11institutetext: Department of Astrophysics/IMAPP, Radboud University, P.O. Box 9010, 6500 GL Nijmegen, The Netherlands
e-mail: paul.disberg@gmail.com
22institutetext: Department of Physics, University of Warwick, Coventry CV4 7AL, UK

Kinematic constraints on the ages and kick velocities of Galactic neutron star binaries

P. Disberg 11    N. Gaspari 11    A. J. Levan 1122
(July 18, 2024)
Abstract

Context. The systems creating binary neutron stars (BNSs) experience systemic kicks when one of the components goes supernova. The combined magnitude of these kicks is still a topic of debate, and has implications for the eventual location of the transient resulting from the merger of the binary. For example, the offsets of short-duration gamma-ray bursts (SGRBs) resulting from BNS mergers depend on the BNS kicks.

Aims. We investigated Galactic BNSs, and traced their motion through the Galaxy. This enabled us to estimate their kinematic ages and construct a BNS kick distribution, based on their Galactic trajectories.

Methods. We used the pulsar periods and their derivatives to estimate the characteristic spin-down ages of the binaries. Moreover, we used a Monte Carlo estimation of their present-day velocity vector in order to trace back their trajectory and estimate their kinematic ages. These trajectories, in turn, were used to determine the eccentricity of their Galactic orbit. Based on simulations of kicked objects in the Galactic potential, we investigated the relationship between this eccentricity and kick velocity, in order to constrain the kicks imparted to the binaries at birth.

Results. We find that the Galactic BNSs are likely older than 40similar-toabsent40\sim 40∼ 40 Myr, which means their current (scalar) galactocentric speeds are not representative of their initial kicks. However, we find a close relationship between the eccentricity of a Galactic trajectory and the experienced kick. Using this relation, we constrained the kicks of the Galactic BNSs, depending on the kind of isotropy assumed in estimating their velocity vectors. These kick velocities are well-described by a log-normal distribution peaking around 4050similar-toabsent4050\sim 40-50∼ 40 - 50 km/s, and coincide with the peculiar velocities of the binaries at their last disc crossing.

Conclusions. We conclude that BNSs receive kicks following a distribution that peaks at kick velocities lower than found in isolated pulsars. However, we find no tension between this distribution and literature on SGRB offsets.

Key Words.:
stars: kinematics and dynamics – Galaxy: stellar content – binaries: general

1 Introduction

Neutron star binaries – and in particular their mergers – can provide insight into a variety of astrophysical phenomena. For example, mergers of these binary neutron stars (BNSs), consisting of two neutron stars (NSs) orbiting each other, have produced gravitational waves that have been detected (e.g. Abbott et al., 2017b) and are used to gain insight into the structure of NSs (e.g. Radice et al., 2018). Moreover, BNS mergers are thought to produce short-duration gamma-ray bursts (SGRBs; Eichler et al., 1989; Narayan et al., 1992), which is why their rates and locations have been topics of research (for a review see Berger, 2014). Furthermore, the mergers of such binaries are a prime source of heavy element (rlimit-from𝑟r-italic_r -process) enrichment throughout the Universe, and so far the only site for which direct evidence of heavy element production has been found (Berger et al., 2013; Tanvir et al., 2013; Arcavi et al., 2017; Kasen et al., 2017; Metzger, 2017; Pian et al., 2017; Smartt et al., 2017; Tanvir et al., 2017; Villar et al., 2017; Rastinejad et al., 2022; Troja et al., 2022; Levan et al., 2023, for a review see Nakar 2020). Understanding both the details of the mergers (e.g. component masses, ejecta masses, composition) and also the locations of the mergers relative to host galaxies is then critical to determining their role in cosmic chemical evolution.
After all, the merger locations of BNS with respect to their host galaxies are determined by the velocity kicks these systems receive on formation (e.g., Fryer et al., 1999; Bloom et al., 1999; Bulik et al., 1999; Belczynski et al., 2002; Perna & Belczynski, 2002; Voss & Tauris, 2003; Belczynski et al., 2006; Church et al., 2011; Abbott et al., 2017a; Zevin et al., 2020). That is, both NSs are formed in a supernova, and because of this receive a natal kick, whose magnitude has been estimated by analysing the velocities of young pulsars (Lyne & Lorimer, 1994; Hobbs et al., 2005; Faucher-Giguère & Kaspi, 2006; Arzoumanian et al., 2002; Verbunt et al., 2017; Igoshev, 2020; Igoshev et al., 2021). The total systemic kick the binary receives is then a combination of these two natal kicks and a Blaauw kick due to mass loss at each supernova (Blaauw, 1961; Hills, 1983; Van den Heuvel et al., 2000; Zhang et al., 2013; Tauris et al., 2017; Andrews & Zezas, 2019; Zhao et al., 2023). The magnitude of the BNSs’ systemic kicks has been debated: high kicks can, for instance, explain SGRBs that appear to be host-less (Berger, 2010; Fong & Berger, 2013; Tunnicliffe et al., 2014; O’Connor et al., 2022; Fong et al., 2022). In particular, while generically large kicks like those seen in pulsars can recreate the observed offsets to short-GRBs reasonably well (Church et al., 2011), Behroozi et al. (2014) find that the galaxy-offset distribution of SGRBs, as found by Fong & Berger (2013), can be explained by a SGRB progenitor kick distribution where only 19%percent1919\%19 % of objects are kicked with velocities >150absent150>150> 150 km/s. Indeed, large kicks may not be necessary to explain SGRBs (Perets & Beniamini, 2021) and Galactic BNSs appear also to have experienced small kicks, as suggested by their orbital parameters (Beniamini & Piran, 2016) and their small peculiar velocities (Gaspari et al., 2024).
We are interested in investigating whether we can constrain the systemic kicks of the Galactic BNSs, for instance by analysing their current speeds. Disberg et al. (2024) show, however, that kicked objects near the Solar System that are older than a few tens of Myr will obtain galactocentric speeds that are not representative of their kicks, due to their motion within the Galactic potential. For this reason we are interested in estimating the ages of the Galactic BNSs, and aim to give an overview of their spin-down ages, based on braking in rotational speed of the NS in the binary that is observed as pulsar, and their kinematic ages, based on their distance from the Galactic plane. For some of the Galactic BNSs there already exist estimates of the spin-down ages (e.g. Lorimer et al., 2005; Kargaltsev et al., 2006; Andrews & Mandel, 2019) and the kinematic ages (e.g. Arzoumanian et al., 1999; Wex et al., 2000; Willems et al., 2004, 2006), but we aim to give a complete overview for the Galactic BNSs systems that have known location, proper motion and no association with a globular cluster (as e.g. listed by Ding et al., 2024).
In order to investigate the kicks of the BNSs, we analysed the magnitude of their current galactocentric velocities (Monte Carlo estimated following Gaspari et al., 2024), taking into account the estimates for their ages and the deceleration found by Disberg et al. (2024). However, our estimates contain the complete velocity vectors, not just their magnitude, which provides additional information that can potentially be used to constrain their kicks. Atri et al. (2019), for instance, use estimated velocity vectors to trace back trajectories of black hole X-ray binaries to the Galactic plane, where they are thought to be formed, and determine the peculiar velocity at the disc crossing as a potential kick velocity (cf. O’Doherty et al., 2023, who apply this method to binaries with a NS component).
Likewise, we used the estimated velocity vectors in order to trace back the trajectories of the Galactic BNSs, but we are interested in using the shape of the complete trajectory in order to estimate their kicks. That is, Disberg et al. (2024) find that the initially circular Galactic orbits of kicked objects are disturbed by the kick, resulting in more eccentric Galactic orbits (as also noted by e.g. Hoang et al., 2022). We are interested in investigating whether we can use this fact to constrain kicks based on the eccentricity of the Galactic orbits of the BNS systems.
This paper is structured as follows. In Sect. 2 we investigate the ages of the Galactic BNSs. Then, in Sect. 3, we analyse the current speeds of the BNSs and use the work of Disberg et al. (2024) to show that these harbour little information about their kicks. As an alternative method to constrain kicks, we show in Sect. 4 that the eccentricity of a Galactic orbit is a better indication of kick velocity, and estimate these eccentricities for the Galactic BNSs. In Sect. 5, we determine the relationship between kick velocity and eccentricity of the Galactic orbit, and use this relationship to kinematically constrain the kicks of the BNSs. Finally, in Sect. 6, we summarise our findings and their implications.

2 Ages

We are interested in investigating the ages of the Galactic BNSs, since a young population of kicked objects might look different from an old population of kicked objects, due to their motion through the Galaxy (Disberg et al., 2024). In order to do this, we used two quantities that can tell us something about their ages: spin-down ages (Sect. 2.1) and kinematic ages (Sect. 2.2).

Table 1: Spin properties of Galactic BNS pulsars: the period (P𝑃Pitalic_P), derivative of the period (P˙˙𝑃\dot{P}over˙ start_ARG italic_P end_ARG), frequency (ν𝜈\nuitalic_ν), derivative of the frequency (ν˙˙𝜈\dot{\nu}over˙ start_ARG italic_ν end_ARG), and the references (Ref.) for the listed values.
BNS pulsar P𝑃Pitalic_P P˙˙𝑃\dot{P}over˙ start_ARG italic_P end_ARG ν𝜈\nuitalic_ν ν˙˙𝜈\dot{\nu}over˙ start_ARG italic_ν end_ARG Ref.
[103ssuperscript103𝑠10^{-3}\ s10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT italic_s] [1018ss1superscript1018𝑠superscript𝑠110^{-18}\ ss^{-1}10 start_POSTSUPERSCRIPT - 18 end_POSTSUPERSCRIPT italic_s italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT] [s1superscript𝑠1s^{-1}italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT] [1015s2superscript1015superscript𝑠210^{-15}\ s^{-2}10 start_POSTSUPERSCRIPT - 15 end_POSTSUPERSCRIPT italic_s start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT]
J0737–3039Aa𝑎aitalic_aa𝑎aitalic_aa𝑎aitalic_aTwo components of the double-pulsar binary J0737–3039A/B. (…) (…) 44.05406864196281(17)44.054068641962811744.05406864196281(17)44.05406864196281 ( 17 ) 3.4158071(11)3.415807111-3.4158071(11)- 3.4158071 ( 11 ) 1
J0737–3039Ba𝑎aitalic_aa𝑎aitalic_aa𝑎aitalic_aTwo components of the double-pulsar binary J0737–3039A/B. (…) (…) 0.36056035506(1)0.3605603550610.36056035506(1)0.36056035506 ( 1 ) 0.116(1)0.1161-0.116(1)- 0.116 ( 1 ) 2
B1913+16 (…) (…) 16.940537785677(3)16.940537785677316.940537785677(3)16.940537785677 ( 3 ) 2.4733(1)2.47331-2.4733(1)- 2.4733 ( 1 ) 3
J1913+1102 24.2850068680286(19)24.28500686802861924.2850068680286(19)24.2850068680286 ( 19 ) 0.15672(7)0.1567270.15672(7)0.15672 ( 7 ) (…) (…) 4
J0509+3801 76.5413487220(1)76.5413487220176.5413487220(1)76.5413487220 ( 1 ) 7.931(2)7.93127.931(2)7.931 ( 2 ) (…) (…) 5
J1756–2251 (…) (…) 35.1350727145469(6)35.1350727145469635.1350727145469(6)35.1350727145469 ( 6 ) 1.256079(3)1.2560793-1.256079(3)- 1.256079 ( 3 ) 6
B1534+12 (…) (…) 26.38213277689397(11)26.382132776893971126.38213277689397(11)26.38213277689397 ( 11 ) 1.686097(2)1.6860972-1.686097(2)- 1.686097 ( 2 ) 7
J1829+2456 (…) (…) 24.384401411040(6)24.384401411040624.384401411040(6)24.384401411040 ( 6 ) 0.029395(14)0.02939514-0.029395(14)- 0.029395 ( 14 ) 8
J1411+2551 62.452895517590(2)62.452895517590262.452895517590(2)62.452895517590 ( 2 ) 0.0956(51)0.0956510.0956(51)0.0956 ( 51 ) (…) (…) 9
J0453+1559 45.7818168729515(33)45.78181687295153345.7818168729515(33)45.7818168729515 ( 33 ) 0.18612(8)0.1861280.18612(8)0.18612 ( 8 ) (…) (…) 10
J1518+4904 (…) (…) 24.4289793809236(3)24.4289793809236324.4289793809236(3)24.4289793809236 ( 3 ) 0.0162263(12)0.016226312-0.0162263(12)- 0.0162263 ( 12 ) 11
J1930–1852 185.52016047926(8)185.520160479268185.52016047926(8)185.52016047926 ( 8 ) 18.001(6)18.001618.001(6)18.001 ( 6 ) (…) (…) 12
111 We follow the literature in giving either P𝑃Pitalic_P and P˙˙𝑃\dot{P}over˙ start_ARG italic_P end_ARG or ν𝜈\nuitalic_ν and ν˙˙𝜈\dot{\nu}over˙ start_ARG italic_ν end_ARG for each pulsar, even though these are essentially the same quantities. The values in parentheses show the 1σ1𝜎1\sigma1 italic_σ uncertainty in the preceding digits.
\tablebib

(1) Kramer et al. (2021); (2) Kramer et al. (2006); (3) Weisberg & Huang (2016); (4) Ferdman et al. (2020); (5) Lynch et al. (2018); (6) Ferdman et al. (2014); (7) Fonseca et al. (2014); (8) Haniewicz et al. (2021); (9) Martinez et al. (2017); (10) Martinez et al. (2015); (11) Janssen et al. (2008); (12) Swiggum et al. (2015).

Table 2: Kinematic properties of the Galactic BNSs: the right ascension (R.A.) and declination (Dec.), the proper motion in right ascension (μαsubscript𝜇𝛼\mu_{\alpha}italic_μ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT) and declination (μδsubscript𝜇𝛿\mu_{\delta}italic_μ start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT), the distance from the Sun (D𝐷Ditalic_D), and the references (Ref.) for the listed values.
BNS pulsar R.A. Dec. μαsubscript𝜇𝛼\mu_{\alpha}italic_μ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT μδsubscript𝜇𝛿\mu_{\delta}italic_μ start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT D𝐷Ditalic_D Ref.
(J2000) (J2000) [mas yr-1] [mas yr-1] [kpc]
J0737–3039A/B 07h37m51.s248115(10)superscript07hsuperscript37msuperscriptitalic-.𝑠512481151007^{\rm h}37^{\rm m}51\aas@@fstack{s}248115(10)07 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 37 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 51 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 248115 ( 10 ) 30°3940.70485(17)-30\degr 39^{\prime}40\aas@@fstack{\prime\prime}70485(17)- 30 ° 39 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 40 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 70485 ( 17 ) 2.57(3)2.573-2.57(3)- 2.57 ( 3 ) 2.08(4)2.0842.08(4)2.08 ( 4 ) 0.74(6)0.7460.74(6)0.74 ( 6 ) 1
B1913+16 19h15m27.s99942(3)superscript19hsuperscript15msuperscriptitalic-.𝑠2799942319^{\rm h}15^{\rm m}27\aas@@fstack{s}99942(3)19 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 15 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 27 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 99942 ( 3 ) +16°0627.3868(5)+16\degr 06^{\prime}27\aas@@fstack{\prime\prime}3868(5)+ 16 ° 06 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 27 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 3868 ( 5 ) 0.77(11)0.7711-0.77(11)- 0.77 ( 11 ) 0.01(14)0.0114-0.01(14)- 0.01 ( 14 ) 4.10.7+2.0subscriptsuperscript4.12.00.74.1^{+2.0}_{-0.7}4.1 start_POSTSUPERSCRIPT + 2.0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.7 end_POSTSUBSCRIPT    a𝑎aitalic_aa𝑎aitalic_aa𝑎aitalic_aNon-Gaussian probability distributions (similar to e.g. Verbiest et al., 2012). 2
J1913+1102 19h13m29.s05365(9)superscript19hsuperscript13msuperscriptitalic-.𝑠2905365919^{\rm h}13^{\rm m}29\aas@@fstack{s}05365(9)19 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 13 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 29 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 05365 ( 9 ) +11°0205.7045(22)+11\degr 02^{\prime}05\aas@@fstack{\prime\prime}7045(22)+ 11 ° 02 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 05 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 7045 ( 22 ) 3.0(5)3.05-3.0(5)- 3.0 ( 5 ) 8.7(10)8.710-8.7(10)- 8.7 ( 10 ) 7.3(15)7.3157.3(15)7.3 ( 15 )b𝑏bitalic_bb𝑏bitalic_bb𝑏bitalic_bWeighted means of the DM-based distances from the Galactic free-electron density models of Yao et al. (2017) and Cordes & Lazio (2002), as computed by Ding et al. (2024). 3
J0509+3801 05h09m31.s788(1)superscript05hsuperscript09msuperscriptitalic-.𝑠31788105^{\rm h}09^{\rm m}31\aas@@fstack{s}788(1)05 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 09 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 31 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 788 ( 1 )c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cConservatively adopted uncertainties. +38°0118.10(1)+38\degr 01^{\prime}18\aas@@fstack{\prime\prime}10(1)+ 38 ° 01 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 18 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 10 ( 1 )c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cConservatively adopted uncertainties. 2.9(1)2.912.9(1)2.9 ( 1 ) 5.9(3)5.93-5.9(3)- 5.9 ( 3 ) 4.20.9+1.6subscriptsuperscript4.21.60.94.2^{+1.6}_{-0.9}4.2 start_POSTSUPERSCRIPT + 1.6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.9 end_POSTSUBSCRIPT    a𝑎aitalic_aa𝑎aitalic_aa𝑎aitalic_aNon-Gaussian probability distributions (similar to e.g. Verbiest et al., 2012). 4, 5
J1756–2251 17h56m46.s633812(15)superscript17hsuperscript56msuperscriptitalic-.𝑠466338121517^{\rm h}56^{\rm m}46\aas@@fstack{s}633812(15)17 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 56 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 46 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 633812 ( 15 ) 22°5159.35(2)-22\degr 51^{\prime}59\aas@@fstack{\prime\prime}35(2)- 22 ° 51 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 59 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 35 ( 2 ) 2.42(8)2.428-2.42(8)- 2.42 ( 8 ) 0(20)0200(20)0 ( 20 ) 0.730.15+0.38subscriptsuperscript0.730.380.150.73^{+0.38}_{-0.15}0.73 start_POSTSUPERSCRIPT + 0.38 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.15 end_POSTSUBSCRIPT    a,d𝑎𝑑a,ditalic_a , italic_da,d𝑎𝑑a,ditalic_a , italic_dfootnotemark: a,d𝑎𝑑a,ditalic_a , italic_d 6, 7
B1534+12 15h37m09.s961730(3)superscript15hsuperscript37msuperscriptitalic-.𝑠09961730315^{\rm h}37^{\rm m}09\aas@@fstack{s}961730(3)15 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 37 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 09 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 961730 ( 3 ) +11°5555.43387(6)+11\degr 55^{\prime}55\aas@@fstack{\prime\prime}43387(6)+ 11 ° 55 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 55 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 43387 ( 6 ) 1.482(7)1.48271.482(7)1.482 ( 7 ) 25.285(12)25.28512-25.285(12)- 25.285 ( 12 ) 1.05(5)1.0551.05(5)1.05 ( 5 )c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cConservatively adopted uncertainties. 8
J1829+2456 18h29m34.s66838(6)superscript18hsuperscript29msuperscriptitalic-.𝑠3466838618^{\rm h}29^{\rm m}34\aas@@fstack{s}66838(6)18 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 29 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 34 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 66838 ( 6 ) +24°5618.2007(12)+24\degr 56^{\prime}18\aas@@fstack{\prime\prime}2007(12)+ 24 ° 56 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 18 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 2007 ( 12 ) 5.51(6)5.516-5.51(6)- 5.51 ( 6 ) 7.82(8)7.828-7.82(8)- 7.82 ( 8 ) 1.1(3)1.131.1(3)1.1 ( 3 )b𝑏bitalic_bb𝑏bitalic_bb𝑏bitalic_bWeighted means of the DM-based distances from the Galactic free-electron density models of Yao et al. (2017) and Cordes & Lazio (2002), as computed by Ding et al. (2024). 9
J1411+2551 14h11m18.s866(3)superscript14hsuperscript11msuperscriptitalic-.𝑠18866314^{\rm h}11^{\rm m}18\aas@@fstack{s}866(3)14 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 11 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 18 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 866 ( 3 ) +25°5108.39(7)+25\degr 51^{\prime}08\aas@@fstack{\prime\prime}39(7)+ 25 ° 51 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 08 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 39 ( 7 ) 3(12)312-3(12)- 3 ( 12 ) 4(9)49-4(9)- 4 ( 9 ) 1.1(2)1.121.1(2)1.1 ( 2 )b𝑏bitalic_bb𝑏bitalic_bb𝑏bitalic_bWeighted means of the DM-based distances from the Galactic free-electron density models of Yao et al. (2017) and Cordes & Lazio (2002), as computed by Ding et al. (2024). 10
J0453+1559 04h53m45.s41368(5)superscript04hsuperscript53msuperscriptitalic-.𝑠4541368504^{\rm h}53^{\rm m}45\aas@@fstack{s}41368(5)04 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 53 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 45 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 41368 ( 5 ) +15°5921.3063(59)+15\degr 59^{\prime}21\aas@@fstack{\prime\prime}3063(59)+ 15 ° 59 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 21 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 3063 ( 59 ) 5.5(5)5.55-5.5(5)- 5.5 ( 5 ) 6.0(42)6.042-6.0(42)- 6.0 ( 42 ) 0.6(4)0.640.6(4)0.6 ( 4 )b𝑏bitalic_bb𝑏bitalic_bb𝑏bitalic_bWeighted means of the DM-based distances from the Galactic free-electron density models of Yao et al. (2017) and Cordes & Lazio (2002), as computed by Ding et al. (2024). 11
J1518+4904 15h18m16.s79817(2)superscript15hsuperscript18msuperscriptitalic-.𝑠1679817215^{\rm h}18^{\rm m}16\aas@@fstack{s}79817(2)15 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 18 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 16 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 79817 ( 2 )c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cConservatively adopted uncertainties. +49°0434.1132(2)+49\degr 04^{\prime}34\aas@@fstack{\prime\prime}1132(2)+ 49 ° 04 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 34 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 1132 ( 2 )c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cConservatively adopted uncertainties. 0.69(3)0.693-0.69(3)- 0.69 ( 3 ) 8.53(8)8.538-8.53(8)- 8.53 ( 8 ) 0.81(2)0.8120.81(2)0.81 ( 2 ) 12
J1930–1852 19h30m29.s7156(7)superscript19hsuperscript30msuperscriptitalic-.𝑠297156719^{\rm h}30^{\rm m}29\aas@@fstack{s}7156(7)19 start_POSTSUPERSCRIPT roman_h end_POSTSUPERSCRIPT 30 start_POSTSUPERSCRIPT roman_m end_POSTSUPERSCRIPT 29 start_POSTFIX SUPERSCRIPTOP italic_. italic_s end_POSTFIX 7156 ( 7 ) 18°5146.27(6)-18\degr 51^{\prime}46\aas@@fstack{\prime\prime}27(6)- 18 ° 51 start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT 46 start_POSTFIX SUPERSCRIPTOP italic_. ′ ′ end_POSTFIX 27 ( 6 ) 4.3(2)4.324.3(2)4.3 ( 2 ) 5.2(4)5.24-5.2(4)- 5.2 ( 4 ) 4.61.4+2.4subscriptsuperscript4.62.41.44.6^{+2.4}_{-1.4}4.6 start_POSTSUPERSCRIPT + 2.4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1.4 end_POSTSUBSCRIPT    a𝑎aitalic_aa𝑎aitalic_aa𝑎aitalic_aNon-Gaussian probability distributions (similar to e.g. Verbiest et al., 2012). 5, 13
222 The values in parentheses are the 1σ1𝜎1\sigma1 italic_σ uncertainties in the preceding digits. If the distribution is not symmetric, we show the 68%percent6868\%68 % uncertainty in the upper and lower half of the distribution (cf. appendix A of Disberg & Nelemans, 2023).
$d$$d$footnotetext: Determined using \hrefhttp://psrpop.phys.wvu.edu/LKbias/http://psrpop.phys.wvu.edu/LKbias/.
\tablebib

(1) Kramer et al. (2021); (2) Deller et al. (2018); (3) Ferdman et al. (2020); (4) Lynch et al. (2018); (5) Ding et al. (2024); (6) Ferdman et al. (2014); (7) Verbiest et al. (2012); (8) Fonseca et al. (2014); (9) Haniewicz et al. (2021); (10) Martinez et al. (2017); (11) Martinez et al. (2015); (12) Ding et al. (2023); (13) Swiggum et al. (2015).

Table 3: Estimated time-scales of the Galactic BNS pulsars: the characteristic age (τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT, Eq. 2), kinematic age (τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT) for the first (χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT) and second (χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) disc crossing, assuming either GC-isotropy or LSR-isotropy, and the merger time (τgwsubscript𝜏gw\tau_{\text{gw}}italic_τ start_POSTSUBSCRIPT gw end_POSTSUBSCRIPT).
BNS pulsar τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT τgwsubscript𝜏gw\tau_{\text{gw}}italic_τ start_POSTSUBSCRIPT gw end_POSTSUBSCRIPTa𝑎aitalic_aa𝑎aitalic_aa𝑎aitalic_aLower bounds of the values determined following Peters (1964), similarly to Gaspari et al. (2024).
χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT
GC-isotropy LSR-isotropy GC-isotropy LSR-isotropy
[Myr] [Myr] [Myr] [Myr] [Myr] [Myr]
J0737–3039Ab𝑏bitalic_bb𝑏bitalic_bb𝑏bitalic_bTwo components of the double-pulsar binary J0737-3039A/B, which share the same τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT and τgwsubscript𝜏gw\tau_{\text{gw}}italic_τ start_POSTSUBSCRIPT gw end_POSTSUBSCRIPT. 204.49247(7)204.492477204.49247(7)204.49247 ( 7 ) 6638+41subscriptsuperscript66413866^{+41}_{-38}66 start_POSTSUPERSCRIPT + 41 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 38 end_POSTSUBSCRIPT 405+4subscriptsuperscript404540^{+4}_{-5}40 start_POSTSUPERSCRIPT + 4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 5 end_POSTSUBSCRIPT 503338+285subscriptsuperscript503285338503^{+285}_{-338}503 start_POSTSUPERSCRIPT + 285 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 338 end_POSTSUBSCRIPT 9614+14subscriptsuperscript96141496^{+14}_{-14}96 start_POSTSUPERSCRIPT + 14 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 14 end_POSTSUBSCRIPT 85858585
J0737–3039Bb𝑏bitalic_bb𝑏bitalic_bb𝑏bitalic_bTwo components of the double-pulsar binary J0737-3039A/B, which share the same τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT and τgwsubscript𝜏gw\tau_{\text{gw}}italic_τ start_POSTSUBSCRIPT gw end_POSTSUBSCRIPT. 49.3(4)49.3449.3(4)49.3 ( 4 )
B1913+16 108.596(4)108.5964108.596(4)108.596 ( 4 ) 102+2subscriptsuperscript102210^{+2}_{-2}10 start_POSTSUPERSCRIPT + 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT 71+1subscriptsuperscript7117^{+1}_{-1}7 start_POSTSUPERSCRIPT + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT 282+6subscriptsuperscript286228^{+6}_{-2}28 start_POSTSUPERSCRIPT + 6 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT 377+11subscriptsuperscript3711737^{+11}_{-7}37 start_POSTSUPERSCRIPT + 11 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 7 end_POSTSUBSCRIPT 301301301301
J1913+1102 2457(1)245712457(1)2457 ( 1 ) 4515+45subscriptsuperscript45451545^{+45}_{-15}45 start_POSTSUPERSCRIPT + 45 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 15 end_POSTSUBSCRIPT 5116+46subscriptsuperscript51461651^{+46}_{-16}51 start_POSTSUPERSCRIPT + 46 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 16 end_POSTSUBSCRIPT 6821+69subscriptsuperscript68692168^{+69}_{-21}68 start_POSTSUPERSCRIPT + 69 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 21 end_POSTSUBSCRIPT 10539+109subscriptsuperscript10510939105^{+109}_{-39}105 start_POSTSUPERSCRIPT + 109 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 39 end_POSTSUBSCRIPT 465465465465
J0509+3801 153.02(4)153.024153.02(4)153.02 ( 4 ) 41+1subscriptsuperscript4114^{+1}_{-1}4 start_POSTSUPERSCRIPT + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT 41+1subscriptsuperscript4114^{+1}_{-1}4 start_POSTSUPERSCRIPT + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT 7219+27subscriptsuperscript72271972^{+27}_{-19}72 start_POSTSUPERSCRIPT + 27 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 19 end_POSTSUBSCRIPT 6315+37subscriptsuperscript63371563^{+37}_{-15}63 start_POSTSUPERSCRIPT + 37 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 15 end_POSTSUBSCRIPT 579579579579
J1756–2251 443.494(1)443.4941443.494(1)443.494 ( 1 ) 32+31subscriptsuperscript33123^{+31}_{-2}3 start_POSTSUPERSCRIPT + 31 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT 22+34subscriptsuperscript23422^{+34}_{-2}2 start_POSTSUPERSCRIPT + 34 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT 5825+63subscriptsuperscript58632558^{+63}_{-25}58 start_POSTSUPERSCRIPT + 63 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 25 end_POSTSUBSCRIPT 5915+27subscriptsuperscript59271559^{+27}_{-15}59 start_POSTSUPERSCRIPT + 27 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 15 end_POSTSUBSCRIPT 1457145714571457
B1534+12 248.0794(3)248.07943248.0794(3)248.0794 ( 3 ) 6446+19subscriptsuperscript64194664^{+19}_{-46}64 start_POSTSUPERSCRIPT + 19 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 46 end_POSTSUBSCRIPT 5540+13subscriptsuperscript55134055^{+13}_{-40}55 start_POSTSUPERSCRIPT + 13 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 40 end_POSTSUBSCRIPT 13645+73subscriptsuperscript1367345136^{+73}_{-45}136 start_POSTSUPERSCRIPT + 73 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 45 end_POSTSUBSCRIPT 10517+67subscriptsuperscript1056717105^{+67}_{-17}105 start_POSTSUPERSCRIPT + 67 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 17 end_POSTSUBSCRIPT 2735273527352735
J1829+2456 13 152(6)13152613\,152(6)13 152 ( 6 ) 324+1subscriptsuperscript321432^{+1}_{-4}32 start_POSTSUPERSCRIPT + 1 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 4 end_POSTSUBSCRIPT 133+4subscriptsuperscript134313^{+4}_{-3}13 start_POSTSUPERSCRIPT + 4 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 3 end_POSTSUBSCRIPT 514+35subscriptsuperscript5135451^{+35}_{-4}51 start_POSTSUPERSCRIPT + 35 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 4 end_POSTSUBSCRIPT 627+18subscriptsuperscript6218762^{+18}_{-7}62 start_POSTSUPERSCRIPT + 18 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 7 end_POSTSUBSCRIPT 55 3755537555\,37555 375c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cExceeds a Hubble time.
J1411+2551 10 300(600)1030060010\,300(600)10 300 ( 600 ) 4235+84subscriptsuperscript42843542^{+84}_{-35}42 start_POSTSUPERSCRIPT + 84 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 35 end_POSTSUBSCRIPT 3217+28subscriptsuperscript32281732^{+28}_{-17}32 start_POSTSUPERSCRIPT + 28 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 17 end_POSTSUBSCRIPT 16479+140subscriptsuperscript16414079164^{+140}_{-79}164 start_POSTSUPERSCRIPT + 140 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 79 end_POSTSUBSCRIPT 10732+76subscriptsuperscript1077632107^{+76}_{-32}107 start_POSTSUPERSCRIPT + 76 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 32 end_POSTSUBSCRIPT 465 446465446465\,446465 446c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cExceeds a Hubble time.
J0453+1559 3900(2)390023900(2)3900 ( 2 ) 108+138subscriptsuperscript10138810^{+138}_{-8}10 start_POSTSUPERSCRIPT + 138 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 8 end_POSTSUBSCRIPT 157+20subscriptsuperscript1520715^{+20}_{-7}15 start_POSTSUPERSCRIPT + 20 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 7 end_POSTSUBSCRIPT 8632+137subscriptsuperscript861373286^{+137}_{-32}86 start_POSTSUPERSCRIPT + 137 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 32 end_POSTSUBSCRIPT 8115+16subscriptsuperscript81161581^{+16}_{-15}81 start_POSTSUPERSCRIPT + 16 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 15 end_POSTSUBSCRIPT 1 456 72114567211\,456\,7211 456 721c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cExceeds a Hubble time.
J1518+4904 23 870(2)23870223\,870(2)23 870 ( 2 )c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cExceeds a Hubble time. 6949+27subscriptsuperscript69274969^{+27}_{-49}69 start_POSTSUPERSCRIPT + 27 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 49 end_POSTSUBSCRIPT 249+13subscriptsuperscript2413924^{+13}_{-9}24 start_POSTSUPERSCRIPT + 13 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 9 end_POSTSUBSCRIPT 15866+67subscriptsuperscript1586766158^{+67}_{-66}158 start_POSTSUPERSCRIPT + 67 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 66 end_POSTSUBSCRIPT 852+11subscriptsuperscript8511285^{+11}_{-2}85 start_POSTSUPERSCRIPT + 11 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT 8 826 53988265398\,826\,5398 826 539c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cExceeds a Hubble time.
J1930–1852 163.40(6)163.406163.40(6)163.40 ( 6 ) 112+3subscriptsuperscript113211^{+3}_{-2}11 start_POSTSUPERSCRIPT + 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 2 end_POSTSUBSCRIPT 91+2subscriptsuperscript9219^{+2}_{-1}9 start_POSTSUPERSCRIPT + 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT 4914+89subscriptsuperscript49891449^{+89}_{-14}49 start_POSTSUPERSCRIPT + 89 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 14 end_POSTSUBSCRIPT 6623+77subscriptsuperscript66772366^{+77}_{-23}66 start_POSTSUPERSCRIPT + 77 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 23 end_POSTSUBSCRIPT 108superscript10810^{8}10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT    c𝑐citalic_cc𝑐citalic_cc𝑐citalic_cExceeds a Hubble time.
333 Quantities are determined through a Monte Carlo estimation, as described in Sect. 2. The median values of the resulting distributions and their uncertainties are given similar to Table 2.

2.1 Spin-down ages

The first quantity related to the ages of the BNS systems uses the spin properties of the observed pulsars in the binaries. During their formation pulsars get spun-up to high frequencies, corresponding to an initial period P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The spin-down time it takes for the pulsar to slow down to a period P𝑃Pitalic_P, given a certain braking-index n𝑛nitalic_n, is then given by

τsd=P(n1)P˙[1(P0P)n1].subscript𝜏sd𝑃𝑛1˙𝑃delimited-[]1superscriptsubscript𝑃0𝑃𝑛1\tau_{\text{sd}}=\dfrac{P}{(n-1)\dot{P}}\left[1-\left(\dfrac{P_{0}}{P}\right)^% {n-1}\right]\quad.italic_τ start_POSTSUBSCRIPT sd end_POSTSUBSCRIPT = divide start_ARG italic_P end_ARG start_ARG ( italic_n - 1 ) over˙ start_ARG italic_P end_ARG end_ARG [ 1 - ( divide start_ARG italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_P end_ARG ) start_POSTSUPERSCRIPT italic_n - 1 end_POSTSUPERSCRIPT ] . (1)

The characteristic age of the pulsar (τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT) is determined by assuming the braking is caused by pure dipole radiation, meaning n=3𝑛3n=3italic_n = 3 – while multipole radiation results in values of n>3𝑛3n>3italic_n > 3, which could reach values op to n=5𝑛5n=5italic_n = 5 if the braking is dominated by gravitational radiation (e.g. Camilo et al., 1994; Ferrario & Wickramasinghe, 2007) – and the initial period is small compared to the current period. This gives, as a function of P𝑃Pitalic_P and P˙˙𝑃\dot{P}over˙ start_ARG italic_P end_ARG or as a function of the frequency ν=P1𝜈superscript𝑃1\nu=P^{-1}italic_ν = italic_P start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and its derivative ν˙=P˙P2˙𝜈˙𝑃superscript𝑃2\dot{\nu}=-\dot{P}P^{-2}over˙ start_ARG italic_ν end_ARG = - over˙ start_ARG italic_P end_ARG italic_P start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT (e.g. Shapiro & Teukolsky, 1983; Arzoumanian et al., 1999):

τc=P2P˙=ν2ν˙.subscript𝜏c𝑃2˙𝑃𝜈2˙𝜈\tau_{\text{c}}=\dfrac{P}{2\dot{P}}=-\dfrac{\nu}{2\dot{\nu}}\quad.italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT = divide start_ARG italic_P end_ARG start_ARG 2 over˙ start_ARG italic_P end_ARG end_ARG = - divide start_ARG italic_ν end_ARG start_ARG 2 over˙ start_ARG italic_ν end_ARG end_ARG . (2)

For some pulsars, τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT seems to approximate their true age, for example because it aligns with an “expansion age”  for a supernova remnant based on ejecta velocities (e.g. Wyckoff & Murray, 1977), though it is not always a good estimate of the true age of a pulsar (Jiang et al., 2013; Zhang et al., 2022). For a millisecond pulsar (MSP) in a BNS, for instance, the assumptions that P0Pmuch-less-thansubscript𝑃0𝑃P_{0}\ll Pitalic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≪ italic_P and n=3𝑛3n=3italic_n = 3 may be inaccurate. These MSPs are thought to have been spun-up by accreting mass from their companion, through which they were “recycled” to millisecond periods (e.g. Bhattacharya & Van den Heuvel, 1991).

Refer to caption
Figure 1: Age-related quantities for the Galactic BNSs as discussed in Sect. 2: the characteristic age τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT (brown, through Eq. 2) and the kinematic ages τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT assumming GC-isotropy (dark green for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and light green for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT), and kinematic ages assuming LSR-isotropy (dark purple for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and light purple for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT). The boxes extend from the first to third quartile (25%percent2525\%25 % to 75%percent7575\%75 % percentile) of the distributions, while the whiskers show the 5%percent55\%5 % and 95%percent9595\%95 % percentiles and the different coloured lines show the medians. The figure also contains lines corresponding to τgwsubscript𝜏gw\tau_{\text{gw}}italic_τ start_POSTSUBSCRIPT gw end_POSTSUBSCRIPT (dash-dotted lines), the Hubble time τH=13.8subscript𝜏H13.8\tau_{\text{H}}=13.8italic_τ start_POSTSUBSCRIPT H end_POSTSUBSCRIPT = 13.8 Gyr (dotted line, Komatsu et al., 2011), as well as a background line at t=40𝑡40t=40italic_t = 40 Myr (dashed line, relevant for Sect. 3). The median values of these distributions and their uncertainties are listed in Table 3. For J0737–3039A/B we show τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT of J0737–3039B.

In order to estimate τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT, we constructed Gaussian distributions corresponding to the observed values of P𝑃Pitalic_P and P˙˙𝑃\dot{P}over˙ start_ARG italic_P end_ARG or ν𝜈\nuitalic_ν and ν˙˙𝜈\dot{\nu}over˙ start_ARG italic_ν end_ARG and their uncertainties (as given in Table 1). Then, we sampled these distributions 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT times for each BNS and for each instance determined τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT through Eq. 2. The resulting distributions of the characteristic ages are shown in Fig. 1, and their median values and uncertainties are given in Table 3. The distributions in Fig. 1 show that τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT is well constrained, due to the precision of P𝑃Pitalic_P and P˙˙𝑃\dot{P}over˙ start_ARG italic_P end_ARG measurements, and does not differ from values that can be found in literature (e.g. Kargaltsev et al., 2006; Andrews & Mandel, 2019). It remains, however, difficult to determine the true ages of these pulsars based on a spin-down age, in particular due to the (binary) evolutionary history of the MSPs. This is, for instance, reflected by the fact that the median τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT of J1518+4904 exceeds a Hubble time. Nevertheless, Maoz & Nakar (2024) argue that τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT is a “reliable indicator of age,” and that corrections for binary evolution are expected to be relatively small (based on Kiziltan & Thorsett, 2010).

2.2 Kinematic ages

As an additional way to investigate the ages of the Galactic BNSs, we estimated their kinematic ages. That is, a BNS should be formed through supernovae in the thin disc, at z0𝑧0z\approx 0italic_z ≈ 0 kpc, after which it is kicked into an orbit containing heights |z|>0𝑧0|z|\ >0| italic_z | > 0. We can use the motion of the binaries away from or towards the disc to determine the last time they crossed z=0𝑧0z=0italic_z = 0. This could potentially give a lower bound for their true age, since the binaries may have crossed the disc multiple times since their birth. However, even if a BNS has not crossed the disc multiple times in its kinematic history, there are many reasons why its kinematic age might differ (significantly) from its actual age. After all, a BNS receives two systemic kicks, after the SNe of each component, which means that it could already start to migrate through the Galaxy before receiving its final systemic kick. In addition, the kinematic ages are made uncertain by the unknown radial velocity with respect to the Solar System (vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT) and the unknown birth location of the binaries (which could be at small but non-zero |z|𝑧|z|| italic_z |). For these reasons we do not consider the kinematic ages of the BNSs as accurate estimates of their actual ages, but since there are several studies concerning the kinematic ages of individual BNSs, we determined them here to give a complete overview.
We estimated the galactocentric speeds for the Galactic BNSs in our sample. For these binaries, the sky locations, proper notions, and distances are given in Table 2. In order to determine the magnitude of their galactocentric velocity vector, we estimated the unknown velocity in the radial direction (vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT) through two different isotropy assumptions (following the method of Gaspari et al., 2024). Firstly, we assumed that the full galactocentric velocity is oriented isotropically. We refer to this kind of isotropy as “GC-isotropy.” For a BNS where the proper motion and distance translates into a transverse velocity vtsubscript𝑣𝑡v_{t}italic_v start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, the GC-isotropic estimate of vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT equals:

vr=vtcotθ,subscript𝑣𝑟subscript𝑣𝑡𝜃v_{r}=v_{t}\cot\theta\quad,italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_v start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT roman_cot italic_θ , (3)

where θ=arccosu𝜃arccosine𝑢\theta=\arccos uitalic_θ = roman_arccos italic_u and u𝑢uitalic_u is uniformly sampled between 00 and 1111. We sampled u𝑢uitalic_u 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT times, with which we created a distribution of vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT for each BNS. Secondly, we assumed that the peculiar velocity of the BNSs is oriented isotropically in the local standard of rest (LSR), which we refer to as “LSR-isotropy.” If a BNS is located at a position where the velocity vector of the LSR equals vLSRsubscript@vecvLSR\@vec{v}_{\text{LSR}}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT, then the unknown vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is estimated by subtracting the 2D transverse part of the LSR velocity vector (vLSR,tsubscript@vecvLSR𝑡\@vec{v}_{\text{LSR},\,t}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT LSR , italic_t end_POSTSUBSCRIPT) from the BNS’ vtsubscript@vecv𝑡\@vec{v}_{t}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT and adding the radial part of the LSR motion.

vr=vtvLSR,tcotθ+vLSR,r,subscript𝑣𝑟normsubscript@vecv𝑡subscript@vecvLSR𝑡𝜃subscript𝑣LSR𝑟v_{r}=\left|\left|\@vec{v}_{t}-\@vec{v}_{\text{LSR},\,t}\right|\right|\hskip 0% .85358pt\cot\theta+v_{\text{LSR},\,r}\quad,italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = | | start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT - start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT LSR , italic_t end_POSTSUBSCRIPT | | roman_cot italic_θ + italic_v start_POSTSUBSCRIPT LSR , italic_r end_POSTSUBSCRIPT , (4)

where θ𝜃\thetaitalic_θ is defined similar to Eq. 3, meaning this also resulted in a distribution of vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT containing 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT values. The LSR-isotropy assumption is appropriate if the BNS receives low kicks, and its velocity is therefore dominated by the initial circular velocity of the LSR (as determined by the Galactic potential).
We obtained a vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT distribution from assuming either GC-isotropy or LSR-isotropy, and then used the BNS sky-locations, distances, and 2D proper motions of the BNS systems (which are given in Table 2) to construct Gaussian distributions based on their mean and uncertainties, for each binary. We sampled these distributions 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT times and determined the 3D velocity vectors for each instance. Having obtained 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT sets of positions and velocities for each BNS, we used the GALPY444\hrefhttp://github.com/jobovy/galpyhttp://github.com/jobovy/galpy v.1.9.0 package (Bovy, 2015) to integrate the orbits backwards in the Milky Way (MW) potential of McMillan (2017), assuming that the circular velocity at the position of the Sun equals its azimuthal velocity (i.e. vLSR(8.122 kpc)=245.6subscript𝑣LSR8.122 kpc245.6v_{\text{LSR}}(8.122\text{ kpc})=245.6italic_v start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT ( 8.122 kpc ) = 245.6 km/s, GRAVITY Collab. et al., 2018). We evaluated these orbits every Myr, up to 1111 Gyr, in order to determine the points at which they cross z=0𝑧0z=0italic_z = 0. If z(ti)z(ti+1)<0𝑧subscript𝑡𝑖𝑧subscript𝑡𝑖10z(t_{i})\cdot z(t_{i+1})<0italic_z ( italic_t start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ⋅ italic_z ( italic_t start_POSTSUBSCRIPT italic_i + 1 end_POSTSUBSCRIPT ) < 0, meaning one has a positive z𝑧zitalic_z and the other a negative one, we set the time of crossing as the average between these consecutive timestamps. Using this method, we determined the distributions of τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for the first (χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT) and second (χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) time each BNS crosses the disc (when tracing back their kinematic history), where we neglect orbits that do not cross within 1111 Gyr, or do so at galactocentric radii R>30𝑅30R>30italic_R > 30 kpc.
In Fig. 1 we show the distributions of τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, assuming either GC-isotropy or LSR-isotropy. For most BNSs, except perhaps J0737–3039A/B and J1518+4904, the distributions of τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for GC-isotropy and LSR-isotropy agree relatively well. For J0737–3039A/B, the characteristic age overlaps with τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT assuming GC-isotropy, and χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT assuming LSR-isotropy. In general, the characteristic ages of the binaries exceed their kinematic ages signficantly. While it is difficult to estimate the reliability of the spin-down ages, the difference of one or more orders of magnitude between the spin-down and the kinematic ages does imply they were likely formed at χ3subscript𝜒absent3\chi_{\geq 3}italic_χ start_POSTSUBSCRIPT ≥ 3 end_POSTSUBSCRIPT. In Table 3 we list the median kinematic ages and their uncertainties. Also, the boxplots in Fig. 1 hide structures such as potential bimodality in the τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT distributions, which is why we show the disc-crossings in Appendix A. For most binaries, and particularly for LSR-isotropy, the τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT estimates describe the disc crossings relatively well.
Willems et al. (2004) discussed the kinematic age of J0737–3039A/B, depending on assumed values of vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and the angle of the 1D proper motion on the sky (ΩΩ\Omegaroman_Ω). They conclude that for almost all values of vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT and ΩΩ\Omegaroman_Ω, τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is less than 100100100100 Myr and τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT exceeds 20202020 Myr, both of which match our findings. In addition, Lorimer et al. (2005) employed spin-down models to constrain the age of J0747–3039A/B to 3070307030-7030 - 70 Myr, which also coincides with the region of overlap between τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT and τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT we find. Willems et al. (2006) found that certain values of vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT give rise to τkin5similar-tosubscript𝜏kin5\tau_{\text{kin}}\sim 5italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT ∼ 5 Myr for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. These solutions can also be seen in Appendix A. For B1534+12, Willems et al. (2004) found that for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT spans a range between 1111 and 210210210210 Myr, and for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT the kinematic age is at least 90909090 Myr, both of which are compatible with our τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT distributions. Arzoumanian et al. (1999), in turn, found that B1534+12 has likely crossed z=0𝑧0z=0italic_z = 0 more than once, which is supported by the fact that τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT likely exceeds τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. Arzoumanian et al. (1999) also noted that the BNS B1913+16 can be constrained well kinematically, since it has a low altitude (z𝑧zitalic_z) and is moving away from the disc. They estimated τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT to be 5similar-toabsent5\sim 5∼ 5 Myr, which is only slightly below our estimates, and τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT to be 6080608060-8060 - 80 Myr, which exceeds our estimates for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT but aligns better with the characteristic age. The estimates of Willems et al. (2004) are similar, with τkin24similar-to-or-equalssubscript𝜏kin24\tau_{\text{kin}}\simeq 2-4italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT ≃ 2 - 4 Myr for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and τkin55greater-than-or-equivalent-tosubscript𝜏kin55\tau_{\text{kin}}\gtrsim 55italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT ≳ 55 Myr for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (and are also in agreement with the analysis of Wex et al., 2000).
Moreover, we considered the merger times (τgwsubscript𝜏gw\tau_{\text{gw}}italic_τ start_POSTSUBSCRIPT gw end_POSTSUBSCRIPT) of the binaries. These are of course not age-estimates themselves, since they give the amount of time needed for these binaries to merge in the future. However, the rate at which the orbital separation of the binaries shrink – due to the emission of gravitational waves – strongly depends on the current orbital separation, meaning the binaries spend most of their time close to their initial orbital separation. This means that, on a population level (but not necessarily on an individual level), we expect binaries with longer merger times to be older than systems with younger merger times (for further discussion see Maoz & Nakar, 2024). We show the τgwsubscript𝜏gw\tau_{\text{gw}}italic_τ start_POSTSUBSCRIPT gw end_POSTSUBSCRIPT values in Fig. 1 and list them in Table 3 (which agree with values given by e.g. Faulkner et al., 2005; Tauris et al., 2017). The merger times of several BNSs exceed a Hubble time and are significantly longer than the merger times of the other binaries. With the exception of J1930–1852, these binaries indeed have the largest values of τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT.

3 Speeds

We are interested in the magnitude of kicks the Galactic BNSs experienced at their formation. As a first attempt to determine the kick velocities of the binaries, we investigated their current speeds, as estimated through the method discussed in Sect. 2.2. For each binary, we obtained 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT positions and velocities, resulting in a distribution of the present-day galactocentric speeds for each binary. In Fig. 2, we show the combined speed distributions of the BNS systems in our sample (for the velocity distributions for some of the individual BNSs, see Gaspari et al., 2024). For both GC-isotropy and LSR-isotropy, the speed distributions peak around 200250200250200-250200 - 250 km/s. Since this is comparable to the circular velocity of the Solar System, one could argue that this means the binaries are likely to have experienced low kicks. However, Disberg et al. (2024) show that this is not necessarily the case. Here, we will briefly recreate and expand their argument for why it is difficult to infer kicks based on current (galactocentric) speeds. In order to do this, we simulated kicked objects in the Galactic potential (Sect. 3.1) and compared the results to the Galactic BNSs (Sect. 3.2).

3.1 Simulation

The argument of Disberg et al. (2024) is based on a Monte Carlo simulation of objects moving through the Galactic potential after receiving a kick. First, they seeded point-masses in an assumed prior distribution which is described by an exponential disc (convolved with a prescription for the spiral arms, using the work of Chrimes et al., 2021). Here, we created a similar simulation, using the prior distribution from Disberg et al. (2024), and compared it with a simulation using a different prior. That is, we compared it with a simulation using a Gaussian annulus as prior distribution for the positions of the objects. Such a distribution was first proposed by Faucher-Giguère & Kaspi (2006), and adopted by Sartore et al. (2010) to fit the pulsar distribution found by Yusifov & Küçük (2004). They adopted the following Gaussian annulus distribution:

ρ(R)exp((R7.04 kpc)22(1.83 kpc)2),proportional-to𝜌𝑅superscript𝑅7.04 kpc22superscript1.83 kpc2\rho(R)\propto\exp\left(-\dfrac{\left(R-7.04\text{ kpc}\right)^{2}}{2\left(1.8% 3\text{ kpc}\right)^{2}}\right)\quad,italic_ρ ( italic_R ) ∝ roman_exp ( - divide start_ARG ( italic_R - 7.04 kpc ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 ( 1.83 kpc ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (5)

where R𝑅Ritalic_R is the galactocentric radius. In Fig. 3, we seeded 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT points in both the exponential disc prior and the Gaussian annulus prior, and show their positions (together with the positions of the BNSs). The figure shows the significant difference between these priors: in the exponential disc most points are located near the Galactic centre, while in the Gaussian annulus most points are located at RR𝑅subscript𝑅R\approx R_{\sun}italic_R ≈ italic_R start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT.
Similarly to Fig. 3, we seeded 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT point-masses in each prior (determining their initial positions), gave them an initial velocity corresponding to the circular velocity at their initial R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT – as dictated by the MW potential of McMillan (2017) – and added a kick velocity which is isotropically oriented (for a more detailed description, see Disberg et al., 2024). We sampled the magnitude of the kick velocity from a Maxwellian distribution, and repeated this simulation for several Maxwellians, choosing values for σ𝜎\sigmaitalic_σ of 10101010, 20202020, 50505050, 100100100100, 200200200200, and 500500500500 km/s (cf. Hobbs et al., 2005, who estimated the σ𝜎\sigmaitalic_σ for isolated pulsars to be 265265265265 km/s). Having established the initial position and initial velocity vector of the 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT point masses, we computed their trajectories using GALPY v.1.9.0 (Bovy, 2015) and the MW potential of McMillan (2017), evaluating the positions and speeds of the point-masses every Myr for 200200200200 Myr. Since we want to compare the speeds of the kicked objects to the observed speeds of the BNSs, we selected the objects that are, at a certain time, in the solar neighbourhood. We adopted the solar neighbourhood definition of Disberg et al. (2024), who – motivated by the cylindrical symmetry of the system – initialised the orbit of the Sun 12121212 times, each one azimuthally rotated by 2π/122𝜋122\pi/122 italic_π / 12, and evolved these together with the point-masses. Then, at each point in time, they evaluated the speeds of objects that are within 2222 kpc of one of the solar obits.

Refer to caption
Figure 2: Total distributions of the magnitudes of the Monte Carlo estimated present-day galactocentric velocity vectors of the BNSs (v𝑣vitalic_v), shown in normalised histograms with bins of 20202020 km/s, for GC-isotropy (green) and LSR-isotropy (purple), together with dotted lines showing the corresponding median speeds.
Refer to caption
Figure 3: Priors for the initial positions in our simulation, shown by seeding 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT points and displaying their density in normalised 2D histograms with bins of 0.50.50.50.5 kpc. The left panel shows the exponential disc prior from Disberg et al. (2024), and the right panel shows the Gaussian annulus prior as defined in Eq. 5. The black stars show the positions of the Galactic BNSs, determined as the mean of the Monte Carlo distributions described in Sect. 2.2 (cf.  Gaspari et al., 2024). The dotted lines show R=R±2𝑅plus-or-minussubscript𝑅2R=R_{\sun}\pm 2italic_R = italic_R start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT ± 2 kpc, which trace the edges of the solar neighbourhood (as defined by Disberg et al., 2024).
Refer to caption
Figure 4: Evolution of the galactocentric speeds (v𝑣vitalic_v) of the objects in the solar neighbourhood, for the different Maxwellian kick distributions (σ=10, 20, 50, 100, 200,and 500𝜎102050100200and 500\sigma=10,\ 20,\ 50,\ 100,\ 200,\ \text{and }500italic_σ = 10 , 20 , 50 , 100 , 200 , and 500 km/s, for each column), shown in 2D histograms with velocity bins of 30303030 km/s and time bins of 1111 Myr. The top row shows the results using the exponential disc prior (cf. Appendix B of Disberg et al., 2024), while the bottom row shows the results using the Gaussian annulus prior. The black lines correspond to the median speed at each time bin. We note that the colour-scale differs for each panel, because the number of objects in the solar neighbourhood differs for each kick distribution.

3.2 Deceleration

We repeated the simulation described above for each Maxwellian (σ=10𝜎10\sigma=10italic_σ = 10, 20202020, 50505050, 100100100100, 200200200200, and 500500500500 km/s), and obtained the galactocentric speeds shown in Fig. 4. In accordance with the results of Disberg et al. (2024), we find that after a certain amount of time the objects in the solar neighbourhood have decreased speeds: they have been decelerated by the Galactic potential. That is, for higher kicks the objects that we see in the solar neighbourhood have more eccentric orbits, while we are more likely to observe them closer to their apocentre where they have a lower speed (corresponding to the asymmetric drift found by Hansen & Phinney, 1997). Because of this, the speeds of older objects in the solar neighbourhood are reduced, and have a median value of 150250similar-toabsent150250\sim 150-250∼ 150 - 250 km/s, independent of the kick distribution. The speed distributions of these older objects do differ in their spread, paradoxically resulting in the lowest speeds being observed for the highest kicks. For young objects, the speeds do depend on the kicks, but these can include objects that become unbound, and are therefore not seen in the solar neighbourhood at later times. Moreover, we find no significant difference between the exponential disc and the Gaussian annulus prior, when it comes to these speeds. After all, orbits with identical eccentricity can be formed at different values of R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (albeit with a different kick magnitude and orientation).
We do find that in general there tend to be more objects in the solar neighbourhood for the Gaussian annulus prior, because in this prior more objects are formed near or in the solar neighbourhood. However, the relative amount of objects that cross the solar neighbourhood in our simulation may not be representative of the observable BNSs in the Galaxy, because these are only observable during the limited time they are visible as a radio pulsar. Also, after a certain amount of time the BNS merges and is no longer observable as a BNS, and the merger times may – to some degree – depend on the kick magnitude. Nevertheless, while these effects may be important in estimating the formation rates of BNSs from the observed Milky Way population, it has little impact on our work, which seeks to infer the kicks.

Refer to caption
Figure 5: Medians of the galactocentric speeds (vmedsubscript𝑣medv_{\text{med}}italic_v start_POSTSUBSCRIPT med end_POSTSUBSCRIPT) as shown in Fig. 4, for kick distributions described by σ=10𝜎10\sigma=10italic_σ = 10 km/s (light dotted), 20202020 km/s, (dotted), 50505050 km/s (dark dotted), 100100100100 km/s (light), 200200200200 km/s, and 500500500500 km/s (dark).
Refer to caption
Figure 6: Accuracy of the velocity distributions determined through isotropy assumptions (visosubscript𝑣isov_{\text{iso}}italic_v start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT) for the objects in our simulation (of which the speeds are shown in Fig. 4) that are in the solar neighbourhood at a certain point in time. This accuracy is estimated by dividing these distributions by the actual speeds (v𝑣vitalic_v) and evaluating viso/vsubscript𝑣iso𝑣v_{\text{iso}}/vitalic_v start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT / italic_v every 50505050 Myr, for GC-isotropy (top row, through Eq. 3) and LSR-isotropy (bottom row, through Eq. 4), and for Maxwellian kick distributions with different values of σ𝜎\sigmaitalic_σ (columns), shown in normalised 2D histogram with viso/vsubscript𝑣iso𝑣v_{\text{iso}}/vitalic_v start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT / italic_v bins of 0.10.10.10.1 and time bins of 50505050 Myr, in units of (500 Myr)1superscript500 Myr1\left(500\text{ Myr}\right)^{-1}( 500 Myr ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The black lines show the medians of the distributions.
Refer to caption
Figure 7: Evolution of the eccentricities (e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, Eq. 6) of the objects in the solar neighbourhood, for the different Maxwellian kick distributions (σ=10, 20, 50, 100, 200,and 500𝜎102050100200and 500\sigma=10,\ 20,\ 50,\ 100,\ 200,\ \text{and }500italic_σ = 10 , 20 , 50 , 100 , 200 , and 500 km/s, for each column), shown in 2D histograms with eccentricity bins of 0.050.050.050.05 and time bins of 1111 Myr. The top row shows the results using the exponential disc prior, while the bottom row shows the results using the Gaussian annulus prior. The black dashed lines show t=40𝑡40t=40italic_t = 40 Myr, approximating the timescale after which the median velocities have been decelerated (as shown in Fig. 5). Similarly to Fig. 4, we note that the colour-scale differs for each panel, because the number of objects in the solar neighbourhood differs for each kick distribution.

Figure 4 shows how the median speeds decrease rapidly for the different kick distributions, obtaining similar values independent of their initial kicks. In Fig. 5, we plot these medians together, for the exponential disc and the Gaussian annulus prior, to show the timescales of this effect. The figure shows that for the exponential disc prior, the medians for all σ𝜎\sigmaitalic_σ obtain similar values sometime between 20202020 and 30303030 Myr (as also found by Disberg et al., 2024), and that for the Gaussian annulus prior the medians become similar around 40404040 Myr. In this work, we conservatively adopt 40404040 Myr as the timescale above which we expect the objects in the solar neighbourhood to be decelerated and have similar median speeds (150250similar-toabsent150250\sim 150-250∼ 150 - 250 km/s), and therefore provide little information about the kick distribution.
In Fig. 1, we show a line at 40404040 Myr. We argue that for the Galactic BNSs, it is reasonable to estimate them all to be older than 40similar-toabsent40\sim 40∼ 40 Myr, since Fig. 1 and Table 3 show that (1) the characteristic ages of the BNSs are well above 40404040 Myr, meaning that their true ages retain values above this limit even if τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT is a considerable overestimation (with the only exception being τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT for J0737–3039B, but this is not a MSP), and (2) the kinematic ages of the binaries indicate that there is only one likely disc-crossings below 40404040 Myr (exception being B1913+16, which could have two). Because the Galactic BNSs are all likely to be older than 40404040 Myr, we expect them to have median speeds around 150250similar-toabsent150250\sim 150-250∼ 150 - 250 km/s, regardless of the kick distribution. This corresponds to the galactocentric speeds we find, as shown in Fig. 2, with median speeds 200250similar-toabsent200250\sim 200-250∼ 200 - 250 km/s. If we could estimate vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT more precisely and get more accurate estimates for the BNSs velocities, we could perhaps differentiate between the speed distributions shown in Fig. 4, but currently the uncertainty in the BNS velocity estimates prevents us from doing so. We therefore conclude that, based on the galactocentric speeds of the BNSs alone, we are not able to constrain their kicks.
Moreover, the simulation of which we show the speeds in Fig. 4 allows us to investigate the accuracy of the GC-isotropy and LSR-isotropy assumptions. We evaluated the simulated trajectories at t=0, 50, 100, 150𝑡050100150t=0,\,50,\,100,\,150italic_t = 0 , 50 , 100 , 150 and 200200200200 Myr, and considered the velocity vectors of the objects that are at these times in the solar neighbourhood (with speeds equal to v𝑣vitalic_v). These vectors are decomposed in order to obtain a vtsubscript@vecv𝑡\@vec{v}_{t}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT vector, which we combined with Eqs. 3 and 4 to determine the GC-isotropic and LSR-isotropic estimated distributions of vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. We obtained the speeds estimated through assumption of isotropy (visosubscript𝑣isov_{\text{iso}}italic_v start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT) by combining the vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT estimates with the vtsubscript@vecv𝑡\@vec{v}_{t}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT vectors. The viso/vsubscript𝑣iso𝑣v_{\text{iso}}/vitalic_v start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT / italic_v distributions are then a proxy for the accuracy of the isotropy assumptions. In Fig. 6 we show how the viso/vsubscript𝑣iso𝑣v_{\text{iso}}/vitalic_v start_POSTSUBSCRIPT iso end_POSTSUBSCRIPT / italic_v distributions evolve over time, for the different Maxwellian kick distributions and for both GC-isotropy and LSR-isotropy. For σ100𝜎100\sigma\geq 100italic_σ ≥ 100 km/s the isotropies do not differ significantly in their accuracy, since high velocities are less affected by subtracting the LSR velocity. For low velocities (σ50𝜎50\sigma\leq 50italic_σ ≤ 50 km/s), however, the velocity vectors of the simulated objects are more or less aligned with vLSRsubscript@vecvLSR\@vec{v}_{\text{LSR}}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT, due to their initial circular velocity, because of which the GC-isotropic estimates underestimate the true speeds while the LSR-isotropic estimates remain remarkably accurate. This is why we are more confident in the LSR-isotropic velocity estimates shown in Fig. 2.

4 Orbits

Despite the galactocentric speeds of the BNSs harbouring little information about their kicks, we explored whether we can use the fact that the Galactic trajectories of the BNSs appear to be more or less circular (Gaspari et al., 2024) in order to constrain their systemic kicks. In order to do this, we determined the eccentricity of the Galactic orbits we simulated in Fig. 4 and discuss the relationship between this eccentricity and kick distribution (Sect. 4.1). Moreover, we analysed the Monte Carlo estimated BNS trajectories and estimated their eccentricity (Sect. 4.2).

Refer to caption
Figure 8: Kinematic histories of the Galactic BNSs (columns), assuming either GC-isotropy (green) or LSR-isotropy (purple) in the Monte Carlo estimation, integrated backwards in time for 200200200200 Myr. These trajectories are extended to 1111 Gyr when used in Sect. 2.2 to determine the kinematic ages (and obtain z𝑧zitalic_z values as shown in Appendix A). The dashed circles show R=R=8.122𝑅subscript𝑅8.122R=R_{\sun}=8.122italic_R = italic_R start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT = 8.122 kpc (GRAVITY Collab. et al., 2018).

4.1 Eccentricity

In Sect. 3 we analysed the galactocentric velocities of the BNSs. However, when it comes to inferring kicks it may be more useful to look at their peculiar velocities (which are obtained by subtracting the local vLSRsubscript@vecvLSR\@vec{v}_{\text{LSR}}start_ID start_ARG italic_v end_ARG end_ID start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT, as determined by the MW potential). After all, for small kicks the orbits of the objects remain more or less circular, which means they have a lower peculiar velocity (vpecsubscript𝑣pecv_{\text{pec}}italic_v start_POSTSUBSCRIPT pec end_POSTSUBSCRIPT) than objects that receive high kicks, and retain these low velocities after 40similar-toabsent40\sim 40∼ 40 Myr. Nevertheless, for higher kicks vpecsubscript𝑣pecv_{\text{pec}}italic_v start_POSTSUBSCRIPT pec end_POSTSUBSCRIPT decreases similarly to v𝑣vitalic_v (Disberg et al., 2024), meaning it may still be difficult to differentiate between large kicks using the magnitude of vpecsubscript𝑣pecv_{\text{pec}}italic_v start_POSTSUBSCRIPT pec end_POSTSUBSCRIPT. Thus, for objects older than 40similar-toabsent40\sim 40∼ 40 Myr the current magnitude of vpecsubscript𝑣pecv_{\text{pec}}italic_v start_POSTSUBSCRIPT pec end_POSTSUBSCRIPT is only useful for inferring kicks insofar as it can reveal how circular the Galactic trajectory of an object is.
We therefore aim to investigate directly how the kick distribution affects the resulting Galactic orbits and their eccentricity, and whether we can use this to infer kicks. In particular, we are interested in the eccentricity of the orbits of the objects in Fig. 4, which received kicks from vastly different kick distributions. However, these Galactic orbits are not Keplerian, so they do not have an eccentricity as defined in Keplerian dynamics. Nevertheless, we define an analogue of eccentricity for Galactic orbits, by analysing the orbits of the kicked objects in our simulation and defining the minimum galactocentric radius an object obtains in their orbit as Rminsubscript𝑅R_{\min}italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT, and the maximum value of R𝑅Ritalic_R as Rmaxsubscript𝑅R_{\max}italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. Then, we compute our eccentricity-analogue:

e~=RmaxRminRmax+Rmin,~𝑒subscript𝑅subscript𝑅subscript𝑅subscript𝑅\tilde{e}=\dfrac{R_{\max}-R_{\min}}{R_{\max}+R_{\min}}\quad,over~ start_ARG italic_e end_ARG = divide start_ARG italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_ARG start_ARG italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT + italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT end_ARG , (6)

which we will refer to as the eccentricity of the Galactic orbit, or just the BNS’ eccentricity (we note that this is not related to the eccentricity within the binary). Similarly to a Keplerian eccentricity: if Rmin=Rmaxsubscript𝑅subscript𝑅R_{\min}=R_{\max}italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT the eccentricity equals 00 and the orbit is perfectly circular, and if Rmin=0subscript𝑅0R_{\min}=0italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0 or RmaxRminmuch-greater-thansubscript𝑅subscript𝑅R_{\max}\gg R_{\min}italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ≫ italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT the eccentricity goes to 1111.

Refer to caption
Figure 9: Eccentricities of the Galactic orbits of the BNSs, determined by applying Eq. 6 to the Monte Carlo estimated orbits of each binary, assuming either GC-isotropy (left panel) or LSR-isotropy (right panel), and shown in 2D histograms with e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG bins of 0.0250.0250.0250.025. For each BNS, we compute e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG for the trajectories integrated back up to tmax=200subscript𝑡200t_{\max}=200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 200 Myr (as shown in Fig. 8), and also for the trajectories with tmax=1subscript𝑡1t_{\max}=1italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1 Gyr (as used in Sect. 2.2), and compare both e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG-distributions (tmax=200subscript𝑡200t_{\max}=200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 200 Myr first, and then tmax=1subscript𝑡1t_{\max}=1italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1 Gyr below).

If we compute the Galactic orbits of kicked objects long enough, the values we find for Rminsubscript𝑅R_{\min}italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT and Rmaxsubscript𝑅R_{\max}italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT will correspond to the peri- and apocentres of the Galactic orbits, respectively. However, it is possible that within our simulations some objects do not have enough time to cross their peri- and apocentres, meaning Rminsubscript𝑅R_{\min}italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT and Rmaxsubscript𝑅R_{\max}italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT can depend – to some degree – on the computation time of the simulation (tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT). Nevertheless, even if tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT does not allow for the simulated objects to cross their peri- and apocentres, our estimates of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG will still be an approximation of the Galactic eccentricity and tell us something about the circularity of their orbits. For example, if an object received a large kick, it might start moving radially outwards in such a way that Rmin=R(t=0)subscript𝑅𝑅𝑡0R_{\min}=R(t=0)italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = italic_R ( italic_t = 0 ) and Rmax=R(t=tmax)subscript𝑅𝑅𝑡subscript𝑡R_{\max}=R(t=t_{\max})italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = italic_R ( italic_t = italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ), for a certain tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. Nevertheless, this will result in high values of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, even if this value will increase slightly if we increase tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT.
For the simulations shown in Fig. 4 (where tmax=200subscript𝑡200t_{\max}=200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 200 Myr), we determined e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG for each of the 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT objects. In Fig. 7, we show for each simulation the eccentricities of the objects that are at a certain time in the solar neighbourhood. For each object, their value of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG does not change, but at each point in time there are different objects in the solar neighbourhood, resulting in different values of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG being observed. The figure shows that for Maxwellian kick distributions with increasing σ𝜎\sigmaitalic_σ, we find objects in the solar neighbourhood with more eccentric orbits. That is, larger kicks perturb the initially circular orbits of the objects more, resulting in higher values of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG. Moreover, the difference between t40𝑡40t\leq 40italic_t ≤ 40 Myr and t>40𝑡40t>40italic_t > 40 Myr we observe in the galactocentric velocities of these objects (Fig. 4) is less prominent in their eccentricities. For t40𝑡40t\leq 40italic_t ≤ 40 Myr, the objects in the solar neighbourhood are slightly more eccentric, possibly due to objects receiving a significant kick because of which they do not return to the solar neighbourhood within tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. However, the median eccentricities do depend on σ𝜎\sigmaitalic_σ, contrary to the velocities in Fig. 4, meaning e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG can potentially allow us to distinguish between different kick distributions. Lastly, we note that in Fig. 7, similarly to Fig. 4, we find no significant differences between the results that use exponential disc and the ones that use the Gaussian annulus prior.

4.2 Trajectories

Since eccentricity can potentially provide insight in kicks, we investigated the eccentricities of the Galactic orbits of the BNSs. Similarly to Sect. 2.2, we traced the BNS trajectories back by estimating their present-day velocity vector and assuming either GC-isotropy or LSR-isotropy. In Fig. 8, we show these trajectories traced back for 200200200200 Myr (corresponding to tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT in Fig. 7). The figure shows that, despite the assumption of isotropy, most orbits are well-constrained, and some of these orbits (such as the LSR-isotropic orbits of J0737–3039A/B, J1829+2456, J1411+2551, J0453+1559, and J1518+4804) appear to have a remarkably low eccentricity, while others – and in particular the GC-isotropic orbits – seem to be more eccentric (such as the orbits of B1913+16, J0509+3801, B1534+12, and J1930–1852). However, not all orbits shown in Fig. 8 show plausible kinematic histories of the BNSs. In particular, the GC-isotropic orbits of J0737–3039A/B mostly trace back the binary to large galactocentric radii, and do not return closer to the Galactic centre within 200200200200 Myr. Since J0737–3039B has a characteristic age of 50similar-toabsent50\sim 50∼ 50, and it is unlikely that this binary was formed this far away from the Galactic centre, we do not deem these particular trajectories an accurate representation of the actual kinematic history of this binary. This may suggest that, at least in this case, LSR-isotropic orbits represent a more physical scenario. Nevertheless, the other orbits appear to be plausible, allowing us to estimate the eccentricities of these Galactic orbits.

Refer to caption
Figure 10: Number of eccentricities found in the solar neighbourhood for each vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT, shown in a 2D histogram with e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG bins of 0.050.050.050.05 and vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT bins of 10101010 km/s, on a logarithmic colour scale. For each vkick=n10subscript𝑣kick𝑛10v_{\text{kick}}=n\cdot 10italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT = italic_n ⋅ 10 km/s, we simulate 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT objects using either the exponential disc (top row) or Gaussian annulus (bottom row) prior, and evaluate their orbits in between 40404040 Myr and tmax=200subscript𝑡200t_{\max}=200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 200 Myr. The resulting eccentricities are divided into prograde (left column) and retrograde (right column) orbits. The dashed lines show 1111 and 2222 times the circular velocity (vLSRsubscript𝑣LSRv_{\text{LSR}}italic_v start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT) at a solar radius.

We used the BNS trajectories to determine Rminsubscript𝑅R_{\min}italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT and Rmaxsubscript𝑅R_{\max}italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, and then calculated e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG (through Eq. 6). This resulted in an e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG-distribution consisting of 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT eccentricities, for each binary. We determined these distributions for tmax=200subscript𝑡200t_{\max}=200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 200 Myr and for tmax=1subscript𝑡1t_{\max}=1italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1 Gyr, and repeated this for both isotropy assumptions. Figure 9 shows the resulting e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG-distributions. For the LSR-isotropic orbits, the eccentricity (1) is well-constrained, (2) obtains values where e~0.5less-than-or-similar-to~𝑒0.5\tilde{e}\lesssim 0.5over~ start_ARG italic_e end_ARG ≲ 0.5 (with the exception of J0509+3801), and (3) does not change significantly between tmax=200subscript𝑡200t_{\max}=200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 200 Myr and tmax=1subscript𝑡1t_{\max}=1italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1 Gyr. This means that (1) the eccentricity distributions appear suitable for constraining the BNS kicks, (2) these kicks will probably be relatively small (considering the results from Fig. 7), and (3) within 200200200200 Myr most of these orbits cross their peri- and apocentres (making e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG mostly independent of tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT if tmax200subscript𝑡200t_{\max}\geq 200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT ≥ 200 Myr). For the GC-isotropic eccentricities in Fig. 9, the distributions are broader and obtain higher values of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, meaning that if we deduce kicks from these distributions the GC-isotropic kicks will be larger. In fact, for some binaries the LSR-isotropic values of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG are below 0.50.50.50.5, while the GC-isotropic counterparts approach 1111 (e.g. B1913+16, J1913+1102, J1829+2456, and J1930–1852). Moreover, some of the GC-isotropic distributions shift – to a certain degree – to higher values between tmax=200subscript𝑡200t_{\max}=200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 200 Myr and tmax=1subscript𝑡1t_{\max}=1italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1 Gyr. After all, for low kicks the assumption of GC-isotropy underestimates v𝑣vitalic_v and therefore overestimates e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG.
In particular, the GC-isotropic eccentricities of J0737–3039A/B are dependent on tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. After all, these orbits are moving away from the Galactic centre (as shown in Fig. 8), meaning that increasing tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT will also increase Rmaxsubscript𝑅R_{\max}italic_R start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT and therefore e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG as well. The other GC-isotropic orbits show a similar effect, albeit less strong: there is a slight pile-up at e~0.8greater-than-or-equivalent-to~𝑒0.8\tilde{e}\gtrsim 0.8over~ start_ARG italic_e end_ARG ≳ 0.8 for tmax=1subscript𝑡1t_{\max}=1italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1 Gyr. Nevertheless, the distributions are mostly stable, due to most trajectories crossing their peri- and apocentres within 200200200200 Myr. The eccentricity of B1534+12 is particularly well-constrained, showing a narrow-peak at e~0.4~𝑒0.4\tilde{e}\approx 0.4over~ start_ARG italic_e end_ARG ≈ 0.4, for both isotropies and both values of tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT.

5 Kicks

Since we are interested in estimating the BNS kicks based on the eccentricities of their Galactic orbits, we investigated the relationship between kick and eccentricity (Sect. 5.1). Then, we applied this relationship to the eccentricity distributions we found for the Galactic BNSs in order to kinematically constrain their kicks (Sect. 5.2).

5.1 Kicks versus eccentricity

In Fig. 7 we already found a relationship between kicks and eccentricities: higher kicks cause more eccentric orbits. In order to show the exact relationship between kick magnitude and eccentricity, we expanded this simulation. That is, we repeated our simulation (as described in Sect. 3.1), but instead of a Maxwellian kick distribution we used a delta function to describe the kicks, and repeated the simulation for values of vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT between 00 km/s and 650650650650 km/s, with steps of 10101010 km/s (and for both the exponential disc and Gaussian annulus prior). For these simulations we chose tmax=200subscript𝑡200t_{\max}=200italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 200 Myr, since we found in Fig. 9 that the e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG-values of the BNSs do not change significantly beyond 200200200200 Myr. Increasing tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT beyond this value would therefore not affect our results meaningfully. Then, we evaluated the trajectories of the simulated objects every 1111 Myr, and determined the eccentricities of the orbits that are in the solar neighbourhood at that point in time. Because young objects show slightly higher values of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG (as found in Fig. 7), we left out these objects and summed all eccentricities we find in between 40404040 Myr and tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, effectively determining the relationship between vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT and e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG.

Refer to caption
Figure 11: Posterior kick distributions resulting from our eccentricity-constrained estimates, for GC-isotropy (green lines) and LSR isotropy (purple lines), and for the exponential disc (dark lines) and Gaussian annulus (light lines) prior. The panels show the normalised results from integrating the posteriors in Appendix B over e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, meaning the lines trace the values of a histogram with bins of 10101010 km/s. Each panel shows the distributions for an individual BNS, except for the lower right panel which combines all distributions.

However, if we only consider the eccentricity of an orbit without taking its direction into account, this relationship will be bimodal. After all, if we observe an object with e~=0~𝑒0\tilde{e}=0over~ start_ARG italic_e end_ARG = 0 moving along with the Solar System, we estimate that vkick=0subscript𝑣kick0v_{\text{kick}}=0italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT = 0 km/s: its initial circular orbit has not been disturbed. If, on the other hand, we observe an object with e~=0~𝑒0\tilde{e}=0over~ start_ARG italic_e end_ARG = 0 orbiting the Galactic centre in opposite direction compared to the Solar System, this would mean that it received a large kick which reversed its initial circular orbit, coincidentally resulting in a circular orbit in opposite direction. This shows that in addition to e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, we need to take into account whether the Galactic orbit of the objects are in the same or opposite direction of the Solar System. We call these two cases “prograde” and “retrograde” orbits, respectively. For the BNS trajectories in Fig. 8 and the trajectories of the objects in our simulation, we know whether they are prograde or retrograde, since we can compare their initial velocity vector to the solar velocity, and if vϕ/vϕ,0subscript𝑣italic-ϕsubscript𝑣italic-ϕ0v_{\phi}/v_{\phi,\sun}\geq 0italic_v start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT / italic_v start_POSTSUBSCRIPT italic_ϕ , ☉ end_POSTSUBSCRIPT ≥ 0 the entire orbit is prograde (due to conservation of angular momentum). For each BNS, 90%greater-than-or-equivalent-toabsentpercent90\gtrsim 90\%≳ 90 % of their LSR-isotropic trajectories are prograde. The GC-isotropic trajectories are >90%absentpercent90>90\%> 90 % prograde, except for the GC-isotropic orbits of B1913+16 (54%percent5454\%54 %), J1913+1102 (61%percent6161\%61 %), J1829+2456 (77%percent7777\%77 %), and J1930–1852 (73%percent7373\%73 %).
In Fig. 10, we show the results of these simulations, separated in prograde and retrograde orbits, for the exponential disc and the Gaussian annulus prior. If vkick=0subscript𝑣kick0v_{\text{kick}}=0italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT = 0 km/s, all orbits are prograde with e~=0~𝑒0\tilde{e}=0over~ start_ARG italic_e end_ARG = 0: the objects retain their initial, circular orbit. Then, when we increase vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT, the Galactic orbits become more eccentric. If vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT equals the circular velocity (vLSRsubscript𝑣LSRv_{\text{LSR}}italic_v start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT) at a solar radius, which is approximately vLSRsubscript𝑣LSRv_{\text{LSR}}italic_v start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT of the solar neighbourhood, it becomes possible for the kick to be of equal magnitude but opposite direction compared to the initial velocity vector. This results in a net velocity of 00 km/s, causing the object to fall to the Galactic centre, obtain Rmin=0subscript𝑅0R_{\min}=0italic_R start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 0 kpc and therefore e~=1~𝑒1\tilde{e}=1over~ start_ARG italic_e end_ARG = 1. If vkick>vLSR(R)subscript𝑣kicksubscript𝑣LSRsubscript𝑅v_{\text{kick}}>v_{\text{LSR}}\left(R_{\sun}\right)italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT > italic_v start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT ), it is possible for the kick to be pointed in opposite direction of the initial circular velocity, and cause retrograde orbits in the solar neighbourhood. At this point, a larger kick means that the retrograde orbits can obtain velocities closer to vLSR(R)subscript𝑣LSRsubscript𝑅v_{\text{LSR}}\left(R_{\sun}\right)italic_v start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT ), resulting in a decrease of the retrograde values of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG. This continues up until vkick=2vLSR(R)subscript𝑣kick2subscript𝑣LSRsubscript𝑅v_{\text{kick}}=2v_{\text{LSR}}\left(R_{\sun}\right)italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT = 2 italic_v start_POSTSUBSCRIPT LSR end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT ☉ end_POSTSUBSCRIPT ), where the kick can cause perfectly circular, retrograde orbits. Increasing vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT beyond this point will result again in more eccentric orbits, approaching limvkicke~=1subscriptsubscript𝑣kick~𝑒1\lim_{v_{\text{kick}}\to\infty}\tilde{e}=1roman_lim start_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT → ∞ end_POSTSUBSCRIPT over~ start_ARG italic_e end_ARG = 1.
We do not find significant differences between the results that use the exponential disc and the ones that use the Gaussian annulus prior, in particular for e~0.6less-than-or-similar-to~𝑒0.6\tilde{e}\lesssim 0.6over~ start_ARG italic_e end_ARG ≲ 0.6 the distributions are remarkably similar. For e~0.6greater-than-or-equivalent-to~𝑒0.6\tilde{e}\gtrsim 0.6over~ start_ARG italic_e end_ARG ≳ 0.6, there are a few differences between the two rows in Fig. 9. The prograde orbits which use the Gaussian annulus prior and are extremely eccentric tend to be the results of slightly lower kicks than their exponential disc counterparts. This can be explained by the fact that the exponential disc prior seeds the objects closer the Galactic centre and thus deeper in the Galactic potential, meaning that it takes a larger kick for them to obtain orbits that travel to extremely large galactocentric radii and do not cross the solar neighbourhood between 40404040 Myr and 200200200200 Myr. In other words, we expect that for highly eccentric orbits (i.e. e~0.8greater-than-or-equivalent-to~𝑒0.8\tilde{e}\gtrsim 0.8over~ start_ARG italic_e end_ARG ≳ 0.8), the results start to depend – to some degree – on tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. This can also be found in the retrograde orbits for the Gaussian annulus: vkick500greater-than-or-equivalent-tosubscript𝑣kick500v_{\text{kick}}\gtrsim 500italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ≳ 500 km/s results in the most eccentric orbits to not be found in the solar neighbourhood (these kicks are also large enough to cause objects to become unbound, depending on the alignment between kick and initial velocity vector). However, for our purposes the relationship between vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT and e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG found in Fig. 10 suffices, since the trajectories of the BNSs mostly contain prograde orbits with e~<0.8~𝑒0.8\tilde{e}<0.8over~ start_ARG italic_e end_ARG < 0.8, except for a few GC-isotropic orbits (which we are less confident in).

5.2 Kick velocities

Now that we have estimated the distributions of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG for the BNSs (Fig. 9) and have determined the relationship between vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT and e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG (Fig. 10), we combined the two in order to estimate the BNS kicks. That is, for the BNSs we took each value of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG in their eccentricity distribution, determined whether the corresponding orbit is prograde or retrograde, and matched the value to the corresponding column in Fig. 10. Then, we took out that column, normalised it, and put it in a posterior distribution. We repeated this (1) for each e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG in the distribution of Fig. 9, (2) for the exponential disc and the Gaussian annulus prior, and (3) for the GC-istropy and LSR-isotropy assumption. Effectively, this means we weighted each column in Fig. 10 based on the e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG distributions we found in Fig. 9 (and combined the prograde and retrograde orbits). In Appendix B we show the posterior distributions for each binary. We integrated these posteriors over e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG and normalised the result, which gave the posterior probability distributions of vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT (as discussed in Appendix B).
In Fig. 11 we show our estimated distributions of vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT for each individual BNS, as well as the combined kick distribution. Due to the LSR-isotropic orbits being more circular, they result in smaller estimated kicks compared to the GC-isotropic ones. GC-isotropy, after all, tends to underestimate low velocities (Fig. 6), resulting in an overestimated eccentricity (Fig. 9), and therefore also overestimate the corresponding kick magnitude. Nevertheless, for some BNSs (i.e. J0509+3801, B1534+12 and J1411+2551) the LSR-istropic and GC-isotropic estimates agree well. Moreover, the exponential disc and Gaussian annulus prior result in similar kick estimates, with the GC-isotropic estimates for J0737–3039A/B, B1913+16, J1913+1102, and J1930–1852 showing the largest differences. This is caused by their highly eccentric orbits (Fig. 9) and the fact that the major differences between the two priors occur at high eccentricities (Fig. 10). Because of this, together with the fact that (1) some GC-isotropic orbits do not appear to be plausible kinematic histories of the BNSs, (2) the highly eccentric orbits in Fig. 10 start to be dependent on tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, and (3) LSR-isotropic velocity estimates are more accurate for low velocities, we are more confident in the LSR-isotropic kick estimates.
The lower right panel in Fig. 11 contains the total kick distributions for the Galactic BNSs, showing that the LSR-isotropic estimates peak at 4050similar-toabsent4050\sim 40-50∼ 40 - 50 km/s and the GC-isotropic estimates peak at 150200similar-toabsent150200\sim 150-200∼ 150 - 200 km/s. In Appendix C we fit a log-normal and a Maxwellian distribution to these estimates, and show that the LSR-isotropic estimates – which we are more confident in than their GC-isotropic counterparts – peak only slightly below the kick distribution found by O’Doherty et al. (2023), who estimate the kicks of NSs with low-mass (non-NS) companions using the method of Atri et al. (2019). We find that our LSR-isotropic estimates (i.e. the average between the results from the exponential disc and Gaussian annulus prior) are well-described by a log-normal distribution that peaks at 43±2plus-or-minus43243\pm 243 ± 2 km/s, and note that vkick50subscript𝑣kick50v_{\text{kick}}\leq 50italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ≤ 50 km/s is usually seen as the region of low or no kicks, since it equals the typical escape velocity of globular clusters (e.g. Atri et al., 2019; Mandel & Müller, 2020; Willcox et al., 2021; O’Doherty et al., 2023). In our log-normal fit to the LSR-isotropic estimates, 23%percent2323\%23 % of the kick velocities are 50absent50\leq 50≤ 50 km/s.
The method we used to constrain the BNS kicks is to some degree similar to the method Atri et al. (2019) applied to black hole X-ray binaries. They Monte Carlo estimated the 3D velocity vector of these binaries and then traced their trajectories back in time. Then, instead of analysing the eccentricity of the entire orbit, they looked at the peculiar velocities at the instances when the binaries crossed the disc and considered these as potential kick velocities. Using this method, they obtain a potential kick distribution for each binary (O’Doherty et al. 2023 note that this method is slightly biased towards low kicks, since these produce more disc crossings, which can be corrected for by considering the same amount of disc crossings for each trajectory). As a comparison between our method and the method of Atri et al. (2019), we considered the first disc crossings of the BNS trajectories (χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, used to estimate τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT in Sect. 2.2 and displayed in Appendix A), and determined their peculiar velocities at these points. We compare these peculiar velocities to our kick estimates in Appendix D, and find that they match remarkably well. Our method does, however, differ from the method of Atri et al. (2019), meaning different assumptions were made in order to estimate the kicks. For example, our method assumes a prior distribution of objects before they are kicked (Fig. 3), whereas their method assumes that averaging peculiar velocities over disc crossings approximates the kick velocity. Also, we note that both methods assume that objects are initially located at z=0𝑧0z=0italic_z = 0 with no peculiar velocity, but this can be potentially be altered in order to describe objects for which this is not an accurate description.

6 Conclusion

In this work, we determined the characteristic spin-down and kinematic ages of the Galactic BNSs (Sect. 2), and examined whether their galactocentric speeds are representative of their kicks (Sect. 3). As a novel method to constrain the BNS kicks, we determined the eccentricity of their Galactic orbits (Sect. 4) and estimated their kicks through these eccentricities (Sect. 5). Below, we summarise our findings for each section:

  • The characteristic ages (through Table 1 and Eq. 2) and the kinematic ages (through the velocity vectors estimated with the method of Gaspari et al., 2024, which assumes either GC-isotropy or LSR-isotropy) of the Galactic BNSs indicate that they are likely to have ages 40greater-than-or-equivalent-toabsent40\gtrsim 40≳ 40 Myr (Fig. 1), even though it remains difficult to relate these estimates to the true ages of the BNSs.

  • The Galactic BNSs have median galactocentric speeds of 200250similar-toabsent200250\sim 200-250∼ 200 - 250 km/s (Fig. 2). However, using the findings of Disberg et al. (2024), and expanding them with a different prior spatial distribution (Fig. 3), we showed that vastly different kick distributions all lead to median galactocentric speeds in the solar neighbourhood of 150250similar-toabsent150250\sim 150-250∼ 150 - 250 km/s (Fig. 4) after 40similar-toabsent40\sim 40∼ 40 Myr (Fig. 5). Since the BNS ages likely exceed this value, this means we cannot constrain their kicks based on these speeds. Also, we find that GC-istotropic estimates tend to underestimate low velocities, while LSR-isotropic estimates remain accurate (Fig. 6).

  • However, we find that despite the fact that different kick distributions result in similar galactocentric speeds, the eccentricities (Eq. 6) of their Galactic trajectories do depend on the kick distribution, also for t40greater-than-or-equivalent-to𝑡40t\gtrsim 40italic_t ≳ 40 Myr (Fig. 7). This indicates that if we investigate the Galactic orbits of the BNSs (Fig. 8), and determine their eccentricity (Fig. 9), we can potentially infer their kick velocities.

  • In order to do this, we simulated populations of kicked objects and determine the eccentricities of the Galactic orbits of the objects that are in the solar neighbourhood between 40404040 Myr and 200200200200 Myr. We repeated this for different kick values, which resulted in a relationship between kick and eccentricity (Fig. 10). Combining this relationship with the BNS eccentricities we found, we constructed kick estimates for each individual BNS, as well as a total kick distribution (Fig. 11). We find that the exponential disc and Gaussian annulus prior both give similar results, whereas the GC-isotropy results in higher kick estimates than the LSR-isotropy counterparts. We are more confident in the LSR-isotropy estimates, and find that they are well-described by a log-normal distribution peaking at 4050similar-toabsent4050\sim 40-50∼ 40 - 50 km/s.

Moreover, we note that (1) O’Doherty et al. (2023) find a kick distribution for NSs with low-mass companions which peaks at slightly higher values compared to our (LSR-isotropic) distribution (Appendix C), and (2) applying the method of Atri et al. (2019) to the Galactic BNSs gives results similar to ours, even if we just limit this to the first disc crossing (Appendix D).
Our results are made uncertain by several factors. For instance, in our simulation we assume that when the BNSs are kicked, they are on perfectly circular Galactic orbits (i.e. they have no peculiar velocity) at z=0𝑧0z=0italic_z = 0 kpc. This means that our results are uncertain to the degree that BNSs can have non-zero peculiar velocities pre-kick, and are located at |z|>0𝑧0|z|>0| italic_z | > 0 kpc when they are kicked. It is likely that before the final kick at the time of the second supernova, the progenitors have already acquired peculiar velocities due to the previous SN event. Moreover, the BNS kick distribution we find is a description of eleven observed Galactic BNSs. Because of this, it is not obvious that the distribution we find is representative of all BNS kicks. It is, for example, conceivable that BNS merger times depend on the magnitude of their kick (e.g. Beniamini & Piran, 2024), possibly resulting in a bias against high kicks, or are perhaps influenced by their Galactic trajectories (Stegmann et al., 2024). One could also suspect that highly kicked objects are less likely to be observed close to the Solar System. However, we argue that Fig. 10 captures kicks up to at least 500500500500 km/s: the figure shows that objects that received kicks in between our Galactic BNS kick estimates and 500greater-than-or-equivalent-toabsent500\gtrsim 500≳ 500 km/s do cross the solar neighbourhood within the time frame of the simulations. Since these kicks do not seem to be present in our BNS sample, albeit a relatively small sample, we argue that our simulation accounts for this bias up to vkick500subscript𝑣kick500v_{\text{kick}}\approx 500italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ≈ 500 km/s. Nevertheless, we cannot decisively exclude the possibility of an undetected BNS population that receives extremely high kicks (500greater-than-or-equivalent-toabsent500\gtrsim 500≳ 500 km/s). We do note that these uncertainties apply to our method as well as the method of Atri et al. (2019).
Behroozi et al. (2014) find that in order to explain the observed offsets of SGRBs, one needs a SGRB progenitor kick distribution where one in five objects receives a kick with a magnitude >150absent150>150> 150 km/s. In the log-normal fit to our (LSR-isotropic) results, the region where vkick>150subscript𝑣kick150v_{\text{kick}}>150italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT > 150 km/s contains approximately 30%percent3030\%30 % of the probability density, which is more or less consistent with the percentage required by Behroozi et al. (2014). We therefore state that, while the BNS kick distribution we find may peak at low kicks (4050similar-toabsent4050\sim 40-50∼ 40 - 50 km/s), it still contains the amount of high kicks required by Behroozi et al. (2014) to match the galaxy-offset distribution found by Fong & Berger (2013).
Future research could expand on several aspects of our analysis, but it would be particularly interesting to expand the simulations behind Fig. 10. That is, it would be interesting to increase (1) the size of the simulations, (2) the value of tmaxsubscript𝑡t_{\max}italic_t start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT, and (3) the range of kicks being sampled, in order to estimate the relationship between kick and the eccentricity of the Galactic orbit more robustly. This could give a more accurate estimate for the kick corresponding to a highly eccentric orbit (e~0.9greater-than-or-equivalent-to~𝑒0.9\tilde{e}\gtrsim 0.9over~ start_ARG italic_e end_ARG ≳ 0.9), after which it could be applied to objects that are thought to receive higher kicks (such as isolated pulsars). After all, our estimation does not use the fact that the objects we describe are BNSs: it can potentially be applied to all kicked objects, in order to shed more light on kicks and kick dynamics. Moreover, it would be interesting to investigate how the fact that BNS kicks consist of two separate kicks – due to the two SNe in the binary – affects our findings. These kicks, in turn, could possibly affect the merger times or radio lifetimes of the binaries, which may provide additional constraints on the estimated BNS kicks. This way, our general method for inferring kicks can be tailored for neutron star binaries.

Acknowledgements.
We thank Fiorenzo Stoppa for useful discussions regarding this Master’s project, as well as the referee for comments that helped to improve this paper. AJL was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 725246), and NG acknowledges studentship support from the Dutch Research Council (NWO) under the project number 680.92.18.02. In this work, we made use of NUMPY (Harris et al., 2020), SCIPY (Virtanen et al., 2020), MATPLOTLIB (Hunter, 2007), GALPY (Bovy, 2015), and ASTROPY, a community-developed core Python package and an ecosystem of tools and resources for astronomy (Astropy Collaboration et al., 2013, 2018, 2022).

References

  • Abbott et al. (2017a) Abbott, B. P., Abbott, R., Abbott, T. D., et al. 2017a, \hrefhttp://dx.doi.org/10.3847/2041-8213/aa93fcApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2017ApJ…850L..40A850, L40
  • Abbott et al. (2017b) Abbott, B. P., Abbott, R., Abbott, T. D., et al. 2017b, \hrefhttp://dx.doi.org/10.1103/PhysRevLett.119.161101Phys. Rev. Lett., \hrefhttps://ui.adsabs.harvard.edu/abs/2017PhRvL.119p1101A119, 161101
  • Andrews & Mandel (2019) Andrews, J. J. & Mandel, I. 2019, \hrefhttp://dx.doi.org/10.3847/2041-8213/ab2ed1ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2019ApJ…880L…8A880, L8
  • Andrews & Zezas (2019) Andrews, J. J. & Zezas, A. 2019, \hrefhttp://dx.doi.org/10.1093/mnras/stz1066MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2019MNRAS.486.3213A486, 3213
  • Arcavi et al. (2017) Arcavi, I., Hosseinzadeh, G., Howell, D. A., et al. 2017, \hrefhttp://dx.doi.org/10.1038/nature24291Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/2017Natur.551…64A551, 64
  • Arzoumanian et al. (2002) Arzoumanian, Z., Chernoff, D. F., & Cordes, J. M. 2002, \hrefhttp://dx.doi.org/10.1086/338805ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2002ApJ…568..289A568, 289
  • Arzoumanian et al. (1999) Arzoumanian, Z., Cordes, J. M., & Wasserman, I. 1999, \hrefhttp://dx.doi.org/10.1086/307482ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/1999ApJ…520..696A520, 696
  • Astropy Collaboration et al. (2022) Astropy Collaboration, Price-Whelan, A. M., Lim, P. L., et al. 2022, \hrefhttp://dx.doi.org/10.3847/1538-4357/ac7c74ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2022ApJ…935..167A935, 167
  • Astropy Collaboration et al. (2018) Astropy Collaboration, Price-Whelan, A. M., Sipőcz, B. M., et al. 2018, \hrefhttp://dx.doi.org/10.3847/1538-3881/aabc4fAJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2018AJ….156..123A156, 123
  • Astropy Collaboration et al. (2013) Astropy Collaboration, Robitaille, T. P., Tollerud, E. J., et al. 2013, \hrefhttp://dx.doi.org/10.1051/0004-6361/201322068A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2013A&A…558A..33A558, A33
  • Atri et al. (2019) Atri, P., Miller-Jones, J. C. A., Bahramian, A., et al. 2019, \hrefhttp://dx.doi.org/10.1093/mnras/stz2335MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2019MNRAS.489.3116A489, 3116
  • Behroozi et al. (2014) Behroozi, P. S., Ramirez-Ruiz, E., & Fryer, C. L. 2014, \hrefhttp://dx.doi.org/10.1088/0004-637X/792/2/123ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2014ApJ…792..123B792, 123
  • Belczynski et al. (2002) Belczynski, K., Kalogera, V., & Bulik, T. 2002, \hrefhttp://dx.doi.org/10.1086/340304ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2002ApJ…572..407B572, 407
  • Belczynski et al. (2006) Belczynski, K., Perna, R., Bulik, T., et al. 2006, \hrefhttp://dx.doi.org/10.1086/505169ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2006ApJ…648.1110B648, 1110
  • Beniamini & Piran (2016) Beniamini, P. & Piran, T. 2016, \hrefhttp://dx.doi.org/10.1093/mnras/stv2903MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2016MNRAS.456.4089B456, 4089
  • Beniamini & Piran (2024) Beniamini, P. & Piran, T. 2024, \hrefhttp://dx.doi.org/10.3847/1538-4357/ad32cdApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2024ApJ…966…17B966, 17
  • Berger (2010) Berger, E. 2010, \hrefhttp://dx.doi.org/10.1088/0004-637X/722/2/1946ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2010ApJ…722.1946B722, 1946
  • Berger (2014) Berger, E. 2014, \hrefhttp://dx.doi.org/10.1146/annurev-astro-081913-035926ARA&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2014ARA&A..52…43B52, 43
  • Berger et al. (2013) Berger, E., Fong, W., & Chornock, R. 2013, \hrefhttp://dx.doi.org/10.1088/2041-8205/774/2/L23ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2013ApJ…774L..23B774, L23
  • Bhattacharya & Van den Heuvel (1991) Bhattacharya, D. & Van den Heuvel, E. P. J. 1991, \hrefhttp://dx.doi.org/10.1016/0370-1573(91)90064-SPhys. Rep, \hrefhttps://ui.adsabs.harvard.edu/abs/1991PhR…203….1B203, 1
  • Blaauw (1961) Blaauw, A. 1961, Bull. Astron. Inst. Netherlands, \hrefhttps://ui.adsabs.harvard.edu/abs/1961BAN….15..265B15, 265
  • Bloom et al. (1999) Bloom, J. S., Sigurdsson, S., & Pols, O. R. 1999, \hrefhttp://dx.doi.org/10.1046/j.1365-8711.1999.02437.xMNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/1999MNRAS.305..763B305, 763
  • Bovy (2015) Bovy, J. 2015, \hrefhttp://dx.doi.org/10.1088/0067-0049/216/2/29ApJS, \hrefhttps://ui.adsabs.harvard.edu/abs/2015ApJS..216…29B216, 29
  • Bulik et al. (1999) Bulik, T., Belczyński, K., & Zbijewski, W. 1999, \hrefhttp://dx.doi.org/10.1046/j.1365-8711.1999.02878.xMNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/1999MNRAS.309..629B309, 629
  • Camilo et al. (1994) Camilo, F., Thorsett, S. E., & Kulkarni, S. R. 1994, \hrefhttp://dx.doi.org/10.1086/187176ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/1994ApJ…421L..15C421, L15
  • Chrimes et al. (2021) Chrimes, A. A., Levan, A. J., Groot, P. J., Lyman, J. D., & Nelemans, G. 2021, \hrefhttp://dx.doi.org/10.1093/mnras/stab2676MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2021MNRAS.508.1929C508, 1929
  • Church et al. (2011) Church, R. P., Levan, A. J., Davies, M. B., & Tanvir, N. 2011, \hrefhttp://dx.doi.org/10.1111/j.1365-2966.2011.18277.xMNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2011MNRAS.413.2004C413, 2004
  • Cordes & Lazio (2002) Cordes, J. M. & Lazio, T. J. W. 2002, \hrefhttps://ui.adsabs.harvard.edu/abs/2002astro.ph..7156C\hrefhttp://dx.doi.org/10.48550/arXiv.astro-ph/0207156arXiv e-prints, astro
  • Deller et al. (2018) Deller, A. T., Weisberg, J. M., Nice, D. J., & Chatterjee, S. 2018, \hrefhttp://dx.doi.org/10.3847/1538-4357/aacf95ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2018ApJ…862..139D862, 139
  • Ding et al. (2023) Ding, H., Deller, A. T., Stappers, B. W., et al. 2023, \hrefhttp://dx.doi.org/10.1093/mnras/stac3725MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2023MNRAS.519.4982D519, 4982
  • Ding et al. (2024) Ding, H., Deller, A. T., Swiggum, J. K., et al. 2024, \hrefhttps://ui.adsabs.harvard.edu/abs/2024arXiv240503914D\hrefhttp://dx.doi.org/10.48550/arXiv.2405.03914arXiv e-prints, arXiv:2405.03914
  • Disberg et al. (2024) Disberg, P., Gaspari, N., & Levan, A. J. 2024, \hrefhttp://dx.doi.org/10.1051/0004-6361/202449996A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2024arXiv240506436D687, A272
  • Disberg & Nelemans (2023) Disberg, P. & Nelemans, G. 2023, \hrefhttp://dx.doi.org/10.1051/0004-6361/202245693A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2023A&A…676A..31D676, A31
  • Eichler et al. (1989) Eichler, D., Livio, M., Piran, T., & Schramm, D. N. 1989, \hrefhttp://dx.doi.org/10.1038/340126a0Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/1989Natur.340..126E340, 126
  • Faucher-Giguère & Kaspi (2006) Faucher-Giguère, C.-A. & Kaspi, V. M. 2006, \hrefhttp://dx.doi.org/10.1086/501516ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2006ApJ…643..332F643, 332
  • Faulkner et al. (2005) Faulkner, A. J., Kramer, M., Lyne, A. G., et al. 2005, \hrefhttp://dx.doi.org/10.1086/427776ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2005ApJ…618L.119F618, L119
  • Ferdman et al. (2020) Ferdman, R. D., Freire, P. C. C., Perera, B. B. P., et al. 2020, \hrefhttp://dx.doi.org/10.1038/s41586-020-2439-xNature, \hrefhttps://ui.adsabs.harvard.edu/abs/2020Natur.583..211F583, 211
  • Ferdman et al. (2014) Ferdman, R. D., Stairs, I. H., Kramer, M., et al. 2014, \hrefhttp://dx.doi.org/10.1093/mnras/stu1223MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2014MNRAS.443.2183F443, 2183
  • Ferrario & Wickramasinghe (2007) Ferrario, L. & Wickramasinghe, D. 2007, \hrefhttp://dx.doi.org/10.1111/j.1365-2966.2006.11365.xMNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2007MNRAS.375.1009F375, 1009
  • Fong & Berger (2013) Fong, W. & Berger, E. 2013, \hrefhttp://dx.doi.org/10.1088/0004-637X/776/1/18ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2013ApJ…776…18F776, 18
  • Fong et al. (2022) Fong, W.-f., Nugent, A. E., Dong, Y., et al. 2022, \hrefhttp://dx.doi.org/10.3847/1538-4357/ac91d0ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2022ApJ…940…56F940, 56
  • Fonseca et al. (2014) Fonseca, E., Stairs, I. H., & Thorsett, S. E. 2014, \hrefhttp://dx.doi.org/10.1088/0004-637X/787/1/82ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2014ApJ…787…82F787, 82
  • Fryer et al. (1999) Fryer, C. L., Woosley, S. E., & Hartmann, D. H. 1999, \hrefhttp://dx.doi.org/10.1086/307992ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/1999ApJ…526..152F526, 152
  • Gaspari et al. (2024) Gaspari, N., Levan, A. J., Chrimes, A. A., & Nelemans, G. 2024, \hrefhttp://dx.doi.org/10.1093/mnras/stad3259MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2024MNRAS.527.1101G527, 1101
  • GRAVITY Collab. et al. (2018) GRAVITY Collab., Abuter, R., Amorim, A., et al. 2018, \hrefhttp://dx.doi.org/10.1051/0004-6361/201833718A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2018A&A…615L..15G615, L15
  • Haniewicz et al. (2021) Haniewicz, H. T., Ferdman, R. D., Freire, P. C. C., et al. 2021, \hrefhttp://dx.doi.org/10.1093/mnras/staa3466MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2021MNRAS.500.4620H500, 4620
  • Hansen & Phinney (1997) Hansen, B. M. S. & Phinney, E. S. 1997, \hrefhttp://dx.doi.org/10.1093/mnras/291.3.569MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/1997MNRAS.291..569H291, 569
  • Harris et al. (2020) Harris, C. R., Millman, K. J., Van der Walt, S. J., et al. 2020, \hrefhttp://dx.doi.org/10.1038/s41586-020-2649-2Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/2020Natur.585..357H585, 357
  • Hills (1983) Hills, J. G. 1983, \hrefhttp://dx.doi.org/10.1086/160871ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/1983ApJ…267..322H267, 322
  • Hoang et al. (2022) Hoang, B.-M., Naoz, S., & Sloneker, M. 2022, \hrefhttp://dx.doi.org/10.3847/1538-4357/ac7787ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2022ApJ…934…54H934, 54
  • Hobbs et al. (2005) Hobbs, G., Lorimer, D. R., Lyne, A. G., & Kramer, M. 2005, \hrefhttp://dx.doi.org/10.1111/j.1365-2966.2005.09087.xMNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2005MNRAS.360..974H360, 974
  • Hunter (2007) Hunter, J. D. 2007, \hrefhttp://dx.doi.org/10.1109/MCSE.2007.55Computing in Science and Engineering, \hrefhttps://ui.adsabs.harvard.edu/abs/2007CSE…..9…90H9, 90
  • Igoshev (2020) Igoshev, A. P. 2020, \hrefhttp://dx.doi.org/10.1093/mnras/staa958MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2020MNRAS.494.3663I494, 3663
  • Igoshev et al. (2021) Igoshev, A. P., Chruslinska, M., Dorozsmai, A., & Toonen, S. 2021, \hrefhttp://dx.doi.org/10.1093/mnras/stab2734MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2021MNRAS.508.3345I508, 3345
  • Janssen et al. (2008) Janssen, G. H., Stappers, B. W., Kramer, M., et al. 2008, \hrefhttp://dx.doi.org/10.1051/0004-6361:200810076A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2008A&A…490..753J490, 753
  • Jiang et al. (2013) Jiang, L., Zhang, C.-M., Tanni, A., & Zhao, H.-H. 2013, \hrefhttp://dx.doi.org/10.1142/S2010194513011124International Journal of Modern Physics Conference Series, \hrefhttps://ui.adsabs.harvard.edu/abs/2013IJMPS..23…95J23, 95
  • Kargaltsev et al. (2006) Kargaltsev, O., Pavlov, G. G., & Garmire, G. P. 2006, \hrefhttp://dx.doi.org/10.1086/504837ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2006ApJ…646.1139K646, 1139
  • Kasen et al. (2017) Kasen, D., Metzger, B., Barnes, J., Quataert, E., & Ramirez-Ruiz, E. 2017, \hrefhttp://dx.doi.org/10.1038/nature24453Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/2017Natur.551…80K551, 80
  • Kiziltan & Thorsett (2010) Kiziltan, B. & Thorsett, S. E. 2010, \hrefhttp://dx.doi.org/10.1088/0004-637X/715/1/335ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2010ApJ…715..335K715, 335
  • Komatsu et al. (2011) Komatsu, E., Smith, K. M., Dunkley, J., et al. 2011, \hrefhttp://dx.doi.org/10.1088/0067-0049/192/2/18ApJS, \hrefhttps://ui.adsabs.harvard.edu/abs/2011ApJS..192…18K192, 18
  • Kramer et al. (2006) Kramer, M., Stairs, I. H., Manchester, R. N., et al. 2006, \hrefhttp://dx.doi.org/10.1126/science.1132305Sci, \hrefhttps://ui.adsabs.harvard.edu/abs/2006Sci…314…97K314, 97
  • Kramer et al. (2021) Kramer, M., Stairs, I. H., Manchester, R. N., et al. 2021, \hrefhttp://dx.doi.org/10.1103/PhysRevX.11.041050Phys. Rev. X, \hrefhttps://ui.adsabs.harvard.edu/abs/2021PhRvX..11d1050K11, 041050
  • Levan et al. (2023) Levan, A. J., Malesani, D. B., Gompertz, B. P., et al. 2023, \hrefhttp://dx.doi.org/10.1038/s41550-023-01998-8Nature Astronomy, \hrefhttps://ui.adsabs.harvard.edu/abs/2023NatAs…7..976L7, 976
  • Lorimer et al. (2005) Lorimer, D. R., Burgay, M., Freire, P. C. C., et al. 2005, in Astronomical Society of the Pacific Conference Series, Vol. 328, Binary Radio Pulsars, ed. F. A. Rasio & I. H. Stairs, \hrefhttps://ui.adsabs.harvard.edu/abs/2005ASPC..328..113L113
  • Lynch et al. (2018) Lynch, R. S., Swiggum, J. K., Kondratiev, V. I., et al. 2018, \hrefhttp://dx.doi.org/10.3847/1538-4357/aabf8aApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2018ApJ…859…93L859, 93
  • Lyne & Lorimer (1994) Lyne, A. G. & Lorimer, D. R. 1994, \hrefhttp://dx.doi.org/10.1038/369127a0Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/1994Natur.369..127L369, 127
  • Mandel & Müller (2020) Mandel, I. & Müller, B. 2020, \hrefhttp://dx.doi.org/10.1093/mnras/staa3043MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2020MNRAS.499.3214M499, 3214
  • Maoz & Nakar (2024) Maoz, D. & Nakar, E. 2024, \hrefhttps://ui.adsabs.harvard.edu/abs/2024arXiv240608630M\hrefhttp://dx.doi.org/10.48550/arXiv.2406.08630arXiv e-prints, arXiv:2406.08630
  • Martinez et al. (2015) Martinez, J. G., Stovall, K., Freire, P. C. C., et al. 2015, \hrefhttp://dx.doi.org/10.1088/0004-637X/812/2/143ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2015ApJ…812..143M812, 143
  • Martinez et al. (2017) Martinez, J. G., Stovall, K., Freire, P. C. C., et al. 2017, \hrefhttp://dx.doi.org/10.3847/2041-8213/aa9d87ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2017ApJ…851L..29M851, L29
  • McMillan (2017) McMillan, P. J. 2017, \hrefhttp://dx.doi.org/10.1093/mnras/stw2759MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2017MNRAS.465…76M465, 76
  • Metzger (2017) Metzger, B. D. 2017, \hrefhttps://ui.adsabs.harvard.edu/abs/2017arXiv171005931M\hrefhttp://dx.doi.org/10.48550/arXiv.1710.05931arXiv e-prints, arXiv:1710.05931
  • Nakar (2020) Nakar, E. 2020, \hrefhttp://dx.doi.org/10.1016/j.physrep.2020.08.008Phys. Rep, \hrefhttps://ui.adsabs.harvard.edu/abs/2020PhR…886….1N886, 1
  • Narayan et al. (1992) Narayan, R., Paczynski, B., & Piran, T. 1992, \hrefhttp://dx.doi.org/10.1086/186493ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/1992ApJ…395L..83N395, L83
  • O’Connor et al. (2022) O’Connor, B., Troja, E., Dichiara, S., et al. 2022, \hrefhttp://dx.doi.org/10.1093/mnras/stac1982MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2022MNRAS.515.4890O515, 4890
  • O’Doherty et al. (2023) O’Doherty, T. N., Bahramian, A., Miller-Jones, J. C. A., et al. 2023, \hrefhttp://dx.doi.org/10.1093/mnras/stad680MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2023MNRAS.521.2504O521, 2504
  • Perets & Beniamini (2021) Perets, H. B. & Beniamini, P. 2021, \hrefhttp://dx.doi.org/10.1093/mnras/stab794MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2021MNRAS.503.5997P503, 5997
  • Perna & Belczynski (2002) Perna, R. & Belczynski, K. 2002, \hrefhttp://dx.doi.org/10.1086/339571ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2002ApJ…570..252P570, 252
  • Peters (1964) Peters, P. C. 1964, \hrefhttp://dx.doi.org/10.1103/PhysRev.136.B1224Phys. Rev., \hrefhttps://ui.adsabs.harvard.edu/abs/1964PhRv..136.1224P136, 1224
  • Pian et al. (2017) Pian, E., D’Avanzo, P., Benetti, S., et al. 2017, \hrefhttp://dx.doi.org/10.1038/nature24298Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/2017Natur.551…67P551, 67
  • Radice et al. (2018) Radice, D., Perego, A., Zappa, F., & Bernuzzi, S. 2018, \hrefhttp://dx.doi.org/10.3847/2041-8213/aaa402ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2018ApJ…852L..29R852, L29
  • Rastinejad et al. (2022) Rastinejad, J. C., Gompertz, B. P., Levan, A. J., et al. 2022, \hrefhttp://dx.doi.org/10.1038/s41586-022-05390-wNature, \hrefhttps://ui.adsabs.harvard.edu/abs/2022Natur.612..223R612, 223
  • Sartore et al. (2010) Sartore, N., Ripamonti, E., Treves, A., & Turolla, R. 2010, \hrefhttp://dx.doi.org/10.1051/0004-6361/200912222A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2010A&A…510A..23S510, A23
  • Shapiro & Teukolsky (1983) Shapiro, S. L. & Teukolsky, S. A. 1983, Black holes, white dwarfs and neutron stars: The physics of compact objects (Wiley-VCH)
  • Smartt et al. (2017) Smartt, S. J., Chen, T. W., Jerkstrand, A., et al. 2017, \hrefhttp://dx.doi.org/10.1038/nature24303Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/2017Natur.551…75S551, 75
  • Stegmann et al. (2024) Stegmann, J., Vigna-Gómez, A., Rantala, A., et al. 2024, \hrefhttps://ui.adsabs.harvard.edu/abs/2024arXiv240502912S\hrefhttp://dx.doi.org/10.48550/arXiv.2405.02912arXiv e-prints, arXiv:2405.02912
  • Swiggum et al. (2015) Swiggum, J. K., Rosen, R., McLaughlin, M. A., et al. 2015, \hrefhttp://dx.doi.org/10.1088/0004-637X/805/2/156ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2015ApJ…805..156S805, 156
  • Tanvir et al. (2013) Tanvir, N. R., Levan, A. J., Fruchter, A. S., et al. 2013, \hrefhttp://dx.doi.org/10.1038/nature12505Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/2013Natur.500..547T500, 547
  • Tanvir et al. (2017) Tanvir, N. R., Levan, A. J., González-Fernández, C., et al. 2017, \hrefhttp://dx.doi.org/10.3847/2041-8213/aa90b6ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2017ApJ…848L..27T848, L27
  • Tauris et al. (2017) Tauris, T. M., Kramer, M., Freire, P. C. C., et al. 2017, \hrefhttp://dx.doi.org/10.3847/1538-4357/aa7e89ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2017ApJ…846..170T846, 170
  • Troja et al. (2022) Troja, E., Fryer, C. L., O’Connor, B., et al. 2022, \hrefhttp://dx.doi.org/10.1038/s41586-022-05327-3Nature, \hrefhttps://ui.adsabs.harvard.edu/abs/2022Natur.612..228T612, 228
  • Tunnicliffe et al. (2014) Tunnicliffe, R. L., Levan, A. J., Tanvir, N. R., et al. 2014, \hrefhttp://dx.doi.org/10.1093/mnras/stt1975MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2014MNRAS.437.1495T437, 1495
  • Van den Heuvel et al. (2000) Van den Heuvel, E. P. J., Portegies Zwart, S. F., Bhattacharya, D., & Kaper, L. 2000, \hrefhttp://dx.doi.org/10.48550/arXiv.astro-ph/0005245A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2000A&A…364..563V364, 563
  • Verbiest et al. (2012) Verbiest, J. P. W., Weisberg, J. M., Chael, A. A., Lee, K. J., & Lorimer, D. R. 2012, \hrefhttp://dx.doi.org/10.1088/0004-637X/755/1/39ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2012ApJ…755…39V755, 39
  • Verbunt et al. (2017) Verbunt, F., Igoshev, A., & Cator, E. 2017, \hrefhttp://dx.doi.org/10.1051/0004-6361/201731518A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2017A&A…608A..57V608, A57
  • Villar et al. (2017) Villar, V. A., Guillochon, J., Berger, E., et al. 2017, \hrefhttp://dx.doi.org/10.3847/2041-8213/aa9c84ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2017ApJ…851L..21V851, L21
  • Virtanen et al. (2020) Virtanen, P., Gommers, R., Oliphant, T. E., et al. 2020, \hrefhttp://dx.doi.org/10.1038/s41592-019-0686-2Nature Methods, \hrefhttps://ui.adsabs.harvard.edu/abs/2020NatMe..17..261V17, 261
  • Voss & Tauris (2003) Voss, R. & Tauris, T. M. 2003, \hrefhttp://dx.doi.org/10.1046/j.1365-8711.2003.06616.xMNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2003MNRAS.342.1169V342, 1169
  • Weisberg & Huang (2016) Weisberg, J. M. & Huang, Y. 2016, \hrefhttp://dx.doi.org/10.3847/0004-637X/829/1/55ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2016ApJ…829…55W829, 55
  • Wex et al. (2000) Wex, N., Kalogera, V., & Kramer, M. 2000, \hrefhttp://dx.doi.org/10.1086/308148ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2000ApJ…528..401W528, 401
  • Willcox et al. (2021) Willcox, R., Mandel, I., Thrane, E., et al. 2021, \hrefhttp://dx.doi.org/10.3847/2041-8213/ac2cc8ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2021ApJ…920L..37W920, L37
  • Willems et al. (2004) Willems, B., Kalogera, V., & Henninger, M. 2004, \hrefhttp://dx.doi.org/10.1086/424812ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2004ApJ…616..414W616, 414
  • Willems et al. (2006) Willems, B., Kaplan, J., Fragos, T., Kalogera, V., & Belczynski, K. 2006, \hrefhttp://dx.doi.org/10.1103/PhysRevD.74.043003Phys. Rev. D, \hrefhttps://ui.adsabs.harvard.edu/abs/2006PhRvD..74d3003W74, 043003
  • Wyckoff & Murray (1977) Wyckoff, S. & Murray, C. A. 1977, \hrefhttp://dx.doi.org/10.1093/mnras/180.4.717MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/1977MNRAS.180..717W180, 717
  • Yao et al. (2017) Yao, J. M., Manchester, R. N., & Wang, N. 2017, \hrefhttp://dx.doi.org/10.3847/1538-4357/835/1/29ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2017ApJ…835…29Y835, 29
  • Yusifov & Küçük (2004) Yusifov, I. & Küçük, I. 2004, \hrefhttp://dx.doi.org/10.1051/0004-6361:20040152A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2004A&A…422..545Y422, 545
  • Zevin et al. (2020) Zevin, M., Kelley, L. Z., Nugent, A., et al. 2020, \hrefhttp://dx.doi.org/10.3847/1538-4357/abc266ApJ, \hrefhttps://ui.adsabs.harvard.edu/abs/2020ApJ…904..190Z904, 190
  • Zhang et al. (2022) Zhang, C.-M., Cui, X.-H., Li, D., et al. 2022, \hrefhttp://dx.doi.org/10.3390/universe8120628Universe, \hrefhttps://ui.adsabs.harvard.edu/abs/2022Univ….8..628Z8, 628
  • Zhang et al. (2013) Zhang, Z., Gilfanov, M., & Bogdán, Á. 2013, \hrefhttp://dx.doi.org/10.1051/0004-6361/201220685A&A, \hrefhttps://ui.adsabs.harvard.edu/abs/2013A&A…556A…9Z556, A9
  • Zhao et al. (2023) Zhao, Y., Gandhi, P., Dashwood Brown, C., et al. 2023, \hrefhttp://dx.doi.org/10.1093/mnras/stad2226MNRAS, \hrefhttps://ui.adsabs.harvard.edu/abs/2023MNRAS.525.1498Z525, 1498

Appendix A Disc crossings

Refer to caption
Figure 12: Disc-crossings of the first six Galactic BNSs, estimated using the method described in Sect. 2.2, for GC-isotropy (left column, through Eq. 3) and LSR-isotropy (right column, through Eq. 4). The black lines show the overplotted trajectories of the BNSs, where the dark green and dark purple regions correspond to the 68%percent6868\%68 % intervals of χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and the light green and light purple regions correspond to the 68%percent6868\%68 % intervals for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (as given in Table 3). The dashed lines of the same colour indicate the medians in these regions. The brown plusses show τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT, where the horizontal line extends to the 68%percent6868\%68 % interval and the vertical line is located at the median (which are also given in Table 3).
Refer to caption
Figure 13: Disc-crossings of the latter five Galactic BNSs, similarly to Fig. 12, estimated using the method described in Sect. 2.2, for GC-isotropy (left column, through Eq. 3) and LSR-isotropy (right column, through Eq. 4). The black lines show the overplotted trajectories of the BNSs, where the dark green and dark purple regions correspond to the 68%percent6868\%68 % intervals of χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and the light green and light purple regions correspond to the 68%percent6868\%68 % intervals for χ2subscript𝜒2\chi_{2}italic_χ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (as given in Table 3). The dashed lines of the same colour indicate the medians in these regions. The brown plusses show τcsubscript𝜏c\tau_{\text{c}}italic_τ start_POSTSUBSCRIPT c end_POSTSUBSCRIPT, where the horizontal line extends to the 68%percent6868\%68 % interval and the vertical line is located at the median.

In Sect. 2, we discuss different quantities that can tell us something about the ages of the Galactic BNSs (i.e. characteristic ages and kinematic ages), and show the estimated distributions in Fig. 1 and Table 3. The kinematic ages (τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT) shown in the table are determined through a Monte Carlo estimation of the BNS orbits, assuming either GC-isotropy or LSR-isotropy (Sect. 2.2). Fig. 1 shows the distributions of τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT, but in Figs. 12 and 13 we also display the trajectories of the BNS orbits in the z𝑧zitalic_z dimension, in order to show how the estimated τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT and their 68%percent6868\%68 % intervals correspond to the actual trajectories of the BNSs and their disc crossings. For some binaries, and mostly in the case of LSR-isotropy, the orbits are well-constrained and τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT is a good estimate of the disc-crossings. For others, the trajectories are less constrained by the isotropy assumptions, resulting in a larger uncertainty on τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT (on top of the methodological uncertainties in these kinematic ages). In Fig. 12, the τkin5subscript𝜏kin5\tau_{\text{kin}}\approx 5italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT ≈ 5 Myr solution for J0747–3039A/B by Willems et al. (2006) can be seen in the GC-isotropic case.

Appendix B Posteriors

Refer to caption
Figure 14: Posterior distributions of vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT versus e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, for the first six BNSs (rows), following either the GC-isotropic (two left columns) or LSR-isotropic (two right columns) estimations of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, as shown in Fig. 9. For each isotropy assumption, we compute the posterior using the results from the exponential disc (left column) or Gaussian annulus (right column) prior, as shown in Fig. 10. The posteriors are shown in 2D histograms with e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG bins of 0.050.050.050.05 and vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT bins of 10101010 km/s, and are normalised so that the (linear) colour scale shows the probability density (ρ𝜌\rhoitalic_ρ) in units of 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT s/km.
Refer to caption
Figure 15: Posterior distributions of vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT versus e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG (similar to 14), for the latter five BNSs (rows), following either the GC-isotropic (two left columns) or LSR-isotropic (two right columns) estimations of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, as shown in Fig. 9. For each isotropy assumption, we compute the posterior using the results from the exponential disc (left column) or Gaussian annulus (right column) prior, as shown in Fig. 10. The posteriors are shown in 2D histograms with e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG bins of 0.050.050.050.05 and vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT bins of 10101010 km/s, and are normalised so that the (linear) colour scale shows the probability density (ρ𝜌\rhoitalic_ρ) in units of 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT s/km.

In Sect. 5, we describe how we combine our estimates of the BNSs’ eccentricities (Fig. 9) and the relationship we find between vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT and e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG (Fig. 10), to determine the posterior kick estimates (Fig. 11). In order to infer vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT based on e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG, this method employs Bayes’ theorem:

P(vkick|e~)=P(e~|vkick)P(vkick)P(e~).𝑃conditionalsubscript𝑣kick~𝑒𝑃conditional~𝑒subscript𝑣kick𝑃subscript𝑣kick𝑃~𝑒P(v_{\text{kick}}\hskip 1.42262pt|\hskip 1.42262pt\tilde{e})=\dfrac{P(\tilde{e% }\hskip 1.42262pt|\hskip 1.42262ptv_{\text{kick}})P(v_{\text{kick}})}{P(\tilde% {e})}\quad.italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT | over~ start_ARG italic_e end_ARG ) = divide start_ARG italic_P ( over~ start_ARG italic_e end_ARG | italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) end_ARG start_ARG italic_P ( over~ start_ARG italic_e end_ARG ) end_ARG . (7)

Since we want to relate this theorem to the results of our simulations as shown in Fig. 10, we define each cell in this figure to have a value equal to N(e~|vkick)𝑁conditional~𝑒subscript𝑣kickN(\tilde{e}\hskip 1.42262pt|\hskip 1.42262ptv_{\text{kick}})italic_N ( over~ start_ARG italic_e end_ARG | italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) and a size equals to Δe~ΔvkickΔ~𝑒Δsubscript𝑣kick\Delta\tilde{e}\cdot\Delta v_{\text{kick}}roman_Δ over~ start_ARG italic_e end_ARG ⋅ roman_Δ italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT, while the amount of objects in each column and row equal N(e~)𝑁~𝑒N(\tilde{e})italic_N ( over~ start_ARG italic_e end_ARG ) and N(vkick)𝑁subscript𝑣kickN(v_{\text{kick}})italic_N ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ), respectively. Using these quantities, we determine that P(e~|vkick)=N(e~|vkick)/N(vkick)𝑃conditional~𝑒subscript𝑣kick𝑁conditional~𝑒subscript𝑣kick𝑁subscript𝑣kickP(\tilde{e}\hskip 1.42262pt|\hskip 1.42262ptv_{\text{kick}})=N(\tilde{e}\hskip 1% .42262pt|\hskip 1.42262ptv_{\text{kick}})/N(v_{\text{kick}})italic_P ( over~ start_ARG italic_e end_ARG | italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) = italic_N ( over~ start_ARG italic_e end_ARG | italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) / italic_N ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ). Then, we define the prior P(vkick)𝑃subscript𝑣kickP(v_{\text{kick}})italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) as proportional to the size of each row (i.e. P(vkick)N(vkick)proportional-to𝑃subscript𝑣kick𝑁subscript𝑣kickP(v_{\text{kick}})\propto N(v_{\text{kick}})italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) ∝ italic_N ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT )), effectively accounting for the fact that the fraction of objects that cross the solar neighbourhood differ for each vkicksubscript𝑣kickv_{\text{kick}}italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT, independent of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG. These two relations allow us to determine P(e~)𝑃~𝑒P(\tilde{e})italic_P ( over~ start_ARG italic_e end_ARG ):

P(e~)=P(e~|vkick)P(vkick)𝑑vkickvkickN(e~|vkick)=N(e~).𝑃~𝑒𝑃conditional~𝑒subscript𝑣kick𝑃subscript𝑣kickdifferential-dsubscript𝑣kickproportional-tosubscriptsubscript𝑣kick𝑁conditional~𝑒subscript𝑣kick𝑁~𝑒P(\tilde{e})=\int P(\tilde{e}\hskip 1.42262pt|\hskip 1.42262ptv_{\text{kick}})% P(v_{\text{kick}})dv_{\text{kick}}\propto\sum_{v_{\text{kick}}}N(\tilde{e}% \hskip 1.42262pt|\hskip 1.42262ptv_{\text{kick}})=N(\tilde{e})\quad.italic_P ( over~ start_ARG italic_e end_ARG ) = ∫ italic_P ( over~ start_ARG italic_e end_ARG | italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) italic_d italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ∝ ∑ start_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_N ( over~ start_ARG italic_e end_ARG | italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) = italic_N ( over~ start_ARG italic_e end_ARG ) . (8)

Combining these probabilities with Eq. 7 gives, then:

P(vkick|e~)N(vkick|e~)N(vkick)N(vkick)N(e~)=N(e~|vkick)N(e~),proportional-to𝑃conditionalsubscript𝑣kick~𝑒𝑁conditionalsubscript𝑣kick~𝑒𝑁subscript𝑣kick𝑁subscript𝑣kick𝑁~𝑒𝑁conditional~𝑒subscript𝑣kick𝑁~𝑒P(v_{\text{kick}}\hskip 1.42262pt|\hskip 1.42262pt\tilde{e})\propto\dfrac{N(v_% {\text{kick}}\hskip 1.42262pt|\hskip 1.42262pt\tilde{e})}{N(v_{\text{kick}})}% \dfrac{N(v_{\text{kick}})}{N(\tilde{e})}=\dfrac{N(\tilde{e}\hskip 1.42262pt|% \hskip 1.42262ptv_{\text{kick}})}{N(\tilde{e})}\quad,italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT | over~ start_ARG italic_e end_ARG ) ∝ divide start_ARG italic_N ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT | over~ start_ARG italic_e end_ARG ) end_ARG start_ARG italic_N ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) end_ARG divide start_ARG italic_N ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) end_ARG start_ARG italic_N ( over~ start_ARG italic_e end_ARG ) end_ARG = divide start_ARG italic_N ( over~ start_ARG italic_e end_ARG | italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT ) end_ARG start_ARG italic_N ( over~ start_ARG italic_e end_ARG ) end_ARG , (9)

which corresponds to Fig. 10 with each column normalised individually. This relation allows us to estimate kicks based on an observed eccentricity. However, we do not have a single value of e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG but a distribution (De~subscript𝐷~𝑒D_{\tilde{e}}italic_D start_POSTSUBSCRIPT over~ start_ARG italic_e end_ARG end_POSTSUBSCRIPT), which is why we combine Eq. 9 with the eccentricity distribution P(e~|De~)𝑃conditional~𝑒subscript𝐷~𝑒P(\tilde{e}\hskip 1.42262pt|\hskip 1.42262ptD_{\tilde{e}})italic_P ( over~ start_ARG italic_e end_ARG | italic_D start_POSTSUBSCRIPT over~ start_ARG italic_e end_ARG end_POSTSUBSCRIPT ), which are displayed in Fig. 9, in order to determine the desired kick estimate posteriors from Fig. 11:

P(vkick|De~)=P(vkick|e~)P(e~|De~)𝑑e~,𝑃conditionalsubscript𝑣kicksubscript𝐷~𝑒𝑃conditionalsubscript𝑣kick~𝑒𝑃conditional~𝑒subscript𝐷~𝑒differential-d~𝑒P(v_{\text{kick}}\hskip 1.42262pt|\hskip 1.42262ptD_{\tilde{e}})=\int P(v_{% \text{kick}}\hskip 1.42262pt|\hskip 1.42262pt\tilde{e})P(\tilde{e}\hskip 1.422% 62pt|\hskip 1.42262ptD_{\tilde{e}})d\tilde{e}\quad,italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT | italic_D start_POSTSUBSCRIPT over~ start_ARG italic_e end_ARG end_POSTSUBSCRIPT ) = ∫ italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT | over~ start_ARG italic_e end_ARG ) italic_P ( over~ start_ARG italic_e end_ARG | italic_D start_POSTSUBSCRIPT over~ start_ARG italic_e end_ARG end_POSTSUBSCRIPT ) italic_d over~ start_ARG italic_e end_ARG , (10)

for which we show the product P(vkick|e~)P(e~|De~)𝑃conditionalsubscript𝑣kick~𝑒𝑃conditional~𝑒subscript𝐷~𝑒P(v_{\text{kick}}\hskip 1.42262pt|\hskip 1.42262pt\tilde{e})P(\tilde{e}\hskip 1% .42262pt|\hskip 1.42262ptD_{\tilde{e}})italic_P ( italic_v start_POSTSUBSCRIPT kick end_POSTSUBSCRIPT | over~ start_ARG italic_e end_ARG ) italic_P ( over~ start_ARG italic_e end_ARG | italic_D start_POSTSUBSCRIPT over~ start_ARG italic_e end_ARG end_POSTSUBSCRIPT ) in Figs. 14 and 15. By collecting normalised columns from Fig. 10, corresponding to the distributions in Fig. 9, integrating the posterior over e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG and normalising the result, we effectively compute the kick posteriors following Eqs. 9 and 10.

Appendix C Fits

Refer to caption
Figure 16: Total kick distribution as determined through e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG (dots, average of the results using the exponential disc prior and the ones using the Gaussian annulus prior, tracing the tops of normalised histograms with bins of 10101010 km/s), together with the fitted log-normal (dashed lines) and Maxwellian (dotted lines) distributions, for GC-isotropy (left panel) and LSR-isotropy (right panel). The fitted parameters of the distributions are given in Table 4. We also show the kick distribution that O’Doherty et al. (2023) find for NSs with low-mass companions (black dash-dotted lines).

In Fig. 11 we show our kinematically constrained kick estimates for the individual BNSs, as well as a total distribution for all BNSs. To the total BNSs kicks, we fit a log-normal distribution:

ρ(x|μ,σ)=1xσ2πexp((ln(x)μ)22σ2),𝜌conditional𝑥𝜇𝜎1𝑥𝜎2𝜋superscript𝑥𝜇22superscript𝜎2\rho(x\hskip 1.42262pt|\hskip 1.42262pt\mu,\sigma)=\dfrac{1}{x\sigma\sqrt{2\pi% }}\exp\left(-\dfrac{\left(\ln(x)-\mu\right)^{2}}{2\sigma^{2}}\right)\quad,italic_ρ ( italic_x | italic_μ , italic_σ ) = divide start_ARG 1 end_ARG start_ARG italic_x italic_σ square-root start_ARG 2 italic_π end_ARG end_ARG roman_exp ( - divide start_ARG ( roman_ln ( start_ARG italic_x end_ARG ) - italic_μ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (11)

and a Maxwellian:

ρ(x|σ)=2πx2σ3exp(x22σ2).𝜌conditional𝑥𝜎2𝜋superscript𝑥2superscript𝜎3superscript𝑥22superscript𝜎2\rho(x\hskip 1.42262pt|\hskip 1.42262pt\sigma)=\sqrt{\dfrac{2}{\pi}}\dfrac{x^{% 2}}{\sigma^{3}}\exp\left(-\dfrac{x^{2}}{2\sigma^{2}}\right)\quad.italic_ρ ( italic_x | italic_σ ) = square-root start_ARG divide start_ARG 2 end_ARG start_ARG italic_π end_ARG end_ARG divide start_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_σ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG roman_exp ( - divide start_ARG italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) . (12)

We use a non-linear least-squares fit and obtain the optimal parameter values given in Table 4. In Fig. 16, we show the fitted distributions, as well as the results from O’Doherty et al. (2023), who estimate the kicks of NSs with low-mass companions, and describe this with a beta distribution. The figure shows that the log-normal distribution seems to fit our results significantly better than the Maxwellians.

Table 4: Parameters of the distributions fitted to our results, for a log-normal distribution (Eq. 11) and a Maxwellian distribution (Eq. 12).
Log-normal GC-isotropy LSR-isotropy
μ𝜇\muitalic_μ 5.36(1)5.3615.36(1)5.36 ( 1 ) 4.57(2)4.5724.57(2)4.57 ( 2 )
σ𝜎\sigmaitalic_σ 0.62(1)0.6210.62(1)0.62 ( 1 ) 0.89(2)0.8920.89(2)0.89 ( 2 )
modea𝑎aitalic_aa𝑎aitalic_aa𝑎aitalic_aEquals exp(μσ2)𝜇superscript𝜎2\exp\left(\mu-\sigma^{2}\right)roman_exp ( italic_μ - italic_σ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). 146(3)1463146(3)146 ( 3 ) 43(2)43243(2)43 ( 2 )
Maxwellian
σ𝜎\sigmaitalic_σ 135(2)1352135(2)135 ( 2 ) 71(3)71371(3)71 ( 3 )
modeb𝑏bitalic_bb𝑏bitalic_bb𝑏bitalic_bEquals 2σ2𝜎\sqrt{2}\sigmasquare-root start_ARG 2 end_ARG italic_σ. 191(3)1913191(3)191 ( 3 ) 100(4)1004100(4)100 ( 4 )
555 Values in brackets are the 68%percent6868\%68 % uncertainty on preceding digits.

Appendix D Peculiar velocities

Refer to caption
Figure 17: Comparison between our kick estimates as constrained through e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG (solid lines) and peculiar velocities (vpecsubscript𝑣pecv_{\text{pec}}italic_v start_POSTSUBSCRIPT pec end_POSTSUBSCRIPT) at the first disc crossing (dashed lines), for GC-isotropy (green lines) and LSR-isotropy (purple lines). For both isotropies, we show the average between the results using the exponential disc prior and the ones using the Gaussian annulus prior (as shown in Fig. 11). The lines trace normalised histograms with bins of 10101010 km/s.

As a comparison between our method and the method proposed by Atri et al. (2019), we compare our kick estimates to the peculiar velocities at the first disc crossings of the BNSs. For each BNS, we take the disc crossings that we used in order to estimate τkinsubscript𝜏kin\tau_{\text{kin}}italic_τ start_POSTSUBSCRIPT kin end_POSTSUBSCRIPT for χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (in Sect. 2.2), and determine the peculiar velocity at that point as a potential kick velocity. Now, Atri et al. (2019) do not limit themselves to the first disc crossing, but trace back the trajectories of the black hole X-ray binaries in their sample for 10101010 Gyr and analyse the disc crossings during this entire period. Nevertheless, we find that even if we limit ourselves to the first disc crossing, our kick estimates match the peculiar velocities at χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT well (as we show in Fig. 17). In particular, the total kick distribution for the BNS sample looks nearly indistinguishable if determined through peculiar velocities at χ1subscript𝜒1\chi_{1}italic_χ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, compared to our estimates through e~~𝑒\tilde{e}over~ start_ARG italic_e end_ARG. The most notable differences in the kick estimates occur for J0737–3039A/B, J0509+3801, B1534+12, and J1930–1852.