are needed to track the behaviorof each object over a range of timescales and luminositystates. This will, in turn, help improve our understand-ing of the global RBLR − LAGN relation.

6. SUMMARY

We present the results of an optical spectroscopic mon-itoring program in 2014 targeting the galaxy NGC 5548as part of the AGN STORM project. Our campaignspanned six months and observed the AGN with an al-most daily cadence. Our main findings are as follows.

(1) We determined Hβ and He II λ4686 emission-linelags relative to the far-UV and optical continua, andfound that the lag measured against the UV continuumis ∼ 2 days longer than that measured against the opti-cal continuum, consistent with the lag between the UVand optical continua. Given that past measurements ofthe Hβ lag against the optical continuum for this objectrange from ∼ 4 to ∼ 25 days and assuming that this2-day lag difference is constant over time, then the char-acteristic size of the BLR inferred from previous data isbiased low by 10%–50%. Depending on how the ratio ofUV and optical Hβ lags scales with other AGN proper-ties, the RM black hole mass scale and the RBLR−LAGN

relation may be affected, which would, in turn, impact

Page 21: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

21

single-epoch MBH estimates for high-redshift AGNs.(2) We measured velocity-resolved lags for the broad

Hβ line and found a double-peaked lag profile as a func-tion of line-of-sight velocity, with shorter lags in the high-velocity wings. The overall shape of the lag profile isqualitatively similar to those of Keplerian models (e.g.,Horne et al. 2004), and is very similar to what is foundfor Lyα (De Rosa et al. 2015).

(3) Both the Hβ and He II λ4686 emission lines ex-hibit significant changes in their response to UV con-tinuum variations halfway through our monitoring cam-paign. The line light curves decorrelate from that of thecontinuum and remain in a suppressed state until nearthe end of the campaign. The same anomalous behavioris observed for all the UV emission lines (De Rosa et al.2015; Goad et al. 2016). Further investigation into thesimultaneous UV and optical line responses during thiscampaign may elucidate the cause of this anomaly. De-pending on how frequently this phenomenon occurs inthe AGN population as a whole, this effect could con-tribute to the scatter in both single-object and globalRBLR − LAGN relations. This type of anomalous linebehavior is likely only detectable with monitoring cam-paigns that have a combination of high cadence, longduration, and high data quality.

(4) Given the optical luminosity of NGC 5548 duringour campaign, the Hβ lag measured against the opticalcontinuum is a factor of five shorter than the expectedvalue based on the RBLR −LAGN relation for NGC 5548from past monitoring campaigns. Our results, combinedwith other recent Hβ lag measurements, suggest that thisobject does not follow a simple power-law RBLR −LAGN

relation at all times.

We thank the staffs at the various observatories usedto obtain the data in this paper. Support for HSTprogram number GO-13330 was provided by NASAthrough a grant from the Space Telescope Science In-stitute, which is operated by the Association of Uni-versities for Research in Astronomy, Inc., under NASAcontract NAS5-26555. L.P. and A.J.B. have been sup-ported by National Science Foundation (NSF) grantAST-1412693. M.M.F., G.D.R., B.M.P., C.J.G., andR.W.P. are grateful for the support of NSF grant AST-1008882 to The Ohio State University. M.C. Bentz grate-fully acknowledges support through NSF CAREER grantAST-1253702 to Georgia State University. A.V.F.’sgroup at UC Berkeley is grateful for financial assistancefrom NSF grant AST-1211916, the TABASGO Founda-tion, and the Christopher R. Redlich Fund. C.S.K. issupported by NSF grant AST-1515876. V.N.B. grate-fully acknowledges assistance from a NSF Research atUndergraduate Institutions (RUI) grant AST-1312296.M.C. Bottorff acknowledges HHMI for support throughan undergraduate science education grant to South-

western University. K.D.D. is supported by an NSFFellowship awarded under grant AST-1302093. M.E.thanks the members of the Center for Relativistic As-trophysics at Georgia Tech and the Department of As-tronomy at the University of Washington, where he wasbased during the observing campaign, for their warmhospitality. R.E. gratefully acknowledges support fromNASA under awards NNX13AC26G, NNX13AC63G,and NNX13AE99G. J.M.G. gratefully acknowledges sup-port from NASA under award NNH13CH61C. P.B.H.is supported by NSERC. K.H. acknowledges supportfrom the UK Science and Technology Facilities Councilthrough grant ST/M001296/1. TW-SH is supported bythe DOE Computational Science Graduate Fellowship,grant number DE-FG02- 97ER25308. M.I. acknowledgessupport from the Creative Initiative program, No. 2008-0060544, of the National Research Foundation of Ko-rea (NRFK) funded by the Korean government (MSIP).M.D.J. acknowledges NSF grant AST-0618209. SRONis financially supported by NWO, the Netherlands Orga-nization for Scientific Research. B.C.K. is partially sup-ported by the UC Center for Galaxy Evolution. C.S.K.acknowledges the support of NSF grant AST-1009756.D.C.L. acknowledges support from NSF grants AST-1009571 and AST-1210311. P.L. acknowledges supportfrom Fondecyt grant No. 1161184. A.P. acknowledgessupport from a NSF graduate fellowship and a UCSBDeans Fellowship. J.S.S. acknowledges CNPq, NationalCouncil for Scientific and Technological Development(Brazil) for partial support and The Ohio State Uni-versity for warm hospitality. N.T. acknowledges sup-port from CONICYT PAI/82140055. T.T. has been sup-ported by NSF grant AST-1412315. T.T. and B.C.K. ac-knowledge support from the Packard Foundation in theform of a Packard Research Fellowship to T.T. Also, T.T.thanks the American Academy in Rome and the Obser-vatory of Monteporzio Catone for kind hospitality. TheDark Cosmology Centre is funded by the Danish NationalResearch Foundation. M.V. gratefully acknowledges sup-port from the Danish Council for Independent Researchvia grant no. DFF 4002-00275. J.-H.W. acknowledgessupport by the National Research Foundation of Ko-rea (NRF) grant funded by the Korean government (No.2010-0027910). This work is based partly on observationsobtained with the Apache Point Observatory 3.5-m tele-scope, which is owned and operated by the AstrophysicalResearch Consortium. Research at Lick Observatory ispartially supported by a generous gift from Google. Thisresearch has made use of the NASA/IPAC ExtragalacticDatabase (NED), which is operated by the Jet Propul-sion Laboratory, California Institute of Technology, un-der contract with the National Aeronautics and SpaceAdministration. We thank the anonymous referee forthe thorough review and helpful comments that helpedto improve the clarity of this manuscript.

REFERENCES

Antonucci, R. R. J., & Cohen, R. D. 1983, ApJ, 271, 564Baldwin, J. A. 1977, ApJ, 214, 679Barth, A. J., et al. 2015, ApJS, 217, 26Barth, A. J., & Bentz, M. C. 2016, MNRAS, 458, L109Barth, A. J., et al. 2011, ApJ, 732, 121—. 2013, ApJ, 769, 128Bentz, M. C., et al. 2007, ApJ, 662, 205

—. 2013, ApJ, 767, 149—. 2010a, ApJ, 720, L46Bentz, M. C., & Katz, S. 2015, PASP, 127, 67Bentz, M. C., Peterson, B. M., Netzer, H., Pogge, R. W., &

Vestergaard, M. 2009a, ApJ, 697, 160Bentz, M. C., Peterson, B. M., Pogge, R. W., Vestergaard, M., &

Onken, C. A. 2006, ApJ, 644, 133

Page 22: arXiv:1702.01177v1 [astro-ph.GA] 3 Feb 2017

22

Bentz, M. C., et al. 2009b, ApJ, 705, 199—. 2010b, ApJ, 716, 993Blandford, R. D., & McKee, C. F. 1982, ApJ, 255, 419Boroson, T. A., & Green, R. F. 1992, ApJS, 80, 109Bruzual, G., & Charlot, S. 2003, MNRAS, 344, 1000Clavel, J., et al. 1991, ApJ, 366, 64Collier, S., & Peterson, B. M. 2001, ApJ, 555, 775Collier, S. J., et al. 1998, ApJ, 500, 162Collin, S., Kawaguchi, T., Peterson, B. M., & Vestergaard, M.

2006, A&A, 456, 75Crenshaw, D. M., & Blackwell, Jr., J. H. 1990, ApJ, 358, L37Czerny, B., Doroshenko, V. T., Niko lajuk, M.,

Schwarzenberg-Czerny, A., Loska, Z., & Madejski, G. 2003,MNRAS, 342, 1222

Czerny, B., Schwarzenberg-Czerny, A., & Loska, Z. 1999,MNRAS, 303, 148

De Rosa, G., et al. 2015, ApJ, 806, 128Denney, K. D., et al. 2006, ApJ, 653, 152—. 2010, ApJ, 721, 715—. 2009, ApJ, 704, L80Dietrich, M., et al. 1993, ApJ, 408, 416—. 1998, ApJS, 115, 185Done, C., & Krolik, J. H. 1996, ApJ, 463, 144Du, P., et al. 2015, ApJ, 806, 22—. 2016a, ArXiv e-prints—. 2016b, ArXiv e-printsEdelson, R., et al. 2015, ApJ, 806, 129Emmering, R. T., Blandford, R. D., & Shlosman, I. 1992, ApJ,

385, 460Fausnaugh, M. M. 2016, ArXiv e-printsFausnaugh, M. M., et al. 2016, ApJ, 821, 56Filippenko, A. V. 1982, PASP, 94, 715Gaskell, C. M. 1988, ApJ, 325, 114Gaskell, C. M., & Goosmann, R. W. 2013, ApJ, 769, 30—. 2016, Ap&SS, 361, 67Gaskell, C. M., & Sparke, L. S. 1986, ApJ, 305, 175Gilbert, K. M., & Peterson, B. M. 2003, ApJ, 587, 123Goad, M., & Wanders, I. 1996, ApJ, 469, 113Goad, M. R., & Korista, K. T. 2014, MNRAS, 444, 43Goad, M. R., et al. 2016, ApJ, 824, 11Goad, M. R., Korista, K. T., & Knigge, C. 2004, MNRAS, 352,

277Goad, M. R., Korista, K. T., & Ruff, A. J. 2012, MNRAS, 426,

3086Graham, A. W., Onken, C. A., Athanassoula, E., & Combes, F.

2011, MNRAS, 412, 2211Grier, C. J., et al. 2013, ApJ, 773, 90Ho, L. C., & Kim, M. 2015, ApJ, 809, 123Horne, K. 1986, PASP, 98, 609Horne, K., Peterson, B. M., Collier, S. J., & Netzer, H. 2004,

PASP, 116, 465Horne, K., Welsh, W. F., & Peterson, B. M. 1991, ApJ, 367, L5Kaspi, S., Brandt, W. N., Maoz, D., Netzer, H., Schneider, D. P.,

& Shemmer, O. 2007, ApJ, 659, 997Kaspi, S., Maoz, D., Netzer, H., Peterson, B. M., Vestergaard,

M., & Jannuzi, B. T. 2005, ApJ, 629, 61Kaspi, S., Smith, P. S., Netzer, H., Maoz, D., Jannuzi, B. T., &

Giveon, U. 2000, ApJ, 533, 631Kelly, B. C., Bechtold, J., & Siemiginowska, A. 2009, ApJ, 698,

895Kilerci Eser, E., Vestergaard, M., Peterson, B. M., Denney, K. D.,

& Bentz, M. C. 2015, ApJ, 801, 8King, A. L., et al. 2015, MNRAS, 453, 1701Kollatschny, W. 2003, A&A, 407, 461Korista, K. T., et al. 1995, ApJS, 97, 285Kormendy, J., & Ho, L. C. 2013, ARA&A, 51, 511Kovacevic, J., Popovic, L. C., & Dimitrijevic, M. S. 2010, ApJS,

189, 15Koz lowski, S., et al. 2010, ApJ, 708, 927Krolik, J. H., Horne, K., Kallman, T. R., Malkan, M. A., Edelson,

R. A., & Kriss, G. A. 1991, ApJ, 371, 541Laor, A. 1998, ApJ, 505, L83Leighly, K. M., & Moore, J. R. 2004, ApJ, 611, 107Li, Y.-R., Wang, J.-M., Ho, L. C., Du, P., & Bai, J.-M. 2013,

ApJ, 779, 110Lu, K.-X., et al. 2016, ApJ, 827, 118

MacLeod, C. L., et al. 2010, ApJ, 721, 1014

McHardy, I. M., et al. 2014, MNRAS, 444, 1469McLure, R. J., & Jarvis, M. J. 2002, MNRAS, 337, 109Mejıa-Restrepo, J. E., Trakhtenbrot, B., Lira, P., Netzer, H., &

Capellupo, D. M. 2016, MNRAS, 460, 187Miller, J. S., & Stone, R. P. S. 1993, Lick Obs. Tech. Rep., Tech.

rep., Santa Cruz, CA: Lick ObservatoryMurray, N., & Chiang, J. 1997, ApJ, 474, 91Onken, C. A., Ferrarese, L., Merritt, D., Peterson, B. M., Pogge,

R. W., Vestergaard, M., & Wandel, A. 2004, ApJ, 615, 645Pancoast, A., Brewer, B. J., & Treu, T. 2011, ApJ, 730, 139Pancoast, A., Brewer, B. J., Treu, T., Park, D., Barth, A. J.,

Bentz, M. C., & Woo, J.-H. 2014, MNRAS, 445, 3073Park, D., Kelly, B. C., Woo, J.-H., & Treu, T. 2012a, ApJS, 203, 6Park, D., et al. 2012b, ApJ, 747, 30Peterson, B. M. 1993, PASP, 105, 247Peterson, B. M. 2006, in Astronomical Society of the Pacific

Conference Series, Vol. 360, Astronomical Society of the PacificConference Series, ed. C. M. Gaskell, I. M. McHardy, B. M.Peterson, & S. G. Sergeev, 191

—. 2014, Space Sci. Rev., 183, 253Peterson, B. M., et al. 1991, ApJ, 368, 119—. 1999, ApJ, 510, 659—. 2002, ApJ, 581, 197—. 2013, ApJ, 779, 109—. 2004, ApJ, 613, 682Peterson, B. M., & Horne, K. 2004, Astronomische Nachrichten,

325, 248Peterson, B. M., Pogge, R. W., Wanders, I., Smith, S. M., &

Romanishin, W. 1995, PASP, 107, 579Peterson, B. M., Wanders, I., Bertram, R., Hunley, J. F., Pogge,

R. W., & Wagner, R. M. 1998, ApJ, 501, 82Pogge, R. W., & Peterson, B. M. 1992, AJ, 103, 1084Reichert, G. A., et al. 1994, ApJ, 425, 582Rosenblatt, E. I., & Malkan, M. A. 1990, ApJ, 350, 132Santos-Lleo, M., et al. 1997, ApJS, 112, 271Schlafly, E. F., & Finkbeiner, D. P. 2011, ApJ, 737, 103Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ, 500,

525Sergeev, S. G., Doroshenko, V. T., Golubinskiy, Y. V.,

Merkulova, N. I., & Sergeeva, E. A. 2005, ApJ, 622, 129Shappee, B. J., et al. 2014, ApJ, 788, 48Shen, Y., et al. 2016, ApJ, 818, 30Skielboe, A., Pancoast, A., Treu, T., Park, D., Barth, A. J., &

Bentz, M. C. 2015, MNRAS, 454, 144Stirpe, G. M., et al. 1994, ApJ, 425, 609Ulrich, M. H., et al. 1984, MNRAS, 206, 221Ulrich, M.-H., & Horne, K. 1996, MNRAS, 283, 748van der Marel, R. P., & Franx, M. 1993, ApJ, 407, 525van Dokkum, P. G. 2001, PASP, 113, 1420van Groningen, E., & Wanders, I. 1992, PASP, 104, 700Vasudevan, R. V., Fabian, A. C., Gandhi, P., Winter, L. M., &

Mushotzky, R. F. 2010, MNRAS, 402, 1081Veron-Cetty, M.-P., Joly, M., & Veron, P. 2004, A&A, 417, 515Vestergaard, M., & Peterson, B. M. 2006, ApJ, 641, 689Wandel, A., Peterson, B. M., & Malkan, M. A. 1999, ApJ, 526,

579Wanders, I., et al. 1995, ApJ, 453, L87—. 1997, ApJS, 113, 69Watson, L. C., Mathur, S., & Grupe, D. 2007, AJ, 133, 2435Welsh, W. F. 1999, PASP, 111, 1347Welsh, W. F., & Horne, K. 1991, ApJ, 379, 586Welsh, W. F., Martino, D. L., Kawaguchi, G., & Kollatschny, W.

2007, in Astronomical Society of the Pacific Conference Series,Vol. 373, The Central Engine of Active Galactic Nuclei, ed.L. C. Ho & J.-W. Wang, 29

White, R. J., & Peterson, B. M. 1994, PASP, 106, 879Whittle, M. 1992, ApJS, 79, 49Woo, J.-H., et al. 2010, ApJ, 716, 269Woo, J.-H., Yoon, Y., Park, S., Park, D., & Kim, S. C. 2015,

ApJ, 801, 38Zu, Y., Kochanek, C. S., Koz lowski, S., & Udalski, A. 2013, ApJ,

765, 106Zu, Y., Kochanek, C. S., & Peterson, B. M. 2011, ApJ, 735, 80