Detecting Double Neutron Stars with LISADetecting DNSs with LISA 3 for these DNSs to increase their...

13
MNRAS 000, 113 (2019) Preprint 7 January 2020 Compiled using MNRAS L A T E X style file v3.0 Detecting Double Neutron Stars with LISA Mike Y. M. Lau, 1 ,2 ? Ilya Mandel 1 ,2 ,3 , Alejandro Vigna-G´ omez 4 ,1 ,2 ,3 , Coenraad J. Neijssel 3 ,1 ,2 , Simon Stevenson 5 ,2 , and Alberto Sesana 6 1 Monash Centre for Astrophysics, School of Physics and Astronomy, Monash University, Clayton, Victoria 3800, Australia 2 OzGrav: The ARC Centre of Excellence for Gravitational Wave Discovery, Australia 3 Birmingham Institute for Gravitational Wave Astronomy and School of Physics and Astronomy, University of Birmingham, Birmingham, B15 2TT, United Kingdom 4 DARK, Niels Bohr Institute, University of Copenhagen, Blegdamsvej 17, 2100, Copenhagen, Denmark 5 Centre for Astrophysics and Supercomputing, Swinburne University of Technology, Hawthorn, VIC 3122, Australia 6 Dipartimento di Fisica“G. Occhialini”, Universit´ a degli Studi Milano Bicocca, Piazza della Scienza 3, I-20126 Milano, Italy Accepted XXX. Received YYY; in original form ZZZ ABSTRACT We estimate the properties of the double neutron star (DNS) population that will be observable by the planned space-based interferometer LISA. By following the gravita- tional radiation driven evolution of DNSs generated from rapid population synthesis of massive binary stars, we estimate that around 35 DNSs will accumulate a signal- to-noise ratio above 8 over a four-year LISA mission. The observed population mainly comprises Galactic DNSs (94 per cent), but detections in the LMC (5 per cent) and SMC (1 per cent) may also be expected. The median orbital frequency of detected DNSs is expected to be 0.8 mHz, and many of them will be eccentric (median eccen- tricity of 0.11). LISA is expected to localise these DNSs to a typical angular resolution of 2 . We expect the best-constrained DNSs to have eccentricities known to a few parts in a thousand, chirp masses measured to better than 1 per cent fractional un- certainty, and sky localisation at the level of a few arcminutes. The orbital properties will provide insights into DNS progenitors and formation channels. The localisations may allow neutron star natal kick magnitudes to be constrained through the Galactic distribution of DNSs, and make it possible to follow up the sources with radio pulsar searches. LISA is also expected to resolve 10 4 Galactic double white dwarfs, many of which may have binary parameters that resemble DNSs; we discuss how the com- bined measurement of binary eccentricity, chirp mass, and sky location may aid the identification of a DNS. Key words: gravitational waves – binaries: close 1 INTRODUCTION The LIGO Scientific Collaboration made the first direct de- tection of gravitational waves (GWs) in 2015 from the bi- nary black hole (BBH) merger GW150914 (Abbott et al. 2016). Since then, eleven GW events were recorded in the Gravitational-Wave Transient Catalog (GWTC-1) (Abbott et al. 2019). The start of the Advanced LIGO and Advanced Virgo third observing run (O3) on 1 April, 2019 with im- proved detector sensitivity has given almost weekly public alerts to credible GW candidates on the Gravitational Wave Candidate Event Database (GraceDB). The exploration of double compact object (DCO) population statistics will be integral to constraining the relative importance of different ? E-mail: [email protected] (MYML) formation channels and reducing the large uncertainties that characterise key stages of isolated binary evolution. Double neutron star (DNS) coalescences are of particu- lar interest as they may produce electromagnetic counter- parts, including gamma ray bursts, their afterglows, and kilonovae. GW170817, detected during the second observ- ing run (O2) of Advanced LIGO and Virgo (Abbott et al. 2017a), was associated with EM counterparts GRB 170817A and AT 2017gfo (Abbott et al. 2017c,b), marking the first multi-messenger event involving GWs. Neutron stars can receive supernova (SN) natal kicks of several hundred kms -1 and lose significant fractions of mass during SNe, and so DNSs forming from isolated bina- ries may possess significant eccentricities at birth. However, by the time these DNSs evolve to the 10–1000 Hz sensitivity window of the LIGO-Virgo advanced detectors, gravitational radiation reaction circularises the orbit of isolated binaries © 2019 The Authors arXiv:1910.12422v2 [astro-ph.HE] 6 Jan 2020

Transcript of Detecting Double Neutron Stars with LISADetecting DNSs with LISA 3 for these DNSs to increase their...

  • MNRAS 000, 1–13 (2019) Preprint 7 January 2020 Compiled using MNRAS LATEX style file v3.0

    Detecting Double Neutron Stars with LISA

    Mike Y. M. Lau,1,2? Ilya Mandel1,2,3, Alejandro Vigna-Gómez4,1,2,3,

    Coenraad J. Neijssel3,1,2, Simon Stevenson5,2, and Alberto Sesana61 Monash Centre for Astrophysics, School of Physics and Astronomy, Monash University, Clayton, Victoria 3800, Australia2 OzGrav: The ARC Centre of Excellence for Gravitational Wave Discovery, Australia3 Birmingham Institute for Gravitational Wave Astronomy and School of Physics and Astronomy, University of Birmingham,

    Birmingham, B15 2TT, United Kingdom4 DARK, Niels Bohr Institute, University of Copenhagen, Blegdamsvej 17, 2100, Copenhagen, Denmark5 Centre for Astrophysics and Supercomputing, Swinburne University of Technology, Hawthorn, VIC 3122, Australia6 Dipartimento di Fisica “G. Occhialini”, Universitá degli Studi Milano Bicocca, Piazza della Scienza 3, I-20126 Milano, Italy

    Accepted XXX. Received YYY; in original form ZZZ

    ABSTRACTWe estimate the properties of the double neutron star (DNS) population that will beobservable by the planned space-based interferometer LISA. By following the gravita-tional radiation driven evolution of DNSs generated from rapid population synthesisof massive binary stars, we estimate that around 35 DNSs will accumulate a signal-to-noise ratio above 8 over a four-year LISA mission. The observed population mainlycomprises Galactic DNSs (94 per cent), but detections in the LMC (5 per cent) andSMC (1 per cent) may also be expected. The median orbital frequency of detectedDNSs is expected to be 0.8 mHz, and many of them will be eccentric (median eccen-tricity of 0.11). LISA is expected to localise these DNSs to a typical angular resolutionof 2◦. We expect the best-constrained DNSs to have eccentricities known to a fewparts in a thousand, chirp masses measured to better than 1 per cent fractional un-certainty, and sky localisation at the level of a few arcminutes. The orbital propertieswill provide insights into DNS progenitors and formation channels. The localisationsmay allow neutron star natal kick magnitudes to be constrained through the Galacticdistribution of DNSs, and make it possible to follow up the sources with radio pulsarsearches. LISA is also expected to resolve ∼ 104 Galactic double white dwarfs, manyof which may have binary parameters that resemble DNSs; we discuss how the com-bined measurement of binary eccentricity, chirp mass, and sky location may aid theidentification of a DNS.

    Key words: gravitational waves – binaries: close

    1 INTRODUCTION

    The LIGO Scientific Collaboration made the first direct de-tection of gravitational waves (GWs) in 2015 from the bi-nary black hole (BBH) merger GW150914 (Abbott et al.2016). Since then, eleven GW events were recorded in theGravitational-Wave Transient Catalog (GWTC-1) (Abbottet al. 2019). The start of the Advanced LIGO and AdvancedVirgo third observing run (O3) on 1 April, 2019 with im-proved detector sensitivity has given almost weekly publicalerts to credible GW candidates on the Gravitational WaveCandidate Event Database (GraceDB). The exploration ofdouble compact object (DCO) population statistics will beintegral to constraining the relative importance of different

    ? E-mail: [email protected] (MYML)

    formation channels and reducing the large uncertainties thatcharacterise key stages of isolated binary evolution.

    Double neutron star (DNS) coalescences are of particu-lar interest as they may produce electromagnetic counter-parts, including gamma ray bursts, their afterglows, andkilonovae. GW170817, detected during the second observ-ing run (O2) of Advanced LIGO and Virgo (Abbott et al.2017a), was associated with EM counterparts GRB 170817Aand AT 2017gfo (Abbott et al. 2017c,b), marking the firstmulti-messenger event involving GWs.

    Neutron stars can receive supernova (SN) natal kicksof several hundred kms−1 and lose significant fractions ofmass during SNe, and so DNSs forming from isolated bina-ries may possess significant eccentricities at birth. However,by the time these DNSs evolve to the 10–1000 Hz sensitivitywindow of the LIGO-Virgo advanced detectors, gravitationalradiation reaction circularises the orbit of isolated binaries

    © 2019 The Authors

    arX

    iv:1

    910.

    1242

    2v2

    [as

    tro-

    ph.H

    E]

    6 J

    an 2

    020

  • 2 M. Y. M. Lau et al.

    to eccentricities e . 10−5 regardless of their formation eccen-tricity if they are formed at orbital frequencies . 10−4 Hz.On the other hand, ESA’s proposed space-based Laser Inter-ferometer Space Antenna (LISA) (Amaro-Seoane et al. 2017)is anticipated to observe inspiral GWs in the 10−4–10−2 Hzwindow, and so may detect residual eccentricity in inspi-ralling DNSs. Eccentricity measurements by LISA may pro-vide constraints on the physics of isolated binary evolution(Nelemans et al. 2001b; Belczynski et al. 2010; Vigna-Gómezet al. 2018; Tauris 2018; Kyutoku et al. 2019) or dynamicalformation (Kremer et al. 2018; Hamers & Thompson 2019;Andrews & Mandel 2019).

    The detection of GW170817 and the selection of LISAas ESA’s third L-class mission in January 2017 has led torecent interest in LISA DNSs sources. Seto (2019) estimatedthe frequency distribution of DNSs in local group galaxies byextrapolating the comoving volumetric DNS merger rate in-ferred from GW170817. Kyutoku et al. (2019) demonstratedthat LISA-informed observations can enhance the efficiencyof radio pulsar searches with the Square Kilometre Array(SKA). Thrane et al. (2019) showed that constraints maybe placed on the neutron star equation of state by mea-suring the Lense-Thirring precession with multi-messengerobservations with LISA and SKA.

    In this paper, we predict the detection rate, distribu-tion of source parameters (eccentricity, signal-to-noise ratio,distance), and uncertainty in source parameters (eccentric-ity uncertainty, sky-localisation accuracy, chirp mass uncer-tainty) of DNSs in the Milky Way (MW) and in nearbygalaxies. We generate a population of synthetic DNSs usingthe Compact Object Mergers: Population Astrophysics andStatistics (COMPAS) suite (Stevenson et al. 2017; Barrettet al. 2018; Vigna-Gómez et al. 2018; Neijssel et al. 2019),and follow the evolution of these DNSs through the LISAband driven by gravitational radiation reaction, for which weuse the leading quadrupole order expressions (Peters 1964).Starting from an initial population of zero-age main se-quence (ZAMS) binary stars, COMPAS performs single-starevolution using the fitting formulae in Hurley et al. (2000)and calculates changes in stellar and orbital properties dueto wind-driven mass loss, mass transfer, common-envelopeevents, and SNe, until the formation of a DCO.

    This paper is structured as follows. Section 2 describeshow the DNS detection rate is calculated; it begins by high-lighting important features of COMPAS’s Fiducial modelof binary evolution (2.1), then discusses the general proce-dure for estimating the LISA DNS detection rate (2.2), theDNS formation rate within the detector’s sensitive volume(2.3), and the detector sensitivity (2.4). Section 3 presentsour predictions for the distribution of LISA DNS binary pa-rameters and their uncertainties. In particular, we discusshow the eccentricity distribution may constrain binary evo-lution physics. Section 4 discusses strategies for distinguish-ing LISA DNSs from resolved Galactic double white dwarfs(DWDs) and neutron star-white dwarf (NS-WD) binaries.We summarise our results and discuss the validity of ourassumptions in Section 5.

    2 METHODS

    2.1 Population Synthesis

    This work uses a synthetic population of DNSs evolved byVigna-Gómez et al. (2018) using the rapid population syn-thesis element of COMPAS. A total of 106 binary stars wereevolved, with 0.13 per cent becoming DNSs, of which 73 percent merge within the age of the Universe. We highlight dis-tinctive features of the assumed Fiducial model of binaryevolution, and refer the reader to Stevenson et al. (2017)and Vigna-Gómez et al. (2018) for details.

    The mass of the primary star is drawn from the Kroupainitial mass function (Kroupa 2001) in the mass range[5, 100]M� (the full mass range was used for normalisation),while the mass ratio q = m2/m1 is drawn from a uniformdistribution in [0.1, 1] (Sana et al. 2012). All binaries are as-sumed to be circular at ZAMS with solar metallicity. Thebinary separation is drawn from a log-uniform distributionin [0.1, 1000] AU, following Öpik (1924).

    A common-envelope phase follows dynamically-unstable mass transfer, and is described by the αλ-formalism(Webbink 1984; de Kool 1990) with α = 1 and λ determinedby the fits of Xu & Li (2010).

    The Fiducial model distinguishes between core-collapse, ultra-stripped, and electron-capture supernovae.The natal kick direction is randomly drawn from the unitsphere while the kick magnitude follows a bimodal distri-bution. The core-collapse supernova kick magnitude is dis-tributed by a Maxwellian with σhigh = 265 kms−1, followingHobbs et al. (2005). Ultra-stripped and electron-capture su-pernova kicks follow a low-kick Maxwellian with σlow = 30kms−1 (Pfahl et al. 2002; Podsiadlowski et al. 2004; Ver-bunt et al. 2017). The ‘rapid’ explosion model in Fryer et al.(2012) is used to calculate the compact remnant mass fromthe pre-supernova core mass.

    A discussion of the Fiducial model’s two dominantDNS formation channels (accounting for 91 per cent of allDNSs formed) can be found in Vigna-Gómez et al. (2018).

    To illustrate the sensitivity of our results to uncertain-ties in binary evolution prescriptions, we compare resultsobtained with the Fiducial model assumptions to the fol-lowing three variants:

    • Case BB unstable: Case BB mass transfer from a posthelium-main-sequence star (see Section 3.2.1) is assumed toalways be dynamically unstable, whereas it is always stablein the Fiducial model.• Single SN mode: The distribution of natal kick magni-

    tude is a Maxwellian with σhigh = 265 kms−1 for all types ofSNe, as opposed to the bimodal distribution in the Fiducialmodel.• α = 0.1: The common-envelope efficiency parameter (see

    Section 3.2.3) is set to α = 0.1.

    We use the DNS populations simulated with these vari-ation models1 by Vigna-Gómez et al. (2018).

    1 The case BB unstable, single SN mode, and α = 0.1 models arethe (02), (05), and (10) variations respectively in Vigna-Gómez

    et al. (2018).

    MNRAS 000, 1–13 (2019)

  • Detecting DNSs with LISA 3

    2.2 Detection Rate

    In Monte Carlo population synthesis, each DNS synthe-sised by COMPAS represents a sample population labelledby a set of binary parameters θi = (e0,i, a0,i,m1,i,m2,i) fori = 1, 2, ..., NDNS, where e0,i and a0,i are the eccentricity andsemi-major axis at the formation of the ith DNS, m1,i andm2,i are the component masses, and NDNS is the total num-ber of simulated DNS. For each DNS θi , we denote by f itsstarting orbital frequency, the orbital frequency of the DNSat the start of its observation by LISA. We also denote bydN( f ) = (dN/df )df the number of detections this DNS pop-ulation contributes to the bin [ f , f + df ] of starting orbitalfrequencies. The total contribution is therefore

    N =NDNS∑i=1

    ∫ ∞0

    dNi( f )df

    df , (1)

    where the subscript i denotes a quantity evaluated for theparameters θi . The integrand can be written in a more ex-plicit form:

    N =NDNS∑i=1

    ∫ ∞0

    dNi(t)dt

    dti( f )df

    df , (2)

    which involves (i) dNi(t)/dt, the formation rate of LISA-detectable DNSs with θ = θi and (ii) dti( f ), the time takenfor these DNSs to increase their orbital frequencies from f tof + df . The time interval dti is calculated by integrating theorbit-averaged, quadrupole-level expression for [(de/dt)−1](e)given in Peters (1964):

    dedt= −19

    12β

    c40

    e−29/19(1 − e2)3/2

    [1 + (121/304)e2]1181/2299(3)

    where β = 645 G3m1m2(m1 + m2)/c5 and c0 = a0(1 −

    e20)e−12/190 (1+

    121304 e

    20)−870/2299 are constants that depend only

    on the initial binary parameters θi . The lower and upperintegration limits elower = e( f ) and eupper = e( f + df ), arecalculated by inverting the Keplerian expression

    f (e) = 12π

    √G(m1 + m2)

    a(e)3, (4)

    where the orbit-averaged, quadrupole-approximated expres-sion for a = a(e) is also given in Peters (1964):

    a(e) = c0e12/19

    1 − e2

    (1 +

    121304

    e2)870/2299

    . (5)

    2.3 DNS Formation Rate

    We now discuss how the DNS formation rate dNi/dt is cal-culated. Since DNSs are produced by massive stars, theirformation rate traces that of massive stars. Significant delaytimes are possible between star formation and DNS merger2.However, the delay time distribution favours a significantpopulation with short delays, falling off more steeply than

    2 Equation 17 gives the merger time of a m1 = m2 = 1.4 M�circular DNS at the characteristic LISA GW frequency fGW = 2mHz as tmerge = 240,000 yr, so LISA typically observes DNSshortly before merger relative to the overall DNS evolutionary

    time-scale.

    t−1delay

    (e.g., Vigna-Gómez et al. 2018). Moreover, our rates

    are dominated by the MW, which does not show evidenceof significant star formation rate variations over time.3.

    Therefore, we use the DNS formation rate as a proxyfor the DNS merger rate, and use blue light, which tracesthe massive star formation rate, as a proxy for both (e.g.,Kopparapu et al. 2008). To account for long delay times,other approaches that focus on the total mass as a proxyfor the DNS merger rate are also possible (e.g., Artale et al.2019).

    Thus, we take a galaxy’s DNS formation rate to be pro-portional to its extinction-corrected blue light luminosity LB(Phinney 1991; Kalogera et al. 2001; Kopparapu et al. 2008).Then, the total formation rate of a DNS within some detec-tion volume is proportional to the total blue-light luminositycontained in that volume. As we use a sky-averaged signal-to-noise ratio (SNR) in our study, this detection volume isspherical, with radius set by a SNR detectability thresh-old (see Section 2.4). With this assumption, the total DNSformation rate within distance d is simply the MW DNSformation rate reweighted by the total blue light luminositywithin d:

    dN(< d)dt

    =LB(< d)LB,MW

    dNMWdt

    . (6)

    The cumulative blue light luminosity LB(< d) as a functionof distance d is derived from the Gravitational Wave GalaxyCatalogue (GWGC) (White et al. 2011), which containsthe extinction-corrected absolute blue magnitude of 53,255galaxies. In particular, we calculate the MW blue light lumi-nosity from the GWGC to be LB,MW = 1.07 × 1010 LB,�, inunits of solar blue light luminosity LB,�. For the syntheticDNS population in this study, the MW DNS formation rate4

    is 33 Myr−1 (Vigna-Gómez et al. 2018), assuming contin-uous star formation at 2.0 M�yr−1 with solar metallicityZ� = 0.0142. In Equation 2, the formation rate contributedby the ith simulated binary is equal to the total formationrate within distance dmax,i , dN(< dmax,i)/dt, reweighted bythe number of simulated DNSs, NDNS:

    dNidt=

    1NDNS

    dN(< dmax, i)dt

    . (7)

    where dmax,i is the horizon distance of the ith DNS, the max-imum distance the DNS may be located to be detectable byLISA. It is a function of θi and Section 2.4 explains how itis calculated. This prescription assumes the fraction of mas-sive binary stars that become DNSs in different galaxies issame as in the MW, neglecting variations in, for example,metallicity, binary fraction, and initial mass function. It alsoneglects variations in star formation rate over cosmic history.Finally, in the MW, we focus on DNSs produced by isolated

    3 The star formation histories of the LMC and SMC show signifi-cant variations over time (Harris & Zaritsky 2009), with historicalstar formation rate up to a factor of 10 lower than the current

    value, probably leading us to overestimates the DNS formation

    rates in the LMC and SMC.4 Table 2 in Vigna-Gómez et al. (2018) lists the merger rateof Galactic DNSs to be 24 Myr−1 with a merger fraction offmerger = 0.73. The total formation rate of Galactic DNSs is there-fore 24 Myr−1/0.73 = 33Myr−1.

    MNRAS 000, 1–13 (2019)

  • 4 M. Y. M. Lau et al.

    binary evolution in the Galactic disc and do not consider dy-namical formation in globular clusters. The validity of theseassumptions is discussed in Section 5.

    We consider two models as limiting cases for the distri-bution of DNSs within the Galaxy. The first model assumesnegligible kicks or dynamical evolution, so that DNSs aredistributed in the same way as today’s massive star birthsites. The second assumes the limit of very large kicks, underwhich the DNSs distribution follows the mass distributionof the dark matter halo.

    In the first model, we spatially distribute Galactic DNSsto the plane-projected Galactic disc density profile. We usea disc profile inferred from the disc gravitational potentialproposed in Miyamoto & Nagai (1975),

    φd(r, z) = −GMd√

    r2 +(ad +

    √z2 + b2

    d

    )2 , (8)where Md is the total disc mass, ad and bd are the radial andvertical scales, and (r, z) are Galactocentric cylindrical coor-dinates. The density profile ρd(r, z) is obtained by solvingPoisson’s equation ∇2φd = 4πGρd:

    ρd(r, z) =b2d

    Md4π

    adr2 +(ad + 3

    √z2 + b2

    d

    ) (ad +

    √z2 + b2

    d

    )2(z2 + b2

    d

    )3/2 [r2 +

    (ad +

    √z2 + b2

    d

    )2]5/2 .(9)

    Finally, we obtain the plane-projected Galactic disc densityprofile from the integral

    ∫ρd(r, z)dz.

    In reality, the DNS distribution will not trace the birthsite distribution because of a combination of natal kicks fromasymmetric supernovae, Blaauw kicks produced by symmet-ric mass loss accompanying supernovae (Blaauw 1961), andsubsequent dynamical evolution in the Galaxy’s potential.We therefore also consider the opposite extreme: natal kickmagnitudes being large enough to eject DNSs into the darkhalo potential. Considering this as a boundary case, we takethe extreme limit in which these DNSs are allowed to relaxand virialize, and so trace the dark halo mass distribution.For the MW dark halo, we use the density profile in Wilkin-son & Evans (1999),

    ρh(R) =Mh4π

    a2h

    R2(R2 + a2h)3/2

    , (10)

    where Mh is the total halo mass, ah is a characteristic fall-offradius, and R is the Galactocentric distance. We intention-ally do not cut off the halo mass distribution in this modelin order to consider it as an extreme limiting case. In Equa-tions 9 and 10, we use parameters given in ‘Model II’ ofIrrgang et al. (2013), obtained by a χ2-fit to observationalconstraints: Md = 2829 Mgal, ad = 4.85 kpc, bd = 0.184kpc, Mh = 69, 725 Mgal, ah = 200 kpc, and the solar dis-placement r� = 8.35 kpc from the Galactic Centre, whereMgal = 2.325× 107 M� is the Galactic mass unit. We expectthe true distribution of Galactic DNSs to be between theGalactic disc and the dark halo scenarios.

    Figure 1 plots the MW DNS formation rate dN(< d)/dtcontained in a spherical detection volume (centred upon thesolar system) as a function of the sphere radius d for our

    Figure 1. Cumulative fraction of the MW DNS formation withina given distance from the solar system according to three DNS

    spatial distribution prescriptions: (i) DNS distributed accordingto the plane-projected Galactic disc density profile (blue); (ii)

    DNS distributed according to the MW dark halo density profile

    (orange); (iii) DNS distributed uniformly on a flat disc with radius12 kpc (black).

    two prescriptions. A third, toy prescription has also beenincluded for comparison, where the MW is modelled as auniform flat disc of radius 12 kpc. In the toy model, the de-tection volume contains the entire disc-like ‘MW’ at d ≈ 20kpc, beyond which the curve flattens out sharply. The for-mation rate grows more gently with distance for the Galacticdisc potential, where more than 95 per cent of the DNS for-mation rate is contained in d < 100 kpc. In the dark haloprescription, the diffuse halo stretches out to large distanceswith scale radius ah = 200 kpc, and only 45 per cent of theDNS formation rate is contained in d < 100 kpc.

    2.4 Signal-to-Noise Ratio

    We use an SNR expression that is averaged over sky location(θ, φ), GW polarisation ψ, and source inclination ι. Then, theaveraged SNR of the nth GW harmonic depends only on thetotal energy per unit frequency carried by GWs emitted inthe nth harmonic, dEn/d(n f ) (see, e.g., Flanagan & Hughes1998)5:

    〈ρ2n〉 =2G

    π2c3d2

    ∫ nffnfi

    |dEn/d(n f )|(n f )2〈Sn(n f )〉(θ,φ)

    d(n f ). (11)

    Here, 〈ρ2n〉 is the squared SNR of the nth GW harmonic aver-aged over (θ, φ, ι, ψ), d is the source distance, and 〈Sn(n f )〉(θ,φ)

    5 Note that 〈ρ2n 〉 in Equation 11 is larger by a factor of 5 com-pared to the corresponding expression in LIGO literature, becauseof the convention in LISA to include the signal response func-

    tion R in the noise spectral density as 1/RLISA, rather than inthe strain power spectral density as RLISA. Converting from theLIGO to the LISA SNR expression requires dividing by a factor

    of RLIGO = 1/5.

    MNRAS 000, 1–13 (2019)

  • Detecting DNSs with LISA 5

    is the sky-averaged LISA one-sided noise power spectral den-sity. The upper and lower integration limits n ff and n fi arethe nth harmonic GW frequency of the source at the startand end of LISA observation, respectively. In this study, weassume a four-year LISA mission duration as put forwardin the LISA ESA L3 mission proposal (Amaro-Seoane et al.2017).

    We approximate the LISA sensitivity curve Sn(n f ) withthe analytically-fitted expression of Robson et al. (2019).This is plotted in Figure 2, along with a Monte Carlo pop-ulation of detectable DNSs. Below GW frequencies of 1–3mHz, the noise spectrum is dominated by confusion noisedue to unresolved Galactic binaries, mainly comprising ∼ 108DWDs (Nelemans et al. 2001b; Farmer & Phinney 2003;Ruiter et al. 2010). As the LISA mission progresses, theconfusion noise reduces since resolved binaries can be re-moved. We use the set of parameters in Robson et al. (2019)that assume signal subtraction over a four-year LISA mis-sion. The total SNR associated with a (possibly eccentric)source is obtained by summing the SNRs for each harmonicin quadrature:

    〈ρ2〉 =∞∑n=1〈ρ2n〉. (12)

    In the actual computation, we truncate the sum at the har-monic number

    ncutoff =

    5√1 + e(1 − e)3/2

    , (13)

    where ‖k ‖ denotes the nearest integer to k. The error inthe GW luminosity due to this truncation is less than 10−3(O’Leary et al. 2009). We switch to e as the integrationvariable as dEn/dt is an explicit function of eccentricity. Thisis achieved by rewriting in Equation 11 |dEn/d(n f )|d(n f ) =|dEn/dt | |de/dt |−1de and substituting the equations for orbit-averaged |dEn/dt | and de/dt from Peters & Mathews (1963)and Peters (1964):

    〈ρ2n〉 =4819

    Gm1m2a20(1 − e0)2

    c3d2M

    ∫ e fei

    g(n, e)u(e, e0)n2〈Sn(n f (e))〉(θ,φ)

    dee

    (14)

    where

    u(e, e0) =(

    ee0

    ) 2419

    (1 + 121304 e

    2

    1 + 121304 e20

    ) 17402299 (1 + e20)(1 − e

    2)3/2

    1 − 183304 e2 −121304 e

    4(15)

    and g(n, e) determines the relative contribution of each har-monic to the total GW luminosity, whose expression is givenin Peters & Mathews (1963). The eccentricity at the end ofthe mission lifetime, e f , is found by integrating Equation3 over the four-year mission lifetime, or set to zero if thebinary merges during the mission.

    Given the total SNR (Equation 12) of a DNS and athreshold SNR ρmin for detection, one may calculate thehorizon distance dmax of the source, which is the maximumdistance at which this DNS is detectable. Using the inverserelationship between the SNR ρ and distance d (see Equa-tion 14), the horizon distance is

    dmaxkpc

    =ρ(d = 1 kpc)

    ρmin. (16)

    This defines the radius of the spherical detection volume in

    Figure 2. Plot of the total LISA noise amplitude spectral den-sity,

    √Sn( fGW) (solid line), the amplitude spectral density of the

    Galactic background confusion noise,√Sc ( fGW) (dashed line) as-

    suming signal subtraction over a four-year LISA mission, and

    2h(t)√τobs for 35 Monte Carlo realisations of LISA DNS sources(filled circles) with frequencies drawn from Figure 3 and distancesdrawn from Figure 5. The height of a dot above the solid curve

    gives the SNR of the DNS. The green circles correspond to LMC

    sources.

    Equation 7 that is required to calculate the formation rateof DNSs. We assume a four-year LISA mission duration andan SNR threshold ρmin = 8 in this study.

    3 RESULTS

    We anticipate that LISA will be able to detect GWs fromseveral tens of locals DNS binaries. Evaluating Equation 2yields 35 detectable DNSs assuming Galactic sources dis-tributed according to the plane-projected disc density pro-file (our default for the rest of the paper) and 8.4 DNSsassuming Galactic sources distributed according to the MWdark matter halo density profile. Figure 3 shows the cumula-tive number of DNS detections as a function of the startingorbital frequency f .

    Although the LISA sensitivity curve is limited by theGalactic confusion noise below GW frequencies of 1–3 mHz,LISA DNSs are detected with 1mHz characteristic orbitalfrequency (17-minute period), or a gravitational-wave fre-quency of 2 f = 2 mHz for a circular DNS. There are fewhigh-frequency DNSs in the sample due to the shorter timeper unit frequency interval at higher frequencies (dt/df ∝f −11/3). This is shown in Figure 4, which illustrates howour rate estimate is developed using a circular DNS withm1 = m2 = 1.4 M�. The DNS can be observed to the great-est distance d = 1160 kpc (blue curve in top panel) at 17.5mHz, yielding the largest DNS formation/merger rate withinthe sensitive volume at that orbital frequency (orange curvein top panel–the apparent steps on this curve correspond toadditional galaxies coming into view). However, because ofthe steep decrease in the time spent by binaries at higherfrequencies (orange line in bottom panel), the distribution

    MNRAS 000, 1–13 (2019)

  • 6 M. Y. M. Lau et al.

    Figure 3. The cumulative number of expected DNS detections

    by LISA over a four-year mission lifetime, as a function of the

    DNS orbital frequency at the start of observation. Blue: GalacticDNSs distributed according to the plane-projected disc profile.

    Orange: Galactic DNSs distributed according to the MW darkmatter halo density profile.

    of expected DNS detections per unit logarithmic frequencypeaks at 0.6 mHz (blue curve in bottom panel).

    Figure 5 shows the distance distribution of the de-tectable DNSs. We predict 33 (94 per cent) of sources tobe Galactic, 1.7 (5 per cent) from the LMC and 0.3 (1 percent) from the SMC. The number of detectable DNSs in M31(d = 780 kpc) and beyond is negligible (N < 0.01), as DNSsspend too little time at orbital frequencies above 10 mHzwhere they can be observed out to M31 and M33 (see Fig-ure 4). This is broadly consistent with Seto (2019), who finds∼ 3 and ∼ 0.5 detectable DNSs in the LMC and SMC respec-tively, using a slightly higher SNR threshold ρmin = 10 buta larger intrinsic DNS merger rate inferred from GW170817.Seto (2019) also expects ∼ 1 detection in M31 and no signif-icant number of detections in M33.

    Below, we discuss how accurately DNS parameters suchas sky location, eccentricity, and chirp mass can be measuredfrom LISA observations. In general, parameter estimationimproves with the accumulated SNR of a GW source, withthe typical uncertainty in individual parameters scaling as1/ρ in a regime where the linear signal approximation isvalid (Cutler & Flanagan 1994; Poisson & Will 1995).

    Figure 6 shows the cumulative SNR distribution ofLISA DNSs for both the disc and the dark matter halo dis-tribution of Galactic DNSs. For the disc prescription, themedian SNR is 16.8, ∼ 15 DNSs can accumulate ρ > 20,2.5 can accumulate ρ > 100, and the highest expected SNR(set by N(> ρ) = 1) is ∼ 180. The bottom panel is a log-logplot of dN/dρ labelled with an approximate slope obtainedby a least-squares fit. Because the Galactic disc density iscentrally concentrated, dN/dρ falls off more gently than theexpected ρ−3 scaling for a uniform disc. Likewise, for theMW dark matter halo prescription, dN/dρ falls off moregently than the expected ρ−4 scaling for a uniform sphere.This behaviour also causes the more centrally-concentratedGalactic disc prescription to have a lower characteristic DNS

    Figure 4. Top: The horizon distance dmax (blue) and the DNS

    formation rate dN/dt within that distance (orange) as func-tions of starting orbital frequency f , with the horizontal dotted

    lines marking the distances of four nearby galaxies: the LMC,SMC, M31, and M33. Bottom: The DNS frequency distribu-

    tion (dN/dt)(dt/d log f ) (blue), which is the DNS formation rateweighted by the time spent by the evolving DNS per frequencybin dt/d log f (orange) as functions of f . This figure assumes acircular DNS with m1 = m2 = 1.4 M�.

    frequency in Figure 3, as closer DNSs produce stronger sig-nals that can be detected at lower frequencies.

    3.1 Sky Localisation

    Recent works have discussed the importance of sky localisa-tion of LISA DNSs for multi-messenger follow-ups of Galac-tic systems, including radio pulsar observations, to constrainthe neutron star equation of state and test general relativ-ity (Kyutoku et al. 2019; Thrane et al. 2019). Accurate skylocalisation by LISA can reduce the search time for pulsarsurveys such as the SKA Phase 2, which may coincide withLISA’s expected launch in the 2030s. LISA triggers wouldthus allow a longer signal accumulation time, which leadsto higher detection significance and detections of fainter bi-nary pulsars. Kyutoku et al. (2019) show that LISA mea-surements of orbital frequency and other binary parameterscan allow a computationally efficient correction of Dopplersmearing associated with tight radio pulsars, where the sig-

    MNRAS 000, 1–13 (2019)

  • Detecting DNSs with LISA 7

    Figure 5. The expected number of LISA DNS detections plotted

    as a function of the distance d from the solar system. The ver-tical dashed lines mark the positions of the Milky Way nucleus,

    LMC, and SMC, and the bracketed number gives the number of

    detections expected in each galaxy.

    Figure 6. Top: Cumulative SNR distribution of detectable DNSs.Bottom: log-log plot of the SNR distribution labelled with the

    power-law index obtained by a least-squares fit.

    nal integration time is a significant fraction of the orbitalperiod. Moreover, since our calculations show that a total of∼ 2 DNSs may be detected in the LMC and SMC, sky local-isation is also needed for host galaxy identification. Three-dimensional localisation is possible if, in addition to sky po-sition, the distance is also well constrained. For GalacticDNSs, measuring the sky distribution, particularly the dis-placement from the Galactic plane, will place constraints onDNSs kick magnitudes.

    A GW source is triangulated using differences in thesignal arrival time in a detector network. For long-lived GWsources, this may be accomplished by a single detector ob-serving the source at different points along its orbit aroundthe Sun. The timing error of a GW source observed with adetector network is inversely-proportional to the SNR andthe detector frequency bandwidth σfGW through which thesource evolves (Fairhurst 2009). However, LISA DNSs withorbital frequencies f ∼ 1 mHz are approximately monochro-matic, since their merger time for a circular binary,

    τmerge = 240, 000(Mc

    1.2 M�

    )−5/3 ( fGW0.002 Hz

    )−8/3yr, (17)

    is much longer than the fiducial four-year LISA mission du-ration. Then, following Mandel et al. (2018), the timing ac-curacy instead scales as 1/(ρ f ), since the GW phase is de-termined down to 1/ρ of the wave cycle. LISA will completemultiple heliocentric orbits as it observes a DNS, and sogives rise to an effective detector baseline of 2 AU. The un-certainty σθ in the source angular coordinate in one planeis approximately

    σθ ≈ 2.9(ρ

    10

    )−1 ( fGW2 mHz

    )−1 ( L2AU

    )−1deg, (18)

    which is just the timing accuracy divided by the light traveltime L/c across the effective detector baseline, and we haveused a characteristic SNR of 10. This corresponds to localisa-tion within a sky patch of solid angle ∆Ω ∼ πσ2θ ≈ 26.4 deg

    2.We use the approximation of Equation 18 to plot the distri-bution of σθ (Figure 7) for our synthetic DNS population,finding that most DNSs can be localised to within σθ ≈ 2◦.

    An angular resolution of 2◦ = 0.035 rad allows the verti-cal displacement of a Galactic DNS above the Galactic planeat d = 10 kpc to be resolved to (0.035 rad)(10 kpc) = 0.35kpc, roughly the thickness of the old thin stellar disc itself.The size of a pencil beam for a 15 m diameter SKA dish ob-serving at 1.4 GHz is approximately 0.67 deg2 (Smits et al.2009; Kyutoku et al. 2019). It then follows from Figure 7that ≈ 6 DNSs, if containing a radio pulsar, can be coveredby a single pointing with the SKA.

    3.2 Eccentricity

    Significant orbital eccentricities may be imparted to DNSsby supernova kicks (e.g., Tauris et al. 2017) or Blaauwkicks (Blaauw 1961). Short-period DNSs may also be formedthrough dynamical hardening interactions in globular clus-ters until the binary is ejected into the field, presents toosmall of a cross-section for further interactions, or mergesthrough GW emission (Kulkarni et al. 1990; Phinney & Sig-urdsson 1991), or in hierarchical triple-star systems (Hamers& Thompson 2019). The typical separation of the ejected

    MNRAS 000, 1–13 (2019)

  • 8 M. Y. M. Lau et al.

    Figure 7. Cumulative distribution of the uncertainty in the skyangle (in one plane) of the simulated detectable DNSs.

    DNSs depends on the globular cluster properties, but mayfall in the range of a few solar radii, or orbital frequenciesof a few times 10−5 Hz (Andrews & Mandel 2019). Theseejected systems sample a thermal eccentricity distributionp(e) = 2e (Heggie 1975), thereby producing high-frequency,eccentric GW sources. However, DCOs typically circularisebefore reaching the 10-1000 Hz GW sensitivity window of theAdvanced LIGO and Virgo detector networks due to grav-itational radiation reaction (Peters 1964), though some dy-namical channels may yield observable eccentricities in theground-based detector frequency band (e.g., Samsing et al.2014). On the other hand, even field DNSs possess measur-able residual eccentricities in LISA’s millihertz GW window,giving important insights into DNS formation channels andtheir progenitor properties.

    The blue solid line of Figure 8 shows the expected ec-centricity distribution of the DNS population observed byLISA, assuming the isolated binary evolution channel as pre-dicted by the COMPAS Fiducial model of Vigna-Gómezet al. (2018). We find that this model predicts a significantnumber of eccentric DNSs in the LISA band with medianeccentricity of 0.11 at detection and several highly eccen-tric systems, e.g. N(e > 0.6) = 3.6. To illustrate how binaryphysics may be imprinted onto the LISA DNS distribution,we also include the eccentricity distributions of DNSs simu-lated with variations in binary evolution prescription.

    3.2.1 Case BB Mass Transfer Stability

    Case BB mass transfer refers to Roche lobe overflow froma post helium-main-sequence star (a helium Hertzsprung-gap star) (e.g., Delgado & Thomas 1981; Ivanova et al.2003). In the DNS formation channels identified by Vigna-Gómez et al. (2018), this is initiated by a secondary that haspreviously been stripped of its hydrogen envelope during acommon-envelope event. Case BB mass transfer leads to fur-ther stripping of the helium envelope down to a metal core,resulting in an ‘ultra-stripped’ star (Tauris et al. 2013, 2015).

    This stripping may leave a thin carbon and helium layer,which allows the ensuing ultra-stripped SN to receive a lowbut non-zero supernova natal kick, along with the Blaauwkick from symmetric mass loss. This allows the DNS to be-come eccentric despite previously going through a common-envelope.

    The COMPAS Fiducial model assumes that case BBmass transfer is always stable, which is justified a posterioriby the better match to the observed Galactic DNS period–eccentricity distribution. Moreover, all simulated systemsundergoing case BB mass transfer meet the mass ratio-period stability criterion of Tauris et al. (2015) and morethan 90 per cent meet the mass ratio stability criterion ofClaeys et al. (2014).

    The orange dashed curve of Figure 8 shows the eccen-tricity distribution of DNSs detectable by LISA under the as-sumption that case BB mass transfer is always dynamicallyunstable instead. With this model variation, case BB masstransfer always leads to a common-envelope phase, whichsignificantly tightens the orbit and produces DNSs with ∼ 1mHz orbital frequencies. This is right in the detectabilityregion of LISA, and so these DNSs undergo little to no cir-cularisation by gravitational radiation by the time they aredetected. This is reflected by the higher median eccentric-ity of 0.36. However, unstable case BB mass transfer leadsto fewer overall detections (see Appendix A). Although thetotal DNS merger rate in the unstable case BB variation issimilar to that in the Fiducial model with stable case BBmass transfer, unstable case BB produces tighter, higher-frequency binaries that evolve rapidly through the LISAsensitive frequency window, leading to fewer observable sys-tems at a given time. Both stable and unstable case BB masstransfer could occur in reality, so the Fiducial (blue solidcurve) and case BB unstable (orange dashed curve) modelsrepresent boundary cases.

    3.2.2 Natal Kick Magnitude Distribution

    The distribution of neutron star natal kicks is another un-certainty in binary population synthesis. Hobbs et al. (2005)proposed a Maxwellian distribution with scale parameterσ = 265 kms−1 based on the observed 2-d pulsar velocitydistribution, while Verbunt et al. (2017) suggest that a bi-modal Maxwellian produces a better agreement because itbetter fits the low-speed pulsar subpopulation. Populationsynthesis studies also suggest that the bimodal distributionis needed to match the observed wide Galactic DNSs, whichare overwhelmingly disrupted by a natal kick drawn from asingle, high-velocity mode.

    Figure 8 shows that the single high SN natal kick vari-ation (dotted purple curve) produces a moderately more ec-centric population than the Fiducial bimodal distribution,with median e = 0.23. However, the most significant differ-ence relative to the Fiducial model is an overall decrease inthe number of detectable DNS systems by almost a factorof 3, as more binaries are disrupted by the greater SN natalkicks (see Figure A1).

    3.2.3 Common-Envelope Efficiency

    The common-envelope efficiency parameter α (Webbink1984; de Kool 1990) is the ratio of the binding energy of

    MNRAS 000, 1–13 (2019)

  • Detecting DNSs with LISA 9

    Figure 8. Normalised cumulative eccentricity distribution for

    LISA-detectable DNSs for the COMPAS Fiducial model (bluesolid curve) of binary evolution, and for variations with: always

    dynamically unstable case BB mass transfer (orange dashed),

    a single SN kick magnitude (purple dotted), and a common-envelope efficiency of α = 0.1 (green dash-dot).

    the common envelope to the difference in orbital energy be-fore and after the common-envelope phase. The Fiducialmodel’s default value of α = 1 assumes perfectly efficienttransfer of orbital energy into unbinding the envelope, while0 < α < 1 assumes that this energy transfer is not fully effi-cient. We consider a variation with α = 0.1. The green dash-dot curve of Figure 8 shows the corresponding eccentricitydistribution, which has a moderately less eccentric popula-tion than the Fiducial model, with a median eccentricityof 0.071.

    The examples above highlight the value of LISA eccen-tricity measurements to constraining the physics of binaryevolution. Figure A1 shows that the same model variationsdo not significantly affect the frequency distribution of DNSsat the moment of detection by LISA, which is mainly drivenby the LISA sensitivity; it also highlights the differences inthe overall rates between variations.

    3.2.4 Eccentricity Measurement

    We consider a conservative threshold for the detection ofmultiple harmonics by testing whether individual harmonicspass the SNR detection threshold (e.g. Willems et al. 2007);in practice, this condition may be relaxed with the aid ofa matched-filtering search for eccentric signals. The uncer-tainty in measured eccentricity depends strongly on whethertwo or more harmonics are individually detected, or only asingle harmonic is detected, and so we consider these casesseparately.

    If two or more harmonics are detected, the source ec-centricity can be determined from the ratio of GW ampli-tudes of these harmonics. We denote the SNR and harmonicnumber of the loudest (largest SNR) harmonic by ρα andα respectively, and denote the respective quantities for thesecond-loudest harmonic by ρβ and β. The harmonic num-

    Figure 9. The SNR ratio of the second-loudest to the most loud-

    est GW harmonic, ρβ/ρα (black), as a function of eccentricity fora DNS with m1 = m2 = 1.4 M� with starting orbital frequencyf = 1 mHz. Also plotted are the harmonic numbers n correspond-ing to the loudest GW harmonic, α (orange), and second-loudestGW harmonic, β (blue).

    bers α and β and the orbital frequency f can be determinedfrom the observed harmonic frequencies α f and β f with theadditional knowledge that the two loudest harmonics areneighbouring, β = α ± 1. Then the SNR ratio ρβ/ρα ∈ (0, 1)can be mapped uniquely to the source eccentricity e. We plotρβ/ρα as a function of e in Figure 9 for a typical DNS ( f = 1mHz and m1 = m2 = 1.4 M�) observed by LISA. The upwardtrend in ρβ/ρα with increasing eccentricity reflects the dis-persal of GW luminosity across a larger range of frequencyharmonics for a more eccentric source. However, there arealso spiked structures in the plot originating from α and βinterchanging values as a DNS’s eccentricity decreases.

    In Figure 9, α and β drop abruptly at e = 0.93 to α = 4and β = 3. This occurs because although the peak GW lumi-nosity shifts to higher harmonics as eccentricity increases, itis also emitted at increasingly larger frequencies away fromthe trough of the LISA noise curve and so is suppressed. Forthe 1 mHz orbital frequency chosen for this example, thissuppression becomes sufficiently large at e = 0.93 that then = 4 harmonic becomes loudest because its frequency fallsin the region of minimum noise, ∼ 4 mHz. This shows thatvery eccentric DNSs may be detected as systems with domi-nant harmonics that have similar SNRs but small harmonicnumbers.

    Meanwhile, for a DNS with only one detectable GWharmonic, an upper constraint may be placed on the ec-centricity based on the fact that the harmonic with thesecond-largest SNR is below the detection threshold: ρβ <ρmin =⇒ ρβ/ρα < ρmin/ρα. This maximum SNR ratio canthen be mapped to a maximum eccentricity. For example,in Figure 9, constraining the eccentricity to e < 0.1 requiresρβ/ρα < 0.5, i.e., the SNR in the n = 2 harmonic would needto be at least a factor of two above the detection thresh-old. Therefore, eccentricity is relatively poorly constrainedfor DNSs with only one detectable harmonic.

    The uncertainty in measured eccentricity e is further

    MNRAS 000, 1–13 (2019)

  • 10 M. Y. M. Lau et al.

    Figure 10. Cumulative distribution of eccentricity uncertaintyfor LISA DNSs with one detectable GW harmonic (orange) and

    with two or more detectable GW harmonics (blue). The total

    distribution (black) contains fewer than 35 systems because forsome systems the total SNR exceeds the detection threshold but

    no individual harmonics do so.

    compounded by fluctuations in the SNR due to noise. WhileFigure 9 shows the ratio of expected SNRs, actual SNRsfluctuate at the level of ±1 for different noise realisations.Consequently, in the limit of large SNR, the uncertainty onthe SNR ratio ρβ/ρα is at the level 1/ρβ + 1/ρα.

    Figure 10 shows the distribution of eccentricity uncer-tainties based on ρβ/ρα vs. e such as Figure 9 for each start-ing DNS frequency. We find that there are 9 (26 per cent)DNSs with two or more detectable harmonics, for which ec-centricity is determined to within a few times 10−3 to a fewtimes 10−2, and 14 (40 per cent) DNSs with only one de-tectable harmonic, for which eccentricity is determined towithin 0.1–0.2. The remaining 11.7 DNSs (33 per cent) passthe total SNR detection threshold (Equation 12) but with-out any individually detectable harmonics.

    Measuring the eccentricity distribution would providean important probe of binary evolution physics, e.g., distin-guishing between the two models shown in Figure 8.

    3.3 Mass Measurement

    For circular binaries, the chirp mass Mc = m3/51 m3/52 (m1 +

    m2)−1/5 can be directly inferred from the frequency and itsrate of evolution in time. For an eccentric binary, the fre-quency evolution depends on both the chirp mass and theeccentricity:

    n Ûf (Mc, f , e) =965

    (2πn

    )8/3(n f )11/3

    (GMc

    c3

    )5/3F(e), (19)

    where

    F(e) =1 + 7324 e

    2 + 3796 e4

    (1 − e2)7/2(20)

    is the enhancement factor, and setting n = 1 gives the ex-pression for the orbital frequency chirp, Ûf . Therefore, the im-prints of the eccentricity and chirp mass are correlated and

    they must be measured simultaneously, although the limitF(e) ≥ 1 on the enhancement factor implies that an upperlimit on the chirp mass can be safely obtained by settingF(e) = 1.

    Once f , Ûf , and e are measured from the GW signal, thechirp massMc may be determined from Equation 19. It alsofollows from Equation 19 that the fractional uncertainty inchirp mass is

    ∆McMc

    =115∆ ff+

    35∆ ÛfÛf+

    35∆F(e)F(e) . (21)

    For a DNS that is observed over time τobs by LISA and hasSNR ρ, the uncertainties in f and Ûf are ∆ f ≈ 2.2/(ρτobs)and ∆ Ûf ≈ 4.3/(ρτ2

    obs) (Takahashi & Seto 2002). From this

    and using Equation 19 for Ûf , we have the scalings

    ∆ ff= 8.7 × 10−7

    (fGW

    2 mHz

    )−1 ( ρ10

    )−1 ( τobs4 yr

    )−1, (22)

    ∆ ÛfÛf= 0.26

    (fGW

    2 mHz

    )−11/3 ( ρ10

    )−1 ( τobs4 yr

    )−2 ( Mc1.2 M�

    )−5/3(23)

    for a circular DNSs. This suggests that the contribution ofthe frequency measurement uncertainty to the chirp massmeasurement uncertainty can be neglected. The contribu-tion to chirp mass error due to eccentricity, ∆F(e), can becalculated directly for known ∆e using Equation 20 for F(e),while the uncertainty in e may be calculated as described inSection 3.2.

    We plot the cumulative distribution of the chirp massrelative error of detectable LISA DNSs in Figure 11. Forsome sources, particularly low-frequency detections whichdo not appreciably evolve over the observation time (seeEquations 19 and 23), the fractional chirp mass measure-ment uncertainty exceeds 1, meaning that LISA measure-ments alone cannot constrain the chirp mass. We excludesuch sources from Figure 11. Among the ≈ 15 DNSs withmeaningful chirp mass constraints, those with two or moredetectable harmonics only have marginally tighter mass con-straints (median ∆Mc/Mc ≈ 0.02) than those with only onedetectable harmonic (median ∆Mc/Mc ≈ 0.05). AlthoughDNSs with only one detectable harmonic have poorer con-strained absolute values of eccentricity (see Figure 10), theytend to be less eccentric compared to sources with two de-tectable harmonics, and ∆F(e)/F(e) ∝ e∆e for e → 0, so thecontribution of the eccentricity uncertainty to the chirp massmeasurement error is small for low-eccentricity sources. A to-tal of ∼ 8 DNSs in our simulated population will have chirpmasses constrained to better than 10 per cent in fractionaluncertainty, which should be sufficient for the purpose ofidentifying the GW source. The best-measured LISA DNSswill yield chirp masses with . 1% fractional uncertainty.

    4 IDENTIFYING A DNS WITH LISA

    Binary population synthesis studies estimate a populationof ∼ 108 DWDs to exist in the MW (Marsh 2011, and ref-erences therein), most of which are expected to be detachedDWDs (Nelemans et al. 2001b). As discussed in Section 2.4,GWs emitted by unresolved Galactic DWDs form a confu-sion noise below 1–2 mHz, which has been included in the

    MNRAS 000, 1–13 (2019)

  • Detecting DNSs with LISA 11

    Figure 11. Cumulative distribution of chirp mass relative uncer-tainty for LISA DNSs (black), separated into those with one de-

    tectable GW harmonic (orange) and with two or more detectable

    GW harmonics (blue). We exclude sources with ∆Mc/Mc ≥ 1,for which LISA alone cannot measure the chirp mass.

    sensitivity curve used in this study (Robson et al. 2019).However, ∼ 104 binaries from this Galactic DWDs popu-lation are estimated to be detectable by LISA (Farmer &Phinney 2003; Nelemans et al. 2001a; Ruiter et al. 2010;Korol et al. 2017), significantly outnumbering our estimated∼ 30 Galactic DNSs. Here, we discuss methods of positivelyidentifying a DNS with LISA observations.

    The chirp mass is the primary means of differentiatingDWD and DNS systems, with chirp masses above ≈ 1.2 M�indicating that at least one component exceeds the Chan-drasekhar limit for the maximum white dwarf mass. How-ever, given the size of the DWD population, a high-masstail of Mc . 1.2 M� (but sub-Chandrasekhar) DWD bina-ries could still cause confusion with DNSs, as could neutronstar-white dwarf binaries.

    The detection of a source with non-zero eccentric-ity favours a DNS interpretation. The disc population ofDWDs is thought to have formed via isolated binary evo-lution, where the progenitors are expected to have tidally-circularised from multiple mass transfer episodes (Nelemanset al. 2001b). Observationally, there are no known eccentricGalactic DWDs, although there are observations of an eccen-tric Galactic pulsar-WD binary (Antoniadis et al. 2016) anda WD-main sequence (Siess et al. 2014) binary in the MW.On the other hand, DNSs may have significant eccentricitiesfrom supernova and Blaauw kicks: in our model, half of LISADNSs will have e > 0.1, and ∼ 10 will have measurable sec-ond GW harmonics, which allow eccentricity to be measuredwith ∆e . 0.02 accuracy. Yet, dynamical formation channelsin MW globular clusters (Willems et al. 2007) or Lidov-Kozai oscillations in hierarchical triple systems (Thompson2011) may produce eccentric DWDs. Kremer et al. (2018)estimate that ejected binaries will only comprise a few MWsources with ρ ≥ 2, but given the very large DWD popu-

    lation, even rare systems could be responsible for confusionwith DNSs.

    The identification of an eccentric source as a DNS iseven more confident if a chirp mass measurement is possible.Above a chirp mass of ≈ 1.2 M�, the DWD interpretationbecomes highly unlikely. In fact, the chirp mass distributionof eccentric DWDs formed in globular clusters is expectedto strong peak at 0.3–0.4 M� (Willems et al. 2007).

    Finally, sky localisation may also aid source identifi-cation. Since eccentric DWDs dynamically formed in MWglobular clusters are ejected into the Galactic halo, we ex-pect eccentric disc binaries to be DNSs, though the lat-ter may also be found far from the disc due to dynami-cal formation or kicks (see, e.g., Figure C1 of Vigna-Gómezet al. 2018). Accurate sky localisation will also enhance theprospects for electromagnetic follow-up, which could defini-tively distinguish DNS and DWD systems (Kyutoku et al.2019; Thrane et al. 2019).

    5 CONCLUSIONS & DISCUSSION

    We estimated that around 35 inspiralling DNSs will be de-tectable over a four-year LISA mission with SNR ρ > 8 us-ing a mock population of isolated binaries synthesised withCOMPAS. Of those, 94 per cent are expected to be Galac-tic DNSs, with the remainder in the LMC (5 per cent) andSMC (1 per cent). These DNSs are detected when the or-bital frequency is typically 1 mHz, despite the presence ofconfusion-limited noise below GW frequencies of 1–3 mHzfrom unresolved Galactic DWD binaries.

    Half of the detectable DNSs retain significant residualeccentricities, e > 0.11, imparted mostly by the Blaauw kickat the second supernova in the COMPAS population synthe-sis models. Around a quarter of the LISA DNSs will havetwo or more individually detectable GW harmonics and ∼ 40per cent have only a single resolvable harmonic, while theremaining third will have GW harmonics that combine toexceed the SNR threshold, but are not individually resolv-able. When two or more harmonics are observed, eccentrici-ties may be accurately estimated to ∆e . 0.02 by measuringSNR ratios of different GW harmonics. If only one GW har-monic is observed for a DNS, only an upper constraint onthe eccentricity is placed at a typical level of e . 0.1.

    A population of DNSs with well measured periods andeccentricities places valuable constraints on binary evolu-tion physics. With a merger time of ∼ 2.4 × 105 years froma GW frequency of 2 f = 2 mHz, the DNSs evolve slowly infrequency over the four year LISA mission, only changingtheir frequency by parts in 105. This makes accurate chirpmass measurements challenging, which is compounded bythe correlation between chirp mass and eccentricity in driv-ing orbital frequency evolution. We find that ≈ 15 DNSswill have useful chirp mass constraints from the LISA sig-nal, with median fractional chirp mass uncertainties of 0.04,dropping to below 1% for the best-measured sources. Thesechirp mass and eccentricity measurements will make it pos-sible to distinguish at least a fraction of the better-measuredeccentric DNSs from the much larger Galactic DWD popula-tion. They can also elucidate the origin of the DNS systems:although the isolated binary channel is generally assumed todominate DNS formation, with globular clusters expected to

    MNRAS 000, 1–13 (2019)

  • 12 M. Y. M. Lau et al.

    contribute less than 10 per cent of all merging DNSs (Phin-ney 1991; Grindlay et al. 2006; Ivanova et al. 2008; Kre-mer et al. 2018), recent work has suggested that dynamicalor three-body formation channels may be relevant (Hamers& Thompson 2019; Andrews & Mandel 2019). Moreover,LISA’s measurement of the eccentricity distribution in theearly DNS evolutionary history could shed light on uncer-tainties in models of isolated binary evolution, such as thestability of case BB mass transfer.

    LISA’s heliocentric orbit produces an effective detec-tor baseline of 2 AU for source triangulation, allowing foraccurate sky localisation. We find that most DNSs will belocalised with an angular resolution σθ . 2 deg. This issufficient to measure the height of Galactic DNSs relativeto the Galactic plane to within ∼ 0.35 kpc, which providesa constraint on the DNS natal kick distribution. Around 6DNSs will be localised sufficiently well to be covered by asingle pointing of the SKA, giving rise to an efficient, LISA-informed follow-up of possible radio pulsars.

    The best-constrained LISA DNSs–the golden binaries–will be localised to a few arc-minutes with eccentricity in-ferred at an accuracy of a few parts in a thousand and thechirp mass to better than 1 per cent fractional uncertainty.

    While this paper was under review, the manuscript ofAndrews et al. (2019) (hereafter A19) became available.A19 study the population of LISA DNSs by sampling DNSmerger times and positions in the MW. They assume thatall systems have periods and eccentricities set by the for-ward evolution of PSR B1913+16. They further assume aMW merger rate of 210 Myr−1 inferred from the DNS GWevent GW170817 (Abbott et al. 2017a), which is ≈ 6 timeshigher than our assumed rate of 33 Myr−1 under our Fidu-cial model. Therefore, A19 predict approximately 6 timesmore detections over a four-year LISA mission than we do.With the larger merger rate, A19 further predict ∼ 1 detec-tions in M31. A19 also find eccentricity uncertainties thatare roughly consistent with ours, based on measuring theSNR ratio of the n = 2, 3 harmonics. As shown in Figure9, for a typical LISA DNS, the second and third harmonicshave the highest SNRs only for e . 0.3. For more eccentricDNSs, A19’s approach overestimates the uncertainty.

    ACKNOWLEDGEMENTS

    We thank Floor Broekgaarden, Philipp Podsiadlowski andAlberto Vecchio for discussions and suggestions. M. Y. M.L. acknowledges support by an Australian Government Re-search Training Program (RTP) Scholarship. A. V.-G. ac-knowledges funding support from Consejo Nacional de Cien-cia y Tecnoloǵıa (CONACYT). S. S. is supported by theAustralian Research Council Centre of Excellence for Grav-itational Wave Discovery (OzGrav), through project numberCE170100004.

    REFERENCES

    Abbott B. P., et al., 2016, Phys. Rev. Lett., 116, 061102

    Abbott B. P., et al., 2017a, Phys. Rev. Lett., 119, 161101

    Abbott B. P., et al., 2017b, ApJ, 848, L12

    Abbott B. P., et al., 2017c, ApJ, 848, L13

    Abbott B., et al., 2019, Physical Review X, 9

    Amaro-Seoane P., et al., 2017, arXiv e-prints, p. arXiv:1702.00786

    Andrews J. J., Mandel I., 2019, ApJ, 880, L8

    Andrews J. J., Breivik K., Pankow C., D’Orazio D. J., Safarzadeh

    M., 2019, arXiv e-prints, p. arXiv:1910.13436

    Antoniadis J., Tauris T. M., Ozel F., Barr E., Champion D. J.,

    Freire P. C. C., 2016, arXiv e-prints, p. arXiv:1605.01665

    Artale M. C., Mapelli M., Giacobbo N., Sabha N. B., Spera M.,

    Santoliquido F., Bressan A., 2019, MNRAS, 487, 1675

    Barrett J. W., Gaebel S. M., Neijssel C. J., Vigna-Gómez A.,Stevenson S., Berry C. P. L., Farr W. M., Mandel I., 2018,

    MNRAS, 477, 4685

    Belczynski K., Benacquista M., Bulik T., 2010, ApJ, 725, 816

    Blaauw A., 1961, Bull. Astron. Inst. Netherlands, 15, 265

    Claeys J. S. W., Pols O. R., Izzard R. G., Vink J., VerbuntF. W. M., 2014, A&A, 563, A83

    Cutler C., Flanagan É. E., 1994, Phys. Rev. D, 49, 2658

    Delgado A. J., Thomas H. C., 1981, A&A, 96, 142

    Fairhurst S., 2009, New Journal of Physics, 11, 123006

    Farmer A. J., Phinney E. S., 2003, MNRAS, 346, 1197

    Flanagan É. É., Hughes S. A., 1998, Phys. Rev. D, 57, 4535

    Fryer C. L., Belczynski K., Wiktorowicz G., Dominik M.,

    Kalogera V., Holz D. E., 2012, ApJ, 749, 91

    Grindlay J., Portegies Zwart S., McMillan S., 2006, Nature

    Physics, 2, 116

    Hamers A. S., Thompson T. A., 2019, ApJ, 883, 23

    Harris J., Zaritsky D., 2009, ApJ, 138, 1243

    Heggie D. C., 1975, MNRAS, 173, 729

    Hobbs G., Lorimer D. R., Lyne A. G., Kramer M., 2005, MNRAS,

    360, 974

    Hurley J. R., Pols O. R., Tout C. A., 2000, MNRAS, 315, 543

    Irrgang A., Wilcox B., Tucker E., Schiefelbein L., 2013, A&A,

    549, A137

    Ivanova N., Belczynski K., Kalogera V., Rasio F. A., Taam R. E.,

    2003, ApJ, 592, 475

    Ivanova N., Heinke C. O., Rasio F. A., Belczynski K., FregeauJ. M., 2008, MNRAS, 386, 553

    Kalogera V., Narayan R., Spergel D. N., Taylor J. H., 2001, ApJ,

    556, 340

    Kopparapu R. K., Hanna C., Kalogera V., O’Shaughnessy R.,

    González G., Brady P. R., Fairhurst S., 2008, ApJ, 675, 1459

    Korol V., Rossi E. M., Groot P. J., Nelemans G., Toonen S.,

    Brown A. G. A., 2017, Monthly Notices of the Royal Astro-nomical Society, 470, 1894–1910

    Kremer K., Chatterjee S., Breivik K., Rodriguez C. L., Larson

    S. L., Rasio F. A., 2018, Phys. Rev. Lett., 120, 191103

    Kroupa P., 2001, Monthly Notices of the Royal Astronomical So-

    ciety, 322, 231

    Kulkarni S. R., Narayan R., Romani R. W., 1990, ApJ, 356, 174

    Kyutoku K., Nishino Y., Seto N., 2019, MNRAS, 483, 2615

    Mandel I., Sesana A., Vecchio A., 2018, Classical and Quantum

    Gravity, 35, 054004

    Marsh T. R., 2011, Classical and Quantum Gravity, 28, 094019

    Miyamoto M., Nagai R., 1975, Publications of the Astronomical

    Society of Japan, 27, 533

    Neijssel C. J., et al., 2019, Monthly Notices of the Royal Astro-

    nomical Society

    Nelemans G., Yungelson L. R., Portegies Zwart S. F., Verbunt F.,

    2001a, A&A, 365, 491

    Nelemans G., Yungelson L. R., Portegies Zwart S. F., 2001b,A&A, 375, 890

    O’Leary R. M., Kocsis B., Loeb A., 2009, MNRAS, 395, 2127

    Öpik E., 1924, Publications of the Tartu Astrofizica Observatory,25

    Peters P. C., 1964, Phys. Rev., 136, B1224

    Peters P. C., Mathews J., 1963, Phys. Rev., 131, 435

    Pfahl E., Rappaport S., Podsiadlowski P., 2002, ApJ, 573, 283

    Phinney E. S., 1991, ApJ, 380, L17

    Phinney E. S., Sigurdsson S., 1991, Nature, 349, 220

    MNRAS 000, 1–13 (2019)

    http://dx.doi.org/10.1103/PhysRevLett.116.061102http://adsabs.harvard.edu/abs/2016PhRvL.116f1102Ahttp://dx.doi.org/10.1103/PhysRevLett.119.161101http://dx.doi.org/10.3847/2041-8213/aa91c9http://adsabs.harvard.edu/abs/2017ApJ...848L..12Ahttp://dx.doi.org/10.3847/2041-8213/aa920chttp://adsabs.harvard.edu/abs/2017ApJ...848L..13Ahttp://dx.doi.org/10.1103/physrevx.9.031040https://ui.adsabs.harvard.edu/abs/2017arXiv170200786Ahttp://dx.doi.org/10.3847/2041-8213/ab2ed1https://ui.adsabs.harvard.edu/abs/2019ApJ...880L...8Ahttps://ui.adsabs.harvard.edu/abs/2019arXiv191013436Ahttps://ui.adsabs.harvard.edu/abs/2016arXiv160501665Ahttp://dx.doi.org/10.1093/mnras/stz1382https://ui.adsabs.harvard.edu/abs/2019MNRAS.487.1675Ahttp://dx.doi.org/10.1093/mnras/sty908https://ui.adsabs.harvard.edu/abs/2018MNRAS.477.4685Bhttp://dx.doi.org/10.1088/0004-637X/725/1/816https://ui.adsabs.harvard.edu/abs/2010ApJ...725..816Bhttps://ui.adsabs.harvard.edu/abs/1961BAN....15..265Bhttp://dx.doi.org/10.1051/0004-6361/201322714https://ui.adsabs.harvard.edu/abs/2014A&A...563A..83Chttp://dx.doi.org/10.1103/PhysRevD.49.2658http://adsabs.harvard.edu/abs/1994PhRvD..49.2658Chttps://ui.adsabs.harvard.edu/abs/1981A&A....96..142Dhttp://dx.doi.org/10.1088/1367-2630/11/12/123006http://adsabs.harvard.edu/abs/2009NJPh...11l3006Fhttp://dx.doi.org/10.1111/j.1365-2966.2003.07176.xhttps://ui.adsabs.harvard.edu/abs/2003MNRAS.346.1197Fhttp://dx.doi.org/10.1103/PhysRevD.57.4535https://ui.adsabs.harvard.edu/abs/1998PhRvD..57.4535Fhttp://dx.doi.org/10.1088/0004-637X/749/1/91http://adsabs.harvard.edu/abs/2012ApJ...749...91Fhttp://dx.doi.org/10.1038/nphys214http://dx.doi.org/10.1038/nphys214https://ui.adsabs.harvard.edu/abs/2006NatPh...2..116Ghttp://dx.doi.org/10.3847/1538-4357/ab3b06https://ui.adsabs.harvard.edu/abs/2019ApJ...883...23Hhttp://dx.doi.org/10.1088/0004-6256/138/5/1243https://ui.adsabs.harvard.edu/abs/2009AJ....138.1243Hhttp://dx.doi.org/10.1093/mnras/173.3.729https://ui.adsabs.harvard.edu/abs/1975MNRAS.173..729Hhttp://dx.doi.org/10.1111/j.1365-2966.2005.09087.xhttps://ui.adsabs.harvard.edu/abs/2005MNRAS.360..974Hhttp://dx.doi.org/10.1046/j.1365-8711.2000.03426.xhttps://ui.adsabs.harvard.edu/abs/2000MNRAS.315..543Hhttp://dx.doi.org/10.1051/0004-6361/201220540http://adsabs.harvard.edu/abs/2013A%26A...549A.137Ihttp://dx.doi.org/10.1086/375578https://ui.adsabs.harvard.edu/abs/2003ApJ...592..475Ihttp://dx.doi.org/10.1111/j.1365-2966.2008.13064.xhttps://ui.adsabs.harvard.edu/abs/2008MNRAS.386..553Ihttp://dx.doi.org/10.1086/321583https://ui.adsabs.harvard.edu/abs/2001ApJ...556..340Khttp://dx.doi.org/10.1086/527348https://ui.adsabs.harvard.edu/abs/2008ApJ...675.1459Khttp://dx.doi.org/10.1093/mnras/stx1285http://dx.doi.org/10.1093/mnras/stx1285http://dx.doi.org/10.1103/PhysRevLett.120.191103https://ui.adsabs.harvard.edu/abs/2018PhRvL.120s1103Khttp://dx.doi.org/10.1046/j.1365-8711.2001.04022.xhttp://dx.doi.org/10.1046/j.1365-8711.2001.04022.xhttp://adsabs.harvard.edu/abs/2001AS.322..231Khttp://dx.doi.org/10.1086/168828https://ui.adsabs.harvard.edu/abs/1990ApJ...356..174Khttp://dx.doi.org/10.1093/mnras/sty3322https://ui.adsabs.harvard.edu/abs/2019MNRAS.483.2615Khttp://dx.doi.org/10.1088/1361-6382/aaa7e0http://dx.doi.org/10.1088/1361-6382/aaa7e0http://adsabs.harvard.edu/abs/2018CQGra..35e4004Mhttp://dx.doi.org/10.1088/0264-9381/28/9/094019http://adsabs.harvard.edu/abs/1975PASJ...27..533Mhttp://dx.doi.org/10.1093/mnras/stz2840http://dx.doi.org/10.1093/mnras/stz2840http://dx.doi.org/10.1051/0004-6361:20000147https://ui.adsabs.harvard.edu/abs/2001A&A...365..491Nhttp://dx.doi.org/10.1051/0004-6361:20010683https://ui.adsabs.harvard.edu/abs/2001A&A...375..890Nhttp://dx.doi.org/10.1111/j.1365-2966.2009.14653.xhttps://ui.adsabs.harvard.edu/abs/2009MNRAS.395.2127Ohttps://ui.adsabs.harvard.edu/abs/1924PTarO..25f...1Ohttp://dx.doi.org/10.1103/PhysRev.136.B1224http://dx.doi.org/10.1103/PhysRev.131.435http://dx.doi.org/10.1086/340494http://dx.doi.org/10.1086/186163http://adsabs.harvard.edu/abs/1991ApJ...380L..17Phttp://dx.doi.org/10.1038/349220a0https://ui.adsabs.harvard.edu/abs/1991Natur.349..220P

  • Detecting DNSs with LISA 13

    Podsiadlowski P., Langer N., Poelarends A. J. T., Rappaport S.,

    Heger A., Pfahl E., 2004, ApJ, 612, 1044

    Poisson E., Will C. M., 1995, Phys. Rev. D, 52, 848

    Robson T., Cornish N. J., Liu C., 2019, Classical and Quantum

    Gravity, 36, 105011

    Ruiter A. J., Belczynski K., Benacquista M., Larson S. L.,

    Williams G., 2010, Astrophys. J., 717, 1006

    Samsing J., MacLeod M., Ramirez-Ruiz E., 2014, ApJ, 784, 71

    Sana H., et al., 2012, Science, 337, 444

    Seto N., 2019, MNRAS, 489, 4513

    Siess L., Davis P. J., Jorissen A., 2014, A&A, 565, A57

    Smits R., Kramer M., Stappers B., Lorimer D. R., Cordes J.,

    Faulkner A., 2009, A&A, 493, 1161

    Stevenson S., Vigna-Gómez A., Mandel I., Barrett J. W., Neijssel

    C. J., Perkins D., de Mink S. E., 2017, Nature Communica-tions, 8, 14906

    Takahashi R., Seto N., 2002, ApJ, 575, 1030

    Tauris T. M., 2018, Phys. Rev. Lett., 121, 131105

    Tauris T. M., Langer N., Moriya T. J., Podsiadlowski P., YoonS. C., Blinnikov S. I., 2013, ApJ, 778, L23

    Tauris T. M., Langer N., Podsiadlowski P., 2015, MNRAS, 451,

    2123

    Tauris T. M., et al., 2017, ApJ, 846, 170

    Thompson T. A., 2011, ApJ, 741, 82

    Thrane E., Os lowski S., Lasky P., 2019, arXiv e-prints, p.

    arXiv:1910.12330

    Verbunt F., Igoshev A., Cator E., 2017, A&A, 608, A57

    Vigna-Gómez A., et al., 2018, MNRAS, 481, 4009

    Webbink R. F., 1984, ApJ, 277, 355

    White D. J., Daw E. J., Dhillon V. S., 2011, Classical and Quan-

    tum Gravity, 28, 085016

    Wilkinson M. I., Evans N. W., 1999, MNRAS, 310, 645

    Willems B., Kalogera V., Vecchio A., Ivanova N., Rasio F. A.,

    Fregeau J. M., Belczynski K., 2007, ApJ, 665, L59

    Xu X.-J., Li X.-D., 2010, ApJ, 716, 114

    de Kool M., 1990, ApJ, 358, 189

    APPENDIX A: MODEL VARIATIONS IN THEDNS FREQUENCY DISTRIBUTION

    Figure A1 shows the distribution of the orbital frequenciesof detectable DNSs at the start of LISA observation for theFiducial model and the three model variations discussed inSection 3.2. The characteristic detection frequency is simi-lar across the four models as it is mainly set by the LISAsensitivity.

    The overall normalisation is, however, sensitive tochanges in the binary evolution prescription. The Fiducialmodel yields the most DNS detections among the consideredvariations. The single SN mode causes fewer systems to bedetected, as the higher natal kick scale parameter for ultra-stripped SNe, σhigh = 265 kms−1, compared to σlow = 30kms−1 in the Fiducial model, is more likely to disrupt thebinary. The α = 0.1 variation makes it approximately tentimes more difficult to satisfy the global energy criterion forenvelope ejection, Ebind > αEorb, compared to the Fidu-cial model where α = 1, thereby decreasing the surviv-ability of the common-envelope phase. On the other hand,the unstable case BB variation actually gives rise to a DNSmerger rate that is similar to the Fiducial model with sta-ble case BB mass transfer, but with fewer detections becauseDNSs are produced at higher frequencies and quickly evolvethrough the LISA sensitivity window.

    Figure A1. The cumulative number of LISA DNS detections as

    a function of orbital frequency at the start of the observation forthe COMPAS Fiducial model (solid blue curve) and for vari-

    ations with: always dynamically unstable case BB mass transfer

    (orange dashed), a single SN kick magnitude (purple dotted), anda common-envelope efficiency of α = 0.1 (green dash-dot).

    This paper has been typeset from a TEX/LATEX file prepared by

    the author.

    MNRAS 000, 1–13 (2019)

    http://dx.doi.org/10.1086/421713https://ui.adsabs.harvard.edu/abs/2004ApJ...612.1044Phttp://adsabs.harvard.edu/abs/1995PhRvD..52..848Phttp://dx.doi.org/10.1088/1361-6382/ab1101http://dx.doi.org/10.1088/1361-6382/ab1101https://ui.adsabs.harvard.edu/abs/2019CQGra..36j5011Rhttp://dx.doi.org/10.1088/0004-637X/717/2/1006http://dx.doi.org/10.1088/0004-637X/784/1/71http://adsabs.harvard.edu/abs/2014ApJ...784...71Shttp://dx.doi.org/10.1126/science.1223344https://ui.adsabs.harvard.edu/abs/2012Sci...337..444Shttp://dx.doi.org/10.1093/mnras/stz2439https://ui.adsabs.harvard.edu/abs/2019MNRAS.489.4513Shttp://dx.doi.org/10.1051/0004-6361/201423397https://ui.adsabs.harvard.edu/abs/2014A&A...565A..57Shttp://dx.doi.org/10.1051/0004-6361:200810383https://ui.adsabs.harvard.edu/abs/2009A&A...493.1161Shttp://dx.doi.org/10.1038/ncomms14906http://dx.doi.org/10.1038/ncomms14906http://adsabs.harvard.edu/abs/2017NatCo...814906Shttp://dx.doi.org/10.1086/341483https://ui.adsabs.harvard.edu/abs/2002ApJ...575.1030Thttp://dx.doi.org/10.1103/PhysRevLett.121.131105https://ui.adsabs.harvard.edu/abs/2018PhRvL.121m1105Thttp://dx.doi.org/10.1088/2041-8205/778/2/L23https://ui.adsabs.harvard.edu/abs/2013ApJ...778L..23Thttp://dx.doi.org/10.1093/mnras/stv990https://ui.adsabs.harvard.edu/abs/2015MNRAS.451.2123Thttps://ui.adsabs.harvard.edu/abs/2015MNRAS.451.2123Thttp://dx.doi.org/10.3847/1538-4357/aa7e89https://ui.adsabs.harvard.edu/abs/2017ApJ...846..170Thttp://dx.doi.org/10.1088/0004-637X/741/2/82https://ui.adsabs.harvard.edu/abs/2011ApJ...741...82Thttps://ui.adsabs.harvard.edu/abs/2019arXiv191012330Thttps://ui.adsabs.harvard.edu/abs/2019arXiv191012330Thttp://dx.doi.org/10.1051/0004-6361/201731518https://ui.adsabs.harvard.edu/abs/2017A&A...608A..57Vhttp://dx.doi.org/10.1093/mnras/sty2463http://adsabs.harvard.edu/abs/2018MNRAS.481.4009Vhttp://dx.doi.org/10.1086/161701https://ui.adsabs.harvard.edu/abs/1984ApJ...277..355Whttp://dx.doi.org/10.1088/0264-9381/28/8/085016http://dx.doi.org/10.1088/0264-9381/28/8/085016http://adsabs.harvard.edu/abs/2011CQGra..28h5016Whttp://dx.doi.org/10.1046/j.1365-8711.1999.02964.xhttps://ui.adsabs.harvard.edu/abs/1999MNRAS.310..645Whttp://dx.doi.org/10.1086/521049https://ui.adsabs.harvard.edu/abs/2007ApJ...665L..59Whttp://dx.doi.org/10.1088/0004-637X/716/1/114https://ui.adsabs.harvard.edu/abs/2010ApJ...716..114Xhttp://dx.doi.org/10.1086/168974https://ui.adsabs.harvard.edu/abs/1990ApJ...358..189D

    1 Introduction2 Methods2.1 Population Synthesis2.2 Detection Rate2.3 DNS Formation Rate2.4 Signal-to-Noise Ratio

    3 Results3.1 Sky Localisation3.2 Eccentricity3.3 Mass Measurement

    4 Identifying a DNS with LISA5 Conclusions & DiscussionA Model variations in the DNS frequency distribution