arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

11
DRAFT VERSION MAY 22, 2019 Preprint typeset using L A T E X style emulateapj v. 12/16/11 THE NANOGRAV 11-YEAR DATA SET: LIMITS ON GRAVITATIONAL WAVES FROM INDIVIDUAL SUPERMASSIVE BLACK HOLE BINARIES K. AGGARWAL 1,2 , Z. ARZOUMANIAN 3 , P. T. BAKER 1,2 , A. BRAZIER 4,5 , M. R. BRINSON 6 , P. R. BROOK 1,2 , S. BURKE-SPOLAOR 1,2 , S. CHATTERJEE 4 , J. M. CORDES 4 , N. J. CORNISH 7 , F. CRAWFORD 8 , K. CROWTER 9 , H. T. CROMARTIE 10 , M. DECESAR 11 , P. B. DEMOREST 12 , T. DOLCH 13 , J. A. ELLIS 1,2,14 , R. D. FERDMAN 15 , E. FERRARA 16 , E. FONSECA 17 , N. GARVER-DANIELS 1,2 , P. GENTILE 1,2 , J. S. HAZBOUN 18 , A. M. HOLGADO 19 , E. A. HUERTA 19 , K. I SLO 6 , R. J ENNINGS 4 , G. J ONES 20 , M. L. J ONES 1,2 , A. R. KAISER 1,2 , D. L. KAPLAN 6 , L. Z. KELLEY 21 , J. S. KEY 18 , M. T. LAM 1,2 , T. J. W. LAZIO 22,23 , L. LEVIN 24 , D. R. LORIMER 1,2 , J. LUO 17 , R. S. LYNCH 25 , D. R. MADISON 1,2 , M. A. MCLAUGHLIN 1,2 , S. T. MCWILLIAMS 1,2 , C. M. F. MINGARELLI 26 , C. NG 9 , D. J. NICE 11 , T. T. PENNUCCI 1,2,27,28 , N. S. POL 1,2 , S. M. RANSOM 10,29 , P. S. RAY 30 , X. SIEMENS 6 , J. SIMON 22,23 , R. SPIEWAK 6,31 , I. H. STAIRS 9 , D. R. STINEBRING 32 , K. STOVALL 12 , J. SWIGGUM 6 , S. R. TAYLOR 22,23 , J. E. TURNER 1,2,6 , M. VALLISNERI 22,23 , R. VAN HAASTEREN 22 , S. J. VIGELAND §6 , C. A. WITT 1,2 , W. W. ZHU 33 (THE NANOGRAV COLLABORATION) Author order alphabetical by surname 1 Department of Physics and Astronomy, West Virginia University, P.O. Box 6315, Morgantown, WV 26506, USA 2 Center for Gravitational Waves and Cosmology, West Virginia University, Chestnut Ridge Research Building, Morgantown, WV 26505, USA 3 X-Ray Astrophysics Laboratory, NASA Goddard Space Flight Center, Code 662, Greenbelt, MD 20771, USA 4 Department of Astronomy, Cornell University, Ithaca, NY 14853, USA 5 Cornell Center for Advanced Computing, Ithaca, NY 14853, USA 6 Center for Gravitation, Cosmology and Astrophysics, Department of Physics, University of Wisconsin-Milwaukee, P.O. Box 413, Milwaukee, WI 53201, USA 7 Department of Physics, Montana State University, Bozeman, MT 59717, USA 8 Department of Physics and Astronomy, Franklin & Marshall College, P.O. Box 3003, Lancaster, PA 17604, USA 9 Department of Physics and Astronomy, University of British Columbia, 6224 Agricultural Road, Vancouver, BC V6T 1Z1, Canada 10 University of Virginia, Department of Astronomy, P.O. Box 400325, Charlottesville, VA 22904, USA 11 Department of Physics, Lafayette College, Easton, PA 18042, USA 12 National Radio Astronomy Observatory, 1003 Lopezville Rd., Socorro, NM 87801, USA 13 Department of Physics, Hillsdale College, 33 E. College Street, Hillsdale, Michigan 49242, USA 14 Infinia ML, 202 Rigsbee Avenue, Durham, NC 27701, USA 15 Department of Physics, University of East Anglia, Norwich, UK 16 NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA 17 Department of Physics, McGill University, 3600 University St., Montreal, QC H3A 2T8, Canada 18 University of Washington Bothell, 18115 Campus Way NE, Bothell, WA 98011, USA 19 NCSA and Department of Astronomy, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, USA 20 Department of Physics, Columbia University, New York, NY 10027, USA 21 Center for Interdisciplinary Exploration and Research in Astrophysics (CIERA), Northwestern University, Evanston, IL 60208 22 Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109, USA 23 Theoretical AstroPhysics Including Relativity (TAPIR), MC 350-17, California Institute of Technology, Pasadena, California 91125, USA 24 Jodrell Bank Centre for Astrophysics, University of Manchester, Manchester, M13 9PL, United Kingdom 25 Green Bank Observatory, P.O. Box 2, Green Bank, WV 24944, USA 26 Center for Computational Astrophysics, Flatiron Institute, 162 Fifth Avenue, New York, NY 10010, USA 27 Institute of Physics, Eötvös Loránd University, Pázmány P. s. 1/A, 1117 Budapest, Hungary 28 Hungarian Academy of Sciences MTA-ELTE Extragalactic Astrophysics Research Group, 1117 Budapest, Hungary 29 National Radio Astronomy Observatory, 520 Edgemont Road, Charlottesville, VA 22903, USA 30 Naval Research Laboratory, Washington DC 20375, USA 31 Centre for Astrophysics and Supercomputing, Swinburne University of Technology, PO Box 218, Hawthorn, VIC 3122, Australia 32 Department of Physics and Astronomy, Oberlin College, Oberlin, OH 44074, USA 33 CAS Key Laboratory of FAST, Chinese Academy of Science, Beijing 100101, China and NANOGrav Physics Frontiers Center Postdoctoral Fellow Draft version May 22, 2019 ABSTRACT Observations indicate that nearly all galaxies contain supermassive black holes (SMBHs) at their centers. When galaxies merge, their component black holes form SMBH binaries (SMBHBs), which emit low-frequency gravitational waves (GWs) that can be detected by pulsar timing arrays (PTAs). We have searched the North American Nanohertz Observatory for Gravitational Waves (NANOGrav) 11-year data set for GWs from indi- vidual SMBHBs in circular orbits. As we did not find strong evidence for GWs in our data, we placed 95% upper limits on the strength of GWs from such sources. At f gw = 8nHz, we placed a sky-averaged upper limit of h 0 < 7.3(3) × 10 -15 . We also developed a technique to determine the significance of a particular signal in each pulsar using “dropout” parameters as a way of identifying spurious signals. From these upper limits, we ruled out SMBHBs emitting GWs with f gw = 8nHz within 120 Mpc for M = 10 9 M , and within 5.5 Gpc for M = 10 10 M at our most-sensitive sky location. We also determined that there are no SMBHBs with M > 1.6 × 10 9 M emitting GWs with f gw =2.8 - 317.8nHz in the Virgo Cluster. Finally, we compared our strain upper limits to simulated populations of SMBHBs, based on galaxies in Two Micron All-Sky Survey (2MASS) and merger rates from the Illustris cosmological simulation project, and found that only 34 out of arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

Transcript of arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

Page 1: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

DRAFT VERSION MAY 22, 2019Preprint typeset using LATEX style emulateapj v. 12/16/11

THE NANOGRAV 11-YEAR DATA SET: LIMITS ON GRAVITATIONAL WAVES FROM INDIVIDUAL SUPERMASSIVEBLACK HOLE BINARIES

K. AGGARWAL1,2 , Z. ARZOUMANIAN

3 , P. T. BAKER1,2 , A. BRAZIER

4,5 , M. R. BRINSON6 , P. R. BROOK

1,2 , S. BURKE-SPOLAOR1,2 ,

S. CHATTERJEE4 , J. M. CORDES

4 , N. J. CORNISH7 , F. CRAWFORD

8 , K. CROWTER9 , H. T. CROMARTIE

10 , M. DECESAR†11 ,

P. B. DEMOREST12 , T. DOLCH

13 , J. A. ELLIS†1,2,14 , R. D. FERDMAN

15 , E. FERRARA16 , E. FONSECA

17 , N. GARVER-DANIELS1,2 ,

P. GENTILE1,2 , J. S. HAZBOUN

†18 , A. M. HOLGADO19 , E. A. HUERTA

19 , K. ISLO6 , R. JENNINGS

4 , G. JONES20 , M. L. JONES

1,2 ,A. R. KAISER

1,2 , D. L. KAPLAN6 , L. Z. KELLEY

21 , J. S. KEY18 , M. T. LAM

†1,2 , T. J. W. LAZIO22,23 , L. LEVIN

24 , D. R. LORIMER1,2 ,

J. LUO17 , R. S. LYNCH

25 , D. R. MADISON1,2 , M. A. MCLAUGHLIN

1,2 , S. T. MCWILLIAMS1,2 , C. M. F. MINGARELLI

26 , C. NG9 ,

D. J. NICE11 , T. T. PENNUCCI

†1,2,27,28 , N. S. POL1,2 , S. M. RANSOM

10,29 , P. S. RAY30 , X. SIEMENS

6 , J. SIMON22,23 , R. SPIEWAK

6,31 ,I. H. STAIRS

9 , D. R. STINEBRING32 , K. STOVALL

†12 , J. SWIGGUM†6 , S. R. TAYLOR

†22,23 , J. E. TURNER1,2,6 , M. VALLISNERI

22,23 ,R. VAN HAASTEREN

22 , S. J. VIGELAND§†6 , C. A. WITT

1,2 , W. W. ZHU33 (THE NANOGRAV COLLABORATION)⋆

⋆Author order alphabetical by surname1Department of Physics and Astronomy, West Virginia University, P.O. Box 6315, Morgantown, WV 26506, USA

2Center for Gravitational Waves and Cosmology, West Virginia University, Chestnut Ridge Research Building, Morgantown, WV 26505, USA3X-Ray Astrophysics Laboratory, NASA Goddard Space Flight Center, Code 662, Greenbelt, MD 20771, USA

4Department of Astronomy, Cornell University, Ithaca, NY 14853, USA5Cornell Center for Advanced Computing, Ithaca, NY 14853, USA

6Center for Gravitation, Cosmology and Astrophysics, Department of Physics, University of Wisconsin-Milwaukee,P.O. Box 413, Milwaukee, WI 53201, USA

7Department of Physics, Montana State University, Bozeman, MT 59717, USA8Department of Physics and Astronomy, Franklin & Marshall College, P.O. Box 3003, Lancaster, PA 17604, USA

9Department of Physics and Astronomy, University of British Columbia, 6224 Agricultural Road, Vancouver, BC V6T 1Z1, Canada10University of Virginia, Department of Astronomy, P.O. Box 400325, Charlottesville, VA 22904, USA

11Department of Physics, Lafayette College, Easton, PA 18042, USA12National Radio Astronomy Observatory, 1003 Lopezville Rd., Socorro, NM 87801, USA

13Department of Physics, Hillsdale College, 33 E. College Street, Hillsdale, Michigan 49242, USA14Infinia ML, 202 Rigsbee Avenue, Durham, NC 27701, USA

15Department of Physics, University of East Anglia, Norwich, UK16NASA Goddard Space Flight Center, Greenbelt, MD 20771, USA

17Department of Physics, McGill University, 3600 University St., Montreal, QC H3A 2T8, Canada18University of Washington Bothell, 18115 Campus Way NE, Bothell, WA 98011, USA

19NCSA and Department of Astronomy, University of Illinois at Urbana-Champaign, Urbana, Illinois 61801, USA20Department of Physics, Columbia University, New York, NY 10027, USA

21Center for Interdisciplinary Exploration and Research in Astrophysics (CIERA), Northwestern University, Evanston, IL 6020822Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109, USA

23Theoretical AstroPhysics Including Relativity (TAPIR), MC 350-17, California Institute of Technology, Pasadena, California 91125, USA24Jodrell Bank Centre for Astrophysics, University of Manchester, Manchester, M13 9PL, United Kingdom

25Green Bank Observatory, P.O. Box 2, Green Bank, WV 24944, USA26Center for Computational Astrophysics, Flatiron Institute, 162 Fifth Avenue, New York, NY 10010, USA

27Institute of Physics, Eötvös Loránd University, Pázmány P. s. 1/A, 1117 Budapest, Hungary28Hungarian Academy of Sciences MTA-ELTE Extragalactic Astrophysics Research Group, 1117 Budapest, Hungary

29National Radio Astronomy Observatory, 520 Edgemont Road, Charlottesville, VA 22903, USA30Naval Research Laboratory, Washington DC 20375, USA

31Centre for Astrophysics and Supercomputing, Swinburne University of Technology, PO Box 218, Hawthorn, VIC 3122, Australia32Department of Physics and Astronomy, Oberlin College, Oberlin, OH 44074, USA

33CAS Key Laboratory of FAST, Chinese Academy of Science, Beijing 100101, China and† NANOGrav Physics Frontiers Center Postdoctoral Fellow

Draft version May 22, 2019

ABSTRACT

Observations indicate that nearly all galaxies contain supermassive black holes (SMBHs) at their centers. Whengalaxies merge, their component black holes form SMBH binaries (SMBHBs), which emit low-frequencygravitational waves (GWs) that can be detected by pulsar timing arrays (PTAs). We have searched the NorthAmerican Nanohertz Observatory for Gravitational Waves (NANOGrav) 11-year data set for GWs from indi-vidual SMBHBs in circular orbits. As we did not find strong evidence for GWs in our data, we placed 95%upper limits on the strength of GWs from such sources. At fgw = 8nHz, we placed a sky-averaged upper limitof h0 < 7.3(3)× 10−15. We also developed a technique to determine the significance of a particular signal ineach pulsar using “dropout” parameters as a way of identifying spurious signals. From these upper limits, weruled out SMBHBs emitting GWs with fgw = 8nHz within 120 Mpc for M = 109 M⊙, and within 5.5 Gpcfor M = 1010 M⊙ at our most-sensitive sky location. We also determined that there are no SMBHBs withM > 1.6× 109 M⊙ emitting GWs with fgw = 2.8 − 317.8nHz in the Virgo Cluster. Finally, we compared ourstrain upper limits to simulated populations of SMBHBs, based on galaxies in Two Micron All-Sky Survey(2MASS) and merger rates from the Illustris cosmological simulation project, and found that only 34 out of

arX

iv:1

812.1

1585v3 [a

stro

-ph.G

A]

21 M

ay 2

019

Page 2: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

2 THE NANOGRAV COLLABORATION

75,000 realizations of the local Universe contained a detectable source.Keywords: Gravitational waves – Methods: data analysis – Pulsars: general

1. INTRODUCTION

Pulsar timing arrays (PTAs) seek to detect gravitationalwaves (GWs) by searching for correlations in the timing ob-servations of a collection of millisecond pulsars (MSPs). Thestability of MSPs over long timescales (∼ decades) makesPTAs ideal detectors for long-wavelength GWs (see Cordes2013). Currently, there are three PTA experiments in op-eration: the North American Observatory for GravitationalWaves (NANOGrav; McLaughlin 2013), the European PulsarTiming Array (EPTA; Desvignes et al. 2016), and the ParkesPulsar Timing Array (PPTA; Hobbs 2013). Together thesegroups form the International Pulsar Timing Array (IPTA;Verbiest et al. 2016). The NANOGrav collaboration has re-leased three data sets based on five years of observations (De-morest et al. 2013; hereafter NG5a), nine years of observa-tions (Arzoumanian et al. 2015; hereafter NG9a), and 11 yearsof observations (Arzoumanian et al. 2018a; hereafter NG11a).

Potential GW sources in the PTA band include supermas-sive black hole binaries (SMBHBs; see Sesana et al. 2004;Sesana 2013; Burke-Spolaor et al. 2018), primordial GWs(Grishchuk 2005; Lasky et al. 2016), cosmic strings and su-perstrings (Damour & Vilenkin 2001; Ölmez et al. 2010;Blanco-Pillado et al. 2018), and bubble collisions during cos-mological phase transitions (Caprini et al. 2010). Histori-cally, analyses have focused on the stochastic gravitationalwave background (GWB) formed by the ensemble of a cos-mic population of SMBHBs, as models predict that this signalis expected to be detected first (Rosado et al. 2015). In theabsence of a detection, constraints have been placed on theGWB, most recently with the NANOGrav 11-year data set(Arzoumanian et al. 2018b, hereafter NG11b). These limitshave been used to narrow the viable parameter space for bi-nary evolution in dynamic galactic environments (e.g. Tayloret al. 2017; Chen et al. 2017; Middleton et al. 2018) and makestatements about SMBHB population statistics (e.g. Sesanaet al. 2018, Holgado et al. 2018).

PTAs are also sensitive to GWs emitted from nearby in-dividual SMBHBs with periods on the order of months toyears, total masses of ∼ 108

− 1010M⊙, and orbital separa-tions of ∼ 10−3

− 10−1 pc, depending on the total mass of thebinary. SMBHBs that are emitting in the PTA band havenearly-constant orbital frequencies, and hence the GWs fromthese sources are referred to as “continuous waves” (CWs).However, we do account for the evolution of their orbits overthe span of our observations in our analyses. Although therehas not yet been a detection of GWs from individual sourceswith PTAs, they have already been used to place limits on themasses of candidate SMBHBs (e.g. Jenet et al. 2004; Schutz& Ma 2016). Simulations predict that individual sources willbe observed by PTAs within the next 10 – 20 years (Rosadoet al. 2015; Mingarelli et al. 2017; Kelley et al. 2018).

In this paper, we present the results of searches forGWs from individual circular SMBHBs performed on theNANOGrav 11-year data set. This is an extension of Arzou-manian et al. 2014 (hereafter NG5b), which performed a sim-ilar analysis on the NANOGrav 5-year data set. Our approachwas also inspired by searches performed by the PPTA and

§ Corresponding author email: [email protected]

EPTA. The first all-sky search for GWs from individual SMB-HBs was performed by the PPTA in Yardley et al. (2010),and a later analysis was published in Zhu et al. (2014). Themost recent limits on GWs from individual SMBHBs comesfrom the EPTA (Babak et al. 2016), which performed bothBayesian and frequentist searches for GWs, and placed upperlimits as a function of GW frequency and sky location.

The paper is organized as follows. In Sec. 2, we review thepulsar observations and data reduction techniques used in thecreation of the data sets. In Sec. 3, we describe the GW signalmodel and noise models used in our search pipelines. We alsodescribe the Bayesian and frequentist methods and software.In Sec. 4, we present the results of detection searches. Aswe did not find evidence for GWs in the 11-year data set, weplaced upper limits on the GW strain for fgw = 2.8−317.8nHz.We also discuss a new analysis technique for identifying spu-rious signals in PTA data. In Sec. 5 we present limits onthe distances to individual SMBHBs, and limits on the chirpmasses of potential SMBHBs in the nearby Virgo Cluster.We also compare our current sensitivity to simulations ofSMBHB populations, and estimate the expected number ofdetectable sources. We conclude in Sec. 6. Throughout thispaper, we use units where G = c = 1.

2. THE 11-YEAR DATA SET

We analyzed the NANOGrav 11-year data set, which waspublished in NG11a and consisted of times of arrival (TOAs)for 45 pulsars with observations made between 2004 and2015. Some of these data were previously published as theNANOGrav 5-year data set in NG5a and the NANOGrav 9-year data set in NG9a. We briefly review the observations anddata reduction techniques here – further details can be foundin NG11a.

We made observations using two radio telescopes: the 100-m Robert C. Byrd Green Bank Telescope (GBT) of the GreenBank Observatory in Green Bank, West Virginia; and the 305-m William E. Gordon Telescope (Arecibo) of Arecibo Obser-vatory in Arecibo, Puerto Rico. Since Arecibo is more sensi-tive than GBT, all pulsars that could be observed from Arecibo(0 < δ < 39) were observed with it, while those outside ofArecibo’s declination range were observed with GBT. Twopulsars were observed with both telescopes: PSR J1713+0747and PSR B1937+21. We observed most pulsars once a month.In addition, we started a high-cadence observing campaign in2013, in which we made weekly observations of two pulsarswith GBT (PSR J1713+0747 and PSR J1909−3744) and fivepulsars with Arecibo (PSR J0030+0451, PSR J1640+2224,PSR J1713+0747, PSR J2043+1711, and PSR J2317+1439).This high-cadence observing campaign was specifically de-signed to increase the sensitivity of our PTA to GWs fromindividual sources (Burt et al. 2011; Christy et al. 2014).

In most cases, we observed pulsars at every epoch with tworeceivers at different frequencies in order to measure the pulsedispersion due to the interstellar medium (ISM). At GBT, themonthly observations used the 820 MHz and 1.4 GHz re-ceivers. The weekly observations used only the 1.4 GHz re-ceiver, which has a wide enough bandwidth to measure thedispersion. At Arecibo, four receivers were used for thisproject (327 MHz, 430 MHz, 1.4 GHz, and 2.3 GHz); eachpulsar was observed with two different receivers, which were

Page 3: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

NANOGRAV 11-YEAR CONTINUOUS WAVE CONSTRAINTS 3

chosen based on the spectral index and timing characteristicsof that pulsar. Backend instrumentation was upgraded aboutmidway through our project. Initially, data at Arecibo andGBT were recorded using the ASP and GASP systems, re-spectively, which had bandwidths of 64 MHz. Between 2010and 2012, we transitioned to the wideband systems PUPPI andGUPPI, which had bandwidths up to 800 MHz. Instrumentaloffsets between the data acquisition systems at each observa-tory were measured with high precision and were removedfrom the data to allow for seamless data sets (see NG9a fordetails).

For each pulsar, the observed TOAs were fit to a timingmodel that described the pulsar’s spin period and spin periodderivative, sky location, proper motion, and parallax. Thetiming model also included terms describing pulse dispersionalong the line of sight. Additionally, for those pulsars in bina-ries the timing model also included five Keplerian parametersthat described the binary orbit, and additional post-Keplerianparameters that described relativistic binary effects if they im-proved the timing fit. In the GW analyses, we used a lin-earized timing model centered around the best-fit parametervalues.

3. DATA ANALYSIS METHODS

PTAs are sensitive to GWs through their effect on the tim-ing residuals. We can write the residuals for each pulsar δtas

δt = Mǫ+ nwhite + nred + s , (1)

where M is the design matrix, which describes the linearizedtiming model, ǫ is a vector of the timing model parameter off-sets, nwhite is a vector describing white noise, nred is a vectordescribing red noise, and s is a vector of the residuals inducedby a GW. In this section, we briefly discuss the signal model,likelihood, and methods used in our analyses. These are allsimilar to those used in NG5b – in the discussion that fol-lows, we emphasize areas in which this analysis differs fromprevious ones.

3.1. Signal and noise models

Consider a GW source whose location in equatorial coor-dinates is given by declination δ and right ascension α. It isconvenient to write the sky position in terms of the polar an-gle θ and azimuthal angle φ, which are related to δ and α byθ = π/2 − δ and φ = α. The emitted GWs can be written interms of two polarizations:

hab(t, Ω) = e+

ab(Ω) h+(t, Ω) + e×ab(Ω) h×(t, Ω) , (2)

where Ω is a unit vector from the GW source to the Solar Sys-tem barycenter (SSB), h+,× are the polarization amplitudes,and e

+,×ab are the polarization tensors. The polarization tensors

can be written in the SSB frame as (Wahlquist 1987)

e+

ab(Ω) = ma mb − na nb , (3)

e×ab(Ω) = ma nb + na mb , (4)

where

Ω= −sinθ cosφ x − sinθ sinφ y − cosθ z , (5)

m = −sinφ x + cosφ y , (6)

n = −cosθ cosφ x − cosθ sinφ y + sinθ z . (7)

The response of a pulsar to the source is described by the an-tenna pattern functions F+ and F× (Sesana & Vecchio 2010;

Ellis et al. 2012; Taylor et al. 2016),

F+(Ω) =1

2

(m · p)2− (n · p)2

1 + Ω · p, (8)

F×(Ω) =(m · p)(n · p)

1 + Ω · p, (9)

where p is a unit vector pointing from the Earth to the pulsar.The effect of a GW on a pulsar’s residuals can be written as

s(t, Ω) = F+(Ω) ∆s+(t) + F×(Ω) ∆s×(t) , (10)

where ∆s+,× is the difference between the signal induced atthe Earth and at the pulsar (the so-called “Earth term” and“pulsar term”),

∆s+,×(t) = s+,×(tp) − s+,×(t) , (11)

where t is the time at which the GW passes the SSB and tp isthe time at which it passes the pulsar. From geometry, we canrelate t and tp by

tp = t − L(1 + Ω · p) , (12)

where L is the distance to the pulsar.For a circular binary, at zeroth post-Newtonian (0-PN) or-

der, s+,× is given by (Wahlquist 1987; Lee et al. 2011; Corbin& Cornish 2010)

s+(t) =M5/3

dLω(t)1/3

[

−sin2Φ(t)(

1 + cos2 i)

cos2ψ

−2cos2Φ(t) cos i sin2ψ] , (13)

s×(t) =M5/3

dLω(t)1/3

[

−sin2Φ(t)(

1 + cos2 i)

sin2ψ

+2cos2Φ(t) cos i cos2ψ] , (14)

where i is the inclination angle of the SMBHB, ψ is the GWpolarization angle, dL is the luminosity distance to the source,and M ≡ (m1m2)3/5/(m1 + m2)1/5 is a combination of theblack hole masses m1 and m2 called the “chirp mass.” Notethat the variables M and ω are the observed redshifted values,which are related to the rest-frame values Mr and ωr accord-ing to

Mr =M

1 + z, (15)

ωr =ω(1 + z) . (16)

Currently PTAs are only sensitive to sources in the local Uni-verse for which (1 + z) ≈ 1.

For a circular binary, the orbital angular frequency is relatedto the GW frequency by ω0 = π fgw, where ω0 = ω(t0). For oursearch, we defined the reference time t0 as 31 December 2015(MJD 57387), which corresponded to the last day data weretaken for the 11-year data set. The orbital phase and frequencyof the SMBHB are given by (NG5b)

Φ(t) =Φ0 +

1

32M−5/3

[

ω−5/30 −ω(t)−5/3

]

, (17)

ω(t) =ω0

(

1 −

256

5M5/3ω

8/30 t

)

−3/8

, (18)

where Φ0 and ω0 are the initial orbital phase and frequency,respectively. Unlike in NG5b, we used the full expression forω(t) in our signal model rather than treating the GW frequency

Page 4: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

4 THE NANOGRAV COLLABORATION

at the Earth as a constant, as high-chirp-mass binaries willevolve significantly over the timescale of our observations.

Our noise model for individual pulsars included both whitenoise and red noise. We used the same white noise modelas NG5b, which has three parameters: EFAC, EQUAD, andECORR. The EFAC parameter scales the TOA uncertain-ties, and the EQUAD parameter adds white noise in quadra-ture. The ECORR parameter describes additional white noiseadded in quadrature that is correlated within the same observ-ing epoch, such as pulse jitter (Dolch et al. 2014; Lam et al.2017). We used the improved implementation of ECORRdescribed in NG11b. To model the red noise, we dividedthe noise spectrum into 30 bins spaced linearly between f =1/Tobs and f = 30/Tobs, where Tobs is the total observation timefor a particular pulsar,1 and then fit the power spectral density(PSD) to a power-law model,

P( f ) = A2red

(

f

fyr

)

−γ

, (19)

where fyr ≡ 1/(1 yr), Ared is the amplitude, and γ is the spec-tral index. There are many possible sources of red noisein pulsar timing residuals, including spin noise, variationsin pulse shape, pulsar mode changes, and errors in model-ing pulse dispersion from the ISM (Cordes 2013; Lam et al.2017; Jones et al. 2017). We model time-variations in theISM through DMX parameters, which measure the dispersionat almost every observing epoch (NG9a, Lam et al. 2016).

3.2. Bayesian methods and software

We used Bayesian inference to determine posterior dis-tributions of GW parameters from our data. The proce-dure followed closely that of NG5b, with the addition ofthe BAYESEPHEM model for the uncertainty in the SSB in-troduced in NG11b. Pulsar timing uses a Solar Systemephemeris (SSE) to transform from individual observatories’reference frames to an inertial reference frame centered at theSSB. We used DE436 (Folkner & Park 2016) to perform thistransformation, plus the BAYESEPHEM model. Uncertainty inthe SSE has a significant impact on the computation of GWupper limits from PTA data. The BAYESEPHEM model mit-igates this by marginalizing over perturbations in the outerplanets’ masses and Jupiter’s orbit. This approach removessystematic uncertainty in the position of the SSB by introduc-ing statistical uncertainty through the addition of new parame-ters. Another approach to differentiating between GW signalsand uncertainty in the SSE, which we do not explore in thispaper, is to exploit the fact that the two have different spatialcorrelations (Tiburzi et al. 2016). A detailed analysis of howerrors in the SSE effect PTAs can be found in Caballero et al.(2018).

We used the same likelihood as in NG5b. We implementedthe likelihood and priors and performed the searches usingNANOGrav’s new software package enterprise.2 Weconfirmed the accuracy of this package by also performingsome searches using the software package PAL2,3 which hasbeen used for previous NANOGrav GW searches. Both pack-

1 van Haasteren & Vallisneri (2015) introduced a better method for choos-ing the frequency basis for red noise, which reduces the computational cost.However, in this work chose a linear frequency basis to make it easier tocompare these results with the results in NG5b.

2 https://github.com/nanograv/enterprise3 https://github.com/jellis18/PAL2

ages used the Markov Chain Monte Carlo (MCMC) samplerPTMCMCSampler4 to explore the parameter space.

For detection and upper-limit runs, we described the Earth-term contribution to the GW signal by eight parameters:

λ0 = θ,φ,Φ0,ψ, i,M, fgw,h0 , (20)

where the characteristic strain h0 is related to M, fgw, and dL

according to

h0 =2M5/3(π fgw)2/3

dL

. (21)

We used log-uniform priors on h0 for detection analyses,and a uniform prior on h0 to compute upper limits on thestrain. For both types of analyses, we searched over log10 h0 ∈[−18,−11].

We used isotropic priors on the sky position of the source(θ,φ), source inclination angle i, GW polarization angle ψ,and GW phase Φ0. We searched over log10M with a uniformprior log10(M/M⊙) ∈ [7,10]. For high fgw, we truncated theprior on log10M to account for the fact that high-chirp-masssystems will have merged before emitting high-frequencyGWs. Assuming binaries merge when the orbital frequency isequal to the innermost stable circular orbit (ISCO) frequency,M must satisfy

M≤ 1

63/2π fgw

[

q

(1 + q)2

]3/5

, (22)

where q is the mass ratio. For our analyses, we used the chirp-mass cutoff with q = 1. This change to the prior on M onlyaffected fgw ≥ 191.3nHz.

We performed searches at fixed values of fgw. The min-imum GW frequency was set by the total observation time,fgw = 1/(11.4 yrs) = 2.8nHz. The maximum GW frequencywas set by the observing cadence. Because of the high-cadence observing campaign, the 11-year data set can detectGWs with frequencies up to 826.7nHz; however, the data arenot very sensitive at high frequencies. Also, we do not ex-pect to find any SMBHBs with orbital periods of weeks be-cause high-chirp-mass systems would have already mergedbefore emitting at those frequencies, and low-chirp-mass sys-tems would be evolving through the PTA band very quicklyat that point. Therefore, we only searched for GWs with fre-quencies up to 317.8nHz, which corresponded to the high-frequency-cutoff adopted in NG5b.

The pulsar-term contributions to the GW signal used thepulsar distances to compute the light-travel-time betweenwhen the GW passed the pulsars and when it passed the SSB(see Eq. (12)). We used a Gaussian prior on the distanceswith the measured mean and uncertainty from Verbiest et al.(2012); for the pulsars not included in that paper, we used amean of 1 kpc and error of 20%. The use of approximate dis-tances for some pulsars did not seem to affect our results –we found that the pulsar distance posteriors were identical tothe priors, indicating that the data are insensitive to the pulsardistances. Furthermore, we only used approximate pulsar dis-tances for pulsars that had only been observed for a few years,and therefore did not contribute significantly to the GW sen-sitivity. The phase at the pulsar can be written as

Φ(t) = Φ0 +Φp +

1

32M−5/3

[

ω(tp,0)−5/3−ω(tp)−5/3

]

, (23)

4 https://github.com/jellis18/PTMCMCSampler

Page 5: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

NANOGRAV 11-YEAR CONTINUOUS WAVE CONSTRAINTS 5

where Φp is the phase difference between the Earth and thepulsar. The pulsar phase parameters Φp can be computed fromthe pulsar distances and chirp mass as

Φp =1

32M−5/3

[

ω−5/30 −ω(tp,0)−5/3

]

; (24)

however, in most cases the pulsar distance uncertainties(∆L ∼ 10 − 100 pc) are significantly greater than the GWwavelengths (λgw ∼ 0.1 − 10 pc), and so the phase differencesbetween the Earth terms and pulsar terms are effectively ran-dom. Therefore, following the approach of Corbin & Cornish(2010), we treated Φp as an independent parameter with a uni-form prior Φp ∈ [0,2π].

We fixed the white noise parameters to their best-fit val-ues, as determined from noise analyses performed on in-dividual pulsars. In the GW analyses, we simultaneouslysearched over the individual pulsars’ red noise using a power-law model with uniform priors on log10 Ared ∈ [−20,−11] andγ ∈ [0,7]. In order to burn-in the red noise and BAYESEPHEM

parameters efficiently, we introduced jump proposals thatdrew proposed samples from empirical distributions based onthe posteriors from an initial Bayesian analysis with only thepulsars’ red noise and BAYESEPHEM (i.e., excluding a GWsignal). For more details, see Appendix A.

We computed Bayes factors for the presence of a GW signalusing the Savage-Dickey formula (Dickey 1971),

B10 ≡evidence[H1]

evidence[H0]=

p(h0 = 0|H1)

p(h0 = 0|D,H1), (25)

where H1 is the model with a GW signal plus individual pul-sar red noise, H0 is the model with only individual pulsarred noise, p(h0 = 0|H1) is the prior volume at h0 = 0, andp(h0 = 0|D,H1) is the posterior volume at h0 = 0. We wereable to use the Savage-Dickey formula because H1 and H0are nested models, i.e., H0 is H1 : h0 = 0. We approximatedp(h0 = 0|D,H1) as the fraction of quasi-indepdent samples inthe lowest-amplitude bin of a histogram of h0. We found thequasi-independent samples by thinning the chain by the auto-correlation chain length, which is a measure of how far aparttwo samples in the chain must be in order to be statisticallyindependent. We computed the error in the Bayes factor as

σ =B10√

n, (26)

where n is the number of samples in the lowest-amplitude bin.For upper limits, following the approach of NG11b, we

computed the standard error as

σ =

x(1 − x)/Ns

p(h0 = h95%0 |D)

, (27)

where x = 0.95 and Ns is the number of effective samples in thechain. This definition of σ is the error in the computed 95%upper limit due to using a finite number of samples. We es-timated the number of effective samples by dividing the totalnumber of samples by the autocorrelation chain length, whichis a measure of how far apart two samples in the chain mustbe in order to be statistically independent.

3.3. Fp-statistic

As in NG5b, we also performed a frequentist analysis withthe Fp-statistic, which we computed using the software pack-age enterprise. The Fp-statistic is an incoherent detec-

tion statistic that is derived by maximizing the log of the like-lihood ratio (Ellis et al. 2012). Essentially, it is the weightedsum of the power spectrum of the residuals, summed overall pulsars. This statistic assumes the SMBHB’s orbital fre-quency is not evolving significantly over the timescale of ourobservations. In the absence of a signal, 2Fp follows a chi-squared distribution with 2Np degrees of freedom, where Np

is the number of pulsars. The corresponding false-alarm-probability (FAP) is

PF (Fp,0) = exp(−Fp,0)

Np−1∑

k=0

F kp,0

k!. (28)

In performing GW searches over our entire frequency range,we compute the Fp statistic N f times, where N f is the numberof independent frequencies, i.e., the number of frequenciesseparated by 1/Tobs = 2.7nHz. The FAP for the entire searchis

PTF (Fp,0) = 1 −

[

1 − PF (Fp,0)]N f

. (29)

For the analysis in this paper, N f = 115.

4. RESULTS

In this section we report the results of both detection andupper limit analyses of the NANOGrav 11-year data set forGWs from individual circular SMBHBs. We used the data toplace upper limits as a function of frequency and sky loca-tion, and to compare upper limits from the 11-year data setto those from the 5- and 9-year data sets. We also discuss anew Bayesian technique to determine how much each pulsarin a PTA contributes to a common signal in order to diagnosespurious signals. Following the approach of NG11b, our anal-yses of the 11-year data set only used the 34 pulsars which hadbeen observed for at least three years. Our analyses of the 5-and 9-year data sets used the same subset of pulsars that wereused in the corresponding analyses for the GWB (NG5a, Ar-zoumanian et al. 2016), which included 17 and 18 pulsars,respectively.

4.1. Detection analyses

We performed detection searches for GWs from individualcircular SMBHBs on the 11-year data set. Figure 1 showsthe Bayes factors for each frequency, marginalized over thesky location. We did not find strong evidence for GWs inthe 11-year data set. The largest Bayes factor was at fgw =109nHz, for which B10 = 15(6). For all other frequencies, theBayes factors were between B10 = 0.449(4) and B10 = 1.4(3),indicating no evidence of GWs in the data.

We also used the Fp-statistic to determine the significanceof a GW signal. Figure 2 shows the Fp-statistic as a func-tion of fgw, and the corresponding FAP as computed fromEq. (28). There are no frequencies for which the FAP liesbelow our detection threshold of 10−4. At the GW frequencythat maximizes Fp, the total FAP for the search as computedfrom Eq. (29) is PT

F (Fp,0) = 0.543. Thus we concluded thatthe frequentist analyses also found that the 11-year data setdoes not contain significant evidence for GWs.

Although the detection search at fgw = 109nHz found ahigher Bayes factor than any of the other values of fgw, weemphasize that the Bayes factor is not high enough to claima detection. A Bayes factor of 15 means 15:1 odds for thepresence of a GW signal; similarly, at this frequency the Fp-statistic corresponds to PF (Fp,0) = 0.235, or signal-to-noise

Page 6: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019
Page 7: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019
Page 8: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019
Page 9: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019
Page 10: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

10 THE NANOGRAV COLLABORATION

jump proposals based on empirical distributions. SJV, KI,JET, KA, PRB, AMH, and NSP ran the searches. MRB, JAE,and SRT performed an initial analysis of the 9-year data set.CMFM, KI, TJWL, JS, and SJV did the astrophysical inter-pretation. SJV, KI, and CMFM wrote the paper and generatedthe plots.

Acknowledgments. We thank the anonymous referee forhelpful comments and suggestions which improved thismanuscript. The NANOGrav project receives support fromNational Science Foundation (NSF) Physics Frontier Centeraward number 1430284. NANOGrav research at UBC is sup-ported by an NSERC Discovery Grant and Discovery Ac-celerator Supplement and by the Canadian Institute for Ad-vanced Research. Portions of this research were carried outat the Jet Propulsion Laboratory, California Institute of Tech-nology, under a contract with the National Aeronautics andSpace Administration. MV and JS acknowledge support fromthe JPL RTD program. SRT was partially supported by anappointment to the NASA Postdoctoral Program at the JetPropulsion Laboratory, administered by Oak Ridge Associ-ated Universities through a contract with NASA. JAE was par-tially supported by NASA through Einstein Fellowship grantsPF4-150120. SBS and CAW were supported by NSF award#1815664. WWZ is supported by the Chinese Academy ofScience Pioneer Hundred Talents Program, the Strategic Pri-ority Research Program of the Chinese Academy of Sciencesgrant No. XDB23000000, the National Natural Science Foun-dation of China grant No. 11690024, and by the AstronomicalBig Data Joint Research Center, co-founded by the NationalAstronomical Observatories, Chinese Academy of Sciencesand the Alibaba Cloud. Portions of this work performed atNRL are supported by the Chief of Naval Research. The Flat-iron Institute is supported by the Simons Foundation.

We thank Logan O’Beirne for helping to uncover a bug inan early implementation of the search algorithms, and for pro-viding comments on the algorithms and manuscript. We thankCasey McGrath for performing some of the CW analyses. Wethank Tingting Liu for helpful discussions. We thank AlbertoSesana, Xinjiang Zhu, and Siyuan Chen for providing usefulcomments on the manuscript.

We are grateful for computational resources provided by theLeonard E Parker Center for Gravitation, Cosmology and As-trophysics at the University of Wisconsin-Milwaukee, whichis supported by NSF Grants 0923409 and 1626190. Thisresearch made use of the Super Computing System (SpruceKnob) at WVU, which is funded in part by the National Sci-ence Foundation EPSCoR Research Infrastructure Improve-ment Cooperative Agreement #1003907, the state of WestVirginia (WVEPSCoR via the Higher Education Policy Com-mission) and WVU. This research also made use of the Illi-nois Campus Cluster, a computing resource that is operatedby the Illinois Campus Cluster Program (ICCP) in conjunc-tion with the National Center for Supercomputing Applica-tions (NCSA) and which is supported by funds from the Uni-versity of Illinois at Urbana-Champaign. Data for this projectwere collected using the facilities of the Green Bank Observa-tory and the Arecibo Observatory. The Green Bank Observa-tory is a facility of the National Science Foundation operatedunder cooperative agreement by Associated Universities, Inc.The Arecibo Observatory is a facility of the National ScienceFoundation operated under cooperative agreement by the Uni-versity of Central Florida in alliance with Yang Enterprises,Inc. and Universidad Metropolitana.

Software: enterprise (Ellis et al. 2017), PAL2 (Ellis

& van Haasteren 2017a), PTMCMCSampler (Ellis & vanHaasteren 2017b)

APPENDIX

A. JUMP PROPOSALS FROM EMPIRICALDISTRIBUTIONS

In our data analysis pipelines, we computed the posteriordistributions using MCMC algorithms, which explore param-eter space through a random walk. The CW search for the11-year data set included 154 parameters: 7 common GWparameters, 68 pulsar-term GW parameters, 68 pulsar rednoise parameters, and 11 BAYESEPHEM parameters. Explor-ing such a large parameter space is computationally intensive,and many iterations are required to burn-in the parameters andensure the chains have converged. Jump proposals determinehow proposed samples are generated, and using particularlygood jump proposals can significantly decrease the burn-inand convergence time. Appendix C of NG5b discusses thejump proposals used in the CW search of the NANOGrav 5-year data set, which were also used in the analyses describedin this paper.

In the course of analyzing the 11-year data set, we intro-duced a new type of jump proposal for the pulsars’ red noiseparameters and the BAYESEPHEM parameters. These jumpproposals chose new parameter values by drawing from em-pirical distributions based on the posteriors from an initialBayesian analysis searching over all of the pulsars that in-cluded only the pulsars’ red noise and BAYESEPHEM; i.e.excluding any GW term. These jump proposals do not alterthe likelihood or the priors – they ensure that the sampler ischoosing new parameters that have a high probability of im-proving the fit, but they do not affect the probability that thenew parameter values will be accepted or rejected.

This initial pilot run included only 79 parameters, andtherefore the red noise and BAYESEPHEM parameters burned-in relatively quickly. We constructed the empirical distribu-tions from histograms of the posteriors, adding one sample toall bins so that the probability density function was nonzeroat every point in the prior. For the red noise parameters,we constructed 2-dimensional empirical distributions for theamplitude log10 A and spectral index γ for each pulsar. Forthe BAYESEPHEM parameters, we constructed 1-dimensionalempirical distributions for each of the six Jupiter orbital ele-ments, which describe perturbations to Jupiter’s orbit.

We have found that including jumps that draw from the em-pirical distributions to the MCMC dramatically reduces thenumber of samples needed for the chains to burn-in and con-verge because the red-noise and BAYESEPHEM parametersconverge almost immediately. Efficiently sampling the pul-sars’ red noise parameters will become increasingly impor-tant as the number of pulsars in our PTA increase, as eachpulsar added to the PTA adds two red noise parameters andtwo pulsar-term parameters to the model.

REFERENCES

Arzoumanian, Z., Brazier, A., Burke-Spolaor, S., et al. 2014, ApJ, 794, 141[1, 3, 3.1, 3.1, 3.2, 1, 3.2, 3.3, A]

—. 2015, ApJ, 813, 65 [1, 2, 3.1]—. 2016, ApJ, 821, 13 [4]—. 2018a, ApJS, 235, 37 [1, 2]Arzoumanian, Z., Baker, P. T., Brazier, A., et al. 2018b, ApJ, 859, 47 [1,

3.1, 3.2, 3.2, 4]Babak, S., Petiteau, A., Sesana, A., et al. 2016, MNRAS, 455, 1665 [1, 4.2,

5]

Page 11: arXiv:1812.11585v3 [astro-ph.GA] 21 May 2019

NANOGRAV 11-YEAR CONTINUOUS WAVE CONSTRAINTS 11

Blanco-Pillado, J. J., Olum, K. D., & Siemens, X. 2018, Physics Letters B,778, 392 [1]

Burke-Spolaor, S., Taylor, S. R., Charisi, M., et al. 2018, arXiv e-prints,arXiv:1811.08826 [1]

Burt, B. J., Lommen, A. N., & Finn, L. S. 2011, ApJ, 730, 17 [2]Caballero, R. N., Guo, Y. J., Lee, K. J., et al. 2018, MNRAS, 481, 5501

[3.2]Caprini, C., Durrer, R., & Siemens, X. 2010, Phys. Rev. D, 82, 063511 [1]Chen, S., Middleton, H., Sesana, A., Del Pozzo, W., & Vecchio, A. 2017,

MNRAS, 468, 404 [1]Christy, B., Anella, R., Lommen, A., et al. 2014, ApJ, 794, 163 [2]Corbin, V., & Cornish, N. J. 2010, ArXiv e-prints, arXiv:1008.1782

[astro-ph.HE] [3.1, 3.2]Cordes, J. M. 2013, Classical and Quantum Gravity, 30, 224002 [1, 3.1]Damour, T., & Vilenkin, A. 2001, Phys. Rev. D, 64, 064008 [1]Demorest, P. B., Ferdman, R. D., Gonzalez, M. E., et al. 2013, ApJ, 762, 94

[1, 2, 4]Desvignes, G., Caballero, R. N., Lentati, L., et al. 2016, MNRAS, 458, 3341

[1, 4.2]Dickey, J. M. 1971, The Annals of Mathematical Statistics, 42, 204 [3.2]Dolch, T., Lam, M. T., Cordes, J., et al. 2014, ApJ, 794, 21 [3.1]Efron, B., & Stein, C. 1981, Ann. Statist., 9, 586 [4.3]Ellis, J., & van Haasteren, R. 2017a, jellis18/PAL2: PAL2 [6]—. 2017b, jellis18/PTMCMCSampler: Official Release [6]Ellis, J. A., Siemens, X., & Creighton, J. D. E. 2012, ApJ, 756, 175 [3.1,

3.3]Ellis, J. A., Vallisneri, M., Taylor, S. R., & Baker, P. T. 2017,

https://github.com/nanograv/enterprise: enterprise [6]Folkner, W. M., & Park, R. S. 2016, JPL planetary and Lunar ephemeris

DE436, Tech. rep., Jet Propulsion Laboratory, Pasadena, CA,https://naif.jpl.nasa.gov/pub/naif/JUNO/kernels/spk/de436s.bsp.lbl [3.2]

Gebhardt, K., Adams, J., Richstone, D., et al. 2011, ApJ, 729, 119 [5]Genel, S., Vogelsberger, M., Springel, V., et al. 2014, MNRAS, 445, 175 [5]Grishchuk, L. P. 2005, Physics Uspekhi, 48, 1235 [1]Hobbs, G. 2013, Classical and Quantum Gravity, 30, 224007 [1]Holgado, A. M., Sesana, A., Sandrinelli, A., et al. 2018, MNRAS, 481, L74

[1]Jenet, F. A., Lommen, A., Larson, S. L., & Wen, L. 2004, ApJ, 606, 799 [1]Jones, M. L., McLaughlin, M. A., Lam, M. T., et al. 2017, ApJ, 841, 125

[3.1]Kelley, L. Z., Blecha, L., Hernquist, L., Sesana, A., & Taylor, S. R. 2018,

MNRAS, 477, 964 [1]

Lam, M. T., Cordes, J. M., Chatterjee, S., et al. 2016, ApJ, 821, 66 [3.1]—. 2017, ApJ, 834, 35 [3.1, 3.1]Lasky, P. D., Mingarelli, C. M. F., Smith, T. L., et al. 2016, Physical Review

X, 6, 011035 [1]Lee, K. J., Wex, N., Kramer, M., et al. 2011, MNRAS, 414, 3251 [3.1]Manchester, R. N., Hobbs, G., Bailes, M., et al. 2013, Publications of the

Astronomical Society of Australia, 30, e017 [4.2]McLaughlin, M. A. 2013, Classical and Quantum Gravity, 30, 224008 [1]Middleton, H., Chen, S., Del Pozzo, W., Sesana, A., & Vecchio, A. 2018,

Nature Communications, 9, 573 [1]Mingarelli, C. M. F., Lazio, T. J. W., Sesana, A., et al. 2017, Nature

Astronomy, 1, 886 [1, 5, 11, 6]Ölmez, S., Mandic, V., & Siemens, X. 2010, Phys. Rev. D, 81, 104028 [1]Rodriguez-Gomez, V., Genel, S., Vogelsberger, M., et al. 2015, MNRAS,

449, 49 [5]Rosado, P. A., Sesana, A., & Gair, J. 2015, MNRAS, 451, 2417 [1]Rusli, S. P., Thomas, J., Saglia, R. P., et al. 2013, AJ, 146, 45 [5]Schutz, K., & Ma, C.-P. 2016, MNRAS, 459, 1737 [1]Sesana, A. 2013, Classical and Quantum Gravity, 30, 224014 [1]Sesana, A., Haardt, F., Madau, P., & Volonteri, M. 2004, ApJ, 611, 623 [1]Sesana, A., Haiman, Z., Kocsis, B., & Kelley, L. Z. 2018, ApJ, 856, 42 [1]Sesana, A., & Vecchio, A. 2010, Phys. Rev. D, 81, 104008 [3.1]Shen, J., & Gebhardt, K. 2010, ApJ, 711, 484 [5]Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ, 131, 1163 [5]Srivastava, N., Hinton, G., Krizhevsky, A., Sutskever, I., & Salakhutdinov,

R. 2014, Journal of Machine Learning Research, 15, 1929 [4.3]Taylor, S. R., Huerta, E. A., Gair, J. R., & McWilliams, S. T. 2016, ApJ,

817, 70 [3.1]Taylor, S. R., Simon, J., & Sampson, L. 2017, Physical Review Letters, 118,

181102 [1]Tiburzi, C., Hobbs, G., Kerr, M., et al. 2016, MNRAS, 455, 4339 [3.2]van Haasteren, R., & Vallisneri, M. 2015, Mon. Not. Roy. Astron. Soc., 446,

1170 [1]Verbiest, J. P. W., Weisberg, J. M., Chael, A. A., Lee, K. J., & Lorimer, D. R.

2012, ApJ, 755, 39 [3.2]Verbiest, J. P. W., Lentati, L., Hobbs, G., et al. 2016, MNRAS, 458, 1267 [1]Vigeland, S. J., Vallisneri, M., & Taylor, S. R. 2019, in preparation [4.3]Wahlquist, H. 1987, General Relativity and Gravitation, 19, 1101 [3.1, 3.1]Yardley, D. R. B., Hobbs, G. B., Jenet, F. A., et al. 2010, MNRAS, 407, 669

[1]Zhu, X.-J., Hobbs, G., Wen, L., et al. 2014, MNRAS, 444, 3709 [1, 4.2]